1. Introduction
Semi-symmetric spaces are a broad and exciting class of Riemannian manifolds, and they have applications in various areas of mathematics, particularly in the study of homogeneous spaces and differential geometry. They serve as an essential class of examples for understanding the interplay between curvature, symmetry, and geometry on manifolds. Researchers in Riemannian and differential geometry have studied these spaces to understand better their geometric properties and applications in physics, such as in the study of Einstein’s field equations in general relativity [
1]. Nomizu introduced the notion of semi-symmetric manifolds. A Riemannian manifold
is called semi-symmetric if
for all vector fields
and
on
,
acts as a derivation on
[
2].
A Riemannian manifold
, which is not necessarily complete, is locally symmetric if its curvature tensor is parallel, i.e.,
. In other words,
is locally symmetric if and only if there exists a symmetric space
such that
is locally isometric to
. Nomizu proved that if
is a complete, connected semi-symmetric hypersurface of Euclidean space
, then
is locally symmetric. Then, Sekigawa and Tanno showed that the manifold is locally symmetric if the Riemannian curvature tensor provided some conditions related to the covariant derivatives for
[
3]. For the case of a compact Kaehler manifold, Ogawa proved that if it is semi-symmetric, it must be locally symmetric [
4]. In the case of contact structures, Tanno showed no proper semi-symmetric or Ricci semi-symmetric K-contact manifold [
5]. Moreover, Szabó gave a complete intrinsic classification of these spaces [
6].
It is well known that semi-symmetric manifolds include the set of locally symmetric manifolds as a proper subset. Semi-symmetric spaces are the natural generalization of locally symmetric spaces. Such a space is called semi-symmetric since is the same as the curvature tensor of a symmetric space at a point of . Namely, locally symmetric spaces are semi-symmetric, but the converse is generally untrue.
Semi-symmetric contact metric manifolds have been studied by numerous authors [
7,
8,
9]. In particular, Takahashi proved that the constant sectional curvature of a semi-symmetric Sasakian manifold is
. In addition, semi-symmetric contact manifolds satisfying (
-nullity condition for dimensions greater than 3 were investigated by Papantoniu. Moreover, if
is semi-symmetric and the tensor field
is
-parallel, then
is either smooth or has a constant curvature of
. Then, Perrone proved that a semi-symmetric contact Riemannian three-manifold is flat or has constant sectional curvature of
. On the other hand, Blair and Sharma proved that the constant curvature of a locally symmetric contact metric three-manifold is
or
[
10].
Later, Calvaruso and Perrone investigated semi-symmetric contact three-manifolds [
11]. Under some additional conditions, they obtained several classification results. Then, conformally flat semi-symmetric spaces were investigated by Calvaruso [
12]. The author obtained that a conformally flat semi-symmetric space
of dimension greater than 2 is either locally symmetric or irreducible and isometric to a semi-symmetric real cone. In [
13], if
is a locally symmetric contact metric manifold with dimensions 3 and 5, it is either Sasakian and has constant curvature of 1 or locally isometric to the unit tangent sphere bundle of Euclidean space.
Almost contact metric structure has a special subclass called almost cosymplectic manifold. It was first introduced to the literature by Goldberg and Yano [
14]. An almost contact metric manifold is said to be an almost cosymplectic manifold if
and
. Here,
is the exterior differential operator. An almost cosymplectic manifold with constant curvature is cosymplectic if and only if it is locally flat. A comprehensive study of almost cosymplectic manifolds has been undertaken by Olszak [
15,
16]. The author obtained some sufficient conditions and proved that no almost cosymplectic manifolds of non-vanishing constant curvature exist in dimensions greater than 5. In addition, Perrone classified simply connected homogeneous almost cosymplectic three-manifolds [
17]. The author showed that if an almost cosymplectic three-manifold is locally symmetric, then its structure is cosymplectic and it is locally a product of a one-dimensional manifold and a Kaehler surface of constant curvature
. After this study, the author classified connected homogeneous dimensional almost
-coKaehler structures [
18].
Kenmotsu manifolds were first introduced by Kenmotsu [
19]. A Kenmotsu manifold can be defined as a normal almost contact metric manifold. Kenmotsu showed that a locally symmetric Kenmotsu manifold has constant curvature of
. Therefore, local symmetry is an essential restriction for Kenmotsu manifolds. The author obtained that if the Kenmotsu structure satisfies the semi-symmetric condition, it has constant negative curvature. Furthermore, if the Kenmotsu manifold
is conformally flat, then
is a space of constant negative curvature of −1 for
. A
-dimensional almost contact metric manifold is said to be an almost
-Kenmotsu manifold if
and
, where
is a non-zero real constant [
20]. The geometric properties and examples of these manifolds were studied [
16,
19,
20]. Remark that almost
-Kenmotsu structures are related to certain local conformal deformations of almost cosymplectic structures [
16,
21]. If we consider these two classes jointly, we introduce a new notion called an almost
-cosymplectic manifold for any real constant
, which is given by
and
[
22].
On the other hand, a systematic study of semi-symmetric almost contact metric manifolds still needs to be undertaken. In [
23], the authors studied certain classification results related to the nullity condition for an almost Kenmotsu manifold
with the characteristic vector field
belonging to the
-nullity distribution. They showed that if
is
-Riemannian semi-symmetric, then
is locally isometric to the Riemannian product of an
-dimensional manifold of constant sectional curvature of
and a flat
-dimensional manifold. Furthermore, if
is a
-Riemannian semi-symmetric almost Kenmotsu manifold such that
belongs to the null distribution, then
has constant sectional curvature of
. In [
24], Öztürk studied semi-symmetric conditions for
-Kenmotsu manifolds. In addition, many authors on these topics have studied almost Kenmotsu manifolds [
25,
26,
27,
28].
The paper is organized in the following way: In
Section 2, we recall the concept of almost
-cosymplectic manifolds. In
Section 3, we give some basic formulas on almost
-cosymplectic manifolds. In
Section 4, we obtain several results for three-dimensional almost
-cosymplectic manifolds.
Section 5 obtains the results of the semi-symmetric almost
α-cosymplectic three-manifolds. In
Section 6, we give illustrative examples of almost
-Kenmotsu manifolds. The last section of the paper is devoted to the discussion.
2. Preliminaries
Let
be a
-dimensional smooth manifold. Then,
is said to be an almost contact manifold if its structure group is reducible to
This corresponds to an almost contact structure defined by a triple
satisfying the following conditions
which yield
Here, the
is called the Reeb vector field or characteristic vector field. Then, we have a compatible Riemannian metric
on
defined by [
29]
for arbitrary vector fields on
. Such
is said to be an almost contact metric manifold [
30]. The fundamental two-form
of
is defined by
. Additionally, if
holds the condition
, then
is said to be a contact metric manifold. It is well known that Tanno classified the structures into three classes using their automorphism groups [
5]. Blair analyzed the contact metric structure, which also includes the Sasakian structure for class (1). Cosymplectic structures characterize the geometrical relations of class (2). The first simple example that comes to mind for class (2) is local products of a real line or a circle and a Kaehler manifold. Class (3) was extended by Kenmotsu, which is expressed locally by a warped product of an open interval and a Kaehler manifold [
31]. This type of manifold is called Kenmotsu and has normal structure. We have noted that every orientable surface admits a Kaehler metric. If we take a warped product metric on the product space
, then we have a cosymplectic or a Kenmotsu three-manifold, respectively. A cosymplectic or a Kenmotsu structure satisfies the normallity and CR-integrability [
32]. An almost complex structure
on
is defined by [
29].
Here,
is a vector field tangent to
,
is the standart coordinate of
, and
is a function on
. If (5) is integrable, then
is called normal. In addition, it is well known that
is normal if and only if
satisfies
where
is the Nijenhuis torsion tensor field of
. We recall that we have much broader classes without normality. Note that a normal almost
-cosymplectic manifold is said to be an
-cosymplectic manifold. An
-cosymplectic manifold is either cosymplectic (when
) or
-Kenmotsu (when
) [
22].
We denote by
the Levi Civita connection of
, by
the corresponding Riemannian curvature tensor for a Riemannian manifold
defined by
by
the Ricci tensor, and by
the Ricci operator given by
For an almost contact manifold, the (1,1)-tensor field
is defined by
where
denotes the Lie derivative in the direction of
[
30].
Lemma 1. Ref. [33]. Let be a -dimensional almost contact metric manifold. Then, is normal if and only if the tensor field identically vanishes. Throughout the paper, we shall denote by and the Lie algebra of all tangent vector fields on and the Levi Civita connection of Riemannian metric respectively.
4. Almost -Cosymplectic Three-Manifolds
Let
be an almost
-cosymplectic three-manifold. Let us consider the the open subsets
Then, the union set is an open dense subset of . There exists a local orthonormal basis of smooth eigenvectors of in a neighborhood of for any point . This basis is called the -basis of . Let on , where is a positive non-vanishing smooth function. Next, using (14) and (15), we have . Thus, we can state the following lemma:
Lemma 3. Let be an almost -cosymplectic three-manifold. Then, we have on : Here,
is a smooth function and
are defined by
and
respectively.
Proof. For any
using the definition of covariant derivation, it follows that
Moreover, we have
Here, if we set we obtain . In a similar way, we assume that and , then the other covariant derivatives can be obtained.
It is well known that Weyl conformal curvature tensor vanishes in dimension 3. That is to say, we have:
Replacing
, and
in (26), we find
Since
, we obtain
Then, using (28), it follows that
From (28) and (29), we have
Hence, the smooth functions
and
take the form
Thus, it completes the proof. □
Proposition 5. Let be a -dimensional locally symmetric almost -cosymplectic manifold. Then, we have [37]. Proposition 6. Let be an almost -cosymplectic three-manifold. On , we have:where is a (1,1)
-tensor field such that , and Proof. Taking the covariant derivative of
with respect to
, we have
Here, we remark that
. In view of (31) and (32), we deduce
where
. In addition, since
we obtain
on
. □
Proposition 7. Let be an almost -cosymplectic three-manifold. Then, we have:where
is the Jacobi operator. Proof. Following from (24), we have
To complete the proof, let us calculate their values according to the components of the basis. In fact, we have:
From (34), the proof is clear.
Since
, (26) turns out to be
Therefore, the last formula gives
On the other hand, we have
Moreover, we obtain
with respect to the
-basis. Taking into account (37) and (38), we obtain
Next, arranging the above equation, we have
Hence, if we set
then (40) turns into
Thus, we state the following result: □
Lemma 4. Let be an almost -cosymplectic three-manifold. Then, the Ricci operator satisfies the following: Here, the functions
and
are defined by
respectively [
37].
Proposition 8. Let be an almost -cosymplectic three-manifold. Then, the components of the Ricci operator with respect to the -basis are given as follows:
where
and
.
Proof. Taking
in (41), it yields that
Using
and
, the above equation becomes
Now, putting
in (41), it follows that
Then, the last equation reduces to
where
ve
. Analogously, putting
in (41), then we obtain
This completes the proof. Note that and . □
Proposition 9. Let be an almost -cosymplectic three-manifold. Then, considering (43), we have: Proof. From the first equation of (43), we have
where
and
. Then, the first equation clears from (47).
Similarly, from (25) and (43), we obtain
Taking into account (43) and (48), we compute
which gives the second equation of (46).
Finally, considering (43) with respect to
, we obtain
Therefore, if we proceed similarly, we complete the last part of the proof. □
Proposition 10. Let be an almost -cosymplectic three-manifold. Then, the components of with respect to the -basis are as follows: Proof. Let us consider (26) and (43). Putting
and
in (26), we have
Putting again
and
in (26), we obtain
Using (52) and (53), the first two equations can be seen. Usage of the same methodology (51) is clear. Here, we recall that
In addition, the equations given in (51) are all the possible non-zero components of the Riemannian curvature . They depend on the changes in the order of the vector fields. □
5. Semi-Symmetric Almost -Cosymplectic Three-Manifolds
In this section, we study semi-symmetric almost -cosymplectic three-manifolds. Then, we prove the following:
Theorem 1. Let be an almost -cosymplectic three-manifold. Then, is semi-symmetric if and only if Proof. According to the hypothesis,
is an almost
-cosymplectic three-manifold. We note that (1) is equivalent to
, for all
on
. In other words, we have
for all
on
.
Putting
, and
in (59), then we have
From (51) and (60), we obtain
where
and
. Therefore, (61) turns into
Hence, this ends the proof of (54) and (55).
Using a similar methodology, putting
,
,
ve
in (59) and (51), then we obtain
where
and so (56) and (57) satisfy (63). Finally, we take
and
in (59) and, taking account of (51), we deduce
Thus, the proof of (58) is clear. We also note that all the other possible choices of the vector fields in the -basis are given again (54)–(58). Therefore, if (54)–(58) is satisfied, then (59) is also satisfied, which means is semi-symmetric. □
Theorem 2. Let be a semi-symmetric almost cosymplectic three-manifold. If the structure is cosymplectic and the Ricci curvature is constant along the characteristic vector field , then is locally symmetric. Otherwise, is not locally symmetric if the structure is almost cosymplectic under the same condition.
Proof. Suppose that is a semi-symmetric almost cosymplectic three-manifold. Therefore, (54)–(58) satisfy . Now, we shall classify our arguments under the following two conditions:
Case 1. If
, then the structure is cosymplectic [
38,
39]. According to the hypothesis, because of (43), the Ricci curvature
constant along the characteristic vector field
means exactly
. Hence, from (31), we have
. In this case, whether the smooth function
is different from zero will be independent. Thus, the conclusion follows from Proposition 5.
Case 2. In this case, let us consider
and
. Note that (54)–(58) satisfy on
. Then, as it follows from (31), if
, we obtain
. The result can be seen in Proposition 5. To end the proof, we shall obtain that the case of
cannot take place. If so, we suppose that
and consider a point
at
, where
. Thus, there exists a neighbourhood
of a point
such that
on
. First, we multiply (54) by
and (55) by
. Then, we have:
and
Then, we subtract (65) from (66) and take into account (57) and (58) for expressing
and
, respectively. It follows that
where
Since
, (67) can be written as
It is noted that if the other equation holds, we proceed in the same manner, and since , the two equations cannot satisfy simultaneously. However, (69) shows that the function cannot vanish. In this case, we are unlikely to find a contradiction in our assumption.
Let us continue the calculation with the thought that our assumption is true. (56) holds since and . Namely, we have locally either or Let us suppose and we shall prove that is constant and . If we suppose the other case (), we proceed in the same way.
Differentiating (69) with respect to
, we have
, where
Then, again differentiating with respect to
, we obtain
To obtain whether
is constant or not, let us remember the well-known formula
for any
where
is an arbitrary orthonormal basis. Applying (46) and (71) to calculate
and
, then making use of (25), (46), (69), and (71), we observe that
is not necessarily constant and
does not have to vanish. In fact, using (25), we calculate
as follows:
Then, comparing with (51), we obtain
From (69), by a direct calculation, we deduce
Then, taking account of (57), (69), and (74), we obtain
and so we have
. Moreover, a result of Olszak verifies our proof by (74) [
38]. □
Theorem 3. Let be a semi-symmetric almost -Kenmotsu three-manifold with the Ricci curvature constant along the characteristic vector field . If the structure is normal, then is locally symmetric only when . Moreover, is locally symmetric if is given by a constant scalar curvature .
Proof. The geometry of almost -Kenmotsu manifolds differs in two cases, with the tensor field being zero or non-zero.
Case 1. Assume that
. Then, an almost
-Kenmotsu three-manifold is an
-Kenmotsu manifold. A result of Dileo is that if an almost
-Kenmotsu three-manifold has a constant curvature, then the structure is normal, and the constant curvature is
when it is locally symmetric [
25]. Furthermore, Öztürk showed that a semi-symmetric
-Kenmotsu manifold is not of constant curvature. From Corollary 4.3 in [
24], semi-symmetry implies local symmetry for
. In fact, using the hypothesis and (54)–(58), Theorem 1 is not verified except in the case α = 1. This completes the proof.
Case 2. Suppose that
. Then, applying the same technique as in Theorem 2 for
,
, and
, it follows that
and we take
In addition, by virtue of (56), we have
where
For the last equation vanishes, that is, locally
or
if and only if
Next, we assume
and we prove that
is constant and
(the other case proceeds in the same way). Differentiating (75) with respect to
, we obtain
, and then again differentiating with respect to
, we obtain
Making use of (46) and (76), we have
Then, differentiating (78) by
, since
, we obtain
Now, taking account of (46) and (71) to calculate
, we obtain
On the other hand, using (25) and
, we also have
Differentiating (75) by
, we find
Moreover, from (75), we obtain
Then, using (78)–(83), and have to vanish. Thus, is constant. In addition, is identically zero. Hence, we find a contradiction in our assumption. We conclude that . This ends the proof. □
Corollary 1. Let be a semi-symmetric -cosymplectic manifold with the Ricci curvature constant along the characteristic vector field . It is locally symmetric if it is cosymplectic when or Kenmotsu when
6. Examples
Example 1. Let us consider the manifold such that . Here, are the standart coordinates in The vector fields are as follows: Let
be the metric tensor product given by
where
are defined by
with
for constants
, and
It is obvious that
are linearly independent at each point of
Therefore, we have
for any
.
From the above relations, there exists an almost contact metric structure
on
. Now, we check if the structure is almost
-Kenmotsu metric or not. Hence, we obtain
Since we deduce on Moreover, we notice that Thus, is an -Kenmotsu manifold, and with constant curvature Consequently, Theorem 1 and Theorem 3 are verified.
Example 2. Consider the three-dimensional manifold , where are the standart coordinates in The vector fields are Here,
and
are given by
with
for constants
, and
. It is sufficient to check that the only non-zero components of the second fundamental form
are
The above equation gives that
We notice that the structure is not normal; the given structure is almost
-cosymplectic. In addition, by simple calculation, the Riemannian curvature tensor components are as follows:
Thus, Theorem 2 and Theorem 3 hold.