1. Introduction
Over the past two decades, geometric flows have established themselves as powerful and versatile tools in the analysis of geometric structures on Riemannian manifolds. Among these, flows such as the Ricci flow and the mean curvature flow have proven particularly effective in deforming metrics or submanifolds in a manner that reveals deep structural and topological information. A central theme in the study of geometric flows is the development and classification of singularities, which often encode significant geometric and analytic data. In this context, a distinguished class of special solutions—those in which the metric evolves purely under the action of diffeomorphisms and possible scaling—plays a fundamental role. These are referred to as soliton solutions (e.g., Ricci solitons in the Ricci flow setting), and they represent self-similar solutions to the flow equations. Solitons frequently arise as singularity models in blow-up analyses, capturing the local geometry near singular points and serving as canonical forms toward which more general solutions may converge under appropriate rescaling. Their study is therefore essential for understanding both the local and global behavior of geometric flows, particularly in the vicinity of singularities and in the long-time asymptotic regime.
The Yamabe flow, introduced by Hamilton [
1] alongside the Ricci flow, admits distinguished solutions known as solitons—specifically, Ricci solitons for the Ricci flow and Yamabe solitons for the Yamabe flow. In two-dimensional settings
, these solitons are equivalent, meaning that a Ricci soliton also satisfies the conditions of a Yamabe soliton. However, in higher dimensions
, this equivalence no longer holds, and the respective solitons exhibit distinct behaviors and properties.
Recent developments in geometric analysis have highlighted the increasing importance of curvature-driven flows, with the Ricci flow and Yamabe flow occupying central positions in current research. These evolution equations have proven particularly valuable for understanding the dynamic behavior of geometric structures and their convergence properties under curvature constraints. In 2019, Güler and Crasmareanu [
2] introduced a new geometric flow, referred to as the Ricci–Yamabe map, which is defined as a scalar linear combination of the Ricci and Yamabe flows. This flow, also known as the
-type Ricci–Yamabe flow, governs the evolution of metrics on the Riemannian manifold
introduced in [
2] and is given by the equation
where
S is the Ricci tensor,
r denotes the scalar curvature, and
, respectively. The sign and magnitude of the parameters
and
allow this flow to be interpreted as a Riemannian, singular Riemannian, or semi-Riemannian flow. Such flexibility is particularly valuable in modeling diverse geometric and physical phenomena, including those appearing in general relativity and other relativistic theories. In this framework, RYSs naturally emerge as self-similar solutions—essentially fixed points under the flow modulo diffeomorphisms and scalings.
An additional motivation for studying RYS lies in the fact that, although Ricci and Yamabe solitons coincide in two dimensions, they differ substantially in higher dimensions. This highlights the significance of the combined Ricci–Yamabe structure in exploring more general geometric behavior in higher-dimensional settings.
An RYS on a Riemannian manifold
is a quintuple
satisfying the following equation
where
being the Lie derivative of the metric
g along the vector field
,
S is the Ricci tensor, and
r denotes the scalar curvature. The parameters
are real constants. Moreover, an RYS is said to be expanding, shrinking, or steady depending on whether the constant
is positive, negative, or zero, respectively [
3,
4]. If the parameters
and
are allowed to vary smoothly over the manifold, that is, they are smooth real-valued functions rather than constants, then Equation (
1) defines what is referred to as an almost RYS.
In 2020, Siddiqi et al. [
5] introduced a further generalization of the RYS, referred to as the
-RYS. This structure is defined on a Riemannian manifold
by a data set
satisfying the equation
where
is a smooth 1-form and
is a constant.
In 2021, Sarkar et al. [
6] introduced and studied a new generalization of the RYS, termed the conformal RYS. This structure extends the classical RYS by incorporating an additional conformal term and is defined by the equation
where
l is referred to as a non-dynamical scalar field, meaning it is a time-dependent function not governed by the geometric flow itself. The parameter
is the dimension of the underlying manifold. Similar to other geometric solitons, the conformal RYS is classified based on the sign of the term
: if this expression vanishes, then the soliton is said to be steady, if it is negative, then the soliton is shrinking, and if it is positive, then the soliton is expanding. Moreover, if the vector field
is the gradient of a smooth function
, that is,
(where
D is the gradient opeartor), then Equation (
3) describes a conformal gradient RYS. In this case, the structure captures a natural extension of gradient Ricci solitons, adapted to the conformal deformation of the metric.
The concept of a conformal
-Ricci–Yamabe soliton (conformal
-RYS) was formulated in [
6]. It is characterized by the condition
where
S and
r, respectively, denote the Ricci tensor and the scalar curvature of the manifold,
l is a non-dynamical scalar field (possibly time-dependent), and
m represents the dimension of the manifold
. The quantities
and
are real constants, while
stands for the Lie derivative of the metric tensor
g along the vector field
. Depending on the value of the soliton constant, the conformal
-RYS is described as steady when the soliton constant is zero, shrinking when it is negative, and expanding when it is positive.
In 2020, Dey and Roy [
7] introduced the notion of a *-
-Ricci soliton, a generalization of the classical
-Ricci soliton, defined by
where
denotes the *-Ricci tensor. The concept of the *-Ricci tensor was first proposed in 1959 by Tachibana [
8] in the study of almost Hermitian manifolds and was subsequently extended in 2002 by Hamada [
9] to real hypersurfaces of non-flat complex space forms by
for any vector fields
on
, where
is the
*-Ricci operator, and
R is the Riemannian curvature tensor.
Building upon these ideas, the authors of [
6] defined the *-conformal
-Ricci soliton by the relation
Later, in 2021, Zhang et al. [
10] extended this framework to introduce the *-conformal
-Ricci–Yamabe soliton (*-conformal
-RYS) on a Riemannian manifold
, defined by the equation
where
denotes the *-scalar curvature. If the vector field
is chosen to be the gradient of a smooth function
f, i.e.,
, then Equation (
7) reduces to its gradient form, known as a
gradient *-η-RYS, expressed as
where
represents the Hessian of the smooth potential function
f.
The study of connections on differentiable manifolds has undergone substantial development over time, shaped by several foundational contributions. In 1924, Friedman and Schouten [
11] introduced the concept of a semi-symmetric linear connection, marking a significant milestone in the evolution of differential geometry. Subsequently, Hayden [
12] expanded this framework by proposing a metric connection with non-zero torsion tensor on Riemannian manifolds, thereby extending the classical Levi-Civita paradigm. Further advancement came through the work of [
13], who introduced the notion of a semi-symmetric non-metric connection in Riemannian geometry, enriching the diversity of connection types studied within the field.
These connections represent only a few key milestones in the broader study of manifold structures. The generalized symmetric metric connection has been extensively investigated by researchers such as [
14,
15,
16,
17], leading to deeper insights into its geometric properties. Motivated by these studies, in the present work, we introduce a new type of connection, called the GSNMC on an
. This connection is uniquely characterized as a combination of both semi-symmetric non-metric and quarter-symmetric non-metric connections, integrating aspects of both structures.
, along with different types of solitions and curvature tensors, has been studied by geometers like [
18,
19,
20,
21,
22,
23,
24] and many more authors.
This research article is organized as follows:
Section 1 introduces the topic and sets the stage for the study.
Section 2 presents the foundational concepts of an
. In
Section 3, we defined a new type of connection called the GSNMC on an
and proved the existence of such connection.
Section 4 discusses certain curvature properties of an
with regard to the new connection.
Section 5 discussed a *-conformal
-RYS with respect to the GSNMC on an
.
Section 6 presents several applications of a *-conformal
-RYS with respect to the GSNMC on an
, whereas harmonic aspects of a *-conformal
-RYS with respect to GSNMC on an
have been considered in
Section 7. Finally, this is followed by an example of a five-dimensional
-cosymplectic metric as a *-conformal
-RYS admitting the GSNMC.
2. Preliminaries
A smooth manifold
of dimension
m is called an almost contact metric manifold if it is endowed with an almost contact metric structure
, where
is a
-tensor field,
is a vector field,
is a 1-form, and
g is a Riemannian metric compatible with
, subject to the conditions [
25]
for all vector fields
, where
denotes the set of all smooth vector fields of
.
On such a manifold, the fundamental form
of
is defined as
for all vector fields
.
In 1967, Blair [
26] described a cosymplectic structure as a particular case of a quasi-Sasakian structure for which
. It is important to note that this concept differs from the cosymplectic manifolds previously introduced by Libermann [
27]. Later, an almost contact metric manifold
was termed almost cosymplectic [
28] whenever the conditions
and
are satisfied, where
d denotes the exterior differential operator. A basic example of such a manifold is given by
, with
N being an almost Kähler manifold and
being the real line [
29]. Furthermore, an almost contact manifold
is called normal if the corresponding Nijenhuis torsion
vanishes for any vector fields
and
. A normal almost cosymplectic manifold is a cosymplectic manifold.
An almost contact metric manifold is said to be almost -Kenmotsu if and , with being a non-zero real constant.
Kim and Pak [
30] introduced a unified framework that brings together the structures of
-Kenmotsu and almost cosymplectic manifolds into a broader class, referred to as almost
, where
is a scalar parameter. This generalized structure is defined on an almost contact metric manifold
by the conditions
for any real number
, where
is the fundamental 2-form. An almost
-cosymplectic manifold is said to be
-cosymplectic if it is normal, i.e., if the associated Nijenhuis torsion vanishes. This class naturally interpolates between known geometric structures: when
, the manifold reduces to a cosymplectic manifold; when
, it corresponds to an
-Kenmotsu manifold. Thus, almost
provide a unified setting that generalizes both cosymplectic and
-Kenmotsu geometries.
In an
-cosymplectic manifold, we have [
31,
32]
where
∇ is the Levi-Civita connection associated with
g.
If we let
be an
m-dimensional
-cosymplectic manifold (in short,
), then the following relations also hold [
31]:
for all vector fields
, where
S is the Ricci tensor,
Q is the Ricci operator defined by
, and
R is the Riemannian curvature tensor of
, respectively.
In paper [
31], Lemma 2.2, the authors give the proof of the *-Ricci tensor on an
m-dimensional
as given by
for any vector field
on
, where
is the *-Ricci tensor for type (0,2) on
, and
S is the Ricci tensor for type (0,2) on
.
We take
in (
24), where
is the orthonormal basis of
, for
. Therefore, we obtain
where
is the *-scalar curvature and
r is the scalar curvature.
Definition 1. If we let be an m-dimensional , then it is classified as a generalized η-Einstein manifold if its Ricci tensor S satisfies the relation where , and are scalar functions defined on . If , then (26) becomes an η-Einstein manifold. Furthermore, if both and vanish, then Equation (26) reduces to that of an Einstein manifold. 7. Harmonic View of a *-Conformal -RYS on an Admitting the GSNMC
This section concerns a smooth function
, which is said to be harmonic if it satisfies
, where
denotes the Laplacian operator on the manifold
[
36]. Now, suppose that the vector field
is given by the gradient of such a harmonic function
f, i.e.,
. Under this assumption, and in light of Lemma 1, we can deduce the following results:
Lemma 2. If we let be an m-dimensional that admits a *-conformal η-RYS with respect to the GSNMC and suppose that the soliton vector field ξ is of gradient type, i.e., , where f is the harmonic function on , then the nature of a *-conformal η-RYS is expanding, steady, or shrinking, on the basis of where , and , respectively.
Proof. If we let
in (
61), then we get
Thus, we obtained the result. □
Lemma 3. If we consider an m-dimensional α-cosymplectic manifold endowed with a GSNMC and suppose that admits a *-conformal η-Ricci–Yamabe soliton whose potential vector field ξ is given by , where f is a harmonic function on , then under these assumptions, the corresponding *-conformal η-Ricci soliton represents a shrinking soliton.
Proof. Upon substituting
into (
62), we get
Thus, the proof is fully established. □
Lemma 4. If we consider an m-dimensional α- manifold endowed with a GSNMC and suppose that admits a *-conformal η-Ricci–Yamabe soliton whose associated vector field ξ is given by , where f is a harmonic function on , then the nature of the corresponding *-conformal η-Yamabe soliton—whether expanding, steady, or shrinking—is determined by the following conditions: Proof. From (
63), the result can be easily obtained. □
Definition 2. A vector field Ψ is called a conformal Killing vector field if and only if it satisfies the following condition: where denotes the Lie derivative of the metric tensor g along Ψ, and β is a smooth scalar function on the manifold, referred to as the conformal scalar.
Furthermore, the nature of a conformal Killing vector field
is characterized by the behavior of the conformal scalar
in Equation (
65). Specifically,
- (i)
If is a non-constant, then is referred to as a proper conformal Killing vector field. This represents the most general case of conformal symmetry, where the metric is preserved up to a position-dependent scaling.
- (ii)
If is a constant, then is called a homothetic vector field. In this case, the metric is preserved up to a uniform scaling across the manifold.
- (iii)
When is a non-zero constant, the vector field is further classified as a proper homothetic vector field, distinguishing it from the trivial case.
- (iv)
If
, Equation (
65) reduces to
, in which case
is a Killing vector field, generating an isometry of the manifold.
These classifications are fundamental in differential geometry and play important roles in the study of symmetries of geometric structures, conservation laws in physics, and the behavior of solutions to geometric flow equations.
Theorem 4. If we let be an m-dimensional that admits a *-conformal η-RYS with respect to the GSNMC and if Ψ is a conformal Killing vector field, then Equation (70) holds. Proof. Let us consider that is an m-dimensional admitting a *-conformal -RYS. Here, .
From (
7), (
24) and (
25), we obtain
Now, if we let
be an
m-dimensional
that admits a *-conformal
-RYS
in regard to the GSNMC
, then according to (
66), we achieve
In view of (
39), (
40), (
51) and (
67), we obtain
Using (
65) in (
68), we get
By setting
and using (
9), () and () in (
69), we get
Therefore, Equation (
70) can be written as
where
Thus, the proof of the theorem is hereby completed. □
8. Example
Let
be a five-dimensional manifold, where
are regarded as the standard coordinates in
. Let
be a set of linearly independent global frame fields on the five-dimensional manifold
, defined as follows:
Let us suppose that
g is a Riemannian metric, defined by the following expression:
where
. We may define the 1-form
on
by
for any vector field
on
and introduce the
-type tensor field
as follows:
After exploiting the linearity of both
and
g, we obtain the following result:
for all
on
. Thus,
, and the structure
defines an almost contact metric structure on
.
If we let
∇ denote the Levi-Civita connection
associated with the Riemannian metric
g, then we obtain the non-zero Lie brackets as
The Koszul formula, which characterizes the Levi-Civita connection
∇ in terms of the metric
g, is expressed as
From the above formula, we can easily calculate
One readily observes that the manifold satisfies
Thus, the manifold
is an
. With the help of (
27) and the above results, the calculation for the non-zero values become
Based on the preceding relations, it is evident that
Consequently, the above equation confirms the validity of Equation (
28). Furthermore, it is readily verified that
, indicating that Equation (
29) holds and that
is indeed a non-metric connection. Hence, the linear connection
defined in (
27) qualifies as the GSNMC on
.
Now, the following values are the non-zero entries of the Riemannian curvature tensor
R under the
∇From the above computations, the components of the Ricci tensor
S with respect to the
, can be obtained as follows:
The scalar curvature
r with respect to
∇ can be calculated as
Furthermore, under the GSNMC
, the non-vanishing components of the Riemannian curvature tensor
yields
The distinct non-vanishing elements of the Ricci tensor
corresponding to
are as follows:
On an
, the scalar curvature with respect to
can be evaluated by using the above results as follows:
Now, plugging
in (
52), we get
By substituting
and
with
, and using (
72) and (
74) in (
50), we arrive at
Thus, the last equation verifies that the relation between the soliton constants
and
satisfy Equation (
57) for
. Accordingly, the provided example corresponds perfectly with the outcomes of our study.