Fact-checked by Grok 2 weeks ago

Euclidean plane

The Euclidean plane is a two-dimensional affine space equipped with an inner product on its vector space of translations, enabling the definition of distances, angles, and rigid motions that form the foundation of classical plane geometry. Analytically, it can be modeled as the set \mathbb{R}^2 paired with the Euclidean distance function d(\mathbf{x}, \mathbf{y}) = \|\mathbf{x} - \mathbf{y}\|_2 = \sqrt{(x_1 - y_1)^2 + (x_2 - y_2)^2}, where the norm arises from a positive definite quadratic form, ensuring properties like the Pythagorean theorem hold for orthogonal vectors. This structure satisfies key axioms, including the existence of unique lines between points, congruence of segments and angles, and the parallel postulate, which states that through a point not on a given line, exactly one parallel line can be drawn. In synthetic terms, the Euclidean plane consists of points and lines without reference to coordinates, where geometric figures are defined by incidence, order, , and relations, as formalized in modern axiomatizations like Hilbert's. It distinguishes itself from non-Euclidean planes by the symmetry of and the congruence of all right angles, leading to characteristic theorems such as the sum of angles in a equaling 180 degrees. The plane's isometries—translations, rotations, reflections, and glide reflections—preserve distances and orientations, making it a model for motions in physics and . Historically rooted in Euclid's Elements (circa 300 BCE), the Euclidean plane provides the axiomatic basis for much of , influencing fields from to , while serving as a benchmark for contrasting geometries like or elliptic spaces.

Fundamentals

Definition and axioms

The Euclidean plane is a two-dimensional flat space that satisfies the axioms of and serves as the ambient space for the study of plane geometry, where points, lines, and figures are defined without intrinsic . It can be formalized as the set \mathbb{R}^2 of ordered pairs of real numbers, equipped with the Euclidean metric that measures distances between points. The distance d between two points (x_1, y_1) and (x_2, y_2) in this space is defined by the formula d((x_1, y_1), (x_2, y_2)) = \sqrt{(x_2 - x_1)^2 + (y_2 - y_1)^2}, which induces the standard topology and geometry on the plane. The initial axiomatic foundation for the Euclidean plane was provided by Euclid in his treatise Elements (circa 300 BCE), where five postulates specifically govern plane constructions and relations. These postulates are:
  1. A straight line can be drawn between any two points.
  2. Any terminated straight line can be extended indefinitely.
  3. A circle can be drawn with any given center and radius.
  4. All right angles are equal to each other.
  5. If a straight line intersects two other straight lines such that the sum of the interior angles on one side is less than two right angles, then the two lines, if extended, will meet on that side.
The fifth postulate, known as the parallel postulate, asserts that given a line and a point not on it, exactly one line through the point is parallel to the given line; this ensures the plane's flatness by preventing or elliptic curvatures that arise in non-Euclidean geometries. To resolve ambiguities and gaps in Euclid's system—such as undefined terms like "between" and incomplete assumptions— presented a rigorous axiomatization in his 1899 work Grundlagen der Geometrie (), comprising 20 independent axioms that fully characterize plane . These are grouped as follows:
  • Incidence axioms: These define points and lines as primitive elements, with relations like: two distinct points determine a unique line; every line contains at least two points; there exist three points not all on the same line. They establish the basic combinatorial structure without or measurement.
  • Order axioms (4 axioms): Introducing betweenness (II, 1–4), these specify that for any three collinear points, exactly one lies between the other two; they prevent cycles and ensure linear ordering on lines, foundational for defining segments and rays.
  • Congruence axioms (5 axioms): These define equality of segments and angles (III, 1–5), such as congruence being an equivalence relation for segments (III, 1) and the ability to superimpose congruent figures (III, 4–5); they enable the measurement and comparison central to Euclidean constructions.
  • Parallelism axiom (1 axiom): Stating that through a point not on a line, there exists one and only one parallel line (IV, 1), this is equivalent to Euclid's fifth postulate and ensures the plane's flatness.
  • Continuity axioms (2 axioms): The Archimedean axiom (V, 1) guarantees that the real numbers densely embed into the plane's lengths, while the (V, 2) ensures every bounded nonempty set of points has a least upper bound, providing the full for the plane.
Hilbert's framework proves the consistency and completeness of Euclidean plane geometry relative to the real numbers, resolving foundational issues in Euclid's original postulates.

Basic properties

The Euclidean plane, founded on axioms of incidence, , , parallelism, and , possesses core properties that follow directly from these foundations. Betweenness, defined via the axioms, establishes a on points along any line: for distinct points A and B, any third point C on the line through A and B is either between A and B (lying in the open segment they determine), or A between B and C, or B between A and C. This relation ensures the plane's lines are Dedekind-complete chains, preventing gaps in linear . Congruence of segments equates those of equal length, while equates those superimposable by rigid motion; these are axiomatized such that if two segments are congruent and adjacent segments from their endpoints are congruent, the included angles are congruent (SAS criterion). The existence of midpoints follows from these axioms: for any segment AB, there exists a point M such that AM \cong MB, constructed via the circle axiom and properties. A hallmark plane-specific consequence is the , which holds due to the parallel postulate. In a with legs of lengths a and b and hypotenuse c, a^2 + b^2 = c^2, proved by dropping an altitude to the hypotenuse, creating similar triangles whose ratios yield the relation through area equivalence or similarity arguments. This theorem fails in non-Euclidean geometries without the parallelism axiom. The isometries of the Euclidean plane—distance-preserving transformations—comprise translations (rigid shifts along a ), rotations (about a fixed point by an ), reflections (over a line), and glide reflections (reflection followed by translation parallel to the line). Every non-identity isometry is exactly one of these types, and they generate the E(2), a of the translation group \mathbb{R}^2 and the O(2) of rotations and reflections, with order 2 elements being orientation-reversing (reflections and glides) and others orientation-preserving. Underlying these structures, the Euclidean plane is coordinatized by the real numbers \mathbb{R}, an Archimedean ordered field where, for any positive x, y \in \mathbb{R}, there exists a natural number n such that nx > y. Completeness arises from the construction: every partition of \mathbb{R} into non-empty lower and upper sets with no greatest lower element defines a unique as the cut's supremum, ensuring every bounded non-empty subset has a least upper bound, which underpins the plane's . Topologically, the Euclidean plane \mathbb{R}^2 is simply connected, as every closed curve can be continuously shrunk to a point without leaving the space, reflecting its contractibility. It exhibits zero Gaussian curvature everywhere, enabling the parallel postulate—through a point not on a line, exactly one parallel exists—unlike positively curved elliptic planes (where parallels intersect) or negatively curved hyperbolic planes (with multiple parallels).

Historical development

Ancient origins

The concepts underlying the Euclidean plane emerged from practical needs in ancient civilizations, particularly for land measurement and . In around 2000 BCE, geometry developed empirically to survey fields after floods, using tools like the for alignment and ropes for right angles. The , dating to circa 1650 BCE, contains problems on calculating areas of triangles and circles, reflecting these surveying techniques without formal proofs. Similarly, in around the same period, cuneiform tablets demonstrate knowledge of Pythagorean triples and approximations for circular areas, applied to and astronomy. Greek thinkers in the 6th century BCE built on these influences, shifting toward . (c. 624–546 BCE), often credited as the founder of Greek geometry, introduced theorems such as the equality of base angles in isosceles triangles and the intercept theorem for , likely inspired by Egyptian and Babylonian practices during his travels. His work emphasized proofs, marking a transition from empirical to logical methods. The Pythagorean school, founded by (c. 570–495 BCE) in Croton, , advanced studies of right triangles, formalizing the relationship now known as the , which states that in a right-angled triangle, the square of the hypotenuse equals the sum of the squares of the other two sides. This school integrated geometry with philosophy, viewing numbers and shapes as fundamental to the cosmos. Euclid of Alexandria (c. 300 BCE) synthesized these developments in his seminal work, Elements, providing the first comprehensive axiomatic treatment of plane geometry in Books I–IV. Book I covers basic constructions like equilateral triangles and ; Book II addresses ; Book III deals with circles and inscribed angles; and Book IV constructs regular polygons. This compilation drew from prior Greek sources while establishing a rigorous framework that influenced for over two millennia. The preservation and extension of these ideas occurred in the during the 8th–9th centuries . (c. 780–850 ), working in Baghdad's , authored The Compendious Book on Calculation by Completion and Balancing, incorporating geometric methods inspired by into algebraic solutions for plane figures and applying them to inheritance and problems. His works, along with translations of texts by other scholars, bridged ancient traditions, ensuring the transmission of Euclidean plane concepts to medieval .

Modern axiomatization

In the 19th century, mathematicians identified several gaps in Euclid's ancient for plane geometry, including the lack of explicit axioms for (betweenness) on lines, incidence relations between points and lines, and to guarantee the existence of intersections in constructions such as drawing circles. These deficiencies allowed for ambiguities in proofs and potential inconsistencies, prompting efforts to develop rigorous, complete, and independent sets of axioms that could derive all Euclidean theorems without hidden assumptions. A key early contribution came from Moritz Pasch in 1882, who introduced axioms for betweenness to formalize the ordering of points on a line and the separation properties of lines in the . Pasch's axiom states that if a line intersects one side of a , it must also intersect exactly one of the other two sides, ensuring that lines divide the into distinct half-planes and preventing pathological configurations absent in . This addressed Euclid's implicit reliance on intuitive notions of "inside" and "outside" without proof, providing a foundation for ordered that influenced subsequent axiomatizations. Building on such work, Mario Pieri advanced the axiomatization of incidence in 1895, proposing a system for that treated points and lines as primitive concepts with minimal assumptions about their mutual relations. Pieri's axioms emphasized the independence of incidence from metric properties, allowing Euclidean plane geometry to be derived as a special case by adding order and conditions, thus clarifying the foundational structure detached from coordinate or measurement assumptions. David Hilbert's seminal 1899 monograph Grundlagen der Geometrie provided a comprehensive modern foundation by organizing s into five groups: incidence, order (incorporating Pasch's betweenness), congruence, parallelism, and . Hilbert demonstrated the independence of each by constructing models where specific ones fail while others hold, and ensured through Archimedean and axioms that model the real numbers, resolving Euclid's gaps such as the required for constructions to intersect properly. For instance, his axioms guarantee that any line through a point inside a intersects the at two points, enabling rigorous proofs of existence in geometric constructions. In 1932, Garrett Birkhoff proposed a streamlined metric-based axiomatization with just four postulates, integrating real numbers directly to define distance and angle measures via ruler and protractor operations. Birkhoff's system posits points as pairs of real coordinates, lines by equations, a positive definite distance function, and a congruence axiom for angles modulo $2\pi, while assuming the ruler-compass constructibility aligns with real arithmetic, thus deriving Euclidean properties from analytic foundations without separate incidence or order primitives. These developments profoundly influenced the foundations of by establishing as a amenable to logical , with Hilbert's framework linking to through models over the reals and to via categoricity and completeness results. For the Euclidean plane, this axiomatization confirms unique up-to-isometry realization in \mathbb{R}^2, underpinning applications in and while highlighting 's dependence on the .

Coordinate systems

Cartesian coordinates

The provides an algebraic framework for representing points and geometric objects in the Euclidean plane, transforming geometric problems into equations solvable by algebraic methods. Invented by in his 1637 treatise , this system marked a pivotal advancement by systematically linking algebraic notation to geometric constructions, enabling the resolution of complex figures through coordinate assignments rather than purely synthetic proofs. In , Descartes assigned coordinates to points using intersecting lines as references, laying the groundwork for and influencing subsequent mathematical developments. In this system, the Euclidean plane is modeled as the set \mathbb{R}^2, where each point is uniquely specified by an of real numbers (x, y), corresponding to distances along two axes: the x- and vertical y-, intersecting at the (0, 0). The axes form a right-handed orthogonal frame, with positive directions extending rightward and upward from the origin, allowing precise location of any point via signed distances from these references. lines in the plane are represented by linear equations of the form ax + by + c = 0, where a, b, and c are real constants (with a and b not both zero), encapsulating all points satisfying the relation. Common transformations within the Cartesian system include translations and scalings, which preserve the plane's structure while shifting or resizing coordinates. A by (h, k) maps a point (x, y) to (x + h, y + k), effectively relocating the without altering orientations or relative distances. by factors s_x and s_y transforms (x, y) to (s_x x, s_y y), stretching or compressing along the axes, though non-uniform scalings may distort angles unless s_x = s_y. Curves beyond lines are often parameterized as x = f(t), y = g(t), where t varies over an interval, tracing paths like parabolas or ellipses through functional dependence on a single parameter. The primary advantage of Cartesian coordinates lies in converting geometric inquiries into algebraic manipulations, such as finding line intersections by solving simultaneous linear equations, which yields exact coordinates without ruler-and-compass constructions. For instance, the intersection of lines a_1 x + b_1 y + c_1 = 0 and a_2 x + b_2 y + c_2 = 0 is determined via or , providing a systematic algebraic to what was previously a synthetic challenge. This approach not only simplifies computations but also extends to higher-degree equations for conics and beyond, underpinning much of modern geometry and its applications.

Polar and other systems

In the Euclidean plane, polar coordinates provide an alternative representation to Cartesian coordinates, specifying each point by a radial r \geq 0 from a fixed called the and an angle \theta measured counterclockwise from the positive x-axis. This system leverages , making it particularly useful for problems involving circles, rotations, or radial patterns. The angle \theta is typically expressed in radians and can take any real value, though representations are not unique since adding $2\pi k to \theta (for k) or negating r with \theta + \pi yields equivalent points. Conversions between polar and Cartesian coordinates are given by the equations x = r \cos \theta, \quad y = r \sin \theta and inversely, r = \sqrt{x^2 + y^2}, \quad \theta = \atantwo(y, x), where \atantwo accounts for the correct quadrant. These relations derive from the unit circle definitions of sine and cosine. In polar form, geometric objects often simplify: a circle centered at the origin has equation r = a for radius a > 0, while the Archimedean spiral, which winds outward at a constant rate, follows r = a \theta for \theta \geq 0 and scaling constant a > 0. Other coordinate systems build on or extend polar coordinates for specific purposes. Cylindrical coordinates (r, \theta, z) generalize polar to three dimensions, but restricted to the plane (where z = 0), they coincide exactly with polar coordinates, emphasizing the radial and angular components without height. , used in , represent points in the Euclidean plane as triples (x : y : w) with w \neq 0, where finite points correspond to (x/w, y/w) in affine (; scaling the triple leaves the point unchanged, allowing compact handling of lines, intersections, and points at infinity. Polar coordinates prove advantageous in applications exploiting , such as describing : a counterclockwise by \phi transforms (r, \theta) simply to (r, \theta + \phi), preserving radial . This facilitates analysis in contexts like or periodic phenomena. Additionally, polar forms simplify certain operations, such as integrating over circular regions, by aligning with the natural geometry and reducing through .

Geometric structures

Lines, angles, and distances

In the Euclidean plane, the distance between two points P_1(x_1, y_1) and P_2(x_2, y_2) is given by the [formula d](/page/Formula_D) = \sqrt{(x_2 - x_1)^2 + (y_2 - y_1)^2}, which follows directly from the applied to the formed by the horizontal and vertical segments connecting the points. This defines the straight-line and underpins the plane's uniformity. A line in the Euclidean plane can be characterized by its m = \frac{y_2 - y_1}{x_2 - x_1}, where (x_1, y_1) and (x_2, y_2) lie on the line, provided the line is not vertical. Two lines with slopes m_1 and m_2 are if m_1 \cdot m_2 = -1, a condition arising from the fact that the angles they form with the differ by 90 degrees. The acute \phi between two non- lines satisfies \tan \phi = \left| \frac{m_2 - m_1}{1 + m_1 m_2} \right|, derived from the subtraction formula using the slopes as tangents of inclination angles. Angles in the Euclidean plane are measured in degrees or radians, with one full rotation corresponding to 360 degrees or $2\pi radians. In any triangle, the sum of the interior angles equals \pi radians (or 180 degrees), a consequence of the exterior angle theorem, which states that an exterior angle equals the sum of the two remote interior angles. Triangle congruence criteria include side-angle-side (SAS), where two sides and the included angle determine congruence, and angle-side-angle (ASA), where two angles and the included side suffice. The of a joining points (x_1, y_1) and (x_2, y_2) is \left( \frac{x_1 + x_2}{2}, \frac{y_1 + y_2}{2} \right), which bisects the segment into two equal parts. The midpoint theorem states that the segment joining the midpoints of two sides of a is to the third side and half its , provable using properties of similar triangles or averages in the plane.

Circles and conics

In the Euclidean plane, a is defined as the locus of all points equidistant from a fixed point called , with this denoted as the r. In Cartesian coordinates, with the center at (h, k), the equation of the is given by (x - h)^2 + (y - k)^2 = r^2. This equation represents the set of points (x, y) satisfying the constant r from , derived from the distance formula in the plane. The circumference of the , or the of its boundary, is $2\pi r, while its area is \pi r^2; these measures follow from integrating the and applying the , though they can also be established geometrically using limits of polygonal approximations in Euclidean constructions. Key properties of the circle include the perpendicularity of the line to the at the point of tangency, ensuring that any line touching the circle at exactly one point forms a with the line from the center to that point. Additionally, an subtending a given measures half the subtending the same , a relation that holds because angles in the same segment are equal and the angle in a is a . The power of a point further characterizes intersections: for a point P outside the circle, if two secants from P intersect the circle at points A, B and C, D respectively, then PA \cdot PB = PC \cdot PD, reflecting an invariant product of segment lengths. Conic sections encompass a family of curves in the , including ellipses, parabolas, and , obtained as intersections of a with a right circular ; the is a special limiting case. Their general in Cartesian coordinates is ax^2 + bxy + cy^2 + dx + ey + f = 0, where the coefficients determine the type and orientation of the curve, with the b^2 - 4ac distinguishing ellipses (<0), parabolas (=0), and hyperbolas (>0) for non-degenerate cases. Conics are classified by their e, a dimensionless measuring deviation from circularity: e = 0 for a , $0 < e < 1 for an ellipse, e = 1 for a parabola, and e > 1 for a . A unifying focus-directrix definition characterizes each conic as the locus of points P such that the ratio of the distance from P to a fixed point (focus F) to the distance from P to a fixed line (directrix l) equals the constant eccentricity e; this property, originally explored by , provides an intrinsic geometric description independent of the cone intersection. For the parabola (e=1), this yields equal distances to focus and directrix, leading to a notable reflective property: incoming rays parallel to the axis of symmetry reflect off the curve and pass through the focus, as the tangent at any point bisects the angle between the axis-parallel ray and the line to the focus. Similar reflective behaviors hold for ellipses and hyperbolas, where rays from one focus reflect toward the other focus.

Polygons and polytopes

A in the Euclidean plane is a closed figure bounded by a finite chain of line segments connected end-to-end, forming an n-sided shape for n ≥ 3. A has all sides of equal length and all interior angles equal, with each interior angle measuring \frac{(n-2)\pi}{n} radians. Any simple with n vertices admits a , partitioning its interior into n-2 non-overlapping triangles using n-3 non-intersecting diagonals, and for the resulting (including the outer face), holds: V - E + F = 2, where V is the number of vertices, E the number of edges, and F the number of faces. Convex polygons are those where all interior angles are less than \pi radians and any connecting two points inside the lies entirely within it. The of any collection of sets, including polygons, is itself , preserving this under operations. In two dimensions, every point of a has a supporting line (the analog of a ) that touches the at that point and leaves the entire on one side of the line, as guaranteed by the supporting hyperplane theorem for sets. In the context of polytopes, a 2-polytope is precisely a in the Euclidean plane, defined as the of finitely many points in \mathbb{R}^2 or the bounded intersection of finitely many half-planes. Non-convex polygons include those with interior angles exceeding \pi radians or self-intersecting edges, such as star polygons like the , a regular five-pointed figure denoted by the {5/2}, formed by connecting every second vertex of a regular pentagon. These self-intersections create intersecting line segments within the boundary. Regarding tilings, only three types of regular s—equilateral triangles, squares, and regular hexagons—can tile the Euclidean plane without gaps or overlaps in a monohedral fashion, due to the requirement that interior angles sum to $2\pi radians at each vertex.

Vector and algebraic aspects

Vectors and operations

In the Euclidean plane, vectors can be conceptualized as directed line segments, or arrows, originating from a point, or more formally as ordered pairs (a, b) where a, b \in \mathbb{R}, representing elements of the \mathbb{R}^2. Position vectors specifically denote those arrows starting from the , providing a coordinate-based representation of in the plane. Vector addition follows the : for vectors \mathbf{u} = (u_1, u_2) and \mathbf{v} = (v_1, v_2), the sum is \mathbf{u} + \mathbf{v} = (u_1 + v_1, u_2 + v_2), geometrically obtained as the diagonal of the parallelogram formed by \mathbf{u} and \mathbf{v} sharing a common initial point. by a k scales the vector: k\mathbf{u} = (k u_1, k u_2), which stretches or compresses the arrow while preserving direction if k > 0, or reversing it if k < 0. These operations enable linear combinations, such as \alpha \mathbf{u} + \beta \mathbf{v} for scalars \alpha, \beta \in \mathbb{R}, and the span of a set of vectors is the collection of all such combinations, forming a of \mathbb{R}^2./04%3A_R/4.10%3A_Spanning_Linear_Independence_and_Basis_in_R) The vectors \mathbf{i} = (1, [0](/page/0)) and \mathbf{j} = ([0](/page/0), 1) provide a fundamental framework for \mathbb{R}^2, as any (a, [b](/page/List_of_punk_rap_artists)) can be uniquely expressed as a\mathbf{i} + [b](/page/List_of_punk_rap_artists)\mathbf{j}. These vectors are linearly , meaning the equation \alpha \mathbf{i} + \beta \mathbf{j} = ([0](/page/0), [0](/page/0)) holds only for \alpha = \beta = [0](/page/0), ensuring they form a basis that spans the entire without redundancy./04%3A_R/4.10%3A_Spanning_Linear_Independence_and_Basis_in_R) Affine combinations extend linear combinations by restricting the scalars \lambda_1, \lambda_2, \dots, \lambda_n such that \sum \lambda_i = 1, allowing points in the to be expressed as weighted averages of other points without shifting the . In particular, barycentric coordinates arise from affine combinations of three non-collinear points forming a , where the coefficients represent areal weights summing to 1, facilitating the location of any interior point as a .

Dot product, norms, and angles

In the Euclidean plane, the provides a fundamental way to measure the interaction between two vectors, imposing a structure that allows for the computation of lengths and angles. For vectors \mathbf{u} = (u_1, u_2) and \mathbf{v} = (v_1, v_2), the is defined algebraically as \mathbf{u} \cdot \mathbf{v} = u_1 v_1 + u_2 v_2. This operation, introduced in the context of modern vector analysis, is bilinear, meaning it is linear in each argument separately: (\alpha \mathbf{u} + \beta \mathbf{w}) \cdot \mathbf{v} = \alpha (\mathbf{u} \cdot \mathbf{v}) + \beta (\mathbf{w} \cdot \mathbf{v}) and similarly for the second argument, and symmetric such that \mathbf{u} \cdot \mathbf{v} = \mathbf{v} \cdot \mathbf{u}. Geometrically, the equals the product of the vectors' magnitudes times the cosine of \theta between them: \mathbf{u} \cdot \mathbf{v} = \|\mathbf{u}\| \|\mathbf{v}\| \cos \theta. These properties make the the standard inner product on the Euclidean plane, enabling projections and other geometric interpretations. The , or , of a \mathbf{u} is derived from the as \|\mathbf{u}\| = \sqrt{\mathbf{u} \cdot \mathbf{u}}. This Euclidean norm satisfies the properties of a norm, including positivity (\|\mathbf{u}\| \geq 0 with equality only for the zero ), scalability (\|\alpha \mathbf{u}\| = |\alpha| \|\mathbf{u}\|), and the . A , or vector of 1, has \|\mathbf{u}\| = 1; any nonzero can be normalized to a by \hat{\mathbf{u}} = \mathbf{u} / \|\mathbf{u}\|, which preserves direction while standardizing magnitude for applications like basis constructions. The angle \theta between two nonzero vectors \mathbf{u} and \mathbf{v} is determined via the by \cos \theta = \frac{\mathbf{u} \cdot \mathbf{v}}{\|\mathbf{u}\| \|\mathbf{v}\|}, where \theta ranges from 0 to \pi radians. Vectors are orthogonal if \mathbf{u} \cdot \mathbf{v} = 0, corresponding to \theta = \pi/2 and \cos \theta = 0; this condition defines perpendicularity in the Euclidean sense, crucial for coordinate systems and decompositions. A key inequality arising from the is the Cauchy-Schwarz inequality, which states that |\mathbf{u} \cdot \mathbf{v}| \leq \|\mathbf{u}\| \|\mathbf{v}\|, with equality if and only if \mathbf{u} and \mathbf{v} are linearly dependent ( or anti-parallel). This bound, first proved for sums in the context of , limits how aligned vectors can be and underpins many theorems in and .

Analytic and calculus applications

Functions and gradients

In the Euclidean plane, a scalar f: \mathbb{R}^2 \to \mathbb{R} assigns a real value to each point (x, y), enabling the representation of quantities such as height, temperature, or across the plane. The level sets of f are the curves where f(x, y) = c for a constant c, forming contours that illustrate regions of equal value and revealing the function's topological structure, such as hills and valleys in a . These contours are particularly useful for visualizing how the function varies spatially, with denser spacing indicating steeper changes. Partial derivatives measure the rate of change of f along the coordinate axes: the partial derivative with respect to x, denoted \partial f / \partial x, is the \lim_{h \to 0} [f(x+h, y) - f(x, y)] / h, treating y as constant, while \partial f / \partial y is defined analogously by varying y. These quantities form the components of the vector \nabla f = (\partial f / \partial x, \partial f / \partial y), which points in the direction of the function's steepest increase at a point and whose magnitude |\nabla f| quantifies that rate. Notably, \nabla f is perpendicular to the level sets of f, as the of \nabla f with any to a level curve is zero. The directional derivative of f in the direction of a unit vector \mathbf{u} = (u_1, u_2) is given by D_{\mathbf{u}} f = \nabla f \cdot \mathbf{u}, representing the instantaneous rate of change along that direction; this leverages the to project the onto \mathbf{u}. The maximum directional derivative occurs when \mathbf{u} aligns with \nabla f, corresponding to the steepest ascent, with value |\nabla f|, while the minimum (steepest descent) is -|\nabla f| in the opposite direction. A classic example is the distance function to a fixed point (a, b), f(x, y) = \sqrt{(x - a)^2 + (y - b)^2}, whose is \nabla f = \frac{(x - a, y - b)}{f(x, y)}, the unit from (a, b) to (x, y), illustrating how the normalizes for radial increase. Another key case involves functions, which satisfy \Delta f = \partial^2 f / \partial x^2 + \partial^2 f / \partial y^2 = 0, where the Laplacian \Delta f is the of \nabla f; these functions, such as the real or imaginary parts of holomorphic functions, model steady-state phenomena like electrostatic potentials in the plane.

Integrals and theorems

In the Euclidean plane, line integrals provide a means to compute quantities along curves, essential for applications in physics such as work done by field. For a scalar f(x, y) and a smooth C parametrized by \mathbf{r}(t) = (x(t), y(t)) for t \in [a, b], the \int_C f \, ds is defined as \int_a^b f(x(t), y(t)) \|\mathbf{r}'(t)\| \, dt, where \|\mathbf{r}'(t)\| is the speed of the parametrization, representing the integral of f with respect to . For a vector field \mathbf{F}(x, y) = (P(x, y), Q(x, y)), the \int_C \mathbf{F} \cdot d\mathbf{r} is given by \int_a^b \mathbf{F}(\mathbf{r}(t)) \cdot \mathbf{r}'(t) \, dt = \int_a^b [P(x(t), y(t)) x'(t) + Q(x(t), y(t)) y'(t)] \, dt, which measures the circulation or flux along the path. Double integrals extend this to areas, allowing computation of volumes under surfaces or masses of regions in the plane. Over a bounded D in \mathbb{R}^2, the double \iint_D f(x, y) \, dA represents the signed beneath the of f, evaluated via iterated integrals such as \int_a^b \int_{g(x)}^{h(x)} f(x, y) \, dy \, dx for type I regions. To simplify evaluation over irregular D, a x = x(u, v), y = y(u, v) transforms the to \iint_{D'} f(x(u, v), y(u, v)) |J| \, du \, dv, where D' is the image and J = \det \begin{pmatrix} \frac{\partial x}{\partial u} & \frac{\partial x}{\partial v} \\ \frac{\partial y}{\partial u} & \frac{\partial y}{\partial v} \end{pmatrix} is the of the , accounting for the scaling of area elements under the transformation. The fundamental theorem of line integrals connects path integrals to scalar potentials for conservative fields. If \mathbf{F} = \nabla f is the gradient of a scalar potential f (i.e., conservative, with \frac{\partial P}{\partial y} = \frac{\partial Q}{\partial x}), then for any piecewise smooth curve C from point P to Q, \int_C \mathbf{F} \cdot d\mathbf{r} = f(Q) - f(P), independent of the path taken, provided the domain is simply connected. This theorem generalizes the one-dimensional fundamental theorem of calculus to the plane, enabling efficient computation without explicit parametrization. Green's theorem relates line integrals around closed curves to double integrals over enclosed regions, a cornerstone for planar vector analysis. For a positively oriented, piecewise smooth, simple closed curve C bounding region D, with continuously differentiable P and Q, \oint_C (P \, dx + Q \, dy) = \iint_D \left( \frac{\partial Q}{\partial x} - \frac{\partial P}{\partial y} \right) dA, this equates circulation to the integral of the , facilitating conversions between boundary and interior computations. Originally derived in George Green's 1828 essay on electricity and magnetism, the theorem applies to diverse fields like . A key application computes areas: for a closed curve C, the area of D is \frac{1}{2} \oint_C (x \, dy - y \, dx) = \iint_D dA, using P = -y, Q = x.

Topological and graph-theoretic views

Topological properties

The Euclidean plane, denoted \mathbb{R}^2, is a of dimension 2. It is Hausdorff, meaning that for any two distinct points, there exist disjoint open neighborhoods separating them. Additionally, \mathbb{R}^2 is second-countable, possessing a countable basis for its , which ensures that the space is separable and allows for manageable coverings in proofs of topological properties. As a locally Euclidean space, every point in \mathbb{R}^2 has an open neighborhood homeomorphic to an open subset of \mathbb{R}^2 itself, typically via coordinate charts. The Euclidean plane is path-connected, as any two points can be joined by a continuous path, such as a straight line segment. It is also simply connected, meaning it is path-connected and has a trivial fundamental group, \pi_1(\mathbb{R}^2) = \{e\}, where loops based at any point can be continuously contracted to a point within the space. This triviality of the fundamental group distinguishes \mathbb{R}^2 from spaces with "holes," such as the punctured plane \mathbb{R}^2 \setminus \{0\}, whose fundamental group is isomorphic to \mathbb{Z}, generated by loops winding around the origin. A key consequence is the Jordan curve theorem, which states that every simple closed curve in \mathbb{R}^2 separates the plane into two distinct connected components: a bounded interior region and an unbounded exterior region, with the curve itself forming the boundary between them. Subsets of the Euclidean plane exhibit compactness properties characterized by the Heine-Borel : a of \mathbb{R}^2 is compact it is closed and bounded. This equivalence holds because \mathbb{R}^2 is a where closed sets contain all points and bounded sets fit within a ball of finite radius, ensuring that every open cover of a closed and bounded set has a finite subcover. The topology of the Euclidean plane is preserved under homeomorphisms, which are continuous bijections with continuous inverses. Such maps maintain the plane's key invariants, including its simply connectedness and lack of holes, as evidenced by the non-homeomorphism between \mathbb{R}^2 and the punctured plane, due to differing fundamental groups. This invariance underscores the plane's topological uniqueness among 2-dimensional manifolds without punctures or boundaries.

Planar graphs and embeddings

A is a graph that can be embedded in the Euclidean plane such that no two edges cross except possibly at vertices. This embedding divides the plane into faces, including an unbounded outer face. For a connected with V vertices, E edges, and F faces (counting the outer face), states that V - E + F = 2. This relation holds for any maximal planar embedding, where adding any edge would require a crossing or violate planarity. Embeddings of s can use curved edges or straight-line segments. Fáry's theorem asserts that every simple admits a straight-line in the without crossings, preserving the combinatorial structure of a given planar . Such straight-line drawings position vertices at distinct points in the and connect them with line segments that do not intersect except at endpoints. A subclass of s is the s, which can be embedded such that all vertices lie on the boundary of the outer face, with internal edges not crossing. s satisfy a stricter bound from : for a connected with V \geq 2, E \leq 2V - 3. Kuratowski's theorem provides a characterization of planarity: a finite graph is planar if and only if it contains no subdivision of the complete graph K_5 (five vertices all connected) or the complete bipartite graph K_{3,3} (two sets of three vertices, each connected to all in the other set). A subdivision replaces edges with paths, preserving the graph's topological structure. The graph K_{3,3} exemplifies non-planarity and arises in the utility graph problem, where three houses and three utilities cannot be connected pairwise without crossings in the plane. Similarly, K_5 cannot be embedded without crossings. Planar graphs have significant applications, notably in . The states that any is 4-colorable, meaning its vertices can be colored with at most four colors such that no adjacent vertices share the same color; this was proved in 1976 using a computer-assisted discharge method on unavoidable configurations. This theorem directly implies that four colors suffice to color any in the so that adjacent regions differ in color.

References

  1. [1]
    Univ at Albany: William F. Hammond: Math 331
    Definition. A synthetic n-dimensional Euclidean space is an affine space for which the n-dimensional vector space of translations has a given inner product. In ...
  2. [2]
    [PDF] Plane Geometry from VECTOR NORMS AND THE Pythagorean ...
    Aug 29, 2025 · Euclidean planes are traditionally defined by a set of axioms. A strongly normed plane satisfies the Euclidean axioms if and only if the its ...
  3. [3]
    Euclidean geometry | Definition, Axioms, & Postulates - Britannica
    Euclidean geometry is the study of plane and solid figures on the basis of axioms and theorems employed by the ancient Greek mathematician Euclid.Missing: authoritative | Show results with:authoritative
  4. [4]
    Euclidean Metric -- from Wolfram MathWorld
    The Euclidean metric is the function d:R^n×R^n->R that assigns to any two vectors in Euclidean n-space x=(x_1,...,x_n) and y=(y_1,...,y_n) the number d(x ...
  5. [5]
    [PDF] Euclid's Elements of Geometry - Richard Fitzpatrick
    The aim of the translation is to make the mathematical argument as clear and unambiguous as possible, whilst still adhering closely to the meaning of the ...Missing: primary source
  6. [6]
    Euclid's Postulates -- from Wolfram MathWorld
    1. A straight line segment can be drawn joining any two points. · 2. Any straight line segment can be extended indefinitely in a straight line. · 3. Given any ...Missing: plane | Show results with:plane
  7. [7]
    Parallel postulate | Euclidean, Non-Euclidean, Axiom | Britannica
    It states that through any given point not on a line there passes exactly one line parallel to that line in the same plane.
  8. [8]
    [PDF] Foundations of Geometry - UC Berkeley math
    As a basis for the analysis of our intuition of space, Professor Hilbert commences his discus- sion by considering three systems of things which he calls points ...
  9. [9]
    [PDF] Universal Foundations
    Hilbert uses five groups of axioms to form the foundation of Eu- clidean geometry: axioms of incidence, betweenness, congruence, conti- nuity, and parallelism.
  10. [10]
    [PDF] Axioms and Basic Results in Plane Geometry
    Hilbert's Parallel Axiom: For every line ` and every point P not on ` there is at most one line m through P and parallel to `. Prop 2.1: If ` and m are ...
  11. [11]
    [PDF] PRACTICE MIDTERM MATH 130 1. Show that on a Hilbert plane ...
    Show that on a Hilbert plane every segment has a midpoint. 2. Let Φ be a rigid motion of a Hilbert plane such that Φ2 is the identical map but Φ is not.
  12. [12]
    [PDF] Euclidean Geometry
    Theorem 2.3 (SSS) If the corresponding sides of two triangles 4ABC and 4DEF have equal lengths, then the two triangles are congruent.
  13. [13]
    [PDF] isometries of the plane - UChicago Math
    Jul 21, 2009 · Every isometry of the plane, other than the identity, is either a translation, a rotation, a reflection, or a glide-reflection. Proof. Let α be ...
  14. [14]
    What are the 'real numbers,' really?
    Dedekind completeness turns out to be crucial in analysis, because it enables us to take limits. Some ordered fields are Dedekind complete, and some aren't.
  15. [15]
    [PDF] Lecture Notes in Modern Geometry 1 The euclidean plane
    The Euclidean plane, as defined by Euclid, uses points, lines, and circles. Modern geometry uses sets and numbers, and the Euclidean distance is defined as p(x ...
  16. [16]
    [PDF] Surveying in Ancient Egypt
    Apr 21, 2005 · The Egyptians also had measures of area, the most common one being the or , (setat) which was one square khet, or 10,000 square cubits. This ...
  17. [17]
    Diagrams in ancient Egyptian geometry: Survey and assessment
    By far the most important of our hieratic sources is the Rhind Mathematical Papyrus (RMP). This treatise was compiled by a scribe named Ahmose in the 33rd year ...
  18. [18]
    [PDF] Chapter 1 - The Origins of Geometry - Mathematics
    This began with Thales of Miletus (624–547 BC). He was familiar with the computations, right or wrong, handed down from Egyptian and Babylonian mathematics.
  19. [19]
    Thales of Miletus (624 BC - 547 BC) - Biography - MacTutor
    In many textbooks on the history of mathematics Thales is credited with five theorems of elementary geometry:- A circle is bisected by any diameter. The ...
  20. [20]
    Thales of Miletus | Internet Encyclopedia of Philosophy
    Thales of Miletus (c. 620 B.C.E.—c. 546 B.C.E.). thales The ancient Greek philosopher Thales was born in Miletus in Greek Ionia. Aristotle, the major source ...The Writings of Thales · Thales's Astronomy · The Possible Travels of Thales
  21. [21]
    Pythagoras - Biography - MacTutor - University of St Andrews
    The theorem now known as Pythagoras's theorem was known to the Babylonians 1000 years earlier but he may have been the first to prove it. Thumbnail of ...
  22. [22]
    Pythagoreanism - Stanford Encyclopedia of Philosophy
    Mar 29, 2006 · Pythagoreanism is the philosophy of the ancient Greek philosopher Pythagoras (ca. 570–ca. 490 BCE), which prescribed a highly structured way of life.
  23. [23]
    Euclid's Elements, Introduction - Clark University
    This dynamically illustrated edition of Euclid's Elements includes 13 books on plane geometry, geometric and abstract algebra, number theory, ...Missing: plane | Show results with:plane
  24. [24]
    Al-Khwarizmi (790 - 850) - Biography - MacTutor
    He composed the oldest works on arithmetic and algebra. They were the principal source of mathematical knowledge for centuries to come in the East and the West.Missing: preservation | Show results with:preservation
  25. [25]
    Contribution of Al-Khwarizmi to Mathematics and Geography
    Dec 27, 2006 · He synchronized Greek and Indian mathematical knowledge but was the first mathematician to distinguish clearly between algebra and geometry and ...Missing: preservation | Show results with:preservation
  26. [26]
    [PDF] Formalization of Geometry - Michael Beeson's
    May 25, 2019 · ▷ Pasch's axiom, introduced in 1882: a line that meets one side of ... axioms, use strict betweenness and put back the original betweenness axioms ...
  27. [27]
    [PDF] HILBERT'S AXIOMS
    His construction tacitly assumes that if a line passes through a point inside a circle, then the line intersects the circle in two points-an assumption you can ...
  28. [28]
    [PDF] Axiomatic Systems for Geometry - University of Illinois
    Birkhoff's four postulates for Euclidean geometry appear compact enough for us not to lose our way. We interpret the points A, B, C... as ordered pairs, (x, y), ...
  29. [29]
    [PDF] BURTON DREBEN and AKIHIRO KANAMORI - HILBERT AND SET ...
    Nov 9, 1993 · However, for Gödel the axiom would be considered both the basis of com prehension axioms in set theory as well as the antecedent to his argument.
  30. [30]
    Descartes' Mathematics - Stanford Encyclopedia of Philosophy
    Nov 28, 2011 · In La Géométrie, Descartes details a groundbreaking program for geometrical problem-solving—what he refers to as a “geometrical calculus” ( ...
  31. [31]
    Decartes - Berkeley Learning Platform
    His treatise, La Géométrie, published in 1637, introduced the notion of coordinates and the use of algebra to solve geometric problems. This innovative ...
  32. [32]
    CARTESIAN COORDINATE SYSTEM - AE Resources
    A Cartesian coordinate system specifies a point on a plane by a pair of numerical coordinates that are fixed in an orthogonal reference frame, which is defined ...
  33. [33]
    [PDF] 4 The Cartesian Coordinate System - Pictures of Equa- tions
    A linear equation in x and y is an equation that is equivalent to an equation of the form. Ax + By + C = 0 where A, B, and C are constants. Example 4.37.
  34. [34]
    CoordinateTransformations - Intelligent Motion Lab
    Coordinate frame: A collection of an origin point and orthogonal axis vectors (e.g., X, Y, and Z) that may be used to represent points, directions, and ...<|control11|><|separator|>
  35. [35]
    [PDF] §10.1 - Parametric Equations Definition. A cartesian ... - Linda Green
    A cartesian equation for a curve is an equation in terms of x and y only. Definition. Parametric equations for a curve give both x and y as functions of a third.
  36. [36]
    construction of polar coordinates - PlanetMath.org
    Mar 22, 2013 · Polar coordinates are an alternative way of describing points in E E . Like the Cartesian coordinates, polar coordinates of a point are also ...
  37. [37]
    Archimedes' Spiral -- from Wolfram MathWorld
    Archimedes' spiral is an Archimedean spiral with polar equation r=atheta, studied by Archimedes, and used for angle division and circle squaring.
  38. [38]
    [PDF] Homogeneous Coordinates - UNCHAINED GEOMETRY
    Homogeneous coordinates are used in computer graphics for projecting 3D scenes onto 2D planes, and are scale invariant, with points at infinity.
  39. [39]
    Calculus II - Polar Coordinates - Pauls Online Math Notes
    ١٣‏/١١‏/٢٠٢٣ · In this section we will introduce polar coordinates an alternative coordinate system to the 'normal' Cartesian/Rectangular coordinate system ...ناقصة: Euclidean | عرض نتائج تتضمّن:Euclidean
  40. [40]
    Euclid's Elements, Book I, Proposition 47 - Clark University
    This proposition, I.47, is often called the Pythagorean theorem, called so by Proclus and others centuries after Pythagoras and even centuries after Euclid. The ...
  41. [41]
    Slope -- from Wolfram MathWorld
    A quantity which gives the inclination of a curve or line with respect to another curve or line. For a line in the xy ; where Deltax ; For a plane curve specified ...
  42. [42]
    Euclid's Elements, Book I, Proposition 32 - Clark University
    In any triangle, if one of the sides is produced, then the exterior angle equals the sum of the two interior and opposite angles.
  43. [43]
    Circle -- from Wolfram MathWorld
    ### Summary of Circle from Wolfram MathWorld
  44. [44]
    Euclid's Elements, Book III, Proposition 31 - Clark University
    The part of this proposition which says that an angle inscribed in a semicircle is a right angle is often called Thale's theorem. This proposition is used in ...
  45. [45]
    Power of a Point Theorem
    The power of a point inside the circle is negative, whereas that of a point outside the circle is positive. This is exactly what one obtains from the algebraic ...
  46. [46]
    Conic Section -- from Wolfram MathWorld
    The conic sections are the nondegenerate curves generated by the intersections of a plane with one or two nappes of a cone.
  47. [47]
    Eccentricity -- from Wolfram MathWorld
    The eccentricity of a conic section is a parameter that encodes the type of shape and is defined in terms of semimajor a and semiminor axes b as follows.Missing: source | Show results with:source
  48. [48]
    Focus | conic section | Britannica
    Oct 30, 2025 · In Conics, Apollonius defined the parabola in terms of its focus and directrix, laying the foundation for its modern geometric understanding.
  49. [49]
    Ellipse -- from Wolfram MathWorld
    An ellipse is a curve that is the locus of all points in the plane the sum of whose distances r_1 and r_2 from two fixed points F_1 and F_2 (the foci) ...Missing: Euclidean | Show results with:Euclidean<|separator|>
  50. [50]
    Hyperbola -- from Wolfram MathWorld
    A hyperbola (plural "hyperbolas"; Gray 1997, p. 45) is a conic section defined as the locus of all points P in the plane the difference of whose distances ...<|separator|>
  51. [51]
    Regular Polygon -- from Wolfram MathWorld
    alpha is the interior (vertex) angle, beta is the exterior angle, r is the inradius, R is the circumradius, and A is the area (Williams 1979, p. 33). polygon ...
  52. [52]
    [PDF] Polygon Triangulation - UCSB Computer Science
    Polygon triangulation partitions a polygon into non-overlapping triangles using diagonals, reducing complex shapes to simpler ones.
  53. [53]
    Planar Graphs and Euler's Formula
    The equation v−e+f=2 v − e + f = 2 is called Euler's formula for planar graphs . To prove this, we will want to somehow capture the idea of building up more ...
  54. [54]
    [PDF] Chapter 3 Basic Properties of Convex Sets - CIS UPenn
    It is obvious that the intersection of any family (finite or infinite) of convex sets is convex. Then, given any (nonempty) subset S of E, there is a smallest ...
  55. [55]
    [PDF] Lecture 4. Supporting and Separating Hyperplane Theorem
    Roughly speaking, we would like to show that every convex set C ⊆ Rd can be characterized by its 'supporting hyperplanes', and every two convex sets can be.
  56. [56]
    [PDF] What is a Polytope? - Harvard Mathematics Department
    A d-polytope P is the convex hull of finitely many points in Rd. 2. A d-polytope P is the bounded intersection of finitely many half-spaces in Rd.
  57. [57]
    [PDF] Stereographic Projection
    Stereographic projection maps a sphere's point to where the line from that point to a chosen point on the sphere intersects a plane through the sphere's center.Missing: polygons | Show results with:polygons
  58. [58]
    Pentagram -- from Wolfram MathWorld
    The pentagram, also called the five-point star, pentacle, pentalpha, or pentangle, is the star polygon {5/2}. It is a pagan religious symbol.
  59. [59]
    [PDF] Tilings by Regular Polygons - University of Washington
    Aug 3, 2005 · Another variant deserving attention deals with regular convex polygons tiling the sphere or the hyperbolic (non-Euclidean Lobačevski) plane.
  60. [60]
    [PDF] The geometry of Euclidean Space
    Mar 1, 2022 · A vector in Rn,n = 2,3 is an ordered pair(triple) of real numbers, such as. (a1,a2), or (a1,a2,a3). a1, a2 are called x coordinate, y coordinate ...
  61. [61]
    [PDF] Euclidean Spaces and Their Geometry
    Jan 7, 2004 · Thus vector addition (1.1) agrees with the classical way of defining ad- dition. The Parallelogram Law in R2 by showing that the line through. ( ...
  62. [62]
    [PDF] Vectors in R" and C", Spatial Vectors - Cimat
    (i) Addition: The resultant u+v of two vectors u and v is obtained by the so-called parallelogram law, i.e., u + v is the diagonal of the parallelogram formed ...
  63. [63]
    Basis Vector - an overview | ScienceDirect Topics
    Any vector in a plane can be expressed in terms of the vectors i and j; for instance, (2, 3) = 2(1,0) + 3(0,1) = 2i + 3j. (2,3) = 2(1,0) + 3(0, 1) = 2i + 3j.
  64. [64]
    [PDF] Basics of Affine Geometry - UPenn CIS
    Since an affine map preserves barycenters, and since an affine subspace V is closed under barycentric combinations, the image f(V ) of V is an affine subspace ...
  65. [65]
    [PDF] Day 11 Barycentric Coordinates and de Casteljau's algorithm
    Dec 16, 2004 · To calculate the barycentric coordinates of a point P in a plane with respect to an affine frame P0,P1,P2, one can proceed as follows. First ...
  66. [66]
    [PDF] MATH 202, CALCULUS 3 1. The geometry of Euclidean space 1.1 ...
    Definition 2.15 (Partial derivatives). Let U ⊆ Rn be an open set and suppose that f : U ⊂ Rn → R is a real-valued function. Then the partial derivatives ...
  67. [67]
    [PDF] Vector Calculus in Two Dimensions
    Level Sets and Gradient. Theorem 5.7. The gradient ∇u of a scalar field u is everywhere orthogonal to its level sets {u = c}.
  68. [68]
    [PDF] Vector Calculus
    [Df(u)] = ∂f. ∂x1. (u). ∂f. ∂x2. (u) ···. ∂f. ∂xn. (u) . Definition 4.3.2 (Gradient). Suppose that the partial derivatives of a scalar field f : U → R exist at ...
  69. [69]
    Calculus III - Directional Derivatives - Pauls Online Math Notes
    Nov 16, 2022 · In other words, we can write the directional derivative as a dot product and notice that the second vector is nothing more than the unit vector ...
  70. [70]
    [PDF] Approximate Distance Fields with Non-Vanishing Gradients
    This definition is consistent with the fact that the gradient vector of a distance function points in the normal direction in the neighborhood of all boundary ...
  71. [71]
    [PDF] helein.harmonic.pdf - Johns Hopkins University
    1 Harmonic functions on Euclidean spaces. Harmonic functions on an open domain Ω of Rm are solutions of the Laplace equation. ∆f = 0, where ∆ := ∂2. (∂x1)2.
  72. [72]
    Line Integral -- from Wolfram MathWorld
    The line integral of a vector field F(x) on a curve sigma is defined by int_(sigma)F·ds=int_a^bF(sigma(t))·sigma^'(t)dt, (1) where a·b denotes a dot product ...
  73. [73]
    4.3: Line Integrals - Mathematics LibreTexts
    Oct 27, 2024 · A line integral takes two dimensions, combines it into s , which is the sum of all the arc lengths that the line makes, and then integrates the ...Missing: Euclidean | Show results with:Euclidean
  74. [74]
    Double Integral -- from Wolfram MathWorld
    A double integral is a two-fold multiple integral. Examples of definite double integrals evaluating to simple constants include ...
  75. [75]
    Change of Variables Theorem -- from Wolfram MathWorld
    double integral · indefinite integrals · 3,2 TM rule 2139050. References. Jeffreys, H. and Jeffreys, B. S. "Change of Variable in an Integral." §1.1032 in ...
  76. [76]
    Gradient Theorem -- from Wolfram MathWorld
    Gradient Theorem int_a^b(del f)·ds=f(b)-f(a), where del is the gradient, and the integral is a line integral.
  77. [77]
    16.3: Conservative Vector Fields - Mathematics LibreTexts
    Jan 17, 2025 · In this section, we continue the study of conservative vector fields. We examine the Fundamental Theorem for Line Integrals, which is a useful generalization ...
  78. [78]
    Green's Theorem -- from Wolfram MathWorld
    in the plane with boundary partialD , Green's theorem states. ∮_(partialD)P(x,y)dx+Q(x. (1). where the left side is a line integral and the right side is a ...
  79. [79]
    An essay on the application of mathematical analysis to the theories ...
    Jun 21, 2022 · An essay on the application of mathematical analysis to the theories of electricity and magnetism. by: Green, George, 1793-1841. Publication ...Missing: Green's | Show results with:Green's
  80. [80]
    16.4: Green's Theorem - Mathematics LibreTexts
    Jan 17, 2025 · Green's theorem relates a line integral around a simply closed plane curve C and a double integral over the region enclosed by C.Learning Objectives · Circulation Form of Green's... · Flux Form of Green's Theorem
  81. [81]
    [PDF] Smooth Manifolds - UC Homepages
    Show that the disjoint union of uncountably many copies of R is locally Euclidean and Hausdorff, but not second countable. 1-3. Let M be a nonempty topological ...
  82. [82]
    [PDF] FUNDAMENTAL GROUPS - UChicago Math
    Aug 31, 2014 · The fundamental group of the punctured plane R2 − {0} is isomorphic to Z. Proof. We will show that the fundamental group of R2 − {0} is ...
  83. [83]
    [PDF] The Jordan Curve Theorem, Formally and Informally
    Dec 2, 2007 · The Jordan curve theorem states that every simple closed pla- nar curve separates the plane into a bounded interior region and an unbounded ...
  84. [84]
    Heine-Borel theorem in nLab
    ### Statement of Heine-Borel Theorem for Euclidean Spaces
  85. [85]
  86. [86]
    [PDF] Chapter 7 Planar graphs - UCSD Math
    Note: The quantity 2(1 − γ) is the Euler characteristic. It's usually denoted χ, which conflicts using χ(G) for chromatic number. Prof. Tesler. Ch. 7 ...
  87. [87]
    [PDF] Chapter 18 Planar Graphs
    A planar graph is a graph which can be drawn in the plane without any edges crossing. Some pictures of a planar graph might have crossing edges,.
  88. [88]
    Planar Graph - an overview | ScienceDirect Topics
    A planar graph is defined as a graph that can be drawn in a plane such that no edges cross each other, with intersections occurring only at the endpoints of ...
  89. [89]
    [PDF] On straight line representation of planar graphs.
    In the present note I shall prove that if a finite graph can be represented on the plane at all, it can be represented with straight segments as edges too, ...
  90. [90]
    [PDF] Planar Straight-Line Drawing Algorithms - Brown CS
    Planar straight-line drawings map edges to straight-line segments. Algorithms use shift and realizer methods to construct grid drawings. Every planar graph ...Missing: source | Show results with:source<|control11|><|separator|>
  91. [91]
    Outerplanar Graph -- from Wolfram MathWorld
    An outerplanar graph is a graph that can be embedded in the plane such that all vertices lie on the outer face.
  92. [92]
    [PDF] Structure and properties of maximal outerplanar graphs. - ThinkIR
    Aug 11, 2009 · In this section, the reader is provided basic graph theory definitions and notations used throughout this dissertation. A graph G is a finite ...
  93. [93]
    Sur le problème des courbes gauches en Topologie - EuDML
    Sur le problème des courbes gauches en Topologie. Casimir Kuratowski · Fundamenta Mathematicae (1930) ... pdf.png Full (PDF). How to cite. top. MLA; BibTeX; RIS.
  94. [94]
    Every planar map is four colorable - Project Euclid
    Every planar map is four colorable. K. Appel, W. Haken. DOWNLOAD PDF + SAVE TO MY LIBRARY. Bull. Amer. Math. Soc. 82(5): 711-712 (September 1976).
  95. [95]
    [PDF] an elementary proof that the utilities puzzle is impossible - Lomont.org
    It is equivalent to the graph K3,3 being nonplanar. Here I present a simple proof that can convince even most non- math majors. 1. The Utilities Puzzle.
  96. [96]
    [PDF] Kuratowski's Theorem - UChicago Math
    Aug 28, 2017 · Abstract. This paper introduces basic concepts and theorems in graph the- ory, with a focus on planar graphs. On the foundation of the basics, ...
  97. [97]
    EVERY PLANAR MAP IS FOUR COLORABLE PART I - Project Euclid
    Appel wishes tothank his teacher, Roger Lyndon, for teaching him how to think about mathematics. 429. Page 2. 430. K. APPEL AND W. HAKEN.
  98. [98]
    EVERY PLANAR MAP IS FOUR COLORABLE - Project Euclid
    EVERY PLANAR MAP IS FOUR COLORABLE. PART I1: REDUCIBILITY. BY. K. APPEL, W. HAKEN, AND J. KOCH. 1. Introduction. In Part I of this paper, a discharging ...