Next Article in Journal
Solving Time-Fractional Partial Differential Equation Using Chebyshev Cardinal Functions
Next Article in Special Issue
On a Surface Associated to the Catalan Triangle
Previous Article in Journal
Optical Solitons in Fiber Bragg Gratings with Dispersive Reflectivity Having Five Nonlinear Forms of Refractive Index
Previous Article in Special Issue
Soft Regular Generalized ω-Closed Sets and Soft ω-T1/2 Spaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On h-Quasi-Hemi-Slant Riemannian Maps

1
Department of Mathematical Sciences, Faculty of Applied Sciences, Umm Al Qura University, Makkah 21955, Saudi Arabia
2
Department of Mathematics, Shri Jai Narain Post Graduate College, University of Lucknow (U.P.), Lucknow 226001, India
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
Department of Mathematics, Dr. Shree Krishna Sinha Women’s College Motihari, Babasaheb Bhimrao Ambedkar University, Muzaffarpur 845401, India
*
Author to whom correspondence should be addressed.
Axioms 2022, 11(11), 641; https://doi.org/10.3390/axioms11110641
Submission received: 19 October 2022 / Revised: 9 November 2022 / Accepted: 11 November 2022 / Published: 14 November 2022
(This article belongs to the Special Issue Differential Geometry and Its Application)

Abstract

:
In the present article, we indroduce and study h-quasi-hemi-slant (in short, h-qhs) Riemannian maps and almost h-qhs Riemannian maps from almost quaternionic Hermitian manifolds to Riemannian manifolds. We investigate some fundamental results mainly on h-qhs Riemannian maps: the integrability of distributions, geometry of foliations, the condition for such maps to be totally geodesic, etc. At the end of this article, we give two non-trivial examples of this notion.

1. Introduction

In Riemannian geometry, there are few appropriate maps among Riemannian manifolds that compare their geometric properties. In this direction, as a generalization of the notions of isometric immersions and Riemannian submersions, Riemannian maps between Riemannian manifolds were initiated by Fischer [1], while isometric immersions and Riemannian submersions were widely studied in [2] and [3], respectively. However, the notion of Riemannian maps is a new research topic for geometers. More precisely, a differentiable map π : ( N 1 , g 1 ) ( N 2 , g 2 ) between Riemannian manifolds ( N 1 , g 1 ) and ( N 2 , g 2 ) is called a Riemannian map ( 0 < r a n k π < min { m , n } , where dim N 1 = m and dim N 2 = n ) if it satisfies the following equation:
g 2 ( π W 1 , π W 2 ) = g 1 ( W 1 , W 2 ) , for W 1 , W 2 Γ ( ker π ) ,
where π is the differentiable map of π .
Consequently, isometric immersions and Riemannian submersions are particular cases of Riemannian maps with ker π = 0 and ( r a n g e π ) = 0 , respectively [1].
The other prominent basic map for comparing geometric structures between Riemannian manifolds is Riemannian submersion, and it was studied by O’Neill [4] and Gray [5]. In 1976, Watson [6] studied Riemannian submersion between Riemannian manifolds equipped with differentiable structures. After that, several kinds of Riemannian submersions were introduced and studied, including Riemannian submersion [3], H-anti-invariant submersion [7], H-semi-invariant submersion [8] and H-semi-slant submersion [9].
Currently, one of the most inventive topics in differential geometry is the theory of Riemannian maps between different Riemannian manifolds. It is well known that differentiable maps between Riemannian manifolds have wide applications in differential geometry as well as in physics, such as in Yang–Mills theory [10], Kaluza–Klein theory [11], and supergravity and superstring theories [12].
We also note that quarternionic manifolds have many applications, including for nonlinear σ models with super symmetry [12], in the theory of harmonic differential forms [13] and obtaining estimates for the Betti numbers of the manifold [14,15]. In this paper, we have for the first time investigated h-qhs Riemannian maps from almost quarternionic manifolds to Riemannian manifolds. Here, we mainly focus on the most fundamental and interesting geometric properties on the fibers and distributions of these maps.
Nowadays, Riemannian maps and related topics have been actively studied by many authors, such as invariant and anti-invariant Riemannian maps [16], semi-invariant Riemannian maps [17], slant Riemannian maps [18], semi-slant Riemannian maps [19,20], hemi-slant Riemannian maps [21], quasi-hemi-slant Riemannian maps [22], almost h-semi-slant Riemannian maps [23], V-quasi-bi-slant Riemannian maps [24] and Clairaut semi-invariant Riemannian maps [25]. As a generalization of h-slant Riemannian maps [26], h-semi-slant Riemannian maps [9] and h-hemi-slant Riemannian maps, we define and study h-qhs Riemannian maps from almost Hermitian manifolds to Riemannian manifolds. In the near future, we plan to work on conformal h-qhs submersions, conformal h-qhs submersions, h-qhs semi-Riemannian submersions, etc.
This paper is structured as follows. In Section 2, we recall basic facts about Riemannian maps and almost Hermitian manifolds. In Section 3, we define h-qhs Riemannian maps and study the geometry of leaves of distributions that are involved in the definition of such maps. We give necessary and sufficient conditions for h-qhs Riemannian maps to be totally geodesic. Finally, we provide two concrete examples of h-qhs Riemannian maps.

2. Preliminaries

Let ( N 1 , g 1 ) and ( N 2 , g 2 ) be Riemannian manifolds and π : ( N 1 , g 1 ) ( N 2 , g 2 ) be a C -Riemannian map [1].
We define O’Neill’s tensors T and A [4] 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 vector fields F 1 , F 2 on N 1 , where ∇ is the Levi-Civita connection of g 1 .
From Equations (2) and (3), we have
Y 1 Y 2 = T Y 1 Y 2 + V Y 1 Y 2 ,
Y 1 U 1 = T Y 1 U 1 + H Y 1 U 1 ,
U 1 Y 1 = A U 1 Y 1 + V U 1 Y 1 ,
U 1 U 2 = H U 1 U 2 + A U 1 U 2 ,
for Y 1 , Y 2 Γ ( ker π ) and U 1 , U 2 Γ ( ker π ) , where H Y 1 U 1 = A U 1 Y 1 and U 1 is basic.
Let π : ( N 1 , E , g 1 ) ( N 2 , g 2 ) be a C map. The second fundamental form of π is given by
( π ) ( V 1 , V 2 ) = V 1 π π ( V 2 ) π ( V 1 N 1 V 2 ) ,
for V 1 , V 2 Γ ( T N 1 ) , where π is the pullback connection [27]. The map π is said to be a total geodesic if ( π ) ( V 1 , V 2 ) = 0 for V 1 , V 2 Γ ( T N 1 ) .
Let ( N 1 , E , g 1 ) be an almost quaternionic Hermitian manifold, where g 1 is a Riemanian metric on the maniifold N 1 and E is a rank 3 subbundle of E n d ( T N 1 ) such that for any point p N 1 within some neighborhood U, there exists a local basis { J 1 , J 2 , J 3 } of sections of E on U satisfying all α { 1 , 2 , 3 } in which
J α 2 = i d , J α J α + 1 = J α + 1 J α = J α + 2 ,
g 1 ( J α X 2 , J α X 1 ) = g 1 ( X 1 , X 2 ) ,
for X 1 , X 2 Γ ( T N 1 ) , where the indices are taken from { 1 , 2 , 3 } modulo 3 and { J 1 , J 2 , J 3 } is called the quaternionic Hermitian basis. The structure ( N 1 , E , g 1 ) is called a quaternionic kähler manifold if there exist locally defined 1 forms ω 1 , ω 2 , ω 3 such that for α { 1 , 2 , 3 } , we have
X 1 J α = ω α + 2 ( X 1 ) J α + 1 ω α + 1 ( X 1 ) J α + 2 ,
for X 1 Γ ( T N 1 ) , where the indices are taken from { 1 , 2 , 3 } modulo 3 . If there exists a global parallel quaternionic Hermitian basis { J 1 , J 2 , J 3 } of sections of E on N 1 , then ( N 1 , E , g 1 ) is called a hyperkähler. The structure { J 1 , J 2 , J 3 , g 1 } , where g 1 , a hyperkähler metric, is called a hyperkähler structure on N 1 .
A map π : ( N 1 , E 1 , g 1 ) ( N 2 , E 2 , g 2 ) is called an ( E 1 , E 2 ) -holomorphic map if for any point p N 1 and J ( E 1 ) p , there exists J ( E 2 ) π ( p ) such that
π J = J π .
A Riemannian submersion between quaternionic kähler manifolds π : ( N 1 , E 1 , g 1 ) ( N 2 , E 2 , g 2 ) , which is an ( E 1 , E 2 ) -holomorphic map, is known as a quaternionic kähler submersion (or a hyperkähler submersion) [9]:
Definition 1
([23]). A Riemannian map π from the almost quaternionic Hermitian manifold ( N 1 , E , g 1 ) to the Riemannian manifold ( N 2 , g 2 ) is called an h-semi-slant Riemannian map if, given a point p N 1 with a neighborhood U, there exists a quaternionic Hermitian basis { I , J , K } of sections of E on U such that for any R { I , J , K } , the following is true:
ker π = D 1 D 2 , R ( D 1 ) = D 1 ,
in which the angle θ R = θ R ( Z 1 ) between R Z 1 and the space ( D 2 ) q is constant for a non-zero Z 1 ( D 2 ) q and q U , where D 2 is an orthogonal complement of D 1 in ker π .
Furthermore, assume we have
θ = θ I = θ J = θ K ,
Then, we call the map π : ( N 1 , E , g 1 ) ( N 2 , g 2 ) a strictly h-semi-slant Riemannian map, the basis { I , J , K } a strictly h-semi-slant basis and the angle θ a strictly h-semi-slant angle.

3. h-Quasi-Hemi-Slant Riemannian Maps

Motivated by the studies given in Section 2, we give the definition of the h-qhs Riemannian map as follows:
Definition 2.
A Riemannian map π from the almost quaternionic Hermitian manifold ( N 1 , E , g 1 ) to the Riemannian manifold ( N 2 , g 2 ) is called an h-qhs Riemannian map if, given a point p N 1 with a neighborhood U, there exists a quaternionic Hermitian basis { I , J , K } of sections of E on U such that for any R { I , J , K } , there is a distribution D ( ker π ) on U such that
ker π = D D 1 D 2 , R ( D ) = D , R ( D 2 ) ( ker π ) ,
and the angle θ R = θ R ( Z 1 ) between R Z 1 and the space ( D 1 ) q is constant for a non-zero Z 1 ( D 1 ) q and q U , where ker π admits three orthogonal complementary distributions D , D 1 and D 2 such that D is invariant, D 1 is a slant with an angle θ R and D 2 is anti-invariant.
We call the basis { I , J , K } an h-qhs basis and the angles { θ I , θ J , θ K } h-qhs angles.
Furthermore, let us say we have
θ = θ I = θ J = θ K ,
Then, we call the map π : ( N 1 , E , g 1 ) ( N 2 , g 2 ) a strictly h-qhs Riemannian map, the basis { I , J , K } a strictly quasi-hemi-slant basis and the angle θ a strictly quasi-hemi-slant angle:
Definition 3.
A Riemannian map π from the almost quaternionic Hermitian manifold ( N 1 , E , g 1 ) to the Riemannian manifold ( N 2 , g 2 ) is called an almost h-qhs Riemannian map if, given a point p N 1 with a neighborhood U, there exists a quaternionic Hermitian basis { I , J , K } of sections of E on U such that for any R { I , J , K } , there is a distribution D R ( ker π ) on U such that
ker π = D R D 1 R D 2 R , R ( D R ) = D R , R ( D 2 R ) ( ker π ) ,
and the angle θ R = θ R ( Z 1 ) between R Z 1 and the space ( D 1 R ) q is constant for a non-zero Z 1 ( D 1 R ) q and q U , where the vertical distribution ker π admits three orthogonal complementary distributions D R , D 1 R and D 2 R such that D R is invariant, D 1 R is a slant with an angle θ R and D 2 R is anti-invariant.
We call the basis { I , J , K } an almost h-qhs basis and the angles { θ I , θ J , θ K } almost h-qhs angles.
Let π : ( N 1 , E , g 1 ) ( N 2 , g 2 ) be an almost h-qhs Riemannian map. We can easily observe the following:
(a)
If dim D R 0 , dim D 1 R 0 , 0 < θ R < π 2 and dim D 2 R = 0 , then π is an almost proper h-semi-slant Riemannian map with a semi-slant angle θ R ;
(b)
If dim D R = 0 , dim D 1 R 0 , 0 < θ R < π 2 and dim D 2 R 0 , then π is an almost h-hemi-slant Riemannian map.
We say that the almost h-qhs Riemannian map π : ( N 1 , E , g 1 ) ( N 2 , g 2 ) is proper if D R { 0 } , D 2 R { 0 } and θ R 0 , π 2 . Thus, one can easily see that the h-hemi-slant Riemannian map, h-semi-invariant Riemannian map and h-semi-slant Riemannian map are examples of h-qhs Riemannian maps.
Thus, we have
( ker π ) = ω R ( D 1 R ) R ( D 2 R ) μ R .
Obviously, μ R is an invariant sub-bundle of ( ker π ) with respect to the complex structure R.
For V 1 Γ ( ker π ) , we have
V 1 = P R V 1 + Q R V 1 + S R V 1 ,
where P R V 1 Γ ( D R ) , Q R V 1 Γ ( D 1 R ) , S R V 1 Γ ( D 2 R ) and R { I , J , K } .
For Z 1 Γ ( ker π ) , we obtain
R Z 1 = ϕ R Z 1 + ω R Z 1 ,
where ϕ R Z 1 Γ ( ker π ) , ω R Z 1 Γ ( ker π ) and R { I , J , K } .
For X 1 Γ ( ker π ) , we have
R X 1 = B R X 1 + C R X 1 ,
where B R X 1 Γ ( ker π ) , C R X 1 Γ ( μ R ) and R { I , J , K } .
We will denote an almost h-qhs Riemannian map from a hyperkähler manifold ( N 1 , I , J , K , g 1 ) onto a Riemannian manifold ( N 2 , g 2 ) such that ( I , J , K ) is an almost h-qhs basis by π .
The following lemmas can be easily obtained:
Lemma 1.
For π : ( N 1 , g 1 , E 1 ) ( N 2 , g 2 , E 2 ) , we get
ϕ R D R = D R , ω R D R = 0 , ϕ R D 2 R = 0 , ω R D 2 R ( ker π ) ,
where R { I , J , K } .
Lemma 2.
For π : ( N 1 , g 1 , E 1 ) ( N 2 , g 2 , E 2 ) , we have
ϕ R 2 Z 1 + B R ω R Z 1 = Z 1 , ω R ϕ R Z 1 + C R ω R Z 1 = 0 , ϕ R B R Z 2 + B R C R Z 2 = 0 , ω R B R Z 2 + C R 2 Z 2 = Z 2 ,
for any Z 1 Γ ( ker π ) , Z 2 Γ ( ker π ) and R { I , J , K } .
Proof. 
Using Equations (9), (13) and (14), we can find all equations of Lemma 2: □
Lemma 3.
With π : ( N 1 , I , J , K , g 1 ) ( N 2 , g 2 ) being an almost h-qhs Riemannian map, we then obtain
V X 1 ϕ R X 2 + T X 1 ω R X 2 = B R T X 1 X 2 + ϕ R V X 1 X 2 ,
T X 1 ϕ R X 2 + H X 1 ω R X 2 = C R T X 1 X 2 + ω R V X 1 X 2 ,
T X 1 B R Z 1 + H X 1 C R Z 1 = C R H X 1 Z 1 + ω R T X 1 Z 1 ,
V X 1 B R Z 1 + T X 1 C R Z 1 = B R H X 1 Z 1 + ϕ T X 1 Z 1 ,
V Z 1 ϕ R X 1 + A Z 1 ω R X 1 = B R A Z 1 X 1 + ϕ R V Z 1 X 1 ,
A Z 1 ϕ R X 1 + H Z 1 ω R X 1 = C R A Z 1 X 1 + ω R V Z 1 X 1 ,
A Z 1 B R Z 2 + H Z 1 C R Z 2 = C R H Z 1 Z 2 + ω R A Z 1 Z 2 ,
V Z 1 B R Z 2 + A Z 1 C R Z 2 = B R H Z 1 Z 2 + ϕ R A Z 1 Z 2 ,
for X 1 , X 2 Γ ( ker π ) , Z 1 , Z 2 Γ ( ker π ) and R { I , J , K } .
Proof. 
Using Equations (4)–(7), (13) and (14), we can easily obtain Equations (15)–(22). □
Now, we define
( X 1 ϕ R ) X 2 = V X 1 ϕ R X 2 ϕ R V X 1 X 2 ,
( X 1 ω R ) X 2 = H X 1 ω R X 2 ω R V X 1 X 2 ,
( Z 1 B R ) Z 2 = V Z 1 B R Z 2 B R H Z 1 Z 2 ,
( Z 1 C R ) Z 2 = H Z 1 C R Z 2 C R H Z 1 Z 2 ,
for X 1 , X 2 Γ ( ker π ) , Z 1 , Z 2 Γ ( ker π ) and R { I , J , K } .
Lemma 4.
For π : ( N 1 , I , J , K , g 1 ) ( N 2 , g 2 ) , we find
( X 1 ϕ R ) X 2 = B R T X 1 X 2 T X 1 ω R X 2 , ( X 1 ω R ) X 2 = C R T X 1 X 2 T X 1 ϕ R X 2 ,
( Z 1 C R ) Z 2 = ω R A Z 1 Z 2 A Z 1 B R Z 2 , ( Z 1 B R ) Z 2 = ϕ R A Z 1 Z 2 A Z 1 C R Z 2 ,
for all X 1 , X 2 Γ ( ker π ) , Z 1 , Z 2 Γ ( ker π ) and R { I , J , K } .
Proof. 
Using Equations (15) and (16) as well as Equations (21)–(26), Lemma 4 follows. □
If the tensors ϕ R and ω R are parallel with respect to the linear connection on N 1 , then
B R T X 1 X 2 = T X 1 ω R X 2 , C R T X 1 X 2 = T X 1 ϕ R X 2 ,
for all X 1 , X 2 Γ ( ker π ) and R { I , J , K } :
Lemma 5.
Let π : ( N 1 , E , g 1 ) ( N 2 , g 2 ) , be an almost h-qhs Riemannian map. Then, we obtain
ϕ R 2 V 1 = cos 2 θ R V 1 ,
for any non-zero vector field V 1 Γ ( D 1 R ) and R { I , J , K } , where { I , J , K } is an almost h-qhs basis with the almost h-qhs angles { θ I , θ J , θ K } .
Proof. 
For any non-zero vector field V 1 Γ ( D 1 R ) and R { I , J , K } , we have
cos θ R = ϕ R V 1 R V 1 ,
and
cos θ R = g 1 ( R V 1 , ϕ R V 1 ) ϕ R V 1 R V 1 ,
where θ R ( V 1 ) is the h-qhs angle.
Using Equations (9) and (13), we obtain
cos θ R = g 1 ( V 1 , ϕ R 2 V 1 ) ϕ R V 1 R V 1 .
From Equations (29) and (30), Equation (27) follows. □
Theorem 1.
Let π be an h-qhs Riemannian map from an almost hyperkahler manifold ( N 1 , I , J , K , g 1 ) to a Riemannian manifold ( N 2 , g 2 ) . Then, the following cases are equivalent:
(a)
D R is integrable;
(b)
g 1 ( T Z 2 I Z 1 T Z 1 I Z 2 , ω I Q I U 1 + I S I U 1 ) = g 1 ( V Z 1 I Z 2 V Z 2 I Z 1 , ϕ I Q I U 1 )
for Z 1 , Z 2 Γ ( D I ) and U 1 Γ ( D 1 I D 2 I ) ;
(c)
g 1 ( T Z 2 J Z 1 T Z 1 J Z 2 , ω J Q J U 1 + J S J U 1 ) = g 1 ( V Z 1 J Z 2 V Z 2 J Z 1 , ϕ J Q J U 1 )
for Z 1 , Z 2 Γ ( D J ) and U 1 Γ ( D 1 J D 2 J ) ;
(d)
g 1 ( T Z 2 K Z 1 T Z 1 K Z 2 , ω K Q K U 1 + K S K U 1 ) = g 1 ( V Z 1 K Z 2 V Z 2 K Z 1 , ϕ K Q K U 1 )
for Z 1 , Z 2 Γ ( D K ) and U 1 Γ ( D 1 K D 2 K ) .
Proof. 
For Z 1 , Z 2 Γ ( D R ) , U 1 Γ ( D 1 R D 2 R ) , U 2 ( ker π ) and R { I , J , K } , since [ Z 1 , Z 2 ] ( ker π ) , we have g 1 ( [ Z 1 , Z 2 ] , U 2 ) = 0 . Thus, D R is integrable ⇔ g 1 ( [ Z 1 , Z 2 ] , U 1 ) = 0 . Now, using Equations (4) and (12)–(14), we have
g 1 ( [ Z 1 , Z 2 ] , U 1 ) = g 1 ( R Z 1 Z 2 , R U 1 ) g 1 ( R Z 2 Z 1 , R U 1 ) , = g 1 ( Z 1 R Z 2 , R U 1 ) g 1 ( Z 2 R Z 1 , R U 1 ) , = g 1 ( T Z 1 R Z 2 T Z 2 R Z 1 , ω R Q R U 1 + J R U 1 ) g 1 ( V Z 1 R Z 2 V Z 2 R Z 1 , ϕ R Q R U 1 ) .
Since D R is R-invariant, we have
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Therefore, we obtain the result. □
Theorem 2.
The following cases are equivalent for the map π defined in Theorem 1:
(a)
D 1 R is integrable;
(b)
g 1 ( T Y 1 ω I ϕ I Y 2 T Y 2 ω I ϕ I Y 1 , V 1 ) = g 1 ( T Y 1 ω I Y 2 T Y 2 ω I Y 1 , ϕ I P I V 1 ) + g 1 ( H Y 1 ω I Y 2 H Y 2 ω I Y 1 , ω I S I V 1 )
for all Y 1 , Y 2 Γ ( D 1 I ) and V 1 Γ ( D I D 2 I ) ;
(c)
g 1 ( T Y 1 ω J ϕ J Y 2 T Y 2 ω J ϕ J Y 1 , V 1 ) = g 1 ( T Y 1 ω J Y 2 T Y 2 ω J Y 1 , ϕ J P J V 1 ) + g 1 ( H Y 1 ω J Y 2 H Y 2 ω J Y 1 , ω J S J V 1 )
for all Y 1 , Y 2 Γ ( D 1 J ) and V 1 Γ ( D J D 2 J ) ;
(d)
g 1 ( T Y 1 ω K ϕ K Y 2 T Y 2 ω K ϕ K Y 1 , V 1 ) = g 1 ( T Y 1 ω K Y 2 T Y 2 ω K Y 1 , ϕ K P K V 1 ) + g 1 ( H Y 1 ω K Y 2 H Y 2 ω K Y 1 , ω K S K V 1 )
for all Y 1 , Y 2 Γ ( D 1 K ) and V 1 Γ ( D K D 2 K ) .
Proof. 
For Y 1 , Y 2 Γ ( D 1 R ) , V 1 Γ ( D R D 2 R ) , V 2 ( ker F ) and R { I , J , K } , since [ Y 1 , Y 2 ] ( ker π ) , we have g 1 ( [ Y 1 , Y 2 ] , V 2 ) = 0 . Thus, D 1 R is integrable ⇔ g 1 ( [ Y 1 , Y 2 ] , V 1 ) = 0 . Using Equations (4), (5), (12) and (13) as well as Lemma 5 , we have
g 1 ( [ Y 1 , Y 2 ] , V 1 ) = g 1 ( Y 1 R Y 2 , R V 1 ) g 1 ( Y 2 R Y 1 , R V 1 ) , = g 1 ( Y 1 ϕ R Y 2 , R V 1 ) + g 1 ( Y 1 ω R Y 2 , R V 1 ) g 1 ( Y 2 ϕ R Y 1 , R V 1 ) g 1 ( Y 2 ω R Y 1 , R V 1 ) , = cos 2 θ R g 1 ( Y 1 Y 2 , V 1 ) cos 2 θ R g 1 ( Y 2 Y 1 , V 1 ) g 1 ( T Y 1 ω R ϕ R Y 2 T Y 2 ω R ϕ R Y 1 , V 1 ) + g 1 ( H Y 1 ω R Y 2 + T Y 1 ω R Y 2 , R P R V 1 + ω R S R V 1 ) g 1 ( H Y 2 ω R Y 1 + T Y 2 ω R Y 1 , R P R V 1 + ω R S R V 1 ) ,
which gives
sin 2 θ 1 g 1 ( [ Y 1 , Y 2 ] , V 1 ) = g 1 ( T Y 1 ω R Y 2 T Y 2 ω R Y 1 , R P R V 1 ) + g 1 ( H Y 1 ω R Y 2 H Y 2 ω R Y 1 , ω R S R V 1 ) g 1 ( T Y 1 ω R ϕ R Y 2 T Y 2 ω R ϕ R Y 1 , V 1 ) .
Since D 1 R is an R-slant distribution, therefore, we obtain
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Therefore, we find the result. □
Theorem 3.
For the h-qhs Riemannian map π defined in Theorem 1, D 2 R is always integrable.
Proof. 
We can easily prove the Theorem as hemi-slant case given in [21]. □
Theorem 4.
For the h-qhs Riemannian map π defined in Theorem 1, any one of the following assertions implies the others:
(a)
( ker π ) defines a totally geodesic foliation on N 1 ;
(b)
g 1 ( A Z 1 Z 2 , P I W 1 + cos 2 θ I Q I W 1 ) = g 1 ( H Z 1 Z 2 , ω I ϕ I P I W 1 + ω I ϕ I Q I W 1 ) g 1 ( A Z 1 B I Z 2 + H Z 1 C I Z 2 , ω I W 1 )
for Z 1 , Z 2 Γ ( ker π ) and W 1 Γ ( ker π ) ;
(c)
g 1 ( A Z 1 Z 2 , P J W 1 + cos 2 θ J Q J W 1 ) = g 1 ( H Z 1 Z 2 , ω J ϕ J P J W 1 + ω J ϕ J Q J W 1 ) g 1 ( A Z 1 B J Z 2 + H Z 1 C J Z 2 , ω J W 1 )
for Z 1 , Z 2 Γ ( ker π ) and W 1 Γ ( ker π ) ;
(d)
g 1 ( A Z 1 Z 2 , P K W 1 + cos 2 θ K Q K W 1 ) = g 1 ( H Z 1 Z 2 , ω K ϕ K P K W 1 + ω K ϕ K Q K W 1 ) g 1 ( A Z 1 B K Z 2 + H Z 1 C K Z 2 , ω K W 1 )
for Z 1 , Z 2 Γ ( ker π ) and W 1 Γ ( ker π ) .
Proof. 
For Z 1 , Z 2 Γ ( ker π ) , W 1 Γ ( ker π ) and R { I , J , K } , using Equations (6), (7) and (12)–(14) as well as Lemma 5, we have
g 1 ( Z 1 Z 2 , W 1 ) = g 1 ( R Z 1 Z 2 , R W 1 ) , = g 1 ( R Z 1 Z 2 , ϕ R P R W 1 + ϕ R Q R W 1 + ω R Q R W 1 + ω R S R W 1 ) , = g 1 ( Z 1 Z 2 , ϕ R 2 P R W 1 + ω R ϕ R P R W 1 + ω R ϕ R Q R W 1 ) + g 1 ( Z 1 B R Z 2 , ω R Q R W 1 + ω R S R W 1 ) + g 1 ( Z 1 C R Z 2 , ω R Q R W 1 + ω R S R W 1 ) , = g 1 ( A Z 1 Z 2 , P R W 1 + cos 2 θ R Q R W 1 ) g 1 ( H Z 1 Z 2 , ω R ϕ R P R W 1 + ω R ϕ R Q R W 1 ) + g 1 ( A Z 1 B R Z 2 , ω R Q R W 1 + ω R S R W 1 ) + g 1 ( H Z 1 C R Z 2 , ω R Q R W 1 + ω R S R W 1 ) .
Thus, we obtain
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Therefore, the result follows. □
Theorem 5.
The following conditions are equivalent for the h-qhs Riemannian map π:
(a)
( ker π ) defines a totally geodesic foliation on N 1 ;
(b)
g 1 ( T X 1 P I X 2 + cos 2 θ I T X 1 Q I X 2 , Y 1 ) = g 1 ( H X 1 ω I ϕ I P I X 2 + H X 1 ω I ϕ I Q I X 2 , Y 1 ) g 1 ( H X 1 ω I Q I X 2 + H X 1 ω I S I X 2 , C I Y 1 ) g 1 ( T X 1 ω I Q I X 2 + T X 1 ω I S I X 2 , B I Y 1 )
for X 1 , X 2 Γ ( ker π ) and Y 1 Γ ( ker π ) ;
(c)
g 1 ( T X 1 P J X 2 + cos 2 θ J T X 1 Q J X 2 , Y 1 ) = g 1 ( H X 1 ω J ϕ J P J X 2 + H X 1 ω J ϕ J Q J X 2 , Y 1 ) g 1 ( H X 1 ω J Q J X 2 + H X 1 ω J S J X 2 , C J Y 1 ) g 1 ( T X 1 ω J Q J X 2 + T X 1 ω J S J X 2 , B J Y 1 )
for X 1 , X 2 Γ ( ker π ) and Y 1 Γ ( ker π ) ;
(d)
g 1 ( T X 1 P K X 2 + cos 2 θ K T X 1 Q K X 2 , Y 1 ) = g 1 ( H X 1 ω K ϕ K P K X 2 + H X 1 ω K ϕ K Q K X 2 , Y 1 ) g 1 ( H X 1 ω K Q K X 2 + H X 1 ω K S K X 2 , C K Y 1 ) g 1 ( T X 1 ω K Q K X 2 + T X 1 ω K S K X 2 , B K Y 1 )
for X 1 , X 2 Γ ( ker π ) and Y 1 Γ ( ker π ) .
Proof. 
For X 1 , X 2 Γ ( ker π ) , Y 1 Γ ( ker π ) and R { I , J , K } , using Equations (4), (5) and (12)–(14) as well as Lemma 5, we have
g 1 ( X 1 X 2 , Y 1 ) = g 1 ( R X 1 X 2 , R Y 1 ) , = g 1 ( X 1 ϕ R P R X 2 , R Y 1 ) + g 1 ( X 1 ϕ R Q R X 2 , R Y 1 ) + g 1 ( X 1 ω R Q R X 2 , R Y 1 ) + g 1 ( X 1 ω R S R X 2 , R Y 1 ) , = g 1 ( T X 1 P R X 2 , Y 1 ) + cos 2 θ R g 1 ( T X 1 Q R X 2 , Y 1 ) g 1 ( H X 1 ω R ϕ R P R X 2 , Y 1 ) g 1 ( H X 1 ω R ϕ R Q R X 2 , Y 1 ) + g 1 ( H X 1 ω R Q R X 2 + H X 1 ω R S R X 2 , C R Y 1 ) + g 1 ( T X 1 ω R Q R X 2 + T X 1 ω R S R X 2 , B R Y 1 ) .
Thus, we obtain
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Therefore, the result follows. □
Theorem 6.
Let π be an h-qhs Riemannian map from an almost hyperkahler manifold ( N 1 , I , J , K , g 1 ) to a Riemannian manifold ( N 2 , g 2 ) . Then, any one of the following assertions implies the others:
(a)
D R defines a totally geodesic foliation on N 1 ;
(b)
g 1 ( T Z 1 I P I Z 2 , ω I Q I Y 1 + ω I S I Y 1 ) = g 1 ( V Z 1 I P I Z 2 , ϕ I Y 1 ) , g 1 ( T Z 1 I P I Z 2 , C I Y 2 ) = g 1 ( V Z 1 I P I Z 2 , B I Y 2 )
for Z 1 , Z 2 Γ ( D I ) , Y 1 Γ ( D 1 I D 2 I ) and Y 2 Γ ( ker π ) ;
(c)
g 1 ( T Z 1 J P J Z 2 , ω J Q J Y 1 + ω J S J Y 1 ) = g 1 ( V Z 1 J P J Z 2 , ϕ J Y 1 ) , g 1 ( T Z 1 J P J Z 2 , C J Y 2 ) = g 1 ( V Z 1 J P J Z 2 , B J Y 2 )
for Z 1 , Z 2 Γ ( D J ) , Y 1 Γ ( D 1 J D 2 J ) and Y 2 Γ ( ker π ) ;
(d)
g 1 ( T Z 1 K P K Z 2 , ω K Q K Y 1 + ω K S K Y 1 ) = g 1 ( V Z 1 K P K Z 2 , ϕ K Y 1 ) , g 1 ( T Z 1 K P K Z 2 , C K Y 2 ) = g 1 ( V Z 1 K P K Z 2 , B K Y 2 )
for Z 1 , Z 2 Γ ( D K ) , Y 1 Γ ( D 1 K D 2 K ) and Y 2 Γ ( ker π ) .
Proof. 
For Z 1 , Z 2 Γ ( D R ) , Y 1 Γ ( D 1 R D 2 R ) , Y 2 Γ ( ker π ) and R { I , J , K } , using Equations (4), (12) and (13), we have
g 1 ( Z 1 Z 2 , Y 1 ) = g 1 ( Z 1 R Z 2 , R Y 1 ) , = g 1 ( Z 1 R P R Z 2 , R Q R Y 1 + R S R Y 1 ) , = g 1 ( T Z 1 ϕ R P R Z 2 , ω R Q R Y 1 + ω R S R Y 1 ) + g 1 ( V Z 1 ϕ R P R Z 2 , ϕ R Q R Y 1 ) .
Moreover, using Equations (4), (12) and (14), we obtain
g 1 ( Z 1 Z 2 , Y 2 ) = g 1 ( Z 1 R Z 2 , R Y 2 ) , = g 1 ( Z 1 R P R Z 2 , B R Y 2 + C R Y 2 ) , = g 1 ( V Z 1 R P R Z 2 , B R Y 2 ) + g 1 ( T Z 1 J P R Z 2 , C R Y 2 ) .
Hence, we have
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Therefore, the result follows. □
Theorem 7.
With π : ( N 1 , I , J , K , g 1 ) ( N 2 , g 2 ) being an h-qhs Riemannian map, the following conditions are equivalent:
(a)
D 1 R defines a totally geodesic foliation on N 1 ;
(b)
g 1 ( T Y 1 ω I ϕ I Y 2 , Z 1 ) = g 1 ( T Y 1 ω I Y 2 , ϕ I P I Z 1 ) + g 1 ( H Y 1 ω I Y 2 , ω I S I Z 1 ) , g 1 ( H Y 1 ω I ϕ I Y 2 , Z 2 ) = g 1 ( H Y 1 ω I Y 2 , C I Z 2 ) + g 1 ( T Y 1 ω I Y 2 , B I Z 2 )
for Y 1 , Y 2 Γ ( D 1 I ) , Z 1 Γ ( D I D 2 I ) and Z 2 Γ ( ker π ) ;
(c)
g 1 ( T Y 1 ω J ϕ J Y 2 , Z 1 ) = g 1 ( T Y 1 ω J Y 2 , ϕ J P J Z 1 ) + g 1 ( H Y 1 ω J Y 2 , ω J S J Z 1 ) , g 1 ( H Y 1 ω J ϕ J Y 2 , Z 2 ) = g 1 ( H Y 1 ω J Y 2 , C J Z 2 ) + g 1 ( T Y 1 ω J Y 2 , B J Z 2 )
for Y 1 , Y 2 Γ ( D 1 J ) , Z 1 Γ ( D J D 2 J ) and Z 2 Γ ( ker π ) ;
(d)
g 1 ( T Y 1 ω K ϕ K Y 2 , Z 1 ) = g 1 ( T Y 1 ω K Y 2 , ϕ K P K Z 1 ) + g 1 ( H Y 1 ω K Y 2 , ω K S K Z 1 ) , g 1 ( H Y 1 ω K ϕ K Y 2 , Z 2 ) = g 1 ( H Y 1 ω K Y 2 , C K Z 2 ) + g 1 ( T Y 1 ω K Y 2 , B K Z 2 )
for Y 1 , Y 2 Γ ( D 1 K ) , Z 1 Γ ( D K D 2 K ) and Z 2 Γ ( ker π ) .
Proof. 
For Y 1 , Y 2 Γ ( D 1 R ) , Z 1 Γ ( D R D 2 R ) , Z 2 Γ ( ker π ) and R { I , J , K } , using Equations (5), (12) and (13) as well as Lemma 5 , we have
g 1 ( Y 1 Y 2 , Z 1 ) = g 1 ( Y 1 R Y 2 , R Z 1 ) , = g 1 ( Y 1 ϕ R Y 2 , R Z 1 ) + g 1 ( Y 1 ω R Y 2 , R Z 1 ) , = cos 2 θ R g 1 ( Y 1 Y 2 , Z 1 ) g 1 ( T Y 1 ω R ϕ R Y 2 , Z 1 ) + g 1 ( T Y 1 ω R Y 2 , ϕ R P R Z 1 ) + g 1 ( H Y 1 ω R Y 2 , ω R S R Z 1 ) ,
which gives
sin 2 θ R g 1 ( Y 1 Y 2 , Z 1 ) = g 1 ( T Y 1 ω R ϕ R Y 2 , Z 1 ) + g 1 ( T Y 1 ω R Y 2 , R P R Z 1 ) + g 1 ( H Y 1 ω R Y 2 , ω R S R Z 1 ) .
Moreover, from Equations (5), (13) and (14) as well as Lemma 5, we have
g 1 ( Y 1 Y 2 , Z 2 ) = g 1 ( Y 1 R Y 2 , R Z 2 ) , = g 1 ( Y 1 ϕ R Y 2 , R Z 2 ) + g 1 ( Y 1 ω R Y 2 , R Z 2 ) , = cos 2 θ R g 1 ( Y 1 Y 2 , Z 2 ) g 1 ( H Y 1 ω R ϕ R Y 2 , Z 2 ) + g 1 ( H Y 1 ω R Y 2 , C R Z 2 ) + g 1 ( T Y 1 ω R Y 2 , B R Z 2 ) .
Thus, we find that
sin 2 θ R g 1 ( Y 1 Y 2 , Z 2 ) = g 1 ( H Y 1 ω R ϕ R Y 2 , Z 2 ) + g 1 ( H Y 1 ω R Y 2 , C R Z 2 ) + g 1 ( T Y 1 ω R Y 2 , B R Z 2 ) .
Hence, we have
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Therefore, the result follows. □
Theorem 8.
For the h-qhs Riemannian map π defined in Theorem 1, any one of the following assertions implies the others:
(a)
D 2 R defines a totally geodesic foliation on N 1 ;
(b)
g 1 ( H Y 1 ω I Y 2 , ω I Q I W 1 ) = g 1 ( T Y 1 ω I S I Y 2 , ϕ I P I W 1 + ϕ I Q I W 1 ) , g 1 ( H Y 1 ω I S I Y 2 , C I W 2 ) = g 1 ( T Y 1 ω I S I Y 2 , B I W 2 )
for Y 1 , Y 2 Γ ( D 2 I ) , W 1 Γ ( D I D 1 I ) and W 2 Γ ( ker π ) ;
(c)
g 1 ( H Y 1 ω J Y 2 , ω J Q J W 1 ) = g 1 ( T Y 1 ω J S J Y 2 , ϕ J P J W 1 + ϕ J Q J W 1 ) , g 1 ( H Y 1 ω J S Y 2 , C J W 2 ) = g 1 ( T Y 1 ω J S Y 2 , B J W 2 )
for Y 1 , Y 2 Γ ( D 2 J ) , W 1 Γ ( D J D 1 J ) and W 2 Γ ( ker π ) ;
(d)
g 1 ( H Y 1 ω K Y 2 , ω K Q K W 1 ) = g 1 ( T Y 1 ω K S Y 2 , ϕ K P K W 1 + ϕ K Q K W 1 ) , g 1 ( H Y 1 ω K S Y 2 , C K W 2 ) = g 1 ( T Y 1 ω K S K Y 2 , B K W 2 )
for Y 1 , Y 2 Γ ( D 2 K ) , W 1 Γ ( D K D 1 K ) and W 2 Γ ( ker π ) .
Proof. 
For Y 1 , Y 2 Γ ( D 2 R ) , W 1 Γ ( D R D 1 R ) , W 2 Γ ( ker π ) and R { I , J , K } , using Equations (5), (12) and (13), we have
g 1 ( Y 1 Y 2 , W 1 ) = g 1 ( Y 1 R Y 2 , R W 1 ) = g 1 ( Y 1 ω R S R Y 2 , ϕ R P R W 1 + ϕ R Q R W 1 + ω R Q R W 1 ) , = g 1 ( T Y 1 ω R S R Y 2 , ϕ R P R W 1 + ϕ R Q R W 1 ) + g 1 ( H Y 1 ω R S R Y 2 , ω R Q R W 1 ) .
Again, using Equations (5), (13) and (14), we have
g 1 ( Y 1 Y 2 , W 2 ) = g 1 ( Y 1 R Y 2 , R W 2 ) = g 1 ( Y 1 ω R S R Y 2 , B R W 2 + C R W 2 ) , = g 1 ( T Y 1 ω R S R Y 2 , B R W 2 ) + g 1 ( H Y 1 ω R R Y 2 , C R W 2 ) .
Hence, we have
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Therefore, the result follows. □
Theorem 9.
Let π be an h-qhs Riemannian map from an almost hyperkahler manifold ( N 1 , I , J , K , g 1 ) to a Riemannian manifold ( N 2 , g 2 ) . Then, the following conditions are equivalent:
(a)
π is a totally geodesic map;
(b)
g 1 ( T Y 1 P I Y 2 + cos 2 θ I T Y 1 Q I Y 2 H Y 1 ω I ϕ I P I Y 2 H Y 1 ω I ϕ I Q I Y 2 , W 1 ) = g 1 ( T Y 1 ω I Q I Y 2 + T Y 1 ω I S I Y 2 , B I W 1 ) + g 1 ( H Y 1 ω I ϕ I Q I Y 2 + H Y 1 ω I ϕ I S I Y 2 , W 1 ) ,
g 1 ( A W 1 P I Y 1 + cos 2 θ I A W 1 Q I Y 1 H W 1 ω I ϕ I P I Y 1 H W 1 ω I ϕ I Q I Y 1 , W 2 ) = g 1 ( A W 1 ω I Q I Y 1 + A W 1 ω I S I Y 1 , B I W 2 ) + g 1 ( H W 1 ω I Q I Y 1 + H W 1 ω I S I Y 1 , C I W 2 )
for Y 1 , Y 2 Γ ( ker π ) and W 1 , W 2 Γ ( ker π ) ;
(c)
g 1 ( T Y 1 P J Y 2 + cos 2 θ J T Y 1 Q J Y 2 H Y 1 ω J ϕ J P J Y 2 H Y 1 ω J ϕ J Q J Y 2 , W 1 ) = g 1 ( T Y 1 ω J Q J Y 2 + T Y 1 ω J S J Y 2 , B J W 1 ) + g 1 ( H Y 1 ω J ϕ J Q J J Y 2 + H Y 1 ω J ϕ J S J Y 2 , W 1 ) ,
g 1 ( A W 1 P J Y 1 + cos 2 θ J A W 1 Q J Y 1 H W 1 ω J ϕ J P J Y 1 H W 1 ω J ϕ J Q J Y 1 , W 2 ) = g 1 ( A W 1 ω J Q J Y 1 + A W 1 ω J S J Y 1 , B J W 2 ) + g 1 ( H W 1 ω J Q J Y 1 + H W 1 ω J S J Y 1 , C J W 2 )
for Y 1 , Y 2 Γ ( ker π ) and W 1 , W 2 Γ ( ker π ) ;
(d)
g 1 ( T Y 1 P K Y 2 + cos 2 θ K T Y 1 Q K Y 2 H Y 1 ω K ϕ K P K Y 2 H Y 1 ω K ϕ K Q K Y 2 , W 1 ) = g 1 ( T Y 1 ω K Q K Y 2 + T Y 1 ω K S K Y 2 , B K W 1 ) + g 1 ( H Y 1 ω K ϕ K Q K Y 2 + H Y 1 ω K ϕ K S K Y 2 , W 1 ) ,
g 1 ( A W 1 P K Y 1 + cos 2 θ K A W 1 Q K Y 1 H W 1 ω K ϕ K P K Y 1 H W 1 ω K ϕ K Q K Y 1 , W 2 ) = g 1 ( A W 1 ω K Q K Y 1 + A W 1 ω K S K Y 1 , B K W 2 ) + g 1 ( H W 1 ω K Q K Y 1 + H W 1 ω K S K Y 1 , C K W 2 )
for Y 1 , Y 2 Γ ( ker π ) and W 1 , W 2 Γ ( ker π ) .
Proof. 
Since π is a Riemannian map, therefore, we have
( π ) ( W 1 , W 2 ) = 0 ,
for W 1 , W 2 Γ ( ker π ) .
For Y 1 , Y 2 Γ ( ker π ) , W 1 , W 2 Γ ( ker π ) and R { I , J , K } , using Equations (4), (5) and (12)–(14) as well as Lemma 5, we have
g 2 ( ( π ) ( Y 1 , Y 2 ) , π ( W 1 ) ) = g 1 ( Y 1 Y 2 , W 1 ) = g 1 ( Y 1 R Y 2 , R W 1 ) = g 1 ( Y 1 R P R Y 2 , R W 1 ) g 1 ( Y 1 R Q R Y 2 , R W 1 ) g 1 ( Y 1 R S R Y 2 , R W 1 ) , = g 1 ( Y 1 ϕ R P R Y 2 , R W 1 ) g 1 ( Y 1 ϕ R Q R Y 2 , R W 1 ) g 1 ( Y 1 ω R Q R Y 2 , R W 1 ) g 1 ( Y 1 ω R S R Y 2 , R W 1 ) , = g 1 ( T Y 1 P R Y 2 + cos 2 θ R T Y 1 Q R Y 2 H Y 1 ω R ϕ R P R Y 2 H Y 1 ω R ϕ R Q R Y 2 , W 1 ) g 1 ( T Y 1 ω R Q R Y 2 + T Y 1 ω R S R Y 2 , B R W 1 ) g 1 ( H Y 1 ω R ϕ R Q R Y 2 + H Y 1 ω R ϕ R S R Y 2 , W 1 ) .
Moreover, using Equations (4), (5) and (12)–(14) as well as Lemma 5, we have
g 2 ( ( π ) ( W 1 , Y 1 ) , π ( W 2 ) ) = g 1 ( W 1 Y 1 , W 2 ) , = g 1 ( W 1 R Y 1 , R W 2 ) , = g 1 ( W 1 R P R Y 1 , R W 2 ) g 1 ( W 1 R Q R Y 1 , R W 2 ) g 1 ( W 1 R S R Y 1 , R W 2 ) , = g 1 ( W 1 ϕ R P R Y 1 , R W 2 ) g 1 ( W 1 ϕ R Q R Y 1 , R W 2 ) g 1 ( W 1 ω R Q R Y 1 , R W 2 ) g 1 ( W 1 ω R S R Y 1 , R W 2 ) , = g 1 ( A W 1 P R Y 1 + cos 2 θ R A W 1 Q R Y 1 H W 1 ω R ϕ R P R Y 1 H W 1 ω R ϕ R Q R Y 1 , W 2 ) g 1 ( A W 1 ω R Q R Y 1 + A W 1 ω R S R Y 1 , B R W 2 ) g 1 ( H W 1 ω R Q R Y 1 + H W 1 ω R S R Y 1 , C R W 2 ) .
Hence, we obtain
( a ) ( b ) , ( a ) ( c ) , ( a ) ( d ) .
Thus, the theorem is proven. □

4. Example

Note that given a Euclidean space R 4 n with coordinates ( x 1 , x 2 , ..... , x 4 n ) , we can canonically choose complex structures I, J and K on R 4 n as follows:
I ( x 4 s + 1 ) = x 4 s + 2 , I ( x 4 s + 2 ) = x 4 s + 1 , I ( x 4 s + 3 ) = x 4 s + 4 , I ( x 4 s + 4 ) = x 4 s + 3 , J ( x 4 s + 1 ) = x 4 s + 3 , J ( x 4 s + 2 ) = x 4 s + 4 , J ( x 4 s + 3 ) = x 4 s + 1 , J ( x 4 s + 4 ) = x 4 s + 2 , K ( x 4 s + 1 ) = x 4 s + 4 , K ( x 4 s + 2 ) = x 4 s + 3 , K ( x 4 s + 3 ) = x 4 s + 2 , K ( x 4 s + 4 ) = x 4 s + 1 ,
for s { 0 , 1 , 2 , .... , , n 1 } .
Then, we can easily check that ( I , J , K , , ) is a hyperkähler structure on R 4 n , where , denotes the Euclidean metric on R 4 n . Throughout this section, we will use these notations.
Example 1.
Define a map π : R 12 R 6 by
π ( x 1 , x 2 , ......... , x 12 ) = ( 2020 , x 2 , x 6 , x 8 x 9 2 , x 10 , 2022 ) .
Then, the map π is an almost h-qhs Riemannian map such that
ker π = x 1 , x 3 , x 4 , x 5 , x 7 , 1 2 ( x 8 + x 9 ) , x 11 , x 12 ,
( ker π ) = x 2 , x 6 , 1 2 ( x 8 x 9 ) , x 10 ,
D I = x 3 , x 4 , x 11 , x 12 , D 1 I = x 7 , 1 2 ( x 8 + x 9 ) , D 2 I = x 1 , x 5 ,
D J = x 1 , x 3 , x 5 , x 7 , D 1 J = 1 2 ( x 8 + x 9 ) , x 11 , D 2 J = x 4 , x 12 ,
D K = x 1 , x 4 , D 1 K = x 5 , 1 2 ( x 8 + x 9 ) , x 12 , D 2 K = x 3 , x 7 , x 11 ,
with the almost h-qhs angles { θ I = θ J = θ K = π 4 } .
Example 2.
Define a map π : R 16 R 8 by
π ( x 1 , x 2 , ......... , x 16 ) = ( 101 , 3 x 5 x 9 2 , x 6 , x 8 , x 11 , x 14 , 202 , x 15 ) .
Then, the map π is an almost h-qhs Riemannian map such that
ker π = x 1 , x 2 , x 3 , x 4 , 1 2 ( x 5 + 3 x 9 ) , x 7 , x 10 , x 12 , x 13 , x 16 ,
( ker π ) = 1 2 ( 3 x 5 x 9 ) , x 6 , x 8 , x 11 , x 14 , x 15 , ,
D I = x 1 , x 2 , x 3 , x 4 , D 1 I = 1 2 ( x 5 + 3 x 9 ) , x 10 , D 2 I = x 7 , x 12 , x 13 , x 16 ,
D J = x 1 , x 2 , x 3 , x 4 , x 10 , x 12 , D 1 J = 1 2 ( x 5 + 3 x 9 ) , x 7 , D 2 J = x 13 , x 16 ,
D K = x 1 , x 2 , x 3 , x 4 , x 13 , x 16 , D 1 K = 1 2 ( x 5 + 3 x 9 ) , x 12 , D 2 K = x 7 , x 10 ,
with the almost h-qhs angles { θ I = π 6 , θ J = π 3 , θ K = π 6 } .

Author Contributions

Conceptualization, M.B., R.P., A.H. and S.K. (Sushil Kumar); methodology, M.B., 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, M.B., A.H., S.K. (Sushil Kumar) and S.K. (Sumeet Kumar); writing—review and editing, M.B., 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

The authors would like to thank the Deanship of Scientific Research at Umm Al-Qura University for supporting this work under Grant Code 22UQU4330007DSR04.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are thankful to the editor and anonymous referees for the constructive comments to improve the quality of this 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. Chen, B.Y. Geometry of Slant Submaniflods; Katholieke Universiteit: Leuven, Belgium, 1990. [Google Scholar]
  3. Sahin, B. Riemannian Submersions, Riemannian Maps in Hermitian Geometry, and Their Applications; Elsevier/Academic Press: Amsterdam, The Netherlands, 2017. [Google Scholar]
  4. O’Neill, B. The fundamental equations of a submersion. Mich. Math. J. 1966, 13, 458–469. [Google Scholar] [CrossRef]
  5. Gray, A. Pseudo-Riemannian almost product manifolds and submersions. J. Math. Mech. 1967, 16, 715–737. [Google Scholar]
  6. Watson, B. Almost Hermitian submersions. J. Differ. Geom. 1976, 11, 147–165. [Google Scholar] [CrossRef]
  7. Park, K.S. H-anti-invariant submersions from almost quaternionic Hermitian manifolds. Czechoslov. Math. J. 2017, 67, 557–578. [Google Scholar] [CrossRef] [Green Version]
  8. Park, K.S. H-semi-invariant submersions. Taiwan. J. Math. 2012, 16, 1865–1878. [Google Scholar] [CrossRef]
  9. Park, K.S. H-semi-slant submersions from almost quaternionic Hermitian manifolds Taiwan. J. Math. 2014, 18, 1909–1926. [Google Scholar]
  10. Bourguignon, J.P.; Awson, H.B. Stability and isolation phenomena for Yang-Mills fields. Commun. Math. Phys. 1981, 79, 189–230. [Google Scholar] [CrossRef]
  11. Bourguignon, J.P. A mathematician’s visit to Kaluza-Klein theory. Rend. Sem. Mat. Univ. Pol. Torino. 1989, 143–163. [Google Scholar]
  12. Cortes, V.; Mayer, C.; Mohaupt, T.; Saueressig, F. Special geometry of Euclidean supersymmetry: Vector multiplets. J. High Energy Phys. 2008, 3, 028. [Google Scholar] [CrossRef] [Green Version]
  13. Kraines, V.Y. Topology of quaternionic manifolds. Trans. Am. Math. Soc. 1966, 122, 357–367. [Google Scholar] [CrossRef]
  14. Guan, D. On Riemann-Roch Formula and Bounds of the Betti Numbers of Irreducible Compact Hyperkähler Manifold-n = 4. 1999; preprint. [Google Scholar]
  15. Guan, D. On the Betti numbers of irreducible compact hyperkähler manifolds of complex dimension four. Math. Res. Lett. 2001, 8, 663–669. [Google Scholar] [CrossRef] [Green Version]
  16. Sahin, B. Invariant and anti-invariant Riemannian maps to Kahler manifolds. Int. J. Geom. Methods Mod. Phys. 2010, 7, 355–377. [Google Scholar] [CrossRef]
  17. Sahin, B. Semi-invariant Riemannian maps from almost Hermitian manifolds. Indag. Math. 2012, 23, 80–94. [Google Scholar] [CrossRef]
  18. Prasad, R.; Kumar, S. Slant Riemannian maps from Kenmotsu manifolds into Riemannian manifolds. Global J. Pure App. Math. 2017, 13, 1143–1155. [Google Scholar]
  19. 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]
  20. Prasad, R.; Kumar, S. Semi-slant Riemannian maps from cosymplectic manifolds into Riemannian manifolds. Gulf J. Math. 2020, 9, 62–80. [Google Scholar]
  21. Sahin, B. Hemi-slant Riemannian maps. Mediterr. J. Math. 2017, 14, 10. [Google Scholar] [CrossRef]
  22. Prasad, R.; Kumar, S.; Kumar, S.; Vanli, A.T. On quasi-hemi-slant Riemannian maps. GU J Sci. 2021, 34, 477–491. [Google Scholar] [CrossRef]
  23. Park, K.S. Almost h-semi-slant Riemannian map. Taiwan. J. Math. 2013, 17, 937–956. [Google Scholar] [CrossRef]
  24. Kumar, S.; Bilal, M.; Prasad, R.; Haseeb, A.; Chen, Z. V-quasi-bi-slant Riemannian maps. Symmetry 2022, 14, 1360. [Google Scholar] [CrossRef]
  25. 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. [Google Scholar] [CrossRef]
  26. Park, K.S. H-slant submersions. Bull. Korean Math. Soc. 2012, 49, 329–338. [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

Bilal, M.; Kumar, S.; Prasad, R.; Haseeb, A.; Kumar, S. On h-Quasi-Hemi-Slant Riemannian Maps. Axioms 2022, 11, 641. https://doi.org/10.3390/axioms11110641

AMA Style

Bilal M, Kumar S, Prasad R, Haseeb A, Kumar S. On h-Quasi-Hemi-Slant Riemannian Maps. Axioms. 2022; 11(11):641. https://doi.org/10.3390/axioms11110641

Chicago/Turabian Style

Bilal, Mohd, Sushil Kumar, Rajendra Prasad, Abdul Haseeb, and Sumeet Kumar. 2022. "On h-Quasi-Hemi-Slant Riemannian Maps" Axioms 11, no. 11: 641. https://doi.org/10.3390/axioms11110641

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