Nonlinear stability of phase transition steady states to a hyperbolic-parabolic system modelling vascular networks
Abstract.
This paper is concerned with the existence and stability of phase transition steady states to a quasi-linear hyperbolic-parabolic system of chemotactic aggregation, which was proposed in [1, 13] to describe the coherent vascular network formation observed in vitro experiment. Considering the system in the half line with Dirichlet boundary conditions, we first prove the existence and uniqueness of non-constant phase transition steady states under some structure conditions on the pressure function. Then we prove that this unique phase transition steady state is nonlinearly asymptotically stable against a small perturbation. We prove our results by the method of energy estimates, the technique of a priori assumption and a weighted Hardy-type inequality.
2020 Mathematics Subject Classification. 35L60, 35L04, 35B40, 35Q92
Keywords. Hyperbolic-parabolic system; vascular network; phase transitions; nonlinear stability; energy estimates
1. Introduction
Experiments of in vitro blood vessel formation demonstrate that endothelial cells randomly dispersing on a gel substrate (matrix) can spontaneously organize into a coherent vascular network (see [1, 13] and figures therein), which is called angiogenesis - a major factor driving the tumor growth. How endothelial cells self-organize geometrically into capillary networks and how separate individual cells cooperate in the formation of coherent patterns remain poorly understood biologically up to date. These networking patterns can not be explained by the macroscopic aggregation models such as Keller-Segel type chemotaxis models that lead to point-wise blowup or rounded aggregates, nor by the microscopic kinetic models that describe individual cell behaviors, as commented in [6]. Strikingly they can be numerically reproduced by a hydrodynamic (hyperbolic-parabolic) models of chemotaxis proposed in [1, 13] as follows
(1.4) |
where denotes the endothelial cell density, the cell velocity and the concentration of chemoattractant secreted by cells; is a pressure function accounting for the fact that closely packed cells resist to compression due to the impenetrability of cellular matter, the parameter measures the intensity of cell response to the chemoattractant and corresponds to a damping (friction) force due to the interaction between cells and underlying substratum with a drag coefficient ; and are positive constants accounting for the growth and death rates of the chemoattractant. The convective term models the cell persistence (inertia effect). We refer more detailed biological interpretations on the model (1.4) to [1, 13]. The hydrodynamic system (1.4) has been (formally) derived from the following kinetic equation in [7] via the mean-field approximation
where denotes cell density distribution at time , position moving with a velocity from a compact set in , and tumbling kernel describes the frequency of changing trajectories from velocity (anterior) to (posterior) in response to a chemical concentration .
At first glance from mathematical point of view, the above hydrodynamic system (1.4) is analogous to the well-known damped Euler-Poisson system where the -equation is elliptic (i.e. ) which appears in various important applications depending on the sign of , such as the propagation of electrons in semiconductor devices (cf. [22]) and the transport of ions in plasma physics (cf. [8]) and the collapse of gaseous stars due to self-gravitation [5]. However the parabolic -equation in (1.4) will bring substantial differences in mathematical analysis and many existing mathematical frameworks developed for the Euler-Poisson system are inapplicable directly to (1.4). Due to the competing interactions between parabolic dissipation and hyperbolic anti-dissipation effect plus nonlinear convection, the global well-posedness and regularity of solutions to (1.4) is very complicated as can be glimpsed from the Euler-Poisson equations for which the understanding of solution behaviors is rather incomplete despite of numerous studies attempted. Up to date, there are only few results obtained for (1.4) in the literature. First when the initial value is a small perturbation of a constant ground state in with sufficiently small, the global existence and stability of solutions with non-vacuum (i.e. ) to (1.4) was established in [11, 9]. The linear stability of constant ground state was obtained under the condition in [16] where an additional viscous term is supplied to the second equation of (1.4). The stationary solutions of (1.4) with vacuum (bump solutions) in a bounded interval with zero-flux boundary condition and in were constructed in [2] and further elaborated in [3]. The model (1.4) with and periodic boundary conditions in one dimension was numerically explored in [12]. These appear to be the only results available to the system (1.4) in the literature and further studies are in demand. For results on some other types of hyperbolic-parabolic chemotaxis models, we refer to [10, 17, 18, 24, 27] and references therein.
Note that the above-mentioned mathematical works on (1.4) prescribe initial data as a small perturbation of constant equilibria and the large-time profile of solutions is also constant, which can not explain the experimental observations of [1, 13] showing prominent phase transition patterns connecting regions clear (or low density) of cells. This motivates us to explore the possible non-constant phase transition profiles of solutions. The aim of this paper is to study the existence and stability of phase transition steady states without vacuum to the system (1.4) in one dimensional half space . For the convenience of presentation in the sequel, we set , namely denotes the momentum of cells, and recast the one-dimensional system (1.4) in as
(1.5a) | |||||
(1.5b) | |||||
(1.5c) |
As in [16], we assume the pressure function depends only on the density and satisfies
(1.6) |
A typical form of is with . We supplement the system (1.5a)–(1.5c) with the following boundary conditions
(1.7) |
and the initial value
(1.8) |
where , , are constants and .
In this paper, we shall first use delicate analysis to show that the system (1.5a)–(1.5c) has a unique non-constant stationary solution without vacuum satisfying
where and are monotone. Then we show that the stationary solution is asymptotically stable, namely the solution of (1.5a)–(1.5c) converges to point-wisely as if the initial value is an appropriate small perturbation of . The monotonicity of and indicates that the steady states have phase transition profiles. To the best of our knowledge, there are not such results available in the literature for (1.5a)–(1.5c) and even for the Euler-Poisson equations (cf. [26, 21, 20, 19, 14]. To prove our results, we fully capture the structure of (1.5a)–(1.5c) where the stationary solution has exponential decay at far field, which is not enjoyed by the Euler-Poisson (or Euler) equations. Though part of the proof of our results is inspired by some ideas of [14, 25, 26, 15] on Euler-Poisson or Euler equations, we have added lof of extra efforts to deal with complicated couplings and boundary effects. The coupling term leads to two linear terms in the linearized system around stationary solutions and how to make these linear terms under control is key to the asymptotic stability against small perturbations. We resolve this issue by the structural assumption (1.6) to take up the dissipation and use the exponentially weighted Hardy inequality in half space to compensate for the lack of dissipation in the hyperbolic equations. Due to the couplings and boundary effects, the energy estimates are very sophisticated, where the lower-order estimates involve higher-order estimates and vice versa. We use the delicate energy estimates along with the technique of a priori assumptions to unravel these tangles and gradually achieve our results.
The rest of this paper is organized as follows. In section 2, we state our main results on the existence of non-constant stationary solutions (Theorem 2.1) and stability of stationary solutions (Theorem 2.2). In section 3, we study the stationary problem and prove Theorem 2.1. The proof of Theorem 2.2 is given in Section 4.
2. Statement of main results
In this section, we shall state the main results of this paper. To be precise, we first introduce some notations used. Throughout the paper, we use , and to denote the norms of usual , and the standard Sobolev space , respectively. We also use to denote for some . We denote by a generic constant that may vary in the context, and by a constant depending on . Occasionally, we simply write if for some constant .
It can be verified that the system (1.5a)–(1.5c) possesses the following energy functional (cf. [2, 7])
which, subject to the boundary condition (1.7), satisfies that
where . Thus the stationary solution satisfying gives rise to and in . Since we are interested in non-constant profile for , is the only (physical) stationary profile for the velocity . Therefore stationary solutions of (1.5a)–(1.5c) without vacuum must possess the form , where satisfies
(2.1a) | |||||
(2.1b) | |||||
(2.1c) | |||||
(2.1d) |
Here the pressure satisfies (1.6) and the constants and are the same as in (1.7) and (1.8).
Then our first result concerning the existence and uniqueness of solutions to the stationary problem (2.1a)–(2.1d) is given below.
Theorem 2.1.
Remark 2.1.
Our second result is the asymptotic stability of the stationary solutions obtained in Theorem 2.1, which is stated in the following theorem.
Theorem 2.2.
Remark 2.2.
With the condition , we can define the initial values of and through the equation for . That is
(2.6) | ||||
(2.7) |
These initial values of time derivatives are of importance in deriving the higher-order estimates in section 4. Furthermore, we always assume that the initial data is compatible with the boundary conditions at .
3. Stationary problem (Proof of Theorem 2.1)
In this section, we shall study the stationary problem (2.1a)–(2.1d) and complete the proof of Theorem 2.1. To this end, we first reformulate our problem (2.1a)–(2.1d), and then prove the existence and uniqueness of solutions. Finally, we derive the monotone and decay properties of solutions.
3.1. Reformulation of our problem
We start by proving the following lemma, which plays a key role in the reformulation of our problem.
Lemma 3.1.
If is a solution to the problem
(3.1) |
where is a continuous function, is a constant. Then we have
Proof.
Since is continuous, and , we have . It remains to show . We proof this by contradiction. Supposing that , without loss of generality, we assume . Thanks to the continuity of and , there exists a constant such that for any , . This along with implies that
which contradicts the fact . Hence, . The proof of Lemma 3.1 is complete.
∎
Define
(3.2) |
for any . Clearly, . We claim that under the conditions , (2.2) and , there exists a unique constant such that
(3.3) |
and
(3.4) |
Indeed, in view of (1.6), we know that
(3.5) |
This implies that the function is strictly monotonically increasing. Furthermore, we have if , and . For the case , since , then there exists a unique constant such that (3.3) holds. For the case , we have . If , we know that , then similar to the case , there exists a unique constant such that . Now it remains to consider the case when and . In this case, since the is continuous, monotonic and bounded below, we can extend by defining . Then the extended function is continuous on . Furthermore, from (2.2), we get . Hence, there exists a unique constant such that . Then (3.3) is proved. Moreover, with the help of (3.3) and (3.5), we immediately get (3.4). We thus finish the proof of the claim.
To proceed, assume that is a classical solution to (2.1a)–(2.1d) with . Dividing (2.1a) by and integrating the resulting equation over , we get
(3.6) |
where is as in (3.2). Sending along with (2.1c), (3.3) and the fact , we get
By using the monotonicity and continuity of , we further have that
Inserting (3.6) into (2.1b), we get
(3.7) |
Multiplying (3.7) by , it follows that
(3.8) |
with
(3.9) |
Thus,
(3.10) |
for some constant and some function with
(3.11) |
By virtue of (1.6) and (3.9), we get
(3.12) |
for any . Thanks to the condition , it holds that . This along with (3.12) yields that
(3.13) |
Then by (3.5), (3.11) and (3.13), we get if and if . This gives
(3.14) |
for any . We claim that . Otherwise, we have or . If , by the continuity of and , there exists a constant such that if ,
Then for . This contradicts to (3.10). If , using (3.14), we get for any . Therefore, for any , it holds that
With the fact and Lemma 3.1, we have . This is a contradiction. Hence, we have and
(3.15) |
This together with (3.5) and (3.14) implies that if and if , and that
(3.16) |
for any . Hence, we can solve from (3.15) that
where we have used (3.4).
Summing up, we have the following lemma.
Lemma 3.2.
Lemma 3.3 (Reformulation).
Proof.
In view of Lemma 3.2, it remains to show that if is a solution to the problem (3.17), then solves the problem (2.1a)–(2.1d). By using and (3.5), one can easily derive (2.1a). Thanks to (3.3) and , we have and . To show (2.1b), by (3.8), (3.15) and , it suffices to show that for any . We prove this for the case , and the proof for the case is similar. Since , we have for any . Denote
We claim that is Lipschitz continuous on . With this claim, we can prove that for , and hence finish the proof. Indeed, if there exists a point such that , then from (3.16), we have . This implies that is a solution to the following problem
(3.18) |
Since is Lipschitz continuous on , the problem (3.18) admits a unique solution on . While is also a solution to (3.18), and obviously, . This is a contradiction. Therefore, for any . Now it remains to prove the claim that is Lipschitz continuous on . With the help of (1.6), (3.5) and (3.14), we know that is differentiable if . Furthermore, a direct computation gives
(3.19) |
By using (3.11), (3.12) and L’Hôpital’s rule, we have
(3.20) |
From (3.13) and (3.14), we get for . This along with (3.20) yields that
Therefore,
Notice that for , and that
due to (3.11), (3.12) and L’Hôpital’s rule, we have
where is the right derivative of at . We thus have . This in combination with (3.19) yields that is continuous on , and thus for some constant depending on and . This implies that is Lipschitz continuous on . The proof of the present lemma is complete. ∎
3.2. Existence and uniqueness of solutions
In this section, we will prove that the problem (2.1a)–(2.1d) admits a unique solution with . Thanks to Lemma 3.3, it now suffices to consider the problem (3.17). As before, we focus only on the case , the proof for the case is similar and so omitted. Let us begin with the following ODE problem
(3.21) |
By the Lipschitz continuity of on , we conclude that the problem (3.21) admits a unique solution on for some . Then by the contradiction argument and discussions in Step 1 on the uniqueness of solutions to (3.18), we get for any . This, along with the standard extension theorem for ordinary differential equations, implies that the solution to the problem (3.21) exists globally in , and for any , . In addition, notice from (3.16) that if and only if , we have that
(3.22) |
and that exists. Denoting , from Lemma 3.1, we obtain . This combined with (3.14) give rises to . With at hand, we can define from . Clearly, is a solution to (3.17) for . Finally, since , the uniqueness of solutions can be proved by the Lipschitz continuity of on .
3.3. Monotonicity and decay properties
Recalling (3.4) and (3.22), we get if . In a manner similar to the derivation of (3.22), we have if . With the help of (3.5), (3.6) and the properties of , we have
This gives (2.3). Now let us turn to the decay properties of the solution. By using (3.3) and (3.5), we get
(3.23) |
Thus implies . In the following, without loss of generality, we assume that , and thus for any continuous function defined on , depends only on . If , recalling (2.3) and (3.17), we have
(3.24) |
for any . Owing to (1.6), (3.5), (3.11) and the condition , we get
(3.25) |
for any , where is a constant depending on . From (3.11) and (3.13), we get . This combined with (3.5), (3.3) and the Taylor expansion implies that
(3.26) |
provided is suitably small, where is a positive constant depending on . Combining (3.3) with (3.24), we get
for some constant depending on , provided is suitably small, where we have used (1.6) and (3.5). Consequently, with (3.23), we have the following decay estimate:
(3.27) |
For the case , it holds that
(3.28) |
Using (3.5), (3.3) and (3.28), we get
that is,
for some constant depending on , provided is suitably small. It thus holds that
Finally, by (2.1b), (3.3), (3.6), (3.7) and (3.27), we get (2.4). The proof is complete.
4. Global existence and asymptotic stability
In this section, we are devoted to studying the asymptotic stability of the unique stationary solution to (1.5a)–(1.5a) obtained in Section 3. To this end, we first reformulate the problem with the technique of taking anti-derivative for .
4.1. Reformulation of problem
Combining (1.5a)–(1.5c) with (1.5a)–(1.5c), we have
(4.1a) | |||||
(4.1b) | |||||
(4.1c) |
It follows from (4.1a) that
which, together with the condition in Theorem 2.2, gives
Defining the perturbation function
(4.2) |
with
we get the reformulated problem:
and its linearized problem around is
(4.3a) | |||||
(4.3b) | |||||
(4.3c) | |||||
(4.3d) | |||||
(4.3e) |
where
(4.4) |
and
(4.5) |
To proceed, we define the solution space of the problem (4.3a)–(4.3d) as follows:
for any .
Since we are interested in the case where the solution has no vacuum, naturally we require that , namely
(4.6) |
For simplicity, we denote
Then by the standard parabolic theory and fixed point theorem (cf. [23]), we have the following local existence result.
Proposition 4.1 (Local existence).
In what follows, we are devoted to proving the following theorem on the global existence and uniqueness of solutions to the problem (4.3a)–(4.3d).
Proposition 4.2.
4.2. Some preliminaries
The proof of Proposition 4.2 is based on the combination of the local existence result in Proposition 4.1 with the a priori estimates given in (4.7)-(4.8). In the sequel, we assume that is a solution to the problem (4.3a)–(4.3d) obtained in Proposition 4.1 for some and derive the a priori estimates (4.7)-(4.8) based on the technique of a priori assumption. That is we first assume that the solution of (4.3a)–(4.3d) satisfies
(4.9) |
where is a constant to be determined later, and then derive the a priori estimates to obtain the global existence of solutions. Finally we justify that the global solutions obtained satisfy the above a priori assumption and thus close our argument.
Using the fact from (4.3e) and the Sobolev inequality, we have
(4.10) |
Denote , by (2.4), one can find a constant depending on , and such that
(4.11) |
provided is suitably small. Combining (4.11) with (4.10), we get
(4.12) |
for some constant depending on , and , provided and are small enough. The boundary condition (4.3d) together with the equation (4.3a) leads to the following boundary conditions on higher-order derivatives:
(4.13) |
Moreover, the following Hardy inequality plays a key role in deriving the a priori estimates.
Lemma 4.1.
Let be a constant, it holds that
(4.14) |
for any , where is a constant depending on but independent of .
4.3. Energy estimates
In this section, we will derive the some estimates for the solution of (4.3a)–(4.3d) under the a priori assumption (4.9) by the method of energy estimates. The estimates for follows from the fact .
We begin with the lower-order estimates.
Lemma 4.2.
Proof.
Multiplying (4.3a) by and integrating the resulting equation over , we get
(4.16) |
By the Taylor expansion, we get
for some . Then it follows from (1.6), (4.5), (4.10) and (4.11) that
(4.17) |
provided and are suitably small. Integrating by parts, we have
(4.18) |
In view of (4.10) and the Cauchy-Schwarz inequality, we deduce
(4.19) |
By the fact from (2.4) and the Hardy inequality (4.14), it holds that
(4.20) |
where the Cauchy-Schwarz inequality has been used. Inserting (4.19) and (4.3) into (4.3) leads to
(4.21) |
For the last term on the right-hand side of (4.3), from (4.4), (4.10), (4.12), integration by parts and Cauchy-Schwarz inequality, we have
(4.22) |
Substituting (4.17), (4.21) and (4.22) into (4.3), we get
(4.23) |
Multiplying (4.3b) by and integrating the resulting equation over , one has
where, due to the fact from (2.4) and the Hardy inequality (4.14), the following inequality holds
Therefore,
(4.24) |
Combining (4.24) with (4.3), we obtain
(4.25) |
for suitably small and . By (1.6) and (4.11), we have
for some constant independent of . Then, for sufficiently small and , we have from (4.3) that
(4.26) |
for some constant , where we have used (4.11). Now let us turn to the estimate for . Multiplying (4.3a) by and integrating the resulting equation over , we get
(4.27) |
Next, we estimate the terms on the right-hand side of (4.27). First, it follows from a direct computation that
Second, due to (4.10), Cauchy-Schwarz inequality and the fact from (2.4), we have
(4.28) |
where we have used the fact by the Hardy inequality (4.14). Finally, using (4.10), (4.12) and integration by parts, one has
(4.29) |
Hence, for suitably small and , we find from (4.27) that
(4.30) |
where the terms on the right-hand side of (4.3) and (4.29) have been absorbed. To proceed, we multiply (4.3b) by and integrate the resulting equation over to get
(4.31) |
where (2.4) and the Cauchy-Schwarz inequality have been used. Combining (4.3) with (4.3) gives
(4.32) |
where
due to (1.6) and (4.11). Given any constant , adding (4.26) with (4.3) multiplied by leads to
(4.33) |
where is as in (4.26). From (4.11), it holds that
for some constant which depends on , and . Taking large enough such that and
for some constant , then for suitably small and , we have from (4.3) that
(4.34) |
where
Applying the Taylor expansion to the function along with (4.10) leads to
(4.35) |
for some constant . With (4.35), integrating (4.3) with respect to , by taking and suitably small, we get (4.15) and hence complete the proof. ∎
Lemma 4.3.
Proof.
Differentiating (4.3a)–(4.3b) with respect to , we get
(4.37a) | |||||
(4.37b) |
Multiplying (4.37a) by , and integrating it over , we get, thanks to (4.13), that
(4.38) |
Recalling the definitions of and in (4.5), using (1.6), (4.11), (4.10), the fact from (2.4) and Cauchy-Schwarz inequality, we have
(4.39) |
and
where we have used the following identity
(4.40) |
due to the Taylor expansion. For the last term on the right hand of (4.3), thanks to (2.4), (4.4), (4.10), (4.12), Cauchy-Schwarz inequality and the Hardy inequality (4.14), it holds for suitably small and that
(4.41) |
We thus conclude from (4.3)–(4.3) that
(4.42) |
Multiplying (4.37b) by , and integrating it to get
(4.43) |
where (2.4), (4.11) and Cauchy-Schwarz inequality have been used. Since at and hence at , recalling (4.3b), we have
(4.44) |
This along with the Sobolev inequality and Young’s inequality implies that
(4.45) |
for any . Substituting (4.3) into (4.3), we get
(4.46) |
Combining (4.3) with (4.3), we get after taking , and suitably small that
(4.47) |
where we have used (4.11) and the following inequality
(4.48) |
due to (1.6). Next, we integrate (4.37a) multiplied by over to get
(4.49) |
A direct computation along with (2.4) and Cauchy-Schwarz inequality gives
(4.50) |
Recalling (4.3), we arrive at
(4.51) |
From (1.6), (2.4), (4.10) and (4.12), it holds that
Therefore, we have
(4.52) |
Similar to (4.3)–(4.3), the third term on the right-hand side of (4.3) can be estimated as follows:
(4.53) |
Noticing
we get, thanks to integration by parts and the boundary condition at , that
(4.54) |
Next, we estimate and . First, we utilize (2.4), (4.10) and (4.12) to get
(4.55) |
provided and are suitably small. For , by using (2.4), (4.10), (4.12), Cauchy-Schwarz inequality and the Hardy inequality (4.14), we obtain
(4.56) |
where we have used the following inequality
(4.57) |
due to (2.4), (4.3a), (4.4), (4.5), (4.10) and (4.12). With (4.3) and (4.3), we update (4.3) as
(4.58) |
Substituting (4.3), (4.3), (4.3) and (4.58) into (4.3), we get
(4.59) |
Multiplying (4.37b) by , and integrating it to get
(4.60) |
where, in view of (2.4), (4.44), the Sobolev inequality and Cauchy-Schwarz inequality, the following inequalities hold:
for any . Then, combining (4.3) with (4.3) gives
(4.61) |
for any , provided and are suitably small, where we have used (4.11).
Finally, similar to the proof of Lemma 4.2, adding (4.3) with (4.3) multiplied by a constant , it follows that
(4.62) |
where , is given by
(4.63) |
Taking large enough such that
for some constant independent of , recalling (4.48), we have
By (1.6), (2.4), (4.10), (4.12), Cauchy-Schwarz inequality and the Taylor expansion, we have
Therefore, for sufficiently small and , we have from (4.3) that
(4.64) |
With (4.64), after integrating (4.3) over and taking , , sufficiently small, we get (4.3) and thus finish the proof of Lemma 4.3. ∎
To close the a priori assumption (4.9), some higher-order estimates of solutions are needed. Let us begin with the estimates on .
Lemma 4.4.
Proof.
Multiplying (4.37a) by followed by an integration over , we obtain
(4.66) |
Denote
then (4.3) can be rewritten as
(4.67) |
Noting that
we utilize (1.6), (2.4), (4.10), (4.11) and the mean value theorem to get
Then by (2.4) and Cauchy-Schwarz inequality, we have
(4.68) |
A direct computation leads to
(4.69) | ||||
(4.70) |
Combining the above identities with (1.6), (2.4), (4.10)–(4.12) and the Taylor expansion yields that
(4.71) | |||
(4.72) |
We thus deduce that
(4.73) |
for suitably small and , and that
(4.74) |
Here (1.6), (4.11) and Cauchy-Schwarz inequality have been used. Thanks to (4.3), (4.3) and (4.74), it follows that
(4.75) |
where we have used (1.6) and the bounds of . Now let us turn to the estimates of . From (4.3a), we get
(4.76) |
A direct computation leads to
(4.77) |
This along with (1.6), (2.4), (4.10) and (4.12) implies that
(4.78) |
for sufficiently small and . Therefore, it holds that
(4.79) |
where we have used (1.6), (2.4), (4.11), (4.71) and (4.78). Resorting to (2.4), (4.79) and Cauchy-Schwarz inequality, we get
For , a direct computation gives
where, due to (1.6), (2.4) and (4.10), can be estimated as follows
Then, owing to (1.6), (4.10)–(4.12), (4.79), Cauchy-Schwarz inequality and the mean value theorem, we have
for any . To deal with , we rearrange in (4.77) as follows
(4.80) |
with
(4.81) |
for suitably small and , due to (2.4), (4.10) and (4.12). Based on (4.3), we split into three parts:
Next, we estimate . Recalling (4.69) and (4.70), we have
where, in view of (1.6), (2.4), (4.10) and (4.11), the following inequality holds
We thus have
and
where we have used (1.6), (2.4), (4.10), (4.12), (4.57), Cauchy-Schwarz inequality and the integration by parts. For , the integration by parts leads to
which combined with (4.4), (4.5), (4.10), (4.81) and Cauchy-Schwarz inequality implies that
Hence, we have
Plugging into (4.67), we now reach
(4.82) |
for any . Consequently, thanks to (1.6), (4.10)–(4.12), (4.75) and the mean value theorem, we obtain (4.4) after integrating (4.3) over and taking , and small enough. The proof of Lemma 4.4 is complete. ∎
In the next lemma, we shall estimate the higher-order terms on the right-hand side of (4.4).
Lemma 4.5.
Proof.
We divide the proof into two steps.
Step 1: Estimates on . Differentiating (4.37b) with respect to , we have
(4.85) |
Multiplying (4.85) by and integrating the resulting equation over , we get by the Hölder’s inequality that
(4.86) |
In view of (4.44), for smooth solution , we have at . Then it holds that
(4.87) |
for any , where the Sobolev inequality and Cauchy-Schwarz inequality have been used. Plugging (4.3) into (4.3), we get after taking suitably small that
(4.88) |
where we have used (2.6). By (4.37b) and (4.3), we have
(4.89) |
Differentiating in (4.3b) with respect to twice gives
(4.90) |
Multiplying (4.90) by , we have
(4.91) |
Thanks to (1.6), (2.4), (4.11), (4.71), (4.72), (4.76), (4.78), Cauchy-Schwarz inequality and integration by parts, one has
(4.92) |
for any . Substituting (4.3) into (4.91), we get after taking , , small enough that
(4.93) |
where we have used (2.7). Combining (4.3) with (4.3) and (4.3), then we get (4.5).
Step 2: Estimates on . Multiplying (4.37a) by with as in (3.3), and integrating the resulting equation over for any , we get
(4.94) |
Thanks to (1.6) and (4.11), one can find a constant such that
Now let us estimates term by term. Recalling (4.72), it holds that
and that
for any , where the Cauchy-Schwarz has been used. Again, by Cauchy-Schwarz inequality, it holds that
From the boundary conditions in (4.13), we get
Thanks to (1.6), (2.4) and (4.10)–(4.12), we further have that
(4.95) |
By (4.95), Cauchy-Schwarz inequality, the Sobolev inequality and the integration by parts, we have
and
For , we utilize (4.71), (4.78) and Cauchy-Schwarz inequality to get
Inserting the estimates for into (4.3), we get
provided and are suitably small. The proof of Lemma 4.5 is complete. ∎
Lemma 4.6.
Under the conditions of Proposition 4.2, for any , we have
(4.96) |
provided and are sufficiently small.
Proof.
First, to control the terms related to on the right-hand side of (4.4), we add (4.4) with (4.5) multiplied by a large positive constant to get
(4.97) |
provided and are suitably small, where the Cauchy-Schwarz inequality has been used. Combining (4.3) with (4.5), for sufficiently small and , we have
This gives rise to (4.6) and thus finishes the proof of Lemma 4.6. ∎
4.4. Proof of Proposition 4.2
By the local existence result in Proposition 4.1 and the standard extension criterion, it suffices to show the estimates (4.7) and (4.8) to prove Proposition 4.2. We first close the a priori assumption (4.9). To this end, we add (4.6) to (4.3) multiplied by a suitable positive constant and get
where the smallness of and has been used. This along with the Cauchy-Schwarz inequality gives
(4.98) |
Furthermore, multiplying (4.15) by a suitable positive constant, and adding the resulting inequality to (4.4), we obtain
(4.99) |
Then by setting
and taking suitably small, we have
which hence closes the a priori assumption (4.9). To complete the proof of (4.7)-(4.8), now it remains to show the following
Collecting (4.37b), (4.57), (4.79) and (4.4), one immediately has
To derive the estimate for , we first deduce from (4.3b) and (4.4) that
(4.100) |
Next, differentiating (4.3b) with respect to leads to
This along with (4.100) and (4.4) yields that
Finally, differentiating (4.37b) with respect to , we have
and thus
The proof of Proposition 4.2 is complete.
4.5. Proof of Theorem 2.2
In view of Proposition 4.2, the problem (1.5a)–(1.5c), (1.7)–(1.8) admits a unique classical solution in . Moreover, thanks to (4.2), (4.7) and (4.8), it holds that
and that
(4.101) |
for any . In the following, we shall prove the large time behavior of as in (2.5). For this, recalling the Sobolev inequality , it suffices to show that
(4.102) |
In fact, with the help of (4.101) and Cauchy-Schwarz inequality, we get
(4.103) |
The estimate (4.101) in combination with (4.5) gives (4.102). Then (2.5) is proved and we complete the proof of Theorem 2.2.
Acknowledgement
G. Hong is partially supported from the CAS AMSS-POLYU Joint Laboratory of Applied Mathematics postdoctoral fellowship scheme. H.Y. Peng was supported from the National Natural Science Foundation of China (No. 11901115), Natural Science Foundation of Guangdong Province (No. 2019A1515010706). Z.-A. Wang was supported in part by the Hong Kong RGC GRF grant No. PolyU 153055/18P (P0005472) and an internal grant No. ZZKN from HKPU (P0031013). C.J. Zhu was supported by the National Natural Science Foundation of China (No. 11771150, 11831003 and 11926346) and Guangdong Basic and Applied Basic Research Foundation (No. 2020B1515310015).
References
- [1] D. Ambrosi, F. Bussolino, and L. Preziosi, A review of vasculogenesis models, J. Theore. Med., 6 (2005), pp. 1–19.
- [2] F. Berthelin, D. Chiron, and M. Ribot, Stationary solutions with vacuum for a one-dimensional chemotaxis model with nonlinear pressure, Comm. Math. Sci., 14 (2016), pp. 147–186.
- [3] J. Carrillo, X. Chen, Q. Wang, Z. Wang, and L. Zhang, Phase transitions and bump solutions of the Keller-Segel model with volume exclusion, SIAM J. Appl. Math., 80 (2020), pp. 232–261.
- [4] J.-A. Carrillo, J.-Y. Li, and Z.-A. Wang, Boundary spike-layer solutions of the singular Keller-Segel system: existence and stability, Proc. London Math. Soc., Doi:10.1112/plms.12319, 2020.
- [5] S. Chandrasekhar, An introduction to the study of stellar structure, vol. 2, Courier Corporation, 1957.
- [6] P.-H. Chavanis, Jeans type instability for a chemotactic model of cellular aggregation, Eur. Phys. J. B, 52 (2006), pp. 433–443.
- [7] P.-H. Chavanis and C. Sire, Kinetic and hydrodynamic models of chemotactic aggregation, Physica A, 384 (2007), pp. 199–222.
- [8] A. Choudhuri, The physics of fluids and plasmas: an introduction for astrophysicists, Cambridge University Press, 1998.
- [9] C. Di Russo, Analysis and numerical approximations of hydrodynamical models of biological movements, Rend. Mat. Appl. (7), 32 (2012), pp. 117–367.
- [10] C. Di Russo, Existence and asymptotic behavior of solutions to a semilinear hyperbolic-parabolic model of chemotaxis, Indiana Univ. Math. J., 65 (2016), pp. 493–533.
- [11] C. Di Russo and A. Sepe, Existence and asymptotic behavior of solutions to a quasi-linear hyperbolic-parabolic model of vasculogenesis, SIAM J. Math. Anal., 45 (2013), pp. 748–776.
- [12] F. Filbet and C.-W. Shu, Approximation of hyperbolic models for chemosensitive movement, SIAM J. Sci. Comput., 27 (2005), pp. 850–872.
- [13] A. Gamba, D. Ambrosi, A. Coniglio, A. De Candia, S. Di Talia, E. Giraudo, G. Serini, L. Preziosi, and F. Bussolino, Percolation, morphogenesis, and burgers dynamics in blood vessels formation, Phys. Rev. Lett., 90 (2003), p. 118101.
- [14] L. Hsiao and T.-P. Liu, Convergence to nonlinear diffusion waves for solutions of a system of hyperbolic conservation laws with damping, Comm. Math. Phys., 143 (1992), pp. 599–605.
- [15] F.-M. Huang, M. Mei, Y. Wang, and H.-M. Yu, Asymptotic convergence to stationary waves for unipolar hydrodynamic model of semiconductors, SIAM J. Math. Anal., 43 (2011), pp. 411–429.
- [16] R. Kowalczyk, A. Gamba, and L. Preziosi, On the stability of homogeneous solutions to some aggregation models, Disc. Cont. Dyn. Syst. Series B, 4 (2004), pp. 203–220.
- [17] D. Li, T. Li, and K. Zhao, On a hyperbolic-parabolic system modeling chemotaxis, Math. Models Meth. Appl. Sci., 21 (2011), pp. 1631–1650.
- [18] T. Li and Z.-A. Wang, Nonlinear stability of traveling waves to a hyperbolic-parabolic system modeling chemotaxis, SIAM J. Appl. Math., 70 (2010), pp. 1522–1541.
- [19] C.-K. Lin, C.-T. Lin, and M. Mei, Asymptotic behavior of solution to nonlinear damped -system with boundary effect, Int. J. Numer. Anal. Model. Ser. B, 1 (2010), pp. 70–92.
- [20] H.-F. Ma and M. Mei, Best asymptotic profile for linear damped -system with boundary effect, J. Differential Equations, 249 (2010), pp. 446–484.
- [21] P. Marcati, M. Mei, and B. Rubino, Optimal convergence rates to diffusion waves for solutions of the hyperbolic conservation laws with damping, J. Math. Fluid Mech., 7 (2005), pp. S224–S240.
- [22] P. Markowich, C. Ringhofer, and C. Schmeiser, Semiconductor equations, Springer Science & Business Media, 2012.
- [23] A. Matsumura, Global existence and asymptotics of the solutions of the second-order quasilinear hyperbolic equations with the first-order dissipation, Publ. Res. Inst. Math. Sci., 13 (1977/78), pp. 349–379.
- [24] M. Mei, H. Peng, and Z.-A. Wang, Asymptotic profile of a parabolic-hyperbolic system with boundary effect arising from tumor angiogenesis, J. Differential Equations, 259 (2015), pp. 5168–5191.
- [25] K. Nishihara, Convergence rates to nonlinear diffusion waves for solutions of system of hyperbolic conservation laws with damping, J. Differential Equations, 131 (1996), pp. 171–188.
- [26] K. Nishihara and T. Yang, Boundary effect on asymptotic behaviour of solutions to the -system with linear damping, J. Differential Equations, 156 (1999), pp. 439–458.
- [27] M. Zhang and C. Zhu, Global existence of solutions to a hyperbolic-parabolic system, Proc. Amer. Math. Soc., 135 (2007), pp. 1017–1027.