es

Dual Spaces

Dual basis

Double Dual

Annihilators

Recall that V is an 𝔽-vector space over field 𝔽 (which is either ℚ or ℝ or ℂ). The set of all linear functionals is denoted by V or V′ (the Riesz representation theorem establishes isomorphism between these two spaces, see Dual transformations in Part 5).
If S is a subset (not need to be a subspace) of a vector space V, then the annihilator S0 is the set of all linear functionals φ such that φ(v) = <φ|v> = 0 for all vS. When V is a Euclidean space, then its annihilator is denoted by S.
The term annihilator is quite descriptive, since S0 consists of all linear functionals that annihilate (send to 0) every vector in S. For a subset S of V, the annihilator S0 of S is defined as the set of all vV such that φ(v) = 0 for all φ ∈ S.

From this definition, we immediately get that {0}0 = V* and V0 = {0}. If V is finite-dimensional and contains a non-zero vector, then Theorem 5 assures us that S0V*.

Lemma 2: Let V be a vector space over a field 𝔽. Then
  1. For any subset S of V, S0 = (span{S})0.
  2. For any subsets S₁ and S₂ of V, if S₁ ⊆ S₂, then S0S0.
  3. For any subset S of V, S0 is a subspace of V and S ⊆ (S0)0.
  1. Since S ⊆ span(S), we find that (span(S))0S0.

    Conversely, suppose that φ ∈ S0, i.e., φ(s) = 0 for all sS. For any linear combination c1v1 + c2v2 + ⋯ + cnvn from span(S), where c1, c2, … , cn ∈ 𝔽 and v1, v2, … , vnS, we have \begin{align*} \varphi \left( c_1 {\bf v}_1 + c_2 {\bf v}_2 + \cdots + c_n {\bf v}_n \right) &= c_1 \varphi \left( {\bf v}_1 \right) + c_2 \varphi \left( {\bf v}_2 \right) + \cdots + c_n \varphi \left( {\bf v}_n \right) \\ &= 0 \end{align*} and hence φ ∈ (span(S))0.

  2. Suppose S₁ ⊆ S₂ and φ ∈ S₂. Then for any vS₁, φ(v) = 0 and consequently, φ ∈ S₁.
  3. Since for any vS, 0(v) = 0, we find that 0 ∈ S0, and therefore S0 ≠ ∅. Let φ, ψ ∈ S0 and c, k ∈ 𝔽. Then for every vS, \[ \left( c\varphi + k\psi \right) {\bf v} = c\varphi ({\bf v}) + k \psi ({\bf v}) = c0 + k0 \\ = 0 , \] which shows that S0 is subspace of V*. Now let vS. Then for every linear functional φ ∈ S0, v*(φ) = φ(v) = 0. So v* ∈ (S0)0; since V can be naturally identifies with V**, v ∈ (S0)0.
   
Example 13: Let S be a set of two vectors in ℝ³ \[ {\bf v} = \left( 3, 2, 1 \right) , \qquad {\bf u} = \left( 1, -2, 1\right) . \] Arbitrary functional has the form: \[ \varphi = b_1 {\bf e}^1 + b_2 {\bf e}^2 + b_3 {\bf e}^3 , \] where b₁, b₂, b₃ are some real numbers and {e¹, e², e³} is a dual basis. Applying functional φ to vectors v and u, we obtain \begin{align*} \varphi ({\bf v}) &= \left( b_1 {\bf e}^1 + b_2 {\bf e}^2 + b_3 {\bf e}^3 \right) \left( 3\,{\bf e}_1 + 2 {\bf e}_2 + {\bf e}_3 \right) \\ &= 3\,b_1 + 2\,b_2 + b_3 = 0 \end{align*} and \begin{align*} \varphi ({\bf u}) &= \left( b_1 {\bf e}^1 + b_2 {\bf e}^2 + b_3 {\bf e}^3 \right) \left( {\bf e}_1 - 2{\bf e}_2 + {\bf e}_3 \right) \\ &= b_1 - 2\,b_2 + b_3 = 0 , \end{align*} because \[ {\bf e}^1 \left( {\bf e}_1 \right) = 1 , \quad {\bf e}^2 \left( {\bf e}_2 \right) = 1 ,\quad {\bf e}^3 \left( {\bf e}_3 \right) = 1 , \] and for other applications we have zeroes: ei(ej) = 0 when ij. Therefore, φ annihilates vectors v and u if and only if the coefficients bk satisfy the system of equations: \[ \begin{split} 3\,b_1 + 2\,b_2 + b_3 &= 0 , \\ b_1 - 2\,b_2 + b_3 &= 0 \end{split} \tag{12.1} \] Subtraction one equation from another gives \[ 2\,b_1 + 4\, b_2 \qquad \iff \qquad b_1 = -2\,b_2 . \] Correspondingly, b₃ = 4b₂. We verify with Mathematica:
Solve[{3*b1 + 2*b2 + b3 == 0, b1 - 2*b2 + b3 == 0}, {b1, b3}]
{b1 -> -2 b2, b3 -> 4 b2}}
Therefore, b₂ is a free variable and S0 is spanned on the vector \[ S^0 = \mbox{span}(-2, 1, 4) , \] which is a one-dimensional space.    ■
End of Example 13
Theorem 8: If S is an m-dimensional subspace of an n-dimensional vector space V, then S0 is an (n−m)-dimensional subspace of V*.
If φ and ψ annihilate the set S, so
\[ \varphi ({\bf v}) = \psi ({\bf v}) = 0, \qquad \forall {\bf v} \in V, \]
then their sum is also an annihilator. Also, if we multiply φ by a constant c, then (cφ)(v) = c·φ(v) = c·0 = 0. So the set of all annihilators for arbitrary set S is a vector space.

Let β = { x1, x2, … , xn } be a basis in V whose first m elements are in S (so they constitute a basis in S as being linearly independent). Let β* = { y1, y2, … , yn } be the dual basis in V*, We denote by RV* the subspace spanned on { ym+1, ym+2, … , yn }. Clearly, R has dimension n−m. We are going to show that R = S0.

If x is any vector in S, then x is a linear combination of the first m elements from the basis β:

\[ {\bf x} = \sum_{i=1}^m a_i {\bf x}_i , \]
and for any j = m+1, m+2, … , n, we have
\[ \langle {\bf y}^j \vert\,{\bf x} \rangle = \sum_{i=1}^m a_i \langle {\bf y}^j \vert\,{\bf x}_i \rangle = 0 . \]
In other words, each yj, j = m+1, m+2, … , n, is in S0. It follows that R is in V*, \[ R \subset S^0 . \] Now we prove the opposite relation and assume that y is any element of S0. Then it is a linear combination of vectors from the dual basis β*: \[ {\bf y} = \sum_{j=1}^n b_j {\bf y}^j . \] Since by assumption, y is in S0, we have, for every i = 1, 2, … , m, \[ 0 = \langle {\bf y} \,|\,{\bf x}_i \rangle = \sum_{j=1}^n b_j \langle {\bf y}^j \,|\,{\bf x}_i \rangle = b_i . \] In other words, y is a linear combination of ym+1 , … , yn. This proves that y is in R and consequently that \[ S^0 \subset R, \] and the theorem follows.
   
Example 14: Let V = ℝ2,2 be the vector space of all 2 × 2 matrices with real entries and let W be the subspace of V consisting of those matrices AV for which A B = B A, where \( \displaystyle{\bf B} = \begin{bmatrix} \phantom{-}1 & -2 \\ -2 & \phantom{-}4 \end{bmatrix} . \)

Let \( \displaystyle{\bf A} = \begin{bmatrix} a & b \\ c & d \end{bmatrix} , \) where 𝑎, b, c, and d are some real numbers to be determined. To find their values, we calculate A BB A: \[ {\bf A}\,{\bf B} - {\bf B}\,{\bf A} = \begin{bmatrix} 2c - 2b & 3b + 2d - 2a \\ 2a - 3c - 2d & 2b - 2c \end{bmatrix} . \]

B = {{1, -2}, {-2, 4}};
A = {{a, b}, {c, d}};
A.B - B.A
{{-2 b + 2 c, -2 a + 3 b + 2 d}, {2 a - 3 c - 2 d, 2 b - 2 c}}
Upon equating A BB A to zero, we obtain four equations \[ \begin{split} 2c - 2b &= 0, \\ 3b + 2d - 2a &= 0, \\ 2a - 3c - 2d &= 0, \\ 2b -2c &= 0, \end{split} \] two of which are the same. Choosing b as a free variable, we find \[ a = 0 , \qquad c = b, \qquad d = \frac{3}{2}\, b . \]
Solve[{-2 a + 3 b + 2 d == 0, a - 3 b - 2 d == 0}, {a, d}]
{{a -> 0, d -> -((3 b)/2)}}
Therefore, matrices from W depend on one real parameter: \[ {\bf A} = b \begin{bmatrix} 0 & 2 \\ 2 & 3 \end{bmatrix} , \qquad b \in \mathbb{R} . \tag{14.1} \] So dimW = 1 whereas dim V = 4.

In order to determine W0, we use the dual basis {E1, E2, E3, E4} that corresponds to the standard basis of ℝ2,2: \[ {\bf E}_1 = \begin{bmatrix} 1 & 0 \\ 0 & 0 \end{bmatrix} , \quad {\bf E}_2 = \begin{bmatrix} 0 & 1 \\ 0 & 0 \end{bmatrix} , \quad {\bf E}_3 = \begin{bmatrix} 0 & 0 \\ 1 & 0 \end{bmatrix} , \quad {\bf E}_1 = \begin{bmatrix} 0 & 0 \\ 0 & 1 \end{bmatrix} . \] Arbitrary element from W0 can be expanded with respect to the dual basis \[ W^{0} \ni \varphi = c_1 {\bf E}^1 + c_2 {\bf E}^2 + c_3 {\bf E}^3 + c_4 {\bf E}^4 , \tag{14.2} \] where coefficients are determined from condition φ(A) = 0. Expansion of A with respect to standard basis is \[ {\bf A} = 2{\bf E}_2 + 2 {\bf E}_3 + 3 {\bf E}_4 . \] Then \[ W^{0} \ni \varphi ({\bf A}) = \left( c_1 {\bf E}^1 + c_2 {\bf E}^2 + c_3 {\bf E}^3 + c_4 {\bf E}^4 \right) \left( 2{\bf E}_2 + 2 {\bf E}_3 + 3 {\bf E}_4 \right) = {\bf 0} . \] Since action of elements from dual basis on standard basis elements are known, we get \[ W^{0} \ni \varphi ({\bf A}) = 2 c_2 + 2 c_3 + 3 c_4 = 0. \tag{14.3} \] Since there is one condition Eq.(14.3) between four coefficients in Eq.(14.2), we conclude that dimW0 = 3.    ■

End of Example 14
Theorem 9: If S is a subspace of a finite-dimensional vector space V, then S00 = S. In general, span{S} ≌ S⁰⁰.
By definition of S0, ⟨φ | x⟩ = 0 for all xS and all φ ∈ S0, it allows that SS00, The desired conclusion follows from the dimension argument, Let S be m-dimensional, then dimension of S0 is n−m, and that of S00 is n − (n − m) = m. Hence S = S00.
   
Example 15: Let V = ℝ≤n[x] be the space of polynomials of degree at most n in variable x. Let us consider its subset (which is actually, subspace) S that consists of all polynomials without a free term, so S = x≤n-1[x]. Then its annihilator becomes \[ S^0 = \left\{ \varphi \in V^{\ast} \, : \ \varphi (p) = p(0), \quad p \in V \right\} . \]

The second annihilator consists of all polynomials p such that φ(p) = p(0) = 0. A polynomial to be zero at the origin should be without a free term. Therefore, S00 = S.

Upon introducing a standard basis in ℝ≤n[x], \[ {\bf e}_0 =1, \ {\bf e}_1 = x, \ {\bf e}_2 = x^2 , \ \ldots , \ {\bf e}_n = x^n , \] we establish a one-to-one and onto (bijection) transofrmation f : ℝ≤n[x] ≌ ℝn+1 such that \[ f \left( a_0 + a_1 x + a_2 x^2 + \cdots + a_n x^n \right) = \left( a_0 , \ a_1 , \ a_2 , \ \ldots , \ a_n \right) . \] Then S is congruent to a subspace of ℝn+1: \[ S \cong \left\{ \left( 0 , x_1, x_2 , \ldots , x_n \right) \ : \ x_k \in \mathbb{R}, \quad k=1,2,\ldots , n \right\} . \]

Similarly, we can consider the subspace Sn of polynomials of degree at most n−1. Its annihilating functional acts as \[ \psi \left( a_0 , a_1 , a_2 , \ldots , a_{n-1} , a_n \right) = \left( a_0 , a_1 , a_2 , \ldots , a_{n-1} , 0 \right) \] In the space of polynomials ℝ≤n[x], this annihilating functional just drp the last term with xn.    ■

End of Example 15
Theorem 10: Let X and Y be subspaces of a vector space V = XY. Then the dual space X* is isomorphic to Y⁰ and Y* is isomorphic to X⁰; moreover, V* = X⁰ ⊕ Y⁰.
Let xX, φ ∈ X*, and x⁰ ∈ X⁰, respectively. Similarly, let yY, ψ ∈ Y*, and y⁰ ∈ Y⁰. The subspaces X⁰ and Y⁰ are disjoint because if κ(x) = κ(y) = 0 for all x and y, then κ(v) = κ(x + y) = 0 for all v

Moreover, if κ is any functional over V and if v = x + y, we write x⁰(v) = κ(y) and y⁰(v) = κ(x). It is not hard to see that functions x⁰ and y⁰ thus defined are linear functionals on V, so x⁰ and y⁰ belong V*; they also belong to X⁰ and Y⁰, respectively. Since κ = x⁰ + y⁰, it follows that V* is indeed the direct sum of X⁰ and Y⁰.

To establish the asserted isomorphisms, we make a correspondence to every x⁰ a functional ψ from Y* defined by ψ(y) = x⁰(y). It can be shown that the correspondence x⁰ ↦ ψ is linear and one-to-one, and therefore an isomorphism between X⁰ and Y*. The corresponding result for Y⁰ and X* follows from symmetry by interchanging x and y.

   
Example 16: Let us find annihilator W0 of the subspace W of ℝ4 spanned by v = (4, −3, 2, −1) and u = (1, 2, −1, 3).

Given vectors v and u are linearly independent because none is a scalar multiple of another. Therefore, W = span(v, u) is two-dimensional vector space.

We build an annihilator W0 using dual basis {e1, e2, e3, e4} to the standard basis e₁ = (1, 0, 0, 0), e₂ = (0, 1, 0, 0), e₃ = (0, 0, 1, 0), e₄ = (0, 0, 0, 1). Then any vector φ from W0 can be expressed as a linear combination of basis vectors: \[ \varphi = b_1 {\bf e}^1 + b_2 {\bf e}^2 + b_3 {\bf e}^3 + b_4 {\bf e}^4 , \] where scalars b₁, b₂, b₃, b₄ should be chosen so that φ(w) = 0 for any wW. So \[ W^{0} \ni \varphi ({\bf w}) = \left( b_1 {\bf e}^1 + b_2 {\bf e}^2 + b_3 {\bf e}^3 + b_4 {\bf e}^4 \right) \left( c_1 {\bf v} + c_2 {\bf u} \right) = 0 , \] where c₁, c₂ ∈ ℝ. Applying φ to basis vectors v and u, we get the system of equations: \begin{align*} \varphi ({\bf v}) &= \left( b_1 {\bf e}^1 + b_2 {\bf e}^2 + b_3 {\bf e}^3 + b_4 {\bf e}^4 \right) \left( 4, -3, 2, -1 \right) \\ &= 4\,b_1 -3\,b_2 + 2\, b_3 - b_1 = 0, \\ \varphi ({\bf u}) &= \left( b_1 {\bf e}^1 + b_2 {\bf e}^2 + b_3 {\bf e}^3 + b_4 {\bf e}^4 \right) \left( 1, 2, -1, 3 \right) \\ &= b_1 + 2 b_2 - b_3 + 3\,b_4 = 0 . \end{align*} Solving system of linear equations \[ \begin{split} 4\,b_1 -3\,b_2 + 2\, b_3 - b_1 &= 0, \\ b_1 + 2 b_2 - b_3 + 3\,b_4 &= 0 , \end{split} \]

Solve[{4*b1 - 3*b2 + 2*b3 -b4 == 0, b1 + 2*b2 - b3 + 3*b4 == 0}, {b1, b2}]
{{b1 -> 1/11 (-b3 - 7 b4), b2 -> 1/11 (6 b3 - 13 b4)}}
we obtain \[ b_1 = \frac{1}{11} \left( b_3 - 7\, b_4 \right) , \qquad b+2 = \frac{1}{11} \left( 6\,b_3 - 13\,b_4 \right) , \] where b₃, b₄ are free variables. Hence W0 is a two-dimensional vector space.    ■
End of Example 16
Theorem 11: If S and T are subspaces of a vector space V, then \[ \left( S \cap T \right)^0 = S^0 + T^0 \qquad \mbox{and} \qquad \left( S + T \right)^0 = S^0 \cap T^0 . \]
It is clear that φ annihilates S + T if and only if φ annihilates both, S and T. Hence, (S + T)⁰ = S⁰ ∩ T⁰. More precisely, since SS + T and TS + T, by Lemma 2, (S + T)⁰ ⊆ S⁰ + T⁰.

Now, on the other hand, suppose that ϕ ∈ (S⁰ ∩ T⁰). Then ϕ annihilates both S and T. If vS + T, then v = s + t, where sS and tT. Now ϕ(v) = ϕ(s + t) = ϕ(s) + ϕ(t) = 0 + 0 = 0. This shows that ϕ annihilates S + T, i.e., ϕ ∈ (S + T)⁰. Therefore, (S⁰ ∩ T⁰) ⊆ (S + T)⁰ and hence (S⁰ ∩ T⁰) = (S + T)⁰.

Also, if φ = ψ + χ ∈ S⁰ + T⁰, where ψ ∈ S⁰ and χ ∈ T⁰, then ψ, χ ∈ (ST)⁰, and so φ ∈ (ST)⁰. Thus, \[ S^0 + T^0 \subseteq \left( S \cap T \right)^0 . \] For the reverse inclusion, suppose that φ ∈ ( ST)⁰. Write \[ V = S' \oplus \left( S \cap T \right) \oplus T' \oplus U , \] where S = S' ⊕ (ST) and T = T' ⊕ (ST). Define χ ∈ V* by \[ \left. \chi \right\vert_{S'} = \phi , \qquad \left. \chi \right\vert_{S \cap T} = \left. \phi \right\vert_{S \cap T} = 0 , \qquad \left. \chi \right\vert_{T'} = 0 , \qquad \left. \psi \right\vert_{U} = \phi , \] and define ψ ∈ V* by \[ \left. \psi \right\vert_{S'} = 0 , \qquad \left. \psi \right\vert_{S \cap T} = \left. \phi \right\vert_{S \cap T} = 0 , \qquad \left. \psi \right\vert_{T'} = \phi , \qquad \left. \psi \right\vert_{U} = 0 . \] It follows that &χ ∈ T⁰, ψ ∈ S⁰ and χ + ψ = φ.

We can also derive the second partof this statement from the first part by replacing S by S⁰ and T by T⁰, and using the identity S⁰⁰ = S, we get (S⁰ + S⁰)⁰ = (S⁰)⁰ ∩ (T⁰)⁰ = ST.

   
Example 17: Let S be a two-dimensional space spanned on vectors i = (1, 0, 0) and j = (0, 1, 0) while T be a two-dimensional space spanned on vectors j and k = (0, 0, 1). Their intersection ST is a one-dimensional space spanned on vector j. The annihilator of the intersection, (ST)⁰ contains all vectors of the form \[ \left( S \cap T \right)^0 = \left\{ \left( x, 0, z \right) \, : \ x, z \in \mathbb{R} \right\} . \] On the other hand, annihilators of sets S and T are \begin{align*} S^0 &= \left\{ \left( 0, 0, z \right) \ : \ z \in \mathbb{R} \right\} , \\ T^0 &= \left\{ \left( x, 0, 0 \right) \ : \ x \in \mathbb{R} \right\} . \end{align*} Then their sum becomes \[ S^0 + T^0 = \left\{ \left( x, 0, z \right) \ : \ x,z \in \mathbb{R} \right\} . \]    ■
End of Example 17
Observe that no dimension argument is employed in the proof of the first identity (S + T)⁰ = S⁰ ∩ T⁰, hence it holds for vector spaces of finite or infinite dimensions.

 

Hyperplanes/Hyperspaces


When a vector space is a Cartesian product 𝔽n = 𝔽 × 𝔽 × ⋯ × 𝔽 of a finite number of copies of field 𝔽, we can give a "geometrical" interpretation to annihilators. A crucial step is to consider the bilinear functional, known as a dot-product:
\[ {\bf a} \bullet {\bf x} = a_1 x_1 + a_2 x_2 + \cdots + a_n x_n = {\bf x} \bullet {\bf a}, \]
which assigns a number for a pair of two vectors a = (𝑎₁, 𝑎₂, … , 𝑎n) and x = (x₁, x₂, … , xn) from 𝔽n. Although dot products and inner products are the topic of Part 5 (Euclidean Spaces), we will not use its properties over here, but only a convenient notation (as dot, "•"). Dot-product provides an example of a linear form over vector space 𝔽n, which is also denoted with bra-ket notation, called Dirac notation: ⟨a | x⟩ = ax. Actually, every linear functional over 𝔽n is generated by a dot product for some bra-vector a0 (see Theorem 6 in section on dual basis).

With dot-product, we can make the following (geometric) definition.

Two vectors a and x from 𝔽n are called orthogonal if their dot-ptoduct is zero: ax = 0. The set of all vectors that are orthogonal (or perpendicular) to a0 is a vector space, denoted by \[ {\bf a}^{\perp} = \left\{ {\bf x} \ : \ {\bf a} \bullet {\bf x} = 0 \right\} . \]

A linear equation ax = b is a constraint on our choice of a point in 𝔽n. With such constraint, we expect to lose a degree of freedom on the solution set for the linear equation. When choosing a point x = (x₁, x₂, … , xn) in 𝔽n, we have n degrees of freedom---the number of coordinates. By satisfying a single linear equation, we expect to have n − 1 degrees of freedom.

A hyperplane in 𝔽n for n ≥ 2 is the solution set of an n-variable linear equation ax = b (or ⟨a | x⟩ = b) with a0. The homogeneous linear equation 𝑎₁x₁ + 𝑎₂x₂ + ⋯ + 𝑎nxn = 0 defines a hyperspace in 𝔽n, which is (n − 1)-dimensional subspace of 𝔽n.

In case of general finite dimensional vector space V, a maximal proper subspace of V is called a hyperspace of V. A hyperplane is a coset of a hyperspace.

Actually, every hyperspace is the solution set of homogeneous linear equation
\begin{equation} \label{EqHyp.1} a_1 x_2 + a_2 x_2 + \cdots + a_n x_n = 0 , \end{equation}
where the coefficient vector a = (𝑎₁, 𝑎₂, … , 𝑎n) ∈ 𝔽n is given (or known in advance) and x = (x₁, x₂, … , xn) is a vector of unknowns. Eq.\eqref{EqHyp.1} has the general solution, which fills an n − 1vector subspace of 𝔽n. When field 𝔽 is real, the linear combination in the left-hand side of Eq.\eqref{EqHyp.1} is a dot product
\begin{equation} \label{EqHyp.2} {\bf a} \bullet {\bf x} = a_1 x_2 + a_2 x_2 + \cdots + a_n x_n \in \mathbb{F} . \end{equation}
Lemma 3: The solution to one linear inhomogeneous equation over 𝔽 in n unknowns \begin{equation} \label{EqHyp.3} {\bf a} \bullet {\bf x} = a_1 x_2 + a_2 x_2 + \cdots + a_n x_n = b \in \mathbb{F} . \end{equation} is a hyperplane in 𝔽n.
Let a = [𝑎₁, [𝑎₂, … , 𝑎n] be the coefficient matrix for a single linear equation in n unknowns over 𝔽. Notice that a is a nonzero row vector in 𝔽n. Since a is nonzero and the corresponding maps Ta : 𝔽n ⇾ 𝔽, the Rank Theorem (rank(A) + nullity(A) = n) for matrices ensures that \[ mbox{rank}({\bf a}) = \dim \mathbb{F} = 1. \] It follows that the nullity of a is n − 1, thus, that any coset of NullSpace(a) is a hyperplane in 𝔽n.
   
Example 18: Consider the solution to x + 2 y − 3 z = 4 in ℝ³. We can take coordinates y = r and z = s arbitrary in ℝ, which forces x = 4 + 3 z − 2 y = 4 + 3 s − 2 r. The solution set is the affine plane in ℝ³ given by \[ \left\{ (4 + 3\,s -2\,r)\ : \ r,s \in \mathbb{R} \right\} = \left\{ (4,0,0) + r\,(0,-2,0) + s\,(0,0,3) \ : \ r,s \in \mathbb{R}\right\} . \]
Roger
The plane in ℝ³ given by x + 2 y − 3 z = 4.
   ■
End of Example 18
If vector x = (x₁, x₂, … , xn) satisfies the equation ax = 0, it is orthogonal (or perpendicular) to vector a = (𝑎₁, 𝑎₂, … , 𝑎n) (by definition, ax iff ax = 0), and therefore, it is perpendicular to all scalar multiples of a. So every solution of homogeneous linear equation ax = 0 is orthogonal to the span(a). We can reformulate the general solution set of homogeneous equation \eqref{EqHyp.1} as
\[ \mbox{span}({\bf a})^{\perp} = \left\{ {\bf x} \in \mathbb{F}^n \ : \ {\bf a} \bullet {\bf x} = 0 \right\} . \]
Every nonzero vector from span{a} is a normal vector to the hyperspace ax = 0. Therefore, solving the linear homogeneous equation ax = 0 corresponds to finding all vectors x orthogonal to a in 𝔽n.

Theorem 12: If ϕ is a nonzero linear functional on a vector space V, then the null space of ϕ is a hyperspace of V. Conversely, every hyperspace of V is the null space of a (not unique) nonzero linear functional on V.
Let φ be a nonzero linear functional on the vector space V and W the null space of φ. We have to show that W is a hyperspace of V. It is obvious that VW. Also W ≠ {0} as dimV > 1 and V is finite-dimensional. This shows that W is a proper subspace of V.

According to the rank nullity theorem. the dimension of the domain of a linear transformation φ is the sum of the rank of φ (the dimension of the image of φ) and the nullity of φ (the dimension of the kernel of φ). Since the dimension of the image of φ is 1 (it is either ℝ or ℂ---both are one dimensional spaces), we conclude that the kernel (null space) of φ has dimension n−1, so it is hyperspace.

Conversely, suppose that W is a hyperspace of V. Then we have to construct a nonzero linear functional ψ on V such that null space of ψ is W, as {0} ≠ WV. There exists a basis u1, u2, … , un-1W because the dimension of a hyperspace is 1 less than dimension of V. Let 0u be a vector in V that does not belong to W. If we set ψ(u) = 1, but ψ annihilates every element from the basis uj (j = 1, 2, … , n−1), then ψ becomes the required linear form.

   
Example 19: Let ϕ be a linear functional on ℝ² defined by \[ \phi ({\bf x}) = \phi (x, y) = 2\,x+y , \qquad x, y \in \mathbb{R} . \] The corresponding homogeneous relation \[ \phi (x, y) = 0 \qquad \iff \qquad 2\,x+y = 0 \tag{19.1} \] defines a line in ℝ². Rewriting Eq.(19.1) as a dot-product \[ {\bf a} \bullet {\bf x} = 0 \qquad \iff \qquad 2\,x_1 + x_2 = 0 \] for a = (2, 1) and x = (x₁, x₂), we see that vector a is perpendicular to the line (18.1).
line = Graphics[{Black, Thick, Line[{{-1, -0.5}, {1, 0.5}}]}];
n = Graphics[{Blue, Thickness[0.01], Arrowheads[0.1], Arrow[{{0, 0}, {-0.5, 1}}]}]; hor = Graphics[{Dashed, Line[{{-1, 0}, {1, 0}}]}];
ver = Graphics[{Dashed, Line[{{0, -0.5}, {0, 1}}]}];
txt1 = Graphics[ Text[Style["2 x + y = 0", 25, Italic, Black], {0.5, 0.5}]];
txt2 = Graphics[Text[Style["n", 25, Bold, Black], {-0.25, 0.95}]];
Show[hor, ver, n, line, txt1, txt2]
Normal vector to hyperspace.

This is a wonderful benefit of numeric/geometric duality: if we have a homogeneous linear equation in two variables, we can visualize its solution as a line.

We can do the same job algebraically too. If we solve 2 x + y = 0 by row- reducing the one-rowed matrix [2 1] ∼ [1, 1/2], we get one free column, and hence one homogeneous generator h = [−1, 2]. Our solution mapping consequently takes the form \[ H(t) = t\,{\bf h} = t\left[ -2, \ 1 \right] . \]

Now consider an inhomogeneous two-variable equation \[ {\bf a} \bullet {\bf x} = b \qquad (b \ne 0) . \tag{18.2} \] Solve this by row-reducing the one-rowed augmented matrix \[ \left[ 2 \ 1 \vert \ b \right] . \] In our example, a = [2, 1] ≠ (0, 0), so we must get a pivot when we row-reduce. That leaves one free column, producing one homogeneous generator h that we found previously. The general solution of the inhomogeneous equation ax = b therefore takes the form \[ {\bf x}(t) = {\bf x}_p + {\bf x}_h = {\bf x}_p + t\,{\bf h} , \] where xp is a particular solution of the nonhomogeneous equation (18.2) and xh is the general solution of the corresponding homogeneous equation. Here h = [−1, 2] is the homogeneous generator for the sequation (18.1).

line = Graphics[{Black, Thick, Line[{{-1, -0.5}, {1, 0.5}}]}];
line2 = Graphics[{Purple, Thick, Line[{{-0.9, -0.3}, {1.1, 0.7}}]}];
n = Graphics[{Blue, Thickness[0.01], Arrowheads[0.1], Arrow[{{0, 0}, {-0.5, 1}}]}]; hor = Graphics[{Dashed, Line[{{-1, 0}, {1, 0}}]}];
ver = Graphics[{Dashed, Line[{{0, -0.5}, {0, 1}}]}];
arr = Graphics[{Blue, Thickness[0.01], , Arrowheads[0.04], Arrow[{{0.0, 0}, {-0.6, -0.15}}]}];
txt1 = Graphics[ Text[Style["2 x + y = b", 25, Italic, Black], {0.5, 0.77}]];
txt2 = Graphics[Text[Style["n", 25, Bold, Black], {-0.25, 0.95}]];
xp = Graphics[ Text[Style[Subscript[x, p], 25, Italic, Black], {-0.77, -0.1}]];
disk = Graphics[{Purple, Disk[{-0.6, -0.15}, 0.02]}];
Show[hor, ver, n, line, line2,arr,txt1, txt2,xp,disk]
Solution set to inhomogeneous equation---hyperplane.
   ■
End of Example 19

Using isomorphisms of vector spaces ℝn ≌ ℝ1×n ≌ ℝn×1, we rewrite the homogeneous equation in matrix/vector form

\[ {\bf a} * {\bf x} = \left[ a_1 , a_2 , \ldots , a_n \right] \cdot \begin{bmatrix} x_1 \\ x_2 \\ \vdots \\ x_n \end{bmatrix} = \left[ 0 \right] \in \mathbb{R}^{1\times 1} . \]
Then the corresponding dot product of these two vectors a (wriiten in row form) and x (written in column form) is the entry of 1 × 1 matrix, which you need to extract.    

 

  1. Axler, Sheldon Jay (2015). Linear Algebra Done Right (3rd ed.). Springer. ISBN 978-3-319-11079-0.
  2. Halmos, Paul Richard (1974) [1958]. Finite-Dimensional Vector Spaces (2nd ed.). Springer. ISBN 0-387-90093-4.
  3. Katznelson, Yitzhak; Katznelson, Yonatan R. (2008). A (Terse) Introduction to Linear Algebra. American Mathematical Society. ISBN 978-0-8218-4419-9.
  4. Treil, S., Linear Algebra Done Wrong.
  5. Wikipedia, Dual space/