
Citation: Alberto Bressan, Yucong Huang. Globally optimal departure rates for several groups of drivers[J]. Mathematics in Engineering, 2019, 1(3): 583-613. doi: 10.3934/mine.2019.3.583
[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] | Glen Wheeler . Convergence for global curve diffusion flows. Mathematics in Engineering, 2022, 4(1): 1-13. doi: 10.3934/mine.2022001 |
[3] | Luca Formaggia, Alessio Fumagalli, Anna Scotti . A multi-layer reactive transport model for fractured porous media. Mathematics in Engineering, 2022, 4(1): 1-32. doi: 10.3934/mine.2022008 |
[4] | Carlo Ciliberto, Massimiliano Pontil, Dimitrios Stamos . Reexamining low rank matrix factorization for trace norm regularization. Mathematics in Engineering, 2023, 5(3): 1-22. doi: 10.3934/mine.2023053 |
[5] | Michiaki Onodera . Linear stability analysis of overdetermined problems with non-constant data. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023048 |
[6] | Feida Jiang . Weak solutions of generated Jacobian equations. Mathematics in Engineering, 2023, 5(3): 1-20. doi: 10.3934/mine.2023064 |
[7] | Serena Della Corte, Antonia Diana, Carlo Mantegazza . Global existence and stability for the modified Mullins–Sekerka and surface diffusion flow. Mathematics in Engineering, 2022, 4(6): 1-104. doi: 10.3934/mine.2022054 |
[8] | 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 |
[9] | Matteo Lapucci, Davide Pucci . Mixed-integer quadratic programming reformulations of multi-task learning models. Mathematics in Engineering, 2023, 5(1): 1-16. doi: 10.3934/mine.2023020 |
[10] | Raúl Ferreira, Arturo de Pablo . A nonlinear diffusion equation with reaction localized in the half-line. Mathematics in Engineering, 2022, 4(3): 1-24. doi: 10.3934/mine.2022024 |
Macroscopic models of traffic flow, first introduced in [26,28], have now become a topic of extensive research. On a single road, the evolution of the traffic density can be described by a scalar conservation law. In order to extend the model to a whole network of roads, additional boundary conditions must be inserted, describing traffic flow at each intersection; see [10,16,20,21,23,24] or the survey [4]. A major eventual goal of these models is to understand traffic patterns, determined by the behavior of a large number of drivers with different origins and destinations.
In a basic setting, one can consider N groups of drivers, say G1,…,GN. Drivers from each group have the same origin and destination, and a cost which depends on their departure and arrival time. For such a model, two kind of solutions are of interest:
− The Nash equilibrium solution, where each driver chooses his own departure time and route to destination, in order to minimize his own cost.
− The global optimization problem, where a central planner seeks to schedule all departures in order to minimize the sum of all costs.
In general, these criteria determine very different traffic patterns. To fix the ideas, let t↦ui(t), i=1,…,N, be the departure rate of drivers of the i-th group, so that
∫t−∞ui(s)ds |
yields the total number of these drivers who depart before time t. We recall that the support of ui, denoted by Supp(ui), is the closure of set of times t where ui(t)>0.
Roughly speaking, the two above solutions can be characterized as follows.
(Ⅰ) In a Nash equilibrium, all drivers within the same group pay the same cost. Namely, there exists constants K1,…,KN such that
● every driver of the i-th group, departing at a time t∈Supp(ui) bears the cost Ki.
● if a driver of the i-th group were to depart at any time t∈IR (possibly outside the support of ui), he would incur in a cost ≥Ki.
(Ⅱ) For a global optima, there exist constants C1,…,Cn (where Ci is the marginal cost for adding one more driver of the i-th group) such that
● If one additional driver of the i-th group is added at any time t∈Supp(ui), then the total cost increases by Ci.
● If one additional driver of the i-th group is added at any time t∈IR (possibly outside the support of ui), then the increase in the total cost is greater or equal to Ci.
At an intuitive level, these conditions are easy to explain. In Figure 1(left), the function Γi(t) denotes the cost to an i-driver departing at time t. If Γi did not attain its global minimum simultaneously at all points t∈[a,b]=Supp(ui), then we could find times t1∈[a,b] and t2∈IR such that Γi(t2)<Γi(t1). In this case, the driver departing at time t1 could lower his own cost choosing to depart at time t2 instead. This contradicts the definition of equilibrium.
In Figure 1(right), the function Λi(t) denotes the marginal cost for inserting one additional driver of the i-th family, departing at time t. This accounts for the additional cost to the new driver, and also for the increase in the cost to all other drivers who are slowed down by the presence of one more car on the road. If Λi did not attain its global minimum at all points in [c,d]=Supp(ui), then we could find times t1∈[c,d] and t2∈IR such that Λi(t2)<Λi(t1). In this case we could consider a new traffic pattern, with one less driver departing at time t1 and one more departing at time t2. This would achieve a smaller total cost, contradicting the assumption of optimality.
While the criterion (Ⅰ) for an equilibrium solution is easy to justify, a rigorous proof of the necessary condition (Ⅱ) for a global optimum faces considerable difficulties. Indeed, to compute the ``marginal cost" for adding one more driver, one should differentiate the solution of a conservation law w.r.t. the initial data (or the boundary data). As it is well known, in general one does not have enough regularity to carry out such a differentiation. To cope with this difficulty one can introduce a ``shift differential", describing how the shock locations change, depending on parameters. See [8,9,13,27,30] for results in this direction.
The first part of this paper contains an introduction to macroscopic models of traffic flow on a network of roads. Section 2 starts by reviewing the classical LWR model for traffic flow on a single road, in terms of a scalar conservation law for the traffic density. We then discuss various boundary conditions, modeling traffic flow at an intersection. Finally, given a cost function depending on the departure and arrival times of each driver, we review the concepts of globally optimal solution and of Nash equilibrium solution.
The second part of paper contains original results. We consider here N groups of drivers traveling along the same road, but with different departure and arrival costs. We seek departure rates u1(⋅),…,uN(⋅) which are globally optimal. Namely, they minimize the sum of all costs to all drivers. A set of necessary conditions for optimality is derived, thus extending the result in [5] to the case where several groups of drivers are present. Relying on these conditions, in the last section we introduce an algorithm that numerically computes such globally optimal solutions.
For an introduction to the general theory of conservation laws we refer to [3,19,29]. A more comprehensive discussion of various models of traffic flow can be found in [1,2,18,20].
According to the classical LWR model [26,28], traffic density on a single road can be described in terms of a scalar conservation law
ρt(t,x)+f(ρ(t,x))x=0. | (2.1) |
Here t is the time, while x∈IR is the space variable along the road. Moreover
● ρ is the traffic density, i.e., the number of cars per unit length of the road.
● v=v(ρ) is the velocity of cars, which we assume depends only on the traffic density.
● f=f(ρ) is the flux, i.e., the number of cars crossing a point x along the road, per unit time. We have the identity
[flux]=[density]×[velocity]=ρ⋅v(ρ) |
As shown in Figure 2, the velocity should be a decreasing function of the car density. Concerning the flux function, a natural set of assumptions is
f∈C2,f″<0,f(0)=f(ρjam)=0. | (2.2) |
Here ρjam is the maximum density of cars allowed on the k-th road. This corresponds to bumper-to-bumper packing, where no car can move.
Smooth solutions of the conservation law (2.1) can be computed by the classical method of characteristics. By the chain rule, one obtains
ρt+f′(ρ)ρx=0. | (2.3) |
Hence, if t↦x(t) is a curve such that
˙x(t)≐ddtx(t)=f′(ρ(t,x(t))), | (2.4) |
then the equation (2.3) yields
ddtρ(t,x(t))=ρt+ρx˙x=0. |
In other words, the density is constant along each characteristic curve satisfying (2.4). Notice that the assumptions (2.2) imply the inequality
[car speed]˙=v(ρ)=f(ρ)/ρ≥f′(ρ)=v(ρ)+ρv′(ρ)=[characteristic speed]. |
With reference to Figure 2, let ρmax be the density at which the flux is maximum. We say that a state ρ is
● free, if ρ<ρmax, hence the characteristic speed f′(ρ) is positive,
● congested, if ρ>ρmax, hence the characteristic speed f′(ρ) is negative.
Due to the non-linearity of the flux function f, it is well known that solutions can develop shocks in finite time. The conservation law (2.1) must thus be interpreted in distributional sense. For the general theory of entropy weak solutions to conservation laws, we refer to [3,29].
To model vehicular traffic on an entire network of roads, the conservation laws describing traffic flow on each road must be supplemented with boundary conditions, describing the behavior at road intersections.
Consider an intersection, say with m incoming roads i∈{1,…,m}=I and n outgoing roads j∈{m+1,…,m+n}=O, see Figure 3. We shall use the space variable x∈]−∞,0] for incoming roads and x∈[0,+∞[ for outgoing roads. Throughout the following we assume that the density of traffic on each road is governed by a conservation law
ρt+fk(ρ)x=0,fk(ρ)=ρvk(ρ), | (2.5) |
where the flux function fk satisfies (2.2), for every k=1,…,m+n.
An appropriate model must depend on various parameters, namely
● ci = relative priority of drivers arriving from road i.
● θij = fraction of drivers from road i that turn into road j.
For example, if the intersection is regulated by a crosslight, ci could measure the fraction of time when drivers from road i get green light, on average. It is natural to assume
ci,θij≥0,∑i∈Ici=1,∑j∈Oθij=1. | (2.6) |
Boundary conditions should determine the limit values of the traffic density on each of the m+n roads meeting at the intersection:
ρi(t,0−)=limx→0−ρi(t,x),ρj(t,0+)=limx→0+ρj(t,x),for alli∈I,j∈O. | (2.7) |
At first sight, one might guess that m+n conditions will be required. However, this is not so, because on some roads the characteristics move toward the intersection. For these roads, the limits in (2.7) are already determined by integrating along characteristics. Boundary conditions are required only for those roads where the characteristics move away from the origin. Recalling the definition of free and congested states, we thus have
[# of boundary conditions needed to determine the flux at the intersection]= [# of incoming roads which are congested ] + [# of outgoing roads which are free]. |
It now becomes apparent that, to assign a meaningful set of boundary conditions, several different cases must be considered.
To circumvent these difficulties, an alternative approach developed by Coclite, Garavello, and Piccoli [16,21,22] relies on the construction of a Riemann Solver. Instead of assigning a variable number of boundary conditions, here the idea is to introduce a rule for solving all Riemann problems (i.e., the initial-value problems where at time t=0 the densities ρk and turning preferences θij are constant along each road). Relying on front-tracking approximations, under suitable conditions one can prove that the solutions with general initial data are also uniquely determined.
We briefly review the main steps of this construction, for the constant initial data
{ρ1,…,ρm,ρm+1,…ρm+n=initial densities on the incoming and outgoing roads, θij=fraction of drivers from road i that turn into road j. |
Step 1. Determine the maximum flux fmaxi that can exit from each incoming road i∈I.
As shown in Figure 4, this is computed by
fmaxi=ˆfi(ρi)={fi(ρi)ifρi≤ρmaxi,fi(ρmaxi)ifρi>ρmaxi. |
Step 2: Determine the maximum flux fmaxj that can enter each outgoing road j∈O.
As shown in Figure 5, this is computed by
fmaxj=ˆfj(ρj)={fj(ρj)ifρi≥ρmaxi,fj(ρmaxj)ifρj<ρmaxj. |
Step 3: Given the maximum incoming and outgoing fluxes fmaxi, fmaxj, and the turning preferences θij, determine the region of admissible incoming fluxes (see Figure 6)
Ω≐{(f1,…,fm);fi∈[0,fmaxi],∑i∈Ifiθij≤fmaxjfor allj∈O}. | (2.8) |
Step 4. To construct a Riemann solver, it now suffices to give a rule for selecting a point ˉω=(f1,…,fm) in the feasible region Ω. In general, this rule will depend on the priority coefficients c1,…,cm assigned to incoming roads. Observe that, as soon as the incoming fluxes fi, i∈I, are given, the outgoing fluxes are uniquely determined by the identities
fj=∑i∈Ifiθij,j∈O. | (2.9) |
Various ways to define a Riemann Solver are illustrated by the following examples.
Example 1: Given priority coefficients c1,…,cm, following [16] one can choose the vector of incoming fluxes
ˉω=(f1,…,fm)≐argmaxω∈Ω∑i∈Icifi. | (2.10) |
In particular, if c1=⋯=cm=1m, this means we are maximizing the total flux through the intersection (see Figure 7(left)).
Since in (2.10) we are maximizing a linear function over a polytope, in some cases the maximum can be attained at multiple points. This somewhat restricts the applicability of this model. An alternative model, with better continuity properties, is considered below.
Example 2: Given positive coefficients c1,…,cm as in (2.6), consider the one-parameter curve
s↦γ(s)=(γ1(s),…,γm(s)), |
where
γi(s)≐min{cis,fmaxi}. |
As shown in Figure 7(right), we then choose the vector of incoming fluxes
¯ω=(f1,…,fm),fi=γi(ˉs), | (2.11) |
where
ˉs=max{s≥0;∑i∈Iγi(s)θij≤fmaxjfor allj∈O}. | (2.12) |
Example 3 : To model an intersection with two incoming and two outgoing roads, where road 2 has a stop sign, we choose the point ¯ω=(f1,f2) according to the following rules (see Figure 8).
f1=max{ω1;(ω1,0)∈Ω}. | (2.13) |
f2={0iff1<fmax1,max{ω2;(f1,ω2)∈Ω}iff1=fmax1 | (2.14) |
According to (2.13), as many cars as possible are allowed to arrive from road 1. According to (2.14), if any available space is left, cars arriving from road 2 are allowed through the intersection.
Having defined a way to solve each Riemann problem, a major issue is whether the Cauchy problem with general initial data is well posed. Assuming that the turning preferences θij remain constant in time, some results in this direction can be found in [16].
We remark, however, that in general these turning preferences may well vary in time. One should thus regard θij=θij(t,x) as variables. Assuming that drivers know in advance their itinerary, the conservation of the number of drivers on road i that will eventually turn into road j is expressed by the additional conservation law
[ρiθij]t+[ρivi(ρ)θij]x=0. | (2.15) |
Combining (2.15) with the conservation law
(ρi)t+[ρivi(ρ)]x=0, |
one obtains a linear transport equation for each of the quantities θij, namely
(θij)t+vi(ρ)(θij)x=0,i∈I,j∈O. | (2.16) |
A surprising counterexample constructed in [14] shows that, for a very general class of Riemann Solvers, one can construct measurable initial data ρi(0,⋅), ρj(0,⋅), and θij(0,⋅), so that the Cauchy problem has two distinct entropy-admissible solutions.
The ill-posedness of these model equations represents a serious obstruction, toward the existence of globally optimal traffic patterns, or Nash equilibria, on a general network of roads. To cope with this difficulty, in [10] an alternative model was proposed, for traffic flow at an intersection. Namely, it is assumed that the junction contains a buffer (say, a traffic circle), as shown in Figure 9. Incoming cars are admitted at a rate depending of the amount of free space left in the buffer, regardless of their destination. Once they have entered the intersection, cars flow out at the maximum rate allowed by the outgoing road of their choice.
More precisely, consider a constant M>0, describing the maximum number of cars that can occupy the intersection at any given time, and constants ci>0, i∈I, accounting for priorities given to different incoming roads. For j∈O, at any time t we denote by qj(t)∈[0,M] the number of cars, already within the buffer, that seek to turn into road j.
As before, let fmaxi and fmaxj the maximum fluxes that can exit from road i∈I, or can enter into road j∈O. We then require that the incoming fluxes fi satisfy
fi=min{fmaxi,ci(M−∑j∈Oqj)},i∈I. | (2.17) |
In addition, the outgoing fluxes fj should satisfy
{if qj>0, then fj=fmaxj, if qj=0, then fj=min{fmaxj,∑i∈Ifiθij}, j∈O. | (2.18) |
Having determined the incoming and outgoing fluxes fi, fj, the time derivatives of the queues qj are then computed by
˙qj=∑i∈Ifiθij−fj,j∈O. | (2.19) |
The well-posedness of the intersection model with buffers, for general L∞ data, was proved in [10].
It is interesting to understand the relation between the intersection model with buffer, and the models based on a Riemann Solver. The analysis in [12] shows that, letting the size of the buffer M→0, the solution of the problem with buffers converges to the solution determined by the Riemann Solver at (2.11) and (2.12), described in Example 2.
Consider a network of roads, with several intersections. We call γk, k=1,…,ˉk the arcs corresponding to the various roads, and A1,…,Aν the nodes corresponding to intersections. It is assumed that, on the k-th road, the flux function has the form fk(ρ)=ρvk(ρ), with vk a decreasing function of the density. As in the previous sections, traffic flow at each intersection can be modeled in terms of a Riemann Solver, or by means of a buffer.
We consider N groups of drivers with different origins and destinations, and possibly different departure and arrival costs. As shown in Figure 10:
● Drivers in the i-th group depart from the node Ad(i) and arrive at the node Aa(i).
● Their cost for departing at time t is φi(t), while their arrival cost is ψi(t).
● They can use different paths Γ1,Γ2,… to reach destination.
In the following, ˉui,p(⋅) will denote the departure rate of drivers of the i-th group, who choose the path Γp to reach destination. Calling Gi the total number of drivers in the i-th group, we say that the departure rates ˉui,p are admissible if, for every i=1,…,N they satisfy the obvious constraints
ˉui,p(t)≥0,∑p∫+∞−∞ˉui,p(t)dt=Gi. | (2.20) |
Given the departure rates, in principle one can then solve the equation of traffic flow on the whole network and determine the arrival times of the various drivers. We call
τp(t)=arrival time of a driver departing at time t, traveling along the path Γp. |
In practice, we always expect τp(t)<+∞. However, it is possible to envision a situation where traffic becomes completely stuck, and τp becomes infinite. See [15] for a discussion of this issue.
With the above notations, we can introduce
Definition 2.1. An admissible family {ˉui,p} of departure rates is globally optimal if it minimizes the sum of the total costs of all drivers
J(ˉu)≐∑i,p∫(φi(t)+ψi(τp(t)))ˉui,p(t)dt. |
Definition 2.2. An admissible family {ˉuk,p} of departure rates is a Nash equilibrium if no driver of any group can lower his own total cost by changing departure time or switching to a different path to reach destination.
From the above definition it follows the existence of constants C1,…,CN such that
φk(t)+ψk(τp(t))=Ckfor allt∈Supp(ˉuk,p),φk(t)+ψk(τp(t))≥Ckfor allt∈IR. |
As remarked in the Introduction, a similar characterization for the globally optimal solution is much harder to justify.
In the above setting, a natural set of assumptions is (see Figure 11)
(A1) On each road k=1,…,ˉk, the flux function fk satisfies
fk∈C2,f″k<0,fk(0)=fk(ρjamk)=0. | (2.21) |
(A2) For each group of drivers i=1,…,N, the cost functions φi, ψi satisfy
φ′i<0,ψi,ψ′i>0,lim|t|→∞(φi(t)+ψi(t))=+∞. | (2.22) |
When all intersections are modeled in terms of a buffer as in (2.17)–(2.19), under the assumptions (A1), (A2) the existence of a globally optimal solution was proved in [11]. Moreover, if an upper bound on the travel time τp(t)−t can be given, then a Nash equilibrium solution also exists.
Consider a single road, where the traffic density is governed by the conservation law
ρt+f(ρ)x=0forx∈[0,L]. | (3.1) |
We assume that N groups of drivers are present, of sizes G1,…,GN, with departure and arrival costs φi, ψi, i=1,…,N. The flux function will be denoted by
u(t,x)=f(ρ(t,x))=N∑i=1ui(t,x). |
Here
ui(t,x)=θi(t,x)u(t,x) | (3.2) |
is the flux of drivers of the i-th group. As in (2.6), we always assume that
θi(t,x)≥0,∑iθi(t,x)=1. | (3.3) |
For each i∈{1,…,N}, the conservation of the number of drivers of the i-th family yields the additional conservation law
(ρθi)t+(ρv(ρ)θi)x=0. | (3.4) |
By (3.1), one obtains the linear equations
θi,t+v(ρ)θi,x=0,i=1,…,N. | (3.5) |
The incoming flux at the beginning of the road is
u(t,0)=ˉu(t)=N∑i=1ˉθi(t)ˉu(t). | (3.6) |
The global optimization problem can be formulated as follows.
(OP) Given the constants Gi>0, i=1,…,N, find departure rates ˉui(t)=ˉθi(t)ˉu(t) which provide an optimal solution to the problem
minimize:J(ˉu1,…,ˉuN)≐N∑i=1∫+∞−∞[ui(t,0)φi(t)+ui(t,L)ψi(t)]dt, | (3.7) |
and satisfy the constraints
ˉu(t)∈[0,M],ˉθi(t)≥0,N∑i=1ˉθi(t)=1,for allt∈IR, | (3.8) |
ˉui(t)≥0,∫+∞−∞ˉui(t)dt=Gi,i=1,…,N. | (3.9) |
We recall that ui(t,L)=θi(t,L)u(t,L) is the rate at which the drivers of the i-th group arrive at the end of the road.
Since no intersections are present, the existence of a globally optimal solution follows as a special case of the result in [11]. Here we briefly recall the main argument in the proof.
1. Let (ˉu(n)1,…,ˉu(n)N)n≥1 be a minimizing sequence of admissible departure rates. Namely
ˉu(n)i(t)≥0,N∑i=1ˉu(n)i(t)≤M,∫+∞−∞ˉu(n)i(t)dt=Gi |
for every n≥1, and moreover
limn→∞J(ˉu(n)1,…,ˉu(n)N)=infJ(ˉu1,…,ˉuN). |
2. By the assumption (A2), as t→±∞ the cost functions φi,ψi become very large. By possibly modifying the functions ˉui, we can thus obtain a minimizing sequence where all departure rates vanish outside a fixed time interval [a,b].
3. By taking a subsequence, we obtain a weak limit (ˉu(n)1,…,ˉu(n)N)⇀(ˉu1,…,ˉuN).
The boundedness of the supports guarantees that these limit departure rates are still admissible (i.e., no mass leaks at infinity).
4. Call u(n)i(t,x), ui(t,x) the corresponding solutions. By the genuine nonlinearity of the conservation law (3.1), after taking a subsequence, one obtains the strong convergence u(n)(⋅,L)→u(⋅,L) in L1(IR), and the weak convergence of the departure and arrival rates
u(n)i(⋅,0)⇀ui(⋅,0),u(n)i(⋅,L)⇀ui(⋅,L). |
Since the cost functional in (3.7) is linear w.r.t. these departure and arrival rates, it is continuous w.r.t. weak convergence. This yields the optimality of the departure rates (ˉu1,…,ˉuN).
In the remainder of this section, we seek necessary conditions for a solution to be optimal. As a first step, we derive an explicit representation of the solution.
Following [5,6], it is convenient to switch the roles of the variables t,x, and write the density ρ as a function of the flux u. The boundary value problem (3.1)–(3.6) thus becomes a Cauchy problem for the conservation law describing the flux u=ρv(ρ), namely
ux+g(u)t=0, | (3.10) |
u(t,0)=ˉu(t). | (3.11) |
As shown in Figure 12, the function u↦g(u)=ρ is defined as a partial inverse of the function ρ↦ρv(ρ)=u, assuming that
0≤u≤M≐maxρ≥0ρv(ρ),0≤ρ≤ρmax. |
To justify this assumption, consider any solution ρ=ρ(t,x) of (2.1), where the initial data satisfy
ρ(0,x)≤ρmaxfor allx∈[0,L]. |
Then, since all cars that reach the end of the road at x=L can exit immediately, we have
ρ(t,x)≤ρmaxfor allx∈[0,L]andt≥0. |
For convenience, we extend g to the entire real line by setting
g(u)≐{g′(0+)uifu<0,+∞ifu>M. | (3.12) |
The solution to (3.10) and (3.11) can now be expressed by means of the Lax formula [19,25]. Namely, call
g∗(p)≐maxu{pu−g(u)} | (3.13) |
the Legendre transform of g. Notice that
g∗(p)=+∞forp<g′(0). |
On the other hand, for p≥g′(0) the strict convexity of g implies that there exists a unique value u=γ(p)≥0 where the maximum in (3.13) is attained, so that
g∗(p)=p⋅γ(p)−g(γ(p)). |
This function γ:[g′(0),+∞[↦[0,M[ is implicitly defined by the relation
g′(γ(p))=p. | (3.14) |
Consider the integrated function
U(t,x)≐∫t−∞u(τ,x)dτ, |
which measures the number of drivers that have crossed the point x along the road before time t. The conservation law (3.10) can be equivalently written as a Hamilton-Jacobi equation
Ux+g(Ut)=0 | (3.15) |
with data at x=0
¯U(t)≐U(t,0)=∫t−∞ˉu(s)ds. | (3.16) |
The solution to (3.10) and (3.11) is now provided by the Lax formula
U(t,x)=minτ{xg∗(t−τx)+¯U(τ)}, | (3.17) |
τ(t,x)≐argminτ{xg∗(t−τx)+¯U(τ)}, | (3.18) |
u(t,x)=γ(t−τ(t,x)x). | (3.19) |
We observe that the function U=U(t,x) is globally Lipschitz continuous. Its values satisfy
U(t,x)∈[0,G],G=G1+…+GN. | (3.20) |
Car trajectories t↦y(t) are defined to be the solutions to the ODE
˙y(t)=v(ρ(t,y(t))). | (3.21) |
In the region where ρ>0, and hence u=ρv(ρ)>0 as well, these trajectories coincide with the level curves of the integral function U. Indeed, observing that v=u/ρ, when ρ=g(u)>0 we can write
˙y(t)=v(ρ(t,y(t)))=u(t,y(t))g(u(t,y(t))=Ut(t,y(t))g(Ut(t,y(t))). |
By (3.15) one has
ddtU(t,y(t))=Ut+Ux˙y(t)=Ut+UxUtg(Ut)=0. | (3.22) |
More generally, consider a car departing at time t0. The solution to the Cauchy problem
˙y(t)=v(ρ(t,y(t))),y(t0)=0 | (3.23) |
can be determined by the formula
y(t;t0)=inf{x;U(t,x)>¯U(t0)orx=(t−t0)⋅v(0)}. | (3.24) |
The arrival time of a driver departing at time t0 is
τa(t0)=sup{t′;U(t′,L)<¯U(t0)ort′=t0+Lv(0)}. | (3.25) |
By (3.5), the functions θi are constant along car trajectories. With the notation introduced in (3.24), in connection with the boundary data (3.6) we thus have the identities
θi(t,y(t;t0))=ˉθi(t0). | (3.26) |
In order to compute the arrival rates ui=θiu at the terminal point of the road x=L, we first observe that the map
t↦¯U(t) |
in (3.16) is nondecreasing. Hence we can define an inverse by setting
¯τ(s)≐inf{t∈IR;¯U(t)≥s}∈[0,∑iGi]. | (3.27) |
We then introduce the functions
Θi(s)≐ˉθi(ˉτ(s)) | (3.28) |
By (3.5), the functions θi=θi(t,x) are constant along car trajectories. The general solution to (3.5) and (3.6) can thus be written as
θi(t,x)=Θi(U(t,x)). | (3.29) |
We shall be mostly interested in the terminal values u(t,L), describing the rate at which cars arrive at the end of the road. Denoting the arrival distribution as Ua(t)≐U(t,L), the total cost can now be written as
J=∫φ(t)d¯U(t)+∑i∫ψi(t)Θi(Ua(t))dUa(t). | (3.30) |
Theorem 3.1. Let the flux function f satisfy the standard assumptions (2.2). Assume that all drivers have the same departure cost φi=φ and possibly different arrival costs ψ1,…,ψN, satisfying (2.22). Let (ˉu1,…,ˉuN) an optimal departure rate, minimizing the total cost to all drivers.
Then the corresponding solution does not contain shocks. Moreover, there exists constants C1,…,CN such that, setting
ψ(t)≐mink(ψk(t)−Ck), | (3.31) |
the following holds:
(i) For any t∈IR, let T(t) be the unique time such that
φ(t)+ψ(T(t))=0. | (3.32) |
Then, for every point (t′,x′) along the segment with endpoints (t,0) and (T(t),L), one has
u(t′,x′)={γ(T(t)−tL)ifT(t)−tL≥g′(0),0otherwise. | (3.33) |
(ii) Calling ui(⋅,L) the arrival rate of drivers of the i-th group, one has
Supp(ui(⋅,L))⊆{s;ψ(s)=ψi(s)−Ci}. | (3.34) |
A proof of Theorem 3.1 will be given in the next section.
Remark 3.2. The above theorem can be regarded as a first step in the analysis of optimality conditions for traffic flow on a general network. A natural further step would be to look at optimality conditions for (ⅰ) two groups of drivers, starting their journey on the same road and then bifurcating on two distinct roads, or (ⅱ) two groups of drivers starting on two distinct roads that merge into a single one.
We remark that the optimal solution constructed in Theorem 3.1 does not change if the cost functions φi,ψi, i=1,…,N, are replaced by φi+ci,ψi+ci respectively, for any constants c1,…,cN. Therefore, the result remains valid if one only assumes that any two departure costs φi,φj differ by a constant.
As a preliminary, we review the basic theory of scalar conservation laws with convex flux [3,19,29]. Notice that in (3.10) the usual role of the variables t,x is reversed, because of the particular meaning of the equations.
Let u=u(t,x) be a weak solution to (3.10), taking values within the interval [0,M]. This solution is entropy admissible if it contains only downward jumps, namely
u(t+,x)≐lims→t+u(s,x)≤lims→t−u(s,x)≐u(t−,x). |
By a generalized characteristic we mean a function x↦t(x) which provides a solution to the differential inclusion
ddxt(x)∈[g′(u(t+,x)),g′(u(t−,x))]. | (4.1) |
For any given point (T,L), there exists a minimal and a maximal backward characteristic. As shown in Figure 13, we denote by (η−(T),0) and (η+(T),0) the initial points of these characteristics. Calling ¯U the integral function in (3.16), the points η−(T) and η+(T) are respectively the minimum and the maximum elements within the set
I(T)≐{t∈IR;t=argminτ{Lg∗(T−τL)+¯U(τ)}} | (4.2) |
where the function Λ(t)≐Lg∗(T−tL)+¯U(t) attains its global minimum.
Two cases can occur:
(ⅰ) The global minimum in (4.2) is attained at a single point t∗=η−(T)=η+(T).
The function u(⋅,L) is then continuous at the point T, and
u(T,L)=γ(T−t∗L). |
(ⅱ) The global minimum in (4.2) is attained at multiple points, hence η−(T)<η+(T).
In this case the solution contains a shock through the point (T,L). Recalling (3.14), the left and right values across the shock are determined by
u(T−,L)=γ(T−η−(T)L),u(T+,L)=γ(T−η+(T)L). |
We observe that characteristics do not cross each other. Indeed, one has the implication
T1<T2⟹η+(T1)≤η−(T2). | (4.3) |
From (4.3) it immediately follows that
(ⅰ) The profile u(⋅,L) can contain at most countably many shocks. Namely, there can be at most countably many points Ti such that η−(Ti)<η+(Ti).
(ⅱ) There can be at most countably many points ti such that
ti=η+(T1)=η−(T2), | (4.4) |
for two distinct points T1<T2.
We recall that, given any function ϕ∈L1loc(IR), almost every point t∈IR is a Lebesgue point of ϕ. By definition, this means
limh→0+1h∫t+ht−h|ϕ(s)−ϕ(t)|ds=0. | (4.5) |
As proved in [5], if t is a Lebesgue point of the initial datum ˉu(⋅), then there exists a unique forward characteristic starting at t. In particular, t cannot be the center of a rarefaction wave, and there exists a unique point T such that
t∈[η−(T),η+(T)]. | (4.6) |
The next lemma is concerned with the stability of the map T↦η±(T), w.r.t. small perturbations in the initial datum ˉu.
Lemma 4.1. Let u=u(t,x) be the unique entropy weak solution of (3.10) and (3.11). Assume that t is a Lebesgue point for the initial datum ˉu, and let T be the unique point such that (4.6) holds. Then, for any ε>0, one can find δ,δ′>0 such that the following holds.
Let ˉu† be a second initial datum, with
‖ˉu†−ˉu‖L1≤δ. | (4.7) |
If we call u† the corresponding solution, and define the maps (η†)± accordingly, then
[t−δ′,t+δ′]⊆[(η†)+(T−ε),(η†)−(T+ε)]. | (4.8) |
Proof. 1. Let ε>0 be given. By the uniqueness assumption, for the solution u the backward characteristics through the points T−ε and T+ε satisfy
η+(T−ε)<t<η−(T+ε). |
Hence we can find δ′>0 such that
η+(T−ε)<t−2δ′<t+2δ′<η−(T+ε). | (4.9) |
2. If the conclusion of the lemma does not hold, we could find a sequence of initial data ˉun with ‖ˉun−ˉu‖L1→0, such that the corresponding maps η±n satisfy
η+n(T−ε)≥t−δ′orη−n(T−ε)≤t+δ′. | (4.10) |
To fix the ideas, assume that the first case holds. Namely, for every n≥1, there exists tn≥t−δ′ such that
Lg∗(T−ε−tnL)+¯Un(tn)=minτ{Lg∗(T−ε−τL)+¯Un(τ)}. | (4.11) |
By possibly taking a subsequence we can assume tn→t∗≥t−δ′. The uniform convergence ¯Un→¯U now yields
Lg∗(T−ε−t∗L)+¯U(t∗)=minτ{Lg∗(T−ε−τL)+¯U(τ)}. | (4.12) |
This implies η+(T−ε)≥t−δ′, reaching a contradiction.
Remark 4.2. As shown in Figure 13, consider a solution u=u(t,x) containing a shock through the point (T,L). Then we can modify the initial data at x=0 inside the interval [η−(T),η+(T)] so that the solution perturbed solution u† contains a centered compression wave which breaks exactly at (T,L). In view of (3.14), This is achieved by taking
ˉu†(t)={γ(T−tL)ift∈[η−(T),η+(T)],ˉu(t)ift∉[η−(T),η+(T)]. |
This ensures that all characteristics starting at a point (t,0) with t∈[η−(T),η+(T)] join together at the point (T,L).
Consider the constant
λ≐Lg∗(T−η−(T)L)+¯U(η−(T))=Lg∗(T−η+(T)L)+¯U(η+(T)). |
Then the corresponding integrated function ¯U† satisfies
Lg∗(T−tL)+¯U†(t)=λfor allt∈[η−(T),η+(T)]. |
By (4.2), this implies
¯U†(t)≤¯U(t)for allt∈IR, | (4.13) |
while the corresponding solutions coincide at x=L, namely
U†(t,L)=U(t,L),u†(t,L)=u(t,L),for allt∈IR. | (4.14) |
The next lemma, analyzing various perturbations to an optimal solution u, provides the key step toward the proof of Theorem 3.1.
Lemma 4.3. Let (ˉu1,…,ˉuN)=(ˉθ1ˉu,…,ˉθNˉu) be optimal departure rates. Assume that t1,t2 are Lebesgue points for all functions ˉu,ˉθ1,…,ˉθN, and
ˉu(t1)<M,ˉui(t2)>0. | (4.15) |
Call τa(t1),τa(t2) the arrival times of a driver departing at times t1,t2, respectively. Moreover, let T(t1),T(t2) be the times where the (unique) generalized forward characteristic starting from t1,t2 reaches the point L. Then
φ(t1)+ψi(τa(t1))+N∑j=1∫T(t1)τa(t1)ψ′j(s)θj(s,L)ds≥φ(t2)+ψi(τa(t2))+N∑j=1∫T(t2)τa(t2)ψ′j(s)θj(s,L)ds. | (4.16) |
Remark 4.4. The left hand side of (4.16) can be interpreted as the cost for inserting an additional i-driver, departing at time t1. In this case, an additional driver arrives at time T(t1), but this is not the same one! Indeed, the new driver arrives at time τa(t1). However, the presence of this additional car slows down all the other cars whose arrival time is T∈[τa(t1),T(t1)]. The delay in the arrival time of all these cars causes a further increase in the total cost, accounted by the integral term on the left hand side of (4.16). Similarly, the right hand side is the amount which can be saved by removing an i-driver departing at time t2.
Proof of Lemma 4.3. 1. Since t1,t2 are Lebesgue points of ˉu, they cannot be the center of a rarefaction wave. Hence there exist unique points T1=T(t1) and T2=T(t2) such that
ti=argminτ{Lg∗(Ti−τL)+¯U(τ)},i=1,2. | (4.17) |
Assuming that (4.16) fails, we shall derive a contradiction. Indeed, we will construct a new initial data ˉu†i which is slightly smaller than ˉui in a neighborhood of t1 and slightly larger than ˉui in a neighborhood of t2, yielding a lower total cost. Various cases can arise, depending on the relative position of τa(ti) and T(ti). To fix the ideas, in the following we assume that
τa(t1)<T(t1)<τa(t2)<T(t2), | (4.18) |
as shown in Figure 14. The other cases are handled in a similar way.
We observe that the above strict inequalities imply that u(⋅,L) is strictly positive on the intervals [τa(t1),T(t1)] and [τa(t2),T(t2)]. Indeed, the car speed is always ≤v(0)=1g′(0). As shown in Figure 15, if
τa(t1)−t1L=1v(0)=g′(0), |
then the car speed would be identically equal to the maximum speed v(0). In this case the car trajectory coincides with a characteristic line, and hence T(t1)=τa(t1), against the assumption (4.18). Therefore, we must have
τa(t1)−t1L>g′(0),τa(t1)−Lv(0)<t1. |
Since characteristics do not cross each other, for every T∈]τa(t1),T(t1)[ the initial point of a characteristic through (T,L) must satisfy
η(T)=T−L⋅g′(u(T,L))≤t1. |
Hence
g′(u(T,L))≥T−t1L≥τa(t1)−t1L>g′(0). | (4.19) |
Since g′ is an increasing function, this yields a lower bound on u(T,L).
2. Consider a perturbed set of initial data of the form (ˉu1,…,ˉu†i,…,ˉuN), where only the component ˉui is modified. The new departure rate for drivers of the i-th group is chosen so that
{ˉu†i(t)=ˉui(t)ift∉[t1,t1+δ′]∪[t2,t2+δ′],ˉu†i(t)≥ˉui(t)ift∈[t1,t1+δ′],0≤ˉu†i(t)≤ˉui(t)ift∈[t2,t2+δ′], | (4.20) |
∫t1+δ′t1[ˉu†i(s)−ˉui(s)]ds=∫t2+δ′t2[ˉui(s)−ˉu†i(s)]ds=δ>0. | (4.21) |
Given ε>0, according to Lemma 4.1, we can choose δ,δ′>0 small enough so that the perturbation in the initial datum
u(t,0)=ˉu(t)=N∑i=1ˉui(t) |
affects the values of u(⋅,L) only in a small neighborhood of the points T(t1),T(t2), namely
u†(t,L)=u(t,L)for allt∉[T(t1)−ε,T(t1)+ε]∪[T(t2)−ε,T(t2)+ε]. | (4.22) |
3. We now consider a sequence of perturbations of the form (4.20) and (4.21), with εn,δn,δ′n→0. Calling τan(t) the corresponding arrival times, we claim that the following holds.
(C) Let t be a Lebesgue point for ˉu, with ˉu(t)>0, and let τa(t) be a Lebesgue point for u(⋅,L). Then u(τa(t),L)>0 and the following implications hold.
τa(t1)<τa(t)<T1⟹limn→∞τan(t)−τa(t)δn=1u(τa(t),L), | (4.23) |
τa(t2)<τa(t)<T2⟹limn→∞τan(t)−τa(t)δn=−1u(τa(t),L), | (4.24) |
τa(t)∉[τa(t1),T1]∪[τa(t2),T2]⟹limn→∞τan(t)−τa(t)δn=0. | (4.25) |
In first approximation, the above limits show that:
● For those drivers who were reaching destination at a time T∈]τa(t1),T1[, the arrival time is delayed by δn/u(τa(t),L).
● For those drivers who were reaching destination at a time T∈]τa(t2),T2[, the arrival time is anticipated by δn/u(τa(t),L).
● For all other drivers, the arrival time does not change.
To prove the above claim we first observe that, if u(τa(t),L)=0, then u(t′,x′)=0 along the backward characteristic
{(t′,x′);x′=L−(τa(t)−t′)v(0)}. |
but in this case, this characteristic coincides with a car trajectory. Hence ˉu(t)=u(0,t)=0 as well, contradicting our first assumption.
To prove (4.23), assume τa(t1)<τa(t)<T1. Then, for all n large enough, the arrival time τan(t) is uniquely determined by the identity
U(τan(t),L)=U(τa(t),L)+δn. | (4.26) |
Observing that the partial derivative is
∂∂τU(τ,L)|τ=τa(t)=u(τa(t),L), |
from (4.26) one obtains (4.23). Notice that here the denominator is uniformly positive, as a consequence of (4.19).
The proof of (4.24) is entirely similar, replacing (4.26) with the identity
U(τan(t),L)=U(τa(t),L)−δn. | (4.27) |
Finally, if the condition on the left hand side of (4.25) holds, then for all n≥1 sufficiently large one has τan(t)=τa(t), and the implication is trivial.
4. By the properties (4.23) it follows
limn→∞1δnN∑j=1∫τa(t)∈]τa(t1),T1[[ψj(τan(t))−ψj(τa(t))]ˉuj(t)dt=N∑j=1∫τa(t)∈]τa(t1),T1[ψ′j(τa(t))⋅limn→∞τan(t)−τa(t)δnˉuj(t)dt=N∑j=1∫T1τa(t1)ψ′j(τ)⋅1u(τ,L)uj(τ,L)dτ=N∑j=1∫T1τa(t1)ψ′j(τ)θj(τ,L)dτ. |
An entirely similar computation can be performed on the interval ]τa(t2),T2[. Combining these estimates, we thus conclude
limn→∞J(ˉu1,…,ˉu†i,…,ˉuN)−J(ˉu1,…,ˉui,…,ˉuN)δn=φ(t1)+ψi(τa(t1))+N∑j=1∫T(t1)τa(t1)ψ′j(s)θj(s,L)ds−[φ(t2)+ψi(τa(t2))+N∑j=1∫T(t2)τa(t2)ψ′j(s)θj(s,L)ds]. | (4.28) |
If the inequality (4.16) does not hold, then the right hand side of (4.28) is negative. This yields a contradiction with the optimality of the departure rates (ˉu1,…,ˉui,…,ˉuN).
5. The above analysis proves the lemma in the case where (4.18) holds. On the other hand, if T(t1)=τa(t1), then a small perturbation of the departure rate on the interval [t1,t1+δn] will modify the arrival rate only in a small neighborhood of τa(t1) (see Figure 15(right)). In this case, one directly proves that the limit (4.28) remains valid, since the integral over the interval [τa(t1),T(t1)] trivially vanishes.
Let ui(t,0)=ˉθi(t)ˉu(t) be optimal departure rates, and let u, θi be the corresponding solutions to
ux+g(u)t=0,θi,t+v(g(u))θi,x=0,i=1,…,N. | (4.29) |
The proof will be worked out in several steps.
1. We begin by showing that an optimal solution cannot contain any shock in the interior of the domain, i.e., for 0≤x<L.
Indeed, assume on the contrary that a shock is present, and let (T,L) be the terminal position of this shock. According to Remark 4.2, we can change the initial datum so that the new solution u† contains a centered compression wave focusing at (T,L). More precisely, define ¯U†, ˉu† as in Remark 4.2. Assuming that the functions Θi(s) are defined by
¯Ui(t)≐Θi(¯U(t))⋅¯U(t), | (4.30) |
define the components (¯U†1,…,¯U†N) by setting
¯U†i(t)≐Θi(¯U†(t))⋅¯U†(t). | (4.31) |
Notice that these definition imply
U†i(t,L)=Ui(t,L)for allt∈IR, |
hence the arrival costs remain the same. On the other hand, we have
¯U†i(t)≤¯Ui(t), | (4.32) |
for all i,t. Observing that there exists some i and some t∈[η−(T),η+(T)] where (4.32) is satisfied as a strict inequality, we claim that the total departure cost for the perturbed solution is strictly smaller.
To see this, introduce a variable ξ∈[0,Gi] labeling drivers of the i-th group. Define the departure times
ti(ξ)≐inf{t;¯Ui(t)>ξ},t†i(ξ)≐inf{t;¯U†i(t)>ξ}. | (4.33) |
By the previous definitions it follows
ti(ξ)≤t†i(ξ), |
with strict inequality holding at least for some index i and some values of ξ∈[0,Gi]. We now compute
∑i∫φ(t)d¯U†i(t)=∑i∫Gi0φ(t†i(ξ))dξ<∑i∫Gi0φ(ti(ξ))dξ=∑i∫φ(t)d¯Ui(t), |
proving our claim.
2. Next, we claim that an optimal departure rate satisfies
ˉu(t)<Mfor allt∈IR. | (4.34) |
Indeed, since the characteristic speed satisfies g′(u)→+∞ as u→M, if ˉu(τ)=M at some point τ then the solution u(⋅,x) would immediately contain a shock, for every x>0. By the previous step, this contradicts the optimality assumption.
3. According to Lemma 4.3, by (4.34), the quantity
ΔJ(i,t)=φ(t)+ψi(τa(t))+N∑j=1∫T(t)τa(t)ψ′j(s)θj(s,L)ds | (4.35) |
is equal to some constant Ci for all t∈Supp(ˉui), and is greater or equal to Ci for all t∈IR. In other words, for each i=1,…,N we have
φ(t)+ψi(τa(t))−Ci+N∑j=1∫T(t)τa(t)ψ′j(s)θj(s,L)ds=0for allt∈Supp(ˉui), | (4.36) |
φ(t)+ψi(τa(t))−Ci+N∑j=1∫T(t)τa(t)ψ′j(s)θj(s,L)ds≥0for allt∈IR. | (4.37) |
This implies
φ(t)+ψi(τa(t))−Ci=φ(t)+minj(ψj(τa(t))−Cj)=−N∑j=1∫T(t)τa(t)ψ′j(s)θj(s,L)ds | (4.38) |
for all t∈Supp(ˉui). We now observe that, for a.e. s∈[τa(t),T(t)] and j∈{1,…,N}, one has the implication
θj(s,L)>0⟹ψj(s)−Cj=mink(ψk(s)−Ck). | (4.39) |
Moreover, for every j,k and a.e. s in the set
{s∈IR;ψj(s)−Cj=ψk(s)−Ck}, |
one has ψ′j(s)=ψ′k(s). Therefore, defining ψ as in (3.31) and recalling that ∑jθj=1, we obtain
N∑j=1∫T(t)τa(t)ψ′j(s)θj(s,L)ds=∫T(t)τa(t)ψ′(s)ds. | (4.40) |
In turn, this implies
φ(t)+ψi(τa(t))−Ci+N∑j=1∫T(t)τa(t)ψ′j(s)θj(s,L)ds=φ(t)+ψi(τa(t))−Ci+∫T(t)τa(t)ψ′(s)ds=φ(t)+ψi(τa(t))−Ci+ψ(T(t))−ψ(τa(t))=φ(t)+ψ(T(t))=0. | (4.41) |
According to (4.41), each characteristic where the solution u is positive must connect two points (t,0) with (T(t),L) with φ(t)+ψ(T(t))=0. This proves part (ⅰ) of Theorem 3.1. Finally, part (ⅱ) follows from (4.39).
Here we illustrate how these necessary conditions can be used to construct optimal solutions. For simplicity, we shall assume that the cost functions ψi are C2 and satisfy the assumption
(A3) For any i≠j one has the implication
ψ′i(t)=ψ′j(t)⟹ψ″(t)≠ψ″j(t). | (5.1) |
Notice that, by (5.1), for any given constants C1,…,CN, the set of times
{t∈IR;ψi(t)−Ci=ψj(t)−Cjfor some i≠j} |
consists only of isolated points, hence it has measure zero.
We remark that the assumption (A3) is generically valid in the space of twice continuously differentiable functions. Indeed, given ε>0 and any N-tuple of twice continuously differentiable functions (ˆψ1,…,ˆψN), by a small perturbation one can construct functions ψ1,…,ψN which satisfy (A3) together with
‖ˆψi−ψi‖C2<ε,i=1,…,N. |
Let now G1,…,GN be the sizes of the N groups of drivers. In order to construct a globally optimal family of departure rates u1(⋅),…,uN(⋅), we introduce the following algorithm.
(ⅰ) Start by guessing N constants C1,…,CN, and define the cost function ψ as in (3.31).
(ⅱ) Let u=u(t,x) be the solution of (3.10) constructed according to (3.32) and (3.33).
(ⅲ) Define the sets
Ai={t∈IR;u(t,L)>0,ψi(t)−Ci=mink(ψk(t)−Ck)}. | (5.2) |
Notice that, by (A3), for a.e. t∈IR the minimum in (5.2) is attained by a unique index k∈{1,…,N}.
(ⅳ) Consider the map Λ:IRN↦IRN, Λ(C1,…,CN)=(κ1,…,κN), where κi is the total number of drivers of the i-th group, defined by
κi≐∫Aiu(t,L)dt. | (5.3) |
Determine values ˆC1,…,ˆCN such that
Λ(ˆC1,…,ˆCN)=(G1,…,GN). | (5.4) |
(ⅴ) Set
ψ(t)=mini(ψi(t)−ˆCi). |
And let u=u(t,x) be the solution of (3.10) whose characteristics satisfy (3.32).
(ⅶ) Finally, for T∈Ai, call η(T) the departure time of the driver that arrives at time T. This is obtained by solving the ODE
˙x(t)=v(t,x(t))=u(t,x)g(u(t,x)), | (5.5) |
with terminal condition
x(T)=L. | (5.6) |
The solution t↦x(t,T) of (3.22) and (5.6) yields the trajectory of a car arriving at the end of the road time T. Its departure time η(T) is defined by the equality x(η(T))=0.
We now consider the sets of departure times
A∗i≐{η(t);t∈Ai}. |
The departure distribution
ˉui(t)˙={ˉu(t)ift∈A∗i,0ift∉A∗i, |
then satisfies all the necessary conditions for optimality.
We remark that these conditions are only necessary, not sufficient for optimality. Since an optimal solution exists, and can obtained by the above method, the previous analysis implies that, if the N-tuple (ˆC1,…,ˆCN) which satisfies (5.4) is unique, then this must yield the optimal solution.
Example 4. We seek a globally optimal departure rate for two groups of drivers, with sizes G1=G2=2.51 on a road with length L=10. The conservation law governing traffic density is
ρt+[ρv(ρ)]x=0,v(ρ)=2−ρ | (5.7) |
The departure and arrival costs for drivers of the two groups are
φ(t)=−tψ1(t)=et−4ψ2(t)=et−7.6. | (5.8) |
The optimal solution, found by the algorithm described above, is shown in Figure 16. The marginal costs for adding one more driver of the first group or of the second group are found to be ˆC1=5.18 and ˆC2=2.10, respectively.
This research was partially supported by NSF, with grant DMS-1411786: "Hyperbolic Conservation Laws and Applications".
The authors declare no conflict of interest.
[1] |
Bellomo N, Delitala M, Coscia V (2002) On the mathematical theory of vehicular traffic flow I: Fluid dynamic and kinetic modeling. Math Models Methods Appl Sci 12: 1801–1843. doi: 10.1142/S0218202502002343
![]() |
[2] |
Bellomo N, Dogbe C (2011) On the modeling of traffic and crowds: A survey of models, speculations, and perspectives. SIAM Rev 53: 409–463. doi: 10.1137/090746677
![]() |
[3] | Bressan A (2000) Hyperbolic Systems of Conservation Laws: The One Dimensional Cauchy Problem. Oxford University Press. |
[4] |
Bressan A, Canic S, Garavello M, et al. (2014) Flow on networks: Recent results and perspectives. EMS Surv Math Sci 1: 47–111. doi: 10.4171/EMSS/2
![]() |
[5] |
Bressan A, Han K (2011) Optima and equilibria for a model of traffic flow. SIAM J Math Anal 43: 2384–2417. doi: 10.1137/110825145
![]() |
[6] |
Bressan A, Han K (2012) Nash equilibria for a model of traffic flow with several groups of drivers. ESAIM Control Optim Calc Var 18: 969–986. doi: 10.1051/cocv/2011198
![]() |
[7] |
Bressan A, Liu CJ, Shen W, et al. (2012) Variational analysis of Nash equilibria for a model of traffic flow. Quarterly Appl Math 70: 495–515. doi: 10.1090/S0033-569X-2012-01304-9
![]() |
[8] |
Bressan A, Marson A (1995) A variational calculus for discontinuous solutions of conservative systems. Commun Part Diff Eq 20: 1491–1552. doi: 10.1080/03605309508821142
![]() |
[9] | Bressan A, Marson A (1995) A maximum principle for optimally controlled systems of conservation laws. Rend Sem Mat Univ Padova 94: 79–94. |
[10] |
Bressan A, Nguyen K (2015) Conservation law models for traffic flow on a network of roads. Netw Heter Media 10: 255–293. doi: 10.3934/nhm.2015.10.255
![]() |
[11] |
Bressan A, Nguyen K (2015) Optima and equilibria for traffic flow on networks with backward propagating queues. Netw Heter Media 10: 717–748. doi: 10.3934/nhm.2015.10.717
![]() |
[12] |
Bressan A, Nordli A (2017) The Riemann Solver for traffic flow at an intersection with buffer of vanishing size. Netw Heter Media 12: 173–189. doi: 10.3934/nhm.2017007
![]() |
[13] |
Bressan A, Shen W (2007) Optimality conditions for solutions to hyperbolic balance laws, In: Ancona, F., Lasieka, I., Littman, W., et al. Editors. Control Methods in PDE - Dynamical Systems, AMS Contemporary Mathematics 426: 129–152. doi: 10.1090/conm/426/08187
![]() |
[14] |
Bressan A, Yu F (2015) Continuous Riemann solvers for traffic flow at a junction. Discr Cont Dyn Syst 35: 4149–4171. doi: 10.3934/dcds.2015.35.4149
![]() |
[15] |
Chitour Y, Piccoli B (2005) Traffic circles and timing of traffic lights for cars flow. Discrete Contin Dyn Syst B 5: 599–630. doi: 10.3934/dcdsb.2005.5.599
![]() |
[16] |
Coclite GM, Garavello M, Piccoli B (2005) Traffic flow on a road network. SIAM J Math Anal 36: 1862–1886. doi: 10.1137/S0036141004402683
![]() |
[17] |
Dafermos C (1972) Polygonal approximations of solutions of the initial value problem for a conservation law. J Math Anal Appl 38: 33–41. doi: 10.1016/0022-247X(72)90114-X
![]() |
[18] | Daganzo C (1997) Fundamentals of Transportation and Traffic Operations. Oxford, UK: Pergamon-Elsevier. |
[19] | Evans LC (2010) Partial Differential Equations. 2 Eds., Providence, RI: American Mathematical Society. |
[20] | Garavello M, Han K, Piccoli B (2016) Models for Vehicular Traffic on Networks. Missouri: AIMS Series on Applied Mathematics, Springfield. |
[21] | Garavello M, Piccoli B (2006) Traffic Flow on Networks. Conservation Laws Models. Missouri: AIMS Series on Applied Mathematics, Springfield. |
[22] |
Garavello M, Piccoli B (2009) Traffic flow on complex networks. Ann Inst H Poincaré Anal Nonlinear 26: 1925–1951. doi: 10.1016/j.anihpc.2009.04.001
![]() |
[23] |
Herty M, Moutari S, Rascle M (2006) Optimization criteria for modeling intersections of vehicular traffic flow. Netw Heterog Media 1: 275–294. doi: 10.3934/nhm.2006.1.275
![]() |
[24] |
Holden H, Risebro NH (1995) A mathematical model of traffic flow on a network of unidirectional roads. SIAM J Math Anal 26: 999–1017. doi: 10.1137/S0036141093243289
![]() |
[25] |
Lax PD (1957) Hyperbolic systems of conservation laws. Comm Pure Appl Math 10: 537–556. doi: 10.1002/cpa.3160100406
![]() |
[26] | Lighthill M, Whitham G (1955) On kinematic waves. II. A theory of traffic flow on long crowded roads. P Roy Soc A Math Phys Eng Sci 229: 317–345. |
[27] |
Pfaff S, Ulbrich S (2015) Optimal boundary control of nonlinear hyperbolic conservation laws with switched boundary data. SIAM J Control Optim 53: 1250–1277. doi: 10.1137/140995799
![]() |
[28] |
Richards PI (1956) Shock waves on the highway. Oper Res 4: 42–51. doi: 10.1287/opre.4.1.42
![]() |
[29] | Smoller J (1994) Shock Waves and Reaction-Diffusion Equations. 2 Eds., New York: Springer-Verlag. |
[30] |
Ulbrich S (2002) A sensitivity and adjoint calculus for discontinuous solutions of hyperbolic conservation laws with source terms. SIAM J Control Optim 41: 740–797. doi: 10.1137/S0363012900370764
![]() |