You are on page 1of 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/358292134

Clairaut conformal submersions

Preprint · February 2022


DOI: 10.48550/arXiv.2202.00393

CITATIONS READS

0 106

2 authors:

Kiran Meena Akhilesh Yadav

15 PUBLICATIONS   46 CITATIONS   
Banaras Hindu University
39 PUBLICATIONS   96 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

PhD Thesis: Geometry of Riemannian Maps View project

All content following this page was uploaded by Kiran Meena on 03 May 2022.

The user has requested enhancement of the downloaded file.


Clairaut conformal submersions
Kiran Meena∗ and Akhilesh Yadav
arXiv:2202.00393v3 [math.DG] 2 May 2022

Abstract

The aim of this paper is to introduce Clairaut conformal submersion be-


tween Riemannian manifolds. First, we find necessary and sufficient conditions
for conformal submersions to be Clairaut conformal submersions. Further, we
find a necessary and sufficient condition for a Clairaut conformal submersion
to be harmonic. Moreover, we calculate the scalar and Ricci curvatures of ver-
tical distribution of M . Finally, we give two non-trivial examples of Clairaut
conformal submersion to illustrate the theory. Out of them, one is given by us-
ing doubly warped product, and another one is given by verifying main result
of the paper.

M. S. C. 2020: 53B20, 53C43.


Keywords: Riemannian manifold, Riemannian submersion, Clairaut Riemannian
submersion, Conformal submersion, Harmonic map.

1 Introduction
Immersions and submersions are important topics in differential geometry which
play a fundamental role in Riemannian geometry. A smooth map F : (M, g) →
(B, g ′ ) between Riemannian manifolds is said to be an isometric immersion if deriva-
tive map F∗ of F is injective and there is an isometry between them, i.e.,

g ′ (F∗ X, F∗ Y ) = g(X, Y ) for X, Y ∈ Γ(T M ).

On the other hand, Riemannian submersion between Riemannian manifolds was


studied by O’Neill [13], Gray [9] and Falcitelli [7]. A smooth map F : (M, g) →
(B, g ′ ) between Riemannian manifolds is said to be a Riemannian submersion if
derivative map F∗ of F is surjective and satisfies

g ′ (F∗ X, F∗ Y ) = g(X, Y ) for X, Y ∈ Γ(kerF∗ )⊥ .

Conformal submersion is a natural generalization of Riemannian submersion [13],


which restricted to vectors orthogonal to it’s fibers is a horizontal conformal sub-
mersion (or horizontal conformal map). A Riemannian submersion F : (M m , g) →
(B n , g ′ ) said to be horizontally conformal submersion (or conformal submersion) at
p ∈ M if there is a smooth function λ : M → R+ (called dilation) such that

g ′ (F∗ X, F∗ Y ) = λ2 (p)g(X, Y ) for X, Y ∈ Γ(kerF∗ )⊥ . (1.1)


∗ Corresponding Author

1
2 K. Meena, A. Yadav

Obviously, every conformal submersion is a Riemannian submersion with λ = 1.


The number λ2 (p) is called the square dilation and it is necessarily non-negative. In
[10], Gundmundsson obtained the fundamental equations for the conformal submer-
sion. Zawadzki studied conformal submersion with totally umbilical, geodesic and
minimal fibers in [17] and [18]. Further, horizontally conformal map was defined
by Fugledge [8] and Ishihara [11] which is useful for characterization of harmonic
morphisms [3] and have applications in medical imaging (brain imaging) and com-
puter graphics. A conformal submersion F : M → B is said to be homothetic if
the gradient of it’s dilation λ is vertical, i.e., H(gradλ) = 0 ∀p ∈ M , where H is a
projection of gradλ on horizontal space (kerF∗ )⊥ .
In elementary differential geometry, if ω is the angle between the velocity vector
of a geodesic and a meridian, and r is the distance to the axis of a surface of
revolution, then Clairaut’s relation ([5], [15]) states that rsinω is constant. In [4],
Bishop introduced Clairaut submersion and gave a necessary and sufficient condition
for a Riemannian submersion to be Clairaut. Further, Clairaut submersions were
studied in [1] and [2].
The concept of harmonic maps and morphisms is a very useful tool for global
analysis and differential geometry. The theory of harmonic maps has been developed
in [6], which is still an active field in differential geometry and it has applications
to many different areas of mathematics and physics. A map between Riemannian
manifolds is called harmonic if the divergence of its differential map vanishes. Har-
monic maps between Riemannian manifolds satisfy a system of quasi-linear partial
differential equations, therefore one would solve partial differential equations on
certain manifolds.
In this paper, we study Clairaut conformal submersion between Riemannian
manifolds. The paper is organised as: in Section 2, we give some basic information
about the conformal submersion which is needed for this paper. In Section 3, we
define Clairaut conformal submersion and obtained necessary and sufficient condi-
tions for conformal submersions to be Clairaut. Further, we find a necessary and
sufficient condition for a Clairaut conformal submersion to be harmonic. Moreover,
we calculate the scalar and Ricci curvature of vertical distribution. Finally, we give
two non-trivial examples of Clairaut conformal submersion.

2 Preliminaries
In this section, we recall the notion of conformal submersion between Riemannian
manifolds and give a brief review of basic facts.
Let F : (M m , g) → (B n , g ′ ) be a conformal submersion between Riemannian
manifolds. The fibers of F over q ∈ B is defined as F −1 (q). The vectors tangent
to fibers form the smooth vertical distribution denoted by νp and its orthogonal
complementary with respect to g is called horizontal distribution denoted by Hp .
Projections onto the horizontal and vertical distributions are denoted by H and ν,
respectively. A vector field E on M is said to be projectable if ∃ a vector field Ẽ
on B such that F∗p (E) = ẼF (p) . Then E and Ẽ are called F -related. For all Ẽ on
B there is a unique vector field E on M such that E and Ẽ are F -related, and the
vector field E is called the horizontal lift of Ẽ.
The O’Neill tensors A and T defined in [13] as
AX1 X2 = H∇HX1 νX2 + ν∇HX1 HX2 , (2.1)
Clairaut conformal submersions 3

TX1 X2 = H∇νX1 νX2 + ν∇νX1 HX2 , (2.2)


∀X1 , X2 ∈ Γ(T M ), where ∇ is the Levi-Civita connection of g. For any X1 ∈
Γ(T M ), TX1 and AX1 are skew-symmetric operators on (Γ(T M ), g) reversing the
horizontal and the vertical distributions. It is also easy to see that T is vertical, i.e.,
TX1 = TνX1 and A is horizontal, i.e., AX1 = AHX1 . We note that the tensor field T
satisfies TU W = TW U, ∀U, W ∈ Γ(kerF∗ ). Now, from (2.1) and (2.2), we have

∇U V = TU V + ν∇U V, (2.3)

∇X U = AX U + ν∇X U, (2.4)
∇X Y = AX Y + H∇X Y, (2.5)
∀X, Y ∈ Γ(kerF∗ )⊥ and U, V ∈ Γ(kerF∗ ). A conformal submersion F is with totally
umbilical fibers if [17]

TU V = g(U, V )H or TU X = −g(H, X)U, (2.6)

∀U, V ∈ Γ(kerF∗ ) and X ∈ Γ(kerF∗ )⊥ , where H is the mean curvature vector field
of fibers of F given by
Xm
(m − n)H = TUi Ui , (2.7)
i=n+1

where {Ui }n+1≤i≤m is an orthonormal basis of fibers of F . The horizontal vector


field H vanishes if and only if any fiber of F is minimal.

Proposition 2.1. [10] Let F : (M, g) → (B, g ′ ) be a conformal submersion with


dilation λ. Then
 
1 1
AX Y = ν[X, Y ] − λ2 g(X, Y )(∇ν 2 ) , ∀X, Y ∈ Γ(kerF∗ )⊥ . (2.8)
2 λ

Moreover, by (2.8) the horizontal space (kerF∗ )⊥ is totally geodesic if and only if λ
is constant on kerF∗ .

The differential F∗ of F can be viewed as a section of bundle Hom(T M, F −1 T B)


→ M , where F −1 T B is the pullback bundle whose fibers at p ∈ M is (F −1 T B)p =
TF (p) B, p ∈ M . The bundle Hom(T M, F −1 T B) has a connection ∇ induced from
the Levi-Civita connection ∇M and the pullback connection ∇F . Then the second
fundamental form of F is given by [12]

(∇F∗ )(X, Y ) = ∇F M
X F∗ Y − F∗ (∇X Y ), ∀X, Y ∈ Γ(T M ), (2.9)

or
(∇F∗ )(X, Y ) = ∇B M
F∗ X F∗ Y − F∗ (∇X Y ), ∀X, Y ∈ Γ(T M ), (2.10)
B
where ∇ is the Levi-Civita connection on B. Note that for the sake of simplicity
we can write ∇M as ∇.

Lemma 2.1. [10] Let F : (M, g) → (B, g ′ ) be a conformal submersion. Then


λ2
F∗ (H∇X Y ) = ∇B X( λ12 )F∗ Y + Y ( λ12 )F∗ X − g(X, Y )F∗ (gradH ( λ12 )) ,

F∗ X F∗ Y + 2

for X, Y basic vector fields and ∇ is the Levi-Civita connection on M .


4 K. Meena, A. Yadav

Now, using (2.9) in Lemma 2.1, we get


2
n o
(∇F∗ )(X, Y ) = − λ2 X( λ12 )Ỹ + Y ( λ12 )X̃ − g(X, Y )F∗ (gradH λ12 )) , (2.11)

where X and Y are horizontal lift of X̃ and Ỹ , respectively.


Now we recall the gradient, divergence and Laplacian [16]. Let f ∈ F (M ) then
the gradient of f , denoted by ∇f or gradf , given by

g(gradf, X) = X(f ), f or X ∈ Γ(T M ). (2.12)

The divergence of X, denoted by div(X), given by


m
X
div(X) = g(∇ek X, ek ), ∀X ∈ Γ(T M ), (2.13)
k=1

where {ek }1≤k≤m be an othonormal basis of Tp M . The Laplacian of f , denoted by


∆f , given by
∆f = div(∇f ). (2.14)

3 Clairaut conformal submersion


In this section, we will find some interesting results which are useful to investigate
the geometry of Clairaut conformal submersion.
Definition 3.1. A conformal submersion F : (M, g) → (B, g ′ ) between Riemannian
manifolds with dilation λ is said to be Clairaut conformal submersion if there is a
function r : M → R+ such that for every geodesic α on M , the function (r◦α)sinω(t)
is constant along α, where for all t, ω(t) is the angle between α̇(t) and the horizontal
space at α(t).

We know that with the help of Clairaut’s theorem we can find all geodesics
on a surface of revolution, and the notion of Clairaut conformal submersion comes
from geodesic curve on a surface of revolution. Therefore, we are going to find
necessary and sufficient conditions for a curve on the total space, and base space to
be geodesic.
Proposition 3.1. Let F : (M, g) → (B, g ′ ) be a conformal submersion. Let α :
I → M be a regular curve on M such that U (t) = ν α̇(t) and X(t) = Hα̇(t). Then
α is geodesic on M if and only if

AX X + ν∇X U + TU X + ν∇U U = 0, (3.1)

and
H∇X X + 2AX U + TU U = 0. (3.2)

Proof. Let α : I → M be a regular curve on M and α̇ = X(t) + U (t), where


X(t) ∈ Γ(kerF∗ )⊥ and U (t) ∈ Γ(kerF∗ ) is basis vector field along α. Then

∇α̇ α̇ = ∇X+U X + U = ∇X X + ∇X U + ∇U X + ∇U U,

which implies

∇α̇ α̇ = ν∇X X + ν∇X U + ν∇U X + ν∇U U + H∇X X + H∇X U + H∇U X + H∇U U.


Clairaut conformal submersions 5

Using (2.2), (2.3), (2.4) and (2.5) in above equation, we get

∇α̇ α̇ = AX X + ν∇X U + TU X + ν∇U U + H∇X X + 2AX U + TU U. (3.3)

We know that α is geodesic curve ⇐⇒ ∇α̇ α̇ = 0 [5]. Then proof follows by


(3.3).
Proposition 3.2. Let F : (M, g) → (B, g ′ ) be a conformal submersion with dilation
λ. Let α : I → M be a geodesic curve on M such that U (t) = ν α̇(t) and X(t) =
Hα̇(t). Then the curve β = F ◦ α is geodesic on B if and only if

λ2
   
2 1 2 1
λ X X̃ + F∗ (2AX U + TU U ) = kXk F∗ ∇H 2 ,
λ2 2 λ

where X is horizontal lift of X̃.

Proof. Let α : I → M be a geodesic on M and α̇ = X(t) + U (t). Then β = F ◦ α


is geodesic on B ⇐⇒ ∇B B
F∗ (α̇) F∗ (α̇) = 0 ⇐⇒ ∇F∗ X F∗ X = 0 ⇐⇒

∇F
X F∗ X = 0. (3.4)

Using (2.9) in (3.4), we get

F∗ (∇X X) + (∇F∗ )(X, X) = 0. (3.5)

Using (3.2) and (2.11) in (3.5), we get

λ2
 
1 1
F∗ (2AX U + TU U ) + 2X( 2 )X̃ − g(X, X)F∗ (gradH 2 )) = 0,
2 λ λ
which implies

λ2
   
1 1
λ2 X X̃ + F∗ (2AX U + TU U ) = kXk2 F∗ ∇H 2 .
λ2 2 λ

Corollary 3.1. Let F : (M, g) → (B, g ′ ) be a conformal submersion with dilation


λ. Let α : I → M be a horizontal geodesic on M such that U (t) = ν α̇(t) and
X(t) = Hα̇(t). Then the curve β = F ◦ α is a horizontal geodesic on B if and only
if
AX X = 0,
and    
1 1 1
X X̃ = kXk2 F∗ ∇H 2 ,
λ2 2 λ
where X is horizontal lift of X̃.

Proof. Let α be a horizontal geodesic on M then α̇ = Hα̇(t) = X(t) ∈ Γ(kerF∗ )⊥ ,


which implies U = ν(α̇(t)) = 0. Then proof follows by (3.1) and Proposition 3.2.
Theorem 3.1. Let F : (M, g) → (B, g ′ ) be a conformal submersion with connected
fibers and dilation λ. Then F is a Clairaut conformal submersion with r = ef if
λ2 1

and only if TU U = −g(U, U )∇f − 2 g ∇ν λ2 , U X, where f be a smooth function
on M and ∇f denotes gradient of f .
6 K. Meena, A. Yadav

Proof. Let α : I → M be a geodesic on M with U (t) = ν α̇(t) and X(t) = Hα̇(t).


Let ω(t) be the angle in [0, π] between α̇(t) and X(t). Since α is geodesic so speed
is constant a = kα̇k2 (say).
If sinω(t0 ) = 0 =⇒ g(α̇(t), V (t)) = 0 =⇒ g(X(t) + V (t), V (t)) = 0 =⇒
g(V (t), V (t)) = 0 =⇒ V (t0 ) = 0, which implies α is horizontal geodesic and
r(α(t)) sinω(t0 ) = 0. Therefore, we consider only non-horizontal geodesic. Now

g(α̇(t), X(t))g(α̇(t), X(t))


cos2 ω(t) = ,
kα̇(t)k2 kX(t)k2
which implies
g(X, X) = acos2 ω(t). (3.6)
Similarly, we can get
g(U, U ) = asin2 ω(t). (3.7)
On differentiation of (3.7), we get
d
g(U, U ) = 2g(∇α̇ U, U ) = 2g(ν∇U U + ν∇X U, U ). (3.8)
dt
Using (3.1) in (3.8), we get
d
g(U, U ) = −2g(AX X + TU X, U ). (3.9)
dt
On the other hand, by (3.7), we have
d dω
g(U, U ) = 2asinω(t)cosω(t) . (3.10)
dt dt
By (3.9) and (3.10), we get

g(AX X + TU X, U ) = −asinω(t)cosω(t) . (3.11)
dt
d
Moreover, F is a Clairaut conformal submersion with r = ef ⇐⇒ dt (ef ◦α sinω) =
dω df
0 ⇐⇒ cosω dt + sinω dt = 0. By multyplying this with non-zero factor asinω, we
get
dω df
− acosωsinω = asin2 ω . (3.12)
dt dt
By (3.7) and (3.12), we get
df dω
g(U, U ) = −acosωsinω . (3.13)
dt dt
By (3.11) and (3.13), we get
df
g(U, U ) = g(AX X + TU X, U ),
dt
which implies
df
g(TU X, U ) = g(U, U ) − g(AX X, U ),
dt
which implies
df
− g(TU U, X) = g(U, U ) − g(AX X, U ). (3.14)
dt
Clairaut conformal submersions 7

Using (2.8) in (3.14), we get

λ2
 
df 1
−g(TU U, X) = g(U, U ) + g(X, X)g U, ∇ν 2 ,
dt 2 λ

which implies

λ2
 
1
g(TU U, X) = −g(U, U )g(α̇, ∇f ) − g(X, X)g U, ∇ν 2 .
2 λ

If we consider any geodesic α on M with initial vertical tangent vector, gradf turns
out to be horizontal. Because, initially at t = 0, ω(t) = 0 =⇒ sinω(t)|t=0 =
0 =⇒ g(U, U )|t=0 = 0 =⇒ U (t)|t=0 = 0, which shows that ν(gradf ) = 0. Thus
the function f is constant on any fiber and the fibers are being connected. Now,
putting α̇ = X + U in above equation, we get

λ2
 
1
g(TU U, X) = −g(U, U )g(X, ∇f ) − g(X, X)g U, ∇ν 2 . (3.15)
2 λ

Now, comparing both sides of (3.15), we get

λ2
 
1
TU U = −g(U, U )∇f − g U, ∇ν 2 X,
2 λ

where ∇f ∈ Γ(kerF∗ )⊥ . This completes the proof.


Corollary 3.2. Let F : (M, g) → (B, g ′ ) be a conformal submersion with totally
umbilical fibers and dilation λ. Then F is a Clairaut conformal submersion with
r = ef if and only if

(i) H = −∇f , and


(ii) (kerF∗ )⊥ is totally geodesic (i.e. λ is constant on kerF∗ ),

where H is the mean curvature vector field of fibers of F .

Proof. Let F : (M, g) → (B, g ′ ) be a conformal submersion with totally umbilical


fibers then using (2.6) in Theorem 3.1, we get

λ2
 
1
g(U, U )H = −g(U, U )∇f − g U, ∇ν 2 X.
2 λ

On compairing both sides, we get H = −∇f and g U, ∇ν λ12 = 0 =⇒ U λ12 =


 

0 =⇒ λ is constant on fibers of F . This completes the proof.


Corollary 3.3. Let F : (M, g) → (B, g ′ ) be a conformal submersion such that
dilation λ is constant on kerF∗ . Then F is a Clairaut conformal submersion with
r = ef if and only if fibers of F are totally umbilical with H = −∇f , where H is
the mean curvature vector field of fibers of F .

Proof. The proof is follows from (2.6) and Theorem 3.1.


Corollary 3.4. Let F : (M, g) → (B, g ′ ) be a Clairaut conformal submersion with
r = ef and dilation λ. Then fibers of F are totally geodesic if and only if
8 K. Meena, A. Yadav

(i) f is constant on (kerF∗ )⊥ , and


(ii) λ is constant on (kerF∗ ).

Proof. The proof is follows from Theorem 3.1.


Proposition 3.3. Let F : (M m , g) → (B n , g ′ ) be a Clairaut conformal submersion
with r = ef and dilation λ. Then
m X n
λ2
 
X 1
H = −∇f − g Ui , ∇ν 2 Xj , (3.16)
2(m − n) i=n+1 j=1 λ

and
m X n
nλ2
  
H
X 1
div(H) = −∆ f − ∇Xj g Ui , ∇ν 2 , (3.17)
(m − n) i=n+1 j=1 λ

where H is the mean curvature vector field of fibers of F and div(H) it’s divergence.
Also, {Ui }n+1≤i≤m and {Xj }1≤j≤n are orthonormal bases of kerF∗ and (kerF∗ )⊥ ,
respectively.

Proof. Let F : (M m , g) → (B n , g ′ ) be a Clairaut conformal submersion then from


Theorem 3.1, we have
λ2
 
1
TU U = −g(U, U )∇f − g U, ∇ν 2 X. (3.18)
2 λ
Taking trace of (3.18), we get
m m m n
λ2 X X
 
X X 1
TUi Ui = − g(Ui , Ui )∇f − g Ui , ∇ν 2 Xj (3.19)
i=n+1 i=n+1
2 i=n+1 j=1 λ

where {Ui }n+1≤i≤m and {Xj }1≤j≤n are orthonormal bases of kerF∗ and (kerF∗ )⊥ ,
respectively. Using (2.7) in (3.19), we get
m n
λ2 X X
 
1
(m − n)H = −(m − n)∇f − g Ui , ∇ν 2 Xj , (3.20)
2 i=n+1 j=1 λ

which implies (3.16). Also, by (3.20), we get


n
P n
P
g(∇Xj H, Xj ) =− g(∇Xj ∇f, Xj )
j=1 j=1
2
m n (3.21)
λ
∇Xj g Ui , ∇ν λ12
P P 
− 2(m−n) g(Xj , Xj ).
i=n+1 j=1

Using (2.13) and (2.14) in (3.21), we get (3.17). This completes the proof.
Definition 3.2. ([16], p. 45) Let F : (M m , g) → (B n , g ′ ) be a smooth map between
Riemannian manifolds. Then F is harmonic if and only if the tension field τ (F ) of
F vanishes at each point p ∈ M .
Theorem 3.2. Let F : (M, g) → (B, g ′ ) be a Clairaut homothetic conformal sub-
mersion with r = ef such that dilation λ is constant on kerF∗ . Then F is harmonic
if and only if f is constant on (kerF∗ )⊥ .
Clairaut conformal submersions 9

Proof. Let F : (M, g) → (B, g ′ ) be a conformal


Pm submersion between Riemannian
manifolds. Then τ (F ) = trace(∇F∗ ) = k=1 (∇F∗ )(ek , ek ) = 0, where {ek }1≤k≤m
is local orthonormal basis around a point p ∈ M . Hence

τ (F ) = τ kerF∗ (F ) + τ (kerF∗ ) (F ). (3.22)

Let {X̃j }1≤j≤n , {λXj }1≤j≤n and {Ui }n+1≤i≤m be orthonormal bases of T B, (kerF∗ )⊥
and kerF∗ , respectively. Now, we have
m
X
τ kerF∗ (F ) = (∇F∗ )(Ui , Ui ).
i=n+1

Using (2.10) and (2.3) in above equation, we get


m
X m
X
τ kerF∗ (F ) = − F∗ (H∇Ui Ui ) = − F∗ (HTUi Ui ). (3.23)
i=n+1 i=n+1

Using (2.7) in (3.23), we get

τ kerF∗ (F ) = −(m − n)F∗ (H). (3.24)

Also, since
n
⊥ X
τ (kerF∗ ) (F ) = (∇F∗ )(λXj , λXj ).
j=1

Using (2.11) in above equation, we get


n 
λ2 X
   
⊥ 1 1
τ (kerF∗ ) (F ) = − 2λ2 Xj X̃ j − g(λX j , λX j )F∗ ∇H 2 . (3.25)
2 j=1 λ2 λ

Using (2.12) in (3.25), we get


n 
λ2 X
   
(kerF∗ )⊥ 2 1 1
τ (F ) = − 2λ g Xj , ∇ 2 X̃j − g(λXj , λXj )F∗ ∇H 2 .
2 j=1 λ λ

Using (1.1) in above equation, we get


n 
λ2 X
    
⊥ 1 1
τ (kerF∗ ) (F ) = − 2g ′ X̃j , F∗ ∇H 2 X̃j − g ′ (X̃j , X̃j )F∗ ∇H 2 ,
2 j=1 λ λ

which implies
λ2
 
⊥ 1
τ (kerF∗ ) (F ) = (n − 2) F∗ ∇H 2 . (3.26)
2 λ
Putting (3.24) and (3.26) in (3.22), we get

λ2
 
1
τ (F ) = (n − 2) F∗ ∇H 2 − (m − n)F∗ (H). (3.27)
2 λ

Since F is Clairaut homothetic conformal submersion therefore putting ∇H λ12 = 0


in (3.27), we get
τ (F ) = −(m − n)F∗ (H). (3.28)
10 K. Meena, A. Yadav

Using (3.16) in (3.28), we get


 
2 Xm X n  
λ 1
τ (F ) = (m − n)F∗ (∇f ) + F∗  g Ui , ∇ν 2 Xj  .
2 i=n+1 j=1 λ

Since λ is constant on kerF∗ therefore ∇ν λ12 = 0, then above equation implies

τ (F ) = (m − n)F∗ (∇f ).

Thus τ (F ) = 0 ⇐⇒ f is constant on (kerF∗ )⊥ . This completes the proof.

Theorem 3.3. Let F : (M, g) → (B, g ′ ) be a Clairaut conformal submersion with


r = ef and dilation λ. Then
λ4
  
1 1
Kν = K̂ − (m − n)(m − n − 1) λ2 kgradH f k2 − kgradH 2 k2 + λ2 g ∇f, ∇H 2
4 λ λ
4 m−n    
λ X 1 1
− (m − n − 1) Xl 2 Ui 2
2 i=1
λ λ

where Kν and K̂ denote the scalar curvature of vertical distribution and fibers, re-
spectively. In addition, {U1 , U2 , . . . , Um−n } and {X1 , X2 , . . . , Xn } are orthonormal
bases of vertical and horizontal distribution, respectively.

Proof. Let g and g be two Riemannian metrics on M which are conformally related
to each other. Let F : (M, g) → (B, g ′ ) be a Clairaut conformal submersion with
r = ef then for g = λ2 g, the submersion F : (M, g) → (B, g ′ ) such that F (p) =
F (p) ∀p ∈ M is a Clairaut Riemannian submersion with r = ef , where f is a
positive function on M . Clearly, we can see that F and F have same vertical and
horizontal distributions. Since F is a Clairaut Riemannian submersion with r = ef ,
then for any vertical vectors U and V , we have [4]

T U V = g(U, V )H = −g(U, V )∇f = −λ2 g(U, V )∇f, (3.29)

where H = −∇f is the mean curvature vector field of fibers and T denotes the tensor
for the Levi-Civita connection ∇ of g. Also, for a Clairaut conformal submersion
F , we have (by Theorem 3.1)

λ2
 
1
TU V = −g(U, V )∇f − g U, ∇ν 2 X, (3.30)
2 λ

where X ∈ Γ(kerF∗ )⊥ and T denotes the tensor for the Levi-Civita connection ∇
of g. Using (3.29), (3.30) and (2.12), we get

λ4
 
2 1
T U V = λ TU V + U X. (3.31)
2 λ2

Let R and R denote the Riemannian curvature tensor with respect to ∇ and ∇,
respectively. Then

g(R(U, V )V, U ) 1
sec(U, V ) = 2
= 2 g(R(U, V )V, U ).
g(U, U )g(V, V ) − g(U, V ) λ
Clairaut conformal submersions 11

Using ([10], p. 14) in above equation, we get


λ2
 
1 1
sec(U, V ) = sec(U, V ) + Hess 2 (U, U ) + Hess 2 (V, V )
2 λ λ
!
2 2
λ4 1 2 λ4
   
1 1
− |grad 2 | − U 2 + V 2 , (3.32)
4 λ 4 λ λ

where sec(U, V ) and sec(U, V ) denote the sectional curvature of the plane span by
orthonormal vectors (with respect to g) U and V on (M, g) and (M, g), respectively.
In addition, Hess λ12 (U, U ) = g(∇U ∇ λ12 , U ). Since F : (M, g) → (B, g ′ ) be a
Clairaut Riemannian submersion therefore for non-zero orthogonal vertical vectors
(with respect to g) at p ∈ M , we have ([7], p. 12)
ˆ
g(R(U, V )V, U ) = g(R(U, V )V, U ) + g(T U V, T V U ) − g(T V V, T U U ),

where Rˆ is the Riemannian curvature tensor of fibers of F . Using (3.29) in above


equation, we get
ˆ
sec(U, V ) = sec(U, V ) − λ2 |gradf |2 , (3.33)
ˆ
where sec denotes the sectional curvature of fibers of F . Since we know that
ˆ
g(R(U, V )V, U ) 1 ˆ
ˆ
sec(U, V)= = 2 g(R(U, V )V, U ).
g(U, U )g(V, V ) − g(U, V )2 λ
Using ([10], p. 14) in above equation, we get
λ2
 
ˆ 1 1
sec(U, V ) = ŝec(U, V ) + g(∇U ∇ν 2 , U ) + g(∇V ∇ν 2 , V )
2 λ λ
4 4
 2  2 !
λ 1 λ 1 1
− |gradν 2 |2 − U 2 + V 2 . (3.34)
4 λ 4 λ λ

By using (3.32), (3.33) and (3.34), we get


λ4 1
ŝec(U, V ) − λ2 |gradf |2 = sec(U, V ) − |gradH 2 |2
4 λ
λ2
 
1 1
− g(H∇U U, ∇ 2 ) + g(H∇V V, ∇ 2 ) .
2 λ λ
Using (2.3) in above equation, we get
λ4 1
ŝec(U, V ) − λ2 |gradf |2 = sec(U, V ) − |gradH 2 |2
4 λ
λ2
 
1 1
− g(TU U, ∇ 2 ) + g(TV V, ∇ 2 ) . (3.35)
2 λ λ
By using (3.30), we get
λ2
    
1 1 1 1 1 1
g(TU U, ∇ ) + g(T V V, ∇ ) = −2g ∇f, ∇ − U + V X (3.36)
λ2 λ2 λ2 2 λ2 λ2 λ2
Using (3.36) in (3.35), we get
λ4
 
2 2 1 2 2 1
sec(U, V ) = ŝec(U, V ) − λ |gradf | + |gradH 2 | − λ g ∇f, ∇H 2
4 λ λ
λ4
  
1 1 1
− U 2 +V 2 X 2 . (3.37)
4 λ λ λ
12 K. Meena, A. Yadav

Taking trace of (3.37), we get


m−n
X m−n
X m−n
X m−n
X
sec(Ui , Uj ) = ŝec(Ui , Uj ) − λ2 (m − n)(m − n − 1)kgradH f k2
i=1 j=1,j6=i i=1 j=1,j6=i

λ4 1
+ (m − n)(m − n − 1)kgradH 2 k2
4  λ 
2 1
− λ (m − n)(m − n − 1)g ∇f, ∇H 2
λ
4 m−n m−n        
λ X X 1 1 1
− Ui 2
+ Uj 2
Xl ,
4 i=1 λ λ λ2
j=1,j6=i

where {U1 , U2 , . . . , Um−n } and {X1 , X2 , . . . , Xn } be orthonormal bases of vertical


and horizontal distribution of F , respectively. This implies the proof.

Corollary 3.5. Let F : (M, g) → (B, g ′ ) be a Clairaut conformal submersion with


r = ef and dilation λ such that fibers of F are totally umbilical. Then

λ4
  
2 2 1 2 2 1
Kν = K̂ − (m − n)(m − n − 1) λ kgradH f k − kgradH 2 k + λ g ∇f, ∇H 2 ,
4 λ λ

where Kν and K̂ denote the scalar curvature of vertical distribution and fibers,
respectively.

Proof. By Corollary 3.2, λ is constant on kerF∗ . Then the proof is follows by


Theorem 3.3.

Theorem 3.4. Let F : (M, g) → (B, g ′ ) be a Clairaut conformal submersion with


r = ef and dilation λ such that fibers of F are Einstein. Then

λ4
 
1 1
Ric(Ui , Ui ) = λf − (m − n − 1)2 λ2 kgradH f k2 − kgradH 2 k2 + λ2 g(∇f, ∇H 2 )
4 λ λ
 
m−n
λ4  1 X 1  1
− (m − n − 1) (m − n − 1)(Ui 2 ) + (Uj 2 ) (Xl 2 ),
4  λ λ  λ
j=1,j6=i

where Ric(Ui , Ui ) denotes the Ricci curvature of vertical distribution at p ∈ M


and λf is constant. In addition, {U1 , U2 , . . . , Um−n } and {X1 , X2 , . . . , Xn } are
orthonormal bases of vertical and horizontal distribution, respectively.

Proof. Taking trace of (3.37), we get


m−n
X m−n
X
sec(Ui , Uj ) = ŝec(Ui , Uj ) − λ2 (m − n − 1)kgradH f k2
j=1,j6=i j=1,j6=i

λ4
 
1 1
+ (m − n − 1)kgradH 2 k2 − λ2 (m − n − 1)g ∇f, ∇H 2
4 λ λ
4 m−n
       
λ X 1 1 1
− Ui 2
+ Uj 2
Xl , (3.38)
4 λ λ λ2
j=1,j6=i
Clairaut conformal submersions 13

where {U1 , U2 , . . . , Um−n } and {X1 , X2 , . . . , Xn } be orthonormal bases of vertical


and horizontal distribution of F , and sums are taken over all vectors of orthonormal
basis except Ui . Since we know that
m−n m−n
X X g(R(Ui , Uj )Uj , Ui )
sec(Ui , Uj ) =
g(Ui , Ui )g(Uj , Uj ) − g(Ui , Uj )2
j=1,j6=i j=1,j6=i
1
= Ric(Ui , Ui ). (3.39)
(m − n − 1)

Using (3.39) in (3.38), we get


1
(m−n−1)
Ric(Ui , Ui ) = 1
(m−n−1)
R̂ic(Ui , Ui ) − λ2 (m − n − 1)kgradH f k2
4
+ λ4 (m − n − 1)kgradH λ12 k2 − λ2 (m − n − ! 1)g ∇f, ∇H λ12


4 m−n
− λ4 (m − n − 1)Ui λ12 + Uj λ12 Xl λ12 .
 P  
j=1,j6=i
(3.40)
Also fibers are Einstein therefore R̂ic(Ui , Uj ) = λf g(Ui , Ui ) = λf (constant). Then
(3.40) implies the proof.

Example 3.1. Let M = M1 λ ×f1 M2 be a doubly warped product [14] of two


Riemannian manifolds (M1 , g1 ) and (M2 , g2 ). Let f1 : M1 → R+ and λ : M2 → R+
be smooth functions. Let π : M1 λ ×f1 M2 → M1 and σ : M1 λ ×f1 M2 → M2 be the
projections. The doubly warped product manifold M = M1 λ ×f1 M2 furnished with
the metric tensor
g = (λ ◦ σ)2 π ∗ (g1 ) + (f1 ◦ π)2 σ ∗ (g2 ).
If X is tangent to M1 λ ×f1 M2 at (p, q), then

1
g(X, X) = λ2 (q)g1 (π∗ X, π∗ X) + f12 (p)g2 (σ∗ X, σ∗ X) = . (3.41)
λ2
It is easy to prove that the first projection π : M1 λ ×f1 M2 → M1 is conformal
submersion onto M1 whose vertical and horizontal spaces at (p, q) are identified by
Tq M2 and Tp M1 , respectively, and second projection σ : M1 λ ×f1 M2 → M2 is a
positive homothetic map onto M2 with scale factor f 21(p) .
1

Here for U, V ∈ L(M2 ) =⇒ [U, V ] ∈ L(M2 ), i.e., H[U, V ] = 0 and for X, Y ∈


L(M1 ) =⇒ [X, Y ] ∈ L(M1 ), i.e., ν[X, Y ] = 0, where L denotes set of lifts. Since
[X, V ] = 0 =⇒ [X, V ] is π-related to 0 and [X, V ] is σ-related to 0. Now, since
leaves are integrable then by using Koszul’s formula [14], we obtain

2g(ν∇X Y, V ) = −V g(X, Y ).

Using (3.41) in above equation, we get


2g(AX Y, V ) = −V (λ2 g1 (π∗ X, π∗ Y )) = −2λ(V λ)g1 (π∗ X, π∗ Y ) = −2 g(X, Y ),
λ
which implies
g(AX Y, V ) = −V (logλ)g(X, Y ). (3.42)
Using (2.12) in (3.42), we get

g(AX Y, V ) = −g(V, ∇ν (logλ))g(X, Y ).


14 K. Meena, A. Yadav

On comparing we get

AX Y = ν∇X Y = −g(X, Y )∇ν (logλ),


which means each leaves is totally umbilical. Also, g(ν∇X Y, V ) 6= 0 and g(∇X V, Y ) 6=
0. Now by using Koszul’s formula, we get
g(H∇U V, Y ) = 12 {U g(V, Y ) + V g(Y, U ) − Y g(U, V )
+g([U, V ], Y ) + g([Y, U ], V ) − g([V, Y ], U )},
which implies
1
g(H∇U V ) = − Y g(U, V ).
2
By using (3.41) in above equation, we get
Y 2
g(H∇U V, Y ) = − f g2 (σ∗ U, σ∗ V ). (3.43)
2 1
Again using (3.41) in (3.43), we get
 
Y f1
g(H∇U V, Y ) = − g(U, V ).
f1
Using (2.3) in above equation, we get
g(TU V, Y ) = −g(Y, ∇H log(f1 ))g(U, V ),

which implies
TU V = −g(U, V )∇H log(f1 ), (3.44)
which means fibers are umbilical. Now, by using (3.44) TU U = −g(U, U )∇f −
λ2 1

2 g ∇ν λ2 , U X ⇐⇒ f = logf1 and λ is constant on fibers. Thus by Corollary
3.2, π is Clairaut conformal submersion with r = elogf1 = f1 .
Example 3.2. Let M = {(x1 , x2 ) ∈ R2 : x1 6= 0, x2 6= 0} be an Euclidean space. We
define a Riemannian metric g = e2x2 dx21 + e2x2 dx22 on M . Let B = {(y1 , y2 ) ∈ R2 }
be a Riemannian manifold with Riemannian metric g ′ = dy12 + dy22 on B. Consider
a map F : (M, g) → (B, g ′ ) such that
x + x 
1
F (x1 , x2 ) = √ 2,0 .
2
Then, we get
kerF∗ = Span{U = e1 − e2 },
and
(kerF∗ )⊥ = Span{X = e1 + e2 },
n o n o
∂ −x2 ∂
where e1 = e−x2 ∂x , e 2 = e ∂x2 , e′1 = ∂y∂ 1 , e′2 = ∂
are bases on Tp M
∂y2
1

and TF (p) B respectively, for all p ∈ M . We see that F∗ (X) = 2e−x2 e′1 and
λ2 g(X, X) = g ′ (F∗ X, F∗ X) for X ∈ Γ(kerF∗ )⊥ , and dilation λ = e−x2 . Thus F is
conformal submersion.
Now we will find a smooth function f on M satisfying TU U = −g(U, U )∇f −
λ2
U, ∇ν λ12 X for all U ∈ Γ(kerF∗ ). Now, we have

2 g
 2x   −2x 
e 2 0 ij e 2 0
gij = , g = . (3.45)
0 e2x2 0 e−2x2
Clairaut conformal submersions 15

By using (3.45), we obtain

Γ111 = 0, Γ211 = −1, Γ122 = 0, Γ222 = 1, Γ112 = 1 = Γ121 , Γ212 = 0 = Γ221 .

Then, by some computations, we obtain


∂ ∂
∇e1 e1 = −e−2x2 , ∇e2 e2 = 0, ∇e2 e1 = 0, ∇e1 e2 = e−2x2 .
∂x2 ∂x1
Therefore
∇U U = ∇e1 −e2 e1 − e2 = −e−x2 X.
Then by (2.3), we get
TU U = −e−x2 X. (3.46)
Since λ = e−x2 , therefore

λ2
 
1
g U, ∇ν 2 X = −e−x2 X. (3.47)
2 λ

In addition
g(U, U ) = 2. (3.48)
For any smooth function f on M , the gradient of f with respect to the metric g is
2
∂f ∂ ∂f ∂ ∂f ∂
g ij ∂x . Therefore ∇f = e−2x2 ∂x +e−2x2 ∂x
P
given by ∇f = i ∂xj 1 ∂x1 2 ∂x2
. Hence,
i,j=1

∇f = e−x2 X, (3.49)

for the function f = x1 + x2 . Then by (3.46), (3.47), (3.48) and (3.49) it is easy
2
to verify that TU U = −g(U, U )∇f − λ2 g U, ∇ν λ12 X. Thus, by Theorem 3.1, F is


Clairaut conformal submersion.

Acknowledgment
Kiran Meena, would like to thank to the Council of Scientific and Industrial Re-
search (C.S.I.R.), New Delhi, India for providing the financial support. [ File No.:
09/013(0887)/2019-EMR-I ].

References
[1] D. Allison, Lorentzian Clairaut submersions, Geom. Dedicata, 63, 309-319,
1996.

[2] K. Aso and S. Yorozu, A generalization of Clairaut’s theorem and umbilical


foliations, Nihonkai Math. J., 2, 139-153, 1991.

[3] P. Baird and J. C. Wood, Harmonic Morphisms between Riemannian Mani-


folds, Clarendon Press, Oxford, 2003.

[4] R. L. Bishop, Clairaut submersions, Differential geometry (in Honor of K.


Yano), Kinokuniya, Tokyo, 21-31, 1972.
16 K. Meena, A. Yadav

[5] M. P. do Carmo, Differential geometry of curves and surfaces, Prentice-Hall,


Inc., Englewood Cliffs, New Jersy, 1976.
[6] J. Eells and H. J. Sampson, Harmonic mappings of Riemannian manifolds,
Amer. J. Math., 86, 109-160, 1964.
[7] M. Falcitelli, S. Ianus and A. M. Pastore, Riemannian Submersions and Related
Topics, World Scientific, River Edge, NJ, 2004.
[8] B. Fugledge, Harmonic morphisms between Riemannian manifolds, Ann. Inst.
Fourier (Grenoble), 28, 107-144, 1978.
[9] A. Gray, Pseudo-Riemannian almost product manifolds and submersions, J.
Math. Mech., 16, 715-737, 1967.
[10] S. Gundmundsson, The Geometry of Harmonic Morphisms, PhD Thesis, Uni-
versity of Leeds, 1992.
[11] T. Ishihara, A mapping of Riemannian manifolds which preserves harmonic
functions, J. Math. Kyoto Univ., 19, 215-229, 1979.
[12] T. Nore, Second fundamental form of a map, Ann. Mat. Pura Appl., 146 (1),
281-310, 1986.
[13] B. O’Neill, The fundamental equations of a submersion, Michigan Math. J., 13
(4), 459-469, 1966.
[14] B. O’Neill, Semi-Riemannian Geometry with Applications to Relativaty, Aca-
demic Press, New York, 1983.
[15] A. Pressley, Elementary Differential Geometry, Springer, London, 2010.
[16] B. Şahin, Riemannian submersions, Riemannian maps in Hermitian geometry,
and their applications, Academic Press, 2017.
[17] T. Zawadzki, Existence conditions for conformal submersions with totally um-
bilical fibers, Differential Geom. Appl., 35, 69-84, 2014.
[18] T. Zawadzki, On conformal submersions with geodesic or minimal fibers, Ann.
Global Anal. Geom., 58, 191-205, 2020.

K. Meena & A. Yadav


Department of Mathematics
Institute of Science
Banaras Hindu University
Varanasi - 221005, India.
E-mail: kirankapishmeena@gmail.com {K. Meena}
akhilesha68@gmail.com {A. Yadav}

View publication stats

You might also like