Metric space


In mathematics, a metric space is a set together with a metric on the set. The metric is a function that defines a concept of distance between any two members of the set, which are usually called points. The metric satisfies a few simple properties. Informally:


  • the distance from a point to itself is zero,

  • the distance between two distinct points is positive,

  • the distance from A to B is the same as the distance from B to A, and

  • the distance from A to B (directly) is less than or equal to the distance from A to B via any third point C.

A metric on a space induces topological properties like open and closed sets, which lead to the study of more abstract topological spaces.


The most familiar metric space is 3-dimensional Euclidean space. In fact, a "metric" is the generalization of the Euclidean metric arising from the four long-known properties of the Euclidean distance. The Euclidean metric defines the distance between two points as the length of the straight line segment connecting them. Other metric spaces occur for example in elliptic geometry and hyperbolic geometry, where distance on a sphere measured by angle is a metric, and the hyperboloid model of hyperbolic geometry is used by special relativity as a metric space of velocities.




Contents





  • 1 History


  • 2 Definition


  • 3 Examples of metric spaces


  • 4 Open and closed sets, topology and convergence


  • 5 Types of metric spaces

    • 5.1 Complete spaces


    • 5.2 Bounded and totally bounded spaces


    • 5.3 Compact spaces


    • 5.4 Locally compact and proper spaces


    • 5.5 Connectedness


    • 5.6 Separable spaces


    • 5.7 Pointed metric spaces



  • 6 Types of maps between metric spaces

    • 6.1 Continuous maps


    • 6.2 Uniformly continuous maps


    • 6.3 Lipschitz-continuous maps and contractions


    • 6.4 Isometries


    • 6.5 Quasi-isometries



  • 7 Notions of metric space equivalence


  • 8 Topological properties


  • 9 Distance between points and sets; Hausdorff distance and Gromov metric


  • 10 Product metric spaces

    • 10.1 Continuity of distance



  • 11 Quotient metric spaces


  • 12 Generalizations of metric spaces

    • 12.1 Metric spaces as enriched categories



  • 13 See also


  • 14 Notes


  • 15 References


  • 16 External links




History


In 1906 Maurice Fréchet introduced metric spaces in his work Sur quelques points du calcul fonctionnel.[1] However the name is due to Felix Hausdorff.



Definition


A metric space is an ordered pair (M,d)displaystyle (M,d)(M,d) where Mdisplaystyle MM is a set and ddisplaystyle dd is a metric on Mdisplaystyle MM, i.e., a function


d:M×M→Rdisplaystyle dcolon Mtimes Mto mathbb R d colon M times M to mathbbR

such that for any x,y,z∈Mdisplaystyle x,y,zin Mx,y,zin M, the following holds:[2]














1.

d(x,y)≥0displaystyle d(x,y)geq 0displaystyle d(x,y)geq 0

non-negativity or separation axiom
2.d(x,y)=0⇔x=ydisplaystyle d(x,y)=0Leftrightarrow x=ydisplaystyle d(x,y)=0Leftrightarrow x=y
identity of indiscernibles
3.d(x,y)=d(y,x)displaystyle d(x,y)=d(y,x)displaystyle d(x,y)=d(y,x)
symmetry
4.d(x,z)≤d(x,y)+d(y,z)displaystyle d(x,z)leq d(x,y)+d(y,z)displaystyle d(x,z)leq d(x,y)+d(y,z)
subadditivity or triangle inequality

The first condition follows from the other three. Since for any x,y∈Mdisplaystyle x,yin Mx, y in M:











d(x,y)+d(y,x)≥d(x,x)displaystyle d(x,y)+d(y,x)geq d(x,x)d(x,y) + d(y,x) ge d(x,x)
by triangle inequality

d(x,y)+d(x,y)≥d(x,x)displaystyle d(x,y)+d(x,y)geq d(x,x)d(x,y) + d(x,y) ge d(x,x)
by symmetry

2d(x,y)≥0displaystyle 2d(x,y)geq 02d(x,y) ge 0
by identity of indiscernibles

d(x,y)≥0displaystyle d(x,y)geq 0d(x,y)geq 0
we have non-negativity

The function ddisplaystyle dd is also called distance function or simply distance. Often, ddisplaystyle dd is omitted and one just writes Mdisplaystyle MM for a metric space if it is clear from the context what metric is used.


Ignoring mathematical details, for any system of roads and terrains the distance between two locations can be defined as the length of the shortest route connecting those locations. To be a metric there shouldn't be any one-way roads. The triangle inequality expresses the fact that detours aren't shortcuts. If the distance between two points is zero, the two points are indistinguishable from one-another. Many of the examples below can be seen as concrete versions of this general idea.



Examples of metric spaces


  • The real numbers with the distance function d(x,y)=|y−x|displaystyle d(x,y)=vert y-xvert d(x,y) = vert y - x vert given by the absolute difference, and, more generally, Euclidean n-space with the Euclidean distance, are complete metric spaces. The rational numbers with the same distance function also form a metric space, but not a complete one.

  • The positive real numbers with distance function d(x,y)=|log⁡(y/x)|displaystyle d(x,y)=vert log(y/x)vert d(x,y) =vert log(y/x) vert is a complete metric space.

  • Any normed vector space is a metric space by defining d(x,y)=‖y−x‖displaystyle d(x,y)=lVert y-xrVert d(x,y) = lVert y - x rVert, see also metrics on vector spaces. (If such a space is complete, we call it a Banach space.) Examples:
    • The Manhattan norm gives rise to the Manhattan distance, where the distance between any two points, or vectors, is the sum of the differences between corresponding coordinates.

    • The maximum norm gives rise to the Chebyshev distance or chessboard distance, the minimal number of moves a chess king would take to travel from xdisplaystyle xx to ydisplaystyle yy.


  • The British Rail metric (also called the “post office metric” or the “SNCF metric”) on a normed vector space is given by d(x,y)=‖x‖+‖y‖displaystyle d(x,y)=lVert xrVert +lVert yrVert d(x,y) = lVert x rVert + lVert y rVert for distinct points xdisplaystyle xx and ydisplaystyle yy, and d(x,x)=0displaystyle d(x,x)=0d(x,x) = 0. More generally ‖.‖displaystyle lVert .rVert lVert . rVert can be replaced with a function fdisplaystyle ff taking an arbitrary set Sdisplaystyle SS to non-negative reals and taking the value 0displaystyle 0displaystyle 0 at most once: then the metric is defined on Sdisplaystyle SS by d(x,y)=f(x)+f(y)displaystyle d(x,y)=f(x)+f(y)d(x,y) = f(x) + f(y) for distinct points xdisplaystyle xx and ydisplaystyle yy, and d(x,x)=0displaystyle d(x,x)=0d(x,x) = 0. The name alludes to the tendency of railway journeys to proceed via London (or Paris) irrespective of their final destination.

  • If (M,d)displaystyle (M,d)(M,d) is a metric space and Xdisplaystyle XX is a subset of Mdisplaystyle MM, then (X,d)displaystyle (X,d)(X,d) becomes a metric space by restricting the domain of ddisplaystyle dd to X×Xdisplaystyle Xtimes XXtimes X.

  • The discrete metric, where d(x,y)=0displaystyle d(x,y)=0d(x,y) = 0 if x=ydisplaystyle x=yx=y and d(x,y)=1displaystyle d(x,y)=1d(x,y) = 1 otherwise, is a simple but important example, and can be applied to all sets. This, in particular, shows that for any set, there is always a metric space associated to it. Using this metric, any point is an open ball, and therefore every subset is open and the space has the discrete topology.

  • A finite metric space is a metric space having a finite number of points. Not every finite metric space can be isometrically embedded in a Euclidean space.[3][4]

  • The hyperbolic plane is a metric space. More generally:
    • If Mdisplaystyle MM is any connected Riemannian manifold, then we can turn Mdisplaystyle MM into a metric space by defining the distance of two points as the infimum of the lengths of the paths (continuously differentiable curves) connecting them.

    • If Xdisplaystyle XX is some set and Mdisplaystyle MM is a metric space, then, the set of all bounded functions f:X→Mdisplaystyle fcolon Xrightarrow Mf colon X rightarrow M (i.e. those functions whose image is a bounded subset of Mdisplaystyle MM) can be turned into a metric space by defining d(f,g)=supx∈Xd(f(x),g(x))displaystyle d(f,g)=sup _xin Xd(f(x),g(x))d(f,g) = sup_x in X d(f(x),g(x)) for any two bounded functions fdisplaystyle ff and gdisplaystyle gg (where supdisplaystyle sup sup is supremum).[5] This metric is called the uniform metric or supremum metric, and If Mdisplaystyle MM is complete, then this function space is complete as well. If X is also a topological space, then the set of all bounded continuous functions from Xdisplaystyle XX to Mdisplaystyle MM (endowed with the uniform metric), will also be a complete metric if M is.

    • If Gdisplaystyle GG is an undirected connected graph, then the set Vdisplaystyle VV of vertices of Gdisplaystyle GG can be turned into a metric space by defining d(x,y)displaystyle d(x,y)d(x,y) to be the length of the shortest path connecting the vertices xdisplaystyle xx and ydisplaystyle yy. In geometric group theory this is applied to the Cayley graph of a group, yielding the word metric.



  • Graph edit distance is a measure of dissimilarity between two graphs, defined as the minimal number of graph edit operations required to transform one graph into another.

  • The Levenshtein distance is a measure of the dissimilarity between two strings udisplaystyle uu and vdisplaystyle vv, defined as the minimal number of character deletions, insertions, or substitutions required to transform udisplaystyle uu into vdisplaystyle vv. This can be thought of as a special case of the shortest path metric in a graph and is one example of an edit distance.

  • Given a metric space (X,d)displaystyle (X,d)(X,d) and an increasing concave function f:[0,∞)→[0,∞)displaystyle fcolon [0,infty )rightarrow [0,infty )f colon [0,infty) rightarrow [0,infty) such that f(x)=0displaystyle f(x)=0f(x) = 0 if and only if x=0displaystyle x=0x=0, then f∘ddisplaystyle fcirc df circ d is also a metric on Xdisplaystyle XX.

  • Given an injective function fdisplaystyle ff from any set Adisplaystyle AA to a metric space (X,d)displaystyle (X,d)(X,d), d(f(x),f(y))displaystyle d(f(x),f(y))d(f(x), f(y)) defines a metric on Adisplaystyle AA.

  • Using T-theory, the tight span of a metric space is also a metric space. The tight span is useful in several types of analysis.

  • The set of all mdisplaystyle mm by ndisplaystyle nn matrices over some field is a metric space with respect to the rank distance d(X,Y)=rank(Y−X)displaystyle d(X,Y)=mathrm rank (Y-X)d(X,Y) = mathrmrank(Y - X).

  • The Helly metric is used in game theory.


Open and closed sets, topology and convergence


Every metric space is a topological space in a natural manner, and therefore all definitions and theorems about general topological spaces also apply to all metric spaces.


About any point xdisplaystyle xx in a metric space Mdisplaystyle MM we define the open ball of radius r>0displaystyle r>0r > 0 (where rdisplaystyle rr is a real number) about xdisplaystyle xx as the set


B(x;r)=y∈M:d(x,y)<r.displaystyle B(x;r)=yin M:d(x,y)<r.B(x;r) = y in M : d(x,y) < r.

These open balls form the base for a topology on M, making it a topological space.


Explicitly, a subset Udisplaystyle UU of Mdisplaystyle MM is called open if for every xdisplaystyle xx in Udisplaystyle UU there exists an r>0displaystyle r>0r > 0 such that B(x;r)displaystyle B(x;r)B(x;r) is contained in Udisplaystyle UU. The complement of an open set is called closed. A neighborhood of the point xdisplaystyle xx is any subset of Mdisplaystyle MM that contains an open ball about xdisplaystyle xx as a subset.


A topological space which can arise in this way from a metric space is called a metrizable space; see the article on metrization theorems for further details.


A sequence (xndisplaystyle x_nx_n) in a metric space Mdisplaystyle MM is said to converge to the limit x∈Mdisplaystyle xin Mxin M iff for every ϵ>0displaystyle epsilon >0epsilon >0, there exists a natural number N such that d(xn,x)<ϵdisplaystyle d(x_n,x)<epsilon d(x_n,x) < epsilon for all n>Ndisplaystyle n>Nn>N. Equivalently, one can use the general definition of convergence available in all topological spaces.


A subset Adisplaystyle AA of the metric space Mdisplaystyle MM is closed iff every sequence in Adisplaystyle AA that converges to a limit in Mdisplaystyle MM has its limit in Adisplaystyle AA.



Types of metric spaces



Complete spaces



A metric space Mdisplaystyle MM is said to be complete if every Cauchy sequence converges in Mdisplaystyle MM. That is to say: if d(xn,xm)→0displaystyle d(x_n,x_m)to 0d(x_n, x_m) to 0 as both ndisplaystyle nn and mdisplaystyle mm independently go to infinity, then there is some y∈Mdisplaystyle yin Myin M with d(xn,y)→0displaystyle d(x_n,y)to 0d(x_n, y) to 0.


Every Euclidean space is complete, as is every closed subset of a complete space. The rational numbers, using the absolute value metric d(x,y)=|x−y|displaystyle d(x,y)=vert x-yvert d(x,y) = vert x - y vert, are not complete.


Every metric space has a unique (up to isometry) completion, which is a complete space that contains the given space as a dense subset. For example, the real numbers are the completion of the rationals.


If Xdisplaystyle XX is a complete subset of the metric space Mdisplaystyle MM, then Xdisplaystyle XX is closed in Mdisplaystyle MM. Indeed, a space is complete iff it is closed in any containing metric space.


Every complete metric space is a Baire space.



Bounded and totally bounded spaces




Diameter of a set.



A metric space M is called bounded if there exists some number r, such that d(x,y) ≤ r for all x and y in M. The smallest possible such r is called the diameter of M. The space M is called precompact or totally bounded if for every r > 0 there exist finitely many open balls of radius r whose union covers M. Since the set of the centres of these balls is finite, it has finite diameter, from which it follows (using the triangle inequality) that every totally bounded space is bounded. The converse does not hold, since any infinite set can be given the discrete metric (one of the examples above) under which it is bounded and yet not totally bounded.


Note that in the context of intervals in the space of real numbers and occasionally regions in a Euclidean space Rndisplaystyle mathbb R ^ndisplaystyle mathbb R ^n a bounded set is referred to as "a finite interval" or "finite region". However boundedness should not in general be confused with "finite", which refers to the number of elements, not to how far the set extends; finiteness implies boundedness, but not conversely. Also note that an unbounded subset of Rndisplaystyle mathbb R ^ndisplaystyle mathbb R ^n may have a finite volume.



Compact spaces


A metric space M is compact if every sequence in M has a subsequence that converges to a point in M. This is known as sequential compactness and, in metric spaces (but not in general topological spaces), is equivalent to the topological notions of countable compactness and compactness defined via open covers.


Examples of compact metric spaces include the closed interval [0,1] with the absolute value metric, all metric spaces with finitely many points, and the Cantor set. Every closed subset of a compact space is itself compact.


A metric space is compact iff it is complete and totally bounded. This is known as the Heine–Borel theorem. Note that compactness depends only on the topology, while boundedness depends on the metric.


Lebesgue's number lemma states that for every open cover of a compact metric space M, there exists a "Lebesgue number" δ such that every subset of M of diameter < δ is contained in some member of the cover.


Every compact metric space is second countable,[6] and is a continuous image of the Cantor set. (The latter result is due to Pavel Alexandrov and Urysohn.)



Locally compact and proper spaces


A metric space is said to be locally compact if every point has a compact neighborhood. Euclidean spaces are locally compact, but infinite-dimensional Banach spaces are not.


A space is proper if every closed ball y : d(x,y) ≤ r is compact. Proper spaces are locally compact, but the converse is not true in general.



Connectedness


A metric space Mdisplaystyle MM is connected if the only subsets that are both open and closed are the empty set and Mdisplaystyle MM itself.


A metric space Mdisplaystyle MM is path connected if for any two points x,y∈Mdisplaystyle x,yin Mx, y in M there exists a continuous map f:[0,1]→Mdisplaystyle fcolon [0,1]to Mfcolon [0,1] to M with f(0)=xdisplaystyle f(0)=xf(0)=x and f(1)=ydisplaystyle f(1)=yf(1)=y.
Every path connected space is connected, but the converse is not true in general.


There are also local versions of these definitions: locally connected spaces and locally path connected spaces.


Simply connected spaces are those that, in a certain sense, do not have "holes".



Separable spaces


A metric space is separable space if it has a countable dense subset. Typical examples are the real numbers or any Euclidean space. For metric spaces (but not for general topological spaces) separability is equivalent to second-countability and also to the Lindelöf property.



Pointed metric spaces


If Xdisplaystyle XX is a nonempty metric space and x0∈Xdisplaystyle x_0in Xx_0in X then (X,x0)displaystyle (X,x_0)(X,x_0) is called a pointed metric space, and x0displaystyle x_0x_0 is called a distinguished point. Note that a pointed metric space is just a nonempty metric space with attention drawn to its distinguished point, and that any nonempty metric space can be viewed as a pointed metric space. The distinguished point is sometimes denoted 0displaystyle 0displaystyle 0 due to its similar behavior to zero in certain contexts.



Types of maps between metric spaces


Suppose (M1,d1) and (M2,d2) are two metric spaces.



Continuous maps



The map f:M1M2 is continuous
if it has one (and therefore all) of the following equivalent properties:


General topological continuity

for every open set U in M2, the preimage f -1(U) is open in M1

This is the general definition of continuity in topology.

Sequential continuity

if (xn) is a sequence in M1 that converges to x in M1, then the sequence (f(xn)) converges to f(x) in M2.

This is sequential continuity, due to Eduard Heine.

ε-δ definition

for every x in M1 and every ε>0 there exists δ>0 such that for all y in M1 we have
d1(x,y)<δ⇒d2(f(x),f(y))<ε.displaystyle d_1(x,y)<delta Rightarrow d_2(f(x),f(y))<varepsilon .d_1(x,y)<delta Rightarrow d_2(f(x),f(y))< varepsilon.

This uses the (ε, δ)-definition of limit, and is due to Augustin Louis Cauchy.

Moreover, f is continuous if and only if it is continuous on every compact subset of M1.


The image of every compact set under a continuous function is compact, and the image of every connected set under a continuous function is connected.



Uniformly continuous maps


The map ƒ : M1 → M2 is uniformly continuous if for every ε > 0 there exists δ > 0 such that


d1(x,y)<δ⇒d2(f(x),f(y))<εfor allx,y∈M1.displaystyle d_1(x,y)<delta Rightarrow d_2(f(x),f(y))<varepsilon quad mboxfor allquad x,yin M_1.d_1(x,y)<delta Rightarrow d_2(f(x),f(y))< varepsilon quadmboxfor allquad x,yin M_1.

Every uniformly continuous map ƒ : M1 → M2 is continuous. The converse is true if M1 is compact (Heine–Cantor theorem).


Uniformly continuous maps turn Cauchy sequences in M1 into Cauchy sequences in M2. For continuous maps this is generally wrong; for example, a continuous map
from the open interval (0,1) onto the real line turns some Cauchy sequences into unbounded sequences.



Lipschitz-continuous maps and contractions



Given a real number K > 0, the map ƒ : M1 → M2 is K-Lipschitz continuous if


d2(f(x),f(y))≤Kd1(x,y)for allx,y∈M1.displaystyle d_2(f(x),f(y))leq Kd_1(x,y)quad mboxfor allquad x,yin M_1.d_2(f(x),f(y))leq K d_1(x,y)quadmboxfor allquad x,yin M_1.

Every Lipschitz-continuous map is uniformly continuous, but the converse is not true in general.


If K < 1, then ƒ is called a contraction. Suppose M2 = M1 and M1 is complete. If ƒ is a contraction, then ƒ admits a unique fixed point (Banach fixed point theorem). If M1 is compact, the condition can be weakened a bit: ƒ admits a unique fixed point if



d(f(x),f(y))<d(x,y)for allx≠y∈M1displaystyle d(f(x),f(y))<d(x,y)quad mboxfor allquad xneq yin M_1 d(f(x), f(y)) < d(x, y) quad mboxfor all quad x ne y in M_1.


Isometries


The map f:M1M2 is an isometry if


d2(f(x),f(y))=d1(x,y)for allx,y∈M1displaystyle d_2(f(x),f(y))=d_1(x,y)quad mboxfor allquad x,yin M_1d_2(f(x),f(y))=d_1(x,y)quadmboxfor allquad x,yin M_1

Isometries are always injective; the image of a compact or complete set under an isometry is compact or complete, respectively. However, if the isometry is not surjective, then the image of a closed (or open) set need not be closed (or open).



Quasi-isometries


The map f : M1 → M2 is a quasi-isometry if there exist constants A ≥ 1 and B ≥ 0 such that


1Ad2(f(x),f(y))−B≤d1(x,y)≤Ad2(f(x),f(y))+B for all x,y∈M1displaystyle frac 1Ad_2(f(x),f(y))-Bleq d_1(x,y)leq Ad_2(f(x),f(y))+Btext for all x,yin M_1frac1A d_2(f(x),f(y))-Bleq d_1(x,y)leq A d_2(f(x),f(y))+B text for all x,yin M_1

and a constant C ≥ 0 such that every point in M2 has a distance at most C from some point in the image f(M1).


Note that a quasi-isometry is not required to be continuous. Quasi-isometries compare the "large-scale structure" of metric spaces; they find use in geometric group theory in relation to the word metric.



Notions of metric space equivalence


Given two metric spaces (M1, d1) and (M2, d2):


  • They are called homeomorphic (topologically isomorphic) if there exists a homeomorphism between them (i.e., a bijection continuous in both directions).

  • They are called uniformic (uniformly isomorphic) if there exists a uniform isomorphism between them (i.e., a bijection uniformly continuous in both directions).

  • They are called isometric if there exists a bijective isometry between them. In this case, the two metric spaces are essentially identical.

  • They are called quasi-isometric if there exists a quasi-isometry between them.


Topological properties


Metric spaces are paracompact[7]Hausdorff spaces[8] and hence normal (indeed they are perfectly normal). An important consequence is that every metric space admits partitions of unity and that every continuous real-valued function defined on a closed subset of a metric space can be extended to a continuous map on the whole space (Tietze extension theorem). It is also true that every real-valued Lipschitz-continuous map defined on a subset of a metric space can be extended to a Lipschitz-continuous map on the whole space.


Metric spaces are first countable since one can use balls with rational radius as a neighborhood base.


The metric topology on a metric space M is the coarsest topology on M relative to which the metric d is a continuous map from the product of M with itself to the non-negative real numbers.



Distance between points and sets; Hausdorff distance and Gromov metric


A simple way to construct a function separating a point from a closed set (as required for a completely regular space) is to consider the distance between the point and the set. If (M,d) is a metric space, S is a subset of M and x is a point of M, we define the distance from x to S as



d(x,S)=infd(x,s):s∈Sdisplaystyle d(x,S)=infd(x,s):sin Sd(x,S) = infd(x,s) : s in S where infdisplaystyle inf inf represents the infimum.

Then d(x, S) = 0 if and only if x belongs to the closure of S. Furthermore, we have the following generalization of the triangle inequality:


d(x,S)≤d(x,y)+d(y,S),displaystyle d(x,S)leq d(x,y)+d(y,S),d(x,S) leq d(x,y) + d(y,S),

which in particular shows that the map x↦d(x,S)displaystyle xmapsto d(x,S)xmapsto d(x,S) is continuous.


Given two subsets S and T of M, we define their Hausdorff distance to be



dH(S,T)=maxsupd(s,T):s∈S,supd(t,S):t∈Tdisplaystyle d_H(S,T)=maxsupd(s,T):sin S,supd(t,S):tin Td_H(S,T) = max supd(s,T) : s in S , sup d(t,S) : t in T where supdisplaystyle sup sup represents the supremum.

In general, the Hausdorff distance dH(S,T) can be infinite. Two sets are close to each other in the Hausdorff distance if every element of either set is close to some element of the other set.


The Hausdorff distance dH turns the set K(M) of all non-empty compact subsets of M into a metric space. One can show that K(M) is complete if M is complete.
(A different notion of convergence of compact subsets is given by the Kuratowski convergence.)


One can then define the Gromov–Hausdorff distance between any two metric spaces by considering the minimal Hausdorff distance of isometrically embedded versions of the two spaces. Using this distance, the class of all (isometry classes of) compact metric spaces becomes a metric space in its own right.



Product metric spaces


If (M1,d1),…,(Mn,dn)displaystyle (M_1,d_1),ldots ,(M_n,d_n)(M_1,d_1),ldots,(M_n,d_n) are metric spaces, and N is the Euclidean norm on Rn, then (M1×…×Mn,N(d1,…,dn))displaystyle Big (M_1times ldots times M_n,N(d_1,ldots ,d_n)Big )Big(M_1times ldots times M_n, N(d_1,ldots,d_n)Big) is a metric space, where the product metric is defined by


N(d1,...,dn)((x1,…,xn),(y1,…,yn))=N(d1(x1,y1),…,dn(xn,yn)),displaystyle N(d_1,...,d_n)Big ((x_1,ldots ,x_n),(y_1,ldots ,y_n)Big )=NBig (d_1(x_1,y_1),ldots ,d_n(x_n,y_n)Big ),N(d_1,...,d_n)Big((x_1,ldots,x_n),(y_1,ldots,y_n)Big) = NBig(d_1(x_1,y_1),ldots,d_n(x_n,y_n)Big),

and the induced topology agrees with the product topology. By the equivalence of norms in finite dimensions, an equivalent metric is obtained if N is the taxicab norm, a p-norm, the max norm, or any other norm which is non-decreasing as the coordinates of a positive n-tuple increase (yielding the triangle inequality).


Similarly, a countable product of metric spaces can be obtained using the following metric


d(x,y)=∑i=1∞12idi(xi,yi)1+di(xi,yi).displaystyle d(x,y)=sum _i=1^infty frac 12^ifrac d_i(x_i,y_i)1+d_i(x_i,y_i).d(x,y)=sum_i=1^infty frac12^ifracd_i(x_i,y_i)1+d_i(x_i,y_i).

An uncountable product of metric spaces need not be metrizable. For example, RRdisplaystyle mathbf R ^mathbf R mathbfR^mathbfR is not first-countable and thus isn't metrizable.



Continuity of distance


In the case of a single space (M,d)displaystyle (M,d)(M,d), the distance map d:M×M→R+displaystyle dcolon Mtimes Mrightarrow R^+dcolon Mtimes M rightarrow R^+ (from the definition) is uniformly continuous with respect to any of the above product metrics N(d,d)displaystyle N(d,d)N(d,d), and in particular is continuous with respect to the product topology of M×Mdisplaystyle Mtimes MMtimes M.



Quotient metric spaces


If M is a metric space with metric d, and ~ is an equivalence relation on M, then we can endow the quotient set M/~ with the following (pseudo)metric. Given two equivalence classes [x] and [y], we define


d′([x],[y])=infd(p1,q1)+d(p2,q2)+⋯+d(pn,qn)displaystyle d'([x],[y])=infd(p_1,q_1)+d(p_2,q_2)+dotsb +d(p_n,q_n)d'([x],[y]) = infd(p_1,q_1)+d(p_2,q_2)+dotsb+d(p_n,q_n)

where the infimum is taken over all finite sequences (p1,p2,…,pn)displaystyle (p_1,p_2,dots ,p_n)(p_1, p_2, dots, p_n) and (q1,q2,…,qn)displaystyle (q_1,q_2,dots ,q_n)(q_1, q_2, dots, q_n) with [p1]=[x]displaystyle [p_1]=[x][p_1]=[x], [qn]=[y]displaystyle [q_n]=[y][q_n]=[y], [qi]=[pi+1],i=1,2,…,n−1displaystyle [q_i]=[p_i+1],i=1,2,dots ,n-1[q_i]=[p_i+1], i=1,2,dots, n-1. In general this will only define a pseudometric, i.e. d′([x],[y])=0displaystyle d'([x],[y])=0d'([x],[y])=0 does not necessarily imply that [x]=[y]displaystyle [x]=[y][x]=[y]. However, for nice equivalence relations (e.g., those given by gluing together polyhedra along faces), it is a metric.


The quotient metric d is characterized by the following universal property. If f:(M,d)⟶(X,δ)displaystyle f:(M,d)longrightarrow (X,delta )f:(M,d)longrightarrow(X,delta) is a metric map between metric spaces (that is, δ(f(x),f(y))≤d(x,y)displaystyle delta (f(x),f(y))leq d(x,y)delta(f(x),f(y))le d(x,y) for all x, y) satisfying f(x)=f(y) whenever x∼y,displaystyle xsim y,xsim y, then the induced function f¯:M/∼⟶Xdisplaystyle overline fcolon M/sim longrightarrow Xoverlinefcolon M/simlongrightarrow X, given by f¯([x])=f(x)displaystyle overline f([x])=f(x)overlinef([x])=f(x), is a metric map f¯:(M/∼,d′)⟶(X,δ).displaystyle overline fcolon (M/sim ,d')longrightarrow (X,delta ).overlinefcolon (M/sim,d')longrightarrow (X,delta).


A topological space is sequential if and only if it is a quotient of a metric space.[9]



Generalizations of metric spaces


  • Every metric space is a uniform space in a natural manner, and every uniform space is naturally a topological space. Uniform and topological spaces can therefore be regarded as generalizations of metric spaces.

  • If we consider the first definition of a metric space given above and relax the second requirement, we arrive at the concepts of a pseudometric space or a dislocated metric space.[10] If we remove the third or fourth, we arrive at a quasimetric space, or a semimetric space.

  • If the distance function takes values in the extended real number line R∪+∞, but otherwise satisfies all four conditions, then it is called an extended metric and the corresponding space is called an ∞displaystyle infty infty -metric space. If the distance function takes values in some (suitable) ordered set (and the triangle inequality is adjusted accordingly), then we arrive at the notion of generalized ultrametric.[10]


  • Approach spaces are a generalization of metric spaces, based on point-to-set distances, instead of point-to-point distances.

  • A continuity space is a generalization of metric spaces and posets, that can be used to unify the notions of metric spaces and domains.

  • A partial metric space is intended to be the least generalisation of the notion of a metric space, such that the distance of each point from itself is no longer necessarily zero.[11]


Metric spaces as enriched categories


The ordered set (R,≥)displaystyle (mathbb R ,geq )(mathbb R ,geq ) can be seen as a category by requesting exactly one morphism a→bdisplaystyle ato bato b if a≥bdisplaystyle ageq bageq b and none otherwise. By using +displaystyle ++ as the tensor product and 0displaystyle 0displaystyle 0 as the identity, it becomes a monoidal category R∗displaystyle R^*R^*.
Every metric space (M,d)displaystyle (M,d)(M,d) can now be viewed as a category M∗displaystyle M^*M^* enriched over R∗displaystyle R^*R^*:


  • Set Ob⁡(M∗):=Mdisplaystyle operatorname Ob (M^*):=Moperatorname Ob(M^*):=M

  • For each X,Y∈Mdisplaystyle X,Yin MX,Yin M set Hom⁡(X,Y):=d(X,Y)∈Ob⁡(R∗)displaystyle operatorname Hom (X,Y):=d(X,Y)in operatorname Ob (R^*)operatornameHom(X,Y):=d(X,Y)in operatornameOb(R^*)

  • The composition morphism Hom⁡(Y,Z)⊗Hom⁡(X,Y)→Hom⁡(X,Z)displaystyle operatorname Hom (Y,Z)otimes operatorname Hom (X,Y)to operatorname Hom (X,Z)operatornameHom(Y,Z)otimes operatornameHom(X,Y)to operatornameHom(X,Z) will be the unique morphism in R∗displaystyle R^*R^* given from the triangle inequality d(y,z)+d(x,y)≥d(x,z)displaystyle d(y,z)+d(x,y)geq d(x,z)d(y,z)+d(x,y)geq d(x,z)

  • The identity morphism 0→Hom⁡(X,X)displaystyle 0to operatorname Hom (X,X)0to operatornameHom(X,X) will be the unique morphism given from the fact that 0≥d(X,X)displaystyle 0geq d(X,X)0geq d(X,X).

  • Since R∗displaystyle R^*R^* is a poset, all diagrams that are required for an enriched category commute automatically.

See the paper by F.W. Lawvere listed below.



See also


  • Space (mathematics)

  • Metric (mathematics)

  • Metric signature

  • Metric tensor

  • Metric tree

  • Norm (mathematics)

  • Normed vector space

  • Measure (mathematics)

  • Hilbert space

  • Hilbert's fourth problem

  • Product metric

  • Aleksandrov–Rassias problem

  • Category of metric spaces

  • Classical Wiener space

  • Glossary of Riemannian and metric geometry


  • Isometry, contraction mapping and metric map

  • Lipschitz continuity

  • Triangle inequality

  • Ultrametric space


Notes




  1. ^ Rendic. Circ. Mat. Palermo 22 (1906) 1–74


  2. ^ B. Choudhary (1992). The Elements of Complex Analysis. New Age International. p. 20. ISBN 978-81-224-0399-2..mw-parser-output cite.citationfont-style:inherit.mw-parser-output .citation qquotes:"""""""'""'".mw-parser-output .citation .cs1-lock-free abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/6/65/Lock-green.svg/9px-Lock-green.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .citation .cs1-lock-limited a,.mw-parser-output .citation .cs1-lock-registration abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/d/d6/Lock-gray-alt-2.svg/9px-Lock-gray-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .citation .cs1-lock-subscription abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/a/aa/Lock-red-alt-2.svg/9px-Lock-red-alt-2.svg.png")no-repeat;background-position:right .1em center.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registrationcolor:#555.mw-parser-output .cs1-subscription span,.mw-parser-output .cs1-registration spanborder-bottom:1px dotted;cursor:help.mw-parser-output .cs1-ws-icon abackground:url("//upload.wikimedia.org/wikipedia/commons/thumb/4/4c/Wikisource-logo.svg/12px-Wikisource-logo.svg.png")no-repeat;background-position:right .1em center.mw-parser-output code.cs1-codecolor:inherit;background:inherit;border:inherit;padding:inherit.mw-parser-output .cs1-hidden-errordisplay:none;font-size:100%.mw-parser-output .cs1-visible-errorfont-size:100%.mw-parser-output .cs1-maintdisplay:none;color:#33aa33;margin-left:0.3em.mw-parser-output .cs1-subscription,.mw-parser-output .cs1-registration,.mw-parser-output .cs1-formatfont-size:95%.mw-parser-output .cs1-kern-left,.mw-parser-output .cs1-kern-wl-leftpadding-left:0.2em.mw-parser-output .cs1-kern-right,.mw-parser-output .cs1-kern-wl-rightpadding-right:0.2em


  3. ^ Nathan Linial. Finite Metric Spaces—Combinatorics, Geometry and Algorithms, Proceedings of the ICM, Beijing 2002, vol. 3, pp573–586 Archived 2018-05-02 at the Wayback Machine


  4. ^ Open problems on embeddings of finite metric spaces, edited by Jirīı Matoušek, 2007 Archived 2010-12-26 at the Wayback Machine


  5. ^ Searcóid, p. 107.


  6. ^ "PlanetMath: a compact metric space is second countable". planetmath.org. Archived from the original on 5 February 2009. Retrieved 2 May 2018.


  7. ^ Rudin, Mary Ellen. A new proof that metric spaces are paracompact Archived 2016-04-12 at the Wayback Machine. Proceedings of the American Mathematical Society, Vol. 20, No. 2. (Feb., 1969), p. 603.


  8. ^ "metric spaces are Hausdorff". PlanetMath.


  9. ^ Goreham, Anthony. Sequential convergence in Topological Spaces Archived 2011-06-04 at the Wayback Machine. Honours' Dissertation, Queen's College, Oxford (April, 2001), p. 14


  10. ^ ab Pascal Hitzler and Anthony Seda, Mathematical Aspects of Logic Programming Semantics. Chapman and Hall/CRC, 2010.


  11. ^ "Partial metrics : welcome". www.dcs.warwick.ac.uk. Archived from the original on 27 July 2017. Retrieved 2 May 2018.




References


  • Victor Bryant, Metric Spaces: Iteration and Application, Cambridge University Press, 1985,
    ISBN 0-521-31897-1.

  • Dmitri Burago, Yu D Burago, Sergei Ivanov, A Course in Metric Geometry, American Mathematical Society, 2001,
    ISBN 0-8218-2129-6.

  • Athanase Papadopoulos, Metric Spaces, Convexity and Nonpositive Curvature, European Mathematical Society, First edition 2004,
    ISBN 978-3-03719-010-4. Second edition 2014,
    ISBN 978-3-03719-132-3.


  • Mícheál Ó Searcóid, Metric Spaces, Springer Undergraduate Mathematics Series, 2006,
    ISBN 1-84628-369-8.

  • Lawvere, F. William, "Metric spaces, generalized logic, and closed categories", [Rend. Sem. Mat. Fis. Milano 43 (1973), 135—166 (1974); (Italian summary)

This is reprinted (with author commentary) at Reprints in Theory and Applications of Categories
Also (with an author commentary) in Enriched categories in the logic of geometry and analysis. Repr. Theory Appl. Categ. No. 1 (2002), 1–37.


  • Weisstein, Eric W. "Product Metric". MathWorld.


External links



  • Hazewinkel, Michiel, ed. (2001) [1994], "Metric space", Encyclopedia of Mathematics, Springer Science+Business Media B.V. / Kluwer Academic Publishers, ISBN 978-1-55608-010-4


  • Far and near — several examples of distance functions at cut-the-knot.









Popular posts from this blog

How to check contact read email or not when send email to Individual?

Bahrain

Postfix configuration issue with fips on centos 7; mailgun relay