In mathematics, any vector space
V
V,
The dual space as defined above is defined for all vector spaces, and to avoid ambiguity may also be called the .When defined for a topological vector space, there is a subspace of the dual space, corresponding to continuous linear functionals, called the continuous dual space.
Dual vector spaces find application in many branches of mathematics that use vector spaces, such as in tensor analysis with finite-dimensional vector spaces.When applied to vector spaces of functions (which are typically infinite-dimensional), dual spaces are used to describe measures, distributions, and Hilbert spaces. Consequently, the dual space is an important concept in functional analysis.
Early terms for dual include polarer Raum [Hahn 1927], espace conjugué, adjoint space [Alaoglu 1940], and transponierter Raum [Schauder 1930] and [Banach 1932]. The term dual is due to Bourbaki 1938.
V
F
V*
V\lor
V'
\varphi:V\toF
\hom(V,F)
V*
F
\begin{align} (\varphi+\psi)(x)&=\varphi(x)+\psi(x)\\ (a\varphi)(x)&=a\left(\varphi(x)\right) \end{align}
\varphi,\psi\inV*
x\inV
a\inF
Elements of the algebraic dual space
V*
The pairing of a functional
\varphi
V*
x
V
\varphi(x)=[x,\varphi]
\varphi(x)=\langlex,\varphi\rangle
\langle ⋅ , ⋅ \rangle:V x V*\toF
See also: Dual basis. If
V
V*
V
\{e1,...,en\}
V
V*
\{e1,...,en\}
V
ei(c1
ne | |
e | |
n) |
=ci, i=1,\ldots,n
ci\inF
i(e | |
e | |
j) |
=
i | |
\delta | |
j |
i | |
\delta | |
j |
Consider
\{e1,...,en\}
\{e1,...,en\}
ei(c1
ne | |
e | |
n) |
=ci, i=1,\ldots,n
These are a basis of
V*
ei,i=1,2,...,n,
x,y\inV
x=\alpha1e1+...+\alphanen
y=\beta1e1+...+\betanen
i(x)=\alpha | |
e | |
i |
i(y)=\beta | |
e | |
i |
x+λy=(\alpha1+λ\beta1)e1+...+(\alphan+λ\betan)en
ei(x+λy)=\alphai+λ\beta
i(x)+λ | |
i=e |
ei(y)
ei\inV*
i=1,2,...,n
λ1e1+ … +λnen=0\inV*
V
λ1=λ2=...=λn=0
ei
λi
\{e1,...,en\}
V*
g\inV*
g(x)=g(\alpha1e1+...+\alphanen)=\alpha1g(e1)+...+\alphang(e
1(x)g(e | |
1) |
+...+
n(x)g(e | |
e | |
n) |
\{e1,...,en\}
V*
V*
For example, if
V
\R2
\{e1=(1/2,1/2),e2=(0,1)\}
e1
e2
1(e | |
e | |
1)=1 |
1(e | |
e | |
2)=0 |
2(e | |
e | |
1)=0 |
2(e | |
e | |
2)=1 |
\begin{bmatrix} e11&e12\\ e21&e22\end{bmatrix} \begin{bmatrix} e11&e21\\ e12&e22\end{bmatrix} = \begin{bmatrix} 1&0\\ 0&1 \end{bmatrix}.
\{e1=(2,0),e2=(-1,1)\}
e1
e2
e1(x,y)=2x
e2(x,y)=-x+y
In general, when
V
\Rn
E=[e1| … |en]
\hat{E}=[e1| … |en]
\hat{E}rm{T} ⋅ E=In,
In
n
x\inV
x=\sumi\langlex,ei\rangleei=\sumi\langlex,ei\rangleei,
\langle ⋅ , ⋅ \rangle
In particular,
\Rn
n
n
\Rn
n
x
y
M
x
n x 1
y
1 x 1
Mx=y
M
1 x n
M
If
V
V*
V
V*
V
V*
V
If
V
e\alpha
A
e\alpha
\alpha\inA
For instance, consider the space
\Rinfty
\N
i\in\N
ei
i
\Rinfty
\R\N
(an)
(xn)
\Rinfty
\sumnanxn,
which is a finite sum because there are only finitely many nonzero
xn
\Rinfty
\R\N
This observation generalizes to any[8] infinite-dimensional vector space
V
F
\{e\alpha:\alpha\inA\}
V
A) | |
(F | |
0 |
f:A\toF
f\alpha=f(\alpha)
\alpha\inA
f
\sum\alpha\inf\alphae\alpha
in
V
f
v\inV
The dual space of
V
FA
A
F
T
V
\theta\alpha=T(e\alpha)
V
\theta:A\toF
\theta(\alpha)=\theta\alpha
T
V
T\left(\sum\alpha\inf\alphae\alpha\right)=\sum\alphaf\alphaT(e\alpha)=\sum\alpha\inf\alpha\theta\alpha.
Again, the sum is finite because
f\alpha
\alpha
The set
A) | |
(F | |
0 |
F
A
V\cong
A) | |
(F | |
0\congoplus |
\alpha\inF.
On the other hand,
FA
F
A
V*\cong\left(oplus\alpha\inF\right)*\cong\prod\alpha\inF*\cong\prod\alpha\inF\congFA
If a vector space is not finite-dimensional, then its (algebraic) dual space is always of larger dimension (as a cardinal number) than the original vector space. This is in contrast to the case of the continuous dual space, discussed below, which may be isomorphic to the original vector space even if the latter is infinite-dimensional.
The proof of this inequality between dimensions results from the following.
If
V
F
dim(V)=|A|<|F||A|=|V\ast|=max(|dim(V\ast)|,|F|),
dim(V)<dim(V*),
|F|\le|dim(V\ast)|,
If V is finite-dimensional, then V is isomorphic to V∗. But there is in general no natural isomorphism between these two spaces. Any bilinear form on V gives a mapping of V into its dual space via
v\mapsto\langlev, ⋅ \rangle
where the right hand side is defined as the functional on V taking each to . In other words, the bilinear form determines a linear mapping
\Phi\langle ⋅ , ⋅ \rangle:V\toV*
defined by
\left[\Phi\langle ⋅ , ⋅ \rangle(v),w\right]=\langlev,w\rangle.
If the bilinear form is nondegenerate, then this is an isomorphism onto a subspace of V∗.If V is finite-dimensional, then this is an isomorphism onto all of V∗. Conversely, any isomorphism
\Phi
\langlev,w\rangle\Phi=(\Phi(v))(w)=[\Phi(v),w].
Thus there is a one-to-one correspondence between isomorphisms of V to a subspace of (resp., all of) V∗ and nondegenerate bilinear forms on V.
If the vector space V is over the complex field, then sometimes it is more natural to consider sesquilinear forms instead of bilinear forms.In that case, a given sesquilinear form determines an isomorphism of V with the complex conjugate of the dual space
\Phi\langle:V\to\overline{V*}.
\overline{V*}
f(\alphav)=\overline{\alpha}f(v).
\Psi
V
V**=\{\Phi:V*\toF:\Phi linear\}
(\Psi(v))(\varphi)=\varphi(v)
v\inV,\varphi\inV*
*\to | |
ev | |
v:V |
F
\varphi\mapsto\varphi(v)
\Psi:V\toV**
v\mapstoevv
\Psi
V
See main article: Transpose of a linear map. If is a linear map, then the transpose (or dual) is defined by
f*(\varphi)=\varphi\circf
\varphi\inW*
f*(\varphi)
V*
\varphi
f
The following identity holds for all
\varphi\inW*
v\inV
[f*(\varphi),v]=[\varphi,f(v)],
The assignment produces an injective linear map between the space of linear operators from V to W and the space of linear operators from W to V; this homomorphism is an isomorphism if and only if W is finite-dimensional.If then the space of linear maps is actually an algebra under composition of maps, and the assignment is then an antihomomorphism of algebras, meaning that .In the language of category theory, taking the dual of vector spaces and the transpose of linear maps is therefore a contravariant functor from the category of vector spaces over F to itself.It is possible to identify (f) with f using the natural injection into the double dual.
If the linear map f is represented by the matrix A with respect to two bases of V and W, then f is represented by the transpose matrix AT with respect to the dual bases of W and V, hence the name.Alternatively, as f is represented by A acting on the left on column vectors, f is represented by the same matrix acting on the right on row vectors.These points of view are related by the canonical inner product on Rn, which identifies the space of column vectors with the dual space of row vectors.
Let
S
V
S
V*
S0
f\inV*
[f,s]=0
s\inS
S0
f:V\toF
S
f|S=0
The annihilator of a subset is itself a vector space.The annihilator of the zero vector is the whole dual space:
\{0\}0=V*
V0=\{0\}\subseteqV*
V
\{0\}\subseteqS\subseteqT\subseteqV
\{0\}\subseteqT0\subseteqS0\subseteqV*.
If
A
B
V
A0+B0\subseteq(A\capB)0.
(Ai)i\in
V
i
I
\left(cupi\inAi\right)0=capi\in
0 | |
A | |
i |
.
A
B
V
(A+B)0=A0\capB0
(A\capB)0=A0+B0.
If
V
W
W00=W
W
V ≈ V**
If
W
V
V/W
f:V\toF
V/W
W
f
(V/W)*\congW0.
V
A
B
V*
A0
B0
The dual space is analogous to a "negative"-dimensional space. Most simply, since a vector
v\inV
\varphi\inV*
\langlex,\varphi\rangle:=\varphi(x)\inF
V ⊕ V*
This arises in physics via dimensional analysis, where the dual space has inverse units.[12] Under the natural pairing, these units cancel, and the resulting scalar value is dimensionless, as expected. For example, in (continuous) Fourier analysis, or more broadly time–frequency analysis:[13] given a one-dimensional vector space with a unit of time, the dual space has units of frequency: occurrences per unit of time (units of). For example, if time is measured in seconds, the corresponding dual unit is the inverse second: over the course of 3 seconds, an event that occurs 2 times per second occurs a total of 6 times, corresponding to
3s ⋅ 2s-1=6
When dealing with topological vector spaces, the continuous linear functionals from the space into the base field
F=\Complex
\R
V*
V'
V
V'
\varphi:V\to{F}
l{D}
l{D}',
l{E}
l{E}',
l{S},
l{S}',
If is a Hausdorff topological vector space (TVS), then the continuous dual space of is identical to the continuous dual space of the completion of .
See main article: Polar topology and Dual system.
There is a standard construction for introducing a topology on the continuous dual
V'
V
l{A}
V
V
l{A},
\|\varphi\|A=\supx\in|\varphi(x)|,
where
\varphi
V
A
l{A}.
This means that a net of functionals
\varphii
\varphi
V'
forallA\inl{A} \|\varphii-\varphi\|A=\supx\in|\varphii(x)-\varphi(x)|\underset{i\toinfty}{\longrightarrow}0.
Usually (but not necessarily) the class
l{A}
x
V
A\inl{A}
forallx\inV thereexistssomeA\inl{A} suchthatx\inA.
A\inl{A}
B\inl{A}
C\inl{A}
forallA,B\inl{A} thereexistssomeC\inl{A} suchthatA\cupB\subseteqC.
l{A}
forallA\inl{A} andallλ\in{F} suchthatλ ⋅ A\inl{A}.
If these requirements are fulfilled then the corresponding topology on
V'
UA~=~\left\{\varphi\inV'~:~ \|\varphi\|A<1\right\}, forA\inl{A}
form its local base.
Here are the three most important special cases.
V'
V
l{A}
V
V
V'
\|\varphi\|=\sup\|x\||\varphi(x)|.
V'
V
l{A}
V
V'
V
l{A}
V
Each of these three choices of topology on
V'
V'
V'
V'
Let 1 < p < ∞ be a real number and consider the Banach space ℓ p of all sequences for which
\|a\|p=\left(
infty | |
\sum | |
n=0 |
p | |
|a | |
n| |
\right)
| ||||
<infty.
Define the number q by . Then the continuous dual of ℓ p is naturally identified with ℓ q: given an element
\varphi\in(\ellp)'
(\varphi(en))
en
\varphi
\varphi(b)=\sumnanbn
for all (see Hölder's inequality).
In a similar manner, the continuous dual of is naturally identified with (the space of bounded sequences).Furthermore, the continuous duals of the Banach spaces c (consisting of all convergent sequences, with the supremum norm) and c0 (the sequences converging to zero) are both naturally identified with .
By the Riesz representation theorem, the continuous dual of a Hilbert space is again a Hilbert space which is anti-isomorphic to the original space.This gives rise to the bra–ket notation used by physicists in the mathematical formulation of quantum mechanics.
By the Riesz–Markov–Kakutani representation theorem, the continuous dual of certain spaces of continuous functions can be described using measures.
See also: Transpose of a linear map.
If is a continuous linear map between two topological vector spaces, then the (continuous) transpose is defined by the same formula as before:
T'(\varphi)=\varphi\circT, \varphi\inW'.
The resulting functional is in . The assignment produces a linear map between the space of continuous linear maps from V to W and the space of linear maps from to .When T and U are composable continuous linear maps, then
(U\circT)'=T'\circU'.
When V and W are normed spaces, the norm of the transpose in is equal to that of T in .Several properties of transposition depend upon the Hahn–Banach theorem.For example, the bounded linear map T has dense range if and only if the transpose is injective.
When T is a compact linear map between two Banach spaces V and W, then the transpose is compact.This can be proved using the Arzelà–Ascoli theorem.
When V is a Hilbert space, there is an antilinear isomorphism iV from V onto its continuous dual .For every bounded linear map T on V, the transpose and the adjoint operators are linked by
iV\circT*=T'\circiV.
When T is a continuous linear map between two topological vector spaces V and W, then the transpose is continuous when and are equipped with "compatible" topologies: for example, when for and, both duals have the strong topology of uniform convergence on bounded sets of X, or both have the weak-∗ topology of pointwise convergence on X.The transpose is continuous from to, or from to .
Assume that W is a closed linear subspace of a normed space V, and consider the annihilator of W in,
W\perp=\{\varphi\inV':W\subseteq\ker\varphi\}.
Then, the dual of the quotient can be identified with W⊥, and the dual of W can be identified with the quotient .Indeed, let P denote the canonical surjection from V onto the quotient ; then, the transpose is an isometric isomorphism from into, with range equal to W⊥.If j denotes the injection map from W into V, then the kernel of the transpose is the annihilator of W:
\ker(j')=W\perp
If the dual of a normed space is separable, then so is the space itself.The converse is not true: for example, the space is separable, but its dual is not.
In analogy with the case of the algebraic double dual, there is always a naturally defined continuous linear operator from a normed space V into its continuous double dual, defined by
\Psi(x)(\varphi)=\varphi(x), x\inV, \varphi\inV'.
As a consequence of the Hahn–Banach theorem, this map is in fact an isometry, meaning for all .Normed spaces for which the map Ψ is a bijection are called reflexive.
When V is a topological vector space then Ψ(x) can still be defined by the same formula, for every, however several difficulties arise.First, when V is not locally convex, the continuous dual may be equal to and the map Ψ trivial.However, if V is Hausdorff and locally convex, the map Ψ is injective from V to the algebraic dual of the continuous dual, again as a consequence of the Hahn–Banach theorem.[14]
Second, even in the locally convex setting, several natural vector space topologies can be defined on the continuous dual, so that the continuous double dual is not uniquely defined as a set. Saying that Ψ maps from V to, or in other words, that Ψ(x) is continuous on for every, is a reasonable minimal requirement on the topology of, namely that the evaluation mappings
\varphi\inV'\mapsto\varphi(x), x\inV,
be continuous for the chosen topology on . Further, there is still a choice of a topology on, and continuity of Ψ depends upon this choice.As a consequence, defining reflexivity in this framework is more involved than in the normed case.
V\lor
( ⋅ )*
F*
F
F*\omega
\omega
uses
V'
V
V'
V*
\R\N
V
V
T-1 | |
V |
VT