Processing math: 100%

Non-local multi-class traffic flow models

  • Received: 01 August 2018 Revised: 01 October 2018
  • Primary: 35L65, 90B20; Secondary: 65M08

  • We prove the existence for small times of weak solutions for a class of non-local systems in one space dimension, arising in traffic modeling. We approximate the problem by a Godunov type numerical scheme and we provide uniform ${{\mathbf{L}}^\infty } $ and BV estimates for the sequence of approximate solutions, locally in time. We finally present some numerical simulations illustrating the behavior of different classes of vehicles and we analyze two cost functionals measuring the dependence of congestion on traffic composition.

    Citation: Felisia Angela Chiarello, Paola Goatin. Non-local multi-class traffic flow models[J]. Networks and Heterogeneous Media, 2019, 14(2): 371-387. doi: 10.3934/nhm.2019015

    Related Papers:

    [1] Martina Grifoni, Francesca Pedron, Gianniantonio Petruzzelli, Irene Rosellini, Meri Barbafieri, Elisabetta Franchi, Roberto Bagatin . Assessment of repeated harvests on mercury and arsenic phytoextraction in a multi-contaminated industrial soil. AIMS Environmental Science, 2017, 4(2): 187-205. doi: 10.3934/environsci.2017.2.187
    [2] Ioannis Panagopoulos, Athanassios Karayannis, Georgios Gouvalias, Nikolaos Karayannis, Pavlos Kassomenos . Chromium and nickel in the soils of industrial areas at Asopos river basin. AIMS Environmental Science, 2016, 3(3): 420-438. doi: 10.3934/environsci.2016.3.420
    [3] Maja Radziemska, Agnieszka Bęś, Zygmunt M. Gusiatin, Jerzy Jeznach, Zbigniew Mazur, Martin Brtnický . Novel combined amendments for sustainable remediation of the Pb-contaminated soil. AIMS Environmental Science, 2020, 7(1): 1-12. doi: 10.3934/environsci.2020001
    [4] Mahidin, Asri Gani, Saiful, Muhammad Irham, Wulan Windari, Erdiwansyah . An overview of the potential risks, sources, and analytical methods for microplastics in soil. AIMS Environmental Science, 2022, 9(2): 185-216. doi: 10.3934/environsci.2022013
    [5] M.A. Rahim, M.G. Mostafa . Impact of sugar mills effluent on environment around mills area. AIMS Environmental Science, 2021, 8(1): 86-99. doi: 10.3934/environsci.2021006
    [6] Tammy M. Milillo, Gaurav Sinha, Joseph A. Gardella Jr. . Determining site-specific background level with geostatistics for remediation of heavy metals in neighborhood soils. AIMS Environmental Science, 2017, 4(2): 323-347. doi: 10.3934/environsci.2017.2.323
    [7] Emitt C. Witt III . Use of lidar point cloud data to support estimation of residual trace metals stored in mine chat piles in the Old Lead Belt of southeastern, Missouri. AIMS Environmental Science, 2016, 3(3): 509-524. doi: 10.3934/environsci.2016.3.509
    [8] Cristina Calderón-Tapia, Edinson Medina-Barrera, Nelson Chuquin-Vasco, Jorge Vasco-Vasco, Juan Chuquin-Vasco, Sebastian Guerrero-Luzuriaga . Exploration of bacterial strains with bioremediation potential for mercury and cyanide from mine tailings in "San Carlos de las Minas, Ecuador". AIMS Environmental Science, 2024, 11(3): 381-400. doi: 10.3934/environsci.2024019
    [9] Santosh Kumar Karn, Xiangliang Pan . Biotransformation of As (III) to As (V) and their stabilization in soil with Bacillus sp. XS2 isolated from gold mine tailing of Xinjiang, China. AIMS Environmental Science, 2016, 3(4): 592-603. doi: 10.3934/environsci.2016.4.592
    [10] Jerry R. Miller, John P. Gannon, Kyle Corcoran . Concentrations, mobility, and potential ecological risks of selected metals within compost amended, reclaimed coal mine soils, tropical South Sumatra, Indonesia. AIMS Environmental Science, 2019, 6(4): 298-325. doi: 10.3934/environsci.2019.4.298
  • We prove the existence for small times of weak solutions for a class of non-local systems in one space dimension, arising in traffic modeling. We approximate the problem by a Godunov type numerical scheme and we provide uniform ${{\mathbf{L}}^\infty } $ and BV estimates for the sequence of approximate solutions, locally in time. We finally present some numerical simulations illustrating the behavior of different classes of vehicles and we analyze two cost functionals measuring the dependence of congestion on traffic composition.



    To Giuseppe Mingione, on the occasion of his 50th birthday, with regard and admiration.

    The aim of this paper is to study a nonlinear and noncoercive parabolic variational inequality with constraint and homogeneous Dirichlet boundary condition. The Lewy-Stampacchia inequality associated with it is addressed. After the first results of H. Lewy and G. Stampacchia [19] concerning inequalities in the context of superharmonic problems, there is by now a large literature concerning the theory of elliptic obstacle problems as well as of elliptic variational inequalities. We refer to [3,16,25] for a classical overview. For a more recent treatment related to nonlinear elliptic operators see also [23]. The obstacle problem for nonlocal and nonlinear operators has been cosidered in [17,26]. An abstract and general version of the Lewy-Stampacchia inequality is given in [13]. Concerning the parabolic case, first existence results related to problems with time independent obstacles have been treated in [20] in the linear case and in [5] for the more general parabolic problems. The case of obstacles functions regular in time has been considered in [2,5]. Existence and regularity theory for solutions of parabolic inequalities involving degenerate operators in divergence form have been established in [4,18]. More recently in [15], the Authors prove Lewy-Stampacchia inequality for parabolic problems related to pseudomonotone type operators. In this paper we study a variational parabolic inequality for noncoercive operators that present singularities in the coeffcients of the lower order terms in the same spirit of [9,12,14].

    Let us state the functional setting and the assumptions on the data.

    Let $ \Omega\subset \mathbb R^N $, $ N \geqslant2 $, be a bounded open Lipschitz domain and let $ \Omega_T: = \Omega \times (0, T) $ be the parabolic cylinder over $ \Omega $ of height $ T > 0 $. We shall denote by $ \nabla v $ and $ \partial _ t v $ (or $ v_t $) the spatial gradient and the time derivative of a function $ v $ respectively. We consider the class

    $ Wp(0,T):={vLp(0,T,W1,p0(Ω)):vtLp(0,T,W1,p(Ω))},
    $
    (1.1)

    where

    $ 2NN+2<p<N.
    $
    (1.2)

    and $ p^\prime $ is the conjugate exponent of $ p $, i.e., $ \frac 1 p + \frac 1 {p^\prime} = 1 $. In (1.1), $ L^p(0, T, W^{1, p}_0(\Omega)) $ and $ L^{p'}\left(0, T, W^{-1, p'}(\Omega)\right) $ denote parabolic Banach spaces defined according to (2.7).

    Given a measurable function $ \psi \colon \, \Omega_T \cup \Omega \times\{0\} \rightarrow \mathbb{R} $, we are interested in finding functions $ u\colon \Omega_T\rightarrow \mathbb R $ in the convex subset $ \mathcal K_{\psi}(\Omega_T) $ of $ W_p(0, T) $ defined as

    $ Kψ(ΩT):={vWp(0,T):vψa.e. in ΩT}
    $

    and satisfying the following variational inequality

    $ T0ut,vudt+ΩTA(x,t,u,u)(vu)dxdtT0f,vudtvKψ(ΩT),
    $
    (1.3)

    where

    $ fLp(0,T,W1,p(Ω))
    $
    (1.4)

    and $ \langle\cdot, \cdot\rangle $ denotes the duality between $ W^{-1, p'}(\Omega) $ and $ W_0^{1, p}(\Omega) $. The vector field

    $ A=A(x,t,u,ξ):ΩT×R×RNRN
    $

    is a Carathéodory function, i.e., $ A $ measurable w.r.t. $ (x, t)\in \Omega_T $ for all $ (u, \xi)\in \mathbb R \times \mathbb{R}^N $ and continuous w.r.t. $ (u, \xi)\in \mathbb R\times \mathbb{R}^N $ for a.e. $ (x, t)\in \Omega_T $, and such that for a.e. $ (x, t)\in \Omega_T $ and for any $ u\in \mathbb R $ and $ \xi, \eta \in \mathbb{R}^N $,

    $ A(x,t,u,ξ)ξα|ξ|p(b(x,t)|u|)pH(x,t)
    $
    (1.5)
    $ [A(x,t,u,ξ)A(x,t,u,η)](ξη)>0if ξη
    $
    (1.6)
    $ |A(x,t,u,ξ)|β|ξ|p1+(˜b(x,t)|u|)p1+K(x,t)
    $
    (1.7)

    hold true. Here $ \alpha, \beta $ are positive constants, while $ H $, $ K $, $ b $ and $ \tilde b $ are nonnegative measurable functions defined on $ \Omega_T $ such that $ H \in L^1(\Omega_T) $, $ K\in L^{p^\prime}(\Omega_T) $ and

    $ b,˜bL(0,T,LN,(Ω)),
    $
    (1.8)

    where $ L^{N, \infty}(\Omega) $ is the Marcinkiewicz space. For definitions of $ L^{N, \infty}(\Omega) $ and $ L^\infty \left(0, T, L^{N, \infty} (\Omega) \right) $ see Sections 2.2 and 2.3, respectively.

