A Space-time Nonlocal Traffic Flow Model: Relaxation Representation and Local Limit††thanks: This rsearch is supported in part by US NSF DMS-1937254, DMS-2012562, and CNS-2038984.
Abstract
We propose and study a nonlocal conservation law modelling traffic flow in the existence of inter-vehicle communication. It is assumed that the nonlocal information travels at a finite speed and the model involves a space-time nonlocal integral of weighted traffic density. The well-posedness of the model is established under suitable conditions on the model parameters and by a suitably-defined initial condition. In a special case where the weight kernel in the nonlocal integral is an exponential function, the nonlocal model can be reformulated as a hyperbolic system with relaxation. With the help of this relaxation representation, we show that the Lighthill-Whitham-Richards model is recovered in the equilibrium approximation limit.
This work has been published in Discrete and Continuous Dynamical Systems, 2023, 43(9): 3456-3484. Please refer to the official publication for citation.
1 Introduction
1.1 The nonlocal space-time traffic flow model
We consider the following nonlocal conservation law modeling traffic flow
(1.1) | |||
(1.2) |
Here, the quantity represents the traffic density, where indicates an empty road ahead and models bumper-to-bumper traffic jam. The nonlocal quantity is a weighted average of along the space-time path
with an averaging kernel . The vehicle velocity depends on the nonlocal traffic density through a decreasing function . The model (1.1)-(1.2) is the evolution associated to a past-time condition given on the half plane .
1.2 Background and motivation
The model (1.1)-(1.2) takes inspiration from the classical Lighthill-Whitham-Richards (LWR) model [35, 36]
(1.3) |
in which the vehicle velocity depends only on the local traffic density . The LWR model (1.3) is a scalar conservation law with the flux function . In the instance of inter-vehicle communication [18], the flux may have a nonlocal dependence on traffic density in order to capture each vehicle’s reaction to downstream traffic conditions. It is useful to incorporate time delays of this traffic density information in the distance [40, 31]. In (1.1)-(1.2), we incorporate both nonlocal fluxes and time delays via velocities that depend on a weighted space-time average of the traffic density, assuming that the traffic density information travels at a constant speed .
If the choice of rescaled weights is made, then formally the equations (1.1)-(1.2) converge to the local equation (1.3) as . The main goal of this paper is to demonstrate this in a rigorous manner via convergence of solutions.
There has recently been much research interest in nonlocal effects in phenomena described by conversation laws; there is a wide variety of applications but a dearth of analytical understanding. Some application areas from which nonlocal conservation laws arise are traffic flows [34, 33, 8, 30, 29, 10, 9, 11], sedimentation [1], pedestrian traffic [16, 5], material flow on conveyor belts [27, 38], and the numerical approximation of local conservation laws [21, 20, 19, 23].
For several traffic flow models, the nonlocal mechanism is introduced in the flux term. One such model that recovers the LWR model (1.3) when the effect is localized was proposed in [2, 26]:
(1.4) | |||
(1.5) |
Various analytical aspects of this model have been investigated, including the existence and uniqueness of solutions [2, 26, 4], existence and stability of traveling wave solutions [37, 39], development of numerical schemes [26, 6, 25], and stability analysis of the model in the case where the domain (road) is a closed ring [28]. Convergence of solutions of (1.4)-(1.5) to its local limit (which is the LWR model (1.3)) was established in [4, 3] by way of an a priori BV estimate and an entropy estimate, both of which were obtained via reformulation of the nonlocal model as a relaxation system in the case of exponential weight kernels. This is not the only mechanism that has been used to investigate the nonlocal-to-local limit; see the works of [13, 32, 12, 14, 15, 24].
1.3 Assumptions on the model
We conduct an analogous study of the nonlocal-to-local limit for the model (1.1)-(1.2) with suitable choices of the functions and the past-time condition. To fix ideas, we make the following assumptions on :
Assumption 1. The velocity function is strictly decreasing with and , where represents the maximum vehicle speed.
Assumption 2. The weight kernel is non-negative and satisfies
(1.6) |
for a constant .
1.4 Main results
Our first main result is the existence of Lipschitz solutions to the past-time value problem (1.1)-(1.2)-(1.7) with Lipschitz past-time data .
Theorem 1.1.
Suppose that Assumption 1 and Assumption 2 are satisfied and that
(1.8) |
Suppose that the past-time data is a bounded Lipschitz function belonging to the class ; see definition (2.2) below. Then the past-time value problem (1.1)-(1.2)-(1.7) admits a solution that is Lipschitz continuous and satisfies (1.1)-(1.2)-(1.7) pointwise. Furthermore, the solution satisfies the uniform bounds
(1.9) |
where and are defined in (2.4)-(2.5) below and depend only on , , and .
Formally, as the time-delay parameter approaches zero, the system (1.1)-(1.2) approaches the nonlocal-in-space system (1.4)-(1.5). This is also true in a qualitative sense; each of the key estimates for (1.1)-(1.2) remain valid as , as the bounding constants neither vanish nor blow up. Analogous statements of all of our results hold for (1.4)-(1.5), see [4], and can be formally recovered from our results by taking . However, quantitatively stronger results hold for (1.4)-(1.5). For example, the main estimates in Proposition 3.1 and Theorem 1.3 concern estimates in space and time, whereas the analogous results for (1.4)-(1.5) hold for in space and in time; see again [4].
The proof of Theorem 1.1 makes up Section 2. We use a fixed point argument combined with the method of characteristics, which is heavily inspired by the proof of [4] for the existence of solutions to the nonlocal-in-space model.
In certain modelling applications, it might only be possible to gather the traffic data at a certain initial time. In such a case, a natural choice of past-time data via the following extension of initial data:
(1.10) |
for a given function . We can then treat (1.1)-(1.2)-(1.7) as an initial-value problem, since the quantity depends only on . To be precise, with (1.10), the equation (1.2) becomes
(1.11) |
Under this consideration, the equations (1.1)-(1.2)-(1.7) where the past-time data is given by the equation (1.10) are equivalent to the Cauchy problem (1.1)-(1.11) with the initial condition .
For the particular choice (1.10) of past-time data, we establish the well-posedness of the Cauchy problem in the setting of weak solutions, which is our second main result.
Theorem 1.2.
Suppose that Assumption 1, Assumption 2 and (1.8) are satisfied, and let be a bounded function with finite total variation belonging to the class ; see (3.1) below. Then there exists a unique satisfying (1.9) that is a weak solution to (1.1)-(1.11) with initial condition ; in other words, satisfies
(1.12) |
for all with defined by (1.11).
The key tool used to prove Theorem 1.2 is the -stability of Lipschitz solutions; once that is established in Proposition 3.1, the existence and uniqueness of weak solutions follows by using Theorem 1.1 and an approximation argument.
With the well-posedness of the problem (1.1)-(1.11) in hand, we analyze the nonlocal-to-local limit. This limit is realized in the following way: consider the rescaled kernels . Taking , the kernel converges to a Dirac delta function, and so – formally – solutions of the nonlocal model (1.1)-(1.11) converge to the entropy admissible solution of the local model (1.3). We make the choice of exponential kernel function for , defined as
(1.14) |
With defined as in (1.14), the model (1.1)-(1.11) (and more generally (1.1)-(1.2)) can be rewritten as a relaxation system:
(1.15) | ||||
(1.16) |
Utilizing the special features of this relaxation system formulation (1.15)-(1.16), a uniform global BV bound on that is independent of the relaxation parameter can be proved, which serves as a key estimate for the compactness theory and guarantees the existence of a limit of the solutions.
Theorem 1.3.
Suppose that Assumption 1, Assumption 2 and (1.8) are satisfied, and let . Assume that the weight kernel is given by the exponential functions as in (1.14). In addition, suppose that the minimum density as defined in (1.9) is positive, and that the following condition holds for and :
(1.17) |
Then the unique weak solution of (1.1)-(1.11) with initial condition satisfies
(1.18) |
where represents the total variation of on , and the constant is independent of .
The choice (1.14) is the same as the one made in [4] to analyze the nonlocal-to-local limit for the nonlocal-in-space model (1.4)-(1.5). Our methods closely follow theirs, but the relaxation system (1.15)-(1.16) is a genuine system of conservation laws in the original -coordinate system, and we additionally take this into account. We remark that, in the case of , the condition (1.8) holds whenever , and (1.17) becomes . These conditions on the functions are the same as the ones proposed in [4] for the nonlocal-in-space model (1.4)-(1.5).
1.5 Organization of the paper
This paper is organized as follows. First, we establish the existence of Lipschitz solutions from Lipschitz past-time data in Section 2 (Theorem 1.1). In Section 3 we establish the stability estimate for Lipschitz solutions and prove Theorem 1.2. Section 4 is devoted to the uniform BV bound estimate of solutions based on the model’s relaxation system formulation (Theorem 1.3), which guarantees the existence of local limit solutions. Section 5 provides the proof of entropy admissibility of the local limit solution and completes the nonlocal-to-local limit theorem (Theorem 1.4).
2 Existence of Lipschitz solutions
This section is devoted to the proof of Theorem 1.1.
2.1 Initial and past-time data
To begin, we make precise the conditions on the past-time data. First, the initial values of and corresponding to a past-time condition are denoted throughout the paper as
(2.1) |
Second, for a given constant we introduce the following notation for a class of functions for past-time data with and given by (2.1) correspondingly.
(2.2) |
Now we define
(2.3) |
Under the Assumption 1, we have and . Moreover, it holds that for provided . In this case the function is monotone and we let denote the inverse function of . We define
(2.4) | |||
(2.5) |
where and are defined in (2.1). It is clear that for any .
2.2 Reformulation as a fixed-point problem
The essential idea in the proof of Theorem 1.1 is to reformulate the model as a fixed point problem and apply the contraction mapping theorem. We first define the fixed point mapping on a proper domain with a finite time horizon, and then show it is contractive through a priori and Lipschitz estimates. The fixed point solution is shown to be a Lipschitz solution to the model and it can be extended to all times .
First let us fix a time horizon where , and suppose , are as defined in (2.4)-(2.5). For any and , we define the domain
Then we introduce a directional derivative operator
where the direction is taken along the line integral paths in (1.2), and an auxiliary variable
With the above definitions, the past-time value problem (1.1)-(1.2)-(1.7) can be reformulated as a system to be solved on .
(2.6) | |||
(2.7) | |||
(2.8) |
This representation motivates the following step-by-step definition of a mapping .
-
1.
With a given and for any , we define for all as in (2.6).
-
2.
We define for all as the solution to the linear Cauchy problem (2.8) with the above and the initial condition
-
3.
With and defined above, we define as
Finally we define the mapping by
The outline of the proof of Theorem 1.1 is to establish the following facts:
-
•
For any , ;
-
•
is a contraction mapping on in the norm;
-
•
The contraction mapping theorem gives the unique fixed point , i.e. ;
- •
- •
We remark here that the map as constructed requires no relation between and at to hold. However, the condition is imposed so that quantities such as are Lipschitz with appropriate constant.
2.3 Proof of Theorem 1.1
The proof consists of six steps. In the proof, we omit the notations and in , and for simplicity, but keep in mind that they both depend on and . In addition, we use the equation (1.2) for to simplify the calculation, but keep in mind that the nonlocal integral for involves and its precise form is (2.6).
Step 1 (Characteristics). We rewrite the linear Cauchy problem (2.8) as
(2.9) |
Given for and for , (2.9) can be solved by the method of characteristics. For a point , the characteristic curve is given by where satisfies
(2.10) |
It is easy to see that by definition of and that
(2.11) |
This implies
for all characteristic curves. Therefore, for any given point one can trace the characteristic curve back to reach a unique point on the -axis, and . Integrating the characteristic ODE
(2.12) |
from the unique satisfying to , one can obtain the value of .
Step 2 ( and directional Lipschitz bounds). We first note that the identity
gives
for all . In addition, integration by parts gives
(2.13) |
hence
for all .
To give a bound on , we note that
By integrating (2.3) and using the uniform bound
we obtain that
when is sufficiently small. This together with (2.11) gives
(2.14) |
Now let us give a bound on . Taking the directional derivative of the equation (2.9), we obtain
(2.15) |
At time , by the equation (2.9) we write
using that we have
We integrate the equation (2.3) along the characteristic curves defined in (2.10). With the uniform bounds
we deduce from a comparison argument that
(2.16) |
where is the solution to the linear ODE
(2.17) |
with constant coefficients given by
By choosing sufficiently small, we obtain that for . Then the identity
implies that
(2.18) |
The equality is clear from the definition. Using the obtained and directional Lipschitz bounds (2.14)-(2.18), we conclude that there exist , depending only on , such that maps to itself.
Step 3 (Contraction). For any , and any , we denote by
the two characteristic curves satisfying
We have
Using Grönwall’s inequality backward in time, we obtain
(2.19) |
with the constant .
Note that and can be solved from
along the characteristic curves , with the same initial condition. Using again Grönwall’s inequality and noticing (2.19), we obtain
with the constant .
Thanks to the above estimates, we finally deduce that
with the constant . Choosing sufficiently small such that , is a contraction mapping in the norm.
By the contraction mapping theorem, there exists such that has a unique fixed point in . From now on we denote as the unique solution in that satisfies (1.1)-(1.2)-(1.7) on . With this definition of we define by (2.7).
Step 4 (Uniform bound). We aim to show that and satisfy the uniform bounds
(2.20) |
We provide a proof for the upper bounds; the lower bounds are obtained in a similar manner.
We denote
It is clear that and .
Let us fix and consider the characteristic curve for such that
where is the intersection of the characteristic curve and the horizontal line . For any , the equation (2.9) gives
Integrating by parts gives
(2.21) |
where
is a new weight kernel satisfying
Noting that
we compute
It yields that
This inequality combined with (2.21) gives
Furthermore, we have
and hence
where . Integrating the above inequality with the initial condition , we obtain that
Noting that and , we have
(2.22) |
where . Now we let run over ; the respective characteristic curves fill the domain and so (2.22) is uniform to the choice of , hence we have
(2.23) |
Now suppose . Then we have
and since is decreasing we have for any
Therefore , and so by definition of and by (2.23)
which contradicts . Therefore we deduce that . Applying this in (2.23) gives , and so the upper bounds in (2.20) are proved.
Step 5 (Final Lipschitz estimates). In Step 2, we obtain bounds on and . Using the equation (2.8), a bound on can also be obtained and we conclude that is Lipschitz continuous on . To show the Lipschitz continuity of , it suffices to show that of since . Given the established bound on , we only need to show the existence of and give a bound on it.
Let us denote
and
For any and , we have:
The equation gives
Therefore we have
for any and .
Using Grönwall’s inequality, we deduce that there exists a constant only depending on such that for any and , which gives the Lipschitz bound for .
Step 6 (Continuation). We iteratively construct the solution on time intervals , , , from . At time (), the past-time data is given by for and for . Thanks to the and Lipschitz bounds obtained in Step 2, Step 4, and Step 5, the constructed solution on the time interval satisfies
and
where is the solution of the ODE (2.17). The above estimates guarantee that as and the solution can be extended to the whole domain .
2.4 Discussion
We now make some remarks on the model assumptions and the proof of Theorem 1.1.
Remark 2.1.
The Assumption 2 requires that the weight kernel has an exponential decay. Such an assumption was also used in [13] to establish the nonlocal-to-local limit of the nonlocal-in-space model (1.4)-(1.5). In Theorem 1.1, it is assumed that , which together with (1.6) gives
(2.24) |
It is worth noting that if satisfies the condition (2.24), the rescaled kernel also satisfies the condition with the same parameters and . For the exponential kernel defined in (1.14), the Assumption 2 is satisfied for all whenever , which is consistent with the sub-characteristic condition under the relaxation system formulation (1.15)-(1.16).
Remark 2.2.
Let us define the function space
(2.25) |
where the velocity is written as . Then for defined via the extension (1.10) for a given function ,
(2.26) |
By the form of we can see that even if for all without an additional condition that the constraint can be violated at some point where and . A sufficient condition on alone to ensure is
In this case, the lower and upper bounds for the solutions given in Theorem 1.1 become
These sufficient conditions and solution bounds may not be the best possible results, we will leave possible improvements for the future research.
3 Existence, uniqueness and stability of weak solutions
For the remainder of the paper we concern ourselves with a class of past-time data extended vertically from given initial data. That is, we assume that the past-time data satisfies (1.10) for a given , where denotes the class
(3.1) |
With this assumption we establish the -stability of Lipschitz solutions, from which Theorem 1.2 follows.
Proposition 3.1 (-stability of Lipschitz solutions).
Under Assumption 1, Assumption 2, and (1.8), assume that two functions for (that is, their Lipschitz constants are possibly different). Let for be Lipschitz solutions to (1.1)-(1.11) with initial conditions respectively.
Then for any there exists a positive constant such that
(3.2) |
Proof.
Let be a parameter; for each value of , define to be a Lipschitz solution to (1.1)-(1.11) satisfying with initial data . At least one such solution exists by Theorem 1.1 and Remark 2.2. Define the first order perturbations for :
Recalling the definition of the quantity in (2.7), define its first order perturbation as
Then
(3.3) |
and satisfies the linearized equation
(3.4) |
where .
From (1.11) the integral defining can be written as
(3.5) |
We also use a consequence of the condition (1.6) on :
(3.6) |
Third, we note a variant of Grönwall’s inequality
(3.7) |
Step 1. We show that along any finite characteristic segment, the perturbed quantity has bounded total variation. To be precise, define
We claim that there exists depending only on and such that
(3.8) |
and
(3.9) |
We will prove only (3.8); (3.9) is obtained using the same procedure. From
we have
Since the second term in the integral can be dropped in the estimate, and so
(3.10) |
We need to estimate the last integral on the right-hand side. We use (1.11) to obtain the identities
from which it follows that
(3.11) |
In the same way, we can obtain the bound
(3.12) |
The estimates (3.11) and (3.12) are applied to majorize the integral on the last line of (3.10) by
(3.13) |
Now, since we have that , and so along with (3.11)
Since by assumption we can absorb the last term into the left-hand side of the estimate to get
(3.14) |
Inserting (3.14) into (3.13), the estimate for the total variation of from (3.10) is now
Then by (3.7)
where depends only on and . The bound (3.8) follows.
Step 2. We prove the main result. The method is similar to Step 1. Define by
We use the linearized equation (3.4) and apply integration by parts to obtain
Since , the second term in the integral can be dropped, and so
(3.15) |
We need to estimate the last integral on the right-hand side. We use (3.5) and (3.6) to obtain the estimates
(3.16) |
Then (3.16), (3.12), (3.8) and (3.9) are applied to majorize the last integral in (3.15) by
(3.17) |
Now, as a consequence of (3.3) we have
(3.18) |
so with (3.16) and the conditions (1.8) on and
Therefore we can absorb the last term into the left-hand side of the estimate to get
(3.19) |
Inserting (3.19) into (3.17), the estimate for the derivative of from (3.15) is now
the bound is easily seen from (3.3) and (3.5). Applying Grönwall’s inequality and changing coordinates, we obtain
(3.20) |
Proof of Theorem 1.2.
Let , and let , , be a sequence of mollified functions in (possibly with ) that converge to in . By virtue of (3.2) the corresponding solutions to (1.1)-(1.11) with initial condition are Cauchy, and hence converge, in to a function . Thus satisfies (1.12), and so is a weak solution. Furthermore, we note that the weak solutions constructed in this way inherit the same stability property (3.2), since the bounding constant in that inequality does not depend on the Lipschitz constant of the solutions, and so uniqueness follows. To complete the proof, given that is a bounded sequence in , and the weak- limits are unique, by noting the sequence is obtained with initial conditions that are mollified approximations of , we can pass through the limits to obtain the bounds (2.4)-(2.5) for the weak solution . ∎
4 Uniform BV bound and existence of limit solutions
Towards the aim of proving the convergence of the solutions of (1.1)-(1.11) as the weight kernel converges to a Dirac delta function, we consider only the exponential kernels as defined in (1.14):
In this case the nonlocal model (1.1)-(1.11) can be reformulated as the relaxation system (1.15)-(1.16), which is recalled here:
The characteristic speeds of the system are
Taking , we expect the solution of (1.15)-(1.16) to converge to that of its equilibrium approximation, which is the LWR model (1.3). The characteristic speed of the limit equation (1.3) is
The condition (1.8) plus ensures the strict sub-characteristic condition .
4.1 Uniform BV bound
Proof of Theorem 1.3.
Let us first assume . In this case, and are Lipschitz continuous and satisfy the reformulated system (1.15)-(1.16) pointwise.
Noting that and stay positive provided , we construct
(4.1) |
One can easily verify that and are Riemann invariants of the system (1.15)-(1.16) corresponding to the system’s characteristic speeds and , respectively. With the new set of variables , the system (1.15)-(1.16) can be diagonalized as
(4.2) | ||||
(4.3) |
where is an increasing function, and
(4.4) |
Note that and are Lipschitz continuous. By the method of characteristics we see that are Lipschitz continuous and compactly supported. We claim that the system (4.2)-(4.3) is total variation diminishing, i.e.,
(4.5) |
Indeed, differentiating (4.2)-(4.3) with respect to gives
from which we obtain that
where
and
A direct calculation gives
and
where the condition (1.17) and the solution bounds are used. With and , the estimate (4.5) follows immediately.
Thanks to the estimate (4.5), we now turn to the uniform BV bound on . At the initial time , we have
Therefore,
Since the total variation of is diminishing, it holds that
for any time . Noting that , we deduce that
Then, using (1.11) and (1.1), we have
for any time . Combining the above inequalities, we obtain
which gives the desired uniform BV bound (1.18).
4.2 Convergence to a weak solution
Now we are in a position to show the existence of limit solutions that satisfy the limit equation (1.3) in the weak sense. To pass the limit we need to establish the following theorem.
Theorem 4.2.
Proof.
By Theorem 1.2 and Theorem 1.3, the family of solutions is uniformly bounded in . As a consequence, the family is precompact in the norm (see [22]). Then we can select a sequence such that in , where the limit function .
5 Entropy admissibility of the limit solution
In this section, we show that the weak solution to the local model (1.3) obtained from the limit as of a sequence of weak solutions to (1.1)-(1.11) is in fact the entropy admissible solution. This completes the theory of nonlocal-to-local limit from (1.1)-(1.11) to (1.3) in the case of exponential kernels.
Proof of Theorem 1.4.
Following a similar approach as in [3], it suffices to establish the entropy inequality for one convex entropy, see also [17]. For this purpose, we introduce the following entropy-entropy flux pair:
(5.1) |
It is straightforward to verify that , and that is strictly convex. We claim the following entropy inequality for the nonlocal solution of (1.1)-(1.11):
(5.2) |
for all nonnegative test functions , where the constant is independent of . Assuming this claim, any limit solution obtained following Theorem 4.2 satisfies the entropy inequality
(5.3) |
for all nonnegative test functions , and thus is the unique entropy admissible solution of (1.3).
Now we prove the inequality (5). Let us first assume that is Lipschitz continuous and show (5) for Lipschitz solutions. For simplicity we omit the superscript in . The equation (1.1) can be rewritten as
(5.4) |
For any nonnegative test function , multiplying on both sides of (5.4) gives
(5.5) |
Using again the directional derivative notation , we obtain the identity . Then (5.5) becomes
(5.6) |
Integrating (5) and using integration by parts, we get
where
and
Repeatedly using the identity and integrating by parts, we compute
with .
Let us make some remarks on entropy pairs for the relaxation system (1.15)-(1.16) and its equilibrium approximation (1.3). In the proof of Theorem 4.2 we base the analysis directly on the nonlocal model (1.1)-(1.11), and do not rely on the rigorous justification of the entropy inequality for the relaxation system (1.15)-(1.16). However, we remark that some intuitive analysis based on the relaxation system (1.15)-(1.16) offers insight to our choice of the entropy pair (5.1).
Following the paradigm described in [7], if is any entropy-entropy flux pair for the limiting conservation law (1.3), one can construct an entropy-entropy flux pair for the relaxation system (1.15)-(1.16) such that
for any test function , and when one has
Therefore, it holds
Assuming and are smooth, the right hand side is when .
Provided any convex , one can construct by solving the following hyperbolic Cauchy problem [7]:
We note that, with the simple choice of convex entropy , the analytic solution may be complicated. Instead, if we choose a different convex entropy function:
we obtain a simple solution for as
This motivates our choice of the entropy-entropy flux pair in (5.1).
6 Concluding remarks
In this paper we propose a space-time nonlocal conservation law modelling traffic flow. The proposed model (1.1)-(1.2) extends the classical LWR model by introducing nonlocal velocities in the flux function. To fit realistic traffic scenarios, the model considers time delays in the long-range inter-vehicle communication, and the model parameter corresponds to the temporal nonlocal effects. In the limit as , our analysis shows that the model recovers a model involving only spatial nonlocality, which has been extensively studied in the literature.
We provide well-posedness theories of the proposed model (1.1)-(1.2) under suitable assumptions on model parameters and the past-time condition. Furthermore, in the special case of exponential weight kernels, we prove convergence from solutions of the nonlocal model to the unique entropy admissible solution of the local limit equation, i.e. the LWR model. The results established in this paper provide a rigorous underpinning in potential implementation of the space-time nonlocal model for the modelling of nonlocal traffic flows.
Let us make some concluding remarks on possible generalizations of the model. An alternative model to (1.1)-(1.2) is to instead take a weighted average of vehicle velocity. To be precise,
where |
For this model, we expect that the well-posedness and nonlocal-to-local limit can be established in a similar fashion. Furthermore, in future works we hope to consider more general cases where the traveling speed of nonlocal traffic information depends on additional quantities in the model.
We would also like to conduct more mathematical analysis. In this paper we show convergence of solutions of the space-time nonlocal model to the entropy admissible solution of the local model in the case of exponential weight kernels. The convergence result may be established on the nonlocal quantity for more general initial data and kernels. Such a result has been established for the nonlocal-in-space model (1.4)-(1.5) in [14]. We hope to show more nonlocal-to-local convergence results for the space-time nonlocal model along that direction. Furthermore, understanding the behavior – such as the existence, uniqueness and stability – of traveling wave solutions of the space-time nonlocal model will shed light on the long time behavior and stability of shock waves. In the case of exponential kernels, this is equivalent to the study of traveling waves for the relaxation system, which could be easier to analyze. For general kernels, an integro-differential equation is satisfied by the traveling wave profiles. In all cases, we expect that traveling waves are local attractors for solutions.
References
- [1] F. Betancourt, R. Bürger, K. H. Karlsen, and E. M. Tory, On nonlocal conservation laws modelling sedimentation, Nonlinearity, 24 (2011), p. 855.
- [2] S. Blandin and P. Goatin, Well-posedness of a conservation law with non-local flux arising in traffic flow modeling, Numerische Mathematik, 132 (2016), pp. 217–241.
- [3] A. Bressan and W. Shen, Entropy admissibility of the limit solution for a nonlocal model of traffic flow, arXiv preprint arXiv:2011.05430, (2020).
- [4] , On traffic flow with nonlocal flux: a relaxation representation, Archive for Rational Mechanics and Analysis, 237 (2020), pp. 1213–1236.
- [5] R. Bürger, P. Goatin, D. Inzunza, and L. M. Villada, A non-local pedestrian flow model accounting for anisotropic interactions and domain boundaries, Mathematical biosciences and engineering, 17 (2020), pp. 5883–5906.
- [6] C. Chalons, P. Goatin, and L. M. Villada, High-order numerical schemes for one-dimensional nonlocal conservation laws, SIAM Journal on Scientific Computing, 40 (2018), pp. A288–A305.
- [7] G.-Q. Chen, C. D. Levermore, and T.-P. Liu, Hyperbolic conservation laws with stiff relaxation terms and entropy, Communications on Pure and Applied Mathematics, 47 (1994), pp. 787–830.
- [8] F. A. Chiarello, J. Friedrich, P. Goatin, and S. Goẗtlich, Micro-macro limit of a nonlocal generalized aw-rascle type model, SIAM Journal on Applied Mathematics, 80 (2020), pp. 1841–1861.
- [9] F. A. Chiarello, J. Friedrich, P. Goatin, S. Göttlich, and O. Kolb, A non-local traffic flow model for 1-to-1 junctions, European Journal of Applied Mathematics, (2019), pp. 1–21.
- [10] F. A. Chiarello and P. Goatin, Global entropy weak solutions for general non-local traffic flow models with anisotropic kernel, ESAIM: Mathematical Modelling and Numerical Analysis, 52 (2018), pp. 163–180.
- [11] , Non-local multi-class traffic flow models, Networks & Heterogeneous Media, 14 (2019), p. 371.
- [12] G. M. Coclite, J.-M. Coron, N. De Nitti, A. Keimer, and L. Pflug, A general result on the approximation of local conservation laws by nonlocal conservation laws: The singular limit problem for exponential kernels, arXiv preprint arXiv:2012.13203, (2020).
- [13] M. Colombo, G. Crippa, E. Marconi, and L. V. Spinolo, Local limit of nonlocal traffic models: convergence results and total variation blow-up, Annales de l’Institut Henri Poincaré C, Analyse non linéaire, 38 (2021), pp. 1653–1666.
- [14] , Nonlocal traffic models with general kernels: singular limit, entropy admissibility, and convergence rate, arXiv preprint arXiv:2206.03949, (2022).
- [15] M. Colombo, G. Crippa, and L. V. Spinolo, On the singular local limit for conservation laws with nonlocal fluxes, Archive for Rational Mechanics and Analysis, 233 (2019), pp. 1131–1167.
- [16] R. M. Colombo, M. Garavello, and M. Lécureux-Mercier, A class of nonlocal models for pedestrian traffic, Mathematical Models and Methods in Applied Sciences, 22 (2012), p. 1150023.
- [17] C. M. Dafermos and C. M. Dafermos, Hyperbolic conservation laws in continuum physics, vol. 3, Springer, 2005.
- [18] K. C. Dey, A. Rayamajhi, M. Chowdhury, P. Bhavsar, and J. Martin, Vehicle-to-vehicle (V2V) and vehicle-to-infrastructure (V2I) communication in a heterogeneous wireless network–Performance evaluation, Transportation Research Part C: Emerging Technologies, 68 (2016), pp. 168–184.
- [19] Q. Du and Z. Huang, Numerical solution of a scalar one-dimensional monotonicity-preserving nonlocal nonlinear conservation law, J. Math. Res. Appl, 37 (2017), pp. 1–18.
- [20] Q. Du, Z. Huang, and P. G. LeFloch, Nonlocal conservation laws. a new class of monotonicity-preserving models, SIAM Journal on Numerical Analysis, 55 (2017), pp. 2465–2489.
- [21] Q. Du, J. R. Kamm, R. B. Lehoucq, and M. L. Parks, A new approach for a nonlocal, nonlinear conservation law, SIAM Journal on Applied Mathematics, 72 (2012), pp. 464–487.
- [22] L. C. Evans and R. F. Garzepy, Measure theory and fine properties of functions, Routledge, 2018.
- [23] U. S. Fjordholm and A. M. Ruf, Second-order accurate tvd numerical methods for nonlocal nonlinear conservation laws, SIAM Journal on Numerical Analysis, 59 (2021), pp. 1167–1194.
- [24] J. Friedrich, S. Göttlich, A. Keimer, and L. Pflug, Conservation laws with nonlocal velocity–the singular limit problem, arXiv preprint arXiv:2210.12141, (2022).
- [25] J. Friedrich and O. Kolb, Maximum principle satisfying cweno schemes for nonlocal conservation laws, SIAM Journal on Scientific Computing, 41 (2019), pp. A973–A988.
- [26] P. Goatin and S. Scialanga, Well-posedness and finite volume approximations of the LWR traffic flow model with non-local velocity, Networks and Hetereogeneous Media, 11 (2016), pp. 107–121.
- [27] S. Göttlich, S. Hoher, P. Schindler, V. Schleper, and A. Verl, Modeling, simulation and validation of material flow on conveyor belts, Applied mathematical modelling, 38 (2014), pp. 3295–3313.
- [28] K. Huang and Q. Du, Stability of a nonlocal traffic flow model for connected vehicles, SIAM Journal on Applied Mathematics, 82 (2022), pp. 221–243.
- [29] W.-L. Jin, Generalized bathtub model of network trip flows, Transportation Research Part B: Methodological, 136 (2020), pp. 138–157.
- [30] I. Karafyllis, D. Theodosis, and M. Papageorgiou, Analysis and control of a non-local PDE traffic flow model, International Journal of Control, (2020), pp. 1–19.
- [31] A. Keimer and L. Pflug, Nonlocal conservation laws with time delay, Nonlinear Differential Equations and Applications NoDEA, 26 (2019), p. 54.
- [32] , On approximation of local conservation laws by nonlocal conservation laws, Journal of Mathematical Analysis and Applications, 475 (2019), pp. 1927–1955.
- [33] Y. Lee, Thresholds for shock formation in traffic flow models with nonlocal-concave-convex flux, Journal of Differential Equations, 266 (2019), pp. 580–599.
- [34] , Traffic flow models with looking ahead-behind dynamics, arXiv preprint arXiv:1903.08328, (2019).
- [35] M. J. Lighthill and G. B. Whitham, On kinematic waves II. A theory of traffic flow on long crowded roads, Proc. R. Soc. Lond. A, 229 (1955), pp. 317–345.
- [36] P. I. Richards, Shock waves on the highway, Operations research, 4 (1956), pp. 42–51.
- [37] J. Ridder and W. Shen, Traveling waves for nonlocal models of traffic flow, Discrete & Continuous Dynamical Systems-A, 39 (2019), p. 4001.
- [38] E. Rossi, J. Weißen, P. Goatin, and S. Göttlich, Well-posedness of a non-local model for material flow on conveyor belts, ESAIM: Mathematical Modelling and Numerical Analysis, 54 (2020), pp. 679–704.
- [39] W. Shen, Traveling waves for conservation laws with nonlocal flux for traffic flow on rough roads, Networks and Heterogeneous Media, 14 (2019), pp. 709–732.
- [40] H. Yang and W.-L. Jin, A control theoretic formulation of green driving strategies based on inter-vehicle communications, Transportation Research Part C: Emerging Technologies, 41 (2014), pp. 48–60.