1. Introduction
Let , and consider an n-dimensional Riemannian manifold , which we assume to be connected throughout this paper. Denote by the set of all smooth vector fields defined on M.
In the present paper, we adopt the curvature sign convention following [
1], which is the reverse of the one appearing in [
2]. Specifically, the Riemann curvature tensor is defined as the
-type tensor field
for all vector fields
.
At any point
, let
be a local orthonormal frame of
. The Ricci tensor
and the scalar curvature
S are then defined, respectively, by
for all
, and
where
denotes the trace operator.
Consider a smooth manifold
M endowed with a one-parameter family of Riemannian metrics
. This family is called a solution of the Ricci flow whenever
where
is the Ricci tensor of the metric
.
A solution
of the Ricci flow is said to be self-similar if there exists a positive function
and a one-parameter family of diffeomorphisms
on
M such that
satisfies (
1). Here,
denotes the pullback of
under
.
For a given Riemannian manifold
, any self-similar solution with
is associated with a vector field
X and a constant
satisfying
where
is the Ricci tensor of the metric
g and
stands for the Lie derivative of
g in the direction of
X.
Conversely, if
X and
solve (
2), then
X generates a one-parameter family of diffeomorphisms
such that
is a self-similar solution to the Ricci flow (
1). For details, see [
3,
4].
A quadruple
satisfying (
2) is called a
Ricci soliton. Depending on the sign of
, the soliton is called shrinking
, steady
, or expanding
. If
X is a Killing vector field (so that
), then the soliton is said to be trivial, and (
2) reduces to
which means that
is Einstein.
When the vector field
X is the gradient of a smooth function
f, denoted by
, it is defined through the relation
for every
.
Under this assumption, Equation (
2) can be rewritten as
with
representing the Hessian of
f, i.e., the
-tensor defined by the formula
A quadruple
satisfying Equation (
3) is called a gradient Ricci soliton.
In the special case where the potential function
f is constant, the soliton is said to be trivial. Then, the metric
g is Einstein, that is,
For any vector field
, the divergence of
X is defined as
where
is a local orthonormal frame on
M.
For a smooth function
f on
M, the Laplacian is defined by
Using (
4) and (
6), we obtain
Recall that for a
-tensor field
T, its covariant derivative is defined as
for all
.
The divergence of
T is the vector field
and if
is chosen to be parallel at a point, this expression simplifies to
The squared norm of a
-tensor field
T on
is defined by
where
is a local orthonormal frame on
M, and
.
In the case where
T is self-adjoint (i.e., symmetric), this expression can equivalently be written as
with
denoting the eigenvalues of
T.
It is well known that every compact Ricci soliton, as well as every noncompact shrinking soliton, must be of a gradient type (see [
5,
6]). Over the past years, gradient Ricci solitons have been intensively studied due to their central role in the Ricci flow, their rigidity properties, and their appearance as models for singularity formation. For further accounts, we refer to [
7,
8,
9,
10,
11,
12,
13,
14].
The outline of the paper is as follows. In
Section 2, we recall the basic notions and formulas that will be needed in the sequel.
Section 3 is concerned with non-steady gradient Ricci solitons satisfying the identity
, where
S denotes the scalar curvature,
the soliton constant, and
f the potential function. We prove that such a soliton is trivial exactly when
Further consequences are obtained in the situation where S satisfies a Poisson-type equation.
In
Section 5 we turn to the study of compact solitons. We show that compact expanding or steady Ricci solitons are always trivial, whereas a compact normalized shrinking gradient Ricci soliton is trivial precisely when the average value of its potential function is at most
. Finally, we investigate compact Ricci solitons that can be realized as hypersurfaces in Euclidean space, proving in particular that such solitons must be shrinking, and that
and the mean curvature of
M satisfy an inequality, with equality if and only if
M is a sphere.
2. Preliminaries
Consider a Ricci soliton
. Taking the trace of Equation (
2) yields
where
S denotes the scalar curvature of the metric
g. In the special case of a gradient Ricci soliton
, this reduces to
with
the Laplacian of the potential
f.
A well-known relation derived from the contracted second Bianchi identity (see [
14]) states that
Using Equations (
2) and (
10) in the case of a gradient soliton
, one obtains
for some constant
c (cf. [
15]).
When the soliton is non-steady, the potential
f can be shifted to eliminate
c, and Equation (
11) simplifies to
In this case, the soliton is said to be normalized.
The squared norm of the Ricci tensor, equivalently of the associated Ricci operator, is defined according to (
7) and can be written as
where
denote the eigenvalues of the Ricci operator.
In a similar manner, the squared norm of the Hessian of a smooth function
f, written as
, is defined as the squared norm of the corresponding self-adjoint endomorphism
(a
-tensor) associated with the
-tensor
through the relation
for all
.
By Cauchy–Schwarz, one has
The following identities from [
14] will be useful later:
and
The latter is simply Bochner’s formula adapted to the setting of gradient Ricci solitons.
5. Compact Shrinking Gradient Ricci Solitons
We begin with a basic observation.
Proposition 1. A compact Ricci soliton is trivial precisely when its scalar curvature is constant.
Proof. Recall that every compact Ricci soliton is automatically a gradient Ricci soliton.
Let be such a soliton with constant scalar curvature S.
Integrating Equation (
9) over
M, and observing that
we obtain
Here, denotes the Riemannian volume element on M.
Since both
and the scalar curvature
S are constants, Equation (
24) simplifies to
or equivalently,
where
stands for the total volume of
M.
Consequently, we deduce that
. Substituting this back into Equation (
9) gives
. Because
M is compact, it follows that
f must be constant.
The converse direction is immediate according to (
9). □
It was shown in [
17] that for compact Ricci solitons of dimension
, the case of non-constant scalar curvature forces the soliton to be shrinking. In contrast, steady and expanding compact Ricci solitons always have constant scalar curvature. From this we deduce the following, for which we give a direct argument.
Proposition 2. Every compact steady or expanding Ricci soliton is trivial.
Proof. Let
be a compact gradient Ricci soliton. By combining (
9) and (
11), we obtain
Case 1. . At a maximum point
of
f, one has
and
It follows by Equation (
25) that
.
At a minimum point
, similarly
and
giving
Thus
, so (
25) reduces to
showing that
f is subharmonic. By the maximum principle,
f is constant.
At a maximum point
of
we have
and
which together with
imply
.
At a minimum point
, the analogous reasoning yields
Therefore, for an arbitrary point
we have
from which we deduce that
, that is
vanishes identically, showing that
f is constant and equal to
. Thus the soliton
is trivial. □
For a function
f on a compact Riemannian manifold
, its average value is defined by
In particular, if
is a compact gradient Ricci soliton, integrating (
9) gives
Moreover, when
is a non-steady compact gradient Ricci soliton, one can normalize by replacing
f with
so that (
11) is equivalent to (
12).
We shall always assume this normalization. From (
27) and (
12) it follows that
Thus, in the shrinking case () we have , while in the expanding case () we obtain . This motivates seeking conditions ensuring triviality in the shrinking case.
Theorem 5. Let be a compact normalized shrinking gradient Ricci soliton. Then it is trivial if and only if .
Proof. From (
9) and (
12), we obtain
Integrating over
M gives
which can be rewritten as
If
, then (
29) forces
, so
f is constant and the soliton is trivial.
Conversely, if the soliton is trivial, then
and (
9), (
12) yield
, hence
and
. □
Remark 2. The assumption in Theorem 5 is equivalent to the condition appearing in Theorem 1.1 of [16]. However, phrasing the condition directly in terms of is often more convenient.
Next we study compact Ricci solitons that can be realized as hypersurfaces in Euclidean space. The following result shows that such solitons must be shrinking.
Theorem 6. If a compact Ricci soliton admits an isometric immersion as a hypersurface in Euclidean space, then it must be shrinking. In other words, no compact steady or expanding Ricci soliton can be immersed in codimension one into .
Proof. Let be a compact gradient Ricci soliton immersed as a hypersurface in with .
By Proposition 2, the soliton is trivial, hence .
On the other hand, every compact hypersurface in Euclidean space contains a point where all principal curvatures are positive (see [
18]), which forces
.
This contradiction shows that cannot be nonpositive, so the soliton must be shrinking. □
Remark 3. Theorem 6 extends a result of [19], where it was proved that if is a compact Ricci soliton realized as a hypersurface in Euclidean space, with X the tangential part of the position vector field, then . We now establish a geometric estimate relating the soliton constant to the mean curvature of the hypersurface. We first recall a standard inequality.
Let
be a hypersurface of a Riemannian manifold
with unit normal vector field
, shape operator
A, and mean curvature
H. Then, the scalar curvature
of
is given by
where
denotes the scalar curvature of
, and
is the squared norm of the shape operator. Recall that the mean curvature
H of
is defined as
Theorem 7. Let be a compact gradient Ricci soliton of dimension n, isometrically immersed in . Then the soliton constant λ and the mean curvature H satisfywith equality if and only if M is isometric to a round sphere. Proof. By Theorem 6, the soliton is shrinking, hence .
Since the ambient space is Euclidean, Formula (
30) reduces to
The Cauchy–Schwarz inequality implies
, hence
From (
9) and (
31), one obtains
Integrating over
M yields
that is,
If equality holds, then integrating (
9) shows that
With (
30), this forces
pointwise, which means
M is totally umbilical. A complete totally umbilical hypersurface in Euclidean space with
is a round sphere, as claimed. □
Corollary 1. Let be a compact gradient Ricci soliton immersed as a hypersurface in . If its mean curvature H satisfiesthen M is isometric to a round sphere. 6. Conclusions and Future Directions
In this work, we obtained new characterization criteria for compact gradient Ricci solitons. We showed that a normalized shrinking gradient Ricci soliton is trivial if and only if the average of the potential function does not exceed
. For non-steady normalized gradient Ricci solitons, triviality is guaranteed precisely when the scalar curvature satisfies the relation
Moreover, in the context of isometric immersions, we proved that a compact gradient Ricci soliton realized as a hypersurface in Euclidean space necessarily corresponds to the shrinking case, with an explicit inequality linking to the mean curvature, attaining equality uniquely on spheres.
These results emphasize the rigidity of compact gradient Ricci solitons and further illuminate the deep interaction between curvature conditions and potential functions.
Several directions remain open for further investigation. It would be natural to explore whether analogous characterizations hold for noncompact settings, especially for complete shrinking or steady gradient Ricci solitons.
Another avenue is to examine the behavior under different ambient spaces: for instance, characterizing Ricci solitons realized as hypersurfaces in space forms or more general ambient manifolds such as Einstein spaces.
It is also worthwhile to investigate gradient Ricci solitons arising on Lorentzian manifolds, particularly when considered as timelike hypersurfaces within a Lorentzian ambient space.
Finally, it would be interesting to analyze stability aspects, such as whether the inequalities obtained persist under perturbations of the metric or immersion, and how these results connect to the broader study of geometric flows and their singularity models.