Loading [MathJax]/jax/output/SVG/jax.js
Research article

A symplectic fission scheme for the association scheme of rectangular matrices and its automorphisms

  • Received: 28 September 2024 Revised: 08 November 2024 Accepted: 11 November 2024 Published: 20 November 2024
  • MSC : 05B25, 05C20, 05E30

  • In this paper, a symplectic fission scheme for the association scheme of m×n rectangular matrices over the finite field Fq, denoted by SMat(m×n,q), is constructed, where q is a power of a prime number. We discuss its association classes and inner automorphism group. In particular, we determine the intersection numbers and automorphism group of SMat(m×n,q) for m=1 and m=2.

    Citation: Yang Zhang, Shuxia Liu, Liwei Zeng. A symplectic fission scheme for the association scheme of rectangular matrices and its automorphisms[J]. AIMS Mathematics, 2024, 9(11): 32819-32830. doi: 10.3934/math.20241570

    Related Papers:

    [1] Abdul Rauf, Nehad Ali Shah, Aqsa Mushtaq, Thongchai Botmart . Heat transport and magnetohydrodynamic hybrid micropolar ferrofluid flow over a non-linearly stretching sheet. AIMS Mathematics, 2023, 8(1): 164-193. doi: 10.3934/math.2023008
    [2] M. A. El-Shorbagy, Waseem, Mati ur Rahman, Hossam A. Nabwey, Shazia Habib . An artificial neural network analysis of the thermal distribution of a fractional-order radial porous fin influenced by an inclined magnetic field. AIMS Mathematics, 2024, 9(6): 13659-13688. doi: 10.3934/math.2024667
    [3] Yellamma, N. Manjunatha, Umair Khan, Samia Elattar, Sayed M. Eldin, Jasgurpreet Singh Chohan, R. Sumithra, K. Sarada . Onset of triple-diffusive convective stability in the presence of a heat source and temperature gradients: An exact method. AIMS Mathematics, 2023, 8(6): 13432-13453. doi: 10.3934/math.2023681
    [4] Bengisen Pekmen Geridonmez . RBF simulation of natural convection in a nanofluid-filled cavity. AIMS Mathematics, 2016, 1(3): 195-207. doi: 10.3934/Math.2016.3.195
    [5] Saleh Mousa Alzahrani . Enhancing thermal performance: A numerical study of MHD double diffusive natural convection in a hybrid nanofluid-filled quadrantal enclosure. AIMS Mathematics, 2024, 9(4): 9267-9286. doi: 10.3934/math.2024451
    [6] Ritu Agarwal, Mahaveer Prasad Yadav, Dumitru Baleanu, S. D. Purohit . Existence and uniqueness of miscible flow equation through porous media with a non singular fractional derivative. AIMS Mathematics, 2020, 5(2): 1062-1073. doi: 10.3934/math.2020074
    [7] M. Hamid, T. Zubair, M. Usman, R. U. Haq . Numerical investigation of fractional-order unsteady natural convective radiating flow of nanofluid in a vertical channel. AIMS Mathematics, 2019, 4(5): 1416-1429. doi: 10.3934/math.2019.5.1416
    [8] Kiran Sajjan, N. Ameer Ahammad, C. S. K. Raju, M. Karuna Prasad, Nehad Ali Shah, Thongchai Botmart . Study of nonlinear thermal convection of ternary nanofluid within Darcy-Brinkman porous structure with time dependent heat source/sink. AIMS Mathematics, 2023, 8(2): 4237-4260. doi: 10.3934/math.2023211
    [9] Suliman Khan, M. Riaz Khan, Aisha M. Alqahtani, Hasrat Hussain Shah, Alibek Issakhov, Qayyum Shah, M. A. EI-Shorbagy . A well-conditioned and efficient implementation of dual reciprocity method for Poisson equation. AIMS Mathematics, 2021, 6(11): 12560-12582. doi: 10.3934/math.2021724
    [10] Reem K. Alhefthi, Irum Shahzadi, Husna A. Khan, Nargis Khan, M. S. Hashmi, Mustafa Inc . Convective boundary layer flow of MHD tangent hyperbolic nanofluid over stratified sheet with chemical reaction. AIMS Mathematics, 2024, 9(7): 16901-16923. doi: 10.3934/math.2024821
  • In this paper, a symplectic fission scheme for the association scheme of m×n rectangular matrices over the finite field Fq, denoted by SMat(m×n,q), is constructed, where q is a power of a prime number. We discuss its association classes and inner automorphism group. In particular, we determine the intersection numbers and automorphism group of SMat(m×n,q) for m=1 and m=2.



    A two-sphere of revolution is a compact Riemannian surface (M,h), which is homeomorphic to the sphere S2R3. If this manifold is endowed with a Randers metric F=α+β, or more generally, with an arbitrary positive defined Finsler metric F, then (M,F) is called a Randers or Finsler two-sphere of revolution, respectively.

    One of the major problems in Differential Geometry (see [14,15]) and Optimal Control (see [5]) is the study of geodesics, conjugate points and cut points of Riemannian or Finsler manifolds. We recall that a vector field J along a unit speed geodesic γ:[0,a]M is said to be a Jacobi field if it satisfies the well-known Jacobi equation (see for instance [3], Chapter 7 for details). A point p is said to be conjugate to q:=γ(0) along γ if there exists a non-zero Jacobi field J along γ which vanishes at p and q. The set of conjugate points of q along all curves γ starting at q is called the conjugate locus of q.

    If γ:[0,l]M is a minimal geodesic on a such manifold, then its end point γ(l)M is called the cut point of the initial point q=γ(0)M, in the sense that any extension of γ beyond γ(l) is not a minimizing geodesic from q anymore. The cut locus Cut(q) is defined as the set of cut points of q, and on Riemannian or Finslerian surfaces, it has the structure of a local tree. Moreover, the cut points pCut(q) of q are characterized by the property that the distance d(q,) from q is not smooth any more at p (see [12] for details). The cut points p along a geodesic γ emanating from the point q=γ(0) can appear either before or at the first conjugate point of q along γ, but not after that (see [3]).

    To determine the precise structure of the cut locus on a Riemannian or Finsler manifold is not an easy task. The majority of known results concern Riemannian or Randers surfaces of revolution (see [15,16] for the Riemannian, and [7,8] for the Randers case).

    A Randers metric F=α+β is a special Finsler metric obtained by the deformation of a Riemannian metric α by a one-form β whose Riemannian α-length is less than one in order to assure that F is positively defined ([10]). These Finsler metrics are intuitive generalizations of the Riemannian ones having most of the geometrical objects relatively easy to compute (see [3]).

    An equivalent characterization of Randers metrics is through the Zermelo's navigation problem. We recall that a Finsler metric F is characterized by its indicatrix {(x,y)TM:F(x,y)=1} (see [3]). In particular, a Randers metric indicatrix is obtained as the rigid translation of the unit sphere {yTxM:h(x,y)=1} of a Riemannian metric (M,h) by a vector field WX(M) whose Riemannian length is less than one. The pair (h,W) will be called the navigation data of the Randers metric F=α+β. Conversely, the Randers metric F=α+β will be called the solution of Zermelo's navigation problem (h,W). In the case when W is an h-Killing field, provided h is not flat, the geodesics, conjugate points and cut points of the Randers metric F=α+β can be obtained by the translation of the corresponding geodesics, conjugate points and cut points, of the Riemannian metric h by the flow of W, respectively (see [8,11]). More generally, new Finsler metrics F can be obtained by the rigid translation of the indicatrix of a given Finsler metric F0 by a vector field W, such that F0(W)<1 (see [6,13]). In this case, the pair (F0,W) will be called the general navigation data of F.

    Another case when the Randers geodesics can be easily related to the Riemannian ones is when the deformation one-form β is closed. Indeed, the Randers metric F=α+β is projectively equivalent to the underlying Riemannian metric α if and only if dβ=0. In this case, the α-geodesics, conjugate points and cut points coincide with the F-geodesics, cut points and conjugate points, respectively (see [3]).

    We combine these two cases of Randers metrics in order to obtain new families of Finsler of Randers type with simple cut locus (see Section 2 for the definition). The originality of our paper lies in the followings:

    (ⅰ) We determine the geodesics behavior, conjugate and cut loci of some families of Finsler metrics of Randers type whose navigation data do not necessarily include a Killing field.

    (ⅱ) We show that the structure of the cut locus of these families can be determined without any sectional or flag curvature restrictions. These are generalizations of the results in [16] to the Randers case.

    (ⅲ) We construct a sequence of Randers metrics whose cut locus structure is simple.

    (ⅳ) We extend some classical results from the case of Randers metrics to β-changes of Finsler metrics and give new proofs to some known results.

    If we start with a Riemannian two-sphere of revolution (MS2,h) and the vector fields V0,V,WX(M). Then the following construction gives the Randers metric F0,F1 and F2 as solutions of the Zermelo's navigation problem with data (h,V0),(F0,V) and (F1,W), respectively,

    which are positively defined provided V0h<1, F0(V)<1, and F1(W)<1, respectively. If we impose conditions that V0 and V to be h- and F0-Killing fields, respectively, and dβ2=0, then the geodesics, conjugate and cut loci of F2 can be determined.

    Remarkably, a shortcut of this construction would be to simply impose V0+V+Wh<1, that guarantees F2 is positively defined, and V0+V+W to be h-Killing. In this case, F2 is also with simple cut locus having the same structure with the cut locus of h. Obviously, the cut loci of these metrics are slightly different as set of points on M.

    This construction can be extended to a sequence of Randers metrics {Fi=αi+βi}i=1,n whose cut loci are simple (see Remark 8).

    Here is the structure of our paper.

    In Section 2, we review the geometry of Riemannian two-spheres of revolution from [15] and [16].

    In Section 3, we describe the geometry of some families of Randers metrics obtained as generalizations to the Finslerian case of the Riemannian metrics in [16]. We use the Hamiltonian formalism for giving and proving some basic results to be used later in the section. Lemma 2 is an important result that generalizes a well-known result [9] for Randers metrics to more general Finsler metrics obtained by β-changes. The relation with F-Killing fields are given in Lemma 3 and the basic properties of our family of Randers metrics are in Lemma 4. Some of these results are indirectly suggested in [6], but here we clarify all the basic aspects and prove them in our specific formalism. Lemma 5 gives the concrete expressions of ˜α and ˜β in the families of our Randers metric, formulas that provide a better understanding of the positive definiteness of these metrics.

    Lemma 6 gives the behavior of geodesics, conjugate and cut points of the β-change of a Randers metric generalizing the results in [11]. Lemma 7 gives the conditions for the one-form β to be closed in terms of the navigation data. Finally, we sum up the results in all these lemmas in Theorem 3.1, which is the main result of the present paper. In Remark 8, we show how an infinite sequence of such Randers metrics can be obtained.

    In Section 4, we construct one example of the Randers metric on the two-sphere of revolution that satisfies the conditions in Theorem 3.1.

    Classically, surfaces of revolution are obtained by rotating a curve (c) in the xz plane around the z axis. More precisely, if the profile curve (c) is given parametrically

    (c):{x=φ(u)z=ψ(u), φ>0, uIR, (2.1)

    then, in the case (φ(u))2+(ψ(u))20, for all uI, the curve can be written explicitly x=f(z) or implicitly by Φ(x,z)=0, where is the derivative with respect to u.

    In the case of the parametric representation (2.1) one obtains an R3-immersed surface of revolution ψ:ΩR2R3, given by

    ψ(u,v)=(φ(u)cosv,φ(u)sinv,ψ(u)), uI, v[0,2π). (2.2)

    Remark 2.1. The immersed surface of revolution (2.2) is called of elliptic type, while

    ψ(u,v)=(φ(u)coshv,φ(u)sinhv,ψ(u))

    is called a hyperbolic type. Since we are interested in compact surfaces of revolution, only the elliptic case will be considered hereafter, leaving the hyperbolic type for a future research.

    Even though the representation (2.2) is quite intuitive, it has two major disadvantages:

    (1) it leads to quite complicated formulas for the induced Riemannian metric, geodesic equations, Gauss curvature, etc.,

    (2) it excludes the case of abstract surfaces of revolution which cannot be embedded in R3.

    The first disadvantage can be easily fixed by taking the curve (c) to be unit speed parameterized in the Euclidean plane xz, i.e.,

    [φ(u)]2+[ψ(u)]2=1,

    which leads to the warped Riemannian metric

    ds2=du2+φ2(u)dv2.

    This simplification suggests the following definition (which also fixes the second disadvantage).

    Definition 2.1. [15] Let (M,h) be a compact Riemannian surface homeomorphic to S2. If M admits a pole pM, and for any two points q1,q2M, such that dh(p,q1)=dh(p,q2), there exists an Riemannian isometry i:MM for which

    i(q1)=q2, i(p)=p,

    then (M,h) is called a two-sphere of revolution. Here dh is the distance function associated to the Riemannian metric h.

    Remark 2.2. One example of compact surface of revolution that cannot be embedded in R3 is the real projective space RP2. It is compact being homeomorphic to S2/, where S2 is the unit sphere in R3 and is the equivalence relation xx, for all xS2. It is a surface of revolution because it can be obtained by rotating the Möbius strip along its center line.

    Finally, it cannot be embedded in R3 because it is non-orientable. More generally, it is known that any embedding of a non-orientable surface in R3 must create self-intersections and this is not allowed. Nevertheless, RP2 can be immersed in R3, and therefore can be locally embedded in R3, but not globally (see [4] for properties of projective spaces).

    Another example is the so-called Lorentz surface, obtained by rotating the hyperbola x2y2=1 around the x-axis. This surface is orientable but cannot embedded in R3 because it has a self-intersection at origin.

    This definition allows to introduce the geodesic polar coordinates (r,θ)(0,2a)×[0,2π) around p, such that the Riemannian metric is given as

    h=dr2+m2(r)dθ2

    on M{p,q}, where q is the unique cut point of p and

    m(r)=h(θ,θ)

    (see [14] or [15] for details).

    Moreover, the functions m(r) and m(2ar) can be extended to smooth odd function around r=0, where dh(p,q)=2a, m(0)=1=m(2a).

    It is well-known that any pole pM must have a unique cut point qM, and that any geodesic starting from p contains q.

    For the sake of simplicity, we consider a=π2, that is m:[0,π][0,) will satisfy m(0)=0, m(0)=1, m(πr)=m(r)>0, for all r(0,π), see Figure 1.

    Figure 1.  A two-sphere of revolution.

    Recall (see [14]) that the equations of an h-unit speed geodesic γ(s):=(r(s),θ(s)) of (M,h) are

    {d2rds2mm(dθds)2=0,d2θds2+2mm(drds)(dθds)=0,

    where s is the arclength parameter of γ with the h-unit speed parameterization condition

    (drds)2+m2(dθds)2=1.

    It follows that every profile curve, or meridian, {θ=θ0} with θ0 constant is an h-geodesic, and that a parallel {r=r0}, with r0(0,2a) constant, is geodesic if and only if m(r0)=0. We observe that the geodesics equations implies

    dθ(s)dsm2(r(s))=ν, where ν  is constant,

    that is, the quantity dθdsm2 is conserved along the h-geodesics.

    Lemma 2.1. (The Clairaut relation) Let γ(s)=(r(s),θ(s)) be an h-unit speed geodesic on (M,h). There exists a constant ν such that

    m(r(s))cosΦ(s)=ν

    holds for any s, where Φ(s) denotes the angle between the tangent vector of γ(s) and profile curve.

    The constant ν is called the Clairaut constant of γ.

    Several characterization of the cut locus of a Riemannian two-sphere of revolution are known (see [5,15,16]).

    We recall the following important result from [16].

    Proposition 2.1. Let h:[0,π]R be a smooth function that can be extended to an odd smooth function on R. If

    (c1) h(πr)=πh(r), for any r[0,π];

    (c2) h(r)>0, for any r[0,π2);

    (c3) h(r)>0, for any r(0,π2),

    then

    (i) the function m:[0,π]R given by m(r):=asinh(r), where a=1h(0), is the warp function of a two-sphere of revolution M.

    (ii) Moreover, if h(r)>0 on (0,π2), then the cut locus of a point q=(r0,0)M coincides with a subarc of the antipodal parallel r=πr0.

    Proof. We give only the proof outline here, for details please consult [16]. It can be seen that conditions (c1), (c2) imply that the function m:[0,π]R is positive, and m(0)=0, m(0)=1, m(πr)=m(r)>0 for r(0,π), hence the two surface of revolution is well-defined.

    Moreover, if (c3) holds good, then it can be proved that the half period function

    φm(ν):=2π2m1(ν)νm(r)m2(r)ν2dr

    is decreasing, where ν is the Clairaut constant, hence the conclusion follows (see Lemma 1 and Proposition 1 in [16]).

    Remark 2.3. Observe that h(0)=0, h(π2)=π2, h(π)=π, and the graph of h looks like in Figure 2.

    Figure 2.  The outline of the graph of h.

    Definition 2.2. A Riemannian (or Finsler) two-sphere of revolution whose cut locus is a subarc of a parallel will be called with simple cut locus.

    Remark 2.4. This naming is related to graph theory in the sense that simple cut locus means that the cut locus is a simple graph with 2 vertices and one edge.

    We recall some examples given in [16].

    Example 2.1. (i) If h(r)=rαsin(2r), for any α(0,12), one can see that

    m(r)=asin(rαsin(2r)).

    It follows that

    m(r)=acos(rαsin(2r))[12αcos(2r)],m(r)=acos(rαsin(2r))[4αcos(2r)]a[12αcos(2r)]sin(rαsin(2r))[12αcos(2r)].

    Observe that the Gaussian curvature is

    G(r)=m(r)m(r)=1asin(rαsin(2r)){acos(rαsin(2r))[4αcos(2r)]+a[12αcos(2r)]sin(rαsin(2r))[12αcos(2r)]}=4αcos(2r)cot(rαsin(2r))+[12αcos(2r)]2,

    which clearly is not monotone on [0,π], see Figure 3. On the other hand, it is easy to check that this h satisfies conditions (c1), (c2), (c3) in Proposition 2.1, hence it results that the Riemannian surface of revolution with the warp function m has simple cut locus.

    Figure 3.  The graph of G(r) in Example 1 (i), r(0,π), where α=14.

    (ii) If h(r)=arcsinsinr1+λcos2r, for any λ0, it follows that

    m(r)=asin(arcsinsinr1+λcos2r)=asinr1+λcos2r,

    therefore

    m(r)=a1+λcos2r[cos1+λcos2r+λcosrsin2r1+λcos2r]=a(1+λ)cosr(1+λcos2r)3/2,m(r)=a(1+λ)(1+λcos2r)3[(1+λcos2r)3/2sinr+3λcos2rsinr(1+λcos2r)1/2]=a(1+λ)(1+λcos2r)5/2[sinr(1+2λcos2r+λ2cos4r)+3λcos2rsinr]=a(1+λ)sinr(1+λcos2r)5/2[1+λcos2rλ2cos4r].

    We obtain the Gaussian curvature as follows

    G(r)=m(r)m(r)=(1+λ)(1λcos2r+λ2cos4r)(1+λcos2r)2

    which again is not monotone on [0,π], see Figure 4.

    Figure 4.  The graph of G(r) in Example 2.1 (ⅱ), r(0,π), where λ=1.

    This second example also satisfies the conditions (c1), (c2), (c3) in Proposition 2.1. Hence it provides a two-sphere of revolution with simple cut locus.

    Remark 2.5. A more complicated sequence of functions hn(r) with simple cut locus is constructed in [16], Theorem 1.

    We will show the existence of Randers two-spheres of revolution with simple cut locus using the following basic construction:

    where (M,h) is a Riemannian manifold, and V0,V,WX(M) are vector fields on M.

    It is known that in general, the navigation data (h,V), where (M,h) is a Riemannian metric and V a vector field on M such that Vh<1, induces the Randers metric

    F=α(x,y)=β(x,y)=λy2h+h(y,V)λh(y,V)λ.

    Here λ:=1V2h and h(y,V)=hijViyj is the h-inner product of the vectors V and y.

    Conversely, the Randers metric F=α+β, where α=aij(x)yiyj is a Riemannian metric and β=bi(x)yi a linear one-form on TM, induces the navigation data (h,V) given by

    h2=ε(α2β2), V=1εβ#.

    Here h2=hij(x)yiyj, ε:=1b2α, and β# is the Legendre transform of β, i.e.,

    β#=biyi=aijbiyj

    (see [1,2,11] for details).

    We recall some definitions for later use.

    A vector field X on TM is called Hamiltonian vector field if there exists a smooth function f:TMR, (x,p)f(x,p) such that

    Xf=fpixifxipi.

    For instance, we can consider the Hamiltonian vector fields of the lift W:=Wi(x)pi of W=Wixi to TM, or of the Hamiltonian K(x,p), the Legendre dual of any Finsler metric F(x,y) on M (see [9]).

    Indeed, on a Finsler manifold (M,F), for any yTxM{0} one can define

    p(y):=12ddt[F2(x,y+tv)]|t=0, vTxM,

    and obtain in this way the map

    L:TMTM,(x,y)(x,p),

    called the Legendre transformation of F.

    The curve ˆγ(t)=(x(t),p(t)):[a,b]TM is called the integral curve (or sometimes the flow) of a Hamiltonian vector field XfX(TM) if

    dˆγ(t)dt=Xf|ˆγ(t).

    More precisely, the mapping ϕ:R×TMTM, (t,(x,p))ϕ(t,(x,p)), denoted also by ϕt(x,p) or ϕ(x,p)t, satisfying the properties

    (ⅰ) ϕ(0,(x,p))=(x,p), for any (x,p)TM;

    (ⅱ) ϕsϕt=ϕs+t, for all s,tR;

    (ⅲ) dϕ(x,p)tdt|t=0=X|(x,p),

    is called the one-parametric group, or simply the flow, of the vector field XX(TM). A given one-parametric group always induces a vector field XX(TM). Conversely, a given vector field XX(TM) induces only locally a one-parametric group, sometimes called the local flow of X.

    A smooth vector field XX(M) on a Finsler manifold (M,F) is called F-Killing field if every local one-parameter transformation group {φt} of M generated by X consists of local isometries of F. The vector field X is F-Killing if and only if LˆXF=0, where L is the Lie derivative, and ˆX:=Xixi+yjXixjyi is the canonical lift of X to TM, or, locally Xi|j+Xj|i+2CpijXp|qyq=0, where "|" is the h-covariant derivative with respect to the Chern connection.

    Moreover, in the Hamiltonian formalism, the vector field X on M is Killing field with respect to F if and only if

    {K,W}=0,

    where K is the Legendre dual of F (see [9]), W=Wi(x)pi and {,} is the Poisson bracket.

    Lemma 3.1. Generalization of Hrimiuc-Shimada's result, see [9]. Let (M,˜F=˜α+˜β) be a Randers metric with general navigation data (F=α+β,W), F(W)<1. Then the Legendre dual of ˜F is ˜K:TMR, ˜K=K+W, where K is the Legendre dual of F and W=Wi(x)pi.

    Proof. Indeed, let F=α+β be a positive defined Randers metric on a differentiable manifold M with indicatrix F(x)={yTxM: F(x,y)=1}TxM, and let WX(M) be a vector field such that F(W)<1.

    Let us denote by ˜(x):=F(x)+W(x) the rigid translation of ΣF(x) by W(x), i.e.,

    ˜(x):={y=u+WTxM: F(u)=1}.

    Firstly, observe that by rigid translation, the tangent vectors to F and ˜ remain parallel, i.e., there exists a smooth function c(u)0 such that

    ˜Yu+Wx=c(u)(Fx),u, (3.1)

    where ˜Yu+Wx is the tangent vector to ˜ at u+Wx, and (Fx),y:Ty(TxM)TRR is the tangent map of Fx:TxM[0,), see Figure 5.

    Figure 5.  The rigid translation of the indicatrix.

    The solution of the Zermelo's navigation problem with data (F,W) is a Finsler metric ˜F such that

    ˜Fx(u+Wx)=1,

    where uTxM, F(x,u)=1, and ˜Fx is the restriction of ˜F to TxM. Since ˜ is the rigid translation of such a Finsler metric must exist.

    Second, with these notations, observe that in TxM we have

    L˜F(u+Wx)=c(u)LF(u), (3.2)

    where L˜F and LF are the Legendre transformations of ˜F and F, respectively. This formula follows directly from (3.1) and the definition of the Legendre transformation.

    Since relation (3.2) is between one-forms, actually this is a relation between linear transformations of the tangent space TxM. If we pair (3.2) with Wx and u, we get

    L˜F(u+Wx),Wx=c(u)LF(u),W (3.3)

    and

    L˜F(u+Wx),u=c(u)LF(u),u=c(u), (3.4)

    respectively, where we have used the fact that F(u)=1 is equivalent to LF(u),u=1. Here, , denotes the usual pairing of a one-form with a vector field.

    Therefore, by the same reason, since ˜F(u+W)=1 we have

    1=L˜F(u+Wx),u+Wx=L˜F(u+Wx),u+L˜F(u+Wx),Wx=c(u)+c(u)LF(u),W,

    where we use (3.3), (3.4). By the way, observe that

    c(u)=11+LF(u),W=11+u,Wxgx(u),

    where ,gx(u) is the inner product in TxM by gx(u), i.e. X,Ygx(u)=gij(x,u)XiYj.

    Next, let us denote by ˜K and K the Legendre dual metrics of ˜F and F, respectively. It follows that

    1=˜K[L˜F(u+Wx)]=c(u)˜K(LF(u)),

    and thus

    ˜K(LF(u))=1c(u)=1+LF(u),W=K(LF(u))+LF(u),W.

    If we denote LF(u)=ωx=(x,p)TM, then

    ˜Kx(p)=Kx(p)+ωx(W), (3.5)

    where Kx is the L-dual of F=α+β.

    Therefore, if ˜F is {the solution of the Zermelo's navigation} (i.e. it is the rigid translation of the indicatrix F by W) with navigation data (F,W), then

    ˜Kx(p)=Kx(p)+Wx(p), (3.6)

    where ˜K and K are the Hamiltonians of ˜F and F, respectively, and W=Wi(x)pi.

    Lemma 3.2. Let (M,F=α+β) be a Randers metric, the vector field WX(M) with flow ψt. Then the Hamiltonian vector field XK on TM is invariant under the flow ψt, of XW if and only if W is an F-Killing field, where K is the Legendre dual of F.

    Proof. Indeed, the invariance condition ψt,(XK)=XK is equivalent to LXWXK=0 by definition, hence [XW,XK]=0, i.e. X{W,K}=0. This shows that W is actually F-Killing field.

    Lemma 3.3. Let (M,F) be a Randers metric and WTM a vector field on M. Then

    (i) The navigation data of ˜F is (h,V+W), where (h,V) is the navigation data of F=α+β, and ˜F is the solution of Zermelo's navigation problem for (F,W).

    (ii) The Randers metric ˜F=˜α+˜β is positive defined if and only if F(W)<1.

    Proof. (ⅰ) Recall that (see [2,11]) the indicatrix of F is obtained by a rigid translation of the h-unit sphere h(x) by V, i.e. for any xM

    F(x)=h(x)+V(x),

    where F(x)={yTxM: F(x,y)=1}, h(x)={yTxM, yh=1}, and Vh<1. Then, if ˜F is the solution of the Zermelo's navigation problem for (F,W), we have

    ˜F(x)=F(x)+W(x)=h(x)+V(x)+W(x),

    i.e., navigation data of ˜F is (h,V+W).

    (ⅱ) If we use (ⅰ), then ˜F is positive defined Randers metric if and only if V+Wh<1. Observe that

    α2(W)=α2(W)=aijWiWj=1λhijWiWj+(ViλWi)2=1λW2h+1λ2V,W2h,

    where λ=1V2h>0, and

    β(W)=β(W)=biWi=ViλWi=1λV,Wh.

    It follows that

    F(W)=1λW2h+1λ2V,W2h+1λV,Wh,

    hence F(W)<1 is equivalent to

    λW2h+V,W2h+V,Wh<λ,

    where we use λ>0 due to the fact that F is positive defined Randers metric. Therefore, we successively obtain

    λW2h+V,W2h<{λV,Wh}2,λW2h+V,W2h<λ22λV,Wh+V,W2h,λW2h<λ22λV,Wh,W2h<λ2V,Wh,W2h<1V2h2V,Wh,

    which is equivalent to V+W<1, hence ˜F is positive defined. The converse implication is trivial.

    Lemma 3.4. If ˜F=˜α+˜β is the Randers metric obtained in Lemma 3.3, then we have

    ˜α2=1η(α2β2)+˜Wη,yα,˜β=˜Wη,y,

    where

    η:=[1+F(W)][1F(W)],˜Wi:=Wibi[1+β(W)], and Wi=aijWj.

    Proof. Since the Zermelo's navigation data for ˜F is (h,U:=V+W), as shown in Lemma 3.3, it follows (see [1,11])

    ˜aij=1σhij+UiσUjσ,˜bi=Uiσ, (3.7)

    where

    Ui=hijUj=hij(Vi+Wi),σ:=1V+W2h.

    Recall that the navigation data (h,V) of a Randers metric F=α+β can be computed by

    hij=ε(aijbibj),Vi=biε,

    where ε:=1b2α, bi=aijbj (see [1], p. 233). Observe that as value ε=1b2α=1V2h=λ.

    We have

    V,Wh=hijViWj=ε(aijbibj)(biε)Wj=(aijbiWjbibibjWj)(bjWjb2αbjWj)=(β(W)b2αβ(W))=εβ(W),

    i.e.,

    V,Wh=εβ(W)

    and

    W2h=ε(aijbibj)WiWj=ε{α2(W)β2(W)}.

    It results

    σ=1U2h=1V2h2V,WhW2h=ε+2εβ(W)ε{α2(W)β2(W)}=ε{1+2β(W)+β2(W)α2(W)}=ε{[1+β(W)]2α2(W)}=ε[1+β(W)+α(W)][1+β(W)α(W)]=ε[1+F(W)][1F(W)],

    i.e.,

    σ=εη, (3.8)

    where η=[1+F(W)][1F(W)].

    Moreover, we have

    Ui=hijUj=hij(Vj+Wj)=ε(aijbibj)Vj+ε(aijbibj)Wj=ε(aijbibj)(bjε)+ε(aijbibj)Wj=[bibib2α]+ε[Wibiβ(W)]=εbi+ε[Wibiβ(W)]=ε{Wibi[1+β(W)]}=ε˜Wi,

    i.e., U=ε˜W.

    With these results, we compute

    ˜aij=1σhij+UiσUjσ=1εηε(aijbibj)+ε˜Wiεηε˜Wjεη=1η(aijbibj)+˜Wiη˜Wjη

    and

    ˜bi=Uiσ=ε˜Wiεη=˜Wiη,

    hence the conclusion follows.

    Remark 3.1. We observe that ˜F=˜α+˜β is positive defined if and only if ˜b˜α<1, i.e., σ=1U2h=1˜b2˜α>0.

    On the other hand, (3.8) implies that

    σ>0  ε[1+F(W)][1F(W)]>0  1F(W)>0,

    since ε>0 due to the fact that F is assumed positive defined and F(W)>0.

    In other words, we have shown that

    F(W)<1  ˜b˜α<1,

    that is another proof and more a intuitive explanation of positive definiteness condition F(W)<1 (compare to [6]).

    We will show a generic result on geodesics, conjugate and cut loci of a Randers metric.

    Lemma 3.5. Let (M,F=α+β) be a not flat Randers metric, let WX(M) be a vector field on M such that F(W)<1 and let ˜F=˜α+˜β be the solution of navigation problem for (F,W). If W is F-Killing field, then

    (i) the ˜F-unit speed geodesics ˜P are given by

    ˜P(t)=ψt(P(t)),

    where P is an F-unit speed geodesic and ψt is the flow of W;

    (ii) the point ˜P(l) is conjugate to q=˜P(0) along the ˜F-geodesic ˜P:[0,l]M if and only if the point P(l) is conjugate to q=P(0) along the corresponding F-geodesic P(t)=ψt(˜P(t)), for t[0,l];

    (iii) the point ˆp is an ˜F-cut point of q if and only if p=ψl(ˆp) is an F-cut point of q,

    where l=d˜F(q,ˆp).

    Proof. We will prove (ⅰ).

    For simplicity, if we also denote by ψt:TMTM the flow of XW, then for a curve P(t) on TM we denote

    ˆP(t)=ψt(P(t)),

    i.e., we map P(t)ˆP(t) by the flow ψt.

    By taking the tangent map

    (ψt,)P(t):TP(t)(TM)TˆP(t)(TM),

    we have

    X|P(t)(ψt,)P(t)(X|P(t))=(ψt,X)ˆP(t),

    for any vector field X on TM.

    If P(t) is an integral curve of the Hamiltonian vector field XK, i.e. dP(t)dt=XK|P(t), where K is the Legendre dual of F, then the derivative formula of a function of two variables give

    ddt(ˆP(t))=ddtψ(t,P(t))=XW|P(t)+ψt,(dP(t)dt)=XW|ˆP(t)+ψt,(XK|P(t))=XW|ˆP(t)+(ψt,XK)ˆP(t)=XW|ˆP(t)+(XK)ˆP(t)=(XW+K)ˆP(t)=(X˜K)ˆP(t),

    where we have used that the Legendre dual of ˜F is ˜K=K+W, and ψt,XK=XK (see Lemmas 3.1 and 3.2), hence (ⅰ) is proved.

    Next, we will prove (ⅱ).

    If we denote by Ps:[0,l]M, ε<s<ε a geodesic variation of the F-geodesic P, such that all curves in the variation are F-geodesics, then we obtain the variation vector field

    J:=Pss|s=0,

    which clearly is an F-Jacobi field.

    Taking now into account (i), which shows that

    ˜J=ψ(J)

    is a Jacobi vector field along ˜P, hence the conjugate points along P and ˜P correspond each other under the flow ψt of W, hence (ⅱ) is proved.

    Finally, we will prove (ⅲ). From (ⅱ) it is easy to see that since W is F-Killing field, the arclength parameter of the F-geodesic P and of the ˜F-geodesic ˜P coincide.

    It can be seen, like in the Riemannian case, that the points where the distance function dF(p,) looses its differentiability coinciding by the flow ψt to the points where the distance function d˜F(p,) looses its differentiability (see [12], Theorem A for the characterization of cut points in terms of differentiability of distance function). Hence, (ⅲ) follows.

    Lemma 3.6. Let (M,F=α+β) be a Randers metric with navigation data (h,W). The followings are equivalent

    (i) dβ=0,

    (ii) dW#=dlogλW#,

    where the one-form W# is the h-Legendre transformation of W and λ=1W2h.

    Proof. Indeed, observe that from the Zermelo's navigation formulas we get (see for instance (3.7), or [1,2,11]) we get

    β=Wiλdxi=1λW#,

    where W#=LhW. Here, Lh is the Legendre transform with respect to h.

    By differentiation, we get

    dβ=d(1λW#)=[1λ2dλW#+1λdW#]=1λ[dlogλW#+dW#],

    hence the desired equivalence follows.

    Summing up, here is our main result.

    Theorem 3.1. Let (M,h) be a Riemannian manifold and let V0,V,WX(M) be vector fields on M. If V0h<1, we denote by F0=α0+β0 the positive defined Randers metric obtained as solution of the Zermelo's navigation problem (h,V0).

    (i) (i.1) If F0(V)<1, then F1=α1+β1 is a positive defined Randers metric, where F1 is the solution of Zermelo's navigation problem (F0,V).

    (i.2) If F1(W)<1, then F2=α2+β2 is a positive defined Randers metric, where F2 is the solution of Zermelo's navigation problem (F1,W).

    (ii) (ii.1) The Randers metric F1=α1+β1 is the solution of Zermelo's navigation problem (h,V0+V).

    (ii.2) The Randers metric F2=α2+β2 is the solution of Zermelo's navigation problem (h,V0+V+W).

    (iii) If the following conditions are satisfied

    (C0) V0 is h-Killing,

    (C1) V is F-Killing,

    (C2) d(V0+V+W)#=dlog˜λ(V0+V+W),

    where (V0+V+W)#=Lh(V0+V+W) is the Legendre transformation of V0+V+W with respect to h, and ˜λ:=1V0+V+W2h, then

    (iii.1) The F0-unit speed geodesics P0, and the F1-unit speed geodesics P1 are given by

    P0(t)=φt(ρ(t)),P1(t)=ψt(P0(t))=ψtφt(ρ(t)), (3.9)

    where ρ(t) is an h-unit speed geodesic and φt and ψt are the flows of V0, and V, respectively.

    The F2-unit speed geodesic P2(t) coincides as points set with P1(t).

    (iii.2) The conjugate points of q=P2(0) along the F2-geodesic P2 coincide to the conjugate points of q=P1(0) along the F1-geodesic P1, up to parameterization. The point P1(l) is conjugate to q=P1(0) along the F1-geodesic P1:[0,l]M if and only if the point P0(l) is conjugate to q=P0(0) along the corresponding F0-geodesic P0(t)=ψt(P1(t)), for t[0,l]. The point P0(l) is conjugate to q=P0(0) along the F0-geodesic P0:[0,l]M if and only if the point ρ(l) is conjugate to q=ρ(0) along the corresponding h-geodesic ρ(t)=φt(P0(t)), for t[0,l], where φt, and ψt are the flows of V0, and V, respectively.

    (iii.3) The F2-cut locus of q coincide as points set with the F1cut locus of q, up to parameterization.

    The point ˆp1 is an F1-cut point of q, if and only if ˆp0=ψl(ˆp1) is an F1-cut point of q, where l=dF1(q,ˆp1). The point ˆp0 is an F0-cut point of q, if and only if p0=φl(ˆp0) is an h-cut point of q, where l=dF0(q,ˆp0).

    Proof of (i), (ii). The proof of (ⅰ.1), (ⅱ.1) follows immediately from Lemma 4 for (F0,V). Likewise, (ⅰ.2) and (ⅱ.2) follows from Lemma 4 for (F1,W).

    Proof of (iii). The proof will be given in two steps.

    Step 1. (Properties of F0, F1) With the notations in hypothesis, conditions (C0), (C1) imply that the geodesics, conjugate points and cut points of the Randers metrics F0, F1 have the properties in (ⅲ) due to Lemma 3.5.

    Step 2. (Properties of F2) By taking into account Lemma 3.6 one can see that condition (C2) is actually equivalent to d˜β=0, that is the Randers metrics F1=α+β and F2=˜α+˜β are projectively related ([3]), therefore having the same geodesics as non-parameterized curves, same conjugate points and same cut points. Hence, the desired properties of F2 follows (see Figure 6).

    Figure 6.  The unit-speed geodesics ρ,ρ0,ρ1 and P2 of h,F0,F1 and F2, respectively.

    Remark 3.2. Under the hypotheses in the Theorem 3.1 we have

    (i) conditions (C0), (C1) are equivalent to (C1), (C2), where

    (C2): V is h-Killing;

    (ii) if we replace conditions (C0), (C1), (C2) with

    (C3): V0+V+W is h-Killing,

    then the F2-geodesics, conjugate locus and cut locus is obtained from the h-geodesics, conjugate locus and cut locus deformed by the flow of V0+V+W, respectively. Observe that in this case, the F2-geodesics, conjugate locus and cut locus are different from these in Theorem 3.1.

    Remark 3.3. The construction presented here can now be extended to a sequence of Finsler metrics.

    Our sequence construction has two steps. Let (M,h) be a Riemannian two-sphere of revolution, and V0,V1,,Vk1,Wk,,WnX(M), n=kl, a sequence of vector fields on M.

    Step 1. A sequence of vector fields: V0,V1,,Vk1, such that all Vi are Fi-Killing fields, i{0,1,,k1},

    (M,h)V0,V0h<1h-Killing (M,F1=α1+β2)V1,F1(V1)<1F1-Killing (M,F2=α2+β2)V2,F2(V2)<1F2-Killing Vk1,Fk1(Vk1)<1Fk1-Killing (M,Fk=αk+βk)

    Step 2. A sequence of vector fields: Wk.,Wl, such that each βj is closed one-form, for j{k,k+1,,l}.

    (M,Fk+αk+βk)Wk,Fk(Wk)<1(M,Fk+1=αk+1+βk+1)Wk+1,Fk+1(Wk+1)<1(M,Fk+2=αk+2+βk+2)                                                                 dβk+1=0                                                 dβk+2=0Wk+2,Fk+2(Wk+2)<1Wn1,Fn1(Wn1)<1(M,Fn=αn+βn).                                                                       dβn=0

    Theorem 3.1 can be naturally extended to the two-step construction above. Indeed, if we start with a Riemannian structure (M,h) and a sequence of vector fields V0,V1,,Vk1X(M), the Zermelo's navigation problems for

    (h,V0) with solution F1=α1+β1,(F1,V1) with solution F2=α2+β2,(Fk1,Vk1) with solution Fk=αk+βk,

    will generate a sequence of positive defined Randers metrics provided V0h<1, Fi(Vi)<1, i{1,,k1}. The Zermelo's navigation data for Fi is also (h,V0++Vi), for all i{1,,k1}, hence Fk is positive defined if and only if V0++Vk1h<1.

    Next, if we start with (M,Fk) and the sequence of vector fields Wk,,Wn1X(M) the Zermelo's navigation problems for

    (Fk=αk+βk,Wk) with solution Fk+1=αk+1+βk+1,(Fk+1=αk+1+βk+1,Wk+1) with solution Fk+2=αk+2+βk+2,(Fn1,Wn1) with solution Fn=αn+βn,

    will generate another sequence of positive defined Randers metrics provided Fk+j(Wk+j)<1, j{0,1,2,,nk1}.

    Observe again that by combining these with the sequence of Randers metrics constructed at first step, we can easily see that the Zermelo's navigation data of Fk+j, j{0,1,,nk} is (h,V0++Vk1+Wk++Wk+j), hence the final Randers metric Fn=αn+βn is positive defined if and only if

    k1i=0Vi+nk1j=0Wj+kh<1.

    Moreover, if we impose conditions

    (C0) V0 is h-Killing;

    (C1i) Vi is Fi-Killing, i{1,,k1};

    (C2j) Wk+j is chosen such that dβk+j=0, j{0,,nk}.

    Clearly the geodesics, conjugate and cut loci of Fn can be obtained from the geodesics, conjugate locus, cut locus of h through the flow of V:=k1i=0Vi, respectively.

    Observe that condition (C2j) are similar to (C2) in Theorem 3.1, but we prefer not to write them here explicitly, for simplicity.

    This is the generalization of Theorem 3.1 to the sequence of Finsler metrics {F1,,Fn}.

    Nevertheless, there is a shortcut in this construction in the spirit of Remark 3.2. Indeed, if V+W is h-Killing, where V=k1i=0Vi, W=nk1j=0Wk+j, then the geodesics, conjugate and cut loci of Fn are obtained from the geodesics, conjugate and cut loci of h through the flow of V+W, respectively.

    We will consider a simple example of the construction described in Theorem 3.1.

    Let us start with the Riemannian two-sphere of revolution (MS2,h=dr2+m2(r)dθ2) given in Section 2, Proposition 2.1. The vector field V0X(M) is h-Killing if and only if it is a rotation, i.e., V0=μ0θ, μ0 constant, where (r,θ) are the h-geodesic coordinates. In order that F0 is positive defined we need the condition V0h<1, i.e. μ20m2(r)<1.

    Next, we consider the vector field VX(M) which is also h-Killing if and only if V=μ1θ with μ1 constant (see Remark 3.2). The Randers metric F1=α1+β1 is positive defined if and only if (μ0+μ1)2m2(r)<1, i.e., we choose μ0,μ1 such that m(r)<1μ0+μ1.

    Finally, we construct a vector field WX(M) such that dβ=0. For instance

    W=A(r)rμθ,

    where μ:=μ0+μ1 is an obvious choice. Observe that V0+V+W=A(r)r, hence β2=A(r)1A2(r)dr. If we impose condition A2(r)<1, then F2 is a positive defined Randers metric.

    We obtain

    Proposition 4.1. Let (MS2,h=dr2+m2(r)dθ2) be the Riemannian two-sphere of revolution described in Proposition 2.1. Let

    V0=μ0θ, V=μ1θ, W=A(r)rμθ,

    be three vector fields on M, where μ=μ0+μ1.

    (i) If m(r)<1μ for all r[0,π] and A:[0,π][0,) is smooth function such that A2(r)<1, then the Finsler metrics F0=α0+β0, F1=α1+β1, F2=α2+β2, obtained as solutions of Zermelo's navigation problem with data (h,V0), (F0,V) and (F1,W), respectively, are positive defined Randers metrics.

    (ii) The Randers metrics F1=α1+β1 and F2=α2+β2 can be obtained as solutions of Zermelo's navigation problem with data (h,μθ) and (h,A(r)r), respectively.

    (iii) iii.1 The unit speed F2-geodesics are given by

    P(t)=ψt(ρ(t)),

    where ρ are unit speed h-geodesics and ψt is the flow of ˜V=V0+V+W=A(r)r.

    iii.2 The point ˆp=P(l) is conjugate to ˆq:=P(0) along the F2-geodesic P:[0,l]M if and only if q=P(0)=ρ(0) is conjugate to p:=ρ(l) along the corresponding h-geodesic ρ(t)=ψt(P(t)), t[0,l].

    iii.3 The cut locus of a point q(M,F2) is a subarc if the antipodal parallel displaced by the flow φt.

    One can describe the Finsler metric F2 in coordinates as follows. If h is given by ds2=dr2+m2(r)dθ2 in the geodesic coordinates (r,θ)(0,π]×[0,2π), then

    α22=1˜λ2(r)dr2+m(r)˜λds2,β2=A(r)˜λ(r)dr, ˜λ(r):=1A2(r).

    More precisely, if m(r)=112αsin(rαsin2r) (see Example 2.1), for any α(0,12) and A(r):=rr2+1, then ˜λ=1r2+1 hence the Finsler metric F2 is given by

    α22=(r2+1)2dr2+112α(rαsin2r)(r2+1)dθ2,β2=rr2+1 dr, r(0,π], θ[0,2π).

    Other examples can be similarly constructed from the Riemannian examples in [16].

    Remark 4.1. Observe that in order to construct Randers metric having same cut locus structure as the Riemannian metric h, another condition is also possible. Indeed, choosing

    V0:=v0(r,θ)r+w0(r,θ)θ,V:=v0(r,θ)r+[μw0(r,θ)]θ,

    will lead to V0+V=μθ which is h-Killing and combined with W=A(r)rμθ the derived conclusion follows, for any smooth functions v0,w0 and constant μ such that m(r)<1μ.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors are grateful to professors H. Shimada and K. Kiyohara for many useful discussions. The first author is financially supported by Office of the Permanent Secretary, Ministry of Higher Education, Science, Research and Innovation. Grant No. RGNS 63-241. The second author is partially supported by the JSPS grant number 20K03595.

    The authors declare no conflict of interest.



    [1] E. Bannai, T. Ito, Algebraic combinatorics Ⅰ: Association schemes, Menlo Park, Calif.: Benjamin/Cummings Pub. Co., 1984.
    [2] R. C. Bose, T. Shimamoto, Classification and analysis of partially balanced incomplete block designs with two associate classes, J. Am. Stat. Assoc., 47 (1952), 151–184. http://doi.org/10.1080/01621459.1952.10501161 doi: 10.1080/01621459.1952.10501161
    [3] Y. Huo, Z. Wan, Non-symmetric association schemes of symmetric matrices, Acta Math. Appl. Sin.-E., 9 (1993), 236–255. http://doi.org/10.1007/BF02032918 doi: 10.1007/BF02032918
    [4] C. Ma, Y. Wang, Automorphisms of association schemes of quadratic forms over a finite field of chatacteristic two, Algebra Colloq., 10 (2003), 63–74. http://doi.org/10.1007/s100110300008 doi: 10.1007/s100110300008
    [5] Y. Wang, Y. Huo, C. Ma, J. Ma, Association schemes of matrices, Beijing: Science Press, 2011.
    [6] Y. Wang, C. Wang, C. Ma, Association schemes of quadratic forms over a finite field of characteristic two, Chin. Sci. Bull., 43 (1998), 1965–1969. http://doi.org/10.1007/BF03186985 doi: 10.1007/BF03186985
    [7] Y. Wang, C. Wang, C. Ma, J. Ma, Association schemes of quadratic forms and symmetric bilinear forms, J. Algebr. Comb., 17 (2003), 149–161. http://doi.org/10.1023/A:1022978613368 doi: 10.1023/A:1022978613368
  • This article has been cited by:

    1. Bengisen Pekmen Geridönmez, Magnetic source effect on EG-CuO nanofluid in a semi-annulus using RBFs, 2019, 20, 1550-2287, 201, 10.1080/15502287.2019.1604583
    2. Bengisen Pekmen Geridonmez, Free Convection in a Wavy Walled Cavity With a Magnetic Source Using Radial Basis Functions, 2019, 141, 0022-1481, 10.1115/1.4042782
    3. Abuzar Abid Siddiqui, Mustafa Turkyilmazoglu, Natural convection in the ferrofluid enclosed in a porous and permeable cavity, 2020, 113, 07351933, 104499, 10.1016/j.icheatmasstransfer.2020.104499
    4. Abuzar Abid Siddiqui, Mustafa Turkyilmazoglu, A New Theoretical Approach of Wall Transpiration in the Cavity Flow of the Ferrofluids, 2019, 10, 2072-666X, 373, 10.3390/mi10060373
    5. B. Pekmen Geridonmez, H.F. Oztop, MHD natural convection in a cavity in the presence of cross partial magnetic fields and Al2O3-water nanofluid, 2020, 80, 08981221, 2796, 10.1016/j.camwa.2020.10.003
    6. Abdelraheem M. Aly, Zehba A.S. Raizah, Coupled fluid-structure interactions of natural convection in a ferrofluid using ISPH method, 2019, 58, 11100168, 1499, 10.1016/j.aej.2019.12.004
    7. Bullo Hindebu Rikitu, Oluwole Daniel Makinde, Lemi Guta Enyadene, Modeling Heat Transfer Enhancement of Ferrofluid (Fe3O4–H2O) Flow in a Microchannel Filled with a Porous Medium, 2021, 10, 2169-432X, 31, 10.1166/jon.2021.1764
    8. Pekmen Geridonmez, Hakan Oztop, Natural convection in a sinusoidally heated cavity filled with ferrofluid in the presence of partial variable magnetic field, 2023, 33, 0961-5539, 411, 10.1108/HFF-01-2022-0053
    9. Aneesa Nadeem, Nejla Mahjoub Said, LSM analysis of thermal enhancement in KKL model-based unsteady nanofluid problem using CCM and slanted magnetic field effects, 2023, 1388-6150, 10.1007/s10973-023-12801-1
    10. Bengisen Pekmen Geridonmez, Hakan Oztop, Magnetotactic bacteria and Fe3O4–water in a wavy walled cavity, 2024, 0961-5539, 10.1108/HFF-08-2023-0465
    11. Bengisen Pekmen Geridonmez, Modeling of Average Nusselt Number by Machine Learning and Interpolation Techniques, 2024, 146, 2832-8450, 10.1115/1.4064562
    12. B. Pekmen Geridonmez, H. F. Oztop, A machine learning approach for entropy due to natural convection flow of a nanofluid under the uniform inclined magnetic field, 2024, 1040-7790, 1, 10.1080/10407790.2024.2338912
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(575) PDF downloads(22) Cited by(0)

Figures and Tables

Tables(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog