검색
Article Search

JMB Journal of Microbiolog and Biotechnology

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

Article

Kyungpook Mathematical Journal 2022; 62(1): 133-166

Published online March 31, 2022

Copyright © Kyungpook Mathematical Journal.

Submanifolds of Codimension 3 in a Complex Space Form with Commuting Structure Jacobi Operator

U-Hang Ki, Hyunjung Song*

The National Academy of Sciences, Seoul 06579, Korea
e-mail : uhangki2005@naver.com

Department of Mathematics Hankuk University of Foreign Studies, Seoul 02450, Korea
e-mail : hsong@hufs.ac.kr

Received: November 12, 2019; Revised: February 22, 2021; Accepted: February 24, 2021

Let M be a semi-invariant submanifold with almost contact metric structure (ϕ,ξ,η,g) of codimension 3 in a complex space form Mn+1(c) for c0. We denote by S and Rξ be the Ricci tensor of M and the structure Jacobi operator in the direction of the structure vector ξ, respectively. Suppose that the third fundamental form t satisfies dt(X,Y)=2θg(ϕX,Y) for a certain scalar θ2c and any vector fields X and Y on M. In this paper, we prove that if it satisfies Rξϕ=ϕRξ and at the same time Sξ=g(Sξ,ξ)ξ, then M is a real hypersurface in Mn(c) (Mn+1(c)) provided that r¯2(n1)c0, where r¯ denotes the scalar curvature of M.

Keywords: semi-invariant submanifold, distinguished normal, complex space form, structure Jacobi operator, Ricci tensor, Hopf hypersurfaces

A submanifold M is called a CR submanifold of a Kaehlerian manifold M˜ with complex structure J if there exists a differentiable distribution :ppMp on M such that △ is J-invariant and the complementary orthogonal distribution is totally real, where Mp denotes the tangent space at each point p in M ([1], [25]). In particular, M is said to be a semi-invariant submanifold provided that dim=1. The unit normal in J is called the distinguished normal to the semi-invariant submanifold ([4], [23]). In this case, M admits an induced almost contact metric structure (ϕ,ξ,η,g). A typical example of a semi-invariant submanifold is real hypersurfaces. New examples of nontrivial semi-invariant submanifolds in a complex projective space Pn are constructed in [13] and [20]. Therefore we may expect to generalize some results which are valid in a real hypersurface to a semi-invariant submanifold.

An n-dimensional complex space form Mn(c) is a Kaehlerian manifold of constant holomorphic sectional curvature 4c. As is well known, complete and simply connected complex space forms are isometric to a complex projective space Pn, or a complex hyperbolic space Hn according as c>0 or c<0.

For the real hypersurface of a complex space form Mn(c), many results are known. One of them, Takagi([21], [22]) classified all the homogeneous real hypersurfaces of Pn as six model spaces which are said to be A1,A2,B,C,D and E, and Cecil-Ryan ([5]) and Kimura ([14]) proved that they are realized as the tubes of constant radius over Kaehlerian submanifolds when the structure vector field ξ is principal.

On the other hand, real hypersurfaces in Hn have been investigated by Berndt ([2]), Montiel and Romero ([15]) and so on. Berndt ([2]) classified all real hypersurfaces with constant principal curvatures in Hn and showed that they are realized as the tubes of constant radius over certain submanifolds. Also such kinds of tubes are said to be real hypersurfaces of type A0,A1,A2 or type B.

Let M be a real hypersurface of type A1 or type A2 in a complex projective space Pn or that of type A0, A1 or A2 in a complex hyperbolic space Hn. Now, hereafter unless otherwise stated, such hypersurfaces are said to be of type (A) for our convenience sake.

Characterization problems for a real hypersurface of type (A) in a complex space form were studied by many authors ([6], [7], [8], [15], [16], [18], etc.).

Two of them, we introduce the following characterization theorems due to Okumura [18] for c>0 and Montiel and Romero [15] for c<0 respectively.

Theorem O. Let M be a real hypersurface of Pn, n ≥ 2. If it satisfies

g((AϕϕA)X,Y)=0

for any vector fields X and Y, then M is locally congruent to a tube of radius r over one of the following Kaehlerian submanifolds :

  • (A1) a hyperplane Pn1, where 0<r<π/2,

  • (A2) a totally geodesic Pk(1kn2), where 0<r<π/2.

Theorem MR. Let M be a real hypersurface of Hn, n2. If it satisfies (1.1), then M is locally congruent to one of the following hypersurface :

  • (A0) a horosphere in Hn, i.e., a Montiel tube,

  • (A1) a geodesic hypersphere, or a tube over a hyperplane Hn1,

  • (A2) a tube over a totally geodesic Hk(ckn2).

Denoting by R the curvature tensor of the submanifold, we define the Jacobi operator Rξ=R(,ξ)ξ with respect to the structure vector ξ. Then Rξ is a self adjoint endomorphism on the tangent space of a CR submanifold.

Using several conditions on the structure Jacobi operator Rξ, characterization problems for real hypersurfaces of type (A) have recently studied. In the previous paper ([7]), Cho and one of the present authors gave another characterization of real hypersurface of type (A) in a complex projective space Pn. Namely they prove the following :

Theorem CK.([7]) Let M be a connected real hypersurface of Pn if it satisfies (1)RξAϕ=ϕARξ or (2) Rξϕ=ϕRξ,RξA=ARξ, then M is of type (A), where A denotes the shape operator of M.

On the other hand, semi-invariant submanifolds of codimension 3 in a complex projective space Pn+1 have been studied in [10], [12], [13] and so onby using properties of induced almost contact metric structure and those of the third fundamental form of the submanifold. In the preceding work, Ki, Song and Takagi ([13]) assert the following:

Theorem KST.([13]) Let M be a real (2n-1)-dimensional semi-invariant submanifold of codimension 3 in a complex projective space Pn+1 with constant holomorphic sectional curvature 4c. If the structure vector ξ is an eigenvector for the shape operator in the direction of the distinguished normal and the third fundamental form t satisfies dt=2θω for a certain scalar θ(<2c), where ω(X,Y)=g(ϕX,Y) for any vectors X and Y on M, then M is a Hopf hypersurface in a complex projective space Pn.

In this paper, we consider a semi-invariant submanifold M of codimension 3 in a complex space form Mn+1(c),c0 which satisfies Rξϕ=ϕRξ and at the same time Sξ=g(Sξ,ξ)ξ such that the third fundamental form t satisfies dt=2θω for a certain scalar θ(2c) and the scalar curvature r¯ of M satisfies r¯2c(n1)0, where S denotes the Ricci tensor of M. In the present paper, we prove that M is a real hypersurface of type (A) in Mn(c) mentioned Theorem O and Theorem MR. Our main theorem stated in section 6.

All manifolds in the present paper are assumed to be connected and of class C and the semi-invariant submanifolds are supposed to be orientable.

Let M˜ be a real 2(n+1)-dimensional Kaehlerian manifold with parallel almost complex structure J and a Riemannian metric tensor G. Let M be a real (2n-1)-dimensional Riemannian manifold isometrically immersed in M˜. We denote by g the Riemannian metric tensor on M from that of M˜.

We denote by ˜ the operator of covariant differentiation with respect to the metric tensor G on M˜ and by ∇ the one on M. Then the Gauss and Weingarten formulas are given respectively by

˜XY=XY+i=13g(A(i)X,Y)C(i),˜XC(i)=A(i)X+j=13lj (i)(X)C(j)

for any vector fields X and Y tangent to M and any vector field C(i) normal to M, where A(i) are called the second fundamental forms with respect to the normal vector C(i).

As is well-known, a submanifold of a Kaehlerian manifold is said to be a CR submanifold ([1], [25]) if it is endowed with a pair of mutually orthogonal and complementary differentiable distribution (T,T) such that for any point p ∈ M we have JTp=Tp, JTpTpM, where TpM denotes the normal space of M at p. In particular, M is said to be semi-invariant submanifold provided that dimT=1 ([4], [23]). Inthis case the unit vector field in JT is called a distinguished normal to the semi-invariant submanifold and denote by C ([4], [23]).

More precisely, we choose an orthonormal basis e1,,e2n2,ξ of Mp in such a way that e1,e2,,e2n2T, where Mp denotes the tangent space to M at each point p in M. Then we see that

G(Jξ,ei)=G(ξ,Jei)=0

for i=1,,2n2.

From now on we consider M is a real (2n-1)-dimensional semi-invariant submanifold of a Kaehlerian manifold M˜ of real dimension 2(n+1). Then we can write ([4], [24])

JX=ϕX+η(X)C,JC=ξ,JD=E,JE=D,

where we have put g(ϕX,Y)=G(JX,Y),η(X)=G(JX,C) for any vector fields X and Y tangent to M, and put C(1)=C, C(2)=D and C(3)=E.

By the Hermitian property of J, we see, using (2.2), that the aggregate (ϕ,ξ,η,g) is an almost contact metric structure on M, that is, we have

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

for any vectors X and Y on M.

We can also write the second equation of (2.1) as

˜XC=AX+l(X)D+m(X)E,˜XD=KXl(X)C+t(X)E,˜XE=LXm(X)Ct(X)D

because C, D and E are mutually orthogonal, where we have put

A(1)=A,A(2)=K,A(3)=L,l=l2(1)=l1(2),m=l3(1)=l1(3),t=l3(2)=l2(3),

In the sequel, we denote the normal components of ˜XC by C. The distinguished normal C is said to be parallel in the normal bundle if we have C=0, that is, l and m vanish identically.

From the Kaehler condition ˜J=0 and take account of the Gauss and Weingarten formulas,we obtain from (2.2)

Xξ=ϕAX,
(Xϕ)Y=η(Y)AXg(AX,Y)ξ,
KX=ϕLXm(X)ξ,KϕX=LXη(X)Lξ,
LX=ϕKX+l(X)ξ,LϕX=KX+η(X)Kξ

for any vectors X and Y on M. The last two relationships give

l(X)=g(Lξ,X),m(X)=g(Kξ,X),
m(ξ)=k,l(ξ)=TrA(3),

where, we have put k=TrA(2).

We notice here that there is no loss of generality such that we may assume TrA(3)=0. In fact, a normal vector v of M we denote by Av the second fundamental tensor of M in the direction of v. Then we have Av=Av. Hence there is a unit normal vector D' of M in the plane spanned by two vectors D and E such that TrAD=0, which proves our assertion. Therefore we have by (2.10)

l(ξ)=0.

Applying (2.8) by ϕ and using (2.7), we find

g(KX,Y)m(X)η(Y)=g(ϕKX,ϕY)η(X)l(ϕY).

If we take the skew-symmetric part of this with respect to X and Y, then we obtain

m(X)η(Y)+m(Y)η(X)=η(X)l(ϕY)η(Y)l(ϕX),

which together with (2.10) gives

l(ϕX)=m(X)+kη(X). 

Similarly we have

m(ϕX)=l(X)

because of (2.10).

Transforming (2.7) by L and using (2.8) and (2.9), we obtain

g(KLX,Y)+g(LKX,Y)=l(X)m(Y)l(Y)m(X).

In the rest of this paper we shall suppose that M˜ is a Kaehlerian manifold of constant holomorphic sectional curvature 4c, which is called a complex space form and denote by Mn+1(c). Then equations of the Gauss and Codazzi are given by

(X,Y)Z=c{g(Y,Z)Xg(X,Z)Y+g(ϕY,Z)ϕXg(ϕX,Z)ϕY2g(ϕX,Y)ϕZ}+g(AY,Z)AXg(AX,Z)AY+g(KY,Z)KXg(KX,Z)KY+g(LY,Z)LXg(LX,Z)LY,
(XA)Y(YA)Xl(X)KY+l(Y)KXm(X)LY+m(Y)LX=c{η(X)ϕYη(Y)ϕX2g(ϕX,Y)ξ},
(XK)Y(YK)X+l(X)AYl(Y)AXt(X)LY+t(Y)LX=0,
(XL)Y(YL)X+m(X)AYm(Y)AX        +t(X)KYt(Y)KX=0,

where R is the Riemann-Christoffel curvature tensor on M, and those of the Ricci by

(Xl)(Y)(Yl)(X)+g(KAX,Y)g(AKX,Y)+m(X)t(Y)m(Y)t(X)=0,
(Xm)(Y)(Ym)(X)+g(LAX,Y)g(ALX,Y)        +t(X)l(Y)t(Y)l(X)=0,
(Xt)(Y)(Yt)(X)+g(LKX,Y)g(KLX,Y)        +l(X)m(Y)l(Y)m(X)=2cg(ϕX,Y).

In what follows, to write our formulas in a convention form, we denote by α=η(Aξ),β=η(A2ξ),TrA=h,TrA(2)=k,Tr(tAA)=h(2) and for a function f we denote by f the gradient vector field of f.

Now, we put ξξ=U in the sequel. Then U is orthogonal to ξ because of (2.5). From now on we put

Aξ=αξ+μW,

where W is a unit vector field orthogonal to ξ. Then we have

U=μϕW

because of (2.5). So, W is orthogonal to U. Further, we have

μ2=βα2.

From (2.22) and (2.23) we have

ϕU=Aξ+αξ,

which together with (2.5) and (2.22) yields

g(Xξ,U)=μg(AW,X),μg(XW,ξ)=g(AU,X)

because W is orthogonal to ξ.

Differentiating (2.25) covariantly along M and using (2.5) and (2.6), we find

(XA)ξ=ϕXU+g(AU+α,X)ξAϕAX+αϕAX,

which enables us to obtain

(ξA)ξ=2AU+α2kLξ.

Because of (2.5), (2.26) and (2.27), we verify that

ξU=3ϕAU+αAξβξ+ϕα2k(Kξkξ).

In the next place, the Jacobi operators Rξ is given by

RξX=R(X,ξ)ξ=c(Xη(X)ξ)+αAXη(AX)Aξ+kKX      m(X)Kξl(X)Lξ,

where we have used (2.9), (2.10) and (2.15).

Suppose that Rξϕ=ϕRξ holds on M. Then from (2.30) we have

α(ϕAXAϕX)=g(Aξ,X)U+g(U,X)Aξ+2kLX      2k{l(X)ξ+η(X)Lξ},

where we have used (2.5), (2.8) and (2.12).

In this section we shall suppose that M is a semi-invariant submanifold of codimension 3 in a complex space form Mn+1(c), c0 and that the third fundamental form t satisfies

dt=2θω,ω(X,Y)=g(ϕX,Y)

for a certain scalar θ and any vector fields X and Y on M, where d denotes the exterior differential operator. Then (2.21) reformed as

g(LKX,Y)g(KLX,Y)+l(X)m(Y)l(Y)m(X)=2(θc)g(ϕX,Y),

or, using (2.14)

g(LKX,Y)+l(X)m(Y)=(θc)g(ϕX,Y),

which together with (2.9)~(2.11) implies that

KLξ=kLξ,LKξ=0.

Differentiating (3.1) covariantly along M and using (2.6) and the first Bianchi identity, we find

(Xθ)ω(Y,Z)+(Yθ)ω(Z,X)+(Zθ)ω(X,Y)=0,

which implies (n2)Xθ=0. Thus θ(c) is constant if n>2.

For the case where θ=c in (3.1) we have dt=2cω. In this case, the normal connection of M is said tobe L-flat([18]).

Lemma 3.1. Let M be a semi-invariant submanifold with L-flat normal connection in Mn+1(c), c0. If Aξ=αξ, then we have C=0 and A(2)=A(3)=0.

Proof. From (3.2) we have

Tr(tA(2)A(2))Kξ2+Lξ2=2(n1)(θc)

because of (2.7), (2.9) and (2.12), which implies

A(2)kηξ2+Lξ2=2(n1)(θc),

where F2=g(F,F) for any vector field F on M. Thus, by our hypothesis θ=c, we have A(2)=kηξ.

In the same way, we see from (2.8), (2.10), (2.13) and (3.2) that A(3)=0. And hence m(X)=kη(X) and l=0 because of (2.9). Therefore, it suffices to show that k=0. Using these facts, (2.19) reformed as

k{η(X)Aξg(Aξ,X)ξ}=k(η(X)tt(X)ξ),

which together with Aξ=αξ gives

k(tt(ξ)ξ)=0.

We also have from (2.18)

k{η(X)(AY+t(Y)ξ)η(Y)(AX+t(X)ξ)}=0,

which implies k(hα)=0. Form this and (3.4) we verify that k=0. This completes the proof.

Applying (3.2) by ϕ and taking account of (2.7) and (2.13), we find

K2X+η(X)K2ξ+l(X)Lξ=(θc)(Xη(X)ξ),

which implies η(X)K2ξg(K2ξ,X)ξ=0. Thus, it follows that

K2ξ=(Kξ2)ξ

by virtue of (2.9). Thus, (3.5) becomes

K2X+l(X)Lξ+Kξ2η(X)ξ=(θc)(Xη(X)ξ).

Putting X=Lξ in (2.8) and taking account of (2.12) and (3.3), we obtain

L2ξ=kKξ+(Kξ2+k2)ξ.

If we put X=Lξ in (3.2) and make use of (2.13) and (3.2), we find

(θcKξ2)Lξ=0.

Similarly, we verify, using (3.2) and (3.7), that

(θcLξ2k2)(Kξ2k2)=0.

Let Lξ0 at every point of M and suppose that this subset does not void. Then we have Kξ2=θc and Lξ2+k2=θc on the subset. Using these facts, we can verify that ( for detail, see (2.22) and (2.24) of [13])

k=2ALξ,
XLξ=t(X)KξAKXkAX

on the set. Differentiating (3.8) covariantly and taking the skew-symmetric part obtained, we find

(θ2c)(η(X)Kξm(X)ξ)=0,

where we have used (2.12), (2.16), (3.3) and (3.9), which shows that (θ2c)(m(X)+kη(X))=0 and hence θ=2c on this subset. Thus, from the first equation of (2.3) we have

Lemma 3.2. Let M be a semi-invariant submanifold of codimension 3 in Mn+1(c), c0 satisfying (3.1).

If θ2c0, then C=kξE on M.

In the following we assume that M satisfies (3.1) with θ2c0. Then we have

Lξ=0,Kξ=kξ

because of (2.9). It is, using (3.10), clear that (2.7), (2.8) and (3.2) are reduced respectively to

ϕLX=KXkη(X)ξ,
L=Kϕ,
g(LKX,Y)+(θc)g(ϕX,Y)=0.

From the last two equations, we obtain

L2X=(θc)(Xη(X)ξ).

Further, if we take account of (3.10), then the other structure equations (2.16)~(2.21) reformed as

(XA)Y(YA)X=k{η(Y)LXη(X)LY}+c{η(X)ϕYη(Y)ϕX2g(ϕX,Y)ξ},
(XK)Y(YK)X=t(X)LYt(Y)LX,
(XL)Y(YL)X=k{η(X)AYη(Y)AX}t(X)KY+t(Y)KX,
KAXAKX=k{η(X)tt(X)ξ},
LAXALX=(Xk)ξη(X)k+k(ϕAX+AϕX),

where we have used (2.5).

Putting X=ξ in (3.18) and using (3.10), we find

KAξ=kAξ+k(tt(ξ)ξ).

Replacing X by ξ in (3.19) and using (2.5), (3.10) and (3.12), we get

KU=(ξk)ξk+kU.

If we apply (3.20) by ϕ and make use of (2.22) (3.11) and (3.12), then we find

KU=k(tϕU),

which together with (3.21) yields

k=(ξk)ξ+k(tϕ+2U).

If we transform (3.19) by ϕ and take account of (2.22), (3.11) and the last equation, then we obtain

ϕALXKAX=k{tt(ξ)ξ}η(X)+2μη(X)W+2g(Aξ,X)ξAXϕAϕX},

which connected to (3.18) gives

ϕAL=LAϕ.

Since θ is constant if n>2, differentiating (3.14) covariantly, we find

(XL2)Y=(cθ){η(Y)ϕAX+g(ϕAX,Y)ξ},

or, using (3.13) and (3.17), it is verified that (see,[13])

2(XL)LY=(θc){2t(X)ϕYη(Y)(ϕA+Aϕ)X+g((AϕϕA)X,Y)ξ      η(X)(ϕAAϕ)Y}k{η(Y)(AL+LA)X      g((AL+LA)X,Y)ξη(X)(LAAL)Y},

which together with (3.10) and (3.22) yields

(θc)(AϕϕA)X+(k2+θc)(u(X)ξ+η(X)U)+k{(AL+LA)X+k{t(ϕX)ξ+η(X)ϕ°t}=0,

where u(X)=g(U,X) for any vector X.

In the following we consider the case where (2.22) with µ=0, that is Aξ=αξ. Differentiating this covariantly and using (2.5), we find

(XA)ξ=AϕAX+αϕAX+(Xα)ξ,

which together with (3.10) and (3.15) gives

2AϕAX+α(ϕA+Aϕ)X+2cϕX=η(X)α(Xα)ξ.

If we put X=ξ in this and using (2.22) with µ=0, then we find

α=(ξα)ξ.

Differentiating the second equation of (3.10) covariantly along M, and using (2.5), we find Xm=(Xk)ξ+kϕAX, from which taking the skew-symmetric part and making use of (2.20) with l=0,

LAXALXk(ϕAX+AϕX)=(Xk)ξη(X)k.

Since Aξ=αξ was assumed, we then have

k=(ξk)ξ

because of (3.10). From the last two equations, it follows that

LAAL=k(ϕA+Aϕ).

If we put X=ξ in (3.18) and remember (2.22) with µ=0 and (3.10), then we get

k(t(X)t(ξ)η(X))=0.

Since we have Aξ=αξ, differentiating (3.28) covariantly, and taking the skew-symmetric part obtained, we get

(ξk)(Aϕ+ϕA)=0.

From this and (3.27) we can write (3.26) as α(A2ϕ+cϕ)=0. By the properties of the almost contact metric structure, it follows that

ξk{h(2)α2+2c(n1)}=0,

which implies ξk=0 if c>0.

We will continue our arguments under the same hypotheses dt=2θω for a scalar θ(2c) as those stated in section 3. Further suppose, throughout this paper, that Rξϕ=ϕRξ, which means that the eigenspace of the structure Jacobi operator Rξ is invariant by the structure operator ϕ. Then (2.31) reformed as

α(ϕAXAϕX)=g(Aξ,X)U+g(U,X)Aξ+2kLX

by virtue of (3.10).

Transforming this by A, and taking the trace obtained, we have g(A2ξ,U)=0 because of (3.25), which together with (2.22) yields

μg(AW,U)=0.

Applying (4.1) by L and using (2.25), (3.11) and (3.19), we find

α{AKXkη(X)AξϕALX}+g(LU,X)Aξ+g(KU,X)U                  =2kL2X,

which together with (3.18) and (3.22) yields

kα{t(X)ξη(X)t+g(Aξ,X)ξη(X)Aξ}+g(LU,X)Aξg(Aξ,X)LUu(X)KU+g(KU,X)U=0,

where u(X)=g(U,X) for any vector X. If we take the inner product with ξ to this and use (3.10), then we get

kα{t(X)t(ξ)η(X)+g(Aξ,X)αη(X)}+αg(LU,X)=0.

Combining the last two equations and taking account of (2.24), we obtain

μ(w(X)LUg(LU,X)W)+u(X)KUg(KU,X)U=0,

where w(X)=g(W,X) for any vector X.

In the previous paper [13] we prove the following proposition.

Proposition 4.1. Let M be a real (2n-1)-dimensional(n>2) semi-invariant submanifold of codimension 3 in a complex space form Mn+1(c), c0. If it satisfies dt=2θω for a scalar θ2c and μ=g(Aξ,W)=0, then we have k=0.

Sketch of Proof. This fact was proved for c>0 (see, Proposition 3.5 of [13]). But, regardless of the sign of c this one is established. However, only ξ k=0 and ξα=0 should be newly certified. We are now going to prove, using (4.1), that ξ k=0.

Now, let Ω1 be a set of points such that ξk0 on M and suppose that Ω1 be nonvoid. Then we have

Aϕ+ϕA=0,LA=AL

on Ω1 because of (3.29) and (3.31). We discuss our arguments on Ω1.

From (4.1) we have αϕA+kL=0 because of µ=0, which together with (3.11) gives αAY+kKY=(α2+k2)η(Y)ξ. Differentiating this covariantly along Ω1 and using (3.27) and (3.28), we find

(Xα)AY+α(XA)Y+(ξk)η(X)KY+k(XK)Y    =2(α(ξα)+k(ξk))η(X)η(Y)+(α2+k2){g(ϕAX,Y)ξ+η(Y)ϕAX},

from which, taking the skew-symmetric part and making use of (3.16), we obtain

(Xα)AY(Yα)AX+α((XA)Y(YA)X)+k(t(X)LYt(Y)LX)        =(α2+k2)(η(Y)ϕAXη(X)ϕAY).

If we take the inner product ξ to this and remember (3.10), (3.15) and the fact that µ=0, then we have cα=0, which together with (4.1) yields kL=0, a contradiction because of (3.14). In the same way we see from (3.27) that ξα=0. This completes the proof.

We set Ω={pM:k(p)0}, and suppose that Ω is nonempty. In the rest of this paper, we discuss our arguments on the open subset Ω of M. So, by Proposition 4.1 we see that µ ≠0 on Ω.

We notice here that the following fact :

Remark 4.2. α0 on Ω.

In fact, if not, then we have α=0 on this subset. We discuss our arguments on such a place. So (4.1) reformed as

μ(w(X)U+u(X)W)+2kLX=0

because of (2.22) with α=0. Putting X=U or W in this we have respectively

LU=μβ2kW,LW=μ2kU

by virtue of (2.24) with α=0. Using this and (3.14), we can write (4.3) as

β22kw(X)W+g(KU,X)U=2k(θc)(Xη(X)ξ).

Taking the inner product with W to this, we obtain β2=4k2(θc).

On the other hand, combining (4.6) and (4.7) to (3.14) we also have β2=4(n1)k2(θc), which implies (n2)(θc)k=0, a contradiction because of our assumption and Lemma 3.1. Thus, α=0 is not impossible on Ω.

Now, putting X=U in (4.4) and remembering Remark 4.2, we find kt(U)+g(LU,U)=0.

By the way, replacing X by U in (4.1) and using (2.22) and (2.25), we find

α(ϕAU+μAW)=μ2Aξ+2kLU.

If we take the inner product with U and make use of (4.2) and Proposition 4.1, then we obtain g(LU,U)=0 and hence t(U)=0.

By putting X=U in (4.5), we then have

KU=τU,

where τ is given by τμ2=g(KU,U) by virtue of Proposition 4.1. Applying this by ϕ and using (3.12), we find

LU=τμW.

It is, using (4.8) and (4.9), seen that

τ2=θc.

because of (3.13).

Remark 4.3. Ω= if θ=c.

Since we have θ=c, then (3.14) gives L=0 and thus KX=kη(X)ξ by virtue of (3.11). Hence, (3.17) reformed as

k{η(X)AYη(Y)AX+η(X)t(Y)ξt(X)η(Y)ξ}=0,

which shows k(t(X)+g(Aξ,X)ση(X))=0, where we have put σ=α+t(ξ).

Thus, the last two equations imply

AX=η(X)Aξ+g(Aξ,X)ξαη(X)ξ.

Since U is orthogonal to ξ and W, it is clear that AU=0 and AW=μξ.

If we put X=µ W in (4.1) and remember (2.23) and the fact that L=0, then we obtain μ2U=0 and hence Aξ=αξ. Owing to Lemma 3.1, we conclude that k=0 and thus Ω=.

By Remark 4.3, we may only consider the case where τ0 on Ω. Because of (3.22) and (4.8) we have

t(ϕX)=(1+τk)g(U,X).

Therefore, by properties of the almost contact metric structure, it is clear that

t=t(ξ)ξμ(1+τk)W.

Using (2.22), we can write (3.20) as

μKW=kμW+k(tt(ξ)ξ),

which together with (4.12) implies that

KW=τW

because of Proposition 4.1.

If we take account of (3.25) and (4.11), then we find

τ2(AϕXϕAX)+τ(τk)(u(X)ξ+η(X)U)+k(ALX+LAX)=0.

From (2.15) the Ricci tensor S of type (1,1) of M is given by

SX=c{(2n+1)X3η(X)ξ}+hAXA2X+kKXK2XL2X

by virtue of (3.10).

By the way, we see, using (3.12)~(3.14), that

K2X=(θc)(Xη(X)ξ)+k2η(X)ξ.

Substituting this and (3.14) into the last equation and using (4.10), we obtain

SX={c(2n+1)2(θc)}X+(2(θc)k23c)η(X)ξ+hAXA2X+kKX,

which connected to (3.10) yields

Sξ=2c(n1)ξ+hAξA2ξ.

Differentiating (4.8) covariantly along Ω, we find

(XK)U+KXU=τXU,

which together with (3.16) and (4.9) yields

μτ(t(X)w(Y)t(Y)w(X))+g(KXU,Y)g(KYU,X)      =τ{g(XU,Y)g(YU,X)}.

By the way, because of (2.22) and (2.24), we can write (2.29) as

ξU=3ϕAU+αμWμ2ξ+ϕα.

Replacing X by ξ in (4.18) and taking account of the last two relationships, we find

μ2(τk)ξ+μτ(t(ξ)2α)W+μ(kτ)AW    +3(LAUτϕAU)=τϕαLα,

where we have used the first equation of (2.26).

In a direct consequence of (3.12) and (4.8), we obtain

μLW=τU

because of µ ≠ 0 on Ω.

In the same way as above, we see from (4.13)

τμ{t(X)u(Y)t(Y)u(X)}+g(KXW,Y)g(KYW,X)      =τ{g(YW,X)g(XW,Y)}.

In the next place, from (2.22) and (2.25) we have ϕU=μW. Differentiating this covariantly and using (2.6), we find

g(AU,X)ξϕXU=(Xμ)W+μXW.

Putting X=ξ in this and making use of (2.29), we get

μξW=3AUαU+α(ξα)ξ(ξμ)W,

which enables us to obtain

Wα=ξμ.

We will continue our arguments under the same hypotheses Rξϕ=ϕRξ and dt=2θω for a scalar θ(2c) as those in section 3. Further, we assume that Sξ=g(Sξ,ξ)ξ is satisfied on a semi-invariant submanifold of codimension 3 in Mn+1(c),c0. Then we have from (4.17)

A2ξ=hAξ+(βhα)ξ.

From this, and (2.22) and (2.24) we see that

AW=μξ+(hα)W.

In the next place, differentiating (5.2) covariantly along Ω, we find

(XA)W+AXW=(Xμ)ξ+μXξ+X(hα)W+(hα)XW.

By taking the inner product with W to this and using (2.26) and (5.2), we obtain

g((XA)W,W)=2g(AU,X)+XhXα

because W is a unit orthogonal vector to ξ.

Applying (5.3) by ξ and using (2.26), we also obtain

μg((XA)W,ξ)=(h2α)g(AU,X)+μ(Xμ),

which connected to (3.15) gives

μ(ξA)W=(h2α)AU+μμkμLWcU,

or, using (3.10), (3.15) and (5.5),

μ(WA)ξ=(h2α)AU2cU+μμ.

Putting X=ξ in (5.4) and taking account of (5.5), we have

Wμ=ξhξα.

Replacing X by ξ in (5.3) and using (5.6), we find

(h2α)AUkμLWcU+μμ+μ(AξW(hα)ξW)      =μ(ξμ)ξ+μ2U+μ(ξhξα)W.

Substituting (4.23) and (4.24) into this and making use of (4.21), we find

3A2U2hAU+(αhβckτ)U+Aα+12βhα      =2μ(Wα)ξ+(2αh)(ξα)ξ+μ(ξh)W.

On the other hand, if we put X=μW in (4.1) and take account of (2.23), (2.24) and (5.2), then we find αAU+(βhα+2kτ)U=0, which shows

AU=λU,

where the function λ is defined, using Remark 4.2, by

αλ=hαβ2kτ.

Differentiating (5.10) covariantly along Ω, we find

(XA)U+AXU=(Xλ)U+λXU.

If we take the skew-symmetric part of this, then we get

μ(kτc)(η(Y)w(X)η(X)w(Y))+g(AXU,Y)g(AYU,X)    =(Xλ)u(Y)(Yλ)u(X)+λ(g(XU,Y)g(YU,X)),

where we have used (2.22), (2.25), (3.15) and (4.9). Replacing X by U in this and using (5.10), we get

AUUλUU=(Uλ)Uμ2λ.

Taking the inner product with W to this and remembering (5.2), we obtain

μg(ξ,UU)+μ2(Wλ)+(hαλ)g(W,UU)=0.

By the way, from KU=τU, we have

(XK)U+KXU=τXU,

which implies that g((XK)U,U)=0. Because of (3.16), (4.9) and the last relationship give (UK)U=0, which connected to (4.13) and (5.14) yields g(W,UU)=0. Thus, (5.13) reformed as

μg(ξ,UU)+μ2(Wλ)=0.

However, the first term of this vanishes identically because of (2.26) and (5.2), which shows μ(Wλ)=0 and hence

Wλ=0.

In the same way, we verify, using (2.26) and (5.2), that

ξλ=0.

Now, differentiating (2.25) covariantly and using (2.5), we find

(XA)ξ+AϕAX=(Xα)ξ+αϕAX+(Xμ)W+μXW.

If we put X=µ W in this and use (5.2), (5.7) and (5.10), then we find

μ2WWμμ=(2hλ3αλ+α2αh2c)Uμ(Wα)ξμ(Wμ)W.

Lemma 5.1. If M satisfies (4.1), (5.2) and dt=2θω for a scalar θ(2c), then we have on Ω

k=(kτ)U.

Proof. Using (3.21) and (4.8) we have

Xk=(ξk)η(X)+(kτ)u(X)

for any vector field X. Differentiating this covariantly along Ω and taking the skew-symmetric part obtained, we find

η(Y)X(ξk)η(X)Y(ξk)+(ξk){η(X)u(Y)η(Y)u(X)      +g(ϕAX,Y)g(ϕAY,X)}+(kτ)du(X,Y)=0,

where we have used (2.5).

Now, we take an orthonormal frame filed {e0=ξ,e1=W,e2,,en1,en=ϕe1=1μU,en+1=ϕe2,,e2n2=ϕen1} of M. Taking the trace of (2.27), we obtain

i=0 2n2g(ϕeiU,ei)=ξαξh.

Putting X=ϕei and Y=ei in (5.19) and summing up for i=1,2,,n1, we have

(kτ) i=0 2n2du(ϕei,ei)=ξk(αh),

where we have used (2.22), (2.25), (5.2) and (5.10). Combining the last two relationships, we get

(hα)ξk=(kτ)(ξhξα). 

By the way, if we put X=µW in (3.25) and take account of (2.22), (3.10) and (5.2), we obtain

(θc){AU(hα)U}+kτ{AU+(hα)U}=0,

which connected to (4.9) and (5.10) yields

λ(k+τ)+(hα)(kτ)=0.

From this we have

(hα+λ)k+(kτ)(hα)+(k+τ)λ=0.

So we have (hα+λ)ξk+(kτ)(ξhξα)=0 with the aid of (5.16). From this and (5.20) we see that (2h2α+λ)ξk=0.

If ξk0 on Ω, then we have λ=2(αh), which together with (5.21) implies that (hα)(k+3τ)=0 on this subset. We discuss our arguments on such a place. So we have hα=0 from the last equation and hence λ=0. Consequently we have μ2+2kτ=0 by virtue of (2.24) and (5.11). Differentiation with respect to ξ gives μ(ξμ)+τ(ξk)=0.

However, if we take the inner product with U to (5.7) and remember (2.24), (5.10) and the fact that hα=0 and λ=0, then we have μμ=(μ2+kτ+c)U and consequently ξµ=0. Hence we have τ(ξk)=0, a contradiction. Thus, we have (5.18). This completes the proof.

Lemma 5.2. Under the same hypotheses as those stated in Lemma 5.1, we have kτ0 on Ω.

Proof. If not, then we have k-τ=0 on an open subset of Ω. We discuss our argument on such a place. Then we have λ=0 because of (5.21) and Remark 4.3. So (5.10) and (5.11) turn out respectively to

AU=0,
βhα+2τ2=0.

We also have from (4.11) t=t(ξ)ξ2ϕU, which shows t(Y)=t(ξ)η(Y)2g(ϕU,Y) for any vector Y. Differentiating this covariantly and using (2.5), (2.6) and (5.22), we find

(Xt)Y=X(t(ξ))η(Y)+t(ξ)g(ϕAX,Y)2g(ϕXU,Y),

from which, taking the skew-symmetric part with respect to X and Y and using (3.1),

2θg(ϕX,Y)=X(t(ξ))η(Y)Y(t(ξ))η(X)+t(ξ){g(ϕAX,Y)g(ϕAY,X)}    +2{g(ϕYU,X)g(ϕXU,Y)}.

On the other hand, we verify from (2.27) that

g(ϕXU,Y)g(ϕYU,X)+(Xα)η(Y)(Yα)η(X)  =2cg(ϕX,Y)2g(AϕAX,Y)+α(g(ϕAX,Y)g(ϕAY,X)).

Combining the last two equations, it follows that

2(θ2c)g(ϕX,Y)+t(ξ){g(ϕAX,Y)g(ϕAY,X)}=X(t(ξ))η(Y)Y(t(ξ))η(X)+2{2g(AϕAX,Y)+α(g(ϕAX,Y)g(ϕAY,X))+(Xα)η(Y)(Yα)η(X)}.

Putting Y=ξ in this and remembering (5.22), we find

X(t(ξ))+2(Xα)={ξ(t(ξ))+2ξα}η(X)+(t(ξ)+2α)u(X).

Substituting this into the last equation, we obtain

2(θ2c)g(ϕX,Y)=(t(ξ)+2α)(u(X)η(Y)u(Y)η(X)      +g(ϕAX,Y)g(ϕAY,X))+4g(AϕAX,Y).

If we put X=μW in this and take account of (2.23), (5.2) and (5.22), then we get

2(θ2c)=(t(ξ)+2α)(hα).

In the next step, differentiating (4.13) covariantly, we find

(XK)W+KXW+τXW=0,

from which, taking the skew-symmetric part and using (3.16) and (4.9),

τμ(t(Y)u(X)t(X)u(Y))+g(KXW,Y)g(KYW,X)      =τ{(YW)X(XW)Y}.

If we put X=ξ in this and make use of (2.26), (4.23) and (5.22), then, we find

Kα+τα=2τ(ξα)ξ+τ(2α+t(ξ))U.

Replacing X by W in (5.26) and making use of (5.17), we have

μ(Kμ+τμ)=2τ(μ2α2+hα+2c)U+2μτ(Wα)ξ.

If we take the inner product with U to this and take account of (4.8), then we obtain μ(Uμ)=(μ2α2+hα+2c)μ2, which together with (2.24) and (5.23) gives

μ(Uμ)=2(μ2+τ2+c)μ2.

On the other hand, differentiating (5.22) covariantly with respect to ξ, we find (ξA)U+AξU=0, which together with (4.19) (5.1) and (5.22) implies that

(ξA)U+(αhβ)Aξα(βhα)ξ+Aϕα=0.

Applying by ϕ, we have

ϕ(ξA)U+(αhβ)U+ϕAϕα=0.

Since we see from (3.15)

(UA)ξ(ξA)U=μ(τ2+c)W

by virtue of (2.25), (3.10) and (4.9), it follows that

ϕ(UA)ξ=ϕ(ξA)U+(τ2+c)U.

We also have from (2.27)

XU+g(A2ξ,X)ξ=ϕ(XA)ξ+ϕAϕAX+αAX,

which connected to (5.22) gives UU=ϕ(UA)ξ. Thus, (5.30) reformed as

UU=ϕ(ξA)U+(τ2+c)U.

Combining this to (5.29) and using (5.23), it follows that

UU=(cτ2)UϕAϕα.

If we apply by A and take account of (5.12) with λ=0 and (5.22), then we have AϕAϕα=0.

Now, taking the inner product with U to (5.30) and making use of (2.22) (2.25) and (5.2), we obtain

μ(Uμ)=(cτ2)μ2+(hα)Uα.

However, applying (5.27) by U and using (4.8), we find 2Uα=(t(ξ)+2α)μ2, which connected to (5.25) gives (hα)Uα=(θ2c)μ2. Substituting (5.28) and this into (5.32), we find 2μ2+3c+3τ2=θ, which together with (4.10) gives μ2+τ2+c=0 and consequently µ is constant. Thus, we see, using (2.24) and (5.23), that

α(hα)=τ2c.

Therefore, α(hα)=const. Differentiation gives

(hα)α+α(hα)=0,

which connected to (5.8) implies that (hα)ξα=0, where we have used μ=const. Accordingly we have ξα=0 by virtue of (5.33) and the fact that θ2c0.

Using (4.10) and (5.33), we can write (5.25) as

2(θ2c)α=(θ2c)(t(ξ)+2α).

Thus, it follows that t(ξ)=0 provided that θ2c0. Hence, (5.24) turns out to be α=αU, which implies du=0. Therefore, it is clear that UU=0 because of μ=const, which connected to (5.31) yields (cτ2)U=αϕAϕU. So we have cτ2=α(hα), where we have used (2.23), (2.25) and (5.2). From this and (5.33) it follows that θ2c=0, a contradiction. Hence, Lemma 5.2 is proved.

Lemma 5.3. Under the same hypotheses as those in Lemma 5.1, we have

α=(h3λ)U.

Proof. Because of Lemma 5.1 and Lemma 5.2, we can write (5.19) as du(X,Y)=0, that is, g(XU,Y)g(YU,X)=0. Putting X=ξ in this, and using (2.26) and (4.19), we find

3ϕAU+αAξβξ+ϕα+μAW=0,

which together with (2.22), (2.25), (5.2) and (5.10) implies that

ϕα+(h3λ)μW=0.

Thus, it follows that

α=(ξα)ξ+(h3λ)U.

We are now going to prove that ξα=0.

Differentiation (5.21) with respect to ξ gives ξhξα=0 with the aid of (5.16), Lemma 5.1 and Lemma 5.2.

Using (5.10), (5.35) and this fact, we can write (5.9) as

12β+(2hλ+αhβckτh2)U={2μ(Wα)+α(ξα)}ξ.

Since we have Wµ=0 because of (5.8), if we take the inner product ξ to the last equation and take account of (2.24), then we obtain α(Wα)=0 and hence Wα=0 by virtue of Remark 4.2.

Differentiating (5.11) with respect to ξ and making use of (5.16), Lemma 5.1 and the fact that ξhξα=0, we find

ξβ=(h+αλ)ξα.

On the other hand, if we differentiate (2.24) with respect to ξ and remember Wα=0 and (4.24), then we have ξβ=2α(ξα). From this and the last relationship we get (λ+αh)ξα=0.

Now, if ξα0 on Ω, the we have λ=hα on this subset. We discuss our arguments on this subset. Then (5.21) yields λk=0 and hence λ=0 and hα=0. So (5.35) and (5.36) are reduced respectively to

α=(ξα)ξ+αU,12β=α(ξα)ξ+(β+kτ+c)U.

We also have from (5.11) β=α22kτ, which together with (5.18) implies that

12β=αατ(kτ)U.

Combining above equations, it follows that τ2=c, that is, θ2c=0, a contradiction. This completes the proof of Lemma 5.3.

First of all, we will prove the following lemma.

Lemma 6.1. Let M be a real (2n-1)-dimensional semi-invariant submanifold of codimension 3 in a complex space form Mn+1(c), c≠0 satisfying dt=2θω for a scalar θ2c. Suppose that M satisfies Rξϕ=ϕRξ and at the same time Sξ=g(Sξ,ξ)ξ. Then the distinguished normal is parallel in the normal bundle, where S denotes the Ricci tensor of M.

Proof. Because of (5.19), Lemma 5.1 and Lemma 5.2, we have du=0. So we have from (5.14)

g(KXU,Y)g(KYU,X)+μτ{t(X)w(Y)t(Y)w(X)}=0,

where we have used (3.16) and (4.9). Putting X=ξ in this and using (2.25), (2.26), (4.19) and (5.10), we find

K(3λμW+αAξβξ+ϕα)+kμAW+μτt(ξ)W=0,

which connected to (2.22), (3.10), (3.12), (4.13), (5.2) and (5.34) gives

τt(ξ)+(hα)(k+τ)=0,

or, using (5.21)

τ(kτ)t(ξ)=λ(k+τ)2.

On the other hand, differentiating (4.12) covariantly along Ω, and taking account of (2.5), (2.6), (5.10) and (5.18), we get

(Xt)Y=X(t(ξ))η(Y)+t(ξ)g(ϕAX,Y)+τk2(kτ)μu(X)w(Y)          (1+τk){λu(X)η(Y)g(ϕXU,Y)+t(XY),

from which taking the skew-symmetric part and using (2.25) and (3.1),

2θg(ϕX,Y)+τk2(kτ)μ(u(Y)w(X)u(X)w(Y))  =X(t(ξ))η(Y)Y(t(ξ))η(X)+t(ξ){g(ϕAX,Y)g(ϕAY,X)}  (1+τk){λ(u(X)η(Y)u(Y)η(X))g(ϕXU,Y)+g(ϕYU,X)}.

By the way, we have from (2.27) and (3.15)

g(ϕXU,Y)g(ϕYU,X)+(h+λ3α)(u(X)η(Y)u(Y)η(X))  =2cg(ϕX,Y)2g(AϕAX,Y)+α(g(ϕAX,Y)g(ϕAY,X)),

where we have used (3.10), (5.10) and (5.34).

Combining the last two equations, we obtain

2θg(ϕX,Y)+τk2(kτ)μ(u(Y)w(X)u(X)w(Y))t(ξ)(g(ϕAX,Y)g(ϕAY,X))    =X(t(ξ))η(Y)Y(t(ξ))η(X)+(1+τk){2cg(ϕX,Y)+(h3λ)(u(X)η(Y)        u(Y)η(X))2g(AϕAX,Y)+α(g(ϕAX,Y)g(ϕAY,X))}.

Putting Y=ξ in this and making use of (2.5) and (5.10), we find

X(t(ξ))=ξ(t(ξ))η(X)+{t(ξ)+(1+τk)(λ+αh)}u(X),

which together with (6.1) yields

X(t(ξ))=ξ(t(ξ))η(X)+(1+τk)(λ+t(ξ))u(X).

Substituting this into the last equation and using (5.21), we find

2θg(ϕX,Y)+τk2μ(kτ)(w(X)u(Y)w(Y)u(X))=(1+τk){(h-2λ+t(ξ))(u(X)η(Y)-u(Y)η(X))+2cg(ϕX,Y)+2g(AϕAX,Y)+(h+t(ξ))(g(ϕAX,Y)-g(ϕAY,X))}.

Differentiating (6.1) covariantly and remembering (5.18), we find

τX(t(ξ))=(αh)(kτ)u(X)+(k+τ)(XαXh),

which connected to (5.21) yields

τX(t(ξ))=(k+τ)(XαXh+λu(X)).

By the way, we see, using (5.20), Lemma 5.1 and Lemma 5.2, that ξhξα=0. Thus, from the last equation, it follows that ξ(t(ξ))=0 and hence (6.4) can be written as

X(t(ξ))={t(ξ)+(1+τk)(λh+α)}u(X),

which together with (6.1) gives

τX(t(ξ))={(k+2τ+τ2k)(αh)+τλ(1+τk)}u(X).

Combining this to (6.6), we get

(k+τ)(αh+λU)=(1+τk){(k+τ)(αh)+τλ}U,

which together with (5.21) gives

k(αh)=2τ(λ+αh)U,

where we have used k+τ0.

If we differentiate (6.2) and take account of Lemma 5.1 and itself, we find

λ(k+τ)2U+τ(kτ)t(ξ)=(k+τ)2λ+2λ(k2τ2)U,

which together with (6.6), and Lemma 5.1 and Lemma 5.2 implies that (k+τ)λ=(kτ)(αh)+2τλU, or using (5.21) and (6.7),

(k+τ)λ=6τλU.

Now, if we put X=U and Y=W in (6.5) and using (2.23), (5.2) and (5.10), then we find

2θ+τk2(kτ)μ2=(1+τk){2c2λ(hα)+(t(ξ)+h)(λ+hα)}.

By the way, it is seen, using (5.11) and (5.21), that (kτ)2μ2+2k(αλ+τkτ2)=0. Thus, the last equation can be written as

θk(kτ)ταλ(kτ)τ2(kτ)2    =c(k2τ2)+λ2(k+τ)2τλ(k+τ)(t(ξ)+h).

If we multiply k-τ to this and take account of (4.10), (5.21) and (6.2), then we obtain

λ2(k+τ)2+2ταλ(kτ)+(kτ)2(τ2c)=0.

Differentiating this covariantly and using (5.18) and (6.8), we find

τ(kτ)(αλ)+6τλ2(k+τ)U=τλ{2λ(k+τ)+α(kτ)}U,

which implies

(kτ)(αλ)=λ{α(kτ)4λ(k+τ)}U.

From this and (5.21) and (5.34), we have

α(kτ)λ+6τλ2U=0,

which together with (6.8) yields λ{α(kτ)+λ(k+τ)}=0. Thus, it follows that α(kτ)+λ(k+τ)=0 by virtue of (6.9), which connected to (5.21) gives h=2α. Further, we have from the last relationship (k+τ)λ+(kτ)α=0, which together with (5.34) and (6.8) gives 6τλ+(kτ)(2α3λ)=0. Thus, it follows that (8τ5k)λ=0, and hence 5k=8τ because of (6.9).

So, we see, using (5.18), that k is a constant on Ω and hence U=0, a contradiction. This completes the proof.

According to Lemma 6.1 we can prove the following :

Lemma 6.2. Under the same hypotheses as those in Lemma 6.1, we have A(2)=A(3)=0 provided that r¯2(n1)c0.

Remark 6.3. This lemma proved in [13] for the case where θ2c<0 and c>0. But, we need the condition r¯2c(n1)0 for the case where c<0, where r¯ is the scalar curvature of M. So we introduce the outline of the proof.

The sketch of Proof. By Lemma 2.2 and Lemma 6.1, we have k=0 and hence m=0 on M because of (3.10). Thus, (3.15)~(3.20) turn out to be

(XA)Y(YA)X=c{η(X)ϕYη(Y)ϕX2g(ϕX,Y)ξ},
(XK)Y(YK)X=t(X)LYt(Y)LX,
(XL)Y(YL)X=0,
KAAK=0,LAAL=0,

Since we have Kξ=0 because of (3.10), differentiating Kξ=0 covariantly along M and using (2.5) and (3.12), we find

(XK)ξ=LAX.

If we take account of Lemma 5.2 and (4.10), then (4.15) reformed as

K2X=τ(Xη(X)ξ),

where τ=θc.

Differentiating (6.15) covariantly along M and using (2.5), we find

(XK)KY+K(XK)Y=τ{η(Y)ϕAX+g(ϕAX,Y)ξ}.

Using the quite same method as those used to (3.26) from (3.14), we can derive from the last equation the following :

2(XK)KY=τ{2t(X)ϕY+η(X)(ϕAAϕ)Y    +g((ϕAAϕ)X,Y)ξ+η(Y)(ϕA+Aϕ)X},

where we have used (3.13) and (6.11).

By the way, if we take the trace of K in (6.11), we have ie iKei=Lt because of (3.10). If we use this fact to (6.16), we obtain

KLt=τ(ϕt+U),

where we have used (2.5), which together with (3.11) gives τU=0 and consequently U=0 on M, that is

Aξ=αξ because of (2.25). Therefore, if we take account of Lemma 5.3 and (3.26), then we obtain

τ(AϕϕA)=0.

In the following, we assume that τ0 on M. Then, from this and (6.10) we can verify the following (cf. [6], [16]) :

A2=αA+c(Iηξ),
(XA)Y=c(η(Y)ϕX+g(ϕX,Y)ξ).

Using (6.17), we can write (6.16) as

K(XK)Y=τ{t(X)ϕY+η(X)ϕAY+g(ϕAX,Y)ξ}.

If we transform this by K and make use of (3.12), (6.11), (6.14) and (6.15), then we have

(XK)Y=t(X)LYη(X)ALXη(Y)LAXg(ALX,Y)ξ.

Differentiating (3.12) covariantly along M and using (2.6) and the last equation, we find

(XL)Y=t(X)KY+η(X)AKY+η(Y)AKX+g(AKX,Y)ξ.

If we take the trace of L in this and remember (3.20) and the fact that TrA(2)=TrA(3)=0 and Aξ=αξ, we verify that

Tr(AA(2))=0,

which connected to (6.18) gives

Tr(A2A(2))=0.

For the orthonormal frame field {e0,e1,,e2n2} already selected, we write g(ej,ei)=gji,g(ϕei,ej)=ϕij,(gji)1=gji,g(Aei,ej)=Aij and eiX=(iXh)eh for any vector X=Xiei. And the Einstein summation convention will be used. Then (6.20) can be written as

kKji=tkLjiξkAjrLirξiAkrLjrξjAirLkr.

Differentiating this covariantly along M and taking account of (2.5), (3.20), (6.18), (6.19) and itself, we find

hkKji=(htk)Ljic(Kjhξkξi+Kkiξjξh+2Kihξjξk)+Bhkji    α(ξjξhAkrKir+ξkξiAjrKhr+2ξjξkAirKhr)    +(Ahsϕjs)(AkrLir)+(Ahsϕks)(AirLjr)+(Ahsϕis)(AjrLkr),

where Bhkji is a certain tensor with Bhkji=Bkhji, from which, taking the skew-symmetric part with respect to h and k, and making use of (3.1), (6.17) and the Ricci identity for Kji (for detail, see (4.17) of [13]),

RkhjrKir+RkhirKjr=2θϕhkLjic{ξj(ξkKihξhKik)+ξi(ξkKjhξhKjk)}α{ξj(ξkAirKhrξhAirKkr)+ξi(ξkAjrKhrξhAjrKkr)}+(Ahsϕjs)(AkrLir)(Aksϕjs)(AhrLir)+(Ahsϕis)(AkrLjr)(Aksϕis)(AhrLjr)+2(Ahsϕks)(AjrLir).

Multiplying (6.24) with ϕkh and summing for k and h, and using (3.1), (3.11), (3.12), (6.17) and (6.18), we find

ϕkh(RkhjrKir+RkhirKjr)=4{c(n1)θ}Lji+2(h+α)AjrLir.

On the other hand, from (2.15) we see, using (3.12), (6.15), (6.17) and (6.18), that

ϕkh(RkhirKjr+RkhjrKir)=4{2θ(2n+3)c}Lji4αAjrLir,

which connected to (6.25) implies that (for detail, see (4.19) of [13])

(h+3α)AL=2{(n+1)θ2(n+2)c}L,

which connected to (3.14) yields

(h+3α)(AXαη(X)ξ)=2{(n+1)θ2(n+2)c}(Xη(X)ξ).

Taking the trace of (6.26), we have

(h+3α)(hα)=4(n1){(n+1)θ2c(n+2)},

which implies

(hα)2+4α(hα)=δ,

where we put

δ=4(n1){(n+1)θ2c(n+2)}.

In the same way as above, by using properties of A and (2.15), (6.22), (6.23) and (6.25), we obtain (for detail, see (4.21) of [13])

(4θ12ch(2)3α2)AK={4cα(θ2c)(hα)}K,

which connected to (6.15) yields

(4θ12ch(2)3α2)(hα)=2(n1){4cα(θ2c)(hα)}.

Since we have h(2)=αh+2c(n1) from (6.18), combining (6.27) to (6.28), we obtain

(θ3c)(hα)=2(n1)α(θ2c).

On the other hand, from (4.16) we verify that the scalar curvature r¯ of M is given by

r¯=4c(n21)4(n1)τ+h2h(2),

which connected to (6.18) gives

r¯=2c(n1)(2n+1)4(n1)τ+h(hα).

By the way, it is seen, using (4.10), that θ3c0 for c<0. We also have θ3c0 for c>0,

In fact, if not, then we have θ3c=0 on this open subset. Thus, it follows, using (6.29), that

α=0,τ=2c.

Hence h2=4(n1)2c on the set by virtue of (6.26) and (6.27). Using this fact and (6.31), we can write (6.30) as r¯=2c(n1)(4n5), a contradiction because of r¯2c(n1)0 and c>0. Therefore θ3c0 is proved.

Thus, we can write (6.29) as

hα=2(n1)θ3c(θ2c)α.

Substituting this into (6.26), we obtain

4(n1)(θ2c){(n+1)θ2(n+2)c}α2=δ(θ3c)2,

which together (6.27) gives

δ{(θ3c)2(θ2c)α2}=0.

We notice here that δ0 if c<0. We also see that δ0 for c>0. In fact, if not, then we have δ=0. Then we have by (6.27)

θc=n+3n+1c.

Using this fact and (6.26), we can write (6.30) as

r¯2(n1)c=4(n1)n+1(n23)c+ε2,

where ε2=0 or 12α2, a contradiction because c>0 and r¯2(n1)c0 was assumed. Therefore (6.32) turns out to be

(θ3c)2=(θ2c)α2.

Accordingly, if we combine (6.29) to (6.33), then we obtain α(hα)=2(n1)(θ3c), which together with (6.26) yields

h(hα)=2(n1)(2n1)τ4n(n1)c.

Using this, we can write (6.30) as

r¯2c(n1)=2(n1)(2n3)τ.

Therefore we have τ=0 if r¯2c(n1)0. This completes the proof of Lemma 6.2.

Let N0(p)={vTp(M):Av=0} and H0(p) be the maximal J-invariant subspace of N0(p). As a consequence of Lemma 6.2, we have A(2)=A(3)=0, the orthogonal complement of H0(p) is invariant under parallel translation with respect to the normal connection because of C=0. Thus, by the reduction theorem for Pn+1([19]) and Hn+1([9], [11]) there exists a totally geodesic complex space form including M in Mn+1(c), we conclude that

Theorem 6.4. Let M be a real (2n-1)-dimensional (n>2) semi-invariant submanifold of codimension 3 in a complex space form Mn+1(c), c0 with constant holomorphic sectional curvature 4c such that the third fundamental form t satisfies dt=2θω for a nonzero scalar θ2c0 and r¯2c(n1)0, where ω(X,Y)=g(ϕX,Y) for any vector fields X and Y on M. If M satisfies Rξϕ=ϕRξ and at the same time Sξ=g(Sξ,ξ)ξ, then M is a real hypersurface in a complex space form Mn(c), c0.

Since we have C=0, we can write (2.16) and (4.1) as

(XA)Y(YA)X=c{η(X)ϕYη(Y)ϕX2g(ϕX,Y)ξ},  α(ϕAXAϕX)g(Aξ,X)Ug(U,X)Aξ=0

respectively. Making use of (2.5), (2.6) and the above equations, it is prove in [16] that g(U,U)=0, that is, M is a Hopf real hypersurface. Hence, we conclude that α(AϕϕA)=0 and hence Aξ=0 or Aϕ=ϕA. Since M is a Hopf hypersurface, Aξ=0 means that α=0. Here, we note that the case α=0 correspond to the case of tube of radius π/4 in Pn([5],[6]). But, in the case Hn it is known that α never vanishes for Hopf hypersurfaces (cf.[3]) Thus, owing to Theorem 6.4, Theorem O and Theorem MR, we have

Main Theorem. Let M be a real (2n-1)-dimensional (n>2) semi-invariant submanifold of codimension $3$ in a complex space form Mn+1(c), c0 with constant holomorphic sectional curvature $4c$ such that the Ricci tensor $S$ satisfies Sξ=g(Sξ,ξ)ξ and the third fundamental form t satisfies dt=2θω for a scalar θ2c(0) and satisfies r¯2c(n1)0, where S and r¯ denote the Ricci tensor and the scalar curvature of M, respectively. Then Rξϕ=ϕRξ holds on M if and only if M is locally congruent to one of the following hypersurfaces :

(I) in case that Mn(c)=Pn,

  • (A1) a geodesic hypersphere of radius $r$, where 0<r<π/2 and rπ/4,

  • (A2) a tube of radius $r$ over a totally geodesic Pk for some k{1,...,n2}, where 0<r<π/2 and rπ/4,

  • (T) a tube of radius π/4 over a certain complex submanifold in Pn;

(II) in case that Mn(c)=Hn,

  • (A0) a horosphere,

  • (A1) a geodesic hypersphere or a tube over a complex hyperbolic hyperplane Hn1,

  • (A2) a tube over a totally geodesic Hk for some k{1,...,n2}.

Remark 6.5. Because of (4.10), it is clear that θ0 if c>0, and θ2c0 if c<0.

  1. A. Bejancu, CR-submanifolds of a Kähler manifold I, Proc. Amer. Math. Soc., 69(1978), 135-142.
    CrossRef
  2. J. Berndt, Real hypersurfaces with constant principal curvatures in a complex hyperbolic space, J. Reine Angew. Math., 395(1989), 132-141.
    CrossRef
  3. J. Berndt and H. Tamaru, Cohomogeneity one actions on non compact symmetric space of rank one, Trans. Amer. Math. Soc., 359(2007), 3425-3438.
    CrossRef
  4. D. E. Blair, G. D. Ludden and K. Yano, Semi-invariant immersion, Kodai Math. Sem. Rep., 27(1976), 313-319.
    CrossRef
  5. T. E. Cecil and P. J. Ryan, Focal sets and real hypersurfaces in complex projective space, Trans. Amer. Math. Soc., 269(1982), 481-499.
    CrossRef
  6. T. E. Cecil and P. J. Ryan. Geometry of Hypersurfaces. Springer; 2015.
    CrossRef
  7. J. T. Cho and U-H. Ki, Real hypersurfaces of a complex projective space in terms of the Jacobi operators, Acta Math. Hungar, 80(1998), 155-167.
    CrossRef
  8. J. T. Cho and U-H. Ki, Real hypersurfaces in complex space forms with Reeb-flows symmetric Jacobi operator, Canadian Math. Bull., 51(2008), 359-371.
    CrossRef
  9. J. Erbacher, Reduction of the codimension of an isometric immersion, J. Differential Geom., 5(1971), 333-340.
    CrossRef
  10. J. I. Her, U-H. Ki and S. -B. Lee, Semi-invariant submanifolds of codimension 3 of a complex projective space in terms of the Jacobi operator, Bull. Korean Math. Soc., 42(2005), 93-119.
    CrossRef
  11. S. Kawamoto, Codimension reduction for real submanifolds of a complex hyperbolic space, Rev. Mat. Univ. Complut. Madrid., 7(1994), 119-128.
    CrossRef
  12. U-H. Ki and H. Song, Jacobi operators on a semi-invariant submanifolds of codimension 3 in a complex projective space, Nihonkai Math J., 14(2003), 1-16.
  13. U-H. Ki, H. Song and R. Takagi, Submanifolds of codimension 3 admitting almost contact metric structure in a complex projective space, Nihonkai Math J., 11(2000), 57-86.
  14. M. Kimura, Real hypersurfaces and complex submanifolds in complex projective space, Trans. Amer. Math. Soc., 296(1986), 137-149.
    CrossRef
  15. S. Montiel and A. Romero, On some real hypersurfaces of a complex hyperbolic space, Geometriae Dedicata., 20(1986), 245-261.
    CrossRef
  16. R. Niebergall and P. J. Ryan. Real hypersurfaces in complex space form, in Tight and Taut submanifolds. In: T. E. Cecil and S. -S. Chern, eds. Cambridge University Press; 1998:233-305.
  17. M. Okumura, On some real hypersurfaces of a complex projective space, Trans. Amer. Math. Soc., 212(1973), 355-364.
    CrossRef
  18. M. Okumura, Normal curvature and real submanifold of the complex projective space, Geom. Dedicata., 7(1978), 509-517.
    CrossRef
  19. M. Okumura, Codimension reduction problem for real submanifolds of complex projective space, Colloz. Math. Soc. Jámos Bolyai, 56(1989), 574-585.
  20. H. Song, Some differential-geometric properties of R-spaces, Tsukuba J. Math., 25(2001), 279-298.
    CrossRef
  21. R. Takagi, On homogeneous real hypersurfaces in a complex projective space, Osaka J. Math., 19(1973), 495-506.
  22. R. Takagi, Real hypersurfaces in a complex projective space with constant principal curvatures I, II, J. Math. Soc. Japan, 27(1975), 507-516, 43-53 pp.
    CrossRef
  23. Y. Tashiro, Relations between the theory of almost complex spaces and that of almost contact spaces (in Japanese), Stildeugaku, 16(1964), 34-61.
  24. K. Yano and U-H. Ki, On (f, g, u, v, w,λ, µ, ν)-structure satisfying λ222 =1,, Kodai Math. Sem. Rep., 56(1978), 574-585.
  25. K. Yano and M. Kon. CR submanifolds of Kaehlerian and Sasakian manifolds. Birkhäuser; 1983.
    CrossRef