U-Hang Ki" /> Soo Jin Kim" /> Hiroyuki Kurihara

" /> U-Hang Ki, Soo Jin Kim, Hiroyuki Kurihara

" /> U-Hang Ki, Soo Jin Kim, Hiroyuki Kurihara

. Kyungpook Mathematical Journal 2016;56:541-75. https://doi.org/10.5666/KMJ.2016.56.2.541">
검색
Article Search

JMB Journal of Microbiolog and Biotechnology

OPEN ACCESS eISSN 0454-8124
pISSN 1225-6951
QR Code

Article

Kyungpook Mathematical Journal 2016; 56(2): 541-575

Published online June 1, 2016

Copyright © Kyungpook Mathematical Journal.

Jacobi Operators with Respect to the Reeb Vector Fields on Real Hypersurfaces in a Nonat Complex Space Form

U-Hang Ki1, Soo Jin Kim2, Hiroyuki Kurihara3

The National Academy of Siences, Seoul 06579, Korea1
Department of Mathematics Chosun University, Gwangju 61452, Korea2
The College of Education, Ibaraki University, Mito 310-8512, Japan3

Received: August 18, 2015; Accepted: January 19, 2016

Let M be a real hypersurface of a complex space form with almost contact metric structure (φ, ξ, η, g). In this paper, we prove that if the structure Jacobi operator Rξ = R(·, ξ)ξ is φξξ-parallel and Rξ commute with the structure tensor φ, then M is a homogeneous real hypersurface of Type A provided that TrRξ is constant.

Keywords: complex space form, real hypersurface, structure Jacobi operator, structure tensor, Ricci tensor.

A complex n-dimensional Kähler manifold of constant holomorphic sectional curvature 4c ≠= 0 is called a complex space form, which is denoted by Mn(c). So naturally there exists a Kähler structure J and Kähler metric on Mn(c). As is well known, complete and simply connected complex space forms are isometric to a complex projective space Pn(ℂ), or complex hyperbolic space Hn(ℂ) as c > 0 or c < 0. Now let us consider a real hypersurface M in Mn(c). Then we also denote by g the induced Riemannian metric of Mand by N a local unit normal vector field of M in Mn(c). Further, A denotes by the shape operator of M in Mn(c). Then, an almost contact metric structure (φ, ξ, η, g) of M is naturally induced from the Kähler structure of Mn(c) as follows:

φX=(JX)T,ξ=-JN,η(X)=g(X,ξ),XTM,

where TM denotes the tangent bundle of M and ( )T the tangential component of a vector. The Reeb vector ξ is said to be principal if = αξ, where α = η(). A real hypersurface is said to a Hopf hypersurface if the Reeb vector ξ of M is principal. Hopf hypersurfaces is realized as tubes over certain submanifolds in Pnℂ, by using its focal map (see Cecil and Ryan [2]). By making use of those results and the mentioned work of Takagi ([17], [18]), Kimura [11] proved the local classification theorem for Hopf hypersurfaces of Pnℂ whose all principal curvatures are constant. For the case Hnℂ, Berndt [1] proved the classification theorem for Hopf hypersurfaces whose all principal curvatures are constant. Among the several types of real hypersurfaces appeared in Takagi’s list or Berndt’s list, a particular type of tubes over totally geodesic Pkℂ or Hkℂ (0 ≤ kn–1) adding a horosphere in Hnℂ, which is called type A, has a lot of nice geometric properties. For example, Okumura [13](resp. Montiel and Romero [12]) showed that a real hypersurface in Pnℂ (resp. Hnℂ) is locally congruent to one of real hypersurfaces of type A if and only if the Reeb flow ξ is isometric or equivalently the structure operator φ commutes with the shape operator A.

The Reeb vector field ξ plays an important role in the theory of real hypersurfaces in a complex space form Mn(c). Related to the Reeb vector field ξ the Jacobi operator Rξ defined by Rξ = R(·, ξ)ξ for the curvature tensor R on a real hypersurface M in Mn(c) is said to be a structure Jacobi operator on M. The structure Jacobi operator has a fundamental role in contact geometry. In [3], Cho and first author started the study on real hypersurfaces in complex space form by using the operator Rξ. In particular the structure Jacobi operator has been studied under the various commutative conditions ([4], [5], [7], [16]). For example, Pérez et al. [16] called that real hypersurfaces M has commuting structure Jacobi operator if RξRX = RXRξ for any vector field X on M, and proved that there exist no real hypersurfaces in Mn(c) with commuting structure Jacobi operator. On the other hand Ortega et al. [14] have proved that there are no real hypersurfaces in Mn(c) with parallel structure Jacobi operator Rξ, that is, ∇XRξ = 0 for any vector field X on M. More generally, such a result has been extended by [15]. In this situation, if naturally leads us to be consider another condition weaker than parallelness. In the preceding work, we investigate the weaker condition ξ-parallelness, that is, ∇ξRξ = 0 (cf. [4], [7], [8]). Moreover some works have studied several conditions on the structure Jacobi operator Rξ ([3], [5], [7] and [8]). The following facts are used in this paper without proof.

Theorem 1.1. (Ki, Kim and Lim [5])

Let M be a real hypersurface in a nonat complex space form Mn(c), c ≠= 0 which satisfies Rξ(φA) = 0. Then M is a Hopf hypersurface in Mn(c). Further, M is locally congruent to one of the following hypersurfaces:

  • (I) In cases that Mn(c) = Pnwith η() ≠= 0,

    • (A1) a geodesic hypersphere of radius r, where 0 < r < π/2 and r ≠= π/4;

    • (A2) a tube of radius r over a totally geodesic Pkfor some k ∈ {1, . . . , n–2}, where 0 < r < π/2 and r ≠= π/4.

  • (II) In cases Mn(c) = Hnℂ,

    • (A0) a horosphere;

    • (A1) a geodesic hypersphere or a tube over a complex hyperbolic hyperplane Hn–1;

    • (A2) a tube over a totally geodesic Hkfor some k ∈ {1, . . . , n – 2}.

Theorem 1.2. (Ki, Nagai and Takagi [9])

Let M be a real hypersurface in a nonat complex space form Mn(c), c ≠= 0 If M satisfies Rξφ = φRξand at the same time RξS = SRξ. Then M is the same types as those in Theorem 1.1, where S denotes the Ricci tensor of M.

In [6], the authors started the study on real hypersurfaces in a complex space form with φξξ-parallel structure Jacobi operator Rξ, that is, ∇φξξRξ = 0 for the vector φξξ orthogonal to ξ. In this paper we invetigate the structure Jacobi operator is φξξ-parallel under the condition that the structure Jacobi operator commute with the structure tensor φ. We prove that if the structure Jacobi operator Rξ is φξξ-parallel and Rξ commute with the structure tensor φ, then M is homogeneous real hypersurfaces of Type A provided that TrRξ is constant.

All manifolds in this paper are assumed to be connected and of class C and the real hypersurfaces are supposed to be oriented.

Let M be a real hypersurface immersed in a complex space form Mn(c), c ≠= 0 with almost complex structure J, and N be a unit normal vector field on M. The Riemannian connection ∇̃ in Mn(c) and ∇ in M are related by the following formulas for any vector fields X and Y on M:

˜XY=XY+g(AX,Y)N,         ˜XN=-AX

where g denotes the Riemannian metric of M induced from that of Mn(c) and A denotes the shape operator of M in direction N. For any vector field X tangent to M, we put

JX=φX+η(X)N,         JN=-ξ.

We call ξ the structure vector field (or the Reeb vector field) and its flow also denoted by the same latter ξ. The Reeb vector field ξ is said to be principal if = αξ, where α = η().

A real hypersurface M is said to be a Hopf hypersurface if the Reeb vector field ξ is principal. It is known that the aggregate (φ, ξ, η, g) is an almost contact metric structure on M, that is, we have

φ2X=-X+η(X)ξ,g(φX,φY)=g(X,Y)-η(X)η(Y),η(ξ)=1,φξ=0,η(X)=g(x,ξ)

for any vector fields X and Y on M. From Kähler condition ∇̃J = 0, and taking account of above equations, we see that

Xξ=φAX,(Xφ)Y=η(Y)AX-g(AX,Y)ξ

for any vector fields X and Y tangent to M.

Since we consider that the ambient space is of constant holomorphic sectional curvature 4c, equations of Gauss and Codazzi are respectively given by

R(X,Y)Z=c{g(Y,Z)X-g(X,Z)Y+g(φY,Z)φX-g(φX,Z)φY-2g(φX,Y)φZ}+g(AY,Z)AX-g(AX,Z)AY,(XA)Y-(YA)X=c{η(X)φY-η(Y)φX-2g(φX,Y)ξ}

for any vector fields X, Y and Z on M, where R denotes the Riemannian curvature tensor of M.

In what follows, to write our formulas in convention forms, we denote by α = η(), β = η(A2ξ) and h = TrA, and for a function f we denote by ∇f the gradient vector field of f.

From the Gauss equation (2.3), the Ricci tensor S of M is given by

SX=c{(2n+1)X-3η(X)ξ}+hAX-A2X

for any vector field X on M.

Now, we put

Aξ=αξ+μW,

where W is a unit vector field orthogonal to ξ. In the sequel, we put U = ∇ξξ, then by (2.1) we see that

U=μφW

and hence U is orthogonal to W. So we have g(U,U) = μ2. Using (2.7), it is clear that

φU=-Aξ+αξ,

which shows that g(U,U) = βα2. Thus it is seen that

μ2=β-α2.

Making use of (2.1), (2.7) and (2.8), it is verified that

μg(XW,ξ)=g(AU,X),g(Xξ,U)=μg(AW,X)

because W is orthogonal to ξ.

Now, differentiating (2.8) covariantly and taking account of (2.1) and (2.2), we find

(XA)ξ=-φXU+g(AU+α,X)ξ-AφAX+αφAX,

which together with (2.4) implies that

(ξA)ξ=2AU+α.

Applying (2.12) by φ and making use of (2.11), we obtain

φ(XA)ξ=XU+μg(AW,X)ξ-φAφAX-αAX+αg(Aξ,X)ξ,

which connected to (2.1), (2.9) and (2.13) gives

ξU=3φAU+αAξ-βξ+φα.

Using (2.3), the structure Jacobi operator Rξ is given by

Rξ(X)=R(X,ξ)ξ=c{X-η(X)ξ}+αAX-η(AX)Aξ

for any vector field X on M. Differentiating this covariantly along M, we find

g((XRξ)Y,Z)=g(X(RξY)-Rξ(XY),Z)=-c(η(Z)g(Xξ,Y)+η(Z)g(Xξ,Z))+(Xα)g(AY,Z)+αg((XA)Y,Z)-η(AZ){g((XA)ξ,Y)+g(AφAX,Y)}-η(AY){g((XA)ξ,Z)+g(AφAX,Z)}.

From (2.5) and (2.16), we have

(RξS-SRξ)(X)=-η(AX)A3ξ+η(A3X)Aξ-η(A2X)(hAξ-cξ)+(hη(AX)-cη(X))A2ξ-ch(η(AX)ξ-η(X)Aξ).

Let Ω be the open subset of M defined by

Ω={pM;Aξ-αξ0}.

At each point of Ω, the Reeb vector field ξ is not principal. That is, ξ is not an eigenvector of the shape operator A of M if Ω ≠= ∅︀.

In what follows we assume that Ω is not an empty set in order to prove our main theorem by reductio ad absurdum, unless otherwise stated, all discussion concerns the set Ω.

Let M be a real hypersurface in Mn(c), c ≠= 0. We suppose that Rξφ = φRξ. Then by using (2.16) we have

α(φAX-AφX)=g(Aξ,X)U+g(U,X)Aξ.

Then, using (3.1), it is clear that α ≠= 0 on Ω. So a function λ given by β = αλ is defined. Because of (2.9), we have

μ2=αλ-α2.

Replacing X by U in (3.1) and taking account of (2.8), we find

φAU=λAξ-A2ξ,

which implies

φA2ξ=AU+λU

because U is orthogonal to . From this and (2.6) we have

μφAW=AU+(λ-α)U,

which together with (2.7) yields

g(AW,U)=0.

Using (2.6) and (3.3), we can write (2.15) as

ξU=(3λ-2α)Aξ-3μAW-αλξ+φα.

Since α ≠= 0 on Ω, (3.1) reformed as

(φA-Aφ)X=η(X)U+u(X)ξ+τ(u(X)W+w(X)U),

where a 1-form u is defined by u(X) = g(U,X) and w by w(X) = g(W,X), where we put

ατ=μ,λ-α=μτ.

Differentiating (3.8) covariantly and taking the inner product with any vector field Z, we find

g(φ(YA)X,Z)+g(φ(YA)Z,X)=-η(AX)g(AY,Z)-g(AX,Y)η(AZ)+g(A2X,Y)η(Z)+η(X)g(A2Y,Z)+(η(X)+τω(X))g(YU,Z)+g(YU,X)(η(Z)+τω(Z))+u(X)g(Yξ,Z)+g(Yξ,X)u(Z)+(Yτ)(u(X)w(Z)+u(Z)w(X))+τ(u(X)g(YW,Z)+g(YW,X)u(Z))

because of (2.1) and (2.2). From this, taking the skew-symmetric part with respect to X and Y, and making use of the Codazzi equation (2.4), we find

c(η(X)g(Y,Z)-η(Y)g(X,Z))+g((XA)φY,Z)-g((YA)φX,Z)=-η(AX)g(AY,Z)+η(AY)g(AX,Z)+η(X)g(A2Y,Z)-η(Y)g(A2X,Z)+(η(X)+τw(X))g(YU,Z)-(η(Y)+τw(Y))g(XU,Z)+(g(YU,X)-g(XU,Y))(η(Z)+τw(Z))+u(X)g(Yξ,Z)-u(Y)g(Xξ,Z)+(g(Yξ,X)-g(Xξ,Y))u(Z)+(Yτ)(u(X)w(Z)+u(Z)w(X))-(Xτ)(u(Y)w(Z)+u(Z)w(Y))+τ{u(X)g(YW,Z)-u(Y)g(XW,Z)}+τ{(g(YW,X)-g(XW,Y))u(Z)}.

Interchanging Y and Z in (3.10), we obtain

g(φ(ZA)X,Y)+g(φ(ZA)Y,X)=-η(AX)g(AY,Z)-g(AX,Z)η(AY)+g(A2X,Z)η(Y)+η(X)g(A2Y,Z)+(η(X)+τw(X))g(ZU,Y)+g(ZU,X)(η(Y)+τw(Y))+u(X)g(Zξ,Y)+g(Zξ,X)u(Y)+(Zτ)(u(X)w(Y)+u(Y)w(X))+τ(u(X)g(ZW,Y)+g(ZW,X)u(Y)),

or, using (2.4)

g(φ(XA)Z,Y)+g(φ(YA)Z,X)+c(η(X)g(Z,Y)+η(Y)g(Z,X)-2η(Z)g(X,Y))=-η(AX)g(AY,Z)-g(AX,Z)η(AY)+g(A2X,Z)η(Y)+η(X)g(A2Y,Z)+(η(X)+τw(X))g(ZU,Y)+g(ZU,X)(η(Y)+τw(Y))+u(X)g(Zξ,Y)+g(Zξ,X)u(Y)+(Zτ)(u(X)w(Y)+u(Y)w(X))+τ(u(X)g(ZW,Y)+g(ZW,X)u(Y)).

Combining this to (3.11), we have

2g(YA)φX,Z)+2c(η(Z)g(X,Y)-η(X)g(Y,Z))+2η(X)g(A2Z,Y)-2η(AX)g(AZ,Y)+(g(ZU,X)-g(XU,Z))(η(Y)+τw(Y))+(g(YU,X)-g(XU,Y))(η(Z)+τw(Z))+(g(ZU,Y)+g(YU,Z))(η(X)+τw(X))+(g(Zξ,X)-g(Xξ,Z))u(Y)+(g(Yξ,X)-g(Xξ,Y))u(Z)+(g(Zξ,Y)+g(Yξ,Z))u(X)+(Yτ)(u(X)w(Z)+u(Z)w(X))+(Zτ)(u(X)w(Y)+u(Y)w(X))-(Xτ)(u(Y)w(Z)+u(Z)w(Y))+τ{u(X)(g(ZW,Y)+g(YW,Z))+u(Z)(g(XW,Y)-g(YW,X))+u(Y)(g(ZW,X))-g(XW,Z))}=0.

If we put X = ξ in (3.12), then we have

g(YU,Z)+g(ZU,Y)+2c(η(Z)η(Y)-g(Z,Y))+2g(A2Y,Z)-2αg(AY,Z)-du(ξ,Z)(η(Y)+τw(Y))-du(ξ,Y)(η(Z)+τw(Z))-2u(Y)u(Z)-(ξτ)(u(Y)w(Z)+u(Z)w(Y))-τ{u(Z)dw(ξ,Y)+u(Y)dw(ξ,Z)}=0,

where d denotes the operator of the exterior derivative.

We will continue our discussions under the same hypothesis Rξφ = φRξ as in Section 3. Furthermore, suppose that ∇φξξRξ = 0 and then ∇WRξ = 0 since we assume that μ ≠= 0. Replacing X by W in (2.17), we find

(Wα)g(AY,Z)-c(η(Z)g(φAW,Y)+η(Y)g(φAW,Z))+αg((WA)Y,Z)-η(AZ){g((WA)ξ,Y)+g(AφAW,Y)}-η(AY){g((WA)ξ,Z)+g(AφAW,Z)}=0

by virtue of ∇WRξ = 0. Putting Y = ξ in this and making use of (2.13) and (3.6), we obtain

αAφAW+cφAW=0

because U and W are mutually orthogonal. From this and (2.16), it is seen that RξφAW = 0 by virtue of (3.6), and hence RξAW = 0 which together with (2.16) implies that

αA2W=-cAW+cμξ+μ(α+g(AW,W))Aξ,

which tells us that

αg(A2W,W)=(μ2-c)g(AW,W)αμ2.

Since α ≠= 0, β = αλ and (3.2), we see that

g(A2W,W)=(λ-α-cα)g(AW,W)+μ2.

Combining (3.5) to (4.2), we get

αA2U=-(μ2+c)AU-c(λ-α)U.

If we apply μW to (3.3) and make use of (2.6), then we find

g(AU,U)=μ2(g(AW,W)+α-λ).

Using (4.2), we see from (4.1)

α(WA)X=-(Wα)AX+η(AX)(WA)ξ+g((WA)ξ,X)Aξ-cσμ(w(X)φAW+g(φAW,X)W)

for any vector field X, which together with (3.5) yields

α(WA)X=-(Wα)AX+η(AX)(WA)ξ+g((WA)ξ,X)Aξ-cα{w(X)AU+u(AX)W+(λ-α)(w(X)U+u(X)W)}.

Now, if we put X = W in (2.12), and make use of (3.5) and (4.2), then we find

(WA)ξ=-φWU+(Wα)ξ+1μ(α+cα){AU+(λ-α)U)}.

Also, if we take the inner product (2.12) with and take account of (2.6), (3.2) and (3.4), then we obtain

α(Xα)+μ(Xμ)=g(αξ+μW,(XA)ξ)-g(A2U+λAU,X),

which together with (2.4), (2.13) and (4.6) yields

μ(WA)ξ=-(α+cα)AU-cα(λ+α)U+μμ.

If we take the inner product (4.10) with ξ and make use of (2.13) and (3.6), then we find

Wα=ξμ.

Using (4.10), we can write (4.8) as

α(WA)X+(Wα)AX+1μη(AX){(α+cα)AU+cα(λ+α)U-μμ}+1μ{(α+cα)u(AX)+cα(λ+α)u(X)-μ(Xμ)}Aξ+cα{w(X)AU+u(AX)W+(λ-α)(w(X)U+u(X)W)}=0.

Putting X = W in this, we get

α(WA)W+(Wα)AW-(Wμ)Aξ+(α+2cα)AU+2cλαU-μμ=0.

Combining (4.9) to (4.10), we obtain

μφWU-μ(Wα)ξ+μμ=2(α+cα)AU+(μ2+2cαλ)U.

If we apply φ to this and make use of (2.8), (2.11) and (3.3), then we find

-μWU-μ2g(AW,W)ξ+μφμ=2(α+cα)(λAξ-A2ξ)-μ(μ2+2cαλ)W,

which together with (2.6) yields

μWU=μφμ+(2c-μ2)Aξ+2μ(α+cα)AW-(αμ2+2cλ+μ2g(AW,W))ξ.

Now, we can take a orthonormal frame field {e0 = ξ, e1 = W, e2, . . . , en, en+1 = φe1 = (1/μ)U, en+2 = φe2, . . . , e2n = φen} of M. Differentiating (2.6) covariantly and making use of (2.1), we find

(XA)ξ+AφAX=(Xα)ξ+αφAX+(Xμ)W+μXW,

which implies

μdivW=μi=02ng(eiW,ei)=ξh-ξα-Wμ.

Taking the inner product with Y to (4.15) and taking the skew-symmetric part, we have

-2cg(φX,Y)+2g(AφAX,Y)=(Xα)η(Y)-(Yα)η(X)+αg((φA+Aφ)X,Y)+(Xμ)w(Y)-(Yμ)w(X)+μ(g(XW,Y)-g(YW,X)).

Putting X = ξ in this and using (2.10) and (4.11), we have

μξW=3AU-αU+α-(ξα)ξ-(Wα)W.

Putting X = μW in (4.15) and taking account of (4.10), we get

-(α+cα)AU-cα(λ+α)U+μμ+μAφAW=μ(Wα)ξ+μ(Wμ)W+μαφAW+μ2WW,

or, using (3.5) and (4.2),

μ2WW=-2(α+cα)AU-(μ2+2cαλ)U+μμ-μ(Wα)ξ-μ(Wμ)W.

Now, putting X = U in (4.17) and making use of (2.6) and (3.3), we have

μ(g)(UW,Y)-g(YW,U))=(2cμ-Uμ)w(Y)-(Uα)η(Y)+μ2η(AY)+2λμw(AY)-2μw(A2Y),

which together with (4.3) gives

μdw(U,Y)=(2cμ-Uμ)w(Y)-{Uα+2c(λ-α)}η(Y)-{μ2+2(λ-α)g(AW,W)}η(AY)+2μ(λ+cα)w(AY).

Because of (2.10) and (4.18), it is verified that

μdw(ξ,X)=2u(AX)-αu(X)-(ξα)η(X)-(Wα)w(X)+Xα.

Using (2.11) and (3.7), we obtain

du(ξ,X)=(3λ-2α)η(AX)-2μw(AX)-αλη(X)+g(φα,X).

Using above two equations, (3.13) is reduced to

g(XU,Y)+g(YU,X))=2c(g(X,Y)-η(X)η(Y))-2g(A2X,Y)+2αg(AX,Y)+(ξτ)(u(X)w(Y)+u(Y)w(X))+1α(2u(AX)+Xα-(ξα)η(X)-(Wα)w(X))u(Y)+1α(2u(AY)+Yα-(ξα)η(Y)-(Wα)w(Y))u(X)+{(3λ-2α)η(AX)-2μw(AX)-αλη(X)+g(φα,X)}(η(Y)+τw(Y))+{(3λ-2α)η(AY)-2μw(AY)-αλη(Y)+g(φα,Y)}(η(X)+τw(X)),

where we have used (4.21) and (4.22). Taking the trace of this and using (4.7), we find

divU=2c(n-1)+αh-TrA2+λ(λ-α).

Replacing X by U in (4.23) and using (4.6) and (4.7), we find

g(UU,Y)+g(YU,U)=(λ-α)(Yα)+2(2λ-α+cα)u(AY)+{Uαα+2cλα+2(λ-α)(g(AW,W)+α-λ)}u(Y)+{μ(Wα)-(λ-α)ξα}η(Y)+μ2(ξτ)w(Y).

Since g(∇XU,U) = μ(), it follows that

du(U,X)=-2μ(Xμ)+(λ-α)(Xα)+2(2λ-α+cα)u(AX)+{Uαα+2cλα+2(λ-α)(g(AW,W)+α-λ)}u(X)+{μ(Wα)-(λ-α)ξα}η(X)+μ2(ξτ)w(X),

which implies that

du(U,W)=-2μ(Wμ)+(λ-α)Wα+μ2(ξτ).

We will continue our discussions under the hypotheses as those stated in Section 4.

Putting Z = U in (3.12), we find

-2μg((YA)X,W)+2cη(X)u(Y)-du(U,X)(η(Y)+τw(Y))-du(U,Y)(η(X)+τw(X))-dη(U,X)u(Y)-dη(U,Y)u(X)+μ2(g(Xξ,Y)+g(Yξ,X))+μ2((Xτ)w(Y)+(Yτ)w(X))+τ{μ2(g(XW,Y)+g(YW,X))-dw(U,Y)u(X)-dw(U,X)u(Y)}-(Uτ)(u(X)w(Y)+u(Y)w(X))=0.

Because of (2.1), (2.11) and (3.3), we see

dη(U,X)=(λ-α)η(AX)-2μw(AX).

Using this and (2.4), above equation reformed as

-2μg((WA)Y,X)-2c(η(Y)u(X)+η(X)u(Y))-du(U,X)(η(Y)+τw(Y))-du(U,Y)(η(X)+τw(X))+μ2((Xτ)w(Y)+(Yτ)w(X))-(Uτ)(u(X)w(Y)+u(Y)w(X))-{(λ-α)η(AX)-2μwAX)}u(Y)-{(λ-α)η(AY)-2μw(AY)}u(X)+μ2(g(Xξ,Y)+g(Yξ,X))+τ{μ2(g(XW,Y)+g(YW,X))-dw(U,Y)u(X)-dw(U,X)u(Y)}=0.

Substituting (4.20) into this, we obtain

2μg((WA)Y,X)=-2c(η(Y)u(X)+η(X)u(Y))-du(U,X)(η(Y)+τw(Y))-du(U,Y)(η(X)+τw(X))+μ2((Xτ)w(Y)+(Yτ)w(X))-(Uτ)(u(X)w(Y)+u(Y)w(X))-{(λ-α)η(AX)-2μw(AX)}u(Y)-{(λ-α)η(AY)-2μw(AY)}u(X)+μ2(g(Xξ,Y)+g(Yξ,X))+τμ2(g(XW,Y)+g(YW,X))-1αu(X){(2cμ-Uμ)w(Y)-(Uα+2c(λ-α))η(Y)-{μ2+2(λ-α)g(AW,W)}η(AY)+2μ(λ+cα)w(AY)}-1αu(Y){(2cμ-Uμ)w(X)-{Uα+2c(λ-α)}η(X)-{μ2+2(λ-α)g(AW,W)}η(AX)+2μ(λ+cα)w(AX)}.

Combining this to (4.12), we have

-2μ(Wα)g(AY,X)+2η(AY){-(α+cα)u(AX)-cα(α+λ)u(X)+μXμ}+2{-(α+cα)u(AY)-cα(α+λ)u(Y)+μ(Yμ)}η(AX)-2cμα{u(AX)w(Y)+u(AY)w(X)+(λ-α)(w(X)u(Y)+w(Y)u(X))}=-2αc(η(Y)u(X)+η(X)u(Y))-αdu(U,X)(η(Y)+τw(Y))-αdu(U,Y)(η(X)+τw(X))+αμ2((Xτ)w(Y)+(Yτ)w(X))-α(Uτ)(u(X)w(Y)+u(Y)w(X))-μ2(η(AX)u(Y)+η(AY)u(X))+2αμ(w(AY)u(X)+w(AX)u(Y))+αμ2(g(Xξ,Y)+g(Yξ,X))+μ3(g(XW,Y)+g(YW,X))-u(X){(2cμ-Uμ)w(Y)-(Uα+2c(λ-α))η(Y)-(μ2+2(λ-α)g(AW,W))η(AY)+2μ(λ+cα)w(AY)}-u(Y){(2cμ-Uμ)w(X)-(Uα+2c(λ-α))η(X)-(μ2+2(λ-α)g(AW,W))η(AX)+2μ(λ+cα)w(AX)}.

If we put Y = W in (5.1) and take account of (2.1), (3.5) and (4.19), then we find

-2μ(Wα)w(AX)+μ2(Xμ)+2μ(Wμ)η(AX)-2cμα{u(AX)+(λ-α)u(X)}=-μdu(U,X)-αdu(U,W)(η(X)+τw(X))+αμ2((Xτ)+(Wτ)w(X))-μ2{(Wα)η(X)+(Wμ)w(X)}+(Uμ-α(Uτ)-2cαμg(AW,W))u(X),

or, using (4.25) and (4.26)

2μ(Wα)AW-2cμU+{μ(λ-α)ξα-3μ2Wα-αμ2(ξτ)}ξ-{μ2(Wμ)+τμ2(Wα)+2μ3(ξτ)}W+μ2μ-μ(λ-α)α-2μ(2λ-α)AU-μ{Uαα+2(λ-α)g(AW,W)-2(λ-α)2}U+αμ2((Wτ)W+τ)+{Uμ-α(Uτ)-2cμαg(AW,W)}U=0.

By the way, since ατ = μ, we find

αμτ=μμ-(λ-α)α.

Using this, above equation is reduced to

μμ-(λ-α)α=(2λ-α)AU+{(λ-α+cα)g(AW,W)-(λ-α)2+c}U-(Wα)AW+{2μ(Wα)-(λ-α)ξα}ξ+(λ-α)(2Wα-τ(ξα))W.

If we take the inner product (5.3) with W, then we get

μ(Wμ)={3(λ-α)-g(AW,W)}Wα-τ(λ-α)ξα.

Also, taking the inner product (5.3) with U and making use of (4.7), we obtain

Uμμ-Uαα=(3λ-2α+cα)g(AW,W)+(λ-α)(2α-3λ)+c.

On the other hand, replacing Y by W in (4.23) and using (4.3), we find

g(XU,W)+g(WU,X)-μαg(φα,X)-(ξτ)u(X)-{μ(3λ-2α)-2μg(AW,W)+g(φα,W)}(η(X)+τw(X))+2(λ-2α-cα)g(AW,X)-2cw(X)+μα(4α-3λ+2g(AW,W))η(AX)+μ(λ+2cα)η(X)=0,

or using (4.14),

g(XU,W)+g(φμ,X)-λ-αμg(φα,X)-(ξτ)u(X)+2(λ-α)w(AX)+{Uαα+(λ-α)(5α-6λ+4g(AW,W))}w(X)+{Uαμ+μ(4α-5λ+3g(AW,W))}η(X)=0.

By the way, applying (5.3) by φ and making use of (2.6), (3.3) and (3.5), we have

μφμ-(λ-α)φα=-1μ(Wα)AU+μ(ξτ)U+μ2(2λ-α)ξ-μ(2λ-α)AW-μ{(λ-α+cα)g(AW,W)-(λ-α)(3λ-2α)+c}W.

Substituting this into (5.6), we find

g(XU,W)=Wαμ2u(AX)+αw(AX)+{3(λ-α)2+cαg(AW,W)+c-Uαα-3(λ-α)g(AW,W)}w(X)+{3μ(λ-α-g(AW,W))-Uαμ}η(X).

On the other hand, (4.12) turns out, using (2.4), to be

α(XA)W=cαμ(η(X)U+2u(X)ξ)-(Wα)AX+1μη(AX){μμ-(α+cα)AU-cα(λ+α)U}+1μ{μ(Xμ)-(α+cα)u(AX)-cα(λ+α)u(X)}Aξ-cα{w(X)AU+u(AX)W+(λ-α)(u(X)W+w(X)U)}.

If we apply by φ to this and make use of (3.3), then we find

-αφ(XA)W=(Wα)φAX+cαη(X)W-(Xμ)U+1μη(AX){(α+cα){(λ-α)Aξ-μAW}-cαμ(λ+α)W-μφμ}+1μ{(α+cα)u(AX)+2cλαu(X)}U+cαw(X)(μ2ξ-μAW).

Now, if we put Z = W in (3.12), then we find

2g(φ(YA)W,X)=2{(w(A2Y)-cw(Y))η(X)-w(AY)η(AX)}+du(W,X)(η(Y)+τw(Y))+τdu(Y,X)+(Wτ)(w(Y)u(X)+w(X)u(Y))+(g(WU,Y)+g(YU,W))(η(X)+τw(X))+2μ{u(AX)+(λ-α)u(X)}u(Y)+(Yτ)u(X)-(Xτ)u(Y)+τ(u(Y)g(WW,X)+u(X)g(WW,Y)).

Using (2.1), (2.10), (3.5) and (3.8), we can write the above equation as

2αg(φ(YA)W,X)=μdu(Y,X)-2cη(X)w(AY)+2μ(c+α2+αg(AW,W))η(X)η(Y)+2(αμ2+μ2g(AW,W)-cα)η(X)w(Y)-2αη(AX)w(AY)+α(Wτ)(w(X)u(Y)+w(Y)u(X))+αg(WU,X)(η(Y)+τw(Y))+αg(WU,Y)(η(X)+τw(X))-g(XU,W)η(AY)+g(YU,W)η(AX)+2αμ{u(AX)+(λ-α)u(X)}u(Y)+α((Yτ)u(X)-(Xτ)u(Y))+μ(u(X)g(WW,Y)+u(Y)g(WW,X)),

or using (5.8),

μdu(X,Y)=(Wα)g((φA+Aφ)X,Y)+2cαμ(w(X)w(AY)-w(Y)w(AX))+η(AX)g(φμ,Y)-η(AY)g(φμ,X)+2cμα(u(X)u(AY)-u(Y)u(AX))-(Xμ)u(Y)+(Yμ)u(X)+α((Xτ)u(Y)-(Yτ)u(X))+g(YU,W)η(AX)-g(XU,W)η(AY)+{2cα-2cλ-μ2(α+g(AW,W))}(η(X)w(Y)-η(Y)w(X)),

which together with (5.2) and (5.7) yields

μdu(X,Y)=(Wα)g((φA+Aφ)X,Y)+2cμα(w(X)w(AY)-w(Y)w(AX))+Wαμ2(η(AX)u(AY)-η(AY)u(AX))+η(AX)g(φμ,Y)-η(AY)g(φμ,X)+α(η(AX)w(AY)-η(AY)w(AX))+2cμα(u(X)u(AY)-u(Y)u(AX))+μα((Xα)u(Y)-(Yα)u(X))+{(μ2+c)g(AW,W)+αμ2-cα+2cλ}(η(X)w(Y)-η(Y)w(X)).

Putting X = φei and Y = ei in this and summing up for i = 1, 2, · · ·, n, we obtain

μi=02ndu(φei,ei)=(h-α-g(AW,W))Wα-μ(Wμ),

where we have used (2.6)–(2.8), (3.5) and (4.7). Taking the trace of (2.12), we obtain

i=02ng(φeiU,ei)=ξα-ξh.

Thus, it follows that

μ(ξh-ξα)=μ(Wμ)+(g(AW,W)+α-h)Wα,

which together with (4.16) gives

μ2(divW)=(g(AW,W)+α-h)Wα,

We notice here that

Remark 5.1

If AU = σU for some function σ on Ω, then AW ∈ span{ξ,W} on Ω, where span{ξ,W} is a linear subspace spanned by ξ and W.

In fact, because of the hypothesis AU = σU, (3.5) reformed as

μφAW=(σ+λ-α)U,

which implies that AW = μξ + (σ + λα)W ∈ span{ξ,W}.

Now, we prepare the following lemma for later use.

Lemma 5.2

Let M be a real hypersurface of Mn(c), c ≠= 0 which satisfies Rξφ = φRξandφξξRξ = 0. If AW ∈ span{ξ,W}, then Ω = ∅︀.

Proof

Since (3.5) and AW = μξ + g(AW,W)W, we have

AU=(g(AW,W)+α-λ)U.

From (4.2) we also have

g(AW,W)(αAU+cU)=0.

Now, suppose that g(AW,W) ≠= 0 on Ω. Then we have αAU + cU = 0 on this subset, which together with (5.12) gives

μ2=αg(AW,W)+c.

From this and (2.16) we have RξW = 0 and consequently Rξ = 0 on the subset because of (2.6) and (2.16). If we take (3.1) by Rξ and using RξU = 0 and Rξ = 0, we obtain Rξ(φA) = 0, that is, Rξ(ℒξg) = 0 on the subset, where ℒξ denotes the operator of the Lie derivative with respect to ξ. Owing to Theorem 5.1 of [5], it is verified that = αξ, a contradiction. Therefore we have the following

g(AW,W)=0

on Ω. So we have

AW=μξ.

From (5.12) and (5.14), we get

AU=(α-λ)U.

Differentiating (5.15) covariantly, we find

(XA)W+AXW=(Xμ)ξ+μXξ.

Taking the inner product with W and making use of (2.11) and (5.16), we have

g((XA)W,W)=2(λ-α)u(X)

Using (2.4) it reformed as

(WA)W=2(λ-α)U.

On the other hand, (4.13) is reduced, using (5.16) and (5.17), to

(μ2+2c)U=-μ(Wα)ξ+(Wμ)Aξ+μμ.

Taking the inner product with W, we have

Wμ=0.

Hence, it follows from (5.18) that

μμ=μ(Wα)ξ+(μ2+2c)U,

which shows that for any vector fields X

μ(Xμ)=μ(Wα)η(X)+(μ2+2c)u(X).

Differentiating this covariantly and using (2.1), we have

(Yμ)(Xμ)+μ(Y(Xμ))=Y(μ(Wα))η(X)+μ(Wα)g(φAY,X)+(2μ(Wα)η(Y)+2(μ2+2c)u(Y))u(X)+(μ2+2c)g(YU,X)+{μ(Wα)η(YX)+(μ2+2c)u(YU)}.

Taking the skew-symmtric part of this, we find

Y(μ(Wα))η(X)-X(μ(Wα))η(Y)+(μ(Wα))g((φA+Aφ)Y,X)+2μ(Wα)(η(Y)u(X)-η(X)u(Y))+(μ2+2c)(g(YU,X)-g(XU,Y))=0.

Replacing Y by ξ in this, and using (2.10) and (5.17), we have

X(μ(Wα))-2μ(Wα)u(X)=ξ(μ(Wα))η(X)+(μ(Wα))u(X)+(μ2+c)(g(ξU,X)-μ2η(X)).

Substituting this into (5.21), we obtain

μ(Wα)(u(Y)η(X)-u(X)η(Y))+(μ2+2c)(g(ξU,Y)η(X)-g(ξU,X)η(Y))+μ(Wα)g((φA+Aφ)Y,X)+(μ2+2c)(g(YU,X)-g(XU,Y))=0.

Putting Y = U in this, and using (2.8), (3.3), (4.11), (5.15) and (5.16), we obtain

(μ2+2c)(g(UU,X)-μ(Xμ))+μ(μ2+2c)(Wα)η(X)+μ2(λ-α)(Wα)w(X)=0.

On the other hand, putting Y = U in (5.9) and making use of (5.14), (5.15) and (5.19), we have

g(UU,X)-μ(Xμ)=2(λ-α)(Wα)w(X)+Uααu(X)-(λ-α)Xα.

Combining this to (5.22), we have

(μ2+2c){2(λ-α)(Wα)w(X)+Uααu(X)-(λ-α)Xα}+μ(μ2+2c)(Wα)η(X)-μ2(λ-α)(Wα)w(X)=0.

If we put X = W in this, then we have

(μ2+2c)(λ-α)Wα=(μ2+4c)(λ-α)Wα,

which, together with λ ≠= α, shows that

Wα=0.

Thus, (5.20) becomes

μμ=(μ2+2c)U,

which implies

φμ=-(μ2+2c)W.

Using (5.23), we can write (5.21) as

(μ2+2c)(g(YU,X)-g(XU,Y))=0.

Now, suppose that μ2 + 2c ≠= 0. Then we have g(∇YU,X) − g(∇XU, Y ) = 0. Using (5.14)–(5.16), (5.20), (5.23) and (5.25), we can write (5.9) as

(μ2+c)(w(X)η(Y)-w(Y)η(X))=0,

which implies μ2 + c = 0. So μ is constant. Thus, (5.23) becomes μ2 + 2c = 0, a contradiction. Therefore, we see that μ2 + 2c = 0.

Accordingly we see that μ is constant, which together with (5.4) yields ξα = 0. Hence (5.3) is reduced to

μ2α={μ2(3λ-2α)-cα}U.

Taking the inner product this to X and differentiating covariantly, we find

μ2(Y(Xα))={μ2(3Yλ-2Yα)-cα}u(X)+{μ2(3λ-2α)-cα}(g(YU,X)+g(U,YX)).

The skew-symmetric part of this is given by

3μ2((Yλ)u(X)-(Xλ)u(Y))+(2μ2+c)((Xα)u(Y)-(Yα)u(X))+{μ2(3λ-2α)-cα}(g(YU,X)-g(XU,Y))=0,

which implies that ∇λ = χU for some function χ, where we have used (4.24) and (5.26). Thus it follows that

{μ2(3λ-α)-cα}(g(YU,X)-g(XU,Y))=0.

If g(∇YU,X) − g(∇XU, Y ) = 0, then similarly as above we have a contradiction. Thus we have μ2(3λ −2α) − = 0, which together with μ2+2c = 0 gives 2λα = 0. i.e. 2μ2 + α2 = 0, a contradiction. Therefore Lemma 5.2 is proved.

Lemma 5.3

α2φ(λ-h)=-4μ(μ2+c)(AW-μξ)+αμ(h-λ)(Wα)U+fW,

for some function f on Ω.

Proof

Putting X = Y = ei in (3.12), summing up for i = 0, 1, · · ·, 2n and using (2.1) and (2.4), we find

Tr(φZA)-2c(n-1)η(Z)+(TrA2)η(Z)-hη(AZ)+g(ξU,Z)-g(ZU,ξ)+τ(g(WU,Z)+g(UW,Z))+(divU)(η(Z)+τw(Z))+g((φA+Aφ)U,Z)+(Wτ)u(Z)+(Uτ)w(Z)+τ(divW)u(Z)=0,

or using (2.10), (3.3), (3.7) and (4.24)

φα-φh+μα(WU+UW)-4μAW+(Wτ+τ(divW))U+(Uτ+τ(divU)+μ(4λ-3α-h))W+(λ-α)(λ+3α)ξ=0.

On the other hand, combining (4.20) to (5.8) and making use of (5.7), we find

UW=1μ{μφμ-(λ-α)φα}-(ξτ)U+2(2λ-α+cα)AW+{(λ-α)(2α-3λ)+c-(λ+cα)g(AW,W)}W+μ{g(AW,W)+3α-5λ-2cα}ξ.

Substituting this and (4.14) into (5.27), we find

αφ(α-h)+2μφμ-(λ-α)φα=4μ(α-λ-cα)AW+(μ(ξτ)-α(Wτ)-μ(divW))U-4(λ-α)(μ2+c)ξ-(α(Uτ)+μ(divW))W-μ{3c+(λ-α)(α-3λ)-(λ+cα)g(AW,W)+4αλ-3α2-hα}W.

From (4.11), (4.16) and (5.2) we have

αμ(μ(ξτ)-α(Wτ)-μ(divW))=2μ2(Wα)-μ(λ-2α)ξα-αμ(ξh).

By the way, using (5.4) and (5.10) we have

μ(λ-2α)ξα+αμ(ξh)=α(3λ-2α-h)Wα.

Thus, we have

αμ(μ(ξτ)-α(Wτ)-μ(divW))=α(h-λ)Wα.

Differentiating (3.2) covariantly, we find

2μμ=(λ-2α)α+αλ.

Using this and (5.29), the equation (5.28) reformed as

α2φ(λ-h)=-4μ(μ2+c)(AW-μξ)+αμ(h-λ)(Wα)U+fW,

where we have put

f=αμ{hα+4α2-8αλ+3λ2-3c+(λ+cα)g(AW,W)-divU-αμ(Uτ)}.

This completes the proof of Lemma 5.3.

We will continue our discussions under the same hypotheses as those in Section 4. Further we assume that TrRξ is constant, that is, g(Sξ, ξ) is constant. Then, from (2.5) we see that β is constant, i.e.

α(h-λ)=C,

where C is some constant. Differentiating this covariantly, we have

(λ-h)α+α(λ-h)=0.

So we have αφ(∇λ − ∇h) = (hλ)φα. Thus, from Lemma 5.3 we find

α(h-λ)μφ(α-(Wα)W)=-4(μ2+c)(AW-μξ)+αμfW,

which tells us that

α(h-λ)μ2(Uα)=4(μ2+c)g(AW,W)-αμf.

Combining the last two equations, it follows that

α(h-λ)μφ(α-(Wα)W-Uαμ2U)=-4(μ2+c)(AW-μξ-g(AW,W)W).

Applying this by φ and using (3.5), we find

α(h-λ)(α-(ξα)ξ-(Wα)W-Uαμ2U)=4(μ2+c){AU+(λ-α)U-g(AW,W)U}.

Taking the inner product with AW to this, and using (4.6), (5.4) and α ≠= 0, we see

(h-λ)(g(AW,α)-μ(ξα)-g(AW,W)(Wα))=0.

First of all, we prove the following:

Lemma 6.1

hλ ≠= 0 on Ω.

Proof

If not, then we have from (6.4)

(μ2+c){AU-(g(AW,W)+α-λ)U}=0

on this subset. Because of Remark 5.1 and Lemma 5.2, it is verified that μ2 +c = 0 on the set and hence μ is constant. Accordingly we see that = 0 because of (4.11) and hence ξα = 0 and ξτ = 0 by virtue of (5.2) and (5.4). Thus, (5.3) reformed as

(λ-α)α+(2λ-α)AU+{c-(λ-α)2}U=0,

which together with μ2 + c = 0 implies that

Xα=λu(X)+ɛg(AU,X)

for any vector field X, where we have put cɛ = α2 − 2c. Differentiating (6.6) covariantly with respect to a vector field Y and taking skew-symmetric part, we get

(Yλ)u(X)-(Xλ)u(Y)+λ(g(YU,X)-g(XU,Y))+(Yɛ)u(AX)-(Xɛ)u(AY)+ɛ{cμ(η(Y)w(X)-η(X)w(Y))+g(AYU,X)-g(AXU,Y)}=0.

where we have used the Codazzi equation (2.4). Since ξα = 0 and (6.1), by replacing X by ξ in this, we get

-λ(g(ξU,Y)+g(Yξ,U))+ɛ(g(YU,αξ+μW)-cμw(Y)-g(ξU,AY))=0,

where we have used (2.6), which together with (2.10) and (5.9) implies that

ɛAξU+λξU+μλAWspan{ξ,W}.

On the other hand, we can write (3.7) as

ξU=-μ(ɛ+3)AW+(λ-α)(ɛ+2)Aξ,

where we have used (3.3) and (6.6), which together with (2.6), (4.3) and the fact that μ2 + c = 0 yields

AξU=-μ(λ-α)AW+{c-(λ-α)(ɛ+3)g(AW,W)}Aξ-c(λ-α)(ɛ+3)ξ.

Combining the last three equations, it is seen that

{(2λ-α)ɛ+2λ}AWspan{ξ,W},

which shows that (2λα)ɛ + 2λ = 0 by Lemma 5.2. So we have (2λα)(α2 − 2c) + 2 = 0, a contradiction because of μ2 + c = 0. This completes the proof.

If we combine (6.2) to (5.30), then we have

2μμ=(h-2α)α+αh.

If we apply this by ξ, then we find

2μ(ξμ)=(h-2α)ξα+α(ξh).

From (4.11), (5.6) and (5.12) we get (hλ)(μ(ξα) − α()) = 0 and hence

μ(ξα)=α(Wα)

by virtue of Lemma 6.1, which together with (6.9) yields

μ(ξh)=(2λ-h)Wα.

From (5.2) and (6.10) we have ξτ = 0. Thus, using (6.8) and (6.10) we verify from (5.3)

12(hα+αh)-λα+(Wα)AW=(2λ-α)AU+{(λ-α+cα)g(AW,W)-(λ-α)2+c}U+(Wα){μξ+(λ-α)W}.

Because of Lemma 6.1, (6.5) implies that

g(AW,α)=(α+g(AW,W))Wα,

with the aid of (6.10). Applying (5.3) by AW and making use of (4.3), (6.10), (6.13) and ξτ = 0, we find

μg(AW,μ)-(λ-α)(α+g(AW,W))Wα+g(A2W,W)Wα={μ2+(λ-α)g(AW,W)}Wα,

which together with (4.5) gives

μαg(AW,μ)={(μ2+c)g(AW,W)+αμ2}Wα.

In the next place, we will prove that

Lemma 6.2

ξα = = = ξh = ξλ = = 0 and ξ(g(AW,W)) = 0 on Ω.

Proof

Differentiating (4.4) covariantly, we get

g(A2W,W)(Xα)+α(X(g(A2W,W)))=2μg(AW,W)(Xμ)+(μ2-c)(X(g(AW,W)))+μ2(Xα)+2μα(Xμ).

Replacing X by ξ in this, and using (4.11) and (6.10), we find

α(ξ(g(A2W,W)))=αμ(μ2-g(A2W,W))Wα+2μ(α+g(AW,W))Wα+(μ2-c)(ξ(g(AW,W))).

By the way, using (4.10), (4.18), (6.10) and (6.13), we verify that ξ(g(AW,W)) = , which together with (5.4) and (6.10) yields

ξ(g(AW,W))=1μ{2(λ-α)-g(AW,W)}Wα.

Substituting this and (4.5) into (6.14), we find

α2ξ(g(A2W,W))={cμg(AW,W)+μα+μα(μ2-c)}Wα.

On the other hand, we have

12(X(g(A2W,W)))=g((XA)W,AW)+g(A2W,XW),

which implies

12α(X(g(A2W,W)))=αg((WA)X,AW)+2cαu(X)-cg(AW,XW)+cu(AX)+α(α+g(AW,W))u(AX),

where we have used (2.6), (2.11) and (4.3).

By the way, putting X = AW in (4.12) and making use of (2.6) and (6.14), we obtain

α(WA)AW=-(Wα)A2W+(α+g(AW,W)){-(α+cα)AU-cα(λ+α)U+μμ}+1μα{(μ2+c)g(AW,W)+αμ2}(Wα)Aξ-cαg(AW,W){AU+(λ-α)U},

which implies

αg((WA)AW,ξ)=1μ{αμ2+(μ2+c)g(AW,W)}

because of (2.6) and (4.11). If we replace X by ξ in (6.16) and make use of (4.11), (4.18) and (6.17), then we obtain

(μ2-c-αg(AW,W))(Wα)=0

because of λα ≠= 0.

Now, suppose that ≠= 0 on Ω. Then since λ ≠= α, we have αg(AW,W) = μ2c, which together with (3.2) and (4.7) gives αg(AU,U) = −2. From this and (4.6) we verify that α2g(A2U,U) = c2μ2. Using the last two equations it is seen that ||αAU + cU||2 = 0 and hence αAU + cU = 0. Thus, (3.5) is reduced to μφAW = (λαc/α)U, which shows that AW = μξ + g(AW,W)W on this subset. According to Lemma 5.2, we have Ω = ∅︀, and hence = 0 on Ω. Thus, it is clear that = 0, ξα = 0, ξh = 0 and ξλ = 0, where we have used (4.11), (5.6), (6.2), (6.9), (6.10) and (6.11). Since (3.2), = 0 and = 0, we have = 0. Hence Lemma 6.2 is proved.

Because of Lemma 6.2, we can write (6.4) as

α(h-λ)(α-Uαμ2U)=4(μ2+c){AU-(λ-α-g(AW,W))U},

which tells us that

14α(h-λ)α=(μ2+c)AU+θU,

where the function θ is defined by

μ2θ=α(h-λ)4(Uα)-(μ2+c)g(AU,U).

We also have from (5.4)

μμ-(λ-α)α=(2λ-α)AU+ρU,

where we have put

ρ=(λ-α+cα)g(AW,W)-(λ-α)2+c.

Remark 6.3

μ2 + c ≠= 0 on Ω.

If not, then we have μ2+c = 0 and hence μ is constant on this subset. So (6.19) and (6.20) are reduced respectively to

μ2α=(Uα)U,(λ-α)α+(2λ-α)AU+{c-(λ-α)2}U=0

because of Lemma 5.2. Combining these two equations, we obtain

(2λ-α)AU={(λ-α)2-c-Uαα}U.

Suppose that 2λα = 0 on this subset. Then, the equation μ2 + c = 0 becomes α2 − 2c = 0, a contradiction. Thus we have 2λα ≠= 0. Owing to Remark 5.1 and Lemma 5.2, above equation produces a contradiction. Hence μ2 + c ≠= 0 on Ω is proved.

Lemma 6.4

(2λα)θ = (μ2 + c)ρ on Ω.

Proof

From (6.17) and (6.18) we have

14α(h-λ)(2λ-α)α-(μ2+c){12μ2-(λ-α)α}={(2λ-α)θ-(μ2+c)ρ}U.

Using the same method as that used to derive (6.7) from (6.6), we can deduce from this that

(2λ-α)(ξθ)U+{(2λ-α)θ-(μ2+c)ρ}(ξU+μAW)=0,

where, we have used (2.10), (6.1) and Lemma 6.2. If we take the inner product with U to this and make use of ξμ = 0, then we get (2λα)ξθ = 0 and hence

{(2λ-α)θ-(μ2+c)ρ}(ξU+μAW)=0.

If (2λα)θ − (μ2 + c)ρ ≠= 0 on Ω, then we have

ξU+μAW=0.

We discuss our arguments on such a place. Using (3.7), the last equation can be written as

φα=2μAW+(2α-3λ)Aξ+αλξ.

Applying this by φ and taking account of (3.5) and Lemma 6.2, we obtain

α=-2AU+λU.

Combining this to (6.19), we obtain

{μ2+c+12α(h-λ)}AU={14αλ(h-λ)-θ}U.

Because of Remark 5.1 and Lemma 5.2, we conclude that μ2+c+(1/2)α(hλ) = 0. Hence it follows from (6.1) that μ is constant. Thus, (6.20) reformed as

(λ-α)α=(α-2λ)AU-ρU,

which together with (6.22) implies that αAU = {λ(αλ) − ρ}U. Therefore we verify that (2λα)θρ(μ2 +c) = 0 by virtue of Remark 5.1 and Lemma 5.2. This completes the proof.

Lemma 6.5

Let span{ξ,W} be the linear subspace spanned by ξ and W. Then there exists P ∈ span{ξ,W} such that

g(AW,XU)=cαw(A2X)-{μ2+(λ-α+cα)g(AW,W)}w(AX)+g(P,X).
Proof

Putting Y = AW in (5.9) and using (3.6), (4.3), (6.13) and Lemma 6.2, we find

μdu(X,AW)=2cαμ{g(A2W,W)w(X)-g(AW,W)w(AX)}+η(AX)g(φμ,AW)-μ(α+g(AW,W))g(φμ,X)+α{g(A2W,W)η(AX)-μ(α+g(AW,W))w(AX)}+{(μ2+c)g(AW,W)+αμ2-cα+2cλ}(g(AW,W)η(X)-μw(X)),

which enables us to obtain

g(AW,XU)-g(AWU,X)=-α(α+g(AW,W)+2cα2g(AW,W))w(AX)-(α+g(AW,W))g(φμ,X)+g(P1,X),

for some P1 ∈ span{ξ,W}. If we replace X by AW in (4.23) and make use of (3.5), (4.3), (6.14) and Lemma 6.2, then we get

g(XU,AW)+g(AWU,X)=2cw(AX)+2αw(A2X)-2w(A3X)+(μ+μαg(AW,W)){(3λ-2α)η(AX)-2μw(AX)-αλη(X)+g(φα,X)}+{μ(3λ-2α)(α+g(AW,W))-2μg(AW,W)-αλμ-1μg(AU+(λ-α)U,α)}(η(X)+τw(X))-2cμη(X),

which shows that

g(XU,AW)+g(AWU,X)=-2w(A3X)+2αw(A2X)+2cw(AX)-2(λ-α)(α+g(AW,W))w(AX)+μα(α+g(AW,W))g(φα,X)+g(P2,X),

for some P2 ∈ span{ξ,W}. Adding to the last two equations, we obtain

2g(AW,XU)=-2w(A3X)+2αw(A2X)+2cw(AX)-2(λ-α)(α+g(AW,W))w(AX)-α(α+g(AW,W)+2cα2g(AW,W))w(AX)-(α+g(AW,W))(φμ-μαφα)+g(P3,X)

for some P3 ∈ span{ξ,W}.

By the way, applying (6.20) by φ, and using (2.8) and (3.4), we find

φμ-μαφα=(2λ-α){-AW+μξ+(λ-α)W}-ρW.

Because of (4.3), we have

A3W=-cαA2W+(λ-α)(α+g(AW,W))AW+μ(α+cα+g(AW,W))Aξ.

Combining the last three equations, we obtain

g(AW,XU)=cαw(A2X)-{μ2+(λ-α+cα)g(AW,W)}w(AX)+g(P4,X)

for some P4 ∈ span{ξ,W}. The completes the proof.

Remark 6.6

= 0 on Ω.

In fact, we have

W(g(AW,W))=g((WA)W,W)+2g(AW,WW),

which together with (4.13) and Lemma 6.2 yields

W(g(AW,W))=2g((AW,WW).

However, if we take the inner product with AW to (4.19) and make use of Lemma 6.2 and (6.14), then we obtain g(AW,WW) = 0. So we have W(g(AW,W)) = 0, which connected to (6.21) and Lemma 6.2 gives = 0.

We will continue our discussions under the same assumptions as those in Section 6. Taking the inner product X to (6.20) and differentiating covariantly, we have

(Yμ)(Xμ)+μ(Y(Xμ))-(Yλ-Yα)(Xα)-(λ-α)(Y(Xα))=(2(Yλ)-Yα)u(AX)+(2λ-α)(g((YA)U,X)+g(AYU,X))+(Yρ)u(X)+ρg(YU,X)+g((2λ-α)AU+ρU,YX).

Taking the skew-symmetric part of this and using (2.4), we find

(Xλ)(Yα)-(Yλ)(Xα)+(2(Xλ)-Xα)u(AY)-(2(Yλ)-Yα)u(AX)=cμ(2λ-α)(η(Y)w(X)-η(X)w(Y))+(2λ-α)(g(AYU,X)-g(AXU,Y))+(Yρ)u(X)-(Xρ)u(Y)+ρ(g(YU,X)-g(XU,Y)),

where we have used (2.4) and (2.8). Differentiating (6.21) covariantly and taking the inner product ξ to this, it follows from Lemma 6.2 that ξρ = 0. Putting Y = ξ in (7.1) and using (2.6) and ξρ = 0, we find

cμ(2λ-α)w(X)-(2λ-α){g(αξ+μW,XU)+g(ξU,AX)}-ρ(g(XU,ξ)-g(ξU,X))=0,

or using (2.10), (5.7) and Lemma 6.2,

(2λ-α)AξU+ρξU+μρAWspan{ξ,W}.

If we put Y = W in (7.1) and take account of Lemma 6.2 and Remark 6.6, then we have

(2λ-α){g(XU,AW)-g(AWU,X)+cμη(X)}+ρ(g(XU,W)-g(WU,X))=0.

By the way, putting Y = W in (5.9), we have

g(XU,W)-g(WU,X)=-(α+2cα)w(AX)-g(φμ,X)+g(P5,X)

for some P5 ∈ span{ξ,W}, which together with Lemma 6.5 and (7.3) implies that

(2λ-α){cαA2W-(μ2+(λ-α+cα)g(AW,W))AW-AWU}-ρ{(α+2cα)AW+φμ}span{ξ,W}.

It follows from this and (4.14) that

(2λ-α)Aφμ+ρφμ+(2λ-α){cαA2W+(λ-α+cα)g(AW,W)AW}+ρ(α+2cα)AWspan{ξ,W}.

If we take account of (4.3), (6.21) and (6.23), then the last equation can be written as

μα(2λ-α)Aφμ+ρφμ+(2λ-α)2(λ-α+cα)AW+(2λ-α){cαA2W+((λ-α)2-c)AW}+ρ(α+2cα)AWspan{ξ,W}.

On the other hand, from (3.7) we have

AξU=μ(3λ-2α)AW-3μA2W+2μ2Aξ+Aφα,

where we have used (2.6). Substituting this into (7.2), we find

(2λ-α)Aφα+ρφα-2μρAW-(2λ-α)μ{3A2W+(2α-3λ)AW}span{ξ,W}.

Combining this to (7.4), we obtain

(λ-α){-ρμφα+2ρAW+(2λ-α)(3A2W+(2α-3λ)AW)}+ρφμ+(2λ-α)2(λ-α+cα)AW+cα(2λ-α)A2W+(2λ-α){(λ-α)2-c}AW+ρ(α+2cα)AWspan{ξ,W},

which together with (4.3) and (6.23) implies that

{2ρα-(2λ-α)(μ2+c)}AWspan{ξ,W},

that is,

{2ρα-(2λ-α)(μ2+c)}(AW-μξ-g(AW,W))=0.

According to Lemma 5.2, we see that

2ρα=(2λ-α)(μ2+c).

From this fact and Lemma 6.4, we see that 2αθ = (μ2+c)2 by virtue of 2λα ≠= 0. Thus, (6.19) is reduced to

κα=2αAU+(μ2+c)U

with the aid of Remark 6.3, where we have put

κ=α2(h-λ)2(μ2+c).

Differentiating this covariantly and taking the inner product with ξ, it follows from (6.1) and Lemma 6.2 that ξκ = 0.

As in the same method as that used from (6.6) to drive (6.7), we can deduce from (7.6) that

2αg(AξU,X)+(μ2+c)g(ξU,X)=μ{-2cαw(X)-2α2w(AX)-(μ2+c)w(AX)+2αg(XU,W)},

which together with (5.7) implies that

2αAξU+(μ2+c)ξU+μ(μ2+c)AWspan{ξ,W}.

On the other hand, applying (7.6) by φ and using (2.6) and (3.3), we find

κμφα=-2αAW+(μ2+c)W+2αμξ,

which together with (4.3) yields

κμAφα=(μ2+c)AW-2μg(AW,W)Aξ-2cμξ.

From Lemma 6.1 we have κ ≠= 0 and hence combining the last two equations, it is verified that

2αAφα+(μ2+c)φαspan{ξ,W}.

By the way, applying (3.7) by A and using (4.3), we find

2αAξU+(μ2+c)ξU-2αμ(3λ-2α+3cα)AW+3μ(μ2+c)AW-2αAφα-(μ2+c)φαspan{ξ,W},

which together with (7.7) and (7.8) gives

(2μ2+α2+2c)(AW-μξ-g(AW,W))=0.

Owing to Lemma 5.2, we see that 2μ2 + α2 + 2c = 0, which implies that 2μμ + αα = 0. Hence (6.20) reformed as

α+2AU+μ2+cαU=0

by virtue of 2λα ≠= 0 on Ω, where we have used (7.5). Combining this to (6.19), we have

{μ2+c+12α(h-λ)}AU=14{4θ+(h-λ)(μ2+c)}U.

According to Remark 5.1, it follows that μ2+c+(1/2)α(hλ) = 0, which together with (6.1) gives μ is constant and hence α is constant. Thus (7.9) becomes AU = −{(μ2 + c)/(2α)}U, a contradiction by virtue of Remark 5.1.

Therefore we verify that Ω = ∅︀, that is, = αξ on M. Thus, from (2.18) we see that RξS = SRξ. Hence from Theorem 1.2 ([9]) M is homogeneous real hypersurfaces of Type A.

Let M be of Type A. Then M always satisfies ∇φξξRξ = 0. Since TrA is constant and (2.16), it is easy to see that φRξ = Rξφ and TrRξ is constant.

Consequently we conclude that

Theorem 7.1

Let M be a real hypersurface of a complex space form Mn(c), c ≠= 0, n ≥ 3 which satisfiesφξξRξ = 0 and TrRξis constant. Then M holds φRξ = Rξφ if and only if Aξ = 0 or M is locally congruent to one of following:

  • (I) In cases that Mn(c) = Pnwith η() ≠= 0,

    • (A1) a geodesic hypersphere of radius r, where 0 < r < π/2 and r ≠= π/4;

    • (A2) a tube of radius r over a totally geodesic Pkfor some k ∈ {1, . . . , n–2}, where 0 < r < π/2 and r ≠= π/4.

  • (II) In cases Mn(c) = Hnℂ,

    • (A0) a horosphere;

    • (A1) a geodesic hypersphere or a tube over a complex hyperbolic hyperplane Hn–1;

    • (A2) a tube over a totally geodesic Hkfor some k ∈ {1, . . . , n – 2}.

  1. Berndt, J (1989). Real hypersurfaces with constant principal curvatures in complex hyperblic spaces. J Reine Angew Math. 395, 132-141.
  2. Cecil, TE, and Ryan, PJ (1982). Focal sets and real hypersurfaces in complex projective space. Trans Amer Math Soc. 269, 481-499.
  3. Cho, JT, and Ki, U-H (1998). Real hypersurfaces in complex projective spaces in terms of Jacobi operators. Acta Math Hungar. 80, 155-167.
    CrossRef
  4. Cho, JT, and Ki, U-H (2008). Real hypersurfaces in complex space form with Reeb ow symmetric Jacobi operator. Canadian Math Bull. 51, 359-371.
    CrossRef
  5. Ki, U-H, Kim, I-B, and Lim, DH (2010). Characterizations of real hypersurfaces of type A in a complex space form. Bull Korean Math Soc. 47, 1-15.
    CrossRef
  6. Ki, U-H, and Kurihara, H (2009). Real hypersurfaces and ξ-parallel structure Jacobi operators in complex space forms. J Korean Academy Sciences, Sciences Series. 48, 53-78.
  7. Ki, U-H, Kurihara, H, Nagai, S, and Takagi, R (2009). Characterizations of real hypersurfaces of type A in a complex space form in terms of the structure Jacobi operator. Toyama Math J. 32, 5-23.
  8. Ki, U-H, Kurihara, H, and Takagi, R (2009). Jacobi operators along the structure ow on real hypersurfaces in a nonat complex space form. Tsukuba J Math. 33, 39-56.
    CrossRef
  9. Ki, U-H, Nagai, S, and Takagi, R (2010). The structure vector field and structure Jacobi operator of real hypersurfaces in nonat complex space forms. Geom Dedicata. 149, 161-176.
    CrossRef
  10. Ki, U-H, and Suh, YJ (1990). On real hypersurfaces of a complex space form. Math J Okayama Univ. 32, 207-221.
  11. Kimura, M (1986). Real hypersurfaces and complex submanifolds in complex projective space. Trans Amer Math Soc. 296, 137-149.
    CrossRef
  12. Montiel, S, and Romero, A (1986). On some real hypersurfaces of a complex hyperblic space. Geom Dedicata. 20, 245-261.
    CrossRef
  13. Okumura, M (1975). On some real hypersurfaces of a complex projective space. Trans Amer Math Soc. 212, 355-364.
    CrossRef
  14. Ortega, M, Pérez, JD, and Santos, FG (2006). Non-existence of real hypersurfaces with parallel structure Jacobi operator in nonat complex space forms. Rocky Mountain J Math. 36, 1603-1613.
    CrossRef
  15. Pérez, JD, Santos, FG, and Suh, YJ (2006). Real hypersurfaces in complex projective spaces whose structure Jacobi operator is D-parallel. Bull Belg Math Soc. 13, 459-469.
  16. Pérez, JD, Santos, FG, and Suh, YJ (2007). Real hypersurfaces in nonat complex space forms with commuting structure Jacobi operator. Houston J Math. 33, 1005-1009.
  17. Takagi, R (1973). On homogeneous real hypersurfaces in a complex projective space. Osaka J Math. 19, 495-506.
  18. Takagi, R (1975). Real hypersurfaces in a complex projective space with constant principal curvatures I, II. J Math Soc. 15, Array-Array.
    CrossRef