Fact-checked by Grok 2 weeks ago

Euclidean distance

The Euclidean distance between two points in Euclidean space is the length of the straight-line segment connecting them, computed as the of the sum of the squared differences between their corresponding coordinates. For points P = (x_1, x_2, \dots, x_n) and Q = (y_1, y_2, \dots, y_n) in n-dimensional , this is given by the formula
d(P, Q) = \sqrt{\sum_{i=1}^n (x_i - y_i)^2}. In two dimensions, it corresponds directly to the hypotenuse in the for the formed by the coordinate differences.
This distance metric satisfies the standard properties of a : it is non-negative (d(P, Q) \geq 0, with equality if and only if P = Q), symmetric (d(P, Q) = d(Q, P)), and obeys the (d(P, R) \leq d(P, Q) + d(Q, R) for any points P, Q, R). Equivalent to the \ell^2- (or L^2-) of the difference P - Q, it generalizes the ordinary in the plane and space to arbitrary finite dimensions, serving as the foundational measure in . The concept underpins key geometric structures, such as circles (sets of points at fixed from a ) and (vectors with zero projection). Beyond , Euclidean distance finds extensive applications across disciplines. In physics, it quantifies displacements and trajectories in . In and , it powers algorithms like k-nearest neighbors for and clustering in . Fields such as localization, molecular conformation determination in chemistry, and in statistics rely on it to reconstruct positions from distance or analyze similarities in high-dimensional spaces. These uses highlight its role in bridging theoretical with practical problem-solving.

Definition and Formulas

One dimension

In one dimension, the Euclidean distance between two points x and y on the real line is defined as the |x - y|. This formulation arises as a special case of the applied to a degenerate where one leg has zero length, reducing the distance to \sqrt{(x - y)^2 + 0^2} = |x - y|. For instance, the Euclidean distance between the points 3 and 7 is |7 - 3| = 4. This one-dimensional distance intuitively represents the straight-line separation between the points along the , providing the simplest measure of how far apart they are without any deviation or curvature. It serves as the foundational concept for extending the Euclidean distance to higher dimensions, such as two dimensions where the fully comes into play.

Two dimensions

In two dimensions, the Euclidean distance measures the straight-line separation between two points in the Cartesian plane, extending the one-dimensional case where equal y-coordinates reduce to a difference along the x-axis. Consider two points P = (x_1, y_1) and Q = (x_2, y_2) in the plane. The Euclidean distance d(P, Q) is given by the formula d(P, Q) = \sqrt{(x_2 - x_1)^2 + (y_2 - y_1)^2}. This expression arises directly from the Pythagorean theorem, which states that in a right-angled triangle, the square of the hypotenuse equals the sum of the squares of the other two sides. To derive the distance, draw a horizontal line segment from P to the point (x_2, y_1), with length \Delta x = |x_2 - x_1|, and a vertical line segment from (x_2, y_1) to Q, with length \Delta y = |y_2 - y_1|. These segments form the legs of a right triangle, where the line segment from P to Q is the hypotenuse. By the Pythagorean theorem, the hypotenuse length is \sqrt{(\Delta x)^2 + (\Delta y)^2}, yielding the distance formula. For example, the distance between (0, 0) and (3, 4) is \sqrt{(3-0)^2 + (4-0)^2} = \sqrt{9 + 16} = \sqrt{25} = 5. The Euclidean distance inherits the units of the , such as meters if x and y represent spatial measurements, ensuring physical consistency. It also scales linearly under uniform scaling of the coordinates by a k > 0, meaning d(kP, kQ) = k \cdot d(P, Q), preserving relative proportions in the plane.

Higher dimensions

In n-dimensional , denoted \mathbb{R}^n, the Euclidean distance between two points, represented as \mathbf{x} = (x_1, x_2, \dots, x_n) and \mathbf{y} = (y_1, y_2, \dots, y_n), is defined using the of their difference \mathbf{x} - \mathbf{y}. The formula is: d(\mathbf{x}, \mathbf{y}) = \|\mathbf{x} - \mathbf{y}\| = \sqrt{\sum_{i=1}^n (x_i - y_i)^2} This expression arises from the vector difference, where the squared differences in each coordinate are summed before taking the . The summation over coordinates generalizes the to multiple dimensions (or axes), treating the space as a multi-axial extension where the total distance is the across all directions. This reduces to the two-dimensional case when n=2. For a concrete illustration in three dimensions, consider the points (1, 2, 3) and (4, 5, 6). The differences in coordinates are $4-1=3, $5-2=3, and $6-3=3. Squaring these gives $9 + 9 + 9 = 27, and the distance is \sqrt{27} = 3\sqrt{3} \approx 5.196. Computing the Euclidean distance requires evaluating the sum of n terms, resulting in O(n) time complexity.

General formula

The Euclidean distance between two points x and y in a is defined as d(x, y) = \sqrt{\langle x - y, x - y \rangle}, where \langle \cdot, \cdot \rangle denotes the inner product on the space. This formula arises from the induced by the inner product, where the norm of a v is given by \|v\| = \sqrt{\langle v, v \rangle}; applying this to the difference v = x - y yields the distance as d(x, y) = \|x - y\|. The definition applies to any finite-dimensional , providing a coordinate-free characterization that unifies distances across dimensions without reliance on explicit coordinates. In such spaces, the standard inner product corresponds to the of coordinate representations. The of the inner product—ensuring \langle v, v \rangle > 0 for all nonzero v—guarantees that the resulting distances are real and nonnegative, with d(x, y) = 0 if and only if x = y.

Geometric and Metric Properties

Geometric interpretation

The Euclidean distance between two points in Euclidean space represents the length of the straight-line segment connecting them, embodying the intuitive notion of the shortest path in a flat, uncurved geometry. This concept originates from classical Euclidean geometry, where space is assumed to be homogeneous and isotropic, allowing direct measurement along the geodesic that is a straight line. In this framework, any deviation from the straight path would increase the total length, as established by the properties of straight lines in plane and solid geometry. Geometrically, the Euclidean distance can be visualized as the of a formed by the differences in the coordinates of the two points. For instance, in two dimensions, consider two lattice points such as (0,0) and (3,4); the horizontal and vertical separations form the legs of the triangle, and the straight-line distance is the spanning these differences, measuring 5 units. This interpretation extends naturally to higher dimensions, where in , the distance between points like (0,0,0) and (1,1,1) traces the space diagonal of a , again as the generalized across the coordinate axes. Such visualizations highlight how the distance captures the direct, minimal separation without intermediate curves or bends. A key property of Euclidean distance is its preservation under isometries, which are rigid motions such as translations, rotations, and reflections that maintain the of the . These transformations do not alter distances between points, ensuring that the geometric interpretation remains invariant; for example, rotating a pair of points around an axis leaves their separation unchanged. This invariance underscores the foundational role of Euclidean distance in defining the of flat .

Metric space axioms

The Euclidean distance, defined on the vector space \mathbb{R}^n by the general formula d(\mathbf{x}, \mathbf{y}) = \sqrt{\sum_{i=1}^n (x_i - y_i)^2} for \mathbf{x} = (x_1, \dots, x_n) and \mathbf{y} = (y_1, \dots, y_n), satisfies the axioms of a metric space, thereby establishing \mathbb{R}^n as a metric space under this distance function. The non-negativity axiom holds: d(\mathbf{x}, \mathbf{y}) \geq 0 for all \mathbf{x}, \mathbf{y} \in \mathbb{R}^n, with equality if and only if \mathbf{x} = \mathbf{y}. This follows directly from the definition, as the expression under the square root is a sum of squares, which is nonnegative and zero precisely when each x_i = y_i. Symmetry is also satisfied: d(\mathbf{x}, \mathbf{y}) = d(\mathbf{y}, \mathbf{x}) for all \mathbf{x}, \mathbf{y} \in \mathbb{R}^n. This is immediate, since (x_i - y_i)^2 = (y_i - x_i)^2 for each i. The states that d(\mathbf{x}, \mathbf{z}) \leq d(\mathbf{x}, \mathbf{y}) + d(\mathbf{y}, \mathbf{z}) for all \mathbf{x}, \mathbf{y}, \mathbf{z} \in \mathbb{R}^n. To verify this, consider the vectors \mathbf{u} = \mathbf{x} - \mathbf{y} and \mathbf{v} = \mathbf{y} - \mathbf{z}, so \mathbf{x} - \mathbf{z} = \mathbf{u} + \mathbf{v}. The inequality d(\mathbf{x}, \mathbf{z}) \leq d(\mathbf{x}, \mathbf{y}) + d(\mathbf{y}, \mathbf{z}) then reduces to \|\mathbf{u} + \mathbf{v}\| \leq \|\mathbf{u}\| + \|\mathbf{v}\|, where \|\cdot\| denotes the Euclidean norm. This follows from the Cauchy-Schwarz inequality: |\mathbf{u} \cdot \mathbf{v}| \leq \|\mathbf{u}\| \|\mathbf{v}\|, which implies \|\mathbf{u} + \mathbf{v}\|^2 = \|\mathbf{u}\|^2 + 2\mathbf{u} \cdot \mathbf{v} + \|\mathbf{v}\|^2 \leq \|\mathbf{u}\|^2 + 2\|\mathbf{u}\| \|\mathbf{v}\| + \|\mathbf{v}\|^2 = (\|\mathbf{u}\| + \|\mathbf{v}\|)^2. Taking square roots yields the desired result, with equality when \mathbf{u} and \mathbf{v} are linearly dependent and point in the same direction. In \mathbb{R}^n, the Euclidean metric is unique up to , meaning any other that induces the standard geometry on the space is equivalent via a distance-preserving .

Squared Euclidean Distance

Definition and computation

The squared Euclidean distance between two points \mathbf{x} = (x_1, \dots, x_n) and \mathbf{y} = (y_1, \dots, y_n) in n-dimensional is defined as the sum of the squared differences of their corresponding coordinates: d^2(\mathbf{x}, \mathbf{y}) = \sum_{i=1}^n (x_i - y_i)^2 = \|\mathbf{x} - \mathbf{y}\|^2, where \|\cdot\|^2 denotes the square of the Euclidean norm. This formulation omits the square root operation required in the standard Euclidean distance, which simplifies both manual and algorithmic computations. The avoidance of the square root provides key computational advantages, particularly in programming and numerical methods, as the square root function introduces additional processing overhead and potential floating-point precision issues. For example, in iterative algorithms like , using the squared form accelerates distance calculations across large datasets without altering the relative ordering of distances. Additionally, it facilitates gradient computations in optimization problems, where the partial derivative with respect to \mathbf{x} is simply \nabla_{\mathbf{x}} d^2(\mathbf{x}, \mathbf{y}) = 2(\mathbf{x} - \mathbf{y}), enabling efficient updates in methods such as . To illustrate, consider the points (0,0) and (3,4) in two dimensions: the squared Euclidean distance is (3-0)^2 + (4-0)^2 = 9 + 16 = 25. The standard Euclidean distance is the of this value.

Relation to Euclidean distance

The squared Euclidean distance d^2(\mathbf{x}, \mathbf{y}) is a strictly increasing of the standard Euclidean distance d(\mathbf{x}, \mathbf{y}) for non-negative values, since the squaring is monotonic on [0, \infty). Consequently, for any points \mathbf{x}, \mathbf{y}, \mathbf{z}, d(\mathbf{x}, \mathbf{y}) < d(\mathbf{x}, \mathbf{z}) if and only if d^2(\mathbf{x}, \mathbf{y}) < d^2(\mathbf{x}, \mathbf{z}), preserving the relative ordering of distances. This monotonic relationship makes minimizing d^2(\mathbf{x}, \mathbf{y}) mathematically equivalent to minimizing d(\mathbf{x}, \mathbf{y}) in optimization problems, as the square root operation does not alter the location of minima. In particular, this equivalence underpins , where the objective function minimizes the sum of squared Euclidean distances (or errors) between observed data and model predictions, facilitating closed-form solutions via linear algebra. Despite these advantages, the squared Euclidean distance fails to qualify as a true metric because it violates the triangle inequality: for some points \mathbf{x}, \mathbf{y}, \mathbf{z}, d^2(\mathbf{x}, \mathbf{z}) > d^2(\mathbf{x}, \mathbf{y}) + d^2(\mathbf{y}, \mathbf{z}). A simple counterexample in one dimension is the points x = 0, y = 0.5, z = 1: here, d^2(x, z) = 1 but d^2(x, y) + d^2(y, z) = 0.25 + 0.25 = 0.5. The squared Euclidean distance is commonly employed in algorithms where absolute scale is irrelevant and only comparative ordering or relative proximity matters, such as nearest-neighbor search or , as it avoids the computational overhead of square roots while yielding identical rankings.

Generalizations and Extensions

To non-Euclidean spaces

In non-Euclidean spaces, distance measures adapt to intrinsic geometry defined by curvature, replacing the flat Euclidean metric with path lengths along . In Riemannian manifolds, the g_{ij} governs local geometry, with the given by
ds = \sqrt{g_{ij} \, dx^i \, dx^j}.
The distance between points p and q is the infimum of \int ds over all smooth curves connecting them, representing the geodesic length.
A key example is the , where positive curvature necessitates great-circle distances for surface travel, as straight lines (chords) underestimate paths. The computes this distance d between points at latitudes \phi_1, \phi_2 and longitudes \lambda_1, \lambda_2 (in radians) on a sphere of R:
a = \sin^2\left(\frac{\phi_2 - \phi_1}{2}\right) + \cos \phi_1 \cos \phi_2 \sin^2\left(\frac{\lambda_2 - \lambda_1}{2}\right),
d = 2R \arcsin(\sqrt{a}).
This avoids numerical instability in cosine-based alternatives for small angles. In practice, global positioning systems (GPS) on employ distances based on ellipsoidal models such as WGS84 for , as distances fail over scales comparable to the planet's , leading to errors in route optimization and positioning.
Hyperbolic spaces, exhibiting negative curvature, further diverge, with the confining the geometry to the unit disk |z| < 1. The between points z_1 and z_2 is
d(z_1, z_2) = 2 \artanh \left| \frac{z_2 - z_1}{1 - \overline{z_1} z_2} \right|,
or equivalently
d(z_1, z_2) = \cosh^{-1} \left( 1 + \frac{2 |z_1 - z_2|^2}{(1 - |z_1|^2)(1 - |z_2|^2)} \right).
These forms arise from the model's conformal ds = \frac{2 |dz|}{1 - |z|^2}, emphasizing exponential expansion away from the origin. Overall, Euclidean approximates these only locally in low-curvature regimes, breaking down globally where geodesics diverge or converge non-linearly.

To other mathematical objects

The Euclidean distance concept extends naturally to distances between more complex objects in , such as lines, compact sets, and functions, by defining appropriate that leverage the point-to-point Euclidean distance while accounting for the structure of the objects. These extensions maintain the geometric intuition of the Euclidean but adapt computations to aggregate or relational properties. In three-dimensional , the shortest between two lines is the length of the common perpendicular segment connecting them, which varies based on whether the lines are or . For with direction \vec{d} and points P on the first line and Q on the second, the is the of the of the \vec{PQ} and the unit direction , given by d = \frac{|\vec{PQ} \times \vec{d}|}{|\vec{d}|}. For , which neither intersect nor are , the shortest lies along the unique line perpendicular to both, and is computed as d = \frac{|(\vec{PQ} \cdot (\vec{d_1} \times \vec{d_2}))|}{|\vec{d_1} \times \vec{d_2}|}, where \vec{d_1} and \vec{d_2} are the direction vectors of the lines and \vec{PQ} connects a point on each. For compact sets in \mathbb{R}^n, the Hausdorff distance quantifies how far two sets are from each other by measuring the maximum deviation needed to cover one set with balls centered on the other, using the Euclidean distance between points. Formally, for nonempty compact sets A and B, the Hausdorff distance is d_H(A, B) = \max\left( \sup_{a \in A} \inf_{b \in B} d(a, b), \sup_{b \in B} \inf_{a \in A} d(a, b) \right), where d is the Euclidean distance; this defines a metric on the space of compact subsets. This distance is zero if and only if the sets coincide, and it is particularly useful for comparing shapes or approximating sets. In the context of functions, the Euclidean distance generalizes to the L^2 space of square-integrable functions over a domain, where the distance between two functions f and g is the L^2 norm of their difference: \|f - g\|_{L^2} = \sqrt{\int (f(x) - g(x))^2 \, dx}. This norm induces a structure on L^2, analogous to the for finite-dimensional vectors, and satisfies the metric axioms with respect to the underlying . A example of extending to lines is the from a point to a line, computed via orthogonal . In three dimensions, for a line passing through point \mathbf{x_1} = (x_1, y_1, z_1) with direction vector \mathbf{v} = (a, b, c) and a point \mathbf{x_0} = (x_0, y_0, z_0), the is d = \frac{|(\mathbf{x_0} - \mathbf{x_1}) \times \mathbf{v}|}{|\mathbf{v}|}, which is the length of the from \mathbf{x_0} to the line. This formula arises from the geometric property that the shortest path is to the line's direction.

Applications

In classical geometry

In classical geometry, the Euclidean distance serves as the foundational measure of separation between points in a , enabling precise definitions and constructions. , a fundamental figure, is defined as the set of all points in a that are from a fixed point called . This , often denoted as the , determines the circle's size and uniformity, with all radii being equal. A key application appears in the study of triangles, where Euclidean distances between vertices form the sides that underpin congruence criteria. Specifically, if two triangles have their three sides equal in length—measured as distances—then the triangles are , meaning they coincide completely when superimposed. This side-side-side (SSS) criterion ensures that equal distances preserve the triangle's shape and size, allowing geometers to identify identical figures without further measurement. The further illustrates the role of Euclidean distance in right-angled triangles, stating that the square of the equals the sum of the squares of the other two sides. In practical terms, this applies to finding diagonals in rectangles, where the diagonal's length is the Euclidean distance between opposite corners, calculated as the of the sum of the squares of the adjacent sides. Geometric constructions rely heavily on the to measure and transfer Euclidean distances accurately. Euclid's third postulate permits drawing a with any given and , using the to mark points at equal distances from the . This tool facilitates tasks like constructing equilateral triangles by transferring distances between points, ensuring precision in replicating lengths without numerical computation.

In modern fields

In , the Euclidean distance serves as the primary metric in the k-nearest neighbors (k-NN) algorithm for classification tasks, where it measures the similarity between a query point and its nearest training samples to assign class labels based on majority voting among the k closest neighbors. This approach, originally proposed by Fix and Hodges in , relies on Euclidean distance to identify neighbors in feature space, enabling non-parametric predictions without assuming data distribution. In practice, is often employed instead to avoid computations, enhancing efficiency while preserving relative ordering. In , Euclidean distance is fundamental for in environments, where it computes the shortest separation between objects to prevent overlaps during animations or simulations. For ray tracing, this distance metric determines intersection points between rays and scene geometry, facilitating realistic rendering by tracing light paths and calculating distances to surfaces for shading and reflection effects. In physics simulations, particularly , Euclidean distance quantifies inter-particle separations to model forces and interactions, such as van der Waals potentials that depend on atomic pairwise distances in . This application allows for the accurate prediction of molecular conformations and by integrating distance-based energy terms over time steps in simulations of chemical systems. In , clustering algorithms like k-means utilize Euclidean distance to minimize intra-cluster variances, assigning data points to the nearest and iteratively updating centroids as the mean of assigned points to form compact groups. Introduced by MacQueen in 1967, this method optimizes the sum of squared Euclidean distances from points to their cluster centers, providing an effective way to uncover patterns in multidimensional datasets.

Historical Development

Ancient origins

The conceptual foundations of Euclidean distance trace back to ancient civilizations' practical needs for measuring land and constructing structures, where the straight-line distance between points was implicitly understood through geometric tools and theorems. In and , surveyors employed ropes knotted at specific lengths to demarcate boundaries and establish right angles for land measurement, a practice essential for agriculture and architecture along the and in Mesopotamian fields. surveyors, known as harpedonaptai or "rope stretchers," used these cords to form 3-4-5 triangles, enabling precise perpendicular alignments without explicit distance formulas but relying on the inherent of straight lines. Similarly, Babylonian scribes divided irregular plots into right-angled triangles and rectangles using plumb bobs and cords, as evidenced in clay tablets documenting field surveys around 1800 BCE. Around 500 BCE in , the emerged as a key insight into distances in right-angled triangles, stating that the square of the equals the sum of the squares of the other two sides, thus providing a method to compute the straight-line distance between points forming such a figure. Attributed to of (c. 570–495 BCE) and his followers, this relation was likely influenced by earlier Egyptian and Babylonian practices but formalized within the Pythagorean school as a philosophical and mathematical principle. Euclid's Elements, composed around 300 BCE in , built upon these ideas by axiomatizing , including postulates that directly imply the concept of Euclidean : the first allows drawing a straight line between any two points, and the third permits constructing a with a given center and , equating to a fixed . These definitions treated as the along the shortest path in a , underpinning all subsequent propositions without algebraic notation. Concurrently in ancient , the Sulba Sutras (c. 800–200 BCE), ritual texts for altar construction, incorporated diagonal calculations akin to the to ensure precise geometric shapes, such as computing the diagonal of a square to double its area using knotted ropes. Texts like the Baudhayana Sulba Sutra (c. 800 BCE) describe these methods for Vedic fire altars, emphasizing practical distances in sacred architecture.

Modern formalization

The modern formalization of Euclidean distance emerged in the 17th century with ' development of , which introduced Cartesian coordinates as a bridge between algebra and geometry. In his 1637 treatise , Descartes assigned numerical coordinates to points in the plane, allowing geometric relations, including distances, to be expressed and computed algebraically through equations derived from the . This innovation transformed the intuitive notion of straight-line distance into a precise, calculable quantity, foundational for subsequent mathematical rigor. By the early 20th century, the concept was abstracted to infinite-dimensional settings through David Hilbert's investigations into integral equations. In works from 1904 to 1910, Hilbert explored spaces of functions where convergence and distance could be defined analogously to finite-dimensional Euclidean spaces, using inner products to induce a that generalized the Euclidean distance. These efforts established the framework for infinite-dimensional Euclidean-like structures, later termed Hilbert spaces, which verified the distance's properties in broader analytical contexts. A pivotal advancement came with the axiomatic treatment of distance in theory. In his 1906 doctoral thesis Sur quelques points du calcul fonctionnel, Maurice Fréchet first articulated the abstract axioms for a —positivity, , and the —applicable to diverse sets, including those equipped with Euclidean distance, thereby confirming its adherence to a general structure for measuring separation. Building on this, Felix Hausdorff's 1914 monograph Grundzüge der Mengenlehre systematized , integrating Fréchet's ideas with set-theoretic foundations to rigorously define and explore distances satisfying these axioms in topological settings. Following these developments, the saw the integration of Euclidean distance into the emerging theory of abstract vector spaces within linear algebra. Works by mathematicians such as and Hans Hahn formalized normed vector spaces, where the Euclidean distance appeared as the derived from an inner product, providing a unified algebraic on finite- and infinite-dimensional cases. This formulation solidified the distance's role as a fundamental metric in modern linear structures, emphasizing its compatibility with vector addition and .

References

  1. [1]
    [PDF] 1: Geometry and Distance - Harvard Mathematics Department
    The Euclidean distance between two points P = (x, y, z) and Q = (a, b, c) in space is defined as d(P, Q) = q (x - a)2 + (y - b)2 + (z - c)2. Note that this is ...
  2. [2]
    [PDF] Lecture Notes in Modern Geometry 1 The euclidean plane
    The euclidean distance is a function d : R2 × R2 → R defined as d((x1,y1),(x2,y2)) = p. (x1 − x2)2 + (y1 − y2)2. This function represents Pythagorean distance ...
  3. [3]
    [PDF] Distance and Metric Spaces
    Distance in the Plane. The (Euclidean) distance d((x1,y1), (2, 2)) between two points (1,1) and (2,2) is the length of the line segment with endpoints (1,9₁) ...
  4. [4]
    [PDF] Euclidean Distance Geometry and Applications - UC Davis Math
    Abstract. Euclidean distance geometry is the study of Euclidean geometry based on the concept of distance. This is useful in several applications where the ...
  5. [5]
    [PDF] Euclidean Distance Matrices
    Other examples of using Euclidean distance geometry in machine learning are the results by Tenenbaum, De Silva, and Langford [23] on image understanding and ...
  6. [6]
    [PDF] Histogram Specification and Kalman Filter with Euclidian distance ...
    In one dimension, the Euclidean distance between a pair of points, x1 and x2, lying on the same line is the absolute difference between the two points as:.
  7. [7]
    Euclidean Distance | Calculation, Formula & Examples - Study.com
    Euclidean Distance Formula. The Euclidean distance formula for one dimension is D = | x 1 − x 2 | , where D is the Euclidean distance, and and are the ...
  8. [8]
    1.2 Distance Between Two Points; Circles
    1. The Pythagorean theorem then says that the distance between the two points is the square root of the sum of the squares of the horizontal and vertical sides: ...
  9. [9]
    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.
  10. [10]
    1.1 Rectangular Coordinate Plane - The Texas A&M University System
    Equation 1.1 Distance Formula. The distance d between the points P\left(x_{0}, y_{0}\right) and Q\ ...
  11. [11]
    Distance Between Two Points - Department of Mathematics at UTSA
    Oct 26, 2021 · These names come from the ancient Greek mathematicians Euclid and ... A paraboloid, the graph of squared Euclidean distance from the origin.
  12. [12]
    Euclidean Distance In 'n'-Dimensional Space
    Euclidean distance is a measure of the true straight line distance between two points in Euclidean space.Missing: definition | Show results with:definition
  13. [13]
    [PDF] Experimental Comparison of Representations and Distance Measures
    method approaches that of Euclidean distance. Recall that. Euclidean distance has a time complexity of O(n) and that a single DTW calculation has a time ...
  14. [14]
    [PDF] Norms and Inner Products - Stanford AI Lab
    Jul 21, 2018 · If an inner product space is complete with respect to the distance metric induced by its inner product, it is said to be a Hilbert space.Missing: sqrt | Show results with:sqrt
  15. [15]
    [PDF] Hilbert spaces
    A Hilbert space H is a pre-Hilbert space which is complete with respect to the norm induced by the inner product. As examples we know that Cn with the usual ...
  16. [16]
    [PDF] Contents 3 Inner Product Spaces - Evan Dummit
    Definition: If V is an inner product space, we define the norm (or length) of a vector v to be ||v|| = phv, vi. ◦ When V = Rn with the standard dot product, the ...
  17. [17]
    Distance formula | Analytic geometry (video) - Khan Academy
    Aug 30, 2016 · We can rewrite the Pythagorean theorem as d=√((x_2-x_1)²+(y_2-y_1)²) to find the distance between any two points. Created by Sal Khan and CK-12 Foundation.Missing: interpretation | Show results with:interpretation
  18. [18]
    Pythagorean Theorem and its many proofs
    122 proofs of the Pythagorean theorem: squares on the legs of a right triangle add up to the square on the hypotenuse.Analog gadgets · The Eye Opener Series · Tesselation · Another one<|control11|><|separator|>
  19. [19]
    Isometries Preserve Distances
    Oct 31, 2009 · Isometries Preserve Distances · 1. Synthetic and Analytic Geometry · 2. Background · 2.1. Length, Norm, and Magnitude · 2.2. Bilinearity Properties ...
  20. [20]
    [PDF] METRIC SPACES 1. Introduction As calculus developed, eventually ...
    On Rm the Euclidean metric is. dE(x,y) = ||x − y|| = p. (x1 − y1)2 + ··· + (xm − ym)2,. This is the usual distance used in Rm, and when we speak about Rm as a ...
  21. [21]
    [PDF] Metric Spaces - ScholarWorks@GVSU
    ... Euclidean distance dE. In our preview activity we saw that the ... Now we prove the Cauchy-Schwarz Inequality. Proof of the Cauchy-Schwarz Inequality.<|control11|><|separator|>
  22. [22]
    [PDF] Cauchy-Schwarz inequality - Williams College
    1 + x2. 2)(y2. 1 + y2. 2). (†). Thus the triangle inequality for the ... Proof of Cauchy-Schwarz. Given x,y ∈ Rn, we have. (x · y)2 = (|x| |y|cosθ)2 ...
  23. [23]
    Euclidean space in nLab
    Sep 5, 2023 · In the strict sense of the word, Euclidean space E n E^n of dimension n n is, up to isometry, the metric space whose underlying set is the ...
  24. [24]
    Euclidean and Euclidean Squared Distances - Oracle Help Center
    Euclidean distance reflects the distance between each of the vectors' coordinates being compared—basically the straight-line distance between two vectors.
  25. [25]
    Optimizing Euclidean Distance Computation - MDPI
    By utilizing the squared Euclidean distance (3), the K-means algorithm achieves computational efficiency and mathematical tractability, even though it may ...
  26. [26]
    [PDF] The limits of squared Euclidean distance regularization
    We also show that algorithms that regularize with the squared Euclidean distance are easily confused by random features. Finally, we conclude by discussing re-.
  27. [27]
    [PDF] Least Squares Optimization - Center for Neural Science
    Feb 15, 1999 · We can express the sum of squared distances of the data to this N − 1 dimensional space as the sum of squared projections onto a unit.
  28. [28]
    Is squared Euclidean distance a metric? - Math Stack Exchange
    Dec 25, 2015 · Is squared Euclidean distance a metric? In particular does it obey triangle inequality? I think no, but cannot find a counterexample.Euclidean distance vs Squared - Mathematics Stack ExchangeEuclidean distance proof - metric spaces - Math Stack ExchangeMore results from math.stackexchange.com
  29. [29]
    3 Introducing Riemannian Geometry‣ General Relativity by David ...
    A general Riemannian metric gives us a way to measure the length of a vector X at each point, It also allows us to measure the angle between any two vectors X ...
  30. [30]
    Computing Distances - NYU Computer Science
    great circle distance d will be in the same units as R. When the two points are antipodal (on opposite sides of the Earth), the Haversine Formula is ill- ...
  31. [31]
    [PDF] 19840009100.pdf - NASA Technical Reports Server (NTRS)
    The principle advantage of RNAV is that it allows the navigator to fly a great circle course between two points. The great circle course is the shortest path ...
  32. [32]
    [PDF] Chapter 9 - Poincaré's Disk Model for Hyperbolic Geometry
    means that our standard distance formula will not work. We introduce a distance metric by dρ = 2dr. 1 − r2 where ρ represents the hyperbolic distance and r ...
  33. [33]
    [PDF] Supplemental-Lecture-4-Surfaces-and-Euclidean-Spherical-and ...
    For distances small compared to 1 √𝐾𝐾. ⁄ the curved spaces are well approximated by a flat Euclidean metric but this is untrue for distances large compared to 1 ...
  34. [34]
    Point-Line Distance--3-Dimensional -- from Wolfram MathWorld
    Let a line in three dimensions be specified by two points x_1=(x_1,y_1,z_1) and x_2=(x_2,y_2,z_2) lying on it, so a vector along the line is given by ...
  35. [35]
    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 ...
  36. [36]
    Grundzüge der Mengenlehre : Hausdorff, Felix, 1868-1942
    Dec 2, 2008 · Grundzüge der Mengenlehre. by: Hausdorff, Felix, 1868-1942 ... PDF download · download 1 file · SCRIBE SCANDATA ZIP download · download 1 ...
  37. [37]
    L^2-Space -- from Wolfram MathWorld
    L^2-space really consists of equivalence classes of functions. Two functions represent the same L^2-function if the set where they differ has measure zero.
  38. [38]
    L^2-Norm -- from Wolfram MathWorld
    The l^2-norm (also written "l^2-norm") |x| is a vector norm defined for a complex vector x=[x_1; x_2; |; x_n] (1) by |x|=sqrt(sum_(k=1)^n|x_k|^2), (2) where ...Missing: distance | Show results with:distance
  39. [39]
    Euclid's Elements, Book I, Definitions 15-18 - Clark University
    Definition 15. A circle is a plane figure contained by one line such that all the straight lines falling upon it from one point among those lying within the ...Missing: source | Show results with:source
  40. [40]
    Euclid's Elements, Book I, Proposition 8 - Clark University
    If two triangles have the two sides equal to two sides respectively, and also have the base equal to the base, then they also have the angles equal.
  41. [41]
    Euclid's Elements, Book I, Postulate 3 - Clark University
    This is the third assumed construction in the Elements. It corresponds to drawing a circle with a compass. Circles were defined in Def.I.15 and Def.I.16 as ...
  42. [42]
    Euclid's Elements, Book I, Proposition 2 - Clark University
    When using a compass and a straightedge to perform this construction there are more circles drawn than shown in the diagram that accompanies the proposition.
  43. [43]
    K-nearest neighbor - Scholarpedia
    Feb 21, 2009 · The k-nearest-neighbor classifier is commonly based on the Euclidean distance between a test sample and the specified training samples.Characteristics of kNN · Between-sample geometric... · Pseudocode (algorithm)
  44. [44]
    [PDF] Distance Metric Learning for Large Margin Nearest Neighbor ...
    Abstract. The accuracy of k-nearest neighbor (kNN) classification depends significantly on the metric used to compute distances between different examples.
  45. [45]
    3D collision detection - Game development - MDN Web Docs
    Jul 11, 2025 · Taking into account that the Euclidean distance between two points A and B is ( A x − B x ) 2 + ( A y − B y ) 2 + ( A z − ...
  46. [46]
    [PDF] 3D Collision Detection: A Survey
    Apr 13, 2000 · The idea is to determine the closest points of two polyhedra, and then compute the euclidean distance between them. Three methods for ...
  47. [47]
    Euclidean Distance Geometry and Applications | SIAM Review
    We survey the theory of Euclidean distance geometry and its most important applications, with special emphasis on molecular conformation problems.
  48. [48]
    A Euclidean transformer for fast and stable machine learned force ...
    Aug 6, 2024 · We propose a transformer architecture called SO3krates that combines sparse equivariant representations (Euclidean variables) with a self-attention mechanism.
  49. [49]
    K-Means Classification Algorithm - an overview | ScienceDirect Topics
    The K-means algorithm (MacQueen, 1967) classifies points or observations into K clusters by minimizing the sum of squared (Euclidean) distances between the ...
  50. [50]
    [PDF] Surveying from Egypt to Greece – 3000 B.C. to 100 A.D.
    The transferal of geometric knowledge from the Egyptians to the Greeks is shown to commence with Thales and Pythagoras travelling to Egypt and transporting ...
  51. [51]
    How ancient Babylonian land surveyors developed a unique form of ...
    Aug 4, 2021 · Babylonian surveyors would use these tools to divide land into manageable shapes: rectangles, right-angled triangles and right trapezoids.
  52. [52]
    Pythagoras - Biography - MacTutor - University of St Andrews
    Pythagoras was a Greek philosopher who made important developments in mathematics, astronomy, and the theory of music. The theorem now known as Pythagoras's ...
  53. [53]
    Pythagoras - Stanford Encyclopedia of Philosophy
    Feb 23, 2005 · Pythagoras, one of the most famous and controversial ancient Greek philosophers, lived from ca. 570 to ca. 490 BCE. He spent his early years ...
  54. [54]
    Indian Sulbasutras - MacTutor History of Mathematics
    The most important of these documents are the Baudhayana Sulbasutra written about 800 BC and the Apastamba Sulbasutra written about 600 BC. Historians of ...Missing: sources | Show results with:sources
  55. [55]
    [PDF] Baudhayana Sulbasutra written about 800 BC
    Baudhayana Sulbasutra: The rope which is stretched across the diagonal of a square produces an area double the size of the original square. Katyayana Sulbasutra ...Missing: calculations sources
  56. [56]
    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.
  57. [57]
    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” ( ...
  58. [58]
    David Hilbert (1862 - 1943) - Biography - MacTutor
    This work also established the basis for his work on infinite-dimensional space, later called Hilbert space, a concept that is useful in mathematical analysis ...Missing: norm | Show results with:norm
  59. [59]
    René Maurice Fréchet, Hon.F.R.S.E. - RSE Obituary - MacTutor
    In a later part of his thesis Fréchet gave for the first time the now familiar axioms for a metric space, giving examples that included various function spaces ...
  60. [60]
    Felix Hausdorff (1868 - 1942) - Biography - MacTutor
    A year later, in 1914, Hausdorff published his famous text Grundzüge der Mengenlehre Ⓣ. (General set theory). which builds on work by Fréchet and others to ...Missing: space | Show results with:space
  61. [61]
    Earliest Known Uses of Some of the Words of Mathematics (M)
    The term METRIC SPACE is due to Felix Hausdorff (1869-1942) who gave axioms for the metrischer Raum in his Grundzüge der Mengenlehre (1914, pp. 211-2).<|separator|>
  62. [62]
    Abstract linear spaces - MacTutor History of Mathematics
    It is possible however to trace the beginning of the vector concept back to the beginning of the 19th Century with the work of Bolzano. In 1804 he published a ...