Skip to main content

Theory and Modern Applications

The base of warped product submanifolds of Sasakian space forms characterized by differential equations

Abstract

In the present paper, we find some characterization theorems. Under certain pinching conditions on the warping function satisfying some differential equation, we show that the base of warped product submanifolds of a Sasakian space form \(\widetilde{M}^{2m+1}(\epsilon )\) is isometric either to a Euclidean space \(\mathbb{R}^{n}\) or a warped product of a complete manifold N and the Euclidean line \(\mathbb{R}\).

1 Introduction and main results

One of the significant motives for the geometric analysis of Riemannian manifolds \((\Omega, g)\) is the study of the impact of differential equations on its geometry, as well as isometric properties. Furthermore, it is well-known that their classification has an extensive influence on the global analysis of a Riemannian manifold with differential equations. It should be noted that [21, 31, 33, 34] gave characterizations of Euclidean spaces by analyzing differential equations. They showed that a nonconstant function ψ on a complete manifold \((\Omega ^{n},g)\) satisfies the following equation:

$$\begin{aligned} \nabla ^{2}\psi +cg=0 \end{aligned}$$
(1)

if and only if \((\Omega ^{n}, g)\) is isometric to some Euclidean space \(\mathbb{R}^{n}\), where c is any positive constant. Another characterization using a differential equation has been discovered by Río, Kupeli, and Unal [21]. They demonstrated that the complete Riemannian manifold \((\Omega ^{n}, g)\) is isometric to the warped product of a complete Riemannian manifold N and a Euclidean line \(\mathbb{R}\) with warping function θ satisfying the differential equation

$$\begin{aligned} \frac{d^{2}\theta }{dt^{2}}+\lambda _{1}\theta =0 \end{aligned}$$
(2)

if and only if there exists a real-valued nonconstant function ψ associated to a negative eigenvalue \(\lambda _{1}\leq 0\), which satisfies the following differential equation:

$$\begin{aligned} \Delta \psi +\lambda _{1}\psi =0. \end{aligned}$$
(3)

Such complete space classifications are extremely attractive and have been studied by several mathematicians (see, e.g., [3, 4, 6, 7, 15, 16, 1820]). For example, by using (1), Al, Dayel, Deshmukh, and Belova [1] showed that a connected and complete Riemannian manifold \((\Omega ^{n}, g)\) is isometric to \(\mathbb{R}^{n}\) if and only if the nontrivial concircular vector field u along the function ψ satisfies \(R(\nabla \psi, \nabla \psi )=0\) or \(\Delta {\mathbf{u}}=0\). In [13], Chen and Deshmukh proved that a complete Riemannian manifold admits a concurrent vector field if and only if it is isometric to a Euclidean space by (1). Similarly, in [14], it has been shown that \((\Omega ^{n}, g)\) is isometric to a Euclidean space if and only if \((\Omega ^{n}, g)\) permits a nontrivial gradient conformal vector field, that is, a Jacobi-type vector field. On the other hand, Matsuyama [24] derived a characterization stating that if the complete totally real submanifold \(\Omega ^{n}\) for the complex projective space \(C\mathbb{P}^{n}\) with bounded Ricci curvature admits a function ψ satisfying (3), for \(\lambda _{1}\leq n\), then \(\Omega ^{n}\) is isometric to the hyperbolic space component that is connected if \((\nabla \psi )_{x}=0\) or if it is isometric to the warped product of a complete Riemannian manifold and the Euclidean line if ψ is nonvanishing, where the warping function θ on \(\mathbb{R}\) satisfies equation (2). Furthermore, similar results have been obtained for generalized Sasakian space forms by Jamali and Shahid [22]. In this study, inspired by [13, 5, 7, 912, 22, 30, 35], we derive a similar characterization for C-totally real warped product submanifolds of Sasakian space forms as rigidity theorems. To prove our main result, the next lemma, which was proved in [26], will be stated.

Lemma 1.1

Suppose \(\widetilde{M}^{2m+1}(\epsilon )\) is a Sasakian space form and let \(\Phi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \widetilde{M}^{2m+1}(\epsilon )\) be a C-totally real immersion of the warped product submanifold \(\Omega ^{n}\) into \(\widetilde{M}^{2m+1}(\epsilon )\) such that the base \(\mathbb{B}\) is minimal. Then, the Ricci inequality is given as

$$\begin{aligned} \mathscr {R}ic(X)+q\Delta \ln f\leq \frac{n^{2}}{4} \Vert H \Vert ^{2}+q \Vert \nabla \ln f \Vert ^{2}+ \frac{\epsilon +3}{4} \{pq+n-1 \}, \end{aligned}$$
(4)

for every unit vector \(X\in T_{x}\Omega ^{n}\), where \(p=\dim \mathbb{B}\) and \(q=\dim \mathbb{F}\). The quantities in the above inequality have been discussed, in detail, in [26].

The following abbreviations are used towards the end of this paper: ‘SSF’ stands for Sasakian space form, ‘WF’ for warping function, and ‘WPS’ for warped product submanifold. More precisely, we give the next theorem:

Theorem 1.1

Let \(\Phi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \mathbb{M}^{2m+1}(\epsilon )\) be a C-totally real isometric immersion from a WPS \(\Omega ^{n}\) into the SSF \(\mathbb{M}^{2m+1}(\epsilon )\) such that the Ricci curvature is bounded below by a positive constant \(K>0\). Then, a complete minimal base \(\mathbb{B}\) is isometric to a Euclidean space \(\mathbb{R}^{p}\) if the following equality holds:

$$\begin{aligned} (\lambda _{1}+q )K=\lambda _{1} \biggl\{ \frac{q\lambda _{1}}{p}+ \frac{n^{2}}{4} \Vert H \Vert ^{2}+ \frac{(\epsilon +3)}{4}(pq+n-1) \biggr\} . \end{aligned}$$
(5)

The next result is motivated by the study of Río, Kupeli, and Unal [21].

Theorem 1.2

Assuming that \(\Phi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \mathbb{M}^{2m+1}(\epsilon )\) is a C-totally real isometric immersion of a complete WPS \(\Omega ^{n}\) into the SSF \(\mathbb{M}^{2m+1}(\epsilon )\) such that its Ricci curvature is bounded below by a positive constant \(K>0\). Let the complete base \(\mathbb{B}\) be minimal in \(\mathbb{M}^{2m+1}(\epsilon )\) and satisfy the following assumption:

$$\begin{aligned} n^{2} \vert \mathbb{H} \vert ^{2}+\frac{4pq}{\lambda _{1}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}=\frac{4p}{\lambda _{1}} \biggl( \frac{\epsilon +3}{4} (1-pq-n )+K \biggr) \end{aligned}$$
(6)

for \(\lambda _{1}<0\). Then, \(\mathbb{B}\) is isometric to a warped product of the form \(\mathbb{R}\times _{\theta }N\) with the warping function θ satisfying the following differential equation:

$$\begin{aligned} \frac{d^{2}\theta }{dt^{2}}+\lambda _{1}\theta =0. \end{aligned}$$

Remark 1.1

The paper deals with ordinary differential equations on C-totally real warped product submanifolds. By optimizing the warping function of a C-totally real warped product submanifold of Sasakian space forms, we studied characterizations theorems for a C-totally real warped product submanifold of Sasakian space forms. Therefore, the paper exhibits an excellent combination of the theory of ordinary differential equations with Riemannian geometry.

2 Notation and formulas

The almost contact metric manifold \((\widetilde{{M}}, g)\) with Riemannian metric g preserves the following conditions:

$$\begin{aligned} & \phi ^{2} =-I+\xi \otimes \eta, \\ &\eta (\xi ) =1, \qquad\phi (\xi )=0, \qquad \eta \circ \phi =0, \\ &g(\phi W_{1}, \phi W_{2}) =g(W_{1}, W_{2})-\eta (W_{1})\eta (W_{2}), \\ &\eta (W_{1}) =g(W_{1}, \xi ), \end{aligned}$$
(7)

for the almost contact structure \((\phi,\eta,\zeta )\) and \(W_{1}, W_{2}\in \Gamma (T\widetilde{{M}})\). A manifold \(\widetilde{{M}}^{2m+1}\) is defined to be a Sasakian manifold if the following relation holds:

$$\begin{aligned} (\widetilde{\nabla }_{W_{1}}\phi )W_{2}=g(W_{1}, W_{2})\zeta -\eta (W_{2})W_{1}. \end{aligned}$$
(8)

It follows that

$$\begin{aligned} \widetilde{\nabla }_{W_{1}}\zeta =-\phi W_{1}, \end{aligned}$$
(9)

for every \(W_{1}, W_{2}\in \Gamma (T\widetilde{{M}})\), where ̃ denotes the Riemannian connection with respect to the metric g. A Sasakian space form is a Sasakian manifold considering constant ϕ-sectional curvature ϵ, which is also defined as \({\widetilde{M}}^{2m+1}(\epsilon )\). Consequently, in [6, Eq. (6)], the Riemannian curvature tensor of \({\widetilde{M}}^{2m+1}(\epsilon )\) is defined in detail, which is also usually defined from . If the structure field ξ is perpendicular to the submanifold \(\Omega ^{n}\) in \({\widetilde{M}}^{2m+1}(\epsilon )\), then \(\Omega ^{n}\) is a C-totally real submanifold of \({\widetilde{M}}^{2m+1}(\epsilon )\). Furthermore, in this case, ϕ maps any tangent space of \(\Omega ^{n}\) into its corresponding normal space (see [2, 8, 23, 25, 32, 36]). Now, we recall the Bochner formula [8] for a differentiable function on a Riemannian manifold \(\Omega ^{n}\), that is, \(\psi: \Omega ^{n}\rightarrow \mathbb{R}\). Then, we have that

$$\begin{aligned} \frac{1}{2}\Delta \vert \nabla \psi \vert ^{2}=\mathrm{Ric}_{\Omega ^{n}}( \nabla \psi, \nabla \psi )+ \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}+g \bigl(\nabla ( \Delta \psi ), \nabla \psi \bigr), \end{aligned}$$
(10)

where the Ricci tensor of \(\Omega ^{n}\) is denoted by Ric.

3 The main results

3.1 Proof of Theorem 1.1

Equation (4) gives

$$\begin{aligned} \mathscr {R}ic(X)+q\Delta \psi \leq \frac{n^{2}}{4} \Vert H \Vert ^{2}+q \Vert \nabla \psi \Vert ^{2}+\frac{\epsilon +3}{4} \{pq+n-1 \}. \end{aligned}$$

Assuming the Ricci curvature is bounded below by a positive constant \(K>0\) (i.e., \(\mathscr {R}ic(X)\geq K\)), we get

$$\begin{aligned} K+q\Delta \psi \leq \frac{n^{2}}{4} \Vert H \Vert ^{2}+q \Vert \nabla \psi \Vert ^{2}+ \frac{\epsilon +3}{4} \{pq+n-1 \}. \end{aligned}$$
(11)

One of the most famous results connecting the curvature and topology of the complete Riemannian manifold \(\Omega ^{n}\) is a famous theorem of Myers [29], which states that if the Ricci curvature with respect to unit vectors on \(\mathbb{B}\) is bounded by a positive constant \(K>0\), then \(\mathbb{B}\) is compact. Then, integrating (11) and using Green’s lemma, we find that

$$\begin{aligned} \mathrm{Vol}(\mathbb{B})K&\leq \frac{n^{2}}{4} \int _{\mathbb{B} \times \{q\}} \vert \mathbb{H} \vert ^{2}\,dV+q \int _{\mathbb{B}\times \{q \}} \vert \nabla \psi \vert ^{2}\,dV+ \int _{\mathbb{B}\times \{q\}} \frac{\epsilon +3}{4} \{pq+n-1 \}\,dV. \end{aligned}$$

This can be written as

$$\begin{aligned} &\int _{\mathbb{B}\times \{q\}} \vert \nabla \psi \vert ^{2}\,dV\geq \frac{K}{q}\mathrm{Vol}(\mathbb{B})-\frac{n^{2}}{4q} \int _{\mathbb{B}\times \{q \}} \vert \mathbb{H} \vert ^{2}\,dV- \frac{1}{q} \int _{\mathbb{B}\times \{q \}}\frac{\epsilon +3}{4} \{pq+n-1 \}\,dV. \end{aligned}$$
(12)

On the other hand, we have

$$\begin{aligned} \bigl\vert \mathrm{Hess}(\psi )-tI \bigr\vert ^{2}= \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}+t^{2} \vert I \vert ^{2}-2tg\bigl(I, \mathrm{Hess}(\psi )\bigr), \end{aligned}$$

which leads to

$$\begin{aligned} \bigl\vert \mathrm{Hess}(\psi )-tI \bigr\vert ^{2}=2t\Delta \psi +t^{2}p+ \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}. \end{aligned}$$

Substituting \(t=\frac{\lambda _{1}}{p}\) and integrating the preceding equation with respect to the volume element dV, we obtain

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )- \frac{\lambda _{1}}{p}I \biggr\vert ^{2}\,dV= \int _{\mathbb{B}\times \{q\}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}\,dV+ \int _{\mathbb{B}\times \{q\}} \frac{\lambda _{1}^{2}}{p}\,dV. \end{aligned}$$
(13)

Using the Bochner formula (10), along with the fact that \(\Delta \psi =\lambda _{1}\psi \), we have

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}\,dV=-\lambda _{1} \int _{\mathbb{B}\times \{q\}}|\nabla \psi \|^{2}\,dV- \int _{\mathbb{B} \times \{q\}}\mathrm{Ric}(\nabla \psi, \nabla \psi )\,dV. \end{aligned}$$
(14)

Combining Eqs. (13) and (14), we derive

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )- \frac{\lambda _{1}}{p}I \biggr\vert ^{2}\,dV&= \int _{\mathbb{B}\times \{q\}} \frac{\lambda _{1}^{2}}{p}\,dV-\lambda _{1} \int _{\mathbb{B}\times \{q \}}|\nabla \psi \|^{2}\,dV \\ &\quad{}-\int _{\mathbb{B}\times \{q\}} \mathrm{Ric}( \nabla \psi, \nabla \psi )\,dV. \end{aligned}$$
(15)

As we assumed that \(\mathrm{Ric}(\nabla \psi, \nabla \psi )\geq K\) for \(K>0\), we have

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )- \frac{\lambda _{1}}{p}I \biggr\vert ^{2}\,dV\leq \int _{\mathbb{B}\times \{q\}} \frac{\lambda _{1}^{2}}{p}\,dV-\lambda _{1} \int _{\mathbb{B}\times \{q \}}|\nabla \psi \|^{2}\,dV-K\mathrm{Vol}( \mathbb{B}). \end{aligned}$$

Inserting Eq. (12) into the above equation, we derive

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )- \frac{\lambda _{1}}{p}I \biggr\vert ^{2}\,dV\leq {}& \int _{\mathbb{B}\times \{q\}} \frac{\lambda _{1}^{2}}{p}\,dV- \int _{\mathbb{B}\times \{q\}} \biggl( \frac{\lambda _{1}K}{q}+K \biggr)\,dV \\ &{}+\frac{\lambda _{1}n^{2}}{4q} \int _{\mathbb{B}\times \{q\}} \vert \mathbb{H} \vert ^{2}\,dV +\frac{\lambda _{1}}{q} \int _{\mathbb{B}\times \{q\}} \frac{\epsilon +3}{4} (pq+n-1 )\,dV. \end{aligned}$$
(16)

If (5) is satisfied, then (16) implies that

$$\begin{aligned} \biggl\vert \mathrm{Hess}(\psi )-\frac{\lambda _{1}}{p}I \biggr\vert ^{2}=0. \end{aligned}$$

Hence, we get

$$\begin{aligned} \mathrm{Hess}(\psi ) (V, V)=cg(V, V), \end{aligned}$$
(17)

for any \(V\in \Gamma (\mathbb{B})\) with constant \(c=\frac{\lambda _{1}}{p}\). Therefore, by applying the Tashiro theorems [31, 34], we obtain that \(\mathbb{B}\) is isometric to a Euclidean space \(\mathbb{R}^{p}\).

3.2 Proof of Theorem 1.2

Let us define the following equation with \(\psi =\ln f\). We have

$$\begin{aligned} \bigl\vert t\psi I+\mathrm{Hess}(\psi ) \bigr\vert ^{2}=&t^{2}( \psi )^{2} \vert I \vert ^{2}+ \bigl\vert \mathrm{Hess}( \psi ) \bigr\vert ^{2}+2t\psi g\bigl(I, \mathrm{Hess}(\psi )\bigr). \end{aligned}$$

However, it is well-known that \(|I |^{2}=\mathrm{tr}(II^{*})=p\), as well as \(g(\mathrm{Hess}(\psi ), I^{*})=\mathrm{tr}(I^{*}\mathrm{Hess}(\psi ))= \mathrm{tr}\mathrm{(Hess}(\psi ))\). Thus, the preceding equation takes the form

$$\begin{aligned} \bigl\vert t\psi I+\mathrm{Hess}(\psi ) \bigr\vert ^{2}= \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}+pt^{2}(\psi )^{2}-2t\psi \Delta \psi. \end{aligned}$$
(18)

If ψ is an eigenfunction associated to the eigenvalue \(\lambda _{1}\) such that \(\Delta \psi =\lambda _{1}\psi \), then we get

$$\begin{aligned} \bigl\vert t\psi I+\mathrm{Hess}(\psi ) \bigr\vert ^{2}= \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}+ \bigl(pt^{2}-2t\lambda _{1} \bigr) (\psi )^{2}. \end{aligned}$$
(19)

On the other hand, we obtain

$$\begin{aligned} \Delta \frac{\psi ^{2}}{2}=\psi \Delta \psi - \vert \nabla \psi \vert ^{2}. \end{aligned}$$

Using \(\Delta \psi =\lambda _{1}\psi \) again, we have

$$\begin{aligned} \Delta \frac{\psi ^{2}}{2}=\psi \lambda _{1}\psi - \vert \nabla \psi \vert ^{2}, \end{aligned}$$

which implies that

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}}(\psi )^{2}\,dV=\frac{1}{\lambda _{1}} \int _{\mathbb{B}\times \{q\}} \vert \nabla \psi \vert ^{2}\,dV. \end{aligned}$$
(20)

It follows, from (19) and (20), that

$$\begin{aligned} &\int _{\mathbb{B}\times \{q\}} \bigl\vert \mathrm{Hess}(\psi )+t\psi I \bigr\vert ^{2}\,dV= \int _{\mathbb{B}\times \{q\}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}\,dV+ \biggl(\frac{pt^{2}}{\lambda _{1}}-2t \biggr) \int _{\mathbb{B}\times \{q\}} \vert \nabla \psi \vert ^{2}\,dV. \end{aligned}$$
(21)

In particular, setting \(t=\frac{\lambda _{1}}{p}\) in (21) and integrating, we get

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )+ \frac{\lambda _{1}}{p}\psi I \biggr\vert ^{2}\,dV= \int _{\mathbb{B}\times \{q\}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}\,dV-\frac{\lambda _{1}}{p} \int _{ \mathbb{B}\times \{q\}} \vert \nabla \psi \vert ^{2}\,dV. \end{aligned}$$
(22)

Again taking the integral of (4) and involving the Green lemma, we have

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}}\mathscr {R}ic_{M}(X)\,dV\leq {}& \frac{n^{2}}{4} \int _{\mathbb{B}\times \{q\}} \vert \mathbb{H} \vert ^{2}\,dV+q \int _{ \mathbb{B}\times \{q\}} \vert \nabla \psi \vert ^{2}\,dV \\ &{}+ \int _{\mathbb{B}\times \{q\}}\frac{\epsilon +3}{4} (pq+n-1 )\,dV. \end{aligned}$$
(23)

From (22) and (23), we can obtain

$$\begin{aligned} \frac{1}{q} \int _{\mathbb{B}\times \{q\}}\mathscr {R}ic_{M}(X)\,dV\leq{}& \frac{n^{2}}{4q} \int _{\mathbb{B}\times \{q\}} \vert \mathbb{H} \vert ^{2}\,dV- \frac{p}{\lambda _{1}} \int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )+ \frac{\lambda _{1}}{n}\psi I \biggr\vert ^{2}\,dV \\ &{}+\frac{p}{\lambda _{1}} \int _{\mathbb{B}\times \{q\}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}\,dV+ \int _{\mathbb{B}\times \{q\}} \frac{\epsilon +3}{4} \biggl(p+1+\frac{p-1}{q} \biggr)\,dV. \end{aligned}$$

As we considered that the Ricci curvature is bounded (i.e., \(\mathscr {R}ic(X)\geq K\) for some \(K>0\)), the preceding equation implies that

$$\begin{aligned} &\int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )+ \frac{\lambda _{1}}{p}\psi I \biggr\vert ^{2}\,dV\\ &\quad\leq \frac{n^{2}\lambda _{1}}{4pq} \int _{\mathbb{B}\times \{q\}} \vert \mathbb{H} \vert ^{2}\,dV+ \int _{\Omega ^{n}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}\,dV-\frac{\lambda _{1}}{pq} \int _{\mathbb{B}\times \{q\}}K\,dV \\ &\qquad{}+\frac{\lambda _{1}}{p} \int _{\mathbb{B}\times \{q\}} \frac{\epsilon +3}{4} \biggl(p+1+\frac{p-1}{q} \biggr)\,dV, \end{aligned}$$

which is equivalent to the following:

$$\begin{aligned} &\int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )+ \frac{\lambda _{1}}{p}\psi I \biggr\vert ^{2}\,dV \\ &\quad\leq \int _{\mathbb{B}\times \{q\}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}\,dV \\ &\qquad{}+\frac{\lambda _{1}}{p} \int _{\mathbb{B}\times \{q\}} \biggl\{ \frac{n^{2}}{4q} \vert \mathbb{H} \vert ^{2}+\frac{(\epsilon +3)}{4} \biggl(p+1+ \frac{p-1}{q} \biggr)- \frac{K}{q} \biggr\} \,dV. \end{aligned}$$
(24)

This gives us the following inequality:

$$\begin{aligned} &\int _{\mathbb{B}\times \{q\}} \biggl\vert \mathrm{Hess}(\psi )+ \frac{\lambda _{1}}{p}\psi I \biggr\vert ^{2}\,dV \\ &\quad \leq \int _{\mathbb{B}\times \{q\}} \biggl\{ \frac{\lambda _{1}}{p} \biggl( \frac{n^{2}}{4q} \vert \mathbb{H} \vert ^{2}+\frac{(\epsilon +3)}{4} \biggl(p+1+\frac{p-1}{q} \biggr)-\frac{K}{q} \biggr) + \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2} \biggr\} \,dV. \end{aligned}$$
(25)

Our assumption is satisfied, that is,

$$\begin{aligned} n^{2} \vert \mathbb{H} \vert ^{2}+ \frac{4pq}{\lambda _{1}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}= \frac{4pq}{\lambda _{1}} \biggl( \frac{(\epsilon +3)}{4q} (1-pq-n )+\frac{K}{q} \biggr). \end{aligned}$$
(26)

Combining (25) and (26), we get

$$\begin{aligned} \biggl\vert \mathrm{Hess}(\psi )+\frac{\lambda _{1}}{p}\psi I \biggr\vert ^{2}\leq 0. \end{aligned}$$

The above equation gives us

$$\begin{aligned} \mathrm{Hess}(\psi )+\frac{\lambda _{1}}{p}\psi I=0. \end{aligned}$$
(27)

Taking the trace of the preceding equation, we can derive

$$\begin{aligned} \Delta \psi +\lambda _{1}\psi =0. \end{aligned}$$
(28)

According to [21], the base \(\mathbb{B}\) is isometric to the connected components of a hyperbolic space if \((\nabla \psi )_{x}=0\). However, \((\nabla \psi )_{x}=0\) leads to a contradiction, as \(\Omega ^{n}\) is a nontrivial warped product. Hence, \(\mathbb{B}\) is isometric to a warped product of the type \(\mathbb{R}\times _{\theta }N\), where N is a complete Riemannian manifold and \(\mathbb{R}\) is the Euclidean line. Moreover, the warping function θ satisfies the following differential equation:

$$\begin{aligned} \frac{d^{2}\theta }{dt^{2}}+\lambda _{1}\theta =0. \end{aligned}$$

Thus, the proof is completed.

Remark 3.1

It is well-known that \(\mathbb{R}^{2m+1}(-3)\) and \(\mathbb{S}^{2m+1}(1)\), considering standard Sasakian structures, may be seen as classical examples of Sasakian space forms with constant sectional curvature \(\epsilon =-3\) and \(\epsilon =1\), respectively [25, 27].

We produce a striking application of Theorem 1.1 and Remark 3.1 by selecting \(\epsilon =1\) (see [2]):

Corollary 3.1

Suppose \(\Psi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \mathbb{S}^{2m+1}(1)\) is a C-totally real isometric immersion of a complete WPS \(\Omega ^{n}\) into the SSF \(\mathbb{S}^{2m+1}(1)\) with Ricci curvature bounded below by a positive constant \(K>0\) and the base \(\mathbb{B}\) is minimal in \(\mathbb{S}^{2m+1}(1)\). Then, \(\mathbb{B}\) is isometric to the Euclidean space \(\mathbb{R}^{p}\) if the following equality holds:

$$\begin{aligned} (\lambda _{1}+q )K=\lambda _{1} \biggl\{ \frac{q\lambda _{1}}{p}+ \frac{n^{2}}{4} \Vert H \Vert ^{2}+(pq+n-1) \biggr\} . \end{aligned}$$
(29)

For the minimal case (i.e., \(\|H\|^{2}=0\)), we give the following corollary:

Corollary 3.2

Let \(\Psi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \mathbb{S}^{2m+1}(1)\) be a C-totally real minimal isometric immersion of a complete WPS \(\Omega ^{n}\) into the SSF \(\mathbb{S}^{2m+1}(1)\) such that the Ricci curvature is bounded below by a positive constant \(K>0\) satisfying the condition

$$\begin{aligned} (\lambda _{1}+q )K=\lambda _{1} \biggl\{ \frac{q\lambda _{1}}{p}+(pq+n-1) \biggr\} . \end{aligned}$$
(30)

Then, \(\mathbb{B}\) is isometric to a Euclidean space \(\mathbb{R}^{p}\).

Following Theorem 1.2, we give the following corollary:

Corollary 3.3

Assume that \(\Phi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \mathbb{S}^{2m+1}(1)\) is a C-totally real isometric immersion of the complete WPS \(\Omega ^{n}\) into the SSF \(\mathbb{S}^{2m+1}(1)\) such that Ricci curvature is bounded below by a positive constant \(K>0\) and the base \(\mathbb{B}\) is minimal in \(\mathbb{S}^{2m+1}(1)\), satisfying the assumption

$$\begin{aligned} n^{2} \vert \mathbb{H} \vert ^{2}+\frac{4pq}{\lambda _{1}} \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}=\frac{4p}{\lambda _{1}} \bigl\{ (1-pq-n )+K \bigr\} . \end{aligned}$$
(31)

Then, \(\mathbb{B}\) is isometric to a warped product of the form \(\mathbb{R}\times _{\theta }N\) with the warping function θ satisfying the following differential equation:

$$\begin{aligned} \frac{d^{2}\theta }{dt^{2}}+\lambda _{1}\theta =0. \end{aligned}$$

Corollary 3.4

Let \(\Phi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \mathbb{S}^{2m+1}(1)\) be a C-totally real minimal isometric immersion of a complete WPS \(\Omega ^{n}\) into the SSF \(\mathbb{S}^{2m+1}(1)\) such that Ricci curvature is bounded below by a positive constant \(K>0\) satisfying the assumption

$$\begin{aligned} q \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}= \bigl\{ (1-pq-n )+K \bigr\} . \end{aligned}$$
(32)

Then, \(\mathbb{B}\) is isometric to a warped product of the form \(\mathbb{R}\times _{\theta }N\) with warping function θ satisfying the differential equation (2).

Substituting the constant sectional curvature \(\epsilon =-3\) into Theorems 1.1 and 1.2, we can directly derive the following:

Corollary 3.5

Let \(\Phi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \mathbb{R}^{2m+1}(-3)\) be a C-totally real isometric immersion for the complete WPS \(\Omega ^{n}\) into the SSF \(\mathbb{R}^{2m+1}(-3)\) such that Ricci curvature is bounded below by a positive constant \(K>0\).

  1. (i)

    If the base \(\mathbb{B}\) is minimal in \(\mathbb{R}^{2m+1}(-3)\), then \(\mathbb{B}\) is isometric to the Euclidean space \(\mathbb{R}^{p}\) if the following equality holds:

    $$\begin{aligned} (\lambda _{1}+q )K=\lambda _{1} \biggl\{ \frac{q\lambda _{1}}{p}+ \frac{n^{2}}{4} \Vert H \Vert ^{2} \biggr\} . \end{aligned}$$
    (33)
  2. (ii)

    If Φ is a minimal isometric immersion in \(\mathbb{R}^{2m+1}(-3)\) and \(p (\lambda _{1}+q )K=q\lambda _{1}^{2}\) is satisfied, then \(\mathbb{B}\) is isometric to the Euclidean space \(\mathbb{R}^{p}\).

Using Theorem 1.2, we obtain the following:

Corollary 3.6

Under the same assumptions of Corollary 3.5, we have the following:

  1. (i)

    If the base \(\mathbb{B}\) is minimal in \(\mathbb{R}^{2m+1}(-3)\) and the following equality holds:

    $$\begin{aligned} \lambda _{1}n^{2} \vert \mathbb{H} \vert ^{2}+4pq \bigl\vert \mathrm{Hess}(\psi ) \bigr\vert ^{2}=4pK, \end{aligned}$$
    (34)

    then \(\mathbb{B}\) is isometric to a warped product of the form \(\mathbb{R}\times _{\theta }N\) with warping function θ satisfying the differential equation (2).

  2. (ii)

    If Φ is the minimal isometric immersion in \(\mathbb{R}^{2m+1}(-3)\) and \(|\mathrm{Hess}(\psi ) |^{2}=\frac{K}{q}\) is satisfied, then \(\mathbb{B}\) is isometric to the warped product of \(\mathbb{R}\times _{\theta }N\) with the warping function θ satisfying the differential equation (2).

Classifying the Dirichlet energy of smooth functions is treated as an integral procedure in the fields of physics and engineering. Moreover, the Dirichlet energy is formulated as an equivalent of kinetic energy. Let ψ be any real-valued smooth function on a compact manifold. Then, the Dirichlet energy of ψ is defined by

$$\begin{aligned} \mathscr {E}(\psi )=\frac{1}{2} \int \Vert \nabla \psi \Vert ^{2}\,dV. \end{aligned}$$
(35)

Using the above formula and Lemma 1.1, we obtain the following theorem:

Theorem 3.1

Suppose \(\widetilde{M}^{2m+1}(\epsilon )\) is a Sasakian space form and let \(\Phi:\Omega ^{n}=\mathbb{B}\times _{f}\mathbb{F}\rightarrow \widetilde{M}^{2m+1}(\epsilon )\) be a C-totally real immersion of the warped product submanifold \(\Omega ^{n}\) into \(\widetilde{M}^{2m+1}(\epsilon )\) such that the base \(\mathbb{B}\) is minimal. Then, the Dirichlet energy inequality is given by

$$\begin{aligned} \int _{\mathbb{B}\times \{q\}}\mathscr {R}ic(X)\,dV\leq {}&\frac{n^{2}}{4} \int _{\mathbb{B}\times \{q\}} \Vert \mathbb{H} \Vert ^{2}\,dV+2q \mathscr {E}( \psi )+\big\{ pq+n-1\big\} \int _{\mathbb{B}\times \{q\}} \frac{\epsilon +3}{4}\,dV, \end{aligned}$$
(36)

for every unit vector \(X\in T_{x}\Omega ^{n}\), where \(p=\dim \mathbb{B}\) and \(q=\dim \mathbb{F}\).

Proof

Taking the integral of Eq. (4) and using the Green lemma, we get the required result (36). This completes the proof of the theorem. □

Availability of data and materials

Not applicable.

References

  1. Al-Dayel, I., Deshmukh, S., Belova, O.: A remarkable property of concircular vector fields on a Riemannian manifold. Mathematics 8, 469 (2020)

    Article  Google Scholar 

  2. Ali, A., Laurian-Ioan, P.: Geometric classification of warped products isometrically immersed in Sasakian space forms. Math. Nachr. 292, 234–251 (2018)

    MATH  Google Scholar 

  3. Ali, A., Laurian-Ioan, P., Alkhaldi, A.H.: Ricci curvature on warped product submanifolds in spheres with geometric applications. J. Geom. Phys. 146, 103510 (2019)

    Article  MathSciNet  Google Scholar 

  4. Ali, A., Laurian-Ioan, P., Alkhaldi, A.H., Alqahtani, L.S.: Ricci curvature on warped product submanifolds of complex space forms and its applications. Int. J. Geom. Methods Mod. Phys. 16(9), 1950142 (2019)

    Article  MathSciNet  Google Scholar 

  5. Ali, A., Mofarreh, F., Mior Othman, W.A., Patra, D.S.: Applications of differential equations to characterize the base of warped product submanifolds of cosymplectic space forms. J. Inequal. Appl. 2020, 241 (2020)

    Article  MathSciNet  Google Scholar 

  6. Ali, R., Mofarreh, F., Alluhaibi, N., Ali, A., Ahmad, I.: On differential equations characterizing Legendrian submanifolds of Sasakian space forms. Mathematics 8, 150 (2020)

    Article  Google Scholar 

  7. Barros, A., Gomes, J.N., Ernani, J.R.: A note on rigidity of the almost Ricci soliton. Arch. Math. 100, 481–490 (2013)

    Article  MathSciNet  Google Scholar 

  8. Berger, M., Gauduchon, P., Mazet, E.: Le spectre d’une variété riemannienne. In: Le Spectre D’une Variété Riemannienne, pp. 141–241. Springer, Germany (1971)

    Chapter  Google Scholar 

  9. Chen, B.Y.: Relations between Ricci curvature and shape operator for submanifolds with arbitrary codimension. Glasg. Math. J. 41, 33–41 (1999)

    Article  MathSciNet  Google Scholar 

  10. Chen, B.Y.: On Ricci curvature of isotropic and Lagrangian submanifolds in complex space form. Arch. Math. (Basel) 74, 154–160 (2000)

    Article  MathSciNet  Google Scholar 

  11. Chen, B.Y.: On isometric minimal immersions from warped products into real space forms. Proc. Edinb. Math. Soc. 45, 579–587 (2002)

    Article  MathSciNet  Google Scholar 

  12. Chen, B.Y.: Differential Geometry of Warped Product Manifolds and Submanifolds. World Scientific, Singapore (2017)

    Book  Google Scholar 

  13. Chen, B.Y., Deshmukh, S.: Some results about concircular vector fields on Riemannian manifolds. Filomat (2021, to appear)

  14. Chen, B.Y., Deshmukh, S., Ishan, A.A.: On Jacobi-type vector fields on Riemannian manifolds. Mathematics 7(12), 1139 (2019)

    Article  Google Scholar 

  15. De, U.C., Yildiz, A., Funda Yalınız A.: Locally ϕ-symmetric normal almost contact metric manifolds of dimension 3. Appl. Math. Lett. 22(5), 723–727 (2009)

    Article  MathSciNet  Google Scholar 

  16. Deshmukh, S.: Characterizing spheres and Euclidean spaces by conformal vector fields. Ann. Mat. Pura Appl. 196(6), 2135–2145 (2017)

    Article  MathSciNet  Google Scholar 

  17. Deshmukh, S.: Almost Ricci solitons isometric to spheres. Int. J. Geom. Methods Mod. Phys. 16(5), 1950073 (2019)

    Article  MathSciNet  Google Scholar 

  18. Deshmukh, S., Al-Solamy, F.: A note on conformal vector fields on a Riemannian manifold. Colloq. Math. 136(1), 65–73 (2014)

    Article  MathSciNet  Google Scholar 

  19. Deshmukh, S., Al-Solamy, F.R.: Conformal gradient vector fields on a compact Riemannian manifold. Colloq. Math. 112(1), 157–161 (2008)

    Article  MathSciNet  Google Scholar 

  20. Erkekoğlu, F., García-Río, E., Kupeli, D.N., Ünal, B.: Characterizing specific Riemannian manifolds by differential equations. Acta Appl. Math. 76(2), 195–219 (2003)

    Article  MathSciNet  Google Scholar 

  21. García-Río, E., Kupeli, D.N., Unal, B.: On differential equations characterizing Euclidean sphere. J. Differ. Equ. 194(2), 287–299 (2003)

    Article  MathSciNet  Google Scholar 

  22. Jamali, M., Shahid, M.H.: Application of Bochner formula to generalized Sasakian space forms. Afr. Math. 29(7–8), 1135–1139 (2018)

    Article  MathSciNet  Google Scholar 

  23. Lee, J.W., Lee, C.W., Vîlcu, G.E.: Classification of Casorati ideal Legendrian submanifolds in Sasakian space forms. J. Geom. Phys. 155, 103768 (2020)

    Article  MathSciNet  Google Scholar 

  24. Matsuyama, Y.: Complete totally real submanifolds of a complex projective space. Differ. Geom.-Dyn. Syst. 20, 119–125 (2018)

    MathSciNet  MATH  Google Scholar 

  25. Mihai, I.: Ricci curvature of submanifolds in Sasakian space forms. J. Aust. Math. Soc. 72, 247–256 (2002)

    Article  MathSciNet  Google Scholar 

  26. Mofarreh, F., Ali, A., Alluhaibi, N., Belova, O.: Ricci curvature for warped product submanifolds of Sasakian space forms and its applications to differential equations. J. Math. 2021, 15 (2021)

    Article  Google Scholar 

  27. Munteanu, M.I., Vrancken, L.: Minimal contact CR-submanifolds in \(\mathbb{S}^{2n+1}\) satifisfying the \(\delta (2)\)-Chen inequality. J. Geom. Phys. 75, 92–97 (2014)

    Article  MathSciNet  Google Scholar 

  28. Mustafa, A., Uddin, S., Al-Solamy, F.R.: Chen-Ricci inequality for warped products in Kenmotsu space forms and its applications. Rev. R. Acad. Cienc. Exactas Fís. Nat., Ser. A Mat. 113, 3585–3602 (2019)

    Article  MathSciNet  Google Scholar 

  29. Myers, S.B.: Riemannian manifolds with positive mean curvature. Duke Math. J. 8(2), 401–404 (1941)

    Article  MathSciNet  Google Scholar 

  30. Obata, M.: Certain conditions for a Riemannian manifold to be isometric with a sphere. J. Math. Soc. Jpn. 14, 333–340 (1962)

    Article  MathSciNet  Google Scholar 

  31. Pigola, S., Rimoldi, M., Setti, A.G.: Remarks on non-compact gradient Ricci solitons. Math. Z. 268, 777–790 (2011)

    Article  MathSciNet  Google Scholar 

  32. Sasahara, T.: A class of biminimal Legendrian submanifolds in Sasakian space forms. Math. Nachr. 287(1), 79–90 (2014)

    Article  MathSciNet  Google Scholar 

  33. Tanno, S.: Some differential equations on Riemannian manifolds. J. Math. Soc. Jpn. 30(3), 509–531 (1978)

    Article  MathSciNet  Google Scholar 

  34. Tashiro, Y.: Complete Riemannian manifolds and some vector fields. Trans. Am. Math. Soc. 117, 251–275 (1965)

    Article  MathSciNet  Google Scholar 

  35. Tripathi, M.M.: Improved Chen–Ricci inequality for curvature-like tensors and its applications. Differ. Geom. Appl. 29, 685–698 (2011)

    Article  MathSciNet  Google Scholar 

  36. Tripathi, M.M.: C-totally real warped product submanifolds. An. Ştiinţ. Univ. ‘Al.I. Cuza’ Iaşi, Mat. 58, 417–436 (2012)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

This article is dedicated to the baby girl (Hania Akram) of Professor Akram Ali which was born on 12/05/2020.

Funding

This research was funded by the Deanship of Scientific Research at Princess Nourah bint Abdulrahman University through the Fast-track Research Funding Program.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed equally to the writing of this manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Fatemah Mofarreh.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ali, A., Agrawal, R.P., Mofarreh, F. et al. The base of warped product submanifolds of Sasakian space forms characterized by differential equations. Adv Differ Equ 2021, 69 (2021). https://doi.org/10.1186/s13662-021-03230-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13662-021-03230-1

Keywords