1. Introduction
One of the richest structures on a Riemannian manifold
is provided by the
Ricci soliton, which comprises a smooth vector field
and a real constant
satisfying
where
is the Lie derivative in the direction of
, and Ric is the Ricci tensor of
g. We denote a Ricci soliton by
. A Ricci soliton
is said to be a
trivial Ricci soliton if the soliton vector field
is a Killing vector field, that is,
, and in this case, the Ricci soliton is an Einstein manifold. This is one of the reasons that a Ricci soliton is considered to be a generalization of an Einstein manifold. Then, the notion of
almost Ricci soliton is introduced in (cf. [
1]) in an attempt to generalize Ricci solitons, by replacing the soliton constant
with a smooth function
. Thus, for an almost Ricci soliton
, we have
The geometry of Ricci solitons and almost Ricci solitons has been a subject of immense interest owing to their elegant geometry as well as to their nice applications (cf. [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11,
12,
13,
14,
15]). Given an almost Ricci soliton
, we call the vector field
the
soliton vector field and the smooth function
the
potential function. An almost Ricci soliton
is said to be a
trivial almost Ricci soliton if it is a Ricci soliton, that is, the potential function
is a real constant, and it is said to be
nontrivial if the potential function
is nonconstant. For an almost Ricci soliton
, we denote by
the smooth 1-form dual to the soliton vector field
. This leads to a skew-symmetric tensor
F defined by
where
is the algebra of smooth vector fields on
M. The skew-symmetric tensor
F is called the
associated tensor to the soliton vector field.
A well-known example of an
n-dimensional nontrivial almost Ricci soliton is provided by the
n-sphere
of constant curvature
c, on taking its eigenfunction
f satisfying
, where
f is a nonconstant function, and these are in abundance (for instance, the height functions on
). This function
f satisfies
where
is the Hessian of
f, and
g is the canonical metric of constant curvature on
. On taking
, the gradient of
f, we see that
, and, consequently, we have
where
and we have used Equation (
2) and the expression for the Ricci tensor
of the sphere
. Since
is a nonconstant function (as
f is), we see that
is a nontrivial almost Ricci soliton. For further examples of compact and noncompact nontrivial almost Ricci solitons, we refer to [
1,
3].
Though it appears as if there are hundreds of results in differential geometry mentioning solitons, the results on the Ricci soliton and almost Ricci soliton are not that many. In order to understand the spirit behind the introduction of the Ricci soliton and its natural generalization, the almost Ricci soliton, we have the following heat equation called Hamilton’s Ricci flow (cf. [
7]):
and a Ricci soliton
is a stable solution of the above heat equation (Hamilton’s Ricci flow) of the form
, satisfying
, where
is a diffeomorphism for
such that the one parameter group
induces the vector field
, and
is the scaling function. Similarly, an almost Ricci soliton
is a stable solution of the above Hamilton’s Ricci flow (
4) of the form
where the diffeomorphisms
are generated by the family of vector fields
, and
is the scaling, and it is a function of
t as well as the local coordinates
. Naturally, the initial conditions
,
imply
. On differentiating Equation (
5) with respect to
t and using Hamilton’s Ricci flow Equation (
4) at
, we find
which, on taking
and the function
, the above equation takes the form of Equation (
1), which defines an almost Ricci soliton. It is in this spirit that both the Ricci soliton and almost Ricci soliton being stable solutions of the heat Equation (
4) is important. It is like heat diffusion, where the Hamilton’s Ricci flow changes the Riemannian metric to yield the desired curvature that results in the uniform curvature distribution across the manifold. Therefore, it is worth mentioning that those solitons, which come through the solution of the heat equation, have a concrete physical significance. It is this power, inherited through the heat equation, that Ricci solitons used to settle the famous Poincare conjecture. They are not limited to this role in geometry, as Ricci solitons are also highly useful in brain mapping (cf. [
16]), as well as having a role to play in the study of gravity in physics (cf. [
17]). Since almost Ricci solitons are a natural generalization of Ricci solitons, they automatically carry an importance similar to that of Ricci solitons.
Two natural questions associated to the geometry of an almost Ricci soliton are the following:
(i) What are the conditions under which an n-dimensional compact almost Ricci soliton is isometric to ?
(ii) What are the conditions under which an almost Ricci soliton is a trivial Ricci soliton?
In this article, we answer these questions by proving two results with respect to question (i) and one result with respect to question (ii). In the first result, we show that, if the Ricci curvature Ric of an
n-dimensional compact and connected almost Ricci soliton
is positive, and for a nonzero constant
c,
satisfies a generic inequality, then these conditions guarantee that
is isometric to
(cf. Theorem 2). Similarly, we use the Hodge decomposition of the soliton vector field
to find another characterization of
(cf. Theorem 4). Finally, with respect to question (ii), we show that, if the affinity tensor of the soliton vector field
(introduced in
Section 3) vanishes and a certain inequality involving the scalar curvature holds, then these conditions imply that a compact and connected almost Ricci soliton is a trivial Ricci soliton (cf. Theorem 3).
As the proofs of our results of using integration on a compact Riemannian manifold, here we would like to add a small note on the use of integration in differential geometry. The use of integral formulas goes back to the year 1848, and it appeared through the following very well-known formula.
Theorem 1. (Gauss–Bonnet) [18]: Suppose that M is a compact two-dimensional Riemannian manifold with boundary . Let K be the Gaussian curvature of M, and let be the geodesic curvature of . Then,where is the area element of M, is the line element of the boundary , and is the Euler characteristic of M. The Gauss–Bonnet Theorem is an outstanding result, as it connects the geometry of
M (the Gaussian curvature
K and the geodesic curvature
) to the topology of
M (the Euler characteristic
). In particular, if
M is a compact surface without boundary, then
from the above formula disappears, and the formula states that the integral of the Gaussian curvature is equal to
times the Euler characteristic
, where
is the genus of the surface. Let us note that this innocent looking statement for a compact surface without boundary is very significant in the following sense. The Euler characteristic
remains the same under diffeomorphic changes, where the Gaussian curvature
K is not preserved under these diffeomorphisms (unless the metric is preserved, that is, unless the diffeomorphisms are isometries). The Gauss–Bonnet Theorem for compact surfaces without boundaries states that though
K is not preserved under a diffeomorphism, the total Gaussian curvature
is preserved. This formula has been generalized to higher dimensions by Chern [
19] and also by Buzano and Nguyen in [
20], and was later developed into an independent branch of differential topology.
Another important integral formula that is widely used is Stokes’s theorem, which, for a compact Riemannian manifold
without boundary, states [
18]
for a smooth vector field
X on
M, and
is the volume form of
M with respect to metric
g. In particular, for a smooth function
, on taking
, the above integral formula takes the form
There are different variants of Stokes’s formula that are used in differential geometry, namely, Minkowski’s formula for closed hypersurfaces in Euclidean spaces as well as in spaces of constant curvature (cf. [
21]). As a consequence of Minkowski’s formula, we derive a global result, namely, that there does not exist a compact minimal hypersurface in a Euclidean space.
3. Almost Ricci Solitons Isometric to a Sphere
Let be an n-dimensional compact almost Ricci soliton. In this section, we find conditions under which is isometric to . Indeed, we prove the following result.
Theorem 2. An n-dimensional compact and connected almost Ricci soliton , where , with scalar curvature τ and positive Ricci curvature, satisfiesfor a nonzero constant c, if and only if and is isometric to . Proof. Let
be an
n-dimensional compact and connected almost Ricci soliton
with positive Ricci curvature that satisfies
We intend to compute
; therefore, we choose a local orthonormal frame
and proceed as follows:
The Hessian operator
is symmetric and
F is skew-symmetric; therefore,
Using the above equation as well as Equation (
11) in Equation (
13), we are led to
Integrating the above equation leads to
Next, we use
in the above equation, and we obtain
that is,
We also have the following Bochner’s formula:
Now, for a nonzero constant
c, we have
which, on integration and using Equations (
14) and (
15), leads to
Rearranging the above equation, we have
that is,
Due to inequality (
12), the right-hand side of the above equation is nonpositive, whereas the almost Ricci soliton has positive Ricci curvature and the second integrand on the left-hand side of the above equation is non-negative owing to the Schwartz’s inequality
and consequently, we conclude
and the inequality in (
16) becomes the equality
. The holding of this inequality requires
Now, differentiating Equation (
17) and using Equation (
6), we reach
that is,
Since the constant
and the left-hand side is symmetric, while the right-hand side is skew-symmetric, we immediately conclude
and
Also, taking the divergence in Equation (
17), while using Equation (
7), we conclude
and inserting it in Equation (
18) leads to
Combining Equations (
19) and (
20), it infers
that is,
and summing the above equation over a local orthonormal frame
gives
Now, using Equation (
10), we have
and, as
, we conclude that
is a constant. We note that the almost Ricci soliton
is nontrivial; therefore, the potential function
is nonconstant. From Equation (
20), we have
We define a function
, which is a nonconstant function (as
is), and we have
; therefore,
. Consequently, Equation (
21) takes the form
Taking the trace in the above equation yields
, that is,
, which, on integration by parts, leads to
and, as
is nonconstant, through the above equation, we conclude
. This confirms that Equation (
22) is Obata’s differential equation (cf. [
22,
23,
24]) and, therefore,
is isometric to
.
Conversely, suppose that
is isometric to
. Then, through Equation (
3), it follows that
is a nontrivial almost Ricci soliton, where the potential function is
,
f is an eigenfunction,
, and the soliton vector field is
. We have
and
. Moreover,
and we have
. Thus, we obtain
and using Equation (
23), we find
Also, as
is a constant, we have
and using
, we obtain
which, in view of Equation (
23), yields
Thus, Equations (
24) and (
25) confirm that one of the given conditions is satisfied. Also, the almost Ricci soliton
has positive Ricci curvature; therefore, the second of the given conditions is also satisfied. Hence, the converse implication holds. ☐
Next, we are interested in finding conditions under which an almost Ricci soliton
is trivial, that is, it is a Ricci soliton, and to achieve this, we will ask if the affinity tensor
of the soliton vector field
vanishes.We note that the affinity tensor of a vector field is a useful tool in studying the geometry of a Riemannian manifold (cf. [
25,
26]). We recall that the affinity tensor of a vector field
is given by (cf. [
25], p. 109)
which is equivalent to
We shall impose the condition that the affinity tensor of the soliton vector field vanishes and some additional inequality for the scalar curvature holds for a compact almost Ricci soliton to force it to not only be a trivial almost Ricci soliton but a trivial Ricci soliton. Indeed, we prove the following result.
Theorem 3. Let be an n-dimensional compact and connected almost Ricci soliton, , with scalar curvature τ, vanishing the affinity tensor of the soliton vector field ξ, satisfying the inequality Then, is a trivial Ricci soliton.
Proof. Suppose that the affinity tensor of the soliton vector field
vanishes. Then, using Equations (
6) and (
26), we conclude
For a local orthonormal frame
, we have
and thus, summing Equation (
27) and using Equation (
10), it gives
Now, combining Equations (
11) and (
28), we arrive at
and that, by virtue of the fact
, makes
a constant; therefore, at this stage,
is a Ricci soliton. Next, we proceed to show that, with the remaining condition
, the soliton vector field
is a Killing vector field. It is trivial to check that
and we have the following equality (cf. [
2,
12])
We use a local orthonormal frame
and compute
with the aid of Equation (
6) as follows:
which, in view of Equation (
10), the symmetry of
Q and the skew-symmetry of
F, yields
On integrating the above equation, we arrive at
Inserting from Equations (
29) and (
30) into the above equation, while using
, it leads to
that is,
The above equation and the condition in the hypothesis
forces the conclusion
Since
, the above equation implies that
is a constant. We have already proved that
is a constant, and feeding this information in Equation (
8), we have
which, on integration, gives
, and Equation (
31) in turn gives
that is,
. Hence,
is a trivial Ricci soliton. ☐
5. Conclusions
In
Section 4, we have considered the Hodge decomposition of the soliton vector field
of an
n-dimensional compact almost Ricci soliton
given by Equation (
32), and, in Theorem 4, we have utilized the Hodge potential field
to find a characterization of the sphere
. A natural component of the Hodge decomposition is the Hodge function
h, which satisfies the Poisson Equation (
34), namely
As we have seen that the Hodge potential field is used to find a characterization of the sphere
, one would be interested in seeing the role of the Hodge function
h. It is known that the Poisson equation plays a crucial role in constraining the geometry of
(cf. [
10]); in particular, it is known that, on a compact almost Ricci soliton
, the solution of the above Poisson equation is unique up to a constant, that is, two solutions differ by a constant. Natural functions defined on
other than
h are the potential function
and the scalar curvature
. It will be worth seeking the following questions:
(i) Assuming, for a positive constant
c, that
is a solution of the Poisson Equation (
49), is a compact nontrivial almost Ricci soliton
isometric to the sphere
?
(ii) Assuming that the scalar curvature
is a solution of the Poisson Equation (
49), then, in this case, the compact nontrivial almost Ricci soliton
will not be isometric to the sphere
, as this will require that
is a constant whereas
is not a constant. So, as a point of interest, what will be the impact of the scalar curvature
being a solution of the Poisson Equation (
49) on the geometry of almost Ricci soliton
?
These two questions are worth exploring, and the answers to them will add new insights to the geometry of the almost Ricci soliton.