Research article Special Issues

Partial domination of network modelling

  • Received: 28 March 2023 Revised: 18 July 2023 Accepted: 20 July 2023 Published: 11 August 2023
  • MSC : 05C07, 05C69

  • Partial domination was proposed in 2017 on the basis of domination theory, which has good practical application background in communication network. Let G=(V,E) be a graph and F be a family of graphs, a subset SV is called an F-isolating set of G if G[VNG[S]] does not contain F as a subgraph for all FF. The subset S is called an isolating set of G if F={K2} and G[VNG[S]] does not contain K2 as a subgraph. The isolation number of G is the minimum cardinality of an isolating set of G, denoted by ι(G). The hypercube network and n-star network are the basic models for network systems, and many more complex network structures can be built from them. In this study, we obtain the sharp bounds of the isolation numbers of the hypercube network and n-star network.

    Citation: Shumin Zhang, Tianxia Jia, Minhui Li. Partial domination of network modelling[J]. AIMS Mathematics, 2023, 8(10): 24225-24232. doi: 10.3934/math.20231235

    Related Papers:

    [1] Zhaoxia Li, Lihua Deng, Haifeng Shang . Global well-posedness and large time decay for the d-dimensional tropical climate model. AIMS Mathematics, 2021, 6(6): 5581-5595. doi: 10.3934/math.2021330
    [2] Jing Yang, Xuemei Deng, Qunyi Bie . Global regularity for the tropical climate model with fractional diffusion. AIMS Mathematics, 2021, 6(10): 10369-10382. doi: 10.3934/math.2021601
    [3] Ikram Ullah, Muhammad Bilal, Javed Iqbal, Hasan Bulut, Funda Turk . Single wave solutions of the fractional Landau-Ginzburg-Higgs equation in space-time with accuracy via the beta derivative and mEDAM approach. AIMS Mathematics, 2025, 10(1): 672-693. doi: 10.3934/math.2025030
    [4] Yunlei Zhan . Large time behavior of a bipolar hydrodynamic model with large data andvacuum. AIMS Mathematics, 2018, 3(1): 56-65. doi: 10.3934/Math.2018.1.56
    [5] Shang Mengmeng . Large time behavior framework for the time-increasing weak solutions of bipolar hydrodynamic model of semiconductors. AIMS Mathematics, 2017, 2(1): 102-110. doi: 10.3934/Math.2017.1.102
    [6] Adel M. Al-Mahdi . The coupling system of Kirchhoff and Euler-Bernoulli plates with logarithmic source terms: Strong damping versus weak damping of variable-exponent type. AIMS Mathematics, 2023, 8(11): 27439-27459. doi: 10.3934/math.20231404
    [7] Abdelkader Moumen, Fares Yazid, Fatima Siham Djeradi, Moheddine Imsatfia, Tayeb Mahrouz, Keltoum Bouhali . The influence of damping on the asymptotic behavior of solution for laminated beam. AIMS Mathematics, 2024, 9(8): 22602-22626. doi: 10.3934/math.20241101
    [8] Yangyang Chen, Yixuan Song . Space-time decay rate of the 3D diffusive and inviscid Oldroyd-B system. AIMS Mathematics, 2024, 9(8): 20271-20303. doi: 10.3934/math.2024987
    [9] Mingyu Zhang . On the Cauchy problem of 3D nonhomogeneous micropolar fluids with density-dependent viscosity. AIMS Mathematics, 2024, 9(9): 23313-23330. doi: 10.3934/math.20241133
    [10] Yasir Nawaz, Muhammad Shoaib Arif, Kamaleldin Abodayeh, Mairaj Bibi . Finite difference schemes for time-dependent convection q-diffusion problem. AIMS Mathematics, 2022, 7(9): 16407-16421. doi: 10.3934/math.2022897
  • Partial domination was proposed in 2017 on the basis of domination theory, which has good practical application background in communication network. Let G=(V,E) be a graph and F be a family of graphs, a subset SV is called an F-isolating set of G if G[VNG[S]] does not contain F as a subgraph for all FF. The subset S is called an isolating set of G if F={K2} and G[VNG[S]] does not contain K2 as a subgraph. The isolation number of G is the minimum cardinality of an isolating set of G, denoted by ι(G). The hypercube network and n-star network are the basic models for network systems, and many more complex network structures can be built from them. In this study, we obtain the sharp bounds of the isolation numbers of the hypercube network and n-star network.



    A well-known and active direction in the study of derivations is the local derivations problem, which was initiated by Kadison [8] and Larson and Sourour [9]. Recall that a linear map φ of an algebra A is called a local derivation if for each xA, there exists a derivation φx of A, depending on x, such that φ(x)=φx(x). The question of determining under what conditions every local derivation must be a derivation has been studied by many authors (see [4,6,7,13,15]). Recently, Brešar [2] proved that each local derivation of algebras generated by all their idempotents is a derivation.

    A linear map φ of an algebra A is called a Lie derivation if φ([x,y])=[φ(x),y]+[x,φ(y)] for all x,yA, where [x,y]=xyyx is the usual Lie product, also called a commutator. A Lie derivation φ of A is standard if it can be decomposed as φ=d+τ, where d is a derivation from A into itself and τ is a linear map from A into its center vanishing on each commutator. The classical problem, which has been studied for many years, is to find conditions on A under which each Lie derivation is standard or standard-like. We say that a linear map φ from A into itself is a local Lie derivation if for each xA, there exists a Lie derivation φx of A such that φ(x)=φx(x). In [3], Chen et al. studied local Lie derivations of operator algebras on Banach spaces. We remark that the methods in [3] depend heavily on rank one operators in B(X). Later, Liu and Zhang [10] proved that each local Lie derivation of factor von Neumann algebras is a Lie derivation. Liu and Zhang [11] investigated local Lie derivations of a certain class of operator algebras. An et al. [1] proved that every local Lie derivation on von Neumann algebras is a Lie derivation.

    It is quite common to study local derivations in algebras that contain many idempotents, in the sense that the linear span of all idempotents is 'large'. The main novelty of this paper is that we shall deal with the subalgebra generated by all idempotents instead of their span. Let M2 be the algebra of 2×2 matrices over L[0,1]. By [6], M2 is generated by, but not spanned by, its idempotents. In what follows, we denote by J(A) the subalgebra of A generated by all idempotents in A. The purpose of the present paper is to study local Lie derivations of a certain class of generalized matrix algebras. Finally we apply the main result to full matrix algebras and unital simple algebras with nontrivial idempotents.

    Let A and B be two unital algebras with unit elements 1A and 1B, respectively. A Morita context consists of A,B, two bimodules AMB and BNA, and two bimodule homomorphisms called the pairings ΦMN:MBNA and ΨNM:NAMB satisfying the following commutative diagrams:

    and

    If (A,B,M,N,ΦMN,ΨNM) is a Morita context, then the set

    G=(AMNB)={(amnb)aA,mM,nN,bB}

    forms an algebra under matrix-like addition and multiplication. Such an algebra is called a generalized matrix algebras. We further assume that M is faithful as an (A,B)-bimodule. The most common examples of generalized matrix algebras are full matrix algebras and triangular algebras.

    Consider algebra G. Any element of the form

    (a00b)G

    will be denoted by ab. Let us define two natural projections πA:GA and πB:GB by

    πA:(amnb)a   and   πB:(amnb)b.

    The center of G is

    Z(G)={abam=mb,na=bn for all mM,nN}.

    Furthermore, πA(Z(G))Z(A) and πB(Z(G))Z(B), and there exists a unique algebra isomorphism η from πB(Z(G)) to πA(Z(G)) such that η(b)m=mb and nη(b)=bn for all mM,nN (see [14]). Set

    e=(1A000),  f=(0001B).

    We immediately notice that e and f are orthogonal idempotents of G and so G may be represented as G=(e+f)G(e+f)=eGe+eGf+fGe+fGf. Then each element x=exe+exf+fxe+fxfG can be represented in the form x=eae+emf+fne+fbf=a+m+n+b, where aA,bB,mM,nN.

    We close this section with a well known result concerning Lie derivations.

    Proposition 1.1. (See [5],Theorem 1) Let G be a generalized matrix algebra. Suppose that

    (1) Z(A)=πA(Z(G)) and Z(B)=πB(Z(G));

    (2) either A or B does not contain nonzero central ideals.

    Then every Lie derivation φ:GG is standard, that is, φ is the sum of a derivation d and a linear central-valued map τ vanishing on each commutator.

    Our main result reads as follows.

    Theorem 2.1. Let G be a generalized matrix algebra. Suppose that

    (1) A=J(A) and B=J(B);

    (2) Z(A)=πA(Z(G)) and Z(B)=πB(Z(G));

    (3) either A or B does not contain nonzero central ideals.

    Then every local Lie derivation φ from G into itself is a sum of a derivation δ and a linear central-valued map h vanishing on each commutator.

    To prove Theorem 2.1, we need some lemmas. In the following, φ is a local Lie derivation and, for any xG, the symbol φx stands for a Lie derivation from G into itself such that φ(x)=φx(x). It follows from A=J(A) that every a in A can be written as a linear combination of some elements p1p2pi (i=1,2,,k), where p1,p2,,pi are idempotents in A.

    Lemma 2.2. Let p,qG be idempotents, then for every xG, there exist linear maps τ1,τ2,τ3,τ4:GZ(G) vanishing on each commutator such that

    φ(pxq)=φ(px)q+pφ(xq)pφ(x)q+pτ1(pxq)qpτ2(pxq)q+pτ3(pxq)qpτ4(pxq)q,

    where p=1p and q=1q.

    Proof. Proposition 1.1 implies that for every idempotents p,qG and xG, there exist derivations d1,d2,d3,d4:GG and linear maps τ1,τ2,τ3,τ4:GZ(G) vanishing on each commutator such that

    φ(pxq)=φpxq(pxq)=d1(pxq)+τ1(pxq), (2.1)
    φ(pxq)=φpxq(pxq)=d2(pxq)+τ2(pxq), (2.2)
    φ(pxq)=φpxq(pxq)=d3(pxq)+τ3(pxq), (2.3)
    φ(pxq)=φpxq(pxq)=d4(pxq)+τ4(pxq). (2.4)

    It follows from (2.1)–(2.4) that

    pφ(pxq)q=pτ1(pxq)q, pφ(pxq)q=pτ2(pxq)q,
    pφ(pxq)q=pτ3(pxq)q, pφ(pxq)q=pτ4(pxq)q.

    Hence

    φ(pxq)q=pφ(pxq)q+pφ(pxq)q=pφ(xq)qpφ(pxq)q+pφ(pxq)q=pφ(xq)q+pτ1(pxq)qpτ2(pxq)q=pφ(xq)pφ(xq)q+pτ1(pxq)qpτ2(pxq)q,
    φ(pxq)q=pφ(pxq)q+pφ(pxq)q=pφ(xq)qpφ(pxq)q+pφ(pxq)q=pφ(xq)qpτ3(pxq)q+pτ4(pxq)q.

    Thus,

    φ(pxq)=φ(pxq)q+φ(pxq)q=φ(pxq)q+φ(px)qφ(pxq)q=φ(px)q+pφ(xq)pφ(x)q+pτ1(pxq)qpτ2(pxq)q+pτ3(pxq)qpτ4(pxq)q.

    It is easy to verify that for each derivation d:GG, we have

    d(e)=d(f)MN, d(A)AMN, d(M)AMB. (2.5)

    Lemma 2.3. eφ(e)e+fφ(e)fZ(G).

    Proof. For any mM, there exists a Lie derivation φe of G such that

    φe(m)=φe([e,m])=[φ(e),m]+[e,φe(m)]=φ(e)mmφ(e)+eφe(m)ffφe(m)e.

    Multiplying the above equality from the left by e and from the right by f, we arrive at

    eφ(e)m=mφ(e)f.

    Similarly, for any nN, we have from φe(n)=φe([n,e])=[φe(n),e]+[n,φ(e)] that

    fφ(e)n=nφ(e)e.

    Hence

    eφ(e)e+fφ(e)fZ(G).

    In the sequel, we define ϕ:GG by ϕ(x)=φ(x)[x,eφ(e)ffφ(e)e]. One can verify that ϕ is also a local Lie derivation. Moreover, by Lemma 2.3, we have ϕ(e)=eφ(e)e+fφ(e)fZ(G).

    Lemma 2.4. ϕ(M)M and ϕ(N)N.

    Proof. Let aA,mM and p1 be any idempotent in A. Taking p=p1, x=a and q=e+m in Lemma 2.2, it follows from the facts pxq and pxq can be written as commutators that τ3(pxq)=τ4(pxq)=0, hence

    ϕ(p1a+p1am)=ϕ(p1a)(e+m)+p1ϕ(a+am)p1ϕ(a)(e+m)+(1p1)τ1(p1a+p1am)(fm)p1τ2(a+amp1ap1am)(fm)=ϕ(p1a)e+ϕ(p1a)m+p1ϕ(a)f+p1ϕ(am)p1ϕ(a)m+τ1(p1a)fτ1(p1a)m+p1τ1(p1a)m+p1τ2(ap1a)m. (2.6)

    Multiplying (2.6) from the right by e, we arrive at

    ϕ(p1am)e=p1ϕ(am)e.

    In particular,

    ϕ(p1m)e=p1ϕ(m)e.

    By the above two equations, then

    ϕ(p1p2pnm)e=p1ϕ(p2pnm)e=p1p2pn1ϕ(pnm)e=p1p2pnϕ(m)e

    for any idempotents p1,,pnA. It follows from A=J(A) that

    ϕ(am)e=aϕ(m)e (2.7)

    for all aA,mM. This implies that fϕ(M)e=0.

    The hypothesis (2), (3) and Proposition 1.1 imply that there exist a derivation d:GG and a linear map τ:GZ(G) vanishing on each commutator such that

    ϕ(e+m)=d(e+m)+τ(e+m)=d(e+m)+τ(e). (2.8)

    It follows from (2.5), (2.8) and the fact fϕ(M)e=0 that

    0=fϕ(e+m)e=fd(e)e

    and hence by (2.5) and (2.8) again,

    eϕ(e)e+eϕ(m)e=ed(m)e+eτ(e)e=ed(mf)e+eτ(e)e=md(f)e+eτ(e)e=md(e)e+eτ(e)e=eτ(e)e

    and

    fϕ(e)f+fϕ(m)f=fd(m)f+fτ(e)f=fd(e)m+fτ(e)f=fτ(e)f.

    Then we have from the fact ϕ(e)=eϕ(e)e+fϕ(e)fZ(G) that

    eϕ(m)e+fϕ(m)f=τ(e)ϕ(e)Z(G). (2.9)

    We assume without loss of generality that A does not contain nonzero central ideals. By (2.7) and (2.9) that eϕ(m)e in the central ideal of A. Thus eϕ(M)e=0. So, by (2.9), we get fϕ(M)f=0. Hence, ϕ(M)M.

    With the same argument, we can obtain that ϕ(N)N.

    Lemma 2.5. There exist a linear map h1 from A into Z(G) such that ϕ(a)h1(a)A for all aA and a linear map h2 from B into Z(G) such that ϕ(b)h2(b)B for all bB.

    Proof. Taking m=0 in (2.6), we have

    eϕ(p1a)f=p1ϕ(a)f  and  fϕ(p1a)f=τp1a(p1a)fπB(Z(G)). (2.10)

    In particular,

    eϕ(p1)f=p1ϕ(e)f=0.

    By the two equations above, we obtain

    eϕ(p1p2pn)f=p1ϕ(p2pn)f=p1p2pn1ϕ(pn)f=0

    for all idempotents pi in A. It follows from A=J(A) that eϕ(a)f=0. Similarly, by taking p=e, x=a and q=p1 in Lemma 2.2, we get

    fϕ(ap1)e=fϕ(a)p1.

    This implies that fϕ(a)e=0. So ϕ(a)AB.

    By the hypothesis (2) of Theorem 2.1, there exists a algebra isomorphism η:Z(B)Z(A) such that η(b)bZ(G) for any bZ(B).

    It follows from (2.10) that fϕ(a)fπB(Z(G))=Z(B). We define h1:AZ(G) by h1(a)=η(fϕ(a)f)fϕ(a)f. It is clear that h1 is linear and

    ϕ(a)h1(a)=eϕ(a)e+fϕ(a)fη(fϕ(a)f)fϕ(a)f=eϕ(a)eη(fϕ(a)f)A.

    With the similar argument, we can define a linear map h2:BZ(G) such that ϕ(b)h2(b)B for all bB.

    Now for any xG, we define two linear maps h:GZ(G) and δ:GG by

    h(x)=h1(exe)+h2(fxf)  and  δ(x)=ϕ(x)h(x).

    It is easy to verify that δ(e)=0. Moreover, we have

    δ(A)A, δ(B)B, δ(M)=ϕ(M)M, δ(N)=ϕ(N)N.

    Lemma 2.6. δ is a derivation.

    Proof. We divide the proof into the following three steps.

    Step 1. We first prove that

    δ(p1p2pnm)=δ(p1p2pn)m+p1p2pnδ(m) (2.11)

    for all idempotents pi in A and mM.

    Let aA, mM and p1 be any idempotent in A. Taking p=p1, x=a and q=e+m in (2.2), we have

    ϕ(a+amp1ap1am)=d2(a+amp1ap1am)+τ2(a+amp1ap1am)=d2(a+amp1ap1am)+τ2(ap1a). (2.12)

    It follows from (2.5) and (2.12) that

    0=fd2(ap1a)e=fd2(e(ap1a))e=fd2(e)(ap1a)

    and hence by (2.5) and (2.12) again,

    fϕ(ap1a)f=fd2(amp1am)f+fτ2(ap1a)f=fd2(e)(ap1a)m+fτ2(ap1a)f=fτ2(ap1a)f. (2.13)

    Multiplying (2.6) by f from both sides, we arrive at

    fϕ(p1a)f=fτ1(p1a)f. (2.14)

    By (2.13) and (2.14), then mτ1(p1a)=mϕ(p1a) and

    p1mτ2(ap1a)=p1mϕ(ap1a)=p1mϕ(a)p1mϕ(p1a)=p1mϕ(a)p1mτ1(p1a).

    Hence (2.6) implies that

    δ(p1am)=ϕ(p1am)=ϕ(p1a)m+p1ϕ(am)p1ϕ(a)mmϕ(p1a)+p1mϕ(a)=(δ(p1a)+h(p1a))m+p1δ(am)p1(δ(a)+h(a))mm(δ(p1a)+h(p1a))+p1m(δ(a)+h(a))=δ(p1a)m+p1δ(am)p1δ(a)m. (2.15)

    Taking a=e in (2.15), we have from δ(e)=0 that

    δ(p1m)=δ(p1)m+p1δ(m).

    This shows that (2.11) is true for n=1. One can verify that Eq (2.11) follows easily by induction based on (2.15). It follows from A=J(A) that δ(am)=δ(a)m+aδ(m).

    Similarly, we can get δ(mb)=δ(m)b+mδ(b), δ(mb)=δ(m)b+mδ(b) and δ(na)=δ(n)a+nδ(a).

    Step 2. Let a,aA. For any mM, on one hand, by Step 1, we have

    δ(aam)=δ(a)am+aδ(am)=δ(a)am+aδ(a)m+aaδ(m).

    On the other hand,

    δ(aam)=δ(aa)m+aaδ(m).

    Comparing these two equalities, we have

    (δ(aa)δ(a)aaδ(a))m=0

    for any mM. Since M is a faithful left A-module, we get

    δ(aa)=δ(a)a+aδ(a).

    Similarly, by considering δ(mbb), we can get

    δ(bb)=δ(b)b+bδ(b).

    Step 3. Let m,mM and nN. Taking p=em, x=n+mn and q=em in Lemma 2.2, we have from pxq=pxq=0 that

    0=(em)ϕ(mnmnm+nnm)(em)ϕ(mn+n)(em)(em)τ2(mnnm)(f+m)+(em)τ3(nm)(em)=ϕ(mnm)eϕ(nm)mϕ(mn)+mϕ(nm)+ϕ(mn)mmϕ(n)m+eτ3(nm)eτ3(nm)m. (2.16)

    This implies that

    eϕ(nm)=eτ3(nm)e.

    Then eϕ(nm)m=τ3(nm)m and hence by (2.16),

    δ(mnm)=ϕ(mnm)=mϕ(mn)+mϕ(nm)+ϕ(mn)mmϕ(n)mϕ(nm)m=mh(mn)+mδ(nm)+mh(nm)+δ(mn)m+h(mn)mmδ(n)mh(nm)m=mδ(nm)+δ(mn)mmδ(n)m.

    Replacing m with m+m, we arrive at

    δ(mnm+mnm)=δ(mn)m+mδ(nm)mδ(n)m+δ(mn)m+mδ(nm)mδ(n)m.

    On the other hand, by Steps 1 and 2, we have

    δ(mnm+mnm)=δ(mn)m+mnδ(m)+δ(m)nm+mδ(nm).

    Comparing these two equalities, we have

    (δ(mn)δ(m)nmδ(n))m=m(δ(nm)nδ(m)δ(n)m). (2.17)

    Set

    f(m,n):=δ(mn)δ(m)nmδ(n)

    and

    g(m,n):=δ(nm)nδ(m)δ(n)m.

    We assume without loss of generality that A does not contain nonzero central ideals. For any aA, by (2.17),

    f(m,n)am=amg(m,n)=af(m,n)m.

    which is equivalent to (f(m,n)aaf(m,n))m=0. Since M is a faithful left A-module, we get f(m,n)a=af(m,n). Then

    f(m,n)Z(A).

    By Steps 1 and 2, we have

    f(am,n)=δ(amn)δ(am)namδ(n)=δ(a)mn+aδ(mn)δ(a)mnaδ(m)namδ(n)=af(m,n).

    The above two equalities show that f(m,n) in the central ideal of A and hence

    f(m,n)=0, (2.18)

    that is

    δ(mn)=δ(m)n+mδ(n)

    for all mM,nN. Since M is a faithful right B-module, it follows from (2.17) that

    δ(nm)=nδ(m)+δ(n)m

    for all mM,nN.

    Lemma 2.7. The map h:GZ(G) vanishes on each commutator.

    Proof. Step 1. Let aA, mM, nN and bB, by the definition of h, we have h([a,m])=h([m,b])=h([n,a])=h([b,n])=0.

    Step 2. Let a,aA, we have ϕ([a,a])=eϕ([a,a])e+fϕ([a,a])fAB. On the other hand, Proposition 1.1 implies that ϕ([a,a])=d([a,a])AMN, where d is a derivation. Thus, fϕ([a,a])f=0. This implies that h([a,a])=h1([a,a])=η(fϕ([a,a])f)+fϕ([a,a])f=0.

    Similarly, we can get h([b,b])=0, for all b,bB.

    Step 3. It follows from (2.18) that

    (ϕ(mn)η(fϕ(mn)f)ϕ(m)nmϕ(n))m=m(ϕ(nm)η1(eϕ(nm)e)nϕ(m)ϕ(n)m). (2.19)

    Since fϕ(a)fπB(Z(G)), eϕ(b)eπA(Z(G)), we get that

    mfϕ(mn)f=η(fϕ(mn)f)m,eϕ(nm)em=mη1(eϕ(nm)e).

    It further follows from (2.19) that

    ϕ(mn)mmfϕ(mn)fϕ(m)nmmϕ(n)m=mϕ(nm)+eϕ(nm)m+mnϕ(m)+mϕ(n)m.

    Hence

    (ϕ(mn)eϕ(nm)ϕ(m)nmϕ(n))m=m(ϕ(nm)+fϕ(mn)f+nϕ(m)+ϕ(n)m).

    Using an argument similar to that in the proof of (2.18), we arrive that

    eϕ(mn)eeϕ(nm)ϕ(m)nmϕ(n)=0, (2.20)

    and

    fϕ(nm)f+fϕ(mn)f+nϕ(m)+ϕ(n)m=0.

    By (2.19) and (2.20), we get that eϕ(nm)e=η(fϕ(mn)f). Note that h([m,n])=h1(mn)h2(nm)=η(fϕ(mn)f)+fϕ(mn)feϕ(nm)eη1(eϕ(nm)e), thus h([m,n])=0.

    Therefore it is easily verify that h vanishing on each commutator.

    Proof of Theorem 1.1 By the definition of δ, we have φ(x)=δ(x)+[x,eφ(e)ffφ(e)e]+h(x) for all xA, where δ is a derivation and h is a linear map from A into its center vanishing on each commutator. The proof is complete.

    Let A be a unital algebra and Mk×m(A) be the set of all k×m matrices over A. For n2 and each 2l<n1, the full matrix algebra Mn(A) can be represented as a generalized matrix algebra of the form

    (Ml×l(A)Ml×(nl)(A)M(nl)×l(A)M(nl)×(nl)(A)).

    Corollary 2.8. Let Mn(A) be a full matrix algebra with n4. Then each local Lie derivation φ on Mn(A) is of the form φ=d+τ, where d is a derivation of Mn(A) and τ is a linear map from Mn(A) into its center Z(A)In vanishing on each commutator.

    Proof. It follows from the example (C) of [2] that the matrix algebras Ml(A) and Mnl(A) are generated by their idempotents for 2l<n1. Since Z(Mn(A))=Z(A)In, Z(Ml(A))=Z(A)Il and Z(Mnl(A))=Z(A)Inl, the condition (2) of Theorem 2.1 is satisfied. By [5,Lemma 1], Mk(A) does not contain nonzero central ideals for k2. Hence by Theorem 2.1, every local Lie derivation of Mn(A) is a sum of a derivation and a linear central-valued map vanishing on each commutator.

    Corollary 2.9. Let R be an unital simple algebra with a nontrivial idempotent. If φ:RR is a local Lie derivation, then there exit a derivation d and a linear central map τ vanishing on each commutator, such that φ=d+τ.

    Proof. Let R be an unital simple algebra with a nontrivial idempotent e0 and let f0 denote the idempotent 1e0. Then R can be represented in the so-called Peirce decomposition form

    R=e0Re0+e0Rf0+f0Re0+f0Rf0,

    where e0Re0 and f0Rf0 are subalgebras with unitary element e0 and f0, respectively, e0Rf0 is an (e0Re0,f0Rf0)-bimodule.

    Next, we will show that

    e0xe0e0Rf0={0} implies e0xe0=0

    and

    e0Rf0f0xf0={0} implies f0xf0=0.

    That is e0Rf0 is faithful as an (e0Re0,f0Rf0)-bimodule. Let e=f0+e0Rf0, then e2=e and [e,R]eR(1e)+(1e)Re. Note that

    (1e)Re=(e0e0Rf0)R(f0+e0Rf0)e0Rf0.

    Furthermore, the assumption e0xe0e0Rf0={0} implies

    e0xe0eR(1e)=e0xe0(f0+e0Rf0)R(e0+e0Rf0)={0}

    and then

    e0xe0[e,R]={0}.

    Let r=[e,y] and z,wR. It follows from

    zrw=[e,z[e,r]w][e,z][e,rw][e,zr][e,w]+2[e,z]r[e,w]

    that e0xe0zrw=0. Then

    e0xe0R[e,R]R=0. (2.21)

    It is clear that I=R[e,R]R is a nonzero ideal of R. R is a simple algebra, which implies I=R. By (2.21), e0xe0R=0. Since 1R, we get e0xe0=0. Similarly, we can show that e0Rf0f0xf0={0} implies f0xf0=0. Now, we can conclude that R can be represented as a generalized matrix algebra of the form R=e0Re0+e0Rf0+f0Re0+f0Rf0.

    It follows from the example (A) of [2] that the unital simple algebra with a nontrivial idempotent is generated by its idempotents, the condition (1) of Theorem 2.1 is satisfied. It is clear that e0Re0 and f0Rf0 satisfy the conditions (2) and (3) of Theorem 2.1. Hence by Theorem 2.1, every local Lie derivation of R is the sum of a derivation and a linear central-valued map vanishing on each commutator.

    Let B(H) be the set of bounded linear operators acting on a complex Hilbert space H, and let K(H) be the ideal of compact operators on H. If H is an infinite-dimensional separable Hilbert space, by [12,Theorem 4.1.16], the Calkin algebra B(H)/K(H) is a simple C-algebra.

    Corollary 2.10. If H is an infinite-dimensional separable Hilbert space, then every local Lie derivation of the Calkin algebra B(H)/K(H) is the sum of a derivation and a linear central map vanishing on each commutator.

    In this paper, we investigate local Lie derivations of a certain class of generalized matrix algebras and show that, under certain conditions every local Lie derivation of a generalized matrix algebra is a sum of a derivation and a linear central-valued map vanishing on each commutator. The main result is then applied to full matrix algebras and unital simple algebras with nontrivial idempotents.

    This research was supported by the National Natural Science Foundation of China (No. 11901248). Moreover, the authors express their sincere gratitude to the referee for reading this paper very carefully and specially for valuable suggestions concerning improvement of the manuscript.

    All authors declare no conflicts of interest in this paper.



    [1] D. B. West, Introduction to graph theory, 2 Eds., Upper Saddle River, NJ: Prentice Hall, 2001.
    [2] C. Berge, The theory of graphs and its applications, Paris: Dunod, 1958.
    [3] Y. Caroa, A. Hansbergb, Partial domination-the isolation number of a graph, Filomath, 31 (2017), 3925–3944. https://doi.org/10.2298/FIL1712925C doi: 10.2298/FIL1712925C
    [4] P. Borg, K. Fenech, P. Kaemawichanurat, Isolation of k-cliques, Discrete Math., 343 (2020), 1–7. https://doi.org/10.1016/j.disc.2020.111879 doi: 10.1016/j.disc.2020.111879
    [5] S. Tokunaga, T. Jiarasuksakun, P. Kaemawichanurat, Isolation number of maximal outerplanar graphs, Discrete Appl. Math., 267 (2019), 215–218. https://doi.org/10.1016/j.dam.2019.06.011 doi: 10.1016/j.dam.2019.06.011
    [6] P. Borg, P. Kaemawichanurat, Partial domination of maximal outerplanar graphs, Discrete Appl. Math., 283 (2020), 306–314. https://doi.org/10.1016/j.dam.2020.01.005 doi: 10.1016/j.dam.2020.01.005
    [7] P. Borg, P. Kaemawichanurat, A generalization of the Art Gallery Theorem, arXiv Press, 2020. https://doi.org/10.48550/arXiv.2002.06014
    [8] M. Lemańska, M. J. Souto-Salorio, A. Dapena, F. J. Vazquez-Araujo, Isolation number versus domination number of trees, Mathematics, 9 (2021), 1–10. https://doi.org/10.3390/math9121325 doi: 10.3390/math9121325
    [9] O. Favaron, P. Kaemawichanurat, Inequalities between the Kk-isolation number and the independent Kk-isolation number of a graph, Discrete Appl. Math., 289 (2021), 93–97. https://doi.org/10.1016/j.dam.2020.09.011 doi: 10.1016/j.dam.2020.09.011
    [10] Y. Saad, M. H. Schultz, Topological properties of hypercubes, IEEE Trans. Comput., 37 (1988), 867–872. https://doi.org/10.1109/12.2234 doi: 10.1109/12.2234
    [11] A. K. Somani, O. Peleg, On diagnosability of large fault sets in regular topology-based computer systems, IEEE Trans. Comput., 45 (1996), 892–903. https://doi.org/10.1109/12.536232 doi: 10.1109/12.536232
    [12] S. B. Akers, D. Horel, B. Krishnamurthy, The star graph: an attractive alternative to the n-cube, Proceedings of International Conference on Parallel Processing, 1987,393–400.
    [13] T. W. Haynes, S. T. Hedetniemi, P. J. Slater, Fundamentals of domination in graphs, New York, NY: Marcel Dekker, 1998.
    [14] O. Ore, Theory of graphs, Americal Mathematial Society, Providence, 1962.
    [15] N. Alon, J. H. Spencer, The probabilistic method, 3 Eds., Hoboken, NJ, USA: Wiley, 2008.
  • Reader Comments
  • © 2023 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(1314) PDF downloads(66) Cited by(0)

Figures and Tables

Figures(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog