0% found this document useful (0 votes)
33 views25 pages

Mô Hình Eulerian

Mô tả mô hình
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
33 views25 pages

Mô Hình Eulerian

Mô tả mô hình
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/3332810

Joint-based control of a new Eulerian network model of air traffic flow

Article in IEEE Transactions on Control Systems Technology · October 2006


DOI: 10.1109/TCST.2006.876904 · Source: IEEE Xplore

CITATIONS
READS
115
199

3 authors, including:

Alexandre M. Bayen
Robin Raffard
University of California, Berkeley
Stanford University
364 PUBLICATIONS 13,228 CITATIONS
23 PUBLICATIONS 829 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Robin Raffard on 16 March 2014.

The user has requested enhancement of the downloaded file.


804 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

Adjoint-Based Control of a New Eulerian Network


Model of Air Traffic Flow
Alexandre M. Bayen, Member, IEEE, Robin L. Raffard, and Claire J. Tomlin, Member, IEEE

Abstract—An Eulerian network model for air traffic flow in Digital Object Identifier 10.1109/TCST.2006.876904
the National Airspace System is developed and used to design
flow control schemes which could be used by Air Traffic
Controllers to optimize traffic flow. The model relies on a
modified version of the Lighthill–Whitham–Richards (LWR)
partial differential equa- tion (PDE), which contains a velocity
control term inside the diver- gence operator. This PDE can be
related to aircraft count, which is a key metric in air traffic
control. An analytical solution to the LWR PDE is constructed
for a benchmark problem, to assess the gridsize required to
compute a numerical solution at a prescribed accuracy. The
Jameson–Schmidt–Turkel (JST) scheme is selected among other
numerical schemes to perform simulations, and evi- dence of
numerical convergence is assessed against this analytical
solution. Linear numerical schemes are discarded because of
their poor performance. The model is validated against actual air
traffic data (ETMS data), by showing that the Eulerian
description en- ables good aircraft count predictions, provided a
good choice of numerical parameters is made. This model is then
embedded as the key constraint in an optimization problem, that
of maximizing the throughput at a destination airport while
maintaining aircraft density below a legal threshold in a set of
sectors of the airspace. The optimization problem is solved by
constructing the adjoint problem of the linearized network
control problem, which provides an explicit formula for the
gradient. Constraints are enforced using a logarithmic barrier.
Simulations of actual air traffic data and control scenarios
involving several airports between Chicago and the U.S. East
Coast demonstrate the feasibility of the method.
Index Terms—Adjoint-based optimization, control of partial
dif- ferential equations, LWR PDE.

I. INTRODUCTION

T HE National Airspace System (NAS) consists of aircraft,


control facilities, procedures, navigation and surveillance
equipment, analysis equipment, decision support tools, and
con- troller pilots who operate the systems. In this article, the
focus is traffic flow management (TFM), which has the goal to
maximize throughput while maintaining safety. This entails
the design of efficient methods to route aircraft, while
preventing the density of aircraft from becoming too large in
regions of airspace, and op- erating efficient reroutes when the
weather does not allow traffic to cross a given region of
airspace. These tasks are not currently

Manuscript received May 11, 2005. Manuscript received in final form


March 27, 2006. Recommended by Associate Editor S. Devasia. This work
was supported in part by NASA under Grant NCC 2-5422, by the Office of
Naval Research (ONR) under MURI Contract N00014-02-1-0720, by the
Defense Advanced Research Projects Agency (DARPA) under the Software
Enabled Control Program (AFRL Contract F33615-99-C-3014), and by a
Graduate Fellowship of the Délégation Générale pour l’Armement, France.
The authors are with the Department of Aeronautics and Astronautics,
Stan- ford University, Stanford, CA 94305-4035 USA and are also with the
Depart- ment of Electrical Engineering and Department of Civil and
Environmental En- gineering, University of California at Berkeley, Berkeley,
CA 94720-1710 USA.
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 805
optimized with respect to throughput or maximal density.
FLOW accurately. Second, we show that fast numerical analysis tools
Rather, they are prescribed by playbooks, which are can be applied efficiently to this problem for simulation
procedures that have been established over time, based on purposes. The main difference be- tween ours and previous
controller experience. work using LWR models of air traffic
The key objective of this article is to design control [24] or highway traffic [14], [26], [34], is that we generate an
strategies in the form of “flow patterns,” that is, to come op- timization problem (with throughput and maximal density
up with ways to route streams of aircraft by generating the as an objective function) using the continuous PDE directly,
corresponding aircraft velocities, rather than optimizing instead of its discretization. We show that the use of linear
local trajectories of aircraft. Ideally, one would like to numerical schemes to approximate the solution of the PDE
automatically generate procedures implementable by air perform very poorly, which unfortunately precludes the use of
traffic control (ATC), of the following kind: “aircraft on standard linear optimization programs to control the system.
airway 148 at 33 000 ft, fly at 450 kn for the next hour and Controlling transportation networks in general is extremely
then accelerate by 10 kn per half hour.” This sug- gests challenging and numerically difficult [15], [26]. In the present
following an Eulerian approach advocated by Menon et al. case, the control consists of speed assignments and routing
[24] and dividing the airspace into line elements poli- cies. We show that we may use flow control techniques
corresponding to portions of airways, on which the density [8], which are directly applicable to PDE-driven systems.
of aircraft as a function of time and of the coordinate along Namely, we pose the optimal control of the network as an
the line, is modeled. A traditional way to describe the optimization program, whose variables are solutions to a set
evolution of the density of vehicles in a network is to use a of PDEs and satisfy additional inequality constraints. The op-
partial differential equation (PDE). This PDE appears timization is performed by updating the control variables in
naturally in highway traffic and is called the Lighthill– the opposite direction of the gradient of the objective function.
Whitham–Richards (LWR) PDE [22], [30]. In this work, The gradient is derived using an adjoint method, specially
we will derive a modified version of the LWR PDE adapted to the case in which the system is described by a set of
specifically applicable to the ATC problem of interest. PDEs cou- pled through the boundary conditions, in the
First, we show that despite the information loss inherent presence of con- straints. This algorithm does not provide
in any Eulerian model, the aircraft count (which is a proofs of convergence to a global optimal. However, this
crucial ATC metric defined in this article) is predicted method, as well as other flow
1063-6536/$20.00 © 2006 IEEE
806 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

control approaches [19], [3], [1], [13], [21], [32], have been
shown to work extremely well in practice in fluid mechanics. Introducing , it is fairly easy to see that
In addition, though we consider networks of PDEs, the dimen- if an aircraft were at location at time , it would be at at
sion of each PDE is one, enabling online implementations, as time . Because of the sign of is
solving a set of one dimensional PDEs may be done extremely invertible, and, therefore, is related to and by
quickly. As such, we demonstrate the feasibility of generating
.
direct, open-loop control solutions to the air traffic flow
control problem using accurate numerical schemes. Consider a point and a distance such that
There are a few benefits of the above outlined approach . The number of aircraft between and at
over Lagrangian methods, which incorporate all trajectories of can be related to the number of aircraft at at locations
all aircraft. and (con-
• Most of the Lagrangian methods pose the control servation of aircraft):
problem as an integer optimization program, which is . In other
intractable in real-time because it is NP complete. In words, assuming that there is no inflow at 0
addition, the solution provided by these methods often
takes advantage of actuating single aircraft individually,
which precludes the derivation of global policies. The
Eulerian framework scales well with the number of
aircraft (the larger the number of aircraft is, the more Some simple algebra (two successive applications of the chain
accurate the model be- comes, without further rule) shows that the space derivative and the time derivative of
computational complexity). are related by
• The method presented in this article is general and can
be very easily adapted to specific classes of controllers
(smooth, continuous, piecewise affine, etc.): it is possible
to use this method to derive a control law in a required
format, which is compatible with aircraft capabilities. We recognize this as a first-order linear hyperbolic PDE, and
• This method can be applied to highway traffic with can now enunciate the following proposition.
minor modifications [6], and we believe can be extended Proposition 1: Let be a func-
to other problems such as networks of irrigation channels
tion with a finite number of discontinuities at
[23].
on . Assume for all . Let
This article is organized as follows. In Section II, we will
and . Then the fol-
first rederive the LWR PDE for the case of interest and
lowing PDE:
generalize it to a network. Then, we determine an analytical
solution for the case of time-invariant velocity control, which,
in
in Section II-C,
will be used for numerical validation purposes. In Section III,
in (1)
we explain how to identify the numerical values of the
in
parameters for the airspace of interest using enhanced traffic
management system (ETMS) data. This model is validated
against ETMS data in Section IV. In Section V, we derive the admits a unique continuous (weak) solution, given by
adjoint system to our problem, and show how to use it to
determine the mean velocity profiles along the links as well as
the routing policy. Finally, in Section VI, we show how to
apply this to control a very busy portion of airspace: the area
enclosed by Chicago, New York, Boston, and the eastern (2)
coast of Canada. if

II. NEW EULERIAN NETWORK MODEL OF AIRSPACE tion of airway . Assuming a static mean velocity profile
defined on represents the mean velocity of air-
A. Modified LWR Model of Air Traffic craft at location , and the motion of an aircraft is described by
In describing the air traffic system, like the road system, the dynamical system .
one has to first look at aircraft (or cars) present in the system
and esti- mate a density of vehicles. Therefore, given a portion
of airspace (airway or sector), one needs to introduce the
aircraft count [9] defined as the number of aircraft in that
region. Let us consider a portion airway of length , described
by a coordinate . The number of aircraft in the
segment at time is called . Thus,
represents the aircraft count on the por-
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 807
FLOW

if

where , and is its inverse.


Proof: See the Appendix.
In (1), represents the inflow at the entrance of the link
(i.e., at ). In highway traffic flow analysis, is
sometimes referred to as cumulative flow. It can be related
to the vehicle density through the integral relation

(3)
808 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

where is the vehicle density. It can be checked that the One can also show that when , and are the same
vehicle density satisfies the following PDE:

(4)

Equation (4) can be related to (1) by a simple integration of


along . Equation (4) is a mass conservation equation.
Note that it is very different from the original LWR PDE [22],
At this stage, we have three quantities: , and . The
[30], [2], [15], [12], which is a first order hyperbolic
meaning of as we know it in fluid mechanics assumes a
conservation law. In particular, (4) does not have a fundamental
large number of particles (i.e., aircraft) per unit volume (the
diagram, i.e., there is no functional relation between and , or
threshold is defined by the Knüdsen number). In the present
between and the flux. In the implementation studied in this
case, the number of aircraft we consider will almost certainly
paper, the function will represent the control input. It is
be below this number, meaning that the fluid approximation is
also possible to rewrite the first equation in (4) as
question- able. This means that instead of using ,
we will use in the PDE. We will justify this
(5) approximation with appropriate validations.

which provides the following corollary. B. Network Model


Corollary 2: The solution of (5) for is given by The model of the previous section describes traffic on a
single portion of airway or line element. As was done earlier
for high- ways [15], this model can be generalized to airway
if networks, i.e., sets of interconnected airways, as shown in Fig. 1.
We now derive a framework to describe unidirectional air
otherwise.
traffic. We describe the topology of the network by a
unidirectional graph , in which is the set of edges or
The interpretation of the corollary follows. The quantity is links, and the set of vertices. We index the links by
conserved along the characteristic curves , rather than by the indices of the two corresponding
. At this stage, is defined by and satisfies vertices. For all , we call
(4). However, unlike for highway traffic, the density might the set of upstream links merging into link , and the
not be the best way to characterize the flow situation at a set of links for which the upstream links are only merging. The
given time. If the number of aircraft in the system is small, number of links merging into a single link is not limited; it is
will be a set of spikes, which is intractable numerically. possible to have . If there is a divergence at the end
Therefore, a more tractable quantity to work with would be of a link , we assume for simplicity that there are only two
, where represents the number of aircraft contained emanating links from the corresponding vertex. We index by
in a finite interval of length . This quantity does not a priori and the two em- anating links (left and right), and call the
satisfy the PDE (4). It is meaningful to introduce an additional portion of the flow going from to , and the proportion
“density-like” quantity called , which satisfies the PDE and of the flow going from to . We call the set of links with a
for which we can suggest a physical interpretation divergence at the end of it. The are not known a priori and
have to be determined. These coefficients might depend on
as well and, therefore, a depen- dence is included in the
model. We call the set of sources in the network and a
sink of the network, at which we might want to perform
where is a reference time. represents the number optimization. We index all variables of the pre- vious section
of aircraft included in a window of time units around lo- by : the aircraft density on link is , the coordi- nate is ,
cation and can be referred to as “time density.” This way of the main velocity profile is , etc. Note that we are not using
accounting for density is meaningful for air traffic control, Einstein's notation; the notation is summarized in Table I. The
since it incorporates a time scale into the density governing PDE system thus reads
computation and thus, provides access to the time separation
between aircraft. It is easy to show that itself satisfies the
same PDE as for any value of

(6)
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 809
FLOW

Fig. 1. Top: Tracks of flights incoming into Chicago (ORD). The upper stream comes from Canada, the lower from New York and Boston (BOS). Additional
streams merge into the network (Detroit and Hartford Bradley). Bottom: Network model for the tracks shown above, with waypoints labeled. The model
includes five links, merging into ORD. The corresponding inflow terms correspond to a single airport as in BOS or Detroit (DTW), or to a set of airports, as in
New York (EWR, JFK, LGA).

TABLE I C. Accuracy of Numerical Solutions


NOTATION FOR THE NETWORK PROBLEM
Even for a single link it is, in general, not possible to solve
the system (6) analytically when depends on time. The solu-
tions of the LWR PDE in the system (6) have very undesir-
able properties for numerical integration: they are by construc-
tion discontinuous;1 they can develop kinks if the velocity pro-
files are discontinuous. Several numerical schemes of the
orig- inal LWR PDE have been the focus of recent research
[16] in order to address similar difficulties encountered in the
original LWR PDE; they have proved extremely efficient in
the case of highway traffic. We have chosen to use three
different schemes to compare their respective benefits.
1) The well-known Lax–Friedrichs scheme [17].
In the previous system, the PDE (first equation) describes the 2) A left-centered scheme, inspired by the Daganzo scheme
[16] in light traffic
evolution of on each link. The notation represents the
LWR operator. The second equation is the initial condition
(i.e., the initial density of aircraft on each link). The third
equation expresses the conservation of aircraft at the merging
points. The fourth and fifth equations express the conservation 3) The Jameson–Schmidt–Turkel
of aircraft at the divergence points. The last equation
expresses the boundary conditions (inflow at the sources of (JST) scheme. This scheme is nonlinear and has very
the network). The sinks of the system are free boundary desirable properties for this work: it captures shocks
conditions and, therefore, do not appear in the previous (which are present in the solutions we
system. The solution of (6) enables the computation of certain 1Unlike n which is its primitive and is continuous.
metrics useful for ATC. For example, one quantity of interest
is the aircraft count per sector.
810 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

Fig. 2. L error due to the discretization method, as a function of the number obtain useful
of grid points for both schemes. Lax-Friedrichs scheme (solid), Jameson-
Schmidt- Turkel scheme ( ), left-centered scheme (- -).

compute, as will be seen), and when the PDE has an en-


tropy solution, which is the case for highway traffic in the
original LWR setting, it converges to the entropy solution
of the problem. Details of this scheme are available in [20].
Even if a numerical scheme is theoretically proven to con-
verge to the analytical solution of a PDE, one usually does not
know a priori the required gridsize to guarantee that the nu-
merical solution is close to the analytical solution. This type
of validation is standard in numerical analysis [17], [16].
We use the method developed earlier to compute the analyt-
ical solution of three benchmark problems solvable by hand,
involving solutions with shocks and kinks (a detailed descrip-
tion of the benchmark examples is available in [4]). For each
of the numerical schemes used, we compute the error due
to the discretization method, as a function of the number of
grid points. The result is shown in Fig. 2. This study leads to
sev- eral conclusions. The Lax–Friedrichs scheme is very
diffusive. Its behavior is representative of linear schemes to
approximate a hyperbolic PDE. Consequently, we do not think
that it is a good idea to use such linear numerical schemes,
even if it would have the advantage of making the constraints
linear in the resulting optimization program. The left centered
scheme is less diffu- sive, but fails to capture the kinks of the
solution. However, it still provides good convergence. The
JST scheme captures shocks accurately because of its anti-
diffusive term, and thus, gives the best results overall. It will
be used for the rest of this study. Additionally, the JST
scheme has the benefit that we can use it both for the direct
problem, and for the adjoint. A detailed study of the
computational time required to solve this class of problems is
out of the scope of this study. For this, we refer the reader to
our ongoing work [33], in which we compare the fol- lowing
three models: the original Menon model [24], the present
model, and a new cell-based model [31].

III. SELECTION OF MODEL PARAMETERS

A. System Identification: Main Velocity Profiles


In this section, we identify the mean velocity profiles
on each link. We use enhanced traffic management system
(ETMS) data, which we can obtain from NASA Ames (see [9]
for a de- scription of ETMS data). From ETMS data, we can
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 811
FLOW

Fig. 3. Example of velocity profile used for the junction LGA-ORD. The
hori- zontal coordinate is the distance from ORD in nm. The
corresponding links are shown, as well as the location of the airspace
fixes between the links.

flight information at a 3 min rate:2 position of each aircraft


in the NAS, altitude, velocity, and flight plan (i.e., set of
airways and waypoints). From this data, we are able to
identify the routes in which traffic is concentrated. Note
that in recent work, Menon et al. [24] focused on creating
an automated tool which performs similar tasks
automatically at a NAS-wide level, using FACET [9], a
tool developed by NASA Ames.
We analyzed 24 hours of ETMS data and selected all
aircraft using the links of the network shown in Fig. 1. We
identified all aircraft which used each of the links, and
recorded all tracks and corresponding speeds between
takeoff and landing. For each of the links shown in Fig. 1,
we identified the mean velocity pro- files as piecewise
affine functions, using a least squares fit. The total number
of aircraft used is 220. The result for the flight New York–
Chicago is displayed in Fig. 3. The curve is a piecewise
affine fit obtained using least squares. As can be seen, once
the En Route altitude is reached, the curve fits are almost
flat, which means that the aircraft are En Route at a high
altitude cruise speed. It can also be seen from Fig. 3 that
the data is relatively broadly spread (standard deviation
19.6 kn). This suggests a re- finement using multilayer
models: dividing the link in sublayers corresponding to
altitudes (with different speed profiles) has the benefit of
being more precise. In this work, we consider a single
layer.

B. Initial and Boundary Conditions


Once the mean velocity profiles are computed, we
identify the initial density of aircraft and the inflow
(boundary conditions) in the network. The initial position
of the aircraft is easy to extract from the ETMS data: at the
prescribed time, all airborne aircraft which are on the
relevant links are selected.
1) For any selected aircraft , at location , on link ,
the classical density is taken to be a
“box” around , of length , where is a
reference length
2Current ETMS data can now be obtained at a higher rate, which was

not available at the time this work was performed.


812 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

each value

Fig. 4. Different predictions obtained by the use of and r for aircraft density.
Above: density propagation through the PDE system (6); below: position update
from ETMS data and from the PDE.

relevant for the scale of the problem. Calling the char-


acteristic function of an interval (equal to 0 outside of
and 1 inside), is .
Taking all aircraft initially airborne on link , the density
is

2) Similarly, the density-like function is computed using


the knowledge of the mean velocity profile along link ,
called , and the parameter

These two equations thus, represent the initial conditions for


the density and the density-like functions, which we extract
from ETMS data.
The inflows (boundary conditions) can also be extracted
from ETMS data: each time an aircraft takes off, it will appear
in the ETMS data as soon as it is airborne. The ETMS data
also shows the filed flight plan, which we select when it
intends to use the links of interest to us. is computed the
following way. At any instant when the data shows a new
aircraft on one of the source links , the track is in general
passed the entrance point of that link (because of the sampling
rate of 3 min). Calling the position of this aircraft on link
at the first time it appears, we compute the time at which it
crossed the location (using the knowledge of the mean
velocity profile on the link). We then use one of the two
definitions above to compute corresponding to either
or .

C. Identifying the Numerical Parameters


As explained in Section III-C, we have two ways of de-
scribing the density of aircraft in the network, in terms of
a density function and a “density-like” function , which,
respectively, account for spatial and temporal distribution of
aircraft. The function depends on the numerical parameter
, which we need to adjust. The value of this parameter is
crucial for predicting aircraft count: Fig. 4 shows how errors
can occur in translating density functions into aircraft count.
We want to determine the choice of parameters leading to the
smallest error in aircraft count prediction.
We first run the following set of experiments. For the link
New York–Chicago, we run a set of simulations involving
aircraft, where successively takes all values
between 1 and 50. We vary between 0 and 120 nm. For
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 813
of
FLOW and , we run 400 experiments. Each
experi- ment corresponds to a uniformly distributed
random density of aircraft along link 1 in
(see Fig. 1). The simu- lation starts at a time , with the
density computed as in the previous section, and
computes the solution of the LWR PDE until the time
. For the experiments, was chosen equal to 1 hr
(note that the duration of the total flight is on the order of
two and a half hours). This solution is compared with the
solution obtained by propagating the Lagrangian trajecto-
ries of each of the aircraft independently from to
and computing the resulting density. In mathematical
terms, we compare the two following quantities:
• computed by the LWR PDE [6];

,
where is the position of aircraft at time
.
In order to characterize the best choice of numerical
param- eters, we compute the following two quantities
(notations refer to Fig. 1):
• relative density error, defined by

This quantity represents the error in density


prediction due to the propagation of by the PDE;
• absolute aircraft count error, defined by

where means number. This quantity is the sum of


count error for all sublinks of links 1, 4, and 5.
Typically, a link is divided into sublinks which
correspond to different airspace sectors. For example,
if link 1 goes through 8 sec- tors, we divide it in 8
sublinks and are interested in the aircraft counts on
these sublinks. This error thus estimates the
difference between the number of aircraft predicted
by the PDE and the number obtained by a Lagrangian
prop- agation of aircraft, where the error is the sum of
all errors on the sublinks.
The computation of both quantities is illustrated in Fig.
5. In this figure, for each of these sublinks, we compare the
number of aircraft predicted by the method (depicted by
arrows, which are computed from the density) with the
number of aircraft obtained by a Lagrangian propagation of
the trajectories. The error is the sum of errors for all
sublinks, i.e., the sum of the errors in sector counts. The
relative density error and absolute aircraft count error are
averaged (over the 400 runs) and plotted for the range of
and considered. The result is shown in Fig. 6. The left
plot shows the relative density error. As expected, the
error decreases when the number of aircraft increases and
increases. The right plot shows the absolute aircraft
count error, averaged over 400 simulations. For this plot,
each of the links 1, 4, and 5 have been divided in sublinks
(20 total), of about 50 nm length. This is a worst case
scneario, i.e., the number of relevant sectors for a flight of
this length is never higher. One can see
814 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

Fig. 5. (a) Illustration of the computation of the relative density error depicted in Fig. 6. The difference between the two density curves (shaded area) is divided
by the area below the curve. (b) Illustration of the computation of the error in aircraft count. The link is divided into sublinks (which can correspond to
sectors).

Fig. 6. (a) Relative density error between the density predicted by the Eulerian PDE propagation of the density. (b) Absolute aircraft count error for the junction
New York-Chicago.

that for and , the average aircraft the method. We use a 6-hr ETMS data set. From this data set,
count error is always extremely small. we extract the position of the aircraft, at the initial time, con-
The best choice for is thus obtained at the intersection struct the corresponding initial aircraft density, and propagate
of the lowest level sets of both plots of Fig. 6, i.e., for a range it through the PDE system. At any given time, we compare the
of and . Fig. 6 can also be in- aircraft count predicted by our method and the aircraft count
terpreted as follows. The region in the top-right corner and the provided by the ETMS data (which is exact, since it provides
bottom-left corner are both regions in which the model might the position of each aircraft). We compute the error in aircraft
not be applicable. As can be seen, the relative error or abso- count for a set of ten sublinks for the network shown in Fig. 1.
lute count error exceeds values that might be realistic for prac- The result is shown in Fig. 7(a). The window width was
tical purposes (15% error and absolute aircraft count error of taken equal to 15 nm. One can see on the left plot that the total
7). These regions are to be avoided. error (for all airborne aircraft in this airspace) is relatively low
(the maximal error is 7 aircraft). In fact, the results are much
IV. VALIDATION OF THE MODEL
better than they seem: most of the errors come from the fact
In the previous section, we have shown that the use of the that the aircraft distribution is such that there is always at least
modified LWR PDE either with (with any ) or one or two aircraft close to a sublink boundary, which will
(with an appropriate choice of ) enables accurate aircraft thus be counted in the wrong sublink. In fact, this is not really
count predictions. In this section, we validate the model a problem, as it is more an artifact of the computation rather
against real data. than a true error (Fig. 11 illustrates that the density unambigu-
ously shows where the aircraft is). Furthermore, some of the
A. Static Validation errors in aircraft count are due to errors present in the ETMS
In the first experiment, we use the static velocity profiles data (some have clearly erroneous data; this fact has also been
determined in the previous sections for the validation of reported in [11]).
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 815
FLOW

Fig. 7. (a) Error in aircraft count for the static validation over a five hour period. (b) Error in aircraft count for the dynamic validation over a 5-hr period.

B. Dynamic Validation tice by using a barrier function as commonly done in optimiza-


We extend the validation to a case in which the velocity 3http://cherokee.stanford.edu/~bayen/TCST06.html.

pro- files are time dependent, i.e., . The details of the


identifi- cation of these profiles are technical and are not
explained here. The comparison is the same as in the static
case. The results are shown in Fig. 7 and are more accurate
than the static results, as expected. The same remarks apply,
and the results are again af- fected by the quality of ETMS
data and the inclusion of the com- putation artifact. The only
weakness of this validation is that the simulation is run using
data from the same day as the data used in identification. A
way to improve this would be to perform the velocity
identification with data of a given day over a 24–hr pe- riod,
and validate it over the next 24–hr period, using the fact that
there is periodicity in the traffic for normal days. This was not
done here due to lack of available data. An animation (.avi
movie file) corresponding to the snapshots of Fig. 7 is
available.3
In both cases, the validation is very encouraging and shows
strong predictive capability for our model. The model was
also tested successfully using data from the western states
(Oakland Center with traffic incoming into Bay Area
airports), though for brevity these results are not included
here.

V. NETWORK CONTROL VIA ADJOINT METHOD


Consider solving the following problem: maximize the
throughput (i.e., flux of landing aircraft) at a destination
airport, while maintaining the density of aircraft everywhere
lower than a given threshold. Let us call the maximal
allowed den- sity on link and the
maximal and minimal achievable speedsonlink
(whichcandependonlocation). Using the notations of Section
II-B, the optimization problem thus reads

(7)

The difficulty posed by the constraints can be avoided in prac-


816 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006
tion [10], in which the cost is augmented by a logarithmic
term, which prohibits violation of the constraints.

(8)

We call the augmented cost function. When


, and are used without indices, it means that they are
vectors, i.e., . Note that the two last
constraints in the optimization program (7) have
disappeared into the cost func- tion. This constrained
optimization problem is easier to solve in practice. It is
asymptotically equivalent to the problem of interest when
. We use an adjoint method to alge- braically
compute the gradient of the cost function. This method was
extensively used [8] in flow control. We now adapt the
adjoint method to the case in which we have a set of PDEs
coupled through the boundary conditions, and subject to
con- straints. The adjoint method computes the gradient of
the cost function when is an implicit function
of and via the dynamics (6). Let us denote the cost
function of the two variables and
,
where is the solution of the PDE system (6). We
compute the linearized (6), which we will use to compute
the gradient of the cost function in the optimization
program (8). We denote by the linearized quantities
around a nominal value denoted by
. We call the linearized LWR
operator and . In order to abbreviate the notation, we
will write and . We omit the time and space
dependence when they are obvious. The linearized (6) reads

(9)
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 817
FLOW

The first variation of is obtained from (8)

Fig. 8. Network model.

us to enforce two other conditions for in order to cancel all


An integration by parts leads to the following identity for any the terms containing . We can choose
two functions and

(12)

These conditions have been chosen by necessity of the


algebraic derivation, in order to cancel appropriate terms in
the perturba- tion of the cost function. After some algebra,
which can be rewritten using the standard inner product using (10)–(12), we are able to express the first variation of
denoted for the domain as a function of the first variations control variables only (
and ), as well as nominal and adjoint quantities, which we
can evaluate. The result reads

(10)

where

We will denote by the standard inner product in


. is called the adjoint operator of . In order to express where again denotes the inner product for the domain
the first variation of as a function of the and only, we and for . The functions and
choose an adjoint density field that cancels all the terms generated by this method might be ill-behaved and, thus
con- taining in the cost function. First, in order to be inappropriate for practical air traffic control applications. We
eliminate the can alleviate this difficulty by projecting the descent direction
term , into
we choose such that a vector space of appropriate functions, for example the set
of continuous functions with bounded derivative, or the set of
continuous piecewise affine functions.
(11)
VI. CONTROLLER DESIGN
This is a first-order linear hyperbolic PDE, which is well In this section, we demonstrate the effectiveness of the
posed if is known and both the boundary conditions at one adjoint method by applying it to the air traffic model. Fig. 8
location and the initial conditions at one time are specified. shows the area which we will control (enclosed by a box). The
This allows
inflows into the box are thus now and as shown in Fig.
8. We want to impose the following constraint: for all links,
the density should be below a threshold which we
818 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006
impose. We allow the flow
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 819
FLOW

Fig. 9. Top left: Decrease of the cost as a function of the iterations for the three scenarios. The increases in M are clearly visible (steps), while the gradient
descent is more subtle. Congested traffic (solid); heavy traffic (- -); normal traffic ( ). Top middle: Decrease of the true cost as a function of the iterations.
The true cost is the cost J without the barrier terms. The method does not guarantee the monotonicity of the decrease but only the convergence. Top right:
Evolution of the parameter as a function of time. Bottom: Evolution of the velocity fields as a function of time for the different links. Each of the plots
corresponds to a link, (see top-left corner). The axis of each subplot are: x (arclength along the link), t (time) and v (x ; t), the velocity distribution.

to be split into a new link (link 6), in order to aid satisfaction Scenario 3: Congested Network. We generate data with
of the maximal density constraints. We call the very high densities of aircraft. This situation does not use
corresponding split factor: is the fraction of the flow ETMS data; it is generated randomly.
which stays on link 1 (called 1 bis); is the fraction Fig. 9 shows the decrease in cost for the three scenarios as
which is routed through link 6. This new link might use a function of the total number of iterations (i.e., iterations on
another arrival into the airport (it enters the arrival airspace and gradient advances). As can be seen in this figure, the
from another direction).4 We simulate the following three more congested the situation is, the higher the cost. The evolu-
scenarios. tion of the cost with iterations exhibits two distinct behaviors,
Scenario 1: Normal Traffic. (Real data) We take ETMS as often with barrier methods [10]: large jumps corresponding
data, from which we extract initial conditions and inflows. We to the increases in , and shallower decreases corresponding
impose a restriction on the density and control the flow. to the gradient advances. Convergence is clearly observed for
Scenario 2: Heavy Traffic. (Modified real data) We take the three scenarios. We display some of the results for the
the same data as for the previous case, and add additional third case. An animation (in form of an .avi movie file)
aircraft in order to more heavily overload the network. corresponding to each of the three scenarios is available.3 We
now describe in detail the scenario corresponding to Case 3.
4Note that using is equivalent to using turning proportions in road traffic We run a one-hour simulation. Fig. 9 shows the aircraft
and might not be the best way to represent network traffic. It could be better density on all links at var- ious instants, in the absence of
to define an assignment proportion, i.e., a coefficient indexed by destination.
This might be implemented in the future (as a part of the control strategy), control: the velocity is the mean velocity profile determined
using a framework such as the one developed by Papageorgiou [25]. for each link, and no aircraft is al-
820 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

Fig. 10. Top 6 subfigures: Evolution of the aircraft density on the different links in the absence of control. Each of the subplot shows the density distribution at
a given time on the corresponding link as in Fig. 9 (the horizontal coordinate represents location, the vertical represents density). The horizontal line represents
the density threshold (all quantities are nondimensionalized by , so that the threshold density is 1). As can be seen, the density threshold is violated in link
5 at t = 27;t = 39 and t = 59. Bottom 6 subfigures: Evolution of the aircraft density with control applied. Note that link 6 is now open and used. This prevents
the second violation of density threshold observed in Fig. 10 (t = 59): some of the flow is directly routed from link 1 to link 6. The first violation seen in the
top 6 subfigures is avoided by speed changes. This figure is also available in form of a .avi file.3

lowed into link 6 (i.e., ). The initial density is shown The velocity profiles are shown in Fig. 9. Each of
in the top-left corner. The inflow into links 1 and 3 is such that the subplots corresponds to one of the links. For links 5 and 6,
at time , the density threshold (represented by the hor- one can clearly see the descent velocity profiles. Also, for link
izontal line on each subplot) is violated until time . At 6 (subfigure below), one can see a ridge. It corresponds to a
time , it is violated again, until the end of the set of aircraft which have to fly at high speed into the airport.
experiments. Fig. 9 shows the same experiment when link 6 is One can also see similar ridges on the other subplots, which
now opened to traffic, and velocity control is enabled. As can have the same interpretation. For any ridge, the Controller
be seen, about half of the traffic incoming into link 1 is command could be to the corresponding set of aircraft: “fly
rerouted into link 6, and the other half into link 1 bis. Fig. 9 direct at 420 kn direct into [the next waypoint].” Note that in
shows the variation of with time. As can be seen, around the absence of con- trol, the first violation of the aircraft
min, there is a peak of about 25% of aircraft routed into density threshold occurs 33 min after the beginning of the
link 6, which settles to 50% at . The routing control experiment, almost at the end of the network, which is not
enables avoidance of violation of maximal density shown in intuitive. This shows the efficiency of the method, which is
Fig. 10. The first violation is avoided by velocity changes. capable of generating the right routing and speed assignments
to prevent undesirable events from hap-
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 821
FLOW

pening much later. Finally, the simulations are also depicted


on a U.S. map in Fig. 12. One can see that before , all
aircraft choose the direct route through link 5 to Chicago (it is
shorter). After , the excessive amount of flow incoming
into links 1 and 3 forces the flow to be split through links 1 bis
and 6.
The expression of the cost function can be replaced by
any arbitrary user-defined cost as long as the integrand is
smooth. The goal of this paper was to prove the feasibility of
the tech- nique (with a particular cost function), but extending
this to any cost function is a straightforward process (the only
thing which is needed is to recompute the expression of the
gradient based on the new cost, following the steps outlined
here). In particular, in the work of [27], the authors use an
integral form with quadratic penalty. This can be interpreted
as penalizing the cumulative delay minutes at each point in
time, and penalizing more se- verely large deviations from the
scheduled flow than small ones.

VII. CONCLUSION
We have derived an Eulerian model of the airspace based on
a modified LWR partial-differential equation. The network
struc- ture of the airspace was modeled as a set of coupled
LWR PDEs. Given initial positions of aircraft and airport
inflows, this system of PDEs enables the prediction of the
aircraft density. An ana- lytical solution was derived for a
single link in the case, in which the mean velocity profiles of
aircraft along airways do not vary with time (just with space).
It can be used for multiple links as well. ETMS data was used
to identify the numerical parame- ters associated with this
model. The data is also used to validate the model, i.e., to
demonstrate good predictive capability of this method.
We first have shown how to use efficient numerical
schemes to simulate the network. We have discarded linear
numerical schemes because of their poor performance. We
have used the Jameson–Schmidt–Turkel scheme as our main
numerical scheme to perform numerical simulations of the
network. We have posed the problem of maximizing
throughput at a destination airport while maintaining the
aircraft density below a certain threshold as an optimization
program. The inequality constraints of this program have been
handled using a log-barrier method. The adjoint problem was
derived and used to compute the gradient of the augmented
cost function. The resulting optimization and control schemes
were applied to a real air traffic case. Simulations show that
this method enables automated control of realistic scenarios as
well as highly congested situations. The output of the code is a
set of time dependent velocity profiles to apply to the network,
and a policy telling how to split the flow in areas of diverging
traffic. These outputs could, thus, be used by the Traffic
Management Unit in charge of managing the flow: they
provide high-level policies to apply to the aircraft streams,
which are directly understandable by human controllers.
Finally, this formulation of the air traffic flow control
Fig. 11. Display of the traffic situation for the static validation. The density
of the links is depicted by the color. The colored rectangles shown in this plot problem as an optimization program of PDEs allows for many
repre- sent the density. The color scale is: white for zero density; black for refine- ments in the control procedure. For instance, gradient
highest den- sity. The actual aircraft positions are superimposed (triangles). descent may be replaced by more sophisticated optimization
Traffic is shown at t (top), t + 8 min;t + 16 min, etc. As can be seen, the
peaks of density corresponds to the actual positions of the aircraft. methods such as approximate Newton method [29] in order to
ensure real time convergence of the algorithm. Furthermore,
822 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006
using this model, a decentralized control policy can also
be derived using
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 823
FLOW

Fig. 12. Aircraft density in the network around Chicago in presence of control and velocity assignment. The density of the links is depicted by the color. The
colored rectangles shown in this plot represent the density. The color scale is: white for zero density; black for highest density. As can be seen and was shown
in Fig. 10, a good portion of the flow is routed into link 6 starting at time t = 51.

decomposition techniques in order to allow different airlines that it is continuous. This solution has been constructed using
to separately optimize the flow of their aircraft, while main- a technique analogous to the algorithm of Bayen and Tomlin
taining safety criteria [28]. This method has since been [7] based on the method of characteristics.
applied to highway traffic as well [5], [18] and looks promising Uniqueness: Let us call and two continuous weak
for other applications involving networks of partial differential solu- tions of (1). Call . satisfies:
equations. a.e. in in and
in . Multiplying this PDE by and integrating
VIII. PROOF OF PROPOSITION 1 from to the first discontinuity of gives

Existence: is well defined because for all


. Its inverse exists because is (strictly) increasing. It
is easy to check that (2) satisfies (1) almost everywhere, and
824 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

from which we deduce [6] A. M. Bayen, R. Raffard, and C. J. Tomlin, “Network congestion
alleviation using adjoint hybrid control: Application to highways,”
in Number 2993 in Lecture Notes in Computer Science. New York:
Springer-Verlag, 2004, pp. 95–110.
[7] A. M. Bayen and C. J. Tomlin, “A construction procedure using
char- acteristics for viscosity solutions of the Hamilton-Jacobi
equation,” in Proc. 40th IEEE Conf. Dec. Contr., 2001, pp. 1657–
Integrating by parts gives 1662.
[8] T. R. Bewley, “Flow control: New challenges for a new renaissance,”
Prog. Aerosp. Sci., vol. 37, no. 1, pp. 1–119, Jan. 2001.
[9] K. Bilimoria, B. Sridhar, G. Chatterji, K. Seth, and S. Graabe,
“FACET: Future ATM concepts evaluation tool,” in Proc. 3rd
USA/Eur. Air Traffic Manage. R&D Seminar, 2001, on CDROM.
[10] S. Boyd and L. Vandenberghe, Convex Optimization. Cambridge,
U.K.: Cambridge Univ. Press, 2004.
[11] G. Chatterji, B. Sridhar, and D. Kim, “Analysis of ETMS data qualiy
for traffic flow management decisions,” in Proc. AIAA Conf. Guid.,
Nav. Contr., 2003, AIAA-2003-5626.
[12] C. Chen, Z. Jia, and P. Varaiya, “Causes and cures of highway
congestion,” IEEE Control Syst. Mag., vol. 21, no. 4, pp. 26–33,
Dec. 2001.
since and . Using the fact that [13] P. D. Christofides, Nonlinear and Robust Control of Partial
Differential Equation Systems: Methods and Applications to
for all . Then, Transport-Reaction Processes. Cambridge, MA: Birkhäuser, 2001.
we use the fact that [14] C. Daganzo, “The cell transmission model: A dynamic representation
of highway traffic consistent with the hydrodynamic theory,” Trans-
port. Res., vol. 28B, no. 4, pp. 269–287, Aug. 1994.
[15] C. Daganzo, “The cell transmission model, Part II: Network traffic,”
Transport. Res., vol. 29B, no. 2, pp. 79–93, Apr. 1995.
[16] C. Daganzo, “A finite difference approximation of the kinematic
wave model of traffic flow,” Transport. Res., vol. 29B, no. 4, pp.
261–276,
so that we can rewrite the inequality as of air traffic flow in congested areas,” in Proc. Amer. Contr. Conf.,
2004,
pp. 5520–5526.

Using Gronwall's lemma, this last inequality implies


almost everywhere in . By continuity,
everywhere in , and, therefore, at . The same
proof applies to since for all .
By induction on and are equal everywhere in
and, therefore, in .

ACKNOWLEDGMENT
The authors would like to thank Dr. P. K. Menon and Dr.
K. Bilimoria for conversations which inspired this work,
Dr. G. Chatterji for his ongoing support and suggestions
which went into modelling this work, and Dr. G. Meyer for
his sup- port in this project. They also thank Prof. T. Bewley
for useful conversations regarding the application of the
adjoint method to flow control, and his help in the original
formulation of the control problem. Prof. T.-P. Liu helped
define the PDE used for this model.

REFERENCES
[1] O. M. Aamo and M. Krstic, Flow Control by Feedback. New York:
Springer-Verlag, 2002.
[2] R. Ansorge, “What does the entropy condition mean in traffic flow
theory?,” Transp. Res., vol. 24B, no. 2, pp. 133–143, Apr. 1990.
[3] B. Bamieh, F. Paganini, and M. A. Daleh, “Distributed control of
spa- tially-invariant systems,” IEEE Trans. Autom. Control, vol. 47,
no. 7,
pp. 1091–1107, Jul. 2002.
[4] A. M. Bayen, “Computational control of networks of dynamical sys-
tems: Application to the National Airspace System,” Ph.D. disserta-
tion, Dept. Aeronautics and Astronautics, Stanford Univ., Stanford,
CA, 2004.
[5] A. M. Bayen, R. Raffard, and C. J. Tomlin, “Eulerian network model
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 825
FLOWAug. 1995.
[17] C. Hirsch, Numerical Computation of Internal and External
Flows. New York: Wiley, 1988.
[18] D. Jacquet, C. Canudas de Wit, and D. Koenig, “Optimal ramp
metering strategy with an extended LWR model: Analysis and
computational methods,” in Proc. 16th IFAC World Congr., 2005,
to be published.
[19] A. Jameson, “Aerodynamic design via control theory,” J. Scientific
Comput., vol. 3, no. 3, pp. 233–260, Sep. 1988.
[20] A. Jameson, “Analysis and design of numerical schemes for gas dy-
namics 1: Artificial diffusion, upwind biasing, limiters and their
effect on accuracy and multigrid convergence,” Int. J.
Computational Fluid Dyn., vol. 4, pp. 171–218, Sep. 1995.
[21] M. Krstic, “On global stabilization of Burgers' equation by
boundary control,” Syst. Control Lett., vol. 37, pp. 123–142, Jul.
1999.
[22] M. J. Lighthill and G. B. Whitham, “On kinematic waves. II. A
theory of traffic flow on long crowded roads,” in Proc. Royal Soc.
London, 1956, pp. 317–345.
[23] X. Litrico, “Robust IMC flow control of SIMO dam-river open-
channel systems,” IEEE Trans. Control Syst. Technol., vol. 10, no.
5, pp. 432–437, May 2002.
[24] P. K. Menon, G. D. Sweriduk, and K. Bilimoria, “New approach for
modeling, analysis, and control of air traffic flow,” AIAA J. Guid.,
Contr., Dyn., vol. 27, no. 5, pp. 731–5090, Sep./Oct. 2004.
[25] A. Messmer and M. Papageorgiou, “Route diversion control in mo-
torway networks via nonlinear optimization,” IEEE Trans. Control
Syst. Technol., vol. 3, no. 1, pp. 144–154, Mar. 1995.
[26] L. Munoz, X. Sun, R. Horowitz, and L. Alvarez, “Traffic density
esti- mation with the cell transmission model,” in Proc. Amer.
Contr. Conf., 2003, pp. 3750–3755.
[27] R. Raffard, S. L. Waslander, A. M. Bayen, and C. J. Tomlin,
“Toward efficient and equitable distributed air traffic flow control,”
presented at the Proc. Amer. Contr. Conf., 2006, Minneapolis, MN.
[28] R. Raffard, S. L. Waslander, A. M. Bayen, and C. J. Tomlin, “A co-
operative, distributed approach to multi-agent Eulerian network
con- trol: Application to air traffic management,” in Proc. AIAA
Guid., Nav. Contr. Conf. Exhibit, 2005, AIAA-2005-6050.
[29] R. L. Raffard and C. J. Tomlin, “Second order optimization of
ordinary and partial differential equations with applications to air
traffic flow,” in Proc. Amer. Contr. Conf., 2005, pp. 798–803.
[30] P. I. Richards, “Shock waves on the highway,” Oper. Res., vol. 4,
no. 1, pp. 42–51, Feb. 1956.
[31] C. Robelin, D. Sun, G. Wu, and A. M. Bayen, “En-route air traffic
modeling and strategic flow management using mixed integer linear
programming,” in INFORMS Annu. Meeting, 2005.
[32] R. C. Smith and M. A. Demetriou, Research Directions in
Distributed Parameter Systems. Philadelphia, PA: SIAM, 2000.
826 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 14, NO. 5, SEPTEMBER 2006

[33] D. Sun, S. Yang, I. S. Strub, B. Sridhar, K. Sheth, and A. M. Bayen,


Robin L. Raffard received the M.S. degree in me-
“Assessment of the respective performance of three Eulerian air
chanical engineering from the Ecole Centrale Paris,
traffic flow models,” presented at the Proc. AIAA Conf. Guid., Nav.
France, in 2002, and the M.S. degree in aeronautics
Contr., Keystone, CO, 2006.
and astronautics from Stanford University,
[34] Y. Wang, M. Papageorgiou, and A. Messmer, “Motorway traffic
Stanford, CA, in 2003. He is currently working
state estimation based on extended Kalman filter,” in Proc. Euro.
towards the Ph.D. degree in aeronautics and
Contr. Conf., 2003. astronautics at the same school.
His research interests include distributed opti-
mization and optimal control of systems governed
by differential equations, with applications in air
traffic flow, systems biology, and stochastic systems.

Claire J. Tomlin (S'93–M'99) received the Ph.D.


degree in electrical engineering from the University
of California at Berkeley, Berkeley, in 1998. She
also received the M.Sc. degree from Imperial
College, London, in 1993, and the B.A.Sc. degree
Alexandre M. Bayen (S'02–M'04) received the from the University of Waterloo, Canada, in 1992,
B.S. degree in applied mathematics from the Ecole both in electrical engineering.
Poly- technique, Palaiseau, France, in 1998, the M She is an Associate Professor in the Department
.S. and Ph.D. degrees in aeronautics and of Electrical Engineering and Computer Sciences
astronautics from Stanford University, Stanford, at the University of California at Berkeley, and is
CA, in 1999 and 2004, respectively. an As- sociate Professor in the Department of
He was a Visiting Researcher at NASA Ames, Aeronautics
Moffett Field, CA, from 2001 to 2003. From 2004 and Astronautics at Stanford University, Stanford, CA, where she also holds
to 2005, he worked for the Department of Defense, the Vance D. and Arlene C. Coffman Faculty Scholarship in the School of
France, where he held the rank of Major. During Engi- neering. She joined Stanford in September 1998, as a Terman Assistant
that time, he was the Research Director of the Pro- fessor, and received tenure at Stanford in November 2004. In July 2005,
Labora- she joined Berkeley as an Associate Professor. She has held visiting research
toire de Navigation Autonome at the Laboratoire de Recherches Balistiques et po- sitions at NASA Ames, Honeywell Labs, and the University of British
Aérodynamiques, Vernon, France. Since March 2005, he has been an Co- lumbia. Her research interests include control systems, specifically
Assistant Professor in the Department of Civil and Environmental hybrid con- trol theory, and she works on air traffic control automation, flight
Engineering at the University of California at Berkeley, Berkeley. His management system analysis and design, and modeling and analysis of
research interests include control of distributed parameter systems, biological cell net- works.
combinatorial optimization, hybrid systems, and air traffic automation. Dr. Tomlin is a recipient of the Eckman Award of the American Automatic
Dr. Bayen is a recipient of the Graduate Fellowship of the Délégation Control Council (2003), MIT Technology Review's Top 100 Young
Générale pour l'Armement (1998–2002) from France, and the Ballhaus Prize Innovators Award (2003), the AIAA Outstanding Teacher Award (2001), an
for best doctoral thesis from the Department of Aeronautics and Astronautics National Sci- ence Foundation (NSF) Career Award (1999), and the Bernard
at Stanford University (2004). Friedman Memo- rial Prize in Applied Mathematics (1998).
BAYEN et al.: ADJOINT-BASED CONTROL OF A NEW EULERIAN NETWORK MODEL OF AIR TRAFFIC 827
FLOW

View publication stats

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy