Jump to content

Euclidean space: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
kept last edit, but reverted second-to-last, then tried to repair
→‎Other geometrical spaces: geometrical -> geometric
(99 intermediate revisions by 8 users not shown)
Line 1: Line 1:
{{short description|Generalization of Euclidean geometry to higher dimensions}}
{{short description|Fundamental space of geometry}}
{{about|Euclidean spaces of all dimensions|3-dimensional Euclidean space|3-dimensional space}}


[[Image:Coord system CA 0.svg|thumb|right|250px|Every point in three-dimensional Euclidean space is determined by three coordinates.]]
[[Image:Coord system CA 0.svg|thumb|right|250px|A point in three-dimensional Euclidean space can be located by three coordinates.]]


In [[geometry]], '''Euclidean space''' encompasses the two-[[dimension (mathematics)|dimensional]] Euclidean [[plane (geometry)|plane]], the [[three-dimensional space]] of [[Euclidean geometry]], and similar spaces of higher dimension. It is named after the [[Greek mathematics|Ancient Greek mathematician]] [[Euclid|Euclid of Alexandria]].<ref>{{cite book|last=Ball|first=W.W. Rouse|authorlink=W. W. Rouse Ball|title = A Short Account of the History of Mathematics|origyear=1908|url=|edition=4th|year=1960|publisher= Dover Publications
'''Euclidean space''' is the fundamental space of [[geometry]]. Originally, this was the [[three-dimensional space]] of [[Euclidean geometry]], but, in modern [[mathematics]], there are Euclidean spaces of any nonnegative integer dimension,{{sfn|Solomentsev|2001}} including the three-dimensional space and the ''Euclidean plane'' ([[dimension (mathematics)|dimension]] two). It has been introduced by the [[Greek mathematics|Ancient Greek]] mathematician [[Euclid|Euclid of Alexandria]],{{sfn|Ball|1960|pages =50–62}} and the qualifier ''Euclidean'' has been added for distinguishing it from other spaces that are considered in [[physics]] and modern mathematics.
|isbn=0-486-20630-0|pages =50–62}}</ref> The term "Euclidean" distinguishes these spaces from [[#Alternatives and generalizations|other types of spaces considered in modern geometry]]. Euclidean spaces also generalize to [[higher dimension]]s.


Ancient [[History of geometry#Greek geometry|Greek geometers]] introduced Euclidean space for modeling the [[physical universe]]. Their great innovation was to ''[[mathematical proof|prove]]'' all properties of the space ([[theorem]]s) by starting from a few fundamental properties, called ''[[postulate]]s'', which either were considered as evidences (for example, there is exactly one [[straight line]] passing through two points), or seemed impossible to prove ([[parallel postulate]]).
Classical [[History of geometry#Greek geometry|Greek geometry]] defined the Euclidean plane and Euclidean three-dimensional space using certain [[axiom|postulates]], while the other properties of these spaces were deduced as [[theorem]]s. Geometric constructions were also used to define [[rational number]]s as ratios of [[commensurability (mathematics)|commensurable lengths]]. When [[algebra]] and [[mathematical analysis]] became developed enough, this relation reversed and now it is more common to define Euclidean spaces from [[vector space]]s, which allows using [[Cartesian coordinates]] and the power of [[algebra]] and [[calculus]]. This means that [[point (geometry)|points]] are specified with [[tuple]]s of [[real number]]s, called [[coordinate vector]]s, and [[geometric shape]]s are defined by [[equation]]s and [[inequality (mathematics)|inequalities]] relating these coordinates. This approach also has the advantage of easily allowing the generalization of geometry to Euclidean spaces of more than three dimensions.


After the introduction at the end of 19th century of [[non-Euclidean geometries]], the old postulates have been formalized for defining Euclidean spaces through an [[axiomatic theory]]. Another definition of Euclidean spaces by mean of [[vector space]]s and [[linear algebra]] has been shown to be equivalent to the axiomatic definitions. This is this definition that is more commonly used in modern mathematics, and detailed in this article.{{sfn|Berger|1987}}
From the modern viewpoint, there is essentially only one Euclidean space of each dimension. While Euclidean space is defined by a set of axioms, these axioms do not specify how the points are to be represented.<ref>{{cite web|last1=Gabi|first1=Aalex|title=What is the difference between Euclidean and Cartesian spaces?|url=https://math.stackexchange.com/q/112076|website=Mathematics Stack Exchange|publisher=Mathematics Stack Exchange}}</ref> Euclidean space can, as one possible choice of representation, be modeled using Cartesian coordinates. In this case, the Euclidean space is then modeled by the [[real coordinate space]] ({{math|'''R'''<sup>''n''</sup>}}) of the same dimension. In one dimension, this is the [[real line]]; in two dimensions, it is the [[Cartesian plane]]; and in higher dimensions it is a [[coordinate space]] with three or more real number coordinates. Mathematicians denote the [[n-dimensional space|{{mvar|n}}-dimensional]] Euclidean space by {{math|'''E'''<sup>''n''</sup>}} if they wish to emphasize its Euclidean nature, but {{math|'''R'''<sup>''n''</sup>}} is used as well since the latter is assumed to have the standard Euclidean structure, and these two [[structure (mathematics)|structures]] are not always distinguished. Euclidean spaces have finite dimension.<ref name="mathen">{{cite web|title=Euclidean space.|url=https://www.encyclopediaofmath.org/index.php/Euclidean_space|work=Encyclopedia of Mathematics|publisher=Springer|accessdate=1 May 2014|author=E.D. Solomentsev|date=7 February 2011}}</ref>


For all definitions, Euclidean spaces consist of points, which are defined only by the properties that they must have for forming a Euclidean space.
==History==

Euclidean space was introduced by [[Greek mathematics|ancient Greeks]] as an abstraction of our physical space. Their great innovation, appearing in [[Euclid's Elements|Euclid's ''Elements'']] was to build and ''[[proof (mathematics)|prove]]'' all geometry by starting from a few very basic properties, which are abstracted from the physical world, and are too basic for being mathematically proved. These properties are called [[postulate]]s, or [[axiom]]s in modern language. This way of defining Euclidean space is still in use under the name of [[synthetic geometry]].
There is essentially only one Euclidean space of each dimension; that is, all Euclidean spaces of a given dimension are [[isomorphism|isomorphic]]. Therefore, in many cases, it is possible to work with a specific Euclidean space, which is generally the [[real n-space|real {{mvar|n}}-space]] <math>\mathbb R^n,</math> equipped with the [[dot product]]. An isomorphism from a Euclidean space to <math>\mathbb R^n</math> associates to each point a [[tuple|{{mvar|n}}-tuple]] of [[real number]]s, which locate them in the Euclidean space and are called ''[[Cartesian coordinates]]''.

==Definition==
===History of the definition===
Euclidean space was introduced by [[Greek mathematics|ancient Greeks]] as an abstraction of our physical space. Their great innovation, appearing in [[Euclid's Elements|Euclid's ''Elements'']] was to build and ''[[proof (mathematics)|prove]]'' all geometry by starting from a few very basic properties, which are abstracted from the physical world, and cannot be mathematically proved because of the lack of more basic tools. These properties are called [[postulate]]s, or [[axiom]]s in modern language. This way of defining Euclidean space is still in use under the name of [[synthetic geometry]].


In 1637, [[René Descartes]] introduced [[Cartesian coordinates]] and showed that this allows reducing geometric problems to algebraic computations with numbers. This reduction of geometry to [[algebra]] was a major change of point of view, as, until then, the [[real number]]s—that is, [[rational number]]s and non-rational numbers together–were defined in terms of geometry, as lengths and distance.
In 1637, [[René Descartes]] introduced [[Cartesian coordinates]] and showed that this allows reducing geometric problems to algebraic computations with numbers. This reduction of geometry to [[algebra]] was a major change of point of view, as, until then, the [[real number]]s—that is, [[rational number]]s and non-rational numbers together–were defined in terms of geometry, as lengths and distance.


Euclidean geometry was not applied in spaces of more than three dimensions until the 19th century. [[Ludwig Schläfli]] generalized Euclidean geometry to spaces of ''n'' dimensions using both synthetic and algebraic methods, and discovered all of the regular [[Polytope|polytopes]] (higher-dimensional analogues of the [[Platonic solid|Platonic solids]]) that exist in Euclidean spaces of any number of dimensions.<ref>{{Cite book|title=Regular Polytopes|title-link=Regular Polytopes (book)|last=Coxeter|first=H.S.M.|publisher=Dover|year=1973|isbn=|edition=3rd|location=|publication-place=New York|pages=141-144|quote="Schläfli ... discovered them before 1853 -- a time when Cayley, Grassman and Möbius were the only other people who had ever conceived of the possibility of geometry in more than three dimensions."|author-link=Harold Scott MacDonald Coxeter|orig-year=1948}}</ref>
Euclidean geometry was not applied in spaces of more than three dimensions until the 19th century. [[Ludwig Schläfli]] generalized Euclidean geometry to spaces of ''n'' dimensions using both synthetic and algebraic methods, and discovered all of the regular [[Polytope|polytopes]] (higher-dimensional analogues of the [[Platonic solid|Platonic solids]]) that exist in Euclidean spaces of any number of dimensions.{{sfn|Coxeter|1973|}}


Despite the wide use of Descartes' approach, which was called [[analytic geometry]], the definition of Euclidean space remained unchanged until the end of 19th century. The introduction of abstract [[vector space]]s allowed their use in defining Euclidean spaces with a purely algebraic definition. This new definition has been shown to be equivalent to the classical definition in terms of geometric axioms. It is this algebraic definition that is now most often used for introducing Euclidean spaces.
Despite the wide use of Descartes' approach, which was called [[analytic geometry]], the definition of Euclidean space remained unchanged until the end of 19th century. The introduction of abstract [[vector space]]s allowed their use in defining Euclidean spaces with a purely algebraic definition. This new definition has been shown to be equivalent to the classical definition in terms of geometric axioms. It is this algebraic definition that is now most often used for introducing Euclidean spaces.


===Motivation of the modern definition===
==Intuitive overview==
One way to think of the Euclidean plane is as a [[set (mathematics)|set]] of [[point (geometry)|point]]s satisfying certain relationships, expressible in terms of distance and angle. For example, there are two fundamental operations (referred to as [[motion (geometry)|motions]]) on the plane. One is [[translation (geometry)|translation]], which means a shifting of the plane so that every point is shifted in the same direction and by the same distance. The other is [[rotation (mathematics)|rotation]] about a fixed point in the plane, in which every point in the plane turns about that fixed point through the same angle. One of the basic tenets of Euclidean geometry is that two figures (usually<!-- not always --> considered as [[subset]]s) of the plane should be considered equivalent ([[congruence (geometry)|congruent]]) if one can be transformed into the other by some sequence of translations, rotations and [[reflection (mathematics)|reflection]]s (see [[#Euclidean group|below]]).
One way to think of the Euclidean plane is as a [[set (mathematics)|set]] of [[point (geometry)|point]]s satisfying certain relationships, expressible in terms of distance and angles. For example, there are two fundamental operations (referred to as [[motion (geometry)|motions]]) on the plane. One is [[translation (geometry)|translation]], which means a shifting of the plane so that every point is shifted in the same direction and by the same distance. The other is [[rotation (mathematics)|rotation]] around a fixed point in the plane, in which all points in the plane turn around that fixed point through the same angle. One of the basic tenets of Euclidean geometry is that two figures (usually<!-- not always --> considered as [[subset]]s) of the plane should be considered equivalent ([[congruence (geometry)|congruent]]) if one can be transformed into the other by some sequence of translations, rotations and [[reflection (mathematics)|reflection]]s (see [[#Euclidean group|below]]).


In order to make all of this mathematically {{visible anchor|precise}}, the theory must clearly define the notions of distance, angle, translation, and rotation for a mathematically described space. Even when used in [[physics|physical]] theories, Euclidean space is an [[abstraction]] detached from actual physical locations, specific [[frame of reference|reference frames]], measurement instruments, and so on. A purely mathematical definition of Euclidean space also ignores questions of [[unit of length|units of length]] and other [[dimensional analysis|physical dimensions]]: the distance in a "mathematical" space is a [[number]], not something expressed in inches or metres. The standard way to define such space, as carried out in the remainder of this article, is to define the Euclidean plane as a two-dimensional [[real number|real]] [[vector space]] equipped with an [[inner product space|inner product]].<ref name="mathen"/> The reason for working with arbitrary vector spaces instead of {{math|'''R'''<sup>''n''</sup>}} is that it is often preferable to work in a ''coordinate-free'' manner (that is, without choosing a preferred basis). For then:
In order to make all of this mathematically precise, the theory must clearly define what is an Euclidean space, and the related notions of distance, angle, translation, and rotation. Even when used in [[physics|physical]] theories, Euclidean space is an [[abstraction]] detached from actual physical locations, specific [[frame of reference|reference frames]], measurement instruments, and so on. A purely mathematical definition of Euclidean space also ignores questions of [[unit of length|units of length]] and other [[dimensional analysis|physical dimensions]]: the distance in a "mathematical" space is a [[number]], not something expressed in inches or metres.
*the [[coordinate vector|vector]]s in the vector space correspond to the points of the Euclidean plane,
*the [[addition]] operation in the vector space corresponds to [[translation (geometry)|translation]], and
*the inner product implies notions of angle and distance, which can be used to define rotation.
Once the Euclidean plane has been described in this language, it is actually a simple matter to extend its concept to arbitrary dimensions. For the most part, the vocabulary, formulae, and calculations are not more difficult by the presence of more dimensions. (However, rotations are more subtle in high dimensions, and visualizing [[high-dimensional space]]s remains difficult, even for experienced mathematicians.)


The standard way to mathematically define an Euclidean space, as carried out in the remainder of this article, is to define an Euclidean space as a set of points on which [[group action (mathematics)|acts]] a [[real vector space]], the ''space of translations'' which is equipped with an [[inner product space|inner product]].{{sfn|Solomentsev|2001}} The action of translations makes the space an [[affine space]], and this allow defining lines, planes, subspaces, dimension, and [[parallelism (geometry)|parallelism]]. The inner product allows defining distance and angles.
A Euclidean space is not technically a [[vector space]] but rather an [[affine space]], on which a vector space [[Group action (mathematics)|acts]] by translations, or, conversely, a [[Euclidean vector]] is the [[subtraction|difference]] ([[displacement (vector)|displacement]]) in an ordered pair of points, not a single point. Intuitively, the distinction says merely that there is no [[canonical form|canonical]] choice of where the [[origin (mathematics)|origin]] should go in the space, because it can be translated anywhere. When a certain point is chosen, it can be declared the origin and subsequent calculations may ignore the difference between a point and its coordinate vector, as said above. See [[point–vector distinction]] for details.


The set <math>\mathbb R^n</math> of {{mvar|n}}-tuples of real numbers equipped with the [[dot product]] is a Euclidean space of dimension {{mvar|n}}. Conversely, the choice of a point called the ''origin'' and an [[orthonormal basis]] of the space of translations is equivalent with defining an [[isomorphism]] between a Euclidean space of dimension {{mvar|n}} and <math>\mathbb R^n</math> viewed as a Euclidean space.
==Technical definition==
{{cleanup|date=August 2019|reason=The definition should be given clearly and early; the current text just starts discussing.}}


It follows that everything that can be said about a Euclidean space can also be said about <math>\mathbb R^n.</math> Therefore, many authors, specially at elementary level, call <math>\mathbb R^n</math> the ''standard Euclidean space'' of dimension {{mvar|n}},{{sfn|Berger|1987|loc=Section 9.1}} or simply ''the'' Euclidean space of dimension {{mvar|n}}.
===Euclidean structure===
These are distances between points and the angles between lines or vectors, which satisfy certain conditions (see [[#conditions|below]]), which makes a set of points a Euclidean space. The natural way to obtain these quantities is by introducing and using the standard inner product (also known as the [[dot product]]) on {{math|'''R'''<sup>''n''</sup>}}.<ref name="mathen"/> The inner product of any two real {{mvar|n}}-vectors {{math|'''x'''}} and {{math|'''y'''}} is defined by


A reason for introducing such an abstract definition of Euclidean spaces, and for working with it instead of <math>\mathbb R^n</math> is that it is often preferable to work in a ''coordinate-free'' and ''origin-free'' manner (that is, without choosing a preferred basis and a preferred origin). Another reason is that there is no origin nor any basis in the physical world.
:<math>\mathbf{x}\cdot\mathbf{y} = \sum_{i=1}^n x_iy_i = x_1y_1+x_2y_2+\cdots+x_ny_n,</math>


===Technical definition===
where {{mvar|x<sub>i</sub>}} and {{mvar|y<sub>i</sub>}} are {{mvar|i}}th coordinates of vectors {{math|'''x'''}} and {{math|'''y'''}} respectively.
A '''Euclidean vector space''' is a finite-dimensional [[inner product space]] over the [[real number]]s.
The result is always a real number.


A '''Euclidean space''' is an [[affine space]] over the [[real number|reals]] such that the associated vector space is a Euclidean vector space. Euclidean spaces are sometimes called ''Euclidean spaces'' for distinguishing them from Euclidean vector spaces.{{sfn|Berger|1987|loc=Chapter 9}}
=== Distance ===

The ''dimension'' of a Euclidean space is the [[dimension (vector space)|dimension]] of its associated vector space.

If {{mvar|E}} is a Euclidean space, its associated vector space is also called its space of [[translation (geometry)|translation]]s, and often denoted <math>\overrightarrow E.</math>

The elements of {{mvar|E}} are called ''points'' and are commonly denoted by capital letters. The elements of <math>\overrightarrow E</math> are called ''translations'', ''[[Euclidean vector]]s'' or ''[[free vector]]s''.

The action of a translation {{mvar|v}} on a point {{mvar|P}} provides a point that is denoted {{math|''P'' + ''v''}}. This action satisfies
:<math>P+(v+w)= (P+v)+w.</math>
(The second {{math|+}} in the left-hand side is a vector addition; all other {{math|+}} denote an action of a vector on a point. This notation is not ambiguous, as, for distinguishing between the two meanings of {{math|+}}, it suffices to look on the nature of its left argument.)

The fact that the action is free and transitive means that for every pair of points {{math|(''P'', ''Q''}} there is exactly one vector {{mvar|v}} such that {{math|''P'' + ''v'' {{=}} ''Q''}}. This vector {{mvar|v}} is denoted {{math|''Q'' − ''P''}} or <math>\overrightarrow {PQ}.</math>

As previously explained, some of the basic properties of Euclidean spaces result of the structure of affine space. They are described in {{slink||Affine structure}} and its subsections. The properties resulting from the inner product are explained in {{slink||Metric structure}} and its subsections.

==Prototypical examples==
For any vector space, the addition acts freely and transitively on the vector space itself. Thus a Euclidean vector space can be viewed as a Euclidean space that has itself as associated vector space.

A typical case of Euclidean vector space is <math>\mathbb R^n</math> viewed as a vector space equipped with the [[dot product]] as an [[inner product]]. The importance of this particular example of Euclidean space lies is the fact that every Euclidean space is [[isomorphism|isomorphic]] to it. More precisely, given a Euclidean space {{mvar|E}} of dimension {{mvar|n}}, the choice of a point, called a ''origin'' and a [[orthonormal basis]] of <math>\overrightarrow E</math> defines an isomorphism of Euclidean spaces from {{mvar|E}} to <math>\mathbb R^n.</math>

As every Euclidean space of dimension {{mvar|n}} is isomorphic to it, the Euclidean space <math>\mathbb R^n</math> is sometimes called the ''standard Euclidean space'' of dimension {{mvar|n}}. {{sfn|Berger|1987|loc=Section 9.1}}

==Affine structure==
{{Main|Affine space}}

Some basic properties of Euclidean spaces depend only of the fact that an Euclidean space is an [[affine space]]. they are called [[affine property|affine properties]] and include the concepts of lines, subspaces, and parallelism. which are detailed in next subsections.

===Subspaces===
{{main|Flat (geometry)}}
Let {{mvar|E}} a Euclidean space, and <math>\overrightarrow E</math> its associated vector space.

A ''flat'', ''Euclidean subspace'' or ''affine subspace'' of {{mvar|E}} is a subset {{mvar|F}} of {{mvar|E}} such that
:<math>\overrightarrow F = \{\overrightarrow {PQ}\mid P\in F, Q\in F\}</math>
is a [[linear subspace]] of <math>\overrightarrow E.</math> An Euclidean subspace {{mvar|F}} is a Euclidean space, with <math>\overrightarrow F</math> as associated vector space. This linear subspace <math>\overrightarrow F</math> is called the ''direction'' of {{mvar|F}}.

If {{mvar|P}} is a point of {{mvar|F}} then
:<math>F=\{P+v\mid v\in \overrightarrow F\}.</math>
Conversely, if {{mvar|P}} is a point of {{mvar|E}}, and {{mvar|V}} is a [[linear subspace]] of <math>\overrightarrow E,</math> then
:<math>P+V=\{P+v\mid v\in V\}</math>
is a Euclidean subspace of direction {{mvar|V}}.

A Euclidean vector space (that is a Euclidean space such that <math>E=\overrightarrow R</math>) has two sorts of subspaces, its Euclidean subspaces and its linear subspaces. Linear subspaces are Euclidean subspaces, and a Euclidean subspace is a linear subspace if and only if it contains the zero vector.

===Lines and segments===
In a Euclidean space, a ''line'' is a Euclidean subspace of dimension one. Since a vector space of dimension one is spanned by any nonzero vector a line is a set of the form
:<math>\{P+\lambda \overrightarrow{PQ} \mid \lambda \in \mathbb R\},</math>
where {{mvar|P}} and {{mvar|Q}} are two distinct points.

It follows that ''there is exactly one line that passes through (contains) two distinct points.'' This implies that two distinct lines intersect in at most one point.

A more symmetric representation of the line passing through {{mvar|P}} and {{mvar|Q}} is
:<math>\{O + (1-\lambda)\overrightarrow{OP} + \lambda \overrightarrow{OQ} \mid \lambda \in \mathbb R\},</math>
where {{mvar|O}} is an arbitrary point (not necessary on the line).

In a Euclidean vector space, the zero vector is usually chosen for {{mvar|O}}; this allows simplifying the preceding formula into
:<math>\{(1-\lambda) P + \lambda Q \mid \lambda \in \mathbb R\}.</math>
A standard convention allows using this formula in every Euclidean space, see {{slink|Affine space|Affine combinations and barycenter}}.

The ''[[line segment]]'', or simply ''segment'', joining the points {{mvar|P}} and {{mvar|Q}} is the subset of the points such that {{math|0 ≤ ''{{lambda}}'' ≤ 1 }} in the preceding formulas. It is denoted {{mvar|PQ}} or {{mvar|QP}}; that is
:<math>PQ = QP = \{P+\lambda \overrightarrow{PQ} \mid 0 \le \lambda \le 1\}.</math>

===Parallelism===
Two subspaces {{mvar|S}} and {{mvar|T}} of the same dimension in a Euclidean space are ''parallel'' if they have the same direction.{{efn|It may depend on the context or the author whether a subspace is parallel to itself}} Equivalently, they are parallel, if there is a translation {{mvar|v}} vector that maps one to the other:
:<math>T= S+v.</math>

Given a point {{mvar|P}} and a subspace {{mvar|S}}, there exists exactly one subspace that contains {{mvar|P}} and is parallel to {{mvar|S}}, which is <math>P + \overrightarrow S.</math> In the case where {{mvar|S}} is a line (subspace of dimension one), this property is [[Playfair's axiom]].

It follows that in a Euclidean plane, two lines either meet in one point or are parallel.

The concept of parallel subspaces has been extended to subspaces of different dimensions: two subspaces are parallel if the direction of one of them is contained in the direction to the other.

==Metric structure==
The vector space <math>\overrightarrow E</math> associated to a Euclidean space {{mvar|E}} is an [[inner product space]]. This implies a [[symmetric bilinear form]]
:<math>\begin{align}
\overrightarrow E \times \overrightarrow E &\to \mathbb R\\
(x,y)&\mapsto \langle x,y \rangle
\end{align}</math>
that is [[positive definite]] (that is <math>\langle u,u \rangle</math> is always positive for {{math|''x'' ≠ 0}}).

The inner product of a Euclidean space is often called ''dot product'' and denoted {{math|''x'' ⋅ ''y''}}. This is specially the case when a [[Cartesian coordinate system]] has been chosen, as, in this case, the inner product of two vectors is the [[dot product]] of their [[coordinate vector]]s. For this reason, and for historical reasons, the dot notation is more commonly used than the bracket notation for the inner product of Euclidean spaces. This article will follow this usage; that is <math>\langle x,y \rangle</math> will be denoted {{math|''x'' ⋅ ''y''}} in the remainder of this article.

The '''Euclidean norm''' of a vector {{mvar|v}} is
:<math>\|x\| = \sqrt {x\cdot x}.</math>

The inner product and the norm allows expressing and proving all [[metric space|metric]] and [[topology|topological]] properties of [[Euclidean geometry]]. The next subsection describe the most fundamental ones. ''In these subsections,'' {{mvar|E}} ''denotes an arbitrary Euclidean space, and <math>\overrightarrow E</math> denotes its vector space of translations.''

=== Distance and length===
{{main article|Euclidean distance}}
{{main article|Euclidean distance}}
The ''distance'' (more precisely the ''Euclidean distance'') between two points of a Euclidean space is the norm of the translation vector that maps one point to the other; that is
The inner product [[square (algebra)|of {{math|'''x'''}} with itself]] is always [[non-negative]]. This product allows us to define the "length" of a vector {{math|'''x'''}} through [[square root]]:
:<math>d(P,Q)=\|\overrightarrow {PQ}\|.</math>
The ''length'' of a segment {{math|''PQ''}} is the distance {{math|''d''(''P'', ''Q'')}} between its endpoints. It is often denoted {{math|{{!}}''PQ''{{!}}}}.


The distance is a [[metric (mathematics)|metric]], as it satisfies the [[triangular inequality]]
:<math>\|\mathbf{x}\| = \sqrt{\mathbf{x}\cdot\mathbf{x}} = \sqrt{\sum_{i=1}^{n}x_i^2}.</math>
:<math>d(P,Q)\le d(P,R) + d(R, Q).</math>
Moreover, the equality is true if and only if {{mvar|R}} belongs to the segment {{math|''PQ''}}.
This inequality means that the length of any edge of a [[triangle]] is smaller than the sum of the lengths of the other edges. This is the origin of the term ''triangular inequality''.


With the Euclidean distance, every Euclidean space is a [[complete metric space]].
This length function satisfies the required properties of a [[norm (mathematics)|norm]] and is called the '''Euclidean norm''' on {{math|'''R'''<sup>''n''</sup>}}.


===Orthogonality===
Finally, one can use the norm to define a [[metric (mathematics)|metric]] (or distance function) on {{math|'''R'''<sup>''n''</sup>}} by
{{main|Perpendicular|Orthogonality}}
Two nonzero vectors {{mvar|u}} and {{mvar|v}} of <math>\overrightarrow E</math> are ''perpendicular'' or ''orthogonal'' if their inner product is zero:
:<math> u\cdot v =0</math>


Two linear subspaces of <math>\overrightarrow E</math> are orthogonal if every nonzero vector of the first one is perpendicular to every nonzero vector of the second one. This implies that the intersection of the linear subspace is reduced to the zero vector.
:<math>d(\mathbf{x}, \mathbf{y}) = \|\mathbf{x} - \mathbf{y}\| = \sqrt{\sum_{i=1}^n (x_i - y_i)^2}.</math>


Two lines, and more generally two Euclidean subspaces are orthogonal if their direction are orthogonal. Two orthogonal lines that intersect are said ''perpendicular''.
This distance function is called the [[Euclidean distance|Euclidean metric]]. This formula expresses a special case of the [[Pythagorean theorem]].


Two segments {{math|''AB''}} and {{math|''AC''}} that share a common endpoint are ''perpendicular'' or ''form a [[right angle]]'' if the vectors <math>\overrightarrow {AB}</math> and <math>\overrightarrow {AC}</math> are orthogonal.
This distance function (which makes a [[metric space]]) is sufficient to define all Euclidean geometry, including the dot product<!-- can be recovered from it through the [[cosine law]]-->. Thus, a real coordinate space together with this Euclidean structure is called '''Euclidean space'''. Its vectors form an [[inner product space]] (in fact a [[Hilbert space]]), and a [[normed vector space]].


If {{math|''AB''}} and {{math|''AC''}} form a right angle, one has
The metric space structure is the main reason behind the use of [[real number]]s {{math|'''R'''}}, not some other [[ordered field]], as the mathematical foundation of Euclidean (and many other) spaces. Euclidean space is a [[complete metric space]]<!-- explain -->, a property which is impossible to achieve operating over [[rational number]]s, for example.
:<math>|BC|^2 = |AB|^2 + |AC|^2.</math>
This is the [[Pythagorean theorem]]. Its proof is easy in this context, as, expressing this in terms of the inner product, one has, using bilinearity and symmetry of the inner product:
:<math>\begin{align}
|BC|^2 &= \overrightarrow {BC}\cdot \overrightarrow {BC}\\
&=\left(\overrightarrow {BA}+\overrightarrow {AC}\right ) \cdot \left(\overrightarrow {BA}+\overrightarrow {AC}\right)\\
&=\overrightarrow {BA}\cdot \overrightarrow {BA}+ \overrightarrow {AC}\cdot \overrightarrow {AC} -2 \overrightarrow {AB}\cdot \overrightarrow {AC}\\
&=\overrightarrow {AB}\cdot \overrightarrow {AB} + \overrightarrow {AC}\cdot\overrightarrow {AC}\\
&=|AB|^2 + |AC|^2.
\end{align}</math>


=== Angle ===
=== Angle ===
{{main article|Angle}}
{{main article|Angle}}
[[Image:45, -315, and 405 co-terminal angles.svg|thumb|right|Positive and negative angles on the oriented plane]]
[[Image:45, -315, and 405 co-terminal angles.svg|thumb|right|Positive and negative angles on the oriented plane]]
The ('''non-reflex''') '''angle''' {{mvar|θ}} ({{math|0° ''θ'' ≤ 180°}}) between vectors {{math|'''x'''}} and {{math|'''y'''}} is then given by
The (non-oriented) ''angle'' {{mvar|θ}} between two nonzero vectors {{mvar|x}} and {{mvar|y}} in <math>\overrightarrow E</math> is
:<math>\theta = \arccos\left(\frac{\mathbf{x}\cdot\mathbf{y}}{\|\mathbf{x}\|\|\mathbf{y}\|}\right)</math>
:<math>\theta = \arccos\left(\frac{x\cdot y}{\|x\|\|y\|}\right)</math>
where {{math|arccos}} is the [[principal value]] of the [[arccosine]] function. By [[Cauchy–Schwarz inequality]], the argument of the arcsine is in the interval {{math|[–1, 1]}}. Therefore {{mvar|{{theta}}}} is real, and {{math|0 ≤ ''θ'' ≤ ''{{pi}}''}} (or {{math|0 ≤ ''θ'' ≤ 180}}} if angles are measured in degrees).
where {{math|arccos}} is the [[arccosine]] function. It is useful only for {{math|''n'' > 1}},<ref group=footnote>On the [[real line]] ({{math|1=''n'' = 1}}) any two non-zero vectors are either parallel or [[antiparallel (mathematics)|antiparallel]] depending on whether their [[sign (mathematics)|signs]] match or oppose. There are no angles between 0 and 180°.</ref> and the case {{math|1=''n'' = 2}} is somewhat special. Namely, on an [[orientation (vector space)|oriented]] Euclidean plane one can define an angle between two vectors as a number defined [[modular arithmetic|modulo]] [[turn (geometry)|1 turn]] (usually denoted as either {{math|2[[pi|π]]}} or 360°), such that {{math|1=∠'''y''' '''x''' = −∠'''x''' '''y'''}}. This oriented angle is equal either to the angle {{mvar|θ}} from the formula above or to {{math|−''θ''}}. If one non-zero vector is fixed (such as the first basis vector), then each non-zero vector is uniquely defined by its magnitude and angle.


Angles are not useful in a Euclidean line, as they can be only 0 or ''{{pi}}''.
The angle does not change if vectors {{math|'''x'''}} and {{math|'''y'''}} are [[scalar multiplication|multiplied by positive numbers]].
<!-- distance on (''n'' − 1) -sphere -->


In an [[orientation (vector space)|oriented]] Euclidean plane, one can define the ''oriented angle'' of two vectors. The oriented angle of two vectors {{mvar|x}} and {{mvar|y}} is then the opposite of the oriented angle of {{mvar|y}} and {{mvar|x}}. In this case, the angle of two vectors can have any value [[modular arithmetic|modulo]] an integer multiple of {{math|2''{{pi}}''}}. In particular, a [[reflex angle]] {{math|''{{pi}}'' < ''{{theta}}'' < 2''{{pi}}''}} equals the negative angle {{math|–''{{pi}}'' < ''{{theta}}'' – 2''{{pi}}'' < 0}}.
Unlike the [[#precise|aforementioned situation]] with distance, the scale of angles is the same in pure mathematics, physics, and computing. It does not depend on the scale of distances; all distances may be multiplied by some fixed factor, and all angles will be preserved. Usually, the angle is considered a [[dimensionless quantity]], but there are different units of measurement, such as [[radian]] (preferred in pure mathematics and theoretical physics) and [[degree (angle)|degree]] (°) (preferred in most applications).


The angle of two vectors does not change if they are [[scalar multiplication|multiplied]] by positive numbers. More precisely, if {{mvar|x}} and {{mvar|y}} are two vectors, and {{mvar|{{lambda}}}} and {{mvar|{{mu}}}} are real numbers, then
=== Rotations and reflections ===
:<math>\operatorname{angle}(\lambda x, \mu y)= \begin{cases}
{{main article|Rotation (mathematics)|Reflection (mathematics)|Orthogonal group}}
\operatorname{angle}(x, y) \qquad\qquad \text{if } \lambda \text{ and } \mu \text{ have the same sign}\\
{{see also|rotational symmetry|reflection symmetry}}
\pi - \operatorname{angle}(x, y)\qquad \text{otherwise}.
Symmetries of a Euclidean space are transformations which preserve the Euclidean metric (called ''[[isometry|isometries]]''). Although [[#Intuitive overview|aforementioned translations]] are most obvious of them, they have the same structure for any affine space and do not show a distinctive character of Euclidean geometry. Another family of symmetries leave one point fixed, which may be seen as the origin without loss of generality. All transformations, which preserves the origin and the Euclidean metric, are [[linear map]]s.<!-- one can, by induction, prove that ℝ¹ can be isometrically embedded only as a line, ℝ² — as a plane, and so on, but should the article explain it? --> Such transformations {{mvar|Q}} must, for any {{math|'''x'''}} and {{math|'''y'''}}, satisfy:
\end{cases}</math>
:<math>Q\mathbf{x} \cdot Q\mathbf{y} = \mathbf{x} \cdot \mathbf{y}</math> ([[real coordinate space#Matrix notation|explain the notation]]),
:<math>|Q\mathbf{x}| = |\mathbf{x}|.</math>


If {{math|''A'', ''B''}} and {{mvar|C}} are three points in a Euclidean space, the angle of the segments {{math|''AB''}} and {{math|''AC''}} is the angle of the vectors <math>\overrightarrow {AB}</math> and <math>\overrightarrow {AC}.</math> As the multiplication of vectors by positive numbers do not change the angle, the angle of two [[half-line]]s with initial point {{mvar|A}} can be defined: it is the angle of the segments {{math|''AB''}} and {{math|''AC''}}, where {{mvar|B}} and {{mvar|C}} are arbitrary points, one on each half-line. Although this is less used, one can define similarly the angle of segments or half-lines that do not share an initial points.
Such transforms constitute a [[group (mathematics)|group]] called the ''[[orthogonal group]]'' {{math|O(''n'')}}. Its elements {{mvar|Q}} are exactly solutions of a [[matrix (mathematics)|matrix]] equation
:<math>Q^\mathsf{T} Q = Q Q^\mathsf{T} = I,</math>
where {{mvar|Q}}<sup>T</sup> is the [[transpose]] of {{mvar|Q}} and {{math|''I''}} is the [[identity matrix]].


The angle of two lines is defined as follows. If {{mvar|{{theta}}}} is the angle of two segments, one on each line, the angle of any two other segments, one on each line, is either {{mvar|{{theta}}}} or {{math|''{{pi}}'' – ''{{theta}}''}}. One of these angles is in the [[interval (mathematics)|interval]] {{math|[0, ''{{pi}}''/2]}}, and the other being in {{math|[''{{pi}}''/2, ''{{pi}}'']}}. The ''non-oriented angle'' of the two lines is the one in the interval {{math|[0, ''{{pi}}''/2]}}. In an oriented Euclidean plane, the ''oriented angle'' of two lines belongs to the interval {{math|[–''{{pi}}''/2, ''{{pi}}''/2]}}.
But a Euclidean space is [[orientation (vector space)|orientable]].<ref group=footnote>It is {{math|'''R'''<sup>''n''</sup>}} which is orient'''ed''' because of the [[total order|ordering]] of elements of the [[standard basis]]. Although an orientation is not an attribute of the Euclidean structure, there are only two [[orientation (vector space)|possible orientations]], and any linear automorphism either keeps orientation or reverses (swaps the two).</ref> Each of these transformations either preserves or reverses orientation [[real coordinate space#Orientation|depending on whether its determinant is +1 or −1]] respectively. Only transformations which preserve orientation, which form the ''special orthogonal'' group {{math|SO(''n'')}}, are considered (proper) [[rotation (mathematics)|rotations]]. This group has, as a [[Lie group]], the same dimension {{math|''n''(''n'' − 1) /2}} and is the [[identity component]] of {{math|O(''n'')}}.
{| align=right class=wikitable style="text-align:center;"
|-
!Group||Diffeo-<br>morphic<br>to||Isomorphic<br>to
|-
|{{math|SO(1)}}
| colspan=2 align=center |[[trivial group|{1}]]
|-
|{{math|SO(2)}}||[[circle|{{math|''S''<sup>1</sup>}}]]||[[circle group|{{math|U(1)}}]]
|-
|[[Rotation group SO(3)|{{math|SO(3)}}]]||[[real projective space|{{math|'''RP'''<sup>3</sup>}}]]||{{math|[[special unitary group|SU]](2) / {±1} }}
|-
|[[SO(4)|{{math|SO(4)}}]]||{{math|([[3-sphere|''S''<sup>3</sup>]] [[Cartesian product|×]] ''S''<sup>3</sup>) [[double cover (topology)|/ {±1}]]}}||{{math|(SU(2) × SU(2)) / {±1} }}
|-
| colspan=3 style="font-size:87%" |Note: elements of {{math|SU(2)}} are also known as [[versor]]s.
|}


===Cartesian coordinates===
Groups {{math|SO(''n'')}} are well-studied for {{math|''n'' ≤ 4}}. There are no non-trivial rotations in 0- and 1-spaces. Rotations of a Euclidean plane ({{math|1=''n'' = 2}}) are parametrized by the [[#Angle|angle (modulo 1 turn)]]. Rotations of a 3-space are parametrized with [[axis–angle representation|axis and angle]], whereas a rotation of a 4-space is a [[function composition|superposition]] of two 2-dimensional rotations around perpendicular planes.
{{See also|Cartesian coordinate system}}


Every Euclidean vector space has an [[orthonormal basis]] (in fact, infinitely many in dimension higher than one, and two in dimension one), that is a [[basis (vector space)|basis]] <math>(e_1, \dots, e_n) </math> of [[unit vector]]s (<math>\|e_i\| = 1</math>) that are pairwise orthogonal (<math>e_i\cdot e_j</math> for {{math|''i'' ≠ ''j''}}). More precisely, given any [[basis (vector space)|basis]] <math>(b_1, \dots, b_n),</math> the [[Gram–Schmidt process]] computes an orthonormal basis such that, for every {{mvar|i}}, the [[linear span]]s of <math>(e_1, \dots, e_n)</math> and <math>(b_1, \dots, b_n)</math> are equal.
Among linear transforms in {{math|O(''n'')}} which reverse the orientation are [[reflection (mathematics)|hyperplane reflections]]. This is the only possible case for {{math|''n'' ≤ 2}}, but starting from three dimensions, such isometry in the [[general position]] is a [[rotoreflection]].


Given a Euclidean space {{mvar|E}}, a ''Cartesian frame'' is a set of data consisting of an orthonormal basis of <math>\overrightarrow E,</math> and a point of {{mvar|E}}, called the ''origin'' and often denoted {{mvar|O}}. A Cartesian frame <math>(O, e_1, \dots, e_n)</math> allows defining Cartesian coordinates for both {{mvar|E}} and <math>\overrightarrow E</math> in the following way.
=== Euclidean group ===
{{main article|Euclidean group}}
The Euclidean [[group (mathematics)|group]] {{math|E(''n'')}}, also referred to as the group of all [[isometry|isometries]] {{math|ISO(''n'')}}, treats [[translation (geometry)|translations]], rotations, and reflections in a uniform way, considering them as [[Group action (mathematics)|group action]]s in the context of [[group theory]], and especially in [[Lie group]] theory. These group actions preserve the Euclidean structure.


The Cartesian coordinates of a vector {{mvar|v}} are the coefficients of {{mvar|v}} on the basis <math>e_1, \dots, e_n.</math> As the basis is orthonormal, the {{mvar|i}}th coefficient is the dot product <math>v\cdot e_i.</math>
As the group of all isometries, {{math|ISO(''n'')}}, the Euclidean group is important because it makes Euclidean geometry a case of [[Klein geometry]], a theoretical framework including many [[#Alternatives and generalizations|alternative geometries]].


The Cartesian coordinates of a point {{mvar|P}} of {{mvar|E}} are the Cartesian coordinates of the vector <math>\overrightarrow {OP}.</math>
The structure of Euclidean spaces – distances, lines, vectors, angles ([[up to]] sign), and so on – is [[invariant (mathematics)|invariant]] under the transformations of their associated Euclidean group. For instance, translations form a [[abelian group|commutative]] [[subgroup]] that [[transitive action|acts freely and transitively]] on {{math|'''E'''<sup>''n''</sup>}}, while the [[stabilizer (group theory)|stabilizer]] of any point there is the [[#Rotations and reflections|aforementioned {{math|O(''n'')}}]].


=== Other coordinates ===
Along with translations, rotations, reflections, as well as the [[identity transformation]], Euclidean motions comprise also [[glide reflection]]s (for {{math|''n'' ≥ 2}}), [[screw axis|screw operations]] and rotoreflections (for {{math|''n'' ≥ 3}}), and even more complex combinations of primitive transformations for {{math|''n'' ≥ 4}}.
[[Image:Repere espace.png|thumb|right|192px|3-dimensional skew coordinates]]
{{main article|Coordinate system}}


As a Euclidean space is an [[affine space]], one can consider an [[affine frame]] on it, which is the same as a Euclidean frame, except that the basis is not required to be orthonormal. This define [[affine coordinates]], sometimes called ''skew coordinates'' for emphasizing that the basis vectors are not pairwise orthogonal.
{{anchor|conditions}}The group structure determines which conditions a [[metric space]] needs to satisfy to be a Euclidean space:


An [[affine basis]] of a Euclidean space of dimension {{mvar|n}} is a set of {{math|''n'' + 1}} points that are not contained in a hyperplan. An affine basis define [[barycentric coordinates]] for every point.
# Firstly, a metric space must be translationally invariant with respect to some (finite-dimensional) real vector space. This means that the space itself is an [[affine space]], that the space is ''flat'', not [[curved space|curved]], and points do not have different properties, and so any point can be translated to any other point.
# Secondly, the metric must correspond [[#Distance|in the aforementioned way]] to some positive-defined [[quadratic form]] on this vector space, because point stabilizers have to be isomorphic to {{math|O(''n'')}}. <!-- this still needs some clarification, because a Euclidean metric whose values are deformed by a convex function is not a Euclidean space, although has the same isometries group -->


Many other coordinates systems can be defined on a Euclidean space {{mvar|E}} of dimension {{mvar|n}}, in the following way. Let {{mvar|f}} be a [[homeomorphism]] (or, more often, a [[diffeomorphism]]) from a [[dense subset|dense]] [[open subset]] of {{mvar|E}} to an open subset of <math>\mathbb R^n.</math> The ''coordinates'' of a point {{mvar|x}} of {{mvar|E}} are the components of {{math|''f''(''x'')}}. The [[polar coordinate system]] (dimension 2) and the [[spherical coordinate system|spherical]] and [[cylindrical coordinate system|cylindrical]] coordinate systems (dimension 3) are defined this way.
== Non-Cartesian coordinates ==
[[Image:Repere espace.png|thumb|right|192px|3-dimensional skew coordinates]]
{{main article|Coordinate system}}


For points that are outside the domain of {{mvar|f}}, coordinates may sometimes be defined as the limit of coordinates of neighbour points, but these coordinates may be not uniquely defined, and may be not continuous in the neighborhood of the point. For example, for the spherical coordinate system, the longitude is not defined at the pole, and on the [[antimeridian]], the longitude passes discontinuousely from –180° to +180°.
Cartesian coordinates are arguably the standard, but not the only possible coordinate system for a Euclidean space.
[[Affine coordinates]] and [[Barycentric coordinate system|barycentric coordinates]] are compatible with the affine structure of {{math|'''E'''<sup>''n''</sup>}}, but make formulae for angles and distances more complicated.


This way of defining coordinates extends easily to other mathematical structures, and in particular to [[manifold]]s.
Another approach, which goes in line with ideas of [[differential geometry]] and [[conformal geometry]], is [[orthogonal coordinates]], where [[coordinate hypersurface]]s of different coordinates are [[orthogonal]], although [[curvilinear coordinates|curved]]. Examples include the [[polar coordinate system]] on Euclidean plane, the second important plane coordinate system.


== Isometries ==
See [[#Curved spaces|below]] about expression of the Euclidean structure in curvilinear coordinates.
An [[isometry]] between two [[metric space]]s is a bijection preserving the distance,{{efn|If the condition of being a bijection is removed, a function preserving the distance is necessarily injective, and is an isometry from its domain to its image.}} that is
{{clear}}
:<math>d(f(x), f(y))= d(x,y).</math>


In the case of a Euclidean vector space, an isometry preserves the norm
== Geometric shapes ==
:<math>\|f(x)\| = \|x\|,</math>
[[Image:Parabolic coords.svg|thumb|240px|[[Parabolic coordinates]]]]
since the norm of a vector is its distance from the zero vector. It preserves also the inner product
[[Image:Quadray.gif|thumb|right|192px|<span style="font-size:90%">[[Barycentric coordinates (mathematics)|Barycentric coordinates]] in 3-dimensional space: four coordinates are related with one [[linear equation]]</span>]]
:<math>f(x)\cdot f(y)=x\cdot y,</math>
{| align=right style="clear:left; margin-top:-8px"
since
| style="font-size:80%" |Polar<br>coordi-<br>nates:<br>see<br>[[#Angle|Angle]]<br>above
:<math>x.y=\frac 12 \left((\|x+y\|^2-\|x\|^2-\|y\|^2\right).</math>
|[[Image:Polar concept introduction.svg|160px]]
|}
[[Image:Linear subspaces with shading.svg|thumb|right|Three mutually [[transversality (mathematics)|transversal]] planes in the 3-dimensional space and their intersections, three lines]]
{{see also|List of mathematical shapes}}


An isometry Euclidean vector spaces is a [[linear isomorphism]].{{efn|Proof: one must prove that <math>f(\lambda x+ \mu y)-\lambda f(x)-\mu f(y)=0</math>. For that, it suffices to prove that the square of the norm of the left-hand side is zero. Using the bilinearity of the inner product, this squared norm can be expanded into a linear combination of <math>\|f(x)\|^2,</math> <math>\|f(y)\|^2,</math> and <math>f(x)\cdot f(y).</math> As {{mvar|f}} is an isometry, this gives a linear combination of <math>\|x\|^2,\|y\|^2,</math> and <math>x\cdot y,</math> which simplifies to zero.}}{{sfn|Berger|1987|loc=Proposition 9.1.3}}
=== Lines, planes, and other subspaces ===
{{main article|Flat (geometry)}}
The simplest (after points) objects in Euclidean space are [[flat (geometry)|flats]], or Euclidean ''subspaces'' of lesser dimension. Points are 0-dimensional flats, 1-dimensional flats are called ''[[line (geometry)|(straight) lines]]'', and 2-dimensional flats are ''[[plane (geometry)|planes]]''. {{math|(''n'' − 1)}}-dimensional flats are called ''[[hyperplane]]s''.


An isometry <math>f\colon E\to F</math> of Euclidean spaces defines an isometry
Any two distinct points lie on exactly one line. Any line and a point outside it lie on exactly one plane. More generally, the properties of flats and their [[incidence (geometry)|incidence]] of Euclidean space are shared with [[affine geometry]], whereas the affine geometry is devoid of distances and angles.
<math>\overrightarrow f\colon \overrightarrow E\to \overrightarrow F</math> of the associated Euclidean vector spaces. This implies that two isometric Euclidean spaces have the same dimension. Conversely, if {{math|E}} and {{mvar|F}} are Euclidean spaces, {{math|''O'' ∈ ''E''}}, {{math|''O''{{prime}} ∈ ''F''}}, and <math>\overrightarrow f\colon \overrightarrow E\to \overrightarrow F</math> is an isometry, then the map <math>f\colon E\to F</math> defined by
:<math>f(P)=O' + \overrightarrow f\left(\overrightarrow{OP}\right)</math>
is an isometry of Euclidean spaces.


It follows from the preceding results that an isometry of Euclidean spaces maps lines to lines, and, more generally Euclidean subspaces to Euclidean subspaces of the same dimension, and that the restriction of the isometry on these subspaces are isometries of these subspaces.
=== Line segments and triangles ===
{| align=right width=288px style="margin-left:1em"
| style="font-size:87%" |[[Image:Triangle angles sum to 180 degrees.svg|thumb|left|145px]]
The [[sum of angles of a triangle]] is an important problem, which exerted a great influence on 19th-century mathematics. In a Euclidean space it invariably equals 180[[degree (angle)|°]], or a half-turn
|}
{{main article|Line segment|Triangle geometry}}
This is not only a line which a pair {{math|(''A'', ''B'')}} of distinct points defines. Points on the line which lie between {{mvar|A}} and {{mvar|B}}, together with {{mvar|A}} and {{mvar|B}} themselves, constitute a [[line segment]] {{math|''A'' ''B''}}. Any line segment has the [[length]], which equals to distance between {{mvar|A}} and {{mvar|B}}. If {{math|1=''A'' = ''B''}}, then the segment is [[degeneracy (mathematics)|degenerate]] and its length equals to 0, otherwise the length is positive.


===Isometry with prototypical examples===
A (non-degenerate) [[triangle]] is defined by [[3 (number)|three]] points not lying on the same line. Any triangle lies on one plane. The concept of triangle is not specific to Euclidean spaces, but Euclidean triangles have numerous special properties<!--, such as [[triangle inequality]]--> and define many derived objects.
If {{mvar|E}} is a Euclidean space, it associated vector space <math>\overrightarrow E</math> can be considered as a Euclidean space. Every point {{math|''O'' ∈ ''E''}} defines an isometry of Euclidean spaces
:<math>P\mapsto \overrightarrow {OP},</math>
which maps {{mvar|O}} to the zero vector and has the identity as associated linear map. The inverse isometry is the map
:<math>v\mapsto O+v.</math>


A Euclidean frame {{tmath|(O, e_1, \dots, e_n)}} allows defining the map
A triangle can be thought of as a 3-gon on a plane, a special (and the first meaningful in Euclidean geometry) case of a [[polygon]].
:<math>\begin{align}
{{clear}}
E&\to \mathbb R^n\\
P&\mapsto \left(e_1\cdot \overrightarrow {OP}, \dots, e_n\cdot\overrightarrow {OP}\right),
\end{align}</math>
which is an isometry of Euclidean spaces. The inverse isometry is
:<math>\begin{align}
\mathbb R^n&\to E \\
(x_1\dots, x_n)&\mapsto \left(O+x_1e_1+ \dots + x_ne_n\right).
\end{align}</math>


''This means that, up to an isomorphism, there is exactly one Euclidean space of a given dimension.''
=== Polytopes and root systems ===
{| align=right class=wikitable width=260px
|-
| colspan="3"| The [[Platonic solid]]s are the five polyhedra that are most regular in a combinatoric sense, but also, their symmetry groups are embedded into {{math|O(3)}}<br>{{further|point groups in three dimensions}}
|- style="vertical-align: top; text-align: center;"
|[[File:Polyhedron pair 4-4.png|128px]]<br>Pair of dual [[tetrahedron|tetrahedra]]
|[[File:Polyhedron pair 6-8.png|128px]]<br>[[Cube]] and [[octahedron]]
|[[File:Polyhedron pair 12-20 max.png|128px]]<br>[[regular dodecahedron|Dodecahedron]] and [[regular icosahedron|icosahedron]]
|}
{{main article|Polytope|Root system}}
{{see also|List of polygons, polyhedra and polytopes|List of regular polytopes}}
Polytope is a concept that generalizes [[polygon]]s on a plane and [[polyhedra]] in 3-dimensional space (which are among the earliest studied geometrical objects). A [[simplex]] is a generalization of a line segment (1-simplex) and a triangle (2-simplex). A [[tetrahedron]] is a 3-simplex.


This justifies that many authors talk of <math>\mathbb R^n</math> as ''the'' Euclidean space of dimension {{mvar|n}}.
The concept of a polytope belongs to [[affine geometry]], which is more general than Euclidean. But Euclidean geometry distinguish ''[[regular polytope]]s''. For example, affine geometry does not see the difference between an [[equilateral triangle]] and a [[right triangle]], but in Euclidean space the former is regular and the latter is not.


=== Euclidean group ===
Root systems are special sets of Euclidean vectors. A root system is often identical to the set of vertices of a regular polytope.
{{main article|Euclidean group|Rigid transformation}}
An isometry from a Euclidean space onto itself is called ''Euclidean isometry'', ''Euclidean transformation'' or ''rigid transformation''. The rigid transformations of a Euclidean space form a group (under [[function composition|composition]]), called the ''Euclidean group'' and often denoted {{math|E(''n'')}} of {{math|ISO(''n'')}}.

The simplest Euclidean transformations are [[translation (mathematics)|translations]]
:<math>P \to P+v.</math>
They are in bijective correspondence with vectors. This is a reason for calling ''space of translations'' the vector space associated to a Euclidean space. The translations form a [[normal subgroup]] of the Euclidean group.

A Euclidean isometry {{mvar|f}} of a Euclidean space {{mvar|E}} defines a linear isometry <math>\overrightarrow f</math> of the associated vector space (by ''linear isometry'', it is meant an isometry that is also a [[linear map]]) in the following way: denoting by {{math|''Q'' – ''P''}} the vector <math>\overrightarrow {PQ}</math>, if {{mvar|O}} is an arbitrary point of {{mvar|E}}, one has
:<math>\overrightarrow f(\overrightarrow {OP})= f(P)-f(O).</math>
It is straightforward to prove that this is a linear map that does not depend from the choice of {{mvar|O.}}

The map <math>f \to \overrightarrow f</math> is a [[group homomorphism]] from the Euclidean group onto the group of linear isometries, called the [[orthogonal group]]. The kernel of this homomorphism is the translation group, showing that it is a normal subgroup of the Euclidean group.

The isometries that fix a given point {{mvar|P}} form the [[stabilizer subgroup]] of the Euclidean group with respect to {{mvar|P}}. The restriction to this stabilizer of above group homomorphism is an isomorphism. So the isometries that fix a given point form a group isomorphic to the orthogonal group.

Let {{mvar|P}} be a point, {{mvar|f}} an isometry, and {{mvar|t}} the translation that maps {{mvar|P}} to {{math|''f''(''P'')}}. The isometry <math>g=t^{-1}\circ f</math> fixes {{mvar|P}}. So <math>f= t\circ g,</math> and ''the Euclidean group is the [[semidirect product]] of the translation group and the orthogonal group.''

The [[special orthogonal group]] is the normal subgroup of the orthogonal group that preserves [[orientation (vector space)|handeness]]. It is a subgroup of [[index (group theory)|index]] two ion the orthogonal group. Its inverse image by the group homomorphism <math>f \to \overrightarrow f</math> is a normal subgroup of index two of the Euclidean group, which is called the ''special Euclidean group'' or the ''displacement group''. Its elements are called ''rigid motions'' or ''displacements''.

Rigid motions include the [[identity function|identity]], translations, [[rotation]]s (the rigid motions that fix at least a point), and also [[screw axis|screw motions]].


Typical examples of rigid transformations that are not rigid motions are [[reflection (mathematics)|reflection]]s, which are rigid transformations that fix a hyperplane and are not the identity. They are also the transformations consisting in changing the sign of one coordinate over some Euclidean frame.
{| class=wikitable width=320
|- valign=top
|[[Image:Root system G2.svg|160px]]<br>The root system [[G2 (mathematics)|G<sub>2</sub>]]
|[[Image:Up2 2 31 t0 E7.svg|160px]]<br>An orthogonal projection of the [[2 31 polytope|2<sub>31</sub> polytope]], whose vertices are elements of the [[E7 (mathematics)|E<sub>7</sub>]] root system
|}


As the special Euclidean group is a subgroup of index two of the Euclidean group, given a reflection {{mvar|r}}, every rigid transformation that is not a rigid motion is the product of {{mvar|r}} and a rigid motion. A [[glide reflection]] is an example of a rigid transformation that is not a rigid motion not a reflection.
=== Curves ===
{{main article|Euclidean geometry of curves}}
{{see also|List of curves}}
{{expand section|date=April 2013}}


All groups that have been considered in this section are [[Lie group]]s and [[algebraic group]]s.
=== Balls, spheres, and hypersurfaces ===
{{main article|Ball (mathematics)|Hypersurface}}
{{see also|n-sphere|List of surfaces}}
{{expand section|date=April 2013}}


==Topology==
==Topology==
{{main article|Real coordinate space #Topological properties}}
{{main article|Real n-space #Topological properties}}
Since Euclidean space is a [[metric space]], it is also a [[topological space]] with the [[natural topology]] induced by the metric. The metric topology on {{math|'''E'''<sup>''n''</sup>}} is called the '''Euclidean topology''', and it is identical to the [[standard topology]] on {{math|'''R'''<sup>''n''</sup>}}. A set is [[open set|open]] [[if and only if]] it contains an [[open ball]] around each of its points; in other words, open balls form a [[base (topology)|base of the topology]]. The [[topological dimension]] of the Euclidean {{mvar|n}}-space equals {{mvar|n}}, which implies that spaces of different dimension are not [[homeomorphic]]. A finer result is the [[invariance of domain]] which proves that any subset of {{mvar|n}}-space that is (with its [[subspace topology]]) homeomorphic to an open subset of {{mvar|n}}-space is itself open.


The Euclidean distance makes a Euclidean space a [[metric space]], and thus a [[topological space]. This topology is called the [[Euclidean topology]]. In the case of <math>\mathbb R^n,</math> this topology is also the [[product topology]].
==Applications==
Aside from countless uses in fundamental mathematics, a Euclidean model of the physical [[space]] can be used to solve many practical problems with sufficient precision. Two usual approaches are a fixed, or ''stationary'' [[frame of reference|reference frame]] (i.e. the description of a [[motion (physics)|motion of objects]] as their [[position (vector)|positions]] that [[continuous function|change continuously]] with [[time]]), and {{anchor|Galilean}}the use of [[Galilean relativity|Galilean space-time symmetry]] (such as in [[Newtonian mechanics]]). To both of them the modern Euclidean geometry provides a convenient formalism; for example, the space of Galilean [[velocity|velocities]] is itself a Euclidean space (see [[relative velocity]] for details).


The [[open set]]s are the subsets that contains an [[open ball]] around each of their points. In other words, open balls form a [[base (topology)|base of the topology]].
[[Topographical map]]s and [[technical drawing]]s are [[two-dimensional space|planar]] Euclidean. An idea behind them is the [[scale invariance]] of Euclidean geometry, that permits to represent large objects in a small sheet of paper, or a screen.


The [[topological dimension]] of a Euclidean space equals its dimension. This implies that Euclidean spaces of different dimensions are not [[homeomorphic]]. Moreover, the theorem of [[invariance of domain]] asserts that a subset of a Euclidean space is open (for the [[subspace topology]]) if and only if it is homeomorphic to an open subset of a Euclidean space of the same dimension.
==Alternatives and generalizations==

<!-- caution: there are two links to this header from above part of the article -->
Euclidean spaces are [[complete metric|complete]] and [[locally compact]]. That is, a closed subset of a Euclidean space is compact if it is [[bounded set|bounded]] (that is, contained in a ball). In particular, closed balls are compact.
Although Euclidean spaces are no longer considered to be the only possible setting for a [[geometry]], they act as prototypes for other geometric objects. Ideas and terminology from Euclidean geometry (both traditional and analytic) are pervasive in modern mathematics, where other geometric objects share many similarities with Euclidean spaces, share part of their structure, or embed Euclidean spaces.

==Axiomatic definitions==
The definition of Euclidean spaces that has been described in this article differs fundamentally of [[Euclid]]'s one. In reality, Euclid did not define formally the space, because it was thought as a description of the physical world that exists independently of human mind. The need of a formal definition appeared only at the end of 19th century, with the introduction of [[non-Euclidean geometries]].

Two different approach have been used. [[Felix Klein]] suggested to define geometries through their [[symmetries]]. The presentation of Euclidean spaces given in this article, is essentially issued from his [[Erlangen program]], with the emphasis given on the groups of translations and isometries.

On the other hand, [[David Hilbert]] proposed a set of [[Hilbert's axioms|axioms]], inspired by [[Euclid's postulates]]. They belong to [[synthetic geometry]], as they do not involve any definition of [[real number]]s. Later [[G. D. Birkhoff]] and [[Alfred Tarski]] proposed simpler sets of axioms, which use [[real number]]s (see [[Birkhoff's axioms]] and [[Tarski's axioms]]).

In ''[[Geometric Algebra]]'', [[Emil Artin]] has proved that all these definitions of a Euclidean space are equivalent.{{sfn|Artin|1988}} It is rather easy to prove that all definitions of Euclidean spaces satisfy Hilbert's axioms, and that those involving real numbers (including the above given definition) are equivalent. The difficult part of Artin's proof is the following. In Hilbert's axioms, [[congruence (geometry)|congruence]] is an [[equivalence relation]] on segments. One can thus define the ''length'' of a segment as its equivalence class. One must thus prove that this length satisfies properties that characterize nonnegative real numbers. It is what did Artin, with axioms that are not Hilbert's ones, but are equivalent.

==Usage==
Since [[Greek mathematics|ancien Greeks]], Euclidean space is used for modeling [[shape]]s in the physical world. It is thus used in many [[science]]s such as [[physics]], [[mechanics]], [[astronomy]], ... It is also widely used in all technical areas that are concerned with shapes, figure, location and position, such as [[architecture]], [[geodesy]], [[topography]], [[navigation]], [[industrial design]], [[technical drawing]], [[robotics]]...

[[Space-time]] of [[general relativity]] is not a Euclidean space, but it is a [[manifold]], and, as such, is locally approximated by Euclidean spaces.

Space of dimensions higher than three occurs in several modern theories of physics; see [[Higher dimension]]. They occur also in [[configuration space (physics)|configuration space]]s of [[physical system]]s.

Beside [[Euclidean geometry]], Euclidean spaces are also widely used in mathematics. [[Tangent space]]s of [[differentiable manifold]]s are Euclidean vector spaces. More generally, a [[manifold]] is a space that is locally a approximated by Euclidean spaces. Most [[non-Euclidean geometries]] can be modeled by a manifold, and [[embedding|embedded]] in a Euclidean space of higher dimension. For example, an [[elliptic space]] can be modeled by an [[ellipsoid]]. It is common to represent in a Euclidean space mathematics objects that are ''a priori'' not of a geometrical nature. An example among many is the usual representation of [[Graph (discrete mathematics)|graphs]].

==Other geometric spaces==
Since the introduction, at the end of 19th century, of [[Non-Euclidean geometries]], many sorts of spaces have been considered, about which one can do geometric reasoning in the same way as with Euclidean spaces. In general, they share some properties with Euclidean spaces, but may also have properties that could appear as rather strange. Some of these spaces use Euclidean geometry for their definition, or can be modeled as subspaces of a Euclidean space of higher dimension. When such a space is defined by geometrical [[axiom]]s, [[embedding]] the space in a Euclidean space is a standard way for proving [[consistency]] of its definition, or, more precisely for proving that its theory is consistent, if [[Euclidean geometry]] is consistent (which cannot be proved).

===Affine space===
{{main|Affine space}}
A Euclidean space is an affine space equipped with a [[metric (mathematics)|metric]]. Affine spaces have many other uses in mathematics. In particular, as they are defined over any [[field (mathematics)|field]], they allow doing geometry in other contexts.

As soon as non-linear questions are considered, it is generally useful to consider affine spaces over the [[complex number]]s as an extension of Euclidean spaces. For example, a [[circle]] and a [[line (geometry)|line]] have always two intersection points (possibly not distinct) in the complex affine space. Therefore, most of [[algebraic geometry]] is built in complex affine spaces and affine spaces over [[algebraically closed field]]s. The shapes that are studied in algebraic geometry in these affine spaces are therefore called [[affine algebraic variety|affine algebraic varieties]].

Affine spaces over the [[rational number]]s and more generally over [[algebraic number field]]s provide a link between (algebraic) goemetry and [[number theory]]. For example, the [[Fermat's Last Theorem]] can be stated "a [[Fermat curve]] of degree higher than two has no point in the affine plane over the rationals."

Geometry in affine spaces over a [[finite fields]] has also been widely studied. For example, [[elliptic curve]]s over finite fields are widely used in [[cryptography]].

===Projective space===
{{main|Projective space}}
Originally, projective spaces have been introduced by adding "[[points at infinity]]" to Euclidean spaces, and, more generally to affine spaces, in order to make true the assertion "two [[coplanar]] lines meet in exactly in one point". Projective space share with Euclidean and affine spaces the property of being [[isotropic]], that is, there is no property of the space that allows distinguishing between two points or two lines. Therefore, a more isotropic definition is commonly used, which consists as defining a projective space as the set of the [[vector line]]s in a [[vector space]] of dimension one more.

As for affine spaces, projective spaces are defined over any [[field (mathematics)|field]], and are fundamental spaces of [[algebraic geometry]].

===Non-Euclidean geometries===
{{main|Non-Euclidean geometry}}
''Non-Euclidean geometry'' refers usually to geometrical spaces where the [[parallel postulate]] is false.
They include [[elliptic geometry]], where the sum of the angles of a triangle is less than 180°, and [[hyperbolic geometry]], where this sum is more than 180°. Their introduction in the second half of 19th century, and the proof that their theory is [[consistency|consistent]] (if Euclidean geometry is not contradictory) is one of the paradoxes that are at the origin of the [[foundational crisis in mathematics]] of the beginning of 20th century, and motivated the systematization of [[axiomatic theory|axiomatic theories]] in mathematics.


=== Curved spaces ===
=== Curved spaces ===
{{main article|Riemannian geometry}}
{{main article|Manifold|Riemannian manifold}}

A [[smooth manifold]] is a [[Hausdorff space|Hausdorff]] topological space that is locally [[diffeomorphic]] to Euclidean space. Diffeomorphism does not respect distance and angle, but if one additionally prescribes a smoothly varying inner product on the manifold's [[tangent space]]s, then the result is what is called a [[Riemannian manifold]]. Put differently, a {{visible anchor|Riemannian}} manifold is a space constructed by deforming and patching together Euclidean spaces. Such a space enjoys notions of distance and angle, but they behave in a [[curvature|curved]], non-Euclidean manner. The simplest Riemannian manifold, consisting of {{math|'''R'''<sup>''n''</sup>}} with a constant inner product, is essentially identical to Euclidean {{mvar|n}}-space itself. Less trivial examples are [[n-sphere|{{mvar|n}}-sphere]] and [[hyperbolic space]]s. Discovery of the latter in the 19th century was branded as the [[non-Euclidean geometry]].
A [[manifold]] is a space that in the neighborhood of each point resemble to a Euclidean space. In technical terms, a manifold is a [[topological space]], such that each point has a [[neighborhood]] that is [[homeomorphic]] to an [[open subset]] of a Euclidean space. Manifold can be classify, by increasing degree of this "resemblance" into [[topological manifold]]s, [[differentiable manifold]]s, [[smooth manifold]]s, and [[analytic manifold]]s. However, none of these types of "resemblance" respect distances and angles, even approximately.

Distances and angles can be defined on a smooth manifold by providing a [[smooth function|smoothly varying]] Euclidean metric on the [[tangent space]]s at the points of the manifold (these tangent are thus Euclidean vector spaces). This results in a [[Riemannian manifold]]. Generally, [[straight line]]s do not exist in a Riemannian manifold, but their role is played by [[geodesic]]s, which are the "shortest paths" between two points. This allows defining distances, which are measured along geodesics, and angles between geodesics, which are the angle of their tangents in the tangent space at their intersection. So, Riemannian manifolds behave locally like a Euclidean that has been bended.

Euclidean spaces are trivially Riemannian manifolds. A well illustrating example is the surface of a [[sphere]]. In this case, geodesics are [[great circle|arcs of great circle]], which are called [[orthodrome]]s in the context of [[navigation]]. More generally, the spaces of [[non-Euclidean geometries]] can be realized as Riemannian manifolds.


=== Pseudo-Euclidean space ===
Also, the concept of a Riemannian manifold permits an expression of the Euclidean structure in any [[diffeomorphism|smooth]] [[coordinate system]], via [[metric tensor]]. From this tensor one can compute the [[Riemann curvature tensor]]. Where the latter equals to zero, the metric structure is locally Euclidean (it means that at least some [[open set]] in the [[coordinate space]] is isometric to a piece of Euclidean space), no matter whether coordinates are affine or curvilinear.
The [[inner product]] that is defined to define Euclidean spaces is a [[positive definite bilinear form]]. If it is replaced by an [[indefinite quadratic form]] which is [[non-degenerate]], one gets a [[pseudo-Euclidean space]].


A fundamental example of such a space is the [[Minkowski space]], which is the [[space-time]] of [[Albert Einstein|Einstein]]'s [[special relativity]]. It is a four-dimensional space, where the metric is defined by the [[quadratic form]]
=== Indefinite quadratic form ===
:<math>x^2+y^2+z^2-t^2,</math>
{{see also|Sylvester's law of inertia}}
where the last coordinates is the time, and the three other are space coordinates.
If one replaces the inner product of a Euclidean space with an indefinite quadratic form, the result is a [[pseudo-Euclidean space]]. Smooth manifolds built from such spaces are called [[pseudo-Riemannian manifold]]s. Perhaps their most famous application is the [[theory of relativity]], where flat [[spacetime]] is a pseudo-Euclidean space called [[Minkowski space]], where rotations correspond to motions of hyperbolic spaces mentioned [[#Riemannian|above]]. Further generalization to curved spacetimes form pseudo-Riemannian manifolds, such as in [[general relativity]].


To take the [[gravity]] into account, [[general relativity]] uses a [[pseudo-Riemannian manifold]] that has Minkowski spaces as [[tangent space]]s. The [[Curvature of Riemannian manifolds|curvature]] of this manifold at a point is a function of the value of the [[gravitational field]] at this point.
=== Infinite dimension ===
{{main article|inner product space|Hilbert space}}
{{expand section|date=April 2013}}
Another line of generalization is to consider other [[field (mathematics)|fields]] than one of real numbers. Over [[complex numbers]], a [[Hilbert space]] can be seen as a generalization of Euclidean dot product structure, although the definition of the inner product becomes a [[sesquilinear form]] for compatibility with metric structure.


== See also ==
== See also ==
{{Portal|Mathematics}}
{{Portal|Mathematics}}
* [[Hilbert space]], a generalization to infinite dimension, used in [[functional analysis]]
* [[Function of several real variables]], a coordinate presentation of a function on a Euclidean space
* [[Geometric algebra]], an alternative algebraic formalism
* [[High-dimensional space]]
* [[Real coordinate space]], a frequently used representation of Euclidean space
* [[Vector calculus]], a standard algebraic formalism
* [[Vector space]]


== Footnotes ==
== Footnotes ==
{{notelist}}
{{reflist|30em|group=footnote}}


== References ==
== References ==

{{Multiple issues|
{{more footnotes|date=April 2013}}
{{refimprove|date=March 2013}}
}}
{{reflist|30em}}
{{reflist|30em}}


* {{citation |first=Emil |last=Artin |title=[[Geometric Algebra]] |series=Wiley Classics Library|publisher=John Wiley & Sons Inc. |place=New York |year=1988 |pages=x+214 |isbn=0-471-60839-4|mr=1009557|origyear=1957 |doi=10.1002/9781118164518|ref=harv}}
==External links==
* {{cite book|last=Ball|first=W.W. Rouse|authorlink=W. W. Rouse Ball|title = A Short Account of the History of Mathematics|origyear=1908|url=|edition=4th|year=1960|publisher= Dover Publications
* {{springer|title=Euclidean space|id=p/e036380}}
|isbn=0-486-20630-0|ref=harv}}
* {{Citation | last1=Berger | first1=Marcel | author1-link=Marcel Berger | title=Geometry I | publisher=Springer | location=Berlin | isbn= 3-540-11658-3 | year=1987|ref=harv}}
* {{Cite book|title=Regular Polytopes|title-link=Regular Polytopes (book)|last=Coxeter|first=H.S.M.|publisher=Dover|year=1973|ref=harv|isbn=|edition=3rd|location=|publication-place=New York|quote="Schläfli ... discovered them before 1853 -- a time when Cayley, Grassman and Möbius were the only other people who had ever conceived of the possibility of geometry in more than three dimensions."|author-link=Harold Scott MacDonald Coxeter|orig-year=1948}}
* {{springer|title=Euclidean space|id=p/e036380|ref=harv|first=E.D.|last=Solomentsev}}


==External links==
{{Functional Analysis}}


{{Authority control}}
{{Authority control}}

Revision as of 14:48, 18 September 2019

A point in three-dimensional Euclidean space can be located by three coordinates.

Euclidean space is the fundamental space of geometry. Originally, this was the three-dimensional space of Euclidean geometry, but, in modern mathematics, there are Euclidean spaces of any nonnegative integer dimension,[1] including the three-dimensional space and the Euclidean plane (dimension two). It has been introduced by the Ancient Greek mathematician Euclid of Alexandria,[2] and the qualifier Euclidean has been added for distinguishing it from other spaces that are considered in physics and modern mathematics.

Ancient Greek geometers introduced Euclidean space for modeling the physical universe. Their great innovation was to prove all properties of the space (theorems) by starting from a few fundamental properties, called postulates, which either were considered as evidences (for example, there is exactly one straight line passing through two points), or seemed impossible to prove (parallel postulate).

After the introduction at the end of 19th century of non-Euclidean geometries, the old postulates have been formalized for defining Euclidean spaces through an axiomatic theory. Another definition of Euclidean spaces by mean of vector spaces and linear algebra has been shown to be equivalent to the axiomatic definitions. This is this definition that is more commonly used in modern mathematics, and detailed in this article.[3]

For all definitions, Euclidean spaces consist of points, which are defined only by the properties that they must have for forming a Euclidean space.

There is essentially only one Euclidean space of each dimension; that is, all Euclidean spaces of a given dimension are isomorphic. Therefore, in many cases, it is possible to work with a specific Euclidean space, which is generally the real n-space equipped with the dot product. An isomorphism from a Euclidean space to associates to each point a n-tuple of real numbers, which locate them in the Euclidean space and are called Cartesian coordinates.

Definition

History of the definition

Euclidean space was introduced by ancient Greeks as an abstraction of our physical space. Their great innovation, appearing in Euclid's Elements was to build and prove all geometry by starting from a few very basic properties, which are abstracted from the physical world, and cannot be mathematically proved because of the lack of more basic tools. These properties are called postulates, or axioms in modern language. This way of defining Euclidean space is still in use under the name of synthetic geometry.

In 1637, René Descartes introduced Cartesian coordinates and showed that this allows reducing geometric problems to algebraic computations with numbers. This reduction of geometry to algebra was a major change of point of view, as, until then, the real numbers—that is, rational numbers and non-rational numbers together–were defined in terms of geometry, as lengths and distance.

Euclidean geometry was not applied in spaces of more than three dimensions until the 19th century. Ludwig Schläfli generalized Euclidean geometry to spaces of n dimensions using both synthetic and algebraic methods, and discovered all of the regular polytopes (higher-dimensional analogues of the Platonic solids) that exist in Euclidean spaces of any number of dimensions.[4]

Despite the wide use of Descartes' approach, which was called analytic geometry, the definition of Euclidean space remained unchanged until the end of 19th century. The introduction of abstract vector spaces allowed their use in defining Euclidean spaces with a purely algebraic definition. This new definition has been shown to be equivalent to the classical definition in terms of geometric axioms. It is this algebraic definition that is now most often used for introducing Euclidean spaces.

Motivation of the modern definition

One way to think of the Euclidean plane is as a set of points satisfying certain relationships, expressible in terms of distance and angles. For example, there are two fundamental operations (referred to as motions) on the plane. One is translation, which means a shifting of the plane so that every point is shifted in the same direction and by the same distance. The other is rotation around a fixed point in the plane, in which all points in the plane turn around that fixed point through the same angle. One of the basic tenets of Euclidean geometry is that two figures (usually considered as subsets) of the plane should be considered equivalent (congruent) if one can be transformed into the other by some sequence of translations, rotations and reflections (see below).

In order to make all of this mathematically precise, the theory must clearly define what is an Euclidean space, and the related notions of distance, angle, translation, and rotation. Even when used in physical theories, Euclidean space is an abstraction detached from actual physical locations, specific reference frames, measurement instruments, and so on. A purely mathematical definition of Euclidean space also ignores questions of units of length and other physical dimensions: the distance in a "mathematical" space is a number, not something expressed in inches or metres.

The standard way to mathematically define an Euclidean space, as carried out in the remainder of this article, is to define an Euclidean space as a set of points on which acts a real vector space, the space of translations which is equipped with an inner product.[1] The action of translations makes the space an affine space, and this allow defining lines, planes, subspaces, dimension, and parallelism. The inner product allows defining distance and angles.

The set of n-tuples of real numbers equipped with the dot product is a Euclidean space of dimension n. Conversely, the choice of a point called the origin and an orthonormal basis of the space of translations is equivalent with defining an isomorphism between a Euclidean space of dimension n and viewed as a Euclidean space.

It follows that everything that can be said about a Euclidean space can also be said about Therefore, many authors, specially at elementary level, call the standard Euclidean space of dimension n,[5] or simply the Euclidean space of dimension n.

A reason for introducing such an abstract definition of Euclidean spaces, and for working with it instead of is that it is often preferable to work in a coordinate-free and origin-free manner (that is, without choosing a preferred basis and a preferred origin). Another reason is that there is no origin nor any basis in the physical world.

Technical definition

A Euclidean vector space is a finite-dimensional inner product space over the real numbers.

A Euclidean space is an affine space over the reals such that the associated vector space is a Euclidean vector space. Euclidean spaces are sometimes called Euclidean spaces for distinguishing them from Euclidean vector spaces.[6]

The dimension of a Euclidean space is the dimension of its associated vector space.

If E is a Euclidean space, its associated vector space is also called its space of translations, and often denoted

The elements of E are called points and are commonly denoted by capital letters. The elements of are called translations, Euclidean vectors or free vectors.

The action of a translation v on a point P provides a point that is denoted P + v. This action satisfies

(The second + in the left-hand side is a vector addition; all other + denote an action of a vector on a point. This notation is not ambiguous, as, for distinguishing between the two meanings of +, it suffices to look on the nature of its left argument.)

The fact that the action is free and transitive means that for every pair of points (P, Q there is exactly one vector v such that P + v = Q. This vector v is denoted QP or

As previously explained, some of the basic properties of Euclidean spaces result of the structure of affine space. They are described in § Affine structure and its subsections. The properties resulting from the inner product are explained in § Metric structure and its subsections.

Prototypical examples

For any vector space, the addition acts freely and transitively on the vector space itself. Thus a Euclidean vector space can be viewed as a Euclidean space that has itself as associated vector space.

A typical case of Euclidean vector space is viewed as a vector space equipped with the dot product as an inner product. The importance of this particular example of Euclidean space lies is the fact that every Euclidean space is isomorphic to it. More precisely, given a Euclidean space E of dimension n, the choice of a point, called a origin and a orthonormal basis of defines an isomorphism of Euclidean spaces from E to

As every Euclidean space of dimension n is isomorphic to it, the Euclidean space is sometimes called the standard Euclidean space of dimension n. [5]

Affine structure

Some basic properties of Euclidean spaces depend only of the fact that an Euclidean space is an affine space. they are called affine properties and include the concepts of lines, subspaces, and parallelism. which are detailed in next subsections.

Subspaces

Let E a Euclidean space, and its associated vector space.

A flat, Euclidean subspace or affine subspace of E is a subset F of E such that

is a linear subspace of An Euclidean subspace F is a Euclidean space, with as associated vector space. This linear subspace is called the direction of F.

If P is a point of F then

Conversely, if P is a point of E, and V is a linear subspace of then

is a Euclidean subspace of direction V.

A Euclidean vector space (that is a Euclidean space such that ) has two sorts of subspaces, its Euclidean subspaces and its linear subspaces. Linear subspaces are Euclidean subspaces, and a Euclidean subspace is a linear subspace if and only if it contains the zero vector.

Lines and segments

In a Euclidean space, a line is a Euclidean subspace of dimension one. Since a vector space of dimension one is spanned by any nonzero vector a line is a set of the form

where P and Q are two distinct points.

It follows that there is exactly one line that passes through (contains) two distinct points. This implies that two distinct lines intersect in at most one point.

A more symmetric representation of the line passing through P and Q is

where O is an arbitrary point (not necessary on the line).

In a Euclidean vector space, the zero vector is usually chosen for O; this allows simplifying the preceding formula into

A standard convention allows using this formula in every Euclidean space, see Affine space § Affine combinations and barycenter.

The line segment, or simply segment, joining the points P and Q is the subset of the points such that 0 ≤ 𝜆 ≤ 1 in the preceding formulas. It is denoted PQ or QP; that is

Parallelism

Two subspaces S and T of the same dimension in a Euclidean space are parallel if they have the same direction.[a] Equivalently, they are parallel, if there is a translation v vector that maps one to the other:

Given a point P and a subspace S, there exists exactly one subspace that contains P and is parallel to S, which is In the case where S is a line (subspace of dimension one), this property is Playfair's axiom.

It follows that in a Euclidean plane, two lines either meet in one point or are parallel.

The concept of parallel subspaces has been extended to subspaces of different dimensions: two subspaces are parallel if the direction of one of them is contained in the direction to the other.

Metric structure

The vector space associated to a Euclidean space E is an inner product space. This implies a symmetric bilinear form

that is positive definite (that is is always positive for x ≠ 0).

The inner product of a Euclidean space is often called dot product and denoted xy. This is specially the case when a Cartesian coordinate system has been chosen, as, in this case, the inner product of two vectors is the dot product of their coordinate vectors. For this reason, and for historical reasons, the dot notation is more commonly used than the bracket notation for the inner product of Euclidean spaces. This article will follow this usage; that is will be denoted xy in the remainder of this article.

The Euclidean norm of a vector v is

The inner product and the norm allows expressing and proving all metric and topological properties of Euclidean geometry. The next subsection describe the most fundamental ones. In these subsections, E denotes an arbitrary Euclidean space, and denotes its vector space of translations.

Distance and length

The distance (more precisely the Euclidean distance) between two points of a Euclidean space is the norm of the translation vector that maps one point to the other; that is

The length of a segment PQ is the distance d(P, Q) between its endpoints. It is often denoted |PQ|.

The distance is a metric, as it satisfies the triangular inequality

Moreover, the equality is true if and only if R belongs to the segment PQ. This inequality means that the length of any edge of a triangle is smaller than the sum of the lengths of the other edges. This is the origin of the term triangular inequality.

With the Euclidean distance, every Euclidean space is a complete metric space.

Orthogonality

Two nonzero vectors u and v of are perpendicular or orthogonal if their inner product is zero:

Two linear subspaces of are orthogonal if every nonzero vector of the first one is perpendicular to every nonzero vector of the second one. This implies that the intersection of the linear subspace is reduced to the zero vector.

Two lines, and more generally two Euclidean subspaces are orthogonal if their direction are orthogonal. Two orthogonal lines that intersect are said perpendicular.

Two segments AB and AC that share a common endpoint are perpendicular or form a right angle if the vectors and are orthogonal.

If AB and AC form a right angle, one has

This is the Pythagorean theorem. Its proof is easy in this context, as, expressing this in terms of the inner product, one has, using bilinearity and symmetry of the inner product:

Angle

Positive and negative angles on the oriented plane

The (non-oriented) angle θ between two nonzero vectors x and y in is

where arccos is the principal value of the arccosine function. By Cauchy–Schwarz inequality, the argument of the arcsine is in the interval [–1, 1]. Therefore 𝜃 is real, and 0 ≤ θπ (or 0 ≤ θ ≤ 180} if angles are measured in degrees).

Angles are not useful in a Euclidean line, as they can be only 0 or π.

In an oriented Euclidean plane, one can define the oriented angle of two vectors. The oriented angle of two vectors x and y is then the opposite of the oriented angle of y and x. In this case, the angle of two vectors can have any value modulo an integer multiple of 2π. In particular, a reflex angle π < 𝜃 < 2π equals the negative angle π < 𝜃 – 2π < 0.

The angle of two vectors does not change if they are multiplied by positive numbers. More precisely, if x and y are two vectors, and 𝜆 and 𝜇 are real numbers, then

If A, B and C are three points in a Euclidean space, the angle of the segments AB and AC is the angle of the vectors and As the multiplication of vectors by positive numbers do not change the angle, the angle of two half-lines with initial point A can be defined: it is the angle of the segments AB and AC, where B and C are arbitrary points, one on each half-line. Although this is less used, one can define similarly the angle of segments or half-lines that do not share an initial points.

The angle of two lines is defined as follows. If 𝜃 is the angle of two segments, one on each line, the angle of any two other segments, one on each line, is either 𝜃 or π𝜃. One of these angles is in the interval [0, π/2], and the other being in [π/2, π]. The non-oriented angle of the two lines is the one in the interval [0, π/2]. In an oriented Euclidean plane, the oriented angle of two lines belongs to the interval [–π/2, π/2].

Cartesian coordinates

Every Euclidean vector space has an orthonormal basis (in fact, infinitely many in dimension higher than one, and two in dimension one), that is a basis of unit vectors () that are pairwise orthogonal ( for ij). More precisely, given any basis the Gram–Schmidt process computes an orthonormal basis such that, for every i, the linear spans of and are equal.

Given a Euclidean space E, a Cartesian frame is a set of data consisting of an orthonormal basis of and a point of E, called the origin and often denoted O. A Cartesian frame allows defining Cartesian coordinates for both E and in the following way.

The Cartesian coordinates of a vector v are the coefficients of v on the basis As the basis is orthonormal, the ith coefficient is the dot product

The Cartesian coordinates of a point P of E are the Cartesian coordinates of the vector

Other coordinates

3-dimensional skew coordinates

As a Euclidean space is an affine space, one can consider an affine frame on it, which is the same as a Euclidean frame, except that the basis is not required to be orthonormal. This define affine coordinates, sometimes called skew coordinates for emphasizing that the basis vectors are not pairwise orthogonal.

An affine basis of a Euclidean space of dimension n is a set of n + 1 points that are not contained in a hyperplan. An affine basis define barycentric coordinates for every point.

Many other coordinates systems can be defined on a Euclidean space E of dimension n, in the following way. Let f be a homeomorphism (or, more often, a diffeomorphism) from a dense open subset of E to an open subset of The coordinates of a point x of E are the components of f(x). The polar coordinate system (dimension 2) and the spherical and cylindrical coordinate systems (dimension 3) are defined this way.

For points that are outside the domain of f, coordinates may sometimes be defined as the limit of coordinates of neighbour points, but these coordinates may be not uniquely defined, and may be not continuous in the neighborhood of the point. For example, for the spherical coordinate system, the longitude is not defined at the pole, and on the antimeridian, the longitude passes discontinuousely from –180° to +180°.

This way of defining coordinates extends easily to other mathematical structures, and in particular to manifolds.

Isometries

An isometry between two metric spaces is a bijection preserving the distance,[b] that is

In the case of a Euclidean vector space, an isometry preserves the norm

since the norm of a vector is its distance from the zero vector. It preserves also the inner product

since

An isometry Euclidean vector spaces is a linear isomorphism.[c][7]

An isometry of Euclidean spaces defines an isometry of the associated Euclidean vector spaces. This implies that two isometric Euclidean spaces have the same dimension. Conversely, if E and F are Euclidean spaces, OE, OF, and is an isometry, then the map defined by

is an isometry of Euclidean spaces.

It follows from the preceding results that an isometry of Euclidean spaces maps lines to lines, and, more generally Euclidean subspaces to Euclidean subspaces of the same dimension, and that the restriction of the isometry on these subspaces are isometries of these subspaces.

Isometry with prototypical examples

If E is a Euclidean space, it associated vector space can be considered as a Euclidean space. Every point OE defines an isometry of Euclidean spaces

which maps O to the zero vector and has the identity as associated linear map. The inverse isometry is the map

A Euclidean frame allows defining the map

which is an isometry of Euclidean spaces. The inverse isometry is

This means that, up to an isomorphism, there is exactly one Euclidean space of a given dimension.

This justifies that many authors talk of as the Euclidean space of dimension n.

Euclidean group

An isometry from a Euclidean space onto itself is called Euclidean isometry, Euclidean transformation or rigid transformation. The rigid transformations of a Euclidean space form a group (under composition), called the Euclidean group and often denoted E(n) of ISO(n).

The simplest Euclidean transformations are translations

They are in bijective correspondence with vectors. This is a reason for calling space of translations the vector space associated to a Euclidean space. The translations form a normal subgroup of the Euclidean group.

A Euclidean isometry f of a Euclidean space E defines a linear isometry of the associated vector space (by linear isometry, it is meant an isometry that is also a linear map) in the following way: denoting by QP the vector , if O is an arbitrary point of E, one has

It is straightforward to prove that this is a linear map that does not depend from the choice of O.

The map is a group homomorphism from the Euclidean group onto the group of linear isometries, called the orthogonal group. The kernel of this homomorphism is the translation group, showing that it is a normal subgroup of the Euclidean group.

The isometries that fix a given point P form the stabilizer subgroup of the Euclidean group with respect to P. The restriction to this stabilizer of above group homomorphism is an isomorphism. So the isometries that fix a given point form a group isomorphic to the orthogonal group.

Let P be a point, f an isometry, and t the translation that maps P to f(P). The isometry fixes P. So and the Euclidean group is the semidirect product of the translation group and the orthogonal group.

The special orthogonal group is the normal subgroup of the orthogonal group that preserves handeness. It is a subgroup of index two ion the orthogonal group. Its inverse image by the group homomorphism is a normal subgroup of index two of the Euclidean group, which is called the special Euclidean group or the displacement group. Its elements are called rigid motions or displacements.

Rigid motions include the identity, translations, rotations (the rigid motions that fix at least a point), and also screw motions.

Typical examples of rigid transformations that are not rigid motions are reflections, which are rigid transformations that fix a hyperplane and are not the identity. They are also the transformations consisting in changing the sign of one coordinate over some Euclidean frame.

As the special Euclidean group is a subgroup of index two of the Euclidean group, given a reflection r, every rigid transformation that is not a rigid motion is the product of r and a rigid motion. A glide reflection is an example of a rigid transformation that is not a rigid motion not a reflection.

All groups that have been considered in this section are Lie groups and algebraic groups.

Topology

The Euclidean distance makes a Euclidean space a metric space, and thus a [[topological space]. This topology is called the Euclidean topology. In the case of this topology is also the product topology.

The open sets are the subsets that contains an open ball around each of their points. In other words, open balls form a base of the topology.

The topological dimension of a Euclidean space equals its dimension. This implies that Euclidean spaces of different dimensions are not homeomorphic. Moreover, the theorem of invariance of domain asserts that a subset of a Euclidean space is open (for the subspace topology) if and only if it is homeomorphic to an open subset of a Euclidean space of the same dimension.

Euclidean spaces are complete and locally compact. That is, a closed subset of a Euclidean space is compact if it is bounded (that is, contained in a ball). In particular, closed balls are compact.

Axiomatic definitions

The definition of Euclidean spaces that has been described in this article differs fundamentally of Euclid's one. In reality, Euclid did not define formally the space, because it was thought as a description of the physical world that exists independently of human mind. The need of a formal definition appeared only at the end of 19th century, with the introduction of non-Euclidean geometries.

Two different approach have been used. Felix Klein suggested to define geometries through their symmetries. The presentation of Euclidean spaces given in this article, is essentially issued from his Erlangen program, with the emphasis given on the groups of translations and isometries.

On the other hand, David Hilbert proposed a set of axioms, inspired by Euclid's postulates. They belong to synthetic geometry, as they do not involve any definition of real numbers. Later G. D. Birkhoff and Alfred Tarski proposed simpler sets of axioms, which use real numbers (see Birkhoff's axioms and Tarski's axioms).

In Geometric Algebra, Emil Artin has proved that all these definitions of a Euclidean space are equivalent.[8] It is rather easy to prove that all definitions of Euclidean spaces satisfy Hilbert's axioms, and that those involving real numbers (including the above given definition) are equivalent. The difficult part of Artin's proof is the following. In Hilbert's axioms, congruence is an equivalence relation on segments. One can thus define the length of a segment as its equivalence class. One must thus prove that this length satisfies properties that characterize nonnegative real numbers. It is what did Artin, with axioms that are not Hilbert's ones, but are equivalent.

Usage

Since ancien Greeks, Euclidean space is used for modeling shapes in the physical world. It is thus used in many sciences such as physics, mechanics, astronomy, ... It is also widely used in all technical areas that are concerned with shapes, figure, location and position, such as architecture, geodesy, topography, navigation, industrial design, technical drawing, robotics...

Space-time of general relativity is not a Euclidean space, but it is a manifold, and, as such, is locally approximated by Euclidean spaces.

Space of dimensions higher than three occurs in several modern theories of physics; see Higher dimension. They occur also in configuration spaces of physical systems.

Beside Euclidean geometry, Euclidean spaces are also widely used in mathematics. Tangent spaces of differentiable manifolds are Euclidean vector spaces. More generally, a manifold is a space that is locally a approximated by Euclidean spaces. Most non-Euclidean geometries can be modeled by a manifold, and embedded in a Euclidean space of higher dimension. For example, an elliptic space can be modeled by an ellipsoid. It is common to represent in a Euclidean space mathematics objects that are a priori not of a geometrical nature. An example among many is the usual representation of graphs.

Other geometric spaces

Since the introduction, at the end of 19th century, of Non-Euclidean geometries, many sorts of spaces have been considered, about which one can do geometric reasoning in the same way as with Euclidean spaces. In general, they share some properties with Euclidean spaces, but may also have properties that could appear as rather strange. Some of these spaces use Euclidean geometry for their definition, or can be modeled as subspaces of a Euclidean space of higher dimension. When such a space is defined by geometrical axioms, embedding the space in a Euclidean space is a standard way for proving consistency of its definition, or, more precisely for proving that its theory is consistent, if Euclidean geometry is consistent (which cannot be proved).

Affine space

A Euclidean space is an affine space equipped with a metric. Affine spaces have many other uses in mathematics. In particular, as they are defined over any field, they allow doing geometry in other contexts.

As soon as non-linear questions are considered, it is generally useful to consider affine spaces over the complex numbers as an extension of Euclidean spaces. For example, a circle and a line have always two intersection points (possibly not distinct) in the complex affine space. Therefore, most of algebraic geometry is built in complex affine spaces and affine spaces over algebraically closed fields. The shapes that are studied in algebraic geometry in these affine spaces are therefore called affine algebraic varieties.

Affine spaces over the rational numbers and more generally over algebraic number fields provide a link between (algebraic) goemetry and number theory. For example, the Fermat's Last Theorem can be stated "a Fermat curve of degree higher than two has no point in the affine plane over the rationals."

Geometry in affine spaces over a finite fields has also been widely studied. For example, elliptic curves over finite fields are widely used in cryptography.

Projective space

Originally, projective spaces have been introduced by adding "points at infinity" to Euclidean spaces, and, more generally to affine spaces, in order to make true the assertion "two coplanar lines meet in exactly in one point". Projective space share with Euclidean and affine spaces the property of being isotropic, that is, there is no property of the space that allows distinguishing between two points or two lines. Therefore, a more isotropic definition is commonly used, which consists as defining a projective space as the set of the vector lines in a vector space of dimension one more.

As for affine spaces, projective spaces are defined over any field, and are fundamental spaces of algebraic geometry.

Non-Euclidean geometries

Non-Euclidean geometry refers usually to geometrical spaces where the parallel postulate is false. They include elliptic geometry, where the sum of the angles of a triangle is less than 180°, and hyperbolic geometry, where this sum is more than 180°. Their introduction in the second half of 19th century, and the proof that their theory is consistent (if Euclidean geometry is not contradictory) is one of the paradoxes that are at the origin of the foundational crisis in mathematics of the beginning of 20th century, and motivated the systematization of axiomatic theories in mathematics.

Curved spaces

A manifold is a space that in the neighborhood of each point resemble to a Euclidean space. In technical terms, a manifold is a topological space, such that each point has a neighborhood that is homeomorphic to an open subset of a Euclidean space. Manifold can be classify, by increasing degree of this "resemblance" into topological manifolds, differentiable manifolds, smooth manifolds, and analytic manifolds. However, none of these types of "resemblance" respect distances and angles, even approximately.

Distances and angles can be defined on a smooth manifold by providing a smoothly varying Euclidean metric on the tangent spaces at the points of the manifold (these tangent are thus Euclidean vector spaces). This results in a Riemannian manifold. Generally, straight lines do not exist in a Riemannian manifold, but their role is played by geodesics, which are the "shortest paths" between two points. This allows defining distances, which are measured along geodesics, and angles between geodesics, which are the angle of their tangents in the tangent space at their intersection. So, Riemannian manifolds behave locally like a Euclidean that has been bended.

Euclidean spaces are trivially Riemannian manifolds. A well illustrating example is the surface of a sphere. In this case, geodesics are arcs of great circle, which are called orthodromes in the context of navigation. More generally, the spaces of non-Euclidean geometries can be realized as Riemannian manifolds.

Pseudo-Euclidean space

The inner product that is defined to define Euclidean spaces is a positive definite bilinear form. If it is replaced by an indefinite quadratic form which is non-degenerate, one gets a pseudo-Euclidean space.

A fundamental example of such a space is the Minkowski space, which is the space-time of Einstein's special relativity. It is a four-dimensional space, where the metric is defined by the quadratic form

where the last coordinates is the time, and the three other are space coordinates.

To take the gravity into account, general relativity uses a pseudo-Riemannian manifold that has Minkowski spaces as tangent spaces. The curvature of this manifold at a point is a function of the value of the gravitational field at this point.

See also

Footnotes

  1. ^ It may depend on the context or the author whether a subspace is parallel to itself
  2. ^ If the condition of being a bijection is removed, a function preserving the distance is necessarily injective, and is an isometry from its domain to its image.
  3. ^ Proof: one must prove that . For that, it suffices to prove that the square of the norm of the left-hand side is zero. Using the bilinearity of the inner product, this squared norm can be expanded into a linear combination of and As f is an isometry, this gives a linear combination of and which simplifies to zero.

References

  1. ^ a b Solomentsev 2001.
  2. ^ Ball 1960, pp. 50–62.
  3. ^ Berger 1987.
  4. ^ Coxeter 1973.
  5. ^ a b Berger 1987, Section 9.1.
  6. ^ Berger 1987, Chapter 9.
  7. ^ Berger 1987, Proposition 9.1.3.
  8. ^ Artin 1988.
  • Artin, Emil (1988) [1957], Geometric Algebra, Wiley Classics Library, New York: John Wiley & Sons Inc., pp. x+214, doi:10.1002/9781118164518, ISBN 0-471-60839-4, MR 1009557 {{citation}}: Invalid |ref=harv (help)
  • Ball, W.W. Rouse (1960) [1908]. A Short Account of the History of Mathematics (4th ed.). Dover Publications. ISBN 0-486-20630-0. {{cite book}}: Invalid |ref=harv (help)
  • Berger, Marcel (1987), Geometry I, Berlin: Springer, ISBN 3-540-11658-3 {{citation}}: Invalid |ref=harv (help)
  • Coxeter, H.S.M. (1973) [1948]. Regular Polytopes (3rd ed.). New York: Dover. Schläfli ... discovered them before 1853 -- a time when Cayley, Grassman and Möbius were the only other people who had ever conceived of the possibility of geometry in more than three dimensions. {{cite book}}: Invalid |ref=harv (help)
  • Solomentsev, E.D. (2001) [1994], "Euclidean space", Encyclopedia of Mathematics, EMS Press {{citation}}: Invalid |ref=harv (help)

External links