Compact operator on Hilbert space

By contrast, the study of general operators on infinite-dimensional spaces often requires a genuinely different approach.As stated above, the techniques used to prove results, e.g., the spectral theorem, in the non-compact case are typically different, involving operator-valued measures on the spectrum.Similar reasoning shows that if T is compact, then T is the uniform limit of some sequence of finite-rank operators.When a closed linear subspace L of H is invariant under T, then the restriction of T to L is a self-adjoint operator on L, and furthermore, the orthogonal complement L⊥ of L is also invariant under T. For example, the space H can be decomposed as the orthogonal direct sum of two T–invariant closed linear subspaces: the kernel of T, and the orthogonal complement (ker T)⊥ of the kernel (which is equal to the closure of the range of T, for any bounded self-adjoint operator).These basic facts play an important role in the proof of the spectral theorem below.The classification result for Hermitian n × n matrices is the spectral theorem: If M = M*, then M is unitarily diagonalizable, and the diagonalization of M has real entries.Theorem For every compact self-adjoint operator T on a real or complex Hilbert space H, there exists an orthonormal basis of H consisting of eigenvectors of T. More specifically, the orthogonal complement of the kernel of T admits either a finite orthonormal basis of eigenvectors of T, or a countably infinite orthonormal basis {en} of eigenvectors of T, with corresponding eigenvalues {λn} ⊂ R, such that λn → 0.Corollary For every compact self-adjoint operator T on a real or complex separable infinite-dimensional Hilbert space H, there exists a countably infinite orthonormal basis {fn} of H consisting of eigenvectors of T, with corresponding eigenvalues {μn} ⊂ R, such that μn → 0.Proving the spectral theorem for a Hermitian n × n matrix T hinges on showing the existence of one eigenvector x.Once this is done, Hermiticity implies that both the linear span and orthogonal complement of x (of dimension n − 1) are invariant subspaces of T. The desired result is then obtained by induction forIn the finite-dimensional case, part of the first approach works in much greater generality; any square matrix, not necessarily Hermitian, has an eigenvector.The spectral theorem for the compact self-adjoint case can be obtained analogously: one finds an eigenvector by extending the second finite-dimensional argument above, then apply induction.Note that while the Lagrange multipliers generalize to the infinite-dimensional case, the compactness of the unit sphere is lost.If f attains its maximum m(T) on B at some unit vector y, then, by the same argument used for matrices, y is an eigenvector of T, with corresponding eigenvalue λ = ⟨λy, y⟩ = ⟨Ty, y⟩ = f(y) = m(T).These two facts imply that f is continuous on B equipped with the weak topology, and f attains therefore its maximum m on B at some y ∈ B.Let T be a compact operator on a Hilbert space H. A finite (possibly empty) or countably infinite orthonormal sequence {en} of eigenvectors of T, with corresponding non-zero eigenvalues, is constructed by induction as follows.Let F = (span{en})⊥, where {en} is the finite or infinite sequence constructed by the inductive process; by self-adjointness, F is invariant under T. Let S denote the restriction of T to F. If the process was stopped after finitely many steps, with the last vector em−1, then F= Hm and S = Tm = 0 by construction.In the infinite case, compactness of T and the weak-convergence of en to 0 imply that Ten = λnen → 0, therefore λn → 0.The fact that S = 0 means that F is contained in the kernel of T. Conversely, if x ∈ ker(T) then by self-adjointness, x is orthogonal to every eigenvector {en} with non-zero eigenvalue.If F ≠ {0}, it is a non-trivial invariant subspace of T, and by the initial claim, there must exist a norm one eigenvector y of T in F. But then U ∪ {y} contradicts the maximality of U.It follows that F = {0}, hence span(U) is dense in H. This shows that U is an orthonormal basis of H consisting of eigenvectors of T. If T is compact on an infinite-dimensional Hilbert space H, then T is not invertible, hence σ(T), the spectrum of T, always contains 0.The functional calculus map Φ is defined in a natural way: let {en} be an orthonormal basis of eigenvectors for H, with corresponding eigenvalues {λn}; for f ∈ C(σ(T)), the operator Φ(f), diagonal with respect to the orthonormal basis {en}, is defined by settingConversely, any homomorphism Ψ satisfying the requirements of the theorem must coincide with Φ when f is a polynomial.By the Weierstrass approximation theorem, polynomial functions are dense in C(σ(T)), and it follows that Ψ = Φ.The more general continuous functional calculus can be defined for any self-adjoint (or even normal, in the complex case) bounded linear operator on a Hilbert space.-invariant non-trivial closed subspace; and thus by the lemma above, therein would lie a common eigenvector for the operators (necessarily orthogonal to Q).Since the operators restricted to the eigenspaces (which are finite-dimensional) are automatically all compact, we can apply Theorem 1 to each of these, and find orthonormal bases Qσ for theThen if G were compact then there is a unique decomposition of H into a countable direct sum of finite-dimensional, irreducible, invariant subspaces (this is essentially diagonalisation of the family of operatorsIf G were not compact, but were abelian, then diagonalisation is not achieved, but we get a unique continuous decomposition of H into 1-dimensional invariant subspaces.However, if U is identity plus a compact perturbation, U has only a countable spectrum, containing 1 and possibly, a finite set or a sequence tending to 1 on the unit circle.
functional analysisHilbert spacecompact operatorsfinite-rank operatorstopologyoperator normspectral theory of compact operatorsBanach spacesJordan canonical formnormalspectral theoremmeasuresspectrumbounded operatorsrelatively compactsequentially continuousweak topologystrong operator topologyCalkin algebraself-adjointorthogonal complementdirect sumkernelorthonormal basiscountably infiniteseparablecountableLagrange's multiplierRayleigh quotientpolarization identityBanach–Alaoglu theoremZorn's lemmafunctional calculusC*-algebraWeierstrass approximation theoremcontinuous functional calculushyponormal compact operatorsubnormal operatorunitary operatorL2([0, 1])Volterra operatorVolterra integral equationsHilbert-Schmidt integral operatorHilbert–Schmidt operatorMercer's theoremCompact operatorDecomposition of spectrum (functional analysis)Fredholm operatorStrictly singular operatorAdvances in MathematicstopicsglossaryBanachFréchetHilbertHölderNuclearOrliczSchwartzSobolevTopological vectorBarrelledCompleteLocally convexReflexiveHahn–BanachRiesz representationClosed graphUniform boundedness principleKrein–MilmanMin–maxGelfand–NaimarkBanach–AlaogluAdjointBoundedCompactHilbert–SchmidtTrace classTransposeUnboundedUnitaryBanach algebraSpectrum of a C*-algebraOperator algebraGroup algebra of a locally compact groupVon Neumann algebraInvariant subspace problemMahler's conjectureHardy spaceSpectral theory of ordinary differential equationsHeat kernelIndex theoremCalculus of variationsIntegral linear operatorJones polynomialTopological quantum field theoryNoncommutative geometryRiemann hypothesisDistributionGeneralized functionsApproximation propertyBalanced setChoquet theoryBanach–Mazur distanceTomita–Takesaki theoryHilbert spacesInner productL-semi-inner productPrehilbert spaceBessel's inequalityCauchy–Schwarz inequalityHilbert projection theoremParseval's identityDensely definedSesquilinear formCn(K) with K compact & n<∞Segal–Bargmann F