Citation: Selene Tomassini, Annachiara Strazza, Agnese Sbrollini, Ilaria Marcantoni, Micaela Morettini, Sandro Fioretti, Laura Burattini. Wavelet filtering of fetal phonocardiography: A comparative analysis[J]. Mathematical Biosciences and Engineering, 2019, 16(5): 6034-6046. doi: 10.3934/mbe.2019302
[1] | Aymard Christbert Nimi, Daniel Moukoko . Global attractor and exponential attractor for a Parabolic system of Cahn-Hilliard with a proliferation term. AIMS Mathematics, 2020, 5(2): 1383-1399. doi: 10.3934/math.2020095 |
[2] | Dieunel DOR . On the modified of the one-dimensional Cahn-Hilliard equation with a source term. AIMS Mathematics, 2022, 7(8): 14672-14695. doi: 10.3934/math.2022807 |
[3] | Joseph L. Shomberg . Well-posedness and global attractors for a non-isothermal viscous relaxationof nonlocal Cahn-Hilliard equations. AIMS Mathematics, 2016, 1(2): 102-136. doi: 10.3934/Math.2016.2.102 |
[4] | Saulo Orizaga, Maurice Fabien, Michael Millard . Efficient numerical approaches with accelerated graphics processing unit (GPU) computations for Poisson problems and Cahn-Hilliard equations. AIMS Mathematics, 2024, 9(10): 27471-27496. doi: 10.3934/math.20241334 |
[5] | Harald Garcke, Kei Fong Lam . Global weak solutions and asymptotic limits of a Cahn–Hilliard–Darcy system modelling tumour growth. AIMS Mathematics, 2016, 1(3): 318-360. doi: 10.3934/Math.2016.3.318 |
[6] | Shihe Xu, Zuxing Xuan, Fangwei Zhang . Analysis of a free boundary problem for vascularized tumor growth with time delays and almost periodic nutrient supply. AIMS Mathematics, 2024, 9(5): 13291-13312. doi: 10.3934/math.2024648 |
[7] | Hyun Geun Lee . A mass conservative and energy stable scheme for the conservative Allen–Cahn type Ohta–Kawasaki model for diblock copolymers. AIMS Mathematics, 2025, 10(3): 6719-6731. doi: 10.3934/math.2025307 |
[8] | Jean De Dieu Mangoubi, Mayeul Evrard Isseret Goyaud, Daniel Moukoko . Pullback attractor for a nonautonomous parabolic Cahn-Hilliard phase-field system. AIMS Mathematics, 2023, 8(9): 22037-22066. doi: 10.3934/math.20231123 |
[9] | Alain Miranville . The Cahn–Hilliard equation and some of its variants. AIMS Mathematics, 2017, 2(3): 479-544. doi: 10.3934/Math.2017.2.479 |
[10] | Martin Stoll, Hamdullah Yücel . Symmetric interior penalty Galerkin method for fractional-in-space phase-field equations. AIMS Mathematics, 2018, 3(1): 66-95. doi: 10.3934/Math.2018.1.66 |
In this paper, we consider the following Cahn-Hilliard systems with dynamic boundary conditions and time periodic conditions, say (P), which consists of the following equations:
∂u∂t−Δμ=0in Q:=Ω×(0,T), | (1.1) |
μ=−κ1Δu+ξ+π(u)−f,ξ∈β(u)in Q, | (1.2) |
uΓ=u|Γ,μΓ=μ|Γon Σ:=Γ×(0,T), | (1.3) |
∂uΓ∂t+∂νμ−ΔΓμΓ=0on Σ, | (1.4) |
μΓ=κ1∂νu−κ2ΔΓuΓ+ξΓ+πΓ(uΓ)−fΓ,ξΓ∈βΓ(uΓ)on Σ, | (1.5) |
u(0)=u(T)in Ω,uΓ(0)=uΓ(T)on Γ | (1.6) |
where 0<T<+∞, Ω is a bounded domain of Rd (d=2,3) with smooth boundary Γ:=∂Ω, κ1,κ2 are positive constants, ∂ν is the outward normal derivative on Γ, u|Γ,μ|Γ stand for the trace of u and μ to Γ, respectively, Δ is the Laplacian, ΔΓ is the Laplace-Beltrami operator (see, e.g., [21]), and f:Q→R, fΓ:Σ→R are given data. Moreover, β,βΓ:R→2R are maximal monotone operators and π,πΓ:R→R are Lipschitz perturbations.
The Cahn-Hilliard equation [8] is a description of mathematical model for phase separation, e.g., the phenomenon of separating into two phases from homogeneous composition, the so-called spinodal decomposition. In (1.1)–(1.2), u is the order parameter and μ is the chemical potential. Moreover, it is well known that the Cahn-Hilliard equation is characterized by the nonlinear term β+π. It plays an important role as the derivative of the double-well potential W. The well-known example of nonlinear terms is W(r)=(1/4)(r2−1)2, namely W′(r)=r3−r for r∈R, this is called the prototype double well potential. Other examples are stated later. As the abstract mathematical result, Kenmochi, Niezgódka and Pawłow study the Cahn-Hilliard equation with constraint by subdifferential operator approach [24] (see also [25]). Essentially we apply the same method in this paper.
In terms of (1.3)–(1.5), we consider the dynamic boundary condition as being uΓ,μΓ unknown functions on the boundary. The dynamic boundary condition is treated in recent years, for example, for the Stefan problem [1,2,14], wider the degenerate parabolic equation [3,15,16] and the Cahn-Hilliard equation [11,12,17,18,19,20,22,29]. To the best our knowledge, the type of dynamic boundary conditions on the Cahn-Hilliard equation like (P) is formulated in [17,20]. Recently, the well-posedness with singular potentials is discussed in [11]; the maximal Lp regularity in bounded domains is treated in [22]; the related new model is also introduced in [29]. Based on the result [11], we also used the property of dynamic boundary conditions, more precisely, we set up the function space which satisfies that the total mass is equal to 0. At the sight of (P), we consider the same type of equations (1.1)–(1.2) on the boundary. In other words, (1.1)–(1.5) is a transmission problem connecting Ω and Γ. The nonlinear term βΓ+πΓ on boundary is also the derivative of the double-well potential WΓ, that is, we treat different nonlinear terms W′ and W′Γ in Ω and on Γ, respectively. In this case, it is necessary to assume some compatibility condition (see, e.g., [9,11]), stated (A4).
Focusing on (1.6), the study of time periodic problems of the Cahn-Hilliard equation is treated in [26,27,28,31]. In particular, Wang and Zheng discuss the existence of time periodic solutions of the Cahn-Hilliard equation with the Neumann boundary condition [31]. The authors employ the method of [4]. Note that the authors impose two assumptions for a maximal monotone graph, specifically, a restriction of effective domains and the following growth condition for the maximal monotone graph β:
ˆβ(r)≥cr2for all r∈R, |
for some positive constant c. However, the above assumption is too restrictive for some physical applications. In this paper, we follow the method of [31] and apply the abstract theory of evolution equations by using the viscosity approach and the Schauder fixed point theorem in the level of approximate problems. Moreover, by virtue of the viscosity approach, we also can apply the abstract result [4]. Note that, the growth condition is not need to solve the Cahn-Hilliard equation (see, e.g., [11]), therefore, setting the appropriate convex functional and using the Poincaré-Wirtinger inequality, we can relax the growth condition for the time periodic problem. Thanks to this, we can choose various kinds of nonlinear diffusion terms β+π and βΓ+πΓ. On the other hand, a restriction of effective domains is essential to show the existence of solutions of (P).
The present paper proceeds as follows.
In Section 2, a main theorem and a definition of solutions are stated. At first, we prepare the notation used in this paper and set appropriate function spaces. Next, we introduce the definition of periodic solutions of (P) and the main theorems are given there. Also, we give examples of double-well potentials.
In Section 3, in order to pass to the limit, we set convex functionals and consider approximate problems. Next, we obtain the solution of (P)ε by using the Schauder fixed point theorem. Finally, we deduce uniform estimates for the solution of (P)ε.
In Section 4, we prove the existence of periodic solutions by passing to the limit ε→0.
A detailed index of sections and subsections follows.
1. Introduction
2. Main results
2.1. Notation
2.2. Definition of the solution and main theorem
3. Approximate problems and uniform estimates
3.1. Abstract formulation
3.2. Approximate problems for (P)
3.3. Uniform estimates
4. Proof of convergence theorem
We introduce the spaces H:=L2(Ω), HΓ:=L2(Γ), V:=H1(Ω), VΓ:=H1(Γ) with standard norms |⋅|H, |⋅|HΓ, |⋅|V, |⋅|VΓ and inner products (⋅,⋅)H, (⋅,⋅)HΓ, (⋅,⋅)V, (⋅,⋅)VΓ, respectively. Moreover, we set H:=H×HΓ and
V:={z:=(z,zΓ)∈V×VΓ : z|Γ=zΓ a.e. on Γ}. |
H and V are then Hilbert spaces with inner products
(u,z)H:=(u,z)H+(uΓ,zΓ)HΓfor all u:=(u,uΓ),z:=(z,zΓ)∈H,(u,z)V:=(u,z)V+(uΓ,zΓ)VΓfor all u:=(u,uΓ),z:=(z,zΓ)∈V. |
Note that z∈V implies that the second component zΓ of z is equal to the trace of the first component z of z on Γ, and z∈H implies that z∈H and zΓ∈HΓ are independent. Throughout this paper, we use the bold letter u to represent the pair corresponding to the letter; i.e., u:=(u,uΓ).
Let m:H→R be the mean function defined by
m(z):=1|Ω|+|Γ|{∫Ωzdx+∫ΓzΓdΓ}for all z∈H, |
where |Ω|:=∫Ω1dx,|Γ|:=∫Γ1dΓ. Then, we define H0:={z∈H:m(z)=0}, V0:=V∩H0. Moreover, V∗,V∗0 denote the dual spaces of V,V0, respectively; the duality pairing between V∗0 and V0 is denoted by ⟨⋅,⋅⟩V∗0,V0. We define the norm of H0 by |z|H0:=|z|H for all z∈H0 and the bilinear form a(⋅,⋅):V×V→R by
a(u,z):=κ1∫Ω∇u⋅∇zdx+κ2∫Γ∇ΓuΓ⋅∇ΓzΓdΓfor all u,z∈V. |
Then, for all z∈V0, |z|V0:=√a(z,z) becomes a norm of V0. Also, we let F:V0→V∗0 be the duality mapping, namely,
⟨Fz,˜z⟩V∗0,V0:=a(z,˜z)for all z,˜z∈V0. |
We note that the following the Poincaré-Wirtinger inequality holds: There exists a positive constant cP such that
|z|2V≤cP|z|2V0for all z∈V0 | (2.1) |
(see [11,Lemma A]). Moreover, we define the inner product of V∗0 by
(z∗,˜z∗)V∗0:=⟨z∗,F−1˜z∗⟩V∗0,V0for all z∗,˜z∗∈V∗0. |
Also, we define the projection P:H→H0 by
Pz:=z−m(z)1for all z∈H, |
where 1:=(1,1). Then, since P is a linear bounded operator, the following property holds: Let {zn}n∈N be a sequence in H such that zn→z weakly in H for some z, then we infer that
Pzn→Pzweakly in H0as n→∞. | (2.2) |
Then, we have V0↪↪H0↪↪V∗0, where "↪↪" stands for compact embedding (see [11,Lemmas A and B]).
In this subsection, we define our periodic solutions for (P) and then we state the main theorem.
Firstly, from (1.1) and (1.4), the following total mass conservation holds:
m(u(t))=m(u(0))for all t∈[0,T]. |
Therefore, for any given m0∈intD(βΓ), we define the periodic solution satisfying the total mass conservation m(u(t))=m0 for all t∈[0,T]. We use the following notation: the variable v:=u−m01; the datum f:=(f,fΓ); the function π(z):=(π(z),πΓ(zΓ)) for z∈H. Moreover, we set the space W:=H2(Ω)×H2(Γ).
Definition 2.1. For any given m0∈intD(βΓ), the triplet (v,μ,ξ) is called the periodic solution of (P) if
v∈H1(0,T;V∗0)∩L∞(0,T;V0)∩L2(0,T;W),μ∈L2(0,T;V),ξ=(ξ,ξΓ)∈L2(0,T;H), |
and they satisfy
⟨v′(t),z⟩V∗0,V0+a(μ(t),z)=0for all z∈V0, | (2.3) |
(μ(t),z)H=a(v(t),z)+(ξ(t)−m(ξ(t))1+π(v(t)+m01)−f(t),z)Hfor all z∈V | (2.4) |
for a.a. t∈(0,T), and
ξ∈β(v+m0)a.e. in Q,ξΓ∈βΓ(vΓ+m0)a.e. on Σ |
with
v(0)=v(T)in H0. | (2.5) |
Remark 2.1. We can see that μ:=(μ,μΓ) satisfies
μ=−κ1Δu+ξ−m(ξ)+π(u)−fa.e. in Q,μΓ=κ1∂νu−κ2ΔΓuΓ+ξΓ−m(ξ)+πΓ(uΓ)−fΓa.e. on Σ, |
where u=v+m0 and uΓ=vΓ+m0, because of the regularity v∈L2(0,T;W).
Remark 2.2. In (2.4), this is different from the following definition of [11,Definition 2.1]:
(μ(t),z)H=a(v(t),z)+(ξ(t)+π(v(t)+m01)−f(t),z)Hfor all z∈V | (2.6) |
for a.a. t∈(0,T). However, by setting ˜μ:=μ+m(ξ)1, ˜μ satisfies ˜μ∈L2(0,T;V) and (2.6). Hence, in other words, we can employ (2.6) as definition of (P) replaced by (2.4).
We assume that
(A1) f∈L2(0,T;V) and f(t)=f(t+T) for a.a. t∈[0,T];
(A2) π,πΓ:R→R are locally Lipschitz continuous functions;
(A3) β,βΓ:R→2R are maximal monotone operators, which is the subdifferential
β=∂Rˆβ,βΓ=∂RˆβΓ |
of some proper lower semicontinuous convex functions ˆβ,ˆβΓ:R→[0,+∞] satisfying ˆβ(0)=ˆβΓ(0)=0 with domains D(β) and D(βΓ), respectively;
(A4) D(βΓ)⊆D(β) and there exist positive constants ρ and c0 such that
|β∘(r)|≤ρ|β∘Γ(r)|+c0for all r∈D(βΓ); | (2.7) |
(A5) D(β),D(βΓ) are bounded domains with non-empty interior, i.e., ¯D(β)=[σ∗,σ∗] and ¯D(βΓ)=[σΓ∗,σ∗Γ] for some constants σ∗,σ∗,σΓ∗ and σ∗Γ with −∞<σ∗≤σΓ∗<σ∗Γ≤σ∗<∞.
The minimal section β∘ of β is defined by β∘(r):={q∈β(r):|q|=mins∈β(r)|s|} for r∈R. Also, β∘Γ is defined similarly. In particular, (A3) yields 0∈β(0). The assumption (A5) is not imposed in [11]. However, it is essential to obtain uniform estimates in Section 3. This is a difficulty of time periodic problems. Also, the assumption of compatibility of β and βΓ (A4) is the same as in [9,11].
Now, we give some examples of the nonlinear perturbation terms which satisfies the above assumptions:
● β(r)=βΓ(r)=(α1/2)ln((1+r)/(1−r)), π(r)=πΓ(r)=−α2r for all r∈D(β)=D(βΓ)=(−1,1) and 0<α1<α2 for the logarithmic double well potential W(r)=WΓ(r)=(α1/2){(1−r)ln((1−r)/2)+(1+r)ln((1+s)/2)}+(α2/2)(1−r2). The condition α1<α2 ensures that W,WΓ have double-well forms (see, e.g., [10]).
● β(r)=βΓ(r)=∂I[−1,1](r), π(r)=πΓ(r)=−r for all r∈D(β)=D(βΓ)=[−1,1] for the singular potential W(r)=WΓ(r)=I[−1,1](r)−r2/2, where ∂I[−1,1] is the subdifferential of the indicator function I[−1,1] of the interval [−1,1] (namely, I[−1,1](r)=0 if r∈[−1,1] and I[−1,1](r)=+∞ otherwise).
● β(r)=βΓ(r)=∂I[−1,1](r)+r3, π(r)=πΓ(r)=−r for all r∈D(β)=D(βΓ)=[−1,1] for the modified prototype double well potential W(r)=WΓ(r)=I[−1,1](r)+(1/4)(r2−1)2−r2/2.
Our main theorem is given now.
Theorem 2.1. Under the assumptions (A1)– (A5), for any given m0∈intD(βΓ), there exist at least one periodic solution of (P) such that m(u(t))=m0 for all t∈[0,T].
Remark 2.3. We note that periodic solutions of (P) is not uniquely determined. It is due to the usage of the Gronwall inequality. Indeed, in [11,Theorem 2.1], the continuous dependent on the data is proved, that is, the uniqueness of the solution to a Cauchy problem is obtained. However, in this periodic problem (P), even if we use the same method, the continuous dependent can not be obtained because of Lipschitz perturbations π and πΓ. Without the perturbations, we can obtain the uniqueness (see Section 3).
In this section, we consider approximate problems and obtain uniform estimates to show the existence of periodic solutions of (P). Hereafter, we fix a given constant m0∈intD(βΓ).
In order to prove the main theorem, we apply the abstract theory of evolution equations. To do so, we define a proper lower semicontinuous convex functional φ:H0→[0,+∞] by
φ(z):={κ12∫Ω|∇z|2dx+κ22∫Γ|∇ΓzΓ|2dΓ +∫Ωˆβ(z+m0)dx+∫ΓˆβΓ(zΓ+m0)dΓ ifz∈V0 with ˆβ(z+m0)∈L1(Ω), ˆβΓ(zΓ+m0)∈L1(Γ),+∞ otherwise. |
Next, for each ε∈(0,1], we define a proper lower semicontinuous convex functional φε:H0→[0,+∞] by
φε(z):={κ12∫Ω|∇z|2dx+κ22∫Γ|∇ΓzΓ|2dΓ +∫Ωˆβε(z+m0)dx+∫ΓˆβΓ,ε(zΓ+m0)dΓifz∈V0,+∞otherwise, |
where ˆβε,ˆβΓ,ε are Moreau-Yosida regularizations of ˆβ,ˆβΓ defined by
ˆβε(r):=infs∈R{12ε|r−s|2+ˆβ(s)}=12ε|r−Jε(r)|2+ˆβ(Jε(r)),ˆβΓ,ε(r):=infs∈R{12ερ|r−s|2+ˆβΓ(s)}=12ερ|r−JΓ,ε(r)|2+ˆβΓ(JΓ,ε(r)), |
for all r∈R, where ρ is a constant as in (2.7) and Jε,JΓ,ε:R→R are resolvent operators given by
Jε(r):=(I+εβ)−1(r),JΓ,ε(r):=(I+ερβΓ)−1(r) |
for all r∈R. Moreover, βε,βΓ,ε:R→R are Yosida approximations for maximal monotone operators β,βΓ, respectively:
βε(r):=1ε(r−Jε(r)),βΓ,ε(r):=1ερ(r−JΓ,ε(r)) |
for all r∈R. Then, we easily see that βε(0)=βΓ,ε(0)=0 holds from the definition of the subdifferential. It is well known that βε,βΓ,ε are Lipschitz continuous with Lipschitz constants 1/ε,1/(ερ), respectively. Here, we have following properties:
0≤ˆβε(r)≤ˆβ(r),0≤ˆβΓ,ε(r)≤ˆβΓ(r)for all r∈R. |
Hence, 0≤φε(z)≤φ(z) holds for all z∈H0. These properties of Yosida approximation and Moreau-Yosida regularizations are as in [5,6,23]. Moreover, thanks to [9,Lemma 4.4], we have
|βε(r)|≤ρ|βΓ,ε(r)|+c0for all r∈R | (3.1) |
with the same constants ρ and c0 as in (2.7).
Now, for each ε∈(0,1], we also define two proper lower semicontinuous convex functionals ˜φ,ψε:H0→[0,+∞] by
˜φ(z):={κ12∫Ω|∇z|2dx+κ22∫Γ|∇ΓzΓ|2dΓifz∈V0,+∞otherwise |
and
ψε(z):=∫Ωˆβε(z+m0)dx+∫ΓˆβΓ,ε(zΓ+m0)dΓ |
for all z∈H0, respectively. Then, from [11,Lemma C], the subdifferential A:=∂H0˜φ on H0 is characterized by
Az=(−κ1Δz,κ1∂νz−κ2ΔΓzΓ)with z=(z,zΓ)∈D(A)=W∩V0. |
Moreover, the representation of the subdifferential ∂H0ψε is given by
∂H0ψε(z)=Pβε(z+m01)for all z∈H0, |
where βε(z+m01):=(βε(z+m0),βΓ,ε(zΓ+m0)) for z=(z,zΓ)∈H0. This is proved by the same way as in [16,Lemma 3.2]. Noting that it holds that D(∂H0ψε)=H0 and A is a maximal monotone operator; indeed it follows from the abstract monotonicity methods (see, e.g., [5,Sect. 2.1]) that A+∂H0ψε is also a maximal monotone operator. Moreover, by a simple calculation, we deduce that (A+∂H0ψε)⊂∂H0φε. Hence,
∂H0φε(z)=(A+∂H0ψε)(z) | (3.2) |
for any z∈H0 (see, e.g., [13]).
Now, we consider the following approximate problem, say (P)ε: for each ε∈(0,1] find vε:=(vε,vΓ,ε) satisfying
εv′ε(t)+F−1v′ε(t)+∂H0φε(vε(t)) +P(˜π(vε(t)+m01))=Pf(t)in H0for a.a. t∈(0,T), | (3.3) |
vε(0)=vε(T)in H0. | (3.4) |
where, for all z∈H, ˜π(z):=(˜π(z),˜πΓ(zΓ)) is a cut-off function of π,πΓ given by
˜π(r):={0if r≤σ∗−1,π(σ∗)(r−σ∗+1)if σ∗−1≤r≤σ∗,π(r)if σ∗≤r≤σ∗,−π(σ∗)(r−σ∗−1)if σ∗≤r≤σ∗+1,0if r≥σ∗+1 | (3.5) |
and
˜πΓ(r):={0if r≤σΓ∗−1,πΓ(σΓ∗)(r−σ∗+1)if σΓ∗−1≤r≤σΓ∗,πΓ(r)if σΓ∗≤r≤σ∗Γ,−πΓ(σ∗Γ)(r−σ∗−1)if σ∗Γ≤r≤σ∗Γ+1,0if r≥σ∗Γ+1 | (3.6) |
for all r∈R, respectively. We establish the above cut-off function by referring to [31].
From now, we show the next proposition of the existence of the periodic solution for (P)ε.
Proposition 3.1. Under the assumptions (A1)–(A5), for each ε∈(0,1], there exist at least one function
vε∈H1(0,T;H0)∩L∞(0,T;V0)∩L2(0,T;W) |
such that vε satisfies (3.3) and (3.4).
The proof of Proposition 3.1 is given later. In order to show the Proposition 3.1, we use the method in [31], that is, we employ the fixed point argument. To do so, we consider the following problem: for each ε∈(0,1] and g∈L2(0,T;V0),
(F−1+εI)v′ε(t)+∂φε(vε(t))=g(t)in H0for a.a. t∈(0,T), | (3.7) |
vε(0)=vε(T)in H0. | (3.8) |
Now, we can apply the abstract theory of doubly nonlinear evolution equations respect to the time periodic problem [4] for (3.7), (3.8) because the operator εI+F−1 and ∂φε are coercive in H0. It is an important assumption to apply Theorem 2.2 in [4]. Moreover, the function vε satisfying (3.7) and (3.8) is uniquely determined. Indeed, let v1ε,v2ε be periodic solutions of the problem (3.7) and (3.8). Then, at the time t∈(0,T), taking the difference (3.7) for v1ε and v2ε, respectively, we have
ε(v1ε(t)−v2ε(t))+F−1(v1ε(t)−v2ε(t))+∂φε(v1ε(t))−∂φε(v2ε(t))=0in H0 | (3.9) |
for a.a. t∈(0,T). Now, we test (3.9) at time t∈(0,T) by v1ε(t)−v2ε(t). Then, we deduce that
12ddt(ε|v1ε(t)−v2ε(t)|2H0+|v1ε(t)−v2ε(t)|2V∗0)+12|v1ε(t)−v2ε(t)|2V0≤0 |
for a.a. t∈(0,T), because of (3.2) and the monotonicity of β,βΓ. Therefore, by integrating it over [0,T] with respect to t, it follows from (2.1) that
∫T0|v1ε(t)−v2ε(t)|2Vdt≤0. |
It implies that the function vε satisfying (3.7) and (3.8) is unique.
Hence, we obtain the next proposition.
Proposition 3.2. For each ε∈(0,1] and g∈L2(0,T;V0), there exists a unique function vε such that (3.7) and (3.8) are satisfied.
We apply the Schauder fixed point theorem to prove Proposition 3.1. To this aim, we set
Y1:={ˉvε∈H1(0,T;H0)∩L∞(0,T;V0):ˉvε(0)=ˉvε(T)}. |
Firstly, for each ˉvε∈Y1, we consider the following problem, say (Pε;ˉvε):
εv′ε(s)+F−1v′ε(s)+∂φε(vε(s))+P(˜π(ˉvε(s)+m01))=Pf(s)in H0 | (3.10) |
for a.a. s∈(0,T), with
vε(0)=vε(T)in H0. |
Next, we obtain estimates of the solution of (Pε;ˉvε) to apply the Schauder fixed point theorem. Note that we can allow the dependent of ε∈(0,1] for estimates of Lemma 3.1 because we use the Schauder fixed point theorem in the level of approximation.
Lemma 3.1. Let vε be the solution of problem (Pε;ˉvε). Then, there exist positive constants C1ε,C2,C3ε such that
ε∫T0|v′ε(s)|2H0ds+∫T0|v′ε(s)|2V∗0ds≤C1ε, | (3.11) |
∫T0|vε(s)|2V0ds+∫T0∫Ωˆβε(vε(s)+m0)dxds+∫T0∫ΓˆβΓ,ε(vΓ,ε(s)+m0)ds≤C2 | (3.12) |
and
12|vε(t)|2V0+∫Ωˆβε(vε(t)+m0)dx+∫ΓˆβΓ,ε(vΓ,ε(t)+m0)dΓ≤C3ε | (3.13) |
for all t∈[0,T].
Proof. At first, for each ˉvε∈Y1, there exists a positive constant M, depending only on σ∗,σΓ∗,σ∗ and σ∗Γ, such that
|˜π(ˉvε(t)+m01)|2H0≤Mfor all t∈[0,T]. | (3.14) |
Now, testing (3.10) at time s∈(0,T) by v′ε(s) and using the Young inequality, we infer that
ε|v′ε(s)|2H0+|v′ε(s)|2V∗0+ddsφε(vε(s))=(Pf(s)−P(˜π(ˉvε(s)+m01)),v′ε(s))H0≤12|f(s)|2V+12|v′ε(s)|2V∗0+M2ε+ε2|v′ε(s)|2H0 |
for a.a. s∈(0,T). Therefore, we have that
ε|v′ε(s)|2H0+|v′ε(s)|2V∗0+2ddsφε(vε(s))≤|f(s)|2V+Mε | (3.15) |
for a.a. s∈(0,T). Then, integrating it over (0,T) with respect to s and using the periodic property, we see that
ε∫T0|v′ε(s)|2H0ds+∫T0|v′ε(s)|2V∗0ds≤∫T0|f(s)|2Vds+MTε, |
which implies the first estimate (3.11).
Next, testing (3.10) at time s∈(0,T) by vε(s) and from (3.1), we deduce that
12dds|vε(s)|2V∗0+ε2dds|vε(s)|2H0+φε(vε(s))≤(Pf(s)−P(˜π(ˉvε(s)+m01)),vε(s))H0+φε(0)≤2cP|f(s)|2H0+14cP|vε(s)|2H0+2cPM+φ(0)≤2cP|f(s)|2H0+14|vε(s)|2V0+2cPM+φ(0)≤2cP|f(s)|2H0+12φε(vε(s))+2cPM+φ(0) |
for a.a. s∈(0,T), thanks to the definition of the subdifferential. From the definition of φε, it follows that
12dds|vε(s)|2V∗0+ε2dds|vε(s)|2H0+14|vε(s)|2V0+12∫Ωˆβε(vε(s)+m0)dx+12∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓ≤2cP|f(s)|2H0+2cPM+∫Ωˆβε(m0)dx+∫ΓˆβΓ,ε(m0)dΓ |
for a.a. s∈(0,T). Integrating it over (0,T) and using the periodic property, we see that
12∫T0|vε(s)|2V0ds+∫T0∫Ωˆβε(vε(s)+m0)dxds+∫T0∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓds≤4cP|f|2L2(0,T;H0)+4cPTM+T|Ω||ˆβ(m0)|+T|Γ||ˆβΓ(m0)|. |
Hence, there exist a positive constant C2 such that the second estimate (3.12) holds.
Next, for each s,t∈[0,T] such that s≤t, we integrate (3.15) over [s,t] with respect to s. Then, by neglecting the first two positive terms, we have
φε(vε(t))≤φε(vε(s))+12∫T0|f(s)|2Vds+MT2ε |
for all s,t∈[0,T], namely,
12|vε(t)|2V0+∫Ωˆβε(vε(t)+m0)dx+∫ΓˆβΓ,ε(vΓ,ε(t)+m0)dΓ≤12|vε(s)|2V0+∫Ωˆβε(vε(s)+m0)dx+∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓ+12∫T0|f(s)|2Vds+MT2ε | (3.16) |
for all s,t∈[0,T]. Now, integrating it over (0,t) with respect to s, we deduce that
t2|vε(t)|2V0+t∫Ωˆβε(vε(t)+m0)dx+t∫ΓˆβΓ,ε(vΓ,ε(t)+m0)dΓ≤12∫T0|vε(s)|2V0ds+∫T0∫Ωˆβε(vε(s)+m0)dxds+∫T0∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓds+T2∫T0|f(s)|2Vds+MT22ε | (3.17) |
for all t∈[0,T]. In particular, putting t:=T and dividing (3.17) by T, it follows that
12|vε(T)|2V0+∫Ωˆβε(vε(T)+m0)dx+∫ΓˆβΓ,ε(vΓ,ε(T)+m0)dΓ≤12T∫T0|vε(s)|2V0ds+1T∫T0∫Ωˆβε(vε(s)+m0)dxds+1T∫T0∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓds+12∫T0|f(s)|2Vds+MT2ε. | (3.18) |
Hence, combining the second estimate (3.12) and (3.18), we see that
12|vε(T)|2V0+∫Ωˆβε(vε(T)+m0)dx+∫ΓˆβΓ,ε(vΓ,ε(T)+m0)dΓ≤C2T+12∫T0|f(s)|2Vds+MT2ε. |
Moreover, from the periodic property, we infer that
12|vε(0)|2V0+∫Ωˆβε(vε(0)+m0)dx+∫ΓˆβΓ,ε(vΓ,ε(0)+m0)dΓ≤C2T+12∫T0|f(s)|2Vds+MT2ε. | (3.19) |
Now, let s be 0 in (3.16). Then, owing to (3.19), we deduce that
12|vε(t)|2V0+∫Ωˆβε(vε(t)+m0)dx+∫ΓˆβΓ,ε(vΓ,ε(t)+m0)dΓ≤C2T+|f|2L2(0,T;V)+MTε |
for all t∈[0,T]. Thus, there exists a positive constant C3ε such that the final estimate (3.13) holds.
In terms of (3.11), one key point to prove the estimate is exploiting (3.14). The estimate (3.14) is arised from the form of cut-off functions (3.5) and (3.6). The form of cut-off functions depends on the assumption (A5) essentially. However, considered the same estimate in [11,Lemma 4.1], it is not imposed the assumption. They use the Gronwall inequality to obtain the estimate because the initial value is given data. On the other hand, we can not obtain it even though we use the Gronwall inequality, because the initial value is not given. For this reason, it is necessary to impose (A5). This is a difficult point to solve this time periodic problem (P).
Now, we show the existence of solutions of the approximate problem (P)ε.
Proof of Proposition 3.1. We apply the Schauder fixed point theorem. To do so, we set
Y2:={ˉvε∈Y1:supt∈[0,T]|ˉvε(t)|2V0+ε|ˉvε|2H1(0,T;H0)≤Mε}, |
where Mε is a positive constant and be determined by Lemma 3.1. Then, the set Y2 is non-empty compact convex on C([0,T];H0). Now, from Proposition 3.2, for each ˉvε∈Y2, there exists a unique solution vε of (Pε;ˉvε). Moreover, from Lemma 3.1, it holds vε∈Y2. Here, we define the mapping S:Y2→Y2 such that, for each ˉvε∈Y2, corresponding ˉvε to the solution vε of (Pε;ˉvε). Then, the mapping S is continuous on Y2 with respect to topology of C([0,T];H0). Indeed, let {ˉvε,n}n∈N⊂Y2 be ˉvε,n→ˉvε in C([0,T];H0) and {vε,n}n∈N be the sequence of the solution of (Pε;ˉvε,n). From Lemma 3.1, there exist a subsequence {nk}k∈N, with nk→∞ as k→∞, and vε∈H1(0,T;H0)∩L∞(0,T;V0) such that
vε,nk→vεweakly star in H1(0,T;H0)∩L∞(0,T;V0). | (3.20) |
Hence, from (3.20) and the Ascoli-Arzelà theorem (see, e.g., [30]), there exists a subsequence (not relabeled) such that
vε,nk→vεin C([0,T];H0) | (3.21) |
as k→∞. Also, we have
v′ε,nk→v′εweakly in L2(0,T;H0) | (3.22) |
as k→∞. Because we have vε,nk(0)=vε,nk(T), it implies vε(0)=vε(T) in H0. Hereafter, we show that vε is the solution of (Pε;ˉvε). Since vε,nk is the solution of (Pε;ˉvε,nk), we see that
∫T0(Pf(s)−P(˜π(ˉvε,nk(s)+m01))−εv′ε,nk(s)−F−1v′ε,nk(s),η(s)−vε,nk(s))H0ds≤∫T0φε(η(s))ds−∫T0φε(vε,nk(s))ds | (3.23) |
for all η∈L2(0,T;H0), thanks to the definition of the subdifferential ∂φε. Moreover, it follows from ˉvε,nk→ˉvε in C([0,T];H0) that
P(˜π(ˉvε,nk+m01))→P(˜π(ˉvε+m01))in C([0,T];H0). | (3.24) |
Thus, on account of (3.20)–(3.24), taking the upper limit as k→∞ in (3.23) and using
lim infk→∞∫T0φε(vε,nk(s))ds≥∫T0φε(vε(s))ds, |
we infer that
∫T0(Pf(s)−P(˜π(ˉvε(s)+m01))−εv′ε(s)−F−1v′ε(s),η(s)−vε(s))H0ds≤∫T0φε(η(s))ds−∫T0φε(vε(s))ds |
for all η∈L2(0,T;H0). Hence, we see that the function vε is the solution of (Pε;ˉvε). As a result, it follows from the uniqueness of the solution of (Pε;ˉvε) that
S(ˉvε,nk)=vε,nk→vε=S(ˉvε)in C([0,T];H0) |
as k→∞. Therefore, the mapping S is continuous with respect to C([0,T];H0). Thus, from the Schauder fixed point theorem, there exists a fixed point on Y2, namely, the problem (P)ε admits a solution vε. Finally, from the fact that ∂φε(vε)∈L2(0,T;H0), which implies vε∈L2(0,T;W).
Now, we consider the chemical potential μ:=(μ,μΓ) by approximating. For each ε∈(0,1], we set the approximate sequence
με(s):=εv′ε(s)+∂φε(vε(s))+˜π(vε(s)+m01)−f(s) | (3.25) |
for a.a. s∈(0,T). From (3.2), we can rewrite (3.25) as
με(s)=εv′ε(s)+Avε(s)+Pβε(vε(s)+m01)+˜π(vε(s)+m01)−f(s) | (3.26) |
for a.a. s∈(0,T). Then, we rewrite (3.3) as
F−1v′ε(s)+με(s)−ωε(s)1=0in V |
for a.a. s∈(0,T), where
ωε(s):=m(˜π(vε(s)+m01)−f(s)) |
for a.a. s∈(0,T). Therefore, we have Pμε=με−ωε1∈L2(0,T;V0) and ωε∈L2(0,T). Then, it holds με∈L2(0,T;V) and
v′ε(s)+FPμε(s)=0in V∗0 | (3.27) |
for a.a. s∈(0,T).
In this subsection, we obtain uniform estimates independent of ε∈(0,1]. We refer to [31] to obtain uniform estimates.
Lemma 3.2. There exists a positive constant M1, independent of ε∈(0,1], such that
12∫T0|vε(s)|2V0ds+∫T0∫Ωˆβε(vε(s)+m0)dxds+∫T0∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓds≤M1. | (3.28) |
Proof. From (3.5), (3.6) and the assumption (A3), note that ˜π,˜πΓ are globally Lipschitz continuous on R. We denote Lipschitz constants of ˜π,˜πΓ by ˜L,˜LΓ, respectively. Moreover, we can take the primitive function ˆ˜π of ˜π satisfying
∫Ωˆ˜π(vε(s))dx≥0 |
for a.a. s∈(0,T). Analogously, we define ˆ˜πΓ. Now, we test (3.3) at time s∈(0,T) by vε(s) and use the Young inequality. Then, we deduce that
12dds|vε(s)|2V∗0+ε2dds|vε(s)|2H0+φε(vε(s))≤(Pf(s)−P(˜π(vε(s)+m01)),vε(s))H0+φε(0)≤cP|f(s)|2H+14cP|vε(s)|2H0+cPM+φ(0)≤14|vε(s)|2V0+cP|f(s)|2H+cPM+φ(0)≤12φε(vε(s))+cP|f(s)|2H+cPM+φ(0) |
for a.a. s∈(0,T). Namely, we have
12dds|vε(s)|2V∗0+ε2dds|vε(s)|2H0+14|vε(s)|2V0+12∫Ωˆβε(vε(s)+m0)dx+12∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓ≤cP|f(s)|2H+cPM+φ(0) |
for a.a. s∈(0,T). Integrating it over (0,T) and using the periodic property, we see that
12∫T0|vε(s)|2V0+∫T0∫Ωˆβε(vε(s)+m0)dxds+∫T0∫ΓˆβΓ,ε(vΓ,ε(s)+m0)dΓds≤2cP|f|2L2(0,T;H)+2cPTM+2Tφ(0). |
This yields that the estimate (3.28) holds.
Lemma 3.3. There exists a positive constant M_{2}, independent of \varepsilon \in (0, 1], such that
\varepsilon \int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{H}_{0}}^{2}ds +\frac{1}{2}\int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{V}_{0}^{*}}^{2}ds \leq M_{2}. |
Proof. We test (3.3) at time s\in (0, T) by \boldsymbol{v}_{\varepsilon }'(s). Then, by using the Young inequality, we see that
\begin{eqnarray*} &&\varepsilon \bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{H}_{0}}^{2}+ \bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{V}_{0}^{*}}^{2} +\frac{d}{ds}\varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)\bigr) \\ &&\quad \quad \quad +\frac{d}{ds}\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dx +\frac{d}{ds}\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma \\ &&\quad = \bigl(\boldsymbol{Pf}(s), \boldsymbol{v}_{\varepsilon }'(s)\bigr)_{\boldsymbol{H}_{0}} \\ &&\quad \leq \frac{1}{2}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}}^{2} +\frac{1}{2}\bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{V}_{0}^{*}}^{2} \end{eqnarray*} |
for a.a. s\in (0, T). This implies that
\begin{eqnarray} &&\varepsilon \bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{H}_{0}}^{2} +\frac{1}{2}\bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{V}_{0}^{*}}^{2} +\frac{d}{ds}\varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)\bigr) \nonumber \\ &&\quad \quad \quad +\frac{d}{ds}\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dx +\frac{d}{ds}\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma \nonumber \\ &&\quad \leq \frac{1}{2}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}}^{2} \label{mulvd} \end{eqnarray} | (3.29) |
for a.a. s\in (0, T). Therefore, by integrating it over (0, T) with respect to s and using the periodic property, we can conclude.
Lemma 3.4. There exists a positive constant M_{3}, independent of \varepsilon \in (0, 1], such that
\frac{1}{2}\bigl|\boldsymbol{v}_{\varepsilon }(t)\bigr|_{\boldsymbol{V}_{0}}^{2} +\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(t)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(t)+m_{0}\bigr)d\Gamma \leq M_{3} | (3.30) |
for all t\in [0, T].
Proof. For each s, t\in [0, T] such that s\leq t, we integrate (3.29) over [s, t]. Then, by neglecting the first two positive terms, we see that
\begin{eqnarray*} &&\varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(t)\bigr) +\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(t)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(t)+m_{0}\bigr)d\Gamma \\ &&\quad \leq \varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)\bigr) +\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma +\frac{1}{2}\int_{0}^{T}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds \end{eqnarray*} |
for all s, t\in [0, T]. Now, integrating it over (0, t) with respect to s, it follows that
\begin{eqnarray} &&\frac{t}{2}\bigl|\boldsymbol{v}_{\varepsilon }(t)\bigr|_{\boldsymbol{V}_{0}}^{2} +t\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(t)+m_{0}\bigr)dx +t\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(t)+m_{0}\bigr)d\Gamma \nonumber \\ &&\quad \leq \frac{1}{2}\int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\int_{0}^{T}\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dxds +\int_{0}^{T}\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma ds \nonumber \\ &&\quad \quad \quad +\int_{0}^{T}\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dxds +\int_{0}^{T}\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma ds \nonumber \\ &&\quad \quad \quad \quad \quad +\frac{T}{2}\int_{0}^{T}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds \label{ints} \end{eqnarray} | (3.31) |
for all t\in [0, T]. Here, note that we have
\begin{eqnarray} \bigl|\widehat{\widetilde{\pi }}(r)\bigr|&\leq &\int_{0}^{{ r}}\bigl|\widetilde{\pi }(\tau )\bigr|d\tau \nonumber \\ &\leq &\widetilde{L}\int_{0}^{{ |r|}}|\tau |d\tau +\int_{0}^{{ r}}\bigl|\widetilde{\pi }(0)\bigr|d\tau \nonumber \\ &{ \leq }&\frac{\widetilde{L}}{2}r^{2}+\bigl|\widetilde{\pi }(0)\bigr|{ |r|} \label{primitive} \end{eqnarray} | (3.32) |
for r{ >0}. Then we can easily show that (3.32) holds for any r\in \mathbb{R}. Similarly, we have
\bigl|\widehat{\widetilde{\pi }}_{\Gamma }(r)\bigr| \leq \frac{\widetilde{L}_{\Gamma }}{2}r^{2}+\bigl|\widetilde{\pi }_{\Gamma }(0)\bigr|{ |r|} \quad {\rm for~all~}r\in \mathbb{R}. |
Then, by using the Young inequality, we infer that
\begin{eqnarray} \int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dx &\leq &\int_{\Omega }\left(\frac{\widetilde{L}}{2}\bigl|v_{\varepsilon }(s)+m_{0}\bigr|^{2}+ \bigl|\widetilde{\pi }(0)\bigr|\bigl|v_{\varepsilon }(s)+m_{0}\bigr|\right)dx \nonumber \\ &\leq &\widetilde{L}\int_{\Omega }\bigl|v_{\varepsilon }(s)+m_{0}\bigr|^{2}dx +\frac{1}{2\widetilde{L}}\bigl|\widetilde{\pi }(0)\bigr|^{2}|\Omega | \nonumber \\ &\leq &2\widetilde{L}\int_{\Omega }\bigl|v_{\varepsilon }(s)\bigr|^{2}dx +2m_{0}^{2}|\Omega |+\frac{1}{2\widetilde{L}}\bigl|\widetilde{\pi }(0)\bigr|^{2}|\Omega | \label{pihat} \end{eqnarray} | (3.33) |
for a.a. s\in [0, T]. Similarly, we have
\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma \leq 2\widetilde{L}_{\Gamma }\int_{\Gamma }\bigl|v_{\Gamma, \varepsilon }(s)\bigr|^{2}d\Gamma +2m_{0}^{2}|\Gamma |+\frac{1}{2\widetilde{L}_{\Gamma }}\bigl|\widetilde{\pi }_{\Gamma }(0)\bigr|^{2}|\Gamma | \label{pighat} | (3.34) |
for a.a. s\in [0, T]. Thus, on account of (3.31)–(3.34), we deduce that
\begin{eqnarray*} &&\frac{t}{2}\bigl|\boldsymbol{v}_{\varepsilon }(t)\bigr|_{\boldsymbol{V}_{0}}^{2} +t\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(t)+m_{0}\bigr)dx +t\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(t)+m_{0}\bigr)d\Gamma \\ &&\quad \leq \frac{1}{2}\int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\int_{0}^{T}\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dxds +\int_{0}^{T}\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma ds \\ &&\quad \quad \quad +2\widetilde{L}\int_{\Omega }\bigl|v_{\varepsilon }(s)\bigr|^{2}dx +2\widetilde{L}_{\Gamma }\int_{\Gamma }\bigl|v_{\Gamma, \varepsilon }(s)\bigr|^{2}d\Gamma +\frac{T}{2}\int_{0}^{T}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\tilde{M}_{4} \\ &&\quad \leq \left(\frac{1}{2}+\widehat{L}c_{{\rm P}}\right)\int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\int_{0}^{T}\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dxds \\ &&\quad \quad \quad +\int_{0}^{T}\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma ds +\frac{T}{2}\int_{0}^{T}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\tilde{M}_{4} \end{eqnarray*} |
for all t\in [0, T], where \widehat{L}: = \max \{2\widetilde{L}, 2\widetilde{L}_{\Gamma }\} and
\tilde{M}_{4}: = 2m_{0}^{2}|\Omega |+\frac{1}{2\widetilde{L}}\bigl|\widetilde{\pi }(0)\bigr|^{2}|\Omega | +2m_{0}^{2}|\Gamma |+\frac{1}{2\widetilde{L}_{\Gamma }}\bigl|\widetilde{\pi }_{\Gamma }(0)\bigr|^{2}|\Gamma |. |
In particular, putting t: = T and dividing it by T, it follows that
\begin{eqnarray} &&\frac{1}{2}\bigl|\boldsymbol{v}_{\varepsilon }(T)\bigr|_{\boldsymbol{V}_{0}}^{2} +\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(T)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(T)+m_{0}\bigr)d\Gamma \nonumber \\ &&\quad \leq \frac{1}{T}\left(\frac{1}{2}+\widehat{L}c_{{\rm P}}\right)\int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\frac{1}{T}\int_{0}^{T}\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)dxds \nonumber \\ &&\quad \quad \quad +\frac{1}{T}\int_{0}^{T}\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma ds +\frac{1}{2}\int_{0}^{T}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\frac{\tilde{M}_{4}}{T}. \label{estiT} \end{eqnarray} | (3.35) |
Combining (3.28) and (3.35), there exists a positive constant \tilde{M}_{3} such that
\frac{1}{2}\bigl|\boldsymbol{v}_{\varepsilon }(T)\bigr|_{\boldsymbol{V}_{0}}^{2} +\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(T)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(T)+m_{0}\bigr)d\Gamma \leq \tilde{M}_{3}. |
From the periodic property, we have
\varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(0)\bigr) = \frac{1}{2}\bigl|\boldsymbol{v}_{\varepsilon }(0)\bigr|_{\boldsymbol{V}_{0}}^{2} +\int_{\Omega }\widehat{\beta }_{\varepsilon }\bigl(v_{\varepsilon }(0)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\beta }_{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(0)+m_{0}\bigr)d\Gamma \leq \tilde{M}_{3}. | (3.36) |
Now, integrating (3.29) by (0, t) with respect to s, it follows from (3.33)–(3.34) that
\begin{eqnarray} &&\varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(t)\bigr) +\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(t)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(t)+m_{0}\bigr)d\Gamma \nonumber \\ &&\quad \leq \varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(0)\bigr) +\int_{\Omega }\widehat{\widetilde{\pi }}\bigl(v_{\varepsilon }(0)+m_{0}\bigr)dx +\int_{\Gamma }\widehat{\widetilde{\pi }}_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(0)+m_{0}\bigr)d\Gamma +\frac{1}{2}\int_{0}^{T}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds \nonumber \\ &&\quad \leq \bigl(1+2\widehat{L}c_{{\rm P}}\bigr)\varphi _{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(0)\bigr) +\frac{1}{2}\int_{0}^{T}\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +\tilde{M}_{4} \label{esv0} \end{eqnarray} | (3.37) |
for all t\in [0, T]. Therefore, by virtue of (3.36)–(3.37), there exists a positive constant M_{3} such that the estimate (3.30) holds.
Lemma 3.5. There exists a positive constant M_{4}, independent of \varepsilon \in (0, 1], such that
\delta _{0}\int_{0}^{T}\bigl|\beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr|_{L^{1}(\Omega )}^{2}ds +\delta _{0}\int_{0}^{T}\bigl|\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{L^{1}(\Gamma )}^{2}ds \leq M_{4} | (3.38) |
for some positive constants \delta _{0}.
Proof. We employ the method of [11,Lemmas 4.1,4.3], indeed we impose same assumptions as [11] for \beta, \beta _{\Gamma } and being m_{0}\in {\rm int}D(\beta _{\Gamma }). Therefore, we can also exploit the following inequalities stated in [18,Sect. 5]: for each \varepsilon \in (0, 1], there exist two positive constants \delta _{0} and c_{1} such that
\beta _{\varepsilon }(r)(r-m_{0})\geq \delta _{0}\bigl|\beta _{\varepsilon }(r)\bigr|-c_{1}, \quad \beta _{\Gamma, \varepsilon }(r)(r-m_{0})\geq \delta _{0}\bigl|\beta _{\Gamma, \varepsilon }(r)\bigr|-c_{1} |
for all r\in \mathbb{R}. Hence, it follows that
\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{u}_{\varepsilon }(s)\bigr), \boldsymbol{v}_{\varepsilon }(s)\bigr)_{\boldsymbol{H}} \geq \delta _{0}\int_{\Omega }\bigl|\beta _{\varepsilon }\bigl(u_{\varepsilon }(s)\bigr)\bigr|dx-c_{1}|\Omega | +\delta _{0}\int_{\Gamma }\bigl|\beta _{\Gamma, \varepsilon }\bigl(u_{\Gamma, \varepsilon }(s)\bigr)\bigr|d\Gamma -c_{1}|\Gamma | | (3.39) |
for a.a. s\in (0, T). On the other hand, we test (3.3) at time s\in (0, T) by \boldsymbol{v}_{\varepsilon }(s). Then, from (3.2), we see that
\begin{eqnarray} &&\bigl(\varepsilon \boldsymbol{v}_{\varepsilon }'(s), \boldsymbol{v}_{\varepsilon }(s)\bigr)_{\boldsymbol{H}_{0}} +\bigl(\boldsymbol{v}_{\varepsilon }'(s), \boldsymbol{v}_{\varepsilon }(s)\bigr)_{\boldsymbol{V}_{0}^{*}} +\bigl(\boldsymbol{A}\boldsymbol{v}_{\varepsilon }(s), \boldsymbol{v}_{\varepsilon }(s)\bigr)_{\boldsymbol{H}_{0}} +\bigl(\boldsymbol{P}\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{u}_{\varepsilon }(s)\bigr), \boldsymbol{v}_{\varepsilon }(s)\bigr)_{\boldsymbol{H}_{0}} \nonumber \\ &&\quad \leq \bigl(\boldsymbol{f}(s)-\widetilde{\boldsymbol{\pi }}\bigl(\boldsymbol{u}_{\varepsilon }(s)\bigr), \boldsymbol{v}_{\varepsilon }(s)\bigr)_{\boldsymbol{H}}. \label{inev} \end{eqnarray} | (3.40) |
Hence, from (3.39)–(3.40) and the maximal monotonicity of \boldsymbol{A}, by squaring we have
\begin{eqnarray*} &&\left(\delta _{0}\int_{\Omega }\bigl|\beta _{\varepsilon }\bigl(u_{\varepsilon }(s)\bigr)\bigr|dx+ \delta _{0}\int_{\Gamma }\bigl|\beta _{\Gamma, \varepsilon }\bigl(u_{\Gamma, \varepsilon }(s)\bigr)\bigr|d\Gamma \right)^{2} \leq 3c_{1}^{2}(|\Omega |+|\Gamma |)^{2} \nonumber \\ &&\quad +9\bigl(\bigl|\boldsymbol{f}(s)\bigr|_{\boldsymbol{H}}^{2}+\bigl|\boldsymbol{\pi }\bigl(\boldsymbol{u}_{\varepsilon }(s)\bigr)\bigr|_{\boldsymbol{H}}^{2} +\varepsilon ^{2}\bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{H}_{0}}^{2}\bigr)\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{H}_{0}}^{2} +3\bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{V}_{0}^{*}}^{2}\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{V}_{0}^{*}}^{2} \end{eqnarray*} |
for a.a. s\in (0, T). Therefore, from the Lipschitz continuity of \widetilde{\pi }, \widetilde{\pi }_{\Gamma } and Lemma 3.4, by integrating it over (0, T) with respect to s, there exists a positive constant M_{4} such that the estimate (3.38) holds.
Lemma 3.6. There exists a positive constants M_{5}, independent of \varepsilon \in (0, 1], such that
\int_{0}^{T}\bigl|\boldsymbol{\mu }_{\varepsilon }(s)\bigr|_{\boldsymbol{V}}^{2}ds\leq M_{5}. | (3.41) |
Proof. Firstly, by using the Lipschitz continuity of \widetilde{\pi }, \widetilde{\pi }_{\Gamma } and the Hölder inequality, it follows from (2.1) and Lemma 3.4 that there exists a positive constant M_{5}^{*} such that
\begin{eqnarray}\label{volpi} &&\bigl|m\bigl(\widetilde{\boldsymbol{\pi }}\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr)\bigr| \nonumber \\ &&\quad \leq \frac{1}{|\Omega |+|\Gamma |}\left\{\int_{\Omega }\bigl|\widetilde{\pi }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr|dx +\int_{\Gamma }\bigl|\widetilde{\pi }_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|d\Gamma \right\} \nonumber \\ &&\quad \leq \frac{1}{|\Omega |+|\Gamma |}\left\{\widetilde{L}|\Omega |^{\frac{1}{2}}\bigl|v_{\varepsilon }(s)\bigr|_{H}^{2} +\widetilde{L}|\Omega ||m_{0}|+|\Omega |\bigl|\widetilde{\pi }(0)\bigr| \right. \nonumber \\ &&\left. \quad \quad \quad +\widetilde{L}_{\Gamma }|\Gamma |^{\frac{1}{2}}\bigl|v_{\Gamma, \varepsilon }(s)\bigr|_{H_{\Gamma }}^{2} +\widetilde{L}_{\Gamma }|\Gamma ||m_{0}|+|\Gamma |\bigl|\widetilde{\pi }_{\Gamma }(0)\bigr|\right\} \nonumber \\ &&\quad \leq \frac{1}{|\Omega |+|\Gamma |}M_{5}^{*}\left\{\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{V}_{0}}^{2}+1\right\} \nonumber \\ &&\quad \leq \frac{1}{|\Omega |+|\Gamma |}M_{5}^{*}(M_{3}+1) = :\tilde{M}_{5} \end{eqnarray} | (342) |
for a.a. s\in (0, T). Therefore, owing to (3.42) we deduce that
\begin{eqnarray*} \bigl|m\bigl(\boldsymbol{\mu }_{\varepsilon }(s)\bigr)\bigr|^{2} & = &\bigl|m\bigl(\widetilde{\boldsymbol{\pi }}\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)-\boldsymbol{f}(s)\bigr)\bigr|^{2} \\ &\leq &2\tilde{M}_{5}^{2}+\frac{4}{(|\Omega |+|\Gamma |)^{2}}\bigl(\bigl|f(s)\bigr|_{L^{1}(\Omega )}+\bigl|f_{\Gamma }(s)\bigr|_{L^{1}(\Gamma )}\bigr) = :\hat{M}_{5} \end{eqnarray*} |
for a.a. s\in (0, T). Next, from (2.1), (3.27) and the fact \boldsymbol{P}\boldsymbol{\mu }_{\varepsilon }(s) = \boldsymbol{\mu }_{\varepsilon }(s)-m(\boldsymbol{\mu }_{\varepsilon }(s))\boldsymbol{1} for a.a. s\in (0, T), we deduce that
\begin{eqnarray*} \int_{0}^{T}\bigl|\boldsymbol{\mu }_{\varepsilon }(s)\bigr|_{\boldsymbol{V}}^{2}ds &\leq &2\int_{0}^{T}\bigl|\boldsymbol{P}\boldsymbol{\mu }_{\varepsilon }(s)\bigr|_{\boldsymbol{V}}^{2}ds +2\int_{0}^{T}\bigl|m\bigl(\boldsymbol{\mu }_{\varepsilon }(s)\bigr)\boldsymbol{1}\bigr|_{\boldsymbol{V}}^{2}ds \\ &\leq &2c_{{\rm P}}\int_{0}^{T}\bigl|\boldsymbol{P}\boldsymbol{\mu }_{\varepsilon }(s)\bigr|_{\boldsymbol{V}_{0}}^{2}ds +2(|\Omega |+|\Gamma |)\int_{0}^{T}\bigl|m\bigl(\boldsymbol{\mu }_{\varepsilon }(s)\bigr)\bigr|^{2}ds \\ &\leq &2c_{{\rm P}}\int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }'(s)\bigr|_{\boldsymbol{V}_{0}^{*}}^{2}ds +2T(|\Omega |+|\Gamma |)\hat{M}_{5}^{2}. \\ \end{eqnarray*} |
Thus, from Lemma 3.3, there exists a positive constant M_{5} such that the estimate (3.41) holds.
Lemma 3.7. There exists a positive constant M_{6}, independent of \varepsilon \in (0, 1], such that
\frac{1}{2}\int_{0}^{T}\bigl|\beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr|_{H}^{2}ds +\frac{1}{4\rho }\int_{0}^{T}\bigl|\beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2}ds\leq M_{6}. | (3.43) |
Proof. From the definition of \boldsymbol{\mu }_{\varepsilon }, we can infer that
\mu _{\varepsilon } = \varepsilon \partial _{t}v_{\varepsilon }-\kappa _{1}\Delta v_{\varepsilon } +\beta _{\varepsilon }(v_{\varepsilon }+m_{0}) -m\bigl(\boldsymbol{\beta }_{\varepsilon }(\boldsymbol{v}_{\varepsilon }+m_{0}\boldsymbol{1})\bigr)+\widetilde{\pi }(v_{\varepsilon }+m_{0})-f \quad {\rm a.e.~in~}Q, | (3.44) |
\begin{eqnarray} &&\mu _{\Gamma, \varepsilon } = \varepsilon \partial _{t}v_{\Gamma, \varepsilon }+\kappa _{1}\partial _{\boldsymbol{\nu }}v_{\varepsilon } -\kappa _{2}\Delta _{\Gamma }v_{\Gamma, \varepsilon } +\beta _{\Gamma, \varepsilon }(v_{\Gamma, \varepsilon }+m_{0}) -m\bigl(\boldsymbol{\beta }_{\varepsilon }(\boldsymbol{v}_{\varepsilon }+m_{0}\boldsymbol{1})\bigr) \nonumber \\ &&\quad \quad \quad +\widetilde{\pi }_{\Gamma }(v_{\Gamma, \varepsilon }+m_{0})-f_{\Gamma } \quad {\rm a.e.~on~}\Sigma . \end{eqnarray} | (3.45) |
Now, it follows from (3.38) that there exists a positive constant \tilde{M}_{6} such that
\begin{eqnarray}\label{volbeta} &&\bigl|m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr)\bigr|^{2} \nonumber \\ &&\quad \leq \frac{2}{(|\Omega |+|\Gamma |)^{2}}\bigl(\bigl|\beta _{\varepsilon }\bigl(v_{\varepsilon }+m_{0}\bigr)\bigr|_{L^{1}(\Omega )} +\bigl|\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{L^{1}(\Gamma )}\bigr) \nonumber \\ &&\quad \leq \tilde{M}_{6} \end{eqnarray} | (3.46) |
for a.a. s\in (0, T). Moreover, we test (3.44) at time s\in (0, T) by \beta _{\varepsilon }(v_\varepsilon (s)+m_{0}) and exploit (3.45). Then, on account of the fact (\beta _{\varepsilon }(v_{\varepsilon }+m_{0}))_{|_{\Gamma }} = \beta _{\varepsilon }(v_{\Gamma, \varepsilon }+m_{0}), by integrating over \Omega we deduce that
\begin{eqnarray} &&\kappa _{1}\int_{\Omega }\beta _{\varepsilon }'\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigl|\nabla v_{\varepsilon }(s)\bigr|^{2}dx +\kappa _{2}\int_{\Gamma }\beta _{\varepsilon }'\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigl|\nabla _{\Gamma }v_{\Gamma, \varepsilon }(s)\bigr|^{2}d\Gamma \nonumber \\ &&\quad \quad \quad +\bigl|\beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr|_{H}^{2}+\int_{\Gamma }\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma \nonumber \\ &&\quad \leq \bigl(f(s)+\mu _{\varepsilon }(s)-\varepsilon v_{\varepsilon }'(s)-\widetilde{\pi }\bigl(v_{\varepsilon }(s)+m_{0}\bigr), \beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr)_{H} \nonumber \\ &&\quad \quad \quad +\bigl(m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr), \beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr)_{H} \nonumber \\ &&\quad \quad \quad \quad \quad +\bigl(f_{\Gamma }(s)+\mu _{\Gamma, \varepsilon }(s)-\varepsilon v_{\Gamma, \varepsilon }'(s)-\widetilde{\pi }_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr), \beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr)_{H_{\Gamma }} \nonumber \\ &&\quad \quad \quad \quad \quad \quad \quad +\bigl(m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr), \beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr)_{H_{\Gamma }} \label{intomg} \end{eqnarray} | (3.47) |
for a.a. s\in (0, T). Now, from (3.1), since the both signs of \beta _{\varepsilon }(r) and \beta _{\Gamma, \varepsilon }(r) are same for all r\in \mathbb{R}, we infer that
\begin{eqnarray} \int_{\Gamma }\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)d\Gamma & = &\int_{\Gamma }\bigl|\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|\bigl|\beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|d\Gamma \nonumber \\ &\geq &\frac{1}{2\rho }\int_{\Gamma }\bigl|\beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|^{2}d\Gamma -\frac{c_{0}^{2}}{2\rho }|\Gamma |. \label{sign} \end{eqnarray} | (3.48) |
Also, it holds
\int_{\Omega }\beta _{\varepsilon }'\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigl|\nabla v_{\varepsilon }(s)\bigr|^{2}dx\geq 0, \quad \int_{\Gamma }\beta _{\Gamma, \varepsilon }'\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigl|\nabla _{\Gamma }v_{\Gamma, \varepsilon }(s)\bigr|^{2}d\Gamma \geq 0. | (3.49) |
Moreover, by using the Young inequality, the Lipschitz continuity of \widetilde{\pi }, \widetilde{\pi }_{\Gamma } and (3.46), there exists a positive constant \hat{M}_{6} such that
\begin{eqnarray} &&\bigl(f(s)+\mu _{\varepsilon }(s)-\varepsilon v_{\varepsilon }'(s)-\widetilde{\pi }\bigl(v_{\varepsilon }(s)+m_{0}\bigr), \beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr)_{H} \nonumber \\ &&\quad \quad \quad +\bigl(m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr), \beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr)_{H} \nonumber \\ &&\quad \leq \frac{1}{2}\bigl|\beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr|_{H}^{2} +4\bigl|f(s)\bigr|_{H}^{2}+4\bigl|\mu _{\varepsilon }(s)\bigr|_{H}^{2} +4\varepsilon ^{2}\bigl|v_{\varepsilon }'(s)\bigr|_{H}^{2} +4\bigl|\widetilde{\pi }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr|_{H}^{2} \nonumber \\ &&\quad \quad \quad +\bigl|m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr)\bigr|_{H}^{2} \nonumber \\ &&\quad \leq \frac{1}{2}\bigl|\beta _{\varepsilon }\bigl(v_{\varepsilon }(s)+m_{0}\bigr)\bigr|_{H}^{2} +\hat{M}_{6}\bigl(\bigl|f(s)\bigr|_{H}^{2}+\bigl|\mu _{\varepsilon }(s)\bigr|_{H}^{2} +\varepsilon ^{2}\bigl|v_{\varepsilon }'(s)\bigr|_{H}^{2}+\bigl|v_{\varepsilon }(s)\bigr|_{H}^{2}+1\bigr) \nonumber \\ &&\quad \quad \quad +|\Omega |\tilde{M}_{6} \label{homg} \end{eqnarray} | (3.50) |
and
\begin{eqnarray} &&\bigl(f_{\Gamma }(s)+\mu _{\Gamma, \varepsilon }(s)-\varepsilon v_{\Gamma, \varepsilon }'(s)-\widetilde{\pi }_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr), \beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr)_{H_{\Gamma }} \nonumber \\ &&\quad \quad \quad +\bigl(m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr), \beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr)_{H} \nonumber \\ &&\quad \leq \frac{1}{4\rho }\bigl|\beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2} +2\rho \bigl|f_{\Gamma }(s)\bigr|_{H}^{2}+2\rho \bigl|\mu _{\Gamma, \varepsilon }(s)\bigr|_{H_{\Gamma }}^{2} +2\rho \varepsilon ^{2}\bigl|v_{\Gamma, \varepsilon }'(s)\bigr|_{H_{\Gamma }}^{2} \nonumber \\ &&\quad \quad \quad +2\rho \bigl|\widetilde{\pi }_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2} +2\rho \bigl|m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr)\bigr|_{H_{\Gamma }}^{2} \nonumber \\ &&\quad \leq \frac{1}{4\rho }\bigl|\beta _{\varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2} +2\rho |\Gamma |\tilde{M}_{6} \nonumber \\ &&\quad \quad \quad +\rho \hat{M}_{6}\bigl(\bigl|f_{\Gamma }(s)\bigr|_{H_{\Gamma }}^{2}+\bigl|\mu _{\Gamma, \varepsilon }(s)\bigr|_{H_{\Gamma }}^{2} +\varepsilon ^{2}\bigl|v_{\Gamma, \varepsilon }'(s)\bigr|_{H_{\Gamma }}^{2} +\bigl|v_{\Gamma, \varepsilon }(s)\bigr|_{H_{\Gamma }}^{2}+1\bigr) \label{hgomg} \end{eqnarray} | (3.51) |
for a.a. s\in (0, T). Thus, from Lemmas 3.3, 3.4 and (2.1), by combining from (3.47)–(3.51) and integrating it over (0, T), we can conclude the existence of the constant M_{6} satisfying (3.43).
Lemma 3.8. There exists a positive constant M_{7}, independent of \varepsilon \in (0, 1], such that
\kappa _{1}\int_{0}^{T}\bigl|\Delta v_{\varepsilon }(s)\bigr|_{H}^{2}ds +\int_{0}^{T}\bigl|v_{\varepsilon }(s)\bigr|_{H^{\frac{3}{2}}(\Omega )}^{2}ds +\int_{0}^{T}\bigl|\partial _{\boldsymbol{\nu }}v_{\varepsilon }(s)\bigr|_{H_{\Gamma }}^{2}ds \leq M_{7}. | (3.52) |
This lemma is proved exactly the same as in [11, Lemmas 4.4] because the necessary uniform estimates to prove it is obtained by Lemmas 3.3, 3.4, 3.6 and 3.7. Sketching simply, comparing in (3.44) we deduce that |\Delta v_{\varepsilon }|_{L^{2}(0, T; H)} is uniformly bounded. Moreover, by using the theory of the elliptic regularity (see, e.g., [7,Theorem 3.2,p. 1.79]), we see that |v_{\varepsilon }|_{L^{2}(0, T; H^{3/2}(\Omega))} is also uniformly bounded. Thus, using both uniformly boundeds, we can conclude that (3.52) holds.
Lemma 3.9. There exists a positive constant M_{8}, independent of \varepsilon \in (0, 1], such that
\int_{0}^{T}\bigl|\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2}ds \leq M_{8}. | (3.53) |
Proof. We test (3.45) at time s\in (0, T) by \beta _{\Gamma, \varepsilon }(v_{\Gamma, \varepsilon }(s)+m_{0}) and integrating it over \Gamma . Then, by using the Young inequality and the Lipschitz continuity of \widetilde{\pi }_{\Gamma }, there exists a positive constant \tilde{M}_{8} such that
\begin{eqnarray} &&\kappa _{2}\int_{\Gamma }\beta _{\Gamma, \varepsilon }'\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigl|\nabla _{\Gamma }v_{\Gamma, \varepsilon }(s)\bigr|^{2}d\Gamma +\bigl|\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2} \nonumber \\ &&\quad = \bigl(f_{\Gamma }(s)+\mu _{\Gamma }(s)-\varepsilon v_{\Gamma, \varepsilon }'(s)-\partial _{\boldsymbol{\nu }}v_{\varepsilon }(s) -\widetilde{\pi }_{\Gamma }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr), \beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr)_{H_{\Gamma }} \nonumber \\ &&\quad \leq \frac{1}{2}\bigl|\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2} +\bigl|m\bigl(\boldsymbol{\beta }_{\varepsilon }\bigl(\boldsymbol{v}_{\varepsilon }(s)+m_{0}\boldsymbol{1}\bigr)\bigr)\bigr|_{H_{\Gamma }}^{2} \nonumber \\ &&\quad \quad \quad +\tilde{M}_{8}\bigl(\bigl|f_{\Gamma }(s)\bigr|_{H_{\Gamma }}^{2} +\bigl|\mu _{\Gamma }(s)\bigr|_{H_{\Gamma }}^{2} +\varepsilon ^{2}\bigl|v_{\Gamma, \varepsilon }'(s)\bigr|_{H_{\Gamma }}^{2} +\bigl|\partial _{\boldsymbol{\nu }}v_{\varepsilon }(s)\bigr|_{H_{\Gamma }}^{2} +\bigl|v_{\Gamma, \varepsilon }(s)\bigr|_{H_{\Gamma }}^{2}+1\bigr) \nonumber \\ &&\quad \leq \frac{1}{2}\bigl|\beta _{\Gamma, \varepsilon }\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigr|_{H_{\Gamma }}^{2} +|\Gamma |\tilde{M}_{6} \nonumber \\ &&\quad \quad \quad +\tilde{M}_{8}\bigl(\bigl|f_{\Gamma }(s)\bigr|_{H_{\Gamma }}^{2} +\bigl|\mu _{\Gamma }(s)\bigr|_{H_{\Gamma }}^{2} +\varepsilon ^{2}\bigl|v_{\Gamma, \varepsilon }'(s)\bigr|_{H_{\Gamma }}^{2} +\bigl|\partial _{\boldsymbol{\nu }}v_{\varepsilon }(s)\bigr|_{H_{\Gamma }}^{2} +\bigl|v_{\Gamma, \varepsilon }(s)\bigr|_{H_{\Gamma }}^{2}+1\bigr) \nonumber \\ \label{es8ine} \end{eqnarray} | (3.54) |
for a.a. s\in (0, T). Note that it holds
\kappa _{2}\int_{\Gamma }\beta _{\Gamma, \varepsilon }'\bigl(v_{\Gamma, \varepsilon }(s)+m_{0}\bigr)\bigl|\nabla _{\Gamma }v_{\Gamma, \varepsilon }(s)\bigr|^{2}d\Gamma \geq 0. |
Thus, on account of Lemmas 3.3, 3.4, 3.6 and 3.8, by integrating (3.54) over (0, T), we can find a positive constant M_{7} such that the estimate (3.53) holds.
Lemma 3.10. There exists a positive constant M_{9}, independent of \varepsilon \in (0, 1], such that
\int_{0}^{T}\bigl|\boldsymbol{v}_{\varepsilon }(s)\bigr|_{\boldsymbol{W}}^{2}ds \leq M_{9}. |
This lemma is also proved the same as in [11,Lemmas 4.5]. The key point to prove it is that we can obtain the uniform estimate of |\Delta _{\Gamma }v_{\Gamma, \varepsilon }|_{L^{2}(0, T; H_{\Gamma })} by comparing in (3.45). We omit the proof.
In this section, we obtain the existence of periodic solutions of (P) by performing passage to the limit for the approximate problem (P)_{\varepsilon }. The convergence theorem is also nearly the same [11,Sect. 4]. The different point from [11] is that the component of the periodic solution of (P) satisfies (2.4) and the periodic property (2.5).
Thanks to the previous estimates in Lemmas from 3.3 to 3.10, there exist a subsequence \{\varepsilon _{k}\}_{k\in \mathbb{N}} with \varepsilon _{k}\to 0 as k\to \infty and some limits functions \boldsymbol{v}\in H^{1}(0, T; \boldsymbol{V}_{0}^{*})\cap L^{\infty }(0, T; \boldsymbol{V}_{0})\cap L^{2}(0, T; \boldsymbol{W}), \boldsymbol{\mu }\in H^{1}(0, T; \boldsymbol{V}), \xi \in L^{2}(0, T; H) and \xi _{\Gamma }\in L^{2}(0, T; H_{\Gamma }) such that
\boldsymbol{v}_{\varepsilon _{k}}\to \boldsymbol{v} \quad {\rm weakly~star~in~} H^{1}(0, T; \boldsymbol{V}_{0}^{*})\cap L^{\infty }(0, T; \boldsymbol{V}_{0})\cap L^{2}(0, T; \boldsymbol{W}), | (4.1) |
\varepsilon _{k}\boldsymbol{v}_{\varepsilon _{k}}\to 0 \quad {\rm strongly~in~} H^{1}(0, T; \boldsymbol{H}_{0}), |
\boldsymbol{\mu }_{\varepsilon _{k}}\to \boldsymbol{\mu } \quad {\rm weakly~in~} L^{2}(0, T; \boldsymbol{V}), |
\beta _{\varepsilon _{k}}(u_{\varepsilon _{k}})\to \xi \quad {\rm weakly~in~} L^{2}(0, T; H), | (4.2) |
\beta _{\Gamma, \varepsilon _{k}}(u_{\Gamma, \varepsilon _{k}})\to \xi _{\Gamma } \quad {\rm weakly~in~} L^{2}(0, T; H_{\Gamma }) | (4.3) |
as k\to \infty . Owing to (4.1) and a well-known compactness results (see, e.g., [30]), we obtain
\boldsymbol{v}_{\varepsilon _{k}}\to \boldsymbol{v} \quad {\rm strongly~in~} C([0, T]; \boldsymbol{H}_{0})\cap L^{2}(0, T; \boldsymbol{V}_{0}) | (4.4) |
as k\to \infty . This yeilds that
\boldsymbol{u}_{\varepsilon _{k}}\to \boldsymbol{u}: = \boldsymbol{v}+m_{0}\boldsymbol{1} \quad {\rm strongly~in~} C([0, T]; \boldsymbol{H}_{0})\cap L^{2}(0, T; \boldsymbol{V}_{0}) | (4.5) |
as k\to \infty . Therefore, from (4.5) and the Lipschitz continuity of \widetilde{\pi }, \widetilde{\pi }_{\Gamma }, we deduce that
\widetilde{\boldsymbol{\pi }}(\boldsymbol{u}_{\varepsilon _{k}})\to \widetilde{\boldsymbol{\pi }}(\boldsymbol{u}) \quad {\rm strongly~in~} C([0, T]; \boldsymbol{H}) |
as k\to \infty . Hence, by passing to the limit in (3.26) and (3.27), we obtain (2.3) and the following weak formulation:
\bigl(\boldsymbol{\mu }(t), \boldsymbol{z}\bigr)_{\boldsymbol{H}} = a\bigl(\boldsymbol{v}(t), \boldsymbol{z}\bigr) +\bigl(\boldsymbol{\xi }(t)-m\bigl(\boldsymbol{\xi }(t)\bigr)\boldsymbol{1} +\widetilde{\boldsymbol{\pi }}\bigl(\boldsymbol{u}(t)\bigr)-\boldsymbol{f}(t), \boldsymbol{z}\bigr)_{\boldsymbol{H}} \quad {\rm for~all~}z\in \boldsymbol{V} | (4.6) |
for a.a. t\in (0, T), where \boldsymbol{\xi }: = (\xi, \xi _{\Gamma }), because of the property (2.2) of linear bounded operator \boldsymbol{P}. Now, we can infer v+m_{0}\in D(\beta) and v_{\Gamma }+m_{0}\in D(\beta _{\Gamma }). Hence, from the form (2.5) and (3.6), we deduce that \widetilde{\pi }(v+m_{0}) = \pi (v+m_{0}) a.e. in Q and \widetilde{\pi }_{\Gamma }(v_{\Gamma }+m_{0}) = \pi _{\Gamma }(v_{\Gamma }+m_{0}) a.e. on \Sigma . This implies that we obtain (2.4) replaced by (4.6). Moreover, it follows from (4.4) that
\boldsymbol{v}(0) = \boldsymbol{v}(T) \quad {\rm in~}\boldsymbol{H}_{0}. |
Also, due to (4.2), (4.3), (4.5) and the monotonicity of \beta , from the fact [5,Prop. 2.2,p. 38] we obtain
\xi \in \beta (v+m_{0}) \quad {\rm a.e.~in~}Q, \quad \xi _{\Gamma }\in \beta _{\Gamma }(v_{\Gamma }+m_{0}) \quad {\rm a.e.~on~}\Sigma . |
Thus, we complete the proof of Theorem 2.1.
The author is deeply grateful to the anonymous referees for reviewing the original manuscript and for many valuable comments that helped to clarify and refine this paper.
The author declares no conflicts of interest in this paper.
[1] | M. W. Trierweiler, R. K. Freeman and J. James, Baseline fetal heart rate characteristics as anindicator of fetal status during the antepartum period, Am. J. Obstet. Gynecol., 125 (1976),618–623. |
[2] | V. S. Chourasia, A. K. Tiwari and R. Gangopadhyay, A novel approach for phonocardiographicsignals processing to make possible fetal heart rate evaluations, Digit. Signal Process., 30 (2014),165–183. |
[3] | I. Habermajer, F. Kovács and M. Török, A rule-based phonocardiographic method for long-termfetal heart rate monitoring, IEEE Trans. Biomed. Eng., 47 (2000), 124–130. |
[4] | M. Ruffo, M. Cesarelli, M. Romano, et al., An algorithm for FHR estimation from foetalphonocardiographic signals, Biomed. Signal Process. Control, 5 (2010), 131–141. |
[5] | Y. Song, W. Xie, J. F. Chen, et al., Passive acoustic maternal abdominal fetal heart ratemonitoring using wavelet transform, Comput. Cardiol., 33 (2006), 581–584. |
[6] | H. Tang, T. Li, T. Qiu, et al., Fetal heart rate monitoring from phonocardiograph signal usingrepetition frequency of heart sounds, JECE, (2016), 2404267. |
[7] | J. Chen, K. Phua, Y. Song, et al., A portable phonocardiographic fetal heart rate monitor, ISCAS, (2006), 2141–2144. |
[8] | J. P. Phelan and P. E. Lewis, Fetal heart rate decelerations during a nonstress test, Obstet.Gynecol., 57 (1981), 228–232. |
[9] | F. Kovács, C. Horváth, A. T. Balogh, et al., Fetal phonocardiography-past and futurepossibilities, Comput. Methods Programs Biomed., 104 (2011), 19–25. |
[10] | H. E. Bessil and J. H. Dripps, Real-time processing and analysis of fetal phonocardiographicsensor, Clin. Phys. Physiol. Meas., 10 (1989), 67–74. |
[11] | P. C. Adithya, R. Sankar, W. A. Moreno, et al., Trends in fetal monitoring throughphonocardiography: challenges and future directions, Biomed. Signal Process. Control, 33 (2017),289–305. |
[12] | V. S. Chourasia and A. K. Tiwari, Design methodology of a new wavelet basis function for fetalphonocardiographic signals, Sci. World J., (2013), Article ID 505840. |
[13] | A. N. Pelech, The physiology of cardiac auscultation, Pediatr. Clin. North. Am., 51 (2004), 1515–1535. |
[14] | G. Anastasi, S. Capitani and M. Carnazza, in Trattato di anatomia umana, Edi-Ermes, 4 (2010), 321–324. |
[15] | Z. Comert and A. F. Kocamaz, A study of artificial neural network training algorithms for classification of cardiotocography signals, Bitlis Eren. Univ. J. Sci. Technol., 7(2017), 93–103. |
[16] | R. Sameni and G. Clifford, A review of fetal ECG signal processing; issues and promising directions, Electrophysiol. Ther. J., 3 (2010), 4–20. |
[17] | K. M. J. Verdurmen, C. Lempersz, R. Vullings, et al., Normal ranges for fetal electrocardiogram values for the healthy fetus of 18–24 weeks of gestation: A prospective cohort study, BMC Pregnancy Childbirth, 16 (2016), 227. |
[18] | A. Agostinelli, M. Grillo, A. Biagini, et al., Noninvasive fetal electrocardiography: An overview of the signal electrophysiological meaning, recording procedures, and processing techniques, Ann. Noninvasive Electrocardiol., 20 (2015), 303–313. |
[19] | M. Sato, Y. Kimura, S. Chida, et al., A novel extraction method of fetal electrocardiogram from the composite abdominal signal, IEEE Trans. Biomed. Eng., 54 (2007), 49–58. |
[20] | V. S. Chourasia and A. K. Tiwari, A review and comparative analysis of recent advancements in fetal monitoring techniques, Crit. Rev. Biomed. Eng., 36 (2008), 335–373. |
[21] | A. Sbrollini, A. Strazza, M. Caragiuli, et al., Fetal phonocardiogram denoising by wavelet transformation: robustness to noise, Comput. Cardiol., 44 (2017), 1–4. |
[22] | M. Samieinasab and R. Sameni, Fetal phonocardiogram extraction using single channel blind source separation, ICEE 2015, (2015), 78–83. |
[23] | A. Khandoker, E. Ibrahim, S. Oshio, et al., Validation of beat by beat fetal heart signals acquired from four-channel fetal phonocardiogram with fetal electrocardiogram in healthy late pregnancy, Sci. Rep., 8 (2018), 13635. |
[24] | A. Jimenez, M. R. Ortiz, M. A. Pena, et al., The use of wavelet packets to improve the detection of cardiac sounds from the fetal phonocardiogram, Comput. Cardiol., 26 (1999), 463–466. |
[25] | E. Koutsiana, L. J. Hadjileontiadis, I. Chouvarda, et al., Detecting fetal heart sounds by means of fractal dimension analysis in the wavelet domain, EMBC 2017, (2017), 2201–2204. |
[26] | D. Messer, S. Agzarian and J. Abbott, Optimal wavelet denoising for phonocardiograms, Microelectron. J., 32 (2001), 931–941. |
[27] | S. Vaisman, S.Y. Salem and G. Holcberg, Passive fetal monitoring by adaptive wavelet denoising method, Comput. Biol. Med., 42 (2012), 171–179. |
[28] | A. Strazza, A. Sbrollini, V. Di Battista, et al., PCG-Delineator: an efficient algorithm for automatic heart sounds detection in fetal phonocardiography, Comput. Cardiol., 45 (2018), 1–4. |
[29] | V. S. Chourasia and A. K. Mittra, Selection of mother wavelet and denoising algorithm for analysis of foetal phonocardiographic signals, J. Med. Eng. Technol., 33 (2009), 442–448. |
[30] | V. S. Chourasia and A. K. Mittra, Most suitable mother wavelet for fetal phonocardiographic signal analysis, IJFET, 4 (2009), 23–29. |
[31] | A. L. Goldberger, L. A. Amaral, L. Glass, et al., PhysioBank, PhysioToolkit, and PhysioNet: components of a new research resource for complex physiologic signals, Circulation, 101 (2000), 215–220. |
[32] | M. Cesarelli, M. Ruffo, M. Romano, et al., Simulation of foetal phonocardiographic recordings for testing of FHR extraction algorithms, Comput. Methods. Programs Biomed., 107 (2012), 513–523. |
[33] | C. Liu, D. Springer, Q. Li, et al., An open access database for the evaluation of heart sound algorithms, Physiol. Meas., 37 (2016), 2181–2213. |
[34] | A. Misal, G. R. Sinha, R. M. Potdar, et al., Comparison of wavelet transforms for denoising and analysis of PCG signal, JCS, 1 (2012), 1–5. |
[35] | B. Ergen, Comparison of wavelet types and thresholding methods on wavelet based denoising of heart sounds, JSIP, 4 (2013), 164–167. |
1. | Stefan Metzger, An Efficient and Convergent Finite Element Scheme for Cahn--Hilliard Equations with Dynamic Boundary Conditions, 2021, 59, 0036-1429, 219, 10.1137/19M1280740 | |
2. | Takeshi Fukao, Hao Wu, Separation property and convergence to equilibrium for the equation and dynamic boundary condition of Cahn–Hilliard type with singular potential, 2020, 18758576, 1, 10.3233/ASY-201646 | |
3. | Harald Garcke, Patrik Knopf, Weak Solutions of the Cahn--Hilliard System with Dynamic Boundary Conditions: A Gradient Flow Approach, 2020, 52, 0036-1410, 340, 10.1137/19M1258840 | |
4. | Patrik Knopf, Kei Fong Lam, Chun Liu, Stefan Metzger, Phase-field dynamics with transfer of materials: The Cahn–Hillard equation with reaction rate dependent dynamic boundary conditions, 2021, 55, 0764-583X, 229, 10.1051/m2an/2020090 | |
5. | Patrik Knopf, Andrea Signori, On the nonlocal Cahn–Hilliard equation with nonlocal dynamic boundary condition and boundary penalization, 2021, 280, 00220396, 236, 10.1016/j.jde.2021.01.012 | |
6. | Harald Garcke, Patrik Knopf, Sema Yayla, Long-time dynamics of the Cahn–Hilliard equation with kinetic rate dependent dynamic boundary conditions, 2022, 215, 0362546X, 112619, 10.1016/j.na.2021.112619 | |
7. | Keiichiro Kagawa, Mitsuharu Ôtani, The time-periodic problem of the viscous Cahn–Hilliard equation with the homogeneous Dirichlet boundary condition, 2023, 25, 1661-7738, 10.1007/s11784-022-01044-6 |