Next Article in Journal
A New 3-Parameter Bounded Beta Distribution: Properties, Estimation, and Applications
Next Article in Special Issue
Soft Regular Generalized ω-Closed Sets and Soft ω-T1/2 Spaces
Previous Article in Journal
Statistical Inference for Odds Ratio of Two Proportions in Bilateral Correlated Data
Previous Article in Special Issue
Classification of Surfaces of Coordinate Finite Type in the Lorentz–Minkowski 3-Space
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Study of Clairaut Semi-Invariant Riemannian Maps from Cosymplectic Manifolds

1
School of Mathematics, Hangzhou Normal University, Hangzhou 311121, China
2
Department of Mathematics and Astronomy, University of Lucknow (U.P.), Lucknow 226007, India
3
Department of Mathematics, College of Science, Jazan University, Jazan 45142, Saudi Arabia
4
Shri Jai Narain Post Graduate College, University of Lucknow (U.P.), Lucknow 226001, India
5
Department of Mathematics, Dr. Shree Krishna Sinha Women’s College Motihari, Babasaheb Bhimrao Ambedkar University, Bihar 845401, India
*
Author to whom correspondence should be addressed.
Axioms 2022, 11(10), 503; https://doi.org/10.3390/axioms11100503
Submission received: 7 August 2022 / Revised: 17 September 2022 / Accepted: 21 September 2022 / Published: 26 September 2022
(This article belongs to the Special Issue Differential Geometry and Its Application)

Abstract

:
In the present note, we characterize Clairaut semi-invariant Riemannian maps from cosymplectic manifolds to Riemannian manifolds. Moreover, we provide a nontrivial example of such a Riemannian map.

1. Introduction

The theory of Riemannian maps between Riemannian manifolds is widely used to compare the geometric structures between two Riemannian manifolds, initiated by Fischer [1]. Let ( M 1 , g 1 ) and ( M 2 , g 2 ) be two Riemannian manifolds of dimensions m and n, respectively. Let a Riemannian map Π : ( M 1 , g 1 ) ( M 2 , g 2 ) be a differentiable map between ( M 1 , g 1 ) and ( M 2 , g 2 ) such that 0 < r a n k Π * < min { m , n } , where Π * represents a differential map of Π . 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 M 1 , then the T M 1 has the following orthogonal decomposition:
T M 1 = ker Π * ( ker Π * ) .
We denote the range of Π * by r a n g e Π * and for a point q M 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 ) M 2 . The tangent space T Π ( q ) M 2 has the following orthogonal decomposition:
T Π ( q ) M 2 = ( r a n g e Π * Π ( q ) ) ( r a n g e Π * Π ( q ) ) .
A differentiable map Π : ( M 1 , g 1 ) ( M 2 , g 2 ) is called a Riemannian map at q M 1 if the horizontal restriction Π * q h : ( ker Π * q ) ( r a n g e Π * Π ( q ) ) is a linear isometric between the inner product spaces ( ( 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 ) ) .
Further, the notion of the Riemannian map has been studied from different perspectives, such as invariant and anti-invariant Riemannian maps [2], semi-invariant Riemannian maps [3], slant Riemannian maps [4,5,6], semi-slant Riemannian maps [7,8,9], hemi-slant Riemannian maps [10], quasi-hemi-slant Riemannian maps [11] etc.
On the other side, in the theory of the geodesics upon a surface of revolution, the prestigious Clairaut’s theorem states that for any geodesic c ( c : I 1 R M 1 on M 1 ) on the revolution surface M 1 the product r s i n θ is constant along c, where θ ( s ) is the angle between c ( s ) and the meridian curve through c ( s ) , s I 1 . This means that it is independent of s. In 1972, Bishop [12] studied Riemannian submersions which are a generalization of Clairaut’s theorem. According to him, a submersion Π : M 1 M 2 is said to be a Clairaut submersion if there is a function r : M 1 R + such that for every geodesic making an angle θ with the horizontal subspaces, r s i n θ is constant. This notion has also been studied in Lorentzian spaces, time-like and space-like spaces, by the authors [13,14,15]. Later, in [16], it was shown that such submersions have their applications in static spacetimes.
Moreover, Clairaut submersions were further generalized in [17]. We recommend the papers [18,19,20,21,22,23,24,25,26,27,28,29,30,31,32] and the references therein for more details about the further related studies.
In this paper, we are interested in studying the above idea in contact manifolds. Throughout the manuscript, we denote semi-invariant Riemannian maps by SIR maps and Clairaut semi-invariant Riemannian maps by CSIR maps. The article is organized as follows: In Section 2, we gather some basic facts that are needed for this paper. In Section 3, we define a CSIR map from an almost contact metric manifold to a Riemannian manifold and study its geometry. In Section 4, we give a nontrivial example of the CSIR map from cosymplectic manifolds to Riemannian manifolds.

2. Preliminaries

An odd-dimensional smooth manifold M 1 is said to have an almost contact structure [33] if there exist on M 1 a tensor field ϕ of type ( 1 , 1 ) , a vector field ξ , and 1-form η such that
ϕ 2 V 1 = V 1 + η ( V 1 ) ξ , η ϕ = 0 , ϕ ξ = 0 ,
η ( ξ ) = 1 .
If there exists a Riemannian metric g 1 on an almost contact manifold M 1 satisfying:
g 1 ( ϕ V 1 , ϕ V 2 ) = g 1 ( V 1 , V 2 ) η ( V 1 ) η ( V 2 ) , g 1 ( V 1 , ϕ V 2 ) = g 1 ( ϕ V 1 , V 2 ) ,
g 1 ( V 1 , ξ ) = η ( V 1 ) ,
where V 1 , V 2 are any vector fields on M 1 , then M 1 is called an almost contact metric manifold [34] with an almost contact structure ( ϕ , ξ , η , g 1 ) and is represented by ( M 1 , ϕ , ξ , η , g 1 ) .
An almost contact structure ( ϕ , ξ , η ) is said to be normal if the almost complex structure J on the product manifold M 1 × R is given by
J ( V 1 , F d d t ) = ( ϕ V 1 F ξ , η ( V 1 ) d d t ) ,
where J 2 = I and F is a differentiable function on M 1 × R that has no torsion, i.e., J is integrable. The condition for normality in terms of ϕ , ξ , and η is given by [ ϕ , ϕ ] + 2 d η ξ = 0 on M 1 , where [ ϕ , ϕ ] is the Nijenhuis tensor of ϕ . Further, the fundamental 2-form Φ is defined by Φ ( V 1 , V 2 ) = g 1 ( V 1 , ϕ V 2 ) .
A manifold M 1 with the structure ( ϕ , ξ , η , g 1 ) is said to be cosymplectic [33] if
( V 1 ϕ ) V 2 = 0 ,
for any vector fields V 1 , V 2 on M 1 , where ∇ stands for the Riemannian connection of the metric g 1 on M 1 . For a cosymplectic manifold, we have
V 1 ξ = 0
for any vector field V 1 on M 1 .
O’Neill’s tensors [35] T and A are given by
A F 1 F 2 = H H F 1 V F 2 + V H F 1 H F 2 ,
T F 1 F 2 = H V F 1 V F 2 + V V F 1 H F 2 ,
for any F 1 , F 2 on M 1 . It is easy to see that T F 1 and A F 1 are skew-symmetric operators on the tangent bundle of M 1 reversing the vertical and the horizontal distributions. In addition, for any vertical vector fields X 1 and X 2 , the tensor field T has the symmetry property, i.e.,
T X 1 X 2 = T X 2 X 1 ,
while for horizontal vector fields Z 1 , Z 2 , the tensor field A has alternation property, i.e.,
A Z 1 Z 2 = A Z 2 Z 1 .
From Equations (9) and (10) we have
U 1 U 2 = T U 1 U 2 + V U 1 U 2 ,
U 1 W 1 = T U 1 W 1 + H U 1 W 1 ,
W 1 U 1 = A W 1 U 1 + V W 1 U 1 ,
W 1 W 2 = H W 1 W 2 + A W 1 W 2
for all U 1 , U 2 Γ ( ker Π * ) and W 1 , W 2 Γ ( ker Π * ) , where H U 1 W 1 = A W 1 U 1 and W 1 is basic. It can be easily seen that T acts on the fibers as the second fundamental form, while A acts on the horizontal distribution and measures the obstruction to the integrability of the distribution.
It is noticed that for p M 1 , Z 1 V p and X 1 H p the linear operators
A X 1 , T Z 1 : T p M 1 T p M 1
are skew-symmetric, i.e.,
g 1 ( A X 1 F 1 , F 2 ) = g 1 ( F 1 , A X 1 F 2 ) and g 1 ( T Z 1 F 1 , F 2 ) = g 1 ( F 1 , T Z 1 F 2 )
for each F 1 , F 2 T P M 1 . Since T Z 1 is skew-symmetric, we observe that Π has totally geodesic fibres if and only if T 0 .
The map Π between two Riemannian manifolds is totally geodesic if
( Π * ) ( V 1 , V 2 ) = 0 V 1 , V 2 Γ ( T M 1 ) .
A totally umbilical map is a Riemannian map with totally umbilical fibers [36] if
T Y 1 Y 2 = g 1 ( Y 1 , Y 2 ) H
for all Y 1 , Y 2 Γ ( ker Π * ) , where H denotes the mean curvature vector field of fibers.
The map Π * can be observed as a section of the bundle H o m ( T M 1 , Π 1 T M 2 ) M 1 , where Π 1 T M 2 is the bundle which has fibers ( Π 1 T M 2 ) x = T Π ( x ) M 2 and has a connection ∇ induced from the Riemannian connection M 1 and the pullback connection Π , then the second fundamental form of Π is given by
( Π * ) ( W 1 , W 2 ) = W 1 Π Π * ( W 2 ) Π * ( W 1 M 1 W 2 )
for the vector fields W 1 , W 2 Γ ( T M 1 ) . We know that the second fundamental form is symmetric.
Now, we have the following lemma [2]:
Lemma 1.
Let Π : ( M 1 , g 1 ) ( M 2 , g 2 ) be a map between Riemannian manifolds. Then
g 2 ( ( Π * ) ( Y 1 , Y 2 ) , Π * ( Y 3 ) ) = 0 Y 1 , Y 2 , Y 3 Γ ( ker Π * ) .
As a result of above Lemma, we obtain
( Π * ) ( Z 1 , Z 2 ) ( Γ ( r a n g e Π * ) ) Z 1 , Z 2 Γ ( ker Π * ) .

3. CSIR Map from Cosymplectic Manifolds

Let S be a revolution surface in R 3 with rotation axis L. For any q S, we denote the distance from q to L by r ( q ) . Given a geodesic α : I R S on S, let θ ( t ) be the angle between α ( t ) and the meridian curve through α ( t ) , t I . A well-known Clairaut’s theorem says that for any geodesic α on S, the product r sin θ ( t ) is constant along α , i.e., it is independent of t.
Recently, Sahin [30] initiated the study of Clairaut Riemannian maps. He defined a map Π : ( M 1 , g 1 ) ( M 2 , g 2 ) called a Clairaut Riemannian map if there exists a positive function r on M 1 , such that for any geodesic α on M 1 , the function ( r α ) sin θ is constant, where for any t , θ ( t ) is the angle between α . ( t ) and the horizontal space at α ( t ) . Moreover, he obtained the following necessary and sufficient condition for a Riemannian map to be a Clairaut Riemannian map:
Theorem 1
([30]). Let Π : ( M 1 , g 1 ) ( M 2 , g 2 ) be a Riemannian map with connected fibers. Then, Π is a Clairaut Riemannian map with r = e f if each fiber is totally umbilical and has the mean curvature vector field H = f , where f is the gradient of the function f with respect to g 1 .
Definition 1
([3]). Let Π be a Riemannian map from an almost contact metric manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) . Then, we say that Π is an SIR map if there is a distribution D 1 ker Π * such that
ker Π * = D 1 D 2 , ϕ ( D 1 ) = D 1 , ϕ ( D 2 ) ( ker Π * ) ,
where D 1 and D 2 are mutually orthogonal distributions in ( ker Π * ) .
Let μ denote the complementary orthogonal subbundle to ϕ ( D 2 ) in ( ker Π * ) . Then, we have
( ker Π * ) = ϕ ( D 2 ) μ .
Obviously, μ is an invariant subbundle of ( ker Π * ) with respect to the contact structure ϕ . We say that an SIR map Π : M 1 M 2 admits a vertical Reeb vector field ξ if it is tangent to ( ker Π * ) and it admits a horizontal Reeb vector field ξ if it is normal to ( ker Π * ) . It is easy to see that μ contains the Reeb vector field in case the Riemannian map admits horizontal Reeb vector field.
Now, we define the notion of the CSIR map in contact manifolds as follows:
Definition 2.
An SIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) is called a CSIR map if it satisfies the condition of a Clairaut Riemannian map.
For any vector field Z 1 Γ ( ker Π * ) , we input
Z 1 = P Z 1 + Q Z 1 ,
where P and Q are projection morphisms [36] of ker Π * onto D 1 and D 2 , respectively.
For any V 1 ( ker Π * ) , we obtain
ϕ V 1 = ψ V 1 + ω V 1 ,
where ψ V 1 Γ ( D 1 ) and ω V 1 Γ ( ϕ D 2 ) . In addition, for V 2 Γ ( ker Π * ) , we have
ϕ V 2 = B V 2 + C V 2 ,
where B V 2 Γ ( D 2 ) and C V 2 Γ ( μ ) .
Definition 3
([14]). Let Π be an SIR map from an almost contact metric manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) . If μ = { 0 } or μ = < ξ > , i.e., ( ker Π * ) = ϕ ( D 2 ) or ( ker Π * ) = ϕ ( D 2 ) < ξ > , respectively. Then we call ϕ a Lagrangian Riemannian map. In this case, for any horizontal vector field V 1 , we have
B V 1 = ϕ V 1 and C V 1 = 0 .
Lemma 2.
Let Π be an SIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) admitting vertical or horizontal Reeb vector field. Then, we obtain
V Y 1 ψ Y 2 + T Y 1 ω Y 2 = B T Y 1 Y 2 + ψ V Y 1 Y 2 ,
T Y 1 ψ Y 2 + H Y 1 ω Y 2 = C T Y 1 Y 2 + ω V Y 1 Y 2 ,
V V 1 B V 2 + A V 1 C V 2 = B H V 1 V 2 + ψ A V 1 V 2 ,
A V 1 B V 2 + H V 1 C V 2 = C H V 1 V 2 + ω A V 1 V 2 ,
V Y 1 B V 1 + T Y 1 C V 1 = ψ T Y 1 V 1 + B H Y 1 V 1 ,
T Y 1 B V 1 + H Y 1 C V 1 = ω T Y 1 V 1 + C H Y 1 V 1 ,
V V 1 ψ Y 1 + A V 1 ω Y 1 = B A V 1 Y 1 + ψ V V 1 Y 1 ,
A V 1 ψ Y 1 + H V 1 ω Y 1 = C A V 1 Y 1 + ω V V 1 Y 1 ,
where Y 1 , Y 2 Γ ( ker Π * ) and V 1 , V 2 Γ ( ker Π * ) .
Proof. 
Using Equations (7), (13)–(16), (23) and (24), we obtain Lemma 2. □
Corollary 1.
Let Π be a Lagrangian Riemannian map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) admitting vertical or horizontal Reeb vector field. Then, we obtain
V X 1 ψ X 2 + T X 1 ω X 2 = B T X 1 X 2 + ψ V X 1 X 2 , T X 1 ψ X 2 + H X 1 ω X 2 = ω V X 1 X 2 ,
V Z 1 B Z 2 = B H Z 1 Z 2 + ψ A Z 1 Z 2 , A Z 1 B Z 2 = ω A Z 1 Z 2 ,
V X 1 B Z 1 = ψ T X 1 Z 1 + B H X 1 Z 1 , T X 1 B Z 1 = ω T X 1 Z 1 ,
V Z 1 ψ X 1 + A Z 1 ω X 1 = B A Z 1 X 1 + ψ V Z 1 X 1 , A Z 1 ψ X 1 + H Z 1 ω X 1 = ω V Z 1 X 1 ,
where X 1 , X 2 Γ ( ker Π * ) and Z 1 , Z 2 Γ ( ker Π * ) .
Lemma 3.
Let Π be an SIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) admitting vertical or horizontal Reeb vector field. Then, we have
T U 1 ξ = 0 ,
A U 2 ξ = 0
for U 1 Γ ( ker Π * ) and U 2 Γ ( ker Π * ) .
Proof. 
Using Equations (8), (14) and (16) we obtain Lemma 3. □
Lemma 4.
Let Π be an SIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) . If α : I 2 R M 1 is a regular curve and Y 1 ( t ) and Y 2 ( t ) are the vertical and horizontal components of the tangent vector field α . = E of α ( t ) , respectively, then α is a geodesic if and only if along α the following relations hold:
V α . B Y 2 + V α . ψ Y 1 + ( T Y 1 + A Y 2 ) C Y 2 + ( T Y 1 + A Y 2 ) ω Y 1 = 0 ,
H α . C Y 2 + H α . ω Y 1 + ( T Y 1 + A Y 2 ) B Y 2 + ( A Y 2 + T Y 1 ) ψ Y 1 = 0 .
Proof. 
Let α : I 2 M 1 be a regular curve on M 1 . Since Y 1 ( t ) and Y 2 ( t ) are the vertical and horizontal parts of the tangent vector field α . ( t ) , i.e., α . ( t ) = Y 1 ( t ) + Y 2 ( t ) , from Equations (2), (7), (13)–(16), (23) and (24) we obtain
ϕ α . α . = α . ϕ α . = Y 1 ψ Y 1 + Y 1 ω Y 1 + Y 2 ψ Y 1 + Y 2 ω Y 1 + Y 1 B Y 2 + Y 1 C Y 2 + Y 2 B Y 2 + Y 2 C Y 2 = V α . B Y 2 + V α . ψ Y 1 + ( T Y 1 + A Y 2 ) C Y 2 + ( T Y 1 + A Y 2 ) ω Y 1 + H α . C Y 2 + H α . ω Y 1 + ( T Y 1 + A Y 2 ) B Y 2 + ( A Y 2 + T Y 1 ) ψ Y 1 .
Taking the vertical and horizontal components in the above equation, we have
V ϕ α . α . = V α . B Y 2 + V α . ψ Y 1 + ( T Y 1 + A Y 2 ) C Y 2 + ( T Y 1 + A Y 2 ) ω Y 1 ,
H ϕ α . α . = H α . C Y 2 + H α . ω Y 1 + ( T Y 1 + A Y 2 ) B Y 2 + ( A Y 2 + T Y 1 ) ψ Y 1 .
Thus, α is a geodesic on M 1 if and only if V ϕ α . α . = 0 and H ϕ α . α . = 0 ; this completes the proof. □
Theorem 2.
Let Π be an SIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) . Then, Π is a CSIR map with r = e f if and only if
g 1 ( f , V 2 ) | | V 1 | | 2 = g 1 ( V α . B V 2 , ψ V 1 ) + g 1 ( T V 1 + A V 2 ) C V 2 , ψ V 1 ) + g 1 ( H α . C V 2 , ω V 1 ) + g 1 ( ( T V 1 + A V 2 ) B V 2 , ω V 1 ) ,
where α : I 2 M 1 is a geodesic on M 1 ; V 1 ( t ) and V 2 ( t ) are vertical and horizontal components of α . ( t ) , respectively.
Proof. 
Let α : I 2 M 1 be a geodesic on M 1 with V 1 ( t ) = V α . ( t ) and V 2 ( t ) = H α . ( t ) . We denote the angle in [ 0 , π ] between α . ( t ) and V 2 ( t ) by θ ( t ) . Assuming υ = | | α . ( t ) | | 2 , then we obtain
g 1 ( V 1 ( t ) , V 1 ( t ) ) = υ sin 2 θ ( t ) ,
g 1 ( V 2 ( t ) , V 2 ( t ) ) = υ cos 2 θ ( t ) .
Now, differentiating (38), we obtain
d d t g 1 ( V 1 ( t ) , V 1 ( t ) ) = 2 υ sin θ ( t ) cos θ ( t ) d θ d t , g 1 ( α . V 1 ( t ) , V 1 ( t ) ) = υ sin θ ( t ) cos θ ( t ) d θ d t .
Using Equations (4) and (7) in the above equation, we obtain
g 1 ( α . ϕ V 1 ( t ) , ϕ V 1 ( t ) ) = υ sin θ ( t ) cos θ ( t ) d θ d t .
Thus, we obtain
g 1 ( α . ϕ V 1 , ϕ V 1 ) = g 1 ( V α . ψ V 1 , ψ V 1 ) + g 1 ( H α . ω V 1 , ω V 1 ) + g 1 ( ( T V 1 + A V 2 ) ψ V 1 , ω V 1 ) + g 1 ( ( T V 1 + A V 2 ) ω V 1 , ψ V 1 ) .
Using Equations (36) and (37) in (41), we have
g 1 ( α . ϕ V 1 , ϕ V 1 ) = g 1 ( V α . B V 2 , ψ V 1 ) g 1 ( T V 1 + A V 2 ) C V 2 , ψ V 1 ) g 1 ( H α . C V 2 , ω V 1 ) g 1 ( ( T V 1 + A V 2 ) B V 2 , ω V 1 ) .
From Equations (40) and (42), we have
υ sin θ ( t ) cos θ ( t ) d θ d t = g 1 ( V α . B V 2 , ψ V 1 ) g 1 ( T V 1 + A V 2 ) C V 2 , ψ V 1 ) g 1 ( H α . C V 2 , ω V 1 ) g 1 ( ( T V 1 + A V 2 ) B V 2 , ω V 1 ) .
Moreover, π is a Clairaut semi-invariant Riemannian map with r = e f if and only if d d t ( e f α sin θ ) = 0 , i.e., e f α ( cos θ d θ d t + sin θ d f d t ) = 0 , which, by multiplying with nonzero factor υ sin θ , gives
υ cos θ sin θ d θ d t = υ sin 2 θ d f d t υ cos θ sin θ d θ d t = g 1 ( V 1 , V 1 ) d f d t υ cos θ sin θ d θ d t = g 1 ( f , α . ) | | V 1 | | 2 υ cos θ sin θ d θ d t = g 1 ( f , V 2 ) | | V 1 | | 2 .
Thus, from Equations (43) and (44) we have
g 1 ( f , V 2 ) | | V 1 | | 2 = g 1 ( V α . B V 2 , ψ V 1 ) + g 1 ( T V 1 + A V 2 ) C V 2 , ψ V 1 ) + g 1 ( H α . C V 2 , ω V 1 ) + g 1 ( ( T V 1 + A V 2 ) B V 2 , ω V 1 ) .
Hence, Theorem 2 is proved. □
Corollary 2.
Let Π be an SIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) admitting horizontal Reeb vector field. Then, we obtain
g 1 ( f , ξ ) = 0 .
Theorem 3.
Let Π be a CSIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) with r = e f , then at least one of the following statement is true:
( i ) f is constant on ϕ ( D 2 ) ;
( i i ) The fibers are one-dimensional;
( i i i ) Π ϕ U 1 Π * ( Z 1 ) = Z 1 ( f ) Π * ( ϕ U 1 ) , for all U 1 Γ ( D 2 ) , Z 1 Γ ( μ ) and ξ Z 1 .
Proof. 
Let Π be a CSIR map from a cosymplectic manifold to a Riemannian manifold. Then, for V 1 , V 2 Γ ( D 2 ) , using Equation (18) and Theorem 1 we obtain
T V 1 V 2 = g 1 ( V 1 , V 2 ) g r a d f .
Taking the inner product of Equation (45) with ϕ U 1 , we have
g 1 ( T V 1 V 2 , ϕ U 1 ) = g 1 ( V 1 , V 2 ) g 1 ( g r a d f , ϕ U 1 )
for all U 1 Γ ( D 2 ) .
From Equations (4), (7), (13) and (46) we obtain
g 1 ( V 1 ϕ V 2 , U 1 ) = g 1 ( V 1 , V 2 ) g 1 ( g r a d f , ϕ U 1 ) .
Since ∇ is a metric connection, by using Equations (14) and (45) in the above equation, we obtain
g 1 ( V 1 , U 1 ) g 1 ( g r a d f , ϕ V 2 ) = g 1 ( V 1 , V 2 ) g 1 ( g r a d f , ϕ U 1 ) .
Taking U 1 = V 2 and interchanging the role of V 1 and V 2 , we obtain
g 1 ( V 2 , V 2 ) g 1 ( g r a d f , ϕ V 1 ) = g 1 ( V 1 , V 2 ) g 1 ( g r a d f , ϕ V 2 ) .
From Equations (47) and (48), we obtain
g 1 ( g r a d f , ϕ V 1 ) = ( g 1 ( V 1 , V 2 ) ) 2 | | V 1 | | 2 | | V 2 | | 2 g 1 ( g r a d f , ϕ V 1 ) .
If g r a d f Γ ( ϕ ( D 2 ) ) , then Equation (49) and the condition of equality in the Schwarz inequality imply that either f is constant on ϕ ( D 2 ) or the fibers are one-dimensional. This implies the proof of ( i ) and ( i i ) .
Now, from Equations (13) and (45), we obtain
g 1 ( V 1 U 1 , Z 1 ) = g 1 ( V 1 , U 1 ) g 1 ( g r a d f , Z 1 ) ,
for all Z 1 Γ ( μ ) and ξ Z 1 . Using Equations (4), (7) and (50) we have
g 1 ( V 1 ϕ U 1 , ϕ Z 1 ) = g 1 ( V 1 , U 1 ) g 1 ( g r a d f , Z 1 ) ,
which implies
g 1 ( ϕ U 1 V 1 , ϕ Z 1 ) = g 1 ( V 1 , U 1 ) g 1 ( g r a d f , Z 1 ) .
Since ∇ is a metric connection, then by using Equations (47) and (51) we have
g 1 ( H ϕ U 1 Z 1 , ϕ V 1 ) = g 1 ( ϕ V 1 , ϕ U 1 ) g 1 ( g r a d f , Z 1 ) .
In addition, for the Riemannian map Π , we have
g 2 ( Π * ( ϕ U 1 M 1 Z 1 ) , Π * ( ϕ V 1 ) ) = g 2 ( Π * ( ϕ V 1 ) , Π * ( ϕ U 1 ) ) g 1 ( g r a d f , Z 1 ) .
Again, using Equations (19), (21) and (52) we obtain
g 2 ( Π ϕ U 1 Π * ( Z 1 ) , Π * ( ϕ V 1 ) ) = g 2 ( Π * ( ϕ V 1 ) , Π * ( ϕ U 1 ) ) g 1 ( g r a d f , Z 1 ) ,
which implies.
Π ϕ U 1 Π * ( Z 1 ) = Z 1 ( f ) Π * ( ϕ U 1 ) .
If g r a d f Γ ( μ ) { ξ } , then (53) implies ( i i i ) . This completes the proof. □
Corollary 3.
Let Π be a CSIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) with r = e f and dim ( D 2 ) > 1 . Then, the fibers of Π are totally geodesic if and only if Π ϕ U 1 Π * ( Z 1 ) = 0 U 1 Γ ( D 2 ) and Z 1 Γ ( μ ) .
Lemma 5.
Let Π be a CSIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) with r = e f and dim ( D 2 ) > 1 . Then, Π Z 1 Π * ( ϕ X 1 ) = Z 1 ( f ) Π * ( ϕ X 1 ) X 1 Γ ( D 2 ) and Z 1 Γ ( ker Π * ) { ξ } .
Proof. 
Let Π be a CSIR map from a cosymplectic manifold to a Riemannian manifold. From Theorem 1 , fibers are totally umbilical with mean curvature vector field H = g r a d f , then we have
g 1 ( X 1 Z 1 , X 2 ) = g 1 ( X 1 X 2 , Z 1 ) g 1 ( X 1 Z 1 , X 2 ) = g 1 ( X 1 , X 2 ) g 1 ( g r a d f , Z 1 )
for all X 1 , X 2 Γ ( D 2 ) and Z 1 Γ ( ker Π * ) { ξ } .
Using Equations (4) and (7) in the above equation, we obtain
g 1 ( Z 1 ϕ X 1 , ϕ X 2 ) = g 1 ( ϕ X 1 , ϕ X 2 ) g 1 ( g r a d f , Z 1 ) .
Since Π is an SIR map, by using Equation (54) we have
g 2 ( Π * ( Z 1 Π ϕ X 1 ) , Π * ( ϕ X 2 ) ) = g 2 ( Π * ( ϕ X 1 ) , Π * ( ϕ X 2 ) ) g 1 ( g r a d f , Z 1 ) .
From (19) and (55) we obtain
g 2 ( Π Z 1 Π * ( ϕ X 1 ) , Π * ( ϕ X 2 ) ) = g 2 ( Π * ( ϕ X 1 ) , Π * ( ϕ X 2 ) ) g 1 ( g r a d f , Z 1 ) ,
which implies Π Z 1 Π * ( ϕ X 1 ) = Z 1 ( f ) Π * ( ϕ X 1 ) X 1 Γ ( D 2 ) and Z 1 Γ ( ker Π * ) { ξ } .
Corollary 4.
Let Π be a CSIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) with r = e f and dim ( D 2 ) > 1 . Then, Π Z 1 Π * ( ϕ X 1 ) = 0 X 1 Γ ( D 2 ) and Z 1 = ξ Γ ( ker Π * ) .
Theorem 4.
Let Π be a CSIR map with r = e f from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) . If T is not identically zero, then the invariant distribution D 1 does not define a totally geodesic foliation on M 1 .
Proof. 
For V 1 , V 2 Γ ( D 1 ) and Z 1 Γ ( D 2 ) , using Equations (4), (7), (13) and (18) we obtain
g 1 ( V 1 V 2 , Z 1 ) = g 1 ( V 1 ϕ V 2 , ϕ Z 1 ) = g 1 ( T V 1 ϕ V 2 , ϕ Z 1 ) = g 1 ( V 1 , ϕ V 2 ) g 1 ( g r a d f , ϕ Z 1 ) .
Thus, the assertion can be seen from the above equation and the fact that g r a d f ϕ ( D 2 ) .
Theorem 5.
Let Π be a CSIR map with r = e f from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) . Then, the fibers of Π are totally geodesic, or anti-invariant distribution D 2 is one-dimensional.
Proof. 
If the fibers of Π are totally geodesic, it is obvious. For the second one, since Π is a Clairaut proper semi-invariant Riemannian map, then either dim ( D 2 ) = 1 or dim ( D 2 ) > 1 . If dim ( D 2 ) > 1 , then we can choose V 1 , V 2 Γ ( D 2 ) such that { V 1 , V 2 } is orthonormal. From Equations (14), (23) and (24) we obtain
T V 1 ϕ V 2 + H V 1 ϕ V 2 = V 1 ϕ V 2 T V 1 ϕ V 2 + H V 1 ϕ V 2 = B T V 1 V 2 + C T V 1 V 2 + ψ V V 1 V 2 + ω V V 1 V 2 .
Taking the inner product of the above equation with V 1 , we obtain
g 1 ( T V 1 ϕ V 2 , V 1 ) = g 1 ( B T V 1 V 2 , V 1 ) + g 1 ( ψ V V 1 V 2 , V 1 ) .
From Equation (7) we have
g 1 ( T V 1 V 1 , ϕ V 2 ) = g 1 ( T V 1 ϕ V 2 , V 1 ) = g 1 ( T V 1 V 2 , ϕ V 1 ) .
Now, using Equations (18) and (58) we obtain
g 1 ( T V 1 V 1 , ϕ V 2 ) = g 1 ( g r a d f , ϕ V 2 ) .
From Equations (18), (58) and (59) we obtain
g 1 ( g r a d f , ϕ V 2 ) = g 1 ( T V 1 V 1 , ϕ V 2 ) = g 1 ( T V 1 ϕ V 2 , V 1 ) = g 1 ( T V 1 V 2 , ϕ V 1 ) ,
from which we obtain
g 1 ( g r a d f , ϕ V 2 ) = g 1 ( T V 1 V 2 , ϕ V 1 ) g 1 ( g r a d f , ϕ V 2 ) = g 1 ( V 1 , V 2 ) g 1 ( g r a d f , ϕ V 1 ) g 1 ( g r a d f , ϕ V 2 ) = 0 .
Thus, we obtain
g r a d f ϕ ( D 2 ) .
Therefore, the dimension of D 2 must be one. □
Theorem 6.
Let Π be a CSIR map from a cosymplectic manifold ( M 1 , ϕ , ξ , η , g 1 ) to a Riemannian manifold ( M 2 , g 2 ) with r = e f and dim ( D 2 ) > 1 . Then, we obtain
κ = 1 ω g 1 ( A X 1 x κ , A X 1 x κ ) = κ = 1 ω g 2 ( X 1 Π Π * ( ϕ x κ ) , X 1 Π Π * ( ϕ x κ ) ) ,
i = 1 β + s g 2 ( ( Π * ) ( F i , X 1 ) , ( Π * ) ( X 1 , F i ) ) = l = 1 s g 2 ( ( Π * ) ( ϑ l , X 1 ) , ( Π * ) ( X 1 , ϑ l ) ) ,
j = 1 β g 1 ( A X 1 w j , A X 1 w j ) = ( X 1 ( f ) ) 2 j = 1 β g 1 ( w j , w j ) ,
X 1 Γ ( ker Π * ) { ξ } , where { x 1 , x 2 , . . . . , x ω } , { w 1 , w 2 , . . . . , w β } , { F 1 , F 2 , . . . . , F β + s } and
{ ϑ 1 , ϑ 2 , . . . . ϑ s } are orthonormal frames of D 1 , D 2 , ϕ ( D 2 ) μ and μ, respectively.
Proof. 
Let Π : ( M 1 , ϕ , ξ , η , g 1 ) ( M 2 , g 2 ) be a CSIR map, then for all X 1 Γ ( ker Π * ) { ξ } , we have
κ = 1 ω g 1 ( A X 1 x κ , A X 1 x κ ) = κ = 1 ω g 1 ( H X 1 ϕ x κ , H X 1 ϕ x κ ) .
Since Π is a Riemannian map, in view of Equation (19), Equation (64) transforms to
κ = 1 ω g 1 ( A X 1 x κ , A X 1 x κ ) = κ = 1 ω g 2 ( Π * ( X 1 M 1 ϕ x κ ) , Π * ( X 1 M 1 ϕ x κ ) ) = κ = 1 ω g 2 ( X 1 Π Π * ( ϕ x κ ) , X 1 Π Π * ( ϕ x κ ) ) .
Now, for all X 1 Γ ( ker Π * ) { ξ } , we obtain
i = 1 β + s g 2 ( ( Π * ) ( F i , X 1 ) , ( Π * ) ( X 1 , F i ) ) = j = 1 β l = 1 s g 2 ( ( Π * ) ( ϕ w j + ϑ l , X 1 ) , ( Π * ) ( X 1 , ϕ w j + ϑ l ) ) .
Since ϕ w j Γ ( ker Π * ) and ( Π * ) is linear, from the above equation, we have
i = 1 β + s g 2 ( ( Π * ) ( F i , X 1 ) , ( Π * ) ( F i , X 1 ) ) = j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) + j = 1 β l = 1 s g 2 ( ( Π * ) ( ϑ l , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) + j = 1 β l = 1 s g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϑ l ) ) + l = 1 s g 2 ( ( Π * ) ( ϑ l , X 1 ) , ( Π * ) ( X 1 , ϑ l ) ) .
Thus, (61) holds.
On the other side, using (19) in the first term of the right-hand side of (65), we have
j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) = j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , X 1 Π Π * ( ϕ w j ) Π * ( X 1 M 1 ϕ w j ) ) ,
which, by using Equations (4), (7) and (65), turns into
j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) = j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , X 1 Π Π * ( ϕ w j ) ) .
Now, by using Lemma 4 in (66) we obtain
j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) = j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , X 1 ( f ) Π * ( ϕ w j ) ) .
This implies that
j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) = j = 1 β X 1 ( f ) g 2 ( ( Π * ) ( ϕ w j , X 1 ) , Π * ( ϕ w j ) ) .
By using Equation (20) in the above equation, it follows that
j = 1 β g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) = 0 .
Similarly, we find
j = 1 β l = 1 s g 2 ( ( Π * ) ( ϑ l , X 1 ) , ( Π * ) ( X 1 , ϕ w j ) ) = 0 ,
j = 1 β l = 1 s g 2 ( ( Π * ) ( ϕ w j , X 1 ) , ( Π * ) ( X 1 , ϑ l ) ) = 0 .
Thus, by using Equations (67)–(69) in Equation (65) we obtain (62).
Further, for X 1 Γ ( ker Π * ) { ξ } , we obtain
j = 1 β g 1 ( A X 1 w j , A X 1 w j ) = j = 1 β g 1 ( H X 1 w j , H X 1 w j ) = j = 1 β g 1 ( H X 1 ϕ w j , H X 1 ϕ w j ) .
Since Π is a Riemannian map, in view of Equation (19), the above equation becomes
j = 1 β g 1 ( A X 1 w j , A X 1 w j ) = j = 1 β { g 2 ( ( Π * ) ( X 1 , ϕ w j ) , ( Π * ) ( X 1 , ϕ w j ) ) 2 g 2 ( ( Π * ) ( X 1 , ϕ w j ) , Π X 1 Π * ( ϕ w j ) ) + g 2 ( Π X 1 Π * ( ϕ w j ) , X 1 Π Π * ( ϕ w j ) ) } ,
which, by using Lemma 4 and Equations (21) and (67) in (70) we obtain
j = 1 β g 1 ( A X 1 w j , A X 1 w j ) = j = 1 β g 2 ( X 1 ( f ) Π * ( ϕ w j ) , X 1 ( f ) Π * ( ϕ w j ) ) = ( X 1 ( f ) ) 2 j = 1 β g 2 ( Π * ( ϕ w j ) , Π * ( ϕ w j ) ) .
Since ϕ w j Γ ( ker Π * ) and Π is a Riemannian map, from (71) we obtain
j = 1 β g 1 ( A X 1 w j , A X 1 w j ) = ( X 1 ( f ) ) 2 j = 1 β g 1 ( w j , w j ) .
Thus, from Equations (4) and (72) we obtain (63). □

4. Example

Let M 1 be a differentiable manifold given by M 1 = { ( x 1 , x 2 , x 3 , x 4 , x 5 , x 6 , x 7 ) R 7 : x 7 > 0 } . We define the Riemannian metric g 1 on M 1 by g 1 = e 2 x 7 d x 1 2 + e 2 x 7 d x 2 2 + e 2 x 7 d x 3 2 + e 2 x 7 d x 4 2 + e 2 x 7 d x 5 2 + e 2 x 7 d x 6 2 + d x 7 2 , and the cosymplectic structure ( ϕ , ξ , η , g 1 ) on M 1 is defined as
ϕ ( x 1 x 1 + x 2 x 2 + x 3 x 3 + x 4 x 4 + x 5 x 5 + x 6 x 6 + x 7 x 7 ) = ( x 4 x 1 + x 5 x 2 + x 6 x 3 x 1 x 4 x 2 x 5 x 3 x 6 ) ,
ξ = x 7 , η = d x 7 , and g 1 was earlier defined.
Let M 2 = { ( v 1 , v 2 , v 3 , v 4 ) R 4 } be a Riemannian manifold with Riemannian metric g 2 on M 2 given by g 2 = e 2 x 7 d v 1 2 + e 2 x 7 d v 2 2 + e 2 x 7 d v 3 2 + d v 4 2 . Define a map Π : R 7 R 4 by
Π ( x 1 , x 2 , x 3 , x 4 , x 5 , x 6 , x 7 ) = ( x 2 x 5 2 , 101 , x 6 , x 7 ) .
Then, we have
ker Π * = D 1 D 2 ,
where
D 1 = < V 1 = e 1 , V 2 = e 4 > , D 2 = < V 3 = e 2 + e 5 , V 4 = e 3 > ,
and
( ker Π * ) = < H 1 = e 2 e 5 , H 2 = e 6 , H 3 = e 7 > ,
where { e 1 = e x 7 x 1 , e 2 = e x 7 x 2 , e 3 = e x 7 x 3 , e 4 = e x 7 x 4 , e 5 = e x 7 x 5 , e 6 = e x 7 x 6 , e 7 = x 7 } , { e 1 * = v 1 , e 2 * = v 2 , e 3 * = v 3 , e 4 * = v 4 } are bases on T p M 1 and T Π ( p ) M 2 , respectively, for all p M 1 . By direct computations, we can see that Π * ( H 1 ) = 2 e x 7 e 1 * , Π * ( H 2 ) = e x 7 e 2 * , Π * ( H 3 ) = e 3 * and g 1 ( H i , H j ) = g 2 ( Π * H i , Π * H j ) for all H i , H j Γ ( ker Π * ) , i , j = 1 , 2 , 3 . Thus, Π is a Riemannian map with ( r a n g e Π * ) = < e 4 * > . Moreover, it is easy to see that ϕ V 3 = H 1 and ϕ V 4 = H 2 . Therefore, Π is an SIR map.
Now, we will find the smooth function f on M 1 satisfying T V V = g 1 ( V , V ) f V Γ ( ker Π * ) . The covariant derivative for the vector fields E = E i x i , F = F j x j on M 1 is defined as
E F = E i F j x i x j + E i F j x i x j ,
where the covariant derivative of basis vector fields x j and x i is defined by
x i x j = Γ i j k x k ,
and Christoffel symbols are defined by
Γ i j k = 1 2 g k l ( g 1 j l x i + g 1 i l x j g 1 i j x l ) .
Thus, we obtain
g 1 i j = e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 1 , g 1 i j = e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 e 2 x 7 0 0 0 0 0 0 0 1 .
By using Equations (75) and (76) we find
Γ 11 1 = Γ 11 2 = Γ 11 3 = Γ 11 4 = Γ 11 5 = Γ 11 6 = 0 , Γ 11 7 = e 2 x 7 , Γ 22 1 = Γ 22 2 = Γ 22 3 = Γ 22 4 = Γ 22 5 = Γ 22 6 = 0 , Γ 22 7 = e 2 x 7 , Γ 33 1 = Γ 33 2 = Γ 33 3 = Γ 33 4 = Γ 33 5 = Γ 33 6 = 0 , Γ 33 7 = e 2 x 7 , Γ 44 1 = Γ 44 2 = Γ 44 3 = Γ 44 4 = Γ 44 5 = Γ 44 6 = 0 , Γ 44 7 = e 2 x 7 , Γ 55 1 = Γ 55 2 = Γ 55 3 = Γ 55 4 = Γ 55 5 = Γ 55 6 = 0 , Γ 55 7 = e 2 x 7 , Γ 12 1 = Γ 12 2 = Γ 12 3 = Γ 12 4 = Γ 12 5 = Γ 12 6 = Γ 12 7 = 0 , Γ 21 1 = Γ 21 2 = Γ 21 3 = Γ 21 4 = Γ 21 5 = Γ 21 6 = Γ 21 7 = 0 , Γ 13 1 = Γ 13 2 = Γ 13 3 = Γ 13 4 = Γ 13 5 = Γ 13 6 = Γ 13 7 = 0 , Γ 31 1 = Γ 31 2 = Γ 31 3 = Γ 31 4 = Γ 31 5 = Γ 31 6 = Γ 31 7 = 0 , Γ 14 1 = Γ 14 2 = Γ 14 3 = Γ 14 4 = Γ 14 5 = Γ 14 6 = Γ 14 7 = 0 , Γ 41 1 = Γ 41 2 = Γ 41 3 = Γ 41 4 = Γ 41 5 = Γ 41 6 = Γ 41 7 = 0 , Γ 15 1 = Γ 15 2 = Γ 15 3 = Γ 15 4 = Γ 15 5 = Γ 15 6 = Γ 15 7 = 0 , Γ 51 1 = Γ 51 2 = Γ 51 3 = Γ 51 4 = Γ 51 5 = Γ 51 6 = Γ 51 7 = 0 , Γ 23 1 = Γ 23 2 = Γ 23 3 = Γ 23 4 = Γ 23 5 = Γ 23 6 = Γ 23 7 = 0 , Γ 32 1 = Γ 32 2 = Γ 32 3 = Γ 32 4 = Γ 32 5 = Γ 32 6 = Γ 32 7 = 0 , Γ 24 1 = Γ 24 2 = Γ 24 3 = Γ 24 4 = Γ 24 5 = Γ 24 6 = Γ 24 7 = 0 , Γ 42 1 = Γ 42 2 = Γ 42 3 = Γ 42 4 = Γ 42 5 = Γ 42 6 = Γ 42 7 = 0 , Γ 25 1 = Γ 25 2 = Γ 25 3 = Γ 25 4 = Γ 25 5 = Γ 25 6 = Γ 25 7 = 0 , Γ 52 1 = Γ 52 2 = Γ 52 3 = Γ 52 4 = Γ 52 5 = Γ 52 6 = Γ 52 7 = 0 , Γ 34 1 = Γ 34 2 = Γ 34 3 = Γ 34 4 = Γ 34 5 = Γ 34 6 = Γ 34 7 = 0 , Γ 43 1 = Γ 43 2 = Γ 43 3 = Γ 43 4 = Γ 43 5 = Γ 43 6 = Γ 43 7 = 0 , Γ 35 1 = Γ 35 2 = Γ 35 3 = Γ 35 4 = Γ 35 5 = Γ 35 6 = Γ 35 7 = 0 , Γ 53 1 = Γ 53 2 = Γ 53 3 = Γ 53 4 = Γ 53 5 = Γ 53 6 = Γ 53 7 = 0 , Γ 45 1 = Γ 45 2 = Γ 45 3 = Γ 45 4 = Γ 45 5 = Γ 45 6 = Γ 45 7 = 0 , Γ 54 1 = Γ 54 2 = Γ 54 3 = Γ 54 4 = Γ 54 5 = Γ 54 6 = Γ 54 7 = 0 .
Using Equations (73), (74) and (77) we calculate
e 1 e 1 = e 2 e 2 = e 3 e 3 = e 4 e 4 = x 7 , e 1 e 2 = e 1 e 3 = e 1 e 4 = e 2 e 1 = e 2 e 3 = e 2 e 4 = 0 , e 3 e 1 = e 3 e 2 = e 3 e 4 = e 4 e 1 = e 4 e 2 = e 4 e 3 = 0 .
Thus, we find
V 1 V 1 = e 1 e 1 = x 7 , V 2 V 2 = e 4 e 4 = x 7 , V 3 V 3 = e 2 + e 5 e 2 + e 5 = 2 x 7 , V 4 V 4 = e 3 e 3 = x 7 .
Now, from
T V V = T λ 1 V 1 + λ 2 V 2 + λ 3 V 3 + λ 4 V 4 λ 1 V 1 + λ 2 V 2 + λ 3 V 3 + λ 4 V 4 , λ 1 , λ 2 , λ 3 , λ 4 R ,
we lead to
T V V = λ 1 2 T V 1 V 1 + λ 2 2 T V 2 V 2 + λ 3 2 T V 3 V 3 + λ 4 2 T V 4 V 4 + 2 λ 1 λ 2 T V 1 V 2 + 2 λ 1 λ 3 T V 1 V 3 + 2 λ 1 λ 4 T V 1 V 4 + 2 λ 2 λ 3 T V 2 V 3 + 2 λ 2 λ 4 T V 2 V 4 + 2 λ 3 λ 4 T V 3 V 4 .
From Equations (13) and (79), we obtain
T V 1 V 1 = x 7 , T V 2 V 2 = x 7 , T V 3 V 3 = 2 x 7 , T V 4 V 4 = x 7 , T V 1 V 2 = 0 , T V 1 V 3 = 0 , T V 1 V 4 = 0 , T V 2 V 3 = 0 , T V 2 V 4 = 0 , T V 3 V 4 = 0 .
Thus, by using Equations (80) and (81) we obtain
T V V = ( λ 1 2 + λ 2 2 + 2 λ 3 2 + λ 1 4 ) x 7 .
Since V = λ 1 V 1 + λ 2 V 2 + λ 3 V 3 + λ 4 V 4 , g 1 ( λ 1 V 1 + λ 2 V 2 + λ 3 V 3 + λ 4 V 4 , λ 1 V 1 + λ 2 V 2 + λ 3 V 3 + λ 4 V 4 ) = λ 1 2 + λ 2 2 + 2 λ 3 2 + λ 1 4 . For any smooth function f on R 7 , the gradient of f with respect to the metric g 1 is given by f = i , j = 1 7 g 1 i j f x i x j . Hence, f = e 2 x 7 f x 1 x 1 + e 2 x 7 f x 2 x 2 + e 2 x 7 f x 3 x 3 + e 2 x 7 f x 4 x 4 + e 2 x 7 f x 5 x 5 + e 2 x 7 f x 6 x 6 + f x 7 x 7 . Hence, f = x 7 for the function f = x 7 . Then, it is easy to see that T V V = g 1 ( V , V ) f ; thus, by Theorem 1, Π is a CSIR map from cosymplectic manifold onto Riemannian manifold.

5. Conclusions

In the last few years, Riemannian maps have been extensively studied between different kinds of the manifolds. Recently, a special type of Riemannian map, namely, the "Clairaut Riemannian map" was introduced and studied by Sahin [30]; moreover, he, in [37], gave an open problem to find characterizations for Clairaut Riemannian maps. As a continuation of this study, we tried to study Clairaut semi-invariat Riemannian maps in contact geometry. Here, we investigated the various most fundamental geometric properties on the fibers and distributions of these maps. In the future, we plan to focus on studying Clairaut’s semi-slant Riemannian maps, Clairaut’s hemi-slant Riemannian maps, and Clairaut’s bi-slant Riemannian maps between different kinds of the manifolds.

Author Contributions

Conceptualization, Y.L., R.P., A.H. and S.K. (Sushil Kumar); methodology, Y.L., R.P., A.H. and S.K. (Sumeet Kumar); investigation, R.P., A.H., S.K. (Sushil Kumar) and S.K. (Sumeet Kumar); writing—original draft preparation, Y.L., A.H., S.K. (Sushil Kumar) and S.K. (Sumeet Kumar); writing—review and editing, Y.L., R.P., S.K. (Sushil Kumar) and S.K. (Sumeet Kumar). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (Grant No. 12101168) and Zhejiang Provincial Natural Science Foundation of China (Grant No. LQ22A010014).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully thank the reviewers for the constructive comments to improve the quality of the paper.

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. Invariant and anti-invariant Riemannian maps to Kahler manifolds. Int. J. Geom. Methods Mod. Phys. 2010, 7, 337–355. [Google Scholar] [CrossRef]
  3. Şahin, B. Semi-invariant Riemannian maps from almost Hermitian manifolds. Indag. Math. 2012, 23, 80–94. [Google Scholar] [CrossRef]
  4. Prasad, R.; Kumar, S. Slant Riemannian maps from Kenmotsu manifolds into Riemannian manifolds. Glob. J. Pure Appl. Math. 2017, 13, 1143–1155. [Google Scholar]
  5. Prasad, R.; Pandey, S. Slant Riemannian maps from an almost contact manifold. Filomat 2017, 31, 3999–4007. [Google Scholar] [CrossRef]
  6. Şahin, B. Slant Riemannian maps from almost Hermitian manifolds. Quaest. Math. 2013, 36, 449–461. [Google Scholar] [CrossRef]
  7. Kumar, S.; Prasad, R. Semi-slant Riemannian maps from Cosymplectic manifolds into Riemannian manifolds. Gulf J. Math. 2020, 9, 62–80. [Google Scholar] [CrossRef]
  8. Park, K.S.; Şahin, B. Semi-slant Riemannian maps into almost Hermitian manifolds. Czechoslov. Math. J. 2014, 64, 1045–1061. [Google Scholar] [CrossRef]
  9. Prasad, R.; Kumar, S. Semi-slant Riemannian maps from almost contact metric manifolds into Riemannian manifolds. Tbilisi Math. J. 2018, 11, 19–34. [Google Scholar] [CrossRef]
  10. Şahin, B. Hemi-slant Riemannian Maps. Mediterr. J. Math. 2017, 14, 10. [Google Scholar] [CrossRef]
  11. Prasad, R.; Kumar, S.; Kumar, S.; Vanli, A.T. On quasi-hemi-slant Riemannian maps. Gazi Univ. J. Sci. 2021, 34, 477–491. [Google Scholar] [CrossRef]
  12. Bishop, R.L. Clairaut submersions. In Differential Geometry (In Honor of Kentaro Yano); Kinokuniya: Tokyo, Japan, 1972; pp. 21–31. [Google Scholar]
  13. Lee, J.; Park, J.H.; Şahin, B.; Song, D.Y. Einstein conditions for the base of anti-invariant Riemannian submersions and Clairaut submersions. Taiwan. J. Math. 2015, 19, 1145–1160. [Google Scholar] [CrossRef]
  14. Taştan, H.M. Lagrangian submersions from normal almost contact manifolds. Filomat 2017, 31, 3885–3895. [Google Scholar] [CrossRef]
  15. Taştan, H.M.; Aydin, S.G. Clairaut anti-invariant submersions from cosymplectic manifolds. Honam Math. J. 2019, 41, 707–724. [Google Scholar]
  16. Allison, D. Lorentzian Clairaut submersions. Geom. Dedicata 1996, 63, 309–319. [Google Scholar] [CrossRef]
  17. Aso, K.; Yorozu, S. A generalization of Clairaut’s theorem and umbilic foliations. Nihonkai Math. J. 1991, 2, 139–153. [Google Scholar]
  18. Gauchman, H. On a decomposition of Riemannian manifolds. Houst. J. Math. 1981, 7, 365–372. [Google Scholar]
  19. Haseeb, A.; Prasad, R.; Chaubey, S.K.; Vanli, A.T. A note on *-conformal and gradient *-conformal η-Ricci solitons in α-cosymplectic manifolds. Honam Math. J. 2022, 44, 231–243. [Google Scholar]
  20. Kumar, S.; Prasad, R.; Kumar, S. Clairaut semi-invariant Riemannian maps from almost Hermitian manifolds. Turk. J. Math. 2022, 46, 1193–1209. [Google Scholar] [CrossRef]
  21. Li, Y.; Abolarinwa, A.; Azami, S.; Ali, A. Yamabe constant evolution and monotonicity along the conformal Ricci flow. AIMS Math. 2022, 7, 12077–12090. [Google Scholar] [CrossRef]
  22. Li, Y.; Ali, A.; Mofarreh, F.; Abolarinwa, A.; Ali, R. Some eigenvalues estimate for the ϕ-Laplace operator on slant submanifolds of Sasakian space forms. J. Funct. Space 2021, 2021, 6195939. [Google Scholar]
  23. Li, Y.; Ali, A.; Mofarreh, F.; Alluhaibi, N. Homology groups in warped product submanifolds in hyperbolic spaces. J. Math. 2021, 2021, 8554738. [Google Scholar] [CrossRef]
  24. Li, Y.; 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]
  25. Li, Y.; Khatri, M.; Singh, J.P.; Chaubey, S.K. Improved Chen’s Inequalities for Submanifolds of Generalized Sasakian-Space-Forms. Axioms 2022, 11, 324. [Google Scholar] [CrossRef]
  26. Li, Y.; Lone, M.A.; Wani, U.A. Biharmonic submanifolds of Kähler product manifolds. AIMS Math. 2021, 6, 9309–9321. [Google Scholar] [CrossRef]
  27. Li, Y.; Mofarreh, F.; Dey, S.; Roy, S.; Ali, A. General Relativistic Space-Time with η1-Einstein Metrics. Mathematics 2022, 10, 2530. [Google Scholar] [CrossRef]
  28. Li, Y.; Wang, Z.G.; Zhao, T.H. Geometric Algebra of Singular Ruled Surfaces. Adv. Appl. Clifford Algebr. 2021, 31, 19. [Google Scholar] [CrossRef]
  29. Prasad, R.; Shukla, S.S.; Haseeb, A.; Kumar, S. Quasi hemi-slant submanifolds of Kaehler manifolds. Honam Math. J. 2020, 42, 795–809. [Google Scholar]
  30. Şahin, B. Circles along a Riemannian map and Clairaut Riemannian maps. Bull. Korean Math. Soc. 2017, 54, 253–264. [Google Scholar] [CrossRef]
  31. Taştan, H.M.; Aydin, S.G. Clairaut anti-invariant submersions from Sasakian and Kenmotsu manifolds. Mediterr. J. Math. 2017, 14, 235–249. [Google Scholar] [CrossRef]
  32. Yadav, A.; Meena, K. Clairaut anti-invariant Riemannian maps from Kahler Manifolds. Mediterr. J. Math. 2022, 19, 97. [Google Scholar] [CrossRef]
  33. Şahin, B. Riemannian Submersions, Riemannian Maps in Hermitian Geometry, and Their Applications; Elsevier/Academic Press: Amsterdam, The Netherlands, 2017. [Google Scholar]
  34. Blair, D.E. Contact Manifolds in Riemannian Geometry; Lecture Notes in Math 509; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 1976. [Google Scholar]
  35. O’Neill, B. The fundamental equations of a submersion. Mich. Math. J. 1966, 13, 458–469. [Google Scholar] [CrossRef]
  36. Baird, P.; Wood, J.C. Harmonic Morphism Between Riemannian Manifolds; Oxford Science Publications: Oxford, UK, 2003. [Google Scholar]
  37. Sahin, B. A survey on differential geometry of Riemannian maps between Riemannian manifolds. Sci. Ann. Alexandru Ioan Cuza Univ. Iasi (New Ser.) Math. 2017, 63, 151–167. [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

Li, Y.; Prasad, R.; Haseeb, A.; Kumar, S.; Kumar, S. A Study of Clairaut Semi-Invariant Riemannian Maps from Cosymplectic Manifolds. Axioms 2022, 11, 503. https://doi.org/10.3390/axioms11100503

AMA Style

Li Y, Prasad R, Haseeb A, Kumar S, Kumar S. A Study of Clairaut Semi-Invariant Riemannian Maps from Cosymplectic Manifolds. Axioms. 2022; 11(10):503. https://doi.org/10.3390/axioms11100503

Chicago/Turabian Style

Li, Yanlin, Rajendra Prasad, Abdul Haseeb, Sushil Kumar, and Sumeet Kumar. 2022. "A Study of Clairaut Semi-Invariant Riemannian Maps from Cosymplectic Manifolds" Axioms 11, no. 10: 503. https://doi.org/10.3390/axioms11100503

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