The Riemann problem for a generalised Burgers equation with spatially decaying sound speed. II General qualitative theory
Abstract
We establish that the initial value problem for a generalised Burgers equation considered in part I of this paper, [3], is well-posed. We also establish several qualitative properties of solutions to the initial value problem utilised in [3].
MSC2020: 35K58, 35K15, 35A01, 35A02.
Keywords: Burgers equation, Cauchy problem, well-posed, classical solution.
1 Introduction
In this paper, we establish that the initial value problem for the generalised Burgers equation considered in [3], is well-posed (specifically, see Theorem 2.1). To establish the existence result, we adopt the approach used in [5], and note, that related standard existence results for classical solutions in [1], and similar sources, cannot be applied due to insufficient regularity of solutions to the initial value problem as . The approach is centered on establishing sufficient regularity on solutions to an implicit integral equation, to establish that they are equivalent to classical solutions to the initial value problem. We subsequently establish uniqueness and continuous dependence results for solutions to the initial value problem, via maximum principles in [2], and, the Grönwall inequality in [4], respectively.
1.1 The Initial Value Problem
Let , and . We consider the Cauchy problem for given by:
(1.1) | ||||
(1.2) | ||||
(1.3) | ||||
(1.4) |
Here and is the prescribed initial data, which is Lebesgue measurable, with . We denote the Cauchy problem given by (1.1)-(1.4) as [IVP]. It should be noted that, via (1.4), at each point at which is continuous, then as . Moreover, when is continuous for , then uniformly for . We later consider the specific case of [IVP] with given by,
(1.5) |
for . For initial data given by (1.5), observe that we can replace (1.4) with , on , and .
2 Qualitative Properties of [IVP]
We introduce the fundamental solution to the heat equation on , as , given by
(2.1) |
with . Properties of which are used to establish the existence of solutions to [IVP] are given in Appendix A.
To establish global existence and uniqueness of solutions to [IVP], and local well-posedness in time, we consider an alternative to [IVP]. By applying a Duhamel principle, it follows that if is a solution to [IVP] then
(2.2) | ||||
(2.3) |
We will now demonstrate that there exists a local solution to (2.2) and (2.3). The existence and regularity results for solutions to (2.2) and (2.3), follow a similar approach to that developed in [5]. To begin, we have
Proposition 2.1.
Proof.
Consider the set of functions which satisfy (2.3) on and are such that
(2.5) |
Next, consider the mapping given by for where
(2.6) |
for all . Observe that the first term in the right hand side of (2.6) is the solution to the heat equation with measurable initial data , and in particular, is contained in , with bound on . Also, using (A.1) and (A.2), it follows that the integrand of the second term on the right hand side of (2.6) is absolutely integrable, and hence the integral is well-defined, and bounded on , for each . Moreover, via (A.5), (A.6), (A.8) and (A.9), it follows that the second term in the right hand side of (2.6) is continuous on for each . Furthermore, for each , and all , observe, via (A.1) and (A.2), that
(2.7) |
Consequently it follows from (2.5)-(2.7) that . Now for we have
(2.8) |
for all , again using (A.1) and (A.2). It follows from (2.8) and (2.4) that
(2.9) |
for all , and hence, is a contraction mapping. Since the metric space is complete, it follows that there exists such that , i.e. is a solution to (2.2) and (2.3), as required. ∎
We now establish that the solution to (2.2) and (2.3) given in Proposition 2.1 is twice (once) continuously differentiable with repsect to (), and hence, is a local solution to [IVP]. To begin, for a solution to (2.2) and (2.3), we define the sequence of functions to be
(2.10) |
for all and with . Observe that for each , we have , via (A.5), (A.6), (A.8) and (A.9). Moreover, as , converges to uniformly on compact subsets of . Next we have
Proposition 2.2.
Proof.
Let be the solution to (2.2) and (2.3) given by Proposition 2.1. Then via (A.5), for each , it follows that
(2.12) |
for all . Moreover, it follows from (A.5), (A.6) and (2.4) that
(2.13) | ||||
(2.14) |
for all . Inequality (2.11) follows from (2.12) and (2.14), as required.
∎
Consequently we have
Proposition 2.3.
Proof.
For given by (2.10), since is measurable and , exists, is continuous, and is given by
(2.17) |
for all . After a change of variables , the second integral in (2.17) can be expressed as
(2.18) |
for all . Now, the second integral in (2.18) can be expressed as
(2.19) |
for all . Using Proposition 2.2, , and the mean value theorem, it follows that for each , there exists a constant such that
(2.20) |
for all . Therefore, the absolute value of the integral in (2.19) is bounded above by
(2.21) |
for all . Therefore, via (2.17), (2.18), (2.19) and (2.21), it follows from (A.2) that
(2.22) |
for all . We now demonstrate that converges uniformly on compact subsets of , to a the continuous limit , as . It follows from (A.2) and (2.20) that
for all . Therefore, via (2.17), it follows that is uniformly convergent on (compact subsets of) . It thus follows that there exists a continuous limit of on , which coincides with the derivative . The bound in (2.15) follows immediately from (2.22). As a consequence can be represented, alternatively, as
(2.23) |
for all . Finally, from (2.23), (2.15) and (A.6) it follows that
(2.24) |
for all , from which we arrive at (2.16), as required. ∎
We can now further extend the regularity in the next result.
Proposition 2.4.
Proof.
For given by (2.10), since is measurable and , exists, is continuous, and is given by
(2.26) |
for all . After a change of variables , the second integral in (2.26) can be expressed as
(2.27) |
for all . From (A.4) it follows that the second integral in (2.18) can be expressed as
(2.28) |
for all . Using Propositions 2.2 and 2.3, , and the mean value theorem, it follows that for each , there exists a constant such that
(2.29) |
for all . Therefore, the absolute value of the integral in (2.28) is bounded above by
(2.30) |
for all . Therefore, via (2.26), (2.27), (2.28) and (2.30), it follows from (A.3) that
(2.31) |
for all . It follows, as in the proof of Proposition 2.3 that converges uniformly on compact subsets of , to the continuous limit , as with the bound on in (2.25) following immediately from (2.31). Consequently can be represented, as
(2.32) |
for all . Furthermore, given by (2.10), since is measurable and , exists, is continuous, and is given by
(2.33) |
for all . From (A.1), it follows that forms a -sequence as , and since and are continuous on it follows that
(2.34) |
for all . Moreover, on any compact subset of the convergence in (2.34) is uniform. Finally, it follows, as in the proof of Proposition 2.3 that converges uniformly on compact subsets of , to the continuous limit , as . As a consequence is continuous and can be represented, as
(2.35) |
for all . The bound on in (2.25) now follows from Propositions 2.1 and 2.3, and (2.35), as required. ∎
Corollary 2.1.
Proof.
Remark 2.1.
Suppose for the solution to [IVP] constructed in Proposition 2.1, that . Then following the arguments in Propositions 2.2, 2.3 and 2.4, it follows that can be naturally extended onto , with for all , and we conclude that . Moreover, , and are bounded on by a constant , which is independent of , recalling (2.4).
Before we can establish the existence of global solutions to [IVP] we require a priori bounds on solutions to [IVP].
Proposition 2.5.
When is a solution to [IVP] then
Proof.
Let and . Via (1.2) it follows that satisfies
(2.36) |
Additionally note that there exist positive constants as , such that
(2.37) |
via condition (1.4). From (2.36), (1.1) and (2.37), it follows from the comparison theorem for second order linear parabolic partial differential inequalies (see, for example [2, Theorem 4.4]) by considering:
as regular supersolutions and regular subsolutions respectively, that
(2.38) |
The result follows from (2.38) by letting . ∎
We can now establish
Proposition 2.6.
There exists a global solution to [IVP].
Proof.
We next establish local in time continuous dependence on the initial data, of global solution to [IVP].
Proposition 2.7.
Let and for , suppose that are solutions to [IVP] with constant , and initial data , respectively. Then,
(2.39) |
Proof.
Let and set to be
(2.40) |
Via (2.2)-(2.3), for given and there exist constants as , such that
(2.41) |
for all . We note, via the continuity and bounds on and given in Proposition 2.4, it follows that is a continuous and bounded function of on , and hence the integral in (2.41) is well-defined. It follows immediately that
(2.42) |
for all . Therefore, via a generalisation of Gronwall’s inequality (see [4, Corollary 2]), and the a priori bounds in Proposition 2.5, we conclude that
(2.43) |
for all . On recalling (2.40), (2.39) follows by letting in (2.43), as required. ∎
In summary, we have
Theorem 2.1.
There exists a unique global solution to [IVP]. Moreover, for each and Lebesgue measurable , there exists such that for all Lebesgue measurable such that then the corresponding global solutions to [IVP] given by satisfy
Proof.
We conclude this section by establishing some qualitative properties of solutions to [IVP] for initial data of the form (1.5). First we have
Remark 2.2.
Suppose the initial data in Proposition 2.6 satisfies for some with . Then it follows from Remark 2.1 that the global solution to [IVP] can be extended continuously onto . Moreover, the global solution has partial derivatives with respect to which are continuous on and bounded on for any . This follows from an induction argument based on the derivative estimates in Propositions 2.3 and 2.4 with the identity
for all and , used to bound the first integrals in (2.17) and (2.26). As a consequence, it follows that has partial derivatives with respect to on which are bounded on for any .
We now have
Proposition 2.8.
Suppose that the initial data for [IVP] is given by (1.5) and the corresponding solution is . When () then () for all .
Proof.
Consider the sequence of functions for such that
(2.44) | ||||
(2.45) | ||||
(2.46) |
and is given by (1.5). It follows from Proposition 2.6 that there exists a unique solution to [IVP] with initial data . Moreover, it follows from (2.44)-(2.46), Proposition 2.5 and Remark 2.2 that given by on satisfies:
(2.47) | ||||
(2.48) | ||||
(2.49) | ||||
(2.50) |
Properties (2.47)-(2.50) ensure that we can apply the minimum (maximum) principle (see [2, Theorem 3.3] to to establish that when .
Recalling (2.44)-(2.46), it follows from Proposition 2.5 that are uniformly a priori bounded on for . Moreover, via Propositions 2.3 and 2.4, and are bounded on compact subsets of uniformly for . Therefore forms a uniformly bounded equicontinuous sequence of functions on compact subsets of . Hence, there exists a subsequence which converges uniformly as to a continuous bounded function on each compact subset of . Since the global solution to [IVP] with initial data given by (1.5) is unique, it follows that on compact subsets of , converges uniformly to . Therefore, if .
We next have the far field result
Proposition 2.9.
Suppose the initial data for [IVP] is given by (1.5). Then the solution satisfies
Proof.
Let be the unique global solution to [IVP] with initial condition (1.5) and let . Since is continuous on , it follows for that can be extended onto so that . Consider the following pairs of functions with domain and co-domain .
-
•
For and each we define:
(2.51) (2.52) -
•
For and each we define:
(2.53) (2.54)
It follows that the pairs in (2.51)-(2.54) are all regular subsolution and regular supersolutions on for the second order linear parabolic partial differential operator given by
(2.55) |
It follows from (2.51)-(2.55), Proposition 2.5, and the comparison theorem [2, Theorem 4.4] that as uniformly for for each . The corresponding result for the limit as follows from a symmetrical argument. ∎
3 Conclusion
In this note we have established a well-posedness result for [IVP] to complement the large- asymptotic analysis for solutions to [IVP] contained in [3]. Further work to establish convergence and qualitative properties of the finite difference approximation utilised in [3] is of interest to the authors. Moreover, the development of methods to rigorously establish more results the theory in [3] illustrates, is also of interest to the authors.
Appendix A Properties of
We note several properties of the fundamental solution to the heat equation on , given by (2.1), here:
(A.1) | ||||
(A.2) | ||||
(A.3) | ||||
(A.4) |
Moreover, for any there exist constants such that:
(A.5) | ||||
(A.6) | ||||
(A.7) |
for all and ; and
(A.8) | ||||
(A.9) | ||||
(A.10) |
for all and . These properties can be derived via the approach described in [1, Ch. 1, Lemma 3].
References
- [1] Avner Friedman. Partial differential equations of parabolic type. Courier Dover Publications, 2008.
- [2] John Christopher Meyer and David John Needham. Extended weak maximum principles for parabolic partial differential inequalities on unbounded domains. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences, 470(2167):20140079, 2014.
- [3] David John Needham, John Billingham, John Christopher Meyer, and Catherine Drysdale. The Riemann problem for a generalised Burgers equation with spatially decaying sound speed. I Large-time asymptotics. Submitted to Studies in Applied Mathematics, 2022.
- [4] Haiping Ye, Jianming Gao, and Yongsheng Ding. A generalized gronwall inequality and its application to a fractional differential equation. Journal of Mathematical Analysis and Applications, 328(2):1075–1081, 2007.
- [5] Hui Zhang. Global existence and asymptotic behaviour of the solution of a generalized Burger’s equation with viscosity. Computers & Mathematics with Applications, 41(5-6):589–596, 2001.