Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Sturmian comparison theorem for hyperbolic equations on a rectangular prism

  • In this paper, new Sturmian comparison results were obtained for linear and nonlinear hyperbolic equations on a rectangular prism. The results obtained for linear equations extended those given by Kreith [Sturmian theorems on hyperbolic equations, Proc. Amer. Math. Soc., 22 (1969), 277-281] in which the Sturmian comparison theorem for linear equations was obtained on a rectangular region in the plane. For the purpose of verification, an application was described using an eigenvalue problem.

    Citation: Abdullah Özbekler, Kübra Uslu İşler, Jehad Alzabut. Sturmian comparison theorem for hyperbolic equations on a rectangular prism[J]. AIMS Mathematics, 2024, 9(2): 4805-4815. doi: 10.3934/math.2024232

    Related Papers:

    [1] Jianan Qiao, Guolin Hou, Jincun Liu . Analytical solutions for the model of moderately thick plates by symplectic elasticity approach. AIMS Mathematics, 2023, 8(9): 20731-20754. doi: 10.3934/math.20231057
    [2] Murat O. Mamchuev . Non-local boundary value problem for a system of fractional partial differential equations of the type I. AIMS Mathematics, 2020, 5(1): 185-203. doi: 10.3934/math.2020011
    [3] Naeem Saleem, Salman Furqan, Mujahid Abbas, Fahd Jarad . Extended rectangular fuzzy $ b $-metric space with application. AIMS Mathematics, 2022, 7(9): 16208-16230. doi: 10.3934/math.2022885
    [4] Ting Huang, Shu-Xin Miao . More on proper nonnegative splittings of rectangular matrices. AIMS Mathematics, 2021, 6(1): 794-805. doi: 10.3934/math.2021048
    [5] Ahmed M.A. El-Sayed, Eman M.A. Hamdallah, Hameda M. A. Alama . Multiple solutions of a Sturm-Liouville boundary value problem of nonlinear differential inclusion with nonlocal integral conditions. AIMS Mathematics, 2022, 7(6): 11150-11164. doi: 10.3934/math.2022624
    [6] Nihan Turan, Metin Başarır, Aynur Şahin . On the solutions of the second-order $ (p, q) $-difference equation with an application to the fixed-point theory. AIMS Mathematics, 2024, 9(5): 10679-10697. doi: 10.3934/math.2024521
    [7] Zhengang Zhao, Yunying Zheng, Xianglin Zeng . Finite element approximation of fractional hyperbolic integro-differential equation. AIMS Mathematics, 2022, 7(8): 15348-15369. doi: 10.3934/math.2022841
    [8] Jinming Cai, Shuang Li, Kun Li . Matrix representations of Atkinson-type Sturm-Liouville problems with coupled eigenparameter-dependent conditions. AIMS Mathematics, 2024, 9(9): 25297-25318. doi: 10.3934/math.20241235
    [9] Efruz Özlem Mersin . Sturm's Theorem for Min matrices. AIMS Mathematics, 2023, 8(7): 17229-17245. doi: 10.3934/math.2023880
    [10] Haifa Bin Jebreen, Beatriz Hernández-Jiménez . Pseudospectral method for fourth-order fractional Sturm-Liouville problems. AIMS Mathematics, 2024, 9(9): 26077-26091. doi: 10.3934/math.20241274
  • In this paper, new Sturmian comparison results were obtained for linear and nonlinear hyperbolic equations on a rectangular prism. The results obtained for linear equations extended those given by Kreith [Sturmian theorems on hyperbolic equations, Proc. Amer. Math. Soc., 22 (1969), 277-281] in which the Sturmian comparison theorem for linear equations was obtained on a rectangular region in the plane. For the purpose of verification, an application was described using an eigenvalue problem.



    The classical Sturmian comparison theory of second-order ordinary differential equations and their oscillatory behaviors is acknowledged as the foundation for investigating many fundamental features of their solutions [1,2]. The theory of Sturm comparison on partial differential equations has grown rapidly in the last few decades. Comparison theorems have been explored on various types of partial differential equations and have made significant contributions to the literature [3,4]. Some of remarkable contributions can be counted as for elliptic-type self-adjoint equations by Hartman et al. [5], second-order elliptic equations by Shimoda [6], fourth-order elliptic systems by Kusano et al. [7], a genus of higher-order elliptic systems by Yoshida [8], a genus of second-order half-linear partial differential equations by Kusano et al. [9], a genus of half-linear elliptic equations by Yoshida [10], quasilinear elliptic equations with mixed nonlinearities via Picone-type inequality by Yoshida [11], half-linear elliptic operators with p(x)-Laplacians by Yoshida [12], quasilinear elliptic operators with p(x)-Laplacians by Yoshida [13], and a genus of partial differential equations of order 4m by Jaroš [14] (see also [15,16]). We direct the readers to the monograph by Yoshida [17] for historical development in the oscillation and comparison theory of partial differential equations.

    In 1969, Kreith [18] considered Sturmian comparison theory on hyperbolic differential equations, which was the prime and very likely the sole publication on the topic; see also [19, pp. 24–26]. He considered the pair of hyperbolic equations, which are inspired by the simplistic harmonic motion perception of the Sturm comparison theorem as

    uttuxx+p(x,t)u=0 (1.1)

    and

    vttvxx+q(x,t)v=0 (1.2)

    illustrating throwing two vibrating strings with the same density and elastic constant movement as they oscillate, regarding the equilibrium lines v=0 and u=0 under the effect of continuous restorative forces p(x,t) and q(x,t), respectively. Presume that if

    q(x,t)p(x,t),

    then in some sense Eq (1.2) should oscillate faster than Eq (1.1). Experimentation with straightforward cases that allow variable separation indicates that in the absence of an auxiliary condition, this is not the case. Physically, it is important to analyze finite strings that are elastically bounded at the ends, with the string that is to oscillate more faster being firmly bound. When viewed mathematically, Kreith established an analogue of the Sturm comparison theorem for the pair hyperbolic initial value problems of the form

    uttuxx+p(x,t)u=0,ux(xk,t)+(1)kσk(t)u(xk,t)=0(k=1,2) (1.3)

    and

    vttvxx+q(x,t)v=0,vx(xk,t)+(1)kτk(t)v(xk,t)=0(k=1,2) (1.4)

    on the rectangular domain:

    D={(x,t):x(x1,x2),t(t1,t2)};

    see [18, Theorem 1] and [19, Theorem 3.8].

    Theorem 1.1. Let u be a positive solution of problem (1.3) satisfying the boundary conditions

    u(x,t1)=u(x,t2)=0,x1xx2

    on [x1,x2]×(t1,t2). If q(x,t)p(x,t) on D and τk(t)σk(t) (k=1,2) for t[t1,t2], then every solution v of problem (1.4) has a zero in

    ˉD={(x,t):x[x1,x2],t[t1,t2]}.

    For our purpose, we fix x0,y0,t0R. Let I=(x1,x2)[x0,), J=(y1,y2)[y0,) and K=(t1,t2)[t0,) be three nondegenerate intervals and define the domain (a rectangular prism) as

    Ω=I×J×K. (1.5)

    In this paper, we are attempting to arrange some analogical comparison results for the continuous solutions of a couple of hyperbolic equations

    uttΔu+f(x,y,t)u=0, (1.6)
    vttΔv+g(x,y,t)v=0 (1.7)

    for (x,y,t)Ω, satisfying the initial conditions

    ux(xk,y,t)+(1)krk(t)u(xk,y,t)=0,(y,t)ˉJ×ˉK,uy(x,yk,t)+(1)krk+2(t)u(x,yk,t)=0,(x,t)ˉI×ˉK (1.8)

    and

    vx(xk,y,t)+(1)ksk(t)v(xk,y,t)=0,(y,t)ˉJ×ˉK,vy(x,yk,t)+(1)ksk+2(t)v(x,yk,t)=0,(x,t)ˉI×ˉK (1.9)

    for k=1,2, respectively, where f,gC(ˉΩ,R), rk,skC(ˉK,R) (k=1,2,3,4), and Δ is the usual Laplace operator in R2, i.e.,

    Δ=22x+22y.

    A nontrivial function z(x,y,t) is claimed to be the solution of problem (1.6)-(1.8) if

    i. z:ˉΩRC(ˉΩ,R);

    ii. for each (x,y,t)Ω, it has second-order partial derivatives zxx, zyy, ztt, and z satisfies Eq (1.6);

    iii. z satisfies the initial conditions (1.8) on ˉΩ.

    Solution z(x,y,t)=0 of problem (1.6)-(1.8) has a zero at t=t if z(x,y,t)=0. A solution z of problem (1.6)-(1.8) is said to be an oscillatory if there exists a sequence {ζn}n=1 of real numbers such that z(x,y,ζn)=0 with

    limnζn=.

    Otherwise, z(x,y,t) is said to be nonoscillatory. Moreover, problem (1.6)-(1.8) is said to be oscillatory if all the solutions of it are oscillatory. The similar definition and properties given above are also valid for the solution v of problem (1.7)-(1.9).

    Motivated by the Kreith's comparison result obtained on the rectangular domain in the plane (i.e., Theorem 1.1), we attempt to give an analogous result for continuous solutions of the pair of hyperbolic initial value problems (1.6)-(1.8) and (1.7)-(1.9) on a rectangular prism. The main findings for the linear hyperbolic problems (1.6)-(1.8) and (1.7)-(1.9) are extended to nonlinear hyperbolic problems in Section 3.

    The paper is structured as follows: Sections 2 and 3 are devoted the Sturmian comparison results for linear and nonlinear hyperbolic initial value problems, respectively. The last section deals with an interesting Sturm oscillation result via separation of variables under the assumption that one of the corresponding ordinary differential equations is oscillatory.

    In this section we provide Sturm comparison results for linear hyperbolic initial value problems.

    The first linear comparison consequence of this section is as follows.

    Theorem 2.1. (Sturm comparison theorem) Let u>0 be a solution of problem (1.6)-(1.8) satisfying the boundary conditions

    u(x,y,t1)=u(x,y,t2)=0,(x,y)ˉI×ˉJ (2.1)

    on ˉI×ˉJ×K. If the inequalities

    g(x,y,t)f(x,y,t),(x,y,t)Ω (2.2)

    and

    sk(t)rk(t),tˉK(k=1,2,3,4) (2.3)

    hold, then every solution v of problem (1.7)-(1.9) has a zero in ˉΩ.

    Proof. Assume that a solution v of problem (1.7)-(1.9) has no zero in ˉΩ, then without loss of generality, we may assume that v>0 in ˉΩ. Multiplying Eqs (1.6) and (1.7) by v and u, respectively, and then subtracting, it can be verified that the identity

    [uvxvux]x+[uvyvuy]y+[vutuvt]t=[g(x,y,t)f(x,y,t)]uv (2.4)

    holds for all (x,y,t)ˉΩ.

    Integrating both sides of identity (2.4) over Ω, we obtain

    Ω[g(x,y,t)f(x,y,t)]uvdV=Ω{[uvxvux]x+[uvyvuy]y+[vutuvt]t}dV. (2.5)

    Note that Ω is a simple, solid region with the piece-wise smooth boundary S, so by applying divergence theorem to the smooth vector field F on Ω defined by

    F(x,y,t):=(uvxvux)i+(uvyvuy)j+(vutuvt)k, (2.6)

    the righthand side of (2.5) turns out to be

    Ω{[uvxvux]x+[uvyvuy]y+[vutuvt]t}dV=ΩFdV(=ΩdivFdV)=SFˆNdS, (2.7)

    where ˆN is the unit outward normal on the surface S (=Ω) and the is the usual nabla (gradient) operator defined by

    =ix+jy+kt.

    Note that S is the union of six rectangular regions, that is

    S=6j=1Sj,

    where each Sj are disjoint, oriented, closed surfaces and defined by

    S1={(x,y,t):x=x1,(y,t)J×K},S2={(x,y,t):x=x2,(y,t)J×K},S3={(x,y,t):y=y1,(x,t)I×K},S4={(x,y,t):y=y2,(x,t)I×K},S5={(x,y,t):t=t1,(x,y)I×J}

    and

    S6={(x,y,t):t=t2,(x,y)I×J}.

    The last (surface) integral in (2.7) can be expressed as

    SFˆNdS=6j=1SjFˆNjdS, (2.8)

    where the vectors ˆNj are the unit outward normal vectors on the surfaces Sj, j=1,,6, and, hence, we have

    SFˆNdS=S1FˆN1dS+S2FˆN2dS+S3FˆN3dS+S4FˆN4dS+S5FˆN5dS+S6FˆN6dS. (2.9)

    Since ˆN1=i, ˆN2=i, ˆN3=j, ˆN4=j, ˆN5=k and ˆN6=k, the integrals on the righthand side of (2.9) become

    S1FˆN1dS=S1FidS=t2t1y2y1[uvxvux](x1,y,t)dydt, (2.10)
    S2FˆN2dS=S2FidS=t2t1y2y1[uvxvux](x2,y,t)dydt, (2.11)
    S3FˆN3dS=S3FjdS=x2x1t2t1[uvyvuy](x,y1,t)dtdx, (2.12)
    S4FˆN4dS=S4FjdS=x2x1t2t1[uvyvuy](x,y2,t)dtdx, (2.13)
    S5FˆN5dS=S5FkdS=y2y1x2x1[vutuvt](x,y,t1)dxdy (2.14)

    and

    S6FˆN6dS=S6FkdS=y2y1x2x1[vutuvt](x,y,t2)dxdy. (2.15)

    Imposing the initial conditions (1.8) and (1.9) in the integrals on the righthand sides of (2.10)–(2.13), we get that

    S1FˆN1dS=t2t1y2y1[r1(t)s1(t)]u(x1,y,t)v(x1,y,t)dydt, (2.16)
    S2FˆN2dS=t2t1y2y1[r2(t)s2(t)]u(x2,y,t)v(x2,y,t)dydt, (2.17)
    S3FˆN3dS=x2x1t2t1[r3(t)s3(t)]u(x,y1,t)v(x,y1,t)dtdx (2.18)

    and

    S4FˆN4dS=x2x1t2t1[r4(t)s4(t)]u(x,y2,t)v(x,y2,t)dtdx. (2.19)

    On the other hand, boundary conditions (2.1) imply that (2.14) and (2.15) reduce to

    S5FˆN5dS=y2y1x2x1v(x,y,t1)ut(x,y,t1)dxdy (2.20)

    and

    S6FˆN6dS=y2y1x2x1v(x,y,t2)ut(x,y,t2)dxdy. (2.21)

    Since u and v are positive solutions on ˉΩ, conditions (2.3) of the theorem imply that (2.9) turns out to be

    SFˆNdSy2y1x2x1{v(x,y,t2)ut(x,y,t2)v(x,y,t1)ut(x,y,t1)}dxdy. (2.22)

    Since u(x,y,t1)=u(x,y,t2)=0 and u>0 on ˉI×ˉJ×K, we have that ut(x,y,t1)0 and ut(x,y,t2)0 for all (x,y)ˉI×ˉJ. This implies that the righthand side of (2.22) is nonpositive, and we have

    SFˆNdS0. (2.23)

    Finally, (2.5), (2.7) and inequality (2.23) imply that

    Ω[g(x,y,t)f(x,y,t)]uvdV0 (2.24)

    which contradicts with condition (2.2). This contradiction yields that v cannot be a positive solution of problem (1.7)-(1.9) on ˉΩ. The same proof can be repeated under the assumption that v<0 on ˉΩ. Therefore, v has a zero in ˉΩ. Theorem 2.1 has been proved.

    Remark 2.2. If inequalities (2.2) and (2.3) in Theorem 2.1 are replaced by the strict ones

    g(x,y,t)>f(x,y,t),(x,y,t)Ω (2.25)

    and

    sk(t)>rk(t),tˉK(k=1,2,3,4), (2.26)

    then it can be easily proved that v has a zero in interior of ˉΩ.

    Proposition 2.3. (Sturm comparison theorem) Let u>0 be a solution of problem (1.6)-(1.8) satisfying the boundary condition (2.1) on ˉI×ˉJ×K. If inequalities (2.25) and (2.26) hold, then every solution v of problem (1.7)-(1.9) has a zero in Ω.

    Remark 2.4. Inequalities (2.25) and (2.26) in Proposition 2.3 can be weakened and Proposition 2.3 can be commuted by the following conclusion.

    Proposition 2.5. (Sturm comparison theorem) Let u>0 be a solution of problem (1.6)-(1.8) satisfying the boundary condition (2.1) on ˉI×ˉJ×K, and assume that inequalities (2.2) and (2.3) hold.

    If either

    meas{(x,y,t)Ω:g(x,y,t)f(x,y,t)>0}>0 (2.27)

    or

    meas{tˉK:sk(t)rk(t)>0,k=1,2,3,4}>0, (2.28)

    then every solution v of problem (1.7)-(1.9) has a zero in Ω.

    The following oscillation criterion is immediate.

    Corollary 2.6. (Sturm oscillation theorem) If the inequalities

    g(x,y,t)f(x,y,t),(x,y,t)I×J×(t,) (2.29)

    and

    sk(t)rk(t),t[t,)(k=1,2,3,4) (2.30)

    hold for every tt0, then every solution of problem (1.7)-(1.9) is oscillatory whenever problem (1.6)-(1.8) is oscillatory.

    The results obtained for linear equations in the previous section can be extended to the nonlinear hyperbolic equations of the form

    uttΔu+F(u,x,y,t)=0,(x,y,t)Ω (3.1)

    and

    vttΔv+G(v,x,y,t)=0,(x,y,t)Ω (3.2)

    satisfying the initial conditions (1.8) and (1.9), respectively. The functions rk(t) and sk(t), k=1,2,3,4, are as previously defined, and F,G:R×ˉΩR are continuous functions satisfying

    μF(μ,x,y,t)p(t)μ2;(μ,x,y,t)R×ˉΩ

    and

    μG(μ,x,y,t)q(t)μ2;(μ,x,y,t)R×ˉΩ

    for which p,q:ˉKR are continuous functions.

    The second primary conclusion of this paper is as follows.

    Theorem 3.1. (Sturm comparison theorem) Let u>0 be a solution of problem (1.8)-(3.1) satisfying the boundary condition (2.1) on ˉI×ˉJ×K. If the inequalities

    q(t)p(t)andsk(t)rk(t)(k=1,2,3,4) (3.3)

    hold for tˉK, then every solution v of problem (1.9)-(3.2) has a zero in ˉΩ.

    The proof of Theorem 3.1 is based on the inequality

    [uvxvux]x+[uvyvuy]y+[vutuvt]t=[uG(v,x,y,t)vF(u,x,y,t)][q(t)p(t)]uv

    for uC(¯Ω,R+{0}), vC(ˉΩ,R+), and can be done following the same steps as those in Theorem 2.1. Therefore, it is left to the reader.

    Remark 3.2. If the inequalities given in (3.3) are replaced by the strict ones

    q(t)>p(t)andsk(t)>rk(t)(k=1,2,3,4), (3.4)

    then the comparison conclusion is as follows.

    Proposition 3.3. (Sturm comparison theorem) Let u>0 be a solution of problem (1.8)-(3.1) satisfying the boundary condition (2.1) ˉI×ˉJ×K. If the inequalities in (3.4) hold for tˉK, then every solution v of problem (1.9)-(3.2) has a zero in Ω.

    As mentioned in Remark 2.4, Proposition 3.3 can be alternated by the following result.

    Proposition 3.4. (Sturm comparison theorem) Let u>0 be a solution of problem (1.8)-(3.1) satisfying the boundary condition (2.1) on ˉI×ˉJ×K, and assume that the inequalities in (3.3) hold for tˉK.

    If either

    meas{tˉK:q(t)p(t)>0}>0

    or

    meas{tˉK:sk(t)rk(t)>0,k=1,2,3,4}>0,

    then every solution v of problem (1.9)-(3.2) has a zero in Ω.

    The following oscillation criteria is immediate.

    Corollary 3.5. (Sturm oscillation theorem) If the inequalities given in (3.4) hold for t[t,) for every tt0, then every solution of problem (1.9)-(3.2) is oscillatory whenever problem (1.8)-(3.1) is oscillatory.

    Consider the hyperbolic equation (1.6) with only the time dependent potential

    wttΔw+ˆf(t)w=0,(x,y,t)I×J×(ˆt,) (4.1)

    satisfying the initial conditions

    wx(xk,y,t)+(1)kαkw(xk,y,t)=0,(y,t)ˉJ×[ˆt,),wy(x,yk,t)+(1)kαk+2w(x,yk,t)=0,(x,t)ˉI×[ˆt,) (4.2)

    for k=1,2, where ˆt[t0,) and αk's are real constants.

    As Eq (4.1) allows a separation of variables, we set w(x,y,t)=H(x,y)T(t). Solving (4.1), we get the eigenvalue problem

    ΔH=λH,(x,y)I×J (4.3)

    with the initial conditions

    Hx(xk,y)+(1)kαkH(xk,y)=0,yˉJ,Hy(x,yk)+(1)kαk+2H(x,yk)=0,xˉI (4.4)

    for k=1,2, and

    T+ˆf(t)T=λT,t[ˆt,).

    Applying Corollary 2.6, we can derive an interesting oscillation criterion for a class of hyperbolic equations.

    Theorem 4.1. (Sturm oscillation theorem) Let w be a nontrivial solution of problem (4.1)-(4.2) and assume that equation

    T+[ˆf(t)λ0]T=0 (4.5)

    is oscillatory, where λ0 is the first eigenvalue of problem (4.3)-(4.4).

    If the inequalities

    g(x,y,t)ˆf(t),(x,y,t)I×J×[˜t,) (4.6)

    and

    sk(t)αk,t[˜t,)(k=1,2,3,4) (4.7)

    hold for every ˜tˆt, then every solution v of problem (1.7)-(1.9) has a zero in ˉI×ˉJ×[˜t,).

    For the elliptic case, analogous results of Theorem 4.1 can be found in a paper by Kreith [20].

    Remark 4.2. When the potential does not depend on time variable, the technique of Theorem 4.1 also can be applied to the equation

    νttΔν+Q(x,y)ν=0,(x,y,t)I×J×(ˆt,)

    under the analogous initial conditions with (4.2).

    The paper presented novel Sturmian comparison results for linear and nonlinear hyperbolic equations on a rectangular prism. The results for linear equations provided an extension to those obtained by Kreith in [18], which were founded within a rectangular region in the plane. The results were verified by considering a certain class of hyperbolic equations that were converted to an eigenvalue problem, which enabled us to draw a new and interesting oscillation criterion.

    It will be of great interest for the reader to obtain all the results given in this paper on the (n+1)-orthotope (hyperrectangle) for the hyperbolic equations of the form

    uttΔu+H(x,t)u=0,(x,t)ΓRn×(t0,)

    under the proper boundary conditions, where x=(x1,,xn) and Δ is the usual Laplace operator in Rn, i.e.,

    Δ=nj=122xj,

    and Γ is the hyperrectangle defined by

    Γ:=((x1)1,(x1)2)×((x2)1,(x2)2)××((xn)1,(xn)2)×(t1,t2)

    for (xj)1,(xj)2R are points on the xj-axis, and j=1,2,,n. The details are left for future consideration.

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

    J. Alzabut would like to thank Prince Sultan University and OSTİM Technical University for their endless support.

    The authors declare no conflicts of interest.



    [1] Z. Kayar, Sturm-Picone type theorems for second order nonlinear impulsive differential equations, AIP Conf. Proc., 1863 (2017), 140007. https://doi.org/10.1063/1.4992314 doi: 10.1063/1.4992314
    [2] O. Moaaz, G. E. Chatzarakis, T. Abdeljawad, C. Cesarano, A. Nabih, Amended oscillation criteria for second-order neutral differential equations with damping term, Adv. Differ. Equ., 2020 (2020), 1–12. https://doi.org/10.1186/s13662-020-03013-0 doi: 10.1186/s13662-020-03013-0
    [3] H. Ahmad, T. A. Khan, P. S. Stanimirovic, W. Shatanawi, T. Botmart, New approach on conventional solutions to nonlinear partial differential equations describing physical phenomena, Results Phys., 41 (2022), 105936. https://doi.org/10.1016/j.rinp.2022.105936 doi: 10.1016/j.rinp.2022.105936
    [4] A. Palanisamy, J. Alzabut, V. Muthulakshmi, S. S. Santra, K. Nonlaopon, Oscillation results for a fractional partial differential system with damping and forcing terms, AIMS Math., 8 (2023), 4261–4279. https://doi.org/10.3934/math.2023212 doi: 10.3934/math.2023212
    [5] P. Hartman, A. Wintner, On a comparison theorem for self-adjoint partial differential equations of elliptic type, Proc. Amer. Math. Soc., 6 (1955), 862–865.
    [6] S. Shimoda, Comparison theorems and various related principles in the theory of second order partial differential equations of elliptic type, Sūgaku, 9 (1957), 153–166.
    [7] T. Kusano, N. Yoshida, Comparison and nonoscillation theorems for fourth order elliptic systems, Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend., 59 (1975), 328–337.
    [8] N. Yoshida, Nonoscillation and comparison theorems for a class of higher order elliptic systems, Japan. J. Math., 2 (1976), 419–434. https://doi.org/10.4099/math1924.2.419 doi: 10.4099/math1924.2.419
    [9] T. Kusano, J. Jaroš, N. Yoshida, A Picone-type identity and Sturmian comparison and oscillation theorems for a class of half-linear partial differential equations of second order, Nonlinear Anal., 40 (2000), 381–395. https://doi.org/10.1016/S0362-546X(00)85023-3 doi: 10.1016/S0362-546X(00)85023-3
    [10] N. Yoshida, Sturmian comparison and oscillation theorems for a class of half-linear elliptic equations, Nonlinear Anal., 71 (2009), e1354–e1359. https://doi.org/10.1016/j.na.2009.01.141 doi: 10.1016/j.na.2009.01.141
    [11] N. Yoshida, Sturmian comparison and oscillation theorems for quasilinear elliptic equations with mixed nonlinearities via Picone-type inequality, Toyama Math. J., 33 (2010), 21–41.
    [12] N. Yoshida, Picone identities for half-linear elliptic operators with p(x)-Laplacians and applications to Sturmian comparison theory, Nonlinear Anal., 74 (2011), 5631–5642. https://doi.org/10.1016/j.na.2011.05.048 doi: 10.1016/j.na.2011.05.048
    [13] N. Yoshida, Picone-type inequality and Sturmian comparison theorems for quasilinear elliptic operators with p(x)-Laplacians, Electron. J. Differ. Equ., 2012 (2012), 1–9.
    [14] J. Jaroš, Picone-type identity and comparison results for a class of partial differential equations of order 4m, Opuscula Math., 33 (2013), 701–711. https://doi.org/10.7494/OpMath.2013.33.4.701 doi: 10.7494/OpMath.2013.33.4.701
    [15] J. B. Diaz, J. R. McLaughlin, Sturm comparison theorems for ordinary and partial differential equations, Bull. Amer. Math. Soc., 75 (1969), 335–339.
    [16] J. B. Diaz, J. R. McLaughlin, Sturm separation and comparison theorems for ordinary and partial differential equations, Atti Accad. Naz. Lincei Mem. Cl. Sci. Fis. Mat. Natur. Sez. I, 9 (1969), 135–194.
    [17] N. Yoshida, Oscillation theory of partial differential equations, World Scientific, 2008.
    [18] K. Kreith, Sturmian theorems for hyperbolic equations, Proc. Amer. Math. Soc., 22 (1969), 277–281.
    [19] K. Kreith, Oscillation theory, Springer, 1973. https://doi.org/10.1007/BFb0067537
    [20] K. Kreith, Oscillation theorems for elliptic equations, Proc. Amer. Math. Soc., 15 (1964), 341–344.
  • This article has been cited by:

    1. Abdullah Ozbekler, Kübra Uslu İşler, A Sturm comparison criterion for impulsive hyperbolic equations on a rectangular prism, 2024, 74, 1303-5991, 79, 10.31801/cfsuasmas.1454695
  • Reader Comments
  • © 2024 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(1052) PDF downloads(51) Cited by(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog