Reproducing kernel Hilbert space

In functional analysis (a branch of mathematics), a reproducing kernel Hilbert space (RKHS) is a Hilbert space of functions in which point evaluation is a continuous linear functional. Roughly speaking, this means that if two functions f and g in the RKHS are close in norm, i.e., ‖f-g‖ is small, then f and g are also pointwise close, i.e., |f(x)-g(x)| is small for all x. The reverse need not be true. It is not entirely straightforward to construct a Hilbert space of functions which is not an RKHS.[1] Note that L2 spacesare not Hilbert spaces of functions (and hence not RKHSs), but rather Hilbert spaces of equivalence classes of functions (for example, the functions f and g defined by f(x)=0 and g(x)=1Q\scriptstyle\mathbb{Q} are equivalent in L2). However, there are RKHSs in which the norm is an L2-norm, such as the space of band-limited functions (see the example below). An RKHS is associated with a kernel that reproduces every function in the space in the sense that for any xx in the set on which the functions are defined, "evaluation at xx" can be performed by taking an inner product with a function determined by the kernel. Such a reproducing kernel exists if and only if every evaluation functional is continuous.

The reproducing kernel was first introduced in the 1907 work of Stanisław Zaremba concerning boundary value problems for harmonic and biharmonic functionsJames Mercer simultaneously examined functions which satisfy the reproducing property in the theory of integral equations. The idea of the reproducing kernel remained untouched for nearly twenty years until it appeared in the dissertations of Gábor SzegőStefan Bergman, and Salomon Bochner. The subject was eventually systematically developed in the early 1950s by Nachman Aronszajn and Stefan Bergman.[2]

These spaces have wide applications, including complex analysisharmonic analysis, and quantum mechanics. Reproducing kernel Hilbert spaces are particularly important in the field of statistical learning theory because of the celebrated representer theorem which states that every function in an RKHS that minimises an empirical risk function can be written as a linear combination of the kernel function evaluated at the training points. This is a practically useful result as it effectively simplifies the empirical risk minimization problem from an infinite dimensional to a finite dimensional optimization problem.

For ease of understanding, we provide the framework for real-valued Hilbert spaces. The theory can be easily extended to spaces of complex-valued functions and hence include the many important examples of reproducing kernel Hilbert spaces that are spaces of analytic functions.[3]

Definition[edit]

Let XX be an arbitrary set and HH a Hilbert space of real-valued functions on XX. The evaluation functional over the Hilbert space of functions HH is a linear functional that evaluates each function at a point xx,

Lx:f↦f(x) ∀f∈H.L_{x}:f\mapsto f(x){\text{   }}\forall f\in H.

We say that H is a reproducing kernel Hilbert space if, for all xx in XXLxL_{x} is continuous at any ff in HH or, equivalently, if LxL_{x} is a bounded operator on HH, i.e. there exists some M > 0 such that

|Lx(f)|:=|f(x)|≤M∥f∥H ∀f∈H.{\displaystyle |L_{x}(f)|:=|f(x)|\leq M\|f\|_{H}{\text{   }}\forall f\in H.\,}

 

 

 

 

(1)

While property (1) is the weakest condition that ensures both the existence of an inner product and the evaluation of every function in HH at every point in the domain, it does not lend itself to easy application in practice. A more intuitive definition of the RKHS can be obtained by observing that this property guarantees that the evaluation functional can be represented by taking the inner product of ff with a function KxK_{x} in HH . This function is the so-called reproducing kernel for the Hilbert space HH from which the RKHS takes its name. More formally, the Riesz representation theorem implies that for all xx in XX there exists a unique element KxK_{x} of HH with the reproducing property,

f(x)=Lx(f)=⟨f, Kx⟩∀f∈H.f(x)=L_{x}(f)=\langle f,\ K_{x}\rangle \quad \forall f\in H.

 

 

 

 

(2)

Since KyK_{y} is itself a function in HH we have that for each y in XX

Ky(x)=⟨Ky, Kx⟩.{\displaystyle K_{y}(x)=\langle K_{y},\ K_{x}\rangle .}

This allows us to define the reproducing kernel of HH as a function K:X×X→RK:X\times X\to \mathbb {R}  by

K(x,y)=⟨Kx, Ky⟩.{\displaystyle K(x,y)=\langle K_{x},\ K_{y}\rangle .}

From this definition it is easy to see that K:X×X→RK:X\times X\to \mathbb {R}  is both symmetric and positive definite, i.e.

∑i,j=1ncicjK(xi,xj)≥0\sum _{i,j=1}^{n}c_{i}c_{j}K(x_{i},x_{j})\geq 0

for any n∈N,x1,…,xn∈X, and c1,…,cn∈R.n\in \mathbb {N} ,x_{1},\dots ,x_{n}\in X,{\text{ and }}c_{1},\dots ,c_{n}\in \mathbb {R} .[4] The Moore-Aronszajn theorem (see below) is a sort of converse to this: if a function KK satisfies these conditions then there is a Hilbert space of functions on XX for which it is a reproducing kernel.

Example[edit]

The space of bandlimited functions HH is a RKHS. Fix some a<∞{\displaystyle a<\infty } and define

H={f∈L2(R)|supp⁡(ϕ)⊂[−a,a]}{\displaystyle H=\{f\in L_{2}(\mathbb {R} )|\operatorname {supp} (\phi )\subset [-a,a]\}}

where ϕ(ω)=∫f(x)e−iωxdx\phi (\omega )=\int f(x)e^{-i\omega x}dx is the Fourier transform of ff. One can show that if f∈Hf\in H then

f(x)=12π∫−aaϕ(ω)eixωdωf(x)={\frac {1}{2\pi }}\int _{-a}^{a}\phi (\omega )e^{ix\omega }d\omega

for some ϕ∈L2[−a,a]\phi \in L_{2}[-a,a]. It then follows by the Cauchy–Schwarz inequality and Plancherel's Theorem that

|f(x)|≤12π∫−aa1dω12π∫−aa|ϕ(ω)|2dω=aπ∥f∥.|f(x)|\leq {\sqrt {​{\frac {1}{2\pi }}\int _{-a}^{a}1d\omega }}{\sqrt {​{\frac {1}{2\pi }}\int _{-a}^{a}|\phi (\omega )|^{2}d\omega }}={\sqrt {\frac {a}{\pi }}}\|f\|.

As this inequality shows that the evaluation functional is bounded and HH is also a Hilbert space, HH is indeed a RKHS.

The kernel function KxK_{x} in this case is given by

Kx(y)=aπsinc⁡(a(y−x))=sin⁡(a(y−x))π(y−x)K_{x}(y)={\frac {a}{\pi }}\operatorname {sinc} (a(y-x))={\frac {\sin(a(y-x))}{\pi (y-x)}}.

Note, that KxK_{x} in this case is the "bandlimited version" of the Dirac delta distribution and that KxK_{x} converges to δ(⋅−x)\delta (\cdot -x) in the weak sense, as explained in the entry for the sinc function.

Moore–Aronszajn theorem[edit]

We have seen how a reproducing kernel Hilbert space defines a reproducing kernel function that is both symmetric and positive definite. The Moore-Aronszajn theorem goes in the other direction; it states that every symmetric, positive definite kernel defines a unique reproducing kernel Hilbert space. The theorem first appeared in Aronszajn's Theory of Reproducing Kernels, although he attributes it to E. H. Moore.

Theorem. Suppose K is a symmetric, positive definite kernel on a set X. Then there is a unique Hilbert space of functions on X for which K is a reproducing kernel.

Proof. For all x in X, define Kx = K(x, ⋅ ). Let H0 be the linear span of {Kx : x ∈ X}. Define an inner product on H0 by

⟨∑j=1nbjKyj,∑i=1maiKxi⟩=∑i=1m∑j=1naibjK(yj,xi).\left\langle \sum _{j=1}^{n}b_{j}K_{y_{j}},\sum _{i=1}^{m}a_{i}K_{x_{i}}\right\rangle =\sum _{i=1}^{m}\sum _{j=1}^{n}{a_{i}}b_{j}K(y_{j},x_{i}).

The symmetry of this inner product follows from the symmetry of K and the non-degeneracy follows from the fact that K is positive definite.

Let H be the completion of H0 with respect to this inner product. Then H consists of functions of the form

f(x)=∑i=1∞aiKxi(x)f(x)=\sum _{i=1}^{\infty }a_{i}K_{x_{i}}(x)

where ∑i=1∞ai2K(xi,xi)<∞\sum _{i=1}^{\infty }a_{i}^{2}K(x_{i},x_{i})<\infty . The fact that the above sum converges for every x follows from the Cauchy–Schwarz inequality.

Now we can check the reproducing property (2):

⟨f,Kx⟩=⟨∑i=1∞aiKxi,Kx⟩=∑i=1∞aiK(xi,x)=f(x).\langle f,K_{x}\rangle =\left\langle \sum _{i=1}^{\infty }a_{i}K_{x_{i}},K_{x}\right\rangle =\sum _{i=1}^{\infty }a_{i}K(x_{i},x)=f(x).

To prove uniqueness, let G be another Hilbert space of functions for which K is a reproducing kernel. For any x and y in X, (2) implies that

⟨Kx,Ky⟩H=K(x,y)=⟨Kx,Ky⟩G.\langle K_{x},K_{y}\rangle _{H}=K(x,y)=\langle K_{x},K_{y}\rangle _{G}.\,

By linearity, ⟨⋅,⋅⟩H=⟨⋅,⋅⟩G\langle \cdot ,\cdot \rangle _{H}=\langle \cdot ,\cdot \rangle _{G} on the span of {Kx : x ∈ X}. Then H \subset  G because G is complete and contains H0 and hence contains its completion.

Now we need to prove that every element of G is in H. Let ff be an element of G. Since H is a closed subspace of G, we can write f=fH+fH⊥{\displaystyle f=f_{H}+f_{H^{\bot }}} where fH∈H{\displaystyle f_{H}\in H}and fH⊥∈H⊥{\displaystyle f_{H^{\bot }}\in H^{\bot }}. Now if x∈Xx\in X then, since K is a reproducing kernel of G :

f(x)=⟨Kx,f⟩=⟨Kx,fH⟩=fH(x){\displaystyle f(x)=\langle K_{x},f\rangle =\langle K_{x},f_{H}\rangle =f_{H}(x)}

Which shows that fH⊥=0{\displaystyle f_{H^{\bot }}=0} in G and concludes the proof.

Integral operators and Mercer's theorem[edit]

We may characterize a symmetric positive definite kernel KK via the integral operator using Mercer's theorem and obtain an additional view of the RKHS. Let XX be a compact space equipped with a strictly positive finite Borel measure μ\mu  and K:X×X→RK:X\times X\to \mathbb {R}  a continuous, symmetric, and positive definite function. Define the integral operator TK:L2(X)→L2(X)T_{K}:L_{2}(X)\rightarrow L_{2}(X) as

[TKf](⋅)=∫XK(⋅,t)f(t)dμ(t)[T_{K}f](\cdot )=\int _{X}K(\cdot ,t)f(t)\,d\mu (t)

where L2(X)L_{2}(X) is the space of square integrable functions with respect to μ\mu .

Mercer's theorem states that the spectral decomposition of the integral operator TKT_{K} of KK yields a series representation of KK in terms of the eigenvalues and eigenfunctions of TKT_{K}. This then implies that KK is a reproducing kernel so that the corresponding RKHS can be defined in terms of these eigenvalues and eigenfunctions. We provide the details below.

Under these assumptions TKT_{K} is a compact, continuous, self-adjoint, and positive operator. The spectral theorem for self-adjoint operators implies that there is an at most countable decreasing sequence (σi)i≥0(\sigma _{i})_{i}\geq 0 such that limi→∞σi=0\lim _{i\to \infty }\sigma _{i}=0 and TKϕi(x)=σiϕi(x)T_{K}\phi _{i}(x)=\sigma _{i}\phi _{i}(x), where the {ϕi}\{\phi _{i}\} form an orthonormal basis of L2(X)L_{2}(X). By the positivity TKT_{K}σi>0 ∀i\sigma _{i}>0{\text{  }}\forall i. One can also show that TKT_{K} maps continuously into the space of continuous functions C(X)C(X) and therefore we may choose continuous functions as the eigenvectors, that is, ϕi∈C(X) ∀i\phi _{i}\in C(X){\text{  }}\forall i. Then by Mercer's theorem KK may be written in terms of the eigenvalues and continuous eigenfunctions as

K(x,y)=∑j=1∞σjϕj(x)ϕj(y)K(x,y)=\sum _{j=1}^{\infty }\sigma _{j}\,\phi _{j}(x)\,\phi _{j}(y)

for all x,yx,y in XX such that limn→∞supu,v|K(u,v)−∑j=1nσjϕj(u)ϕj(v)|=0.\lim _{n\to \infty }\sup _{u,v}|K(u,v)-\sum _{j=1}^{n}\sigma _{j}\,\phi _{j}(u)\,\phi _{j}(v)|=0. This above series representation is referred to as a Mercer kernel or Mercer representation of KK.

Furthermore, it can be shown that the RKHS HH of KK is given by

H={f∈L2(X)|∑i=1∞⟨f,ϕi⟩2σi<∞}H=\left\{f{\in }L_{2}(X){\mathrel {\Bigg |}}\sum _{i=1}^{\infty }{\frac {\left\langle f,\phi _{i}\right\rangle ^{2}}{\sigma _{i}}}<\infty \right\}

where the inner product of HH given by ⟨f,g⟩H=∑i=1∞⟨f,ϕi⟩L2⟨g,ϕi⟩L2σi.\left\langle f,g\right\rangle _{H}=\sum _{i=1}^{\infty }{\frac {\left\langle f,\phi _{i}\right\rangle _{L_{2}}\left\langle g,\phi _{i}\right\rangle _{L_{2}}}{\sigma _{i}}}. This representation of the RKHS has application in probability and statistics, for example to the Karhunen-Loeve representation for stochastic processes and kernel PCA.

Feature maps[edit]

feature map is a map φ:X→F\varphi :X\rightarrow F, where FF is a Hilbert space which we will call the feature space. The first sections presented the connection between bounded/continuous evaluation functions, positive definite functions, and integral operators and in this section we provide another representation of the RKHS in terms of feature maps.

We first note that every feature map defines a kernel via

K(x,y)=⟨φ(x),φ(y)⟩.K(x,y)=\langle \varphi (x),\varphi (y)\rangle .

 

 

 

 

(3)

Clearly KK is symmetric and positive definiteness follows from the properties of inner product in FF. Conversely, every positive definite function and corresponding reproducing kernel Hilbert space has infinitely many associated feature maps such that (3) holds.

For example, we can trivially take F=HF=H and φ(x)=Kx\varphi (x)=K_{x} for all x∈Xx\in X. Then (3) is satisfied by the reproducing property. Another classical example of a feature map relates to the previous section regarding integral operators by taking F=ℓ2F=\ell ^{2} and φ(x)=(σiϕi(x))i\varphi (x)=({\sqrt {\sigma _{i}}}\phi _{i}(x))_{i}.

This connection between kernels and feature maps provides us with a new way to understand positive definite functions and hence reproducing kernels as inner products in HH. Moreover, every feature map can naturally define a RKHS by means of the definition of a positive definite function.

Lastly, feature maps allow us to construct function spaces that reveal another perspective on the RKHS. Consider the linear space

Hφ={f:X→R|∃w∈F,f(x)=⟨w,φ(x)⟩F,∀ x∈X}.H_{\varphi }=\{f:X\to \mathbb {R} |\exists w\in F,f(x)=\langle w,\varphi (x)\rangle _{F},\forall {\text{  }}x\in X\}.

We can define a norm on H_{\varphi } by

∥f∥φ=inf{∥w∥F:w∈F,f(x)=⟨w,φ(x)⟩F,∀ x∈X}.{\displaystyle \|f\|_{\varphi }={\text{inf}}\{\|w\|_{F}:w\in F,f(x)=\langle w,\varphi (x)\rangle _{F},\forall {\text{  }}x\in X\}.}

It can be shown that H_{\varphi } is a RKHS with kernel defined by K(x,y)=⟨φ(x),φ(y)⟩K(x,y)=\langle \varphi (x),\varphi (y)\rangle . This representation implies that the elements of the RKHS are inner products of elements in the feature space and can accordingly be seen as hyperplanes. This view of the RKHS is related to the kernel trick in machine learning.[5]

Properties[edit]

The following properties of RKHSs may be useful to readers.

  • Let (Xi)i=1p(X_{i})_{i=1}^{p} be a sequence of sets and (Ki)i=1p(K_{i})_{i=1}^{p} be a collection of corresponding positive definite functions on (Xi)i=1p(X_{i})_{i=1}^{p}. It then follows that
K((x1,…,xp),(y1,…,yp))=K1(x1,y1)…Kp(xp,yp)K((x_{1},\dots ,x_{p}),(y_{1},\dots ,y_{p}))=K_{1}(x_{1},y_{1})\dots K_{p}(x_{p},y_{p})

is a kernel on X=X1×⋯×XpX=X_{1}\times \dots \times X_{p}.

  • Let X0⊂XX_{0}\subset X, then the restriction of KK to X0×X0X_{0}\times X_{0} is also a reproducing kernel.
  • Consider a normalized kernel KK such that K(x,x)=1K(x,x)=1 for all x∈Xx\in X. Define a pseudo-metric on X as
dK(x,y)=∥Kx−Ky∥H2=2(1−K(x,y)) ∀x∈X{\displaystyle d_{K}(x,y)=\|K_{x}-K_{y}\|_{H}^{2}=2(1-K(x,y)){\text{  }}\forall x\in X}.

By the Cauchy–Schwarz inequality,

K(x,y)2≤K(x,x)K(y,y) ∀x,y∈X.K(x,y)^{2}\leq K(x,x)K(y,y){\text{  }}\forall x,y\in X.

This inequality allows us to view KK as a measure of similarity between inputs. If x,y∈Xx,y\in X are similar then K(x,y)K(x,y) will be closer to 1 while if x,y∈Xx,y\in X are dissimilar then K(x,y)K(x,y) will be closer to 0.

  • The closure of the span of {Kx|x∈X}\{K_{x}|x\in X\} coincides with HH.[6]

Examples[edit]

Common examples of kernels include:

  • Linear Kernel:
K(x,y)=⟨x,y⟩K(x,y)=\langle x,y\rangle
The RKHS  HH corresponding to this kernel is the dual space, consisting of functions  f(x)=⟨x,β⟩{\displaystyle f(x)=\langle x,\beta \rangle } with squared norm  ∥f∥H2=∥β∥2{\displaystyle \|f\|_{H}^{2}=\|\beta \|^{2}}
  • Polynomial Kernel:
K(x,y)=(α⟨x,y⟩+1)d,α∈R,d∈NK(x,y)=(\alpha \langle x,y\rangle +1)^{d},\alpha {\in }\mathbb {R} ,d{\in }\mathbb {N}

Other common examples are kernels which satisfy K(x,y)=K(∥x−y∥)K(x,y)=K(\|x-y\|). These are the radial basis function kernels.

  • Radial Basis Function Kernels:
  • Gaussian Kernel:
Sometimes referred to as the  Radial basis function kernel, or squared exponential kernel
K(x,y)=e−∥x−y∥22σ2,σ>0K(x,y)=e^{-{\frac {\|x-y\|^{2}}{2\sigma ^{2}}}},\sigma >0
  • Laplacian Kernel:
K(x,y)=e−∥x−y∥σ,σ>0K(x,y)=e^{-{\frac {\|x-y\|}{\sigma }}},\sigma >0
The squared norm of a function  ff in the RKHS  HH with this kernel is  ∥f∥H2=∫f(x)2dx+∫f′(x)2dx{\displaystyle \|f\|_{H}^{2}=\int f(x)^{2}dx+\int f'(x)^{2}dx}. [7]

We also provide examples of Bergman kernels. Let X be finite and let H consist of all complex-valued functions on X. Then an element of H can be represented as an array of complex numbers. If the usual inner product is used, then Kx is the function whose value is 1 at x and 0 everywhere else, and K(x,y) can be thought of as an identity matrix since K(x,y)=1 when x=y and K(x,y)=0 otherwise. In this case, H is isomorphic to Cn.

The case of X = D (where D denotes the unit disc) is more sophisticated. Here the Bergman space H2(D) is the space of square-integrable holomorphic functions on D. It can be shown that the reproducing kernel for H2(D) is

K(x,y)=1π1(1−xy¯)2.K(x,y)={\frac {1}{\pi }}{\frac {1}{(1-x{\overline {y}})^{2}}}.

Lastly, the space of band limited functions ff in L2(R)L^{2}(\mathbb {R} ) with bandwidth 2a{\displaystyle 2a} are a RKHS with reproducing kernel

K(x,y)=sin⁡a(x−y)π(x−y).{\displaystyle K(x,y)={\frac {\sin a(x-y)}{\pi (x-y)}}.}

Extension to vector-valued functions[edit]

In this section we extend the definition of the RKHS to spaces of vector-valued functions as this extension is particularly important in multi-task learning and manifold regularization. The main difference is that the reproducing kernel Γ\Gamma  is a symmetric function that is now a positive semi-definite matrix for any x,yx,y in XX. More formally, we define a vector-valued RKHS (vvRKHS) as a Hilbert space of functions f:X→RTf:X\to \mathbb {R} ^{T} such that for all c∈RTc\in \mathbb {R} ^{T} and x∈Xx\in X

Γxc(y)=Γ(x,y)c∈H for y∈X\Gamma _{x}c(y)=\Gamma (x,y)c\in H{\text{ for }}y\in X

and

⟨f,Γxc⟩H=f(x)⊺c.{\displaystyle \langle f,\Gamma _{x}c\rangle _{H}=f(x)^{\intercal }c.}

This second property parallels the reproducing property for the scalar-valued case. We note that this definition can also be connected to integral operators, bounded evaluation functions, and feature maps as we saw for the scalar-valued RKHS. We can equivalently define the vvRKHS as a vector-valued Hilbert space with a bounded evaluation functional and show that this implies the existence of a unique reproducing kernel by the Riesz Representation theorem. Mercer's theorem can also be extended to address the vector-valued setting and we can therefore obtain a feature map view of the vvRKHS. Lastly, it can also be shown that the closure of the span of {Γxc:x∈X,c∈RT}\{\Gamma _{x}c:x\in X,c\in \mathbb {R} ^{T}\} coincides with HH, another property similar to the scalar-valued case.

We can gain intuition for the vvRKHS by taking a component-wise perspective on these spaces. In particular, we find that every vvRKHS is isometrically isomorphic to a scalar-valued RKHS on a particular input space. Let Λ={1,…,T}\Lambda =\{1,\dots ,T\}. Consider the space X×ΛX\times \Lambda  and the corresponding reproducing kernel

γ:X×Λ×X×Λ→R.\gamma :X\times \Lambda \times X\times \Lambda \to \mathbb {R} .

 

 

 

 

(4)

As noted above, the RKHS associated to this reproducing kernel is given by the closure of the span of {γ(x,t):x∈X,t∈Λ}\{\gamma _{(x,t)}:x\in X,t\in \Lambda \} where  γ(x,t)(y,s)=γ((x,t),(y,s))\ \gamma _{(x,t)}(y,s)=\gamma ((x,t),(y,s)) for every set of pairs (x,t),(y,s)∈X×Λ(x,t),(y,s)\in X\times \Lambda .

The connection to the scalar-valued RKHS can then be made by the fact that every matrix-valued kernel can be identified with a kernel of the form of (4) via

Γ(x,y)(t,s)=γ((x,t),(y,s)).\Gamma (x,y)_{(t,s)}=\gamma ((x,t),(y,s)).

Moreover, every kernel with the form of (4) defines a matrix-valued kernel with the above expression. Now letting the map D:HΓ→HγD:H_{\Gamma }\to H_{\gamma } be defined as

(Df)(x,t)=⟨f(x),et⟩RT(Df)(x,t)=\langle f(x),e_{t}\rangle _{\mathbb {R} ^{T}}

where ete_{t} is the ttht^{th} component of the canonical basis for RT\mathbb {R} ^{T}, one can show that DD is bijective and an isometry between H_{\Gamma } and H_{\gamma }.

While this view of the vvRKHS can be quite useful in multi-task learning, it should be noted that this isometry does not reduce the study of the vector-valued case to that of the scalar-valued case. In fact, this isometry procedure can make both the scalar-valued kernel and the input space too difficult to work with in practice as properties of the original kernels are often lost.[8][9][10]

An important class of matrix-valued reproducing kernels are separable kernels which can factorized as the product of a scalar valued kernel and a TT-dimensional symmetric positive semi-definite matrix. In light of our previous discussion these kernels are of the form

γ((x,t),(y,s))=K(x,y)KT(t,s)\gamma ((x,t),(y,s))=K(x,y)K_{T}(t,s)

for all x,yx,y in XX and t,st,s in TT. As the scalar-valued kernel encodes dependencies between the inputs, we can observe that the matrix-valued kernel encodes dependencies among both the inputs and the outputs.

We lastly remark that the above theory can be further extended to spaces of functions with values in function spaces but obtaining kernels for these spaces is a more difficult task.[11]

See also[edit]

Notes[edit]

  1. Jump up ^ Alpay, D., and T. M. Mills. "A family of Hilbert spaces which are not reproducing kernel Hilbert spaces." J. Anal. Appl. 1.2 (2003): 107-111.
  2. Jump up ^ Okutmustur
  3. Jump up ^ Paulson
  4. Jump up ^ Durrett
  5. Jump up ^ Rosasco
  6. Jump up ^ Rosasco
  7. Jump up ^ Berlinet, Alain and Thomas, Christine. Reproducing kernel Hilbert spaces in Probability and Statistics, Kluwer Academic Publishers, 2004
  8. Jump up ^ De Vito
  9. Jump up ^ Zhang
  10. Jump up ^ Alvarez
  11. Jump up ^ Rosasco

References[edit]

评论 1
添加红包

请填写红包祝福语或标题

红包个数最小为10个

红包金额最低5元

当前余额3.43前往充值 >
需支付:10.00
成就一亿技术人!
领取后你会自动成为博主和红包主的粉丝 规则
hope_wisdom
发出的红包
实付
使用余额支付
点击重新获取
扫码支付
钱包余额 0

抵扣说明:

1.余额是钱包充值的虚拟货币,按照1:1的比例进行支付金额的抵扣。
2.余额无法直接购买下载,可以购买VIP、付费专栏及课程。

余额充值