Review Special Issues

Spectrum and analytic functional calculus in real and quaternionic frameworks: An overview

  • An approach to the elementary spectral theory for quaternionic linear operators was presented by the author in a recent paper, quoted and discussed in the Introduction, where, unlike in works by other authors, the construction of the analytic functional calculus used a Riesz-Dunford-Gelfand type kernel, and the spectra were defined in the complex plane. In fact, the present author regards the quaternionic linear operators as a special class of real linear operators, a point of view leading to a simpler and a more natural approach to them. The author's main results in this framework are summarized in the following, and other pertinent comments and remarks are also included in this text. In addition, a quaternionic joint spectrum for pairs of operators is discussed, and an analytic functional calculus which uses a Martinelli type kernel in two variables is recalled.

    Citation: Florian-Horia Vasilescu. Spectrum and analytic functional calculus in real and quaternionic frameworks: An overview[J]. AIMS Mathematics, 2024, 9(1): 2326-2344. doi: 10.3934/math.2024115

    Related Papers:

    [1] Ai-qun Ma, Lin Chen, Zijie Qin . Jordan semi-triple derivations and Jordan centralizers on generalized quaternion algebras. AIMS Mathematics, 2023, 8(3): 6026-6035. doi: 10.3934/math.2023304
    [2] Fengxia Zhang, Ying Li, Jianli Zhao . A real representation method for special least squares solutions of the quaternion matrix equation $ (AXB, DXE) = (C, F) $. AIMS Mathematics, 2022, 7(8): 14595-14613. doi: 10.3934/math.2022803
    [3] Anli Wei, Ying Li, Wenxv Ding, Jianli Zhao . Three special kinds of least squares solutions for the quaternion generalized Sylvester matrix equation. AIMS Mathematics, 2022, 7(4): 5029-5048. doi: 10.3934/math.2022280
    [4] Wei Xu, Elvis Aponte, Ponraj Vasanthakumar . The property $ (\omega{ \pi }) $ as a generalization of the a-Weyl theorem. AIMS Mathematics, 2024, 9(9): 25646-25658. doi: 10.3934/math.20241253
    [5] Muhammad Tariq, Asif Ali Shaikh, Sotiris K. Ntouyas, Jessada Tariboon . Some novel refinements of Hermite-Hadamard and Pachpatte type integral inequalities involving a generalized preinvex function pertaining to Caputo-Fabrizio fractional integral operator. AIMS Mathematics, 2023, 8(11): 25572-25610. doi: 10.3934/math.20231306
    [6] Vladislav N. Kovalnogov, Ruslan V. Fedorov, Denis A. Demidov, Malyoshina A. Malyoshina, Theodore E. Simos, Vasilios N. Katsikis, Spyridon D. Mourtas, Romanos D. Sahas . Zeroing neural networks for computing quaternion linear matrix equation with application to color restoration of images. AIMS Mathematics, 2023, 8(6): 14321-14339. doi: 10.3934/math.2023733
    [7] Kahraman Esen Özen . A general method for solving linear matrix equations of elliptic biquaternions with applications. AIMS Mathematics, 2020, 5(3): 2211-2225. doi: 10.3934/math.2020146
    [8] Haiyan Zhou, K. A. Selvakumaran, S. Sivasubramanian, S. D. Purohit, Huo Tang . Subordination problems for a new class of Bazilevič functions associated with $ k $-symmetric points and fractional $ q $-calculus operators. AIMS Mathematics, 2021, 6(8): 8642-8653. doi: 10.3934/math.2021502
    [9] Dong Wang, Ying Li, Wenxv Ding . The least squares Bisymmetric solution of quaternion matrix equation $ AXB = C $. AIMS Mathematics, 2021, 6(12): 13247-13257. doi: 10.3934/math.2021766
    [10] Najat Almutairi, Sayed Saber . Chaos control and numerical solution of time-varying fractional Newton-Leipnik system using fractional Atangana-Baleanu derivatives. AIMS Mathematics, 2023, 8(11): 25863-25887. doi: 10.3934/math.20231319
  • An approach to the elementary spectral theory for quaternionic linear operators was presented by the author in a recent paper, quoted and discussed in the Introduction, where, unlike in works by other authors, the construction of the analytic functional calculus used a Riesz-Dunford-Gelfand type kernel, and the spectra were defined in the complex plane. In fact, the present author regards the quaternionic linear operators as a special class of real linear operators, a point of view leading to a simpler and a more natural approach to them. The author's main results in this framework are summarized in the following, and other pertinent comments and remarks are also included in this text. In addition, a quaternionic joint spectrum for pairs of operators is discussed, and an analytic functional calculus which uses a Martinelli type kernel in two variables is recalled.



    The aim of the present work is to exhibit an overview of the main results from the author's work [1], and partially from [2], adding some relevant comments and new remarks. In the article [1], in order to construct an analytic functional calculus for quaternionic linear operators, the class of the quaternionic slice regular functions (or slice holomorphic functions; see [3]), defined on subsets of the quaternionic algebra, is replaced by a class of vector-valued holomorphic functions, called stem functions (see Remark 2), defined on subsets of the complex plane. In fact, these two classes are isomorphic via a Cauchy type transform (see Theorem 4.2 below, or Theorem 6 from [2]), and we use the latter to construct an analytic functional calculus for what are called quaternionic linear operators. Nevertheless, the images of the analytic functional calculi obtained by these two methods are the same, as one might expect (see Remark 8).

    Our motivation comes from the following remark. Regarding the quaternions as elements of the real Hamilton algebra H, and imbedding H into its complexification M, the quaternions have a spectrum in the complex plane, and the quaternionic slice regular functions can be obtained via an analytic functional calculus with M-valued stem functions, which are holomorphic functions, symmetric with respect to the real axis, used in [2] (see also [13] for a more general definition). It is this idea which led us to shortcat the construction of the analytic functional for quaternionic operators, using directly M-valued analytic functions instead of quaternionic slice regular ones. At the same time, we have noticed that quaternionic linear operators are particular cases of real linear operators, allowing us to refine and adapt some techniques used in the study of the latter to get significant results valid for the former.

    The spectral theory for quaternionic linear operators has already been discussed in numerous works, in particular, in the monographs [3,4], where the construction of an analytic functional calculus (called S-analytic functional calculus) mounts to associating a quaternionic linear operator and a given function from the class of slice regular functions with another quaternionic linear operator, via an integral formula, using a certain non-commutative kernel. Unlike in these works, in the paper [1] one first considers the case of linear operators on real Banach spaces, whose spectrum is in the complex plane, and one sketches the construction of an analytic functional calculus for them, using some classical ideas (see Theorem 3.2 below). Then, regarding the quaternionic operators as particular cases of the real ones, this framework is extended to the quaternionic case, showing that the approach from the real case can be adapted to that more intricate situation. Unlike in [3] or [4], the functional calculus is obtained via a Riesz-Dunford-Gelfand formula, defined in a partially commutatative framework, rather than the non-commutative Cauchy type formula used by some previous authors. This is possible because the S-spectrum, introduced by F. Colombo and I. Sabadini (see [3]), can be replaced by a spectrum in the complex plane. As already mentioned above, one can show that the analytic functional calculus obtained with M-valued stem functions is equivalent to the analytic functional calculus obtained with slice holomorphic functions in [3] or [4], in the sense that the images of these functional calculi coincide, as shown in Remark 8. Unlike in the original paper [1], answering a question of the referee, we have added a complete proof of equality (4.9).

    The analytic functional calculus can actually be defined for a class of analytic operator valued functions, whose definition extends that of stem functions, and it applies, in particular, to a large family of quaternionic linear operators.

    In Subsection 4.2, we recall the construction of the quaternionic Cauchy transform from [2], showing later that the quaternionic Cauchy kernel from [3] can be obtained from the Riesz-Dunford kernel in M, via the Cauchy transform, which is a new result (see Example 3).

    The case of pairs of commuting real operators is also exhibited, following [1], and some connections with the quaternionic case are indicated. In fact, we define a quaternionic spectrum for them and construct an analytic functional calculus using a Martinelli type formula (see [5] for the original approach).

    If A is an arbitrary unital real Banach algebra (see [6] for details), the (complex) spectrum of an element aA may be defined by the equality

    σC(a)={u+iv;(ua)2+v2isnotinvertible,u,vR}, (2.1)

    where i is the imaginary unit.

    This definition goes back to Kaplansky (see [7]), and it applies in particular, to linear operators, acting on real vector spaces.

    The set σC(a) is conjugate symmetric, meaning that u+ivσC(a) if and only if uivσC(a).

    Fixing a unital real Banach algebra A, we denote by AC the complexification of A, that is, AC=CRA, written simply as A+iA, via the identification of the element 1a+ib with the element a+ib, for all a,bA.

    This tensor product AC is a unital complex Banach algebra, with the product of two elements given by (a+ib)(c+id)=acbd+i(ad+bc) for all a,b,c,dA, and with a (not necessarily unique) norm, which we fix to be given by a+ib=a+b, where is the norm of A. It is important to emphasize that in the algebra AC, the complex numbers commute with all elements of A.

    Another useful property of the algebra AC is a the existence of a conjugation, given by

    ACa+ibaibAC,a,bA,

    which is a unital conjugate-linear automorphism, whose square is the identity. Note also that an arbitrary element a+ib is invertible if and only if aib is invertible.

    The ordinary spectrum, defined for every element aAC, will be denoted by σ(a). Identifying the algebra A with a real subalgebra of AC via the map a1a, one gets the following.

    Lemma 1. For every aA we have the equality σC(a)=σ(a).

    Indeed, as we have the obvious identity

    (ua)2+v2=(u+iva)(uiva)

    for an arbitrary complex number u+iv with u,vR, the assertion follows easily.

    We will apply the discussion from above to the context of linear operators. In what follows, we consider real, complex and quaternionic linear operators, that is, R-, C- and H-linear operators, respectively.

    The spectral theory for real linear operators is not a new subject. A pertinent discussion exists actually in the framework of linear relations (see [8]). Moreover, the slightly different approach to real linear operators developed in [1], can be applied, with minor changes, to the case of some quaternionic operators.

    For a real or complex Banach space V, we denote by B(V) the algebra of all bounded R-(respectively C-)linear operators on V. As before, the multiples of the identity will be identified with the corresponding scalars.

    Let V be a real Banach space, and let VC=CRV be its complexification, which is identified with the direct sum V+iV. Every operator TB(V) has a unique extension to an operator TCB(VC), given by TC(x+iy)=Tx+iTy,x,yV and the map B(V)TTCB(VC) is unital, R-linear and multiplicative. Particularly, TB(V) is invertible if and only if TCB(VC) is invertible.

    Given an operator SB(VC), we define the operator SB(VC) to be equal to CSC, where C:VCVC is the (natural) conjugation of VC given by x+iyxiy,x,yV. The map B(VC)SSB(VC) is a unital conjugate-linear automorphism, whose square is the identity on B(VC). We have S=S if and only if S(V)V, and so TC=TC.

    Noting that (S+S)(V)V,i(SS)(V)V, we deduce that the algebras B(VC) and B(V)C are isomorphic and they will often be identified. So B(V) may be regarded as a (real) subalgebra of B(V)C. In particular, if S=U+iV, with U,VB(V), we have S=UiV, implying that the map SS is the conjugation of the complex algebra B(V)C induced by the conjugation C of VC.

    For every operator SB(VC), we denote by σ(S) its usual spectrum. As B(V) is a real algebra, the (complex) spectrum of an operator TB(V) is given by the equality (2.1):

    σC(T)={u+iv;(uT)2+v2isnotinvertible,u,vR}.

    Corollary 1. For every TB(V) we have the equality σC(T)=σ(TC).

    Using the concept of spectrum for real operators, an important step for further development is the construction of an analytic functional calculus (see [8]). As in [2], in what follows, we shall present a similar construction for real linear operators. Unlike in [8], we perform our construction using a class of operator valued analytic functions instead of scalar valued analytic functions. Moreover, this approach looks simpler for this case, and it is a model to get an analytic functional calculus for quaternionic linear operators.

    If V is a real Banach space, and so each operator TB(V) has a complex spectrum σC(T), which is compact and nonempty, one can use the classical Riesz-Dunford functional calculus, in a slightly generalized form (that is, replacing the scalar-valued analytic functions by operator-valued analytic ones, which is a well known idea difficult to track).

    To use various versions of the Cauchy formula, we adopt the following definition. Let UC be open. An open subset ΔU will be called a Cauchy domain (in U) if the boundary of Δ is in U and it consists of a finite family of closed curves, piecewise smooth, positively oriented. A Cauchy domain is bounded but not necessarily connected.

    Remark 1. If V is a real Banach space, and TB(V), we have the usual analytic functional calculus for the operator TCB(VC) (see [9]). That is, in a slightly generalized form, and for later use, if Uσ(TC) is an open set in C and F:UB(VC) is analytic, we set

    F(TC)=12πiΓF(ζ)(ζTC)1dζ,

    where Γ is the boundary of a Cauchy domain Δ containing σ(TC) in U. In fact, because σ(TC) is conjugate symmetric, we may and shall assume that both U and Γ are conjugate symmetric. Because the function ζF(ζ)(ζTC)1 is analytic in Uσ(TC), the integral does not depend on the particular choice of the Cauchy domain Δ containing σ(TC).

    A natural question is to find an appropriate condition to infer that F(TC)=F(TC), which would imply the invariance of V under F(TC).

    With the notation of Remark 1, we have the following.

    Theorem 1. Let UC be open and conjugate symmetric. If F:UB(VC) is analytic and F(ζ)=F(ˉζ) for all ζU, then F(TC)=F(TC) for all TB(V) with σC(T)U.

    This statement, as well as its proof, can be found in [1], as Theorem 1.

    Remark 2. If A is a unital real Banach algebra, AC its complexification, and UC is open, we denote by O(U,AC) the algebra of all analytic AC-valued functions. If U is conjugate symmetric, and ACaˉaAC is its natural conjugation, we denote by Os(U,AC) the real subalgebra of O(U,AC) consisting of those functions F with the property F(ˉζ)=¯F(ζ) for all ζU. Adapting a known terminology, such functions will be called (AC-valued ) stem functions (see also [10] or [13] for a more general definition).

    When A=R, so AC=C, the space Os(U,C) will be denoted by Os(U), which is a real algebra. Note that Os(U,AC) is also a bilateral Os(U)-module.

    Definition 1. For an operator TB(V) and for a given function FOs(U,B(V)C), we shall write

    F(T)=12πi(ΓF(ζ)(ζTC)1dζ)|V,

    and the map FF(T), will be called the (left) analytic functional calculus of T.

    We note that the right hand side of the integral from above belongs to B(V), by the previous theorem. The analytic functional calculus given by Definition 1 has the following properties.

    Theorem 2. Let V be a real Banach space, let UC be a conjugate symmetric open set, and let TB(V), with σC(T)U. Then the assignment

    Os(U,B(V)C)FF(T)B(V)

    is an R-linear map, and the assignment

    Os(U)ff(T)B(V)

    is a unital real algebra morphism.

    Moreover, the following properties are true:

    (1) For all FOs(U,B(V)C),fOs(U), we have (Ff)(T)=F(T)f(T), where (Ff)(ζ)=F(ζ)f(ζ) for all ζU.

    (2) For every polynomial P(ζ)=mn=0Anζn,ζC, with AnB(V) for all n=0,1,,m, we have P(T)=mn=0AnTnB(V).

    This result, and its proof as well, based on arguments that are more or less standard (see [9]), can be found in [1] as Theorem 2.

    Some known definitions and elementary facts about quaternions can be found in [3], Section 4.6.

    The abstract algebra of quaternions H is the four-dimensional R-algebra generated by the unit 1, and by the "imaginary units" {j,k,l}, satisfying

    jk=kj=l,kl=lk=j,lj=jl=k,jj=kk=ll=1.

    We have HR by identifying every number xR with the element x1H.

    The natural (multiplicative) norm on H is given by

    x=x20+x21+x22+x20,x=x0+x1j+x2k+x3l,x0,x1,x2,x3R,

    while the map

    Hx=x0+x1j+x2k+x3lx=x0x1jx2kx3lH

    is a natural involution.

    As we have xx=xx=x2, it follows, in particular, that every element xH{0} is invertible, and x1=x2x.

    For an arbitrary quaternion x=x0+x1j+x2k+x3l,x0,x1,x2,x3R, we set x=x0=(x+x)/2, and x=x1j+x2k+x3l=(xx)/2, that is, the real and imaginary part of x, respectively.

    The complexification CRH of the R-algebra H (appearing also in [10]) will be, as above, identified with the direct sum M=H+iH, which is an algebra containing the complex field C. Clearly, in the algebra M, the elements of H, in particular the "imaginary units" j,k,l, commute with all complex numbers.

    The natural conjugation in the algebra M is given by ˉa=bic, where a=b+ic is arbitrary in M, with b,cH (see also [10]). Note that ¯a+b=ˉa+ˉb, and ¯ab=ˉaˉb, ¯ra=rˉa in particular, for all a,bM and rR. In addition, ˉa=a if and only if aH, which characterizes the elements of H among those of M.

    Remark 3. In the algebra M we have the identities

    (λx)(λx)=(λx)(λx)=λ2λ(x+x)+x2C,

    for all λC and xH. If the complex number λ22λx+x2 is nonnull, then both elements λx,λx are invertible. Conversely, if λx is invertible, we must have λ22λx+x2 nonnull. In this way, the spectrum of a quaternion xH is given by the equality σ(x)={s±(x)}, where s±(x)=x±ix are the eigenvalues of x (see also [1,2]).

    We also need the concept of resolvent set of a quaternion xH, denoted by ρ(x), and given by Cσ(x).

    The polynomial Px(λ)=λ22λx+x2 is the minimal polynomial of x. In fact, the equality σ(y)=σ(x) for some x,yH is an equivalence relation in the algebra H, which holds if and only if Px=Py.

    Let S={s=x1j+x2k+x3l;x1,x2,x3R,x21+x22+x23=1}, that is, the unit sphere of "purely imaginary" quaternions. It is clear that s=s, and so s2=1,s1=s, and s=1 for all sS.

    Representing an arbitrary quaternion x under the form x0+y0κ0, with x0,y0R and κ0S, a quaternion y is equivalent to x if and only if it is of the form x0+y0κ for some κS (see [2] for some details).

    Remark 4. As in [2], a subset ΩH is said to be spectrally saturated (called axially symmetric in [3]) if for every xH with σ(x)=σ(q) for some qΩ, we also have xΩ. For a conjugate symmetric set UC we put UH={qH;σ(q)U}, which is a spectrally saturated set. Conversely, if ΩH is a spectrally saturated set, we put S(Ω)={ζC;qΩ,ζσ(q)}, which is a conjugate symmetric set in C. If U=S(Ω), we have UH=Ω and S(Ω)=U. Moreover, U is open in C if and only if UH is open in H.

    For M-valued functions defined on subsets of H, the concept of slice regularity is defined as follows.

    Let ΩH be an open set, and let F:ΩM be a differentiable function. Following [3], Definition 4.1.1 (see also [13] for a more general context), we say that F is slice right regular on Ω if for all sS,

    ˉsF(x+ys):=12(x+Rsy)F(x+ys)=0,

    on the set Ω(R+Rs), where Rs is the right multiplication of the elements of M by s.

    A slice left regularity can also be defined via the left multiplication of the elements of M by elements from S. In what follows, we use only the slice right regularity, which will be simply called slice regularity. See also [4,10,11,12] for other connections in the "slice" context.

    Let us note that the convergent series of the form k0akqk, on quaternionic balls {qH;q<r}, with r>0 and akH for all k0, are H-valued slice regular on their domain of definition. In fact, if actually akM, such functions are (M-valued) slice regular on their domain of definition.

    Following [2], Definition 1, we consider the following concept.

    Definition 2. The M-valued Cauchy kernel on the open set ΩH is given by

    ρ(q)×Ω(ζ,q)(ζq)1M. (4.1)

    As shown in [2], Example 2, the M-valued Cauchy kernel on the open set ΩH is slice regular.

    Using the M-valued Cauchy kernel, we may define a concept of quaternionic Cauchy transform (see [2], Section 5).

    As before, for a given open set UC, we denote by O(U,M) the complex algebra of all M-valued analytic functions on U. If UC is open and conjugate symmetric, let Os(U,M) be the real subalgebra of O(U,M) consisting of all M-valued stem functions from O(U,M).

    Because CM, we have O(U)O(U,M), where O(U) is the complex algebra of all complex-valued analytic functions on the open set U. Similarly, when UC is open and conjugate symmetric, Os(U)Os(U,M), where Os(U) is the real subalgebra consisting of all functions f from O(U) which are complex stem functions.

    For example, if ΔC is an open disk centered at 0, each function FOs(Δ,M) can be represented as a convergent series F(ζ)=k0akζk,ζΔ, with akH for all k0.

    Definition 3. Let UC be a conjugate symmetric open set, and let FO(U,M). For every qUH we set

    C[F](q)=12πiΓF(ζ)(ζq)1dζ, (4.2)

    where Γ is the boundary of a Cauchy domain in U containing the spectrum σ(q). The function C[F]:UHM is called the (quaternionic) Cauchy transform of the function FO(U,M). Clearly, the function C[F] does not depend on the choice of Γ because the function

    Uσ(q)ζF(ζ)(ζq)1M

    is analytic.

    We put

    R(UH,M)={C[F];FO(U,M)}. (4.3)

    Proposition 1. Let UC be open and conjugate symmetric, and let FO(U,M). Then function C[F]R(UH,M) is slice regular on UH.

    The proof of this assertion can be found in [2], as Proposition 1.

    A condition insuring that the Cauchy transform is actually H-valued is obtained from the following result, proved as Theorem 4 in [2].

    Theorem 3. Let UC be open and conjugate symmetric, and let F be in O(U,M). The Cauchy transform C[F] is H-valued if and only if F belongs to Os(U,M).

    If ΩH be a spectrally saturated open set, we put

    Rs(Ω,H)={C[F];FOs(U,M)}, (4.4)

    where U=S(Ω).

    Theorem 4. Let ΩH be a spectrally saturated open set, and let Φ:ΩH. The function Φ is slice regular if and only if we have ΦRs(Ω,H), with U=S(Ω). Moreover, the assignment

    Os(U,M)FC[F]Rs(Ω,H)

    is a linear isomorphism.

    This statement follows from Theorems 5 and 6 from [2].

    Remark 5. Following [3], a right H-vector space V is a real vector space having a right multiplication with the elements of H, such that (x+y)q=xq+yq,x(q+s)=xq+xs,x(qs)=(xq)s for all x,yV and q,sH.

    If V is also a real Banach space, the operator TB(V) is right H-linear if T(xq)=T(x)q for all xV and qH. The set of right H linear operators will be denoted by Br(V), which is, in particular, a unital real algebra.

    In a similar way, one defines the concept of a left H-vector space. A real vector space V will be said to be an H-vector space if it is simultaneously a right H- and a left H-vector space. As noticed in [3], the framework of H-vector spaces is an appropriate one for the study of right H-linear operators.

    If V is an H-vector space which is also a real Banach space, then V is said to be a Banach H-space. In this case, we also assume that RqB(V), and the map HqRqB(V) is norm continuous, where Rq is the right multiplication of the elements of V by a given quaternion qH. Similarly, if Lq is the left multiplication of the elements of V by the quaternion qH, we assume that LqB(V) for all qH, and that the map HqLqB(V) is norm continuous. Note also that

    Br(V)={TB(V);TRq=RqT,qH}.

    To adapt the discussion regarding the real algebras to this case, we first consider the complexification VC of V. Because V is an H-bimodule, the space VC is actually an M-bimodule, via the multiplications

    (q+is)(x+iy)=qxsy+i(qy+sx),(x+iy)(q+is)=xqys+i(yq+xs),

    for all q+isM,q,sH,x+iyVC,x,yV. Moreover, the operator TC is right M-linear, that is TC((x+iy)(q+is))=TC(x+iy)(q+is) for all q+isM,x+iyVC, via a direct computation.

    Let C be the natural conjugation of VC. As in the real case, for every SB(VC), we put S=CSC. The left and right multiplication with the quaternion q on VC will also be denoted by Lq,Rq, respectively, as elements of B(VC). We set

    Br(VC)={SB(VC);SRq=RqS,qH},

    which is a unital complex algebra containing all operators Lq,qH. Note that if SBr(VC), then SBr(VC). Indeed, because CRq=RqC, we also have SRq=RqS. In fact, as we have (S+S)(V)V and i(SS)(V)V, it folows that the algebras Br(VC),Br(V)C are isomorphic, and they will often be identified, where Br(V)C=Br(V)+iBr(V) is the complexification of Br(V), which is also a unital complex Banach algebra.

    Inspired by the Definition 4.8.1 from [3] (see also [4]), we used in [1] the following.

    Definition 4. For a given operator TBr(V), the set

    σH(T):={qH;T22(q)T+q2)notinvertible}

    is called the quaternionic spectrum (or simply the Q-spectrum) of T.

    The complement ρH(T)=HσH(T) is called the quaternionic resolvent (or simply the Q-resolvent) of T.

    We decided to call Q-spectrum the concept given by previous definition rather than S-spectrum (as in [3], and earlier introduced by F. Colombo and I. Sabadini), to stress its connection with the quaternionic algebra. In fact, in the construction of the analytic functional calculus for quaternionic linear operators, and unlike in [3,4], this spectrum will be replaced by a strongly related one (see Lemma 2), nevertheless defined in the complex plane (see Theorem 7).

    Note that, if qσH(T)), then {sH;σ(s)=σ(q)}σH(T). In other words, the spectrum σH(T) is a spectrally saturated set.

    Assuming that V is a Banach H-space, then Br(V) is a unital real Banach H-algebra (that is, a Banach algebra which also a Banach H-space), via the algebraic operations (qT)(x)=qT(x), and (Tq)(x)=T(qx) for all qH and xV. Hence the complexification Br(V)C is, in particular, a unital complex Banach algebra. Also note that the complex numbers, regarded as elements of Br(V)C, commute with the elements of Br(V). For this reason, for each TBr(V) we have the resolvent set

    ρC(T)={λC;(T22(λ)T+|λ|2)1Br(V)}=
    {λC;(λTC)1Br(VC)}=ρ(TC),

    and the associated spectrum σC(T)=σ(TC).

    One can see that there exists a strong connection between σH(T) and σC(T). Specifically, one can prove the following.

    Lemma 2. For every TBr(V) we have the equalities

    σH(T)={qH;σC(T)σ(q)}, (4.5)

    and

    σC(T)={λσ(q);qσH(T)}. (4.6)

    This statement, and its proof as well, can be found in [1] as Lemma 2. Moreover, as in [14], Remark 14, we can also prove that

    σH(T)={(λ)+|(λ)|;λσC(T),sS}.

    Remark. As expected, the set σH(T) is nonempty and bounded, which follows easily from Lemma 2. It is also compact, as a consequence of Definition 4, because the set of invertible elements in Br(V) is open.

    If V is a real Banach H-space, because Br(V) is also a real Banach space, each operator TBr(V) has a complex spectrum σC(T). Therefore, applying the corresponding result for real operators, we may construct an analytic functional calculus using the classical Riesz-Dunford functional calculus, in a slightly generalized form, as done in Theorem 2. In this new case, our basic complex algebra is Br(V)C, endowed with the conjugation Br(V)CSSBr(V)C.

    Theorem 5. Let UC be open and conjugate symmetric. If F:UBr(VC) is analytic and F(ζ)=F(ˉζ) for all ζU, then F(TC)=F(TC) for all TBr(V) with σC(T)U.

    This assertion coincides with that of Theorem 3 from [1].

    As in the real case, we may identify the algebra Br(V) with a subalgebra of Br(V)C. So, when FOs(U,Br(V)C)={FO(U,Br(V)C);F(ˉζ)=F(ζ)ζU} (see also Remark 2), we can write, via Theorem 5, that

    F(T)=12πiΓF(ζ)(ζT)1dζBr(V),

    for a suitable choice of Γ.

    The next result provides an analytic functional calculus for operators from the real algebra Br(V).

    Theorem 6. Let V be a real Banach H-space, let UC be a conjugate symmetric open set, and let TBr(V), with σC(T)U. Then, the map

    Os(U,Br(V)C)FF(T)Br(V)

    is R-linear, and the map

    Os(U)ff(T)Br(V)

    is a unital real algebra morphism.

    Moreover, the following properties are true:

    (1) For all FOs(U,Br(V)C),fOs(U), we have (Ff)(T)=F(T)f(T), where (Ff)(ζ)=F(ζ)f(ζ) for all ζU.

    (2) For every polynomial P(ζ)=mn=0Anζn,ζC, with AnBr(V) for all n=0,1,,m, we have P(T)=mn=0AnTnBr(V).

    This statement can be found in [1] as Theorem 4.

    Remark 6. That Theorems 5 and 6 have practically the same proofs as Theorems 1 and 2 (respectively) is due to the fact that all of them can be obtained as particular cases of more general results. Indeed, considering a unital real Banach algebra A, and its complexification AC, identifying A with a real subalgebra of AC, for a function FOs(U,AC), where UC is open and conjugate symmetric, the element F(b)A for each bA with σC(b)U. The assertion follows as in the proof of Theorem 1. The other results also have their counterparts. We omit the details.

    Remark 7. The algebra H is clearly a Banach H-space. Moreover, the left multiplications Lq,qH, are elements of Br(H), and the map HqLqBr(H) is an injective morphism of real algebras allowing the identification of H with a subalgebra of Br(H).

    If ΩH is a spectrally saturated open set, let U=S(Ω). Because Os(U)Os(U,M), we must have

    Rs(Ω):={C[f];fOs(U)}Rs(Ω,H).

    The next theorem, which is an analytic functional calculus for quaternions (see [2], Theorem 5), can be obtained as a particular case of Theorem 6.

    Theorem 7. Let ΩH be a spectrally saturated open set, and let U=S(Ω). The space Rs(Ω) is a unital commutative real algebra, the space Rs(Ω,H) is a right Rs(Ω)-module, the map

    Os(U,M)FC[F]Rs(Ω,H)

    is a right module isomorphism, and its restriction

    Os(U)fC[f]Rs(Ω)

    is a real algebra isomorphism.

    Moreover, for every polynomial P(ζ)=mn=0anζn,ζC, with anH for all n=0,1,,m, we have C[P](q)=mn=0anqnH for all qH.

    For other details concerning this assertion see Theorem 5 from [1].

    Remark 8. It is shown in [2], Theorem 6, that the space Rs(Ω,H) coincides with the space of all H-valued functions which are slice regular, independently defined in [3], Definition 4.1.1. Such functions are used in [3] to define a quaternionic functional calculus for quaternionic linear operators (see also [4]). This construction is largely explained in the fourth chapter of [3].

    Our Theorem 6 constructs an analytic functional calculus with functions from Os(U,Br(V)C), where U is a a neighborhood of the complex spectrum of a given quaternionic linear operator, leading to another quaternionic linear operator, replacing formally the complex variable with that operator.

    Identifying the algebra H with a subalgebra of Br(V)C by regarding the elements of H as left multiplication operators, we can show that the functional calculus from [3] is equivalent to our functional calculus with functions from Os(U,M) in the sense that they have the same images. This is a consequence of the fact that the class of regular quaternionic-valued functions Rs(Ω,H) used by the construction in [3] is isomorphic to the class of analytic functions Os(U,M), via Theorem 7. Our approach looks simpler and there is a stronger connection with the classical approach, because we use spectra defined in the complex plane, and our Cauchy type kernels are partially commutative.

    As in [1], Remark 8, a direct argument concerning the equivalence of those analytic functional calculi can be given. For an operator TBr(V), the right S-resolvent is defined via the formula

    S1R(s,T)=(Ts)(T22(s)T+s)1,sρH(T) (4.7)

    (see [3], formula (4.47)). Fixing an element κS, and a spectrally saturated open set ΩH, for ΦRs(Ω,H) one puts

    Φ(T)=12π(Σκ)Φ(s)dsκS1R(s,T), (4.8)

    where ΣΩ is a spectrally saturated open set containing σH(T), such that Σκ={u+vκΣ;u,vR} is a subset whose boundary (Σκ) consists of a finite family of closed curves, piecewise smooth, positively oriented, and dsκ=κdudv. Formula (4.8) is a (right) quaternionic functional calculus, as defined in [3], Section 4.10.

    Because the space VC is also an H-space, we may extend these formulas to the operator TCBr(VC), replacing the operator T by TC in formulas (4.7) and (4.8). For the function ΦRs(Ω,H) there exists a function FOs(U,M) such that C[F]=Φ. Denoting by Γκ the boundary of a Cauchy domain in UC containing the compact set {σ(s);s¯Σκ}, we have

    Φ(TC)=12π(Σκ)(12πiΓκF(ζ)(ζs)1dζ)dsκS1R(s,TC)=12πiΓκF(ζ)(12π(Σκ)(ζs)1dsκS1R(s,TC))dζ. (4.9)

    The intertwining of the integral in formula (4.9) can be explained in the following way. We consider the parametrizations ϕ:[0,1]C and ψ:[0,1]Cκ:={(a+bκ;a,bR} of the curves Γκ and (Σκ) respectively, having continuous derivatives except a finite number of points. We also put ψκ=κψ. Note that these functions commute, because, in our framework, complex numbers and quaternions commute. As dζ=ϕ(u)du,dsκ=ψκ(v)dv,u,v[0,1], we have

    12π(Σκ)(12πiΓκF(ζ)(ζs)1dζ)dsκS1R(s,TC)=12π10(12πi10F(ϕ(u)))(ϕ(u)ψ(v))1ϕ(u)du)ψκ(v)S1R(ψ(v),TC)dv=12π10(12πi10F(ϕ(u)))(ϕ(u)ψ(v))1ϕ(u)ψκ(v)S1R(ψ(v),TC)du)dv=12π10(12πi10F(ϕ(u)))(ϕ(u)ψ(v))1ψκ(v)S1R(ψ(v),TC)ϕ(u)du)dv=12πi10(12π10F(ϕ(u)))(ϕ(u)ψ(v))1ψκ(v)S1R(ψ(v),TC)dv)ϕ(u)du=12πiΓκF(ζ)(12π(Σκ)(ζs)1dsκS1R(s,TC))dζ,

    via Fubini's theorem. Then we have

    12πiΓκF(ζ)(12π(Σκ)(ζs)1dsκS1R(s,TC))dζ12πiΓκF(ζ)(ζTC)1dζ. (4.10)

    To obtain formula (4.10), we use an argument from [1], Remark 9. Specifically, it follows from the complex linearity of S1R(s,TC), and from formula (4.49) in [3], that

    (ζs)S1R(s,TC)=S1R(s,TC)(ζTC)1,

    whence

    (ζs)1S1R(s,TC)=S1R(s,TC)(ζTC)1+(ζs)1(ζTC)1,

    and therefore,

    12π(Σκ)(ζs)1dsκS1R(s,TC)=12π(Σκ)dsκS1R(s,TC)(ζTC)1+12π(Σκ)(ζs)1dsκ(ζTC)1=(ζTC)1,

    because

    12π(Σκ)dsκS1R(s,TC)=1and12π(Σκ)(ζs)1dsκ=0,

    as in Theorem 4.8.11 from [3], since the M-valued function s(ζs)1 is analytic in a neighborhood of the set ¯ΣκCκ for each ζΓκ, respectively.

    Consequently, Φ(T)=Φ(TC)|V=F(TC)|V=F(T)

    In particular, regarding a quaternion qH as a left multiplication operator, formula (4.8), written as

    Φ(q)=12π(Σκ)Φ(s)dsκS1R(s,q), (4.11)

    is formula (2.39) from [3] in the quaternionic case, via Proposition 2.7.20, with a more direct proof.

    Remark 9. Our approach permits to prove a version of the spectral mapping theorem in a classical style, via direct arguments (see also [3,4] for a different context). For every operator TBr(V) and each function ΦRs(Ω) one has σH(Φ(T))=Φ(σH(T)), via Theorem 3.5.9 from [3]. Using our approach, for every function fOs(U), one has f(σC(T))=σC(f(T)), directly from the corresponding (classical) spectral mapping theorem in [9]. This result is parallel to that from [3] mentioned above, also giving an explanation for the former, via the isomorphism of the spaces Os(U) and Rs(Ω) by Theorem 7.

    Example 1. This is Example 2 from [1]. The space V=H is itself a (simple) Banach H-space. Because VC=M, for a fixed element qH, we may consider the operator LqBr(H), whose complex spectrum is given by σC(Lq)=σ(q)={q±iq}. If UC is a conjugate symmetric open set containing σC(Lq), and FOs(U,M), then we have

    F(Lq)=F(s+(q))ι+(s˜q)+F(s(q))ι(s˜q)M, (5.1)

    where we have s±(q)=q±iq, ˜q=q,s˜q=˜q˜q1, and ι±(s˜q)=21(1is˜q), provided ˜q0, via [2], Remark 3. The case ˜q=0 is trivial.

    Of course, this formula can be extended to a larger class of functions.

    Example 2. This is Example 3 from [1]. Let C(X,M) be the space of M-valued continuous functions on the compact space X. Then the space C(X,H), consisting of H-valued functions, is the real subspace of C(X,M), which is also a Banach H-space with respect to the operations (qF)(x)=qF(x) and (Fq)(x)=F(x)q for all FC(X,H) and xX. Moreover, C(X,H)C=C(X,HC)=C(X,M).

    Fixing a function ΘC(X,H), we define the operator TB(C(X,H)) by the equality (TF)(x)=Θ(x)F(x) for all FC(X,H) and xX. Note that (T(Fq))(x)=Θ(x)F(x)q=((TF)q)(x) for all FC(X,H),qH, and xX. In other words, TBr(C(X,H)). Note also that the operator T is invertible if and only if the function Θ has no zero in X.

    According to Definition 4, we have

    ρH(T)={qH;(T22qT+q2)1Br(C(X,H))}.

    Consequently, qσH(T) if and only if zero is in the range of the function

    τ(q,x):=Θ(x)22qΘ(x)+q2,xX.

    Similarly,

    ρC(T)={λC;(T22λT+λ2)1Br(C(X,H))},

    and so λσC(T) if and only if zero is in the range of the function

    τ(λ,x):=Θ(x)22λΘ(x)+|λ|2,xX.

    Looking for solutions u+iv,u,vR, of the equation (uΘ(x))2+v2=0, a direct calculation shows that u=Θ(x) and v=±Θ(x). Hence,

    σC(T)={Θ(x)±iΘ(x);xX}=xXσ(Θ(x)).

    Moreover, for every open conjugate symmetric subset UC containing σC(T), and for every function ΦOc(U,Br(C(X,M))), we may construct the operator Φ(T)Br(C(X,H)), via Theorem 6.

    Example 3. Now, we show that the non-commutative Cauchy kernel from [3] is the Cauchy transform of the complex Cauchy kernel associated to a quaternion.

    Let s,qH with σ(s)σ(q)=, and so sq. In particular, the quaternion s22(q)s+q2 is invertible. Indeed, if ζ=q+iqσ(q), we have in M

    s22(q)s+q2=(sζ)(sˉζ),

    and both sζ,sˉζ are invertible because ζσ(s).

    Let us consider the function

    S1R(q,s)=(sq)(s22(q)s+q2)1,

    which is the right noncommutative Cauchy kernel, as defined in [3], as formula (2.33).

    Note also that the function ρ(q)ζ(ζq)1M is in the space Os(ρ(q),M).

    We can show the equality

    S1R(q,s)=12πiΓs(ζq)1(ζs)1dζ,

    where Γs surrounds a Cauchy domain containing σ(s), whose closure is disjoint of σ(q). Indeed,

    12πiΓs(ζq)1(ζs)1dζ=
    12πiΓs[(ζq)(ζq)]1(ζq)(ζs)1dζ=
    12πiΓs(ζ22ζ(q)+q2)1(ζq)(ζs)1dζ=
    (sq)(s22(q)s+q2)1,

    showing that the kernel S1R(s,q) is the Cauchy transform of the function ρ(q)ζ(ζq)1M.

    Sometimes, especially in applications, it is more convenient to work with matrix quaternions rather than with abstract quaternions. Specifically, one considers the injective unital algebra morphism

    Hx1+y1j+x2k+y2l(x1+iy1x2+iy2x2+iy2x1iy1)M2,

    with x1,y1,x2,y2R, where M2 is the complex algebra of 2×2 matrices, whose image, denoted by H2 is the real algebra of matrix quaternions. The elements of H2 can be also written as matrices of the form

    Q(z)=(z1z2ˉz2¯z1),z=(z1,z2)C2.

    A natural connection between the spectral theory of pairs of commuting operators in a complex Hilbert space and the algebra of quaternions has firstly been noticed in [15]. Another connection will be presented in the following.

    If V is an arbitrary vector space, we denote by V2 the Cartesian product V×V.

    Let V be a real Banach space, and let T=(T1,T2)B(V)2 be a pair of commuting operators. The extended pair TC=(T1C,T2C)B(VC)2 also consists of commuting operators. For simplicity, we set

    Q(TC):=(T1CT2CT2CT1C)

    which acts on the complex Banach space V2C.

    One can define the quaternionic resolvent set and spectrum for the case of a pair of operators (see [1], Definition 2), inspired by the case of a single operator (see [15]).

    Definition 5. Let V be a real Banach space. For a given pair T=(T1,T2)B(V)2 of commuting operators, the set of those Q(z)H2,z=(z1,z2)C2, such that the operator

    T21+T222(z1)T12(z2)T2+|z1|2+|z2|2

    is invertible in B(V) is said to be the quaternionic joint resolvent (or simply the Q-joint resolvent) of T, and it is denoted by ρH(T).

    The complement σH(T)=H2ρH(T) is called the quaternionic joint spectrum (or simply the Q-joint spectrum) of T.

    For every pair TC=(T1C,T2C)B(VC)2 we put TcC=(T1C,T2C)B(VC)2, and for every pair z=(z1,z2)C2 we put zc=(ˉz1,z2)C2.

    Lemma 3. A matrix quaternion Q(z) (zC2) is in the set ρH(T) if and only if the operators Q(TC)Q(z),Q(TcC)Q(zc) are invertible in B(V2C).

    The complete proof of this assertion can be found in [1], as Lemma 3, and it is based on the equalities

    (T1Cz1T2Cz2T2C+ˉz2T1Cˉz1)(T1Cˉz1T2C+z2T2Cˉz2T1Cz1)=
    (T1Cˉz1T2C+z2T2Cˉz2T1Cz1)(T1Cz1T2Cz2T2C+ˉz2T1Cˉz1)=
    [(T1Cz1)(T1Cˉz1)+(T2Cz2)(T2Cˉz2)]I,

    for all z=(z1,z2)C2, where I is the identity of B(VC)2.

    Lemma 3 shows that we have the property Q(z)σH(T) if and only if Q(zc)σH(Tc). Putting

    σC2(T):={zC2;Q(z)σH(T)},

    the set σC2(T) has a similar property, specifically, zσC2(T) if and only if zcσC2(Tc).

    Remark 10. The extended pair TC=(T1C,T2C)B(VC)2 of the commuting pair T=(T1,T2)B(V) can be discussed in connection with the joint spectral theory of J. L. Taylor (see [16,17]; see also [19]). Namely, if the operator T21C+T22C2(z1)T1C2(z2)T2C+|z1|2+|z2|2 is invertible, then the point z=(z1,z2) belongs to the joint resolvent of TC (see Remark 9 from [1]). We also have that the point zc belongs to the joint resolvent of TcC. In addition, if σ(TC) designates the Taylor spectrum of TC, we have the inclusion σ(TC)σC2(T). In particular, for every complex-valued function f analytic in a neighborhood of σC2(T), the operator f(TC) can be computed via Taylor's analytic functional calculus. In fact, we have a Martinelli type formula for the analytic functional calculus:

    Theorem 8. Let V be a real Banach space, let T=(T1,T2)B(V)2 be a pair of commuting operators, let UC2 be an open set, let DU be a bounded domain containing σC2(T), with piecewise-smooth boundary Σ, and let fO(U). Then, we have

    f(TC)=1(2πi)2Σf(z)L(z,TC)2[(ˉz1T1C)dˉz2(ˉz2T2C)dˉz1]dz1dz2

    where

    L(z,TC)=T21C+T22C2(z1)T1C2(z2)T2C+|z1|2+|z2|2.

    The proof of this result can be found in [1], Theorem 6.

    Remark 11. (1) The previous functional calculus can be extended to B(VC)-valued analytic functions, setting, for such a function F, and with the notation from above,

    F(TC)=1(2πi)2ΣF(z)L(z,TC)2[(ˉz1T1C)dˉz2(ˉz2T2C)dˉz1]dz1dz2

    (see [1] Remark 10(1)).

    In particular, if F(z)=j,k0AjkCzj1zk2, with Aj,kB(V), where the series is convergent in a neighbourhood of σC2(T), we obtain

    F(T):=F(TC)|V=j,k0AjkTj1Tk2B(V).

    (2) In the case of Hilbert spaces, there is a stronger connection between the spectral theory of pairs and the algebra of quaternions (see [1], Remark 10(2)). Specifically, if H is a complex Hilbert space and V=(V1,V2) is a commuting pair of bounded linear operators on H, a point z=(z1,z2)C2 is in the joint resolvent of V if and only if the operator Q(V)Q(z) is invertible in H2, where

    Q(V)=(V1V2V2V1),

    (see also [15] for other details). In this case, there is also a Martinelli type formula which can be used to construct the associated analytic functional calculus (see [18,19]). An approach to such a construction in Banach spaces, by using a so-called splitting joint spectrum, can be found in [20].

    A parallel theory valid for operators acting in a Cliffordian context is presented in the article [14] (see also [21]). Inspired by this general approach, another contribution of the present author, developing the spectral decompositions of normal operators in a quaternionic setting by extending results from the real context appears in a work published in arXiv: 2103.16266v2.

    The approach to the elementary spectral theory for the class of quaternionic linear operators regarded as a special class of real linear ones leads to a simpler and more natural construction of the analytic functional calculus, avoiding many of the difficulties which appear when working with the spectrum defined in the quaternionic algebra. This analytic functional calculus with stem functions defined in the complex plane replaces the analytic functional calculus with slice holomorphicc functions, which are defined in the quaternionic algebra.

    The author declares he has not used Artificial Intelligence (AI) tools in the creation of this article.

    The author is indebted to the referee for several useful remarks, which improved the original version of this work.

    The author declares no conflict of interest in this paper



    [1] F. H. Vasilescu, Spectrum and analytic functional calculus in real and quaternionic frameworks, Pure Appl. Funct. Anal., 7 (2022), 389–407.
    [2] F. H. Vasilescu, Quaternionic regularity via analytic functional calculus, Integr. Equat. Oper. Th., 92 (2020), 18. http://dx.doi.org/10.1007/s00020-020-2574-7 doi: 10.1007/s00020-020-2574-7
    [3] F. Colombo, I. Sabadini, D. C. Struppa, Noncommutative functional calculus, theory and applications of slice hyperholomorphic functions, In: Progress in Mathematics, Birkhäuser/Springer Basel AG, Basel, 28 (2011). http://dx.doi.org/10.1007/978-3-0348-0110-2
    [4] F. Colombo, J. Gantner, D. P. Kimsey, Spectral theory on the S-spectrum for quaternionic operators, Birkhäuser, 2018. https://doi.org/10.1007/978-3-030-03074-2-3
    [5] E. Martinelli, Alcuni teoremi integrali per le funzioni analitiche di più variabili complesse, Accad. Ital. Mem. Cl. Sci. Fis. Mat. Nat., 9 (1938), 269–283.
    [6] L. Ingelstam, Real Banach algebras, Ark. Mat., 5 (1964), 239–270. https://doi.org/10.1007/BF02591126 doi: 10.1007/BF02591126
    [7] I. Kaplansky, Normed algebras, Duke Math. J., 16 (1949), 399–418. https://doi.org/10.1215/S0012-7094-49-01640-3 doi: 10.1215/S0012-7094-49-01640-3
    [8] A. G. Baskakov, A. S. Zagorskii, Spectral theory of linear relations on real Banach spaces, Math. Notes, 81 (2007), 15–27. https://doi.org/10.1134/S0001434607010026 doi: 10.1134/S0001434607010026
    [9] N. Dunford, J. T. Schwartz, Linear operators, part I: General theory, Interscience Publishers, New York, London, 1958.
    [10] R. Ghiloni, A. Perotti, Slice regular functions on real alternative algebras, Adv. Math., 226 (2011), 1662–1691. https://doi.org/10.1016/j.aim.2010.08.015 doi: 10.1016/j.aim.2010.08.015
    [11] G. Gentili, D. C. Struppa, A new theory of regular functions of a quaternionic variable, Adv. Math., 216 (2007), 279–301. https://doi.org/10.1016/j.aim.2007.05.010 doi: 10.1016/j.aim.2007.05.010
    [12] R. Ghiloni, V. Moretti, A. Perotti, Continuous slice functional calculus in quaternionic Hilbert spaces, Rev. Math. Phys., 25 (2013), 83. https://doi.org/10.1142/S0129055X13500062 doi: 10.1142/S0129055X13500062
    [13] R. Ghiloni, V. Recupero, Slice regular semigroups, Trans. Amer. Math. Soc., 370 (2018). https://doi.org/10.1090/tran/7354 doi: 10.1090/tran/7354
    [14] F. H. Vasilescu, Spectrum and analytic functional calculus for Clifford operators via stem functions Concr. Oper., 8 (2021), 90–113. https://doi.org/10.1515/conop-2020-0115 LicenseCC BY 4.0
    [15] F. H. Vasilescu, On pairs of commuting operators, Stud. Math., 62 (1978), 203–207.
    [16] J. L. Taylor, A joint spectrum for several commuting operators, J. Funct. Anal., 6 (1970), 172–191. https://doi.org/10.1016/0022-1236(70)90055-8 doi: 10.1016/0022-1236(70)90055-8
    [17] J. L. Taylor, The analytic functional calculus for several commuting operators, Acta Math., 125 (1970), 1–38. https://doi.org/10.1007/BF02392329 doi: 10.1007/BF02392329
    [18] F. H. Vasilescu, A Martinelli type formula for the analytic functional calculus, Rev. Roum. Math. Pures Appl., 23 (1978), 1587–1605.
    [19] F. H. Vasilescu, Analytic functional calculus and spectral decompositions, D. Reidel Publishing Co., Dordrecht and Editura Academiei R. S. R., Bucharest, 1982.
    [20] V. Müller, V. Kordula, Vasilescu-Martinelli formula for operators in Banach spaces, Stud. Math., 113 (1995), 127–139.
    [21] F. H. Vasilescu, Functions and operators in real, quaternionic, and cliffordian contexts, Complex Anal. Oper. Th., 16 (2022), 117. https://doi.org/10.1007/s11785-022-01292-x doi: 10.1007/s11785-022-01292-x
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1144) PDF downloads(61) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog