Research article Topical Sections

Steady-state bifurcation and regularity of nonlinear Burgers equation with mean value constraint

  • Received: 17 February 2025 Revised: 12 April 2025 Accepted: 27 April 2025 Published: 15 May 2025
  • In this paper, we focus on the steady-state bifurcation problem of the nonlinear Burgers equation within a bounded domain, considering both homogeneous Dirichlet boundary conditions and homogeneous Neumann boundary conditions with a mean value constraint. Unlike previous studies, we develop an enhanced turbulence model by incorporating nonlinear higher-order terms (such as u2 and u3) and linear source terms λu into the one-dimensional Burgers equation. Our steady-state bifurcation analysis establishes for the first time how the coupled forward energy cascade and inverse energy transfer mechanisms collectively govern the dynamics of initial flow instability. By combining the spectral theorem for a linear compact operator with the normalized Lyapunov–Schmidt reduction method and the implicit function theorem, we derive the complete criterion for the critical bifurcation condition, the explicit form of the bifurcation solution, and its regularity.

    Citation: Qingming Hao, Wei Chen, Zhigang Pan, Chao Zhu, Yanhua Wang. Steady-state bifurcation and regularity of nonlinear Burgers equation with mean value constraint[J]. Electronic Research Archive, 2025, 33(5): 2972-2988. doi: 10.3934/era.2025130

    Related Papers:

    [1] G. Nandini, M. Venkatachalam, Raúl M. Falcón . On the r-dynamic coloring of subdivision-edge coronas of a path. AIMS Mathematics, 2020, 5(5): 4546-4562. doi: 10.3934/math.2020292
    [2] Bader Alshamary, Milica Anđelić, Edin Dolićanin, Zoran Stanić . Controllable multi-agent systems modeled by graphs with exactly one repeated degree. AIMS Mathematics, 2024, 9(9): 25689-25704. doi: 10.3934/math.20241255
    [3] Jin Cai, Shuangliang Tian, Lizhen Peng . On star and acyclic coloring of generalized lexicographic product of graphs. AIMS Mathematics, 2022, 7(8): 14270-14281. doi: 10.3934/math.2022786
    [4] Shahbaz Ali, Muhammad Khalid Mahmmod, Raúl M. Falcón . A paradigmatic approach to investigate restricted hyper totient graphs. AIMS Mathematics, 2021, 6(4): 3761-3771. doi: 10.3934/math.2021223
    [5] Yang Yang, Yanyan Song, Haifeng Fan, Haiyan Qiao . A note on the generalized Gaussian Estrada index and Gaussian subgraph centrality of graphs. AIMS Mathematics, 2025, 10(2): 2279-2294. doi: 10.3934/math.2025106
    [6] Xiaodi Song, Jianping Li, Jianbin Zhang, Weihua He . Trees with the second-minimal ABC energy. AIMS Mathematics, 2022, 7(10): 18323-18333. doi: 10.3934/math.20221009
    [7] Yunfeng Tang, Huixin Yin, Miaomiao Han . Star edge coloring of $ K_{2, t} $-free planar graphs. AIMS Mathematics, 2023, 8(6): 13154-13161. doi: 10.3934/math.2023664
    [8] Tariq A. Alraqad, Igor Ž. Milovanović, Hicham Saber, Akbar Ali, Jaya P. Mazorodze, Adel A. Attiya . Minimum atom-bond sum-connectivity index of trees with a fixed order and/or number of pendent vertices. AIMS Mathematics, 2024, 9(2): 3707-3721. doi: 10.3934/math.2024182
    [9] Shama Liaqat, Zeeshan Saleem Mufti, Yilun Shang . Newly defined fuzzy Misbalance Prodeg Index with application in multi-criteria decision-making. AIMS Mathematics, 2024, 9(8): 20193-20220. doi: 10.3934/math.2024984
    [10] Shumin Zhang, Tianxia Jia, Minhui Li . Partial domination of network modelling. AIMS Mathematics, 2023, 8(10): 24225-24232. doi: 10.3934/math.20231235
  • In this paper, we focus on the steady-state bifurcation problem of the nonlinear Burgers equation within a bounded domain, considering both homogeneous Dirichlet boundary conditions and homogeneous Neumann boundary conditions with a mean value constraint. Unlike previous studies, we develop an enhanced turbulence model by incorporating nonlinear higher-order terms (such as u2 and u3) and linear source terms λu into the one-dimensional Burgers equation. Our steady-state bifurcation analysis establishes for the first time how the coupled forward energy cascade and inverse energy transfer mechanisms collectively govern the dynamics of initial flow instability. By combining the spectral theorem for a linear compact operator with the normalized Lyapunov–Schmidt reduction method and the implicit function theorem, we derive the complete criterion for the critical bifurcation condition, the explicit form of the bifurcation solution, and its regularity.



    Let G be a simple graph with vertex set V(G)={1,2,,n} and edge set E(G). The adjacency matrix of G is an n×n matrix A(G)=(aij), where aij=1 if vertex i is adjacency to vertex j, and 0 otherwise. We use the notation ij (or ijE(G)) to indicate that vertices i,j are adjacent in G. The adjacency eigenvalues of G are just the eigenvalues of A(G). For an eigenvalue μ, let E(μ) be the eigenspace {xRn|A(G)x=μx }. For more details on graph spectra, see [4].

    Let μ be an eigenvalue of G with multiplicity k. A star set for μ in G is a subset X of V(G) such that |X|=k and μ is not an eigenvalue of GX, where GX is the subetaaph of G induced by ¯X=V(G)X. In this situation H=GX is called a star complement for μ. Star sets and star complements exist for any eigenvalue of a graph, and they need not to be unique. The basic properties of star sets are established in Chapter 7 of [5] and Chapter 5 of [7].

    There is another equivalent geometric definition for star sets and star complements. Let G be a graph with vertex set V(G)={1,,n} and adjacency matrix A. Let {e1,,en} be the standard orthonormal basis of Rn and P be the matrix which represents the orthogonal projection of Rn onto the eigenspace E(μ)={xRn:A(G)x=μx } of A with respect to {e1,,en}. Since E(μ) is spanned by the vectors Pej(j=1,,n), there exists XV(G) such that the vectors Pej(jX) form a basis for E(μ). Such a subset X of V(G) is called a star set for μ in G. In this situation H=GX is called a star complement for μ ([5,7]).

    For any graph G of order n with distinct eigenvalues λ1,,λm, there exists a partition V(G)=V1Vm such that Vi is a star set for eigenvalue λi (i=1,,m). Such a partition is called a star partition of G. For any graph G, there exists at least one star partition ([5]). Each star partition determines a basis for Rn consisting of eigenvectors of an adjacency matrix. It provides a strong link between graph structure and linear algebra.

    In [7], it was proved that if YX then XY is a star set for μ in GY. Thus the induced subetaaph GY also has H as a star complement for μ. If G has H as a star complement for μ, and G is not a proper induced subetaaph of some other graph with H as a star complement for μ, then G is a maximal graph with H as a star complement for μ, or it is an H- maximal graph for μ. Accordingly, in determining all the graphs with H as a star complement for μ, it suffices to describe the maximal graphs which arise.

    There are a lot of literatures about using star complements to construct and characterize certain graphs. Regular graphs with a prescribed graph such as K1,s, K2,s, K1,1,t, K1,1,1,t, ¯sK1Kt, Pt, Kr,r,r or Kr,s+tK1 as a star complement are discussed in the literature ([1,8,10,12,15,16,17,18]). Bipartite graphs with complete bipartite Kr,s as a star complement is considered by Rowlinson in 2014 ([9]). Maximal graphs with a prescribed graph such as Sm, Km, Sm,n, K2,5, K1,1,t, Ct, Pt, ¯L(Rt), ¯L(Qt) or unicyclic graph as a star complement for given eigenvalue are well studied in the literature ([2,3,6,11,13,14,17,18]). In this paper, we introduce some results on the theory of star complements in Section 2, characterize the maximal graphs with the bipartite graph K2,s as a star complement for eigenvalues μ=2,1 in Section 3, and study the cases of other eigenvalues for further research in Section 4.

