Skip to main content

Section 6.1 Coordinate vectors and isomorphisms

Subsection 6.1.1 Coordinate vectors

Before we can define coordinate vectors we need to define an ordered basis. As the name suggests this is nothing more than a basis along with a particular choice of ordering of its elements: i.e. first element, second element, etc.. In other words, an ordered basis will be a sequence of vectors, as opposed to a set of vectors.

Definition 6.1.1. Ordered bases.

Let \(V\) be a finite-dimensional vector space. An ordered basis of \(V\) is a sequence of vectors \(B=(\boldv_1, \boldv_2, \dots, \boldv_n)\) whose underlying set \(\{\boldv_1, \boldv_2, \dots, \boldv_n\}\) is a basis of \(V\text{.}\)

Similarly, an ordered basis is orthogonal (resp. orthonormal) if its underlying set is orthogonal (resp. orthonormal).

Remark 6.1.2.

A single (unordered) basis \(B=\{\boldv_1, \boldv_2, \dots, \boldv_n\}\) of a vector space gives rise to \(n!\) different ordered bases: you have \(n\) choices for the first element of the ordered basis, \((n-1)\) choices for the second element, etc..

For example the standard basis \(B=\{\bolde_1, \bolde_2, \bolde_3\}\) of \(\R^3\) gives rise to \(3!=6\) different ordered bases of \(\R^3\text{:}\)

\begin{align*} B_1\amp =(\bolde_1, \bolde_2, \bolde_3) \amp B_2\amp =(\bolde_1, \bolde_3, \bolde_2) \\ B_3\amp=(\bolde_2, \bolde_1, \bolde_3) \amp B_4\amp =(\bolde_2, \bolde_3, \bolde_1) \\ B_5 \amp =(\bolde_3, \bolde_1, \bolde_2) \amp B_6\amp =(\bolde_3, \bolde_2, \bolde_1)\text{.} \end{align*}

By a slight abuse of language we will use ‘standard basis’ to describe both one of our standard unordered bases and the corresponding ordered basis obtained by choosing the implicit ordering of the set descriptions in Remark 4.5.2. For example, \(\{x^2, x, 1\}\) and \((x^2, x, 1)\) will both be called the standard basis of \(P_2\text{.}\)

Definition 6.1.3. Coordinate vectors.

Let \(B=(\boldv_1, \boldv_2,\dots , \boldv_n)\) be an ordered basis for the vector space \(V\text{.}\) According to Theorem 4.5.7, for any \(\boldv\in V\) there is a unique choice of scalars \(c_i\in \R\) satisfying

\begin{equation*} \boldv=c_1\boldv_1+c_2\boldv_2\cdots +c_n\boldv_n\text{.} \end{equation*}

We call the corresponding \(n\)-tuple \((c_1,c_2,\dots, c_n)\) the coordinate vector of \(\boldv\) relative to the basis \(B\), and denote it \([\boldv]_B\text{:}\) i.e.,

\begin{equation*} [\boldv]_B=(c_1,c_2,\dots, c_n)\text{.} \end{equation*}

To compute the coordinate vector of an element \(\boldv\in V\) relative to a given ordered basis \(B=(\boldv_1,\boldv_2,\dots, \boldv_n)\) we must find the scalars \(c_1, c_2, \dots, c_n\) that satisfy the vector equation

\begin{equation*} \boldv=c_1\boldv_1+c_2\boldv_2+\cdots +c_n\boldv_n\text{.} \end{equation*}

Usually this is accomplished by reducing this vector equation to an equivalent system of linear equations in the unknowns \(c_i\) and solving using the method of Gaussian elimination. However, there are some cases when we can easily produce the \(c_i\) by inspection: for example when computing with standard bases as the first example below illustrates.

Furthermore, if the given basis happens to be orthogonal (or orthonormal) with respect to some inner product, Theorem 6.1.7 provides an inner product formula for computing the \(c_i\text{.}\)

Computing coordinate vectors relative to one of our standard bases for \(\R^n\text{,}\) \(M_{mn}\text{,}\) or \(P_{n}\) amounts to just listing the coefficients or entries used to specify the given vector. The examples below serve to illustrate the general method in this setting.

  1. Let \(V=\R^3\) and \(B=(\bolde_1, \bolde_2, \bolde_3)\text{.}\) For any \(\boldv=(a,b,c)\in \R^3\) we have \([\boldv]_{B}=(a,b,c)\text{,}\) since \((a,b,c)=a\bolde_1+b\bolde_2+c\bolde_3\text{.}\)

  2. Let \(V=M_{22}\) and \(B=(E_{11}, E_{12}, E_{21}, E_{22})\text{.}\) For any \(A=\begin{amatrix}[rr]a\amp b\\ c\amp d \end{amatrix}\) we have \([A]_B=(a,b,c,d)\) since

    \begin{equation*} A=aE_{11}+bE_{12}+cE_{21}+dE_{22}\text{.} \end{equation*}

If we choose an alternate ordering of one of the standard bases, the entries of the coordinate vector are reordered accordingly, as illustrated by the examples below.

  1. Let \(V=\R^3\) and \(B=(\bolde_2, \bolde_1, \bolde_3)\text{.}\) Given \(\boldv=(a,b,c)\in \R^3\) we have \([\boldv]_B=(b,a,c)\text{,}\) since

    \begin{equation*} \boldv=b\bolde_2+a\bolde_1+c\bolde_3\text{.} \end{equation*}
  2. Let \(P=P_3\) and \(B=(1,x,x^2, x^3)\text{.}\) Given \(p(x)=ax^3+bx^2+cx+d\) we have \([p(x)]_B=(d, c, b, a)\text{,}\) since

    \begin{equation*} p(x)=d(1)+cx+bx^2+ax^3\text{.} \end{equation*}

For a nonstandard ordered basis, we compute coordinate vectors by solving a relevant system of linear equations, as the examples below illustrate.

  1. Let \(V=\R^2\text{,}\) \(B=((1,2),(1,1))\text{,}\) and \(\boldv=(3,3)\text{.}\) To compute \([\boldv]_B\) we must find the unique pair \((c_1, c_2)\) satisfying

    \begin{equation*} (3,3)=c_1(1,2)+c_2(1,1)\text{.} \end{equation*}

    By inspection, we see that

    \begin{equation*} (3,3)=3(1,1)=0(1,2)+3(1,1)\text{.} \end{equation*}

    We conclude that

    \begin{equation*} [\boldv]_{B}=(0,3)\text{.} \end{equation*}

    More generally, to compute \([\boldv]_B\) for an arbitrary \(\boldv=(a,b)\in \R^2\text{,}\) we must find the pair \((c_1,c_2)\) satisfting \((a,b)=c_1(1,2)+c_2(1,1)\text{,}\) or equivalelenty

    \begin{equation*} \begin{linsys}{2} c_1\amp +\amp c_2 \amp =\amp a\\ 2c_1\amp +\amp c_2\amp =\amp b \end{linsys}\text{.} \end{equation*}

    The usual techniques yield the unique solution \((c_1,c_2)=(-a+b,2a-b)\text{,}\) and thus

    \begin{equation*} [\boldv]_B=(-a+b, 2a-b) \end{equation*}

    for \(\boldv=(a,b)\text{.}\)

  2. Let \(V=P_2\text{,}\) \(B=(x^2+x+1, x^2-x, x^2-1)\text{,}\) and \(p(x)=x^2\text{.}\) To compute \([p(x)]_B\) we must find the unique triple \((c_1,c_2,c_3)\) satisfying

    \begin{equation*} x^2=c_1(x^2+x+1)+c_2(x^2-x)+c_3(x^2-1)\text{.} \end{equation*}

    The equivalent linear system once we combine like terms and equate coefficients is

    \begin{equation*} \begin{linsys}{3} c_1\amp +\amp c_2\amp +\amp c_3\amp =\amp 1\\ c_1\amp -\amp c_2\amp \amp \amp =\amp 0\\ c_1\amp \amp \amp -\amp c_3\amp =\amp 0\\ \end{linsys}\text{.} \end{equation*}

    The unique solution to this system is \((c_1,c_2,c_3)=(1/3, 1/3, 1/3)\text{.}\) We conclude

    \begin{equation*} [x^2]_B=\frac{1}{3}(1, 1, 1)\text{.} \end{equation*}

    The same reasoning shows that more generally, given polynomial \(p(x)=ax^2+bx+c\text{,}\) we have

    \begin{equation*} [p(x)]_B=\frac{1}{3}(a+b+c, a-2b+c, a+b-2)\text{.} \end{equation*}

When \(B=(\boldv_1, \boldv_2, \dots, \boldv_n)\) is an orthogonal basis with respect to some inner product \(\langle , \rangle \text{,}\) then by Theorem 5.2.6 we have

\begin{equation*} \boldv=\frac{\langle \boldv,\boldv_1 \rangle }{\langle \boldv_1, \boldv_1\rangle }\boldv_1+\frac{\langle \boldv,\boldv_1 \rangle }{\langle \boldv_1, \boldv_1\rangle }\boldv_1+\cdots +\frac{\langle \boldv,\boldv_n \rangle }{\langle \boldv_n, \boldv_n\rangle }\boldv_n\text{,} \end{equation*}

and thus

\begin{equation*} [\boldv]_B=\left(\frac{\langle \boldv,\boldv_1 \rangle }{\langle \boldv_1, \boldv_1\rangle }, \frac{\langle \boldv,\boldv_2 \rangle }{\langle \boldv_2, \boldv_2\rangle },\dots, \frac{\langle \boldv,\boldv_n \rangle }{\langle \boldv_n, \boldv_n\rangle }\right)\text{.} \end{equation*}

We have thus proved the following theorem.

Let \(V=\R^2\) and \(B=((1,1),(-1,2))\text{.}\) Find a general formula for \([(a,b)]_B\text{.}\) Note: \(B\) is orthogonal with respect to the weighted dot product

\begin{equation*} \langle (x_1,x_2), (y_1,y_2)\rangle =2x_1y_1+x_2y_2\text{.} \end{equation*}
Solution.

Applying Theorem 6.1.7 to \(B\) and the dot product with weights \(2, 1\text{,}\) for any \(\boldv=(a,b)\) we compute

\begin{align*} [(a,b)]_B \amp =\left(\frac{\langle (a,b), (1,1)\rangle }{\langle (1,1),(1,1) \rangle }, \frac{\langle (a,b), (-1,2)\rangle }{\langle (-1,2),(-1,2) \rangle }\right)\\ \amp=\left(\frac{1}{3}(2a+b),\frac{1}{3}(-a+b) \right) \text{.} \end{align*}

Let's check our formula with \(\boldv=(3,-3)\text{.}\) The formula yields \([(3,-3)]_B=(1,-2)\text{,}\) and indeed we see that

\begin{equation*} (3,-3)=1(1,1)-2(-1,2)\text{.} \end{equation*}

The next theorem is the key to understanding the tremendous computational value of coordinate vectors. The map

\begin{equation*} \boldv\in V \longmapsto [\boldv]_B\in \R^n \end{equation*}

is a one-to-one correspondence, allowing us identify the vector space \(V\) with \(\R^n\text{.}\) Furthermore, since this map is linear, any question regarding the vector space structure of \(V\) can be translated to an equivalent question about the vector space \(\R^n\text{.}\) The great virtue of this is that we have a wealth of computational procedures that allow us to answer such questions for \(\R^n\text{.}\) As a result, given any “exotic” vector space \(V\) of finite dimension, once we choose an ordered basis \(B\) of \(V\text{,}\) questions about \(V\) can be answered by taking coordinate vectors with respect to \(B\) and answering the corresponding question in the familiar setting of \(\R^n\text{.}\)

  1. Suppose \(T(\boldv)=[\boldv]_B=(a_1,a_2,\dots, a_n), T(\boldw)=[\boldw]_B=(b_1, b_2, \dots, b_n)\text{.}\) By definition this means

    \begin{align*} \boldv \amp =\sum_{i=1}^na_i\boldv_i \amp \boldw\amp =\sum_{i=1}^nb_i\boldv_i\text{.} \end{align*}

    Then

    \begin{equation*} c\boldv+d\boldw=\sum_{i=1}^n(ca_i+db_i)\boldv_i\text{,} \end{equation*}

    and hence

    \begin{align*} T(c\boldv+d\boldw)\amp =[c\boldv+d\boldw]_B \\ \amp =(ca_1+db_1,ca_2+db_2,\dots, ca_n+db_n) \\ \amp =c(a_1,a_2,\dots, a_n)+d(b_1,b_2,\dots, b_n)\\ \amp =c[\boldv]_B+d[\boldw]_B\\ \amp =cT(\boldv)+dT(\boldw) \text{.} \end{align*}

    This proves \(T\) is linear.

  2. Suppose \(T(\boldv)=T(\boldw)=(c_1,c_2,\dots, c_n)\text{,}\) then by definition of \([\hspace{.5pt}]_B\) we have

    \begin{equation*} \boldv=\boldw=c_1\boldv_1+c_2\boldv_2+\cdots +c_n\boldv_n\text{.} \end{equation*}
  3. Given any \(\boldb=(b_1,b_2,\dots, b_n)\in \R^n\text{,}\) we have \(\boldb=T(\boldv)\text{,}\) where

    \begin{equation*} \boldv=b_1\boldv_1+b_2\boldv_2+\cdots +b_n\boldv_n\text{.} \end{equation*}

    This proves \(\im T=\R^n\text{.}\)

  4. This follows from (1)-(3) and Theorem 6.1.12.

As an illustration of the comments before Theorem 6.1.9, we describe a general method of contracting and extending sets to bases.

The set

\begin{equation*} S=\left \{ A_1=\begin{bmatrix}2\amp 1\\ 0\amp -2 \end{bmatrix} , A_2=\begin{bmatrix}1\amp 1\\ 1\amp -1 \end{bmatrix} , A_3=\begin{bmatrix}0\amp 1\\ 2\amp 0 \end{bmatrix} , A_4=\begin{bmatrix}-1\amp 0\\ 1\amp 1 \end{bmatrix} \right\} \end{equation*}

is a subset of the space \(W=\{ A\in M_{22}\colon \tr A=0\}\text{.}\) Let \(W'=\Span S\text{.}\) Contract \(S\) to a basis of \(W'\) and determine whether \(W'=W\text{.}\)

Hint.

Choose an ordered basis \(C\) of \(M_{22}\) and use the coordinate vector map to translate to a question about subspaces of \(\R^4\text{.}\) Answer this question and translate back to \(M_{22}\text{.}\)

Solution.

Let \(C=(E_{11}, E_{12}, E_{21}, E_{22})\) be the standard basis of \(M_{22}\text{.}\) Apply \([\hspace{10pt}]_C\) to the elements of the given \(S\) to get a corresponding set \(S'\subseteq\R^4\text{:}\)

\begin{equation*} S'=\left\{ [A_1]_C=(2,1,0,-2), [A_2]_C=(1,1,1,-1), [A_3]_C=(0,1,2,0), [A_4]_C=(-1,0,1,1) \right\}\text{.} \end{equation*}

Apply the column space procedure of Procedure 4.6.9 to contract \(S'\) to a basis \(B'\) of \(\Span S'\text{.}\) This produces the subset

\begin{equation*} B'=\{[A_1]_C=(2,1,0,-2), [A_2]_C=(1,1,1,-1)\} \end{equation*}

Translating back to \(V=M_{22}\text{,}\) we conclude that the corresponding set

\begin{equation*} B=\{A_1, A_2\} \end{equation*}

is a basis for \(W'=\Span S\text{.}\) We conclude that \(\dim W'=2 \text{.}\)

Lastly the space \(W\) of all trace-zero matrices is easily seen to have basis

\begin{equation*} \left\{ \begin{amatrix}[rr]1\amp 0\\ 0 \amp -1 \end{amatrix}, \begin{amatrix}[rr]0\amp 1\\ 0\amp 0 \end{amatrix}, \begin{amatrix}[rr]0 \amp 0\\ 1\amp 0 \end{amatrix} \right\}\text{,} \end{equation*}

and hence \(\dim W=3\text{.}\) Since \(\dim W'\lt\dim W\text{,}\) we conclude that \(W'\ne W\text{.}\)

Subsection 6.1.2 Isomorphisms

We spoke of the coordinate vector map as a means of “translating” questions about an abstract vector space \(V\) to equivalent ones about the more familiar vector space \(\R^n\text{.}\) Properties (1)-(3) of Theorem 6.1.9 are what guarantee that nothing is lost in this translation. Axiomitizing these properties, we obtain an important family of linear transformations called isomorphisms. Before getting to this definition, we recall some basic properties of general set functions.

The following omnibus result is useful for deciding whether a linear transformation is an isomorphism, and lists a few of the properties of a vector space that are preserved by isomorphisms: namely, dimension, span, and linear independence.

  1. Assume \(T\) is injective. Then

    \begin{align*} T(\boldv)=\boldzero_W\amp \implies T(\boldv)=T(\boldzero_V) \\ \amp\implies \boldv=\boldzero_V \text{.} \end{align*}

    It follows that \(\NS T=\{\boldzero_V\}\text{.}\)

    Now assume \(\NS T=\{\boldzero_V\}\text{.}\) Then

    \begin{align*} T(\boldv)=T(\boldv') \amp\implies T(\boldv)-T(\boldv')=\boldzero_W \\ \amp\implies T(\boldv-\boldv')=\boldzero_W \\ \amp \implies \boldv-\boldv'\in \NS T=\{\boldzero_V\}\\ \amp\implies \boldv-\boldv'=\boldzero_V \\ \amp \implies \boldv=\boldv'\text{.} \end{align*}

    Thus \(T\) is injective.