    We assume that the obstacle function fulfills

    $ ψC0([0,T],L2(Ω))Lp(0,T,W1,p(Ω))
    $
    (1.9)
    $ ψ0a.e. in Ω×(0,T)
    $
    (1.10)
    $ ψtLp(ΩT)
    $
    (1.11)
    $ ψ(,0)W1,p0(Ω).
    $
    (1.12)

    For

    $ u0L2(Ω)
    $
    (1.13)

    we impose the following compatibility condition

    $ u0ψ(,0)a.e. in Ω.
    $
    (1.14)

    In the following, we will refer to a function $ u\in \mathcal K_{\psi}(\Omega_T) $ satisfying (1.3) and such that $ u(\cdot, 0) = u_0 $ as a solution to the variational inequality in the strong form with initial value $ u_0 $.

    Under previous assumptions the existence of a solution in the weak form can be proved, see [12]. However the existence of a solution in the sense stated above is not guaranteed even in simpler cases. Then we assume that the source term and the obstacle function are such that

    $ g:=fψt+div A(x,t,ψ,ψ)=g+gwithg+,gLp(0,T,W1,p(Ω))+.
    $
    (1.15)

    Here $ L^{p'}(0, T, W^{-1, p'}(\Omega))^+ $ denotes the non-negative elements of $ L^{p'}(0, T, W^{-1, p'} (\Omega)) $. Following the terminology of [7] or [15], (1.15) is equivalent to say that $ g $ is an element of the order dual $ L^p(0, T, W_0^{1, p}(\Omega))^\ast $ defined as

    $ L^p(0, T, W_0^{1, p}( \Omega))^\ast: = \{ g = g^+-g^- , \, g^\pm \in L^{p'}(0, T, W^{-1, p'}(\Omega))^+ \}. $

    Then, our main result reads as follows

    Theorem 1.1. Let (1.2) and (1.4)–(1.15) be in charge. Assume further that

    $ Db:=distL(0,T,LN,(Ω))(b,L(ΩT))<α1/pSN,p,
    $
    (1.16)

    where $ S_{N, p} = \omega_N^{-1/N}\frac{p }{ N-p } $ and $ \omega_N $ denotes the measure of the unit ball of $ \mathbb{R}^N $. Then, there exists at least a solution $ u \in \mathcal K_{\psi}(\Omega_T) $ of the strong form of the variational inequality (1.3) satisfying $ u(\cdot, 0) = u_0 $. Moreover, the following Lewy-Stampacchia inequality holds

    $ 0tudiv A(x,t,u,u)fg=(ftψ+div A(x,t,ψ,ψ)).
    $
    (1.17)

    In (1.16), $ \mathscr D_b $ denotes the distance of $ b $ from $ L^\infty(\Omega_T) $ in the space $ L^\infty(0, T, L^{N, \infty}(\Omega)) $ defined in (2.8) below.

    Assumptions (1.8) on the coefficients of the lower order terms allow us to consider diffusion models in which the boundedness of the convective field with respect to the spatial variable is too restrictive (see [8]). The corresponding bounded case has been treated in [15].

    We discuss condition (1.16) through an example. It's easy to verify that the operator

    $ A(x, t, u, \xi) = |\xi|^{p-2}\xi+ e^{-t} |u|^{p-2}u \left(\frac \gamma {|x|} +\frac 1 \gamma \arctan |x| \right)^{p-1} \frac x {|x|} $

    satisfies (1.5)–(1.8). According to (2.2) and (2.3) below, we get that

    $ \mathscr D_b = \left(1-\frac 1 p\right)^{1/p} \omega_N^{1/N}\gamma $

    and so (1.16) holds true whenever $ \gamma $ is small enough. On the other hand, we notice that (1.16) does not imply smallness of the norm of the coefficient $ b $. Indeed

    $ \|b\|_{L^\infty(0, T, L^{N, \infty}( \Omega))} \geqslant \frac C \gamma $

    for a constant $ C $ independent of $ \gamma $.

    Theorem 1.1 also applies in the case $ b $ and $ \tilde b $ lie in a functional subspace of weak–$ L^N $ in which bounded functions are dense. For more details see also [10]. For other examples of operators satisfying conditions above we refer to [12].

    We remark that for $ f, \psi_t, \text{div } A(x, t, \psi, \nabla \psi) \in L^{p^\prime}(\Omega_T) $ condition (1.15) is satisfied. Then, Theorem 1.1 is comparable with the existence result of Lemma 3.1 in [4]. In order to prove our result, we consider a sequence of suitable penalization problems for which an existence result holds true (see [12]). Then we are able to construct a solution u to (1.3) as limit of solutions of such problems despite the presence of unbounded coefficients in the lower order terms.

    In this section we provide the notation and several preliminary results that will be fundamental in the sequel.

    The symbol $ C $ (or $ C_1, C_2, \dots $) will denote positive constant, possibly varying from line to line. For the dependence of $ C $ upon parameters, we will simply write $ C = C(\cdot, \dots, \cdot) $. The positive and the negative part of a real number $ z $ will be denoted by $ z^+ $ and $ z^- $, respectively, and are defined by $ z^+: = \max\{z, 0\} $ and $ z^-: = -\min\{z, 0\} $. Given $ z_1, z_2 \in \mathbb R $, we often use the notation $ z_1 \wedge z_2 $ and $ z_1 \vee z_2 $ in place of $ \min\{z_1, z_2\} $ and $ \max\{z_1, z_2\} $ respectively.

    Let $ \Omega $ be a bounded domain in $ \mathbb{R}^N $. For any $ 1 < p < \infty $ and $ 1\leq q < \infty $, the Lorentz space $ L^{p, q}(\Omega) $ is the set of real measurable functions $ f $ on $ \Omega $ such that

    $ fqLp,q:=p0[λf(k)]qpkq1dk<.
    $

    Here $ \lambda_f(k): = \left| \left\{ x\in \Omega: |f(x)| > k \right\} \right| $ is the distribution function of $ f $. When $ p = q $, the Lorentz space $ L^{p, p}(\Omega) $ coincides with the Lebesgue space $ L^p(\Omega) $. When $ q = \infty $, the space $ L^{p, \infty}(\Omega) $ is the set of measurable functions $ f $ on $ \Omega $ such that

    $ fpLp,:=supk>0kpλf(k)<.
    $

    This set coincides with the Marcinkiewicz space weak-$ L^p(\Omega) $. The expressions above do not define a norm in $ L^{p, q} $ or $ L^{p, \infty} $ respectively, in fact triangle inequality generally fails. Nevertheless, they are equivalent to a norm, which make $ L^{p, q}(\Omega) $ and $ L^{p, \infty}(\Omega) $ Banach spaces when endowed with them. An important role in the potential theory is played by these spaces as pointed out in [22].

    For $ 1\leq q < p < r\leq \infty $, the following inclusions hold

    $ L^r ( \Omega)\subset L^{p, q}( \Omega)\subset L^{p, r} ( \Omega) \subset L^{p, \infty}( \Omega)\subset L^q( \Omega). $

    For $ 1 < p < \infty $, $ 1\leq q \leq \infty $ and $ \frac 1 p+\frac 1 {p'} = 1 $, $ \frac 1 q+\frac 1 {q'} = 1 $, if $ f\in L^{p, q}(\Omega) $, $ g\in L^{p', q'}(\Omega) $ we have the Hölder–type inequality

    $ Ω|f(x)g(x)|dxfLp,qgLp,q.
    $
    (2.1)

    Since $ L^\infty(\Omega) $ is not dense in $ L^{p, \infty}(\Omega) $, for $ f \in L^{p, \infty} (\Omega) $ in [6] the Authors stated the following

    $ distLp,(Ω)(f,L(Ω)):=infgL(Ω)fgLp,(Ω).
    $
    (2.2)

    As already observed in [10,11], we have

    $ distLp,(Ω)(f,L(Ω))=limm+fχ{|f|>m}Lp,
    $
    (2.3)

    and

    $ distLp,(Ω)(f,L(Ω))=limm+fTmfLp,,
    $

    where, for all $ m > 0 $, $ \mathcal T_m $ is the truncation operator at levels $ \pm m $, i.e.,

    $ Tmy:=min{m,max{m,y}}for yR.
    $
    (2.4)

    Another useful estimate is provided by the following sort of triangle inequality

    $ f+εgLp,(1+ε)fLp,+ε(1+ε)gLp,
    $
    (2.5)

    which holds true for $ f, g\in L^{p, \infty}(\Omega) $ and $ \varepsilon > 0 $.

    For $ 1\leq q < \infty $, any function in $ L^{p, q}(\Omega) $ has zero distance to $ L^\infty(\Omega) $. Indeed, $ L^\infty(\Omega) $ is dense in $ L^{p, q}(\Omega) $, the latter being continuously embedded into $ L^{p, \infty}(\Omega) $.

    Assuming that $ 0\in\Omega $, $ b(x) = \gamma/|x| $ belongs to $ L^{N, \infty}(\Omega) $, $ \gamma > 0 $. For this function, we have

    $ distLN,(Ω)(b,L(Ω))=γω1/NN.
    $

    The Sobolev embedding theorem in Lorentz spaces [1,24] reads as

    Theorem 2.1. Let us assume that $ 1 < p < N $, $ 1\leq q\leq p $, then every function $ u\in W_0^{1, 1}(\Omega) $ verifying $ |\nabla u|\in L^{p, q}(\Omega) $ actually belongs to $ L^{p^*, q}(\Omega) $, where $ p^*: = \frac{Np}{N-p} $ is the Sobolev conjugate exponent of $ p $ and

    $ uLp,qSN,puLp,q,
    $
    (2.6)

    where $ S_{N, p} $ is the Sobolev constant given by $ S_{N, p} = \omega_N^{-1/N} \frac{p }{ N-p } $.

    Let $ T > 0 $ and $ X $ be a Banach space endowed with a norm $ \|\cdot\|_X $. Then, the space $ L^p\left(0, T, X\right) $ is defined as the class of all measurable functions $ u\colon [0, T] \rightarrow X $ such that

    $ uLp(0,T,X):=(T0u(t)pXdt)1/p<
    $
    (2.7)

    whenever $ 1\leq p < \infty $, and

    $ \|u\|_{ L^\infty\left(0, T, X\right)} : =\mathop {{\rm{ess}}\;{\rm{sup}}}\limits_{0 < t < T } \|u(t)\| _X < \infty $

    for $ p = \infty $. The space $ C^0\left([0, T], X\right) $ represents the class of all continuous functions $ u\colon [0, T] \rightarrow X $ with the norm

    $ \|u\|_{ C^0\left([0, T], X\right)} : = \max\limits_{0\leq t \leq T} \|u(t)\|_X. $

    We essentially consider the case where $ X $ is either a Lorentz space or Sobolev space $ W^{1, p}_0(\Omega) $. This space will be equipped with the norm $ \|g\| _{W^{1, p}_0(\Omega)} : = \| \nabla g \| _{L^p(\Omega)} $ for $ g\in W^{1, p}_0(\Omega) $.

    For $ f\in L^\infty(0, T, L^{p, \infty}(\Omega)) $ we define

    $ distL(0,T,Lp,(Ω))(f,L(ΩT))=infgL(ΩT)fgL(0,T,Lp,(Ω))
    $
    (2.8)

    and as in (2.3) we find

    $ distL(0,T,Lp,(Ω))(f,L(ΩT))=limm+fχ{|f|>m}L(0,T,Lp,(Ω)).
    $
    (2.9)

    In the class $ W_p(0, T) $ defined in (1.1) and equipped with the norm

    $ \| u \|_{W_p(0, T)} : = \| u \|_{L^p(0, T, W^{1, p}( \Omega))} + \| u_t \|_{ L^{p'}(0, T, W^{-1, p'}(\Omega))}, $

    the following inclusion holds (see [27,Chapter III, page 106]).

    Lemma 2.2. Let $ p > 2N /(N+2) $. Then $ W_p(0, T) $ is contained into the space $ C^0\left([0, T], L^2(\Omega)\right) $ and any function $ u \in W_p(0, T) $ satisfies

    $ \| u \|_{ C^0\left([0, T], L^2(\Omega)\right)} \leq C \| u \|_{W_p(0, T)} $

    for some constant $ C > 0 $.

    Moreover, the function $ t\in[0, T]\mapsto\|u(\cdot, t)\|^2_{L^2(\Omega)} $ is absolutely continuous and

    $ 12ddtu(,t)2L2(Ω)=ut(,t),u(,t)for a.e. t[0,T].
    $

    The compactness result due to Aubin–Lions reads as follows.

    Lemma 2.3. Let $ X_0, X, X_1 $ be Banach spaces with $ X_0 $ and $ X_1 $ reflexive. Assume that $ X_0 $ is compactly embedded into $ X $ and $ X $ is continuously embedded into $ X_1 $. For $ 1 < p, q < \infty $ let

    $ W: = \{u \in L^p(0, T, X_0)\colon \partial_t u \in L^q(0, T, X_1)\} . $

    Then $ W $ is compactly embedded into $ L^p(0, T, X) $.

    As an example, we choose $ q = p' $, $ X_0 = W^{1, p}_0(\Omega) $, $ X_1 = W^{-1, p^\prime} (\Omega) $ and $ X = L^p(\Omega) $ if $ p\geq 2 $ or $ X = L^2(\Omega) $ for $ \frac {2N}{N+2} < p < 2 $. Therefore, we deduce

    Lemma 2.4. If $ p > 2N /(N+2) $ then $ W_p(0, T) $ is compactly embedded into $ L^p(\Omega_T) $ and into $ L^2(\Omega_T) $.

    Let $ \delta > 0 $. We introduce the following initial–boundary value problem

    $ {tuδdiv [A(x,t,max{uδ,ψ},uδ)]=1δ[(ψuδ)+]q1+fin ΩT,uδ=0on Ω×(0,T),uδ(,0)=u0in Ω,
    $
    (3.1)

    where

    $ q: = \min\{2, p\}. $

    Moreover, in this section we assume that

    $ ψ0a.e. in ΩT.
    $
    (3.2)

    We introduce the notation

    $ \tilde A(x, t, w, \xi): = A(x, t, \max \{w, \psi\}, \xi). $

    By the elementary inequality

    $ |aa||a|aRa(,0]
    $
    (3.3)

    and recalling (1.5), (1.6) and (1.7), we easily deduce

    $ ˜A(x,t,u,ξ)ξα|ξ|p(b(x,t)|u|)pH(x,t)[˜A(x,t,u,ξ)˜A(x,t,u,η)](ξη)>0if ξη|˜A(x,t,u,ξ)|β|ξ|p1+(˜b(x,t)|u|)p1+K(x,t)
    $

    for a.e. $ (x, t)\in \Omega_T $ and for any $ u\in \mathbb R $ and $ \xi, \eta \in \mathbb{R}^N $.

    For $ u_0 \in L^2(\Omega) $ and $ f \in L^{p'}(0, T, W^{-1, p'}(\Omega)) $, a solution to problem (3.1) is a function

    $ u_\delta \in C^0\left([0, T], L^2(\Omega)\right) \cap L^p(0, T, W^{1, p}_0(\Omega)) $

    such that

    $ ΩTuδφtdxds+ΩT˜A(x,s,uδ,uδ)φdxds=1δΩT[(ψuδ)+]q1φdxds+Ωu0φ(x,0)dx+T0f,φds
    $

    for every $ \varphi \in C^\infty(\bar \Omega_T) $ such that $ {\rm supp }\, \varphi \subset [0, T)\times \Omega $.

    By using the elementary inequality

    $ (a+a)θaθ+aθa,a[0,+)θ(0,1)
    $

    and Young inequality we see that

    $ p < 2 \quad \Longrightarrow \quad [ ( \psi-u)^+]^{p-1} \leqslant |\psi|^{p-1} + |u|^{p-1} \leqslant (p-1) \left(|u| +|\psi|\right) +2(2-p) . $

    Hence, by Theorem 4.2 and Remark 4.5 in [12] we get the following existence result.

    Proposition 3.1. Let (1.2), (1.4)–(1.16) and (3.2) be in charge. For every fixed $ \delta > 0 $, problem (3.1) admits a solution $ u_{\delta} \in C^0\left([0, T], L^2(\Omega)\right) \cap L^p(0, T, W^{1, p}_0(\Omega)) $.

    The arguments of [12] lead to some estimates for the sequence $ \{u_\delta\}_{\delta > 0} $. We propose here a proof that carefully keeps trace of the constants in the estimates.

