Abstract

In many previous papers, an integral transform was just considered as a transform on appropriate function spaces. In this paper we deal with the integral transform as an operator on a function space. We then apply various operator theories to . Finally we give an application for the spectral representation of a self-adjoint operator which plays a key role in quantum mechanics.

1. Introduction and Definitions

It is a well-known fact that the spectral theory is one of the main subjects of modern functional analysis and applications. It arises quite naturally in connection with the problems of solving equations. In particular, the spectrum of bounded linear operators on a normed and Banach space is the most important concept to understand the spectral theories. Furthermore, the spectral representation is used widely to apply theories in many fields.

Let denote one-parameter Wiener space, that is, the space of continuous real-valued functions on with . Let denote the class of all Wiener measurable subsets of , and let denote Wiener measure. is a complete measure space, and we denote the Wiener integral of a Wiener integrable functional by

A subset of is said to be scale-invariant measurable provided is measurable for all , and a scale-invariant measurable set is said to be a scale-invariant null set provided for all . A property that holds except on a scale-invariant null set is said to hold scale invariant almost everywhere (s-a.e.) [1].

In [2], for each pair of nonzero complex numbers and , Lee introduced an integral transform of a functional on abstract Wiener space. For certain values of the parameters and and for certain classes of functionals, the Fourier-Wiener transform [3], the modified Fourier-Wiener transform [4], the Fourier-Feynman transform, and the Gauss transform are special cases of Lee's integral transform . These transforms play an important role in studying stochastic processes and functional integrals on infinite dimensional spaces.

In many papers [59], the authors studied the integral transform with related topics of functionals in several classes. Recently, in [10, 11], the authors established the existence of a generalized integral transform via a Gaussian process of functionals in a class and then obtained various relationships involving the convolution product and the first variation of them.

However, in all previous works, the authors have considered just as a transform of a functional on function space and then they obtained various relationships as seen in Remark 2.5 below. In this paper, we consider the integral transform as an operator on a Banach space. We then apply various operator theories to . Furthermore, we obtain various theorems of the spectral theory for involving the spectral representation.

Now we are ready to define the integral transform  (IT) of a functional on , the space of all complex-valued continuous functions defined on which vanish at , used in [59, 12].

Definition 1.1. Let be a functional defined on . For each pair of nonzero complex numbers and , the IT of is defined by if it exists.

Remark 1.2. When and , is the Fourier-Wiener transform introduced by Cameron and Martin in [3], and when and , is the modified Fourier-Wiener transform used by Cameron and Martin in [4].

For and , let denote the Paley-Wiener-Zygmund (PWZ) stochastic integral. Then we have the following assertions. (1)For each , exists for a.e. . (2)If is a function of bounded variation on , equals the Riemann-Stieltjes integral for s-a.e. . (3)The PWZ stochastic integral has the expected linearity property. (4)The PWZ stochastic integral is a Gaussian process with mean and variance . For a more detailed study of the PWZ stochastic integral, see [5, 10, 1214].

Let be an orthonormal set in . The following formula is a well-known Wiener integration formula. Let be Borel measurable and let . Then in the sense that if either side of (1.3) exists, both sides exist and equality holds.

2. Some Results as a Transform

In this section, we establish the existence of the IT of functionals in a class , as seen in Theorem 2.4 below. We then give the inverse IT of our IT. Finally we state possible relationships for the IT with related topics.

We start this section by describing the class of functionals that we work with in this paper. Let be an orthonormal set in . Let be the space of all functionals of the form for some positive integer (throughout this paper, is fixed), where is an entire function of the complex variables and To simplify the expressions, we use the following notation:

Remark 2.1. For any and in , we can always express by (2.1) and by using the same positive integer , where is an entire function and

Note that is a very rich class of functionals because contains the Schwartz class . These functionals are of interest in Feynman integration theories and quantum mechanics.

Now, we will introduce a notation. It will be convenient to express for the type of limiting integral that occurs in our paper. For appropriate functions and on , if then we say that the integral is to be interpreted as an -limiting integral, see [15].

The following lemma is due to Cameron and Storvick in [15, Lemma H].

Lemma 2.2. Let be nonzero complex number with . For , let Then , and If , the integral is to be interpreted as an -limiting integral; moreover, in this case

The following lemma is very useful in establishing the existence of the IT.

Lemma 2.3. Let , and be as in Lemma 2.2 and let be a nonzero complex number with . Let . Then and If and , the integral is to be interpreted as an -limiting integral; moreover, in this case

Proof. First note that for all nonzero real numbers with , it follows that But each side of the above expression is an analytic function of throughout the region . Hence, by the uniqueness theorem for analytic functions, the above equality holds for all with . Since , using Lemma 2.2 we have Furthermore, this means that is an element of and so we complete the proof of Lemma 2.3 as desired.

In our first theorem, we establish the existence of the IT of a functional in .

Theorem 2.4. Let and be as in Lemma 2.3 and let be given by (2.1). Then the IT of exists, belongs to , and is given by the formula for , where

Proof. We first note that (2.15) follows from (1.2) and (1.3). Clearly the function is an entire function since is an entire function. What is left to show is that the left-hand side of (2.16) is an element of . Now, we note that for all nonzero real values of and , As mentioned in the proof of Lemma 2.3, each side of the above expression is an analytic function of throughout the region and throughout the region . Hence, by the uniqueness theorem for analytic functions, the above equality holds for all and with and . Using Lemma 2.3, the function is an element of . In fact, Moreover, if and , which completes the proof of Theorem 2.4 as desired.

Remark 2.5. (1) Under the appropriate conditions for and , we could establish the existences of the convolution product (CP) and the first variation of functionals in as in Theorem 2.4. We will state just formulas without proof because the main purpose of this paper is to concern as an operator on Hilbert space.
(2) In [58, 10], the authors established various basic formulas for the IT involving the CP and the first variation of functionals in various classes. Like these, we can obtain various basic relationships for the IT with related topic of functionals in under appropriate conditions for and . We list some relationships as follows.(i)The IT of a CP is the product of ITs, (ii)A relationship among the CP, the IT, and the Inverse IT, (iii)A relationship between the IT, and the first variation, (iv)A relationship among the CP, the IT, and the first variation, (v)A relationship among the CP, the inverse IT, and the first variation, for , where the CP of and is defined by and the first variation is defined by formula if they exist.

In our next theorem, we establish the inverse IT of our IT of functionals in .

Theorem 2.6. Let and be as in Theorem 2.4 with and , and let be given by (2.1). Then for all . That is to say, is the inverse IT of the IT.

Proof. Since and , for all by Theorem 2.4. Also, since and , for all . By using the similar method, we can show that for all . In [2], the author showed that for a integrable functional , for all nonzero complex numbers and . Using this formula and (1.2), we have Hence we complete the proof of Theorem 2.6.

3. A Bounded Linear Operator

In previous Section 2, we have considered as a transform of functionals in . From now on we will consider as an operator from into and then apply various operator theories to the IT. In particular, we obtain various spectral theorems for an operator .

For and in , let denote the inner product on and .

Remark 3.1. One can show that is a complex normed linear space. Also, from the fact that is complete, one can easily show that the space is also complete.

In our first theorem in this section, we show that the operator is well defined on .

Theorem 3.2. Let and be as in Lemma 2.3. Then the is a well defined operator from into and Moreover, if and , then the operator preserves the norm, namely,

Proof. In Section 2, we showed that for each , exists, belongs to and where Furthermore, we obtained that This tells us that for all and satisfy the conditions in Lemma 2.3. Moreover, if and , which completes the proof of Theorem 3.2 as desired.

Next, we give a simple example to illustrate our results and formulas in Theorem 3.2.

Example 3.3. Let and be are nonzero real numbers with . Let be the Schwartz space of infinitely differentiable functions decaying at infinity together with all its derivatives faster than any polynomial of . For nonzero real values of , let Then and hence is an element of . Also, using (2.15), (2.16), and (1.3), we have Furthermore, we note that In fact, All expressions in Example 3.3 are valid for nonzero real values of and . But and hence they are still valid for nonzero complex values of and which satisfy the conditions in Lemma 2.3.

Now, we establish some basic operator theories for the operator . First, in our next theorem, we show that the operator is a bounded linear operator on .

Theorem 3.4. Let and be as in Lemma 2.3. Then is a bounded operator on and hence it is continuous on . Furthermore, if and , then is injective from into .

Proof. We first note that where is the operator norm of an operator . Hence is bounded and so it is continuous. Furthermore if and , then (3.3) tells us that preserves the norm and hence it is injective from into . So we complete the proof of Theorem 3.4.

The following corollary follows from Theorem 3.4 and some basic properties for bounded linear operators on Hilbert space.

Corollary 3.5. Let and be as in Theorem 3.4. Then we have the following assertions. (1)The null space of is closed. (2)Let be a sequence in with as for some . Then as . That is to say, it is closed. (3)For all and in , we have the Cauchy-Schwartz inequality

Remark 3.6. As mentioned in Section 1, the Fourier-Wiener transform, the modified Fourier-Wiener transform, the Fourier-Feynman transform, and the Gauss transform are also well-defined operators on . In particular, from the definition of analytic Fourier-Feynman transform, it is an injective operator on . Hence all those transforms can be applied to our main results and formulas in this paper. In particular, the authors studied that for and , the IT is an element of and That is to say, the IT is injective [7, 9]. This result is a special case of our result in this paper. In addition, in [3, 4], the authors showed that the Fourier Wiener transform acts as a unitary operator on .

We finish this section by stating that the operator is invertible.

Theorem 3.7. Let and be as in Theorem 3.4 with and . Then the inverse operator of the IT exists and is given by Furthermore, the null space consists of the zero vector only.

Proof. From Theorem 3.4, the operator is continuous from into . Since and is in , exists and is in . Also, since and is in , exists and is in . Now, using (2.27), for , which completes the proof of Theorem 3.7 as desired.

We have some observations for the inverse operator of .

Remark 3.8. (1) If and , then and always satisfy the hypotheses of Theorems 3.2 and 3.4. In fact, there are many pairs satisfying the hypotheses of Theorems 3.2 and 3.4.
(2) The operator might not be bijective. Hence we should consider that the domain of the inverse operator is the range of .
(3) The operator is an homeomorphism from into , where is the range of .

4. Some Spectral Theorems for the Bounded Linear Operator

In this section we will apply some spectral theories to the . To do this, we need some concepts related to the spectral theory on a Banach space.

With we associate the operator where is a complex number and is the identity operator on . If has an inverse, we denote it by ; that is, and call it the resolvent operator of or, simply, the resolvent of .

Definition 4.1. A regular value of is a complex number such that (1) exists, (2) is bounded, (3) is defined on a set which is dense in . The resolvent set of is the set of all regular values of . Its complement is called the spectrum of , and a is called a spectral value of . For more details, see [16].

Remark 4.2. (1) The spectrum is partitioned into three disjoint sets as follows. (i)The point spectrum or discrete spectrum is the set such that does not exist. A is called an eigenvalue of . (ii)The continuous spectrum is the set such that exists and satisfies (3) but not (2) in Definition 4.1; that is to say, is unbounded. (iii)The residual spectrum is the set such that exists (and it may be bounded or not) but does not satisfy (3) in Definition 4.1. That is to say, the domain of is not dense in .
(2) We know that .

From now on, if what operator refers to is clear, we will write instead of .

In our next theorem, we apply the spectral theory to the operator .

Theorem 4.3. Let and be as in Theorem 3.4. Then the resolvent set of is open and hence the spectrum is closed. Furthermore, for every , the resolvent has the representation where the series is absolutely convergent for every in the open disk given by in the complex plane. This disk is a subset of .

Proof. Theorem 4.3 immediately follows the fact that the is a bounded linear operator on .

Next, we note that the spectral radius of is the radius of the smallest closed disk centered at the origin of the complex -plane and containing .

Theorem 4.4. Let and be as in Theorem 3.4. Then the spectrum of is compact and lies in the disk given by Hence the resolvent set of is not empty. Furthermore, the spectral radius and

Proof. From Theorem 3.4, the is a bounded linear operator on . Using a basic property for the spectrum, we establish (4.6), and hence the resolvent set is not empty. Furthermore, using (4.6), it is obvious that for the spectral radius of a bounded linear operator we have Also, we can easily obtain (4.7) as desired.

In our next theorem, we give a spectral mapping theorem for polynomials of . The proof of Theorem 4.5 is omitted because it immediately follows the spectral mapping theorem for polynomial on a Banach space.

Theorem 4.5. Let and be as in Theorem 3.4. Let Then This implies that the spectrum of the operator consists precisely of all those values which the polynomial assumes on the spectrum of .

Next, we will explain that our study is meaningful to obtain the solution to a differential equation.

Let be a real separable infinite-dimensional Hilbert space with the inner product and norm . Let be a measurable norm on with respect to the Gaussian cylinder set measure on . Let denote the completion of with respect to . Let denote the natural injection from to . The adjoint operator of is one to one and maps continuously onto a dense subset , where and are topological duals of and , respectively. By identifying with and with , we have a triple with for all in and in , where denotes the natural dual pairing between and . By a well-known result of Gross [17], has a unique countably additive extension to the Borel -algebra of . The triple is called an abstract Wiener space. The classical Wiener space is one of the examples of abstract Wiener space.

For an appropriate functional on , let be an operator defined by the formula where denotes the second Fréchet derivative and denotes the trace of an operator. In [2], Lee showed that the integral transform , forms the solution of a differential equation which is called a Cauchy problem where is an -dimensional polynomial function with respect to . In addition, let and in (4.13). Then the solution of the Cauchy problem is given by formula or, equivalently, where . This showed that the family of measures serves as the “fundamental solution” of the operator . For more details see [2, 18]. Hence our discussions as a bounded linear operator in this paper have some meaningful subjects. That is to say, from Theorems 3.2 through 4.5, if we take a complex number such that , then the transform is a well-defined bounded linear operator on and could be applied to all the results and formulas in this paper.

5. Applications for the Spectral Theory

In Sections 3 and 4, we treated the IT as a bounded linear operator on . Also, we applied spectral theorems to the IT to obtain various useful formulas and results. In this section we will show that the operator is self-adjoint under appropriate parameters and . We then apply the spectral theory to a self-adjoint operator on a Banach space. In particular, we obtain the spectral representation for IT .

In our next theorem, we show that the operator is a self-adjoint operator on under an appropriate condition for and .

Theorem 5.1. Let be as in Theorem 3.4 with and let . Then the operator is a self-adjoint operator on .

Proof. Let denote the adjoint operator on an operator . For all and in , we note that which completes the proof of Theorem 5.1 as desired.

Remark 5.2. We gave the conditions for and in Sections 2 and 3, and Theorem 5.1. We note that these conditions imply that is real and only. But, a self-adjoint operator may not have eigenvalues. So if it has eigenvalues, then it must be real. Hence these are very natural conditions.

Throughout the next corollary, we give some results of the spectral theories for self-adjoint operator .

Corollary 5.3. Let and be as in Theorem 5.1. Then we have the following assertions.(1)For , there exists a positive real number such that and so there exists a sequence in with , such that (2)For all and ,

In our next theorem, we apply the spectral theory to the operator as a self-adjoint operator.

Theorem 5.4. Let and be as in Theorem 5.1. Then the spectrum of is real and it lies in the closed interval where Furthermore, and are spectral values of and .

Proof. First, since is a self-adjoint operator on , the spectrum of must be real. Next we recall that for each bounded self-adjoint operator on a complex Hilbert space , on the real axis and , where and is an inner product on . Hence the spectrum lies in the closed interval . Furthermore, and are spectral values of and . Hence we complete the proof of Theorem 5.4.

We finish this paper by giving an application for the spectral representation of the self-adjoint operator which is one of very important subjects in the fields of the quantum mechanics and physical theories.

In our last theorem, we give the spectral representation for the self-adjoint operator. To do this, we need some concepts for the spectral theory.

Let be the set of eigenvalues of with for and let be the set of eigenfunctions corresponding to . Then we note that where . For each , define an (orthogonal) projection on by . Then we also note that Now, for , define an operator on by . In this case, is called the spectral family of .

Theorem 5.5. Let and be as in Theorem 5.1. Then we have the spectral representation for as follows: where and are given by (5.5) and (5.6), and is the spectral family of .

Proof. First, we note that for and in Theorem 5.1, is a self-adjoint operator on . Furthermore, . Using the spectral representation of the self-adjoint operator, we establish (5.10).

Remark 5.6. (1) In view of Theorem 5.5, for all real-valued continuous functions on and for and in , where and the integral is an ordinary Riemann-Stieltjes integral.
(2) An alternative formulation of the spectral theorem expresses the operator as an integral of the coordinate function over the operator's spectrum with respect to a projection-valued measure . When the normal operator in question is compact, this version of the spectral theorem reduces to the finite-dimensional spectral theorem, except that the operator is expressed as a linear combination of possibly infinitely many projections.
(3) In Sections 3, 4, and 5, we considered the IT as an operator. Like this, we expect that the convolution product could be dealt with as an operator. Furthermore, we could obtain various relations between the IT and the convolution product as a composition of operators.

Acknowledgments

The authors would like to express their gratitude to the referees for their valuable comments and suggestions which have improved the original paper. This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (2012R1A1A1004774:2011-0014552).