Linear Maps - I

An important technique to study structured sets is to study functions between such sets that preserve their structure. In the current context, the structure inherent to vector spaces is linearity. Maps between vector spaces that preserve this linear structure, called linear maps are studied now. Throughout this section, and represent finite dimensional inner product spaces of dimension and , respectively.

Basic definitions

Let us first consider the case when and are real vector spaces, not necessarily endowed with an inner product. We call a map of the form a vector space homomorphism, or more simply a linear map, if, for any , and for any , It is conventional to write as just when is a linear map. Notice how linear maps preserve the linear structure: the action of a linear map on a linear combination of vectors is the linear combination of the action of the linear map on the individual vectors.

Remark

Note that, in the definition above, the vector addition and scalar multiplication in the term are those defined in , while the ones in the term are those in . The same notation is used just to keep the notation simple.

Example

Consider the map defined as follows: for any , It is easy to verify that is a linear map.

We will now present a few definitions associated with linear maps that are useful in practice. The kernel of the linear map is the set of all elements in that are mapped to by : is also called the null space of , and is a linear subspace of . Note that the kernel is always non-empty since .

The image of the linear map between two vector spaces and , written as is defined as the set of all those elements in which are obtained by the action of on some element of : The image of a linear map is a linear subspace of its codomain. The dimension of the image of the linear map is called the rank of .

The following result holds for all linear maps between finite dimensional vector spaces: This is known as the rank-nullity theorem.

Example

Consider the map that was defined earlier as for any . In this case the null space of is seen to be Note that . The range of is easily seen to be the whole of . We therefore obtain the relation . From these results, we see that thereby verifying the rank-nullity theorem.

If a linear map is a bijection, then is called a vector space isomorphism, or just an isomorphism. In this case and are said to be isomorphic - written as .

Example

The map defined as for any , is easily seen to be an isomorphism on .

Example

Consider the set of all real-valued polynomials of one real variable of degree less than or equal to two: It is easy to check that is a real vector space with addition and scalar multiplication defined as follows: given , , and , is defined as Consider now the following map defined as follows: for any such that for any , , It is easy to check that is indeed an isomorphism.

Finally, if is a vector space isomorphism, then is also a linear map. To show this note that for any , and , Here, are the unique elements of such that and . Note that the existence and uniqueness of these vectors follows from the fact that is invertible.

Representation of linear maps

Let us now specialize the discussion to the case when and are finite dimensional inner product spaces of dimension and , respectively. For notational simplicity, we will denote the inner products in both and with the same symbol , with the meaning assumed to be evident from the context. Let be a linear map from into , and let and be orthonormal bases of and , respectively. Then, for any , we see from the linearity of that Since , we can express it in terms of the basis of as for some constants for every and . Note the order in which the indices are placed! We can now exploit the availability of an inner product in both and to see that We now collect together all the constants as an matrix whose entry is . The elements are called the components of with respect to the chosen bases. The matrix is called the matrix representation of with respect to the bases of and of .

Remark

Occasionally, the notation will be used to denote the matrix corresponding to a given linear map with respect to bases of and of . When the choice of bases is evident from the context, the simpler notation will also be used.

Example

Let us revisit the linear map . With respect to the standard basis of and , has the following matrix representation: as can be checked easily with a simple calculation.

To see the advantage of representing a linear map as a matrix, let be the effect of the action of on . We can write this equation in component form, with respect to the orthonormal bases and of and , respectively, as In matrix notation, the component form of the equation reads We thus see that The special choice of the placements of indices for is done so as to ensure a neat matrix equation of the form for the components of the various quantities with respect to chosen bases.

Remark

The foregoing discussion can be summarized through the following commutative diagram,

Representation of linear maps Commutative diagram illustrating the basis representation of linear maps

Notice how we have interpreted the matrix as a map of the form that acts on a column vector of size to produce a column vector of size . This map is called the representation of with respect to the chosen bases.

Let us now consider the case when and are equipped with general bases. Given two general bases and of and , respectively, the linear map can be represented as follows: fo any , note that Since , it can be expressed in terms of the basis of as Defining , it follows that where . Notice the similarity and difference with the corresponding expression for the components of with respect to orthonormal bases in and .

Example

Let and represent real-valued polynomials of orders and , respectively, defined on the closed interval . Given any , define their inner product as The inner product on is similarly defined. Consider the bases and of and , respectively, defined as follows: for any , It is not hard to check that and are orthonormal bases of and , respectively.

Remark

The basis is obtained by applying the Gram-Schmidt orthogonalization procedure to the basis defined as follows: for any , , , and . The reader is invited to verify this.

Let us now consider the map as follows: for any defined as , It is left as an easy exercise to verify that is a linear map.

We can work out the representation of with respect to the orthonormal bases defined earlier as follows: , for and . Doing the computation, we see that As an illustration, the computation of and are shown below: The other components are computed similarly.

To check that this representation is valid, let us verify that if and , then . Suppose, without loss of generality, that for any , , for some . We can compute the representation of with respect to the orthonormal basis of as follows: A straightforward calculation shows that It is likewise easy to compute the representation of with respect to the basis of as We therefore verify by simple matrix multiplication that :

Remark

When learning linear algebra for the first time, it is strongly recommended to work out all the details of this example step by step since it covers many of the concepts introduced earlier.

Given three vector spaces , let us consider the successive action of two linear maps and on a vector . We define the product map as . It can be shown that this translates to in matrix notation, with respect to any choice of bases in . To see this, let be general bases of , respectively. Then, for any , we see that The matrix product in the right hand side of this equation thus corresponds to the familiar matrix multiplication. Combining the expression just derived with the definition , we see that In fact, the rationale for defining matrix multiplication in the specific way it is defined is to ensure that the matrix representation of the product map is the product of the matrix representations of the individual maps.

Remark

Some authors write for the product map which we denote as . We will reserve the symbol almost exclusively for the inner product, and hence will write the product of and as . It’s a good idea to be conscious of the specific notational choices whenever you consult other references.

Change of basis for linear maps

The representation of a linear map between finite dimensional inner product spaces and , of dimension and , respectively, as a map depends on the choice of bases for both and . Specifically, recall that if and are general bases of and , respectively, then the matrix representation of is computed using the relation Suppose now that and are another choice of general bases for and , respectively. Then the components of the matrix representation of , written , are computed as The equations relating and are now worked out. Though these calculations take a much simpler form with respect to a choice of orthonormal bases for both and , the slightly more involved case involving general bases is presented below as a good algebraic exercise. Let the new bases of and of depend on the old bases of and of as where and . It follows then from an easy computation that where and . Noting that and that it follows at once that Using matrix notation, we can write the foregoing equation in matrix form as where and are the matrices whose component is and , respectively.

Remark

The foregoing discussion is summarized in the following commutative diagram:

Change of basis for linear maps Commutative diagram illustrating change of basis rules for linear maps

Notice how the commutative diagram neatly summarizes the change in the representation of a linear map upon change of bases in the domain and codomain vector spaces.

In the special case when all the bases are orthonormal, the transformation matrices are orthogonal. In this case, we can write the transformation rule as follows: This special case will turn out to be very useful in later applications.

Example

Let us revisit the projection map . We saw earlier that with respect to the standard bases of and , the linear map has the following representation: Let us now consider a different set of bases of and of , where , and is obtained by rotating by an angle about . Thus, we have the following relations: The components of with respect to the bases of and are computed using A simple calculation shows that The change of basis can also be computed using the relation where and are the matrices that related the new and old bases in and , respectively. It is left as a simple exercise to check that this yields the same representation as shown above.

Notice that in this special case, the representation of does not change for any value of . This is not true for a general linear map. Can you think of a simple geometric interpretation of this invariance of the representation of ?

Tensor product basis for $L(V,W)$

Suppose that and are inner product spaces of dimension and , respectively. Let us focus on the set , of all linear maps from into . We will now show that the set of all linear maps from into is also a vector space, and study a particularly useful basis called the tensor product basis for .

Defining addition and scalar multiplication as for any and , it is easy to check that the set has the structure of a real vector space - the two operations introduced above satisfy all the axioms of a real vector space listed earlier.

Remark

It is possible to define a norm on as , for any and . This is also called the sup norm. It is easy to show that for any and . An inner product for will be introduced later using the trace of a linear map.

What is the dimension of ? To answer this question, it is helpful to introduce the notion of a tensor product map. Given vectors and , the tensor product map is defined as follows: for any , It is easily checked that this is in fact a linear map. It is convenient to consider first the special case when and are equipped with orthonormal bases. Let and be orthonormal bases of and , respectively. Let us study the linear maps for every and . Note that for any , It is easily checked that the maps are linearly independent. Indeed, if for real numbers , where and , it is the case that then, for every , This shows that the linear maps are linearly independent. Notice how the orthonormality of the basis of and the linear independence of the basis of is used in proving this.

Suppose now that is any linear map. Then for any , we have Since this is true for any , we get the following identity: This informs us that .

The preceding two facts show that the maps indeed form a basis of , called the tensor product basis of . Since there are such maps, we see that the dimension of is : Thus, is a finite dimensional vector space with dimension equal to the product of the dimensions of and .

Remark

The change of basis rule derived earlier for the components of a linear map can also be derived using the tensor product representation. It is a simple exercise to verify this.

Example

Consider the identity map defined as follows: for any , It is trivial to check that this is a linear map. The components of with respect to the standard basis of is obtained as follows: We thus see that This representation of the identity map is very useful in applications.

Example

Considering again the projection map introduced earlier, note that, with respect to the standard bases of and , we have Compare the form of this representation with the representation of the identity map in the previous example. Notice how we can read off the fact that projects the first two components of the vector it acts on based on this analogy.

Let us briefly look at the representation of any with respect to general bases of and . Given any any , Since this is true for any , it is evident that Notice how the reciprocal basis shows up when using general bases.

Remark

Note that the linear map can be represented in a number of equivalent ways with respect to general bases of and of as follows: The default representation will be chosen in these notes as , but this is merely a matter of convention.

Transpose of a linear map

Given a linear map between finite dimensional inner product spaces, we will now construct an important linear map, , called the transpose of as follows: for any and , To get a handle on this definition and relate it to the more elementary notion of the transpose of a matrix, let us consider the representation of with respect to orthonormal bases of and . Given an orthonormal basis of and of , we can easily compute the representation with respect to these bases as follows: if and , then Here and . In matrix notation, this amounts to the following equation: We thus recover the familiar expression for the transpose of a matrix. The fact that also leads to the following representation of with respect to the bases and of and , respectively: In the special case of the linear map , where and , we see from this equation that Note that the equation is valid in general, even though orthonormal bases were used used to prove the fact. The reason is that the bases do not appear in the final form of this equation. Alternatively, the fact that can be directly checked using the definition of the transpose.

Let us now compute the representation of the transpose of a linear map with respect to general bases for and for . With respect to these choice of basis, we have: Since this equation holds true for any choice of and , it follows immediately that In the special case when the bases of and of are orthonormal, the use of the relations and yields the familiar expression for the components of the transpose: .

Example

The properties of the transpose of a linear map parallel that of the transpose of matrices. For instance, if , and , then Given linear maps and , we have the following relation: These properties are easy consequences of the definition of the transpose - it is left as an exercise to verify these claims.