    Lemma 3.2. Let (1.2), (1.4)–(1.16) and (3.2) be in charge. Any solution $ u_\delta \in C^0\left([0, T], L^2(\Omega)\right) \cap L^p(0, T, W^{1, p}_0(\Omega)) $ to problem (3.1) satisfies the following estimate

    $ uδ2L(0,T,L2(Ω))+uδpLp(ΩT)C(b,N,p,α)[u02L2(Ω)+fpLp(0,T,W1,p(Ω))+HL1(ΩT)+(u02L2(Ω)+fpLp(0,T,W1,p(Ω))+bpLp(ΩT))pbpLp(ΩT)].
    $
    (3.4)

    Proof. We fix $ t \in (0, T) $ and we set $ \Omega_t: = \Omega\times (0, t) $. We choose $ \varphi: = \mathcal T_1(u_\delta)\chi_{(0, t)} $ as a test function. If we let $ \Phi(z): = \int_0^z \mathcal T_1(\zeta) \, \mathrm d \zeta $ for $ z\in\mathbb R $, we have

    $ ΩΦ(uδ(x,t))dx+Ωt˜A(x,s,uδ,uδ)T1(uδ)dxds=1δΩt[(ψuδ)+]q1T1(uδ)dxds+ΩΦ(u0)dx+t0f,T1(uδ)ds.
    $

    Assumption (3.2) implies that $ [(\psi- u_{\delta})^+]^{q-1} \mathcal T_1(u_\delta) \leqslant 0 $ a.e. in $ \Omega_T $, so we have

    $ ΩΦ(uδ(x,t))dx+Ωt{|uδ|1}˜A(x,s,uδ,uδ)uδdxdsΩΦ(u0(x,0))dx+t0f,T1(uδ)ds.
    $

    By (1.5) and (1.7) we deduce

    $ ΩΦ(uδ(x,t))dx+αΩt{|uδ|1}|uδ|pdxdsΩΦ(u0)dx+t0f,T1(uδ)ds+Ωt{|uδ|1}(b|uδψ|)pdxds+Ωt{|uδ|1}Hdxds.
    $
    (3.5)

    Now, as $ 0 \leqslant \Phi(z) \leqslant \frac {z^2} 2 $ for all $ z\in \mathbb R $, we have

    $ ΩΦ(u0)dx12u02L2(Ω).
    $
    (3.6)

    By Hölder and Young inequality we get

    $ t0f,T1(uδ)dsfLp(0,T,W1,p(Ω))T1(uδ)Lp(Ωt)=fLp(0,T,W1,p(Ω))(Ωt{|uδ|1}|(uδ)|pdxds)1/pα2Ωt{|uδ|1}|uδ|pdxds+C(α,p)fpLp(0,T,W1,p(Ω)).
    $
    (3.7)

    Finally, by (3.3)

    $ Ωt{|uδ|1}(b|uδψ|)pdxdsΩt{|uδ|1}(b|uδ|)pdxdsbpLp(ΩT).
    $
    (3.8)

    Gathering (3.6), (3.7), and (3.8) and using Hölder inequality, by (3.5) we have

    $ ΩΦ(uδ(x,t))dxM0,
    $

    where

    $ M0:=C(N,p,α)[u02L2(Ω)+fpLp(0,T,W1,p(Ω))+bpLp(ΩT).]
    $
    (3.9)

    It is easily seen that

    $ |u|2Φ(u) for |u|1
    $

    and so

    $ sup0<t<T|{xΩ:|uδ(x,t)|>k}|C(N,p,α,β)M0kk1.
    $
    (3.10)

    We fix $ t \in (0, T) $ and choose $ \varphi: = u_\delta \chi_{(0, t)} $ as a test function in (3.1). Again, assumption (3.2) implies that $ [(\psi- u_{\delta})^+]^{q-1} u_\delta \leqslant 0 $ a.e. in $ \Omega_T $, then

    $ 12uδ(,t)2L2(Ω)+Ωt˜A(x,s,uδ,uδ)uδdxds12u02L2(Ω)+t0f,uδds.
    $

    By Young inequality for $ \varepsilon > 0 $

    $ t0f,uδdsεΩt|uδ|pdxds+p1ppε1pfpLp(0,T,W1,p(Ω)).
    $

    Then, by (1.5) we further have

    $ uδ(,t)2L2(Ω)+αΩt|uδ|pdxdsu02L2(Ω)+εΩt|uδ|pdxds+C(ε,p)fpLp(0,T,W1,p(Ω))+Ωt(b|uδψ|)pdxds+ΩtHdxds.
    $
    (3.11)

    For $ m > 0 $ to be chosen later, we have from (3.3)

    $ Ωt(b|uδψ|)pdxdsΩt(b|uδ|)pdxds=Ωt(bχ{bm}|uδ|)pdxds+Ωt(bχ{b>m}|uδ|)pdxds.
    $
    (3.12)

    We estimate separately the two terms in the right–hand side of (3.12). For $ k > 1 $ fixed, we obtain

    $ Ωt(bχ{bm}|uδ|)pdxdsmpt0ds{|uδ(,s)|>k}|uδ|pdx+kpt0dsΩb(x,s)pdx.
    $
    (3.13)

    Now we apply Hölder inequality (2.1), estimates (2.6) and (3.10) to get

    $ t0ds{|uδ(,s)|>k}|uδ|pdx=t0dsΩ|uδχ{|uδ(,s)|>k}|pdxt0χ{|uδ(,s)|>k}pLN,(Ω)uδpLp,p(Ω)dsSpN,pMp/N0kp/NΩt|uδ|pdxds,
    $
    (3.14)

    where $ M_0 $ is the constant in (3.9). On the other hand, using again Hölder inequality (2.1) and estimate (2.6)

    we have

    $ Ωt(bχ{b>m}|uδ|)pdxdsSpN,pbχ{b>m}pL(0,T,LN,(Ω))Ωt|uδ|pdxds.
    $
    (3.15)

    Inserting (3.13), (3.14) and (3.15) into (3.12) we obtain

    $ Ωt(b|uδψ|)pdxds[mpSpN,pMp/N0kp/N+SpN,pbχ{b>m}pL(0,T,LN,(Ω))]uδpLp(Ωt)+kpt0dsΩb(x,s)pdx.
    $
    (3.16)

    Observe that (3.11) and (3.16) imply

    $ 12uδ(,t)2L2(Ω)+αuδpLp(Ωt)12u02L2(Ω)+kpbpLp(ΩT)+p1ppε1pfpLp(0,T,W1,p(Ω))+HL1(ΩT)+[ε+mpSpN,pMp/N0kp/N+SpN,pbχ{b>m}pL(0,T,LN,(Ω))]uδpLp(Ωt).
    $

    Now we choose $ m > 0 $ so large to guarantee

    $ SpN,pbχ{b>m}pL(0,T,LN,(Ω))<α.
    $

    The existence of such a value of $ m $ is a direct consequence of (1.16) and the characterization of distance in (2.9). It is also clear that $ m $ is a positive constant depending only on $ b $, $ N $, $ p $ and $ \alpha $. So we get

    $ 12uδ(,t)2L2(Ω)+α1uδpLp(Ωt)12u02L2(Ω)+kpbpLp(ΩT)+p1ppε1pfpLp(0,T,W1,p(Ω))+HL1(ΩT)+[ε+mpSpN,pMp/N0kp/N]uδpLp(Ωt)
    $

    for some $ \alpha_1 = \alpha_1(b, N, p, \alpha) $. We may also choose $ \varepsilon = \frac{\alpha_1}{2} $. Then the latter relation becomes

    $ 12uδ(,t)2L2(Ω)+α12uδpLp(Ωt)12u02L2(Ω)+kpbpLp(ΩT)+C1(b,N,p,α)fpLp(0,T,W1,p(Ω))+HL1(ΩT)+C2(b,N,p,α)(M0k)p/NuδpLp(Ωt).
    $

    We choose $ k = M_0 \left(\frac{\alpha_1}{4 C_2} \right)^{N/p} $ so that $ C_2\left(\frac {M_0} k\right)^{p/N} = \frac {\alpha_1} 4 $ and therefore

    $ 12uδ(,t)2L2(Ω)+α14uδpLp(Ωt)12u02L2(Ω)+C3(b,N,p,α)Mp0bpLp(ΩT)+C1(b,N,p,α)fpLp(0,T,W1,p(Ω))+HL1(ΩT).
    $

    Taking into account the definition of $ M_0 $, the latter leads to the estimate (3.4).

    Lemma 3.3. Let (1.2), (1.4)–(1.16) and (3.2) be in charge. Assume further that $ g^- $ defined in (1.15) is such that

    $ gLq(ΩT).
    $
    (3.17)

    Then, for every $ \delta > 0 $, every solution $ u_\delta $ of problem (3.1) satisfies

    $ (uδψ)q1Lq(ΩT)δgLq(ΩT).
    $
    (3.18)

    Moreover, there exists a positive constant C depending only on the data and independent on $ \delta $ such that

    $ tuδLp(0,T;W1,p(Ω))C.
    $
    (3.19)

    Proof. We use the function $ \phi = (\psi-u_\delta)^+ $ as a test function in the equation of Problem (3.1). Then, we get

    $ T0tuδ,(ψuδ)+dt+ΩTA(x,t,max{uδ,ψ},u)(ψuδ)+dxdt=1δΩT[(ψuδ)+]qdxdt+T0f,(ψuδ)+dt.
    $

    Recalling (1.15), this implies

    $ 1δΩT[(ψuδ)+]qdxdt=ΩTg(ψuδ)+dxdtT0g+,(ψuδ)+dtT0t(ψuδ),(ψuδ)+dtΩT{ψ>uδ}[A(x,t,ψ,ψ)A(x,t,ψ,uδ)](ψuδ)dxdt.
    $

    By (1.14) we observe that

    $ \int_0^T \langle \partial _t (\psi-u_\delta), (\psi-u_\delta )^+ \rangle\, dt = \frac12 \| (u_\delta -\psi)^- (T)\|^{2}_{L^2( \Omega) } $

    hence, by (1.6) we get

    $ 1δΩT[(ψuδ)+]qΩTg(ψuδ)+dxdt.
    $

    Then, using Hölder inequality and dividing both sides of the inequality by $ \| (\psi- u_\delta)^+)\|_{L^q((\Omega_T)} $ we obtain (3.18). To obtain (3.19) we fix $ \varphi \in L^p(0, T; W_0^{1, p}(\Omega)) $ and then we observe that

    $ |T0tuδ,φdt|(A(,,max{uδ,ψ},uδ)Lp(ΩT)+fLp(ΩT))φLp(0,T;W1,p0(Ω))+1δ(ψuδ)+q1Lq(ΩT)φLq(ΩT).
    $

    At this point we observe that the definition of $ q $ and Holder inequality imply

    $ \|\varphi\|_{ L^q ( \Omega_T)} \leq C(p, | \Omega|, T) \|\varphi\|_{ L^p ( \Omega_T)} . $

    Finally, using (3.18) and Poncaré inequality slicewise, we conclude that

    $ |T0tuδ,φdt|C(p,|Ω|,T)φLp(0,T;W1,p0(Ω)),
    $

    where $ C $ is a positive constant independent of $ \delta. $ This immediately leads to (3.19).

    We proceed step by step. We first prove the result under regularity assumptions on $ g $ and sign conditon (3.2) on the obstacle function $ \psi $. Then we address the general case.

    Proposition 4.1. Let (1.2), (1.4)–(1.16), (3.2) and (3.17) be in charge. There exists at least solution $ u \in \mathcal K_{\psi}(\Omega_T) $ to the variational inequality (1.3) such that $ u(\cdot, 0) = u_0 $ in $ \Omega $ and satisfying the following estimate

    $ u2L(0,T,L2(Ω))+upLp(ΩT)C(b,N,p,α)[u02L2(Ω)+fpLp(0,T,W1,p(Ω))+HL1(ΩT)+(u02L2(Ω)+fpLp(0,T,W1,p(Ω))+bpLp(ΩT))pbpLp(ΩT)].
    $
    (4.1)

    Proof. By Proposition 3.1, for every $ \delta > 0 $ there exists a solution $ u_{\delta} \in C^0\left([0, T], L^2(\Omega)\right) \cap L^p(0, T, W^{1, p}_0(\Omega)) $ to problem (3.1) satisfying (3.4). Hence we have that, by Lemma 3.3 and Lemma 2.2, there exists $ u\in C^0\left([0, T], L^2(\Omega)\right) \cap L^p(0, T, W^{1, p}_0(\Omega)) $ such that

    $ uδustrongly in Lp(ΩT)
    $
    (4.2)
    $ uδuweakly in Lp(ΩT,RN)
    $
    (4.3)
    $ uδuweakly  in L(0,T;L2(Ω))tuδtuweakly in Lp(0,T,W1,p(Ω))
    $

    as $ \delta \rightarrow 0^+ $. By semicontinuity, (3.4) implies (4.1)

    We claim that the limit function $ u $ solves the variational inequality (1.3) in the strong form.

    It is immediate to check that

    $ u(,0)=u0a.e. in Ω,
    $
    (4.4)
    $ uψ a.e. inΩT.
    $
    (4.5)

    Indeed, (4.4) holds since $ u_\delta(\cdot, 0) = u_0 $ a.e. in $ \Omega $ for every $ \delta > 0 $. On the other hand, if we pass to the limit as $ \delta\rightarrow 0^+ $ in (3.18) and take into account (4.2) we have $ \|(u-\psi)^-\|_{L^{ 2 \wedge p } (\Omega_T)} = 0 $ which clearly implies (4.5).

    Our next goal is to prove that

    $ uδua.e. in ΩT
    $
    (4.6)

    as $ \delta \rightarrow 0^+ $. We test the penalized equation by $ \mathcal T_1(u_\delta-u) $ and since condition (4.5) implies

    $ \int_{ \Omega_T} [ ( \psi- u_{\delta})^+]^{q-1} \mathcal T_1(u_\delta-u) \, \mathrm d x \, \mathrm d t \leqslant 0 $

    we get the following inequality

    $ T0tuδ,T1(uδu)dt+ΩTA(x,t,uδψ,uδ)T1(uδu)dzT0f,T1(uδu)dt.
    $
    (4.7)

    If we set $ \Phi(z): = \int_0^z \mathcal T_1(\zeta) \, \mathrm d \zeta $, by (4.4) we obtain

    $ \int_0^T \langle \partial_t u_\delta , \mathcal T_1(u_\delta-u) \rangle dt = \int_{ \Omega}\Phi(u_\delta-u)(x, T) \, \mathrm d x + \int_0^T \langle \partial_t u , \mathcal T_1(u_\delta-u) \rangle dt. $

    Because of (4.3), the latter term in the last inequality vanishes in the limit as $ \delta \rightarrow 0 $. So, as $ \Phi $ is nonnegative, we get

    $ \limsup\limits_{\delta\rightarrow 0} \int_0^T \langle \partial_t u_\delta , \mathcal T_1(u_\delta-u) \rangle dt \geqslant 0. $

    Again by (4.3), the right hand side of (4.7) vanishes in the limit as $ \delta \rightarrow 0 $, and so (4.7) implies

    $ lim supδ0ΩT{|uδu|1}A(x,t,uδψ,uδ)(uδu)dxdt0.
    $
    (4.8)

    By (1.7), (3.2) and (3.3) we have

    $ |A(x,t,uδψ,u)|χ{|uδu|1}β|u|p1+(˜b|uδ|)p1χ{|uδu|1}+Kβ|u|p1+C(p)˜bp1+C(p)(˜b|u|)p1+K
    $

    therefore, by the dominated convergence theorem and by (4.2), we get

    $ limδ0ΩT{|uδu|1}A(x,t,uδψ,u)(uδu)dxdt=0.
    $
    (4.9)

    Combining (4.8) and (4.9) and by (1.6) we get

    $ limδ0ΩT[A(x,t,uδψ,uδ)A(x,t,uδψ,u)]T1(uδu)dxdt=0.
    $
    (4.10)

    Using again (1.6), relation (4.10) gives

    $ \left[ A(x, t, u_\delta\vee \psi, \nabla u_\delta) - A(x, t, u_\delta\vee \psi, \nabla u ) \right] \cdot\nabla (u_\delta-u) \chi_{ \{ | u_\delta-u | \leqslant 1 \}} \rightarrow 0 \qquad \text{a.e. in } \Omega_T $

    and so by (4.2) we get

    $ \left[ A(x, t, u_\delta\vee \psi, \nabla u_\delta) - A(x, t, u_\delta\vee \psi, \nabla u ) \right] \cdot\nabla (u_\delta-u) \rightarrow 0 \qquad \text{a.e. in } \Omega_T $

    as $ \delta \rightarrow 0 $. By Lemma 3.1 in [21] we deduce that (4.6) holds.

    We let $ v \in \mathcal K_{\psi}(\Omega_T) $. It is clear that $ \left[(\psi-u_\delta)^+\right]^{q-1} \mathcal T_\lambda(u_\delta-v) \leqslant 0 $ a.e. in $ \Omega_T $ and for every $ \lambda > 0 $. For this reason, if we use $ \mathcal T_\lambda(u_\delta-v) $ as a test function in (3.1) we deduce

    $ \int_0^T \langle \partial _t u_\delta, \mathcal T_\lambda(u_\delta-v)\rangle \, \mathrm d t +\int_{ \Omega_T} \left[A(x, t, u_\delta\vee \psi,\\ \nabla u_\delta)-A(x, t, u_\delta\vee \psi, \nabla v) \right]\cdot \nabla\mathcal T_\lambda(u_\delta-v) \, \mathrm d x \, \mathrm d t \\ \leqslant \int_0^T \langle f, \mathcal T_\lambda(u_\delta-v)\rangle \, \mathrm d t -\int_{ \Omega_T} A(x, t, u_\delta\vee \psi, \nabla v)\cdot \nabla\mathcal T_\lambda(u_\delta-v) \, \mathrm d x \, \mathrm d t. $ (4.11)

    We set $ \Phi_\lambda(z): = \int_0^z \mathcal T_\lambda(\zeta) \, \mathrm d \zeta $ and we have

    $ \int_0^T \langle \partial _t u_\delta, \mathcal T_\lambda(u_\delta-v)\rangle \, \mathrm d t =\\ \int_0^T \langle \partial _t v, \mathcal T_\lambda(u_\delta-v)\rangle \, \mathrm d t + \int_0^T \langle \partial _t u_\delta-\partial_t v, \mathcal T_\lambda(u_\delta-v)\rangle \, \mathrm d t \\ = \int_0^T \langle \partial _t v, \mathcal T_\lambda(u_\delta-v)\rangle \, \mathrm d t + \int_ \Omega \Phi_\lambda(u_\delta-v)(x, T) \, \mathrm d x - \int_ \Omega \Phi_\lambda(u_0-v (x, 0)) \, \mathrm d x. $ (4.12)

    We observe that Lemma 2.2 applies because of (3.4) and (3.19), so

    $ u_\delta(\cdot, t)\rightharpoonup u(\cdot, t) \quad \text{weakly in } L^2( \Omega) \text{ for all }t \in [0, T] . $

    This convergence and the Lipschitz continuity of $ \Phi_\lambda $ gives $ \Phi_\lambda(u_\delta-v)(\cdot, T) \rightharpoonup \Phi_\lambda(u-v)(\cdot, T) $ weakly in $ L^2(\Omega) $, then

    $ limδ0ΩΦλ(uδv)(x,T)dx=ΩΦλ(uv)(x,T)dx.
    $
    (4.13)

    On the other hand, by Fatou lemma, we are able to pass to the limit as $ \delta\rightarrow 0 $ in the third term on the left–hand side of (4.11). Indeed, for this term we know by the monotonicity condition (1.6) that the integrand is nonnegative and we have already observed that $ u_\delta $ and $ \nabla u_\delta $ converge a.e. according to (4.2) and (4.6) respectively. We only need to handle the term

    $ \int_{ \Omega_T} A(x, t, u_\delta \vee \psi, \nabla v)\cdot \nabla\mathcal T_\lambda(u_\delta-v) \, \mathrm d x \, \mathrm d t. $

    This can be done arguing similarly as for the case $ \lambda = 1 $. By (1.7) we have

    $ |A(x, t, u_\delta\vee \psi, \nabla v)|\, \chi_{\{|u_\delta-v| \leqslant \lambda\}} \leqslant \beta|\nabla v|^{p-1}+K+C(p) \lambda^{p-1} \left(\tilde b^{p-1} + (\tilde b|v|)^{p-1}\right). $

    By (4.2) and (4.5) we obtain $ A(x, t, u_\delta \vee \psi, \nabla v)\to A(x, t, u, \nabla v) $ a.e. in $ \Omega_T $, Therefore, by the dominated convergence theorem, $ A(x, t, u_\delta \vee \psi, \nabla v)\to A(x, t, u, \nabla v) $ strongly in $ L^{p^\prime}(\Omega_T, \mathbb{R}^N) $, and this yields

    $ \lim\limits_{\delta\rightarrow 0}\int_{ \Omega_T} A(x, t, u_\delta, \nabla v)\cdot \nabla\mathcal T_\lambda(u_\delta-v) \, \mathrm d x \, \mathrm d t = \int_{ \Omega_T} A(x, t, u, \nabla v)\cdot \nabla\mathcal T_\lambda(u-v) \, \mathrm d x \, \mathrm d t. $

    Taking into account the latter relation and also (4.12) and (4.13), we can now pass to the limit as $ \delta\rightarrow 0 $ in (4.11) and obtain

    $ T0tv,Tλ(uv)dt+ΩΦλ(uv)(x,T)dxΩΦλ(u0v(x,0))dx+ΩTA(x,t,u,u)Tλ(uv)dxdtT0f,Tλ(uv)dt.
    $

    Since

    $ Tλ(uv)uvstrongly in Lp(0,T,W1,p0(Ω)) as λ,Φλ(uv)(,T)12|u0v(,0)|2strongly in L1(Ω) as λΦλ(u0v(,0))12|u(,0)v(,0)|2strongly in L1(Ω) as λ
    $

    and also observing that

    $ \int_0^T \langle \partial _t v, u -v \rangle \, \mathrm d t = \int_0^T \langle \partial _t u, u -v \rangle \, \mathrm d t + \frac12\int_\Omega |u_0-v(\cdot, 0)|^2 \, \mathrm d x - \frac12\int_\Omega |u(\cdot, T)-v(\cdot, T)|^2 \, \mathrm d x $

    we conclude that (1.3) holds.

    Next result shows that a Lewy–Stampacchia inequality can be derived under some suitable assuption, that we are going to remove later.

    Proposition 4.2. Let (1.2), (1.4)–(1.16), (3.2) and (3.17) be in charge. If we also assume that

    $ gLp(ΩT)Lp(0,T,W1,p0(Ω))g0a.e. in ΩTtgLq(ΩT)
    $

    the solution $ u $ of the obstacle problem constructed in Proposition 4.1 satisfies the Lewy–Stampacchia inequality (1.17).

    Proof. We define

    $ z_\delta: = g^- - \frac 1 \delta \left[ (\psi-u_\delta)^+ \right]^{q-1}. $

    For $ k \geqslant 1 $ we also define

    $ ηk(y):=(q1)y+0min{k,sq2}dsΨk(x,t,λ):=(g1δηk(λ))Λk(x,t,λ):=λ0Ψk(x,t,σ)dσ.
    $

    Thanks to Lemma 4.3 in [15] we are able to test (3.1) by $ \Psi_k(x, s, u_\delta-\psi) \chi_{(0, t)} $ for $ t\in (0, T) $, obtaining

    $ ΩttΛk(x,s,uδψ)dxds+ΩΛk(x,t,(uδψ)(x,t))dxΩΛk(x,0,(uδψ)(x,0))dxΩt[A(x,s,uδψ,uδ)A(x,s,ψ,ψ)](g1δηk((uδψ)))dxdsΩtzδ(g1δηk((uδψ)))dxds=t0g+,(g1δηk((uδψ)))ds0.
    $
    (4.14)

    By (1.14) we have

    $ \int_{ \Omega} \Lambda_k (x, 0, (u_\delta-\psi)(x, 0)) \, \mathrm d x = 0. $

    We also have

    $ ΩttΛk(x,s,uδψ)dxds=Ωttguδψ0χ{g1δηk(τ)<0}dτdxds=Ωttg(uδψ)0χ{g1δηk(τ)<0}dτdxdsΩt|tg||(uδψ)|dxds.
    $

    So, taking into account (4.14), we have

    $ - \int_{ \Omega_t} | \partial_t g^- | |(u_\delta-\psi)^-| \, \mathrm d x \, \mathrm d s + \\ \int_{ \Omega} \Lambda_k (x, t, (u_\delta-\psi)(x, t)) \, \mathrm d x - \int_{ \Omega_t} z_\delta \left(g^- - \frac 1 \delta \eta_k\left((u_\delta-\psi)^-\right) \right)^- \, \mathrm d x \, \mathrm d s \\ -\int_{ \Omega_t} \left[A(x, s, u_\delta\vee \psi, \nabla u_\delta)-A(x, s, \psi, \nabla \psi) \right]\cdot \nabla \left(g^- - \frac 1 \delta \eta_k((u_\delta-\psi)^-) \right)^- \, \mathrm d x \, \mathrm d s \leqslant 0 . $ (4.15)

    We remark that

    $ Ωtzδ(g1δηk((uδψ)))dxds=Ωt(g1δ[(ψuδ)+]q1)(g1δηk((uδψ)))dxds.
    $

    Since we have $ \left\{ g^- - \frac 1 \delta \eta_k((u_\delta-\psi)^-) < 0\right\} \subset \{ u_\delta < \psi \} $ then

    $ - \int_{ \Omega_t} \left[A(x, s, u_\delta\vee \psi, \nabla u_\delta)-A(x, s, \psi, \nabla \psi) \right]\cdot \nabla \left(g^- - \frac 1 \delta \eta_k((u_\delta-\psi)^-) \right)^- \, \mathrm d x \, \mathrm d s \\ \quad = \int_{ \Omega_t} \chi_{ \left\{ g^- - \frac 1 \delta \eta_k((u_\delta-\psi)^- < 0 \right\} } \left[A(x, s, \psi, \nabla u_\delta)-A(x, s, \psi, \nabla \psi) \right]\cdot \nabla \\ \left(g^- - \frac 1 \delta \eta_k((u_\delta-\psi)^-)) \right) \, \mathrm d x \, \mathrm d s. $

    By (1.6) it follows that

    $ [A(x,s,ψ,uδ)A(x,s,ψ,ψ)](g1δηk((uδψ))))1δηk((uδψ))[A(x,s,ψ,uδ)A(x,s,ψ,ψ)](uδψ)|[A(x,s,ψ,uδ)A(x,s,ψ,ψ)]||g||A(x,s,ψ,uδ)A(x,s,ψ,ψ)||g|.
    $

    Hence, we deduce from (4.15)

    $ Ωt|tg||(uδψ)|dxds+ΩΛk(x,t,(uδψ)(x,t))dxΩt(g1δ[(ψuδ)+]q1)(g1δηk((uδψ)))dxdsΩt|A(x,s,ψ,uδ)A(x,s,ψ,ψ)||g|dxds0.
    $

    Now, we pass to the limit as $ k\rightarrow \infty $. In particular, by using the monotone convergence theorem, we have

    $ \lim\limits_{k\rightarrow \infty} \int_{ \Omega} \Lambda_k (x, t, (u_\delta-\psi)(x, t)) \, \mathrm d x = -\int_ \Omega \, \mathrm d x \int_0^{(u_\delta-\psi)(x, t)} \left(g^- - \frac 1 \delta \left[\sigma^-\right]^{q-1}\right)^- \, \mathrm d\sigma \geqslant0 $

    and also

    $ -\lim\limits_{k\rightarrow \infty} \int_{ \Omega_t} \left( g^- - \frac 1 \delta \left[ (\psi-u_\delta)^+ \right]^{q-1}\right)\left(g^- - \frac 1 \delta \eta_k\left((u_\delta-\psi)^-\right) \right)^- \, \mathrm d x \, \mathrm d s = \|z_\delta^-\|^2_{L^2( \Omega_t)} $

    We gather the previous relations, and (since $ t\in (0, T) $ is arbitrary) we get

    $ -\int_{ \Omega_T} | \partial_t g^- | |(u_\delta-\psi)^-| \, \mathrm d x \, \mathrm d s + \|z_\delta^-\|^2_{L^2( \Omega_T)} \\ \leqslant \int_{ \Omega_T} \chi_{ \{ \psi > u_\delta \} } \left| A(x, t, \psi, \nabla u_\delta)-A(x, t, \psi, \nabla \psi) \right||\nabla g^-| \, \mathrm d x \, \mathrm d s. $

    Since it is clear that

    $ \lim\limits_{\delta\to 0} \int_{ \Omega_T} | \partial_t g^- | |(u_\delta-\psi)^-| \, \mathrm d x \, \mathrm d s = 0 $

    we obtain

    $ lim supδ0zδ2L2(ΩT)lim supδ0Ωtχ{ψ>uδ}|A(x,t,ψ,uδ)A(x,t,ψ,ψ)||g|dxds.
    $
    (4.16)

    Observing that (4.2), (4.5) and (4.6) hold, then

    $ F_\delta: = \chi_{ \{ \psi > u_\delta \} } \left| A(x, t, \psi, \nabla u_\delta)-A(x, t, \psi, \nabla \psi) \right| \rightarrow 0 \qquad \text{a.e. in } \Omega_T $

    as $ \delta \rightarrow 0 $. By (1.7), (3.2) and (3.4), $ F_\delta $ is also bounded in $ L^{p^\prime}(\Omega_T) $, hence $ F_\delta\rightharpoonup 0 $ in $ L^{p^\prime}(\Omega_T) $. We deduce

    $ \lim\limits_{\delta\to 0} \int_{ \Omega_T} \chi_{ \{ \psi > u_\delta \} } \left| A(x, t, \psi, \nabla u_\delta)-A(x, t, \psi, \nabla \psi) \right||\nabla g^-| \, \mathrm d x \, \mathrm d s = 0. $

    By (4.16) we obtain

    $ \lim\limits_{\delta\to 0} \|z_\delta^-\|^2_{L^2( \Omega_T)} = 0. $

    Hence we have

    $ 0 \leqslant \frac 1 \delta \left[ (u_\delta-\psi)^- \right]^{q-1} = \partial_t u_\delta - { \rm div } A(\cdot, \cdot, u_\delta\vee \psi, \nabla u_\delta)-f $

    and so

    $ 0 \leqslant \partial_t u - { \rm div } A(\cdot, \cdot, u , \nabla u )-f. $

    Similarly, rewriting (3.1) as follows

    $ z_\delta^+ + \partial_t u_\delta - { \rm div } A(\cdot, \cdot, u_\delta\vee \psi, \nabla u_\delta)-f = g^-+z_\delta^- $

    then

    $ \partial_t u - { \rm div } A(\cdot, \cdot, u , \nabla u )-f \leqslant g^- $

    and the proof is completed.

    Next result provides the one of Theorem 1.1 under the assumption (3.2) but removing condition (3.17).

    Proposition 4.3. Let (1.2), (1.4)–(1.16) and (3.2) be in charge. There exists at least solution $ u \in \mathcal K_{\psi}(\Omega_T) $ to the variational inequality (1.3) satisfying $ u(\cdot, 0) = u_0 $ in $ \Omega $, the estimate (4.1) and the Lewy–Stampacchia inequality (1.17).

    Proof. We know that

    $ g:=fψt+div A(x,t,ψ,ψ)=g+g,
    $

    where $ g^\pm $ are nonnegative elements of $ L^{p'}(0, T, W^{-1, p'}(\Omega)). $ By using a regularization procedure, due to [7] Lemma p. 593, and Lemma 4.1 in [15], we find a sequence $ \{g_n^-\}_{n\in \mathbb N} $ of nonnegative functions such that

    $ gnLp(ΩT)Lp(0,T,W1,p0(Ω))gn0a.e. in ΩTtgnLq(ΩT)
    $

    and

    $ g_n^- \rightarrow g^- \quad \text{in } L^{p'}(0, T, W^{-1, p'}(\Omega)) \text{ as }n\to \infty . $

    We define

    $ fn=ψtdiv A(x,t,ψ,ψ)+g+gn.
    $

    It is clear that

    $ fnfin Lp(0,T,W1,p(Ω))
    $

    as $ n\to \infty $. Due to the regularity assumptions on $ g_n^- $, we get the existence of $ u_n \in \mathcal K_{\psi}(\Omega_T) $ with $ u_n(\cdot, 0) = u_0 $ in $ \Omega $ such that for every $ v \in \mathcal K_{\psi}(\Omega_T) $ we have

    $ T0tun,vundt+ΩTA(x,t,un,un)(vun)dxdtT0fn,vundt.
    $
    (4.17)

    Moreover, the subsequent estimate holds

    $ un(,t)2L2(Ω)+unpLp(Ωt)C(b,N,p,α)[u02L2(Ω)+fnpLp(0,T,W1,p(Ω))+HL1(ΩT)+(u02L2(Ω)+fnpLp(0,T,W1,p(Ω))+bpLp(ΩT))pbpLp(ΩT)]
    $

    and the following Lewy-Stampacchia inequality holds

    $ 0tundiv A(x,t,un,un)fngn.
    $
    (4.18)

    Since the sequence $ \{f_n\}_{n \in \mathbb N} $ is strongly converging (and hence bounded) in $ L^{p'}(0, T, W^{-1, p'}(\Omega)) $, we obtain

    $ sup0<t<TΩ|un(,t)|2dx+ΩT|un|pdxdtC
    $

    for some positive constant $ C $ independent of $ n $. Moreover, the Lewy–Stampacchia inequality (4.18) implies a uniform bound of this kind

    $ tunLp(0,T;W1,p(Ω))C
    $

    again for some positive constant $ C $ independent of $ n $. Therefore, there exists $ u\in C^0\left([0, T], L^2(\Omega)\right) \cap L^p(0, T, W^{1, p}_0(\Omega)) $ with $ u(\cdot, 0) = u_0 $ in $ \Omega $ such that

    $ unustrongly in Lp(ΩT)unuweakly in Lp(ΩT,RN)unuweakly in L(0,T;L2(Ω))tuntuweakly in Lp(0,T,W1,p(Ω))
    $
    (4.19)

    as $ n\to \infty $. Obviously (4.19) implies $ u \geqslant \psi $ a.e. in $ \Omega_T $. If we summarize, we have $ u \in \mathcal K_{\psi}(\Omega_T) $ and then $ v_n: = u_n-\mathcal T_1(u_n-u) \in \mathcal K_{\psi}(\Omega_T). $ Hence, we use $ v_n $ as a test function in (4.17) and, arguing as in the proof of Proposition 4.1, we obtain

    $ \nabla u_n \rightarrow \nabla u \qquad \text{a.e. in } \Omega_T $

    as $ n\to \infty $. For fixed $ \lambda > 0 $ and $ v \in \mathcal K_{\psi}(\Omega_T) $ we also have $ v_{n, \lambda}: = u_n-\mathcal T_\lambda(u_n-v) \in \mathcal K_{\psi}(\Omega_T). $ Arguing again as in the proof of Proposition 4.1, we get (1.3) passing to the limit (first as $ n \rightarrow \infty $ and then as $ \lambda\rightarrow \infty $) in the inequality obtained by testing (4.17) by $ v_{n, \lambda} $.

    Finally, we remove condition (3.2), i.e., we are able to prove Theorem 1.1.

    Proof of Theorem 1.1. The convex set $ \mathcal K_{\psi}(\Omega_T) $ is nonempty and one can find $ w \in \mathcal K_{\psi}(\Omega_T) $ such that $ w(\cdot, 0) = \psi(\cdot, 0) $ in $ \Omega $ (see for details Remark 2.1 in [15]). Let us define

    $ ˆA(x,t,u,η):=A(x,t,u+w,η+w)ˆf:=ftwˆψ:=ψwˆu0:=u0w(,0).
    $

    Hence $ \hat f \in L^{p'}(0, T, W^{-1, p'}(\Omega)) $ and $ \hat \psi $ and $ \psi $ share the same trace on $ \partial \Omega \times (0, T) $. Therefore, one can conclude

    $ ˆψ0a.e. in ΩTˆψ(,0)=0a.e. in Ω.
    $

    Moreover, the vector field $ \hat A $ enjoys similar properties as $ A $. This is trivial for conditions (1.6) and (1.7). As in [12], properties of $ A $ and Young inequality, we have for $ \varepsilon > 0 $

    $ \hat A(x, t, u, \xi)\cdot \xi \geqslant (\alpha-\beta\, \varepsilon^p)\, |\xi+\nabla w|^p - \left( b^p+ \varepsilon^p\, \tilde b^p\right) |u+w| ^p - H_1 $

    with a suitable $ H_1\in L^1(\Omega_T) $. Moreover, as an elementary consequence of the convexity of $ |\; |^p $, for $ 0 < \vartheta < 1 $ we find a constant $ C = C(\vartheta, p) > 0 $ such that

    $ |\xi+\nabla w|^p \geqslant \vartheta^p\, |\xi|^p-C\, |\nabla w|^p\, , \qquad |u+w|^p \leqslant \vartheta^{-p}\, |u|^p+C\, |w|^p. $

    Hence, we get coercivity condition for $ \hat A $:

    $ \hat A(x, u, \xi)\cdot \xi \geqslant \hat\alpha\, |\xi|^p-(\hat b\, |u|)^p-\hat H, $

    where we set

    $ \hat\alpha = (\alpha-\beta\, \varepsilon^p)\, \vartheta^p\, , \qquad \hat b = \frac{b+ \varepsilon\, \tilde b}{\vartheta} $

    and denoted by $ \hat H $ a suitable nonnegative function in $ L^1(\Omega_T) $. Obviously, we can make $ \hat\alpha $ arbitrarily close to $ \alpha $, by choosing $ \varepsilon $ close to $ 0 $ and $ \vartheta $ close to $ 1 $. Using inequality (2.5) for $ b $ and $ \tilde b $ in place of $ f $ and $ g $, respectively, we can easily show that also $ \mathscr D_{\hat b} $ is arbitrarily close to $ \mathscr D_{ b} $, again by choosing $ \varepsilon $ close to $ 0 $ and $ \vartheta $ close to $ 1 $. Indeed, we have

    $ distL(0,T,LN,(Ω))(ˆb,L(ΩT))1+εϑdistL(0,T,LN,(Ω))(b,L(ΩT))+ε(1+ε)ϑ˜bL(0,T,LN,(Ω)).
    $

    By (1.16) we can also have

    $ Dˆb<ˆα1/pSN,p.
    $

    We observe that

    $ ˆfˆψt+div A(x,t,ˆψ,ˆψ)=fψt+div ˆA(x,t,ψ,ψ).
    $

    We can apply Proposition 4.3 for the operator $ \hat A $. Therefore, we obtain the existence of a function $ \hat u \in \mathcal K_{\hat \psi} (\Omega_T) $ such that

    $ ˆu(,0)=ˆu0in Ω
    $
    (4.20)

    and the following parabolic variational inequality

    $ T0ˆut,ˆvˆudt+ΩTˆA(x,t,ˆu,ˆu)(ˆvˆu)dxdtT0ˆf,ˆvˆudt
    $

    holds true for every admissible function $ \hat v \in \mathcal K_{\hat \psi} (\Omega_T) $. Since any $ v \in \mathcal K_{ \psi} (\Omega_T) $ can be rewritten as $ v = \hat v +w $ for some $ \hat v \in \mathcal K_{\hat \psi} (\Omega_T) $, by (4.20), by the definitions of $ \hat A $, $ \hat f $ and $ \hat \psi $, we see that the variational inequality (1.3) holds true with $ u: = \hat u +w $ and for any admissible function $ v \in \mathcal K_{\psi}(\Omega_T) $. The fact that $ u \in \mathcal K_{\psi}(\Omega_T) $ and $ u(\cdot, 0) = u_0 $ in $ \Omega $ is obvious, and this concludes the proof.

    The authors are members of the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM). F. Farroni also acknowledges support by project Starplus 2020 Unina Linea 1 "New challenges in the variational modeling of continuum mechanics'' from the University of Naples Federico II and Compagnia di San Paolo. G. Zecca also acknowledges support by Progetto FRA 2022 "Groundwork and OptimizAtion Problems in Transport'' from the University of Naples Federico II.

    The authors declare no conflict of interest.



    [1] Nonlocal systems of conservation laws in several space dimensions. SIAM J. Numer. Anal. (2015) 53: 963-983.
    [2] On the numerical integration of scalar nonlocal conservation laws. ESAIM M2AN (2015) 49: 19-37.
    [3] A. Aw and M. Rascle, Resurrection of "second order" models of traffic flow, SIAM J. Appl. Math., 60 (2000), 916–938 (electronic), URL http://dx.doi.org/10.1137/S0036139997332099. doi: 10.1137/S0036139997332099
    [4] S. Benzoni-Gavage and R. M. Colombo, An $n$-populations model for traffic flow, European J. Appl. Math., 14 (2003), 587–612, URL https://doi.org/10.1017/S0956792503005266. doi: 10.1017/S0956792503005266
    [5] S. Blandin and P. Goatin, Well-posedness of a conservation law with non-local flux arising in traffic flow modeling, Numer. Math., 132 (2016), 217–241, URL http://dx.doi.org/10.1007/s00211-015-0717-6. doi: 10.1007/s00211-015-0717-6
    [6] Minimising stop and go waves to optimise traffic flow. Appl. Math. Lett. (2004) 17: 697-701.
    [7] F. A. Chiarello and P. Goatin, Global entropy weak solutions for general non-local traffic flow models with anisotropic kernel, ESAIM: M2AN, 52 (2018), 163–180, URL https://doi.org/10.1051/m2an/2017066. doi: 10.1051/m2an/2017066
    [8] J. Friedrich, O. Kolb and S. Göttlich, A Godunov type scheme for a class of scalar conservation laws with non-local flux, Netw. Heterog. Media, 13 (2018), 531–547, URL http://dx.doi.org/10.3934/nhm.2018024.
    [9] P. Goatin and S. Scialanga, Well-posedness and finite volume approximations of the LWR traffic flow model with non-local velocity, Netw. Heterog. Media, 11 (2016), 107–121, URL http://dx.doi.org/10.3934/nhm.2016.11.107. doi: 10.3934/nhm.2016.11.107
    [10] R. J. LeVeque, Finite Volume Methods for Hyperbolic Problems, Cambridge Texts in Applied Mathematics, Cambridge University Press, Cambridge, 2002. doi: 10.1017/CBO9780511791253
    [11] On kinematic waves. Ⅱ. A theory of traffic flow on long crowded roads. Proc. Roy. Soc. London. Ser. A. (1955) 229: 317-345.
    [12] H. Payne, Models of Freeway Traffic and Control, Simulation Councils, Incorporated, 1971.
    [13] Shock waves on the highway. Operations Res. (1956) 4: 42-51.
    [14] A. Sopasakis and M. A. Katsoulakis, Stochastic modeling and simulation of traffic flow: Asymmetric single exclusion process with Arrhenius look-ahead dynamics, SIAM J. Appl. Math., 66 (2006), 921–944 (electronic). doi: 10.1137/040617790
    [15] G. Whitham, Linear and Nonlinear Waves, Pure and applied mathematics, Wiley, 1974.
    [16] H. Zhang, A non-equilibrium traffic model devoid of gas-like behavior, Transportation Research Part B: Methodological, 36 (2002), 275–290, URL http://www.sciencedirect.com/science/article/pii/S0191261500000503. doi: 10.1016/S0191-2615(00)00050-3
  • This article has been cited by:

    1. Paola Lunetti, René Massimiliano Marsano, Rosita Curcio, Vincenza Dolce, Giuseppe Fiermonte, Anna Rita Cappello, Federica Marra, Roberta Moschetti, Yuan Li, Donatella Aiello, Araceli del Arco Martínez, Graziantonio Lauria, Francesco De Leonardis, Alessandra Ferramosca, Vincenzo Zara, Loredana Capobianco, The mitochondrial aspartate/glutamate carrier (AGC or Aralar1) isoforms in D. melanogaster: biochemical characterization, gene structure, and evolutionary analysis, 2021, 1865, 03044165, 129854, 10.1016/j.bbagen.2021.129854
    2. Paola Lunetti, Ruggiero Gorgoglione, Rosita Curcio, Federica Marra, Antonella Pignataro, Angelo Vozza, Christopher L. Riley, Loredana Capobianco, Luigi Palmieri, Vincenza Dolce, Giuseppe Fiermonte, Drosophila melanogaster Uncoupling Protein-4A (UCP4A) Catalyzes a Unidirectional Transport of Aspartate, 2022, 23, 1422-0067, 1020, 10.3390/ijms23031020
    3. J. Dandurand, E. Dantras, C. Lacabanne, A. Pepe, B. Bochicchio, V. Samouillan, Thermal and dielectric fingerprints of self-assembling elastin peptides derived from exon30, 2021, 8, 2377-9098, 236, 10.3934/biophy.2021018
    4. Jany Dandurand, Magnus Monné, Valérie Samouillan, Martina Rosa, Alessandro Laurita, Alessandro Pistone, Donatella Bisaccia, Ilenia Matera, Faustino Bisaccia, Angela Ostuni, The 75–99 C-Terminal Peptide of URG7 Protein Promotes α-Synuclein Disaggregation, 2024, 25, 1422-0067, 1135, 10.3390/ijms25021135
    5. Ilaria Nigro, Rocchina Miglionico, Monica Carmosino, Andrea Gerbino, Anna Masato, Michele Sandre, Luigi Bubacco, Angelo Antonini, Roberta Rinaldi, Faustino Bisaccia, Maria Francesca Armentano, Neuroprotective Effect of Antiapoptotic URG7 Protein on Human Neuroblastoma Cell Line SH-SY5Y, 2023, 25, 1422-0067, 481, 10.3390/ijms25010481
  • Reader Comments
  • © 2019 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(3365) PDF downloads(265) Cited by(28)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog