This paper investigates the well-posedness of contact discontinuity solutions and the vanishing pressure limit for the Aw–Rascle traffic flow model with general pressure functions. The well-posedness problem is formulated as a free boundary problem, where initial discontinuities propagate along linearly degenerate characteristics. To address vacuum degeneracy, a condition at density jump points is introduced, ensuring a uniform lower bound for density. The Lagrangian coordinate transformation is applied to fix the contact discontinuity.The well-posedness of contact discontinuity solutions is established, showing that compressive initial data leads to finite-time blow-up of the velocity gradient, while rarefactive initial data ensures global existence. For the vanishing pressure limit, uniform estimates of velocity gradients and density are derived via level set argument. The contact discontinuity solutions of the Aw–Rascle system are shown to converge to those of the pressureless Euler equations, with matched convergence rates for characteristic triangles and discontinuity lines. Furthermore, under the conditions of pressure, enhanced regularity in non-discontinuous regions yields convergence of blow-up times.
Citation: Zijie Deng, Wenjian Peng, Tian-Yi Wang, Haoran Zhang. Well-posedness of contact discontinuity solutions and vanishing pressure limit for the Aw–Rascle traffic flow model[J]. Mathematics in Engineering, 2025, 7(3): 316-349. doi: 10.3934/mine.2025014
[1] | Takeyuki Nagasawa, Kohei Nakamura . Asymptotic analysis for non-local curvature flows for plane curves with a general rotation number. Mathematics in Engineering, 2021, 3(6): 1-26. doi: 10.3934/mine.2021047 |
[2] | Alberto Bressan, Yucong Huang . Globally optimal departure rates for several groups of drivers. Mathematics in Engineering, 2019, 1(3): 583-613. doi: 10.3934/mine.2019.3.583 |
[3] | Paola F. Antonietti, Chiara Facciolà, Marco Verani . Unified analysis of discontinuous Galerkin approximations of flows in fractured porous media on polygonal and polyhedral grids. Mathematics in Engineering, 2020, 2(2): 340-385. doi: 10.3934/mine.2020017 |
[4] | Jason Howell, Katelynn Huneycutt, Justin T. Webster, Spencer Wilder . A thorough look at the (in)stability of piston-theoretic beams. Mathematics in Engineering, 2019, 1(3): 614-647. doi: 10.3934/mine.2019.3.614 |
[5] | Daniele Del Santo, Francesco Fanelli, Gabriele Sbaiz, Aneta Wróblewska-Kamińska . On the influence of gravity in the dynamics of geophysical flows. Mathematics in Engineering, 2023, 5(1): 1-33. doi: 10.3934/mine.2023008 |
[6] | Hitoshi Ishii . The vanishing discount problem for monotone systems of Hamilton-Jacobi equations. Part 1: linear coupling. Mathematics in Engineering, 2021, 3(4): 1-21. doi: 10.3934/mine.2021032 |
[7] | Stefano Almi, Giuliano Lazzaroni, Ilaria Lucardesi . Crack growth by vanishing viscosity in planar elasticity. Mathematics in Engineering, 2020, 2(1): 141-173. doi: 10.3934/mine.2020008 |
[8] | François Murat, Alessio Porretta . The ergodic limit for weak solutions of elliptic equations with Neumann boundary condition. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021031 |
[9] | Stefano Biagi, Serena Dipierro, Enrico Valdinoci, Eugenio Vecchi . A Hong-Krahn-Szegö inequality for mixed local and nonlocal operators. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023014 |
[10] | Yuji Hamada, Takeshi Sato, Takashi Taniguchi . Multiscale simulation of a polymer melt flow between two coaxial cylinders under nonisothermal conditions. Mathematics in Engineering, 2021, 3(6): 1-22. doi: 10.3934/mine.2021042 |
This paper investigates the well-posedness of contact discontinuity solutions and the vanishing pressure limit for the Aw–Rascle traffic flow model with general pressure functions. The well-posedness problem is formulated as a free boundary problem, where initial discontinuities propagate along linearly degenerate characteristics. To address vacuum degeneracy, a condition at density jump points is introduced, ensuring a uniform lower bound for density. The Lagrangian coordinate transformation is applied to fix the contact discontinuity.The well-posedness of contact discontinuity solutions is established, showing that compressive initial data leads to finite-time blow-up of the velocity gradient, while rarefactive initial data ensures global existence. For the vanishing pressure limit, uniform estimates of velocity gradients and density are derived via level set argument. The contact discontinuity solutions of the Aw–Rascle system are shown to converge to those of the pressureless Euler equations, with matched convergence rates for characteristic triangles and discontinuity lines. Furthermore, under the conditions of pressure, enhanced regularity in non-discontinuous regions yields convergence of blow-up times.
In this paper, we consider the Aw–Rascle traffic flow model proposed by Aw and Rascle[1]:
{ρt+(ρu)x=0,(ρ(u+P))t+(ρu(u+P))x=0, | (1.1) |
where t>0 and x represent time and space, ρ, u, P are density, velocity, and pressure respectively. The pressure P is the function of density ρ and the small parameter ε>0, satisfying limε→0P(ρ,ε)=0. Here we consider the general pressure
P(ρ,ε)=ε2p(ρ). | (1.2) |
In the paper, for ρ>0, p∈C2(R+) satisfies the following conditions:
p′(ρ)>0,2p′(ρ)+ρp″(ρ)>0,limρ→+∞p(ρ)=+∞,limρ→0p(ρ)=k. | (1.3) |
And for different k we have the following two cases:
Case 1: If k is a finite constant, by (1.1)2−ε2k(1.1)1, (1.1) are equivalent to
{ρt+(ρu)x=0,(ρ(u+ε2(p(ρ)−k)))t+(ρu(u+ε2(p(ρ)−k)))x=0. |
Therefore, we can take k=0, otherwise we can let ˜p(ρ)=p(ρ)−k, then the form of (1.1) is unchanged. In this case, we have limρ→0p(ρ)=0, which includes γ-law: p(ρ)=ργ with γ≥1.
Case 2: If k=−∞, then limρ→0p(ρ)=−∞, which includes p(ρ)=lnρ. To define the inital date, we need to introduce Lips(R) to denote the set of piecewise Lipschitz functions:
Definition 1.1. For each f is belongs to Lips(R), there exists a finite set {xi}ni=1 of the first kind of discontinuity points, in which xi<xi+1 for any i=1,⋯,n−1, such that f is Lipschitz function on (xi,xi+1). Respectively, f is belongs to C1s, if f is C1 function on each (xi,xi+1).
Here, {xi}ni=1 is called partition points set. By the Lipschitz continuous at each subinterval, at each partition point xi, there exists the left limit limx→x−if(x)=f(x−i) and the right limit limx→x+if(x)=f(x+i), but may not be equal. And, [f](xi)=f(x+i)−f(x−i) is the jump of f at xi.
The initial data of (1.1) can be given as
(ρ,u)|t=0=(ρ0,u0), | (1.4) |
where both ρ0 and u0 are bounded, ρ0∈LipS(R), u0∈Lip(R) and ρ0≥ρ0_>0. At each partition point xi of density ρ0, following conditions are introduce for the low bounded of density:
● ε-condition: If ρ0(x−i)<ρ0(x+i) at xi, for x>xi,
u0(xi)+ε2p(ρ0)(x−i)>u0(x). | (1.5) |
● 0-condition: If ρ0(x−i)<ρ0(x+i) at xi, for x>xi,
u0(xi)>u0(x). | (1.6) |
The necessary and sufficient of above conditions will be discussed in Sections 3 and 4. And, 0-condition is the joint of ε-condition for ε>0.
Remark 1.2. The discontinuous points of ρ are of the first kind. If the function f∈Lips(R), then f can be decomposed into
f=fJ+fC, | (1.7) |
where fJ represents the jump part of function f; fC represents the absolutely continuous part of the function f and is a Lipschitz continuous function.
Remark 1.3. If the curve x(t;x0,0) is from the initial discontinuity point (x0,0), the discontinuity will propagate along this curve. On the other hand, according to the Rankine-Hugoniot condition, [u]=0 leads to
dxdt=[ρu][ρ]=u. | (1.8) |
Then, the discontinuous curve x(t;x0,0) satisfies
{dx(t;x0,0)dt=u(x(t;x0,0),t)x(0;x0,0)=x0. | (1.9) |
For fixed ε>0, the two eigenvalues of Aw–Rascle model consist: the genuinely nonlinear one, and the linearly degenerate one. For the well-posedness of Aw–Rascle model, we start with local existence and gradient blow-up.
For the local existence, if the solution is C1 function without vacuum, (1.1) is equivalent to strict hyperbolic system. In this case, the local existence of various types of function spaces, including C1 function class, Hs function class and BV function class, has been studied. Under the condition of ρ>0 in the whole space, Li and Yu [13] could provide the local well-posedness of one-dimensional C1 solutions on each compact characteristic triangle.
For gradient blow-up of conservation laws, in 1964, Lax[12] proved this conclusion for one-dimensional 2×2 genuinely nonlinear hyperbolic system. The results show that for a strictly hyperbolic system, if the initial value is a small smooth perturbation near a constant state, then the initial compression in any genuinely nonlinear characteristic field will produce a gradient blow-up in finite time. John[11], Li et al. [14,15], and Liu[16] have proven the generation of shock waves for n×n conservation law equations under different conditions.
The Riemann problem of (1.1) with p(ρ)=ργ was solved in [1]. Pan and Han[18] introduced the Chaplygin pressure function into the Aw–Rascle traffic model and solved the respective Riemann problem. It is noticeable that under the generalized Rankine–Hugoniot condition and entropy condition, they establish the existence and uniqueness of δ-wave. Cheng and Yang [6] solved the Riemann problem for the Aw–Rascle model with the modified Chaplygin gas pressure. Godvik and Hanche-Olsen[8] proved the existence of the weak entropy solution for the Cauchy problem with vacuum. By the compensated compactness method, Lu [17] proved the global existence of bounded entropy weak solutions for the Cauchy problem of general nonsymmetric systems of Keyfitz–Kranzer type. When the parameter n=1, the system is the Aw–Rascle model.
When the pressure P(ρ)≡0 in the Aw–Rascle model, the equations are simplified to pressureless Euler equations:
{ρt+(ρu)x=0,(ρu)t+(ρu2)x=0. | (1.10) |
The system (1.3) was used to describe both the process of the motion of free particles sticking under collision[3] and the formation of large scale in the universe [21,25]. The research on the pressureless Euler equations mainly focuses on the well-posedness of weak solutions. Brenier and Grenier [3] and Weinan et al. [25] independently obtained the existence of global weak solutions, and E-Sinai obtained the explicit expression of weak solutions by using the generalized variational principle. Wang et al. [24] prove the global existence of generalized solution to the L∞ initial data. Boudin[2] proved that the weak solution is the limit of the solution of the viscous pressureless Euler equations. Furthermore, Huang[9] proved the existence of entropy solutions for general pressureless Euler equations. Wang and Ding [23] proved the uniqueness of the weak solution of the Cauchy problem satisfying the Oleinik entropy condition when the initial value ρ0 is a bounded measurable function. Bouchut and James[7] also obtained similar results. Huang and Wang[10] proved the uniqueness of the weak solution when the initial value is the Radon measure.
To consider the relationship between the Aw–Rascle model and the pressureless Euler equations, a natural idea is the vanishing pressure limit, which considers limits of ε→0 as the pressure in the form of (1.2). The first study of vanishing pressure limit is on the Riemann solution of the isentropic Euler equations by Chen and Liu[4], the limiting solution in which includes δ-wave by concentration and vacuum by cavitation. Later, they extended the above result to the full Euler case[5]. Recently, Peng and Wang[19] studied the case of C1 solutions by a new level set argument. They showed that: For compressive initial data, the continuous solutions converge to a mass-concentrated solution of the pressureless Euler system; For rarefaction initial data, the solutions instead converge globally to a continuous solution. In [20], the authors studied the hypersonic limitation for C1 solution, and showed the convergence of blow-up time. For the Aw–Rascle model, Shen and Sun [22] proved that as ε→0, the Riemann solutions of the perturbed Aw–Rascle system converge to the ones of the pressureless Euler equations (1.10). Pan and Han[18] proved that for the Riemann problem, as the Chaplygin pressure P(ρ)=−ερ vanishes, the Riemann solutions of the Aw–Rascle traffic model converge to the respective solutions of the pressureless gas dynamics model (1.10).
This paper addresses two fundamental problems for the Aw–Rascle traffic flow model: (ⅰ) the well-posedness of contact discontinuity solutions with initial velocity u0 and piecewise Lipschitz initial density ρ0 satisfying the ε-condition, and (ⅱ) the vanishing pressure limit as ε→0. For problem (ⅰ), the analysis begins by establishing a strictly positive lower bound for the density through a partition of the domain, and Lagrangian coordinate transformations. This lower bound leads to a dichotomy: Compressive initial data induce finite time velocity gradient blow-up, whereas rarefactive initial data guarantee global existence of solutions. For problem (ⅱ), uniform density estimates independent of ε are derived via level set argument. It is proven that Aw–Rascle solutions converge to pressureless Euler solutions as ε→0: compressive data lead to mass-concentrated solutions, while rarefactive data yield globally regular solutions, with matching O(ε2) convergence rates for velocity fields and characteristic triangles. Furthermore, under the conditions of pressure, convergence of the blow-up time is established through enhanced regularity analysis in non-discontinuous regions.
We have the following two theorems.
Theorem 1.4. For fixed ε>0, and (1.1)–(1.4), if limρ→0p(ρ) satisfies one of the following two cases:
Case 1: limρ→0p(ρ)=0 and the initial data satisfies the ε-condition;
Case 2: limρ→0p(ρ)=−∞;
then there exists a time Tεb such that on R×[0,Tεb), there exists contacted discontinuity solution (ρε,uε) satisfying:
(1) If infx∈Ru′0(x)≥0, Tεb=+∞, the solution exists globally,
ρε∈Lips(R×[0,+∞)),uε∈Lip(R×[0,+∞)). |
(2) If infx∈Ru′0(x)<0, Tεb is finite and there exists at least one Xεb such that as (x,t)→(Xεb,Tεb), uεx(x,t)→−∞ while ρε(x,t) is upper and lower bounded. And the solution stands
ρε∈Lips(R×[0,Tεb)),uε∈Lip(R×[0,Tεb)). |
And, Γε={(x,t)∈R×[0,Tεb):x=xε2(t)} is the discontinuous curve and (xε2)′(t) is a Lipschitz function with respect to t, satisfying
dxε2(t)dt=uε(xε2(t),t). |
Theorem 1.5. For ε>0, (ρε,uε) are the unique solution of (1.1)–(1.4) with 0−condition on R×[0,Tεb). And, (ˉρ,ˉu) are the unique solution of (1.10) with initial data (1.4) on R×[0,Tb).
(1) For any 0<T∗<Tb, there exists a ε∗>0, such that, for 0<ε<ε∗, T∗<Tεb. As ε→0,
ρε→ˉρinM(R×[0,T∗]),uε→ˉuinLip(R×[0,T∗]). |
And, Tb≤lim_ε→0Tεb. Furthermore, there are the following convergence rates: For i=1,2,
|uε−ˉu|∼O(ε2),|λεi−ˉu|∼O(ε2),|xε2−ˉx|∼O(ε2), |
where λi are the eigenvalues of (1.1); xε2 and ˉx are the discontinuity lines of (1.1) and (1.10) respectively.
(2) If ρ0∈C1s(R), u0∈C1(R), I(ρ):=2ρp′(ρ)+(ρ)2p″(ρ)(ρp′(ρ))2 satisfies following conditions:
(a) I(ρ) is an increasing function with respect to ρ;
(b) There exists a small δ such that ∫δ0I(s)s2ds=+∞.
Then, the blow-up time convergences: limε→0Tεb=Tb.
Comparing with classical solutions of the compressible Euler equations, contact discontinuity solutions of the Aw–Rascle model and their vanishing pressure limits pose unique analytical challenges. First, the characteristic structure of the Aw–Rascle system, comprising both a linearly degenerate field and a genuinely nonlinear field, induces distinct regularity properties in Riemann invariants, with initial discontinuities propagating along linearly degenerate characteristic curves. Second, the coupling between evolving discontinuity curves and the solution itself characterizes the problem as a free boundary problem. Third, density-dependent degeneracy in the Riccati-type equations governing the system necessitates new density lower-bound estimation, distinct from those for the compressible Euler equations. Fourth, discontinuities inherently reduce solution regularity, mandating uniform estimates in tailored function spaces to rigorously establish convergence. Finally, precise analysis of blow-up times requires delicate estimates in non-discontinuity regions, where enhanced regularity can be exploited.
To overcome these challenges, three key ideas are introduced: (ⅰ) Lagrangian coordinate transformations: that map evolving discontinuity curves to the fixed boundaries; (ⅱ) Density Lower bound analysis: the derivatives of Riemann invariants in smooth regions and the jump condition analysis at discontinuities; (ⅲ) Time-directional derivative techniques: avoiding spatial discontinuities in Lagrangian. Level set argument are further developed to track invariant derivatives and establish uniform estimates for the vanishing pressure limit.
The paper is organized as follows: Section 2 proves the existence of classical solutions to the Aw–Rascle system through Lagrangian coordinate transformations, on avoiding vacuum formation. Section 3 establishes the well-posedness of contact discontinuity solutions using uniform L∞ estimates for time-directional derivatives in Lagrangian coordinates. Section 4 studies the vanishing pressure limit, while Section 5 quantifies blow-up time convergence via modulus of continuity estimates in non-discontinuity domains. Finally, Section 6 determines sharp convergence rates (O(ε2)) for characteristic curves and Riemann invariants.
In this section, we consider the Lagrangian transformations for fixed ε>0 that does not involve estimations, so we drop ε of (ρε,uε).
The eigenvalues of (1.1) are
{λ1=u−ε2ρp′(ρ),λ2=u, | (2.1) |
and Riemann invariants are
{w=u,z=u+ε2p(ρ), | (2.2) |
while p(ρ) can be represented by Riemann invariants as p(ρ)=z−uε2. Then, for ρ>0, (1.1) is equivalent to
{ut+(u−ε2ρp′(ρ))ux=0,(u+ε2p(ρ))t+u(u+ε2p(ρ))x=0. | (2.3) |
The derivatives along the charismatics are: For i=1,2,
Di=∂t+λi∂x, | (2.4) |
and the respective characteristic lines passing through (˜x,˜t) are defined as:
{dxi(t;˜x,˜t)dt=λi(xi(t;˜x,˜t),t),xi(˜t;˜x,˜t)=˜x. | (2.5) |
Combining (2.1) and (2.3), we have
{D1u=0,D2z=0. | (2.6) |
Since the eigenvalue λ1 is genuinely nonlinear and λ2 is linearly degenerate, there will be contact discontinuity in density along the second family of eigenvalues. So we consider the following Lagrangian transformation: Let τ=t and
{ddτx(τ;y)=u(x(τ;y),τ),x(0;y)=y. | (2.7) |
Let u(x(τ;y),τ)=v(y,τ), J=∂x∂y is the Jacobian determinant of the coordinate transformation satisfying:
{∂∂τJ=vy(y,τ),J(y,0)=1. | (2.8) |
In Lagrangian coordinates, we denote
g(y,τ):=ρ(x(τ;y),τ),Z(y,τ):=z(x(τ;y),τ), | (2.9) |
and Z(y,τ)=v(y,τ)+ε2p(g(y,τ)).
Next, for the relationship between J and g, in Lagrangian coordinates, (1.1)1 is equivalent to
gτ+gJ−1Jτ=0, | (2.10) |
which equals to (ln(gJ))τ=0. For g>0, integrating on τ leads to
J(y,τ)=g0(y)g(y,τ), | (2.11) |
where g0(y)=g(y,0) is the initial density. Combining with (2.8), in Lagrangian coordinates, (2.3) is equivalent to
{Jτ=vy,vτ+μvy=0,Zτ=0, | (2.12) |
where μ=−ε2gp′(g)J−1=−ε2g0g2p′(g). And, the initial data are
(J,v,Z)|τ=0=(1,v0(y),Z0(y)). | (2.13) |
In the Lagrangian coordinates, one could introduce following direction derivative:
D=∂τ+μ∂y, | (2.14) |
and the characteristic line passing through (˜y,˜τ)
{dy1(τ;˜y,˜τ)dτ=μ(y1(τ;˜y,˜τ),τ),y1(˜τ;˜y,˜τ)=˜y. | (2.15) |
Along the characteristic line, (2.12)2 is equivalent to
Dv=0, | (2.16) |
which leads to
v(y,τ)=v0(y1(0;y,τ)). | (2.17) |
And, by (2.12)3, one could have
Z(y,τ)=Z0(y). | (2.18) |
By the expression of p we have
p(g(y,τ))=Z(y,τ)−v(y,τ)ε2=Z0(y)−v0(y1(0;y,τ))ε2. | (2.19) |
If g(y,τ)>0,
g=p−1(p(g0(y))+v0(y)−v(y,τ)ε2). | (2.20) |
Remark 2.1. If absence of vacuum (g(y,τ)>0), the density g are uniquely determined by the velocity field v(y,τ) and Lagrangian coordinate y. Consequently, the velocity field v is governed by the transport equation:
vτ+μ(v,y)vy=0. | (2.21) |
For initial data containing jump, the coefficient μ(v,y) becomes discontinuous along y-direction. This necessitates estimating v through time τ-derivatives via Lagrangian evolution instead of spatial y-derivatives, thereby avoiding the jumps. Since the ε-condition in this paper ensures that the density has a lower bound at jump. The structure of (2.21) is the basis of the regularity of the solution in this paper. Even in the presence of contact discontinuities, its structure enables deriving the appropriate gradient estimates.
And in Lagrangian coordinates, ε-condition and 0-condition equivalence to:
● ε-condition: If g0(y−i)<g0(y+i) at jump point yi, for y>yi,
v0(yi)+ε2p(g0)(y−i)>v0(y). | (2.22) |
● 0-condition: If g0(y−i)<g0(y+i) at jump point yi, for y>yi
v0(yi)>y0(y). | (2.23) |
In the paper, we do not distinguish the ε-condition and 0-condition in the Eulerian coordinates or Lagrangian coordinates.
In this section, we consider v0, Z0∈C1, by [13], we could have the local existence of C1 solution with g(y,τ)>0 for (2.12)-(2.13). Next, we further consider the sharp life-span of C1 solution of (1.1)–(1.3). First, we have the following lemma.
Lemma 2.2. For the C1 solution (Jε,vε,Zε) of (2.12)-(2.13), gε is upper bounded with respect to ε; vε is uniformly bounded with respect to ε.
Proof. By (2.17),
miny∈Rv0(y)≤vε(y,τ)≤maxy∈Rv0(y). | (2.24) |
On the other hand, for p(gε)
p(gε(y,τ))=Zε(y,τ)−vε(y,τ)ε2=ε2p(g0(y))+v0(y)−v0(yε1(0;y,τ))ε2≤maxy∈Rp(g0)+2maxy∈Rv0ε2≤C(ε). | (2.25) |
Combining (1.3), gε has an upper bound with respect to ε, denoted by ˉgε.
We have the following lower bound estimate for density gε.
Proposition 2.3. For the C1 solutions (Jε,vε,Zε) of (2.12)-(2.13), gε has a uniform lower bound
gε(y,τ)≥A11+A2Bτ, | (2.26) |
where A1=miny∈Rg0, A2=maxy∈Rg0, B=maxy∈R((Zε0)′)+g0 with (f)+=max(f,0).
Proof. To estimate gε
p′(gε)Dgε=Dp(gε)=1ε2DZε=1ε2(Zετ+μεZεy)=1ε2μεZεy=−g−10(gε)2p′(gε)(Zε0)y. | (2.27) |
Then, we have
Dgε=−g−10(gε)2(Zε0)y. | (2.28) |
Dividing both sides of the above equation by (gε)2, by the definition of B, we have
D(1gε)=(Zε0)yg0≤B. | (2.29) |
Integrating s from 0 to τ along the characteristic line yε1(τ;y,τ)
gε(y,τ)≥g0(yε1(0;y,τ))1+g0(yε1(0;y,τ))Bτ≥A11+A2Bτ. | (2.30) |
So for the C1 initial data, we have the following proposition.
Proposition 2.4. For fixed ε>0 and (1.1)–(1.3), (ρ0,u0)∈(C1(R))2, there exists a time Tεb such that on R×[0,Tεb):
(1) If infx∈Ru′0(x)≥0, Tεb=+∞, the solution exists globally, (ρε,uε)∈(C1(R×[0,+∞)))2.
(2) If infx∈Ru′0(x)<0, Tεb is finite and there exists at least one Xεb such that as (x,t)→(Xεb,Tεb), uεx(x,t)→−∞ while ρε(x,t) is upper and lower bounded. And the solution stands (ρε,uε)∈(C1(R×[0,Tεb)))2.
Proof. Since the lower bound of density, by Remark 2.1, we just need to consider the gradient of vε. Taking ∂τ on (2.12)2 leads to
D(vετ)=(vετ)τ+με(vετ)y=−μετvεy. | (2.31) |
To compute μετvεy, we have
gετ=−gε(Jε)−1vεy=−(gε)2vεyg0=(gε)2vετg0με=−vετε2p′(gε), | (2.32) |
which indicates
μετvεy=−ε2g0((gε)2p′(gε))τvεy=−ε2g0(2gεp′(gε)+(gε)2p″(gε))gετvεy=2gεp′(gε)+(gε)2p″(gε)ε2(gεp′(gε))2(vετ)2. | (2.33) |
By
I(g):=2gp′(g)+(g)2p″(g)(gp′(g))2, |
one could get
D(vετ)=−I(gε)ε2(vετ)2. | (2.34) |
Since D(vετ)≤0, vετ is upper bounded:
vετ(y,τ)≤maxy∈Rvετ(y,0). |
Next, we focus on the lower bound of vετ. Dividing both sides of the above formula by (vετ)2, we have
D(−1vετ)=D(−1ε2gεp′(gε)(Jε)−1vεy)=−I(gε)ε2. | (2.35) |
Integrating s from 0 to τ along the characteristic line yε1(τ;ξ,0)
1(gεp′(gε)(Jε)−1vεy)(yε1(τ;ξ,0),τ)−1(g0p′(g0)v′0)(ξ)=1ε2∫τ0I(gε(yε1(s;ξ,0),s))ds. | (2.36) |
By (2.36), we have
(gεp′(gε)(Jε)−1vεy)(yε1(τ;ξ,0),τ)=(g0p′(g0)v′0)(ξ)1+(g0p′(g0)v′0)(ξ)∫τ0I(gε(yε1(s;ξ,0),s))ds. | (2.37) |
If v′0(ξ)≥0, by (2.37), we have
(Jε)−1vεy(yε1(τ;ξ,0),τ)≥0 |
for all τ≥0. If v′0(ξ)<0, we need to estimate ∫τ0I(gε(yε1(s;ξ,0),s))ds. By (2.25) and (2.26), for gε, we have
A11+A2Bτ≤gε(y,τ)≤ˉgε. | (2.38) |
Then, since p(gε)∈C2(R+), for gε∈[A11+A2Bτ,ˉgε], by (1.3), a positive lower bound I_ε such that
I(gε)≥I_ε>0. | (2.39) |
Then we have the following inequality:
∫τ0I(gε(yε1(s;ξ,0),s))ds≥∫τ0I_εds. | (2.40) |
Thus, if v′0(ξ)<0, when τ increases from 0 to some Tεb(ξ), we have
1+(g0p′(g0)v′0)(ξ)∫Tεb(ξ)0I(gε(yε1(s;ξ,0),s))ds=0, | (2.41) |
which indicate (Jε)−1vεy→−∞ as τ→Tεb(ξ).
For ξ run over all the points satisfying v′0(ξ)<0, we could have the minimum life-span
Tεb:=inf{Tεb(ξ)}>0. | (2.42) |
In Lagrangian coordinates, for fixed ε>0, gε is upper and lower bounded in [0,Tεb), so the global solution exists if and only if, for y∈R
v′0(y)≥0. | (2.43) |
On the other side, if v′0(y)<0, (Jε)−1vεy will goes to −∞ in the finite time. Then, there exists at least one point (Yεb,Tεb) such that on R×[0,Tεb), as (y,τ)→(Yεb,Tεb),
(Jε)−1vεy→−∞. | (2.44) |
Next, we want to discuss the transformation of the solution between the Lagrangian coordinates and the Eulerian coordinates. For y∈[−L,L], there is a characteristic triangle
{(y,τ)∣−L≤y≤yε1(s;L,0),0≤s≤τ}, | (2.45) |
where yε1(τ;L,0)=−L. And, in the characteristic triangle, each (y,τ) has a unique (x,t) satisfying
{ddtx(t;y)=vε(y,t),x(0;y)=y, | (2.46) |
which could be expressed as
x(y,τ)=y+∫τ0vε(y,s)ds. | (2.47) |
We let L→+∞, for any y∈R, each (y,τ) has a unique (x,t) by (2.46). According to the transformation of Eulerian coordinates and Lagrangian coordinates, we have
{uεx=(Jε)−1vεy,uεt=−(vε−ε2gεp′(gε))(Jε)−1vεy. | (2.48) |
And, for ρε we have
{ρεx=(Jε)−1gεy=(Jε)−1v′0+ε2p′(g0)g′0−vεyε2p′(gε),ρεt=−vε(Jε)−1gεy−gε(Jε)−1vεy. | (2.49) |
In Eulerian coordinates, for fixed ε>0, uε and ρε is upper and lower bounded in [0,Tεb), so the global solution exists if and only if, for x∈R
u′0(x)≥0. | (2.50) |
On the contrary, if u′0(x)<0, uεx will goes to −∞ in the finite time. Then, respecting to (Yεb,Tεb), there exists (Xεb,Tεb) such that on R×[0,Tεb), as (x,t)→(Xεb,Tεb), uεx(x,t)→−∞.
Further, modulus of continuity estimates of the solution see [13].
In this section we consider the Eq (1.1) with piecewise Lipshcitzs initial data (2.13), where v0∈Lip(R) and Z0∈Lips(R). Without loss of generality, we can assume that Z0 only has a jump discontinuity at y=0.
First, we would to clarify the influence of the jump. Here, we denote the characteristic line starting from y=0 as yε1(τ), which is defined in (2.15). And the characteristic line leading back from (˜y,˜τ) to the initial data is denoted as yε1(s;˜y,˜τ). For whether the characteristic line yε1(s;˜y,˜τ) crosses the discontinuity line or not, we can divide Ω:=R×[0,+∞) into the following regions:
Ω=Ω+∪ΩεI∪Ωε⨿∪{y=0}, | (3.1) |
where
Ω+={(y,τ)∣y>0,τ≥0},ΩεI={(y,τ)∣yε1(τ)≤y<0,τ≥0}, |
and
Ωε⨿={(y,τ)∣y<yε1(τ),τ≥0}. |
For (˜y,˜τ)∈Ω+∪Ωε⨿, yε1(s;˜y,˜τ) does not cross the discontinuity line. From the proof of Proposition 2.4, we can obtain the local existence of Lipschitz solutions in the characteristic triangle of domain Ω+∪Ωε⨿. For (˜y,˜τ)∈ΩεI, its backward characteristic line yε1(s;˜y,˜τ) must reach (0,τ0). So for v(˜y,˜τ) and Zε(˜y,˜τ) we have
v(˜y,˜τ)=v(0,τ0),Zε(˜y,˜τ)=Zε0(˜y). | (3.2) |
Therefore, if g>0, then g in ΩεI can be expressed as
gε(˜y,˜τ)=p−1(Zε0(˜y)−v(0,τ0)ε2). | (3.3) |
Without loss of generality, for different jump cases where Zε0 has only one discontinuity at y=0, we discuss the lower bound estimate of density for different regions. The key point is based on the results of Proposition 2.3, we have the following proposition.
Proposition 3.1. For fixed ε>0 and the solutions (Jε,vε,Zε) of (2.12)-(2.13):
Case 1: limg→0p(g)=0 and the initial data satisfies the ε-condition;
Case 2: limg→0p(g)=−∞ (ε-condition is not required);
then gε has a lower bound with respect to ε.
Proof. (1) Region Ω+∪Ωε⨿: For any point (˜y,˜τ) in Ω+∪Ωε⨿, by the method in Proposition 2.3, we can get the uniform lower bound estimate of density:
gε(y,τ)≥A11+A2Bτ. | (3.4) |
where A1=miny∈Rg0, A2=maxy∈Rg0, and B can be defined as
B=(Lip(Zε0))+g0, | (3.5) |
here (Zε0)C=Zε0−(Zε0)J is the absolute continuous part of Z0, and (Lip(Zε0))+ is the Lipschitzs constant of (Zε0)C without decreasing.
(2) The discontinuous curve y=0: If g0(0−)>g0(0+) at y=0, since v0 is continuous, this means Z0(0−)>Z0(0+). By (2.29), similar to the case in Ω+∪Ωε⨿, there exists a constant B such that the density has a lower bound (3.4).
Otherwise, if g0(0−)<g0(0+), which means Z0(0−)<Z0(0+), for (0,˜τ) and p(gε(0−,˜τ)), we have
p(gε(0−,˜τ))=Zε0(0−)−v0(yε1(0;0,˜τ))ε2. | (3.6) |
For (3.6), we need to discuss the following two cases:
Case 1: limg→0p(g)=0 and the initial data satisfies the ε-condition. By (1.5), there exists a constant Aε which may depend on ε such that
p(gε(0−,˜τ))≥Aε>0. | (3.7) |
So we have
gε(0+,˜τ)>gε(0−,˜τ)≥p−1(Aε)>0. | (3.8) |
Case 2: limg→0p(g)=−∞. By (3.6), for p(g) we have
p(gε(0−,˜τ))=p(g0(0−))+v0(0)−v0(yε1(0;0,˜τ))ε2≥p(g0_)−2maxy∈R|v0(y)|ε2>−∞. | (3.9) |
By (1.3), gε(0−,˜τ) has a lower bound with respect to ε.
(3) Region ΩεI: For any point (˜y,˜τ)∈ΩεI, its backward characteristic line must reach (0,τ0),
yε1(τ0;˜y,˜τ)=0, | (3.10) |
by (2.29) we have
D(1gε)=(Zε0)yg0. | (3.11) |
Integrating s from τ0 to ˜τ along the characteristic line yε1(τ;˜y,˜τ), there exists a constant B such that
B=(Lip(Zε0))+g0, | (3.12) |
such that
1gε(˜y,˜τ)−1min0≤τ0≤˜τgε(0−,τ0)≤1gε(˜y,˜τ)−1gε(0−,τ0)=∫˜ττ0(Lip(Zε0))+g0ds≤B(˜τ−τ0). | (3.13) |
So for fixed ε, we have
gε(˜y,τ0)≥Aε11+Aε1Bτ, | (3.14) |
where
Aε1=min0≤τ0≤˜τgε(0−,τ0),B=(Lip(Zε0))+g0. |
Combining the three cases, we obtain that the density has a uniform lower bound in Ω under the Lagrangian coordinate, denoted by g_ε.
g_ε:=min{A11+A2Bτ,p−1(Aε),Aε11+Aε1Bτ}. | (3.15) |
Remark 3.2. For limg→0p(g)=0, if the assumption in (1.5) becomes an equality, there exists at least one point y0>0 such that
Zε0(0−)=v0(y0). | (3.16) |
Then there exists a time Tεy0 such that
yε1(Tεy0;y0,0)=0. | (3.17) |
So we have
p(gε(0−,Tεy0))=Zε(0−,Tεy0)−v(yε1(Tεy0;y0,0),Tεy0)ε2=Zε0(0−)−v0(y0)ε2=0. | (3.18) |
This means there exists a finite time Tεy0 such that
gε(0−,Tεy0)=0. | (3.19) |
Based on the above discussion, (1.5) is a necessary and sufficient condition for the absence of vacuum.
Remark 3.3. For the case of limg→0p(g)=−∞, we do not need to use ε-condition to get the lower bound of density.
Based on the above discussion, we prove Theorem 1.4.
Proof. By (2.37), in Eulerian coordinates, we have
uεx(xε1(t;η,0),t)=1ρεp′(ρε)(xε1(t;η,0),t)⋅(ρ0p′(ρ0)u′0)(η)1+(ρ0p′(ρ0)u′0)(η)∫t0I(ρε(xε1(s;η,0),s))ds. | (3.20) |
If u′0(η)≥0, by (3.20), we have uεx(xε1(t;η,0),t)≥0 for all t≥0. If u′0(η)<0, we need to estimate ∫τ0I(ρε(xε1(s;η,0),s))ds. By Proposition 3.1 and (2.25), for ρε, we have
ρ_ε≤ρε(xε1(t;η,0),t)≤ˉρε. | (3.21) |
Then, since p(ρε)∈C2(R+), for ρε∈[ρ_ε,ˉρε], I(ρε) has a lower bound I_ε such that
I(ρε)≥I_ε. | (3.22) |
Then we have the following inequality:
∫t0I(ρε(xε1(s;η,0),s))ds≥∫t0I_εds. | (3.23) |
Thus, if u′0(η)<0, when t increases from 0 to some Tεb(η), we have
1+(ρ0p′(ρ0)u′0)(η)∫Tεb(η)0I(ρε(xε1(s;η,0),s))ds=0, | (3.24) |
which indicate (Jε)−1vεy→−∞ as τ→Tεb(η).
To consider η∈R, we could introduce the minimum life-span
Tεb:=inf{Tεb(η)}>0, | (3.25) |
where η run over all the points satisfying u′0(η)<0. In Eulerian coordinates, for fixed ε>0, ρε is upper and lower bounded in [0,Tεb), so the global solution exists if and only if, for x∈R
u′0(x)≥0. | (3.26) |
On the other side, if u′0(x)<0, uεx will goes to −∞ in the finite time. Then, there exists (Xεb,Tεb) such that on R×[0,Tεb), as (x,t)→(Xεb,Tεb), uεx(x,t)→−∞. Therefore, by (2.48) and (2.49), ρε and uε satisfy
ρε∈Lips(R×[0,Tεb)),uε∈Lip(R×[0,Tεb)). | (3.27) |
Next, we consider the regularity of discontinuous line. The discontinuous line in Eulerian coordinates is
dxε2(t)dt=uε(xε2(t),t), | (3.28) |
which has the following implicit expression
xε2(t)=x0+∫t0uε(xε2(s),s)ds. | (3.29) |
Therefore, by (3.29), for h>0, t≥0, we have
|(xε2)′(t+h)−(xε2)′(t)|=|uε(xε2(t+h),t+h)−uε(xε2(t),t)|≤Lip(uε)(|xε2(t+h)−xε2(t)|+h)≤Lip(uε)(‖u‖L∞+1)h. | (3.30) |
So (xε)′ is a Lipschitz function with respect to t.
Remark 3.4. We have proved the well-posedness of the contacted discontinuity solution where ρ0 has one contact discontinuity. For a general piecewise Lipschitz functions function ρ0, we can follow the above idea. Since the discontinuity points where [ρ0]>0 are separable, we can repeat the above procedure for each characteristic triangle containing only one such discontinuity. Then we can glue the fragments together to establish local existence. By repeating this procedure, we can extend the solution's existence time forward until a nonlinear singularity appears. Therefore we proved the well-posedness for general piecewise Lipschitz functions ρ0.
For fixed ε>0, if the initial data are ρ0,u0∈C1, by the method in [13], for a point (x,t) in the domain R×[0,Tεb), we can obtain the uniform modulus of continuity estimation of ux and ut, and then obtain the existence of C1 solution. However, for the case where there is a discontinuity, since the backward characteristic line of the point in the region Ω+∪Ωε⨿ does not pass through the discontinuity, the regularity of the solution in the region can be improved to C1 by the uniform modulus of continuity estimation. However, the backward characteristic line of the points in the region ΩεI will pass through the discontinuity, we cannot obtain the uniform modulus of continuity estimation of ux and ut along the spatial direction, so the Lipschitz regularity is optimal.
The blow-up of pressureless fluid in Eulerian coordinates can refer to [19], and we have similar argument in Lagrangian coordinates. For a smooth solution (ˉρ,ˉu) of (1.10), (1.10)2 is equivalent to Burgers' equation
ˉut+ˉuˉux=0. | (3.31) |
For pressureless fluid, we introduce the following derivative
D0=∂t+ˉu∂x. | (3.32) |
And the characteristic line passing through (˜x,˜t) is defined as
{dˉx(t;˜x,˜t)dt=ˉu(ˉx(t;˜x,˜t),t),ˉx(˜t;˜x,˜t)=˜x. | (3.33) |
Under the Lagrangian transformation, the Eq (1.10) is equivalent to
{ˉJτ=ˉvy,ˉvτ=0. | (3.34) |
and the respective initial data are
(ˉJ,ˉv)|τ=0=(1,v0(y)). | (3.35) |
Next, for the smooth solution (ˉJ,ˉv) of (3.34). By (3.34)2 we have
ˉv(y,τ)=v0(y)andˉvy(y,τ)=v′0(y). | (3.36) |
Integrating the first equation in (3.34), we have
ˉJ(y,τ)=1+v′0(y)τ. | (3.37) |
Combining (2.11),
ˉg(y,τ)=g0(y)J(y,τ)=g0(y)1+v′0(y)τ. | (3.38) |
Then, by (3.36) and (3.37), one could get
ˉJ−1ˉvy(y,τ)=v′0(y)1+v′0(y)τ. | (3.39) |
According to (3.39), if v′0(y)<0, when τ increases from 0 to some Tb(y), such that
1+v′0(y)Tb(y)=0, | (3.40) |
which indicate ˉJ−1ˉvy→−∞ and ˉg→+∞ as τ→Tb(y).
For y run over all the points such that v′0(y)<0, we could have the minimum life-span
Tb:=inf{Tb(y)}=inf{−1v′0(y)}>0. | (3.41) |
Thus, from (3.39), we see that singularity of ˉvy first happens at τ=Tb. When τ→Tb, there is a point such that ˉg goes to +∞, which corresponds to the mass concentration.
According to the transformation of Eulerian coordinates and Lagrangian coordinates, we have
{ˉux=ˉJ−1ˉvy,ˉut=−ˉvˉJ−1ˉvy. | (3.42) |
Therefore, in Eulerian coordinates, we have
ˉρ(ˉx(t;x,0),t)=ρ0(x)1+u′0(x)t, | (3.43) |
and
ˉux(ˉx(t;x,0),t)=u′0(x)1+u′0(x)t. | (3.44) |
According to the above discussion, for t∈[0,Tb), if
ρ0∈Lips(R),u0∈Lip(R), | (3.45) |
then ˉρ∈Lips(R×[0,Tb)), ˉu∈Lip(R×[0,Tb)).
Based on the above discussion, for the case of pressureless fluid, similar to Theorem 1.4, we have the following proposition.
Proposition 3.5. For (1.10) with initial data (3.45), there exists a time Tb such that on R×[0,Tb), there exists a solution (ˉρ,ˉu) satisfying
(1) If infx∈Ru′0(x)≥0, the solution exists globally. For ˉρ and ˉu, there are
ˉρ∈Lips(R×[0,+∞)),ˉu∈Lip(R×[0,+∞)). | (3.46) |
(2) If infx∈Ru′0(x)<0, there exists a finite Tb and at least one Xb such that as (x,t)→(Xb,Tb), ˉux(x,t)→−∞ and ˉρ(x,t)→+∞. And the solution stands
ˉρ∈Lips(R×[0,Tb)),ˉu∈Lip(R×[0,Tb)). | (3.47) |
And, ˉΓ={(x,t)∈R×[0,Tb):x=ˉx(t)} is the discontinuous curve and ˉx′(t) is a Lipschitz function with respect to t, satisfying
dˉx(t)dt=ˉu(ˉx(t),t). | (3.48) |
From the above discussion, we know that for Aw–Rascle model, if u′0<0 initially, ux will goes to −∞ in finite time. In this section, for any fixed T, as ε→0, we consider the convergence of vε and gε on R×[0,T] in Lagrangian coordinates. Without loss of generality, we assume that ε<1. First, we introduce the level set on the lower bound of (Jε)−1vεy:
mε(τ):=infy∈R{(Jε)−1vεy(y,τ)}. | (4.1) |
For any fixed ε>0 and the compressive initial data, there exists a finite life-span Tεb defined in Proposition 2.4, which is +∞ for the rarefaction initial data. And we have
limτ↑Tεbmε(τ)=−∞. | (4.2) |
Further, for M>0, we define τεM
τεM=sup{s:−M≤infτ∈[0,s]mε(τ),s≤T}. | (4.3) |
According to the definition, {τεM} is a monotone sequence with respect to M, and
limM→+∞τεM=min{Tεb,T}=:τεb. | (4.4) |
First, we need to proof that the density has a uniform lower bound with respect to ε, and we have the following proposition.
Proposition 4.1. For any ε>0 and the solution (Jε,vε,Zε) of (2.12)-(2.13) with 0−condition, gε has a uniform lower bound with respect to ε.
Proof. (1) Region Ω+∪Ωε⨿: For any point (˜y,˜τ) in Ω+∪Ωε⨿, by Proposition 3.1, the density has a uniform lower bound:
gε(y,τ)≥A11+A2Bτ. | (4.5) |
where A1=miny∈Rg0, A2=maxy∈Rg0, and B can be defined as
B=(Lip(Zε0))+g0, | (4.6) |
here (Zε0)C=Zε0−(Zε0)J is the absolute continuous part of Z0, and (Lip(Zε0))+ is the Lipschitzs constant of the continuous part of Zε0 without decreasing.
(2) The discontinuous curve y=0: If g0(0−)>g0(0+) at y=0. Since v0 is continuous, Z0(0−)>Z0(0+). By (2.29), similar to the case in Ω+∪Ωε⨿, there exists a constant B such that the density has a lower bound (3.4).
If g0(0−)<g0(0+) at y=0, for (0,˜τ) and p(gε(0−,˜τ)), we have
p(gε(0−,˜τ))=Zε0(0−)−v0(yε1(0;0,˜τ))ε2=p(g0(0−))+v0(0−)−v0(yε1(0;0,˜τ))ε2. | (4.7) |
By (1.6), we have
p(gε(0−,˜τ))≥p(g0(0−))≥p(g0_). | (4.8) |
So we have
gε(0+,˜τ)>gε(0−,˜τ)≥g0_. | (4.9) |
(3) Region ΩεI: For any point (˜y,˜τ)∈ΩεI, by Proposition 3.1, we have
gε(˜y,˜τ)≥mingε(0−,τ0)1+mingε(0−,τ0)B(˜τ−τ0)≥A′11+A′1Bτ, | (4.10) |
where
A′1=mingε(0−,τ0),B=(Lip(Zε0))+g0. |
Combining the above cases, we obtain that the density has a lower bound in Ω under the Lagrangian coordinate, denoted by g_:
g_:=min{A11+A2Bτ,g0_,A′11+A′1Bτ}. | (4.11) |
Remark 4.2. In order to get a uniform lower bound of gε, for constant M, we need
p(gε(y,τ))=p(g0(y))+v0(y)−v0(yε1(0;y,τ))ε2≥M>−∞. | (4.12) |
The above equation is equivalent to
v0(y)≥v0(yε1(0;y,τ))+ε2(M−p(g0(y))). | (4.13) |
As ε→0, the above equation is equivalent to the 0-condition, which is necessary and sufficient for both limg→0p(g)=0 case and limg→0p(g)=−∞ case.
For fixed sufficiently large M, let τ_M:=lim_ε→0τεM, see Lemma 4.5 for specific proof. Next, we have the following L∞ estimates lemma.
Lemma 4.3 (L∞ uniform estimations). On R×[0,τ_M], for any ε>0, gε has uniform bound with respect to ε in L∞(R×[0,τ_M]); and vε has uniform bound with respect to ε in Lip(R×[0,τ_M]).
Proof. On R×[0,τ_M], we have
lngε(y,τ)g0(y)=∫τ0(lngε(y,s))τds=−∫τ0(Jε)−1vεy(y,s)ds≤MT. | (4.14) |
Therefore, on R×[0,τ_M], let ˉg0:=maxy∈Rg0 we obtain the uniform upper bound of gε.
gε≤ˉg0eMT=:ˉgM. | (4.15) |
Combining (4.11), then gε has uniform bound with respect to ε on R×[0,τ_M]:
g_≤gε≤ˉgM. | (4.16) |
By (2.17), vε has uniform bound with respect to ε
miny∈Rv0(y)≤vε(y,τ)≤maxy∈Rv0(y). | (4.17) |
Next, we estimate the bound of (Jε)−1vεy, due to
D(vετ)=−I(gε)ε2(vετ)2≤0. | (4.18) |
So we have
ε2(gεp′(gε)(Jε)−1vεy)(y,τ)≤ε2(g0p′(g0)v′0)(yε1(0;y,τ)). | (4.19) |
gε is uniformly bounded for any ε and p∈C2(R+), so p′(gε) is uniformly bounded. So we obtain the uniform upper bound of (Jε)−1vεy on R×[0,τ_M]
((Jε)−1vεy)(y,τ)≤(g0p′(g0)v′0)(yε1(0;y,τ))(gεp′(gε))(y,τ)≤C1(M), | (4.20) |
On the other hand, due to
(Jε)−1vεy≥−M, | (4.21) |
on R×[0,τ_M], we have
‖(Jε)−1vεy‖∞≤max{C1(M),M}≤C(M). | (4.22) |
According to the transformation of Eulerian coordinates and Lagrangian coordinates, we have
{uεx=(Jε)−1vεy,uεt=−(vε−ε2gεp′(gε))(Jε)−1vεy. | (4.23) |
Correspondingly we have the following lemma.
Lemma 4.4. On R×[0,t_M], for any ε>0, ρε has uniform bound with respect to ε in L∞(R×[0,t_M]); and uε has uniform bound with respect to ε in Lip(R×[0,t_M]).
Before we discuss the convergence of ρε and uε, we need to estimate the uniform lower bound of tεM with respect to ε, which means that for M large enough, tεM does not go to 0 as ε goes to 0.
Lemma 4.5. For M is large enough, tεM has a uniform positive lower bound.
Proof. This claim is obviously true for u′0≥0, we then prove it by contradiction for u′0<0. If the claim is false, we can find a subsequence {εn} such that tεnM→0 as n→∞. By (2.37), in Eulerian coordinates, we can find a point (xεnM,tεnM) such that
−M=uεnx(xεnM,tεnM)=1(ρεnp′(ρεn))(xεnM,tεnM)⋅(ρ0p′(ρ0)u′0)(xεn1(0;xεnM,tεnM))1+(ρ0p′(ρ0)u′0)(xεn1(0;xεnM,tεnM))∫tεnM0I(ρεn(xεnM,s))ds. | (4.24) |
For εn>0, ρεn has uniform upper and lower bounds
ρ_≤ρεn≤ˉρ0eMtεnM. | (4.25) |
Therefore, for ρεn∈[ρ_,ˉρ0eMtεnM], I(ρ) is a continuous function respect to ρ, I(ρεn) has a uniform upper and lower bounds. By condition (1.3) satisfied by p(ρ), we have
(ρ2p′(ρ))′=2ρp′(ρ)+ρ2p″(ρ)>0. | (4.26) |
Combining the (4.25) and monotonicity of ρ2p′(ρ), we have
ρεnp′(ρεn)=(ρεn)2p′(ρεn)ρεn≥(ρ_)2p′(ρ_)ˉρ0eMtεnM. | (4.27) |
So, as n→∞,
−M≥ˉρ0eMtεnM(ρ_)2p′(ρ_)⋅minx∈R(ρ0p′(ρ0)u′0)1+(ρ0p′(ρ0)u′0)(xεn1(0;xεnM,tεnM))∫tεnM0I(ρεn(xεnM,s))ds→ˉρ0⋅minx∈R(ρ0p′(ρ0)u′0)(ρ_)2p′(ρ_). | (4.28) |
On the other hand, when M is sufficiently large
ˉρ0⋅minx∈R(ρ0p′(ρ0)u′0)(ρ_)2p′(ρ_)≥−M. | (4.29) |
This contradicts with (4.28). Therefore, for each fix M large enough, tεM has a uniformly positive lower bound, denoted by T_.
Based on the uniform estimates discussed above, we then prove Theorem 1.5(1). There is a uniform domain defined as t_M:=lim_ε→0tεM. For the compact set H⊂R×[0,t_M], the estimates we did above are uniform in H, we have
uε→ˉuinLip(H),ρε→ˉρinM(H). | (4.30) |
As ε→0, combine the (2.5) and the convergence of (uε,ρε), we have
xε(t)→ˉx(t), | (4.31) |
where ˉx(t) is a Lipschitz function with respect to t.
Remark 4.6. In Lagrangian coordinates, for τ≤τ_M and any point (y,τ) in ΩεI, when ε is sufficiently small, the backward characteristic line of this point is included in Ω−. That is, when ε→0, Ωε⨿→Ω− and Ω=ˉΩ+∪Ω−.
Next, we consider the consistency. For ε>0 and φ∈C∞c(H), we have
∫∞0∫Rρεφt+ρεuεφxdxdt+∫Rρ0φ(x,0)dx=0, | (4.32) |
∫∞0∫Rρε(uε+ε2p(ρε))φt+ρεuε(uε+ε2p(ρε))φxdxdt+∫Rρ0(u0+ε2p(ρ0))φ(x,0)dx=0. | (4.33) |
As ε→0, by the convergence of (ρε,uε), the above equalities turn to
∫∞0∫Rˉρφt+ˉρˉuφxdxdt+∫Rρ0φ(x,0)dx=0, | (4.34) |
∫∞0∫Rˉρˉuφt+ˉρˉu2φxdxdt+∫Rρ0u0φ(x,0)dx=0, | (4.35) |
which show (ˉρ,ˉu) satisfies (1.10) with initial value (1.4) in H. Finally, we consider the convergence of uεx. For φ∈C∞c(H), as ε→0,
∫∞0∫R(uεx−ˉux)φdxdt=−∫∞0∫R(uε−ˉu)φxdxdt→0. | (4.36) |
So we have uεx⇀ˉux. Since for each H, uεx are uniformly bounded respect to ε, then uεx→ˉux in L∞(H).
Then, in Lagrangian coordinates, we could have as ε→0
mε(τ)→m(τ):=infy∈R{ˉJ−1ˉvy(y,τ)}andlimτ↑Tbm(τ)=−∞, | (4.37) |
where Tb is defined in (3.41). Similar to τεM, for M>0, we could introduce
τM=sup{s:−M≤infτ∈[0,s]m(τ),s≤T}. | (4.38) |
According to the definition, {τM} is a monotone sequence with respect to M, and as M→∞
limM→+∞τM=min{Tb,T}=:τb. | (4.39) |
By the convergence of the level set, we have
limε→0τεM=τM. | (4.40) |
Next, we shows that
Tb≤lim_ε→0Tεb. | (4.41) |
Since τεM≤Tεb. So we have
limM→+∞limε→0τεM≤limM→+∞lim_ε→0Tεb. | (4.42) |
The left hand side holds
limM→+∞limε→0τεM=limM→+∞τM=Tb. | (4.43) |
So we get (4.41).
In this section, if ρ0∈C1s(R), u0∈C1(R), we can further obtain the convergence of the blow-up time Tεb. First, we need to introduce the non-discontinuous regions as:
ΩεM=Ω+∪Ωε⨿. |
Similar to the previous definition, we define
ˉmε(τ):=inf(y,τ)∈ΩεM{(Jε)−1vεy(y,τ)}. | (5.1) |
For any fixed ε>0 and the compressive initial data, there exists a finite life-span ˉTεb, which is +∞ for the rarefaction initial data. And we have
limτ↑ˉTεbˉmε(τ)=−∞. | (5.2) |
Further, for M>0, we define ˉτεM
ˉτεM=sup{s:−M≤infτ∈[0,s]ˉmε(τ),s≤T}. | (5.3) |
According to the definition, {ˉτεM} is a monotone sequence with respect to M, and
limM→+∞ˉτεM=min{ˉTεb,T}=:ˉτεb. | (5.4) |
Before discussing the convergence of the blow-up time, we need to discuss the modulus of continuity estimation in ΩεM.
The modulus of continuity of a function f(x,t) is defined by the following non-negative function:
η(f∣ε)(s)=sup{|f(x1,s)−f(x2,s)|:s∈[0,t],|x1−x2|<ε}. | (5.5) |
For modulus of continuity, there are the following properties, for more details please refer to [13].
Lemma 5.1. For modulus of continuity
(1) If ϵ1<ϵ2, η(f∣ϵ1)<η(f∣ϵ2).
(2) For any positive number C, η(f∣Cϵ)≤(C+1)η(f∣ϵ).
(3) For f and g that are two continuous functions, η(f±g∣ϵ)≤η(f∣ϵ)+η(g∣ϵ), and η(fg∣ϵ)≤η(f∣ϵ)‖g‖L∞+η(g∣ϵ)‖f‖L∞.
(4) If f∈Cα, 0<α≤1, if and only if there exists a constant L such that η(f∣ϵ)≤Lϵα.
Define
η(f1,f2,…,fn∣ϵ)(t):=max{η(fi∣ϵ)(t):i=1,…,n}. | (5.6) |
By definition, η(f1,f2,…,fn∣ϵ)(t) has similar properties in Lemma 5.1. So we have the following lemma:
Lemma 5.2. If u0∈C1(R) and g0∈C1(R−∪R+). Then on the ΩεM, the modulus of continuity of (Jε)−1vεy and vετ is uniform, which means (Jε)−1vεy and vετ is equi-continuous.
Proof. By (1.1), we have the following equation
D(vεyJε)=−(2+gεp″(gε)p′(gε))(vεyJε)2+(1+gεp″(gε)p′(gε))gεg0ZεyvεyJε. | (5.7) |
And vτ can be expressed by vεyJε and the lower order terms. Thus we only need to prove the modulus of continuity of vεyJε. To prove this, we first prove the modulus of continuity of the characteristics line. Then combine with the regularity of the lower order terms, we obtain the modulus of continuity of vεyJε.
By the Lemma 4.4, we have that gε and Jε are uniformly Lipschitzs functions on the ΩεM. Thus for the characteristic line, for ϵ>0 and |˜y1−˜y2|<ϵ, without loss of generality, we assume s<τ
|yε1(τ;˜y1,s)−yε1(τ;˜y2,s)|≤|˜y1−˜y2|+∫τs|με(yε1(s′;˜y1,s),s′)−με(yε1(s′;˜y2,s),s′)|ds′≤|˜y1−˜y2|+C(M)∫τs|yε1(s′;˜y1,s)−yε1(s′;˜y2,s)|ds′. | (5.8) |
By Gr¨onwall's inequality, we conclude that
|yε1(τ;˜y1,s)−yε1(τ;˜y2,s)|≤|˜y1−˜y2|eC(M)(τ−s)≤|˜y1−˜y2|eC(M)τ. | (5.9) |
Then we consider the modulus of continuity of vεyJε. We consider the estimation on two directions: characteristic direction and y-direction. For characteristic direction, because gε and vε are uniform bounded in C1(ΩεM),
|vεyJε(yε1(τ1;˜y,˜τ),τ1)−vεyJε(yε1(τ2;˜y,˜τ),τ2)|≤|∫τ2τ1D(vεyJε)ds|≤C(M)|τ1−τ2|. | (5.10) |
Now we consider the y−direction in which we need the (5.9). For convenient, we use the denotation Yi(s)=(yi(s),s)=(yε1(s;˜yi,0),s). Thus for any (˜yi,0)∈ΩεM, i=1,2. Let
G(gε)=gεp″(gε)p′(gε). | (5.11) |
By (5.7) we have
|vεyJε(Y1(τ))−vεyJε(Y2(τ))|≤|v′0(~y1)−v′0(~y2)|+C(M)∫τ0|(Zε0)y(y1(s))−(Zε0)y(y2(s))|ds+C(M)∫τ0(|gε(Y1(s))−gε(Y2(s))|+|G(gε(Y1(s)))−G(gε(Y2(s)))|)ds+C(M)∫τ0|vεyJε(Y1(s))−vεyJε(Y2(s))|ds≤η(v′0||~y1−~y2|)(0)+C(M)∫τ0η((Zε0)y||y1(s)−y2(s)|)(0)ds+C(M)∫τ0(η(gε||Y1(s)−Y2(s)|)(s)+η(G(gε)||Y1(s)−Y2(s)|)(s))ds+C(M)∫τ0η(vεyJε||Y1(s)−Y2(s)|)(s)ds. | (5.12) |
For first term we have
η(v′0||~y1−~y2|)(0)≤η(v′0|ϵ)(0). | (5.13) |
The second term we have
η((Zε0)y||y1(s)−y2(s)|)(0)≤η((Zε0)y|eC(M)sϵ)(0)≤(eC(M)s+1)η((Zε0)y|ϵ)(0). | (5.14) |
For the third term we have
η(gε||Y1(s)−Y2(s)|)(s)≤C(M)|y1(s)−y2(s)|≤C(M)eC(M)sϵ, | (5.15) |
and
η(G(gε)||Y1(s)−Y2(s)|)(s)≤η(G(gε)|C(M)eC(M)sϵ)(s)≤(C(M)eC(M)s+1)η(G|ϵ)(s). | (5.16) |
For the last term, we have
η(vεyJε||Y1(s)−Y2(s)|)(s)≤η(vεyJε|eC(M)sϵ)(s)≤(eC(M)s+1)η(vεyJε|ϵ)(s). | (5.17) |
There exists a constant C′(M) such that
η(vεyJε|ϵ)(s)≤C′(M)((ϵ+η(v′0,(Zε0)y|ϵ)(0))+∫τ0η(G|ϵ)(s)ds)+C′(M)∫τ0η(vεyJε|ϵ)(s)ds. | (5.18) |
For any ε>0, because ˜yi are chosen arbitrary on the ΩεM. Through the Gr¨onwall's inequality, the above inequality equals to
η(vεyJε|ϵ)≤C′(M)eC′(M)T(ϵ+η(v′0,(Zε0)y|ϵ)(0)+∫τ0η(G|ϵ)(s)ds). | (5.19) |
Since G is a continuous function, the modulus of continuity of vεyJε on ΩεM is uniform.
For a compact set H⊂ΩεM, we could lift the convergence as (Jε)−1vεy→ˉJ−1ˉvy in C(H) as ε→0. Then, for (y,τ)∈H, we could have as ε→0,
ˉmε(τ)→m(τ). | (5.20) |
By the convergence of the level set, we have
limε→0ˉτεM=τM. | (5.21) |
Furthermore, we consider the convergence of the blow-up time for I(g) satisfying the following conditions:
(a) I(g) is a increasing function with respect to g;
(b) There exists a small δ such that ∫δ0I(s)s2ds=+∞.
For example, p(g)=lng meets the above conditions. Using the idea in [20], we have the following lemma.
Lemma 5.3. For M>0, if I(g) meet the above conditions (1) and (2), then we have
limε→0Tεb=Tb. | (5.22) |
Proof. Case 1: inf{v′0(y)}≥0.
In this case, Tb=+∞. Let αε:=gεp′(gε)(Jε)−1vεy, by (2.37)
αε(y,0)=α0(y)=g0p′(g0)v′0(y)≥0. | (5.23) |
In this case we want to show limε→0Tεb=+∞. If the claim is false, we can find a subsequence {εn} such that {Tεnb} is bounded. For any εn, there exist yεn1 defined by (2.15) such that when τ↑Tεnb,
limτ↑Tεnbαεn=−∞, | (5.24) |
which
1+αεn(yεn1(0;y,τ),0)∫Tεnb0I(gεn(yεn1(s;y,τ),s))ds=0. | (5.25) |
The above equation is equivalent to
−1αεn(yεn1(0;y,τ))=∫Tεnb0I(gεn(yεn1(s;y,τ),s))ds. | (5.26) |
The right hand side is uniformly bounded by a finite positive constant
∫Tεnb0I(gεn(yεn1(s;y,τ),s))ds≤C. | (5.27) |
For αεn
αεn(yεn1(0;y,τ))≤−1C<0. | (5.28) |
This contradicts with (5.23), so we have
limε→0Tεb=+∞=Tb. | (5.29) |
Case 2: inf{v′0(y)}<0.
For large M, we first proof that ˉTεb is uniformly bounded. If the claim is false, we can find a subsequence {εn} such that ˉTεnb→+∞ as n→+∞. And there exist yεn1 defined by (2.15) such that when τ↑ˉTεnb,
1+αεn(yεn1(0;y,τ),0)∫ˉTεnb0I(gεn(yεn1(s;y,τ),s))ds=0. | (5.30) |
By (4.11) and the monotonicity of I(ρ), there exists a constant C such that
C≥−1αεn(yεn1(0;y,τ))=∫ˉTεnb0I(gεn(yεn1(s;y,τ),s))ds≥∫ˉTεnb0I(A11+A2Bs)ds. | (5.31) |
Let n→∞ we have
C≥∫+∞0I(A11+A2Bs)ds=A2B∫+∞0I(A11+s)ds. | (5.32) |
The definitions of A1, A2 and B, see Proposition 4.1. Let y=A11+s, we have
C≥A1A2B∫A10I(y)y2ds=A2B∫+∞0I(A11+s)ds. | (5.33) |
On the other hand, there exists δ such that
∫δ0I(y)y2ds=+∞. | (5.34) |
This contradicts with (5.33), so ˉTεb is bounded with respect to ε.
We set
−2δ0:=inf{v′0}<0, | (5.35) |
where δ0 is a constant independent from ε. When ε is small enough such that
ε2‖(p(g0))y‖L∞<δ02. | (5.36) |
Then on ΩεM there exists yε1(s;˜y,0)
v′0(˜y)=−2δ0, | (5.37) |
when ε<ε0 is small enough, ˜y and yε1(ˉTεb;˜y,0) are close enough such that for y∈[yε1(ˉTεb;˜y,0),˜y],
v′0(y)≤−32δ0. | (5.38) |
For y∈[yε1(ˉTεb;˜y,0),˜y], we have
(Zε0)y(y)=v′0(y)+ε2(p(g0(y)))y≤−δ0. | (5.39) |
Next, we want to show, for ε small enough, there exists M0 independent of M such that
ˉTεb−ˉτεM≤M0M. | (5.40) |
For gε, we have
∫ˉTεbˉτεMD(1gε)(yε1(s;˜y,0),s)ds=∫ˉTεbˉτεMg−10(Zε0)y(yε1(s;˜y,0),s)ds. | (5.41) |
The above equation is equivalent to
1gε(yε1(ˉτεM;˜y,0),ˉτεM)≥1gε(yε1(ˉτεM;˜y,0),ˉτεM)−1gε(yε1(ˉTεb;˜y,0),ˉTεb)=∫ˉTεbˉτεM−g−10(Zε0)y(yε1(s;˜y,0))≥δ0ˉg0(ˉTεb−ˉτεM). | (5.42) |
When M is large enough, by the convergence of (gε,vε) and the explicit expression of (ˉg,ˉv), we have for τ∈[0,τM]
|ˉJ−1ˉvy(˜y,τM)gε(yε1(ˉτεM;˜y,0),ˉτεM)|→|v′0(y)g0(y)|. | (5.43) |
There exists M0 independent of M such that
|v′0(y)g0(y)|<M02. | (5.44) |
Then, for ε small enough
|ˉJ−1ˉvy(˜y,τM)gε(yε1(ˉτεM,˜y,0),ˉτεM)|<M0. | (5.45) |
Thus
1gε(yε1(ˉτεM,˜y,0),ˉτεM)<M0|ˉJ−1ˉvy(˜y,τM)|=M0M. | (5.46) |
By (5.42) and (5.46) we obtain the following inequality
ˉTεb−ˉτεM≤ˉg0δ0⋅M0M. | (5.47) |
Next, we prove that limε→0ˉTεb=Tb. There exists y such that
τM=−1v′0(y)−1M,Tb=−1v′0(y). | (5.48) |
Thus we have
|τM−Tb|≤1M. | (5.49) |
Then we can choose M large enough such that
|ˉTεb−ˉτεM|+|τM−Tb|<δ2. | (5.50) |
Since (gε,vε) uniformly converges to (ˉg,ˉv) on any level set, there exists ε1 such that for ε∈(0,ε1], we have
|ˉτεM−τM|<δ2. | (5.51) |
Combined with the above discussion, when M is sufficiently large, for any δ>0, there exists ε1 such that for ε∈(0,ε1]
|ˉTεb−Tb|≤|ˉTεb−ˉτεM|+|ˉτεM−τM|+|τM−Tb|<δ. | (5.52) |
Therefore, we have
limε→0ˉTεb=Tb. | (5.53) |
Finally, we need to prove limε→0Tεb=Tb. By the definition of ˉTεb, we have
Tb=limε→0ˉTεb≥¯limε→0Tεb≥lim_ε→0Tεb. | (5.54) |
On the other hand, by (4.41) we have
Tb≤lim_ε→0Tεb. | (5.55) |
Therefore, we have
¯limε→0Tεb=lim_ε→0Tεb=limε→0Tεb=Tb. | (5.56) |
In this section, we will consider the convergence rates of solution on each [0,tM] in Eulerian coordinates. For uε and zε, we have the following lemma.
Lemma 6.1. For M>0, on R×[0,tM], uε∼zε∼λε1∼λε2∼ˉx∼xε2∼ˉu(O(ε2)).
Remark 6.2. λε1∼λε2∼ˉu(O(ε2)), the characteristics triangle vanishing as the ε2 order.
Proof. In order to simplify the calculation, we set
{xε1(s)=xε1(s;˜x,˜t),xε2(s)=xε2(s;˜x,˜t),ˉxε(s)=ˉxε(s;˜x,˜t), | (6.1) |
which are defined at (2.5) and (3.33). On R×[0,tM]
|xε1(s)−xε2(s)|=|∫s˜t(λε1(xε1(t),t)−λε2(xε2(t),t))dt|≤∫s˜t|λε1(xε1(t),t)−λε2(xε2(t),t)|dt≤∫s˜t|uε(xε1(t),t)−uε(xε2(t),t)|dt+C′(M)sε2≤C″(M)∫s˜t|xε1(t)−xε2(t)|dt+C′(M)sε2. | (6.2) |
By Gr¨onwall's inequality
|xε1(s)−xε2(s)|≤C′(M)sε2(eC″(M)(s−˜t)−1)≤C(M)ε2. | (6.3) |
Here, C(M) is a uniform constant with respect to ε. Thus, we can see the different characteristic lines converge with the rate O(ε2) on R×[0,tM]. Then we can have the following estimate of zε and uε
|zε(xε2(t),t)−uε(xε1(t),t)|≤|uε(xε2(t),t)−uε(xε1(t),t)|+C′(M)ε2≤C″(M)|xε1(s)−xε2(s)|+C′(M)ε2≤(C″(M)C(M)+C′(M))ε2≤C‴(M)ε2. | (6.4) |
It means that zε converges to uε with the rate of ε2. By the same way, we can prove that λε1∼λε2∼zε∼uε(O(ε2)) on R×[0,tM]. Next, we consider the convergence rate of uε to ˉu,
|ˉx(s)−xε1(s;ˉx(0),0)|≤∫s0|ˉu(ˉx(t),t)−λε1(xε1(t;ˉx(0),0),t)|dt≤∫s0|ˉu(ˉx(t),t)−uε(xε1(t;ˉx(0),0),t)|dt+C′(M)sε2≤∫s0|u0(ˉx(0))−u0(ˉx(0))|ds+C′(M)sε2≤C(M)ε2. | (6.5) |
Then, we can have the following estimate of uε and ˉu
|uε(˜x,˜t)−ˉu(˜x,˜t)|≤|uε(˜x,˜t)−uε(xε1(˜t;ˉx(0),0),˜t)|+|uε(xε1(˜t;ˉx(0),0),˜t)−ˉu(ˉx(˜t),˜t)|≤|uε(˜x,˜t)−uε(xε1(˜t;ˉx(0),0),˜t)|+|u0(ˉx(0))−u0(ˉx(0))|≤C″(M)|˜x−xε1(˜t;ˉx(0),0)|≤C″(M)|ˉx(˜t)−xε1(˜t;ˉx(0),0)|≤C″(M)C(M)ε2. | (6.6) |
We conclude that zε∼uε∼ˉu(O(ε2)).
Finally, we prove the convergence of discontinuous lines ˉx(t) and xε1(t). Without loss of generality, we consider that there is only one discontinuity at x=0, and discuss the convergence of the discontinuity lines passing through 0. For the case of separable discontinuous points, we have similar results. By (6.6) and the uniform boundedness of uεx, we have
|ˉx(s;0,0)−xε2(s;0,0)|≤∫s0|ˉu(ˉx(t;0,0),t)−uε(xε2(t;0,0),t)|dt≤∫s0|ˉu(ˉx(t;0,0),t)−uε(ˉx(t;0,0),t)|+|uε(ˉx(t;0,0),t)−uε(xε2(t;0,0),t)|dt≤C(M)sε2+∫s0|uε(ˉx(t;0,0),t)−uε(xε2(t;0,0),t)|dt≤C(M)sε2+C′(M)∫s0|ˉx(t;0,0)−xε2(t;0,0)|dt. | (6.7) |
By Gr¨onwall's inequality
|ˉx(s;0,0)−xε2(s;0,0)|≤C(M)sε2eC′(M)s≤C″(M)ε2. | (6.8) |
We conclude that the discontinuous lines converges at the rate of ε2.
This paper investigates the well-posedness of contact discontinuity solutions and the vanishing pressure limit for the Aw-Rascle traffic flow model.
For the well-posedness of contact discontinuity solutions, the Lagrangian coordinate transformation is employed to fix the discontinuity boundary. A novel method establishes the positive lower bound of density at the discontinuity, governed by a necessary and sufficient ε-condition based on the initial density jump. The result show: rarefactive initial data yields global solutions; compressive data causes finite-time singularity. The discontinuity curve is C1 with a Lipschitz tangent. Well-posedness for the pressureless fluid model is also proven.
For the vanishing pressure limit to the pressureless fluid model, a 0-condition ensures the uniform lower bound of density during the limit. By employing level sets in the Lagrangian coordinates, uniform boundedness estimates for density and velocity derivatives are derived, proving convergence of the solutions. For a class of pressure functions, blow-up times away from the discontinuity converge as the pressure vanishes. Finally, we establish the convergence rate of the solution.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Tian-Yi Wang would like to thank Professor Pierangelo Marcati for introducing him to the Aw–Rascle Traffic Flow Model in the Russian seminar. He is deeply grateful for Professor Marcati's positive influence, invaluable help and support in all aspects, and fondly recalls the wonderful time working at GSSI and living in L'Aquila. The research of Tian-Ti Wang was supported in part by the National Science Foundation of China (Grants 12371223, 12061080).
The authors declare no conflicts of interest.
[1] |
A. Aw, M. Rascle, Resurrection of "second order" models of traffic flow, SIAM J. Appl. Math., 60 (2000), 916–938. https://doi.org/10.1137/S0036139997332099 doi: 10.1137/S0036139997332099
![]() |
[2] |
L. Boudin, A solution with bounded expansion rate to the model of viscous pressureless gases, SIAM J. Math. Anal., 32 (2000), 172–193. https://doi.org/10.1137/S0036141098346840 doi: 10.1137/S0036141098346840
![]() |
[3] |
Y. Brenier, E. Grenier, Sticky particles and scalar conservation laws, SIAM J. Numer. Anal., 35 (1998), 2317–2328. https://doi.org/10.1137/S0036142997317353 doi: 10.1137/S0036142997317353
![]() |
[4] |
G. Q. Chen, H. Liu, Formation of δ-shocks and vacuum states in the vanishing pressure limit of solutions to the Euler equations for isentropic fluids, SIAM J. Math. Anal., 34 (2003), 925–938. https://doi.org/10.1137/S0036141001399350 doi: 10.1137/S0036141001399350
![]() |
[5] |
G. Q. Chen, H. Liu, Concentration and cavitation in the vanishing pressure limit of solutions to the Euler equations for nonisentropic fluids, Phys. D, 189 (2004), 141–165. https://doi.org/10.1016/j.physd.2003.09.039 doi: 10.1016/j.physd.2003.09.039
![]() |
[6] |
H. Cheng, H. Yang, Approaching Chaplygin pressure limit of solutions to the Aw–Rascle model, J. Math. Anal. Appl., 416 (2014), 839–854. https://doi.org/10.1016/j.jmaa.2014.03.010 doi: 10.1016/j.jmaa.2014.03.010
![]() |
[7] |
B. François, J. François, Duality solutions for pressureless gases, monotone scalar conservation laws, and uniqueness, Commun. Part. Diff. Eq., 24 (1999), 2173–2189. https://doi.org/10.1080/03605309908821498 doi: 10.1080/03605309908821498
![]() |
[8] |
M. Godvik, H. Hanche-Olsen, Existence of solutions for the Aw–Rascle traffic flow model with vacuum, J. Hyperbol. Differ. Eq., 5 (2008), 45–63. https://doi.org/10.1142/S0219891608001428 doi: 10.1142/S0219891608001428
![]() |
[9] |
F. Huang, Weak solution to pressureless type system, Commun. Part. Differ. Eq., 30 (2005), 283–304. https://doi.org/10.1081/PDE-200050026 doi: 10.1081/PDE-200050026
![]() |
[10] |
F. Huang, Z. Wang, Well-posedness for pressureless flow, Commun. Math. Phys., 222 (2001), 117–146. https://doi.org/10.1007/s002200100506 doi: 10.1007/s002200100506
![]() |
[11] | F. John, Formation of singularities in one-dimensional nonlinear waves propagation, In: J. Moser, Fritz John, Contemporary Mathematicians, Birkhäuser, 1974,469–497. https://doi.org/10.1007/978-1-4612-5406-5_36 |
[12] |
P. D. Lax, Development of singularities of solutions of nonlinear hyperbolic partial differential equations, J. Math. Phys., 5 (1964), 611–613. https://doi.org/10.1063/1.1704154 doi: 10.1063/1.1704154
![]() |
[13] | T. Li, W. Yu, Boundary value problems for quasilinear hyperbolic systems, Mathematics Dept., Duke University, 1985. |
[14] |
T. Li, Y. Zhou, D. X. Kong, Weak linear degeneracy and global classical solutions for general quasilinear hyperbolic systems, Commun. Part. Differ. Eq., 19 (1994), 1263–1317. https://doi.org/10.1080/03605309408821055 doi: 10.1080/03605309408821055
![]() |
[15] | T. Li, Y. Zhou, D. X. Kong, Global classical solutions for general quasilinear hyperbolic systems with decay initial data, Nonlinear Anal., 28 (1997), 1299–1332. |
[16] |
T. Liu, Development of singularities in the nonlinear waves for quasilinear hyperbolic partial differential equations, J. Differ. Equations, 33 (1979), 92–111. https://doi.org/10.1016/0022-0396(79)90082-2 doi: 10.1016/0022-0396(79)90082-2
![]() |
[17] |
Y. Lu, Existence of global bounded weak solutions to nonsymmetric systems of Keyfitz–Kranzer type, J. Funct. Anal., 261 (2011), 2797–2815. https://doi.org/10.1016/j.jfa.2011.07.008 doi: 10.1016/j.jfa.2011.07.008
![]() |
[18] |
L. Pan, X. Han, The Aw–Rascle traffic model with Chaplygin pressure, J. Math. Anal. Appl., 401 (2013), 379–387. https://doi.org/10.1016/j.jmaa.2012.12.022 doi: 10.1016/j.jmaa.2012.12.022
![]() |
[19] |
W. J. Peng, T. Y. Wang, On vanishing pressure limit of continuous solutions to the isentropic Euler equations, J. Hyperbol. Differ. Eq., 19 (2022), 311–336. https://doi.org/10.1142/S0219891622500084 doi: 10.1142/S0219891622500084
![]() |
[20] |
W. Peng, T. Y. Wang, W. Xiang, Hypersonic limit for C1 solution of one dimensional isentropic Euler equations, Commun. Math. Anal. Appl., 3 (2024), 558–581. https://doi.org/10.4208/cmaa.2024-0024 doi: 10.4208/cmaa.2024-0024
![]() |
[21] |
S. F. Shandarin, Y. B. Zeldovich, The large-scale structure of the universe: turbulence, intermittency, structures in a self-gravitating medium, Rev. Mod. Phys., 61 (1989), 185–220. https://doi.org/10.1103/RevModPhys.61.185 doi: 10.1103/RevModPhys.61.185
![]() |
[22] |
C. Shen, M. Sun, Formation of delta-shocks and vacuum states in the vanishing pressure limit of solutions to the Aw–Rascle model, J. Differ. Equations, 249 (2010), 3024–3051. https://doi.org/10.1016/j.jde.2010.09.004 doi: 10.1016/j.jde.2010.09.004
![]() |
[23] |
Z. Wang, X. Ding, Uniqueness of generalized solution for the Cauchy problem of transportation equations, Acta Math. Sci., 17 (1997), 341–352. https://doi.org/10.1016/S0252-9602(17)30852-4 doi: 10.1016/S0252-9602(17)30852-4
![]() |
[24] |
Z. Wang, F. Huang, X. Ding, On the Cauchy problem of transportation equations, Acta Math. Appl. Sin., 13 (1997), 113–122. https://doi.org/10.1007/BF02015132 doi: 10.1007/BF02015132
![]() |
[25] |
E. Weinan, Y. G. Rykov, Y. G. Sinai, Generalized variational principles, global weak solutions and behavior with random initial data for systems of conservation laws arising in adhesion particle dynamics, Commun. Math. Phys., 177 (1996), 349–380. https://doi.org/10.1007/BF02101897 doi: 10.1007/BF02101897
![]() |