Basis (linear algebra)

Set of vectors used to define coordinates
File:3d two bases same vector.svg
The same vector can be represented in two different bases (purple and red arrows).

In mathematics, a set B of vectors in a vector space <span class="texhtml " {{#if:V is called a basis if every element of <span class="texhtml " {{#if:V may be written in a unique way as a finite linear combination of elements of B. The coefficients of this linear combination are referred to as components or coordinates of the vector with respect to B. The elements of a basis are called basis vectors.

Equivalently, a set B is a basis if its elements are linearly independent and every element of V is a linear combination of elements of B.[1] In other words, a basis is a linearly independent spanning set.

A vector space can have several bases; however all the bases have the same number of elements, called the dimension of the vector space.

This article deals mainly with finite-dimensional vector spaces. However, many of the principles are also valid for infinite-dimensional vector spaces.

Definition

A basis <span class="texhtml " {{#if:B of a vector space <span class="texhtml " {{#if:V over a field <span class="texhtml " {{#if:F (such as the real numbers <span class="texhtml " {{#if:R or the complex numbers <span class="texhtml " {{#if:C) is a linearly independent subset of <span class="texhtml " {{#if:V that spans <span class="texhtml " {{#if:V. This means that a subset B of <span class="texhtml " {{#if:V is a basis if it satisfies the two following conditions:

linear independence
for every finite subset of B, if for some in <span class="texhtml " {{#if:F, then ;
spanning property
for every vector <span class="texhtml " {{#if:v in <span class="texhtml " {{#if:V, one can choose in <span class="texhtml " {{#if:F and in B such that .

The scalars are called the coordinates of the vector <span class="texhtml " {{#if:v with respect to the basis <span class="texhtml " {{#if:B, and by the first property they are uniquely determined.

A vector space that has a finite basis is called finite-dimensional. In this case, the finite subset can be taken as <span class="texhtml " {{#if:B itself to check for linear independence in the above definition.

It is often convenient or even necessary to have an ordering on the basis vectors, for example, when discussing orientation, or when one considers the scalar coefficients of a vector with respect to a basis without referring explicitly to the basis elements. In this case, the ordering is necessary for associating each coefficient to the corresponding basis element. This ordering can be done by numbering the basis elements. In order to emphasize that an order has been chosen, one speaks of an ordered basis, which is therefore not simply an unstructured set, but a sequence, an indexed family, or similar; see § Ordered bases and coordinates below.

Examples

File:Basis graph (no label).svg
This picture illustrates the standard basis in R2. The blue and orange vectors are the elements of the basis; the green vector can be given in terms of the basis vectors, and so is linearly dependent upon them.

The set <span class="texhtml " {{#if:R2 of the ordered pairs of real numbers is a vector space under the operations of component-wise addition

and scalar multiplication
where is any real number. A simple basis of this vector space consists of the two vectors <span class="texhtml " {{#if:e1 = (1, 0) and <span class="texhtml " {{#if:e2 = (0, 1). These vectors form a basis (called the standard basis) because any vector <span class="texhtml " {{#if:v = (a, b) of <span class="texhtml " {{#if:R2 may be uniquely written as
Any other pair of linearly independent vectors of <span class="texhtml " {{#if:R2, such as <span class="texhtml " {{#if:(1, 1) and <span class="texhtml " {{#if:(−1, 2), forms also a basis of <span class="texhtml " {{#if:R2.

More generally, if F is a field, the set of n-tuples of elements of F is a vector space for similarly defined addition and scalar multiplication. Let

be the n-tuple with all components equal to 0, except the ith, which is 1. Then is a basis of which is called the standard basis of

A different flavor of example is given by polynomial rings. If F is a field, the collection <span class="texhtml " {{#if:F[X] of all polynomials in one indeterminate X with coefficients in F is an F-vector space. One basis for this space is the monomial basis B, consisting of all monomials:

Any set of polynomials such that there is exactly one polynomial of each degree (such as the Bernstein basis polynomials or Chebyshev polynomials) is also a basis. (Such a set of polynomials is called a polynomial sequence.) But there are also many bases for <span class="texhtml " {{#if:F[X] that are not of this form.

Properties

Many properties of finite bases result from the Steinitz exchange lemma, which states that, for any vector space V, given a finite spanning set S and a linearly independent set L of n elements of V, one may replace n well-chosen elements of S by the elements of L to get a spanning set containing L, having its other elements in S, and having the same number of elements as S.

Most properties resulting from the Steinitz exchange lemma remain true when there is no finite spanning set, but their proofs in the infinite case generally require the axiom of choice or a weaker form of it, such as the ultrafilter lemma.

If V is a vector space over a field F, then:

  • If L is a linearly independent subset of a spanning set <span class="texhtml " {{#if:SV, then there is a basis B such that
  • V has a basis (this is the preceding property with L being the empty set, and <span class="texhtml " {{#if:S = V).
  • All bases of V have the same cardinality, which is called the dimension of V. This is the dimension theorem.
  • A generating set S is a basis of V if and only if it is minimal, that is, no proper subset of S is also a generating set of V.
  • A linearly independent set L is a basis if and only if it is maximal, that is, it is not a proper subset of any linearly independent set.

If V is a vector space of dimension n, then:

  • A subset of V with n elements is a basis if and only if it is linearly independent.
  • A subset of V with n elements is a basis if and only if it is a spanning set of V.

Coordinates

Let V be a vector space of finite dimension n over a field F, and

be a basis of V. By definition of a basis, every <span class="texhtml " {{#if:v in V may be written, in a unique way, as
where the coefficients are scalars (that is, elements of F), which are called the coordinates of <span class="texhtml " {{#if:v over B. However, if one talks of the set of the coefficients, one loses the correspondence between coefficients and basis elements, and several vectors may have the same set of coefficients. For example, and have the same set of coefficients <span class="texhtml " {{#if:{2, 3}, and are different. It is therefore often convenient to work with an ordered basis; this is typically done by indexing the basis elements by the first natural numbers. Then, the coordinates of a vector form a sequence similarly indexed, and a vector is completely characterized by the sequence of coordinates. An ordered basis is also called a frame, a word commonly used, in various contexts, for referring to a sequence of data allowing defining coordinates.

Let, as usual, be the set of the n-tuples of elements of F. This set is an F-vector space, with addition and scalar multiplication defined component-wise. The map

is a linear isomorphism from the vector space onto V. In other words, is the coordinate space of V, and the n-tuple is the coordinate vector of <span class="texhtml " {{#if:v.

The inverse image by of is the n-tuple all of whose components are 0, except the ith that is 1. The form an ordered basis of , which is called its standard basis or canonical basis. The ordered basis B is the image by of the canonical basis of .

It follows from what precedes that every ordered basis is the image by a linear isomorphism of the canonical basis of , and that every linear isomorphism from onto V may be defined as the isomorphism that maps the canonical basis of onto a given ordered basis of V. In other words, it is equivalent to define an ordered basis of V, or a linear isomorphism from onto V.

Change of basis

Let <span class="texhtml " {{#if:V be a vector space of dimension n over a field <span class="texhtml " {{#if:F. Given two (ordered) bases and of <span class="texhtml " {{#if:V, it is often useful to express the coordinates of a vector x with respect to in terms of the coordinates with respect to This can be done by the change-of-basis formula, that is described below. The subscripts "old" and "new" have been chosen because it is customary to refer to and as the old basis and the new basis, respectively. It is useful to describe the old coordinates in terms of the new ones, because, in general, one has expressions involving the old coordinates, and if one wants to obtain equivalent expressions in terms of the new coordinates; this is obtained by replacing the old coordinates by their expressions in terms of the new coordinates.

Typically, the new basis vectors are given by their coordinates over the old basis, that is,

If and are the coordinates of a vector <span class="texhtml " {{#if:x over the old and the new basis respectively, the change-of-basis formula is
for <span class="texhtml " {{#if:i = 1, ..., n.

This formula may be concisely written in matrix notation. Let A be the matrix of the , and

be the column vectors of the coordinates of <span class="texhtml " {{#if:v in the old and the new basis respectively, then the formula for changing coordinates is

The formula can be proven by considering the decomposition of the vector <span class="texhtml " {{#if:x on the two bases: one has

and

The change-of-basis formula results then from the uniqueness of the decomposition of a vector over a basis, here ; that is

for <span class="texhtml " {{#if:i = 1, ..., n.

Related notions

Free module

If one replaces the field occurring in the definition of a vector space by a ring, one gets the definition of a module. For modules, linear independence and spanning sets are defined exactly as for vector spaces, although "generating set" is more commonly used than that of "spanning set".

Like for vector spaces, a basis of a module is a linearly independent subset that is also a generating set. A major difference with the theory of vector spaces is that not every module has a basis. A module that has a basis is called a free module. Free modules play a fundamental role in module theory, as they may be used for describing the structure of non-free modules through free resolutions.

A module over the integers is exactly the same thing as an abelian group. Thus a free module over the integers is also a free abelian group. Free abelian groups have specific properties that are not shared by modules over other rings. Specifically, every subgroup of a free abelian group is a free abelian group, and, if G is a subgroup of a finitely generated free abelian group H (that is an abelian group that has a finite basis), then there is a basis of H and an integer <span class="texhtml " {{#if:0 ≤ kn such that is a basis of G, for some nonzero integers . For details, see Free abelian group § Subgroups.

Analysis

In the context of infinite-dimensional vector spaces over the real or complex numbers, the term Hamel basis (named after Georg Hamel[2]) or algebraic basis can be used to refer to a basis as defined in this article. This is to make a distinction with other notions of "basis" that exist when infinite-dimensional vector spaces are endowed with extra structure. The most important alternatives are orthogonal bases on Hilbert spaces, Schauder bases, and Markushevich bases on normed linear spaces. In the case of the real numbers R viewed as a vector space over the field Q of rational numbers, Hamel bases are uncountable, and have specifically the cardinality of the continuum, which is the cardinal number , where is the smallest infinite cardinal, the cardinal of the integers.

The common feature of the other notions is that they permit the taking of infinite linear combinations of the basis vectors in order to generate the space. This, of course, requires that infinite sums are meaningfully defined on these spaces, as is the case for topological vector spaces – a large class of vector spaces including e.g. Hilbert spaces, Banach spaces, or Fréchet spaces.

The preference of other types of bases for infinite-dimensional spaces is justified by the fact that the Hamel basis becomes "too big" in Banach spaces: If X is an infinite-dimensional normed vector space which is complete (i.e. X is a Banach space), then any Hamel basis of X is necessarily uncountable. This is a consequence of the Baire category theorem. The completeness as well as infinite dimension are crucial assumptions in the previous claim. Indeed, finite-dimensional spaces have by definition finite bases and there are infinite-dimensional (non-complete) normed spaces which have countable Hamel bases. Consider , the space of the sequences of real numbers which have only finitely many non-zero elements, with the norm . Its standard basis, consisting of the sequences having only one non-zero element, which is equal to 1, is a countable Hamel basis.

Example

In the study of Fourier series, one learns that the functions <span class="texhtml " {{#if:{1} ∪ { sin(nx), cos(nx) : n = 1, 2, 3, ... } are an "orthogonal basis" of the (real or complex) vector space of all (real or complex valued) functions on the interval [0, 2π] that are square-integrable on this interval, i.e., functions f satisfying

The functions <span class="texhtml " {{#if:{1} ∪ { sin(nx), cos(nx) : n = 1, 2, 3, ... } are linearly independent, and every function f that is square-integrable on [0, 2π] is an "infinite linear combination" of them, in the sense that

for suitable (real or complex) coefficients ak, bk. But many[3] square-integrable functions cannot be represented as finite linear combinations of these basis functions, which therefore do not comprise a Hamel basis. Every Hamel basis of this space is much bigger than this merely countably infinite set of functions. Hamel bases of spaces of this kind are typically not useful, whereas orthonormal bases of these spaces are essential in Fourier analysis.

Geometry

The geometric notions of an affine space, projective space, convex set, and cone have related notions of basis.[4] An affine basis for an n-dimensional affine space is points in general linear position. A projective basis is points in general position, in a projective space of dimension n. A convex basis of a polytope is the set of the vertices of its convex hull. A cone basis[5] consists of one point by edge of a polygonal cone. See also a Hilbert basis (linear programming).

Random basis

For a probability distribution in <span class="texhtml " {{#if:Rn with a probability density function, such as the equidistribution in an n-dimensional ball with respect to Lebesgue measure, it can be shown that n randomly and independently chosen vectors will form a basis with probability one, which is due to the fact that n linearly dependent vectors <span class="texhtml " {{#if:x1, ..., <span class="texhtml " {{#if:xn in <span class="texhtml " {{#if:Rn should satisfy the equation <span class="texhtml " {{#if:det[x1xn] = 0 (zero determinant of the matrix with columns <span class="texhtml " {{#if:xi), and the set of zeros of a non-trivial polynomial has zero measure. This observation has led to techniques for approximating random bases.[6][7]

File:Random almost orthogonal sets.png
Empirical distribution of lengths N of pairwise almost orthogonal chains of vectors that are independently randomly sampled from the n-dimensional cube <span class="texhtml " {{#if:[−1, 1]n as a function of dimension, n. Boxplots show the second and third quartiles of this data for each n, red bars correspond to the medians, and blue stars indicate means. Red curve shows theoretical bound given by Eq. (1) and green curve shows a refined estimate.[7]

It is difficult to check numerically the linear dependence or exact orthogonality. Therefore, the notion of ε-orthogonality is used. For spaces with inner product, x is ε-orthogonal to y if (that is, cosine of the angle between x and y is less than ε).

In high dimensions, two independent random vectors are with high probability almost orthogonal, and the number of independent random vectors, which all are with given high probability pairwise almost orthogonal, grows exponentially with dimension. More precisely, consider equidistribution in n-dimensional ball. Choose N independent random vectors from a ball (they are independent and identically distributed). Let θ be a small positive number. Then for

 

 

 

 

(Eq. 1)

N random vectors are all pairwise ε-orthogonal with probability <span class="texhtml " {{#if:1 − θ.[7] This N growth exponentially with dimension n and for sufficiently big n. This property of random bases is a manifestation of the so-called measure concentration phenomenon.[8]

The figure (right) illustrates distribution of lengths N of pairwise almost orthogonal chains of vectors that are independently randomly sampled from the n-dimensional cube <span class="texhtml " {{#if:[−1, 1]n as a function of dimension, n. A point is first randomly selected in the cube. The second point is randomly chosen in the same cube. If the angle between the vectors was within <span class="texhtml " {{#if:π/2 ± 0.037π/2 then the vector was retained. At the next step a new vector is generated in the same hypercube, and its angles with the previously generated vectors are evaluated. If these angles are within <span class="texhtml " {{#if:π/2 ± 0.037π/2 then the vector is retained. The process is repeated until the chain of almost orthogonality breaks, and the number of such pairwise almost orthogonal vectors (length of the chain) is recorded. For each n, 20 pairwise almost orthogonal chains were constructed numerically for each dimension. Distribution of the length of these chains is presented.

Proof that every vector space has a basis

Let <span class="texhtml " {{#if:V be any vector space over some field <span class="texhtml " {{#if:F. Let <span class="texhtml " {{#if:X be the set of all linearly independent subsets of <span class="texhtml " {{#if:V.

The set <span class="texhtml " {{#if:X is nonempty since the empty set is an independent subset of <span class="texhtml " {{#if:V, and it is partially ordered by inclusion, which is denoted, as usual, by <span class="texhtml " {{#if:.

Let <span class="texhtml " {{#if:Y be a subset of <span class="texhtml " {{#if:X that is totally ordered by <span class="texhtml " {{#if:, and let <span class="texhtml " {{#if:LY be the union of all the elements of <span class="texhtml " {{#if:Y (which are themselves certain subsets of <span class="texhtml " {{#if:V).

Since <span class="texhtml " {{#if:(Y, ⊆) is totally ordered, every finite subset of <span class="texhtml " {{#if:LY is a subset of an element of <span class="texhtml " {{#if:Y, which is a linearly independent subset of <span class="texhtml " {{#if:V, and hence <span class="texhtml " {{#if:LY is linearly independent. Thus <span class="texhtml " {{#if:LY is an element of <span class="texhtml " {{#if:X. Therefore, <span class="texhtml " {{#if:LY is an upper bound for <span class="texhtml " {{#if:Y in <span class="texhtml " {{#if:(X, ⊆): it is an element of <span class="texhtml " {{#if:X, that contains every element of <span class="texhtml " {{#if:Y.

As <span class="texhtml " {{#if:X is nonempty, and every totally ordered subset of <span class="texhtml " {{#if:(X, ⊆) has an upper bound in <span class="texhtml " {{#if:X, Zorn's lemma asserts that <span class="texhtml " {{#if:X has a maximal element. In other words, there exists some element <span class="texhtml " {{#if:Lmax of <span class="texhtml " {{#if:X satisfying the condition that whenever <span class="texhtml " {{#if:Lmax ⊆ L for some element <span class="texhtml " {{#if:L of <span class="texhtml " {{#if:X, then <span class="texhtml " {{#if:L = Lmax.

It remains to prove that <span class="texhtml " {{#if:Lmax is a basis of <span class="texhtml " {{#if:V. Since <span class="texhtml " {{#if:Lmax belongs to <span class="texhtml " {{#if:X, we already know that <span class="texhtml " {{#if:Lmax is a linearly independent subset of <span class="texhtml " {{#if:V.

If there were some vector <span class="texhtml " {{#if:w of <span class="texhtml " {{#if:V that is not in the span of <span class="texhtml " {{#if:Lmax, then <span class="texhtml " {{#if:w would not be an element of <span class="texhtml " {{#if:Lmax either. Let <span class="texhtml " {{#if:Lw = Lmax ∪ {w}. This set is an element of <span class="texhtml " {{#if:X, that is, it is a linearly independent subset of <span class="texhtml " {{#if:V (because w is not in the span of <span class="texhtml " {{#if:Lmax, and <span class="texhtml " {{#if:Lmax is independent). As <span class="texhtml " {{#if:Lmax ⊆ Lw, and <span class="texhtml " {{#if:Lmax ≠ Lw (because <span class="texhtml " {{#if:Lw contains the vector <span class="texhtml " {{#if:w that is not contained in <span class="texhtml " {{#if:Lmax), this contradicts the maximality of <span class="texhtml " {{#if:Lmax. Thus this shows that <span class="texhtml " {{#if:Lmax spans <span class="texhtml " {{#if:V.

Hence <span class="texhtml " {{#if:Lmax is linearly independent and spans <span class="texhtml " {{#if:V. It is thus a basis of <span class="texhtml " {{#if:V, and this proves that every vector space has a basis.

This proof relies on Zorn's lemma, which is equivalent to the axiom of choice. Conversely, it has been proved that if every vector space has a basis, then the axiom of choice is true.[9] Thus the two assertions are equivalent.

See also

Notes

  1. Halmos, Paul Richard (1987). Finite-Dimensional Vector Spaces (4th ed.). New York: Springer. p. 10. ISBN 978-0-387-90093-3.
  2. Lua error: not enough memory.
  3. Note that one cannot say "most" because the cardinalities of the two sets (functions that can and cannot be represented with a finite number of basis functions) are the same.
  4. Lua error: not enough memory.
  5. Lua error: not enough memory.
  6. Lua error: not enough memory.
  7. 7.0 7.1 7.2 Lua error: not enough memory.
  8. Lua error: not enough memory.
  9. Lua error: not enough memory.

Lua error: not enough memory.

References

General references

  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.

Historical references

  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory., reprint: Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.
  • Lua error: not enough memory.

External links

Template:Linear algebra Lua error: not enough memory.