1. Prologue
Given an
n-dimensional Riemannian manifold
, in general there are two methods for addressing the geometry of
. The first is the intrinsic geometry of
, and the other is the extrinsic geometry of
. In the study of the intrinsic geometry of
, among others, some tools are distance functions, geodesics, and Jacobi fields on
, and these basic tools yield global results on the geometry of
such as the Theorem of Hadamard, the Hopf–Rinow theorem, the Bonnet–Myers theorem, and the Morse Index theorem (cf. [
1,
2,
3,
4]).
An important aspect of the intrinsic geometry of
deals with the existence of certain vector fields on
, such as Killing vector fields, conformal vector fields, potential fields of a Ricci soliton, and almost Ricci solitons and these vector fields influence the geometry as well as the topology of
. These are not only rich due to their elegance but also are influential in physics and general relativity (cf. [
1,
4,
5,
6,
7,
8]). Also, an equally important component of the intrinsic geometry deals with certain partial differential equations such as the Fischer–Marsden equation and Hamilton’s Ricci flow. This component is highly influential, as exhibited by its use in resolving the century-old Poincare conjecture [
9,
10].
Recall that classical differential geometry originated with the study of curves and surfaces in the Euclidean space
, which took a very magnanimous shape after the contributions of Kuiper and Nash (cf. [
4,
7,
11,
12]), which showed that an
n-dimensional Riemannian manifold
can be isometrically embedded in a Euclidean space
for a sufficiently large
. This paved the way for studying extrinsic geometry of
under the title submanifold geometry (cf. [
8,
13,
14,
15,
16,
17,
18,
19,
20,
21,
22,
23,
24]).
The extrinsic geometry under submanifold geometry is vast and it encompasses several global results such as the Chern–Lashof theorem using the notion of total absolute curvature (cf. [
21,
22,
25,
26]) to local results on submanifold geometry (cf. [
8,
13,
14,
15,
16,
17,
18,
19,
23,
24]). The vastness of this subject is also due to the fact that it includes the minimal submanifolds of a unit sphere, which has several open problems (cf. [
27,
28]). An important aspect of extrinsic geometry is the question of analyzing the obstructions to an embedding of the Riemannian manifold
into a Euclidean space
, and there are interesting results such as that of Tompkins [
29], who proved that a flat closed
cannot be isometrically embedded in the Euclidean space
. This finding was further followed by several important contributions that can be found in (cf. [
30,
31]).
One of the most important structures on a Riemannian manifold
is the Ricci soliton; it is given by a smooth vector field
called the potential field and a constant
satisfying:
where
is the Lie derivative with respect to
, and
is the Ricci tensor on
M (cf. [
5,
6,
9]). A Ricci soliton is denoted by
. A Ricci soliton
is a trivial Ricci soliton if the potential field
is a Killing field, that is,
, and we observe that a trivial Ricci soliton
is an Einstein manifold. A Ricci soliton
is a stable solution of the following heat equation known as Hamilton’s Ricci flow:
where the evolving metric
satisfies
, the scaling function
satisfies
, and the one-parameter diffeomorphism group
induces the potential field
It is a well-known fact about the heat equation that it distributes the temperature potential evenly; using this as a clue, Hamilton used Ricci flow to obtain an even distribution of curvature on the Riemannian manifold through a Ricci soliton. It is for this reason that the geometry of Ricci solitons has been subject of immense interest (cf. [
5,
6,
9]). An important aspect of Ricci soliton structure on a Riemannian manifold
is that we see a union of geometry and global analysis through it. Through this union, Perelman conceived of the idea of settling the famous century-old Poincare conjecture (cf. [
9]). We also observe that a Ricci soliton
becomes an Einstein manifold if the potential field
is a Killing vector field; therefore, a Ricci soliton
can be considered to be a generalization of an Einstein manifold. The potential field
and the constant
of a Ricci soliton
is induced through the Ricci flow as a stable solution. A natural question follows: Could there be another way to obtain the vector field
and the constant
on a compact Riemannian manifold
such that it becomes a Ricci soliton
? This article is devoted to answering this question.
Given an
n-dimensional smooth Riemannian manifold
, to measure the appetite of
to acquire the structure of a Ricci soliton
, we require that the potential field
and the constant
satisfy Equation (
1). In this article, we wish to approach this question through Kuiper–Nash isometric embedding
, where
is the Euclidean space for sufficiently large
, and
is the Euclidean metric. We use this embedding to bring a smooth vector field
on
that will assume the role of the potential field for the prospective Ricci soliton
. There are several ways to achieve this vector field. However, our choice is to pick up a constant unit vector field
on the Euclidean space
, which gives us the following decomposition of
where
is tangential to
and
is normal to
. This vector field
is called the Kuiper–Nash vector on
and
is the Kuiper–Nash normal on
. Note that the choice of the pair
,
is not unique, and it changes with the choice of the constant unit vector
on the Euclidean space
.
In this article, we explore the condition under which on an n-dimensional Riemannian manifold and the Kuiper–Nash vector together with a constant makes a Ricci soliton. Indeed, we prove that an n-dimensional compact Riemannian manifold with a scalar curvature and a Kuiper–Nash vector —if the integral of the function has a suitable lower bound containing the constant —is necessarily a Ricci soliton (cf. Theorem 1). We call this Ricci soliton a Kuiper–Nash Ricci soliton and study its properties (cf. Proposition 1). Recall that a trivial Ricci soliton is an Einstein manifold, and it is for this reason that Ricci solitons are considered to be a generalization of an Einstein manifold. Moreover, finding conditions under which a Ricci soliton is trivial is an important challenge in the geometry of Ricci solitons. In this article, we find two results that give conditions under which a Kuiper–Nash Ricci soliton is trivial (cf. Theorems 2 and 3).
Example: Consider the Euclidean spaces
and
, where
are Euclidean metrics. Then, we have the isometric embedding
, given by
It is a totally geodesic embedding with unit normal
, where
are Euclidean coordinates on
. For a constant unit vector
, we express
as
where
and
. Differentiating the above equation with respect to
and equating tangential and normal components, we confirm
where ∇ is the Riemannian connection on the Euclidean space
. As a result, we obtain
and consequently, on the Euclidean space, we have
where
, as the Euclidean space
, is Ricci flat. Thus,
is a Kuiper–Nash Ricci soliton with a Kuiper–Nash vector
.
2. Preliminaries
On an
n-dimensional Riemannian manifold
, let ∇ be the Riemannian connection. Then, the curvature tensor of
is given by
where
is the space of smooth vector fields on
. Contracting the curvature tensor field gives a Ricci tensor
of
and is a symmetric tensor
where
is a local orthonormal frame on
. The Ricci operator
Q of
is a symmetry operator
defined by
and the scalar curvature
of
is given by
The following formula is well known (cf. [
13], trace in Equation (3.1) p. 58)
where
is the gradient of the scalar curvature
and
For an
n-dimensional Riemannian manifold
, we have the Kuiper–Nash isometric embedding
, where
is the Euclidean space for sufficiently large
and
is the Euclidean metric. We denote the Euclidean connection on
with
, denote the normal bundle of this isometric
with
, and denote the space of smooth sections of the normal bundle
with
. Then, we have the following fundamental equations for the isometric embedding
where
h is the second fundamental form,
is the shape operator with respect to the normal vector field
N, and they are related by
Also, we have the following fundamental equations of the isometric embedding
, the curvature tensor of
has the expression
and the Ricci tensor of
has the expression
where
H is the mean curvature vector field defined by
for a local orthonormal frame
.
Given an
n-dimensional smooth Riemannian manifold
, we have the Kuiper–Nash isometric embedding
(cf. [
12,
23]). Fixing a constant unit vector
, we are interested in the tangential and normal parts of
as described in the following:
Definition 1. Given the Kuiper–Nash isometric embedding and a fixed constant unit vector on the Euclidean space expressed as , the tangential vector field ξ is called a Kuiper–Nash vector and the normal component Γ is called the Kuiper–Nash normal.
Definition 2. If ξ is the Kuiper–Nash vector and Γ
is the Kuiper–Nash normal on a Riemannian manifold with respect to the Kuiper–Nash isometric embedding and constant unit vector on the Euclidean space , then the function defined by is called a Kuiper–Nash function on , and the operator defined byis called the Kuiper–Nash operator of the Riemannian manifold . Differentiating the expression
with respect to
and using Equations (8) and (9) while equating like parts, we arrive at
Lemma 1. On an n-dimensional Riemannian manifold with a Kuiper–Nash vector ξ, a Kuiper–Nash function φ, and a Kuiper–Nash operator K, the following equations hold:
(i)
(ii) ,
where is a local orthonormal frame, is the trace of K and is the gradient of the Kuiper–Nash function φ.
Proof. Note that using the definition
, we find
which proves (i). Now, differentiating Equation (
14), we obtain
Consequently, using Equation (
3), we conclude
and using a local orthonormal frame
and the symmetry of the Kuiper–Nash operator
K in the above equation, we arrive at
Note that
Using the fact that the Kuiper–Nash operator
K is symmetric and
, where
are skew-symmetric connection forms, we conclude
and inserting the above equation into the previous equation gives
Combining the above equation with Equation (
15) yields
which proves (ii). □
Recall that the action of the Laplace operator on a smooth function
on a Riemannian manifold
is given by
where
is the gradient of
f and the divergence is defined by
On a compact Riemannian manifold
, the Stokes’s Theorem implies
where
is the volume element of
.
Lemma 2. On an n-dimensional compact Riemannian manifold with a Kuiper–Nash vector ξ and a Kuiper–Nash function φ, the following hold:
(i)
(ii) .
Proof. Using Equation (
14) and (i) in Lemma 1, we obtain
, which leads to (i) upon integration. Now, using
and
, which yields (ii) upon integration. □
Note that if
F is a
tensor field on an
n-dimensional Riemannian manifold
, then we define
Also, for a (0,2)-type symmetric tensor
, we have
3. Transforming a Riemannian Manifold to a Ricci Soliton
Let be an n-dimensional Riemannian manifold with a Kuiper–Nash vector field , a Kuiper–Nash function , and a Kuiper Nash operator K. In this section, we wish to find conditions under which a compact becomes a Ricci soliton and study the properties of this inherited Ricci soliton structure on .
Now, we prove the main result of this section:
Theorem 1. If an n-dimensional compact and connected Riemannian manifold with a scalar curvature τ, a Ricci operator Q, a Kuiper–Nash vector ξ, a Kuiper–Nash function φ, and a Kuiper–Nash operator K satisfiesfor a constant λ, then is a Ricci soliton. Proof. Using symmetry of the Kuiper–Nash operator
K and Equation (
14), we compute
Thus, for a constant
, we have
and treating the above equation with a local orthonormal frame
, using Equation (
19) and (i) in Lemma 1, we have
where we have used
Next, we need to compute the divergence of the vector
, for which we use a local orthonormal frame
and Equations (7) and (14) and arrive at the following
where we used the symmetry of
Q and Equation (
7) to obtain
Inserting the above equation into Equation (
21), we confirm that
Integrating the above equation while using Lemma 2, we conclude
Now, using the condition in the statement, we obtain
which proves
that is,
making
a Ricci soliton. □
Definition 3. The Ricci soliton of Theorem 1 is called the Kuiper–Nash Ricci soliton.
Proposition 1. An n-dimensional compact Kuiper–Nash Ricci soliton satisfies Proof. If
is an
n-dimensional compact Kuiper–Nash Ricci soliton, then Equation (
22) takes the form
Combining Equations (20) and (23), we obtain
which confirms
that is,
Inserting the above equation in Equation (
24), we obtain
which proves the result. □
In the rest of this section, we find the conditions under which the Kuiper–Nash Ricci soliton is a trivial Ricci soliton. First, we prove the following:
Theorem 2. The scalar curvature τ of an n-dimensional compact Kuiper–Nash Ricci soliton satisfiesif and only if is a trivial Ricci soliton. Proof. Suppose that the scalar curvature
of an
n-dimensional compact Kuiper–Nash Ricci soliton
satisfies
Using Equations (14) and (25), we have
We use Equations (7) and (27) and a local orthonormal frame
to compute the divergence of
and find
that is,
Integrating above equation, we arrive at
Combining it with Proposition 1, we obtain
Using the condition in the statement, we conclude
that is,
It shows that the scalar curvature
is a constant and using Equation (
28) in Proposition 1, we conclude
Using the Cauchy–Schwartz inequality
in Equation (
29), we confirm
It being an equality in a Cauchy–Schwartz inequality, it holds if and only if
Combining Equations (28) and (30) with Equation (
27), we confirm that
that is,
is trivial. The converse is trivial, for if
is a trivial Ricci soliton, then
, which implies that
is a constant. Hence, the condition in the statement holds. □
Note that in the statement of Theorem 2, we used the inequality
which ultimately turns out to be an equality. Therefore, it should be noted that a strict inequality cannot occur. Since our goal is to reach a trivial Ricci soliton, which requires
with constant
, that is,
, the considered inequality is justified.
Finally, we prove the following characterization of a compact trivial Ricci soliton using a Kuiper–Nash Ricci soliton.
Theorem 3. An n-dimensional compact Kuiper–Nash Ricci soliton , , with a scalar curvature τ satisfiesif and only if is a trivial Ricci soliton. Proof. Suppose
is an
n-dimensional compact Kuiper–Nash Ricci soliton with a scalar curvature
that satisfies the inequality
Then, using Equations (20) and (23), we have
which in view of Equation (
14) yields
Using a local orthonormal frame
with the above equation, we compute
Also, using Equation (
23), we compute
Note that the trace of Equation (
32) gives
, that is
Also, the Equation (
32) implies
which in view of Equation (
7) gives
Using Equations (36) and (37) in (ii) of Lemma 1, we obtain
that is,
Now, we use Equation (
32) in computing the divergence of the vector field
and arrive at
Using Equation (
7) in the above equation, we conclude
Integrating the above equation while using Equations (18) and (38), we conclude
Next, we recall the following integral formula (cf. [
32])
Inserting Equations (34) and (35) into the above equation yields
Substituting Equation (
39) into the above equation gives
that is,
The above equation is
which in view of the inequality (31), confirms
However, due to the Chauchy–Schwartz inequality
, the above inequality gives
and finally, we obtain
that is,
Summing the above equation over a local orthonormal frame
, we obtain
Combining the above equation with Equation (
7),
and as
, we obtain
, that is,
is a constant. Now, using Equation (
33), we have
, which on integration and the fact that
is a constant gives
. Thus, Equation (
40) yields
and in turn it proves
that is,
is a trivial Ricci soliton. The converse is trivial, as if
is a trivial Ricci soliton, then
; therefore,
, and consequently, the given condition in the statement holds. □