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

Ricci curvature of semi-slant warped product submanifolds in generalized complex space forms

  • Received: 19 July 2021 Revised: 24 January 2022 Accepted: 26 January 2022 Published: 08 February 2022
  • MSC : 53C25, 53C40, 53C42, 53D15

  • The objective of this paper is to achieve the inequality for Ricci curvature of a semi-slant warped product submanifold isometrically immersed in a generalized complex space form admitting a nearly Kaehler structure in the expressions of the squared norm of mean curvature vector and warping function. In addition, the equality case is likewise discussed. We provide numerous physical applications of the derived inequalities. Later, we proved that under a certain condition the base manifold Nn1T is isometric to a n1-dimensional sphere Sn1(λ1n1) with constant sectional curvature λ1n1.

    Citation: Ali H. Alkhaldi, Meraj Ali Khan, Shyamal Kumar Hui, Pradip Mandal. Ricci curvature of semi-slant warped product submanifolds in generalized complex space forms[J]. AIMS Mathematics, 2022, 7(4): 7069-7092. doi: 10.3934/math.2022394

    Related Papers:

    [1] Amira A. Ishan, Meraj Ali Khan . Chen-Ricci inequality for biwarped product submanifolds in complex space forms. AIMS Mathematics, 2021, 6(5): 5256-5274. doi: 10.3934/math.2021311
    [2] Ibrahim Al-Dayel, Meraj Ali Khan . Ricci curvature of contact CR-warped product submanifolds in generalized Sasakian space forms admitting nearly Sasakian structure. AIMS Mathematics, 2021, 6(3): 2132-2151. doi: 10.3934/math.2021130
    [3] Tong Wu, Yong Wang . Super warped products with a semi-symmetric non-metric connection. AIMS Mathematics, 2022, 7(6): 10534-10553. doi: 10.3934/math.2022587
    [4] Lamia Saeed Alqahtani, Akram Ali . The eigenvalues of $ \beta $-Laplacian of slant submanifolds in complex space forms. AIMS Mathematics, 2024, 9(2): 3426-3439. doi: 10.3934/math.2024168
    [5] Ali H. Alkhaldi, Akram Ali, Jae Won Lee . The Lawson-Simons' theorem on warped product submanifolds with geometric information. AIMS Mathematics, 2021, 6(6): 5886-5895. doi: 10.3934/math.2021348
    [6] Meraj Ali Khan, Ali H. Alkhaldi, Mohd. Aquib . Estimation of eigenvalues for the $ \alpha $-Laplace operator on pseudo-slant submanifolds of generalized Sasakian space forms. AIMS Mathematics, 2022, 7(9): 16054-16066. doi: 10.3934/math.2022879
    [7] Aliya Naaz Siddiqui, Mohammad Hasan Shahid, Jae Won Lee . On Ricci curvature of submanifolds in statistical manifolds of constant (quasi-constant) curvature. AIMS Mathematics, 2020, 5(4): 3495-3509. doi: 10.3934/math.2020227
    [8] Noura Alhouiti, Fatemah Mofarreh, Fatemah Abdullah Alghamdi, Akram Ali, Piscoran-Ioan Laurian . Geometric topology of CR-warped products in six-dimensional sphere. AIMS Mathematics, 2024, 9(9): 25114-25126. doi: 10.3934/math.20241224
    [9] Fahad Sikander, Tanveer Fatima, Sharief Deshmukh, Ayman Elsharkawy . Curvature analysis of concircular trajectories in doubly warped product manifolds. AIMS Mathematics, 2024, 9(8): 21940-21951. doi: 10.3934/math.20241066
    [10] Yanlin Li, Yuquan Xie, Manish Kumar Gupta, Suman Sharma . On projective Ricci curvature of cubic metrics. AIMS Mathematics, 2025, 10(5): 11305-11315. doi: 10.3934/math.2025513
  • The objective of this paper is to achieve the inequality for Ricci curvature of a semi-slant warped product submanifold isometrically immersed in a generalized complex space form admitting a nearly Kaehler structure in the expressions of the squared norm of mean curvature vector and warping function. In addition, the equality case is likewise discussed. We provide numerous physical applications of the derived inequalities. Later, we proved that under a certain condition the base manifold Nn1T is isometric to a n1-dimensional sphere Sn1(λ1n1) with constant sectional curvature λ1n1.



    The realization of warped product manifolds came into existence after the approach of R. L. Bishop and B. O'Neill [7] on manifolds of negative curvature. Examining the fact that a Riemannian product of manifolds can not have negative curvature, they construct the model of warped product manifolds for the class of manifolds of negative (or non positive) curvature which is defined as follows:

    Let (N1,g1) and (N2,g2) be two Riemannian manifolds with Riemannian metrics g1 and g2 respectively and ψ be a positive smooth function on N1. If π:N1×N2N1 and η:N1×N2N2 are the projection maps given by π(p,q)=p and η(p,q)=q for every (p,q)N1×N2, then the warped product manifold is the product manifold N1×N2 equipped with the Riemannian structure such that

    g(X,Y)=g1(πX,πY)+(ψπ)2g2(ηX,ηY),

    for all X,YTM. The function ψ is called the warping function of the warped product manifold. If the warping function is constant, then the warped product is trivial, i.e., simply Riemannian product. On the grounds that warped product manifolds admit a number of applications in Physics and theory of relativity [5], this has been a topic of extensive research. Warped products provide many basic solutions to Einstein field equations [5]. The concept of modelling of space-time near black holes adopts the idea of warped product manifolds [19]. Schwartzschild space-time is an example of warped product P×rS2, where the base P=R×R+ is a half plane r>0 and the fibre S2 is the unit sphere. Under certain conditions, the Schwartzchild space-time becomes the black hole. A cosmological model to model the universe as a space-time known as Robertson-Walker model is a warped product [30].

    Some natural properties of warped product manifolds were studied in [7]. B. Y. Chen (see [9,11]) performed an extrinsic study of warped product submanifolds in a Kaehler manifold. Since then, many geometers have explored warped product manifolds in different settings like almost complex and almost contact manifolds and various existence results have been investigated (see the survey article [13]).

    In 1999, Chen [10] discovered a relationship between Ricci curvature and squared mean curvature vector for an arbitrary Riemannian manifold. On the line of Chen, a series of articles have been appeared to formulate the relationship between Ricci curvature and squared mean curvature in the setting of some important structures on Riemannian manifolds (see [3,4,16,20,21,22,26,27,34]). Recently Ali et al. [1] established a relationship between Ricci curvature and squared mean curvature for warped product submanifolds of a sphere and provide many physical applications.

    In this paper our aim is to obtain a relationship between Ricci curvature and squared mean curvature for semi-slant warped product submanifolds in the setting of generalized complex space form admitting a nearly Kaehler structure. Further, we provide some applications in terms of Hamiltonians and Euler-Lagrange equation. In the last we also worked out some applications of Obata's differential equation.

    Let ˉM be an almost Hermitian manifold with an almost complex structure J and Riemannian metric g satisfying the following

    J2=I,g(JX,JY)=g(X,Y), (2.1)

    for all vector fields X,Y on ˉM. If almost Hermitian manifold satisfies the following property

    (ˉXJ)Y+(ˉYJ)X=0 (2.2)

    for all vector fields X,YTˉM, then ˉM is called the nearly Kaehler manifold. The six dimensional sphere S6 is an example of nearly Kaehler manifold which is not a Kaehler manifold. S6 has an almost complex structure J defined by the vector cross product in the space of purely imaginary Cayley numbers which satisfies the tensorial equation of nearly Kaehler manifold. There is a more general class of almost Hermitian manifolds than nearly Kaehler manifold, this class is known as RK-manifold. A generalized complex space form is an RK-manifold of constant holomorphic sectional curvature c and of constant α and is denoted by ˉM(c,α). The sphere S6 endowed with the standard nearly Kaehler structure is an example of generalized complex space form which is not a complex space form. The curvature tensor ˉR of a generalized complex space form ˉM(c,α) is given by

    ˉR(X,Y,Z,W)=c+3α4[g(Y,Z)g(X,W)g(X,Z)g(Y,W)]+cα4[g(X,JZ)g(JY,W)g(Y,JZ)g(JX,W)+2g(X,JY)g(JZ,W)], (2.3)

    for any X,Y,Z,WTˉM.

    Let M be an n-dimensional Riemannian manifold isometrically immersed in a m-dimensional Riemannian manifold ˉM. Then the Gauss and Weingarten formulas are ˉXY=XY+h(X,Y) and ˉXξ=AξX+Xξ respectively, for all X,YTM and ξTM, where is the induced Levi-Civita connection on M, ξ is a vector field normal to M, h is the second fundamental form of M, is the normal connection in the normal bundle TM and Aξ is the shape operator of the second fundamental form. The second fundamental form h and the shape operator are associated by the following formula

    g(h(X,Y),ξ)=g(AξX,Y). (2.4)

    The equation of Gauss is given by

    R(X,Y,Z,W)=ˉR(X,Y,Z,W)+g(h(X,W),h(Y,Z))g(h(X,Z),h(Y,W)), (2.5)

    for all X,Y,Z,WTM, where ˉR and R are the curvature tensors of ˉM and M respectively. For any XTM and NTM, JX and JN can be decomposed as follows

    JX=PX+FX (2.6)

    and

    JN=tN+fN, (2.7)

    where PX (resp. tN) is the tangential and FX (resp. fN) is the normal component of JX (resp. JN).

    For any orthonormal basis {e1,e2,,en} of the tangent space TxM, the mean curvature vector H(x) and its squared norm are defined as follows

    H(x)=1nni=1h(ei,ei),H2=1n2ni,j=1g(h(ei,ei),h(ej,ej)), (2.8)

    where n is the dimension of M. If h=0 then the submanifold is said to be totally geodesic and minimal if H=0. If h(X,Y)=g(X,Y)H for all X,YTM, then M is called totally umbilical.

    The scalar curvature of ˉM is denoted by ˉτ(ˉM) and is defined as

    ˉτ(Mn)=1p<qmˉκpq, (2.9)

    where ˉκpq=ˉκ(epeq) and m is the dimension of the Riemannian manifold ˉM. Throughout this study, we shall use the equivalent version of the above equation, which is given by

    2ˉτ(Mn)=1p,qmˉκpq. (2.10)

    In a similar way, the scalar curvature ˉτ(Lx) of a L-plane is given by

    ˉτ(Lx)=1p<qmˉκpq. (2.11)

    Let {e1,,en} be an orthonormal basis of the tangent space TxM and if er belongs to the orthonormal basis {en+1,,em} of the normal space TM, then we have

    hrpq=g(h(ep,eq),er)) (2.12)

    and

    h2=np,q=1g(h(ep,eq),h(ep,eq)). (2.13)

    Let κpq and ˉκpq be the sectional curvatures of the plane sections spanned by ep and eq at x in the submanifold Mn and in the Riemannian space form ˉMm(c), respectively. Thus by Gauss equation, we have

    κpq=ˉκpq+mr=n+1(hrpphrqq(hrpq)2). (2.14)

    The global tensor field for orthonormal frame of vector field {e1,,en} on Mn is defined as

    ˉS(X,Y)=ni=1{g(ˉR(ei,X)Y,ei)}, (2.15)

    for all X,YTxMn. The above tensor is called the Ricci tensor. If we fix a distinct vector eu from {e1,,en} on Mn, which is governed by χ. Then the Ricci curvature is defined by

    Ric(χ)=np=1puK(epeu). (2.16)

    Consider the warped product submanifold N1×ψN2. Let X be a vector field on M1 and Z be a vector field on M2, then from Lemma 7.3 of [7], we have

    XZ=ZX=(Xψψ)Z (2.17)

    where is the Levi-Civita connection on M. For a warped product M=M1×ψM2 it is easy to observe that

    XZ=ZX=(Xlnψ)Z (2.18)

    for XTM1 and ZTM2.

    ψ is the gradient of ψ and is defined as

    g(ψ,X)=Xψ, (2.19)

    for all XTM.

    Let M be an n-dimensional Riemannian manifold with the Riemannian metric g and let {e1,e2,,en} be an orthogonal basis of TM. Then as a result of (2.19), we get

    ψ2=ni=1(ei(ψ))2. (2.20)

    The Laplacian of ψ is defined by

    Δψ=ni=1{(eiei)ψeieiψ}. (2.21)

    The Hessian tensor for a differentiable function ψ is a symmetric covariant tensor of rank 2 and is defined as

    Δψ=trace Hψ

    For the warped product submanifolds, we have following well known result [14]

    n1p=1n2q=1κ(epeq)=n2Δψψ=n2(Δlnψlnψ2). (2.22)

    Now, we state the Hopf's Lemma.

    Hopf's Lemma. [12] Let M be a m-dimensional connected compact Riemannian manifold. If ψ is a differentiable function on M such that Δψ0 everywhere on M (or Δψ0 everywhere on M), then ψ is a constant function.

    For a compact orientable Riemannian manifold M with or without boundary and as a consequences of the integration theory of manifolds, we have

    MΔψ dV=0, (2.23)

    where ψ is a function on M and dV is the volume element of M.

    The notion of semi-slant submanifolds of a Kaehler manifold is geometrically new and interesting. Infact, the study of differential geometry of semi-slant submanifolds as a generalization of CR-submanifolds and slant submanifolds of a Kaehlerian submanifolds was initiated by N. Papaghiuc [32]. In [23] V. A. Khan and M. A. Khan studied semi-slant submanifolds of a nearly Kaehler manifold and obtained some basic and interesting results. Further, B. Sahin [33] proved the non existence proper semi-slant warped product submanifolds in the setting of Kaehler manifold. So, it was natural to see the existence of semi-slant warped product submanifolds in a more general setting namely nearly Kaehler manifold and in this series V. A. Khan and K. A. Khan [24] studied different types of warped product submanifolds in nearly Kaehler manifolds. Suppose, NT and Nθ be the holomorphic and slant submanifolds of an almost Hermitian manifold ˉM. Now, there are two possibilities of warped product submanifolds of ˉM, these warped products are Nθ×ψNT and NT×ψNθ. In [24] V. A. Khan and K. A. Khan proved the non-existence of the first type of warped product Nθ×ψNT in nearly Kaehler manifolds and they studied the existence of the warped product NT×ψNθ, these warped product submanifolds are called semi-slant warped product submanifolds and studied extensively (see [2,24,25]). Throughout this study, we consider the warped product submanifolds M=Nn1T×ψNn2θ of a nearly Kaehler manifold, where n1 and n2 are the dimensions of the holomorphic and slant submanifolds.

    Now, we have the following initial result:

    Lemma 3.1. Let M=Nn1T×ψNn2θ be a semi-slant warped product submanifold isometrically immersed in a nearly Kaehler manifold ˉM. Then

    (i) g(h(X, Y), JZ) = 0,

    (ii) g(h(JX, JX), N) = - g(h(X, X), N),

    for any X,YTNT, ZTNθ and Nμ, where μ is the invariant subbundle of TM.

    Proof. By using Gauss and Weingarten formulae in Eq (2.2), we have

    XPZ+h(X,PZ)AFZX+XFZJXZJh(X,Z)+ZJX+
    +h(JX,Z)JZXJh(X,Z)=0,

    taking inner product with Y and using (2.4), we get the required result.

    To prove (ii), for any XTNT we have

    ˉXJX=(ˉXJ)X+JˉXX.

    Using Gauss formula and (2.2) in above, we get

    XJX+h(JX,X)=JXX+Jh(X,X).

    Taking inner product with JN, above equation yields

    g(h(JX,X),JN)=g(h(X,X),N). (3.1)

    Interchanging X by JX the above equation gives

    g(h(JX,X),JN)=g(h(JX,JX),N). (3.2)

    From (3.1) and (3.2), we get the required result.

    From the above result it is evident that the isometric immersion Nn1T×ψNn2θ into a nearly Kaehler manifold is DT-minimal. The DT- minimality property provides us a useful relationship between the semi-slant warped product submanifold NT×ψNθ and the equation of Gauss.

    Definition 3.1 The warped product N1×ψN2 isometrically immersed in a Riemannian manifold ˉM is called Ni totally geodesic if the partial second fundamental form hi vanishes identically. It is called Ni-minimal if the partial mean curvature vector Hi becomes zero for i=1,2.

    Let {e1,,ep,ep+1=Je1,,en1=Jep,en1+1=e1,,en1+q=eq,en1+q+1=eq+1=secθPe1,,e(n2=2q)=en2=secθPeq} be a local orthonormal frame of vector fields on the semi-slant warped product submanifold Mn=Nn1T×ψNn2θ such that the set {e1,,ep,ep+1=Je1,,en1=Jep} is tangent to NT and the set {e1,,eq,en2} is tangent to Nθ. Moreover, {en+1=cscθFe1,,en+n2=cscθFeq,en+n2+1=ˉe1,,em=ˉek} is a basis for the normal bundle TM, such that the set {en+1=cscθFe1,,en+n2=cscθFeq} is tangent to FDθ and {ˉe1,,ˉek} is tangent to the complementary invariant subbundle μ with even dimension k.

    From Lemma 3.1, it is easy to conclude that

    mr=n+1n1i,j=1g(h(ei,ej),Jer)=0. (3.3)

    Thus it follows that the trace of h due to NT becomes zero. Hence in view of the Definition 3.1, we obtain the following important result.

    Theorem 3.1. Let Mn=Nn1T×ψNn2θ be a semi-slant warped product submanifold isometrically immersed in a nearly Kaehler manifold. Then Mn is DT-minimal.

    So, it is easy to conclude the following

    H2=1n2mr=n+1(hrn1+1n1+1++hrnn)2, (3.4)

    where H2 is the squared mean curvature.

    In this section, we investigate Ricci curvature in terms of the squared norm of mean curvature and the warping function as follows:

    Theorem 4.1. Let M=Nn1T×ψNn2θ be a semi-slant warped product submanifold isometrically immersed in a generalized complex space form ˉM(c,α) admitting nearly Kaehler structure. Then for each orthogonal unit vector field χTxM, either tangent to NT or Nθ, we have

    (1) The Ricci curvature satisfy the following inequality

    (i) If χ is tangent to Nn1T, then

    Ric(χ)14n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)8. (4.1)

    (ii) If χ is tangent to Nn2θ, then

    Ric(χ)14n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)8cos2θ. (4.2)

    (2) If H(x)=0, then for each point xMn there is a unit vector field X which satisfies the equality case of (1) if and only if Mn is mixed totally geodesic and χ lies in the relative null space Nx at x.

    (3) For the equality case we have

    (a) The equality case of (4.1) holds identically for all unit vector fields tangent to Nn1T at each xMn if and only if Mn is mixed totally geodesic and DT-totally geodesic semi-slant warped product submanifold in ˉMm(c,α).

    (b) The equality case of (4.2) holds identically for all unit vector fields tangent to Nn2θ at each xMn if and only if M is mixed totally geodesic and either Mn is Dθ- totally geodesic semi-slant warped product or Mn is a Dθ totally umbilical in ˉMm(c,α) with dim Dθ=2.

    (c) The equality case of (1) holds identically for all unit tangent vectors to Mn at each xMn if and only if either Mn is totally geodesic submanifold or Mn is a mixed totally geodesic totally umbilical and DT-totally geodesic submanifold with dim Nn2θ=2.

    where n1 and n2 are the dimensions of NT and Nθ respectively.

    Proof. Suppose that M=Nn1T×ψNn2θ be a semi-slant warped product submanifold of a generalized complex space form. From Gauss equation, we have

    n2H2=2τ(Mn)+h22ˉτ(Mn). (4.3)

    Let {e1,,en1,en1+1,,en} be a local orthonormal frame of vector fields on Mn such that {e1,,en1} are tangent to NT and {en1+1,,en} are tangent to Nθ. So, the unit tangent vector χ=eA{e1,,en} can be expanded (4.3) as follows

    n2H2=2τ(Mn)+12mr=n+1{(hr11++hrnnhrAA)2+(hrAA)2}mr=n+11pqnhrpphrqq2ˉτ(Mn). (4.4)

    The above expression can be written as follows

    n2H2=2τ(Mn)+12mr=n+1{(hr11++hrnn)2+(2hrAA(hr11++hrnn))2}+2mr=n+11p<qn(hrpq)22mr=n+11p<qnhrpphrqq2ˉτ(Mn).

    In view of the Lemma 3.1, the preceding expression takes the form

    n2H2=mr=n+1{(hrn1+1n1+1++hrnn)2++(2hrAA(hrn1+1n1+1++hrnn))2}+2τ(Mn)+mr=n+11p<qn(hrpq)2mr=n+11p<qnhrpphrqq+mr=n+1a=1aA(hraA)2+mr=n+11p<qnp,qA(hrpq)2mr=n+11p<qnp,qAhrpphrqq2ˉτ(Mn). (4.5)

    By Eq (2.14), we have

    mr=n+11p<qnp,qA(hrpq)2mr=n+11p<qnp,qAhrpphrqq=1p<qnp,qAˉκpq1p<qnp,qAκpq (4.6)

    Substituting the values of Eq (4.6) in (4.5), we discover

    n2H2=2τ(Mn)+12mr=n+1(2hrAA(hrn1+1n1+1++hrnn))2+mr=n+11p<qn(hrpq)2mr=n+11p<qnhrpphrqq2ˉτ(Mn)+mr=n+1a=1aA(hraA)2+1p<qnp,qAˉκp,q1p<qnp,qAκpq. (4.7)

    Since, Mn=Nn1T×ψNn2θ, then from (2.11), the scalar curvature of Mn can be defined as follows

    τ(Mn)=1p<qnκ(epeq)=n1i=1nj=n1+1κ(eiej)+1r<kn1κ(erek)+n1+1l<onκ(eleo) (4.8)

    The usage of (2.11) and (2.22), we derive

    τ(Mn)=n2Δψψ+τ(Nn1T)+τ(Nn2θ) (4.9)

    Utilizing (4.9) together with (2.14) and (2.3) in (4.7), we have

    12n2H2=n2Δψψ+1p<qnp,qAˉκp,q+ˉτ(Nn1T)+ˉτ(Nn2θ)+mr=n+1{1p<qn(hrpq)21p<qnp,qAhrpphrqq}+mr=n+1a=1aA(hraA)2+mr=n+11ijn1(hriihrjj(hrij)2)+mr=n+1n1+1stn(hrsshrtt(hrst)2)+12mr=n+1(2hrAA(hrn1+1n1+1++hrnn))2c+3α4(n(n1))(cα)4(3n1+3n2cos2θ). (4.10)

    Considering unit tangent vector χ=ea, we have two choices: χ is either tangent to the base manifold Nn1T or to the fibre Nn2θ.

    Case i: If ea is tangent to Nn1T, then fix a unit tangent vector from {e1,,en1} and suppose χ=ea=e1. Then from (4.10) and (2.16), we find

    Ric(χ)12n2H2n2Δψψ12mr=n+1(2hr11(hrn1+1n1+1+hrnn))2mr=n+11p<qn1(hrpq)2+mr=n+1[1i<jn1(hrij)21i<jn1hriihrjj]+mr=n+1n1+1s<tn(hrst)2+mr=n+1[n1+1s<tn(hrij)2n1+1s<tnhrsshrtt]+mr=n+12p<qnhrpphrqq+c+3α4(n(n1))+(cα)4(3n1+3n2cos2θ)2p<qnˉκpqˉτ(Nn1T)ˉτ(Nn2θ). (4.11)

    From (2.3), (2.11) and (2.12), we have

    2p<qnˉκp,q=c+3α8(n1)(n2)+cα8[3(n11)+3n2cos2θ], (4.12)
    ˉτ(Nn1T)=c+3α8n1(n11)+cα83n1, (4.13)
    ˉτ(Nn2θ)=c+3α8n2(n21)+cα83n2cos2θ. (4.14)

    Using (4.19)–(4.21) in (4.11), we have

    Ric(χ)12n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)812mr=n+1(2hr11(hrn1+1n1+1++hrnn))2mr=n+11p<qn(hrpq)2+mr=n+1[1i<jn1(hrij)2+mr=n+1n1+1s<tn(hrst)2]mr=n+1[1i<jn1hriihrjj+n1+1s<tnhrsshrtt]+mr=n+12p<qnhrpphrqq. (4.15)

    Further, the sixth and seventh terms on right hand side of the last inequality can be written as

    mr=n+1[1i<jn1(hrij)2+n1+1s<tn(hrst)2]mr=n+11p<qn(hrpq)2=mr=n+1n1p=1nq=n1+1(hrpq)2.

    Similarly, we have

    mr=n+1[1i<jn1hriihrjj+n1+1stnhrsshrtt2p<qnhrpphrqq]=mr=n+1[n1p=2nq=n1+1hrpphrqqn1j=2hr11hrjj].

    Utilizing above two values in (4.15), we get

    Ric(χ)12n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)812mr=n+1(2hr11(hrn1+1n1+1+hrnn))2mr=n+1[n1p=1nq=n1+1(hrpq)2+n1b=2hr11hrbbn1p=2nq=n1+1hrpphrqq]. (4.16)

    Since Mn=Nn1T×ψNn2θ is Nn1T-minimal then we can observe the following

    mr=n+1n1p=2nq=n1+1hrpphrqq=mr=n+1nq=n1+1hr11hrqq (4.17)

    and

    mr=n+1n1b=2hr11hrbb=mr=n+1(hr11)2. (4.18)

    Simultaneously, we can conclude

    12mr=n+1(2hr11(hrn1+1n1+1++hrnn))2+mr=n+1nq=n1+1hr11hrqq=2mr=n+1(hr11)2+12n2H2. (4.19)

    Using (4.17) and (4.18) in (4.16), after the assessment of (4.19), we finally get

    Ric(χ)12n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)814mr=n+1nq=n1+1(hrqq)2mr=n+1{(hr11)2nq=n1+1hr11hrqq+14(hrn1+1n1+1++hrnn)2}. (4.20)

    Further, using the fact that mr=n+1(hrn1+1n1+1++hrnn)=n2H2, we get

    Ric(χ)14n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)814mr=n+1(2hr11nq=n1+1hrqq)2. (4.21)

    From the above inequality, we can conclude the inequality (4.1).

    Case ii: If ea is tangent to Nn2θ, then we choose the unit vector from {en1+1,,en}, suppose that the unit vector is en, i.e. χ=en. Then from (2.3), (2.11) and (2.12), we have

    1p<qn1ˉκpq=c+3α8(n1)(n2)cα8(3n1+3(n21)cos2θ), (4.22)
    ˉτ(Nn1T)=c+3α8n1(n11)+cα8(3n1), (4.23)
    ˉτ(Nn2θ)=c+3α8n2(n21)+cα8(3n2cos2θ). (4.24)

    Now, in a similar way as in Case i, using (4.22)–(4.24), we have

    Ric(χ)12n2H2n2Δψψ12mr=n+1((hrn1+1n1+1+hrnn)2hrnn)2mr=n+11p<qn1(hrpq)2+mr=n+1[1i<jn1(hrij)21i<jn1hriihrjj]+mr=n+1n1+1s<tn(hrst)2+mr=n+1[n1+1s<tn(hrij)2n1+1s<tnhrsshrtt]+mr=n+11p<qn1hrpphrqq+c+3α4(n+n1n21)+3(cα)8cos2θ. (4.25)

    Using similar steps of Case i, the above inequality takes the form

    Ric(χ)12n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)8cos2θ12mr=n+1{(hrn1+1n1+1++hrnn)2hrnn}2mr=n+1[n1p=1nq=n1+1(hrpq)2+n1b=n1+1hrnnhrbbn1p=1n1q=n1+1hrpphrqq]. (4.26)

    By the Lemma 3.1, one can observe that

    mr=n+1n1p=1n1q=n1+1hrpphrqq=0. (4.27)

    Utilizing this in (4.26), we get

    Ric(χ)12n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)8cos2θ12mr=n+1((hrn1+1n1+1+hrnn)2hrnn)2mr=n+1n1p=1nq=n1+1(hrpq)2mr=n+1n1b=n1+1hrnnhrbb. (4.28)

    The last term of the above inequality can be written as

    mr=n+1n1b=n1+1hrnnhrbb=mr=n+1nb=n1+1hrnnhrbb+mr=n+1(hrnn)2

    Moreover, the fifth term of (4.28) can be expanded as

    12mr=n+1((hrn1+1n1+1++hrnn)2hrnn)2=12mr=n+1(hrn1+1n1+1++hrnn)22mr=n+1(hrnn)2+mr=n+1nj=n1+1hrnnhrjj. (4.29)

    Using last two values in (4.28), we have

    Ric(χ)12n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)8cos2θ12mr=n+1(hrn1+1n1+1+hrnn)22mr=n+1(hrnn)2+2mr=n+1nj=n1+1hrnnhrjjmr=n+1n1p=1nq=n1+1(hrpq)2mr=n+1nb=n1+1hrnnhrbb+mr=n+1(hrnn)2, (4.30)

    or equivalently

    Ric(χ)12n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)8cos2θ12mr=n+1(hrn1+1n1+1+hrnn)2mr=n+1(hrnn)2+mr=n+1nj=n1+1hrnnhrjjmr=n+1n1p=1nq=n1+1(hrpq)2. (4.31)

    On applying similar techniques as in the proof of Case i, we arrive

    Ric(χ)14n2H2n2Δψψ+c+3α4(n+n1n21)+3(cα)8cos2θ14mr=n+1(2hrnn(hrn1+1n1+1++hrnn))2, (4.32)

    which gives the inequality (4.2).

    Next, we explore the equality cases of the inequality (4.1). First, we redefine the notion of the relative null space Nx of the submanifold Mn in the generalized complex space form ˉMm(c,α) at any point xMn, the relative null space was defined by B. Y. Chen [10], as follows

    Nx={XTxMn:h(X,Y)=0,  YTxMn}.

    For A{1,,n} a unit vector field eA tangent to Mn at x satisfies the equality sign of (4.1) identically if and only if

    (i)n1p=1nq=n1+1hrpq=0(ii)nb=1nA=1bAhrbA=0(iii)2hrAA=nq=n1+1hrqq, (4.33)

    such that r{n+1,m} the condition (i) implies that Mn is mixed totally geodesic semi-slant warped product submanifold. Combining statements (ii) and (iii) with the fact that Mn is semi-slant warped product submanifold, we get that the unit vector field χ=eA belongs to the relative null space Nx. The converse is trivial, this proves statement (2).

    For a semi-slant warped product submanifold, the equality sign of (4.1) holds identically for all unit tangent vector belong to Nn1T at x if and only if

    (i)n1p=1nq=n1+1hrpq=0(ii)nb=1n1A=1bAhrbA=0(iii)2hrpp=nq=n1+1hrqq, (4.34)

    where p{1,,n1} and r{n+1,,m}. Since Mn is semi-slant warped product submanifold, the third condition implies that hrpp=0,p{1,,n1}. Using this in the condition (ii), we conclude that Mn is DT-totally geodesic semi-slant warped product submanifold in ˉMm(c,α) and mixed totally geodesicness follows from the condition (i), which proves (a) in the statement (3).

    For a semi-slant warped product submanifold, the equality sign of (4.2) holds identically for all unit tangent vector fields tangent to Nn2θ at x if and only if

    (i)n1p=1nq=n1+1hrpq=0(ii)nb=1nA=n1+1bAhrbA=0(iii)2hrKK=nq=n1+1hrqq, (4.35)

    such that K{n1+1,,n} and r{n+1,,m}. From the condition (iii) two cases emerge, that is

    hrKK=0,K{n1+1,,n}andr{n+1,,m}ordimNn2θ=2. (4.36)

    If the first case of (4.35) satisfies, then by virtue of condition (ii), it is easy to conclude that Mn is a Dθ- totally geodesic semi-slant warped product submanifold in ˉMm(c,α). This is the first case of part (b) of statement (3).

    For the other case, assume that Mn is not Dθ-totally geodesic semi-slant warped product submanifold and dim Nn2θ=2. Then condition (ii) of (4.35) implies that Mn is Dθ-totally umbilical semi-slant warped product submanifold in ˉM(c,α), which is second case of this part. This verifies part (b) of (3).

    To prove (c) using parts (a) and (b) of (3), we combine (4.34) and (4.35). For the first case of this part, assume that dimNn2θ2. Since from parts (a) and (b) of statement (3) we conclude that Mn is DT-totally geodesic and Dθ-totally geodesic submanifold in ˉMm(c,α). Hence Mn is a totally geodesic submanifold in ˉMm(c,α).

    For another case, suppose that first case does not satisfy. Then parts (a) and (b) provide that Mn is mixed totally geodesic and DT-totally geodesic submanifold of ˉMm(c,α) with dimNn2θ=2. From the condition (b) it follows that Mn is Dθ-totally umbilical semi-slant warped product submanifold and from (a) it is DT-totally geodesic, which is part (c). This proves the theorem.

    In view of (2.22), we have another version of the Theorem 4.1 as follows:

    Theorem 4.2. Let M=Nn1T×ψNn2θ be a semi-slant warped product submanifold isometrically immersed in a generalized complex space form ˉM(c,α) admitting nearly Kaehler structure. Then for each orthogonal unit vector field χTxM, either tangent to NT or Nθ, then the Ricci curvature satisfy the following inequalities:

    (i) If χ is tangent to NT, then

    Ric(χ)14n2H2n2Δlnψ+n2lnψ2+c+3α4(n+n1n21)+3(cα)8. (4.37)

    (ii) If χ is tangent to Nθ, then

    Ric(χ)14n2H2n2Δlnψ+n2lnψ2+c+3α4(n+n1n21)+3(cα)8cos2θ. (4.38)

    The equality cases are similar as Theorem 4.1.

    Since, CR-warped product submanifolds are semi-slant submanifolds with the slant angle θ=π2. Therefore, as an example of CR-warped product submanifold, we compile some result of [18] as follows.

    Example 4.1. Let {e0,ei(1i7)} be the canonical basis of Cayley division algebra on R8 over R, and R7 is the subspace of R8 generated by the purely imaginary Cayley numbers ei(1i7). Then

    S6={x1e1+x2e2++x7e7:x21+x22++x27=1}

    is an unit sphere admitting nearly Kaehler structure (ˉ,J,g). Now suppose that S2={x=(x2,x4,x6)R3:x22+x24+x26=1} is an unit sphere. For a real triple P={p1,p2,p3):p1+p2+p3=0 and p1p2p30}, let FP:S2×RS5S6 be a mapping, which is define as follows

    FP(x1,x2,x3,t)=x1(cos(tp1)e1+sin(tp1)e5)+x2(cos(tp2)e2+sin(tp2)e6)+x3(cos(tp3)e3+sin(tp3)e7), (4.39)

    where x21+x22+x23=1 and tR. Then it is clear that FP is an isometric immersion of CR-warped product submanifold S2×fR in to S6. Moreover, induced warped product metric ˉg on S2×fR is given by

    ˉg=π1g0+(3i=1(xipi)2)π2dt2,

    where π1:S2×fRS2 and π2:S2×RR are the natural projections and g0 is the Riemannian metric on S2 and the warping function is given by f=3i=1(xipi)2.

    In this section, we investigate some applications of our attained inequalities, this section is divided in different subsections as follows:

    In this subsection, we shall consider that the submanifold Mn is a compact such that M=ϕ. In the next theorem, we will see the application of Hopf's lemma for semi-slant warped product submanifold.

    Theorem 5.1. Let Mn=Nn1T×ψNn2θ be a semi-slant warped product submanifold isometrically immersed in a generalized complex space form ˉM(c,α) admitting nearly Kaehler structure. If the unit tangent vector χ is tangent to either NT or Nθ, then Mn is simply Riemannian product submanifold if the Ricci curvature satisfy one of the following inequalities:

    (i) the unit vector field χ is tangent to NT and

    Ric(χ)14n2H2+c+3α4(n+n1n21)+3(cα)8. (5.1)

    (ii) the unit vector field χ is tangent to Nθ and

    Ric(χ)14n2H2+c+3α4(n+n1n21)+3(cα)8cos2θ. (5.2)

    Proof. Suppose that inequality (5.1) holds. Then from (4.1), we get Δψψ0, which implies Δψ0, on using Hopf's Lemma, we observe that the warping function is constant and the submanifold Mn is Riemannian product. Similar result can be proved by using inequality (5.2).

    The lower bound of Ricci curvature contains numerous geometric properties. Suppose the submanifold Mn is complete non-compact and x be a any arbitrary point on Mn. For the Riemannian manifold Mn, λ1(Mn) denotes the first eigenvalue of the following Dirichlet boundary value problem.

    Δϕ=λϕinMnandϕ=0onMn, (5.3)

    where Δ denotes the Laplacian on Mn and defined as Δϕ=div(ϕ). By the principle of monotonicity one has r<t which indicates that λ1(Mnr)>λ1(Mnt) and limrλ1(Dr) exists and first eigenvalue is defined as

    λ1(M)=limrλ1(Dr).

    Several geometers have been worked on the analysis of first eigenvalue of the Laplacian operator (see [15,17,31]). For a non-constant warping function the maximum (minimum) principle on the eigenvalue λ1, we have ([6,10])

    λ1Mnϕ2dvMnϕ2dV. (5.4)

    The equality holds if and only if Δϕ=λ1ϕ.

    The relation between Ricci curvature and first eigenvalue is derived in the following theorem:

    Theorem 5.2. Let Mn=Nn1T×ψNn2θ be a semi-slant warped product submanifold isometrically immersed in a generalized complex space form ˉM(c,α) admitting nearly Kaehler structure. Suppose that the warping function lnψ is an eigenfunction of the Laplacian of Mn associated to the first eigenvalue λ1(Mn) of the problem (5.3), then the following inequalities hold:

    (i) If the unit vector field χ is tangent to NT then

    MnRic(χ)dV14n2MnH2dV+n2λ1Mn(lnψ)2dV++[c+3α4(n+n1n21)+3(cα)8]Vol(Mn). (5.5)

    (ii) If the unit vector field χ is tangent to Nθ then

    MnRic(χ)dV14n2MnH2dV+n2λ1Mn(lnψ)2dV++[c+3α4(n+n1n21)+3(cα)8cos2θ]Vol(Mn). (5.6)

    The equality cases are same as in Theorem 4.1.

    Proof. Since Mn is compact that mean it has lower and upper bounds. Let λ1=λ1(M) and lnψ be a solution of Dirichlet boundary problem corresponding to the first eigenvalue λ1(Mn). Suppose χTNT, then the inequality (4.37) can be written as follows

    Ric(χ)n2lnψ214n2H2n2Δlnψ+c+3α4(n+n1n21)+3(cα)8. (5.7)

    Integrating above inequality with respect to volume element dV, we find

    MnRic(χ)dVn2Mnlnψ2dvn24MnH2dV+3(cα)8Vol(Mn)+c+3α4(n+n1n21)Vol(Mn). (5.8)

    Since λ1 is an eigenvalue of the eigenfunction lnψ, such that Δlnψ=λ1lnψ, then equality in (5.4) holds for ϕ=lnψ,

    Mnlnψ2dV=λ1Mn(lnψ)2dV, (5.9)

    using in (5.8), we obtain

    MnRic(χ)dVn2λ1Mn(lnψ)2dVn24MnH2dV+3(cα)8)Vol(Mn)+c+3α4(n+n1n21)Vol(Mn), (5.10)

    which proves the part (i). Similarly, one can proves the part (ii).

    Let Mn be a compact Riemannian manifold and ϕ be a positive differentiable function on Mn. Then formula for Dirichlet energy of a function ϕ is given by [8]

    E(ϕ)=12Mnϕ2dV, (5.11)

    where dV is the volume element of Mn and formula for Lagrangian of the function ϕ on Mn is given in [8]

    Lϕ=12ϕ2. (5.12)

    The Euler-Lagrange equation for Lϕ is given by

    Δϕ=0. (5.13)

    Considering that the semi-slant warped product submanifold Mn=Nn1T×ψNn1θ is a compact orientable without boundary such that Mn=ϕ. Then in the following theorem we have a relation between Dirichlet energy, Ricci curvature and mean curvature vector.

    Theorem 5.3. Let Mn=Nn1T×ψNn2θ be a semi-slant warped product submanifold of a generalized complex space form admitting nearly Kaehler manifold. Then we have the following inequalities for the Dirichlet energy of the warping function lnψ:

    (i) If the unit vector field χ is tangent to NT then

    E(lnψ)12n2MnRic(χ)dVn28n2MnH2dV[(c+3α)8.(n+n1n21)n2+3(cα)16n2]Vol(Mn). (5.14)

    (ii) If the unit vector field χ is tangent to Nθ then

    E(lnψ)12n2MnRic(χ)dVn28n2MnH2dV[(c+3α)8.(n+n1n21)n2+3(cα)16n2cos2θ)]Vol(Mn). (5.15)

    The equality cases are similar as in Theorem 4.1.

    Proof. For a positive valued differentiable function ϕ defined on a compact orientable Riemannian manifold without boundary, by theory of integration on Riemannian manifold we have MnΔϕdV=0. On applying this fact for the warping function lnψ, we have

    MnΔlnψdV=0. (5.16)

    Integrating inequality (4.1) with respect to volume element dV on semi-slant warped product submanifold Mn, which is compact and orientable without boundary, we get

    MnRic(χ)dVn24MnH2dV+n2Mnlnψ2dVn2MnΔlnψdV+[c+3α4(n+n1n21)+3(cα)8]Vol(Mn). (5.17)

    Using the formula (5.11) and after some computation, the required inequality is derived. In a similar method, we can prove the inequality (5.15)

    Further, in the following theorem we will compute the Lagrangian for the warping function lnψ.

    Theorem 5.4. Let Mn=Nn1T×ψNn2θ be a compact orientable semi-slant warped submanifold isometrically immersed in a generalized complex space form admitting nearly Kaehler manifold such that the warping function lnψ satisfies the Euler-Lagrangian equation, then

    (i) If the unit vector field χ is tangent to NT, then

    Llnψ12n2Ric(χ)n28n2H2(c+3α)8.(n+n1n21)n23(cα)16n2. (5.18)

    (ii) If the unit vector field χ is tangent to Nθ, then

    Llnψ12n2Ric(χ)n28n2H2(c+3α)8.(n+n1n21)n23(cα)16n2cos2θ, (5.19)

    where Llnψ is the Lagrangian of the warping function defined in (5.12). The equality cases are same as Theorem 4.1.

    Proof. The proof follows immediately on using (5.12) and (5.13) in Theorem 4.1.

    Further, the Hamiltonian for a local orthonormal frame at any point xMn is expressed as follows [8]

    H(p,x)=12ni=1p(ei)2. (5.20)

    On replacing p by a differential operator dϕ, then from (2.20), we get

    H(dϕ,x)=12ni=1dϕ(ei)2=12ni=1ei(ϕ)2=12ϕ2. (5.21)

    In the next result we obtain a relation between Hamiltonian of warping function, Ricci curvature and squared norm of mean curvature vector.

    Theorem 5.5. Let Mn=Nn1T×ψNn2θ be a semi-slant warped product submanifold isometrically immersed in a generalized complex space form ˉM(c,α) admitting nearly Kaehler structure then the Hamiltonian of the warping function satisfy the following inequalities

    (i) If χTNT, then

    H(dlnψ,x)12n2{Ric(χ)+n2Δlnψn24H2c+3α4(n+n1n21)3(cα)8} (5.22)

    (ii) If χTNθ, then

    H(dlnψ,x)12n2{Ric(χ)+n2Δlnψn24H2c+3α4(n+n1n21)3(cα)8cos2θ} (5.23)

    Proof. By the application of (5.21) in theorem 4.1, we get the required results.

    This subsection is based on the study of Obata [29]. Basically, Obata characterized a Riemannian manifolds by a specific ordinary differential equation and derived that an n-dimensional complete and connected Riemannian manifold (Mn,g) to be isometric to the n-sphere Sn if and only if there exists a non-constant smooth function ϕ on Mn that is the solution of the differential equation Hϕ=cϕg, where Hϕ is the Hessian of ϕ. Inspired by the work of Obata [29], we obtain the following characterization

    Theorem 5.6. Suppose Mn=Nn1T×ψNn2θ is a compact orientable warped product submanifold isometrically immersed in a generalized complex space form Mm(c,α) admitting nearly Kaehler structure with positive Ricci curvature and satisfying one of the following relation

    (i) χTNT and

    Hessϕ2=3λ1n24n1n2H23λ14n1n2[(c+3α)(n+n1n21)+3(cα)2], (5.24)

    (ii) χTNθ and

    Hessϕ2=3λ1n24n1n2H23λ14n1n2[(c+3α)(n+n1n21)3(cα)2cos2θ)], (5.25)

    where λ1>0 is an eigenvalue of the warping function ϕ=lnψ. Then the base manifold Nn1T is isometric to the sphere Sn1(λ1n1) with constant sectional curvature λ1n1.

    Proof. Let χTNT. Consider that ϕ=lnψ and define the following relation as

    HessϕtϕI2=Hessϕ2+t2ϕ2I22tϕg(Hessϕ,I). (5.26)

    But we know that I2=trace(II)=p and

    g(Hess(ϕ),I)=trace(Hessϕ,I)=traceHess(ϕ).

    Then Eq (5.26) transform to

    HessϕtϕI2=Hessϕ2+pt2ϕ22tϕΔϕ. (5.27)

    Assuming λ1 is an eigenvalue of the eigenfunction ϕ then Δϕ=λ1ϕ. Thus we get

    HessϕtϕI2=Hessϕ2+(pt22tλ)ϕ2. (5.28)

    On the other hand, we obtain Δϕ2=2ϕΔϕ+ϕ2 or λ1ϕ2=2λ1ϕ2+ϕ2 which implies that ϕ2=1λ1ϕ2, using this in Eq (5.28), we have

    HessϕtϕI2=Hessϕ2+(2tpt2λ1)ϕ2. (5.29)

    In particular t=λ1n1 on (5.29) and integrating with respect to dV, we get

    MnHessϕ+λ1n1ϕI2dV=MnHessϕ2dV3λ1n1Mnϕ2dV. (5.30)

    Integrating the inequality (4.37) and using the fact MnΔϕdV=0, we have

    MnRic(χ)dVn24MnH2dV+n2Mnϕ2dV++c+3α4(n+n1n21)Vol(Mn)+3(cα)8Vol(Mn). (5.31)

    From (5.30) and (5.31) we derive

    1n2MnRic(χ)dVn24n2MnH2dVn13λ1MnHessϕ+λ1n1ϕI2dV+n13λ1MnHessϕ2dV+c+3α4(n+n1n21)n2Vol(Mn)+3(cα)8n2Vol(Mn). (5.32)

    According to assumption Ric(χ)0, the above inequality gives

    MnHessϕ+λ1n1ϕI2dV3n2λ14n1n2MnH2dV+MnHessϕ2dV+c+3α43λ1(n+n1n21)n1n2Vol(Mn)+9λ1(cα)8n1n2Vol(Mn). (5.33)

    From (5.24), we get

    MnHessϕ+λ1n1ϕI2dV0, (5.34)

    but we know that

    MnHessϕ+λ1n1ϕI2dV0. (5.35)

    Combining last two statements, we get

    MnHessϕ+λ1n1ϕI2dV=0Hessϕ=λ1n1ϕI. (5.36)

    Since the warping function ϕ=lnψ is not constant function on Mn so Eq (5.36) is Obata's [29] differential equation with constant c=λ1n1>0. As λ1>0 and therefore the base submanifold Nn1T is isometric to the sphere Sn1(λ1n1) with constant sectional curvature λ1n1. Similarly, we can prove the theorem by using part (ii).

    In [17] Rio et al. studied another version of Obata's differential equation in the characterization of Euclidean sphere. Basically, they proved that if ϕ is a real valued non constant function on a Riemannian manifold satisfying Δϕ+λ1ϕ=0 such that λ<0, then Mn is isometric to a warped product of the Euclidean line and a complete Riemannian manifold whose warping function ϕ is the solution of the following differential equation

    d2ϕdt2+λ1ϕ=0. (5.37)

    Motivated by the study of Rio et al. [17] and Ali et al. [1] we obtain the following characterization.

    Theorem 5.7. Suppose Mn=Nn1T×ψNn2θ is a compact orientable semi-slant warped product submanifold isometrically immersed in generalized complex space form ˉM(c,α) admitting nearly Kaehler structure with positive Ricci curvature and satisfying one of the following statement:

    (i) χTNT and

    Hessϕ2=3λ1n24n1n2H23λ14n1n2[(c+3α)(n+n1n21)+3(cα)2] (5.38)

    (ii) χTNθ and

    Hessϕ2=3λ1n24n1n2H23λ14n1n2[(c+3α)(n+n1n21)+3(cα)2cos2θ)], (5.39)

    where λ1<0 is a negative eigenvalue of the eigenfunction ϕ=lnψ. Then Nn1T is isometric to a warped product of the Euclidean line and a complete Riemannian manifold whose warping function ϕ=lnψ satisfies the differential equation

    d2ϕdt2+λ1ϕ=0. (5.40)

    Proof. Since we assumed that the Ricci curvature is positive then by the Myers's theorem according to which, a complete Riemannian manifold with positive Ricci curvature is compact that mean Mn is compact semi-slant warped product submanifold with free boundary [28]. Then by (5.32) we get

    1n2MnRic(χ)dVn24n2MnH2dVn13λ1MnHessϕ+λ1n1ϕI2dV+n13λ1MnHessϕ2dV+c+3α4(n+n1n21)n2Vol(Mn)+3(cα)8n2Vol(Mn). (5.41)

    According to hypothesis, Ricci curvature is positive Ric(χ)>0, then we have

    MnHessϕ+λ1n1ϕI2dV<3n2λ14n1n2MnH2dV+MnHessϕ2dV+c+3α4.3λ1(n+n1n21)n1n2Vol(Mn)+9λ1(cα)8n2Vol(Mn). (5.42)

    If Eq (5.38) holds, then from above inequality we get Hessϕ+λ1n1ψI2<0, which is not possible hence Hessϕ+λ1n1ϕI2=0. Since λ<0, then by result of [17], the submanifold Nn1T is isometric to a warped product of the Euclidean line and a complete Riemannian manifold, where the warping function on R is the solution of the differential equation (5.40). This proves the theorem. Similarly by assuming (5.39), we can also prove the theorem.

    In this paper firstly we have obtained a Ricci curvature inequality of a semi-slant warped product submanifold isometrically immersed in a generalized complex space form admitting a nearly Kaehler structure. Then we have given some applications on Hopf's Lemma, Dirichlet energy, Euler-Lagrangian equation, Hamiltonian of warping functions of a semi-slant warped product submanifold isometrically immersed in a generalized complex space form admitting a nearly Kaehler structure. In last we have characterized semi-slant warped product submanifold isometrically immersed in a generalized complex space form admitting a nearly Kaehler structure in the basis of Obata differential equation.

    The first author extends his appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through research groups program under grant number R.G.P.1/206/42.

    The authors declare no conflict of interest in this paper.



    [1] A. Ali, L. I. Piscoran, Ali H. Al-Khalidi, Ricci curvature on warped product submanifolds in spheres with geometric applications, J. Geom. Phys., 146 (2019), 1–17. http://dx.doi.org/10.1016/j.geomphys.2019.103510 doi: 10.1016/j.geomphys.2019.103510
    [2] F. R. Al-Solamy, V. A. Khan, S. Uddin, Geometry of warped product semi-slant submanifolds of Nearly Kaehler manifolds, Results Math., 71 (2017), 783–799. http://dx.doi.org/10.1007/s00025-016-0581-4. doi: 10.1007/s00025-016-0581-4
    [3] K. Arslan, R. Ezentas, I. Mihai, C. ¨Ozgur, Certain inequalities for submanifolds in (k,μ)-contact space form, Bull. Aust. Math. Soc., 64 (2001), 201–212, http://dx.doi.org/10.1017/S0004972700039873. doi: 10.1017/S0004972700039873
    [4] M. Aquib, J. W. Lee, G. E. Vilcu, W. Yoon, Classification of Casorati ideal Lagrangian submanifolds in complex space forms, Differ. Geom. Appl., 63 (2019), 30–49. http://dx.doi.org/10.1016/j.difgeo.2018.12.006 doi: 10.1016/j.difgeo.2018.12.006
    [5] J. K. Beem, P. Ehrlich, T. G. Powell, Warped product manifolds in relativity, selected studies, North-Holland, Amsterdam-New York, 1982.
    [6] M. Berger, Les Varietes riemanniennes (14)-pinces, Ann. Sc. Norm. Super. Pisa CI. Sci., 14 (1960), 161–170.
    [7] R. L. Bishop, B. O'Neil, Manifolds of negative curvature, Trans. Amer. Math. Soc., 145 (1969), 1–49. http://dx.doi.org/10.1090/S0002-9947-1969-0251664-4 doi: 10.1090/S0002-9947-1969-0251664-4
    [8] O. Calin, D. C. Chang, Geometric mechanics on riemannian manifolds: Applications to partial differential equations, Springer Science & Business Media, 2006.
    [9] B. Y. Chen, CR-submanifolds of a Kaehler manifold I, J. Differ. Geom., 16 (1981), 305–323. http://dx.doi.org/ 10.4310/jdg/1214436106 doi: 10.4310/jdg/1214436106
    [10] B.Y. Chen, Relations between Ricci curvature and shape operator for submanifolds with arbitrary codimension, Glasgow Math. J., 41 (1999), 33–41. http://dx.doi.org/10.1017/S0017089599970271 doi: 10.1017/S0017089599970271
    [11] B. Y. Chen, Geometry of warped product CR-submanifolds in Kaehler manifolds I, Monatsh. Math., 133 (2001), 177–195. http://dx.doi.org/10.1007/s006050170019 doi: 10.1007/s006050170019
    [12] B. Y. Chen, Pseudo-Riemannian geometry, δinvariants and applications, World Scientific Publishing Company, Singapore, 2011.
    [13] B. Y. Chen, Geometry of warped product submanifolds: A survey, J. Adv. Math. Stud., 6 (2013), 1–43.
    [14] B. Y. Chen, F. Dillen, L. Verstraelen, L. Vrancken, Characterization of Riemannian space forms, Einstein spaces and conformally flate spaces, Proc. Amer. Math. Soc., 128 (1999), 589–598.
    [15] S. S. Cheng, Spectrum of the Laplacian and its applications to differential geometry, Univ. of California, Berkeley, 1974.
    [16] D. Cioroboiu, B. Y. Chen, Inequalities for semi-slant submanifolds in Sasakian space forms, Int. J. Math., 27 (2003), 1731–1738.
    [17] E. Garcia-Rio, D. N. Kupeli, B. Unal, On a differential equation characterizing Euclidean sphere, J. Differ. Eq., 194 (2003), 287–299.
    [18] H. Hashimoto, K. Mashimo, On some 3-dimensional CR-submanifolds in S6, Nagoya Math. J., 156 (1999), 171–185.
    [19] S. W. Hawkings, G. F. R. Ellis, The large scale structure of space-time, Cambridge Univ. Press, Cambridge, 1973.
    [20] S. K. Hui, T. Pal, J. Roy, Another class of warped product skew CR-submanifolds of Kenmotsu manifolds, Filomat, 33 (2019), 2583–2600.
    [21] S. K. Hui, M. H. Shahid, T. Pal, J. Roy, On two different classes of warped product submanifolds of Kenmotsu manifolds, Kragujevac J. Math., 47 (2023), 965–986.
    [22] S. K. Hui, M. S. Stankovic, J. Roy, T. Pal, A class of warped product submanifolds of Kenmotsu manifolds, Turk. J. Math., 44 (2020), 760–777.
    [23] V. A. Khan, M. A. Khan, Semi-slant submanifolds of a nearly Kaehler manifold, Turk. J. math., 31 (2007), 341–353.
    [24] V. A. Khan, K. A. Khan, Generic warped product submanifolds of nearly Kaehler manifolds, Beitr. Algebra Geom., 50 (2009), 337–352.
    [25] V. A. Khan, K. A. Khan, Semi-slant warped product submanifolds of a nearly Kaehler manifold, Differ. Geom.-Dyn. Sys., 16 (2014), 168–182.
    [26] A. Mihai, Warped product submanifolds in generalized complex space forms, Acta Math. Acad. Paedagog. Nyhazi., 21 (2005), 79–87.
    [27] A. Mihai, C. ¨Ozgur, Chen inequalities for submanifolds of real space forms with a semi-symmetric metric connection, Taiwan. J. Math., 14 (2010), 1465–1477. https://dx.doi.org/10.11650/twjm/1500405961 doi: 10.11650/twjm/1500405961
    [28] S. B. Myers, Riemannian manifolds with positive mean curvature, Duke Math. J., 8 (1941), 401–404.
    [29] M. Obata, Certain conditions for a Riemannian manifold to be isometric with a sphere, J. Math. Soc. Japan, 14 (1962), 333–340.
    [30] B. O'Neill, Semi-Riemannian geometry with application to relativity, Academic Press, 1983.
    [31] B. Palmer, The Gauss map of a spacelike constant mean curvature hypersurface of Minkowski space, Comment. Math. Helv., 65 (1990), 52–57.
    [32] N. Papaghiuc, Semi-slant submanifolds of Kaehler manifold, An. Stiint. U. Al. I-Mat., 40 (1994), 55–61.
    [33] B. Sahin, Non-existence of warped product semi-slant submanifolds of Kaehler manifold, Geometriae Dedicata, 117 (2006), 195–202, https://dx.doi.org/10.1007/s10711-005-9023-2. doi: 10.1007/s10711-005-9023-2
    [34] D. W. Yoon, Inequality for Ricci curvature of slant submanifolds in cosymplectic space forms, Turk. J. Math., 30 (2006), 43–56.
  • This article has been cited by:

    1. Bang-Yen Chen, Adara M. Blaga, 2024, Chapter 1, 978-981-99-9749-7, 1, 10.1007/978-981-99-9750-3_1
    2. Yanlin Li, Meraj Ali Khan, MD Aquib, Ibrahim Al-Dayel, Maged Zakaria Youssef, Chen–Ricci Inequality for Isotropic Submanifolds in Locally Metallic Product Space Forms, 2024, 13, 2075-1680, 183, 10.3390/axioms13030183
    3. Md Aquib, Meraj Khan, Ibrahim Al-Dayel, Ricci tensor of slant submanifolds in locally metallic product space forms, 2024, 38, 0354-5180, 3523, 10.2298/FIL2410523A
  • Reader Comments
  • © 2022 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(2228) PDF downloads(77) Cited by(3)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog