Research article

Solving the system of nonlinear integral equations via rational contractions

  • Received: 22 October 2020 Accepted: 11 January 2021 Published: 21 January 2021
  • MSC : 54H25

  • In this paper, some coupled coincidence point theorems for two mappings established using rational type contractions in the setting of partially ordered Gmetric spaces. Moreover, supporting examples are provided to strengthen our obtained results. By considering Gmetric space, we propose a fairly simple solution for a system of nonlinear integral equations by using fixed point technique.

    Citation: Kumara Swamy Kalla, Sumati Kumari Panda, Thabet Abdeljawad, Aiman Mukheimer. Solving the system of nonlinear integral equations via rational contractions[J]. AIMS Mathematics, 2021, 6(4): 3562-3582. doi: 10.3934/math.2021212

    Related Papers:

    [1] Mohd Danish Siddiqi, Fatemah Mofarreh . Schur-type inequality for solitonic hypersurfaces in (k,μ)-contact metric manifolds. AIMS Mathematics, 2024, 9(12): 36069-36081. doi: 10.3934/math.20241711
    [2] Hamed Faraji, Shahroud Azami, Ghodratallah Fasihi-Ramandi . h-Almost Ricci solitons with concurrent potential fields. AIMS Mathematics, 2020, 5(5): 4220-4228. doi: 10.3934/math.2020269
    [3] Rabia Cakan Akpınar, Esen Kemer Kansu . Metallic deformation on para-Sasaki-like para-Norden manifold. AIMS Mathematics, 2024, 9(7): 19125-19136. doi: 10.3934/math.2024932
    [4] Yusuf Dogru . η-Ricci-Bourguignon solitons with a semi-symmetric metric and semi-symmetric non-metric connection. AIMS Mathematics, 2023, 8(5): 11943-11952. doi: 10.3934/math.2023603
    [5] Yanlin Li, Dipen Ganguly, Santu Dey, Arindam Bhattacharyya . Conformal η-Ricci solitons within the framework of indefinite Kenmotsu manifolds. AIMS Mathematics, 2022, 7(4): 5408-5430. doi: 10.3934/math.2022300
    [6] Yanlin Li, Aydin Gezer, Erkan Karakaş . Some notes on the tangent bundle with a Ricci quarter-symmetric metric connection. AIMS Mathematics, 2023, 8(8): 17335-17353. doi: 10.3934/math.2023886
    [7] Shahroud Azami, Mehdi Jafari, Nargis Jamal, Abdul Haseeb . Hyperbolic Ricci solitons on perfect fluid spacetimes. AIMS Mathematics, 2024, 9(7): 18929-18943. doi: 10.3934/math.2024921
    [8] Rajesh Kumar, Sameh Shenawy, Lalnunenga Colney, Nasser Bin Turki . Certain results on tangent bundle endowed with generalized Tanaka Webster connection (GTWC) on Kenmotsu manifolds. AIMS Mathematics, 2024, 9(11): 30364-30383. doi: 10.3934/math.20241465
    [9] Fahad Sikander, Tanveer Fatima, Sharief Deshmukh, Ayman Elsharkawy . Curvature analysis of concircular trajectories in doubly warped product manifolds. AIMS Mathematics, 2024, 9(8): 21940-21951. doi: 10.3934/math.20241066
    [10] Abdul Haseeb, Fatemah Mofarreh, Sudhakar Kumar Chaubey, Rajendra Prasad . A study of -Ricci–Yamabe solitons on LP-Kenmotsu manifolds. AIMS Mathematics, 2024, 9(8): 22532-22546. doi: 10.3934/math.20241096
  • In this paper, some coupled coincidence point theorems for two mappings established using rational type contractions in the setting of partially ordered Gmetric spaces. Moreover, supporting examples are provided to strengthen our obtained results. By considering Gmetric space, we propose a fairly simple solution for a system of nonlinear integral equations by using fixed point technique.



    In his famous paper on Ricci flows [1], Richard S. Hamilton introduced Ricci solitons, which have played a crucial role in deepening our comprehension of the Ricci flow's long-term dynamics and its relationship with geometric structures. Ricci solitons represent self-similar solutions that provide valuable perspectives on the geometry of manifolds.

    A Ricci soliton on a Riemannian manifold (N,g) is defined by the existence of a vector field η and a constant κ such that the following equation holds:

    Ric+12Lηg=κg, (1.1)

    where Lηg denotes the Lie derivative of the metric g in the direction of η, and Ric represents the Ricci curvature of (N,g). In this context, (N,g,η,κ) is referred to as a Ricci soliton, with η known as the potential vector field and κ as the Ricci soliton constant.

    In [2], S. Pigola introduced the concept of an almost Ricci soliton by allowing the Ricci soliton constant κ to be a function. More specifically, a quadruple (N,g,η,κ) is defined as an almost Ricci soliton (ARS) if Eq (1.1) holds, where κ is a smooth function on N, referred to as the potential function of the ARS.

    If the soliton potential vector field η is Killing, then the ARS (N,g,η,κ) is said to be trivial. We also say that the ARS (N,g,η,κ) is parallel if its potential vector field η is a parallel vector field.

    An almost Ricci soliton (N,g,η,κ) is categorized as steady under the condition κ=0, as shrinking under the condition κ>0, and as expanding under the condition κ<0. When the potential vector field η is gradient, i.e, η=Df for some smooth function f on N, the almost Ricci soliton (N,g,f,κ) is called a gradient. Equation (1.1) becomes

    Ric+Hf=κg, (1.2)

    where Hf is the Hessian of the function f.

    The study of almost Ricci solitons (ARS) provides a natural generalization of Ricci solitons by allowing the soliton constant κ to vary as a smooth function rather than remaining constant. These structures naturally appear in the study of Ricci flow, where they describe shapes of spaces that evolve in a self-similar way over time. Recently, researchers have paid more attention to compact ARS spaces because they often show strong and rigid geometric behavior under certain curvature conditions.

    In [3], Corollary 1 and Theorem 2, it was shown that in dimension 3, if a compact ARS has either constant scalar curvature or a nontrivial conformal vector field, then the space must be isometric to a standard Euclidean sphere. This extends earlier results about Ricci solitons to the more general ARS case, showing that symmetry and compactness strongly influence the geometry.

    In [4], it has been shown that if a nontrivial ARS (N,g,η,κ) has constant scalar curvature, then the vector field η is gradient, and the space (N,g) must again be a sphere. This connects ARS spaces with classical gradient Ricci solitons, suggesting that many of the same rigid properties still apply even in this more general setting.

    It is also known (see [5]) that all compact Ricci solitons are gradient solitons. While this does not always hold for ARS (because κ is not constant), adding the condition of constant scalar curvature forces the ARS to behave like a gradient soliton.

    In two dimensions, Hamilton showed in [6] that all compact Ricci solitons are trivial. However, [3] gave an example of a nontrivial compact two-dimensional almost Ricci soliton, proving that ARS allow for more complex geometry than Ricci solitons, especially in low dimensions.

    For a more detailed study of these topics, including classification results and more examples, readers can refer to the works (see [7,8,9,10,11,12,13]).

    Finally, it is worth noting that many researchers have effectively applied soliton theory across various scientific and engineering disciplines. They have also devised several powerful techniques for obtaining analytic solutions, including the inverse scattering transform [14] and Hirota's bilinear method [15], among others.

    The present work presents several sufficient conditions for a compact almost Ricci soliton to be isometric to a sphere and other conditions under which it becomes trivial or parallel.

    The organization of this paper is as follows: Section 2 contains the fundamental concepts, presenting an overview of the essential principles and key equations within almost Ricci soliton theory. Section 3 determines the sufficient conditions for an almost Ricci soliton on a compact Riemannian manifold to be trivial or parallel. Theorem 3.6 states that given a non-parallel almost Ricci soliton (N,g,η,κ) on an n-dimensional compact and connected Riemannian manifold (N,g), where n3, then, (N,g) is isometric to a Euclidean sphere Sn(c) if and only if the following inequality holds:

    N(Ric(η,η)+(n1)g(η,Dκ))dv0. (1.3)

    Theorem 3.8 demonstrates that for an almost Ricci soliton on an n-dimensional compact Riemannian manifold (N,g), with n2, the condition

    N(Ric(η,η)+(n2)g(η,Dκ))dv0 (1.4)

    leads to the potential vector field being parallel. This extends Theorem 3 in [3], which shows that an almost Ricci soliton on a compact Riemannian manifold, with n3, is trivial (i.e, the potential vector field is Killing) if condition (1.4) is satisfied. This also generalizes Theorem 1.1 in [16], stating that a compact Ricci soliton is trivial whenever

    N(Ric(η,η))dv0.

    On an n-dimensional Riemannian manifold (N,g), where n2, the Riemannian curvature tensor R is given by:

    R(U,V)Z=D[U,V]ZDUDVZ+DVDUZ, (2.1)

    for all U,V,ZX(N), with D being the Levi-Civita connection on (N,g) and X(N) representing the collection of smooth vector fields on (N,g). For more details, see [17].

    If {e1,,en} is a local orthonormal frame on the manifold (N,g), the divergence of a vector field UX(N) is described by the equation:

    div(U)=ni=1g(DDeieiU,ei).

    Similarly, the Laplacian of U is given by:

    ΔU=ni=1DeiDeiUDDeieiU.

    The divergence of the (1,1)-tensor ϕ on (N,g) is expressed as:

    div(ϕ)=ni=1(Deiϕ)(ei).

    The gradient of a smooth function f on N is the vector field Df characterized by:

    g(Df,U)=U(f), (2.2)

    for all UX(N), and the Hessian of f is the symmetric (0,2) tensor field defined by"

    Hf(U,V)=g(DUDf,V), (2.3)

    for all U,VX(N).

    The Laplacian of a smooth function f on N is defined as:

    Δf=div(Df).

    Taking the trace of (1.1) yields

    S+div(η)=nκ, (2.4)

    where S is the scalar curvature of (N,g). Also, we have

    div(κη)=g(η,Dκ)+nκ2κS, (2.5)
    div(Sη)=g(η,DS)+nκSS2, (2.6)

    and

    div((nκS)η)=(nκS)2+η(nκS). (2.7)

    For a gradient Ricci soliton (N,g,f,κ), Equation (2.4) becomes

    S+Δf=nκ. (2.8)

    Take a Riemannian manifold (N,g), and ηX(N). Let θη be the 1-form dual to η (i.e, θη(U)=g(U,η),UX(N)), then the Lie derivative Lη can be written as

    Lηg(U,V)+dθη(U,V)=2g(DUη,V), (2.9)

    for all U,VX(N).

    Now, consider B a symmetric tensor field, and ϕ a skew-symmetric tensor field on (N,g) with the following equations:

    Lηg(U,V)=2g(B(U),V), (2.10)

    and

    dθη(U,V)=2g(ϕ(U),V), (2.11)

    where U,VX(N). We get from (2.9) that

    DUη=B(U)+ϕ(U), (2.12)

    for each UX(N).

    A vector field η is called closed conformal if

    DUη=ΨU, (2.13)

    for each UX(N), where Ψ is a smooth function on (N,g).

    In this subsection, we prove several key lemmas that will be crucial for the rest of the paper.

    Assume (N,g,η,κ) is an almost Ricci soliton on a compact Riemannian manifold (N,g). By substituting (2.10) in (1.1), we deduce that

    B=κIQ, (3.1)

    where Q is the Ricci operator, that is, the operator satisfying

    Ric(U,V)=g(Q(U),V),

    for all U,VX(N). Also, by substituting (3.1) in (2.12), it follows that

    DUη=κUQ(U)+ϕ(U), (3.2)

    for all UX(N).

    The following lemma gives the value of the Ricci curvature in the direction of η.

    Lemma 3.1. Let (N,g,η,κ) be an almost Ricci soliton on an n-dimensional Riemannian manifold (N,g). Then

    Ric(η,V)=(1n)V(κ)+12V(S)g(div(ϕ),V), (3.3)

    for all VX(N).

    Proof. From (2.1), we obtain

    R(U,V)η=U(κ)VV(κ)U(DUQ)(V)+(DVQ)(U)+(DUϕ)(V)(DVϕ)(U),

    for all U,VX(N).

    Assume {e1,,en} is a local orthonormal frame on (N,g) such that it is parallel. Considering equation (1.1) and Corollary 54 in [17], p 88, we obtain

    Ric(V,η)=ni=1g(R(ei,V)η,ei)=ni=1g(ei(κ)VV(κ)ei(DeiQ)(V)+(DVQ)(ei)+(Deiϕ)(V)(DVϕ)(ei),ei)=V(κ)nV(κ)+g(div(Q),V)g(div(ϕ),V)=(1n)V(κ)+12V(S)g(div(ϕ),V).

    Remark 1. From the above lemma, we see that

    Q(η)=(1n)Dκ+12DSdiv(ϕ). (3.4)

    In the following lemma, we compute the divergence of the image of η by the Ricci tensor Q.

    Lemma 3.2. Let (N,g,η,κ) be an almost Ricci soliton on an n-dimensional Riemannian manifold (N,g). Then

    div(Q(η))=12g(DS,η)+κS||Q||2. (3.5)

    Proof. Consider {e1,,en} as a local orthonormal frame on (N,g). By considering (3.4), we deduce

    div(Q(η))=ni=1g(DeiQ(η),ei)=ni=1g((DeiQ)(η),ei)+ni=1g(Q(Deiη),ei)=ni=1g(η,(DeiQ)ei)+ni=1g(Q(Deiη),ei)=g(η,ni=1(DeiQ)ei)+ni=1g(Deiη,Q(ei))=g(η,12DS)+ni=1g(κeiQ(ei)+ϕ(ei),Q(ei))=g(η,12DS)+κni=1g(ei,Q(ei))ni=1g(Q(ei),Q(ei))+ni=1g(Q(ei),ϕ(ei))=g(η,12DS)+κS||Q||2.

    The following formula is well known in the case of Ricci solitons. We extend it here to the case of almost Ricci solitons.

    Lemma 3.3. Let (N,g) be an n-Riemannian manifold. If (N,g,η,κ) is an almost Ricci soliton on (N,g), then

    ||RicSng||2=||Q||2S2n. (3.6)

    Proof. Let λ1,λ2,...,λn be the eigenvalues of the Ricci operator Q. Then

    ||RicSng||2=ni=1(λiSn)2=ni=1(λ2i+S2n22Snλi)=ni=1λ2i+ni=1S2n22Snni=1λi=||Q||2+S2n2S2n=||Q||2S2n.

    The following lemma provides a useful formula for the Laplacian of the potential vector field for an almost Ricci soliton.

    Lemma 3.4. Let (N,g,η,κ) be an almost Ricci soliton on an n-Riemannian manifold (N,g). Then

    Δη=(2n)DκQ(η). (3.7)

    Proof. Consider {e1,,en} as a local orthonormal frame on (N,g). Considering Eq (1.1) and Corollary 54 in [17], p 88, we obtain

    Δη=ni=1DeiDeiηDDeieiη=ni=1Dei(κeiQ(ei)+ϕ(ei))=ni=1(g(κ,ei)eiDeiQ(ei)+Deiϕ(ei))=ni=1(g(κ,ei)ei(DeiQ)(ei)+(Deiϕ)(ei))=Dκ12DS+div(ϕ).

    Inserting (3.4) in the preceding equation yields (3.7).

    Theorem 3.5. Let (N,g,η,κ) be an almost Ricci soliton on an n-dimensional compact Riemannian manifold (N,g), where n2. If the inequality

    N(Ric(η,η)+n(κ2+g(η,Dκ)))dv0 (3.8)

    is satisfied, then the manifold (N,g) is Ricci flat, and η is a closed conformal vector field.

    Proof. By (3.3) and (3.5), we have

    Ric(η,η)=(1n)g(η,Dκ)+12g(η,DS)g(div(ϕ),η)=(1n)g(η,Dκ)+12g(η,DS)+||ϕ||2+div(ϕ(η))=(1n)g(η,Dκ)+div(Q(η))κS+||Q||2+||ϕ||2+div(ϕ(η)).

    Consequently, it follows that

    Ric(η,η)+(n1)g(η,Dκ)=div(Q(η))κS+||Q||2+||ϕ||2+div(ϕ(η)). (3.9)

    By substituting the value of κS from (2.5) in (3.9), we deduce that

    Ric(η,η)=ng(η,Dκ)+div(Q(η))div(κη)nκ2+||Q||2+||ϕ||2+div(ϕ(η)). (3.10)

    By integrating (3.10), one obtains

    N(Ric(η,η)+n(κ2+g(η,Dκ)))dv=N(||Q||2+||ϕ||2)dv. (3.11)

    Now, it is clear that condition (3.8) in (3.11) yields Q=ϕ=0. We deduce from (2.12) that Dη=κI, which means that η is a closed conformal vector field.

    Also, since Q=0 we get Ric=0, meaning that (N,g) is Ricci flat.

    Theorem 3.6. Let (N,g) be an n-dimensional connected and compact Riemannian manifold, where n3. If (N,g,η,κ) is a non-parallel almost Ricci soliton, then (N,g) is isometric to a Euclidean sphere Sn(c) if and only if the subsequent inequality is satisfied

    N(Ric(η,η)+(n1)g(η,Dκ))dv0. (3.12)

    Proof. From (2.6), we have

    κS=1n(div(Sη)+S2g(η,DS)). (3.13)

    By substituting the value of κS in (3.9) and then utilizing (3.6), we obtain

    Ric(η,η)+(n1)g(η,Dκ)=||RicSng||2+div(Q(η)1ndiv(Sη)+ϕ(η))+1ng(η,DS)+||ϕ||2. (3.14)

    On the other hand, Theorem 1 in [4] states that

    N||RicSng||2dv=n22nNg(η,DS)dv. (3.15)

    By integrating Eq (3.14), followed by the substitution of equation (3.15) in the result, we derive

    N(Ric(η,η)+(n1)g(η,Dκ))dv=12N||RicSng||2dv+N||ϕ||2dv.

    Taking into account the hypothesis (1.3), we deduce that Ric=Sng and ϕ=0.

    Given that n3 and Ric=Sng, the Schur Lemma (refer to [17]) can be applied, implying that S must be constant. Also, because ϕ=0, η is a closed conformal vector field with conformal function κSn, that is

    DUη=(κSn)U, (3.16)

    for all UX(N).

    Given that η is not parallel, it follows that κSn meaning that η is a non-Killing conformal vector field. From the fact that Ric=Sng and equation (3.16), we obtain

    LηRic=2Sn(κSn)g. (3.17)

    This relation allows us to apply Theorem 4.2 from [18], p. 54, leading to the conclusion that (N,g) is isometric to the standard Euclidean sphere Sn(c) with radius 1c, where c=Sn(n1).

    Conversely, assume that (N,g,η,κ) is a non-parallel almost Ricci soliton on a connected, compact Riemannian manifold (N,g), where (N,g) is isometric to a Euclidean sphere Sn(c), with c>0. In other words, the Ricci curvature is expressed as Ric=(n1)cg, whereas the scalar curvature is defined by S=n(n1)c. From (1.1), it follows that

    12Lηg=(κ(n1)c)g. (3.18)

    But, by Corollary 2 in [4], η is gradient, i.e, η=Df for some smooth function f on N. Then (3.18) becomes

    Hf=(κ(n1)c)g.

    It necessarily follows that κ(n1)c, as otherwise D2f=0, implying that Df=η is a constant vector field, indicating that η is parallel.

    As demonstrated by Lemma 2.3 in [18] on page 52, Eq (3.18) shows that

    Δκ=nc(κ(n1)).

    Additionally, as indicated by (2.4), it follows that

    Δf=nκS=n(κ(n1)).

    We deduce that Δ(κ+cf)=0. According to the well-known Hopf Lemma, this implies that κ+cf is a constant.

    Consequently, Dκ=cDf=cη, and therefore

    Ric(η,η)+(n1)g(η,Dκ)=c(n1)|η|2c(n1)|η|2=0.

    This demonstrates that the inequality (1.3) holds true as an equality.

    Next, we shall give an example of an almost Ricci soliton on the Euclidean sphere Sn(c), where c>0 satisfies (1.3).

    Example 1. Consider (Sn(c),,), the Euclidean sphere in the Euclidean space (Rn+1,,) of radius 1c, where c>0.

    Let and ¯ be the Levi-Civita connections of Sn(c) and Rn+1, respectively.

    Let ˉZX(Rn+1) be a constant vector field, with Z as its restriction to Sn(c).

    If ψ:Sn(c)Rn+1 represents the position vector field, then consider the unit normal vector field on Sn(c), defined by N=cψ.

    Write

    Z=ZT+θN, (3.19)

    where ZT is the tangential component of Z, and θ=Z,N.

    Employing the Gauss and Weingarten formulas, we obtain

    UZT=cθU, (3.20)

    and

    θ=cZT. (3.21)

    Consequently, ZT serves as a closed conformal vector field associated with the conformal function cθ.

    Assume that (Sn(c),,,ZT,κ) is an almost Ricci soliton on (Sn(c),,). Given that for the Euclidean sphere (Sn(c), we have Ric=(n1)cg, and from (3.20) we know that 12LZTg=cθg, by substituting these expressions into equation (1.1), we deduce that

    cθI+(n1)cI=κI,

    indicating that the function κ is given explicitely by

    κ=(n1)cθc.

    Since Ric(ZT,ZT)=(n1)cZT,ZT and ZT,Dκ=cZT,ZT, it follows that

    Sn(c)(Ric(ZT,ZT)+(n1)ZT,Dκ)dv=0,

    meaning that the inequality (1.3) is satisfied as an equality.

    We close this subsection with the following result, which characterizes Euclidean spheres in terms of the constancy of the scalar curvature along the integral curves of the potential vector field.

    Theorem 3.7. Let (N,g) be an n-dimensional compact oriented Riemannian manifold, n3. If (N,g,η,κ) is a non-trivial almost Ricci soliton on (N,g), then (N,g) is isometric to a Euclidean sphere if and only if η leaves the scalar curvature S constant (i.e, S is constant along the integral curves of η).

    Proof. Assume that η leaves the scalar curvature S constant. This means that g(η,DS)=0.

    Substituting this in Eq (3.15), we obtain

    N||RicSng||2dv=0.

    This implies that Ric=Sng. With n3, Schur's Lemma implies that S must be constant. According to (1.1), we have

    12Lηg=(κSn)g=1n(nκS)g=div(η)ng.

    Given that (N,g,η,κ) is non-trivial, it follows that div(η)0, indicating that η is a non-trivial conformal vector field. It follows that η is a non-homothetic conformal vector field on a compact Einstein Riemannian manifold. Following a similar argument as in the proof of Theorem 3.6 above, and applying Theorem 4.2 from [18], p. 54, we conclude that (N,g) is isometric to a Euclidean sphere. The converse is trivial, since if (N,g) is isometric to a Euclidean sphere, then the scalar curvature is constant.

    The next result extends Theorem 3 from [3], shown when n3 and with the conclusion that η is a Killing vector field. It also generalizes Theorem 1.1 from [16], which proved for a Ricci soliton on a compact Riemannian manifold that satisfies N(Ric(η,η))dv0.

    Theorem 3.8. Let (N,g) be an n-dimensional compact Riemannian manifold, n2. If (N,g,η,κ) is an almost Ricci soliton on (N,g) satisfying the following inequality

    N(Ric(η,η)+(n2)g(η,Dκ))dv0, (3.22)

    then η is a parallel vector field.

    Proof. By Lemma 2 in [3], we have

    12Δ|η|2=|Dη|2Ric(η,η)(n2)g(η,Dκ). (3.23)

    By integrating the above equation, we obtain

    N(Ric(η,η)+(n2)g(η,Dκ))dv=N|Dη|2dv, (3.24)

    and from the hypothesis (3.22), it follows that Dη=0, which means that η is parallel.

    Theorem 3.9. Let (N,g,η,κ) be an almost Ricci soliton on an n-dimensional compact Riemannian manifold (N,g), where n2. If (N,g,η,κ) satisfies the following inequality

    N(Ric(Q(η),η)+(n1)g(DS,Dκ))dv0, (3.25)

    then η is a parallel vector field, and (N,g) is Einstein.

    Proof. By (3.3) and (3.4), we have

    Ric(Q(η),η)+Ric(div(ϕ),η)=12Ric(η,DS)(n1)Ric(η,Dκ)=12((n1)g(DS,Dκ)+12|DS|2g(div(ϕ,DS))(n1)((n1)g(Dκ,η)+12g(DS,Sκ)g(div(ϕ),Dκ))=(n1)g(DS,Dκ)+12div(ϕ(DS))+14|DS|2+(n1)2|Dκ|2(n1)div(ϕ(Dκ)).

    So, we have

    Ric(Q(η),η)+Ric(div(ϕ),η)=(n1)g(DS,Dκ)+12div(ϕ(DS))+14|DS|2+(n1)2|Dκ|2(n1)div(ϕ(Dκ)). (3.26)

    Also, from (3.3), one obtains

    Ric(div(ϕ),η)=(n1)div(ϕ(Dκ)12div(ϕ(DS))|div(ϕ)|2. (3.27)

    Substituting (3.27) in (3.26) and integrating the resulting equation yields

    N(Ric(Q(η),η)+(n1)g(DS,Dκ))dv=N(14|DS|2+(n1)2|Dκ|2+|div(ϕ)|2)dv. (3.28)

    Assuming the hypothesis (3.25), it follows that the scalar curvature S and the potential function κ are constant, and div(ϕ)=0. From equation (2.4), it is deduced that nκ=S.

    From (3.5), it follows

    ||Q||2=κS.

    It follows that Eq (3.6) becomes

    ||RicSng||2=0.

    This implies that (N,g) is an Einstein manifold, and that η is a Killing vector field.

    Furthermore, given that NRic(η,η)dv=0 from (3.28), then (3.24) implies that

    N|Dη|2dv=0,

    which means that Dη=0, and so η is a parallel vector field.

    Remark According to [19], it is proved that an n-dimensional compact almost Ricci soliton (N,g,η,κ), where n3, becomes a trivial Ricci soliton precisely when nκS remains constant along the integral curve of the potential field η. This condition, essentially, is equivalent to η.div(η)=0. We aim to extend this result as follows.

    Theorem 3.10. An almost Ricci soliton (N,g,η,κ) on an n-dimensional compact Riemannian manifold with n3 is trivial if and only if the vector field η is incompressible, meaning that div(η)=0.

    Proof. Assume that (N,g,η,κ) is trivial. Then, η is a Killing vector field, meaning that Lηg=0. Consequently, div(η)=0.

    Conversely, by assuming that div(η)=0, we deduce that nκ=S. By integrating Eq (2.6), we obtain

    Ng(η,DS)dv=0.

    Substituting this in Eq (3.15), it follows that

    N||RicSng||2dv=0.

    This implies that Ric=Sng. With n3, Schur's Lemma indicates that S must be constant. Since nκ=S, it follows that κ is also constant. Therefore, (N,g,η,κ) is trivial.

    Upon reviewing the previous outcome, we derive the following result:

    Corollary 1. A compact almost Ricci soliton of dimension n3 with an affine potential vector field is trivial.

    Proof. By [20], any affine vector field on a compact Riemannian manifold is incompressible.

    Theorem 3.11. Let (N,g,η,κ) be an almost Ricci soliton on an n-dimensional compact Riemannian manifold (N,g), where n2. If (N,g,η,κ) satisfies the following inequality

    N(g(η,DS)2g(η,Dκ))dv0, (3.29)

    then, η is a Killing vector field, and (N,g,η,κ) is trivial.

    Proof. By taking the inner product of (3.7) with η and subsequently integrating, it follows

    N(Ric(η,η)+g(η,Δη)+(n2)g(η,Dκ))dv=0. (3.30)

    According to Proposition 5.10 in [21], it follows that

    N(Ric(η,η)+g(η,Δη)+12|Lηg|2(div(η))2))dv=0. (3.31)

    Using Eq (2.12), we obtain

    |Dη|2=|B|2+|ϕ|2,

    and

    tr(Dη)2=|B|2|ϕ|2,

    where tr(Dη)2 represents the trace of (Dη)2.

    Referencing Lemma 5.9 in [21], we have

    |Lηg|2=4|B|2. (3.32)

    Utilizing (3.31) and considering Eqs (2.7) and (3.32), we obtain

    N(g(η,DS)2g(η,Dκ))dv=2N|B|2dv. (3.33)

    Assuming that N(g(η,S)2g(η,Dκ))dv0, we deduce that B=0, implying that Ric=κg. According to (1.1), η becomes a Killing vector field, and the almost Ricci soliton is trivial.

    In this paper, we extend and refine recent works about Ricci solitons and almost Ricci solitons by using methods based on integral inequalities that involve Ricci curvature and the potential vector field. Unlike previous studies that often focused on specific curvature assumptions or geometric structures, we use integral conditions to set clear criteria for when a compact Riemannian manifold with an almost Ricci soliton is isometric to a sphere. We also explore the potential vector field in more detail, identifying conditions where it becomes either Killing or parallel, making the almost Ricci soliton trivial. By focusing on both the geometry and the properties of the vector field, our work not only broadens the scope of earlier results but also provides a more unified approach for studying solitons with fewer assumptions. Our methods and findings show how effective integral techniques are in analyzing Ricci-type structures and contribute to the ongoing development of Ricci soliton theory. We expect that future research will focus on exploring the properties of almost Ricci solitons in important contexts, like spacelike submanifolds of Lorentzian manifolds, especially within generalized Robertson-Walker (GRW) spacetimes and general Lorentzian warped products.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors would like to extend their sincere appreciation to Ongoing Research Funding Project, (ORF-2025-824), King Saud University, Riyadh, Saudi Arabia.

    The authors declare that they have no conflicts of interest.

    Conceptualization, M.G.; Investigation, N.A. and M.G.; Methodology, N.A. and M.G.; Resources, N.A.; Validation, M.G.; Writing-original draft, N.A.; Writing-review and editing, N.A. and M.G.



    [1] M. Abbas, W. Sintunavarat, P. Kumam, Coupled fixed point of generalized contractive mappings on partially orderedG-metric spaces, Fixed Point Theory Appl., 2012 (2012), 31. doi: 10.1186/1687-1812-2012-31
    [2] T. G. Bhaskar, V. Lakshmikantham, Fixed point theorems in partially ordered metric spaces and applications, Nonlinear Anal., 65 (2006), 1379–1393. doi: 10.1016/j.na.2005.10.017
    [3] S. Chandok, W. Sintunavarat, P. Kumam, Some coupled common fixed points for pair of mappings in partially ordered G-metric spaces, Math. Sci., 7 (2013), 14. doi: 10.1186/2251-7456-7-14
    [4] K. Charkrabarti, Coupled fixed point theorems with rational type contractive condition in a partially ordered G-metric space, J. Math., 2014 (2014), 785357.
    [5] V. S. Chouhan, R. Sharma, Coupled ixed point Theorems for rational contractions in partially ordered G-metric spaces, Int. J. Math. Sci. Appl., 6 (2016), 743–754.
    [6] B. S. Choudhury, P. Maity, Coupled fixed point results in generalized metric spaces, Math. Comput. Modell., 54 (2011), 73–79. doi: 10.1016/j.mcm.2011.01.036
    [7] H. S. Ding, L. Li, Coupled fixed point theorems in partially ordered cone metric spaces, Filomat 25 (2011), 137–149.
    [8] E. Karapinar, A. Ravi Paul, Further Remarks on G-metric spaces, Fixed Point Theory Appl., 154 (2013), 1–19.
    [9] E. Karpinar, P. Kumam, I. M. Erhan, Coupled ixed point theorems on partially ordered G-metric spaces, Fixed Point Thory Appl., 2012 (2012), 174. doi: 10.1186/1687-1812-2012-174
    [10] E. Karapinar, N. V. Luong, N. X. Thuan, T. T. Hai, Coupled coincidence points for mixed monotone operators in partially ordered metric spaces, Arab. J. Math., 1 (2012), 329–339. doi: 10.1007/s40065-012-0027-0
    [11] V. Lakshmikantham, L. Ciric, Coupled fixed point theorems for nonlinearcontractions in partially ordered metric spaces, Nonlinear Anl., 70 (2009), 4341–4349. doi: 10.1016/j.na.2008.09.020
    [12] G. Meenakshi, A. Madhu, C. Renu, Common Fixed Point Results in G-metric spaces and Applications, Int. J. Comput. Appl., 43 (2012), 38–42.
    [13] Z. Mustafa, Some new common fixed point theorems under strict contractive conditions in G-metric spaces, J. Appl. Math., (2012), 248397.
    [14] Z. Mustafa, H. Obiedat, A fixed point theorem of Reich in G-Metric Spaces, CUBO A Math. J., 12 (2010), 83–93.
    [15] Z. Mustafa, H. Obiedat, F. Awawdeh, Some fixed point theorem for mapping on complete G-Metric Spaces, Fixed Point Theory Appl., (2008), 189870.
    [16] Z. Mustafa, W. Shatanawi, M. Bataineh, Existence of Fixed Point Results in G-Metric Spaces, Int. J. Math. Math. Sci., (2009), 283028.
    [17] Z. Mustafa, B. Sims, A new approach to generalized metric spaces, J. Nonlinear Convex Anal., 7 (2006), 289–297.
    [18] Z. Mustafa, B. Sims, Some remarks concerning D-metric spaces, Paperpresented at the international conference on fixed point theory and applications, Valencia Spain, 2004,189–198.
    [19] H. K. Nashine, Coupled common fixed point results in ordered G-metric spaces, J. Nonlinear Sci. Appl., 1 (2012), 1–13.
    [20] T. Abdeljawad, E. Karapınar, S. K. Panda, N, Mlaiki, Solutions of boundary value problems on extended-Branciari b-distance, J. Inequalities Appl., 2020 (2020), 103. doi: 10.1186/s13660-020-02373-1
    [21] C. Ravichandran, S. K. Panda, K. S. Nisar, K. Logeswari, On new approach of fractional derivative by Mittag-Leffler kernel to neutral integro-differential systems with impulsive conditions, Chaos, Solitons Fractals, 139 (2020), 110012. doi: 10.1016/j.chaos.2020.110012
    [22] S. K. Panda, E. Karapınar, A. Atangana, A numerical schemes and comparisons for fixed point results with applications to the solutions of Volterra integral equations in dislocatedextendedb-metricspace, Alexandria Eng. J., 59 (2020), 815–827. doi: 10.1016/j.aej.2020.02.007
    [23] S. K. Panda, T. Abdeljawad, C. Ravichandran, Novel fixed point approach to Atangana-Baleanu fractional and Lp-Fredholm integral equations, Alexandria Eng. J., 59 (2020), 1959–1970. doi: 10.1016/j.aej.2019.12.027
    [24] S. K. Panda, Applying fixed point methods and fractional operators in the modelling of novel coronavirus 2019-nCoV/SARS-CoV-2, Res. Phys., 19 (2020), 103433.
    [25] S. K. Panda, T. Abdeljawad, C. Ravichandran, A complex valued approach to the solutions of Riemann-Liouville integral, Atangana-Baleanu integral operator and non-linear Telegraph equation via fixed point method, Chaos, Solitons Fractals, 130 (2020), 109439. doi: 10.1016/j.chaos.2019.109439
    [26] S. K. Panda, A. Tassaddiq, R. P. Agarwal, A new approach to the solution of non-linear integral equations via various FBe-contractions, Symmetry, 11 (2019), 206. doi: 10.3390/sym11020206
    [27] T. Abdeljawad, R. P. Agarwal, E. Karapınar, P. S. Kumari, Solutions of the nonlinear integral equation and fractional differential equation using the technique of a fixed point with a numerical experiment in extended b-metric space, Symmetry, 11 (2019), 686. doi: 10.3390/sym11050686
    [28] S. K. Panda, O. Alqahtani, E. Karapınar, Some fixed-point theorems in b-dislocated metric space and applications, Symmetry, 10 (2018), 691. doi: 10.3390/sym10120691
    [29] V. Vijayakumar, U. Ramalingam, S. K. Panda, K. S. Nisar, Results on approximate controllability of Sobolev type fractional stochastic evolution hemivariational inequalities, Numer. Methods Partial Differ. Equations, (2020), 22690.
    [30] V. Vijayakumar, S. K. Panda, K. S. Nisar, H. M. Baskonus, Results on approximate controllability results for second-order Sobolev-type impulsive neutral differential evolution inclusions with infinite delay, Numer. Methods Partial Differ. Equations, (2020), 22573.
    [31] S. K. Panda, C. Ravichandran, B. Hazarika, Results on system of Atangana-Baleanu fractional order Willis aneurysm and nonlinear singularly perturbed boundary value problems, Chaos, Solitons Fractals, (2020), 110390.
    [32] M. Asif, I. Khana, N. Haidera, Q. Al-Mdallalb, Legendre multi-wavelets collocation method for numerical solution of linear and nonlinear integral equations, Alexandria Eng. J., 59 (2020), 5099–5109. doi: 10.1016/j.aej.2020.09.040
    [33] A. Babakhani, Q. Al-Mdallal, On the existence of positive solutions for a non-autonomous fractional differential equation with integral boundary conditions, Comput. Methods Differ. Equations, (2020).
    [34] Q. M. Al-Mdallal, Monotone iterative sequences for nonlinear integro-differential equations of second order, Nonlinear Anal.: Real World Appl., 12 (2011), 3665–3673. doi: 10.1016/j.nonrwa.2011.06.023
    [35] Q. M. Al-Mdallal, Boundary value problems for nonlinear fractional integro-differential equations: theoretical and numerical results, Adv. Differ. Equations, 2012 (2012), 18. doi: 10.1186/1687-1847-2012-18
    [36] M. I. Syam, Q. M. Al-Mdallal, M. N. Anwar, An efficient numerical algorithm for solving fractional higher-order nonlinear integrodifferential equations, Abstr. Appl. Anal., 2015 (2015), 616438.
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3015) PDF downloads(176) Cited by(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog