1. Introduction
Let be the Banach algebra of all bounded linear operators defined on a complex Hilbert space with the identity operator in . A bounded linear operator A defined on is self-adjoint if for all . The spectrum of an operator A is the set of all for which the operator does not have a bounded linear inverse operator, and is denoted by . Consider the real vector space of self-adjoint operators on and its positive cone of positive operators on . Additionally, denotes the convex set of bounded self-adjoint operators on the Hilbert space with spectra in a real interval J. A partial order is naturally equipped on by defining if and only if . We write to mean that A is a strictly positive operator, or equivalently, and A is invertible. When , we identify with the algebra of n-by-n complex matrices. Then, is just the cone of n-by-n positive semidefinite matrices.
A linear map is defined to be , which preserves additivity and homogeneity, that is, for any and . A linear map is positive if it preserves the operator order, that is, if then , and in this case we write . Obviously, a positive linear map preserves the order relation, namely implies that and preserves the adjoint operation . Moreover, is said to be unital if it preserves the identity operator, in this case, we write .
A linear map induces another map
in a natural way. If is identified with the -algebra of –matrices with entries in then act as:We say that is k-positive if is a positive map, and is called completely positive if is k-positive for all k.
1.1. Superquadratic Functions
A function is called convex if
(1)
for all points and all . If is convex then we say that f is concave. Moreover, if f is both convex and concave, then f is said to be affine.Geometrically, for two points and on the graph of f are on or below the chord joining the endpoints for all , . In symbols, we write
for any and .Equivalently, given a function , we say that f admits a support line at if there exists a such that
(2)
for all .The set of all such is called the subdifferential of f at x, and it’s denoted by . Indeed, the subdifferential gives us the slopes of the supporting lines for the graph of f. So that if f is convex then at all interior points of its domain.
From this point of view, Abramovich et al. [1] extend the above idea for what they called superquadratic functions. Namely, a function is called superquadratic provided that for all there exists a constant such that
(3)
for all . We say that f is subquadratic if is superquadratic. Thus, for a superquadratic function, we require that f lie above its tangent line plus a translation of f itself. If f is differentiable and satisfies , then we know easily that the appearing in the definition is necessarily (see [2]).At first glance, the superquadratic function looks to be stronger than the convex function itself, but if f takes negative values then it may be considered a weaker function. Therefore, if f is superquadratic and non-negative, then f is convex and increasing [1] (see also [3]).
Moreover, the following result holds for superquadratic functions.
([1]). Let f be a superquadratic function. Then
- (1)
.
- (2)
if f is differentiable and , then for all .
- (3)
if for all , then f is convex and .
The next result gives a sufficient condition when convexity (concavity) implies super(sub)quaradicity.
([1]). If is convex (concave) and , then f is super(sub)quadratic. The converse is not true.
In general, non-negative subquadratic functions do not imply concavity. In other words, there exists a subquadratic function that is convex. For example, , and is subquadratic and convex.
Among others, Abramovich et al. [1] proved that the inequality
(4)
holds for all probability measures and all non-negative, -integrable functions if and only if f is superquadratic. For more details the reader may refer to [3,4,5,6].1.2. Operator Convexity and Jensen’s Inequality
Let f be a real-valued function defined on J. A k-th order divided difference of f at distinct points in J may be defined recursively by
For instance, the first three divided differences are given as follows:
A function is said to be matrix monotone of degree n or n-monotone, if for every , it is true that if and only if . Similarly, f is said to be operator monotone if f is n-monotone for all . Additionally, f is called operator convex if it is matrix convex (n-convex for all n); that is, if for every pair of self-adjoint operators , we have
for all . If the inequality is reversed then f is called operator concave. In case we have general Hilbert space , the above definition holds for every pair of bounded self-adjoint operators A and B in , whose spectra obtained in J. For more details, see [7] and the recent survey [8].In 1955, Bendat and Sherman [9] have shown that f is operator convex on the open interval if and only if it has the following (unique) representation:
for and some probability measure on (it could be the Borel measure). In particular, f must be analytic with , and .We recall that the celebrated Löwner–Heinz inequality reads that:
Let such that . Then for all .
On the other hand the mapping is not operator monotone, for more details, see [10,11,12].
The classical Jensen’s inequality can be formulated as
(5)
is valid for all real-valued convex functions f defined on , for every and every positive real number such that .The inequality (5) would be rephrased under the matrix situation by putting
then the classical Jensen’s inequality (5) is expressed as(6)
which is one of the operator versions of the classical Jensen’s inequality, see [11,13], the recent monograph [14], as well as [15,16,17,18].Kadison [19] established his famous non-commutative version of the previous inequality, where he proved that, for every self-adjoint matrix A, the inequality
(7)
for every positive unital linear map .This inequality was generalized later by Davis in [20], where he obtained that this is true when f is a matrix convex function and is completely positive, that is,
(8)
The latter restriction about complete positivity of was removed by Choi [21] who proved that (7) remains valid for all positive unital linear maps provided f is matrix convex.
Another non-commutative operator version of the classical Jensen’s inequality under the situation that
the classic Jensen’s inequality is expressed as(9)
The inequality (9) was proved by Davis in [20] for all and every isometry V. However, a more informative version was extended by Hansen–Pedersen [12] as follows:
Let and be Hilbert spaces. Let f be a real-valued continuous function on an interval J. Let A and be self-adjoint operators on with spectra contained in I. Then the following conditions are mutual:
-
(1)
f is operator convex on J and .
-
(2)
, for every and contraction , that is, .
-
(3)
, for all and with , .
-
(4)
, for every and projection P.
Here, we give some popular examples of operator convex and concave function [8].
-
(1). For each , is operator concave on .
-
(2). The function is operator convex on .
This work is organized as follows: after this introduction; in Section 2, the operator superquadratic functions for positive Hilbert space operators are introduced and elaborated. Several examples with some important properties together with some observations related to operator convexity are pointed out. A general Bohr’s inequality for positive operators is thus deduced. A Jensen’s type inequality is proved in Section 3. Equivalent statements of a non-commutative version of Jensen’s inequality for operator superquadratic are also established. Finally, several trace inequalities for superquadratic functions (in the ordinary sense) are provided as well in Section 4.
2. Operator Superquadratic Function
Let . A real-valued continuous function on an interval J is said to be an operator superquadratic function if
(10)
holds for all and for every positive operators A and B on a Hilbert space whose spectra are contained in . We say that f is an operator subquadratic function if is an operator superquadratic function. Moreover, if the equality holds in (10), we say that f is an operator quadratic function.It is convenient to note that if f satisfies (10), then with and (two positive scalars), we can obtain the Jensen’s inequality for superquadratic functions and if f is continuous (which is necessary to define an operator functions), then (4) would imply that f is a superquadratic function. Thus, we observe that:
If f is an operator superquadratic function then f is a real superquadratic function.
Let . Then f is operator subquadratic on every bounded interval for all . Indeed, we have
Moreover, is operator superquadratic.
We can easily show that the function is not an operator superquadratic nor an operator subquadratic function. Simply, assume , , and let
thenHowever, the map is a non-negative operator convex on and it is also operator superquadratic on . Indeed, by (10), we have
which is true since , and this proves that is an operator superquadratic function.From the definition of the operator superquadratic function, we have
(11)
for any arbitrary positive operators and each .In particular, by setting in (10), we have
(12)
for each positive operator and all .From this point of view (12), Kian early in [22] and then jointly with Dragomir in [23] proved a finite dimensional operator version of Jensen’s inequality for superquadratic functions (in the ordinary sense) under the interpretation that for and , then we have and it follows that
Therefore, as a matrix Jensen’s inequality for a superquadratic function we have
This result was generalized for positive unital linear maps, as follows:
([23]). Let be a positive operator and be a positive unital linear map. If is a super(sub)quadratic function, then we have
for every with .The above inequality and other consequences were proved later by the first author of this paper in [24], where a different approach is used.
Let f be an operator superquadratic function on J. Then
-
(1)
.
-
(2)
if f is non-negative, then f is operator convex and .
(1) Setting in (12), we get that .
(2) Since f is continuous and non-negative, from (12), we have
which means that f is operator convex. To show that , we have by the part (1) and by the assumption is non-negative, that is, for all . In particular, we have . Thus, . □Let . Then f is non-negative operator convex on . However, f is not an operator superquadratic function on . For instance, let
Applying (13) for , we get
Let f be a real-valued continuous function defined on an interval . If f is operator convex and non-positive, then f is an operator superquadratic function.
Since f is operator convex,
However, f is also non-positive, so that
which means that f is an operator superquadratic function. □Let , . Then it is well known that f is operator convex. Clearly, f is negative for all . Therefore, is an operator superquadratic function for all .
Let f be a real-valued continuous function defined on an interval . If f is operator concave and non-negative, then f is operator subquadratic.
Since f is operator concave,
However, also, f is non-negative, so that
which means that f is operator subquadratic. □Let , given by , . Then f is operator subquadratic on . However, f is also operator concave, so that
which means that f is operator subquadratic on .
On Bohr’s Inequality
The classical Bohr inequality for scalars reads that: if a and b are complex numbers and with , then
An operator version of this inequality was treated by Hirzallah [25] and the latter by many authors. See, for example, [13,26]. Namely, in [25], we find that
is valid for all and with and , where is the absolute value of the operator X.Recently, it is shown in [24] that, for a positive self-adjoint operator and a positive unital linear map , the following inequalities hold:
1.. If is a real superquadratic function, then we have
In particular, for , , , we have
2.. If is a real subquadratic function, then we have
In particular, for , , , we have
Now, set , so that with in (10). If f is an operator superquadratic function, then a general operator Bohr inequality can be given in the form
(13)
for all with . In particular, for , we have(14)
If f is subquadratic then the inequalities (13) and (14) are reversed.
As a direct example, let , , . Then f is operator subquadratic. Hence, by (13), we have
which is equivalent to writing since for all with , for all positive operators .In particular, for , we have
3. Operator Jensen’s Inequality
In order to prove our results, we need the following lemmas:
([11]). If is self-adjoint and U is unitary, that is, , then for every continuous function f on .
([27]). Define a unitary matrix in , where . Then, for each element , we have
([27]). Let P denote the projection in given by for all i and j, so that P is the projection of rank one on the subspace spanned by the vector in , where are the standard basis vectors. Then with E as in Lemma 5 we obtain the pairwise orthogonal projections , for , with .
In order to establish our main first result, we need the following primary result:
Let be positive real numbers such that and let be positive operators on a Hilbert space with spectra contained in a real interval J. If f is an operator superquadratic function on J, then
(15)
In particular, as a useful case, for for all , we have
(16)
Assume f is operator superquadratic. If , then the inequality (15) reduces to (10) with and . Let us suppose that inequality (15) holds for . Then, for n-tuples and , we have
and this is exactly equivalent to writing, for any , which proves the desired result in (15). The particular case follows by setting for all so that . □The result in Lemma 7 was proved by Mond and Pečarić in [28] for all operator convex functions and all bounded self-adjoint operators whose spectra are contained in J. Therefore, in case f is positive, the inequality (15) might be considered as a respective extension and a new refinement of that result proved in [28].
Let be a real-valued continuous function. Let be an n-tuple of positive operators on a Hilbert space with spectra contained in J. Then the following conditions are equivalent:
-
(1)
f is an operator superquadratic function.
-
(2)
The inequality
(17)
holds for every n-tuple of operators on that satisfy the condition . -
(3)
The inequality
(18)
holds for every n-tuple of projections on with .
. We say that is a unitary column if there is a unitary operator matrix , one of whose columns is . Thus, for some j and all i. Assume that we are given a unitary n-column , and choose a unitary in such that . Let as in Lemma 4 and put , both regarded as elements, in . Thus, using the spectral decomposition theorem, we have
We note that, since ,
Using the above facts taking into account Lemmas 4–7 together with the inequality (16), the operator superquadraticity of f implies that
It remains to mention that, when the column is just unital, we extend it to the unitary -column and choose arbitrarily, but with spectrum in J (see [29]). By the first part of the proof, we therefore have
and thus the proof of statement (2) is completely established.. It obviously holds.
. Let A and B be positive and bounded linear operators with spectra in J and .
Consider
Then, C and D are unitary operators on . We have
Thus, we have
Hence, f is operator superquadratic on J by seeing the -components. □
An operator convex version of Theorem 3 was proved by Hansen and Pedersen in [27]. Therefore, in the case of f being positive, the inequalities (17) and (18) could be considered as new refinements of the result proven in [27]; for example the function , , is a nontrivial example that refines the Hansen-Pedersen inequalities in [27].
A refinement of the classical Jensen’s inequality (9) could be elaborated as follows:
Let be a real-valued continuous function. Let A be a positive operator on a Hilbert space with spectra contained in J. If f is an operator superquadratic function, then the inequality
(19)
holds for every operator C on that satisfies the condition .It follows from Theorem 3 by setting . □
Let be a real-valued continuous function. Let A be a positive operator on a Hilbert space with spectra contained in J. If f is an operator subquadratic function, then the inequality
holds for every operator C on that satisfies the condition . Furthermore, by applying the subquadratic function , , then we have
for all .
A generalization of the famous inequality of Davis–Choi (8) and thus (19) to any positive unital linear map.
Let be a Hilbert space. Let be a real-valued continuous function. Let A be a positive operator on a Hilbert space with spectra contained in J and be a positive unital linear map. If f is an operator superquadratic function, then the inequality
(20)
holds. If f is operator subquadratic, then the inequality (20) is reversed. Thus, the following refinement of the celebrated Kadison inequality (7) is valid:
Let be positive. Assume that is the -subalgebra of generated by A and . Without loss of generality, we may assume that is defined on . Since every unital positive linear map on a commutative -algebra is completely positive. It follows that is completely positive. So there exists (by Stinespring’s theorem [30]), some isometry ; and a unital *-homomorphism from into the -algebra such that . Clearly, , for all continuous functions f. Thus,
which proves the required inequality. The last inequality holds by applying (20) to the superquadratic function , . □The inequality (20) can be embodied in multiple versions as stated in the following result:
Let be a Hilbert space. Let be a real-valued continuous function and be a positive linear mappings with . Then, f is an operator superquadratic function if and only if
for all positive operators in .
The proof is obvious, and thus omitted. □
4. Jensen’s Trace Inequality
Let , and recall that the trace of a square matrix equals the sum of the eigenvalues counted with multiplicities. Moreover, the trace of a Hermitian matrix is real. If A is a linear operator represented by a square matrix with real or complex entries and if are the eigenvalues of A, then . This follows from the fact that A is always similar to its Jordan form, an upper triangular matrix having on the main diagonal.
The inner product
which is defined on the space of all complex (or real) matrices, is called the Frobenius norm, which satisfies the submultiplicative property as a matrix norm.If A and B are real positive semi-definite matrices of the same size, using the Cauchy–Schwarz inequality, we have
The concept of trace of a matrix is generalized to the trace class of compact operators on Hilbert spaces, and the analog of the Frobenius norm is called the Hilbert–Schmidt norm, which is can be defined as
over all orthonormal basis of , (see [31]).A Hilbert–Schmidt operator is a bounded operator A on a Hilbert space with finite Hilbert–Schmidt norm
where is the norm of and Tr is the trace of a non-negative self-adjoint operator. This definition is independent of the choice of the basis, and thereforeA bounded linear operator A over a separable Hilbert space is said to be in the trace class if for some (and hence all) orthonormal bases of , the sum of positive terms
is finite. In this case, the trace of A, which is given by the sum is absolutely convergent and is independent of the choice of the orthonormal basis. When is finite-dimensional, every operator is a trace class and this definition of trace of A coincides with the definition of the trace of a matrix.Let be continuous and n be any integer. It is well known that if is convex (monotone increasing), then the trace function is convex (monotone increasing); see [32,33], and the recent work in [34].
In 2003, Hansen and Pedersen [27] proved the following version of Jensen’s inequality:
for every n-tuple of positive matrices with spectra contained in J and every n-tuple of matrices with , where f is assumed to be convex on J.Using the concept of superquadratic functions, we could give the following refinement of Hansen–Pedersen trace inequality:
Let f be a real-valued continuous function defined on an interval J and let m and n be natural numbers. If f is a superquadratic function (in the ordinary sense), then the inequality
(21)
holds for every n-tuple of positive matrices with spectra contained in J and every n-tuple of matrices with . Conversely, if the inequality (21) is satisfied for some n and m, where , then f is a superquadratic function. If f is subquadratic, then the inequality (21) is reversed.Our proof is motivated by [27]. Let denote the spectral resolution of for . Then, is the spectral projection of on the eigenspace corresponding to if is an eigenvalue for , otherwise . For each unit vector in , let us define the probability measure
for any (Borel) set S in . Note that if , thenIf a unit vector is an eigenvector for y, then the corresponding eigenvalue is , and is also an eigenvector for with the corresponding eigenvalue = . In this case, we have
Summing over an orthonormal basis of eigenvectors for y, we get the desired result in (21). □
Let f be a real-valued continuous function defined on an interval J and let m and n be natural numbers. If f is a superquadratic function (in the ordinary sense), then the inequality
(22)
holds for every positive matrix A with spectrum contained in J and every matrix C with . If f is subquadratic, then the inequality (22) is reversed. Furthermore, we have(23)
for every , and(24)
for every .The result follows by setting in Theorem 5. The inequality (23) follows by applying the superquadratic function , . Similarly, the inequality (24) follows by applying the subquadratic function , . □
The inequality (21) could be extended for general positive Hilbert space operators mapped under a positive unital linear map, as follows:
Let f be a real-valued continuous function defined on . Let be positive operators. Let be a positive linear map, such that , where is the identity matrix of . If f is a superquadratic function, then
(25)
holds for every n-tuple of positive matrices with spectra contained in J. Conversely, if the inequality (25) is satisfied for some n and m, where , then f is a superquadratic function. If f is subquadratic, then the inequality (25) is reversed.Firstly, let be positive. Assume that is the -subalgebra of generated by and identity . Let be a positive linear map, such that . Let be n-tuple of matrices with . Without loss of generality, assume that is defined on . Since every unital positive linear map on a commutative -algebra is completely positive. It follows that is completely positive. So, by the Stinespring’s theorem [30], there exists, an isometry (such isometry is valid for all j since each such is -subalgebra can be generated by different for all ), and a unital *-homomorphism from into the -algebra , such that
(26)
Clearly, , for all continuous functions f. Thus, Let denote the spectral resolution of for . Then, is the spectral projection of on the eigenspace corresponding to if is an eigenvalue for , otherwise . For each unit vector in , let us define the probability measure for any (Borel) set S in . Note that, if , then If a unit vector is an eigenvector for w, then the corresponding eigenvalue is , and is also an eigenvector for with corresponding eigenvalue = . In this case, and taking into account the representation (26), we haveSumming over an orthonormal basis of eigenvectors for w, we get the desired result in (25). □
As a special case, we can deduce the following result:
Let f be a real-valued continuous function defined on . Let be a positive operator and be a positive unital linear map with . If f is superquadratic, then we have
(27)
A particular case is the choice , where , is such that . Indeed, this reduces to the inequality (22).The proof follows by setting in (25). □
Conceptualization, M.W.A. and A.A.-K.; methodology, M.W.A., C.C. and A.A.-K.; validation, M.W.A., C.C. and A.A.-K.; formal analysis, M.W.A., C.C. and A.A.-K.; investigation, M.W.A., C.C. and A.A.-K.; writing—original draft preparation, M.W.A.; writing—review and editing, M.W.A., C.C. and A.A.-K. All authors have read and agreed to the published version of the manuscript.
Not applicable.
The referees’ insightful criticism and careful reading of this work’s original manuscript helped the authors make the final version better, so they sincerely thank them.
The authors declare no conflict of interest.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
1. Abramovich, S.; Jameson, G.; Sinnamon, G. Refining Jensen’s inequality. Bull. Math. Soc. Sci. Math. Roumanie; 2004; 47, pp. 3-14.
2. Abramovich, S.; Jamesion, G.; Sinnamon, G. Inequalities for averages of convex and superquadratic functions. J. Inequ. Pure Appl. Math.; 2004; 2004. 5, 91.
3. Abramovich, S. On superquadraticity. J. Math. Inequal.; 2009; 3, pp. 329-339. [DOI: https://dx.doi.org/10.7153/jmi-03-33]
4. Abramovich, S.; Ivelić, S.; Pexcxarixcx, J. Improvement of Jensen-Steffensen’s inequality for superquadratic functions. Banach J. Math. Anal.; 2010; 4, pp. 146-158. [DOI: https://dx.doi.org/10.15352/bjma/1272374678]
5. Olanipekun, P.O.; Mogbademu, A.A. Inequalities of the Jensen type for a new class of convex functions. Nonlinear Funct. Anal. Appl.; 2015; 20, pp. 645-658.
6. Alomari, M.W.; Chesneau, C. On h-superquadratic functions. Afr. Mat.; 2022; 33, 41. [DOI: https://dx.doi.org/10.1007/s13370-022-00984-z]
7. Kraus, F. Über konvexe Matrixfunktionen. Math. Z.; 1936; 41, pp. 18-42. [DOI: https://dx.doi.org/10.1007/BF01180403]
8. Chansangiam, P. A Survey on Operator Monotonicity, Operator Convexity, and Operator Means. Internat J. Anal.; 2015; 2015, 649839. [DOI: https://dx.doi.org/10.1155/2015/649839]
9. Bendat, J.; Sherman, S. Monotone and convex operator functions. Trans. Am. Math. Soc.; 1955; 79, pp. 58-71. [DOI: https://dx.doi.org/10.1090/S0002-9947-1955-0082655-4]
10. Bhatia, R. Matrix Analysis; Springer: New York, NY, USA, 1996.
11. Furuta, T.; Mićixcx, J.; Pexcxarixcx, J.; Seo, Y. Mond-Pečarić Method in Operator Inequalities. Inequalities for Bounded Selfadjoint Operators on a Hilbert Space; Element: Zagreb, Croatia, 2005.
12. Hansen, F.; Pedersen, G.K. Jensen’s inequality for operators and Löwner’s theorem’. Math. Ann.; 1982; 258, pp. 229-241. [DOI: https://dx.doi.org/10.1007/BF01450679]
13. Krnić, M.; Lovrinčević, N.; Pečarić, J.; Perić, J. Superadditivity and Monotonicity of the Jensen-Type Functionals: New Methods for Improving the Jensen-Type Inequalities in Real and in Operator Cases; Element: Zagreb, Croatia, 2016.
14. Ahmed, N.U. Output feedback min-max control problem for a class of uncertain linear stochastic systems on UMD Banach space. Nonlinear Funct. Anal. Appl.; 2019; 24, pp. 127-154.
15. Agarwal, R.P.; Dragomir, S.S. A survey of Jensen type inequalities for functions of selfadjoint operators in Hilbert spaces. Comput. Math. Appl.; 2010; 59, pp. 3785-3812. [DOI: https://dx.doi.org/10.1016/j.camwa.2010.04.014]
16. Barić, J.; Matković, A.; Pečarić, J. A variant of the Jensen–Mercer operator inequality for superquadratic functions. Math. Comput. Model.; 2010; 51, pp. 1230-1239. [DOI: https://dx.doi.org/10.1016/j.mcm.2010.01.005]
17. Banić, S.; Pečarić, J.; Varošanec, S. Superquadratic functions and refinements of some classical inequalities. J. Korean Math. Soc.; 2008; 45, pp. 513-525. [DOI: https://dx.doi.org/10.4134/JKMS.2008.45.2.513]
18. Petz, D. A Survey of Certain Trace Inequalities, Functional Analysis and Operator Theory Center Publications; Institute of Mathematics, Polish Academy od Sciences: Warszawa, Poland, 1994; Volume 30.
19. Kadison, R.V. A generalized Schwarz inequality and algebraic invariants for operator algebras. Ann. Math.; 1952; 56, pp. 494-503. [DOI: https://dx.doi.org/10.2307/1969657]
20. Davis, J.C. A Schwarz inequality for convex operator functions. Proc. Am. Math. Soc.; 1957; 8, pp. 42-44. [DOI: https://dx.doi.org/10.1090/S0002-9939-1957-0084120-4]
21. Choi, M.D. A Schwarz inequality for positive linear maps on C*-algebras. Ill. J. Math.; 1974; 18, pp. 565-574. [DOI: https://dx.doi.org/10.1215/ijm/1256051007]
22. Kian, M. Operator Jensen inequality for superquadratic functions. Linear Algebra Its Appl.; 2014; 456, pp. 82-87. [DOI: https://dx.doi.org/10.1016/j.laa.2012.12.011]
23. Kian, M.; Dragomir, S.S. Inequalities involving superquadratic functions and operators. Mediterr. J. Math.; 2014; 11, pp. 1205-1214. [DOI: https://dx.doi.org/10.1007/s00009-013-0357-y]
24. Alomari, M.W. Operator Popoviciu’s inequality for superquadratic and convex functions of selfadjoint operators in Hilbert spaces. Adv. Pure Appl. Math.; 2019; 10, pp. 313-324. [DOI: https://dx.doi.org/10.1515/apam-2018-0154]
25. Hirzallah, O. Non-commutative operator Bohr inequality. J. Math. Anal. Appl.; 2003; 282, pp. 578-583. [DOI: https://dx.doi.org/10.1016/S0022-247X(03)00185-9]
26. Abramovich, S.; Barić, J.; Pečarič, J. A new proof of an inequality of Bohr for Hilbert space operators. Linear Algebra Appl.; 2009; 430, pp. 1432-1435. [DOI: https://dx.doi.org/10.1016/j.laa.2008.09.041]
27. Hansen, F.; Pedersen, G.K. Jenssen’s operator inequality. Bull. Lond. Math. Soc.; 2003; 35, pp. 553-564. [DOI: https://dx.doi.org/10.1112/S0024609303002200]
28. Mond, B.; Pečarić, J.E. Convex inequalities in Hilbert space. Houst. J. Math.; 1993; 19, pp. 405-420.
29. Araki, H.; Hansen, F. Jensen’s operator inequality for functions of several variables. Proc. Am. Math. Soc.; 2000; 128, pp. 2075-2084. [DOI: https://dx.doi.org/10.1090/S0002-9939-00-05371-5]
30. Stinespring, W.F. Positive functions on C*-algebras. Proc. Am. Math. Soc.; 1955; 6, pp. 211-216.
31. Simon, B. Trace Ideals and Their Applications; 2nd ed. American Mathematical Society: Providence, RI, USA, 2005.
32. Horn, R.A.; Johnson, C.R. Matrix Analysis; 2nd ed. Cambridge University Press: Cambridge, UK, 2013.
33. Lieb, E.H. Convex trace functions and the Wigner–Yanase–Dyson Conjecture. Adv. Math.; 1973; 11, pp. 267-288. [DOI: https://dx.doi.org/10.1016/0001-8708(73)90011-X]
34. Kian, M.; Alomari, M.W. Improvements of trace inequalities for convex functions. Ann. Funct. Anal.; 2022; 13, 64. [DOI: https://dx.doi.org/10.1007/s43034-022-00211-7]
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
In this work, an operator superquadratic function (in the operator sense) for positive Hilbert space operators is defined. Several examples with some important properties together with some observations which are related to the operator convexity are pointed out. A general Bohr’s inequality for positive operators is thus deduced. A Jensen-type inequality is proved. Equivalent statements of a non-commutative version of Jensen’s inequality for operator superquadratic function are also established. Finally, several trace inequalities for superquadratic functions (in the ordinary sense) are provided as well.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details

1 Department of Mathematics, Faculty of Science and Information Technology, Irbid National University, Irbid 21110, Jordan
2 Department of Mathematics, Université de Caen Basse-Normandie, F-14032 Caen, France
3 Department of Information Technology, Faculty of Prince Al-Hussein Bin Abdallah II for Information Technology, The Hashemite University, Zarqa 13133, Jordan; Department of Cyber Security, Faculty of Science and Information Technology, Irbid National University, Irbid 21110, Jordan