Unbounded operator

Summary

In mathematics, more specifically functional analysis and operator theory, the notion of unbounded operator provides an abstract framework for dealing with differential operators, unbounded observables in quantum mechanics, and other cases.

The term "unbounded operator" can be misleading, since

  • "unbounded" should sometimes be understood as "not necessarily bounded";
  • "operator" should be understood as "linear operator" (as in the case of "bounded operator");
  • the domain of the operator is a linear subspace, not necessarily the whole space;
  • this linear subspace is not necessarily closed; often (but not always) it is assumed to be dense;
  • in the special case of a bounded operator, still, the domain is usually assumed to be the whole space.

In contrast to bounded operators, unbounded operators on a given space do not form an algebra, nor even a linear space, because each one is defined on its own domain.

The term "operator" often means "bounded linear operator", but in the context of this article it means "unbounded operator", with the reservations made above. The given space is assumed to be a Hilbert space.[clarification needed] Some generalizations to Banach spaces and more general topological vector spaces are possible.

Short history edit

The theory of unbounded operators developed in the late 1920s and early 1930s as part of developing a rigorous mathematical framework for quantum mechanics.[1] The theory's development is due to John von Neumann[2] and Marshall Stone.[3] Von Neumann introduced using graphs to analyze unbounded operators in 1932.[4]

Definitions and basic properties edit

Let X, Y be Banach spaces. An unbounded operator (or simply operator) T : D(T) → Y is a linear map T from a linear subspace D(T) ⊆ X—the domain of T—to the space Y.[5] Contrary to the usual convention, T may not be defined on the whole space X.

An operator T is said to be closed if its graph Γ(T) is a closed set.[6] (Here, the graph Γ(T) is a linear subspace of the direct sum XY, defined as the set of all pairs (x, Tx), where x runs over the domain of T .) Explicitly, this means that for every sequence {xn} of points from the domain of T such that xnx and Txny, it holds that x belongs to the domain of T and Tx = y.[6] The closedness can also be formulated in terms of the graph norm: an operator T is closed if and only if its domain D(T) is a complete space with respect to the norm:[7]

 

An operator T is said to be densely defined if its domain is dense in X.[5] This also includes operators defined on the entire space X, since the whole space is dense in itself. The denseness of the domain is necessary and sufficient for the existence of the adjoint (if X and Y are Hilbert spaces) and the transpose; see the sections below.

If T : XY is closed, densely defined and continuous on its domain, then its domain is all of X.[nb 1]

A densely defined symmetric operator T on a Hilbert space H is called bounded from below if T + a is a positive operator for some real number a. That is, Tx|x⟩ ≥ −a ||x||2 for all x in the domain of T (or alternatively Tx|x⟩ ≥ a ||x||2 since a is arbitrary).[8] If both T and T are bounded from below then T is bounded.[8]

Example edit

Let C([0, 1]) denote the space of continuous functions on the unit interval, and let C1([0, 1]) denote the space of continuously differentiable functions. We equip   with the supremum norm,  , making it a Banach space. Define the classical differentiation operator d/dx : C1([0, 1]) → C([0, 1]) by the usual formula:

 

Every differentiable function is continuous, so C1([0, 1]) ⊆ C([0, 1]). We claim that d/dx : C([0, 1]) → C([0, 1]) is a well-defined unbounded operator, with domain C1([0, 1]). For this, we need to show that   is linear and then, for example, exhibit some   such that   and  .

This is a linear operator, since a linear combination a f  + bg of two continuously differentiable functions f , g is also continuously differentiable, and

 

The operator is not bounded. For example,

 

satisfy

 

but

 

as  .

The operator is densely defined, and closed.

The same operator can be treated as an operator ZZ for many choices of Banach space Z and not be bounded between any of them. At the same time, it can be bounded as an operator XY for other pairs of Banach spaces X, Y, and also as operator ZZ for some topological vector spaces Z.[clarification needed] As an example let IR be an open interval and consider

 

where:

 

Adjoint edit

The adjoint of an unbounded operator can be defined in two equivalent ways. Let   be an unbounded operator between Hilbert spaces.

First, it can be defined in a way analogous to how one defines the adjoint of a bounded operator. Namely, the adjoint   of T is defined as an operator with the property:

 
More precisely,   is defined in the following way. If   is such that   is a continuous linear functional on the domain of T, then   is declared to be an element of   and after extending the linear functional to the whole space via the Hahn–Banach theorem, it is possible to find some   in   such that
 
since Riesz representation theorem allows the continuous dual of the Hilbert space   to be identified with the set of linear functionals given by the inner product. This vector   is uniquely determined by   if and only if the linear functional   is densely defined; or equivalently, if T is densely defined. Finally, letting   completes the construction of   which is necessarily a linear map. The adjoint   exists if and only if T is densely defined.

By definition, the domain of   consists of elements   in   such that   is continuous on the domain of T. Consequently, the domain of   could be anything; it could be trivial (that is, contains only zero).[9] It may happen that the domain of   is a closed hyperplane and   vanishes everywhere on the domain.[10][11] Thus, boundedness of   on its domain does not imply boundedness of T. On the other hand, if   is defined on the whole space then T is bounded on its domain and therefore can be extended by continuity to a bounded operator on the whole space.[nb 2] If the domain of   is dense, then it has its adjoint  [12] A closed densely defined operator T is bounded if and only if   is bounded.[nb 3]

The other equivalent definition of the adjoint can be obtained by noticing a general fact. Define a linear operator   as follows:[12]

 
Since   is an isometric surjection, it is unitary. Hence:   is the graph of some operator   if and only if T is densely defined.[13] A simple calculation shows that this "some"   satisfies:
 
for every x in the domain of T. Thus   is the adjoint of T.

It follows immediately from the above definition that the adjoint   is closed.[12] In particular, a self-adjoint operator (meaning  ) is closed. An operator T is closed and densely defined if and only if  [nb 4]

Some well-known properties for bounded operators generalize to closed densely defined operators. The kernel of a closed operator is closed. Moreover, the kernel of a closed densely defined operator   coincides with the orthogonal complement of the range of the adjoint. That is,[14]

 
von Neumann's theorem states that   and   are self-adjoint, and that   and   both have bounded inverses.[15] If   has trivial kernel, T has dense range (by the above identity.) Moreover:
T is surjective if and only if there is a   such that   for all   in  [nb 5] (This is essentially a variant of the so-called closed range theorem.) In particular, T has closed range if and only if   has closed range.

In contrast to the bounded case, it is not necessary that   since, for example, it is even possible that   does not exist.[citation needed] This is, however, the case if, for example, T is bounded.[16]

A densely defined, closed operator T is called normal if it satisfies the following equivalent conditions:[17]

  •  ;
  • the domain of T is equal to the domain of   and   for every x in this domain;
  • there exist self-adjoint operators   such that    and   for every x in the domain of T.

Every self-adjoint operator is normal.

Transpose edit

Let   be an operator between Banach spaces. Then the transpose (or dual)   of   is the linear operator satisfying:

 
for all   and   Here, we used the notation:  [18]

The necessary and sufficient condition for the transpose of   to exist is that   is densely defined (for essentially the same reason as to adjoints, as discussed above.)

For any Hilbert space   there is the anti-linear isomorphism:

 
given by   where   Through this isomorphism, the transpose   relates to the adjoint   in the following way:[19]
 
where  . (For the finite-dimensional case, this corresponds to the fact that the adjoint of a matrix is its conjugate transpose.) Note that this gives the definition of adjoint in terms of a transpose.

Closed linear operators edit

Closed linear operators are a class of linear operators on Banach spaces. They are more general than bounded operators, and therefore not necessarily continuous, but they still retain nice enough properties that one can define the spectrum and (with certain assumptions) functional calculus for such operators. Many important linear operators which fail to be bounded turn out to be closed, such as the derivative and a large class of differential operators.

Let X, Y be two Banach spaces. A linear operator A : D(A) ⊆ XY is closed if for every sequence {xn} in D(A) converging to x in X such that AxnyY as n → ∞ one has xD(A) and Ax = y. Equivalently, A is closed if its graph is closed in the direct sum XY.

Given a linear operator A, not necessarily closed, if the closure of its graph in XY happens to be the graph of some operator, that operator is called the closure of A, and we say that A is closable. Denote the closure of A by A. It follows that A is the restriction of A to D(A).

A core (or essential domain) of a closable operator is a subset C of D(A) such that the closure of the restriction of A to C is A.

Example edit

Consider the derivative operator A = d/dx where X = Y = C([a, b]) is the Banach space of all continuous functions on an interval [a, b]. If one takes its domain D(A) to be C1([a, b]), then A is a closed operator which is not bounded.[20] On the other hand if D(A) = C([a, b]), then A will no longer be closed, but it will be closable, with the closure being its extension defined on C1([a, b]).

Symmetric operators and self-adjoint operators edit

An operator T on a Hilbert space is symmetric if and only if for each x and y in the domain of T we have  . A densely defined operator T is symmetric if and only if it agrees with its adjoint T restricted to the domain of T, in other words when T is an extension of T.[21]

In general, if T is densely defined and symmetric, the domain of the adjoint T need not equal the domain of T. If T is symmetric and the domain of T and the domain of the adjoint coincide, then we say that T is self-adjoint.[22] Note that, when T is self-adjoint, the existence of the adjoint implies that T is densely defined and since T is necessarily closed, T is closed.

A densely defined operator T is symmetric, if the subspace Γ(T) (defined in a previous section) is orthogonal to its image J(Γ(T)) under J (where J(x,y):=(y,-x)).[nb 6]

Equivalently, an operator T is self-adjoint if it is densely defined, closed, symmetric, and satisfies the fourth condition: both operators Ti, T + i are surjective, that is, map the domain of T onto the whole space H. In other words: for every x in H there exist y and z in the domain of T such that Tyiy = x and Tz + iz = x.[23]

An operator T is self-adjoint, if the two subspaces Γ(T), J(Γ(T)) are orthogonal and their sum is the whole space  [12]

This approach does not cover non-densely defined closed operators. Non-densely defined symmetric operators can be defined directly or via graphs, but not via adjoint operators.

A symmetric operator is often studied via its Cayley transform.

An operator T on a complex Hilbert space is symmetric if and only if its quadratic form is real, that is, the number   is real for all x in the domain of T.[21]

A densely defined closed symmetric operator T is self-adjoint if and only if T is symmetric.[24] It may happen that it is not.[25][26]

A densely defined operator T is called positive[8] (or nonnegative[27]) if its quadratic form is nonnegative, that is,   for all x in the domain of T. Such operator is necessarily symmetric.

The operator TT is self-adjoint[28] and positive[8] for every densely defined, closed T.

The spectral theorem applies to self-adjoint operators [29] and moreover, to normal operators,[30][31] but not to densely defined, closed operators in general, since in this case the spectrum can be empty.[32][33]

A symmetric operator defined everywhere is closed, therefore bounded,[6] which is the Hellinger–Toeplitz theorem.[34]

Extension-related edit

By definition, an operator T is an extension of an operator S if Γ(S) ⊆ Γ(T).[35] An equivalent direct definition: for every x in the domain of S, x belongs to the domain of T and Sx = Tx.[5][35]

Note that an everywhere defined extension exists for every operator, which is a purely algebraic fact explained at Discontinuous linear map § General existence theorem and based on the axiom of choice. If the given operator is not bounded then the extension is a discontinuous linear map. It is of little use since it cannot preserve important properties of the given operator (see below), and usually is highly non-unique.

An operator T is called closable if it satisfies the following equivalent conditions:[6][35][36]

  • T has a closed extension;
  • the closure of the graph of T is the graph of some operator;
  • for every sequence (xn) of points from the domain of T such that xn → 0 and also Txny it holds that y = 0.

Not all operators are closable.[37]

A closable operator T has the least closed extension   called the closure of T. The closure of the graph of T is equal to the graph of  [6][35] Other, non-minimal closed extensions may exist.[25][26]

A densely defined operator T is closable if and only if T is densely defined. In this case   and  [12][38]

If S is densely defined and T is an extension of S then S is an extension of T.[39]

Every symmetric operator is closable.[40]

A symmetric operator is called maximal symmetric if it has no symmetric extensions, except for itself.[21] Every self-adjoint operator is maximal symmetric.[21] The converse is wrong.[41]

An operator is called essentially self-adjoint if its closure is self-adjoint.[40] An operator is essentially self-adjoint if and only if it has one and only one self-adjoint extension.[24]

A symmetric operator may have more than one self-adjoint extension, and even a continuum of them.[26]

A densely defined, symmetric operator T is essentially self-adjoint if and only if both operators Ti, T + i have dense range.[42]

Let T be a densely defined operator. Denoting the relation "T is an extension of S" by ST (a conventional abbreviation for Γ(S) ⊆ Γ(T)) one has the following.[43]

  • If T is symmetric then TT∗∗T.
  • If T is closed and symmetric then T = T∗∗T.
  • If T is self-adjoint then T = T∗∗ = T.
  • If T is essentially self-adjoint then TT∗∗ = T.

Importance of self-adjoint operators edit

The class of self-adjoint operators is especially important in mathematical physics. Every self-adjoint operator is densely defined, closed and symmetric. The converse holds for bounded operators but fails in general. Self-adjointness is substantially more restricting than these three properties. The famous spectral theorem holds for self-adjoint operators. In combination with Stone's theorem on one-parameter unitary groups it shows that self-adjoint operators are precisely the infinitesimal generators of strongly continuous one-parameter unitary groups, see Self-adjoint operator § Self-adjoint extensions in quantum mechanics. Such unitary groups are especially important for describing time evolution in classical and quantum mechanics.

See also edit

Notes edit

  1. ^ Suppose fj is a sequence in the domain of T that converges to gX. Since T is uniformly continuous on its domain, Tfj is Cauchy in Y. Thus, ( fj , T fj ) is Cauchy and so converges to some ( f , T f ) since the graph of T is closed. Hence, f  = g, and the domain of T is closed.
  2. ^ Proof: being closed, the everywhere defined   is bounded, which implies boundedness of   the latter being the closure of T. See also (Pedersen 1989, 2.3.11) for the case of everywhere defined T.
  3. ^ Proof:   So if   is bounded then its adjoint T is bounded.
  4. ^ Proof: If T is closed densely defined then   exists and is densely defined. Thus   exists. The graph of T is dense in the graph of   hence   Conversely, since the existence of   implies that that of   which in turn implies T is densely defined. Since   is closed, T is densely defined and closed.
  5. ^ If   is surjective then   has bounded inverse, denoted by   The estimate then follows since
     
    Conversely, suppose the estimate holds. Since   has closed range, it is the case that   Since   is dense, it suffices to show that   has closed range. If   is convergent then   is convergent by the estimate since
     
    Say,   Since   is self-adjoint; thus, closed, (von Neumann's theorem),   QED
  6. ^ Follows from (Pedersen 1989, 5.1.5) and the definition via adjoint operators.

References edit

Citations edit

  1. ^ Reed & Simon 1980, Notes to Chapter VIII, page 305
  2. ^ von Neumann 1930, pp. 49–131
  3. ^ Stone 1932
  4. ^ von Neumann 1932, pp. 294–310
  5. ^ a b c Pedersen 1989, 5.1.1
  6. ^ a b c d e Pedersen 1989, 5.1.4
  7. ^ Berezansky, Sheftel & Us 1996, page 5
  8. ^ a b c d Pedersen 1989, 5.1.12
  9. ^ Berezansky, Sheftel & Us 1996, Example 3.2 on page 16
  10. ^ Reed & Simon 1980, page 252
  11. ^ Berezansky, Sheftel & Us 1996, Example 3.1 on page 15
  12. ^ a b c d e Pedersen 1989, 5.1.5
  13. ^ Berezansky, Sheftel & Us 1996, page 12
  14. ^ Brezis 1983, p. 28
  15. ^ Yoshida 1980, p. 200
  16. ^ Yoshida 1980, p. 195.
  17. ^ Pedersen 1989, 5.1.11
  18. ^ Yoshida 1980, p. 193
  19. ^ Yoshida 1980, p. 196
  20. ^ Kreyszig 1978, p. 294
  21. ^ a b c d Pedersen 1989, 5.1.3
  22. ^ Kato 1995, 5.3.3
  23. ^ Pedersen 1989, 5.2.5
  24. ^ a b Reed & Simon 1980, page 256
  25. ^ a b Pedersen 1989, 5.1.16
  26. ^ a b c Reed & Simon 1980, Example on pages 257-259
  27. ^ Berezansky, Sheftel & Us 1996, page 25
  28. ^ Pedersen 1989, 5.1.9
  29. ^ Pedersen 1989, 5.3.8
  30. ^ Berezansky, Sheftel & Us 1996, page 89
  31. ^ Pedersen 1989, 5.3.19
  32. ^ Reed & Simon 1980, Example 5 on page 254
  33. ^ Pedersen 1989, 5.2.12
  34. ^ Reed & Simon 1980, page 84
  35. ^ a b c d Reed & Simon 1980, page 250
  36. ^ Berezansky, Sheftel & Us 1996, pages 6,7
  37. ^ Berezansky, Sheftel & Us 1996, page 7
  38. ^ Reed & Simon 1980, page 253
  39. ^ Pedersen 1989, 5.1.2
  40. ^ a b Pedersen 1989, 5.1.6
  41. ^ Pedersen 1989, 5.2.6
  42. ^ Reed & Simon 1980, page 257
  43. ^ Reed & Simon 1980, pages 255, 256

Bibliography edit

  • Berezansky, Y.M.; Sheftel, Z.G.; Us, G.F. (1996), Functional analysis, vol. II, Birkhäuser (see Chapter 12 "General theory of unbounded operators in Hilbert spaces").
  • Brezis, Haïm (1983), Analyse fonctionnelle — Théorie et applications (in French), Paris: Mason
  • "Unbounded operator", Encyclopedia of Mathematics, EMS Press, 2001 [1994]
  • Hall, B.C. (2013), "Chapter 9. Unbounded Self-adjoint Operators", Quantum Theory for Mathematicians, Graduate Texts in Mathematics, vol. 267, Springer, ISBN 978-1461471158
  • Kato, Tosio (1995), "Chapter 5. Operators in Hilbert Space", Perturbation theory for linear operators, Classics in Mathematics, Springer-Verlag, ISBN 3-540-58661-X
  • Kreyszig, Erwin (1978). Introductory Functional Analysis With Applications. USA: John Wiley & Sons. Inc. ISBN 0-471-50731-8.
  • Pedersen, Gert K. (1989), Analysis now, Springer (see Chapter 5 "Unbounded operators").
  • Reed, Michael; Simon, Barry (1980), Methods of Modern Mathematical Physics, vol. 1: Functional Analysis (revised and enlarged ed.), Academic Press (see Chapter 8 "Unbounded operators").
  • Stone, Marshall Harvey (1932). Linear Transformations in Hilbert Space and Their Applications to Analysis. Reprint of the 1932 Ed. American Mathematical Society. ISBN 978-0-8218-7452-3.
  • Teschl, Gerald (2009). Mathematical Methods in Quantum Mechanics; With Applications to Schrödinger Operators. Providence: American Mathematical Society. ISBN 978-0-8218-4660-5.
  • von Neumann, J. (1930), "Allgemeine Eigenwerttheorie Hermitescher Functionaloperatoren (General Eigenvalue Theory of Hermitian Functional Operators)", Mathematische Annalen, 102 (1), doi:10.1007/BF01782338, S2CID 121249803
  • von Neumann, J. (1932), "Über Adjungierte Funktionaloperatore (On Adjoint Functional Operators)", Annals of Mathematics, Second Series, 33 (2), doi:10.2307/1968331, JSTOR 1968331
  • Yoshida, Kôsaku (1980), Functional Analysis (sixth ed.), Springer

This article incorporates material from Closed operator on PlanetMath, which is licensed under the Creative Commons Attribution/Share-Alike License.