Next Article in Journal
Comparative Study on the Influence of Different Forms of New Tubular Roof Method Construction on Railway Tracks
Next Article in Special Issue
Classification of Lorentzian Lie Groups Based on Codazzi Tensors Associated with Yano Connections
Previous Article in Journal
Solving Quaternion Linear System Based on Semi-Tensor Product of Quaternion Matrices
Previous Article in Special Issue
Singularities of Slant Focal Surfaces along Lightlike Locus on Mixed Type Surfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

V-Quasi-Bi-Slant Riemannian Maps

1
Shri Jai Narain Post Graduate College, University of Lucknow (U.P.), Lucknow 226001, India
2
Department of Mathematical Sciences, Faculty of Applied Sciences, Umm Al Qura University, Makkah 21955, Saudi Arabia
3
Department of Mathematics and Astronomy, University of Lucknow (U.P.), Lucknow 226007, India
4
Department of Mathematics, College of Science, Jazan University, Jazan 45142, Saudi Arabia
5
School of Statistics, Jilin University of Finance an Economics, Changchun 130117, China
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(7), 1360; https://doi.org/10.3390/sym14071360
Submission received: 18 May 2022 / Revised: 9 June 2022 / Accepted: 22 June 2022 / Published: 1 July 2022
(This article belongs to the Special Issue Symmetry and Its Application in Differential Geometry and Topology)

Abstract

:
In this work, we define a v-quasi-bi-slant Riemannian map (in brief, v-QBSR map) from almost Hermitian manifolds to Riemannian manifolds. This notion generalizes both a v-hemi slant Riemannian map and a v-semi slant Riemannian map. The geometry of leaves of distributions that are associated with the definition of such maps is studied. The conditions for v-QBSR maps to be integrable and totally geodesic are also obtained in the paper. Finally, we provide the examples of v-QBSR maps.

1. Introduction

Let π be a C -map from a Riemannian manifold ( N 1 , g 1 ) to another Riemannian manifold ( N 2 , g 2 ) . The notion of a Riemannian map between two Riemannian manifolds was introduced by Fischer [1], and generalizes both the notion of an isometric immersion and a Riemannian submersion. The theory of Riemannian submersion is discussed in [2]. In [1], Fischer also defined the concept of a Riemannian map in the following way: Let π : ( N 1 , g 1 ) ( N 2 , g 2 ) be a differentiable map between Riemannian manifolds such that 0 < r a n k π < min { m , n } ( dim N 1 = m , dim N 2 = n ) . If we denote the k e r n e l space of π by ker π and the orthogonal complementary space of ker π by ( ker π ) in T N 1 , then T N 1 has the following orthogonal decomposition:
T N 1 = ( ker π ) ker π .
Here, if the r a n g e of π is denoted by r a n g e π and for a point q N 1 , the orthogonal complementary space of r a n g e π π ( q ) by ( r a n g e π π ( q ) ) in T π ( q ) N 2 . Then, the tangent space T π ( q ) N 2 has the following orthogonal decomposition:
T π ( q ) N 2 = ( r a n g e π π ( q ) ) ( r a n g e π π ( q ) ) .
A differentiable map π : ( N 1 , g 1 ) ( N 2 , g 2 ) is named a Riemannian map at q N 1 if the horizontal restriction π q h : ( ker π q ) ( r a n g e π π ( q ) ) is a linear isometry between the inner product space ( ( ker π q ) , ( g 1 ) ( q ) | ( ker π q ) ) and ( r a n g e π π ( q ) , ( g 2 ) ( π ( q ) ) | ( r a n g e π q ) ) . We see that a peculiar property of a Riemannian map is that it satisfies the generalized eikonal equation 2 e ( π ) = | | π | | 2 = r a n k π . This equation works as a bridge between physical and geometrical optics [1]. By using Cauchy’s method of characteristics, the eikonal equation of geometrical optics was solved with the help of Cauchy’s method of characteristic.
The Riemannian maps have several important applications in mathematics as well as in physics, especially within the Yang–Mills theory [3], Kaluza–Klein theory [4], supergravity and superstring theories [5], redundant robotic chains [6] etc. General relativity, which relates the gravitational force to the curvature of space-time, provides a respectable theory of gravity on a larger scale. Supergravity theories permit extra dimensions in space-time, beyond the familiar three dimensions of space and one of time. Supergravity models in higher dimensions reduce to the familiar four-dimensional space-time if it is postulated that the extra dimensions are compacted or curled up in such a way that they are not noticeable. The advantage of the extra dimensions is that they allow supergravity theories to incorporate the electromagnetic, weak, and strong forces as well as gravity.
Moreover, Şahin introduced many types of Riemannian maps ([7,8,9,10,11,12]; see also [13,14,15,16,17,18]). One may consult the references [19,20,21] for further studies.
In [22], Park introduced v-semi-slant submersions from almost Hermitian manifolds, and in [23], Sepet and Bozok proposed v-semi-slant submersions from almost product Riemannian manifolds onto Riemannian manifolds. In this paper, we are interested in studying the above idea in the setting of v-QBSR maps. The article is organized as follows: Some basic information about Riemannian maps is given in Section 2. In Section 3, using v-QBSR maps, we obtain some properties, results and decomposition theorems. In Section 4, we present the examples of the v-QBSR maps.

2. Preliminaries

Let N 1 be a C -manifold of even dimension [24]. Let J be a ( 1 , 1 ) tensor field defined on N 1 such that J 2 = I , where I denotes the identity operator; then, J is called an almost complex structure on N 1 , which is an almost complex manifold. We know that an almost complex manifold is orientable. The Nijenhuis tensor N of type ( 1 , 2 ) of an almost complex structure J is defined as:
N ( Z 1 , Z 2 ) = [ J Z 1 , J Z 2 ] [ Z 1 , Z 2 ] J [ J Z 1 , Z 2 ] J [ Z 1 , J Z 2 ] for any Z 1 , Z 2 Γ ( T N 1 ) .
If the Nijenhuis tensor N of an almost complex structure J on the manifold N 1 vanishes then the manifold N 1 becomes a complex manifold.
Let us consider a Riemannian metric g 1 on the manifold N 1 as:
g 1 ( J V 1 , J V 2 ) = g 1 ( V 1 , V 2 ) , for any V 1 , V 2 Γ ( T N 1 ) ,
then g 1 is called a Hermitian metric on the manifold N 1 and N 1 with the metric g 1 called an almost Hermitian manifold. The Levi–Civita connection ∇ on the manifold N 1 can be extended to the whole tensor algebra on the manifold N 1 . The tensor fields ( V 1 J ) are defined as:
( V 1 J ) V 2 = V 1 J V 2 J V 1 V 2
for any V 1 , V 2 Γ ( T N 1 ) .
An almost Hermitian manifold ( N 1 , g 1 , J ) is called a Kähler manifold if:
( V 1 J ) V 2 = 0
for any V 1 , V 2 Γ ( T N 1 ) .
Now, we mention the following definition for further use:
Definition 1.
Ref. [22]: Let π : ( N 1 , g 1 , J ) ( N 2 , g 2 ) be a Riemannian map. Then we say that π is a v-semi-slant Riemannian map if there is a distribution D 1 ( ker π ) such that:
( ker π ) = D 2 D 1 , J ( D 1 ) = D 1 ,
and the angle θ = θ ( Z 1 ) between J Z 1 and the space ( D 2 ) p is constant for non-zero Z 1 ( D 2 ) p and p N 1 , where D 2 is the orthogonal complement of D 1 in ( ker π ) . We call the angle θ a v-semi-slant angle.
Let T and A be O’Neill’s tensors defined as [25]:
A E 1 E 2 = H H E 1 V E 2 + V H E 1 H E 2 ,
T E 1 E 2 = H V E 1 V E 2 + V V E 1 H E 2
For any vector fields E 1 , E 2 on N 1 . It is well known that T E 1 and A E 1 are operators on the tangent bundle of N 1 , which are skew-symmetric and reversing the vertical and horizontal distributions.
Using Equations (3) and (4), we obtain:
W 1 W 2 = T W 1 W 2 + V W 1 W 2 ,
W 1 V 1 = T W 1 V 1 + H W 1 V 1 ,
V 1 W 1 = A V 1 W 1 + V V 1 W 1 ,
V 1 V 2 = H V 1 V 2 + A V 1 V 2
For any W 1 , W 2 Γ ( ker π ) and V 1 , V 2 Γ ( ker π ) , where H W 1 V 1 = A V 1 W 1 , if V 1 is basic.
Let π : ( N 1 , g 1 ) ( N 2 , g 2 ) be a C -map; then the second fundamental form [26] of π is given by:
( π ) ( Z 1 , Z 2 ) = Z 1 π π ( Z 2 ) π ( Z 1 N 1 Z 2 )
For any Z 1 , Z 2 Γ ( T N 1 ) , where π is the pullback connection. Conveniently, the Riemannian connections of the metrics g 1 and g 2 [27] are denoted by ∇.
Furthermore, a C -map π between ( N 1 , g 1 ) and ( N 2 , g 2 ) is called totally geodesic [27] if:
( π ) ( Z 1 , Z 2 ) = 0 , for any Z 1 , Z 2 Γ ( T N 1 ) .

3. V-QBSR Maps

In this section, v-QBSR maps from ( N 1 , g 1 , J ) to ( N 2 , g 2 ) are defined and studied. We now present the notion of a v-QBSR map as follows:
Definition 2.
A Riemannian map
π : ( N 1 , g 1 , J ) ( N 2 , g 2 ) ,
is called a v-QBSR map if there exist orthogonal distributions D , D 1 and D 2 such that:
(i)
( ker π ) = D D 1 D 2 ;
(ii)
D is invariant, that is, J ( D ) = D ;
(iii)
J ( D 1 ) D 2 and J ( D 2 ) D 1 ;
(iv)
for any non-zero Z i ( D i ) q i , q i N 1 , the angle θ i between J Z i and ( D i ) q i is constant and not dependent on points q i and Z i in ( D i ) q i , where i = 1 , 2 . The angles θ 1 and θ 2 are called v-slant angles of the Riemannian map.
Let π be a v-QBSR map from ( N 1 , g 1 , J ) to ( N 2 , g 2 ) . Then we have:
T N 1 = ker π ( ker π ) .
Furthermore, we put
V 1 = P V 1 + Q V 1 + R V 1 , for any V 1 Γ ( ker π ) ,
where P , Q and R are projection morphisms of ( ker π ) onto D , D 1 and D 2 , respectively. For V 2 ( Γ ker π ) , we set
J V 2 = B V 2 + C V 2 ,
where B V 2 ( Γ ker π ) and C V 2 ( Γ ker π ) .
Using Equations (12) and (13), we obtain:
J V 1 = J ( P V 1 ) + J ( Q V 1 ) + J ( R V 1 ) , = B ( P V 1 ) + C ( P V 1 ) + B ( Q V 1 ) + C ( Q V 1 ) + B ( R V 1 ) + C ( R V 1 ) .
Since D is an invariant, we have B P V 1 = 0 . Thus, the above equation reduces to
J V 1 = C ( P V 1 ) + B ( Q V 1 ) + C ( Q V 1 ) + B ( R V 1 ) + C ( R V 1 ) .
Additionally, we can express
J ( ker π ) = D ( B D 1 B D 2 ) ( C D 1 C D 2 ) .
If Z 1 Γ ( D 1 ) and Z 2 Γ ( D 2 ) ; then:
g 1 ( Z 1 , Z 2 ) = 0 .
From Definition 2 ( i i i ) , we have:
g 1 ( J Z 1 , Z 2 ) = 0 and g 1 ( Z 1 , J Z 2 ) = 0 .
Now, we consider:
g 1 ( C Z 1 , Z 2 ) = g 1 ( J Z 1 B Z 1 , Z 2 ) , = g 1 ( J Z 1 , Z 2 ) , = 0 .
In a similarly way, we obtain:
g 1 ( Z 1 , C Z 2 ) = 0 .
Let V 1 Γ ( D ) and V 2 Γ ( D 1 ) . Then, we have:
g 1 ( C V 1 , V 2 ) = g 1 ( J V 1 B V 1 , V 2 ) , = g 1 ( J V 1 , V 2 ) , = g 1 ( V 1 , J V 2 ) , = 0 ,
As D is an invariant, i.e., J V 1 Γ ( D ) .
Similarly, for V 1 Γ ( D ) and V 3 Γ ( D 2 ) , we obtain:
g 1 ( C V 1 , V 3 ) = 0 .
From the above equations, we have:
g 1 ( C Z 1 , C Z 2 ) = 0 ,
and
g 1 ( B Z 1 , B Z 2 ) = 0 ,
for any Z 1 Γ ( D 1 ) and Z 2 Γ ( D 2 ) .
Thus, we can write
C D 1 C D 2 = { 0 } , B D 1 B D 2 = { 0 } .
If θ 2 = π 2 , then C R = 0 and D 2 is anti-invariant, i.e., J ( D 2 ) ( ker π ) . In this case, we denote D 2 by D . We also have
J ( ker π ) = D C D 1 J D .
Since B D 1 ( ker π ) , B D 2 ( ker π ) , we can write
( ker π ) = B D 1 B D 2 V ,
where V denotes the orthogonal complement of ( B D 1 B D 2 ) in ( ker π ) .
Additionally, for any V 1 Γ ( ker π ) , we have
J V 1 = ϕ V 1 + ω V 1 ,
where ϕ V 1 Γ ( ker π ) and ω V 1 Γ ( ker π ) .
Lemma 1.
For π , we have:
ϕ 2 V 1 + B ω V 1 = V 1 , ω ϕ V 1 + C ω V 1 = 0 ,
ω B V 2 + C 2 V 2 = V 2 , ϕ B V 2 + B C V 2 = 0
for any V 1 Γ ( ker π ) and V 2 Γ ( ker π ) .
Proof. 
With the help of Equations (13) and (17) along with the condition J 2 = I , we obtain Lemma 1. □
Lemma 2.
For π , we obtain:
( i ) C 2 V i = ( cos 2 θ i ) V i ;
( i i ) g 1 ( C V i , C U i ) = cos 2 θ i g 1 ( V i , U i ) ;
( i i i ) g 1 ( B V i , B U i ) = sin 2 θ i g 1 ( V i , U i ) , for V i , U i Γ ( D i ) , where i = 1 , 2 .
Proof. 
The proof of Lemma 2 is the same as the one for v-semi-slant submersion (see Proposition (3.5) and Remark (3.6) of [22]). Hence we omit it. □
Lemma 3.
For π , we have:
V Z 1 ϕ Z 2 + T Z 1 ω Z 2 = ϕ V Z 1 Z 2 + B T Z 1 Z 2 ,
T Z 1 ϕ Z 2 + H Z 1 ω Z 2 = ω V Z 1 Z 2 + C T Z 1 Z 2 ,
V W 1 B W 2 + A W 1 C W 2 = ϕ A W 1 W 2 + B H W 1 W 2 ,
A W 1 B W 2 + H W 1 C W 2 = ω A W 1 W 2 + C H W 1 W 2 ,
V Z 1 B W 1 + T Z 1 C W 1 = ϕ T Z 1 W 1 + B H Z 1 W 1 ,
T Z 1 B W 1 + H Z 1 C W 1 = ω T Z 1 W 1 + C H Z 1 W 1 ,
V W 1 ϕ Z 1 + A W 1 ω Z 1 = B A W 1 Z 1 + ϕ V W 1 Z 1 ,
A W 1 ϕ Z 1 + H W 1 ω Z 1 = C A W 1 Z 1 + ω V W 1 Z 1
for any Z 1 , Z 2 Γ ( ker π ) and W 1 , W 2 Γ ( ker π ) .
Proof. 
By Equations (5)–(8), (13) and (17), we obtain Equations (18)–(25). □
Next, we define
( Z 1 ϕ ) Z 2 = V Z 1 ϕ Z 2 ϕ V Z 1 Z 2 ,
( Z 1 ω ) Z 2 = H Z 1 ω Z 2 ω V Z 1 Z 2 ,
( W 1 C ) W 2 = H W 1 C W 2 C H W 1 W 2 ,
( W 1 B ) W 2 = V W 1 B W 2 B H W 1 W 2
for any Z 1 , Z 2 Γ ( ker π ) and W 1 , W 2 Γ ( ker π ) .
Lemma 4.
For π , we obtain:
( Z 1 ϕ ) Z 2 = B T Z 1 Z 2 T Z 1 ω Z 2 ,
( Z 1 ω ) Z 2 = C T Z 1 Z 2 T Z 1 ϕ Z 2 ,
( W 1 C ) W 2 = ω A W 1 W 2 A W 1 B W 2 ,
( W 1 B ) W 2 = ϕ A W 1 W 2 A W 1 C W 2
for any Z 1 , Z 2 Γ ( ker π ) and W 1 , W 2 Γ ( ker π ) .
Proof. 
On the account of Equations (18)–(21) and (26)–(29), we obtain the required results of Lemma 4 .  □
Consequently, if ϕ and ω are parallel tensor with respect to ∇ defined on N 1 , we obtain:
B T Z 1 Z 2 = T Z 1 ω Z 2 , C T Z 1 Z 2 = T Z 1 ϕ Z 2
for any Z 1 , Z 2 Γ ( T N 1 ) .
Theorem 1.
Let π be a v-QBSR map. Then, D is integrable if and only if
g 1 ( H V 2 J V 1 H V 1 J V 2 , C Q Y 1 + C R Y 1 ) = g 1 ( A V 1 J V 2 A V 2 J V 1 , B Q Y 1 + B R Y 1 )
for V 1 , V 2 Γ ( D ) and Y 1 Γ ( D 1 D 2 ) .
Proof. 
For V 1 , V 2 Γ ( D ) and Y 1 Γ ( D 1 D 2 ) , using Equations (2), (8), (12) and (13), we have
g 1 ( [ V 1 , V 2 ] , Y 1 ) = g 1 ( V 1 J V 2 , J Y 1 ) g 1 ( V 2 J V 1 , J Y 1 ) , = g 1 ( V 1 J V 2 , J Q Y 1 + J R Y 1 ) g 1 ( V 2 J V 1 , J Q Y 1 + J R Y 1 ) , = g 1 ( H V 1 J V 2 H V 2 J V 1 , C Q Y 1 + C R Y 1 ) + g 1 ( A V 1 J V 2 A V 2 J V 1 , B Y 1 ) ,
which completes the proof. □
Theorem 2.
Let π be a v-QBSR map. Then, D 1 is integrable if and only if
g 1 ( A X 1 B X 2 A X 2 B X 1 , J P Z 1 + C R Z 1 ) = g 1 ( A X 1 B C X 2 A X 2 B C X 1 , Z 1 ) g 1 ( V X 1 B X 2 V X 2 B X 1 , B R Z 1 )
for X 1 , X 2 Γ ( D 1 ) and Z 1 Γ ( D D 2 ) .
Proof. 
For X 1 , X 2 Γ ( D 1 ) and Z 1 Γ ( D D 2 ) , we have
g 1 ( [ X 1 , X 2 ] , Z 1 ) = g 1 ( X 1 X 2 , Z 1 ) g 1 ( X 1 X 2 , Z 1 ) .
Using Equations (2), (7), (12) and (13) and Lemma 2 , we have:
g 1 ( [ X 1 , X 2 ] , Z 1 ) = g 1 ( X 1 J X 2 , J Z 1 ) g 1 ( X 2 J X 1 , J Z 1 ) , = g 1 ( X 1 B X 2 , J P Z 1 + B R Z 1 + C R Z 1 ) g 1 ( X 2 B X 1 , J P Z 1 + B R Z 1 + C R Z 1 ) + g 1 ( X 1 C X 2 , J Z 1 ) g 1 ( X 2 C X 1 , J Z 1 ) , = cos 2 θ 1 g 1 ( [ X 1 , X 2 ] , Z 1 ) + g 1 ( A X 1 B X 2 A X 2 B X 1 , J P Z 1 + C R Z 1 ) g 1 ( A X 1 B C X 2 A X 2 B C X 1 , Z 1 ) + g 1 ( V X 1 B X 2 V X 2 B X 1 , B R Z 1 ) .
Now, we have:
sin 2 θ 1 g 1 ( [ X 1 , X 2 ] , Z 1 ) = g 1 ( A X 1 B X 2 A X 2 B X 1 , J P Z 1 + C R Z 1 ) g 1 ( A X 1 B C X 2 A X 2 B C X 1 , Z 1 ) + g 1 ( V X 1 B X 2 V X 2 B X 1 , B R Z 1 ) ,
from which the proof follows. □
Theorem 3.
Let π be a v-QBSR map. Then, D 2 is integrable if and only if
g 1 ( A Y 1 B Y 2 A Y 2 B Y 1 , J P V 1 + C R V 1 ) = g 1 ( A Y 1 B C Y 2 A Y 2 B C Y 1 , V 1 ) g 1 ( V Y 1 B Y 2 V Y 2 B Y 1 , B R V 1 ) ,
for Y 1 , Y 2 Γ ( D 2 ) and V 1 Γ ( D D 1 ) .
Proof. 
By considering the similar approach as in the proof of Theorem 2 , we obtain the above result. □
Theorem 4.
Let π be a v-QBSR map. The distribution, ( ker π ) becomes a totally geodesic foliation on N 1 if and only if
g 1 ( A Z 1 P Z 2 + cos 2 θ 1 A Z 1 Q Z 2 + cos 2 θ 2 A Z 1 R Z 2 , V 1 ) = g 1 ( V Z 1 B C P Z 2 + V Z 1 B C Q Z 2 + V Z 1 B C R Z 2 , V 1 ) + g 1 ( A Z 1 B Z 2 , ω V 1 ) + g 1 ( V Z 1 B Z 2 , ϕ V 1 ) ,
for Z 1 , Z 2 Γ ( ker π ) and V 1 Γ ( ker π ) .
Proof. 
For Z 1 , Z 2 Γ ( ker π ) and V 1 Γ ( ker π ) , we have
g 1 ( Z 1 Z 2 , V 1 ) = g 1 ( Z 1 J Z 2 , J V 1 ) .
Using Equations (2), (7), (8), (12), (13) and (17) and Lemma 2 , we have:
g 1 ( Z 1 Z 2 , V 1 ) = g 1 ( Z 1 J P Z 2 , J V 1 ) + g 1 ( Z 1 J Q Z 2 , J V 1 ) + g 1 ( Z 1 J R Z 2 , J V 1 ) , = g 1 ( A Z 1 P Z 2 + cos 2 θ 1 A Z 1 Q Z 2 + cos 2 θ 2 A Z 1 R Z 2 , V 1 ) g 1 ( V Z 1 B C P Z 2 + V Z 1 B C Q Z 2 + V Z 1 B C R Z 2 , V 1 ) + g 1 ( Z 1 B P Z 2 + Z 1 B Q Z 2 + Z 1 B R Z 2 , ϕ V 1 ) + g 1 ( Z 1 B P Z 2 + Z 1 B Q Z 2 + Z 1 B R Z 2 , ω V 1 ) .
Now, since B P Z 2 + B Q Z 2 + B R Z 2 = B Z 2 and B P Z 2 = 0 , we obtain:
g 1 ( Z 1 Z 2 , V 1 ) = g 1 ( A Z 1 P Z 2 + cos 2 θ 1 A Z 1 Q Z 2 + cos 2 θ 2 A Z 1 R Z 2 , V 1 ) g 1 ( V Z 1 B C P Z 2 + V Z 1 B C Q Z 2 + V Z 1 B C R Z 2 , V 1 ) + + g 1 ( A Z 1 B Z 2 , ω V 1 ) + g 1 ( V Z 1 B Z 2 , ϕ V 1 ) .
Theorem 5.
Let π be a v-QBSR map. The distribution ( ker π ) becomes a totally geodesic foliation on N 1 if and only if
g 1 ( T X 1 X 2 , P Z 1 + cos 2 θ 1 Q Z 1 + cos 2 θ 2 R Z 1 ) = g 1 ( V X 1 X 2 , B C P Z 1 + B C Q Z 1 + B C R Z 1 ) + g 1 ( T X 1 ω X 2 , B Z 1 ) + g 1 ( V X 1 ϕ X 2 , B Z 1 ) ,
for X 1 , X 2 Γ ( ker π ) and Z 1 Γ ( ker π ) .
Proof. 
For X 1 , X 2 Γ ( ker π ) ; Z 1 Γ ( ker π ) , with the help of Equations (2) and (12), we have
g 1 ( X 1 X 2 , Z 1 ) = g 1 ( X 1 X 2 , P Z 1 + Q Z 1 + R Z 1 ) .
Now, using Equations (5), (6), (13) and (17) and Lemma 2 , we have:
g 1 ( X 1 X 2 , Z 1 ) = g 1 ( X 1 J X 2 , J P Z 1 ) + g 1 ( X 1 J X 2 , J Q Z 1 ) + g 1 ( X 1 J X 2 , J R Z 1 ) , = g 1 ( T X 1 X 2 , P Z 1 + cos 2 θ 1 Q Z 1 + cos 2 θ 2 R Z 1 ) g 1 ( V X 1 X 2 , B C P Z 1 + B C Q Z 1 + B C R Z 1 ) + g 1 ( X 1 ϕ X 2 + X 1 ω X 2 , B P Z 1 + B Q Z 1 + B R Z 1 ) .
Now, since B P Z 1 + B Q Z 1 + B R Z 1 = B Z 1 and B P Z 1 = 0 , we have:
g 1 ( X 1 X 2 , Z 1 ) = g 1 ( T X 1 X 2 , P Z 1 + cos 2 θ 1 Q Z 1 + cos 2 θ 2 R Z 1 ) g 1 ( V X 1 X 2 , B C P Z 1 + B C Q Z 1 + B C R Z 1 ) + g 1 ( T X 1 ω X 2 + V X 1 ϕ X 2 , B Z 1 ) .
Theorem 6.
Let π be a v-QBSR map. Then, D defines a totally geodesic foliation on N 1 if and only if
g 1 ( A Y 1 P Y 2 , B C Q U 1 + B C R U 1 ) = g 1 ( A Y 1 J P Y 2 , B Q U 1 + B R U 1 ) + g 1 ( H Y 1 P Y 2 , cos 2 θ 1 Q U 1 + cos 2 θ 2 R U 1 ) ,
and
g 1 ( A Y 1 J P Y 2 , ϕ W 1 ) = g 1 ( H Y 1 J P Y 2 , ω W 1 ) ,
for Y 1 , Y 2 Γ ( D ) , U 1 Γ ( D 1 D 2 ) and W 1 Γ ( ker π ) .
Proof. 
For Y 1 , Y 2 Γ ( D ) , U 1 Γ ( D 1 D 2 ) and W 1 Γ ( ker π ) , using Equations (2), (8), (12), (13) and (17) and Lemma 2 , we have:
g 1 ( Y 1 Y 2 , U 1 ) = g 1 ( Y 1 J Y 2 , J U 1 ) , = g 1 ( Y 1 J P Y 2 , J Q U 1 + J R U 1 ) , = g 1 ( A Y 1 J P Y 2 , B Q U 1 + B R U 1 ) + g 1 ( H Y 1 P Y 2 , cos 2 θ 1 Q U 1 + cos 2 θ 2 R U 1 ) g 1 ( A Y 1 P Y 2 , B C Q U 1 + B C R U 1 ) .
Now, again using Equations (2), (8), (12) and (17) we have:
g 1 ( Y 1 Y 2 , W 1 ) = g 1 ( Y 1 J Y 2 , J W 1 ) , = g 1 ( Y 1 J P Y 2 , ϕ W 1 + ω W 1 ) , = g 1 ( A Y 1 J P Y 2 , ϕ W 1 ) + g 1 ( H Y 1 J P Y 2 , ω W 1 ) ,
which completes the proof. □
Theorem 7.
Let π be a v-QBSR map. Then, D 1 defines a totally geodesic foliation on N 1 if and only if
g 1 ( A X 1 B C X 2 , Z 1 ) = g 1 ( V X 1 B X 2 , B R Z 1 ) + g 1 ( A X 1 B X 2 , J P Z 1 + C R Z 1 ) ,
g 1 ( V X 1 B C X 2 , Z 2 ) = g 1 ( A X 1 B Q X 2 , ω Z 2 ) + g 1 ( V X 1 B Q X 2 , ϕ Z 2 ) ,
for X 1 , X 2 Γ ( D 1 ) , Z 1 Γ ( D D 2 ) and Z 2 Γ ( ker π ) .
Proof. 
For X 1 , X 2 Γ ( D 1 ) , Z 1 Γ ( D D 2 ) and Z 2 Γ ( ker π ) , by using Equations (2), (7), (8), (12), (13) and (17) and Lemma 2 , we have:
g 1 ( X 1 X 2 , Z 1 ) = g 1 ( X 1 B X 2 , J Z 1 ) + g 1 ( X 1 C X 2 , J Z 1 ) , = cos 2 θ 1 g 1 ( X 1 X 2 , Z 1 ) g 1 ( A X 1 B C X 2 , Z 1 ) + g 1 ( V X 1 B X 2 , B R Z 1 ) + g 1 ( A X 1 B X 2 , J P Z 1 + C R Z 1 ) .
Now, we have:
sin 2 θ 1 g 1 ( X 1 X 2 , Z 1 ) = g 1 ( A X 1 B C X 2 , Z 1 ) + g 1 ( V X 1 B X 2 , B R Z 1 ) + g 1 ( A X 1 B X 2 , J P Z 1 + C R Z 1 ) .
From Equations (2), (7), (8), (13) and (17) and Lemma 2 , we have:
g 1 ( X 1 X 2 , Z 2 ) = g 1 ( X 1 B X 2 , J Z 2 ) + g 1 ( X 1 C X 2 , J Z 2 ) , = cos 2 θ 1 g 1 ( X 1 X 2 , Z 2 ) g 1 ( V X 1 B C X 2 , Z 2 ) + g 1 ( A X 1 B X 2 , ω Z 2 ) + g 1 ( V X 1 B X 2 , ϕ Z 2 ) .
Therefore,
sin 2 θ 1 g 1 ( X 1 X 2 , Z 2 ) = g 1 ( V X 1 B C X 2 , Z 2 ) + g 1 ( A X 1 B X 2 , ω Z 2 ) + g 1 ( V X 1 B X 2 , ϕ Z 2 ) .
Theorem 8.
Let π be a v-QBSR map. Then, D 2 defines a totally geodesic foliation on N 1 if and only if
g 1 ( A Y 1 B C Y 2 , V 1 ) = g 1 ( V Y 1 B Y 2 , J P V 1 + B R V 1 ) + g 1 ( A Y 1 B Y 2 , C R V 1 ) ,
g 1 ( V Y 1 B C Y 2 , V 2 ) = g 1 ( A Y 1 B Y 2 , ω V 2 ) + g 1 ( V Y 1 B Y 2 , ϕ V 2 )
for Y 1 , Y 2 Γ ( D 2 ) , V 1 Γ ( D D 1 ) and V 2 Γ ( ker π ) .
Proof. 
By considering a similar approach as in the proof of Theorem 7 , we obtain the above result. □
Theorem 9.
Let π be a v-QBSR map. Then, π is a totally geodesic map if and only if
g 1 ( T V 1 V 2 , P X 1 + cos 2 θ 1 Q X 1 + cos 2 θ 2 R X 1 ) ) = g 1 ( V V 1 V 2 , B C P X 1 + B C Q X 1 + B C R X 1 ) g 1 ( V V 1 ϕ V 2 , B X 1 ) g 1 ( T V 1 ω V 2 , B X 1 ) ,
and
g 1 ( A X 1 V 1 , P X 2 + cos 2 θ 1 Q X 2 + cos 2 θ 2 R X 2 ) = g 1 ( V X 1 V 1 , B C P X 2 + B C Q X 2 + B C R X 2 ) g 1 ( V X 1 ϕ V 1 , B X 2 ) g 1 ( A X 1 ω V 1 , B X 2 )
for V 1 , V 2 Γ ( ker π ) and X 1 , X 2 Γ ( ker π ) .
Proof. 
Since π is a Riemannian map, we have
( π ) ( X 1 , X 2 ) = 0 ,
for X 1 , X 2 Γ ( ker π ) .
For V 1 , V 2 Γ ( ker π ) and X 1 , X 2 Γ ( ker π ) , using Equations (2), (9), (12), (13) and (17), we have:
g 2 ( ( π ) ( V 1 , V 2 ) , π ( X 1 ) ) = g 1 ( V 1 V 2 , X 1 ) = g 1 ( V 1 J V 2 , J X 1 ) = g 1 ( V 1 J V 2 , J P X 1 ) g 1 ( V 1 J V 2 , J Q X 1 ) g 1 ( V 1 J V 2 , J R X 1 ) , = g 1 ( V 1 J V 2 , C P X 1 + C Q X 1 + C R X 1 ) g 1 ( V 1 J V 2 , B P X 1 + B Q X 1 + B R X 1 ) , = g 1 ( V 1 V 2 , P X 1 + cos 2 θ 1 Q X 1 + cos 2 θ 2 R X 1 ) + g 1 ( V 1 V 2 , B C P X 1 + B C Q X 1 + B C R X 1 ) g 1 ( V 1 ϕ V 2 , B X 1 ) g 1 ( V 1 ω V 2 , B X 1 ) = g 1 ( T V 1 V 2 , P X 1 + cos 2 θ 1 Q X 1 + cos 2 θ 2 R X 1 ) + g 1 ( V V 1 V 2 , B C P X 1 + B C Q X 1 + B C R X 1 ) g 1 ( V V 1 ϕ V 2 , B X 1 ) g 1 ( T V 1 ω V 2 , B X 1 ) .
Furthermore, using Equations (2), (9), (7), (8), (12), (13) and (17), we obtain:
g 2 ( ( π ) ( X 1 , V 1 ) , π ( X 2 ) ) = g 1 ( X 1 V 1 , X 2 ) , = g 1 ( X 1 J V 1 , J X 2 ) , = g 1 ( X 1 J V 1 , J P X 2 + J Q X 2 + J R X 2 ) , = g 1 ( X 1 J V 1 , B P X 2 + B Q X 2 + B R X 2 ) g 1 ( X 1 J V 1 , C P X 2 + C Q X 2 + C R X 2 ) , = g 1 ( V X 1 ϕ V 1 , B X 2 ) g 1 ( A X 1 ω V 1 , B X 2 ) + g 1 ( V X 1 V 1 , B C P X 2 + B C Q X 2 + B C R X 2 ) g 1 ( A X 1 V 1 , P X 2 + cos 2 θ 1 Q X 2 + cos 2 θ 2 R X 2 ) .

4. Example

Let R 2 t be the Euclidean space. Let ( y 1 , y 2 , , y 2 t 1 , y 2 t ) be the coordinates of R 2 t . Define an almost complex structure J on R 2 t as follows:
J ( a 1 y 1 + a 2 y 2 + + a 2 t 1 y 2 t 1 + a 2 t y 2 t ) = a 2 y 1 + a 1 y 2 + a 2 t y 2 t 1 + a 2 t 1 y 2 t ,
where a 1 , a 2 , , a 2 t are C -functions on R 2 t . Throughout this section, we will use this notation.
Example 1.
Define a map π : R 12 R 8 by:
π ( y 1 , y 2 , , y 12 ) = ( y 1 sin θ 1 + y 3 cos θ 1 , y 4 , 202 , 303 , y 5 sin θ 2 y 7 cos θ 2 , y 8 , y 9 , y 10 ) ,
which is a v-QBSR map such that
ker π = cos θ 1 y 1 sin θ 1 y 3 , y 2 , cos θ 2 y 5 + sin θ 2 y 7 , y 6 , y 11 , y 12
( ker π ) = sin θ 1 y 1 + cos θ 1 y 3 , y 4 , sin θ 2 y 5 + cos θ 2 y 7 , y 8 , y 9 , y 10 ,
( ker π ) = D D 1 D 2 ,
where:
D = y 9 , y 10 , D 1 = sin θ 1 y 1 + cos θ 1 y 3 , y 4 , D 2 = sin θ 2 y 5 + cos θ 2 y 7 , y 8
with v-quasi-bi-slant angles θ 1 and θ 2 .
Example 2.
Define a map π : R 12 R 8 by:
π ( y 1 , y 2 , , y 12 ) = ( y 1 , y 2 , y 3 y 5 2 , y 6 , 2021 , 3 y 7 + y 9 2 , y 8 , 2022 ) ,
which is a v-quasi-bi-slant map such that
( ker π ) = 1 2 ( y 3 + y 5 ) , y 4 , 1 2 ( y 7 3 y 9 ) , y 10 , y 11 , y 12 ,
( ker π ) = y 1 , y 2 , 1 2 ( y 3 y 5 ) , y 6 , 1 2 ( 3 y 7 + y 9 ) , y 8 ,
( ker π ) = D D 1 D 2 ,
where:
D = < y 1 , y 2 > , D 1 = < 1 2 ( y 3 y 5 ) , y 6 > , D 2 = < 1 2 ( 3 y 7 + y 9 ) , y 8 > ,
with v-quasi-bi-slant angles θ 1 = π 4 and θ 2 = π 6 .

Author Contributions

Conceptualization, Z.C., S.K., R.P. and A.H.; methodology, Z.C., S.K., R.P. and M.B.; investigation, Z.C., R.P., A.H. and M.B.; writing—original draft preparation, Z.C., R.P., A.H. and M.B.; writing—review and editing, S.K., R.P. and A.H. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank the Deanship of Scientific Research at Umm Al-Qura University for supporting this work by Grant Code: (22UQU4330007DSR01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fischer, A.E. Riemannian maps between Riemannian manifolds. Contemp. Math. 1992, 132, 331–366. [Google Scholar]
  2. Şahin, B. Riemannian Submersions, Riemannian Maps in Hermitian Geometry, and Their Applications; Elsevier/Academic Press: London, UK, 2017; ISBN 978-0-12-804391-2. [Google Scholar]
  3. Bourguignon, J.P.; Lawson, H.B. Stability and isolation phenomena for Yang-mills fields. Commun. Math. Phys. 1981, 79, 189–230. [Google Scholar] [CrossRef]
  4. Ianus, S.; Visinescu, M. Kaluza-Klein theory with scalar fields and generalized Hopf manifolds. Class. Quantum Gravit. 1987, 4, 1317–1325. [Google Scholar] [CrossRef]
  5. Ianus, S.; Visinescu, M. Space-time compactification and Riemannian submersions. In The Mathematical Heritage of C. F. Gauss; Rassias, G., Ed.; World Scientific: River Edge, NJ, USA, 1991; pp. 358–371. [Google Scholar]
  6. Altafini, C. Redundant robotic chains on Riemannian submersions. IEEE Trans. Robot. Autom. 2004, 20, 335–340. [Google Scholar] [CrossRef]
  7. Park, K.S.; Sahin, B. Semi-slant Riemannian maps into almost Hermitian manifolds. Czechoslov. Math. J. 2014, 64, 1045–1061. [Google Scholar] [CrossRef] [Green Version]
  8. Şahin, B. Invariant and anti-invariant Riemannian maps to Kähler manifolds. Int. J. Geom. Methods Mod. Phys. 2010, 7, 337–355. [Google Scholar] [CrossRef]
  9. Şahin, B. Semi-invariant Riemannian maps from almost Hermitian manifolds. Indag. Math. 2012, 23, 80–94. [Google Scholar] [CrossRef] [Green Version]
  10. Şahin, B. Slant Riemannian maps from almost Hermitian manifolds. Quaest. Math. 2013, 36, 449–461. [Google Scholar] [CrossRef] [Green Version]
  11. Şahin, B. Semi-invariant Riemannian maps to Kähler manifolds. Int. J. Geom. Methods Mod. 2011, 8, 1439–1454. [Google Scholar] [CrossRef]
  12. Şahin, B. Hemi-slant Riemannian Maps. Mediterr. J. Math. 2017, 14, 10. [Google Scholar] [CrossRef]
  13. Park, K.S. Almost h-semi-slant Riemannian maps. Taiwan J. Math. 2013, 17, 937–956. [Google Scholar] [CrossRef]
  14. Park, K.S. Almost h-semi-slant Riemannian maps to almost quaternionic Hermitian manifolds. Commun. Contemp. 2015, 17, 1550008. [Google Scholar] [CrossRef]
  15. Prasad, R.; Kumar, S. Slant Riemannian maps from Kenmotsu manifolds into Riemannian manifolds. Glob. J. Pure Appl. Math. 2017, 13, 1143–1155. [Google Scholar]
  16. Prasad, R.; Kumar, S. Semi-slant Riemannian maps from almost contact metric manifolds into Riemannian manifolds. Tbil. Math. J. 2018, 11, 19–34. [Google Scholar] [CrossRef]
  17. Prasad, R.; Kumar, S.; Kumar, S.; Vanli, T.A. On Quasi-Hemi-Slant Riemannian Maps. Gazi Univ. J. Sci. 2021, 34, 477–491. [Google Scholar] [CrossRef]
  18. Prasad, R.; Kumar, S. Semi-slant Riemannian maps from cosymplectic manifolds into Riemannian manifolds. Gulf J. Math. 2020, 9, 62–80. [Google Scholar]
  19. Li, Y.L.; Alkhaldi, A.H.; Ali, A.; Laurian-Ioan, P. On the topology of warped product pointwise semi-slant submanifolds with positive curvature. Mathematics 2021, 9, 3156. [Google Scholar] [CrossRef]
  20. Li, Y.L.; Lone, M.A.; Wani, U.A. Biharmonic submanifolds of Kähler product manifolds. AIMS Math. 2021, 6, 9309–9321. [Google Scholar] [CrossRef]
  21. Prasad, R.; Pandey, S. Slant Riemannian Maps from an Almost Contact Manifold. Filomat 2017, 31, 3999–4007. [Google Scholar] [CrossRef]
  22. Park, K.S. On the v-semi-slant submersions from almost Hermitian manifolds. Commun. Korean Math. Soc. 2021, 36, 173–187. [Google Scholar]
  23. Sepet, S.A.; Bozok, H.G. V-Semi-slant submersions from almost product Riemannian manifolds. Palest. J. Math 2021, 10, 299–307. [Google Scholar]
  24. Yano, K.; Kon, M. Structures on Manifolds; Series in Pure Mathematics; World Scientific: Singapore, 1984. [Google Scholar]
  25. O’Neill, B. The fundamental equations of a submersion. Michi. Math. J. 1966, 13, 459–469. [Google Scholar] [CrossRef]
  26. Nore, T. Second fundamental form of a map. Ann. Mat. Pura Appl. 1987, 146, 281–310. [Google Scholar] [CrossRef]
  27. Baird, P.; Wood, J.C. Harmonic Morphism between Riemannian Manifolds; Oxford Science Publications: Oxford, UK, 2003. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kumar, S.; Bilal, M.; Prasad, R.; Haseeb, A.; Chen, Z. V-Quasi-Bi-Slant Riemannian Maps. Symmetry 2022, 14, 1360. https://doi.org/10.3390/sym14071360

AMA Style

Kumar S, Bilal M, Prasad R, Haseeb A, Chen Z. V-Quasi-Bi-Slant Riemannian Maps. Symmetry. 2022; 14(7):1360. https://doi.org/10.3390/sym14071360

Chicago/Turabian Style

Kumar, Sushil, Mohd Bilal, Rajendra Prasad, Abdul Haseeb, and Zhizhi Chen. 2022. "V-Quasi-Bi-Slant Riemannian Maps" Symmetry 14, no. 7: 1360. https://doi.org/10.3390/sym14071360

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop