Linear independence

Last updated
Linearly independent vectors in
R
3
{\displaystyle \mathbb {R} ^{3}} Vec-indep.png
Linearly independent vectors in
Linearly dependent vectors in a plane in
R
3
.
{\displaystyle \mathbb {R} ^{3}.} Vec-dep.png
Linearly dependent vectors in a plane in

In the theory of vector spaces, a set of vectors is said to be linearly independent if there exists no nontrivial linear combination of the vectors that equals the zero vector. If such a linear combination exists, then the vectors are said to be linearly dependent. These concepts are central to the definition of dimension. [1]

Contents

A vector space can be of finite dimension or infinite dimension depending on the maximum number of linearly independent vectors. The definition of linear dependence and the ability to determine whether a subset of vectors in a vector space is linearly dependent are central to determining the dimension of a vector space.

Definition

A sequence of vectors from a vector space V is said to be linearly dependent, if there exist scalars not all zero, such that

where denotes the zero vector.

This implies that at least one of the scalars is nonzero, say , and the above equation is able to be written as

if and if

Thus, a set of vectors is linearly dependent if and only if one of them is zero or a linear combination of the others.

A sequence of vectors is said to be linearly independent if it is not linearly dependent, that is, if the equation

can only be satisfied by for This implies that no vector in the sequence can be represented as a linear combination of the remaining vectors in the sequence. In other words, a sequence of vectors is linearly independent if the only representation of as a linear combination of its vectors is the trivial representation in which all the scalars are zero. [2] Even more concisely, a sequence of vectors is linearly independent if and only if can be represented as a linear combination of its vectors in a unique way.

If a sequence of vectors contains the same vector twice, it is necessarily dependent. The linear dependency of a sequence of vectors does not depend of the order of the terms in the sequence. This allows defining linear independence for a finite set of vectors: A finite set of vectors is linearly independent if the sequence obtained by ordering them is linearly independent. In other words, one has the following result that is often useful.

A sequence of vectors is linearly independent if and only if it does not contain the same vector twice and the set of its vectors is linearly independent.

Infinite case

An infinite set of vectors is linearly independent if every nonempty finite subset is linearly independent. Conversely, an infinite set of vectors is linearly dependent if it contains a finite subset that is linearly dependent, or equivalently, if some vector in the set is a linear combination of other vectors in the set.

An indexed family of vectors is linearly independent if it does not contain the same vector twice, and if the set of its vectors is linearly independent. Otherwise, the family is said to be linearly dependent.

A set of vectors which is linearly independent and spans some vector space, forms a basis for that vector space. For example, the vector space of all polynomials in x over the reals has the (infinite) subset {1, x, x2, ...} as a basis.

Geometric examples

Vectores independientes.png

Geographic location

A person describing the location of a certain place might say, "It is 3 miles north and 4 miles east of here." This is sufficient information to describe the location, because the geographic coordinate system may be considered as a 2-dimensional vector space (ignoring altitude and the curvature of the Earth's surface). The person might add, "The place is 5 miles northeast of here." This last statement is true, but it is not necessary to find the location.

In this example the "3 miles north" vector and the "4 miles east" vector are linearly independent. That is to say, the north vector cannot be described in terms of the east vector, and vice versa. The third "5 miles northeast" vector is a linear combination of the other two vectors, and it makes the set of vectors linearly dependent, that is, one of the three vectors is unnecessary to define a specific location on a plane.

Also note that if altitude is not ignored, it becomes necessary to add a third vector to the linearly independent set. In general, n linearly independent vectors are required to describe all locations in n-dimensional space.

Evaluating linear independence

The zero vector

If one or more vectors from a given sequence of vectors is the zero vector then the vector are necessarily linearly dependent (and consequently, they are not linearly independent). To see why, suppose that is an index (i.e. an element of ) such that Then let (alternatively, letting be equal any other non-zero scalar will also work) and then let all other scalars be (explicitly, this means that for any index other than (i.e. for ), let so that consequently ). Simplifying gives:

Because not all scalars are zero (in particular, ), this proves that the vectors are linearly dependent.

As a consequence, the zero vector can not possibly belong to any collection of vectors that is linearly independent.

Now consider the special case where the sequence of has length (i.e. the case where ). A collection of vectors that consists of exactly one vector is linearly dependent if and only if that vector is zero. Explicitly, if is any vector then the sequence (which is a sequence of length ) is linearly dependent if and only if ; alternatively, the collection is linearly independent if and only if

Linear dependence and independence of two vectors

This example considers the special case where there are exactly two vector and from some real or complex vector space. The vectors and are linearly dependent if and only if at least one of the following is true:

  1. is a scalar multiple of (explicitly, this means that there exists a scalar such that ) or
  2. is a scalar multiple of (explicitly, this means that there exists a scalar such that ).

If then by setting we have (this equality holds no matter what the value of is), which shows that (1) is true in this particular case. Similarly, if then (2) is true because If (for instance, if they are both equal to the zero vector ) then both (1) and (2) are true (by using for both).

If then is only possible if and; in this case, it is possible to multiply both sides by to conclude This shows that if and then (1) is true if and only if (2) is true; that is, in this particular case either both (1) and (2) are true (and the vectors are linearly dependent) or else both (1) and (2) are false (and the vectors are linearly independent). If but instead then at least one of and must be zero. Moreover, if exactly one of and is (while the other is non-zero) then exactly one of (1) and (2) is true (with the other being false).

The vectors and are linearly independent if and only if is not a scalar multiple of and is not a scalar multiple of .

Vectors in R2

Three vectors: Consider the set of vectors and then the condition for linear dependence seeks a set of non-zero scalars, such that

or

Row reduce this matrix equation by subtracting the first row from the second to obtain,

Continue the row reduction by (i) dividing the second row by 5, and then (ii) multiplying by 3 and adding to the first row, that is

Rearranging this equation allows us to obtain

which shows that non-zero ai exist such that can be defined in terms of and Thus, the three vectors are linearly dependent.

Two vectors: Now consider the linear dependence of the two vectors and and check,

or

The same row reduction presented above yields,

This shows that which means that the vectors and are linearly independent.

Vectors in R4

In order to determine if the three vectors in

are linearly dependent, form the matrix equation,

Row reduce this equation to obtain,

Rearrange to solve for v3 and obtain,

This equation is easily solved to define non-zero ai,

where can be chosen arbitrarily. Thus, the vectors and are linearly dependent.

Alternative method using determinants

An alternative method relies on the fact that vectors in are linearly independent if and only if the determinant of the matrix formed by taking the vectors as its columns is non-zero.

In this case, the matrix formed by the vectors is

We may write a linear combination of the columns as

We are interested in whether AΛ = 0 for some nonzero vector Λ. This depends on the determinant of , which is

Since the determinant is non-zero, the vectors and are linearly independent.

Otherwise, suppose we have vectors of coordinates, with Then A is an n×m matrix and Λ is a column vector with entries, and we are again interested in AΛ = 0. As we saw previously, this is equivalent to a list of equations. Consider the first rows of , the first equations; any solution of the full list of equations must also be true of the reduced list. In fact, if i1,...,im is any list of rows, then the equation must be true for those rows.

Furthermore, the reverse is true. That is, we can test whether the vectors are linearly dependent by testing whether

for all possible lists of rows. (In case , this requires only one determinant, as above. If , then it is a theorem that the vectors must be linearly dependent.) This fact is valuable for theory; in practical calculations more efficient methods are available.

More vectors than dimensions

If there are more vectors than dimensions, the vectors are linearly dependent. This is illustrated in the example above of three vectors in

Natural basis vectors

Let and consider the following elements in , known as the natural basis vectors:

Then are linearly independent.

Proof

Suppose that are real numbers such that

Since

then for all

Linear independence of functions

Let be the vector space of all differentiable functions of a real variable . Then the functions and in are linearly independent.

Proof

Suppose and are two real numbers such that

Take the first derivative of the above equation:

for all values of We need to show that and In order to do this, we subtract the first equation from the second, giving . Since is not zero for some , It follows that too. Therefore, according to the definition of linear independence, and are linearly independent.

Space of linear dependencies

A linear dependency or linear relation among vectors v1, ..., vn is a tuple (a1, ..., an) with n scalar components such that

If such a linear dependence exists with at least a nonzero component, then the n vectors are linearly dependent. Linear dependencies among v1, ..., vn form a vector space.

If the vectors are expressed by their coordinates, then the linear dependencies are the solutions of a homogeneous system of linear equations, with the coordinates of the vectors as coefficients. A basis of the vector space of linear dependencies can therefore be computed by Gaussian elimination.

Generalizations

Affine independence

A set of vectors is said to be affinely dependent if at least one of the vectors in the set can be defined as an affine combination of the others. Otherwise, the set is called affinely independent. Any affine combination is a linear combination; therefore every affinely dependent set is linearly dependent. Conversely, every linearly independent set is affinely independent.

Consider a set of vectors of size each, and consider the set of augmented vectors of size each. The original vectors are affinely independent if and only if the augmented vectors are linearly independent. [3] :256

Linearly independent vector subspaces

Two vector subspaces and of a vector space are said to be linearly independent if [4] More generally, a collection of subspaces of are said to be linearly independent if for every index where [4] The vector space is said to be a direct sum of if these subspaces are linearly independent and

See also

Related Research Articles

In mathematics, and more specifically in linear algebra, a linear map is a mapping between two vector spaces that preserves the operations of vector addition and scalar multiplication. The same names and the same definition are also used for the more general case of modules over a ring; see Module homomorphism.

In linear algebra, the rank of a matrix A is the dimension of the vector space generated by its columns. This corresponds to the maximal number of linearly independent columns of A. This, in turn, is identical to the dimension of the vector space spanned by its rows. Rank is thus a measure of the "nondegenerateness" of the system of linear equations and linear transformation encoded by A. There are multiple equivalent definitions of rank. A matrix's rank is one of its most fundamental characteristics.

<span class="mw-page-title-main">Affine transformation</span> Geometric transformation that preserves lines but not angles nor the origin

In Euclidean geometry, an affine transformation or affinity is a geometric transformation that preserves lines and parallelism, but not necessarily Euclidean distances and angles.

<span class="mw-page-title-main">Linear span</span> In linear algebra, generated subspace

In mathematics, the linear span (also called the linear hull or just span) of a set S of vectors (from a vector space), denoted span(S), is defined as the set of all linear combinations of the vectors in S. For example, two linearly independent vectors span a plane. The linear span can be characterized either as the intersection of all linear subspaces that contain S, or as the smallest subspace containing S. The linear span of a set of vectors is therefore a vector space itself. Spans can be generalized to matroids and modules.

<span class="mw-page-title-main">Linear subspace</span> In mathematics, vector subspace

In mathematics, and more specifically in linear algebra, a linear subspace or vector subspace is a vector space that is a subset of some larger vector space. A linear subspace is usually simply called a subspace when the context serves to distinguish it from other types of subspaces.

<span class="mw-page-title-main">Gram–Schmidt process</span> Orthonormalization of a set of vectors

In mathematics, particularly linear algebra and numerical analysis, the Gram–Schmidt process or Gram-Schmidt algorithm is a way of making two or more vectors perpendicular to each other.

<span class="mw-page-title-main">Matrix addition</span> Notions of sums for matrices in linear algebra

In mathematics, matrix addition is the operation of adding two matrices by adding the corresponding entries together.

<span class="mw-page-title-main">Cross product</span> Mathematical operation on vectors in 3D space

In mathematics, the cross product or vector product is a binary operation on two vectors in a three-dimensional oriented Euclidean vector space, and is denoted by the symbol . Given two linearly independent vectors a and b, the cross product, a × b, is a vector that is perpendicular to both a and b, and thus normal to the plane containing them. It has many applications in mathematics, physics, engineering, and computer programming. It should not be confused with the dot product.

In mathematics, the dot product or scalar product is an algebraic operation that takes two equal-length sequences of numbers, and returns a single number. In Euclidean geometry, the dot product of the Cartesian coordinates of two vectors is widely used. It is often called the inner product of Euclidean space, even though it is not the only inner product that can be defined on Euclidean space.

In linear algebra, a diagonal matrix is a matrix in which the entries outside the main diagonal are all zero; the term usually refers to square matrices. Elements of the main diagonal can either be zero or nonzero. An example of a 2×2 diagonal matrix is , while an example of a 3×3 diagonal matrix is. An identity matrix of any size, or any multiple of it is a diagonal matrix called scalar matrix, for example, . In geometry, a diagonal matrix may be used as a scaling matrix, since matrix multiplication with it results in changing scale (size) and possibly also shape; only a scalar matrix results in uniform change in scale.

In the mathematical field of differential geometry, a metric tensor is an additional structure on a manifold M that allows defining distances and angles, just as the inner product on a Euclidean space allows defining distances and angles there. More precisely, a metric tensor at a point p of M is a bilinear form defined on the tangent space at p, and a metric tensor on M consists of a metric tensor at each point p of M that varies smoothly with p.

In mathematics, a linear form is a linear map from a vector space to its field of scalars.

In quantum field theory, the Dirac spinor is the spinor that describes all known fundamental particles that are fermions, with the possible exception of neutrinos. It appears in the plane-wave solution to the Dirac equation, and is a certain combination of two Weyl spinors, specifically, a bispinor that transforms "spinorially" under the action of the Lorentz group.

In mathematics, the covariant derivative is a way of specifying a derivative along tangent vectors of a manifold. Alternatively, the covariant derivative is a way of introducing and working with a connection on a manifold by means of a differential operator, to be contrasted with the approach given by a principal connection on the frame bundle – see affine connection. In the special case of a manifold isometrically embedded into a higher-dimensional Euclidean space, the covariant derivative can be viewed as the orthogonal projection of the Euclidean directional derivative onto the manifold's tangent space. In this case the Euclidean derivative is broken into two parts, the extrinsic normal component and the intrinsic covariant derivative component.

In mathematics, the Kronecker product, sometimes denoted by ⊗, is an operation on two matrices of arbitrary size resulting in a block matrix. It is a specialization of the tensor product from vectors to matrices and gives the matrix of the tensor product linear map with respect to a standard choice of basis. The Kronecker product is to be distinguished from the usual matrix multiplication, which is an entirely different operation. The Kronecker product is also sometimes called matrix direct product.

In mathematics, the kernel of a linear map, also known as the null space or nullspace, is the linear subspace of the domain of the map which is mapped to the zero vector. That is, given a linear map L : VW between two vector spaces V and W, the kernel of L is the vector space of all elements v of V such that L(v) = 0, where 0 denotes the zero vector in W, or more symbolically:

In linear algebra, it is often important to know which vectors have their directions unchanged by a linear transformation. An eigenvector or characteristic vector is such a vector. Thus an eigenvector of a linear transformation is scaled by a constant factor when the linear transformation is applied to it: . The corresponding eigenvalue, characteristic value, or characteristic root is the multiplying factor .

In mathematics, especially in linear algebra and matrix theory, the commutation matrix is used for transforming the vectorized form of a matrix into the vectorized form of its transpose. Specifically, the commutation matrix K(m,n) is the nm × mn matrix which, for any m × n matrix A, transforms vec(A) into vec(AT):

In statistics, Bayesian multivariate linear regression is a Bayesian approach to multivariate linear regression, i.e. linear regression where the predicted outcome is a vector of correlated random variables rather than a single scalar random variable. A more general treatment of this approach can be found in the article MMSE estimator.

In linear algebra, eigendecomposition is the factorization of a matrix into a canonical form, whereby the matrix is represented in terms of its eigenvalues and eigenvectors. Only diagonalizable matrices can be factorized in this way. When the matrix being factorized is a normal or real symmetric matrix, the decomposition is called "spectral decomposition", derived from the spectral theorem.

References

  1. G. E. Shilov, Linear Algebra (Trans. R. A. Silverman), Dover Publications, New York, 1977.
  2. Friedberg, Stephen; Insel, Arnold; Spence, Lawrence (2003). Linear Algebra. Pearson, 4th Edition. pp. 48–49. ISBN   0130084514.
  3. Lovász, László; Plummer, M. D. (1986), Matching Theory, Annals of Discrete Mathematics, vol. 29, North-Holland, ISBN   0-444-87916-1, MR   0859549
  4. 1 2 Bachman, George; Narici, Lawrence (2000). Functional Analysis (Second ed.). Mineola, New York: Dover Publications. ISBN   978-0486402512. OCLC   829157984. pp. 3–7