    In this section, we introduce some results of star sets and star complements that will be required in the sequel. The following fundamental result combines the Reconstruction Theorem ([5,Theorem 7.4.1]) with its converse ([5,Theorem 7.4.4]).

    Theorem 2.1. ([5]) Let X be a set of vertices in the graph G. Suppose that G has adjacency matrix

    (AXBTBC),

    where AX is the adjacency matrix of the subetaaph induced by X. Then X is a star set for μ in G if and only if μ is not an eigenvalue of C and

    μIAX=BT(μIC)1B. (2.1)

    Note that if X is a star set for μ, then the corresponding star complement H(=GX) has adjacency matrix C, and (2.1) tells us that G is determined by μ, H and the H-neighbourhood of vertices in X, where the H-neighbourhood of vertex uX, denoted by NH(u), is defined as NH(u)={vvu,vV(H)}.

    It is usually convenient to apply (2.1) in the form m(μ)(μIAX)=BTm(μ)(μIC)1B, where m(x) is the minimal polynomial of C. This is because m(μ)(μIC)1 is given explicitly as follows.

    Proposition 2.2. ([6], Proposition 0.2) Let C be a square matrix with minimal polynomial

    m(x)=xd+1+cdxd+cd1xd1++c1x+c0.

    If μ is not an eigenvalue of C, then

    m(μ)(μIC)1=adCd+ad1Cd1++a1C+a0I,

    where ad=1 and for 0<id, adi=μi+cdμi1+cd1μi2++cdi+1.

    In order to find all the graphs with a prescribed star complement for μ, we need to find, for given μ and C where μ is not an eigenvalue of C, all AX and B satisfying μIAX=BT(μIC)1B. For any x,yRq, where q=|V(H)|, let

    x,y=xT(μIC)1y. (2.2)

    Let bu be the column of B for any uX. By Theorem 2.1, we have

    Corollary 2.3. ([7], Corollary 5.1.8) Suppose that μ is not an eigenvalue of the graph H, where |V(H)|=q. There exists a graph G with a star set X={u1,u2,,uk} for μ such that GX=H if and only if there exist (0,1)-vectors bu1,bu2,,buk in Rq such that

    (1) bu,bu=μ for all uX, and

    (2) bu,bv={1,uv0,uv for all pairs u,v in X.

    Given a graph H, a subset U of V(H) and a vertex u not in V(H), denote by H(U) the graph obtained from H by joining u to all vertices of U. We will say that u (resp. U,H(U)) is a good vertex (resp. good set, good extension) for μ and H, if μ is an eigenvalue of H(U) but is not an eigenvalue of H. By Theorem 2.1, a vertex u, or a subset U, or a graph H(U) is good if and only if bu,bu=μ, and two vertices u and v are good partners if and only if bu,bv{1,0}. It follows that any vertex set X in which all vertices are good and any two vertices are good partners, gives rise to a extensional graph, say G. In this situation, X can be viewed as a star set for μ with H as the corresponding star complement.

    In view of the two conditions in the above corollary, we have

    Lemma 2.4. ([5]) Let X be a star set for μ in G, and H=GX.

    (1) If μ0, then V(H) is a dominating set for G, that is, the H-neighbourhood of any vertex in X are non-empty;

    (2) If μ{1,0}, then V(H) is a location-dominating set for G, that is, the H-neighbourhood of distinct vertices in X are distinct and non-empty.

    It follows from (2) of Lemma 2.4 that there are only finitely maximal graphs with a prescribed star complement for μ{1,0}.

    Let HKt,s (st1), (R,S) be the bipartition of the graph Kt,s with R={i1,i2,,it}, S={j1,j2,,js}. A vertex uX is said to be of type (a,b) if it has a neighbours in R and b neighbours in S, thus (a,b)(0,0) and 0at, 0bs. The following (2.3) and (2.4) have been given in [11], and now we express them in the form of a necessary and sufficient condition and give a proof.

    Proposition 2.5. ([11]) Let HKt,s (st1), R,S defined as above, and μ be not an eigenvalue of the graph H. Then G is a graph with H as a star complement for μ if and only if the vertex set X such that GX=H satisfies the following two conditions:

    (1) for any uX of type (a,b), we have

    (μ2ts)(a+b)+sa2+tb2+2abμ=μ2(μ2ts); (2.3)

    (2) for any two distinct vertices u,vX of type (a,b), (c,d), respectively, ρuv=|NH(u)NH(v)|, and auv={1,uv,0,uv, we have

    (μ2ts)ρuv+acs+bdt+μ(ad+bc)=μ(μ2ts)auv. (2.4)

    Proof. Let C be the adjacency matrix of H. Then C=(Ot×tJt×sJs×tOs×s), where Ot×t and Jt×s denote all-0 matrix of size t×t and all-1 matrix of size t×s, respectively. Thus we have C2=(sJt×tOt×sOs×ttJs×s), and C has minimal polynomial m(x)=x(x2ts).

    Since μ is not an eigenvalue of C, we have μ0 and μ2ts. From Proposition 2.2, we have

    m(μ)(μIC)1=C2+μC+(μ2ts)I. (2.5)

    Let bu=(bTR,bTS)TRt+s, where bR and bS are vectors of size t×1 and size s×1, respectively. For any uX of type (a,b), by bu,bu=μ, (2.2) and (2.5), we have

    μ2(μ2ts)=μm(μ)=bu,bum(μ)=bTum(μ)(μIC)1bu=bTu(C2+μC)bu+(μ2ts)bTubu=(bTR,bTS)(sJt×tμJt×sμJs×ttJs×s)(bRbS)+(μ2ts)(bTR,bTS)(bRbS)=sbTRJt×tbR+tbTSJs×sbS+μbTSJs×tbR+μbTRJt×sbS+(μ2ts)(bTRbR+bTSbS)=sa2+tb2+2abμ+(μ2ts)(a+b).

    Thus we obtain (2.3).

    Similarly, for any two distinct vertices u,vX of type (a,b), (c,d), respectively, ρuv=|NH(u)NH(v)|, by bu,bv=auv, (2.2) and (2.5), we have

    μ(μ2ts)auv=m(μ)auv=m(μ)bu,bv=bTum(μ)(μIC)1bv=bTu(C2+μC)bv+(μ2ts)bTubv=acs+bdt+μ(ad+bc)+(μ2ts)ρuv.

    Thus we obtain (2.4).

    By Corollary 2.3, we complete the proof.

    The following lemma is important for us to establish the location of an eigenvalue of a graph. It is a natural extension of Interlacing Theorem.

    Lemma 2.6. ([8]) Given a graph of order n with eigenvalue μ of multiplicity k1, let H be a star complement for μ in G. Let λr+1(H)<μ<λr(H) for some 0rnk, where λ0(H)=. Then λr+1(G)==λr+k(G)=μ.

    As far as we know, researchers can use the star complement technique to construct graphs with certain spectral properties. In fact, we are usually interested in the eigenvalues with large multiplicity in graphs.

    Maximal graphs with a prescribed graph such as Sm, Km, Sm,n, K2,5, K1,1,t, Ct, Pt, ¯L(Rt) or ¯L(Qt) as a star complement for given eigenvalue μ is well studied in the literature([2,3,6,11,13,17,18]), see Table 1.

    Table 1.  The maximal graphs have been characterized for given star complement and eigenvalue.
    the star complement μ reference
    Ct 2 [2]
    Pt 2 [3]
    ¯L(Rt), ¯L(Qt) 1 [6]
    L(Rt), L(Qt) 2 [6]
    K2,5 1 [11]
    Sm=K1,m1, Km, Sm,n 1 [13]
    K1,1,t (t8,9) 1 [17]
    Sm=K1,m1 2 [18]

     | Show Table
    DownLoad: CSV

    In this section, we study the maximal graphs with K2,s as a star complement for μ=2 and μ=1.

    Let μ=2, H=GXK2,s and (R,S) be the bipartition of the graph HK2,s. In this subsection, we prove that K2,1, K2,10, K2,11, K2,12, K2,18, K2,20 and K2,27 are the only graphs among K2,s which can be star complements for μ=2, and then we take K2,10 as an example to characterize the maximal graphs with it as a star complement for 2.

    Let μ=2. Since 2 is not an eigenvalue of HK2,s, we have μ22s, and then s2. Thus we have the following Proposition.

    Proposition 3.1. Let HK2,s be a star complements for μ=2. Then s2.

    Theorem 3.2. The graphs K2,1, K2,10, K2,11, K2,12, K2,18, K2,20 and K2,27 are the only graphs among K2,s which can be star complements for μ=2.

    Proof. Let uX be a vertex of type (a,b) which means that it has a neighbours in R and b neighbours in S. Then (a,b)(0,0) and a{0,1,2}, 0bs.

    Case 1: a=0.

    Then by (2.3), we have

    b2+(2s)b4(2s)=0. (3.1)

    Since b is an integer, then (2s)2+4×4(2s)=(s10)264 must be a perfect square, so s{2,18,20,27}. Thus only K2,18, K2,20 and K2,27 can be star complements for 2 by Proposition 3.1.

    Case 2: a=1.

    Then by (2.3), we have

    2b22sb+7s12=0. (3.2)

    Since b is an integer, then (2s)24×2×(7s12)=(2s14)2100 must be a perfect square, so s{2,12,20}. Thus only K2,12 and K2,20 can be star complements for 2 by Proposition 3.1.

    Case 3: a=2.

    Then by (2.3), we have

    b2(2+s)b+4(s1)=0. (3.3)

    Since b is an integer, then (2+s)24×4×(s1)=(s6)216 must be a perfect square, so s{1,2,10,11}. Thus only K2,1, K2,10 and K2,11 can be star complements for 2 by Proposition 3.1.

    Combining the above three cases, we complete the proof.

    Recalling the definitions of a good vertex u, a good set U and a good extension H(U) in Section 2, we now proceed to identify all good sets U, i.e., to identify the sets U for which graph H(U) has 2 as an eigenvalue, where H{K2,1,K2,10,K2,11,K2,12,K2,18,K2,20,K2,27}. We denote the a-subset of R by Ra and the b-subset of S by Sb, where (R,S) is the bipartition of the graph K2,s.

    Lemma 3.3. For μ=2, we have

    (1) K2,1(U) is good if and only if U=R2;

    (2) K2,10(U) is good if and only if U=R2S6;

    (3) K2,11(U) is good if and only if U{R2S5,R2S8};

    (4) K2,12(U) is good if and only if U=R1S6;

    (5) K2,18(U) is good if and only if U=S8;

    (6) K2,20(U) is good if and only if U{S6,S12,R1S4,R1S16};

    (7) K2,27(U) is good if and only if U{S5,S20}.

    Proof. The integral solutions of (3.1), (3.2) and (3.3) are shown in Table 2.

    Table 2.  The integral solutions of (3.1), (3.2) and (3.3).
    a (s,b)
    0 (27,5),(27,20),(20,6),(20,12),(18,8)
    1 (20,4),(20,16),(12,6)
    2 (11,5),(11,8),(10,6),(1,0)

     | Show Table
    DownLoad: CSV

    By the definitions of a good vertex u, a good set U and a good extension H(U) in Section 2, Theorem 2.1, Corollary 2.3 and Table 2, we complete the proof.

    Theorem 3.4. A Graph G is a maximal graph with HK2,s as a star complement for μ=2 if and only if the following two conditions hold:

    (1) H{K2,1,K2,10,K2,11,K2,12,K2,18,K2,20,K2,27}.

    (2) The vertex set X such that GX=H satisfies the following (i) (iii):

    (i) for any uX, u is a good vertex, say, NH(u)=U satisfies Lemma 3.3;

    (ii) for any two distinct vertices u,vX of type (a,b), (c,d), respectively, u,v are good partners, say, ρuv=|NH(u)NH(v)| and auv={1,uv0,uv satisfy Table 3.

    Table 3.  Adjacency relations of any two vertices u,vX.
    H (a,b) (c,d) auv ρuv
    K2,1 (2,0) (2,0) 1 0
    K2,10 (2,6) (2,6) 0 4
    1 6
    K2,11 (2,5) (2,5) 0 3
    1 5
    (2,8) (2,8) 0 6
    1 8
    (2,5) (2,8) 0 4
    1 6
    K2,12 (1,6) (1,6) 0 3
    1 5
    K2,18 (0,8) (0,8) 0 4
    1 6
    K2,27 (0,5) (0,5) 0 1
    1 3
    (0,20) (0,20) 0 16
    1 18
    (0,5) (0,20) 0 4
    K2,20 (0,6) (0,6) 0 2
    1 4
    (0,12) (0,12) 0 8
    1 10
    (1,4) (1,4) 0 1
    1 3
    (1,16) (1,16) 0 13
    1 15
    (0,6) (0,12) 0 4
    1 6
    (0,6) (1,4) 0 1
    1 3
    (0,12) (1,4) 0 2
    1 4
    (0,12) (1,16) 0 10
    1 12
    (1,4) (1,16) 0 3
    1 5
    (0,6) (1,16) 0 5

     | Show Table
    DownLoad: CSV

    (iii) XU={U=NH(u)|uX} is a maximal family, say, there is no other family XU satisfies (i),(ii), and XUXU.

    Proof. By Proposition 2.5, Theorem 3.2, Lemma 3.3 and the definition of maximal, we only need to show (2.4) is equivalent to (ⅱ).

    Since the proof is similar, we only show the case of H=K2,10. We note that μ=2,t=2,s=10,a=c=2,b=d=6, then

    (2.4)2auvρuv+4=0ρuv={6,ifuv,4,ifuv.

    By similar way, we can complete the proof.

    Now we want to characterize all the maximal graphs with the star complements H given in Theorem 3.2 for 2. As mentioned in Reference [7], we can invoke an algorithm to find the maximal family by using a computer and thus find the maximal graphs. Now we're just focusing on the cases of H{K2,1,K2,10}.

    Since K2,1S3, we have the following results by Theorem 3.2 in [18].

    Theorem 3.5. ([18]) The 4-cycle C4 is the unique maximal graph with K2,1 as a star complement for μ=2 which is the smallest eigenvalue of C4.

    Let (R,S) be the bipartition of the graph H=K2,10 with R={1,2} and S={1,2,,10}. Now we characterize the maximal graph G with the star complement K2,10 for μ=2.

    Let X={u1,,uk} be the star set for μ=2, and XU={U1,,Uk} be the collection of good sets, where Ui is the corresponding good set of vertex ui. Then for each 1ik, vertex ui is of type (2,6) and Ui={1,2}Vi by Theorem 3.4, where Vi is a 6-subset of S. In addition, each pair of sets in XV={Vi1ik} are compatible, which means |ViVj|=2 or 4 for any 1i<jk by ρuv=4 or 6 from Table 3.

    Therefore, finding the maximal graphs with K2,10 as a star complement for μ=2 is equivalent to finding the maximal family of the 6-subsets of the 10-set S such that any two 6-sets in the family has 2 or 4 common elements. If there exists a bijection of elements between two families, we say these two families are isomorphic.

    Lemma 3.6. Let S={1,2,,10} be a 10-set. Then there exist exactly two non-isomorphic maximal families of 6-subsets of S such that any two 6-sets in each family has 2 or 4 common elements.

    Proof. Let XV={V1,V2,,VkViS,|Vi|=6,1ik} be a maximal family such that each pair of sets in XV are compatible, which means |ViVj|=2 or 4 for any 1i<jk.

    Claim 1: Let Vi,VjXV. If |ViVj|=4, then Vp=(ViVj)(S(ViVj))XV.

    Without loss of generality, we can assume that Vi={1,2,3,4,5,6} and Vj={1,2,3,4,7,8}, then Vp={1,2,3,4,9,10}. Assume to the contrary, VpXV. Then there exist a set VlXV such that |VpVl|=3 or 5 since any two 6-subsets has at least 2 common elements in 10-set S and XV is maximal.

    If |VpVl|=3, then Vl contains three elements in SVp={5,6,7,8} and |{1,2,3,4}Vl|1, which implies one of |ViVl| and |VjVl| must be odd, it is a contradiction with |ViVl|=2 or 4, |VjVl|=2 or 4. If |VpVl|=5, we get a contradiction in a similar way. Thus VpXV. Claim 1 holds.

    Claim 2: There are at least two sets in XV having exactly two elements in common.

    Assume the contrary. Then any two sets in XV have exactly four elements in common. Without loss of generality, we can assume V1={1,2,3,4,5,6}, V2={1,2,3,4,7,8}XV. Then by Claim 1, we have V3={1,2,3,4,9,10}XV.

    The maximality of XV implies that there exist other sets in XV. Let V4={a,b,c,d,e,f}XV, where 1a<b<c<d<e<f10. Then a4, otherwise V4={5,6,7,8,9,10} and |V4Vi|=2 for any i{1,2,3}, it is a contradiction.

    If 1a<b<c<d4<e<f10, then there exist some i{1,2,3} such that |V4Vi|5, it is a contradiction; if 1a<b4<c<d<e<f10, then there exist some i{1,2,3} such that |V4Vi|3, it is a contradiction; if 1a4<b<c<d<e<f10, then for any i{1,2,3}, we have |V4Vi|3, it is a contradiction. Thus 1a<b<c4<d<e<f10.

    Clearly, d{5,6},e{7,8},f{9,10} by the fact |V4Vi|=4 for any i{1,2,3} and 1a<b<c4<d<e<f10. Without loss of generality, we take V4={a,b,c,5,7,9}XV. By Claim 1, we have V5={a,b,c,5,8,10}, V6={a,b,c,6,7,10}, V7={a,b,c,6,8,9}XV, and we can check that any other set V8={a,b,c,d,e,f} such that 1a<b<c4<d<e<f10 cannot have four elements in common with each of V1,,V7, so XV={V1,V2,,V7}.

    Let V9={a,b,g,5,7,10}, g{1,2,3,4}{a,b,c}. We can check that |V9Vi|=4 for 1i6 and |V9V7|=2. Therefore, XV{V9}XV, it implies a contradiction with the maximality of XV. Thus there exists at least two sets in XV having exactly two elements in common. Claim 2 holds.

    Claim 3: Let Vi,VjXV. If ViVj={a,b}, then for any VpXV, {a,b}Vp or {a,b}Vp=ϕ.

    Without loss of generality, we can assume V1={1,2,3,4,5,6}, V2={1,2,7,8,9,10}XV. If V3XV and |{1,2}V3|=1, then one of |V1V3| and |V2V3| must be odd, it is a contradiction. Then Claim 3 holds.

    Now we study the structure of XV. Due to Claim 2, we can assume V1={1,2,3,4,5,6}, V2={1,2,7,8,9,10}XV. By Claim 3, we consider the following two cases.

    Case 1: Each set in XV contains elements 1 and 2.

    In this case, every set in XV has the form {1,2,c,d,e,f} where 3c<d6 and 7e<f10 by |ViVj|=2 or 4 for any Vi,VjXV, and the symmetry between {3,4,5,6} and {7,8,9,10}. If we divide 3, 4, 5, and 6 into two groups of two numbers, there are three ways to divide them, say, {3,4} and {5,6}, {3,5} and {4,6}, {3,6} and {4,5}. For example, we take divide 3, 4, 5, and 6 into {3,4} and {5,6}, then in the sense of isomorphism, we have 2×2=4 ways to take {c,d,e,f}, say, {3,4,7,8}, {3,4,9,10}, {5,6,7,8}, {5,6,9,10}. Then |XV|1+1+3×4=14.

    Without loss of generality, we can assume that V3={1,2,3,4,7,8}, V4={1,2,3,5,7,9}, V5={1,2,3,6,7,10}. By Claim 1, we have V6={1,2,3,4,9,10}XV, V7={1,2,5,6,7,8}XV, V8={1,2,3,5,8,10}XV, V9={1,2,4,6,7,9}XV, V10={1,2,3,6,8,9}XV, V11={1,2,5,6,9,10}XV, V12={1,2,4,6,8,10}XV, V13={1,2,4,5,8,9}XV, V14={1,2,4,5,7,10}XV.

    It is easy to check that every pair Vi,Vj (1i,j14) are compatible. Then |XV|=14, say, XV is a maximal family.

    Case 2: There exists a set in XV does not contain elements 1 and 2.

    Let V3XV be a set that does not contain elements 1 and 2. Then it must have two elements in common with one of sets V1, V2 and four elements in common with the other. In the sense of isomorphism, noting that the symmetry between {3,4,5,6} and {7,8,9,10}, {3,4} and {5,6}, {7,8} and {9,10}, we can assume V3={3,4,5,6,7,8}.

    Subcase 2.1: V4={3,4,7,8,9,10}XV.

    Then by Claim 1, we have V5={3,4,5,6,9,10}, V6={1,2,3,4,9,10}, V7={5,6,7,8,9,10}, V8={1,2,5,6,9,10}, V9={1,2,5,6,7,8}, V10={1,2,3,4,7,8}XV.

    It is easy to check that any two set of {V1,V2,V3,,V10} are compatible, say, having two or four common elements. Thus {V1,V2,V3,,V10}XV.

    Now we show XV={V1,V2,V3,,V10}. Otherwise, we can assume that there exists V11XV. For the five sets W1={1,2},W2={3,4},W3={5,6},W4={7,8},W5={9,10}, any 6-set formed by the union of three 2-sets from {W1,W2,W3,W4,W5} is in {V1,V2,V3,,V10}. Then there exist at least two sets Wi,Wj such that |WiV11|=|WjV11|=1 for 2i<j5 by Claim 3.

    Without loss of generality, we assume that 3,5V11 and 4,6V11, then one of |V11V6|, |V11V7| and |V11V10| is odd, it is a contradiction because any two set of XV are compatible.

    Combining the above arguments, we have XV={V1,V2,V3,,V10} is another maximal family.

    Subcase 2.2: V4={3,4,7,8,9,10}XV.

    Then there exist a set VmXV such that |VmV4|=3 or 5 since any two 6-subsets has at least 2 common elements in 10-set S.

    If |VmV4|=3, then Vm contains three elements in SV4={1,2,5,6}. By Claim 3, {1,2}Vm and thus |Vm{5,6}|=1. Since |Vm|=6, |VmV2|=2 or 4, we have |Vm{7,8,9,10}|=2, |Vm{3,4}|=1. Without loss of generality, we assume that Vm={1,2,3,5,7,8}. By Claim 1 and V1,V2,V3,Vm, we can obtain a new maximal family X1V.

    Let ψ be a mapping from S to S such that ψ(3)=6, ψ(6)=3 and ψ(a)=a for any aS{3,6}. Then ψ is a bijection, thus X1V and {V1,V2,V3,,V10} are isomorphic.

    If |VmV4|=5, then Vm contains one elements in SV4={1,2,5,6}. By Claim 3, {1,2}Vm and thus |Vm{5,6}|=1. Since |Vm|=6, |VmV2|=2 or 4, we have |Vm{7,8,9,10}|=4, |Vm{3,4}|=1. Without loss of generality, we assume that Vm={3,5,7,8,9,10}. By Claim 1 and V1,V2,V3,Vm, we can obtain another new maximal family X2V.

    Let φ be a mapping from S to S such that φ(4)=5, φ(5)=4 and φ(a)=a for any aS{4,5}. Then φ is a bijection, X2V and {V1,V2,V3,,V10} are isomorphic.

    Combining the above two subcases, we obtain another maximal family XV={V1,V2,V3,,V10}.

    By Case 1 and Case 2, we complete the proof.

    Let (R,S) be the bipartition of H=K2,10 with R={1,2}, S={1,2,,10}. We define two graphs G1,G2 as follows:

    (1) X={u1,u2,,u14}, V(G1)=V(H)X, and G1[X]=K14u1u2u3u11u4u12u5u13u6u7u8u9u10u14, and NH(u1)={1,2,1,2,3,4,5,6}, NH(u2)={1,2,1,2,7,8,9,10}, NH(u3)={1,2,1,2,3,4,7,8}, NH(u4)={1,2,1,2,3,5,7,9}, NH(u5)={1,2,1,2,3,6,7,10}, NH(u6)={1,2,1,2,3,4,9,10}, NH(u7)={1,2,1,2,5,6,7,8}, NH(u8)={1,2,1,2,3,5,8,10}, NH(u9)={1,2,1,2,4,6,7,9}, NH(u10)={1,2,1,2,3,6,8,9}, NH(u11)={1,2,1,2,5,6,9,10}, NH(u12)={1,2,1,2,4,6,8,10}, NH(u13)={1,2,1,2,4,5,8,9}, NH(u14)={1,2,1,2,4,5,7,10}.

    (2) X={v1,v2,,v10}, V(G2)=V(H)X, and G2[X]=K10v1v2v1v4v1v7v2v3v2v5v3v6v3v8v4v8 - v4v9v5v9v5v10v6v7v6v9v7v10v8v10, and NH(v1)={1,2,1,2,3,4,5,6}, NH(v2)={1,2,1,2,7,8,9,10}, NH(v3)={1,2,3,4,5,6,7,8}, NH(v4)={1,2,3,4,7,8,9,10}, NH(v5)={1,2,3,4,5,6,9,10}, NH(v6)={1,2,1,2,3,4,9,10}, NH(v7)={1,2,5,6,7,8,9,10}, NH(v8)={1,2,1,2,5,6,9,10}, NH(v9)={1,2,1,2,5,6,7,8}, NH(v10)={1,2,1,2,3,4,7,8}.

    Clearly, G1 has 26 vertices, where 2 vertices of degree 24, 14 vertices of degree 20, 2 vertices of degree 16, and 8 vertices of degree 9, its spectrum is [4.70096,214,02,0.66031,27,18.04065]; G2 has 22 vertices, where 2 vertices of degree 20, 10 vertices of degree 14, and 10 vertices of degree 8, its spectrum is [4.71780,210,06,34,12.71780].

    By Theorem 3.4 and Lemma 3.6, we can conclude that there are exactly two non-isomorphic maximal graphs with K2,10 as a star complement for μ=2.

    Theorem 3.7. Let G be a graph with K2,10 as a star complement for μ=2. Then G is maximal if and only if GG1 or GG2.

    Remark 3.8. Let S={1,2,,10}, S1,S2,,S210 be all the 6-subsets of S. We construct a graph GX of order 210 with V(GX)={v1,v2,,v210} and vivjE(GX) if |SiSj|=2 or 4.

    Then finding the maximal graphs with HK2,10 as a star complement for μ=2 is equivalent to finding the maximal family of the 6-subsets of a 10-set such that the intersection of any two 6-sets in the family has 2 or 4 elements, is equivalent to finding all maximal cliques in GX.

    By Lemma 3.6, GX has two maximal cliques in the sense of isomorphism, one is order 14 and the other is order 10. When H{K2,11,K2,12,K2,18,K2,20,K2,27}, we can study the maximal graphs by a similar way, and now we ignore the characterization.

    Finally, it is obvious to obtain the following result.

    Corollary 3.9. K2,10, K2,11, K2,12, K2,18, K2,20 and K2,27 are the only graphs among K2,s which can be star complements for the second smallest eigenvalue λn1=2.

    Proof. Let G be a graph of order n with H as a star complement for an eigenvalue 2 of multiplicity ns2, where HK2,s (s>2). We know that K2,s has spectrum: 0s, 2s, 2s. If s>2, then

    λs+2(H)=2s<2<0=λs+1(H).

    By Lemma 2.6, we have λs+2(G)=λs+3(G)==λn1(G)=2.

    In this subsection, we study the maximal graphs with K2,s as a star complement for μ=1. Clearly, 1 is not an eigenvalue of K2,s. Then by (2.3), we have

    Theorem 3.10. K2,5 and K2,13 are the only two graphs among K2,s which can be star complements for μ=1.

    Proof. Let uX be a vertex of type (a,b) which means that it has a neighbours in R and b neighbours in S. Then (a,b)(0,0) and a{0,1,2}, 0bs.

    If a=0, then by (2.3), we have

    2b2+(12s)b+2s1=0. (3.4)

    Since b is an integer, then (12s)28×(2s1)=(2s5)216 must be a perfect square, so s=5. Therefore, only K2,5 can be a star complement for μ=1.

    If a=1, then by (2.3), we have

    2b2+(32s)b+s=0. (3.5)

    Since b is an integer, then (32s)28s=(2s5)216 must be a perfect square, so s=5. Therefore, only K2,5 can be a star complement for μ=1.

    If a=2, then by (2.3), we have

    2b2+(52s)b+1+2s=0. (3.6)

    Since b is an integer, then (52s)28×(1+2s)=(2s9)264 must be a perfect square, so s=13. Therefore, only K2,13 can be a star complement for μ=1.

    Combining the above arguments, we complete the proof.

    Recalling the definitions of good vertex u, good set U and good extension H(U) in Section 2, we now proceed to identify all good sets U, i.e., to identify the sets U for which graph H(U) has 1 as an eigenvalue, where H{K2,5,K2,13}. We denote the a-subset of R by Ra and the b-subset of S by Sb, where (R,S) is the bipartition of the graph K2,s.

    Lemma 3.11. For μ=1, we have

    (1) K2,5(U) is good if and only if U{S3,R1S1};

    (2) K2,13(U) is good if and only if U=R2S9.

    Proof. The integral solutions of (3.4), (3.5) and (3.6) are shown in Table 4.

    Table 4.  The integral solutions of (3.4), (3.5) and (3.6).
    a (s,b)
    0 (5,3)
    1 (5,1)
    2 (13,9)

     | Show Table
    DownLoad: CSV

    By the definitions of good vertex u, good set U and good extension H(U) in Section 2, Theorem 2.1, Corollary 2.3 and Table 4, we complete the proof.

    When HK2,5, [11] characterized the unique maximal graph with K2,5 as a star complement for μ=1.

    Theorem 3.12. ([11]) The complement of Schl¨afli graph is the unique maximal graph with K2,5 as a star complement for μ=1.

    Theorem 3.13. Let HK2,13. Then G is the maximal graph with H as a star complement for μ=1 if and only if the vertex set X such that GX=H satisfies the following three conditions:

    (1) for any uX, u is a good vertex, say, NH(u)=U=R2S9;

    (2) for any two distinct vertices u,vX of type (a,b), (c,d), respectively, u,v are good partners, say, ρuv=|NH(u)NH(v)|={9,ifuv,10,ifuv.

    (3) XU={U=NH(u)|uX} is a maximal family, say, there is no other family XU satisfies (1),(2), and XUXU.

    Proof. By Proposition 2.5, Theorem 3.2, Lemma 3.11 and the definition of maximal, we only need to show (2.4) is equivalent to (2).

    We note that μ=1,t=2,s=13,a=c=2,b=d=9, then

    (2.4)auv+ρuv10=0ρuv={9,ifuv,10,ifuv.

    Remark 3.14. Let (R,S) be the bipartition of the graph H=K2,13 with R={1,2} and S={1,2,,13}, X={u1,,uk} be the star set for μ=1, and XU={U1,,Uk} be the collection of good sets, where Ui is the corresponding good set of vertex ui. Then for each 1ik, vertex ui is of type (2,9) and Ui={1,2}Fi by Lemma 3.11, where Fi is a 9-subset of S. In addition, each pair of sets in F={Fi1ik} are compatible, which means |FiFj|=7 or 8 for any 1i<jk by ρuv=9 or 10 from Theorem 3.13.

    Therefore, finding the maximal graphs with K2,13 as a star complement for μ=1 is equivalent to finding the non-isomorphic maximal family of the 9-subsets of the 13-set S such that any two 9-sets in the family has 7 or 8 common elements.

    As mentioned in Reference [7], we can invoke an algorithm to find the maximal family by using a computer and thus find the maximal graphs. Now we only give two examples to illustrate the existence of maximal families.

    Example 3.15. Let S={1,2,,13}, F1={1,2,,9}, ¯F1={10,11,12,13}, and F={F1}{Fi|FiF1|=8,|Fi¯F1|=1} be a family of 9-subsets of S. Then F is a maximal family such that any two sets in F has 7 or 8 common elements.

    Proof. First, we prove that all sets in F are compatible. For any FiFF1, we have |F1Fi|=8. For any two sets Fi,FjFF1, ij, if (Fi¯F1)(Fj¯F1)=, then |(FiF1)(FjF1)|=7 or 8 since |F1|=9, |F1Fi|=8 and |F1Fj|=8, thus |FiFj|=7 or 8; if |(Fi¯F1)(Fj¯F1)|=1, then |(FiF1)(FjF1)|=7 since FiFj, thus |FiFj|=8. Therefore, all sets in F are compatible.

    Next, we prove that the family is maximal. Assume to the contrary, if F is not maximal, then there is a set FF such that |F|=9 and F is compatible with all sets in F, say, |FFi|=7 or 8 for any FiF. Since F1={1,2,,9}F, we have |FF1|=7 or 8.

    If |FF1|=8, then FF, it implies a contradiction. If |FF1|=7, without loss of generality, we assume that F={1,2,3,4,5,6,7,10,11}, then there is a set Fq={2,3,4,5,6,7,8,9,12}F such that |FFq|=6, it is a contradiction.

    Therefore, F is a maximal family such that the intersection of any two sets in F has 7 or 8 elements.

    Example 3.16. Let S={1,2,,13}, F1={1,2,,9}, ¯F1={10,11,12,13}, S1={1,2,,7}, S2={1,2,,7,8}, S3={1,2,,7,9}, and F={F1}{FiS1Fi,|Fi¯F1|=2}{FiS2Fi,|Fi¯F1|=1}{FiS3Fi,|Fi¯F1|=1} be a family of 9-subsets of S. Then F is a maximal family such that the intersection of any two sets in F has 7 or 8 elements.

    Proof. First, we prove that all sets in F are compatible. For any two sets Fp,FqF, we have |FpFq|7 since |S1S2|=|S1S3|=|S2S3|=7. On the other hand, |Fp|=|Fq|=9, and FpFq, thus |FpFq|=7 or 8. Therefore, all sets in F are compatible.

    Next, we prove that the family is maximal. Assume to the contrary, if F is not maximal, then there is a set FF such that |F|=9 and F is compatible with all sets in F, say, |FFi|=7 or 8 for any FiF. Since F1={1,2,,9}F, we have |FF1|=7 or 8.

    Case 1: |FF1|=7.

    Then |FS1|=5,6 or 7. If |FS1|=7, then F{FiS1Fi,|Fi¯F1|=2}F, it is a contradiction. If |FS1|=6, without loss of generality, we assume that F={1,2,3,4,5,6,8,10,11}, then there is a set Fl={1,2,3,4,5,6,7,12,13}{FiS1Fi,|Fi¯F1|=2}F such that |FFl|=6, it is a contradiction. If |FS1|=5, without loss of generality, we assume that F={1,2,3,4,5,8,9,10,11}, then there is a set Fl={1,2,3,4,5,6,7,12,13}F such that |FFl|=5, it is a contradiction.

    Case 2: |FF1|=8.

    Then |FS1|=6 or 7. If |FS1|=7, then F{FiS2Fi,|Fi¯F1|=1}{FiS3Fi,|Fi¯F1|=1}F, it is a contradiction. If |FS1|=6, without loss of generality, we assume that F={1,2,3,4,5,6,8,9,10}, then there is a set Fl={1,2,3,4,5,6,7,12,13}F such that |FFl|=6, it is a contradiction.

    Combining the above arguments, F is a maximal family such that the intersection of any two sets in F has 7 or 8 elements.

    Corollary 3.17. K2,5 and K2,13 are the only graphs among K2,s which can be star complements for the second largest eigenvalue λ2=1.

    Proof. Let G be a graph of order n with H as a star complement for an eigenvalue 1 of multiplicity k, where HK2,s. We know that K2,s has spectrum: 0s, 2s, 2s, then

    λ2(H)=0<1<2s=λ1(H).

    By Lemma 2.6, we have λ2(G)=λ3(G)==λ1+k(G)=1.

    By Table 1, we can see that the known research on the maximal graph with H as a star complement is about a relatively small eigenvalue, such as 1 and 2. We are curious about what will happen to the maximal graph when the eigenvalue becomes larger.

    In this section, we will study the maximal graphs with K2,s as a star complement for other eigenvalues, such as μ=2,3,3,4.

    Proposition 4.1. The only graphs among K2,s which can be star complements for μ=2,3,3,4 are shown in Table 5.

    Table 5.  graphs among K2,s which can be as star complements for μ.
    μ K2,s
    2 K2,1, K2,18, K2,20, K2,26, K2,27, K2,29, K2,34, K2,51, K2,52
    3 K2,3, K2,42, K2,45, K2,65, K2,67, K2,78, K2,126, K2,185, K2,225, K2,317
    -3 K2,3, K2,4, K2,29, K2,36, K2,42, K2,45, K2,65, K2,78, K2,89, K2,117, K2,185
    4 K2,6, K2,7, K2,72, K2,78, K2,80, K2,88, K2,89, K2,98, K2,106, K2,108,
    K2,133, K2,134, K2,152, K2,168, K2,170, K2,250, K2,297, K2,449, K2,656

     | Show Table
    DownLoad: CSV

    Proof. We only prove the case of μ=2. The proofs for μ=3,3,4 are similar, so we omit them.

    Let uX be a vertex of type (a,b) which means that it has a neighbours in R and b neighbours in S. Then (a,b)(0,0) and a{0,1,2}, 0bs.

    By (2.3), μ=2 and t=2, we have

    2b2+(4a+42s)b+sa22sa+4a+8s16=0. (4.1)

    Since b is an integer, the discriminant of (4.1), 4s280s8sa2+16a2+144 must be a perfect square. Table 6 shows the possible values of s and (s,b) when a=0,1,2.

    Table 6.  The integral solutions of (4.1) (0bs).
    a the discriminant of equation (4.1) s (s,b)
    0 4(s10)2256 2,18,20,27 (2,0),(18,8),(20,6),(20,12),(27,5),(27,20)
    1 4(s11)2324 20,26,52 (20,8),(26,5),(26,17),(52,4),(52,44)
    2 4(s14)2576 1,26,27,29,34,51 (1,0),(26,10),(27,8),(27,13),(29,7), (29,16),(34,6),(34,22),(51,5),(51,40)

     | Show Table
    DownLoad: CSV

    Since μ=2 is not the eigenvalue of HK2,s, so μ22s, and then s2. Therefore only K2,1, K2,18, K2,20, K2,26, K2,27, K2,29, K2,34, K2,51 and K2,52 can be star complements for μ=2.

    Similar to the cases of μ=1 and μ=2, we can study the maximum graphs. Now we only characterize the following cases.

    Firstly, we define two graph G3 and G4. Let (R,S) be the bipartition of H=K2,4 with R={1,2}, S={1,2,3,4}. We define a graph G3 as follows: X={u}, V(G3)=V(H)X, and NH(u)={1,2,1} (see Figure 1). The spectrum of G3 is [3,1,03,0.58579,3.41421].

    Figure 1.  G3.

    Let (R,S) be the bipartition of H=K2,6 with R={1,2}, S={1,2,,6}. We define a graph G4 as follows: X={v}, V(G4)=V(H)X, and NH(v)={1,1,2,3} (see Figure 2). The spectrum of G4 is [3.51414,1.57199,05,1.08613,4].

    Figure 2.  G4.

    Theorem 4.2. The graph G shown in Table 7 is the only graph with H as a star complement for eigenvalue μ.

    Table 7.  The only graph with H as a star complement for eigenvalue μ.
    μ H G
    2 K2,1 K2,2
    3 K2,3 K3,3
    -3 K2,3 K3,3
    -3 K2,4 G3
    4 K2,6 G4
    4 K2,7 K2,8

     | Show Table
    DownLoad: CSV

    Proof. When μ=2 and H=K2,1, it is clear that a=c=2, b=d=0 by Proposition 4.1. By Proposition 2.5 and μ=2, t=2,s=1, we have G is the graph with K2,1 as a star complement for μ=2 if and only if the vertex set X satisfies GX=H and the following two conditions:(1) for any uX, u is of type (2,0); (2) for any two vertices u,vX, we have 2auv+ρuv+2=0.

    But ρuv0 and auv=0 or 1 imply |X|=1. Therefore, there is a unique (maximal) graph K2,2 with K2,1 as a star complement for μ=2.

    Similarly, we can prove other cases in Table 7.

    Remark 4.3. Let G be a maximal graph with K2,s as a star complements for μ. In order to study G, the first thing we need to do is to determine what value of s that allows K2,s to be viewed as a star complement for a given eigenvalue. We compare the graphs among K2,s which can be star complements for distinct eigenvalues, for example, μ=1, μ=2, μ=2, μ=3, μ=3 and μ=4 (see Theorem 3.10, Theorem 3.2 and Table 5), and we find that:

    (1) there are only two graphs among K2,s as star complements for μ=1;

    (2) there are only seven graphs among K2,s as star complements for μ=2 and nine for μ=2;

    (3) there are only eleven graphs among K2,s as star complements for μ=3 and ten for μ=3;

    (4) there are nineteen graphs among K2,s as star complements for μ=4;

    (5) as the absolute value of μ gets larger and larger, the value and the number of s get larger and larger; even though the absolute values are the same, it seems that the value of s corresponding to the positive eigenvalues is larger than the one corresponding to the negative eigenvalues.

    It seems that (5) explains why known research choose eigenvalues with small absolute value for study.

    It is well known that for any graph G, there exists at least one star partition ([7]). The fact implies there exist star sets and star complements for any eigenvalue of any graph G. Then what graphs can not be as star complements for some given eigenvalues seems to be an interesting question worth studying.

    The authors would like to thank the referees for their valuable comments, corrections, and suggestions, which lead to an improvement of the original manuscript.

    The research is supported by the National Natural Science Foundation of China (Grant No. 11971180), the Guangdong Provincial Natural Science Foundation (Grant No. 2019A1515012052).

    The authors declare that they have no competing interests.



    [1] H. Bateman, Some recent researches on the motion of fluids, Mon. Weather Rev., 43 (1915), 163–170. https://doi.org/10.1175/1520-0493(1915)43 doi: 10.1175/1520-0493(1915)43
    [2] J. Burgers, A mathematical model illustrating the theory of turbulence, Adv. Appl. Mech., 1 (1948), 171–199. https://doi.org/10.1016/S0065-2156(08)70100-5 doi: 10.1016/S0065-2156(08)70100-5
    [3] J. D. Murray, On Burgers' model equations for turbulence, J. Fluid Mech., 59 (1973), 263–279. https://doi.org/10.1017/S0022112073001564 doi: 10.1017/S0022112073001564
    [4] N. T. Eldabe, E. M. Elghazy, A. Ebaid, Closed form solution to a second order boundary value problem and its application in fluid mechanics, Phys. Lett. A, 363 (2007), 257–259. https://doi.org/10.1016/j.physleta.2006.11.010 doi: 10.1016/j.physleta.2006.11.010
    [5] K. Fujisawa, A. Asada, Nonlinear parametric sound enhancement through different fluid layer and its application to noninvasive measurement, Measurement, 94 (2016), 726–733. https://doi.org/10.1016/j.measurement.2016.09.004 doi: 10.1016/j.measurement.2016.09.004
    [6] Y. L. Chen, T. Zhang, A weak Galerkin finite element method for Burgers' equation, Comput. Appl. Math., 348 (2019), 103–119. https://doi.org/10.1016/j.cam.2018.08.044 doi: 10.1016/j.cam.2018.08.044
    [7] M. A. Shallal, A. H. Taqi, B. F. Jumaa, H. Rezazadeh, M. Inc, Numerical solutions to the 1-D Burgers' equation by a cubic Hermite finite element method, Indian J. Phys., 96 (2022), 3831–3836. https://doi.org/10.1007/s12648-022-02304-4 doi: 10.1007/s12648-022-02304-4
    [8] Y. S. Uçar, N. Yağmurlu, İ. Çelikkaya, Operator splitting for numerical solution of the modified Burgers' equation using finite element method, Numer. Meth. Partial Differ. Equations, 35 (2019), 478–492. https://doi.org/10.1002/num.22309 doi: 10.1002/num.22309
    [9] L. Wang, H. Li, Y. Meng, Numerical solution of coupled Burgers' equation using finite difference and sinc collocation method, Eng. Lett., 29 (2021), 668–674.
    [10] J. Zhang, Q. Yang, The finite volume element method for time fractional generalized Burgers' equation, Fractal Fract., 8 (2024), 1–17. https://doi.org/10.3390/fractalfract8010053 doi: 10.3390/fractalfract8010053
    [11] K. Kaur, G. Singh, An efficient non-standard numerical scheme Coupled with a compact finite difference method to solve the one-dimensional Burgers' equation, Axioms, 12 (2023), 593–609. https://doi.org/10.3390/axioms12060593 doi: 10.3390/axioms12060593
    [12] J. Zhang, J. Yu, A Multi-Quadrics quasi-interpolation scheme for numerical solution of Burgers' equation, Appl. Numer. Math., 208 (2025), 38–44. https://doi.org/10.1016/j.apnum.2024.09.025 doi: 10.1016/j.apnum.2024.09.025
    [13] Y. Shi, X. Yang, A time two-grid difference method for nonlinear generalized viscous Burgers' equation, J. Math. Chem., 62 (2024), 1323–1356.
    [14] R. D. Ortiz, O. M. Nunez, A. M. M. Ramirez, Solving viscous Burgers' equation: Hybrid approach combining Boundary Layer theory and physics-informed neural networks, Mathematics, 12 (2024), 3430. https://doi.org/10.3390/math12213430 doi: 10.3390/math12213430
    [15] T. Mouktonglang, S. Yimnet, N. Sukantamala, B. Wongsaijai, Dynamical behaviors of the solution to a periodic initial-boundary value problem of the generalized Rosenau-RLW-Burgers equation, Math. Comput. Simul., 196 (2022), 114–136. https://doi.org/10.1016/j.matcom.2022.01.004 doi: 10.1016/j.matcom.2022.01.004
    [16] L. Li, K. W. Ong, Dynamic transitions of generalized Burgers equation, J. Math. Fluid Mech., 18 (2016), 89–102. https://doi.org/10.1007/s00021-015-0240-7 doi: 10.1007/s00021-015-0240-7
    [17] Q. Zhang, D. Yan, Z. Pan, Bifurcation from double eigenvalue for nonlinear equation with third-order nondegenerate singularity, Appl. Math. Comput., 220 (2013), 549–559. https://doi.org/10.1016/j.amc.2013.06.024 doi: 10.1016/j.amc.2013.06.024
    [18] D. Wei, S. Guo, Steady-state bifurcation of a nonlinear boundary problem, Appl. Math. Lett., 128 (2022), 107902. https://doi.org/10.1016/j.aml.2021.107902 doi: 10.1016/j.aml.2021.107902
    [19] G. Guo, X. Wang, X. Lin, M. Wei, Steady-state and Hopf bifurcations in the Langford ODE and PDE systems, Nonlinear Anal. Real World Appl., 34 (2017), 343–362. https://doi.org/10.1016/j.nonrwa.2016.09.008 doi: 10.1016/j.nonrwa.2016.09.008
    [20] Z. Pan, L. Jia, Y. Mao, Q. Wang, Transitions and bifurcations in couple stress fluid saturated porous media using a thermal non-equilibrium model, Appl. Math. Comput., 415 (2022), 126727. https://doi.org/10.1016/j.amc.2021.126727 doi: 10.1016/j.amc.2021.126727
    [21] Y. Wang, The Progress of Chemotaxis-fluid Models, Journal of Xihua University (Natural Science Edition), 35 (2016), 30–34, 38. https://doi.org/10.3969/j.issn.1673-159X.2016.04.006
    [22] J. Chen, Y. Han, Global Weak Solvability for a Chemotaxis-Fluid Model with Low Regular Initial Data, Journal of Xihua University (Natural Science Edition), 43 (2024), 97–105. https://doi.org/10.12198/j.issn.1673-159X.5354
    [23] W. Kuang, Y. Dong, Global Well-posedness of 2D Chemotaxis-fluid System, Journal of Xihua University (Natural Science Edition), 43 (2024), 103–108. https://doi.org/10.12198/j.issn.1673-159X.5233
    [24] R. Ma, L. Wei, Z. Chen, Evolution of bifurcation curves for one-dimensional Minkowski-curvature problem, Appl. Math. Lett., 103 (2020), 106176. https://doi.org/10.1016/j.aml.2019.106176 doi: 10.1016/j.aml.2019.106176
    [25] T. Sengul, B. Tiryakioglu, Dynamic transitions and bifurcations of 1D reaction-diffusion equations: The non-self-adjoint case, Math. Anal. Appl., 523 (2023), 127114. https://doi.org/10.1016/j.jmaa.2023.127114 doi: 10.1016/j.jmaa.2023.127114
    [26] Z. Feng, Y. Su, Traveling wave phenomena of inhomogeneous half-wave equation, J. Differ. Equations, 400 (2024), 248–277. https://doi.org/10.1016/j.jde.2024.04.029 doi: 10.1016/j.jde.2024.04.029
    [27] Y. Fan, L. Li, Z. Pan, Q. Wang, On dynamics of double-diffusive magneto-convection in a non-Newtonian fluid layer, Math. Methods Appl. Sci., 46 (2023), 14596–14621. https://doi.org/10.1002/mma.9337 doi: 10.1002/mma.9337
    [28] S. Li, J. Wu, H. Nie, Steady-state bifurcation and Hopf bifurcation for a diffusive Leslie–Gower predator–prey model, Comput. Math. Appl., 70 (2015), 3043–3056. https://doi.org/10.1016/j.camwa.2015.10.017 doi: 10.1016/j.camwa.2015.10.017
    [29] M. Wei, J. Wu, G. Guo, Steady state bifurcations for a glycolysis model in biochemical reaction, Nonlinear Anal. Real World Appl., 22 (2015), 155–175. https://doi.org/10.1016/j.nonrwa.2014.08.003 doi: 10.1016/j.nonrwa.2014.08.003
    [30] L. Ma, Z. Feng, Stability and bifurcation in a two-species reaction–diffusion–advection competition model with time delay, Nonlinear Anal. Real World Appl., 61 (2021), 103327. https://doi.org/10.1016/j.nonrwa.2021.103327 doi: 10.1016/j.nonrwa.2021.103327
    [31] T. Ma, S. Wang, Stability and Bifurcation of Nonlinear Evolution Equations, 1nd edition, Science Press, China, 2007.
    [32] T. Ma, S. Wang, Bifurcation theory and applications, World Sci, Singapore, 2005.
    [33] W. Cao, X. Shan, S. Tang, W. Ouyang, W. Zhang, Solving parametric high-Reynolds-number wall-bounded turbulence around airfoils governed by Reynolds-averaged Navier-Stokes equations using time-stepping-oriented neural network, Phys. Fluids, 37 (2025), 015151. https://doi.org/10.1063/5.0245918 doi: 10.1063/5.0245918
    [34] B. Tripathi, P. W. Terry, A. E. Fraser, E. G. Zweibel, M. J. Pueschel, Three-dimensional shear-flow instability saturation via stable modes, Phys. Fluids, 35 (2023), 105151. https://doi.org/10.1063/5.0167092 doi: 10.1063/5.0167092
    [35] N. Ciola, P. D. Palma, J. C. Robinet, S. Cherubini, Large-scale coherent structures in turbulent channel flow: a detuned instability of wall streaks, Fluid Mech., 997 (2024), A18. https://doi.org/10.1017/jfm.2024.726 doi: 10.1017/jfm.2024.726
    [36] J. Zhu, Y. Peng, W. Zhang, S. Niu, Y. Huang, Y. Li, et al. Research on the high-integration vortex-shape flow channel of a hydraulic servo valve based on additive manufacturing, in CSAA/IET International Conference on Aircraft Utility Systems (AUS 2024), 2024 (2025), 1700–1705. https://doi.org/10.1049/icp.2024.3137
    [37] B. Eckhardt, H. Faisst, A. Schmiegel, T. M. Schneider, Dynamical systems and the transition to turbulence in linearly stable shear flows, Phil. Trans. R. Soc. A, 366 (2008), 1297–1315. https://doi.org/10.1098/rsta.2007.2132 doi: 10.1098/rsta.2007.2132
  • This article has been cited by:

    1. Xiaona Fang, Lihua You, Regular and Maximal Graphs with Prescribed Tripartite Graph as a Star Complement, 2023, 44, 0252-9599, 517, 10.1007/s11401-023-0029-6
  • Reader Comments
  • © 2025 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(198) PDF downloads(35) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog