Skip to content

Functions and Linear Transformations and Matrices

Addendic B: Functions

function

If A and B are sets, then a function f from A to B, written f:AB, is a rule that associates to each element x in A a unique element denoted f(x) in B. The element f(x) is called the image of x (under f), and x is called a preimage of f(x) (under f).

If f:AB, then A is called the domain of f, B is called the codomain of f, and the set {f(x):xA} is called the range of f. Note that the range of f is a subset of B. If SA, we denote by f(S) the set {f(x):xS} of all images of elements of S. Likewise, if TB, we denote by f1(T) the set {xA:f(x)T} of all preimages of elements in T.

Finally, two functions f:AB and g:AB are equal, written f=g, if f(x)=g(x) for all xA.

Example B.1

Suppose that A=[10,10]. Let f:AR be the functio that assignss to each element x in A the element x2+1 in R; that is, f is defined by f(x)=x2+1.

Then A is the domain of f, R is the codmain of f, and [1,101] is the range of f. Since f(2)=5, the image of 2 is 5, and 2 is a preimage of 5. Notice that 2 is another preimage of 5. Moreover, if S=[1,2] and T=[82,101], then f(S)=[2,5] and f1(T)=[10,9][9,10].

one-to-one, onto

As example 1 shows, the preimage of an element in the range need not be unique. Functions s.t. each element of the range has a unique preimage are called one-to-one (injective); that is f:AB is one-to-one if f(x)=f(y) implies x=y or, equivalently, if xy implies f(x)f(y).

If f:AB is a function with range B, that is, if f(A)=B, then f is called onto (surjective).So f is onto iff the range of f equals the codomain of f. [core]

one-to-one + onto = bijective

Let A,B,C be sets and f:AB and g:BC be functios. By following f with g, we obtain a function gf:AC called the composite of g and f. Thus (gf)(x)=g(f(x)),xA.

  • For example, let A=B=C=R,f(x)=sinx, and g(x)=x2+3. Then (gf)(x)=(g(f(x)))=sin2x+3, whereas (fg)(x)=f(g(x))=f(g(x))=sin(x2+3). Hence, gffg.
  • Functionalcomposition is associative, i.e. if h:CD is another function, then h(gf)=(hg)f.

A function f:AB is said to e invertible if there exists a function g:BA s.t. (fg)(y)=y,yB and (gf)(x)=x,xA. If such a function g exists, then it's unique and is called the inverse of f. We denote the inverse of f (when it exists) by f1. It can be shown that f is invertible iff f is both one-to-one and onto.

The following facts about invertible functions are easily proved:

  1. If f:AB is invertible, then f1 is invertible, and (f1)1=f.
  2. If f:AB and g:BC ar invertible, then gf is invertible, and (gf)1=f1g1.

2.1 Linear Transformations, Null Spaces, and Ranges

Linear Transformation

Now it's natural to consider those functions defined on vector spaces that in some sense "preserve" the structure.

Definition: linear transformation

Let V,W be vector spaces (over F). We call a function T:VW a linear transformation from V to W if, x,yV and cF, we have

  • T(x+y)=T(x)+T(y) (additivity)
  • T(cx)=cT(x) (scaling)

Often simply all T linear. Following are properties of a function: T:VW:

Properties

  1. If T is linear, then T(0)=0;
  2. T is linear iff T(cx+y)=cT(x)+T(y),x,yV and cF; (Generally used to prove that a given transformation is linear)[core]
  3. If T is linear, then T(xy)=T(x)T(y),x,yV;
  4. T is linear iff for x1,x2,,xnV and a1,a2,,anF, we have
T(i=1naixi)=i=1naiT(xi).

Example 1

Define

T:R2R2 by T(a1,a2)=(2a1+a2,a1).

To show that T is linear, let cR and x,yR2, where x=(b1,b2) and y=(d1,d2). Since

cx+y=(cb1+d1,cb2+d2)

we have

T(cx+y)=(2(cb1+d1)+cb2+d2,cb1+d1).

Also

c(x)+(y)=c(2b1+b2,b1)+(2d1+d2,d1)=(2cb1+cb2+2d1+d2,cb1+d1)=(2(cb1+d1)+cb2+d2,cb1+d1).

So T is linear.

Example 2: rotation

For any angle θ, define Tθ:R2R2 by the rule: Tθ(a1,a2) is the vector obtained by rotating (a1,a2) counterclockwise by θ if (a1,a2)(0,0), and Tθ(0,0)=(0,0).

Then Tθ:R2R2 is a linear transformation that is called the rotation by θ.

We determine an explicit formula for Tθ. Fix a nonzero vector (a1,a2)R2. Let α be the angle that (a1,a2) makes with the positive x-axis, and let r=a12+a22. Then a1=rcosα and a2=rsinα. Also, Tθ(a1,a2) has length r and makes an angle α+θ with the positive x-axis. It follows that

Tθ(a1,a2)=(rcos(α+θ),rsin(α+θ))=(rcosαcosθrsinαsinθ,rcosαsinθ+rsinαcosθ)=(a1cosθa2sinθ,a1sinθ+a2cosθ)

Finally, observe that this same formula is valid for (a1,a2)=(0,0). It is now easy to show, as in Example 1, that Tθ is linear.

Example 3&4: Reflection

Define T:R2R2 by T(a1,a2)=(a1,a2). T is called the reflection about the x-axis

Define T:R2R2 by T(a1,a2)=(a1,a2). T is called the reflection about the y-axis

Example 5: Polynominal

Define T:Pn(R)Pn1(R) by T(f(x))=f(x), where f(x) denotes the derivative of f(x). To show that T is linear, let g(x),h(x)Pn(R) and aR. Now

T(ag(x)+h(x))=(ag(x)+h(x))=ag(x)+h(x)=aT(g(x))=T(h(x)).

So by property 2 above, T is linear.

Example 6: integration

Let V=C(R), the vector space of continous real-valued functions on R. Let a,bR,a<b. Define T:VR by

T(f)=abf(t)dt.

for all fV. Then T is a linear transformation because the definte integral of a linear combination of functions is the same as the linear combination f he definite integrals of the fnctions.

Special transformations

For vector spaces V,W (over F), we define the identity transformation IV:VV by IV(x)=x,xV and the zero transformationT0:VW by T0(x)=0 for all xV. It is clear that both of these transformations are linear. We often write I instead of IV.

We now turn our attention to two very important sets associated with linear transformations: the range and null space. The determination of these sets allows us to examine more closely the intrinsic properties of a linear transformation.

Definitions: null space/kerne; range / image [core]

Let V,W be vector spaces, and let T:VW be linear. We define the null space (or kernel) N(T) of T to be the set of all vectors x in V s.t. T(x)=0; i.e., N(T)={xV:T(x)=0}.

We define the range (or image) R(T) of T to be the subset of W consisting of all images (under T) of vectors in V; i.e., R(T)={T(x):xV}.

Example 7: identity and zero transformation

Let V,W be vector spaces, and let I:VV and T0:VW be the identity and zero transformation, respectively. Then N(I)={0},R(I)=V,N(T0)=V,R(T0)={0}

Example 8

Let T:R3R2 be the linear transformation defined by

T(a1,a2,a3)=(a1a2,2a3).

It's easy to verify that:

N(T)={(a,a,0):aR}, and R(T)=R2.

Theorem 2.1

Let V,W be vector spaces and T:VW be linear. Then N(T) and R(T) are subspaces of V,W, respectively.

Theorem 2.2

Let V,W be vector spaces and T:VW be linear. If β={v1,v2,vn} is a basis for V, then

R(T)=span(T(β))=span({T(v1),T(v2),,T(vn)}).

Recall Theorem 1.5 in §1 that the relation between subspace and span.

Example 9

Define the linear transformation T:P2(R)M2×2(R) by

T(f(x))=(f(1)f(2)00f(0))

Since β={1,x,x2} is a basis for P2(R), we have

R(T)=span(T(β))=span({T(1),T(x),T(x2)})=span({(0001),(1000),(3000)})=span({(0001),(1000)})

Thus, we have found a basis for R(T), so dim(R(T))=2.

Definition and Theorem

As in §1.4, we measure the "size" of a subspace by its dimension. The null space and range are so important that we attach special names to their respective dimensions.

Definition: nulity; rank [core]

Let V and W be vector spaces, and let T:VW be linear. If N(T) and R(T) are finite-dimensional, then we define:

  • Nullity of T, denoted nullity(T), as the dimension of N(T).
  • Rank of T, denoted rank(T), as the dimension of R(T).

Theorem 2.3: Dimension Theorem [core]

Let V and W be vector spaces, and T:VW linear. If V is finite-dimensional, then

nullity(T)+rank(T)=dim(V)

Theorem 2.4: Null Space and One-to-one [core]

Let V and W be vector spaces, and T:VW linear. Then T is one-to-one iff

N(T)={0}

Theorem 2.5 [core]

Let V and W be finite-dimensional vector spaces of equal and finite dimension, and T:VW linear. The following are equivalent:

  • T is one-to-one.
  • T is onto.
  • rank(T)=dim(V).

Example 10

Let T:P2(R)P3(R) be the linear transformation defined by

T(f(x))=2f(x)+0x3f(t)dt

Now

R(T)=span({T(1),T(x),T(x2)})=span({3x,2+32x2,4x+x3}).
  • Since span({3x,2+32x2,4x+x3}) is linear independent, The rank is 3;
  • Since dim(P3(R))=4,T is not onto;
  • From the dimension theorem, nullity(T)+3=3, so nullity is 0, and therefore N(T)={0}.

So T is one-to-one but not onto.

Example 11

Let T:F2F2 be the linear transformation defined by

T(a1,a2)=(a1+a2,a1)

It's easy to see that N(T)={0}; so T is one-to-one. And Theorem 2.5 tells us that T must be onto.

Example 12

Let T:P2(R)R3 be the linear transformation defined by

T(a0+a1x+a2x2)=(a0,a1,a2)

Clearly T is linear and one-to-one. Let S={1x+3x2,x+x2,12x2}. Then S is linearly independent in P2(R) because

T(S)={(2,1,3),(0,1,1),(1,0,2)}

is linearly independent in R3.

Theorem 2.6

Theorem 2.6

Let V and W be vector spaces over F, and suppose that {v1,v2,,vn} is a basis for V. For w1,w2,,wn in W, there exists exactly one linear transformation: T:VW s.t. T(vi)=wi,for i=1,2,,n.

Corollary

Let V and W be vector spaces, and suppose that V has a finite basis {v1,v2,,vn}. If U,T:VW are linear and U(vi)=T(vi) for i=1,2,,n, then U=T.

Example 13

Let T:R2R2 be the linear transformation defined by

T(a1,a2)=(2a2a1,3a1).

and suppose that U:R2R2 is linear. If we know that U(1,2)=(3,3) and U(1,1)=(1,3), then U=T. This follows from the corollary and from the fact that {(1,2),(1,1)} is the basis for R2.

2.2 The Matrix Representation of a Linear Transformations

coordinate vector

Definition: ordered basis

Let V be a finite-dimensional vector space. An ordered basis for V is a basis for V endowed with a specific order; that is, an ordered basis for V is a finite sequence of linearly independent vectors in V that generates V.

Example 1

In F3,β={e1,e2,e3} can be considered as ordered basis. Also γ={e2,e1,e3} is an ordered basis, but βγ as ordered bases

For the vector space Fn, we call {e1,e2,,en} the standard ordered basis for Fn. Similarly, for the vector space Pn(F), we call {1,x,,xn} the standard ordered basis for Pn(F).

Definition: coordinate vector

Let β={v1,v2,···,vn} be an ordered basis for a finite-dimensional vector space V. For xV, let a1,a2,,an be the unique scalars s.t.

x=i=1naiui

We define the coordinate vecotr of x relative to β, denoted [x]β [core]. by

[x]β=(a1,a2,,an).

Example 2: coordinate vector for polynominal

Let V=P2(R), an let β={1,x,x2} be the standard ordered basis for V. If f(x)=4+6x7x2, then

[f]β=(4,6,7).

matrix representation

Definition: matrix representation [core]

Using the notation above, we call the m×n matrix A defined by Aij=aij the matrix representation of T in the ordered bases β and γ and write A=[T]βγ .

If V=W and β=γ, then we write A=[T]β.

Notice that the jth column of A is simply [T(vj)]γ. Also observe that if U:VW is a linear transformation s.t. [U]βγ=[T]βγ, then U=T by the corollary to theorem 2.6.

Example 3: matrix representation for tuples

Let T:R2R3 be defined by

T(a1,a2)=(a1+3a2,0,2a14a2)

Let β,γ be the standard ordered bases for R2 and R3, respectively. Now

T(1,0)=(1,0,2)=1e1+0e2+2e3.

and

T(0,1)=(3,0,4)=3e1+0e2+4e3.

Hence

[T]βγ=(130024)

If we let γ={e3,e2,e1}, then

[T]βγ=(240013)

Example 4: matrix representation for polynominal

Let T:P3(R)P2(R) be the linear transformation defined by T(f(x))=f(x).Let β,γ be the standard ordered bases for P3(R) and P2(R), respectively. Then

T(1)=01+0x+0x2T(x)=11+0x+0x2T(x2)=01+2x+0x2T(x3)=01+0x+3x2

So

[ddx]=[T]βγ=(010000200003)

i.e.

f(x)=a+bx+cx2+dx3T(f(x))=f(x)=b+2cx+3dx2[b2c3d]=[010000200003][abcd]

Note that when T(xj) is written as a linear combination of the vectors of γ, its coefficients gives the entries of the jth column of [T]βγ.

operations on linear transformation

Definition

Let T,U:VW be arbitary functions, where V,W are vector spaces over F, and let aF. We define:

  • T+U:VW by (T+U)(x)=T(x)+U(x) for all xV;
  • aT:VW by (aT)(x)=aT(x) for all xV;

Theorem 2.7

Let V,W be vector spaces over a field F, and let T,U:VW be linear.

  • For all aF,aT+U is linear.
  • Using the operations of addition andscalar multiplication in the precedingg definition, the collection of all linear transformation from V to W is a vector space over F.

Definition

Let V,W be vector spaces over F. We denote the vector space of all linear transformation from V to W by L(V,W). In the case that V=W, we write L(V) instead of L(V,W).

Theorem 2.8

Let V,W be finite-dimensional vector spaces with ordered bases β,γ, respectively, and let T,U:VW be linear transformations. Then

  • [T+U]βγ=[T]βγ+[U]βγ
  • [aT]βγ=a[T]βγ for all scalars a.

Example 5

Let T:R2R3 and U:R2R3 be the linear transformations respectively defined by

T(a1,a2)=(a1+3a2,0,2a14a2)U(a1,a2)=(a1a2,2a1,3a1+2a2)

Let β,γ be the standard ordered bases of R2R3, respectively. Then

[T]βγ=(130024)

(as completed in Example 3), and

[U]βγ=(112032)

If we compute T+U usingthe proceding definitions, we obtain

(T+U)(a1,a2)=(2a1+2a2,2a1,5a12a2).

So

[T+U]βγ=(222052)

which is simply [T]βγ+[U]βγ, illustrated Theorem 2.8

2.3 Composition of Linear Transformations and Matrix Multiplication

matrix product

Theorem 2.9

Let V,W,Z be vector spaces over the same field F, and let T:VW and U:WZ be linear, then UT:VZ is linear.

Theorem 2.10

Let V be a vector space. Let T,U1,U2L(V). Then

  • T(U1+U2)=TU1+TU2 and (U1+U2)T=U1T+U2T
  • T(U1U2)=(TU1)U2
  • TI=IT=T
  • a(U1U2)=(aU1)U2=U1(aU2) for all scalars a.

A more general result holds for linear transformations that have domains unequal to their codomains.

Definition: product

Let T:VW and U:WZ be linear transformations, and let A=[U]βγ and B=[T]αβ, where α={v1,v2,,vn},β={w1,w2,,wm}, and γ={z1,z2,,zp} are ordered bases for V,W, and Z , respectively. We would like to define the product AB of two matrices so that AB=[UT]αγ. Consider the matrix [UT]αγ. For 1jn, we have

(UT)(vj)=U( T(vj))=U(k=1mBkjwk)=k=1mBkjU(wk)=k=1mBkj(i=1pAikzi)=i=1p(k=1mAikBkj)zi=i=1pCijzi

where

Cij=k=1mAikBkj.

This computation motivates the following definition of matrix multiplication.

why matrix multiplication is defined this way - it perfectly describes the composition of basis transformations. It's not an arbitrary definition, but rather a natural consequence of how basis transformations compose.

Definition: product of Matrix

Let A be an m×n matrix and B be an n×p matrix. We define the product of A and B, denoted AB, to be the m×p matrix s.t.

(AB)ij=k=1nAikBkjfor 1im,1jp

Example 2.3.1

We have

(121041)(425)=(14+22+1504+42+(1)5)=(133)

Notice again the symbolic relationship (2×3)(3×1)=2×1.

As in the case with composition of functions, we have the matrix multiplication is not commutative. It need not be true that AB=BA.

transpose

The transpose At of an m×n matrix A is the n×m matrix obtained from A by interchanging the rows with the columns; that is, (At)ij=Aji. For example,

(123051)t=(102531) and (1223)t=(1223)

show that if A is an m×n matrix and B is an n×p matrix, then (AB)t=BtAt. Since

(AB)ijt=(AB)ji=k=1nAjkBki

and

(BtAt)ij=k=1n(Bt)ik(At)kj=k=1nBkiAjk,

Therefore the transpose of a product is the product of the transposes in the opposite order.

Immediate consequence of our definition of matrix multiplication:

Theorem 2.11 [core]

Let V,W, and Z be finite-dimensional vector spaces with ordered bases α,β, and γ, respectively. Let T:VW and U:WZ be linear transformations. Then

[UT]αγ=[U]βγ[T]αβ.

Corollary

Let V be a finite-dimensional vector space with an ordered basis β. Let T,UL(V). The [UT]β=[U]β[T]β

Example 2.3.2 [core]

Let U:P3(R)P2(R) and T:P2(R)P3(R) be the linear transformations respectively defined by

U(f(x))=f(x) and T(f(x))=0xf(t)dt

Let α and β be the standard ordered bases of P3(R) and P2(R), respectively. From calculus, it follows that UT=I, the identity transformation on P2(R). To illustrate Theorem 2.11, observe that

β=[U]αβ[T]βα=(010000200003)(00010001200013)=(100010001)=[I]β

The preceding 3×3 diagonal matrix is called an identity matrix and is defined next, along with a very useful notation, the Kronecker delta.

Definition and theorem

Definition: identity matrix

We define the Kronecker delta δij by δij=1 if i=j and δij=0 if ij.then n×n identity matrix In is defined by (In)ij=δij.

Theorem 2.12.

Let A be an m×n matrix, B and C be n×p matrices, and D and E be q×m matrices. Then

  • A(B+C)=AB+AC and (D+E)A=DA+EA.
  • a(AB)=(aA)B=A(aB) for any scalar a.
  • ImA=A=AIn.
  • If V is an n-dimensional vector space with an ordered basis β, then [IV]β=In.

Corollary.

Let A be an m×n matrix, B1,B2,,Bk be n×p matrices, C1,C2,,Ck be q×m matrices, and a1,a2,,ak be scalars. Then

A(i=1kaiBi)=i=1kaiABi

and

(i=1kaiCi)A=i=1kaiCiA

For an n×n matrix A, we define A1=A,A2=AA,A3=A2A, and, in general, Ak=Ak1A for k=2,3,. We define A0=In.

With this notation, we see that if

A=(0010)

then A2=O (the zero matrix) even though AO. Thus the cancellation property for multiplication in fields is not valid for matrices. To see why, assume that the cancellation law is valid. Then, from AA=A2=O=AO, we would conclude that A=O, which is false.

Theorem 2.13

Let A be an m×n matrix and B be an n×p matrix. For each j(1jp) let uj and vj denote the j th columns of AB and B, respectively. Then

  • uj=Avj
  • vj=Bej, where ej is the jth standard vector over Fp.

Theorem 2.14 [core]

Let V and W be finite-dimensional vector spaces having ordered bases β and γ, respectively, and let T:VW be linear. Then, for each u V, we have

[T(u)]γ=[T]βγ[u]β.

Example 3

Let T:P3(R)P2(R) be the linear transformation defined by T(f(x))=f(x), and let β and γ be the standard ordered bases for P3(R) and P2(R), respectively. If A=[T]βγ, then, from Example 4 of Section 2.2, we have

A=(010000200003)

We illustrate Theorem 2.14 by verifying that [T(p(x))]γ=[T]βγ[p(x)]β, where p(x)P3(R) is the polynomial p(x)=24x+x2+3x3. Let q(x)=T(p(x)); then q(x)=p(x)=4+2x+9x2. Hence

[T(p(x))]γ=[q(x)]γ=(429)

but also

[T]βγ[p(x)]β=A[p(x)]β=(010000200003)(2413)=(429).

left-multiplication transformation

Definition: left-multiplication transformation

Let A be an m×n matrix with entries from a field F. We denote by La the mapping LA:FnFm defined by LA(x)=Ax (the matrix product of A and x) for each column vector xFn. We call LA a left-multiplication transformation.

Example 4

Let

A=(121012)

Then AM2×3(R) and LA:R3R2. If

x=(131)

then

LA(x)=Ax=(121012)(131)=(61)

Theorem 2.15

Let A be an m×n matrix with entries from F. Then the left-multiplication transformation LA:FnFm is linear. Furthermore, if B is any other m×n matrix (with entries from F) and β and γ are the standard ordered bases for Fn and Fm, respectively, then we have the following properties.

  • [LA]βγ=A.
  • LA=LB if and only if A=B.
  • LA+B=LA+LB and LaA=aLA for all aF.
  • If T:FnFm is linear, then there exists a unique m×n matrix C such that T=Lc. In fact, C=[T]βγ.
  • If E is an n×p matrix, then LAE=LALE.
  • If m=n, then LIn=IFn.

Theorem 2.16: Associative in Matrix Multiplication

Let A,B, and C be matrices such that A(BC) is defined. Then (AB)C is also defined and A(BC)=(AB)C; that is, matrix multiplication is associative.

2.4 Invertibility and Isomorphisms

inverse

Definition: inverse

Let V and W be vector spaces, and let T:VW be linear. A function U:WV is said to be an inverse of T if TU=IW and UT=IV. If T has an inverse, then T is said to be invertible. The inverse of T is unique and is denoted by T1.
The following facts hold for invertible functions T and U:

  1. (TU)1=U1T1.
  2. (T1)1=T; in particular, T1 is invertible.

We often use the fact that a function is invertible if and only if it is both one-to-one and onto. We can therefore restate Theorem 2.5 as follows.

  1. Let T:VW be a linear transformation, where V and W are finite-dimensional spaces of equal dimension. Then T is invertible if and only if rank(T)=dim(V).

Example 1

Let T:P1(R)R2 be the linear transformation defined by T(a+bx)=(a,a+b). The reader can verify directly that T1:R2P1(R) is defined by T1(c,d)=c+(dc)x. Observe that T1 is also linear. As Theorem 2.17 demonstrates, this is true in general.

Theorem

Theorem 2.17

Let V and W be vector spaces, and let T:VW be linear and invertible. Then T1:WV is linear.

Definition: invertible

Let A be an n×n matrix. Then A is invertible if there exists an n×n matrix B such that AB=BA=I.

If A is invertible, then the matrix B such that AB=BA=I is unique. The matrix B is called the inverse of A and is denoted by A1.

Example 2

The reader should verify that the inverse of

(3757) is (3321)

dimension

Lemma

Let T be an invertible linear transformation from V to W. Then V is finite-dimensional if and only if W is finite-dimensional. In this case, dim(V)=dim(W).

Theorem 2.18

Let V and W be finite-dimensional vector spaces with ordered bases β and γ, respectively. Let T:VW be linear. Then T is invertible if and only if [T] is invertible. Furthermore, [T1]γβ=([T]βγ)1.

Example 2.4.3

Let β and γ be the standard ordered bases of P1(R) and R2, respectively.

T:P1(R)R2T(a+bx)=(a,a+b)T1:R2P1(R)T1(c,d)=c+(dc)x

For T as in Example 1, we have

[T]=(1101),[T1]=(1101)

It can be verified by matrix multiplication that each matrix is the inverse of the other.

Corollary 1

Let V be a finite-dimensional vector space with an ordered basis β, and let T:VV be linear. Then T is invertible if and only if [T]β is invertible. Furthermore, [T1]β=([T]β)1.

Corollary 2

Let A be an n×n matrix. Then A is invertible if and only if LA is invertible. Furthermore, (LA)1=LA1.

isomorphism

Definitions: isomorphism

Let V and W be vector spaces. We say that V is isomorphic to W if there exists a linear transformation T:VW that is invertible. Such a linear transformation is called an isomorphism from V onto W.

Example 2.4.4

Define T:F2P1(F) by T(a1,a2)=a1+a2x. It is easily checked that T is an isomorphism; so F2 is isomorphic to P1(F).

Example 2.4.5

Define

T:P3(R)M2×2(R) by T(f)=(f(1)f(2)f(3)f(4))

It is easily verified that T is linear. By the Lagrange interpolation formula (Section 1.6), T(f)=O only when f is the zero polynomial. Thus, T is one-to-one.

Moreover, since dim(P3(R))=dim(M2×2(R)), it follows that T is invertible. We conclude that P3(R) is isomorphic to M2×2(R).

*orphism

a homomorphism preserves the structure, and some types of homomorphisms are:

  • Epimorphism: a homomorphism that is surjectiv (AKA onto)
  • Monomorphism: a homomorphism that is injective (AKA one-to-one, 1-1, or univalent)
  • Isomorphism: a homomorphism that is bijective (AKA 1-1 and onto); isomorphic objects are equivalent, but perhaps defined in different ways
  • Endomorphism: a homomorphism from an object to itself
  • Automorphism: a bijective endomorphism (an isomorphism from an object onto itself, essentially just a re-labeling of elements)

Theorem about isomorphism

Theorem 2.19

Let V and W be finite-dimensional vector spaces (over the same field). Then V is isomorphic to W iff dim(V)=dim(W).

Corollary

Let V be vector space over F. Then V is isomorphic to Fn iff dim(V)=n.

Theorem 2.20

Let V and W be finite-dimensional vector spaces of dimensions n and m, respectively, with ordered bases β and γ. Then the function Φ:L(V,W)Mm×n(F), defined by Φ(T)=[T]βγ for TL(V,W), is an isomorphism.

Corollary

L(V,W) is finite-dimensional with dimension mn.

Definition: standard representation

Let β be an ordered basis for an n-dimensional vector space V over field F. The standard representation of V w.r.t. β is the function ϕβ:VFn defined by ϕβ(x)=[x]β for each xV.

Example 2.4.6

Let β={(1,0),(0,1)} and γ={(1,2),(3,4)} . It's easily observed that both are ordered bases for R2. For x=(1,2), we have

ϕβ(x)=(12),ϕγ(x)=(52)i.e.5(1,2)+2(3,4)=(1,2)

We observed arlier that ϕβ is a linear transformation. the next theorem tells us much more.

Theorem 2.21

For any finite-dimensional vector space V with ordered basis β,ϕβ is an isomorphism.

Relationship

Let V and W be vector spaces of dimension n and m, respectively, and let T:VW be a linear transformation. Define A=[T]βγ, where β and γ are arbitrary ordered bases of V and W, respectively. We are now able to use ϕβ and ϕγ to study the relationship between the linear transformations T and LA:FnFm.

Consider the following two composites of linear transformations that map V into Fm:

  1. Map V into Fn with ϕβ and follow this transformation with LA; this yields the composite LAϕβ.

  2. Map V into W with T and follow it by ϕγ to obtain the composite ϕγT.

diagram

These 2 composites are depictedby the dashed arrows in the diagram. By a simple reformulation of Theorem 2.14, we may conclude that

LAϕβ=ϕγT

([T(u)]γ=[T][u]β)

That is, the diagram "commutes." Heuristically, this relationship indicates that after V and W are identified with Fn and Fm via ϕβ and ϕγ, respectively, we may "identify" T with LA. This diagram allows us to transfer operations on abstract vector spaces to ones on Fn and Fm.

Example 2.4.7

Recall the linear transformation T:P3(R)P2(R) defined by T(f(x))=f(x). Let β and γ be the standard ordered bases for P3(R) and P2(R), respectively, and let ϕβ:P3(R)R4 and ϕγ:P2(R)R3 be the corresponding standard representations.

If A=[T], then

A=(010000200003)

Consider the polynomial p(x)=2+x3x2+5x3. We show that

LAϕβ(p(x))=(010000200003)(2135)=(1615)

But since T(p(x))=p(x)=16x+15x2, we have

ϕγT(p(x))=(1615)

So

LAϕβ(p(x))=ϕγT(p(x))diagram

2.5 The Change of Coordinate Matrix

Theorem 2.22

Let β and β be two ordered bases for a finite-dimensional vector space V, and let Q=[IV]ββ. Then

  • Q is invertible.
  • For any vV,[v]β=Q[v]β

The matrix Q=[IV]ββ is called a change of coordinate matrix. Because of part (b) of the theorem, we say that Q changes β-coordinates into β-coordinates. Observe that if β=x1,x2,...,xn and β=x1,x2,...,xn, then

xj=i=1nQijxi

for j=1,2,,n; that is, the jth column of Q is [xj]β.

Example 2.5.1

In R2, and let

β={(1,1),(1,1)},β={(2,4),(3,1)}

Since

(2,4)=3(1,1)1(1,1),(3,1)=2(1,1)+1(1,1)

the change of coordinate matrix Q changing β-coordinates into β-coordinates is

Q=(3211)

For instance,

[(2,4)]β=Q[(2,4)]β=Q(10)=(31)

For the remainder of this section, we consider only linear transformations that map a vector space V into itself. Such a linear transformation is called a linear operator on V

Theorem 2.23

Let T be a linear operator on a finite-dimensional vector space V, and let β and β be ordered bases for V. Suppose Q is the change of coordinate matrix that changes β-coordinates into β-coordinates. Then

[T]β=Q1[T]βQ

Proof:

Q[T]β=[I]ββ=[T]ββ=[IT]ββ=[TI]ββ=[T]ββ[I]ββ=[I]βQ.

Example 2.5.2: Calculation about coordination change

Let T be the linear operator on R2 defined by

T(ab)=(3aba+3b)

Let

β={(1,1),(1,1)},β={(2,4),(3,1)}

be the ordered bases. It can be known that

[T]β=(3113)

(Here gives details about the result above)

Given that [T]β=Q1[T]γQ, where γ is standard ordered basis, thus

[T]γ=(3113)

And

Q=(1111),Q1=12(1111).

So

β=12(1111)(3113)(1111)=(3113)

The change of coordinate matrix that changes β-coordinate into β-coordinate is

Q=(3211)

and

Q1=15(1213)

hence, by theorem 2.23:

[T]β=Q1[T]βQ=(4122)

To show that this's the correct matrix, we can verify that the image under T of each vector of β is the linear combination of the vectors of β with the entries of thecorresponding column as its coefficients. For example, the image of the 2nd vector in β is

T(31)=(86)=1(24)+2(31)

Note that the coeffficients of the linear combination are the entries of the 2nd column of [T]β

Example 2.5.3: Application about coordination change

Recall the reflection about the x-axis in Example 3 in Section 2.1. The rule (x,y)(x,y) is easy to obtain. Let T be the reflection about the line y=2x. We wish to find an expression for T(a,b) for any (a,b)R2. Since T is linear, it is determined by its values on a basis for R2. For basis vectors,

T(1,2)=(1,2),T(2,1)=(2,1)

Therefore if we let

β={(12),(21)}

Then the matrix representing T in the basis β={(1,2),(2,1)} is

[T]β=(1001)

Let β be the standard ordered basis for R2, and let Q be the matrix that changes β-coordinates into β-coordinates. Then

Q=(1221)

and Q1[T]βQ=[T]β. We can solve this equation for [T]β to obtain that [T]β=Q[T]βQ1. Because

Q1=15(1221)

Then it can be verified that

[T]β=15(3443)

Since β is the standard ordered basis, it follows that T is left-multiplication by [T]β. Thus for any (a,b)R2, we have

T(ab)=15(3443)(ab)=15(3a+4b4a+3b)

Corollary

Let AMn×n(F), and let γ be an ordered basis for Fn. Then

[LA]γ=Q1AQ

where Q is the ntimesn matrix whose jth column is the jth vector of γ.

Example 2.5.4

Let

A=(210113010)

and let

γ={(100),(210),(111)}

which is an ordered basis for R3. Let Q be the 3×3 matrix whose jth column is the jth vector of γ. Then

Q=(121011001) and Q1=(121011001)

So by the preceding corollary,

[LA]γ=Q1AQ=(028146011)

Definition

Let A,B be matrices in Mn×n(F). We say B is similar to A if there exists an invertible matrix Q such that

B=Q1AQ

Released under the MIT License.