0% found this document useful (0 votes)
21 views219 pages

ICAT-2019-04 Mayara Data-Driven Modeling

This report presents a data-driven approach to modeling air traffic flows to enhance air traffic management (ATM) systems, addressing challenges such as congestion and delays. A framework is proposed for analyzing large-scale flight tracking data, which facilitates the identification of traffic patterns and supports decision-making for airport capacity planning. The findings demonstrate significant operational differences among major metroplexes, leading to improved predictions and optimization of airport acceptance rates, resulting in reduced delays.

Uploaded by

2258420069
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
21 views219 pages

ICAT-2019-04 Mayara Data-Driven Modeling

This report presents a data-driven approach to modeling air traffic flows to enhance air traffic management (ATM) systems, addressing challenges such as congestion and delays. A framework is proposed for analyzing large-scale flight tracking data, which facilitates the identification of traffic patterns and supports decision-making for airport capacity planning. The findings demonstrate significant operational differences among major metroplexes, leading to improved predictions and optimization of airport acceptance rates, resulting in reduced delays.

Uploaded by

2258420069
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 219

DATA-DRIVEN MODELING OF AIR TRAFFIC FLOWS FOR

ADVANCED AIR TRAFFIC MANAGEMENT

Mayara Condé Rocha Murça and R. John Hansman

This report is based on the Doctoral Dissertation of Mayara Condé Rocha Murça
submitted to the Department of Aeronautics and Astronautics in partial fulfillment
of the requirements for the degree of Doctor of Philosophy at the
Massachusetts Institute of Technology.

The work presented in this report was also conducted


in collaboration with the members of the Doctoral Committee:

Prof. R. John Hansman (Chair)


Prof. Hamsa Balakrishnan
Dr. Tom G. Reynolds

Report No. ICAT-2019-04


April 2019

MIT International Center for Air Transportation (ICAT)


Department of Aeronautics & Astronautics
Massachusetts Institute of Technology
Cambridge, MA 02139 USA
2
Data-Driven Modeling of Air Traffic Flows for
Advanced Air Traffic Management
by
Mayara Condé Rocha Murça

Submitted to the Department of Aeronautics and Astronautics


on August 23, 2018, in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy in Aeronautics and Astronautics

Abstract
The Air Traffic Management (ATM) system enables air transportation by ensuring a safe and
orderly air traffic flow. As the air transport demand has grown, ATM has become increasingly
challenging, resulting in high levels of congestion, flight delays and environmental impacts.
To sustain the industry growth foreseen and enable more efficient air travel, it is important to
develop mechanisms for better understanding and predicting the air traffic flow behavior and
performance in order to assist human decision-makers to deliver improved airspace design and
traffic management solutions. This thesis presents a data-driven approach to modeling air
traffic flows and analyzes its contribution to supporting system level ATM decision-making.
A data analytics framework is proposed for high-fidelity characterization of air traffic flows
from large-scale flight tracking data. The framework incorporates a multi-layer clustering
analysis to extract spatiotemporal patterns in aircraft movement towards the identification
of trajectory patterns and traffic flow patterns. The outcomes and potential impacts of this
framework are demonstrated with a detailed characterization of terminal area traffic flows
in three representative multi-airport (metroplex) systems of the global air transportation
system: New York, Hong Kong and Sao Paulo.
As a descriptive tool for systematic analysis of the flow behavior, the framework allows
for cross-metroplex comparisons of terminal airspace design, utilization and traffic perfor-
mance. Novel quantitative metrics are created to summarize metroplex efficiency, capacity
and predictability. The results reveal several structural, operational and performance dif-
ferences between the metroplexes analyzed and highlight varied action areas to improve air
traffic operations at these systems.
Finally, the knowledge derived from flight trajectory data analytics is leveraged to de-
velop predictive and prescriptive models for metroplex configuration and capacity planning
decision support. Supervised learning methods are used to create prediction models capa-
ble of translating weather forecasts into probabilistic forecasts of the metroplex traffic flow
structure and airport capacity for strategic time horizons. To process these capacity fore-
casts and assist the design of traffic flow management strategies, a new optimization model
for capacity allocation is developed. The proposed models are found to outperform currently
used methods in predicting throughput performance at the New York airports. Moreover,

3
when used to prescribe optimal Airport Acceptance Rates in Ground Delay Programs, an
overall delay reduction of up to 9.7% is achieved.

Thesis Supervisor: R. John Hansman


Title: T. Wilson Professor of Aeronautics and Astronautics

4
Acknowledgments
The author would like to thank the support from the Aeronautics Institute of Technology
(ITA), the Brazilian Air Force (FAB), the Brazilian Federal Agency for Support and Eval-
uation of Graduate Education (CAPES) and the Latin American Scholarship Program of
American Universities (LASPAU). In addition, the author would like to thank the members
of the doctoral committee, Prof. John Hansman, Prof. Hamsa Balakrishnan and Dr. Tom
Reynolds, for all the guidance and feedback throughout the development of the research,
as well as the thesis readers, Dr. James Jones and Prof. Carlos Müller, for their technical
and editorial comments. The author also thanks all the collaborators, especially Richard
DeLaura and Dr. Lishuai Li, for providing data and valuable insights.

5
THIS PAGE INTENTIONALLY LEFT BLANK

6
Contents

1 Introduction 25
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.2 Thesis Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.3 Methodological Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.4 Background and Related Literature . . . . . . . . . . . . . . . . . . . . . . . 32
1.4.1 An Overview of Air Traffic Management . . . . . . . . . . . . . . . . 32
1.4.2 Challenges in Metroplex Air Traffic Management . . . . . . . . . . . 36
1.4.3 Optimization Models for Air Traffic Management . . . . . . . . . . . 39
1.4.4 Data Analytics for Air Traffic Management . . . . . . . . . . . . . . . 43
1.4.5 Data-Driven Modeling of Air Traffic Flows . . . . . . . . . . . . . . . 45
1.4.6 Data-Driven Modeling of Airport Capacity . . . . . . . . . . . . . . . 49
1.5 Contributions of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
1.6 Organization of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

2 Flight Trajectory Data Analytics Framework for Characterization of Air


Traffic Flows 57
2.1 Overview of Methodological Approach . . . . . . . . . . . . . . . . . . . . . 57
2.2 Dataset Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.3 Clustering at Spatial Scale: Trajectory Clustering . . . . . . . . . . . . . . . 59
2.3.1 Trajectory Representation . . . . . . . . . . . . . . . . . . . . . . . . 62
2.3.2 Trajectory Similarity Assessment . . . . . . . . . . . . . . . . . . . . 62
2.3.3 Trajectory Clustering Method . . . . . . . . . . . . . . . . . . . . . . 63
2.3.4 Example Application . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

7
2.4 Trajectory Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
2.4.1 Matching Flight Trajectories with Learned Trajectory Patterns . . . . 69
2.4.2 Identifying Non-Conforming Trajectories . . . . . . . . . . . . . . . . 72
2.4.3 Flow Identification . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
2.5 Clustering at Temporal Scale: Time-Dependent Flow Vector Clustering . . . 78

3 Comparative Analysis of Terminal Area Operations in Multi-Airport Sys-


tems 81
3.1 Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
3.1.1 New York Multi-Airport System . . . . . . . . . . . . . . . . . . . . . 84
3.1.2 Hong Kong Multi-Airport System . . . . . . . . . . . . . . . . . . . . 86
3.1.3 Sao Paulo Multi-Airport System . . . . . . . . . . . . . . . . . . . . . 87
3.2 Data Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.3 Characterization of Structural Differences through the Analysis of Terminal
Area Route Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
3.3.1 Identification of Trajectory Patterns . . . . . . . . . . . . . . . . . . 89
3.3.2 Identification of Route Intersections . . . . . . . . . . . . . . . . . . . 93
3.4 Characterization of Operational Differences through the Analysis of Traffic
Flow Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
3.4.1 Identification of Metroplex Flow Patterns . . . . . . . . . . . . . . . . 97
3.4.2 Identification of Metroplex Flow Interactions . . . . . . . . . . . . . . 112
3.5 Characterization of Performance Differences . . . . . . . . . . . . . . . . . . 115
3.5.1 Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
3.5.2 Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
3.5.3 Predictability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
3.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

4 Data-Driven Approach for Metroplex Configuration and Capacity Plan-


ning 133
4.1 Metroplex Configuration and Airport Capacity Planning Framework . . . . . 134
4.2 Data Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

8
4.2.1 Terminal Aerodrome Forecast (TAF) . . . . . . . . . . . . . . . . . . 135
4.2.2 Arrival Route Status and Impact (ARSI) Forecast . . . . . . . . . . . 136
4.2.3 Hourly Airport Reports . . . . . . . . . . . . . . . . . . . . . . . . . 136
4.2.4 Flight Tracks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
4.3 Prediction Models and Features . . . . . . . . . . . . . . . . . . . . . . . . . 137
4.4 Prediction of Traffic Flow Patterns . . . . . . . . . . . . . . . . . . . . . . . 140
4.4.1 Classification Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
4.4.2 Performance Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . 144
4.5 Prediction of Airport Arrival Capacity . . . . . . . . . . . . . . . . . . . . . 149
4.5.1 Regression Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4.5.2 Performance Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.6 Prescription of Airport Arrival Rates . . . . . . . . . . . . . . . . . . . . . . 164
4.6.1 Ground Delay Programs . . . . . . . . . . . . . . . . . . . . . . . . . 164
4.6.2 AAR Planning Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
4.6.3 Analysis of Impacts on GDP planning . . . . . . . . . . . . . . . . . 171

5 Conclusions and Future Work 179


5.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
5.2 Future Research Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

A Trajectory Clustering Analysis Results - New York Metroplex 185

B Trajectory Clustering Analysis Results - Hong Kong Metroplex 193

C Trajectory Clustering Analysis Results - Sao Paulo Metroplex 201

9
THIS PAGE INTENTIONALLY LEFT BLANK

10
List of Figures

1-1 Trajectories of arrival flights into JFK during two different time periods on
June 23, 2015 overlaid with the Standard Terminal Arrival Routes. . . . . . 28
1-2 General ways by which data analytics affects decision-making. . . . . . . . . 30
1-3 Ways by which the flight trajectory data analytics framework can impact ATM. 31
1-4 Airport/airspace capacity influencing factors. . . . . . . . . . . . . . . . . . . 35
1-5 ATM functional structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
1-6 Example spatiotemporal patterns in trajectory datasets [1]. . . . . . . . . . . 46
1-7 Example of typical runway configurations at JFK. . . . . . . . . . . . . . . . 50
1-8 Capacity curves for two different runway configurations at JFK (data from
2013-2015). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
1-9 View of thesis contributions in the context of the ATM decision-making process. 55

2-1 Schematic overview of flight trajectory data analytics framework for charac-
terization of air traffic flows. . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2-2 Resampled flight trajectory data. . . . . . . . . . . . . . . . . . . . . . . . . 63
2-3 Illustration of DBSCAN concepts. . . . . . . . . . . . . . . . . . . . . . . . . 65
2-4 Percentage of noise associated with the clustering output obtained with dif-
ferent parameter settings (M inP ts, "). . . . . . . . . . . . . . . . . . . . . . 67
2-5 Silhouette Index associated with the clustering output obtained with different
parameter settings (M inP ts, "). . . . . . . . . . . . . . . . . . . . . . . . . . 68
2-6 Silhouette Index versus number of clusters identified. . . . . . . . . . . . . . 69
2-7 (a) Clusters of arrival trajectories. (b) Silhouette plot. . . . . . . . . . . . . 70
2-8 Supervised learning classification process. . . . . . . . . . . . . . . . . . . . . 71

11
2-9 Example of dataset and resulting decision tree for classification. . . . . . . . 71
2-10 Flow matrix for LGA arrivals on May 24, 2013. . . . . . . . . . . . . . . . . 78

3-1 Location of airports in the New York, Hong Kong and Sao Paulo metroplexes. 82
3-2 Passenger movement by airport at the New York, Hong Kong and Sao Paulo
metroplexes in 2016. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3-3 Aircraft movement by airport at the New York, Hong Kong and Sao Paulo
metroplexes in 2016. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
3-4 Arrival and departure trajectories for one day of terminal area operations in
the New York, Hong Kong and Sao Paulo multi-airport systems. . . . . . . . 83
3-5 Shared airspace regions between JFK and LGA [2]. . . . . . . . . . . . . . . 84
3-6 Runway infrastructure in the New York metroplex. . . . . . . . . . . . . . . 85
3-7 Runway infrastructure in the Hong Kong metroplex. . . . . . . . . . . . . . . 87
3-8 Runway infrastructure in the Sao Paulo metroplex. . . . . . . . . . . . . . . 88
3-9 Clusters of arrival trajectories; each color represents one cluster; grey bar
represents the percentage of noise observations. . . . . . . . . . . . . . . . . 90
3-10 Centroids of metroplex arrival trajectory clusters. . . . . . . . . . . . . . . . 91
3-11 Centroids of metroplex departure trajectory clusters. . . . . . . . . . . . . . 92
3-12 Example of different trajectory patterns from arrival fix to runway threshold
22 at LGA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3-13 Characterization of trajectory tube intersections. . . . . . . . . . . . . . . . . 94
3-14 Example of intersection between JFK and LGA arrival trajectory tubes. . . . 96
3-15 Example of intersection between SZX and MFM arrival trajectory tubes. . . 97
3-16 Example of intersection between GRU and CGH departure trajectory tubes. 97
3-17 Cumulative percentage of observations for the clusters of hourly flow vectors
identified. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3-18 Most frequently observed New York MFPs. . . . . . . . . . . . . . . . . . . . 102
3-19 Convention on flow direction. . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3-20 New York MFP frequency of occurrence by time of day. . . . . . . . . . . . . 105
3-21 Daily New York demand patterns, based on data from years 2013-2015. . . . 105

12
3-22 Distribution of metroplex arrival and departure throughput for the New York
metroplex MFPs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
3-23 New York MFP frequency of occurrence under VMC and IMC. . . . . . . . . 107
3-24 Most frequently observed Hong Kong MFPs. . . . . . . . . . . . . . . . . . . 108
3-25 Hong Kong MFP frequency of occurrence by time of day. . . . . . . . . . . . 109
3-26 Hong Kong MFP frequency of occurrence under VMC and IMC. . . . . . . . 109
3-27 Most frequently observed Sao Paulo MFPs. . . . . . . . . . . . . . . . . . . . 110
3-28 Sao Paulo MFP frequency of occurrence by time of day. . . . . . . . . . . . . 111
3-29 Daily demand pattern at GRU, based on data of 2017. . . . . . . . . . . . . 112
3-30 Sao Paulo MFP frequency of occurrence under VMC and IMC. . . . . . . . . 112
3-31 Trajectory tube intersections associated with flow dependencies in the New
York metroplex. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
3-32 Example of path stretch calculation. . . . . . . . . . . . . . . . . . . . . . . 116
3-33 Structural path stretch for arrivals. . . . . . . . . . . . . . . . . . . . . . . . 117
3-34 Structural path stretch for arrivals by MFP. . . . . . . . . . . . . . . . . . . 117
3-35 Lateral and temporal traffic flow efficiency. . . . . . . . . . . . . . . . . . . . 119
3-36 Distribution of traffic flow efficiency by MFP. . . . . . . . . . . . . . . . . . 119
3-37 Average daily percentage of non-conforming arrival trajectories. . . . . . . . 120
3-38 Example non-conforming arrival trajectories identified for a convective weather
day at JFK, HKG and GRU. . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
3-39 New York metroplex arrival throughput as a function of the arrival demand
in the terminal area. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
3-40 Excess terminal area transit time as a function of the arrival demand. . . . . 122
3-41 Throughput-demand curves for the New York MFPs. . . . . . . . . . . . . . 124
3-42 Delay-demand curves for the New York MFPs. . . . . . . . . . . . . . . . . . 124
3-43 Throughput-demand curves for the Hong Kong MFPs. . . . . . . . . . . . . 125
3-44 Delay-demand curves for the Hong Kong MFPs. . . . . . . . . . . . . . . . . 125
3-45 Throughput-demand curves for the Sao Paulo MFPs. . . . . . . . . . . . . . 125
3-46 Delay-demand curves for the Sao Paulo MFPs. . . . . . . . . . . . . . . . . . 125
3-47 Metroplex arrival rates under persistent demand and associated level of delay. 126

13
4-1 Metroplex configuration and airport capacity planning framework. . . . . . . 135
4-2 Histogram of hourly airport arrival rates. . . . . . . . . . . . . . . . . . . . . 137
4-3 Forecasting procedure throughout the planning horizon. . . . . . . . . . . . . 140
4-4 Parameter sensitivity of RF model based on 5-fold cross validation perfor-
mance. (a) Impacts of the minimum size of leaf node on the classification
error. (b) Impacts of the number of predictors to sample at each split on the
classification error. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
4-5 Statistically derived crosswind thresholds by MFP. . . . . . . . . . . . . . . . 145
4-6 Statistically derived tailwind thresholds by MFP. . . . . . . . . . . . . . . . 145
4-7 Multi-way classification accuracy of NN, RF and SVM traffic flow pattern
prediction models for forecast horizons of 1 h to 6 h. . . . . . . . . . . . . . 146
4-8 Brier score of NN, RF and SVM traffic flow pattern prediction models for
forecast horizons of 1 h to 6 h. . . . . . . . . . . . . . . . . . . . . . . . . . . 147
4-9 Accuracy of traffic flow pattern forecasts for RF model and static model as a
function of the look-ahead time of the prediction. . . . . . . . . . . . . . . . 148
4-10 Classification confusion matrix for traffic flow pattern predictions for a forecast
horizon of 3 h. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4-11 Parameter sensitivity of RF regression model based on 5-fold cross validation
performance. (a) Impacts of the minimum size of leaf node on the prediction
error. (b) Impacts of the number of predictors to sample at each split on the
prediction error. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
4-12 NRMSE of BR, RF and GP capacity prediction models for forecast horizons
of 1 h to 6 h. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4-13 MAPE of BR, RF and GP capacity prediction models for forecast horizons of
1 h to 6 h. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4-14 Predictor importance estimates from RF model. . . . . . . . . . . . . . . . . 157
4-15 PC of BR, RF and GP capacity prediction models for 50%, 70% and 90%
confidence levels and forecast horizons of 1 h to 6 h. . . . . . . . . . . . . . . 158
4-16 NRMSE of alternative GP capacity prediction models for forecast horizons of
1 h to 6 h. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

14
4-17 MAPE of GP model capacity predictions and baseline capacity estimates for
JFK and forecast horizons of 1 h to 6 h. . . . . . . . . . . . . . . . . . . . . 161
4-18 MAPE of GP model capacity predictions and baseline capacity estimates for
EWR and forecast horizons of 1 h to 6 h. . . . . . . . . . . . . . . . . . . . . 161
4-19 MAPE of GP model capacity predictions and baseline capacity estimates for
LGA and forecast horizons of 1 h to 6 h. . . . . . . . . . . . . . . . . . . . . 162
4-20 Histogram of prediction errors for GP model predictions (for 3-h look-ahead
time) and baseline capacity estimates for JFK. . . . . . . . . . . . . . . . . . 163
4-21 Histogram of prediction errors for GP model predictions (for 3-h look-ahead
time) and baseline capacity estimates for EWR. . . . . . . . . . . . . . . . . 163
4-22 Histogram of prediction errors for GP model predictions (for 3-h look-ahead
time) and baseline capacity estimates for LGA. . . . . . . . . . . . . . . . . 164
4-23 Expected delay costs and computational time as a function of the number of
capacity profiles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
4-24 Planned ground delay costs as a function of robustness parameter ↵. . . . . . 173
4-25 Expected airborne delay costs as a function of robustness parameter ↵. . . . 173
4-26 Actual delay costs as a function of robustness parameter ↵. . . . . . . . . . . 174
4-27 Actual delay costs (averaged over all test cases) as a function of robustness
parameter ↵. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
4-28 GDP delay cost predictability as a function of robustness parameter ↵. . . . 175
4-29 Optimized arrival rates as a function of robustness parameter ↵. . . . . . . . 175
4-30 Difference between the revised and the initially planned ground delays as a
function of robustness parameter . . . . . . . . . . . . . . . . . . . . . . . . 176
4-31 Difference between the revised and the initially expected airborne delays as a
function of robustness parameter . . . . . . . . . . . . . . . . . . . . . . . . 176
4-32 Planned delay costs for revised and initial plans as a function of robustness
parameter (averaged over all test cases). . . . . . . . . . . . . . . . . . . . 176
4-33 Actual delay costs for revised and initial plans as a function of robustness
parameter (averaged over all test cases). . . . . . . . . . . . . . . . . . . . 176
4-34 Comparison between GDP costs resulting from data-driven and baseline AAR. 177

15
4-35 MAPE of GP model capacity predictions (averaged over planning horizon) for
GDP test cases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178

A-1 JFK arrival trajectory clusters; each color represents one cluster. . . . . . . . 186
A-2 JFK arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 186
A-3 Centroids of JFK arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 186
A-4 Distribution of JFK arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
A-5 JFK departure trajectory clusters; each color represents one cluster. . . . . . 187
A-6 JFK departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 187
A-7 Centroids of JFK departure trajectory clusters. . . . . . . . . . . . . . . . . 187
A-8 Distribution of JFK departure trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
A-9 EWR arrival trajectory clusters; each color represents one cluster. . . . . . . 188
A-10 EWR arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 188
A-11 Centroids of EWR arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 188
A-12 Distribution of EWR arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
A-13 EWR departure trajectory clusters; each color represents one cluster. . . . . 189
A-14 EWR departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 189
A-15 Centroids of EWR departure trajectory clusters. . . . . . . . . . . . . . . . . 189
A-16 Distribution of EWR departure trajectories by cluster; grey bar represents
the percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
A-17 LGA arrival trajectory clusters; each color represents one cluster. . . . . . . 190
A-18 LGA arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 190
A-19 Centroids of LGA arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 190
A-20 Distribution of LGA arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
A-21 LGA departure trajectory clusters; each color represents one cluster. . . . . . 191
A-22 LGA departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 191

16
A-23 Centroids of LGA departure trajectory clusters. . . . . . . . . . . . . . . . . 191
A-24 Distribution of LGA departure trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191

B-1 HKG arrival trajectory clusters; each color represents one cluster. . . . . . . 194
B-2 HKG arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 194
B-3 Centroids of HKG arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 194
B-4 Distribution of HKG arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
B-5 HKG departure trajectory clusters; each color represents one cluster. . . . . 195
B-6 HKG departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 195
B-7 Centroids of HKG departure trajectory clusters. . . . . . . . . . . . . . . . . 195
B-8 Distribution of HKG departure trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
B-9 SZX arrival trajectory clusters; each color represents one cluster. . . . . . . . 196
B-10 SZX arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 196
B-11 Centroids of SZX arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 196
B-12 Distribution of SZX arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
B-13 SZX departure trajectory clusters; each color represents one cluster. . . . . . 197
B-14 SZX departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 197
B-15 Centroids of SZX departure trajectory clusters. . . . . . . . . . . . . . . . . 197
B-16 Distribution of SZX departure trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
B-17 MFM arrival trajectory clusters; each color represents one cluster. . . . . . . 198
B-18 MFM arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . 198
B-19 Centroids of MFM arrival trajectory clusters. . . . . . . . . . . . . . . . . . 198
B-20 Distribution of MFM arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
B-21 MFM departure trajectory clusters; each color represents one cluster. . . . . 199

17
B-22 MFM departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 199
B-23 Centroids of MFM departure trajectory clusters. . . . . . . . . . . . . . . . . 199
B-24 Distribution of MFM departure trajectories by cluster; grey bar represents
the percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199

C-1 GRU arrival trajectory clusters; each color represents one cluster. . . . . . . 202
C-2 GRU arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 202
C-3 Centroids of GRU arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 202
C-4 Distribution of GRU arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
C-5 GRU departure trajectory clusters; each color represents one cluster. . . . . . 203
C-6 GRU departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 203
C-7 Centroids of GRU departure trajectory clusters. . . . . . . . . . . . . . . . . 203
C-8 Distribution of GRU departure trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
C-9 CGH arrival trajectory clusters; each color represents one cluster. . . . . . . 204
C-10 CGH arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 204
C-11 Centroids of CGH arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 204
C-12 Distribution of CGH arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
C-13 CGH departure trajectory clusters; each color represents one cluster. . . . . 205
C-14 CGH departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 205
C-15 Centroids of CGH departure trajectory clusters. . . . . . . . . . . . . . . . . 205
C-16 Distribution of CGH departure trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
C-17 VCP arrival trajectory clusters; each color represents one cluster. . . . . . . 206
C-18 VCP arrival trajectories labeled as noise. . . . . . . . . . . . . . . . . . . . . 206
C-19 Centroids of VCP arrival trajectory clusters. . . . . . . . . . . . . . . . . . . 206
C-20 Distribution of VCP arrival trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206

18
C-21 VCP departure trajectory clusters; each color represents one cluster. . . . . . 207
C-22 VCP departure trajectories labeled as noise. . . . . . . . . . . . . . . . . . . 207
C-23 Centroids of VCP departure trajectory clusters. . . . . . . . . . . . . . . . . 207
C-24 Distribution of VCP departure trajectories by cluster; grey bar represents the
percentage of noise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207

19
THIS PAGE INTENTIONALLY LEFT BLANK

20
List of Tables

1.1 Major airport/airspace dependency issues that affect metroplex operations [3]. 38

2.1 Non-conforming trajectory detection performance. . . . . . . . . . . . . . . . 77

3.1 Number of trajectory tube intersections identified by pair of metroplex airports. 95


3.2 Results of the metroplex hourly flow vector clustering analysis. . . . . . . . . 101
3.3 Description of the New York MFPs. . . . . . . . . . . . . . . . . . . . . . . . 103
3.4 Airport runway configurations associated with each New York MFP. . . . . . 104
3.5 Description of the Hong Kong MFPs. . . . . . . . . . . . . . . . . . . . . . . 108
3.6 Airport runway configurations associated with each Hong Kong MFP. . . . . 109
3.7 Description of the Sao Paulo MFPs. . . . . . . . . . . . . . . . . . . . . . . . 111
3.8 Airport runway configurations associated with each Sao Paulo MFP. . . . . . 111
3.9 Number of flow crossings by MFP. . . . . . . . . . . . . . . . . . . . . . . . . 114
3.10 Metroplex arrival rates under persistent demand. . . . . . . . . . . . . . . . 126
3.11 Comparison between empirical and baseline metroplex arrival rates for each
New York MFP. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
3.12 Daily capacity and structural path stretch variation. . . . . . . . . . . . . . . 128
3.13 NRMSE for the MFP throughput-demand curves. . . . . . . . . . . . . . . . 129

4.1 Classification accuracy of NN model for various network architectures based


on 5-fold cross validation performance. . . . . . . . . . . . . . . . . . . . . . 141
4.2 Description of GDP events at JFK used as test cases. . . . . . . . . . . . . . 171

A.1 DBSCAN parameter settings used for trajectory clustering. . . . . . . . . . . 185

21
B.1 DBSCAN parameter settings used for trajectory clustering. . . . . . . . . . . 193

C.1 DBSCAN parameter settings used for trajectory clustering. . . . . . . . . . . 201

22
List of Acronyms and Abbreviations

AAR Airport Acceptance Rate


ADS-B Automatic Dependent Surveillance - Broadcast
ANSP Air Navigation Service Provider
ARSI Arrival Route Status and Impact
ARTCC Air Route Traffic Control Center
ASPM Aviation System Performance Metrics
ATC Air Traffic Control
ATCSCC Air Traffic Control System Command Center
ATCT Air Traffic Control Tower
ATFM Air Traffic Flow Management
ATM Air Traffic Management
CDM Collaborative Decision-Making
CGH Sao Paulo/Congonhas Airport
CNS Communication, Navigation and Surveillance
CTOP Collaborative Trajectory Options Program
EDCT Expected Departure Clearance Time
EWR Newark International Airport
FAA Federal Aviation Administration
FCA Flow Constrained Area
GDP Ground Delay Program
GHP Ground Holding Problem
GPS Global Positioning System

23
GRU Sao Paulo/Guarulhos International Airport
HKG Hong Kong International Airport
IATA International Air Transport Association
ICAO International Civil Aviation Organization
ILS Instrument Landing System
IMC Instrument Meteorological Conditions
JFK John F. Kennedy International Airport
LGA LaGuardia Airport
METAR Meteorological Terminal Aviation Routine Weather Report
MFM Macau International Airport
NAS National Airspace System
NASA National Aeronautics and Space Administration
NextGen Next Generation Air Transportation System
NEXTOR National Center of Excellence for Aviation Operations Research
RAPT Route Availability Planning Tool
RBS Ration-By-Schedule
SESAR Single European Sky ATM Research
STAR Standard Terminal Arrival Route
SZX Shenzhen Bao’an International Airport
TAF Terminal Aerodrome Forecast
TBO Trajectory-Based Operations
TFMS Traffic Flow Management System
TMI Traffic Management Initiative
TRACON Terminal Radar Approach Control
VCP Viracopos International Airport
VMC Visual Meteorological Conditions

24
Chapter 1

Introduction

1.1 Motivation
Air transportation is critical to promoting a nation’s economic growth and social develop-
ment. As the fastest means for worldwide transportation, it is essential for global business
and tourism. In 2017, aviation transported 4.1 billion passengers and 59.9 million tonnes of
freight [4]. According to the latest study on the economic and social benefits of air transport,
the industry supported 62.7 million jobs worldwide and its global economic impact (direct,
indirect, induced and enabling) was estimated at US$ 2.7 trillion, equivalent to 3.5% of the
world’s Gross Domestic Product, in 2014 [5]. The importance of the industry is expected
to continue to increase over the next decades. The International Air Transport Association
(IATA) expects the air traffic demand to double by 2035, reaching 7.2 billion passengers
globally [6]. In the U.S., the Federal Aviation Administration (FAA) forecasts an average
increase in system enplanements of 2.2% per year over the next 20 years [7].
Enabling air transportation over the years, Air Traffic Management (ATM) is the system
of systems responsible for promoting a safe and orderly flow of aircraft through the airspace.
Though serving the aviation needs since its beginning, this system has reached its operational
limits. In 1983, the International Civil Aviation Organization (ICAO) first recognized the
need for new air navigation systems and procedures in order to facilitate the civil aviation
growth at a global scale [8]. Yet, present ATM systems are still limited by technologies and
operational procedures from the past century, which constrain capacity and lead to ineffi-

25
ciencies such as unnecessary delays and emissions. Demand-capacity mismatches currently
impose significant delays for passengers and costs for the airline industry and the economy
as a whole. In the U.S., for example, 16.5% of all domestic flights were delayed by more
than 15 minutes in 2016 [9]. In Europe, the percentage of delayed flights was 18.2% in 2015
[10]. According to a study developed by the FAA National Center of Excellence for Aviation
Operations Research (NEXTOR), the annual cost of flight delays to the U.S. economy was
estimated to be US$ 31.2 billion in 2007, including direct costs for airlines and passengers
and indirect costs associated with demand and Gross Domestic Product losses [11].
In order to enable the foreseen air traffic growth sustainably and address future air trans-
portation challenges, a global effort is currently underway towards the modernization and
harmonization of ATM systems. In the U.S., the Next Generation Air Transportation System
(NextGen) initiatives aim to achieve an in-depth transformation of the National Airspace
System (NAS) that will increase safety and efficiency and reduce the environmental impacts
of aviation [12]. Likewise, the Single European Sky ATM Research (SESAR) initiative co-
ordinates and concentrates all EU research and development activities to define, develop
and deploy the new generation of ATM system in Europe [13]. Similar programs have been
created in Japan (CARATS [14]), Brazil (SIRIUS [15]), China (CAAMS [16]) etc.
Overall, these initiatives envision the deployment of new technologies for Communica-
tion, Navigation and Surveillance (CNS) and ATM automation as well as new operational
procedures that best leverage the on-going technological advancements. For example, digital
data-link systems are expected to replace voice-based systems for faster and more accu-
rate air traffic controller-pilot communication. Satellite-based navigation will enable more
direct and efficient routes and provide more stringent navigational accuracy than ground-
based navaids. Satellite-based surveillance will complement radar-based systems, allowing
for greater coverage and increased precision in the monitoring and control of aircraft. New
decision support systems will increase automation and enhance strategic and tactical traffic
flow and airspace management. Taking advantage of all these capabilities, Trajectory-Based
Operations (TBO) is envisioned to be the new concept of operations for future ATM, bringing
together modern procedures to allow aircraft to follow precise 4D trajectories pre-negotiated
between flight operators and Air Navigation Service Providers (ANSPs) towards increasing

26
the efficiency, predictability and flexibility in trajectory planning and execution [8, 17, 18].
Besides the implementation of new technologies and operational procedures, efficient use
and management of air transport system data and information is key to achieve the desired
transformation of ATM systems. Every day, large amounts of data are generated as air traffic
operations occur, both at the planning and execution stages. Data types include environmen-
tal conditions (e.g., weather reports), system state (e.g., air traffic control facility reports),
flight demand (e.g., airline schedules, flight plans, surveillance tracks), etc. Furthermore,
they are expected to become increasingly available and accessible as new technologies are
deployed. For instance, flight trajectory data has become increasingly accessible with the ad-
vent of new surveillance technologies. Historically, there has been limited sharing of this type
of data by ANSPs for national security reasons. With Automatic Dependent Surveillance-
Broadcast (ADS-B), open-source flight tracking data has become publicly available, making
it possible for the first time to analyze aircraft movement at a global scale. The development
of innovative ways to leverage the system raw data can play a key role in the ATM transfor-
mation by providing means for better assessing and understanding operational performance,
increasing the system predictive power and creating new decision support tools to assist the
planning at strategic and tactical levels. It can help moving from a system that often works
reactively and still much relies on experience, intuition and "golden rules" towards a system
that is more intelligent, proactive and adaptive.
This research seeks to contribute to this goal by exploring the development of high-fidelity
air traffic flow models from large-scale aircraft tracking data and analyzing their potential
towards assisting ATM decision-making. Air traffic flows can be far more diverse and complex
than what published routes and procedures alone suggest. For example, Figure 1-1 shows the
trajectories for arrival flights into John F. Kennedy International airport (JFK) during two
different time periods of one same day overlaid with the published Standard Terminal Arrival
Routes (STAR). It is observed that the spatial distribution of the traffic does not fully match
the STARs and it can vary reasonably over time. Knowledge about these spatiotemporal
characteristics of the traffic is key for accomplishing ATM by informing various airspace
management and traffic flow management related decisions, as exemplified below:

• Strategic airspace design: It supports the assessment of the actual utilization of the

27
airspace for better design of sectors and routes.

• System capacity estimation: It supports the estimation of how much traffic can be
accommodated in airspace resources; airspace capacity is a function of air traffic control
complexity, which is a function of the flow structure and associated difficulty of control.

• Demand estimation: It supports estimating the demand on airspace fixes, sectors or


larger airspace volumes.

• Design of traffic flow management strategies: It supports the design of flow manage-
ment strategies to balance demand with capacity and reduce delays; critical regions
of the airspace (more prone to see demand-capacity imbalances) can be more easily
identified and become a target for flow adjustment programs.

• Tactical airspace management: It supports the allocation of staff and/or the tactical
reconfiguration of the airspace to balance workload, increase capacity and throughput;
periods of higher air traffic control complexity, with complex flow structure, can be
more easily anticipated, indicating that changes in staffing or in the airspace structure
may be necessary to have an expeditious flow.

Figure 1-1: Trajectories of arrival flights into JFK during two different time periods on June
23, 2015 overlaid with the Standard Terminal Arrival Routes.

Currently, this knowledge about the traffic flow behavior is primarily obtained through
human experience. Based on the continuous observation of flight trajectories through air

28
traffic situation displays from surveillance systems, air traffic controllers and managers are
able to form mental models and create abstractions of the traffic flow in a region. Conse-
quently, they create an intuition about the location of major flows and about the time they
tend to occur, which inform their decisions. This reliance on human experience/intuition
sometimes yields inefficiencies as decisions tend to be taken based on local aspects rather
than on a systems perspective. Moreover, it limits the capabilities of automated decision sup-
port systems for air traffic management and hinders the implementation of more advanced
operational concepts that require high-fidelity flow information.
Hence, there is a need to develop automated tools that can best characterize and predict
the behavior of the traffic and be integrated into decision-support tools for ATM. It is worth
mentioning that this need is even more patent in the envisioned TBO environment, where
the less structured airspace resulting from more flexible trajectory planning will require more
efforts to understand traffic flows.

1.2 Thesis Objectives


This thesis has two main objectives:

1. Investigate how operational data in the ATM system can be mined to automatically
identify and predict air traffic flows;

2. Analyze the potential contribution of such knowledge towards obtaining a better un-
derstanding of the traffic flow behavior and performance and supporting system-level
decision-making in ATM.

1.3 Methodological Approach


This thesis leverages data analytics to achieve the objectives aforementioned. Hu et al. [19]
define data analytics as the process of using algorithms to uncover information concealed
in data, such as hidden patterns or unknown correlations, with a variety of purposes, such
as to give advice, to diagnose and infer reasons for fault or to predict what will occur in

29
the future. Generally, data analytics can be classified in three levels according to the ways
by which it can affect decision-making: descriptive, predictive and prescriptive [20]. These
paths are illustrated in Figure 1-2. The first path is the most straightforward; descriptive
analytics exploits historical data to diagnose, identify relevant explanatory factors for an
event outcome and create insights for guiding future decisions. The second path leverages
data to predict the most likely outcome of an event, which can be the state of a variable
affecting the decision or even the decision itself. Finally, the third path leverages insights
from data to develop realistic models that can prescribe actions to control the outcome of the
decision for a given objective. Ultimately, data analytics can be described as the science of
using data to build models, improve decisions, and add value to organizations and individuals
[21].

Figure 1-2: General ways by which data analytics affects decision-making.

In this thesis, a flight trajectory data analytics framework is developed based on the
application of machine learning methods to exploit flight tracking data and discover spatial
and temporal patterns in the movement of aircraft through the airspace in order to pro-
vide a high-fidelity characterization of air traffic flows. This characterization provides the
foundation for obtaining new insights about the traffic behavior and performance as well as
for the development of predictive models to support airspace management and traffic flow

30
management decisions. The approach is illustrated in Figure 1-3.

Figure 1-3: Ways by which the flight trajectory data analytics framework can impact ATM.

In order to demonstrate this data analytics framework and discuss its potential impacts
on ATM decision-making, this thesis focus on terminal area operations, particularly of multi-
airport (metroplex) systems. Large metropolitan regions are typically served by two or more
significant airports, which are often closely located and share most of their surrounding
airspace (terminal area). As a result of the high traffic density and the interdependent uti-
lization of both airport and airspace resources, the metroplex terminal airspace tends to
feature complex traffic dynamics. Besides, the terminal area phase has significant contri-
bution to both individual flight and system level efficiency. The high density of aircraft
converging in the terminal area along with the more constrained airspace structure often
result in sub-optimal flight paths. These inefficiencies are even more pronounced in multi-
airport systems because of the even higher traffic volume and the existence of inter-airport
flow interactions. Therefore, better understanding the dynamics and the performance of
the traffic flows at the terminal area scale, especially for super dense and complex terminal
airspace that characterize metroplex systems, has one of the greatest potential towards sup-
porting airspace and traffic flow management decisions at the systems and enhancing system
level ATM performance.

31
1.4 Background and Related Literature

1.4.1 An Overview of Air Traffic Management

The primary function of Air Traffic Management (ATM) is to ensure a safe and orderly flow
of aircraft through the airspace. ATM can be generally subdivided in three distinct functions
[22]:

• Air Traffic Control;

• Airspace Management;

• Air Traffic Flow Management.

The main purpose of Air Traffic Control (ATC) is to guarantee safe separation between
aircraft and between aircraft and obstacles and avoid collisions. ATC is also responsible
for providing flight information services to aircraft, such as weather reports and updates on
facility conditions, as well as alerting about aircraft in need of search and rescue. In order to
enable the ATC activity, some elementary systems are required: communication, navigation,
surveillance and flight and weather information. A detailed description of these systems can
be found in [23].
Airspace Management concerns the allocation of the airspace to its various users and
its design to provide ATC services. In general, the airspace is subdivided into regions that
are assigned to different control facilities. The standard airspace and control structure [24]
is composed of Control Tower, which is responsible for the traffic in the nearby airspace
around airports and on their surface, Approach Control (called Terminal Radar Approach
Control - TRACON - in the U.S.), which controls the terminal airspace (or terminal area)
containing the arrival and departure procedures of one or more airports, and Area Control
Center (called Air Route Traffic Control Center - ARTCC - in the U.S.), which is responsible
for volumes of the airspace containing airways. These different regions of the airspace are
further subdivided into sectors, which are individually controlled by an air traffic controller.
In the U.S., the NAS contains 22 ARTCCs that are subordinated to a central control unit
called Air Traffic Control System Command Center (ATCSCC). This airspace structure is

32
quite rigid and only shows some level of flexibility in terms of internal sectorization as sectors
can be merged or split depending on traffic conditions.
Finally, Air Traffic Flow Management (ATFM) has the goal of matching demand with
available capacity by adjusting the flows on a national or regional basis when imbalances
are predicted to occur because of weather conditions, traffic volume or special events. In
the NAS, ATFM is accomplished through different types of Traffic Management Initiatives
(TMIs) issued at strategic (i.e., planning horizons of 2 to 8 hours) and tactical (i.e., real
time to planning horizons of 2 hours) time frames. Common strategic TMIs include delaying
aircraft on the ground at the origin airport via a Ground Delay Program (GDP) when
capacity at the destination airport is reduced or via an Airspace Flow Program when the
aircraft is planned to traverse a Flow Constrained Area (FCA), i.e., an airspace region that
is impacted (primarily because of weather). Still at strategic time frame, flights can also
be rerouted pre-departure to avoid congested or weather impacted areas. The National
Playbook contains a set of alternative routes that have been pre-validated and coordinated
with ARTCCs and that can be used in support of strategic rerouting. While airborne, flights
are subject to tactical TMIs implemented in a more reactive way. For example, an aircraft
may be tactically rerouted around weather or may be slowed down with a Miles-in-Trail
restriction, a strategy commonly used to control arrival rates into an area through additional
spacing at airspace fixes. Recently, efforts have been made to develop new ATFM strategies
that integrate existing ones and provide a more flexible and efficient solution for airspace
users. The Collaborative Trajectory Options Program (CTOP) is a new strategic TMI
developed by the FAA that simultaneously assigns ground delays and/or reroutes around
one or more FCAs after explicitly taking into account the priorities of flight operators under
a Collaborative Decision-Making (CDM) approach [25]. The National Aeronautics and Space
Administration (NASA) is also currently working towards the development of the Integrated
Demand Management concept, which attempts to coordinate strategic and tactical flow
scheduling by using a CTOP TMI to pre-condition the flows into the Time-Based Flow
Management system so that an arrival schedule with lower delays can be built [26].
The planning and execution of the three functions described above is a complex task
accomplished by the ATM system. The difficulty of ATM can be generally attributed to

33
three major factors:

High volume and heterogenous demand

In the U.S., an average of more than 42,000 flights operate per day, and more than 5,000
aircraft can be present in the sky at any given time [27]. Large hub airports serving dense
metropolitan regions concentrate a significant part of the traffic; in 2016, 30 airports handled
more than 70% of the revenue passenger enplanements [28]. Besides, the system has to deal
with a variety of aircraft and mission types, including large commercial jets, small general
aviation aircraft, rotorcraft, etc. The rise of unmanned and urban air mobility markets
is expected to further increase the volume and heterogeneity of demand, accentuating the
pressure over the system.

Dynamic system capacity

The difficulty of ATM can also be attributed to the dynamic behavior of system capacity.
Knowledge of capacity is key for allocating demand to ATM resources efficiently, but the large
number of capacity-determining factors and their dynamics makes it difficult to estimate flow
rates precisely, especially for long time horizons. Figure 1-4 provides a decomposition of the
factors that influence airport/airspace capacity and their interrelations. The uncertainty
in system capacity can be largely attributed to the dynamic weather behavior. It can be
observed that weather conditions plays a major role in determining achievable capacity, both
indirectly by influencing the selection of an ATM configuration, and directly by impacting
the throughput performance of a given ATM configuration.

Fragmented, multi-stakeholder and human-based decision-making

Additional complexity arises because ATM decisions are taken by multiple human decision-
makers in multiple sub-systems dealing with different phases of a flight at different time
horizons, requiring lots of coordination. Moreover, they have to accommodate interests from
multiple stakeholders, including air traffic service providers, flight operators and passengers.

34
Figure 1-4: Airport/airspace capacity influencing factors.

The ATM functional structure is summarized in the simplified diagram shown in Figure
1-5. The processes and information flows that occur throughout the life-cycle of a flight
are indicated, as well as the time horizons associated with each particular process. At the
long-term planning horizon ranging from months to years in advance, airline scheduling
processes generates the flight schedules to meet the airline’s business goals, while strategic
Airspace Management is accomplished by the ANSP to design the airspace and the opera-
tional procedures. Closer to the flight departure date, which can range from hours to days in
advance, airlines elaborate their flight plans describing in detail the proposed aircraft flight
to realize the schedule. Flight plans are submitted to the ANSP central ATFM unit, which
will approve or require adjustments, for instance, to mitigate an expected demand-capacity
imbalance at a particular resource. Expected system capacity is a function of expected ATM
configurations and forecast weather conditions at that time. At the tactical time horizon
ranging from real time to hours in advance, tactical ATFM is accomplished at the facility
level to adjust the flow rates in response to dynamic changes in weather conditions, ATM
configuration and capacity. Finally, ATC delivers the final clearances and advisories to as-

35
sure the safe separations, while aircraft guidance systems enable the execution of the agreed
trajectory plan.

Figure 1-5: ATM functional structure.

1.4.2 Challenges in Metroplex Air Traffic Management

A multi-airport system (also referred to as metroplex) can be defined as a set of two or more
significant airports that serve commercial traffic in a metropolitan region, without regard to
ownership or political control of individual airports [29]. Multi-airport systems have emerged
worldwide as a response to congestion problems, allowing the air transportation system to
scale and meet the increasing demand [30]. Bonnefoy (2008) identified two fundamental
mechanisms governing the emergence of a multi-airport system historically: construction of
a new airport in a region served by a single airport system, with partial or total transfer
of traffic; or use of existing airports that were previously sub-utilized in the metropolitan
region [30]. The second mechanism has been especially observed after the emergence of
low-cost carriers. Through this scaling process, multi-airport systems have accommodated a
significant portion of the global traffic and have become key nodes of the air transportation

36
system. Based on data of 2006, Bonnefoy (2008) identified the existence of 59 multi-airport
systems worldwide, which handled 50% of the global passenger traffic.
With the emergence of multi-airport systems, the ATM system has to deal with increas-
ing levels of complexity in the management of the traffic surrounding these systems. The
sharing of the terminal airspace results in operational interdependencies and interactions be-
tween dense arrival and departure routes from the multiple airports in the same geographical
area, creating a complex environment for traffic management. An extensive study developed
by Ren et al. [31] and Clarke et al. [3] identified twelve major general metroplex depen-
dency issues that impact their operations based on site visits, domain expert evaluation and
qualitative analysis of four major U.S. multi-airport systems (Atlanta, Los Angeles, New
York and Miami). These issues are listed in Table 1.1. For example, the use of arrival
and departure routes by an airport tends to be dictated not only by its individual runway
configuration, but also by the neighboring airport configuration. These interdependencies
increase the impact and the workload associated with runway configuration changes (issues
3, 5). Besides, the sharing of airspace resources (arrival and departure fixes, path segments)
creates flow interactions that affect the overall performance of the metroplex (issues 1, 2, 6,
7, 8, 11).
As mentioned previously, the difficulty of ATM can be generally attributed to three
major factors: high volume; dynamic system capacity; and fragmented decision-making. At
multi-airport systems, all of these factors are intensively present: metroplex airspace are the
densest of all; metroplex interdependencies result in high variability in ATM configuration
and performance; metroplex decisions have to be coordinated between facilities managing the
individual airports and airspace. As a result, these systems are characterized by a complex
operating environment for ATM. For instance, at the same time they are more reliant on
ATFM to adjust demand with the more dynamic system capacity, their characteristics also
make the planning of ATFM very challenging. Because there is higher uncertainty in the
ATM configuration and its performance, estimating future capacity precisely is more difficult
and the allocation of airport and airspace resources under uncertainty becomes more subject
to inefficiencies.
The metroplex airspace design and the operating strategies governing the use of the

37
Table 1.1: Major airport/airspace dependency issues that affect metroplex operations [3].

Metroplex issue Definition Impact

1 Multi-airport de- Occurs when flights from at least two separate air- Very high
parture merge over ports are procedurally merged over at least one com-
common departure fix mon departure fix.
2 Major volume-based Occurs when a significant level of ATFM restric- Very high
traffic flow man- tions due to demand-to-capacity overloads exist at
agement (ATFM) airspace fixes or at airports.
restrictions
3 Proximate-airport con- Occurs when an airport configuration change of one High
figuration conflicts of at least two proximate airports puts restrictions
on flights flying to/from other proximate airport(s).
This involves flows from one impacting another air-
port’s flows, causing significant rerouting or delays.
4 Slow inter-airport Occurs when inadequate surface transportation of High
ground connectivity passengers between airports cause significant delays
and consequently limits the efficient use of airports
by passengers.
5 Inefficient/high work- Occurs when any major airport configuration change High
load airport configura- requires significant workload due to reasons, such as
tion changes coordination of a large number of personnel, FAA
facilities, and airports or sector reconfigurations.
6 Inefficient multi- Occurs when departure sequencing of flights from High
airport departure multiple airports requires conservative flight restric-
sequencing tions.
7 Major secondary air- Occurs when conflicts between a primary airport and High
port flow constraints a secondary airport lead to constraints on secondary
airport flows. Typically, secondary airport traffic will
be held below primary airport traffic flows or will be
routed around the primary airport traffic patterns
resulting in longer flight paths.
8 Inefficient flushing of Occurs when ATC uses a flushing technique that con- Medium
airport flows strains other airport traffic flows in order to expedite
one airport’s arrival or departure flights as a way to
solve a particular congestion problem (e.g., airport
arrival gridlock).
9 External special use Occurs when SUA external to the TRACON con- Medium
airspace (SUA) causes stricts TRACON flows into narrow corridors and
flow dependencies forces inter-airport traffic flow dependencies.
10 Terrain causes flow de- Occurs when terrain internal to the TRACON con- Medium
pendencies stricts TRACON flows into narrow corridors and
forces inter-airport traffic flow dependencies.
11 Severe limitations on Occurs when the use of instrument procedures is Low
instrument procedures severely constrained due to the existence of a proxi-
due to proximate air- mate airport.
port
12 Insufficient regional Occurs when there is generally not enough TRACON Low
airport capacity runway capacity to efficiently serve the air traffic de-
mand.

38
airspace are recognized to be key factors affecting the coupling of operations, the traffic
dynamics and the overall performance of a metroplex. Yet, studies that investigate the in-
terplay between these factors and their impacts on the traffic dynamics and performance
are limited. Visual inspection of flight trajectory data has been used to qualitatively eval-
uate airspace design and identify potentially constraining flow interactions [31]. Delay and
fuel burn impacts of decoupled metroplex design and new traffic scheduling strategies have
been assessed through simulation [3]. Donaldson and Hansman presented an empirical study
aimed at quantifying capacity impacts associated with metroplex interactions [32]. Through-
put performance at the airport level and at the system level for particular combinations of
runway configurations in the New York metroplex were evaluated using operational data.
Capacity discrepancies as great as 60 operations per hour were observed, emphasizing that
terminal airspace flow interactions tend to be an important constraining factor driving the
capacity of metroplex airports. In this thesis, we demonstrate the potential of the flight
trajectory data analytics framework to provide a deeper understanding of the metroplex
traffic behavior and performance and discuss how this understanding can be useful for ATM
decision support at these systems.

1.4.3 Optimization Models for Air Traffic Management

There is a vast stream of literature dedicated to developing analytical models for optimization
of ATM decisions towards assisting human decision makers. The area of ATFM has received
particular attention given the complex decision-making nature of this ATM function and its
impacts on system-level efficiency. The problem of adjusting air traffic flows in real time in
order to balance demand with capacity was first modeled by Odoni [33] using optimization
techniques. Since then, several models have been conceptualized for the flow management
problem at different scales and considering various types of capacitated resources and control
mechanisms.
At the national and strategic level, ATFM is typically concerned about regulating flights
destined to a capacity-constrained airport or planned to traverse a capacity-constrained
airspace region. The problem of regulation of traffic towards a capacity-constrained airport
is the most discussed in the literature and it is commonly referred to as the Ground Holding

39
Problem (GHP). When demand exceeds capacity at an airport, the arrival traffic will be
subject to delays. There are two primary mechanisms by which flight delays can be incurred:
delays can be absorbed in the air, for example, when the aircraft is slowed down or put on a
holding pattern; or delays can be absorbed on the ground if the departure time is postponed.
Because airborne delays are typically more expensive than ground delays and there is an
upper bound on the amount of time that a flight can be held in the air, ground delays are
generally preferred from both an operating cost and a safety perspective. Nevertheless, as
ground delays are strategic in nature and capacity cannot be predicted exactly a few hours
in advance, a trade-off between ground and airborne delays will exist when the uncertain
nature of capacity is accounted for. Several mathematical programming formulations for the
GHP have been developed in order to determine the optimal allocation of ground delays that
minimizes the expected overall delay costs. The first model formulations for the GHP were
deterministic and static, i.e., capacity was assumed to be known with certainty and decisions
were made once at the beginning of the planning horizon [34]. Stochastic and static models
were proposed by Richetta and Odoni [35] and by Ball et al. [36] in order to account for
uncertainty in the airport capacity profiles. In order to also incorporate the ability to revise
decisions as updated capacity information becomes available during the planning horizon,
dynamic models were proposed by Richetta and Odoni [37] and Mukherjee and Hansen [38].
Most of the GHP models consider that the planning horizon for capacity allocation is
known a priori and determined by the ANSP in a centralized manner. Liu and Hansen [39]
took an alternative approach and proposed stochastic GHP models based on deterministic
queueing theory in order to optimize the duration of the ATFM plan. Instead of incorporat-
ing uncertainty in the arrival rates, they assumed the existence of deterministically known
normal and low capacity levels and introduced uncertainty on the time when the airport
capacity moves from the low level to the normal level. In their models, the decision on
the planning horizon was made to minimize the expected delay costs, which included unpre-
dictability premiums associated with revisions due to early or late capacity recovery. A body
of research has also been dedicated to developing decentralized (or "airline-driven") GHP
models that incorporate individual flight operator’s preferences with respect to the design
of the ATFM plan using voting mechanisms [40, 41, 42, 43]. The underlying motivation

40
for these studies is the fact that airlines have different recovery capabilities due to their
operational characteristics and, therefore, they can have different risk tolerances regarding
the design of the ATFM plan. For example, an airline that operates under higher flight
frequency, lower load factors and more homogeneous fleet mix may have a higher recovery
potential from disruption since it can more easily perform aircraft swaps or re-accommodate
passengers. Such an airline would tend to be less risk tolerant and prefer a more predictable
ATFM plan with longer duration in order to take advantage of its recovery capability [43].
By contrast, airlines with lower recovery potential might be willing to take more risks in or-
der to take advantage of any extra capacity that may be realized (such as an early capacity
recovery). These airlines would tend to prefer more aggressive ATFM plans with shorter
duration.
The Air Traffic Flow Management Problem (TFMP) extended the GHP by considering
that not only the airport but also the airspace can be capacity constrained and by introducing
additional control mechanisms such as airborne delays, rerouting and cancellations. Two
distinct modeling approaches have been pursued for the TFMP: Lagrangian approaches are
aircraft-based and seeks to control departure times and traversing times throughout the
airspace from origin to destination for each individual flight [44, 45, 46, 47, 48, 49]; whereas
Eulerian approaches are flow-based and seeks to control aggregate flow rates at nodes and
links of the air traffic route network [50, 51, 52, 53, 54].
At the regional and tactical level, ATFM also plays an important role in the regulation of
the traffic at individual airports or terminal areas. Many analytical models have been derived
for more efficient coordination of airport arrival and departure flows, with the objectives of
minimizing delays and maximizing resource throughput while maintaining fairness among
airspace users. Runway sequencing and scheduling models have been developed to determine
the optimal sequence and schedule of runway usage by arrival and departure aircraft based on
the safety requirements between specific aircraft types and their wake turbulence categories
[55, 56, 57, 58]. Runway configuration selection models have been developed to optimally
schedule airport runway configuration changes based on expected demand and meteorological
conditions in order to best balance capacity with the demand of arrivals and departures [59].
Taxiway scheduling models have been created to schedule taxi operations and determine

41
the optimal routing in the taxiway system in order to prevent "stop-and-go" situations and
minimize taxi times [60]. Finally, departure metering models have been developed to manage
the demand of departures by holding aircraft at the gate, with engines off, until the right
time for its release (i.e., the optimal time to leave the gate and reach the runway at its
assigned slot for takeoff) towards minimizing taxi and runway delays and mitigating airport
surface congestion [61, 62].
Besides traffic flow management, there has also been substantial research focused on
developing analytical models for airspace management decision support. Several approaches
have been derived for the problem of determining an efficient airspace sectorization given
a particular air traffic situation. Some of them leverage the existing airspace structure to
determine sector merging or splitting as a function of demand [63, 64, 65], whereas others
focus on redesigning airspace boundaries from scratch [66, 67, 68, 69, 70, 71, 72]. In order to
do so, a variety of methods have been used such as mathematical programming [63, 66, 67],
computational geometry [68, 72], heuristics [69] and graph theory [71]. Common design
objectives for an efficient sectorization were the balancing of monitoring workload among
sectors (i.e., the workload associated with monitoring and conflict-avoidance tasks) and
the minimization of coordination workload (i.e., the workload associated with hand-offs
between sectors). Other design goals included the maximization of sector dwell times and
the maximization of sector capacity.
Research in ATM has produced useful models to optimize the traffic flows and airspace
configurations, showing efficient computational times on practical size instances and signif-
icant potential towards improving the efficiency of air traffic operations. Yet, it is still ob-
served a large gap between theory and practical implementation. One reason for the existing
gap between theory and practice is the fact that many analytical models make assump-
tions that do not hold under the high levels of dynamism and uncertainty that characterize
the ATM operating environment, or require information that is not currently automatically
available in the ATM system. For instance, models for airspace configuration management
typically require detailed traffic information (e.g., dominant flows, critical points) that is not
automatically generated by the current systems that support ATM. Models for large-scale
air traffic flow optimization make simplistic assumptions about the air traffic route network

42
that typically do not match the actual spatial patterns of air traffic flows. Besides, they
often assume airport and airspace capacities are deterministic and can be known in advance
with certainty. Even more realistic models that account for the uncertain behavior of system
capacity face barriers towards implementation. For example, most stochastic GHP models
assume the uncertain airport capacity profiles can be represented in the form of scenario
trees, in which capacity changes between few discrete states assumed to be known to the
decision-maker [37, 38]. Yet, generating such scenarios in practice is not straightforward.
Besides, actual capacity may be better represented as a continuous variable rather than
with few discrete states. Other models assume that nominal and off-nominal capacity val-
ues can be treated deterministically and the time of capacity recovery can be modeled with
probability distributions [39, 43], but they lack of empirical validation.

1.4.4 Data Analytics for Air Traffic Management

Given the increasing amounts of operational data generated and recorded every day by the
ATM system and the advancements in computational science, a field of research has emerged
to investigate how data analytics can be leveraged to create useful information from the
system raw data for advanced ATM. A broad range of studies have exploited the various
types of data available in the ATM system for multiple different tasks, with the overarching
goals of better measuring and predicting ATM performance and facilitating decision-support
tool development. For example, a non-exhaustive list of topics addressed by these studies is
presented below.

Data-driven modeling of air traffic flows

Work has been done to model air traffic flows from flight tracking data for performance
assessment and monitoring, airspace design and traffic flow management purposes. For
example, Gariel et al. [73] developed a framework for terminal airspace monitoring through
the use of clustering techniques to learn nominal spatial trajectory patterns and to assess
the conformance of flight trajectories. Sabhnani et al. [74] also applied clustering methods
on flight trajectory data in order to identify en route sectors with highly structured traffic

43
patterns for airspace redesign purposes.

Data-driven modeling of airport and airspace capacity

Another stream of literature has focused on modeling airport and airspace capacity from
weather forecast data for traffic flow management applications. For example, Provan et al.
[75] used supervised learning to translate raw terminal aerodrome forecasts into probabilistic
airport capacity predictions for strategic planning horizons. Pfeil and Balakrishnan [76] also
applied machine learning methods on raw convective weather forecasts in order to generate
probabilistic predictions of terminal area route blockage because of convective weather.

Data-driven modeling of air traffic delays

Work has also been done to model the dynamics of delay from on-time performance records
for prediction and control of air traffic delays. For example, Rebollo and Balakrishnan [77]
applied clustering methods to identify characteristic delay states for the NAS and used su-
pervised learning to develop prediction models that leverages this network state information
to forecast departure delays for specific origin-destination pairs up to 24 h in the future.

Data-driven modeling of aircraft performance

Another group of studies has been dedicated to model aircraft performance from flight
recorder data for performance monitoring applications. For example, Li et al. [78] applied
clustering techniques on flight recorder data to model nominal aircraft behavior and identify
anomalous flights towards improved safety monitoring. Chati [79] developed data-driven
models of aircraft engine performance also using operational data from flight recorders in
order to predict a flight’s fuel burn directly from its trajectory for enhanced fuel efficiency
assessment and monitoring.
A more detailed review of the literature on the topics of data-driven air traffic flow
modeling and airport capacity modeling, which are relevant in the context of this thesis, are
provided in the next two sections.

44
1.4.5 Data-Driven Modeling of Air Traffic Flows

The characterization of air traffic flows from aircraft tracking data can be framed as a problem
of identification of spatial and temporal trends in the collective movement of aircraft through
the airspace. Such problem appears in a variety of other domains in which the movement of
objects tends to exhibit correlations in both spatial and temporal dimensions (e.g., vehicles
in a road network). Therefore, this section will provide a broader review of works (not
restricted to the air traffic domain) that aimed to investigate spatial and temporal patterns
in the collective behavior of moving targets. Before doing so, we will first make important
distinctions about types of spatial and temporal trends commonly investigated in trajectory
data.
Figure 1-6 exemplifies spatiotemporal patterns in a general trajectory dataset [1]. Using
the same terminology provided in [1], the following patterns are highlighted: a flock of three
entities with similar spatial pattern over five time-steps; a periodic pattern for an individual
entity, i.e., one entity repeats a specific spatial pattern with some periodicity; a meeting place
where three entities remain for some period; and a frequent location where a single entity
remains for some period. In this figure, the flock pattern provides a good illustration of the
spatiotemporal patterns that we seek to mine for flow identification: a group of entities with
similar spatial behavior during a period of time. A small distinction is that the entities do
not have to be at the same location at exactly the same time. By contrast, they traverse the
same sequence of locations within the same time interval. In summary, this thesis considers
the following definition of flow:

Definition 1: A flow is a pattern in which entities move along the same paths/routes
in a spatial dimension, and the movements start/end within the same time interval.

Spatiotemporal Pattern Recognition in General Trajectory Data

The problem of discovering spatiotemporal patterns in actual trajectory datasets has re-
ceived growing attention in the recent literature across a variety of domains. The increased

45
Figure 1-6: Example spatiotemporal patterns in trajectory datasets [1].

use of position sensing technologies (e.g., Global Positioning System (GPS)) has produced
large amounts of data and created an opportunity to gain insights through trajectory data
mining. Examples include tracking datasets of vehicles, people, animals, weather phenom-
ena (e.g., hurricanes) etc, which are exploited for a variety of purposes. Analysis of vehicle
trajectory patterns is used for traffic management in order to discover hot spots in a trans-
portation network and support route planning. Analysis of pedestrian flows can help iden-
tifying suspicious behavior in a monitored environment; it can also support urban planning
and guide infrastructure investments. Scientists investigate animal movement behavior in
order to identify regions frequently visited, investigate social structures within a group of
animals and better understand migration patterns. Investigation of hurricane/cyclone move-
ment patterns is performed to understand their behavior and increase the predictability of
severe-weather events for improved disaster relief management.
Early approaches to spatial pattern data mining involved indexing trajectory databases
and performing basic analysis such as nearest neighbor queries [1]. More recently, trajectory
clustering has been extensively used for the identification of common trajectory patterns
and outlier detection in such datasets. Antonini and Thiran [80] used hierarchical clustering
to group trajectories of redundant targets associated with a person for automatic counting
of pedestrians in video sequences. Fu et al. [81] also analyzed video surveillance data of

46
real-time traffic to identify vehicle motion patterns. With a first-layer spectral clustering,
similar trajectories were grouped together, and with a subsequent hierarchical clustering,
dominant paths and lanes were identified. Gaffney et al. [82, 83] proposed a probabilistic
modeling of trajectories as individual sequences of points generated from regression mixture
models (the spatial position was modeled with a polynomial regression model in which time
is the independent variable) and used the expectation-maximization algorithm to estimate
the parameters of the mixture. They applied the model-based clustering to identify spa-
tial patterns in extra-tropical cyclone tracks over the North Atlantic using meteorological
data. Lee et al. [84] developed a partition-and-group framework to discover common sub-
trajectories as an alternative to clustering trajectories as a whole. They argue that in some
applications (especially when there are regions of special interest for the analysis) it may be
useful to find portions of trajectories that have similar behavior even if they are dissimilar
when compared as a whole. Their proposed algorithm TRACLUS first partitions a trajec-
tory into a set of line segments at characteristic points where the behavior of the trajectory
changes rapidly. Then it groups similar line segments into clusters using a density-based
clustering scheme. They used the framework to identify movement patterns in very noisy
datasets from hurricane and animal tracking.
The works described so far primarily focused on the identification of similar trajectory
patterns in the spatial dimension. A much smaller body of work has incorporated the
temporal dimension in the analysis to investigate trends in the occurrence of spatial clusters
over time. However, such analysis is key for understanding the dynamics of flow behavior
and generating the knowledge necessary for prediction. Kim and Mahmassani [85] identified
spatial travel patterns in a road network through density-based clustering of actual vehicle
trajectory data and then used k-means to cluster daily time series of traffic for each identified
spatial travel pattern and discover daily trends in their use. In a similar way, Wen et al.
[86] developed an algorithm to extract shipping route from vessel’s historical position point
cloud using local polynomial regression and then clustered time series of vessel’s frequency of
occurrence in multiple zones defined along the frequent routes to find daily traffic patterns.

47
Spatiotemporal Pattern Recognition in Flight Trajectory Data

In the aviation domain, clustering techniques have also been extensively used to identify spa-
tial traffic patterns from flight track data for performance assessment, airspace monitoring,
airspace design and traffic flow management purposes [74, 87, 73, 88, 89, 53, 90, 54, 91]. Eck-
stein [87] developed a flight track taxonomy for monitoring aircraft behavior based on filter-
ing, segment identification, track decomposition and clustering in order to evaluate how well
individual flight trajectories are performing their procedures in the terminal airspace. Gariel
et al. [73] also developed a framework for terminal airspace monitoring using a density-based
clustering algorithm to learn typical patterns of operation and to assess the conformance of
flight trajectories. Similarly, Rehm [88] and Enriquez [89] relied on hierarchical and spectral
clustering respectively to identify nominal and abnormal spatial traffic patterns to/from a
specific airport. Trajectory clustering has also been used to identify spatial traffic patterns at
the en route phase. Sabhnani et al. [74] developed a greedy grid-based trajectory clustering
algorithm in order to learn standard flows and critical points and identify en route sectors
with highly structured traffic patterns for the goal of airspace redesign. Arneson et al. [90]
used a density-based clustering algorithm to identify dominant routing structures between
the Forth Worth Center and New York Center and assess the impacts of convective weather
on the flow rate capacity along the commonly used routes. Similarly, Marzuoli et al. [53]
and Bombelli et al. [54, 91] clustered trajectories between various origin and destination
pairs in order to identify common routing structures and model the air traffic route network
towards developing higher-fidelity models for traffic flow management optimization.
Studies that aimed to investigate patterns in the spatial organization of the traffic flows
over time are also limited in the aviation domain. Song et al. [92] manually extracted flow
features from actual flight trajectories in a 15-minute basis, which were aggregated into a
vector-based representation and clustered to identify traffic flow patterns in airspace sectors.
Flights entering and exiting an airspace sector through the same neighboring sectors were
considered to be part of the same flow regardless the shape of their trajectories. With the
methodology, they idealized a framework for sector capacity prediction for longer look-ahead
times, although little discussion was provided on the predictability of the flow features.

48
Sidiropoulos et al. [93] proposed a framework to identify spatiotemporal patterns in traffic
flows crossing the terminal area boundary in metroplex systems using flight plan data. The
approach first identified temporal changes in the local demand detected for discretized zones
of the terminal area boundary using a threshold for the difference in the number of flights
between consecutive time periods. Subsequently, temporal clusters in the aggregate metro-
plex demand were identified using a threshold on the number of zones with detected local
demand change. These thresholds were optimized to account for uncertainty in demand data.
Finally, within each temporal cluster, flights were spatially clustered based on the direction
in which they enter or exit the terminal area boundary using k-means algorithm. With the
framework, their goal was to provide enhanced demand estimates over shared arrival and
departure fixes for supporting metroplex traffic flow management. However, as the spa-
tial analysis was restricted to the location of crossings at the terminal area boundary, their
framework did not provide complete information about terminal area traffic flow patterns.

1.4.6 Data-Driven Modeling of Airport Capacity

As discussed in Section 1.4.1, the ATM system capacity is determined by a number of


structural, operational and environmental factors. The capacity of an airport is usually
determined by the capacity of its most constraining airside element. Typically, this element
is the runway system. A common definition of capacity is given by the expected number of
movements (landings and takeoffs) that can be performed per unit of time (usually, 1 hour), in
the presence of continuous demand and complying with all ATC separation requirements [23].
The capacity of a single runway is primarily determined by the minimum safety separation
requirements between consecutive operations, which vary according to the wake turbulence
categories of the leading and trailing aircraft. As a consequence, the mix of aircraft types
using the runway can significantly affect its capacity. Other important factors that impact
the capacity of a single runway include its geometric characteristics (e.g., the existence of
high-speed exits reduces runway occupancy times), aircraft performance (e.g., faster aircraft
can clear the runway more quickly) and the mix of operations (i.e., whether the runway is
dedicated exclusively for arrivals or departures or for both types of operation simultaneously).
Most airports have multiple runways that can be arranged in various configurations for

49
serving the demand. The layout of these configurations will determine the degree of de-
pendence between operations conducted in different runways. Therefore, a major driver of
airport capacity is the active runway configuration [94], i.e., the set of runways and ac-
tive runway thresholds selected by ATC personnel at any given time to serve the arrival
and departure demand. For example, Figure 1-7 shows two different runway configuration
commonly used at JFK. In the first one, two runways are used, with arrival operations on
31L/31R (blue arrows) and departure operations on 31L (green arrow). In the second one,
three runways are used, with arrivals on 22L/22R and departures on 22R and 31L. The selec-
tion of the runway configuration in which an airport will operate is a subjective human-based
decision-making process, which is per se affected by various factors such as meteorological
conditions (wind speed and direction, ceiling and visibility), demand, noise and workload
related restrictions and terminal airspace constraints.

Figure 1-7: Example of typical runway configurations at JFK.

Finally, another key driver of airport capacity is weather. Not only does weather in-
fluence the selection of runway configurations, but also directly impacts their throughput
performance. Adverse weather conditions such as low ceiling and visibility or high surface
winds can drastically reduce capacity and even cause the closure of the airport at extreme
cases.

50
The problem of airport capacity estimation has been widely addressed using analytical
[95], empirical [94] and simulation [96] models. For example, Gilbo [94] introduced the no-
tion of capacity curves (also called Pareto frontier) to characterize the trade-offs between
arrival and departure operations given a particular runway configuration. Figure 1-8 shows
example curves derived empirically from actual observed throughput for both runway con-
figurations shown in Figure 1-7 under Visual Meteorological Conditions (VMC). The curves
envelope the peak operating points and can provide an estimate of airport capacity under
the assumption that the peak data reflects the airport performance near capacity for that
particular condition. However, during real-time ATFM planning, predicting airport capacity
is a more difficult problem because runway configurations, weather conditions and demand
are not deterministically known a priori, especially for long forecast horizons.

Figure 1-8: Capacity curves for two different runway configurations at JFK (data from
2013-2015).

Efficient planning of airport capacity (typically referred to as Airport Acceptance Rates


(AARs)) is key for the overall efficiency of TMIs such as GDP. An overestimation of capacity
can amplify airborne delays and require reactive undesirable actions (e.g., diversions, exces-
sive holding), whereas an underestimation of capacity can lead to resource sub-utilization
and unnecessary ground delays. Yet, the planning of runway/airspace configurations and
capacity is typically done on the basis of experience and through the use of rules-of-thumb.
Despite the availability of many weather forecast sources, few tools are available to directly
assist traffic managers in the translation of weather forecasts into operational impact.

51
A body of literature has emerged to investigate whether data analytics techniques can
be leveraged to provide capacity-related information for improved ATFM decision support
[97, 98, 75, 99, 100, 101, 102, 103, 104]. Different methods for machine learning have been
used to translate weather forecasts and/or historical arrival rates into probabilistic airport
capacity profiles for strategic planning horizons [97, 98, 75, 99]. Liu et al. [97] used cluster-
ing techniques to create airport capacity profiles from historical AAR data. These profiles
were represented by the centroids of the obtained clusters from daily AAR time series. Buxi
and Hansen [98] adopted a similar approach, but also considered weather forecasts (Terminal
Aerodrome Forecast - TAF) to generate the profiles. Provan et al. [75] developed the Weather
Translation Model for GDP Planning (WTMG), which uses an ensemble of decision trees to
predict the AAR up to 12 hours in the future based on weather forecasts (TAF) and the cur-
rent state of the airport. Cox and Kochenderfer [99] considered the same variables to build
probability distributions for the AAR up to 6 hours in advance using Bayesian networks.
Jones et al. [100, 101] included additional features for capturing compression phenomena us-
ing the High-Resolution Rapid Refresh (HRRR) forecast and developed a boosted regression
tree model to issue capacity predictions in the form of quantiles, which were then applied
to chance-constrained integer programming models in order to prescribe AARs. Given the
significant impact of runway configuration selection on capacity, another group of studies has
used historical airport operational data to model the airport runway configuration selection
process and develop tools for runway configuration prediction [102, 103, 104]. For example,
Ramanujam and Balakrishnan [103] and Avery and Balakrishnan [104] used discrete-choice
models to generate probabilistic forecasts of the runway configuration for look-ahead times
of up to 6 hours.
One caveat of the existing data-driven approaches for airport capacity prediction con-
cerns the fact that airports are usually treated as isolated nodes and system level aspects
that arise in the most complex metroplex operational environment and affect throughput are
not completely incorporated, reducing their performance and applicability in such cases. For
instance, Avery [105] found that predicting the runway configurations individually for the
New York airports and then combining the results to obtain the aggregate metroplex con-
figuration rather than directly predicting the combined multi-airport runway configuration

52
led to much lower accuracy in the configuration forecasts. Donaldson and Hansman [32] also
ratified the existence of metroplex effects after identifying capacity discrepancies for specific
combinations of runway configurations in the New York metroplex through the comparison
between capacity envelopes for the individual airports and for the multi-airport system.

1.5 Contributions of the Thesis


Review of the literature in ATM has shown that many analytical models for optimization of
ATM decisions have been developed in order to assist human decision-makers, but they either
make simplistic assumptions about the ATM operating environment or require information
that is not currently automatically generated by the ATM system, hindering their practical
implementation. The increasing availability of data and the advancements in computational
science have created new opportunities for improving ATM decision-making through the
development of data-driven models that can better assess and predict system performance
and facilitate the design and implementation of decision-support tools. Several studies have
already tapped into this opportunity. For example, trajectory data analytics has been par-
ticularly useful to unravel spatial patterns in air traffic for a wide range of ATM applications.
Yet, existing approaches are still limited in their capabilities.
The main contributions of this thesis are:

1. Development of a flight trajectory data analytics framework for high-fidelity character-


ization of air traffic flows

This thesis develops a data analytics framework to exploit large-scale aircraft track-
ing data in order to discover both spatial and temporal patterns in aircraft movement
and enable a high-fidelity traffic flow characterization. The framework has three main
capabilities: (1) identification of spatial trajectory patterns; (2) identification of air
traffic flows; (3) identification of temporal patterns in air traffic flows. In order to
provide these capabilities, the framework is composed of three modules that perform a
sequential application of machine learning methods. In the first module, a trajectory
clustering scheme enables the identification of spatial trajectory patterns, which define

53
the as-flown route structure in the airspace of interest. In the second module, a tra-
jectory classification scheme enables the assessment of flight trajectory conformance
against the learned airspace structure and the identification of air traffic flows. Finally,
in the third module, a clustering analysis at temporal scale enables the identification
of traffic flow patterns. With these capabilities, the framework allows to automatically
learn the airspace structure, assess the use of the airspace and identify patterns of
usage of the airspace.

2. Development of systematic descriptive approach for metroplex operational behavior and


performance assessment

Through the use of the flight trajectory data analytics framework to characterize air
traffic flows in the terminal area of multi-airport systems, we develop a systematic de-
scriptive approach to analyze metroplex airspace design and utilization and to assess
operational performance. Novel quantitative metrics are created to summarize metro-
plex performance in the areas of efficiency, capacity and predictability. The approach
is used to perform a comparative analysis of terminal area operations in three relevant
multi-airport systems of different regions of the world: New York, Hong Kong and
Sao Paulo. As a result of the analysis, we gain better insights about structural and
operational factors driving metroplex performance, and highlight different areas of ac-
tion to be considered at each multi-airport system towards improving their air traffic
operations.

3. Development of predictive and prescriptive models for metroplex configuration and ca-
pacity management

Through the use of the flight trajectory data analytics framework to exploit patterns
in the metroplex traffic flows, we identify recurrent utilization patterns of runways and
airspace and learn the major configuration modes in which the metroplex collectively
operates as a system as well as key intervening factors. We leverage this knowledge to
develop a data-driven end-to-end approach for metroplex configuration and airport ca-
pacity planning. First, predictive models are developed to deliver probabilistic forecasts
of the metroplex configuration and airport capacity for strategic planning horizons. As

54
they incorporate operational interdependency aspects that arise in metroplexes and
that are not captured by existing approaches, our predictive models are better suited
for this more complex operational environment. Second, we demonstrate how these
predictive models can be used to support the design of strategic traffic flow manage-
ment plans. We develop a novel capacity allocation model that directly incorporates
the probabilistic capacity predictions to prescribe optimal AARs for strategic traffic
flow management based on an user-defined robustness level.

The contributions of the thesis in the context of the ATM decision-making process previ-
ously introduced are illustrated in Figure 1-9. We take a "reverse engineering" approach by
developing a data analytics framework to exploit flight trajectory data and provide a high-
fidelity characterization of air traffic flows from actual operations. This characterization
provides the foundation for obtaining new insights about the traffic behavior and perfor-
mance as well as for the development of predictive models to support airspace management
and traffic flow management decisions.

Figure 1-9: View of thesis contributions in the context of the ATM decision-making process.

55
1.6 Organization of the Thesis
The thesis is organized as follows. Chapter 2 describes the flight trajectory data analytics
framework for characterization of air traffic flows. We present the data processing techniques
and machine learning methods used in each module of the framework. The methodology is
demonstrated with flight tracking data for arrival flights at LaGuardia airport (LGA). Chap-
ter 3 describes the application of the framework for the characterization of air traffic flows
in the terminal area of multi-airport systems and demonstrates its use as a descriptive tool
to analyze metroplex airspace design and utilization and to assess operational performance.
The results of the detailed comparative analysis between New York, Hong Kong and Sao
Paulo metroplex operations are discussed. Chapter 4 describes the development of the data-
driven approach for metroplex configuration and airport capacity planning based on the
knowledge derived from trajectory data analytics. First, we describe the predictive models
for traffic flow pattern and capacity prediction, presenting the datasets used, machine learn-
ing algorithms tested and predictive performance analysis results. Subsequently, we present
the optimization model for arrival rate prescription and discuss its impacts in the planning
of GDPs. Finally, Chapter 5 summarizes the major achievements and discusses potential
future research directions.

56
Chapter 2

Flight Trajectory Data Analytics


Framework for Characterization of Air
Traffic Flows

In this chapter, we describe the flight trajectory data analytics framework for characterization
of air traffic flows. The framework has three main capabilities in order to provide a high-
fidelity traffic flow characterization: (1) it enables the identification of spatial patterns of
aircraft movement; (2) it enables the identification of air traffic flows; (3) it enables the
identification of temporal patterns in air traffic flows. With these capabilities, the framework
can be used to automatically learn the airspace structure, assess the use of the airspace and
identify patterns of usage of the airspace.

2.1 Overview of Methodological Approach


In order to provide these capabilities, the framework is composed of three modules that use
machine learning techniques to mine spatial and temporal patterns in flight tracking data.
The framework is illustrated in Figure 2-1. In the first module, clustering at spatial scale is
performed with a trajectory clustering scheme to identify spatial patterns of aircraft move-
ment, which are referred to as trajectory patterns and define the as-flown route structure
in the airspace of interest. Based on this knowledge, the second module uses a trajectory

57
classification scheme to match new flight trajectories with the learned airspace structure and
classify them as conforming (to one of the learned routes) or non-conforming. With trajec-
tories classified, flows are identified as temporally associated flight trajectories conforming
to the same standard route. The last module of the framework seeks to discover temporal
patterns in air traffic flows. For this, a clustering analysis at temporal scale is performed
to identify similar structures in the set of flows observed during a period of time, which are
referred to as traffic flow patterns. A detailed description of each module of the framework
is provided next.

Figure 2-1: Schematic overview of flight trajectory data analytics framework for characteri-
zation of air traffic flows.

2.2 Dataset Description


The framework aims at providing a high-fidelity traffic flow characterization by exploiting
flight tracking data. This type of data is continuously generated by surveillance systems
responsible for tracking aircraft through the airspace and enabling real-time monitoring and
control of the traffic by ANSPs around the world. In the U.S., the FAA uses the Traffic Flow

58
Management System (TFMS) for real-time traffic monitoring. The TFMS displays and logs
fused flight track records from the network of domestic radars. TFMS records are logged
every minute and provide historical flight trajectory data in the domestic airspace. The data
fields include flight ID, latitude, longitude, altitude, speed, origin airport, destination airport
and aircraft type.
With the advent of Automatic Dependent Surveillance - Broadcast (ADS-B), flight tra-
jectory data can now be more easily obtained through public repositories. ADS-B is a new
type of surveillance technology that relies on aircraft avionics, a constellation of GPS satel-
lites, and a network of ground stations. Briefly explained, aircraft position determined by
on-board satellite navigation systems can be broadcasted by the ADS-B transponder on air-
craft and then picked up by ADS-B receivers on the ground. Platforms such as FlightAware
and FlightRadar24 have a huge network of crowdsourced ADS-B receivers around the world
that receive flight information from ADS-B equipped aircraft and send this information to
their servers in order to provide open-source live flight tracking.
In order to demonstrate each module of the data analytics framework described in this
chapter, we used a sample of trajectory data for arrival flights at LaGuardia airport (LGA).
The sample consists of a set of 16 days of flight tracks from the TFMS.

2.3 Clustering at Spatial Scale: Trajectory Clustering


In the first module of the framework, a trajectory clustering scheme is developed to identify
spatial patterns of aircraft movement. Clustering is an unsupervised learning method that
aims at identifying groups of similar observations in a dataset without prior knowledge about
the existence of these groups or about how the observations are distributed among them.
In the trajectory clustering problem, the goal is to find groups of similar trajectories in the
spatial dimension. We define a group of spatially similar trajectories as a trajectory pattern.
In a general perspective, any clustering methodology requires:

• a data representation;

• a similarity/distance function;

59
• a clustering method.

The dataset representation will ensure the alignment with the clustering goals and the
use of an appropriate format for the subsequent step of similarity assessment. The choice of
similarity/distance function is many times driven by the data representation, but it can also
be determined by the choice of clustering method, as some algorithms require a particular
function to be used. In the latter case, the data representation should satisfy the similar-
ity/distance function requirements. Let us consider the general trajectory clustering problem
to exemplify the interaction between data representation and choice of similarity/distance
function. The trajectory of a moving object is usually given by a time-series of positions
in a two or three-dimensional space. The majority of trajectory datasets has the following
characteristics:

• Different lengths: trajectories may have different number of data points depending on
the duration of the object movement;

• Different sampling rates: the time interval between consecutive points may vary be-
tween trajectories (e.g., the collection of data may be performed every 1 s for one
object and every 2 s for another object) or within the same trajectory (e.g., the sensor
may fail to collect the data for some time period).

Because of these characteristics, the assessment of similarity between trajectories is the


first challenge that arises in the trajectory clustering process. As most of the distance
functions deal with vectors of the same length, one approach is to convert the variable-length
trajectory data into a vector of fixed dimension by constraining the trajectory length of and
using resampling mechanisms. One con of this approach is that we may lose the smoothness
in the original trajectory data if trajectories vary too much in length and the choice of vector
size is significantly smaller than the length of the longer trajectories. Another approach is
to use more complex similarity/distance functions that can handle different lengths. For
example, Vlachos et al. [106] created similarity functions based on the Longest Common
Subsequence (LCS) to discover similar multidimensional trajectories. The LCS first appeared
in the area of string matching and its basic idea is to find the longest common subsequence

60
of two sequences of different lengths by allowing them to stretch without rearranging the
sequence of the elements but allowing some elements to be unmatched [106]. Zhu et al. [107]
used the Fréchet distance (commonly used to measure the dissimilarity between curves) to
assess the dissimilarity between trajectories of different lengths. One major drawback of
using more complex distance functions is the increase in computational time.
The final requirement in the clustering methodology is the method for grouping simi-
lar observations. A number of clustering algorithms exist to perform this task and their
selection is primarily determined by the characteristics of the data in analysis. In general,
they can be categorized in four classes: hierarchical, partitioning, density-based and grid-
based. Hierarchical algorithms create a hierarchical decomposition of the data space based
on similarity/dissimilarity measures between observations/groups. This decomposition is
represented by a tree, also called dendrogram, which progressively splits the database into
smaller subsets from the root to the leaves (divisive approach) or merges individual subsets
into groups from the leaves to the root (agglomerative approach) [108]. The clusters are
then extracted using a distance threshold to horizontally cut the dendrogram. Partitioning
algorithms creates partitions of the database based on the desired number of clusters. They
start with an initial random partition and then iteratively optimize an objective function
such as minimizing the intra-cluster distance. K-means is the most popular algorithm in the
partitioning category [109]. Density-based clustering algorithms are based on the fact that
clusters can be determined by high-density regions separated by low-density regions in the
data space. They are highly used in spatial databases and have the advantage of handling
clusters of arbitrary shape. Finally, grid-based methods divide the data space into a finite
number of cells and perform the clustering based on the population of each cell.
Most of the clustering algorithms produce as output a partition of the given dataset.
However, in many cases, the dataset is not entirely structured. There may be a significant
portion that is unstructured and that can affect the clustering of the structured portion.
Usually, this unstructured part is labeled noise or outlier, although it is not necessarily
composed of observations that deviate too much from the rest of the dataset as in the
common outlier definition. Therefore, the clustering analysis should take into account the
existence of this unstructured part in order to produce reasonable clusters that represent

61
the underlying patterns of the dataset. This can be done either by applying noise detection
algorithms prior the clustering or using clustering algorithms that automatically handles
noise.
The data representation, distance function and clustering method used in the flight tra-
jectory clustering problem are presented below.

2.3.1 Trajectory Representation

A data resampling approach is used for representing the flight trajectories with feature
vectors of fixed dimension. The raw flight trajectory data are time-series of aircraft position
sampled every minute from the origin to the destination airport. First, a filtering procedure
is implemented to extract the portion of the trajectory associated with the scale of operation
under analysis. For the terminal area scale, we consider the trajectory information between
the airport runway threshold and the terminal area boundary, which is modeled as a circle of
60-mile radius with its center at the airport. The filtered flight trajectories are characterized
by time-series of different lengths, depending on the time spent in the airspace regions under
consideration. A data resampling is performed to transform each time-series into a high-
dimensional feature vector of fixed dimension. The resampling approach normalizes the
time stamps for each trajectory into the interval ⌧ = [0, 1], divides ⌧ into a fixed number of
equally sized time intervals and linearly interpolates the spatial position for the fixed number
of normalized time stamps in ⌧ . The result is a feature vector of 2D spatial position evenly
spaced in time:

Fi = (xi1 , yi1 , xi2 , yi2 , ..., xin , yin )

It is worth mentioning that the flight trajectory dataset is not characterized by very large
differences in length and a fair representation of the trajectory can be achieved with resam-
pling. Figure 2-2 exemplifies the post-processed trajectory resulting from this procedure.

2.3.2 Trajectory Similarity Assessment

To allow each feature of the trajectory vector to equally contribute to the distance measure,
the trajectory vectors are standardized so that each individual dimension is centered to have

62
Figure 2-2: Resampled flight trajectory data.

mean 0 and scaled to have standard deviation 1, as follows:

xin x̄n
x̂in = (2.1)
n

where xin is the value of feature n for trajectory i, x̂in is the value of feature n for
trajectory i after standardization, x̄n and n are the sample mean and standard deviation
of feature n.
With trajectory vectors of equal dimension and standardized, Euclidean distance is used
to determine the similarity between flight trajectories:
v
u n
uX
d(Fi , Fj ) = kFi Fj k2 = t (xik xjk )2 (2.2)
k=1

2.3.3 Trajectory Clustering Method

A density-based clustering algorithm - Density-Based Spatial Clustering of Applications


with Noise (DBSCAN) [110] - is used for flight trajectory clustering. As the name of the
algorithm suggests, this method is suitable for datasets with background noise. In flight
trajectory datasets, the standard routes and adaptions produce the core underlying patterns,
yet abnormal trajectories can also occur for a variety of reasons and can be considered
noise. DBSCAN enables the identification of the core trajectory patterns in the presence of
abnormal trajectory profiles. Other advantages of DBSCAN include the ability to discover
non-convex clusters and no need to set the number of clusters a priori.
DBSCAN relies on two input parameters in order to cluster the data space:

• M inP ts: a minimum number of points (observations);

63
• ": a distance threshold.

The algorithm is built on three fundamental concepts (illustrated in Figure 2-3):

1. "-neighborhood

The "-neighborhood of an observation Fi contains all observations that are within a


distance " and it is defined as:

N" (Fi ) = {Fj 2 D/d(Fi , Fj )  ", d(Fi , Fj ) = kFi F j k2 } (2.3)

2. Density-reachability

An observation Fj is directly density-reachable from an observation Fi if:

• Fj 2 N" (Fi )

• |N" (Fi )| M inP ts (core point condition)

An observation Fj is density-reachable from an observation Fi if there exists a chain


Fi , ..., Fj such that each observation is directly density-reachable from the predecessor.

3. Density-connectivity

An observation Fj is density-connected to an observation Fi if there exists another


observation Fk such that both Fj and Fi are density-reachable from Fk .

Based on these concepts, the clustering process works as follows. The algorithm starts
with an arbitrary instance of the database and determines its "-neighborhood. If it contains
at least M inP ts (core point condition), the observation and its neighbors start a cluster. The
"-neighborhood of the neighbors is iteratively retrieved and the same procedure is applied
until all the border points of the cluster are achieved. Otherwise, the observation is labeled
as noise and the algorithm visits the next instance of the database. It is worth mentioning
that an instance labeled noise can later become part of a cluster if it is reachable from a new
discovered core point. A cluster is then defined as a set of density-connected points.

64
Figure 2-3: Illustration of DBSCAN concepts.

The clustering solution provided by DBSCAN is dependent on the input parameters


M inP ts and ". DBSCAN is not particularly sensitive to the M inP ts choice, which will
determine the smallest number of observations to define a cluster. Once M inP ts is chosen,
the choice of " will involve a trade-off between undesirable creation of clusters from minor
variations in a major cluster and undesirable merger of distinct clusters.
Given the absence of a correct answer for establishing a baseline for comparison, the
clustering process is inherently subjective and its success is primarily determined by the
meaningfulness and interpretability of structural patterns extracted from the data. As an
attempt to provide a quantitative measure to evaluate the clustering solution quality, several
clustering validity indices have been developed and can be found in the literature. A detailed
description and comparison of some of these indices is provided in [111]. In general, these
different metrics evaluate similar properties such as the compactness and separability between
the clusters. A good clustering solution is usually associated with high levels of compactness
and separability between the clusters. However, in a practical perspective, these indices
should be primarily used as a validation mechanism rather than a tool to blindly select the
number of clusters.
In this work, analysis of Silhouette plots [112] was used as a decision aid in the cluster-

65
ing process on a case-by-case basis. The Silhouette Index provides a quantitative measure
about the correctness of assignment of an observation to a cluster. Let a(i) be the average
dissimilarity of observation i to the observations in the same cluster, d(i, C) be the average
dissimilarity of observation i to the observations in a different cluster C and b(i) be the
minimum d(i, C), the Silhouette Index s(i) of observation i is then given by Eq. 2.4. From
the Silhouette Index formulation, it is possible to note that it varies between -1 and 1: values
close to 1 indicate the observation is well clustered, values close to -1 indicate the observation
is probably assigned to the wrong cluster.
8
>
>
>
> 1 a(i)/b(i) if a(i) < b(i)
>
< b(i) a(i)
s(i) = 0 if a(i) = b(i) ) s(i) = (2.4)
>
> max{a(i), b(i)}
>
>
>
:b(i)/a(i) 1 if a(i) > b(i)

2.3.4 Example Application

This section shows the results of the application of the trajectory clustering methodology to
the sample data of LGA arrival trajectories.
The sensitivity of the DBSCAN solution with respect to the parameters M inP ts and "
is shown in Figure 2-4. It is noticed the percentage of noise observations is not particularly
sensitive to M inP ts, but very sensitive to ". As " increases, more observations are captured
in the clusters.
The average Silhouette Index for the clustering output obtained with different parameter
settings is shown in Figure 2-5. For smaller values of ", the value of the metric is closer
to zero, indicating a lower clustering solution quality. Figure 2-6 suggests this is driven by
the creation of clusters from minor variations in the data, as revealed by the high number
of clusters when " is small. As " increases, the Silhouette Index increases and starts to
stabilize. Between " = 1 and " = 1.2, the metric stabilizes at its highest level. After that
point, increasing " generates a small degradation in the clustering solution quality. Figure
2-6 reveals this degradation is caused by merger of distinct clusters.
Based on these observations, we used M inP ts = 6 and " = 1.2 to cluster the sample
data. The clustering output for the chosen parameters as well as the detailed Silhouette

66
Figure 2-4: Percentage of noise associated with the clustering output obtained with different
parameter settings (M inP ts, ").

values for each observation in each cluster are shown in Figure 2-7.

2.4 Trajectory Classification


In the second module, a trajectory classification scheme is developed to match new flight
trajectories with the airspace structure identified in the first module, including the ability
to identify non-conforming behaviors (those that do not conform with one of the identified
trajectory patterns). The purpose of the trajectory classification module is two-fold: 1) given
the computational effort associated with the trajectory clustering process, the classification
procedure provides a more efficient way for processing large sets of trajectory data for spatial
pattern identification during offline applications; 2) it is key for online applications that
depend on monitoring of airspace use, as it enables consistent processing of new batches of
trajectory data continuously generated by the surveillance system.
In the trajectory classification problem, the classes are defined by the trajectory patterns
learned with the spatial clustering analysis (training dataset). Given a new flight trajectory
defined by its vector of spatial position, the goal is to predict whether it conforms to one of
the learned trajectory patterns. Obviously, not all flight trajectories will match the identified
airspace structure and the classification scheme should also have the ability to detect these

67
Figure 2-5: Silhouette Index associated with the clustering output obtained with different
parameter settings (M inP ts, ").

non-conforming behaviors.
Differently than clustering, classification is a supervised learning method that aims at
predicting the class/label of an observation given a set of predictors/features based on prior
knowledge extracted from a training dataset [113]. Figure 2-8 provides a schematic represen-
tation of the supervised learning classification process. Given a training set Sn = {(xi , yi ), i =
1, ..., n}, where xi 2 Rd is a feature (input) vector and yi 2 {1, 2, 3, ..., C} is the correspond-
ing label (output) in a multi-way classification problem with C labels, a classifier h is defined
as a mapping h : Rd ! {1, 2, 3, ..., C}. The classifier outputs a decision boundary (linear or
not) that partitions the input space and can therefore classify new instances in a test set.
A fair classifier is expected to replicate its performance in the training set to the test set
and therefore shows both low training (Eq. 2.5) and generalization errors (Eq. 2.6). In order
to achieve these objectives and avoid over-fitting (when a low training error is obtained but
it does not translate into a low generalization error), cross-validation should be performed
during the training phase in order to adjust the settings of the classifier. In this work, we used
the k-fold cross-validation method, with k = 5, for this purpose. It consists of partitioning
the training dataset into k different folds and using (k 1) folds to train the classifier and
one fold to evaluate its performance in k different runs (each run uses a different fold for
validation).

68
Figure 2-6: Silhouette Index versus number of clusters identified.

n
1X
"(h) = Jh(xi ) 6= yi K (2.5)
n i=1

n+n 0
1 X
0
" (h) = 0 Jh(xi ) 6= yi K, n0 ! 1 (2.6)
n i=n+1

2.4.1 Matching Flight Trajectories with Learned Trajectory Pat-


terns

The trajectory classification process starts with the development of a classifier that can
successfully discriminate the learned trajectory patterns. The dataset of learned trajectory
patterns was partitioned into 80%/20% subsets for training and testing the classifier.

Multi-way Trajectory Classification Model

A multi-way classifier using the Random Forests algorithm was created. Random Forests
was selected because it is widely recognized as one of the highest performing methods in
terms of classification accuracy and because it runs efficiently on large databases.
Random Forests is a non-parametric method for both classification and regression consist-
ing of an ensemble of decision tree learners [114]. A decision tree is defined by a hierarchical
structure of decision nodes that progressively partitions the input space using a sequence of

69
Figure 2-7: (a) Clusters of arrival trajectories. (b) Silhouette plot.

binary splits from the root node to the leaves (terminal nodes). Each decision node evaluates
the value of a feature from the input set to decide which branch will be taken from it. The
goodness of a split is quantified with an impurity measure. In a classification setting, the
Gini index is typically used. Given a training set with n observations and nk observations in
a decision node k after a sequence of splits, with nk,c observations belonging to class c such
P
that c nk,c = nk , the Gini index measures the impurity of node k as follows:

X
Gk = 1 p2k,c (2.7)
c

where pk,c = nk,c /nk is an estimate for the probability of belonging to class c.
It follows that the minimum value of Gk is zero and it is achieved when the node is pure,
whereas the maximum value of Gk is 1 1/C and it is achieved when the classes c = 1, ..., C
have equal probabilities. At each decision node, the method searches for the partition that
will minimize the impurity among all input variables and their potential splits. The splitting
process is repeated until a stopping criterion is met (e.g., node is pure, minimum number
of observations is achieved). The result is a rectangular partition of the input space. Every
leaf node l then corresponds to a rectangular subspace Rl . For every training observation
xi , there is only one leaf node such that xi 2 Rl . The prediction for a new observation x0 is

70
Figure 2-8: Supervised learning classification process.

obtained by applying a majority rule for the leaf l(x0 ) that this observation falls into when
it is passed through the tree:

X
ŷ 0 = arg max I(c = yi ) (2.8)
c
i:xi 2l(x0 )

Figure 2-9 illustrates a classification tree, based on a toy dataset characterized by a two-
dimensional input space and two labels. In this fairly simple example, the resulting decision
boundary (solid black line) perfectly partitions the input space into local regions, and can
therefore classify new instances.

Figure 2-9: Example of dataset and resulting decision tree for classification.

As any ensemble method, Random Forests creates multiple decision trees with bootstrap

71
samples of the training data, but it has the unique characteristic of selecting only a random
subset of the features to determine each split. The final response of the model is computed
by aggregating the results of the individual trees:

T
X
0
ŷbag = arg max I(c = ŷt0 ) (2.9)
c
t=1

where ŷt0 is the prediction from tree t in the ensemble.


The algorithm has few parameters to tune: T is the number of trees in the ensemble, M is
the number of features to sample at each split and N is the minimum number of observations
in a leaf node. Based on cross-validation performance, the Random Forests classifier was
created with T = 50, M = 10 and N = 1.

Multi-way Trajectory Classification Performance

The Random Forests trajectory classification accuracy achieved on the test data of LGA
arrival trajectories was 99%, revealing that the classifier is able to precisely distinguish the
spatial patterns learned with trajectory clustering.

2.4.2 Identifying Non-Conforming Trajectories

The Random Forests classifier successfully discriminates the spatial patterns between the
conforming trajectories. However, given a new flight trajectory, it is not known a pri-
ori whether it conforms to the standard routes. In order to enable the identification of
non-conforming behaviors, we developed a Conformal Prediction semi-supervised anomaly
detection framework.
Anomaly detection is an important problem that has been researched within various areas
and application domains. Chandola et al. [115] defines an anomaly as "a pattern that does
not conform to expected normal behavior". Anomaly detection approaches typically operate
in three settings: supervised, semi-supervised and unsupervised. In supervised mode, a
training dataset with representative instances of normal and anomalous behavior is used to
create a predictive model that can discriminate between them. In semi-supervised mode,
representative instances of normal behavior are used for creating a model for the normal

72
data, which is then used for identifying anomalies. Finally, in unsupervised mode, a training
dataset is not required; techniques in this category make the implicit assumption that normal
instances are much more frequent than anomalies.
We assume that the trajectory patterns identified with the spatial clustering analysis
are representative of the normal trajectory behavior in the airspace region analyzed. The
trajectory classifier developed in the previous section models the normal behavior. Therefore,
it can also be used to identify non-conforming behaviors, if a measure of confidence for its
predictions can be obtained. Conformal Prediction provides such measure rigorously, as
described below.

Non-conforming Trajectory Detection Model

Conformal Prediction is a technique that intends to generate confidence intervals for pre-
dictions made by machine learning algorithms [116]. Therefore, it is built upon existing
classification and regression methods. The concept involves the estimation of a p-value for
each candidate label that can be attributed to a test example based on a non-conformity
metric, which measures how different this observation is relative to a set of examples that
have the candidate label. The labels having a p-value less than the specified significance
level " are then excluded from the prediction set.
Formally, given a training dataset Z : (x1 , y1 ), ..., (xn 1 , yn 1 ), where xi 2 X is a vector
of attributes and yi 2 Y is the class label for observation i, non-conformity scores ↵i can be
calculated for each observation based on non-conformity functions. These non-conformity
functions are typically tailored to the machine learning algorithm used to create the mapping
between the input space and the output. Since Random Forests was the underlying classifi-
cation method for conformal prediction, a non-conformity score based on the proportion of
tree votes in the ensemble was used [117, 118, 119]:

|treej 2 ensemble : yi = ŷij |


↵i = 1 (2.10)
|ensemble|
For a new observation zn = (xn , ŷn ) for which we want to predict the class label, a non-
conformity score ↵nŷn can be calculated for each tentative class label ŷn and compared with

73
the non-conformity scores of the training dataset in order to generate an associated p-value:

|zi 2 Z : ↵i ↵nŷn | + 1
p(ŷn ) = (2.11)
|Z| + 1
The p-value is a measure of confidence in the prediction. The higher the p-value, the
higher the confidence. For a significance level ", if p ", ŷn becomes part of the prediction
set with confidence 1 ".
Non-conforming flight trajectories can be identified when the p-value is lower than the
significance level for all possible labels, i.e., when the prediction set is empty. In contrast
to previous methods, a key property of conformal anomaly detection is that it provides a
well-founded approach for tuning the confidence thresholds for anomaly detection that can
be directly related to the expected or desired false alarm rate [120].

Non-conforming Trajectory Detection Performance

An anomaly detection framework should seek to achieve a low rate of false anomaly discovery
(false alarm rate) and a high rate of anomaly detection (true positive rate). Ideally, we
would like to have a dataset of labeled non-conforming arrival trajectories to evaluate the
performance of our model not only in terms of false alarm rate but also true positive rate. In
the absence of this labeled data, we used the noise trajectories identified by DBSCAN during
the spatial clustering process, as the noise typically includes the non-conforming behaviors.
We compared the performance of the Conformal Prediction (CP) non-conforming trajec-
tory detection approach with two other approaches commonly used for anomaly detection.
The first one is a nearest neighbor based approach. A classifier is created to discriminate the
trajectory patterns using the K-Nearest Neighbors (KNN) method. For a new trajectory i,
the KNN based model retrieves the k-closest neighbors on the training dataset of trajectory
patterns using Euclidean distance and computes the posterior probability of belonging to
trajectory pattern j as follows:

P
m2M (i) ⇡j Jym = jK
p(j/i) = PK P (2.12)
j=1 m2M (i) ⇡j Jym = jK

where M (i) is the set of k nearest neighbors of trajectory i, ⇡j is the prior probability of

74
trajectory pattern j and ym is the known trajectory pattern of neighbor observation m.
A non-conforming trajectory is then identified when the maximum posterior probability
is smaller than an user-specified threshold.
The second is a statistical approach. We model the learned trajectory patterns using a
Gaussian Mixture Model (GMM), where each component describes one trajectory pattern.
GMM is a parametric model that assumes the probability density function of the data is
given by a weighted sum of Gaussian component densities:

K
X
P (x; ✓) = ⇡j N (x; µj , ⌃j ) (2.13)
j=1

where x is the set of feature vectors in the training dataset, K is the number of com-
ponents in the mixture, ⇡j are the component weights and N (x; µj , ⌃j ) are the component
Gaussian densities, with mean µj and covariance matrix ⌃j .
The posterior probability that a new observation (trajectory) i has been generated from
component (trajectory pattern) j is given by:

⇡j N (x(i) ; µj , ⌃j )
p(j/i) = PK (2.14)
(i)
j=1 ⇡j N (x ; µj , ⌃j )

As in the KNN based approach, a non-conforming trajectory is identified when the max-
imum posterior probability is smaller than an user-specified threshold.
We evaluated the performance of the three non-conforming trajectory detection models
on the test data of LGA arrival trajectories. We used k = 30 in the KNN model, based on
cross-validation performance. The GMM model was created with 16 components (number of
learned trajectory patterns). In order to detect non-conforming trajectories, a significance
level " = 0.05 was used in the CP model. Since the significance level is directly related to the
expected false alarm rate in CP, we calibrated the thresholds for non-conforming trajectory
detection in the KNN and GMM models in order to achieve false alarm rates close to 5%.
The following performance metrics were calculated and compared:

True positive rate - TPR (or sensitivity or recall)

75
tp
TPR = (2.15)
tp + f n
where tp is the number of true positives (actual non-conforming trajectories predicted as
non-conforming) and f n is the number of false negatives (actual non-conforming trajectories
predicted as conforming).

False positive rate - FPR

fp
FPR = (2.16)
f p + tn
where f p is the number of false positives (actual conforming trajectories predicted as
non-conforming) and tn is the number of true negatives (actual conforming trajectories pre-
dicted as conforming).

Positive predictive value - PPV (or precision)

tp
PPV = (2.17)
tp + f p

Accuracy - ACC

tp + tn
ACC = (2.18)
tp + tn + f p + f n

F1-score - F1

PPV ⇤ TPR
F1 = 2 ⇤ (2.19)
PPV + TPR

76
The performance results are shown in Table 2.1. First, it is observed the CP model with
significance level " = 0.05 indeed led to FPR < 5%, as expected. For similar levels of
FPR, the CP model outperformed the others in detecting non-conforming trajectories, as it
showed the highest TPR. As a result, the CP model also showed the highest precision and
accuracy. In summary, the CP model achieved the best balance between precision and sensi-
tivity, as measured by the F1 score, and therefore was selected for non-conforming trajectory
detection.

Table 2.1: Non-conforming trajectory detection performance.

Model CP KNN GMM


TPR(%) 69.5 49.4 17.3
FPR(%) 4.3 5.0 5.0
PPV(%) 73.0 61.4 35.6
ACC(%) 92.1 88.6 84.1
F1(%) 71.1 54.7 23.2

2.4.3 Flow Identification

The trajectory classification scheme enables the processing of new sets of data for spatial
pattern identification. Once trajectories are classified, flows can be identified as temporally
associated flight trajectories conforming to the same standard route. For this, the output of
the trajectory classification is organized and stored in daily flow matrices W 2 Rn⇥p , with
n rows (number of trajectory patterns) and p columns (number of time periods during the
day). Each matrix element wij indicates the number of trajectories conforming to trajectory
pattern i during time period j. A flow is identified whenever wij > 1. The matrix columns
are vectors {Rt 2 Rn , t = 1, ..., p} that indicate the time-dependent traffic flows. An example
of daily flow matrix is shown in Figure 2-10. For this particular date, two major changes
were observed in the arrival traffic flow structure at LGA.

77
Figure 2-10: Flow matrix for LGA arrivals on May 24, 2013.

2.5 Clustering at Temporal Scale: Time-Dependent Flow


Vector Clustering
The third module in the data-driven framework has the goal of identifying patterns in the
traffic flow structure (group of traffic flows in a given time period). For this, a second-layer
clustering analysis is performed on the set of time-dependent traffic flows Rt . The vector
of time-dependent traffic flows is transformed to a time-dependent characteristic flow vector
F Vt to represent the traffic flow structure:

F Vt = f (Rt )

78
The flow vectors are clustered to identify traffic flow patterns. Since the flow vector
representation is tailored to the application, the details about the clustering process are
presented in the next chapter.

79
THIS PAGE INTENTIONALLY LEFT BLANK

80
Chapter 3

Comparative Analysis of Terminal Area


Operations in Multi-Airport Systems

In this chapter, we demonstrate how the flight trajectory data analytics framework for char-
acterization of air traffic flows developed in this thesis can be used to obtain a deeper under-
standing of metroplex operational behavior and performance based on a systematic approach
backed by operational data. For this, we performed a comparative analysis of terminal area
operations in three representative multi-airport systems of the global air transportation sys-
tem: New York, Hong Kong and Sao Paulo. Through the characterization of terminal area
air traffic flows, we identified structural and operational differences between these systems
and evaluated the impacts of such differences in the key ATM performance areas of efficiency,
capacity and predictability.

3.1 Case Studies


For the purpose of this research, a multi-airport system was defined as a set of two or more
significant airports that serve commercial passenger traffic in a metropolitan region. An
airport was defined as significant if it serves at least 5,000,000 passenger movements per
year and if it is located within 50 miles from the center of the major city in the metropolitan
region.
Three relevant multi-airport systems of different regions of the world were selected for

81
the comparative analysis: New York, Hong Kong and Sao Paulo. All three metropolitan
regions are served by three significant commercial airports: New York is served by John F.
Kennedy International (JFK), Newark International (EWR) and LaGuardia (LGA); Hong
Kong is served by Hong Kong International (HKG), Shenzhen Bao’an International (SZX)
and Macau International (MFM); Sao Paulo is served by Sao Paulo/Guarulhos International
(GRU), Sao Paulo/Congonhas (CGH) and Viracopos International (VCP). Figure 3-1 shows
the location of the significant airports in each metroplex and the distances between them.

Figure 3-1: Location of airports in the New York, Hong Kong and Sao Paulo metroplexes.

The detailed passenger and aircraft movement statistics for the year 2016 [121, 122, 123,
124, 125, 126, 127] at these three multi-airport systems is presented in Figures 3-2 and 3-3.
The New York system is the busiest, followed by Hong Kong and Sao Paulo. The Hong
Kong system has the highest ratio between passenger movements and aircraft movements,
indicating higher participation of larger aircraft in its mix of operations.
Figure 3-4 shows the lateral trajectories of arrival and departure flights during one day
of terminal area operations in these regions and provides a good illustration of the air traffic
density and complexity that characterize metroplex systems.

82
Figure 3-2: Passenger movement by airport at the New York, Hong Kong and Sao Paulo
metroplexes in 2016.

Figure 3-3: Aircraft movement by airport at the New York, Hong Kong and Sao Paulo
metroplexes in 2016.

Figure 3-4: Arrival and departure trajectories for one day of terminal area operations in the
New York, Hong Kong and Sao Paulo multi-airport systems.

83
3.1.1 New York Multi-Airport System

The multi-airport system that serves the New York metropolitan region is composed of three
1
primary commercial airports: JFK, EWR and LGA. Together, they served 128.9 million
passengers in 2016, being the world’s second busiest multi-airport system, after London
[128]. Given the high traffic volume (which is accentuated by the presence of other smaller
airports serving the demand for business and charter flights such as Teterboro airport) and
the proximity between the airports, the New York metroplex is often recognized as one of
the most operationally complex multi-airport systems in the world.
ATC services are provided to the major commercial airports by the individual Air Traffic
Control Towers (ATCT) and one consolidated TRACON (N90), which also provides services
to other smaller airports in the New York metropolitan region. ATCT and TRACON rely
on "Letters of Agreement" to use specific operating procedures and delegate airspace regions
that need to be shared [2]. As an example, Figure 3-5 shows some airspace regions that are
shared between JFK and LGA.

Figure 3-5: Shared airspace regions between JFK and LGA [2].

In terms of runway infrastructure, the New York metroplex has the highest installed
capacity among the three systems, with a total of nine runways. JFK has two pairs of
parallel runways - 4L-22R, 4R-22L, 13L-31R and 13R-31L - aligned at right angles (Figure
3-6). Parallel runways 13-31 are 6,700 ft apart and parallel runways 4-22 are 3,000 ft apart.
While runways 13-31 meet the geometry criteria for independent approaches 2 , they are often
1
A primary airport was defined as a significant airport serving at least 20% of the total passenger traffic
in the multi-airport system, whereas a secondary airport was defined as a significant airport with less than
20% of traffic share.
2
In order to operate simultaneous independent instrument parallel approaches, FAA Order JO 7210.3Z

84
operated with a 2-mile stagger in order to avoid the need of additional final monitor staff
[130]. Runways 4-22 are often operated with a 1-mile stagger.

Figure 3-6: Runway infrastructure in the New York metroplex.

EWR has two parallel runways, 4R-22L and 4L-22R, and a third crosswind runway, 11-29.
The parallel runways are separated by 950 ft. Because of the small separation, they are often
operated in segregated mode, primarily using runway 4R-22L for landings and 4L-22R for
departures. Runway 11-29 is often used as an overflow arrival runway during high demand
requires a minimum separation of 3,600 ft between runway centerlines (if the airport field elevation is 2,000
ft MSL or less) and the use of a final monitor controller for each runway operating independently [129].
Simultaneous independent approaches may be conducted without final monitors if the runway centerlines
are separated by at least 9,000 ft. If parallel approaches cannot be executed independently, simultaneous
dependent operations may take place. FAA Order JO 7110.65W [129]) requires the use of diagonal separations
between each aircraft and the trailing aircraft in the adjacent runway in order to ensure aircraft remain
staggered on parallel approaches and reduce the risk of collision from deviation from the final approach
path. The diagonal separations are defined based on the distance between the runway centerlines as follows:
1.0 NM, if 2500 ft <= distance <= 3600 ft; 1.5 NM, if 3600 ft < distance <= 4300 ft ; 2.0 NM, if 4300 ft <
distance <= 9000 ft.

85
periods, but it also provides flexibility during adverse high wind conditions.
Finally, LGA has two crossing runways, 4-22 and 13-31, and typically dedicates one
runway for landings and one runway for takeoffs. Standard operating procedures dictate
that a departing aircraft can be cleared for takeoff after a preceding arriving aircraft crosses
the intersection after landing [129]. For that reason, configuration 22-13 is optimal for LGA
since the intersection is closer to the active runway thresholds, leading to the minimum
spacing to get departures out.

3.1.2 Hong Kong Multi-Airport System

Hong Kong is part of the Pearl River Delta metropolitan region and is served by two primary
commercial airports, HKG and SZX, and a secondary airport, MFM. As a major gateway
to China and Asia, HKG is one of the top-ten international airports in terms of passenger
movement and it is the world’s largest air cargo hub [128]. Together with SZX and MFM,
the Hong Kong multi-airport system served more than 100 million passengers in 2016 and it
is one of the busiest systems in the world.
Approach control services for HKG and MFM are provided by the Hong Kong Air Traffic
Control Centre [131], which is part of the Civil Aviation Department of the Hong Kong
Special Administrative Region. For SZX, approach control is provided by the Guangzhou
Air Traffic Control Centre, under authority of the Civil Aviation Administration of China.
The airport runway systems are shown in Figure 3-7 and totalize five runways for the
metroplex. HKG has two parallel runways, 07L-25R and 07R-25L, separated by 5,000 ft.
SZX also has two parallel runways, 15-33 and 16-34, separated by 5,000 ft. The large dis-
tance between the parallel runway centerlines at HKG and SZX enable them to be operated
independently. MFM is served by only one runway, 16-34.

86
Figure 3-7: Runway infrastructure in the Hong Kong metroplex.

3.1.3 Sao Paulo Multi-Airport System

The multi-airport system that serves the Sao Paulo metropolitan region is composed of two
primary commercial airports, GRU and CGH, and a secondary airport, VCP. Although VCP
is located 50 miles from the Sao Paulo city center, today it serves an important part of the
metropolitan region air travel demand given the physical limitations for capacity expansion at
the primary airports. The Sao Paulo metroplex is the busiest multi-airport system in Latin
America, with GRU and CGH being the busiest airports in Brazil in terms of passenger
movement.
Approach control services for GRU, CGH and VCP is provided by the Regional Service
of Flight Protection of Sao Paulo (SRPV-SP) through a consolidated Terminal Maneuvering

87
Area (TMA-SP).
Similarly to the Hong Kong metroplex, the airport runway systems totalize five runways
and are shown in Figure 3-8. GRU has two parallel runways, 09L-27R and 09R-27L, which are
separated by 1230 ft. The runways are typically operated dependently in segregated mode,
with runway 09R-27L dedicated for arrivals and runway 09L-27R dedicated for departures.
CGH has two parallel runways, 17L-35R and 17R-35L, separated by 700 ft. It typically
operates in a single runway configuration using the main runway 17R-35L, given the small
separation and length constraints of the auxiliary runway 17L-35R. VCP is served by only
one runway, 15-33.

Figure 3-8: Runway infrastructure in the Sao Paulo metroplex.

88
3.2 Data Description
Flight tracking data was used for the identification and characterization of air traffic flows.
For the New York metroplex region, we used a set of 69 days of radar tracks, within the pe-
riod 2013-2015, from the TFMS. For both the Hong Kong and Sao Paulo metroplex regions,
we used a set of 60 days of flight tracks, within the period 2016-2017, from the FlightRadar24
global flight tracking service [132]. The datasets report one-minute updates of aircraft state,
including flight ID, latitude, longitude, altitude, speed, origin airport, destination airport
and aircraft type. Dates were empirically selected to ensure coverage of a broad set of op-
erational conditions; 55% are fair weather condition days and 45% are weather-impacted
days (further categorized as convective weather impacts - 30% - and non-convective weather
impacts, such as adverse wind conditions or low ceiling/visibility - 15%). Historical Meteo-
rological Terminal Aviation Routine Weather Report (METAR) data was used to categorize
the weather conditions for each day of operations.

3.3 Characterization of Structural Differences through


the Analysis of Terminal Area Route Structures

3.3.1 Identification of Trajectory Patterns

The trajectory clustering methodology described in Section 2.3 was used to identify spatial
trajectory patterns in the terminal area. A subset of the data was used to reduce the
computational effort. For the New York metroplex airports, a set of sixteen days was used.
For the Hong Kong and Sao Paulo metroplex airports, given their lower number of operations,
a set of thirty days was used. As an example, Figure 3-9 shows the identified arrival clusters
for the major airport in each metroplex. The distribution of observations in the clusters is
also presented. For instance, it is observed that both HKG and GRU have one dominant
trajectory pattern that concentrates more than 30% of the observations, while JFK has a
lower concentration of trajectories in one particular pattern. It is also observed that JFK
has a lower percentage of noise observations than HKG and GRU. The clustering results for

89
the other airports are presented in Appendix A, B and C.

Figure 3-9: Clusters of arrival trajectories; each color represents one cluster; grey bar repre-
sents the percentage of noise observations.

90
A total of 50, 29 and 24 arrival trajectory clusters and 55, 28 and 36 departure trajectory
clusters were identified for the New York, Hong Kong and Sao Paulo multi-airport systems,
respectively. The centroids of the arrival and departure clusters are shown in Figures 3-10
and 3-11.

Figure 3-10: Centroids of metroplex arrival trajectory clusters.

91
Figure 3-11: Centroids of metroplex departure trajectory clusters.

At a first glance, it appears the New York metroplex has the most complex airspace
structure. First, there are more trajectory patterns, which can be explained not only because
of the higher number of runways, but also because of the presence of more than one trajectory
pattern for some combinations of arrival/departure gate and runway threshold. An example
is shown in Figure 3-12 for LGA arrival trajectories. Two trajectory patterns were identified

92
for flights entering the terminal area through the west gate and landing at runway 22, which
are driven by different maneuvers to intercept the final approach leg. Second, the trajectory
patterns are more closely located, with a higher number of lateral crossings.

Figure 3-12: Example of different trajectory patterns from arrival fix to runway threshold
22 at LGA.

3.3.2 Identification of Route Intersections

The observed lateral crossings of trajectory patterns in Figures 3-10 and 3-11 do not neces-
sarily indicate they are closely located at the vertical dimension to potentially create flow
interactions. In order to investigate the complexity of the airspace structure in terms of
the potential to create flow interactions between the airports, we introduce the concept of
trajectory tubes. A trajectory tube is defined for each spatial cluster based on the dispersion
of their trajectory members. Since all trajectories are described with the same number of
resampling points i = 1, ..., n, the tubes are discretized in n 1 parts. Each discrete part
i = 1, ..., n 1 is determined by resampling points i and i + 1. For each part i, the width
of the tube is determined by the 95th percentile of the lateral spread of all ith and (i + 1)th
resampled points around the cluster centroid. The height of the tube is determined by the
difference between the 95th and the 5th percentile of all ith and (i+1)th vertical positions. We
calculated the intersection volume between all pairs of trajectory tubes from different air-
ports. They indicate the terminal airspace regions where the metroplex trajectory patterns
intersect laterally and vertically and flows can potentially interact.

93
Table 3.1 shows the number of trajectory tube intersections identified by pair of metroplex
airports and Figure 3-13 shows their location and size. First, it is noticeable the New York
metroplex has the most conflicted airspace structure. JFK and LGA are the pair of airports
with the highest number of intersections. A significant number of the New York metroplex
route intersections are located less than 10 nautical miles from the airports (especially for
JFK and LGA). An example is shown in Figure 3-14. It displays the intersection of a LGA
arrival trajectory tube and a JFK arrival trajectory tube, as the airspace volume colored in
red. It is also observed some clusters of high volume trajectory tube intersections located
close to the terminal area boundary, revealing the sharing of arrival/departure gates.

Figure 3-13: Characterization of trajectory tube intersections.

94
Table 3.1: Number of trajectory tube intersections identified by pair of metroplex airports.

Multi-airport Airport pair Number Detailed airport pair Number


system by type of operation
JFK Arr - EWR Arr 0
JFK Dep - EWR Dep 38
JFK - EWR 42
JFK Arr - EWR Dep 4
JFK Dep - EWR Arr 0
JFK Arr - LGA Arr 34
JFK Dep - LGA Dep 61
New York JFK - LGA 130
JFK Arr - LGA Dep 29
JFK Dep - LGA Arr 6
EWR Arr - LGA Arr 17
EWR Dep - LGA Dep 65
EWR - LGA 91
EWR Arr - LGA Dep 0
EWR Dep - LGA Arr 9
HKG Arr - SZX Arr 0
HKG Dep - SZX Dep 0
HKG - SZX 12
HKG Arr - SZX Dep 12
HKG Dep - SZX Arr 0
HKG Arr - MFM Arr 2
HKG Dep - MFM Dep 10
Hong Kong HKG - MFM 28
HKG Arr - MFM Dep 14
HKG Dep - MFM Arr 2
SZX Arr - MFM Arr 10
SZX Dep - MFM Dep 13
SZX - MFM 30
SZX Arr - MFM Dep 2
SZX Dep - MFM Arr 5
GRU Arr - CGH Arr 7
GRU Dep - CGH Dep 33
GRU - CGH 53
GRU Arr - CGH Dep 9
GRU Dep - CGH Arr 4
GRU Arr - VCP Arr 0
GRU Dep - VCP Dep 2
Sao Paulo GRU - VCP 10
GRU Arr - VCP Dep 6
GRU Dep - VCP Arr 2
CGH Arr - VCP Arr 0
CGH Dep - VCP Dep 6
CGH - VCP 16
CGH Arr - VCP Dep 8
CGH Dep - VCP Arr 2

95
The Hong Kong metroplex shows the less conflicted airspace structure. The airspace
structures of the two primary airports are highly de-conflicted, with the exception of a
small cluster of intersections between HKG arrival trajectory patterns and SZX departure
trajectory patterns located in the south departure area of SZX. Most of the intersections are
located more than 10 nautical miles away from the airports. A few high volume intersections
are observed between 10 and 30 nautical miles from the airports, revealing the sharing of
route segments. An example of a shared route segment between a SZX arrival trajectory
tube and a MFM arrival trajectory tube is shown in Figure 3-15. It is also observed some
clusters of high volume intersections located close to the terminal area boundary, revealing
the sharing of arrival/departure gates, especially between SZX and MFM.
Finally, in the Sao Paulo metroplex, GRU and CGH have the highest number of inter-
sections, especially between departure trajectory patterns. Most of these intersections are
located close to the terminal area boundary, revealing the sharing of arrival/departure gates.
Figure 3-16 shows an example of departure trajectory patterns from GRU and CGH sharing
the same departure gate. A few high volume intersections are also observed between 10 and
30 nautical miles from the airports, revealing the sharing of route segments.

Figure 3-14: Example of intersection between JFK and LGA arrival trajectory tubes.

96
Figure 3-15: Example of intersection between SZX and MFM arrival trajectory tubes.

Figure 3-16: Example of intersection between GRU and CGH departure trajectory tubes.

3.4 Characterization of Operational Differences through


the Analysis of Traffic Flow Dynamics

3.4.1 Identification of Metroplex Flow Patterns

Once the metroplex airspace structures were identified, Random Forests classifiers were cre-
ated to discriminate the trajectory patterns for each type of operation (arrival/departure) at

97
each metroplex airport. All trajectory classifiers showed a multi-way classification accuracy
higher than 98%. The classifiers were used to assign trajectories in the remaining dataset
(not used for the spatial clustering) to the learned routes and the Conformal Prediction
framework was applied with a significance level " = 0.05 to identify non-conforming behav-
iors. Flow matrices were then created for each day of operations. As described in Chapter
2, the matrices were designed so that each element wij indicates the number of flights using
a particular route i during a given hour j of the day. A flow is identified whenever wij > 1.
The matrix columns are vectors {Rt 2 Rn , t = 1, ..., p} that indicate the time-dependent
traffic flows.
In order to investigate the presence of patterns in the metroplex terminal area flow
behavior, we used the third module of the trajectory data analytics framework and performed
a second-layer clustering analysis on the set of time-dependent traffic flows Rt . For this, a
data representation was established for the time-dependent metroplex flow structure. The
m,Arr | m,Dep |
columns {(Rm,Arr )t 2 R|K , t = 1, ..., p} and {(Rm,Dep )t 2 R|K , t = 1, ..., p} of
the daily arrival and departure flow matrices created for each metroplex airport m were
aggregated into flow vectors to represent the hourly terminal area flow structure in the
metroplex:
n ⇣ ⌘ o
F Vt = (Cim,Arr )t , (Cjm,Dep )t , m 2 M, i 2 A, j 2 D, t = 1, ..., p

where:
8
>
> m,Arr
if max {(Rkm,Arr )t } > 1
<arg max{(Rk )t }
k2Kim,Arr
(Cim,Arr )t = k2Kim,Arr (3.1)
>
>
:0 otherwise

8
>
> m,Dep
if max {(Rkm,Dep )t } > 1
<arg max{(Rk )t }
k2Kjm,Dep
(Cjm,Dep )t = l2Kjm,Dep (3.2)
>
>
:0 otherwise

where M is the set of metroplex airports, A is the set of arrival gates in the terminal area
boundary, D is the set of departure gates, Kim,Arr ⇢ K m,Arr is the set of arrival trajectory
patterns between arrival gate i 2 A and airport m 2 M , and Kjm,Dep ⇢ K m,Dep is the

98
set of departure trajectory patterns between airport m 2 M and departure gate j 2 D.
Arrival/departure gates are defined as zones in the terminal area boundary intersected by a
group of trajectory patterns.
It follows that the hourly flow vector F Vt is a categorical vector of dimension |M | ⇥ |A| +
|M | ⇥ |D| and indicates the dominant arrival flows from each arrival gate to each airport
and the dominant departure flows from each airport to each departure gate.
Hierarchical clustering was applied using the Hamming distance and complete linkage for
clustering the set of hourly flow vectors. This clustering algorithm creates a hierarchical de-
composition of the data space (usually represented by a dendrogram) based on a dissimilarity
matrix that progressively splits the database into smaller subsets using either a divisive (from
the root to the leaves) or an agglomerative (from the leaves to the nodes) approach [108].
Under the complete linkage agglomerative approach, distances between groups of observa-
tions are calculated considering the maximum distance between elements in these groups.
A distance threshold is then used to horizontally cut the dendrogram and determine the
clusters. The Hamming distance was used to assess the similarity between the categorical
flow vectors and it corresponds to the number of elements in which they differ. In this case,
zero entries indicate absence of flow between the arrival/departure gate and the airport and
were disregarded during the computation of the Hamming distance.
The clustering was performed for the hourly flow vectors observed between 6:00 am and
11:00 pm (local time). A distance threshold equal to two was used, establishing a maximum
tolerable dissimilarity of two between flow vectors in the same cluster. The following rationale
was used for selecting the distance threshold. Consider a particular flow structure at a
given time represented by its flow vector. If there is a meaningful arrival (departure) flow
change at one airport because of a runway configuration switch such that different arrival
(departure) trajectory patterns are observed between the gates at the terminal area boundary
and the airport, a new flow structure will be observed and the flow vector will change. The
dissimilarity between the new flow vector and the previous one will correspond to the number
of arrival (departure) gates at the terminal area boundary for that particular airport. Since
the minimum number of arrival or departure gates by airport is three across all multi-airport
systems, the distance threshold should be at most two in order to prevent meaningfully

99
dissimilar flow structures to be included in the same cluster.
Figure 3-17 shows the number of clusters identified at each multi-airport system and their
cumulative frequency of occurrence. It is noticeable that a few clusters dominate and capture
the majority of the observations, revealing the primary operational modes at each metroplex.
The remaining clusters are composed of very few members and are characterized by less
frequently used metroplex configurations and abnormal flow structures (primarily associated
with periods of runway configuration changes). We defined a cluster of flow structures as a
Metroplex Flow Pattern (MFP). The clustering results are summarized in Table 3.2. The
New York metroplex stands out with the highest number of operational patterns and the
highest operational variability. For instance, the most frequent New York MFP accounts for
only 15.8% of the observations. At the Hong Kong and Sao Paulo metroplexes, the most
frequent MFP accounts for 31.2% and 53.5% of the observations, respectively. The results
suggest the New York metroplex is the most dynamic system. Indeed, it is observed the New
York terminal area has a higher number of flow structure changes during the course of a day
and the MFP dwell times are lower.

Figure 3-17: Cumulative percentage of observations for the clusters of hourly flow vectors
identified.

A detailed characterization of the dominant patterns for each metroplex was performed
in order to obtain insights about some of the factors that drive the behavior of the metro-
plex flows. For this, we considered the most frequent MFPs that account for 85% of the

100
Table 3.2: Results of the metroplex hourly flow vector clustering analysis.

Multi-airport system New York Hong Kong Sao Paulo


Frequency of occurrence of the most frequent 15.8 31.2 53.5
MFP (%)
Median number of flow structure changes per 5 2 2.5
day
Median dwell time of the most frequent MFP 3.5 8 5
(h)

observations.

New York Metroplex

In the New York metroplex, twelve dominant MFPs were identified. Figure 3-18 shows the
different uses of the airspace structure associated with each MFP; specifically, it shows the
average number of aircraft following each spatial trajectory pattern. MFPs are numbered in
descending order of frequency of observation.
The detailed characterization of the New York MFPs is shown in Table 3.3. The most
evident factor driving the behavior of the metroplex flows is clearly the runway configuration
in use at each airport. Five MFPs were characterized as south flow (which means that all
airports operate in a south landing flow configuration), five MFPs were characterized as
north flow and two MFPs were characterized as mixed of north and south flows (Figure
3-19 illustrates the convention about landing flow direction using JFK as an example). It
is noticeable the tendency of alignment between runway configurations across the airports.
Table 3.4 lists the most frequently observed runway configuration for each MFP. They are
represented in the form "A | B", where "A" indicates the arrival runways and "B" indicates
the departure runways.
Besides the runway configuration, some differences are also driven by the airspace con-
figuration in terms of the delegation of airspace regions that are shared. For instance, MFPs
2 and 4 are characterized by the same runway configuration at LGA. However, restricted
climb procedures from runway 13 are observed for MFP 2. Analysis of current operating
procedures reveals that whenever JFK uses runway 31L for departures, it remains with con-
trol of the shared Coney airspace, restricting the use of multiple climbs by LGA. This fact

101
also explains the differences between MFPs 6 and 12.

Figure 3-18: Most frequently observed New York MFPs.

Figure 3-19: Convention on flow direction.

102
Table 3.3: Description of the New York MFPs.

MFP Description
South Flow 1 South landing flow configuration at all airports; Favors arrival oper-
Arrival Priority ations; Right-hand arrival traffic pattern at LGA determined by de-
(South-AP) parture configuration (runway 13); Primarily observed during VMC.
9 South landing flow configuration at all airports; Favors arrival opera-
tions; Left-hand arrival traffic pattern at LGA determined by depar-
ture configuration (runway 31); Primarily observed during VMC.
4 South landing flow configuration at all airports; Favors arrival oper-
ations; Right-hand arrival traffic pattern at LGA determined by de-
parture configuration (runway 13); Primarily observed during IMC;
Multiple climbs from runway 13 at LGA with Coney Airspace release.
South Flow 2 South landing flow configuration at all airports; Favors departure op-
Departure Priority erations; Right-hand arrival traffic pattern at LGA determined by de-
(South-DP) parture configuration (runway 13); Primarily observed during VMC;
Restricted climbs from runway 13 at LGA.
3 South landing flow configuration at all airports; Favors departure op-
erations; Left-hand arrival traffic pattern at LGA determined by de-
parture configuration (runway 31); Primarily observed during VMC.
North Flow 6 North landing flow configuration at all airports; Favors arrival oper-
Arrival Priority ations; Primarily observed during IMC; Multiple climbs from runway
(North-AP) 13 at LGA with Coney Airspace release.
7 North landing flow configuration at all airports; Favors arrival opera-
tions; Visual approach at LGA runway 31; Primarily observed during
VMC.
10 North landing flow configuration at all airports; Favors arrival oper-
ations; Instrument approach at LGA runway 31; Primarily observed
during IMC.
North Flow 11 North landing flow configuration at all airports; Favors departure
Departure Priority operations; Visual approach at LGA runway 31; Primarily observed
(North-DP) during VMC.
12 North landing flow configuration at all airports; Favors departure
operations; Primarily observed during VMC; Restricted climbs from
runway 13 at LGA.
Mixed Flow 5 Mixed flow configuration with south landing flow dominance (EWR
Arrival Priority and LGA); Favors arrival operations; Primarily observed during
(Mixed-AP) VMC.
8 Mixed flow configuration with north landing flow dominance (JFK
and LGA); Favors arrival operations; Primarily observed during
VMC.

103
Table 3.4: Airport runway configurations associated with each New York MFP.

Runway configuration
MFP
JFK EWR LGA
1 13L, 22L | 13R 22L | 22R 22 | 13
2 22L, 22R | 22R, 31L 22L | 22R 22 | 13
3 22L, 22R | 22R, 31L 22L | 22R 22 | 31
4 22L, 22R | 22R 22L | 22R 22 | 13
5 31L, 31R | 31L 22L | 22R 22 | 31
6 4L, 4R | 4L 4R | 4L 4 | 13
7 31L, 31R | 31L 4R | 4L 31 | 4
8 31L, 31R | 31L 22L | 22R 31 | 4
9 13L, 22L | 13R 22L | 22R 22 | 31
10 31L, 31R | 31L 4R | 4L 31 | 4
11 4L, 4R | 4L, 31L 4R | 4L 31 | 4
12 4L, 4R | 4L, 31L 4R | 4L 4 | 13

Figure 3-20 shows that the time of the day is an important variable determining the
occurrence of MFPs. As an example, MFPs 1 and 4 are more likely to be observed in the
afternoon period, whereas MFPs 2 and 3 have a higher frequency of observations in the
morning and evening periods. This was observed to be correlated with the daily demand
profile in the New York metroplex. Analysis of the daily demand patterns reveals the ex-
istence of banks of arrivals in the afternoon, primarily driven by JFK and EWR demand
profiles, and departures in the morning, as shown in Figure 3-21. Indeed, this demand
profile, with unbalanced mix, is one important factor that drives the selection of runway
configuration at these airports in order to favor arrival or departure operations. Figure 3-22
shows the distribution of arrival and departure throughput with a boxplot for each MFP. A
clear imbalance is observed. MFPs 1, 4, 5, 6, 7, 8, 9 and 10 tend to favor arrival operations
(third quartile was used for comparison). Indeed, they are more frequently observed in the
afternoon period. By contrast, MFPs 2, 3, 11 and 12 tend to favor departure operations,
and they are more frequently observed in the morning and evening periods.

104
Figure 3-20: New York MFP frequency of occurrence by time of day.

Figure 3-21: Daily New York demand patterns, based on data from years 2013-2015.

105
Figure 3-22: Distribution of metroplex arrival and departure throughput for the New York
metroplex MFPs.

Finally, a particular influence of meteorological conditions in the operational mode in


the New York metroplex is also observed. Figure 3-23 shows, for each MFP, the percentage
of observations associated with periods of Visual Meteorological Conditions (VMC) and
Instrument Meteorological Conditions (IMC). For instance, MFPs 4, 6 and 10 are more
likely to be observed during IMC. MFP 10 has 80% of its observations associated with
periods of IMC. If we contrast MFP 10 and MFP 7, which was only observed during VMC,
it is noted that the major difference between these two MFPs is determined by the LGA
arrival flows: in MFP 7, they are tailored to a visual approach to runway 31, whereas a long
and straight Instrument Landing System (ILS) approach to the same runway is noticeable in
MFP 10. In other words, part of the New York metroplex behavior is driven by the existence
of multiple routes tailored to the approach procedure, with use governed by meteorological
conditions.

106
Figure 3-23: New York MFP frequency of occurrence under VMC and IMC.

Hong Kong Metroplex

In the Hong Kong metroplex, four dominant MFPs were identified, which are shown in Figure
3-24 and described in Table 3.5. In the most frequent MFP, all airports operate in a north
landing flow configuration. The other MFPs are characterized by a mixed flow configuration.
Table 3.6 lists the airport runway configurations inferred for each MFP.
Time of the day also appears to play a role in the occurrence of MFPs in the Hong
Kong metroplex. Figure 3-25 shows the frequency of occurrence of each MFP during the
day. While the most frequent MFP 1 is equally likely to be observed throughout the day,
MFP 2 is more frequently observed in the morning and night periods and MFPs 3 and 4 are
more likely to be observed in the afternoon period, revealing a preference for use of runways
25L/R during the afternoon at HKG.
Unlike the New York metroplex, meteorological conditions did not appear to have a
significant influence on the operational mode in the Hong Kong metroplex. Figure 3-26
shows that none of the MFPs showed a higher likelihood of being observed during IMC.

107
Figure 3-24: Most frequently observed Hong Kong MFPs.

Table 3.5: Description of the Hong Kong MFPs.

MFP Description
North Flow 1 North landing flow configuration at all airports.
Mixed Flow - N,S,N 2 North landing flow configuration at HKG and MFM; South landing
flow configuration at SZX.
Mixed Flow - S,S,N 3 South landing flow configuration at HKG and SZX; North landing
flow configuration at MFM.
Mixed Flow - S,N,N 4 South landing flow configuration at HKG; North landing flow config-
uration at SZX and MFM.

108
Table 3.6: Airport runway configurations associated with each Hong Kong MFP.

Runway configuration
MFP
HKG SZX MFM
1 07L | 07R 34 | 33 34 | 34
2 07L | 07R 16 | 15 34 | 34
3 25R | 25L 16 | 15 34 | 34
4 25R | 25L 34 | 33 34 | 34

Figure 3-25: Hong Kong MFP frequency of occurrence by time of day.

Figure 3-26: Hong Kong MFP frequency of occurrence under VMC and IMC.

109
Sao Paulo Metroplex

In the Sao Paulo metroplex, four dominant MFPs were also identified, which are shown in
Figure 3-27 and described in Table 3.7. The most frequent MFP is characterized by a mixed
flow configuration, with arrivals landing north at GRU and south at CGH and VCP. As in
the Hong Kong metroplex, it was not observed a clear tendency of alignment between the
airport runway configurations. Table 3.8 lists the airport runway configurations inferred for
each MFP.

Figure 3-27: Most frequently observed Sao Paulo MFPs.

Figure 3-28 shows the frequency of occurrence of each MFP during the day. MFPs 1 and
3 are more likely to be observed in the morning and night periods, whereas MFPs 2 and 4
are more likely to be observed in the afternoon period. This seems to be correlated with the
daily demand patterns at GRU, as shown in Figure 3-29. There is a preference for using
runways 09L/R during the busiest hours of the day in the morning and night periods, as

110
Table 3.7: Description of the Sao Paulo MFPs.

MFP Description
Mixed Flow - N,S,S 1 North landing flow configuration at GRU; South landing flow config-
uration at CGH and VCP.
Mixed Flow - S,N,N 2 South landing flow configuration at GRU; North landing flow config-
uration at CGH and VCP.
Mixed Flow - N,N,S 3 North landing flow configuration at GRU and CGH; South landing
flow configuration at VCP.
South Flow 4 South landing flow configuration at all airports.

Table 3.8: Airport runway configurations associated with each Sao Paulo MFP.

Runway configuration
MFP
GRU CGH VCP
1 09R | 09L 17R | 17R 15 | 15
2 27L | 27R 35L | 35L 33 | 33
3 09R | 09L 35L | 35L 15 | 15
4 27L | 27R 17R | 17R 15 | 15

configuration 09R | 09L is optimal for GRU 3 .


Finally, the operational mode in the Sao Paulo metroplex was not observed to be par-
ticularly influenced by meteorological conditions. Figure 3-30 shows that none of the MFPs
showed a higher likelihood of being observed during IMC.

Figure 3-28: Sao Paulo MFP frequency of occurrence by time of day.


3
The offset of 590 m between the runway thresholds 09L and 09R has a positive impact on capacity when
arrivals use 09R and departures use 09L, whereas the offset of 1140 m between the runway thresholds 27L
and 27R has a negative impact on capacity when arrivals use 27L and departures use 27R

111
Figure 3-29: Daily demand pattern at GRU, based on data of 2017.

Figure 3-30: Sao Paulo MFP frequency of occurrence under VMC and IMC.

3.4.2 Identification of Metroplex Flow Interactions

We leveraged the knowledge about the flow patterns to identify relevant inter-airport flow
interactions driven by the metroplex airspace design and use. The trajectory tube intersec-
tions identified in Section 3.3.2 can be generally translated into two types of flow interactions,
depending on the use of the trajectory patterns involved in the intersection:

• Flow dependency: It occurs when the use of one trajectory pattern inhibits the use of
the other one; therefore, flows are dependent and not observed simultaneously. From

112
an operational perspective, a flow dependency is expected to make the management of
runway/airspace configuration more challenging, as coordination is required between
the airports.

• Flow crossing: It occurs when the use of one trajectory pattern does not inhibit the
use of the other one; therefore, both flows can be observed simultaneously and they
overlap at their shared airspace region. From an operational perspective, a flow crossing
is expected to increase air traffic control complexity, as it tends to increase workload
associated with conflict detection and avoidance in the shared airspace region.

We used the knowledge about the metroplex flow patterns to identify relevant flow cross-
ings and flow dependencies as follows:
A trajectory tube intersection is translated into a relevant flow crossing if there exists at
least one MFP for which the trajectory patterns are used simultaneously and meaningfully.
We define a meaningful use of a trajectory pattern if the median number of aircraft using
the trajectory pattern is higher than one.
A trajectory tube intersection is translated into a relevant flow dependency if, for all
MFPs, the trajectory patterns are not used simultaneously. For this, we consider that
whenever one trajectory pattern is used meaningfully (median number of aircraft is higher
than one), the median number of aircraft using the other trajectory pattern should be equal
to zero.
For the New York Metroplex, 60 flow dependencies were identified. For the Hong Kong
and Sao Paulo metroplexes, flow dependencies were not observed. Figure 3-31 shows the
location of the trajectory tube intersections characterized as flow dependencies in the New
York metroplex. JFK and LGA arrival flows have the highest number of dependencies. Most
of the flow dependencies are associated with shared airspace regions located within 30 nm
from the airports. Clusters of trajectory tube intersections are observed close to the airports,
revealing the existence of infeasible combinations of runway configurations. The results reveal
the New York metroplex is the most interdependent system, suggesting the management of
runway and airspace configuration tends to be more challenging at this system.

113
Figure 3-31: Trajectory tube intersections associated with flow dependencies in the New
York metroplex.

Table 3.9 shows the number of relevant flow crossings identified by MFP. The New York
metroplex also has the highest number of flow crossings by MFP, revealing the typical ter-
minal area flow structures are potentially characterized by higher levels of air traffic control
complexity.

Table 3.9: Number of flow crossings by MFP.

MFP New York Hong Kong Sao Paulo


1 15 2 7
2 10 7 6
3 17 6 7
4 19 4 7
5 14 - -
6 14 - -
7 12 - -
8 15 - -
9 13 - -
10 13 - -
11 12 - -
12 15 - -

114
3.5 Characterization of Performance Differences
The performance analysis was focused on arrival operations, given they are typically subject
to higher levels of inefficiency as most of the arrival sequencing and scheduling is performed
within the terminal area. We focused on the three performance areas of Efficiency, Ca-
pacity and Predictability, which are defined by ICAO as key areas for ATM performance
measurement [133].

3.5.1 Efficiency

In this performance area, we were primarily concerned with flight trajectory efficiency under
the notion of how actual trajectories compare to reference ideal trajectories. With this
notion, we analyzed the efficiency of metroplex airspace design and use and the efficiency of
traffic flows.

Metroplex Airspace Design and Use Efficiency

We first compared the efficiency of the learned airspace structure for the three multi-airport
systems in terms of trajectory lateral efficiency. Based on the trajectory patterns identified
for each metroplex airport, we established a structural path stretch metric S, which is defined
as the weighted average of the path stretch associated with the representative trajectory of
each learned route:

N
X
S= p i si (3.3)
i=1

Where S is the structural path stretch of an airspace with N trajectory patterns, pi is


the weight of trajectory pattern i and si is the path stretch associated with the centroid of
trajectory pattern i. If the weights are equal for all trajectory patterns, the metric assesses
the efficiency of metroplex airspace design regardless airspace use. If the weights are defined
as the frequency of occurrence of trajectory patterns, the metric evaluates the combined
efficiency of metroplex airspace design and use. The path stretch is defined as the difference
between the actual trajectory total length and the length of the shortest path that connects

115
the initial and the end points of the trajectory, as illustrated in Figure 3-32.

Figure 3-32: Example of path stretch calculation.

From the perspective of lateral performance, an airspace designed with lower structural
path stretch can be considered more efficient as aircraft fly smaller distances between the
terminal area boundary and the runway threshold. The calculated structural path stretch
for arrival operations for each metroplex and each of their airports is shown in Figure 3-33.
The metric was calculated considering both equal weights and use-based weights.
The Sao Paulo metroplex presents the most efficient airspace design, with the lowest
structural path stretch. Interestingly, when airspace use is taken into account, the New
York metroplex outperforms the others and shows the most efficient combination of airspace
design and use. The metric is significantly higher for the Hong Kong metroplex. HKG stands
out as the airport with the lowest efficiency, which is potentially due to the high level of de-
confliction with the neighboring airports. At the other end, VCP stands out as the airport
with the highest efficiency, ratifying the advantages of four-corner post airspace design.
Figure 3-34 shows the structural path stretch associated with different metroplex config-
urations. The metric was calculated with use-based weights, based on the average number of
aircraft following each spatial trajectory pattern observed for each MFP. The results reveal
that the south flow configurations in the New York metroplex (MFPs 1, 2, 3, 4, 5 and 9) are
less efficient as they consistently show higher levels of structural path stretch. In the Hong
Kong metroplex, the most frequently used configurations are also the least efficient. This is

116
particularly driven by the north flow configuration at HKG; trajectory patterns associated
with runway 07L showed high levels of path stretch. Finally, Sao Paulo showed the lowest
variability in MFP structural path stretch.

Figure 3-33: Structural path stretch for arrivals.

Figure 3-34: Structural path stretch for arrivals by MFP.

117
Traffic Flow Efficiency

We compared the overall traffic flow efficiency by analyzing the efficiency of actual trajec-
tories, both at spatial and temporal dimensions. Two metrics were defined. The lateral
traffic flow efficiency is a distance-based efficiency metric. For a set of N flight trajectories
associated with the traffic flow structure observed during a given time period, it is defined
as:

PN
di
Ef flateral = PNi=1 (3.4)
i=1 Di

where di is the length of the shortest path connecting the initial and end points of
trajectory i and Di is the actual length of trajectory i.
The temporal traffic flow efficiency is a time-based efficiency metric and it is defined in
a similar fashion:

PN
ti
Ef ftemporal = PNi=1 (3.5)
i=1 Ti
where ti is the unimpeded flight time associated with trajectory i and Ti is the actual
flight time associated with trajectory i. For each trajectory, its unimpeded flight time is
estimated as the minimum between its actual flight time and the 10th percentile of the
distribution of flight times for trajectories in the same spatial pattern.
Based on their definitions, both metrics range between 0 and 1. Values closer to one
indicate higher efficiency. The lateral and temporal efficiencies were computed on an hourly
basis for the metroplex flows. The median efficiency values for each metroplex are shown in
Figure 3-35. Overall, the New York metroplex presents the highest traffic flow efficiency, both
spatially and temporally. The Hong Kong metroplex shows a significantly lower lateral effi-
ciency, which is in part driven by its least efficient airspace design. It also presents the lowest
temporal efficiency. Under inclement weather conditions, however, the New York metroplex
shows significant drops in efficiency, becoming the lowest performing system temporally.

118
Figure 3-35: Lateral and temporal traffic flow efficiency.

We also calculated the median traffic flow efficiency for each of the dominant operational
modes observed in each metroplex. Figure 3-36 shows the distribution of the MFP efficiency
values in a boxplot. While the New York metroplex on average outperforms the others
both in terms of lateral and temporal efficiency, it exhibits the highest variability in MFP
efficiency.

Figure 3-36: Distribution of traffic flow efficiency by MFP.

Trajectory Conformance Efficiency

We also evaluated the lateral conformance of flight trajectories against the learned airspace
structure. While the lateral performance evaluated by the lateral traffic flow efficiency metric

119
is a result of both airspace design and tactical terminal area traffic flow management, the
analysis of trajectory conformance provides insights about the sole contribution of tactical
traffic flow management. Figure 3-37 shows the average daily percentage of non-conforming
arrival trajectories at each metroplex airport. The results are displayed separately for days
with fair weather conditions and for days with weather impacts. Overall, it is observed the
New York metroplex exhibits the best level of trajectory conformance, both for nominal and
off-nominal conditions. This is consistent with its high lateral traffic flow efficiency. HKG
stands out as the airport with the highest percentage of non-conforming behaviors, well
above the other airports, even for nominal conditions. This reveals that the observed low
lateral traffic flow efficiency for the Hong Kong metroplex is not only a result of its least
efficient airspace design, but also of terminal area traffic flow management. At the Sao Paulo
metroplex, the primary airport GRU also shows the lowest level of trajectory conformance.

Figure 3-37: Average daily percentage of non-conforming arrival trajectories.

It is also observed that, for all airports, the percentage of non-conforming trajectories
increases for days impacted by adverse meteorological conditions such as convection, low
ceiling/visibility or strong winds. This suggests that weather is an important factor affecting
the efficiency of terminal area operations at all three multi-airport systems.
As an example, the non-conforming trajectories identified for a convective weather day
at JFK, HKG and GRU are shown in Figure 3-38. It is observed that most of the trajectory
deviations are caused by airborne holding, excessive vectoring or rerouting, in other words,
tactical ATC instructions that tend to increase the length of the trajectory.

120
Figure 3-38: Example non-conforming arrival trajectories identified for a convective weather
day at JFK, HKG and GRU.

3.5.2 Capacity

Capacity generally refers to an upper bound on the allowable throughput of a facility. Two
common definitions of capacity are the maximum capacity (or saturation capacity) and the
practical capacity [29]. The maximum capacity is defined as the expected number of move-
ments under persistent demand. The practical capacity extends this definition by including
the notion of level of service and specifying a threshold on the expected level of delay ex-
perienced by the users of the facility. It recognizes that a certain level of delay should be
acceptable in order to ensure a steady stream of demand at the facility, but it should also
be reasonable to ensure a sustainable operation of the facility through time. Indeed, this
balance between throughput and delay is key for the capacity declaration process.
We developed an empirical approach for assessing metroplex capacity (maximum and
practical) from historical terminal area throughput and delay performance. Rather than
treating each airport individually, metroplex airports are looked from a systems perspective.
Figures 3-39 and 3-40 shows how the arrival throughput and the excess transit times in the
New York terminal area change as a function of the arrival demand, under VMC. Specifically,
for every minute, the arrival throughput in the next 15-minute period and the excess transit
time for each flight that landed in the same 15-minute period is plotted as a function of
the number of arriving aircraft simultaneously present in the terminal area. Clearly, as the

121
arrival demand increases, the arrival throughput increases in a disproportionate fashion until
reaching a saturation level, i.e., point at which delivering more aircraft to the terminal area
does not significantly increase throughput because the system has reached its capacity. This
behavior is reflected in the excess transit time curve. When the arrival demand increases,
more aircraft are competing for the same resources and will encounter a higher probability
of a delay assignment during the runway sequencing and scheduling process. When the
saturation level is reached, the excess transit time increases much more rapidly. The plots
also reveal significant variability in throughput and delay for the same level of demand. Our
assumption is that part of this variability can be explained by differences in performance
across metroplex configurations.

Figure 3-39: New York metroplex arrival Figure 3-40: Excess terminal area transit
throughput as a function of the arrival de- time as a function of the arrival demand.
mand in the terminal area.

Our goal was to estimate the arrival throughput and the excess terminal area transit
time as a function of the arrival demand in order to obtain performance curves that correlate
throughput and level of delay in the terminal area for each MFP. In that way, the curves
would allow to quantify the expected level of delay associated with a specific throughput
level and could be used to determine arrival rates tailored to a given level of service.
We formalize the throughput estimation problem as a regression problem. Considering
the observed throughput and delay profiles shown in Figures 3-39 and 3-40, we modeled
the arrival throughput and the excess transit time in the terminal area with a piecewise

122
linear function of the arrival demand (number of arriving aircraft in the terminal area), with
additional constraints arising from expected operational behavior: the arrival throughput
was modeled as a non-decreasing and concave function of the arrival demand and the excess
transit time was modeled as a non-decreasing function of the arrival demand. The parameters
of the curves were estimated using quantile regression [134].
The quantile regression problem was formulated as a linear programming model and
solved with Gurobi 6.5.0. Given N observations of arrival demand a, arrival throughput t
and excess transit time d, LP-I determines the p-quantile regression fit of arrival throughput
as a function of arrival demand (t̃n = ↵i + i an ) and LP-II determines p-quantile regression
fit of excess transit time (delay) as a function of arrival demand (d˜n = ↵i + i an ). In LP-I,
Equations 3.6, 3.7 and 3.8 frame the quantile regression problem as a linear programming
model, as described in [134]. Equations 3.9 and 3.10 model the arrival throughput as a
concave and non-decreasing function of the arrival demand, while Equation 3.11 ensures the
continuity of the piecewise function. Similarly, in LP-II, Equations 3.12, 3.13 and 3.14 frame
the quantile regression problem as a linear programming model, Equation 3.15 models the
excess transit time as a non-decreasing function of the arrival demand and Equation 3.16
ensures the continuity of the piecewise function.

LP-I:
N
X
min yn (3.6)
↵, ,y
n=1

S.T.:
yn p tn (↵i + i an ) , i 1  an  i, 8n (3.7)

yn (1 p) (↵i + i an ) tn , i 1  an  i, 8n (3.8)

k+1 k  0, 8k = 1, ..., amax 1 (3.9)

↵k + kk  ↵k+1 + k+1 (k + 1), 8k = 1, ..., amax 1 (3.10)

↵k + kk = ↵k+1 + k+1 k, 8k = 1, ..., amax 1 (3.11)

123
LP-II:
N
X
min yn (3.12)
↵, ,y
n=1

S.T.:
yn p dn (↵i + i an ) , i 1  an  i, 8n (3.13)

yn (1 p) (↵i + i an ) dn , i 1  an  i, 8n (3.14)

↵k + kk  ↵k+1 + k+1 (k + 1), 8k = 1, ..., amax 1 (3.15)

↵k + kk = ↵k+1 + k+1 k, 8k = 1, ..., amax 1 (3.16)

The performance curves obtained with a median regression fit (p = 0.5) for each MFP in
the New York, Hong Kong and Sao Paulo metroplexes are presented in Figures 3-41, 3-42,
3-43 , 3-44, 3-45 and 3-46. The curves were estimated using VMC observations in order
to marginalize out weather related throughput impacts and enable a cleaner assessment of
metroplex configuration performance.
The plots reveal that New York is the highest throughput system, followed by Hong
Kong and Sao Paulo. The median arrival throughput of the dominant MFP saturates at 34
aircraft/15min (or 136 aircraft/hour) in the New York metroplex, 14 aircraft/15min (or 56
aircraft/hour) in the Hong Kong metroplex and 13 aircraft/15min (or 52 aircraft/hour) in
the Sao Paulo metroplex.

Figure 3-41: Throughput-demand curves for Figure 3-42: Delay-demand curves for the
the New York MFPs. New York MFPs.

124
Figure 3-43: Throughput-demand curves for Figure 3-44: Delay-demand curves for the
the Hong Kong MFPs. Hong Kong MFPs.

Figure 3-45: Throughput-demand curves for Figure 3-46: Delay-demand curves for the
the Sao Paulo MFPs. Sao Paulo MFPs.

The maximum throughput as well as the median level of delay associated with it are
presented in Table 3.10 for all MFPs. The maximum throughput normalized by the number
of active arrival runways associated with each particular MFP is also presented in Figure 3-47
for facilitating the comparison between metroplexes. The New York metroplex stands out
with the largest standard deviation in throughput performance, emphasizing its operational
variability, as also observed in the previous section. MFP 1 shows the highest arrival rate
(136 aircraft/hour), while MFP 10 shows the lowest arrival rate (92 aircraft/hour). The
median level of delay at saturation is 3.7 minutes, but the standard deviation is high. MFPs

125
4 and 10 show the highest delay levels of 8.7 minutes and 7.4 minutes, respectively. In the
Hong Kong and Sao Paulo metroplexes, there is smaller variation in throughput performance.
For Hong Kong, the median level of delay at saturation is 4.4 minutes. Sao Paulo stands out
with a relatively higher median delay level of 6.4 minutes.

Table 3.10: Metroplex arrival rates under persistent demand.

New York Hong Kong Sao Paulo


MFP Rate Delay Rate Delay Rate Delay
(aircraft/h) (min) (aircraft/h) (min) (aircraft/h) (min)
1 136 3.1 56 5.8 52 6.8
2 108 4.3 60 4.6 44 4.6
3 116 5.3 60 4.3 48 6.0
4 112 8.7 48 4.3 48 8.1
5 132 3.7 - - - -
6 120 4.2 - - - -
7 120 3.1 - - - -
8 124 3.7 - - - -
9 130 3.0 - - - -
10 92 7.4 - - - -
11 104 3.1 - - - -
12 108 2.8 - - - -
Median 118 3.7 58 4.4 48 6.4
Standard deviation 12.8 1.9 5.7 0.7 3.3 1.4

Figure 3-47: Metroplex arrival rates under persistent demand and associated level of delay.

126
In order to exemplify the use of the performance curves for also assessing the practical
capacity, we obtained the metroplex arrival rates corresponding to a specific level of service
of 4-minute of terminal area delay for each New York MFP. They are presented in Table 3.11
and were compared with baseline metroplex arrival rates obtained by summing the individual
airport capacity declared by the FAA [135] for the runway configurations associated with each
MFP. The comparison reveals that, for some MFPs, the differences between the empirical
and the baseline rates are large, indicating that higher levels of delay should be expected
when the New York system is operating at the declared capacity under these operational
modes. It is also observed that MFPs with the same theoretical runway system capacity can
show very different throughput performance. For instance, MFPs 2, 3 and 4 have the same
declared runway system capacity, but the differences in metroplex practical capacity can be
as great as 20 aircraft per hour. The results highlight the impacts of the terminal area flow
structure on the multi-airport system capacity and emphasize the importance of taking a
systems perspective for metroplex airport capacity and flow management.

Table 3.11: Comparison between empirical and baseline metroplex arrival rates for each New
York MFP.

MFP Empirical rates Baseline rates Difference (baseline - empirical)


1 136 138 2
2 108 126 18
3 100 126 26
4 88 126 38
5 132 139 7
6 120 123 3
7 120 137 17
8 124 137 13
9 132 138 6
10 84 111 27
11 104 123 19
12 108 123 15

3.5.3 Predictability

Predictability is defined by ICAO [133] as the ability of airspace users to provide consistent
and dependable levels of performance. It is typically assessed by measuring the variation in

127
system performance experienced by the users [136]. In this section, we analyze the variability
in performance across MFPs and within a MFP and discuss interrelations between ATM
configuration predictability and ATM performance predictability.

ATM performance variability due to ATM configuration variability

As indicated in Table 3.2, the New York metroplex shows the lowest MFP dwell times and the
highest number of changes in the flow structure per day, suggesting the ATM configuration
is less predictable in this system. We translated this variation in ATM configuration into
daily capacity variation and structural path stretch variation in order to analyze how the
ATM configuration predictability affects ATM performance predictability.
Table 3.12 shows the average accumulated change in capacity and terminal area structural
path stretch per day for each metroplex. For this, we considered the observed MFPs for
each day and their associated maximum achievable throughput (normalized by the number
of runways) and structural path stretch, as shown in Table 3.10 and 3-34. It is observed
that the New York system presents the highest capacity variation per day (11 aircraft). For
Hong Kong and Sao Paulo, there is a small capacity variation of less than 2 aircraft. From
an operational perspective, this means the New York system is more reliant on traffic flow
management in order to adjust demand with the more dynamic system capacity. In terms
of structural path stretch variation, Hong Kong stands out with the highest value, which
suggests lower trajectory lateral performance predictability. It is important to mention that,
in the context of TBO, predictability of the ATM configuration and associated distances to
be flown is important for effective trajectory optimization.

Table 3.12: Daily capacity and structural path stretch variation.

New York Hong Kong Sao Paulo


Capacity variation 11.2 1.8 1.4
Structural path stretch variation (nm) 8.5 9.8 1.4

128
ATM performance variability within an ATM configuration

We investigated the variability in ATM configuration performance by analyzing the pre-


dictability power of the capacity curves empirically derived from the data. For this, we
calculated the Normalized Root Mean Squared Error (NRMSE) associated with the median
regression fit for each MFP throughput-demand curve:
r
1
P
n
n
(ŷi yi ) 2
i=1
N RM SE = (3.17)
sd(y)
where ŷ is the predicted throughput based on the capacity curve, y is the actual observed
throughput, n is the total number of observations and sd(y) is the standard deviation of the
observed throughput.
Table 3.13 shows the NRMSE for the MFP throughput-demand curves for each metroplex.
It is observed that the New York metroplex has both the most predictable MFP (lowest
NRMSE) and the least predictable MFP (highest NRMSE) from a capacity standpoint. In
all metroplexes, it is observed that the least predictable MFPs were those with the lowest
throughput performance.

Table 3.13: NRMSE for the MFP throughput-demand curves.

MFP New York Hong Kong Sao Paulo


1 0.375 0.453 0.434
2 0.412 0.440 0.517
3 0.419 0.534 0.386
4 0.549 0.575 0.481
5 0.414 - -
6 0.356 - -
7 0.348 - -
8 0.423 - -
9 0.527 - -
10 0.692 - -
11 0.435 - -
12 0.394 - -

129
3.6 Discussion
The characterization of terminal area traffic flows in the New York, Hong Kong and Sao Paulo
metroplexes revealed structural and operational differences between these multi-airport sys-
tems. First, the New York metroplex is observed to have a more complex route design, with
a higher number of routes and a higher number of interactions between them. A significant
number of these interactions are located close to the airport and were characterized as flow
dependencies, revealing the existence of impracticable combinations of runway configuration
and routes. On the other hand, for the Hong Kong and Sao Paulo systems, most of the inter-
airport route interactions are located further from the airports and were characterized by
larger airspace volumes resulting from the sharing of both airspace fixes and path segments.
The analysis also showed the New York system is the one with more dynamic changes in
the terminal area flow behavior during daily operations. While the metroplex traffic flows
can be categorized in few patterns in the Hong Kong and Sao Paulo systems, they show
much higher variability in the New York system. This variability can be generally attributed
to a couple of factors: higher variability in the set of runway configurations by airport;
existence of inter-airport flow dependencies that create operationally infeasible combinations
of runway configuration and routes; higher airspace design complexity that enables ATM to
tailor operations for VMC and IMC. The New York metroplex was also found to be the
most interdependent system, with a significant number of flow dependencies, and the typical
terminal area flow structures were found to have a higher number of flow crossings, suggesting
higher levels of air traffic control complexity.
By analyzing the impacts of such structural and operational differences on performance,
we found the more complex New York airspace presents the most efficient design and use in
terms of trajectory lateral performance. The Hong Kong airspace design showed a high level
of structural path stretch, which is potentially the cost of its more decoupled airspace, with
fewer flow interactions with the neighboring airports. On average, the New York metroplex
showed the highest traffic flow efficiency, both spatially and temporally. Yet, it exhibited the
highest variability in traffic flow efficiency as well as more pronounced efficiency drops during
inclement weather conditions. The Hong Kong metroplex showed a significantly lower lateral

130
traffic flow efficiency, which was observed to be a result not only of its least inefficient airspace
design, but also of tactical traffic flow management, as revealed by its lower temporal traffic
flow efficiency and trajectory conformance. In terms of capacity, the New York metroplex
was found to be the highest throughput system, but stood out with the largest variability
in throughput performance. Differences in metroplex arrival capacity across MFPs were as
great as 44 aircraft per hour. Such differences were associated not only with its more diverse
set of runway configurations, but also with differences in the terminal area flow structure.
The more dynamic New York metroplex flow behavior, particularly influenced by mete-
orological conditions, suggests the coupled ATM configuration in this multi-airport system
as well as its performance might be less predictable. Indeed, we observed the New York
system exhibited the highest variability in traffic flow efficiency and capacity by MFP. From
an operational perspective, this means the New York system is more reliant on traffic flow
management in order to adjust demand with the more dynamic system capacity. At the
same time, it also means traffic flow management tends to become more challenging, since
it has to deal with increased levels of uncertainty in the ATM configuration and associated
constraints.
Overall, the results highlight different action areas to be considered at each system to-
wards improving their terminal area operations. The high level of structural path stretch
in the Hong Kong metroplex, particularly for HKG, reveals opportunities for improvement
in the airspace structure in order to increase trajectory lateral efficiency. Enhancements in
airspace design in the Hong Kong and Sao Paulo metroplexes may also be performed to de-
couple routes from different airports that share the same airspace region in order to achieve
a more expeditious flow. The lower level of trajectory conformance and temporal efficiency
in the Hong Kong and Sao Paulo metroplexes also reveals opportunities for improvement
in tactical traffic flow management towards better sequencing and scheduling of arriving
flights. Finally, the higher level of airport interdependency, dynamism, weather dependency
and performance variability in the New York metroplex reveal opportunities for (1) improv-
ing the airspace design in order to de-conflict flows close to the airports and (2) integrating
weather forecasting with decision-making in order to better anticipate weather conditions,
translate them into operational impact and plan traffic flow management efficiently.

131
THIS PAGE INTENTIONALLY LEFT BLANK

132
Chapter 4

Data-Driven Approach for Metroplex


Configuration and Capacity Planning

In Chapter 3, we demonstrated the use of the flight trajectory data analytics framework to
identify and characterize traffic flow patterns in the terminal area of multi-airport systems.
As the observed patterns in the terminal area traffic flow revealed recurrent utilization pat-
terns of runways and airspace, we were able to identify the major configurations in which
a metroplex system collectively operates as well as key intervening factors. The results
also emphasized that metroplex systems can have very dynamic airspace use and show high
variability in throughput performance across different configurations. From an operational
perspective, this has two major implications. First, they are more reliant on ATFM to adjust
demand with the more dynamic system capacity. Second, the planning of ATFM is more
challenging. Because there is higher uncertainty in the ATM configuration and its perfor-
mance, estimating future capacity precisely is more difficult and the allocation of airport and
airspace resources under uncertainty becomes more subject to inefficiencies. In this chapter,
we show how the knowledge generated by the characterization of metroplex flow patterns,
provided by the flight trajectory data analytics framework, can be leveraged to develop de-
scriptive and prescriptive models for metroplex configuration and capacity planning towards
improved traffic flow management decision support. The analysis is focused on the New
York metroplex.

133
4.1 Metroplex Configuration and Airport Capacity Plan-
ning Framework
The data-driven approach for metroplex configuration and airport capacity planning de-
veloped in this chapter has two components: a capacity prediction module and a capacity
allocation module. It is illustrated in Figure 4-1.
The first contribution is the development of predictive models for capacity estimation
that specifically accounts for operational interdependency aspects existing in a metroplex
environment. The capacity prediction module uses machine learning methods to translate
weather forecasts into probabilistic forecasts of the metroplex configuration and airport ca-
pacity for strategic time horizons. A traffic flow pattern prediction model first delivers a
forecast of the most likely flow structure to be seen in the terminal airspace on an hourly
basis. At any given time, the terminal area flow pattern reflects coupled runway and airspace
configuration decisions and interactions between arrival and departure procedures, providing
a complete picture of the multi-airport system state. An airport capacity prediction model
then takes as input the metroplex flow pattern forecast and weather features representing the
forecast environmental conditions in the terminal area to deliver arrival capacity forecasts.
The second contribution is the development of an AAR planning model for airport ca-
pacity allocation that is intended to support the design of strategic TMIs such as GDPs.
The capacity allocation module uses optimization methods in order to prescribe optimal
AARs for strategic traffic flow management. The AAR planning model builds upon existing
GHP optimization models, but instead of relying on theoretical capacity modeling or capac-
ity scenario trees, it directly incorporates the output generated by the capacity prediction
module. Besides, it is designed to incorporate robustness goals. In summary, the model
has the following key features: (1) it is stochastic, i.e., capacity uncertainty is accounted
for in the form of capacity profiles drawn from actual predictive distributions; (2) it enables
revision of the plan in order to take advantage of updated information and more accurate
predictions throughout the planning horizon; (3) it incorporates robustness goals, i.e., the
decision-maker has the ability to select a desired level of robustness for the ATFM plan based
on its relative valuation between efficiency and predictability.

134
Figure 4-1: Metroplex configuration and airport capacity planning framework.

4.2 Data Description


The datasets used for this study covers the same period of 69 days from 2013-2015 (discretized
in hourly periods from 6:00 am to 11:00 pm (local time)) considered for the analysis of traffic
flows in the New York metroplex in Chapter 3; 37 are fair weather days and 32 are weather-
impacted days, which include convective weather impacts and non-convective impacts, such
as adverse wind conditions and low ceiling/visibility. The data was randomly partitioned into
training and test datasets, composed of 70% of the days and 30% of the days, respectively.
The partitioning was performed separately for fair weather days and weather-impacted days
in order to ensure the training and test datasets had a high variability of weather conditions
as in the original dataset. The data sources are described below.

4.2.1 Terminal Aerodrome Forecast (TAF)

TAF is a concise weather report that states the expected meteorological conditions at an
airport during a specified period of time (usually 24 hours). It is routinely issued every 6
hours (at 0, 6, 18 and 24 UTC), but an amended TAF can also be issued at any time if the
expected meteorological conditions change significantly. The following weather features were

135
extracted from TAF: wind direction, wind speed, visibility and ceiling. Although ceiling is
not directly reported in TAF, it can be extracted as the lowest height of broken or overcast
cloud layer, as defined by the FAA [137]. For each time t marking the beginning of a
forecast period, the most recent TAF (issued at t0  t) was considered and the weather
features corresponding to each hourly time period in the forecast horizon were extracted for
prediction. For instance, if t0 = t 1, a 1-hour look-ahead capacity prediction for [t, t + 1]
will use a 2-hour look-ahead weather forecast in TAF.

4.2.2 Arrival Route Status and Impact (ARSI) Forecast

ARSI is the arrival version of the Route Availability Planning Tool (RAPT), a decision
support tool that provides information about departure routes affected by convective weather
[138]. Starting at the current time and extending out 30 minutes into the future in 5-minute
intervals, RAPT assigns a departure route status that indicates the expected convective
weather impact along the route (0 for clear, 1 for insignificant weather encountered, 2 for
partial or uncertain blockage and 3 for blocked). Similarly, ARSI assigns an arrival route
status in 15-minute intervals up to 60 minutes into the future. As the ARSI forecast is made
with respect to the time the flight is expected to reach the transition fix in the arrival route,
its conditions will be reflected at the airport after some lead time (corresponding to the length
of the arrival routes embedded in the tool, which is approximately 1 hour). Therefore, ARSI
forecasts were used as predictive features for up to 2-hour arrival capacity forecasts. For a
given time t, the mean of the ARSI forecast {(t 60min) + 0min, (t 45min) + 0min, (t
30min) + 0min, (t 15min) + 0min} was used as a predictive feature for a 1-hour prediction,
whereas the mean of the ARSI forecast {t + 0min, t + 15min, t + 30min, t + 45min} was used
as a predictive feature for a 2-hour prediction.

4.2.3 Hourly Airport Reports

Hourly airport reports ("Airport Information" and "Airport Efficiency Information") from
the FAA Aviation System Performance Metrics (ASPM) database were used to obtain data
about flight schedules and arrival/departure rates.

136
4.2.4 Flight Tracks

One of the challenges for capacity prediction using supervised learning concerns the deriva-
tion of the knowledge about actual capacity (output "truth"). In this thesis, the airport ar-
rival capacity information (actual output) required for building the machine learning models
was derived from historical arrival throughput after filtering out periods of low demand in
which the airport is not running at capacity. For this, we leveraged flight track data and the
empirical approach for capacity estimation described in Section 3.6.2 of Chapter 3. Specif-
ically, we empirically estimated a threshold N ⇤ for the number of arriving aircraft present
in the terminal area (arrival demand) for which the airport system can be considered to be
under pressure and only used the arrival throughput values for which the demand was equal
or above the threshold (N N ⇤ ). Figure 4-2 shows the histogram of hourly arrival rates
resulting from the filtering process.

Figure 4-2: Histogram of hourly airport arrival rates.

4.3 Prediction Models and Features


The capacity prediction module relies on two models: a traffic flow pattern prediction model
and an airport arrival capacity prediction model. The prediction problems are framed as
supervised learning classification/regression problems. In supervised learning, the goal is
to learn a mapping between a set of input variables and a given output variable based on
knowledge extracted from a training dataset.
A multi-way classification model is first developed to predict the most likely pattern

137
in which the metroplex flows might be for strategic time horizons. For this, we leverage
the knowledge about the metroplex flow patterns obtained with the flight trajectory data
analytics framework. As they reflect utilization patterns of runway configurations and ar-
rival/departure route structure, the traffic flow pattern prediction model enables the fore-
casting of the global metroplex configuration state. The multi-way classification model is
represented as

zt = g(ut , zt 1 ) (4.1)

where zt is the metroplex flow pattern at time t and ut is the input vector with the
relevant factors driving the metroplex flow dynamics, as discussed in Chapter 3:

⇣ ⌘
ut = t, (wspeedk,t )k2M , (wdirk,t )k2M , (imck,t )k2M , arrt , dept (4.2)

where t is the time of the day, (wspeedk,t )k2M is the vector of forecast wind speed at time
t for each airport k in the set M of metroplex airports, (wdirk,t )k2M is the vector of forecast
wind direction, (imck,t )k2M is the vector of forecast meteorological conditions (0, if VMC,
and 1, if IMC) and arrt and dept are the scheduled aggregate arrival and departure demands
for the metroplex, respectively, at time t.
In a subsequent step, a regression model is developed to map forecast weather conditions
and metroplex configuration state into flow rates for each individual metroplex airport. It is
represented as

yk,t = f (t, wk,t , yk,t 1 , zt ) (4.3)

where yk,t is the arrival rate output at time t for airport k, zt is the expected metroplex
configuration state at time t and wk,t is a feature vector encompassing all weather forecast
features at time t for airport k, as follows:

⇣ ⌘
wk,t = ceilk,t , visk,t , headk,t , tailk,t , crossk,t , convecN
k,t , convec S
k,t , convec W
k,t (4.4)

138
where ceilk,t is the ceiling forecast at time t for airport k, visk,t is the visibility forecast,
headk,t , tailk,t and crossk,t are the headwind, tailwind and crosswind forecasts, respectively,
at the dominant arrival runway at airport k for metroplex configuration zt , and, finally,

k,t , conveck,t and conveck,t are the forecast convective blockage status for the dominant
convecN S W

north, south and west arrival routes for airport k.


Given the ATFM planning needs for capacity forecasts with proper quantification of un-
certainty, a key requirement of the prediction framework is to provide the complete predictive
distribution of the output. Besides, it has to be able to generate forecasts for each time pe-
riod throughout the ATFM planning horizon. This is accomplished through an iterative
procedure in which the regression model is sequentially used to make predictions for each
time period based on the output from the previous time period, as illustrated in Figure 4-3.
In this process, a Monte Carlo sampling approach is used to propagate the uncertainty about
the metroplex configuration state and about the arrival rate from the beginning to the end
of the planning horizon. For instance, given the predictive distribution of arrival rates for
the first forecast hour, a random sample is taken from this distribution and used to predict
the arrival rate for the second hour, resulting in a conditioned predictive distribution for the
second hour arrival rate. The process continues for the rest of the planning horizon and,
once completed, is repeated n times. As a result, we also obtain n potential capacity profiles
drawn from the predictive distributions that can be directly used for ATFM planning. The
following pseudo code describes the algorithmic process for arrival rate forecasting through-
out the planning horizon and generation of capacity profiles. The subscript k indicating each
airport was suppressed to not overload the notation.
for test instance i = 1, ..., p
for iteration n = 1, ..., 1000
Draw zt+1
i
from Zt+1 ⇠ gZt+1 uit+1 , zti
Draw yt+1
i
from Yt+1 ⇠ fYt+1 wt+1
i
, yti , zt+1
i

for forecast horizon h = 2, ..., H


Draw zt+h
i
from Zt+h ⇠ gZt+h uit+h , zt+h
i
1

Draw i
yt+h from Yt+h ⇠ fYt+h i
wt+h i
, yt+h i
1 , zt+h

139
Figure 4-3: Forecasting procedure throughout the planning horizon.

4.4 Prediction of Traffic Flow Patterns


The prediction of traffic flow patterns is framed as a supervised multi-way classification
problem. We used state-of-the-art methods for machine learning (Neural Network, Random
Forests and Support Vector Machine) to build a set of candidate classification models for this
purpose, which were then evaluated with respect to predictive performance. The following
sub-sections provide the details about the models created. To simplify the notation, the
traffic flow pattern mapping zt = g(ut , zt 1 ), defined in Equation 4.1, is represented as
z = g(x). All models were developed with MATLAB R2017a.

4.4.1 Classification Models

Neural Network Classification Model (NN)

An artificial neural network is a non-parametric method that can be used for both classifica-
tion and regression. It consists of a sequence of layers containing artificial nodes (neurons)
that compute some function of the previous layer in a feed-forward fashion from the input
layer to the output layer [139].
Different network architectures were tested, with varied number of layers and neurons.
Based on 5-fold cross validation performance (Table 4.1), the neural network model designed

140
for the multi-way classification problem was composed of 1 input layer, 1 hidden layer with
30 neurons (m = 30) and 1 output layer. With this architecture, the output z is mapped
from the input space x 2 Rd as follows:

d
X
aj = xi Aij + A0j (4.5)
i=1

m
X
bk = h(aj )Bjk + B0k (4.6)
j=1

zk = h(bk ) (4.7)

where aj is the output of neuron j in the hidden layer, bk is the output of neuron k in the
output layer, ✓ = Aij , A0j [ Bjk , B0k are the layer parameters (weights and bias), tuned with
back-propagation, and h(aj ) = 1
1+e aj is the logistic activation function used in the hidden
layer. Equation 4.5 maps the input layer to the hidden layer, Equation 4.6 maps the hidden
e bk
layer to the output layer and Equation 4.7 applies a softmax function h(bk ) = PK
e bk
to the
k=1

output layer in order to transform the neural network scores into posterior probabilities.

Table 4.1: Classification accuracy of NN model for various network architectures based on
5-fold cross validation performance.

Number of nodes
5 10 15 20 25 30 35 40
1 79.1 80.9 81.1 79.3 82.4 82.4 82 79.7
Number of hidden layers
2 76.6 81.3 81.4 80.7 81.3 82 80.5 80.9

Random Forests Classification Model (RF)

Random Forests is a non-parametric method for both classification and regression consisting
of an ensemble of decision tree learners [114]. As the method has already been described in
Chapter 2, the reader is referred to Section 2.4.1.
The algorithm has few parameters to tune: T is the number of trees in the ensemble, M is
the number of features to sample at each split and N is the minimum number of observations

141
in a leaf node. Based on cross-validation performance (Figure 4-4), the RF model for traffic
flow pattern prediction was created with T = 100, M = 10 and N = 1.

Figure 4-4: Parameter sensitivity of RF model based on 5-fold cross validation performance.
(a) Impacts of the minimum size of leaf node on the classification error. (b) Impacts of the
number of predictors to sample at each split on the classification error.

Support Vector Machine Classification Model (SVM)

Support Vector Machine outputs a decision boundary for binary classification by finding an
optimal separating hyperplane g(x) = aT x + b such that observations mostly fall on the right
side of the hyperplane (g(x) +1 for z = +1, and g(x)  1 for z = 1) and the distance
from the hyperplane to the observations closest to it on either side (margin) is maximized
for best generalization [113].
This problem can be formulated as an optimization problem:

1 X
min kak2 + ⇠ (4.8)
2
Subject to:

z(aT x + b) 1+⇠ (4.9)

⇠ 0 (4.10)

142
In Equation 4.8, the first term of the objective function seeks to maximize the margin and
the second term introduces a penalty factor for misclassifications. Equation 4.9 ensures that
observations fall on the right side of the hyperplane, but with allowance for misclassifications
based on the introduction of non-negative slack variables (Equation 4.10).
For the multi-way classification problem, SVM was applied with an ensemble of binary
learners using a one-versus-one encoding. A binary SVM learner l = 1, ..., L was created for
each pairwise combination of classes. The final class ẑ of a new observation was determined
by maximizing the posterior probability over the binary learners:

n PL |m |p (z = m ) o
kl l kl
ẑ = arg max l=1
PL (4.11)
k l=1 |m kl |
where mkl = 1 if class k is coded as the positive class in binary learner l, mkl = 1 if class
k is coded as the negative class in binary learner l, and mkl = 0 if class k is not considered
in binary learner l.
For each binary learner, posterior probabilities were estimated by fitting a sigmoid func-
tion to the scores g(x) based on cross-validation [140]:

1
p(z = 1/g(x)) = (4.12)
1 + exp(Ag(x) + B)
In order to compute the final posterior probabilities from the ensemble of binary learners,
we followed the approach of Wu and Lin [141]:

L
X K
X K
X
min [ pl (z = 1) p̂k I(mkl = 1) + (1 pl (z = 1)) p̂k I(mkl = +1)]2 (4.13)
l=1 k=1 k=1

Subject to:

0  p̂k  1 (4.14)

X
p̂k = 1 (4.15)
k

143
We tested both a linear SVM model with a linear kernel K(x, x0 ) = xT x0 as well as a
kx x0 k2
non-linear SVM model with the radial basis function kernel K(x, x0 ) = exp( 2 2
). Based
on 5-fold cross-validation performance, the non-linear SVM model was chosen, as it showed
classification accuracy of 83.0%, whereas the classification accuracy was 78.7% for the linear
SVM model.

4.4.2 Performance Evaluation

The predictive performance of the classification models for traffic flow prediction was evalu-
ated for up to six hours of look-ahead time, which is a typical time horizon for strategic traffic
flow management. In order to have useful predictions for decision-making under uncertainty,
probabilistic forecasts were generated for each time period throughout the planning horizon.
For this, Bayes’ rule was sequentially applied: for the first hour, the prediction model was
run and a probability for each MFP was obtained; for the subsequent hours in the forecast
horizon, the prediction model was run assuming a given outcome for the previous hour and
the final probability was computed conditioned on the probability of the assumed outcome
for the previous hour. For each hour t, the predicted flow pattern zt⇤ then corresponds to
the feasible one with maximum probability pt among all C flow patterns:

zt⇤ = arg max{pt,x } (4.16)


x

where:

C
X
pt,x = p(zt = x/zt 1 = c).pt 1,c (4.17)
c=1

The feasibility of each MFP was determined based on crosswind and tailwind thresholds
derived statistically from the data. Specifically, these thresholds represent the 99th percentile
of the observed crosswind and tailwind values on the major arrival/departure runways at each
airport (Figures 4-5 and 4-6). For any given time, a MFP is considered to be feasible if the
crosswind and tailwind values on the corresponding runways do not exceed the statistically
derived thresholds.

144
Figure 4-5: Statistically derived crosswind thresholds by MFP.

Figure 4-6: Statistically derived tailwind thresholds by MFP.

145
First, the accuracy of point predictions was evaluated. Figure 4-7 shows the multi-way
classification accuracy obtained for each model as a function of the look-ahead time of the
prediction. It is observed that the SVM model slightly outperformed the others for shorter
time horizons, whereas the RF model showed higher classification accuracy for longer look-
ahead times. On average, the accuracy was approximately 83% for a 1-hour prediction, 64%
for a 3-hour prediction, and 50% for a 6-hour prediction.

Figure 4-7: Multi-way classification accuracy of NN, RF and SVM traffic flow pattern pre-
diction models for forecast horizons of 1 h to 6 h.

In order to evaluate the accuracy of the probabilistic forecasts, the Brier score was used:

N C
1 XX
BS = (pic I(zi = c)) (4.18)
N i=1 c=1

where N is the total number of observations in the test set, C is the number of classes, pic
is the posterior probability of observation i belonging to class c and I(zi = c) is an indicator
function that equals to 1, if the actual class of observation i is c, and 0, otherwise.
Based on the formulation of Brier score, the lower the value of this metric, the higher
the accuracy of the probabilistic forecast. Figure 4-8 shows the Brier score obtained for each
model throughout the forecast horizon. First, it is observed that the score increases through-
out the forecast horizon, which is consistent with the decrease in classification accuracy for

146
longer look-ahead times. Overall, the RF model showed the lowest Brier score, followed by
SVM and NN. Based on the accuracy of point predictions and posterior probability estima-
tion, the RF model was chosen for prediction of traffic flow patterns.

Figure 4-8: Brier score of NN, RF and SVM traffic flow pattern prediction models for forecast
horizons of 1 h to 6 h.

We also compared the traffic flow pattern prediction model developed in this thesis with a
static model that assumes the flow pattern observed at the beginning of the planning horizon
remains throughout the entire planning horizon. The results are shown in Figure 4-9. It is
observed that, for short look-ahead times, the RF model outperforms the static model only
by a small amount, revealing that the current flow pattern carries most of the predictive
power for short-term predictions. Indeed, there is a lot of inertia in the runway/airspace
configuration selection process since changes can generate a significant increase in workload.
Therefore, a given flow pattern is expected to be seen until conditions turn it operationally
infeasible. As the prediction horizon increases, the likelihood that current conditions will
remain decreases and inertia starts to have a smaller impact (changes become necessary).
The RF model better predicts these changes as the gap between the models increases with
the prediction horizon. For a six-hour forecast horizon, the performance gap was 13%.

147
Figure 4-9: Accuracy of traffic flow pattern forecasts for RF model and static model as a
function of the look-ahead time of the prediction.

Figure 4-9 also shows the prediction accuracy obtained if the top two most likely flow
patterns in the probabilistic forecast are considered instead of just the most likely one. For
long-term predictions, the accuracy stabilizes at about 77%. The result indicates that, for
many instances in which the model incorrectly predicted the MFP, it was right on the second
best option, emphasizing the value of the probabilistic forecast. Indeed, the classification
confusion matrix for a 3-h forecast horizon (Figure 4-10) reveals the majority of the incorrect
predictions occurred because of confusion between very similar flow patterns, especially for
the south flow patterns and because of incorrect prediction of JFK flows.

148
Figure 4-10: Classification confusion matrix for traffic flow pattern predictions for a forecast
horizon of 3 h.

4.5 Prediction of Airport Arrival Capacity

4.5.1 Regression Models

The prediction of airport arrival capacity is framed as a regression problem, and the complete
predictive distributions are desired. We used different methods for machine learning to build
a set of three candidate regression models for this purpose, which were then evaluated with
respect to predictive performance. We started with the simplest regression model under
a parametric approach by assuming the output is a linear function of the input variables.
The linear regression model was constructed in a Bayesian setting, as we are interested in
the predictive distribution. The other two models were constructed using non-parametric
methods - Random Forests and Gaussian Process - that can learn non-linear mappings
directly from the data. While Random Forests is not probabilistic in nature, it can provide
an empirical predictive distribution based on the Quantile Regression Forests framework
[142]. Besides, it has shown reasonable performance in previous studies in the same domain
[75]. The following sub-sections provide the details about the models created. To simplify
the notation, the arrival capacity mapping yk,t = f (t, wk,t , yk,t 1 , zt ), defined in Equation 4.3,

149
is represented as y = f (x). All models were developed with MATLAB R2017a.

Bayesian linear Regression Model (BR)

In a linear regression model, the output is assumed to be a linear function of the input
variables plus a random disturbance, which is assumed to be independent and identically
distributed Gaussian white noise:

y = wT x + " (4.19)

1
" ⇠ N (0, ) (4.20)

Under a Bayesian approach, a prior probability over the parameters of the model p(✓)
is defined, where ✓ = [w, ], and Bayes’ rule is used to calculate the posterior probability
distribution given the training data (x, y):

p(✓)p(x, y/✓) p(✓)p(x, y/✓)


p(✓/x, y) = =R (4.21)
p(x, y) p(x, y/✓)p(✓)d✓
The posterior distribution is then used to derive a predictive distribution for y 0 given new
data x0 :

Z
0
p(y /x, y) = p(y 0 /✓)p(✓/x, y)d✓ (4.22)

In this study, the Bayesian linear regression model was developed using the normal-
gamma prior over the parameters. In the absence of specific prior knowledge, a vague prior
was given, with p(w/ ) ⇠ N (0, 1002 ) and p( ) ⇠ G(1, 100). If we define a gamma prior over
the noise precision, p( ) ⇠ G(a0 , b0 ), and a normal prior over the input weights conditioned
on the noise precision, p(w/ ) ⇠ N (µ0 , ⌃0 ), such that the joint prior probability is normal-
gamma, p(w, ) = p( )p(w/ ) ⇠ N G(µ0 , ⌃0 , a0 , b0 ), it can be shown that the posterior is
analytically tractable and it is also normal-gamma [113]:

p(w, /x, y) ⇠ N G(µN , ⌃N , aN , bN ) (4.23)

150
where:

⌃N = (xT x + ⌃0 ) 1
(4.24)

µN = ⌃N (xT y + ⌃0 µ0 ) (4.25)

N
aN = a0 + (4.26)
2

1
bN = b0 + (y T y + µT0 ⌃0 µ0 µTN ⌃N µN ) (4.27)
2
The predictive distribution for y 0 given new data x0 can then be easily obtained by sam-
pling from the posterior (w̃, ˜) ⇠ N G(µN , ⌃N , aN , bN ) and subsequently from the likelihood
ỹ ⇠ N (w̃T x0 , 1˜ ).

Random Forests Regression Model (RF)

As described in Section 4.4.1, Random Forests can be used for both classification and re-
gression. In a regression context, the prediction of a decision tree for a new observation x0 is
obtained by averaging over the observed output values in the leaf l(x0 ) that this observation
falls into when it is passed through the tree:

1 X
ŷ 0 = yi (4.28)
|i : xi 2 l(x0 )|
{i:xi 2l(x0 )}

The final response of the ensemble is computed by aggregating the results of the individual
trees:

T
X
1
0
ŷbag = PT ↵t ŷt0 (4.29)
t=1 ↵t t=1

where ŷt0 is the prediction from tree t in the ensemble and ↵t is the weight of tree t.
While predicting for a new observation x0 , if we fully record the observed output values
in the leaf l(x0 ) that this observation falls into instead of averaging over them, an empirical

151
predictive distribution can be built [142].
The parameters of algorithm (T , M and N , as described in Section 4.4.1) were tuned
with 5-fold cross validation. Figure 4-11 shows the parameter sensitivity results for arrival
capacity prediction at JFK. Based on these results, the RF model was created with T = 100,
M = 3 and N = 10.

Figure 4-11: Parameter sensitivity of RF regression model based on 5-fold cross validation
performance. (a) Impacts of the minimum size of leaf node on the prediction error. (b)
Impacts of the number of predictors to sample at each split on the prediction error.

Gaussian Process Regression Model (GP)

The final model was created using a Gaussian Process probabilistic framework, which is also
a non-parametric method. In a Gaussian Process regression model, the output variable y
is assumed to be a function f of the input variables x following a Gaussian Process. A
Gaussian Process is defined as a collection of random variables, any finite number of which
have a joint Gaussian distribution [143]. Alternatively, it can be described as a distribution
over functions and be written as:

f (x) ⇠ GP(m(x), k(x, x0 )) (4.30)

where m(x) is the mean function (typically assumed to be zero):

152
m(x) = E[f (x)] (4.31)

and k(x, x0 ) is the covariance/kernel function:

k(x, x0 ) = E[(f (x) m(x))(f (x0 ) m(x0 ))] (4.32)

The kernel function governs the covariance between functions values and it is the main
feature that enables the modeling of different types of functions flexibly. Different kernel
functions can be used to generate a valid covariance matrix (i.e., symmetric and positive
semi-definite) [143]. In this work, we used the canonical squared exponential kernel function,
with a separate length scale for each predictor:

✓ d ◆
1 X (xp,i xq,i )2
k(xp , xq ) = 2
f exp (4.33)
2 i=1 li2

where 2
f is the signal variance and l is d-dimensional vector of length scales. These kernel
parameters are normally referred to as hyperparameters.
As in a typical regression framework, we assumed the observations to be noisy, with
independent and identically distributed Gaussian white noise with zero mean and variance

n:
2

y = f (x) + " (4.34)

" ⇠ N (0, 2
n) (4.35)

Under the assumption of a zero mean function and Gaussian white noise, the observations
y also follow a Gaussian Process:

y ⇠ GP(0, k 0 (xp , xq )) (4.36)

k 0 (xp , xq ) = k(xp , xq ) + 2
n pq (4.37)

153
where pq is a Kronecker delta.
Finally, the joint distribution of the observed target values y and the function values f⇤
at the test locations is also Gaussian and can be written as
2 3 2 3!
y K(X, X) + n2 I K(X, X⇤ )
4 5 ⇠ N 0, 4 5 (4.38)
f⇤ K(X⇤ , X) K(X⇤ , X⇤ )

where K(X, X⇤ ) denotes the matrix of covariances evaluated at all pairs of n training
points and n⇤ test points (similarly for K(X, X), K(X⇤ , X), K(X⇤ , X⇤ )).
By conditioning the joint Gaussian prior distribution on the observations, we can derive
the joint posterior distribution over function values and obtain the predictive equations:

⇣ ⌘
f⇤ /X, y, X⇤ ⇠ N f¯⇤ , cov(f⇤ ) (4.39)

where:

f¯⇤ = K(X⇤ , X)[K(X, X) + 2 1


n I] y (4.40)

cov(f⇤ ) = K(X⇤ , X⇤ ) K(X⇤ , X)[K(X, X) + 2 1


n I] K(X, X⇤ ) (4.41)

The parameters of the GP model can be estimated using Bayesian inference. If ✓ is the
vector of model parameters (kernel hyperparameters 2
f and l; and noise variance n ),
2
it can
be estimated by maximizing its log posterior probability:

n 1 T 1 n o
✓ˆ = arg max{log p(✓/X, y)} = arg max log p(✓) y Ky 1 y log |Ky | log 2⇡ (4.42)
✓ ✓ 2 2 2

where p(✓) is the prior distribution over the hyperparameters, n is the number of training
observations, y is the vector of observed output values, and Ky is the covariance matrix
obtained by applying the kernel function over pairs of input vectors in X.
GPstuff [144] was used for hyperparameter inference in this study. Due to the absence
of specific prior knowledge, the hyperparameters were given a broad gamma prior G(1, 100).

154
Gradient-based optimization was used for maximizing the log posterior probability.

4.5.2 Performance Evaluation

Evaluating the accuracy of point capacity predictions

The predictive performance of the regression models was also evaluated for up to six hours
of look-ahead time. Two metrics were used to evaluate the accuracy of point capacity
predictions: the Normalized Root Mean Squared Error (NRMSE), as defined in Equation
4.43, and the Mean Absolute Prediction Error (MAPE), as defined in Equation 4.44.
q P
1 n
n i=1 (ŷi yi ) 2
N RM SE = (4.43)
sd(y)
where ŷi is the predicted arrival rate for observation i, yi is the observed arrival rate for
observation i, n is the total number of observations in the test set and sd(y) is the standard
deviation of the vector of observed arrival rates.

n
1 X ŷi yi
M AP E = (4.44)
n i=1 yi

Figures 4-12 and 4-13 shows the NRMSE and MAPE values for each regression model
throughout different forecast horizons. Overall, it is observed that the GP model outper-
formed the RF and BR models for all three airports. The GP and RF models exhibited
more similar performance, while the BR model systematically showed higher error rates. It
is also observed that the error tends to increase slightly with the prediction horizon. This
can be explained by the decreasing contribution of the current arrival rate feature and by
larger uncertainties in weather conditions and metroplex configuration state. On average,
the MAPE obtained with the GP model was 13.6% for JFK, 9.5% for EWR, and 8.6% for
LGA.

155
Figure 4-12: NRMSE of BR, RF and GP capacity prediction models for forecast horizons of
1 h to 6 h.

Figure 4-13: MAPE of BR, RF and GP capacity prediction models for forecast horizons of
1 h to 6 h.

Figure 4-14 shows the predictor importance values generated by the RF model. For any
feature, this measure is the increase in the prediction error if the values of that feature
are permuted across the out-of-bag observations. It is computed for every tree, averaged
over the entire ensemble and divided by the standard deviation over the ensemble. The
results show that the previous rate is the dominant predictive feature for the three airports,
followed by the metroplex configuration feature. This result emphasizes the importance of
anticipating the metroplex configuration state for predicting airport capacity. Among the
weather variables, ceiling seems to be the most important feature for the three airports,
followed by visibility. Headwind showed a relatively high importance value for JFK and

156
EWR. In high wind conditions, the changes in wind direction and speed along the arrival
route can cause a loss of horizontal separation between aircraft (known as compression) and
affect throughput. This compression phenomena is more of a factor during strong headwind
conditions near the ground, as the spacing between consecutive aircraft will tend to decrease
during the approach if they encounter significant wind changes during the descent towards
the airport. Finally, convective weather features are also observed to be relevant for the
three airports and convective impacts seem to be more important when they block the west
flows.

Figure 4-14: Predictor importance estimates from RF model.

Evaluating the accuracy of probabilistic capacity predictions

The Prediction Coverage (PC), as defined in Equation 4.45, was the quantitative metric used
to assess the quality of the predictive distributions generated by the models.
⇣P ⌘
n
i=1 I(l↵ŷi  yi  uŷ↵i )
P C↵ = (4.45)
n
where [l↵ŷi , uŷ↵i ] is the ↵% confidence interval for the predicted arrival rate ŷi and I is an
indicator function that equals to 1 if the argument expression is true (i.e., the actual arrival
rate falls within the confidence interval) and 0, otherwise.

157
For a given confidence level ↵%, the PC indicates the percentage of observations in
the test set for which the actual arrival rates fall within the ↵% confidence interval of
the predicted arrival rates. Figure 4-15 shows the PC obtained for each regression model,
considering three confidence levels: 50%, 70% and 90%.

Figure 4-15: PC of BR, RF and GP capacity prediction models for 50%, 70% and 90%
confidence levels and forecast horizons of 1 h to 6 h.

158
Ideally, one would like to obtain a strong match between the PC and the target con-
fidence level, as it would indicate a good matching between the actual and the estimated
distributions. For JFK and EWR, it is observed that the GP model outperformed the others
in estimating the uncertainty of the predictions since its PC values were closer to the tar-
get confidence levels. The RF model systematically showed PC values that are lower than
the target confidence levels, revealing the predicted distributions tend to be narrower than
the actual. For LGA, the GP and BR model exhibited similar performance. Based on the
accuracy results in terms of point prediction and uncertainty quantification, the Gaussian
Process model was chosen for arrival capacity prediction.

Analysis of metroplex system state predictive value

The prediction framework developed in this thesis uniquely accounts for metroplex conditions
in the airport capacity estimation process through an embedded metroplex flow pattern
prediction model. While Figure 4-14 reveals the metroplex configuration is the second most
important feature for predicting capacity, a natural question that arises is how the predictive
performance of our model compares to a simplified model that does not account for metroplex
effects. For this, an alternative Gaussian Process regression scheme is developed to predict
arrival rates solely based on the current arrival rate and weather forecasts, i.e., metroplex
state information is not included as a feature in the capacity estimation process:

yk,t = f (t, wk,t , yk,t 1 ) (4.46)

where yk,t is the arrival rate output at time t for airport k and wk,t is a feature vector
encompassing all forecast weather features at time t for airport k (as described in Section
4.3). It should be noted that headwind, tailwind and crosswind forecasts in the vector wk,t
were replaced by wind speed and direction forecasts since there is no information about the
metroplex configuration and, consequently, about the dominant arrival runway.
The predictive performance of the models is shown in Figure 4-16. It is observed that the
two-layer prediction framework developed in this thesis outperforms the alternative model in
predicting the capacity at the three New York airports throughout the forecast horizon. It

159
is also observed that the benefits are more expressive for JFK and for longer time horizons.
This can be explained by the fact that JFK has a more diverse set of runway configurations
that can be selected by ATC (from a capacity standpoint) and its throughput performance is
affected more significantly by the metroplex configuration. Besides, because of the power of
the current arrival rate feature (Figure 4-14), removing the metroplex configuration feature
does not much impact the predictive performance for shorter time horizons (our model
outperforms the alternative model only by a small amount for short look-ahead times of one
hour). Yet, for longer time horizons, the inertia factor plays a smaller role as it is less likely
that environmental conditions will remain the same and that the system will stay in the
same state. As a result, predicting the metroplex state and including this information in the
capacity estimation process contributes more to reduce the capacity prediction error.
We also evaluated how the uncertainty in the metroplex configuration state affects the
performance of our model by replacing the predicted metroplex configuration with the actual
metroplex configuration in the feature vector. The results are also shown in Figure 4-16. It
is observed that, with deterministic knowledge about the true metroplex configuration state,
the capacity prediction error tends to reduce, as expected. The impacts of the uncertainty
about the metroplex configuration state on the capacity prediction error were found to be
small, even for long time horizons. This highlights the value of the probabilistic metroplex
flow pattern forecast and its good predictive performance, as shown in Figure 4-9.

Figure 4-16: NRMSE of alternative GP capacity prediction models for forecast horizons of
1 h to 6 h.

160
Comparison between data-driven and baseline capacity predictions

We also compared the data-driven capacity predictions with the baseline AARs retrieved
from historical hourly airport reports provided by the FAA. Figures 4-17, 4-18 and 4-19
shows the MAPE values associated with the baseline capacity estimates and the capacity
predictions generated by the GP model, for JFK, EWR and LGA, respectively. The MAPE
values are presented aggregated for the full test dataset as well as disaggregated by type of
day (with respect to the weather condition).

Figure 4-17: MAPE of GP model capacity predictions and baseline capacity estimates for
JFK and forecast horizons of 1 h to 6 h.

Figure 4-18: MAPE of GP model capacity predictions and baseline capacity estimates for
EWR and forecast horizons of 1 h to 6 h.

161
Figure 4-19: MAPE of GP model capacity predictions and baseline capacity estimates for
LGA and forecast horizons of 1 h to 6 h.

For JFK, an average reduction of 5.4% in the prediction error was obtained with the
data-driven capacity predictions generated by the GP model. The average MAPE value was
13.6% for the GP model predictions and 19.0% for the baseline capacity estimates. For fair
weather days, a reduction of 7.4% in the prediction error was observed, whereas a smaller
reduction of 3.2% was observed for weather impacted days. The histogram of the relative
prediction error for individual observations is presented in Figure 4-20. It is observed that the
distribution of the baseline prediction errors is slightly shift to the right of the distribution of
the GP prediction errors for fair weather days, with more capacity estimates that are higher
than the actual values. On the other hand, for weather impacted days, the distribution of the
baseline prediction errors is slightly shift to the left of the distribution of the GP prediction
errors, with more capacity estimates that are lower than the actual values.
For EWR, an average reduction of 3.0% in the prediction error was obtained with the
data-driven capacity predictions. The average MAPE value was 9.5% for the GP model pre-
dictions and 12.5% for the baseline capacity estimates. Similarly to JFK, weather impacted
days showed a smaller reduction in the prediction error (2.3%) than fair weather days (3.7%).
The histogram of the relative prediction error for individual observations is presented in Fig-
ure 4-21. The distribution of the baseline prediction errors is also slightly shift to the right
of the distribution of the GP prediction errors for fair weather days, whereas it is shift to the
left and exhbitis a more frequent underestimation of capacity for weather impacted days.

162
Finally, for LGA, the average MAPE value was 8.6% for the GP model predictions and
10.1% for the baseline capacity estimates. The data-driven capacity estimation process
generated an average reduction of 1.5% in the prediction error, the smallest value between
the three airports. The reduction was 1.7% for fair weather days and 1.4% for weather
impacted days. Figure 4-22 shows the histogram of the relative prediction error for individual
observations. It is observed that the distributions of the baseline prediction error and of the
GP prediction errors are more similar and narrower, with lower mass in the tails.

Figure 4-20: Histogram of prediction errors for GP model predictions (for 3-h look-ahead
time) and baseline capacity estimates for JFK.

Figure 4-21: Histogram of prediction errors for GP model predictions (for 3-h look-ahead
time) and baseline capacity estimates for EWR.

163
Figure 4-22: Histogram of prediction errors for GP model predictions (for 3-h look-ahead
time) and baseline capacity estimates for LGA.

4.6 Prescription of Airport Arrival Rates


The last module of the metroplex configuration and airport capacity planning framework
uses optimization methods in order to prescribe AARs for traffic flow management based on
the capacity predictions generated by the regression models. We evaluate the impacts of the
framework towards traffic flow management decision support by considering the planning of
Ground Delay Programs (GDPs).

4.6.1 Ground Delay Programs

GDPs are the most common strategic TMI implemented is the NAS and the current mech-
anisms by which the GHP is solved in practice. A GDP can be initiated by the ATCSCC
when there is a predicted demand-capacity imbalance at an airport and its plan contains the
following attributes:

• Duration: it specifies the initial and end times for arrival capacity regulation; flights
scheduled to arrive within the GDP duration are subject to ground delays.

• Scope: it specifies the set of airports/flights for which ground delays can be issued;
typically, trans-continental and international flights are exempted from the program.

164
• Airport Acceptance Rates (AARs): They specify the number of arrival slots available
at each time period (typically, hourly periods) throughout the program duration.

Given a GDP plan designed by the central authority, ground delays are initially computed
to meet the planned arrival rates using the Ration-By-Schedule (RBS) principle (i.e., follow-
ing the arrival order in the original flight schedules), and translated into a revised schedule
in which each affected flight receives an Expected Departure Clearance Time (EDCT). In
a subsequent step, flight operators are allowed to exchange arrival slots internally or with
other flight operators under a CDM framework.

4.6.2 AAR Planning Model

The AAR planning model builds upon existing stochastic optimization models for the GHP,
which aim to determine how much capacity to be allocated and how much ground delay to be
assigned in order to minimize overall delay costs. Rather than modeling uncertain capacity
behavior with scenario trees or theoretical probability distributions, the model accounts for
uncertainty in the form of capacity profiles drawn from actual predictive distributions. In
addition, it incorporates robustness goals, i.e., the decision-maker has the ability to select
a desired level of robustness for the ATFM plan based on its relative valuation between
efficiency and predictability. This is motivated by increased attention to predictability per-
formance by the ATM community [145].
In order to formulate the model, we consider the implementation of a GDP during a
planning horizon T = {Ti , ..., Tf }, where T is a set of hourly time periods, Ti and Tf are
the initial and final hours of the GDP program. If a flight is scheduled to arrive at the
airport during the program duration T , and its scheduled departure time is later than the
time Tb at which the program is filed (Tb  Ti ), it is considered in the capacity allocation
process and becomes subject to ground delays. The number of affected flights scheduled
to depart at time period s and arrive at time period i is given by Nsi . We assume the
demand is deterministically known. Airport capacity during the GDP duration is given by
Q possible capacity profiles drawn from the predictive distributions generated by the airport
capacity prediction model. Each capacity profile q is a time-series of predicted arrival rates

165
{Mqj , j = Ti , ..., Tf }. The proposed optimization model is presented below.

X
Tf Tf Tf +1
XX Q Tf
1 XX
min cg (j i)xsij + ca Wqj + ↵✓ (4.47)
s=Tb i=Ti j=i
Q q=1 j=T
i

Subject to:

Tf +1
X
xsij = Nsi , 8s 2 {Tb , ..., Tf }, 8i 2 {Ti , ..., Tf } (4.48)
j=i

Tf j
X X
xsij + Wq,j 1 Wqj  Mqj , 8j 2 {Ti , ..., Tf }, 8q 2 {1, ..., Q} (4.49)
s=Tb i=Ti

Tf
X
✓ Wqj , 8q 2 {1, ..., Q} (4.50)
j=Ti

Wq,Ti 1 = 0, 8q 2 {1, ..., Q} (4.51)

xsij 2 Z+ , 8s 2 {Tb , ..., Tf }, 8i 2 {Ti , ..., Tf }, 8j 2 {Ti , ..., Tf + 1} (4.52)

Wqj 2 Z+ , 8j 2 {Ti , ..., Tf }, 8q 2 {1, ..., Q} (4.53)

✓ 2 Z+ (4.54)

The decision variables are xsij , which is the number of flights originally scheduled to
depart at time period s and arrive at time period i and rescheduled to arrive at time period
j, and Wqj , which is the number of flights unable to land at time period j if capacity profile
q materializes. The objective function is expressed in Equation 4.47 and seeks to find a
balance between efficiency and predictability. The first term of the function is the cost of

166
planned ground delays and the second term is the expected cost of airborne delays, which is
the average over all possible capacity profiles. Together, they account for the efficiency goal
of minimization of overall delay costs. The unit cost of ground delay is cg and the unit cost of
airborne delay is ca . In order to incorporate robustness, the third term introduces a penalty
term for unpredictability as the maximum unplanned airborne delay (associated with the
worst-case capacity profile). This robustness term has a weight ↵, which can be directly
selected by the decision-maker depending on its relative valuation between efficiency and
predictability. The higher the value of ↵, the higher the level of robustness and the lower the
risk of unpredictable airborne delays. While we could indirectly introduce an unpredictability
penalty by increasing the cost ratio ca /cg , we decided to treat separately the fact that airborne
delays are less desirable because of higher operating costs and lower predictability. First,
although we acknowledge that unpredictable airborne delays may generate additional costs
due to their lower recovery potential (as flight operators are left with a shorter time horizon to
revise flight schedules, aircraft routing, crew schedules and re-accommodate passengers in the
event of a disruption, fewer options are available for recovery, which may end up increasing
costs), the extent of this impact is difficult to quantify. Second, this modeling decision gives
the decision-maker the ability to directly trade between efficiency and predictability with the
robustness parameter ↵. Operational constraints are expressed in Equations 4.48 and 4.49
and they require that all flights must either arrive at the time period they were originally
scheduled to arrive or at a later time period and that the number of arrivals cannot exceed
capacity (flow balance constraint).
In order to allow revision of the plan throughout the planning horizon and take advantage
of updated information and more accurate capacity forecasts, we consider the implementation
of a rolling horizon approach. For example, considering that the plan can be revised n times
at time periods {Tk0 , k = 1, ..., n, Ti < Tk0 < Tf }, the rolling horizon approach would be
implemented as follows: first, the ground holding model would be solved for T = {Ti , ..., Tf }
and implemented for T = {Ti , ..., T10 1}, i.e., ground delays would be incurred as planned
for all flights with controlled departure times earlier than T10 ; at the first revision time T10 ,
the model would be re-solved for the updated planning horizon T = {T10 , ..., Tf }, considering
what was been implemented up to time period T10 1 and the updated capacity forecasts;

167
the same procedure would be repeated iteratively at each revision time Tk0 .
Without loss of generality, we present below the formulation for the updated model to be
solved at the first revision time T 0 . The updated cost function is expressed in Equation 4.55
and introduces an additional robustness term to account for the unpredictability that can
be introduced by a plan revision. The first three terms are similar to those in Equation 4.47.
The last term accounts for the unpredictability introduced by changes in ground delays.
These changes are expressed in Equations 4.61, 4.62, 4.63 and 4.64, where x⇤sij is the optimal
solution for the initial plan. If, for all flights initially scheduled to depart at time period
s and arrive at time period i, the new plan results in more ground delays than what was
previously planned when the model was solved at the previous step, ysi is positive and gives
the total increase in ground delays assigned to these flights. On the other hand, if the new
plan results in less ground delays than what was previously planned, zsi is positive and gives
the total amount of ground delays that will not be incurred by these flights. This robustness
term has a weight , which can be directly selected by the decision-maker to control changes
in ground delays during the revision. The higher the robustness parameter , the lower the
modification of the initial GDP plan. The underlying motivation for penalizing changes in
ground delays during the revision is as follows: if, for a set of flights, the new plan results
in more ground delays in expectation of degraded capacity at the destination airport, the
ANSP may wish to be more adherent to the initial plan in order to give these flights the
best chance to arrive at the time initially planned (especially if these flights are sensitive to
the flight operator, for instance, in terms of connecting flights downstream); on the other
hand, if the new plan results in less ground delays in expectation of improved capacity at
the destination airport, the ANSP may also wish to be more adherent to the initial plan
if resources have already been deployed to adapt to this plan, reducing the chance that a
resource is released and, later on, has to be deployed again (after all, capacity is uncertain
and additional airborne delays may be incurred).
Equations 4.56 - 4.58 are similar to Equations 4.48 - 4.50. In Equation 4.56, the updated
demand Nsi0 is given by all flights that are still on the ground at the time of the revision
and that have earliest possible departure time s and earliest possible arrival time i. If the
previous plan is implemented exactly, the updated demand Nsi0 can be calculated with 4.59

168
and it is given by all flights with controlled departure times later than T 0 1, as determined
by the previous plan, and that can now receive a revised ground delay assignment. Equation
4.59 aggregates the flights originally scheduled to depart from T 0 onwards with the flights
originally scheduled to depart before T 0 but that were assigned a controlled departure time
later than T 0 1 in the previous plan. Those flights are now considered as new flights
scheduled to depart at T 0 . In Equation 4.57, the new term Vj is introduced to ensure that
all flights that have taken off before T 0 and are anticipated to arrive at time period j are
accounted for in the flow balance constraint. If the previous plan is implemented exactly,
Vj is given by Equation 4.60. Still in Equation 4.57, Mqj
0
represent the updated capacity
profiles. Finally, Equation 4.65 ensures that the airborne queue Wq,T

0 1 at the beginning of
the updated planning horizon is accounted for in the flow balance constraint.

X
Tf Tf Tf +1
X X Q Tf Tf Tf
1 XX X X
min cg (j i)x0sij + ca Wqj0 + ↵✓ + (ysi + zsi ) (4.55)
s=T 0 i=T 0 j=i
Q q=1 j=T 0 s=T 0 i=T 0

Subject to:

Tf +1
X
x0sij = Nsi0 , 8s 2 {T 0 , ..., Tf }, 8i 2 {T 0 , ..., Tf } (4.56)
j=i

Tf j
X X
x0sij + Vj + Wq,j
0
1 Wqj0  Mqj
0
, 8j 2 {T 0 , ..., Tf }, 8q 2 {1, ..., Q} (4.57)
s=T 0 i=T 0

Tf
X
✓ Wqj0 , 8q 2 {1, ..., Q} (4.58)
j=T 0

8
> P
>
> N si + xlmn , s = T 0, 8i s
>
< {(l,m,n):l<T 0 ;
Nsi0 = l+n m T 0 ; (4.59)
>
> m l=i s}
>
>
:N , 8s > T 0 , 8i s
si

169
X
Vj = xsij , 8j T0 (4.60)
{(s,i):s+j i<T 0 }

Tf +1 Tf +1
X X
ysi (j i)x0sij (j i)x⇤sij , 8s > T 0 , 8i s (4.61)
j=i j=i

Tf +1
X X
ysi (j i)x0sij (l + n m T 0 )x⇤sij , s = T 0, 8i s (4.62)
j=i {(l,m,n):lT 0 ;
n m;l+n m T 0 ;
m l=i s}

Tf +1 Tf +1
X X
zsi (j i)x⇤sij (j i)x0sij , 8s > T 0 , 8i s (4.63)
j=i j=i

Tf +1
X X
zsi (l + n m T 0
)x⇤sij (j i)x0sij , s = T 0, 8i s (4.64)
{(l,m,n):lT 0 ; j=i
n m;l+n m T 0 ;
m l=i s}

0
Wq,T 0 1

= Wq,T 0 1, 8q 2 {1, ..., Q} (4.65)

x0sij 2 Z+ , 8s 2 {T 0 , ..., Tf }, 8i 2 {T 0 , ..., Tf }, 8j 2 {T 0 , ..., Tf + 1} (4.66)

Wqj0 2 Z+ , 8j 2 {T 0 , ..., Tf }, 8q 2 {1, ..., Q} (4.67)

ysi , zsi 2 Z+ , 8s 2 {T 0 , ..., Tf }, 8i 2 {T 0 , ..., Tf } (4.68)

✓ 2 Z+ (4.69)

170
4.6.3 Analysis of Impacts on GDP planning

Test cases and assumptions

The AAR planning model described in Section 4.6.2 was used to optimize the traffic flow
management plan for five test cases corresponding to historical GDP events at JFK (Table
4.2). The optimization problem was solved with Gurobi optimizer (version 6.5.0). The cost
ratio between airborne and ground delay was assumed at 1.5 (cg = 1 and ca = 1.5), based
on delay costs reported by the industry [146, 147]. The robustness parameters ↵ and were
varied from 0 to 10 in small increments of 0.02 in order to investigate the trade-offs between
efficiency and predictability. Finally, we assumed that the plan could be revised once and
that the revision was performed in the middle of the planning horizon.

Table 4.2: Description of GDP events at JFK used as test cases.

Test case Date Planning horizon (h)


1 20130131 6
2 20130217 5
3 20130522 6
4 20131127 6
5 20140812 5

For each GDP event, 100 capacity profiles representing possible evolutions of capacity
throughout the GDP planning horizon were drawn from the predicted arrival rate distribu-
tions generated by the prediction framework. The choice of number of capacity profiles was
made considering trade-offs between variability in expected delay costs and computational
time. Figure 4-23 illustrates these trade-offs for test case 1 (which was optimized with the
simpler static AAR planning model). As the number of sampled capacity profiles increases,
the variability in total delay costs decreases, while the computational time for obtaining the
optimal solution increases. For more than 100 capacity profiles, the expected costs tend to
converge, revealing that this number of capacity profiles is a reasonable choice for not much
increasing the computational effort.

171
Figure 4-23: Expected delay costs and computational time as a function of the number of
capacity profiles.

Impacts of robustness parameter ↵

We first assessed the trade-offs between efficiency and predictability by varying the robustness
parameter ↵, which controls the impacts of the robustness term that penalizes unpredictable
airborne delays, during the optimization of the GDP plan for each test case. For this analysis,
the robustness parameter was set to zero, i.e., changes in ground delays were not penalized
during the revision.
The planned GDP costs obtained for different values of ↵ are shown in Figures 4-24 and
4-25. The planned ground delay costs are shown in Figure 4-24, and the expected airborne
delay costs are shown in Figure 4-25. As expected, as the decision-maker becomes less
risk tolerant and increases the value of ↵, the robustness term more heavily impacts the
objective function and the expected airborne delay costs decrease at the cost of increased
ground delays. It is also observed that the expected airborne delay costs are very sensitive
to changes in ↵ for small values (less than 1).

172
Figure 4-24: Planned ground delay costs as Figure 4-25: Expected airborne delay costs
a function of robustness parameter ↵. as a function of robustness parameter ↵.

Figure 4-26 shows how the optimized GDP plan performs under the actual capacity
profiles. Specifically, it shows the sum of the planned ground delay costs and the actual
1
airborne delay costs as a function of ↵. When robustness is heavily accounted for in the
objective function (large values of ↵), a decrease in efficiency is observed, as expected. Yet,
for most of the test cases, it is also noted that reductions in the actual delay costs could be
obtained with the introduction of the robustness term for smaller values of ↵. The average
actual delay costs for all test cases is shown in Figure 4-27. For ↵ < 1, introducing robustness
turned out to increase efficiency by protecting against imperfect capacity predictions. The
minimum cost was observed for ↵ = 0.6.
1
The actual airborne delay costs were calculated by running the AAR planning model with the optimized
GDP plan solution frozen (xsij = x⇤sij , where x⇤sij is the optimized ground delay solution) and with the
predicted capacity profiles replaced by the actual capacity profile at the airport. In fact, this corresponds
to a deterministic queueing model governed by the flow balance constraint given by Equation 4.49: for any
time period t 2 T = {Ti , ..., T10 1}, the number of departing flights and the number of ground-held flights
matches the optimized GDP plan solution; if the number of flights planned to arrive at the airport is less
than or equal to the actual capacity during time period t, all flights land; otherwise, the number of flights
that land is equal to the actual capacity and the remaining flights are held in the air until the next time
period; for any t from T 0 onwards, the process is repeated, but with the revised GDP plan solution. In
summary, we refer to actual airborne delays as those associated with the actual capacity profiles, under the
assumptions of the AAR planning model, and without any relation to historical delays incurred by the flights
affected by the actual GDP events.

173
Figure 4-26: Actual delay costs as a function Figure 4-27: Actual delay costs (averaged
of robustness parameter ↵. over all test cases) as a function of robust-
ness parameter ↵.

The main goal of incorporating robustness in the objective function for optimization of
the GDP plan is increased predictability. Figure 4-28 shows the difference between the actual
delay costs and the planned delay costs as a function of ↵. Indeed, the introduction of the
robustness term increases GDP delay cost predictability. This is driven by a reduction in
unexpected airborne delays, as shown in Figure 4-25.
Finally, Figure 4-29 shows the impacts of ↵ on the optimized arrival rates, which are
displayed as the corresponding percentile of the predicted capacity distributions. As ↵
increases, the optimal arrival rate reduces, indicating that fewer flights should be allocated
in order to reduce the risk of unplanned airborne delays. For each test case, the optimal
arrival rate that minimizes the actual delay cost is also plotted. It is observed that it ranges
from the 28th to the 47th percentile of the predicted capacity distributions.

Impacts of robustness parameter

This section shows the results of the GDP plan optimization for various values of , which
controls the impacts of the robustness term that penalizes changes in ground delays during
the revision. For this analysis, the robustness parameter ↵ was set to zero.

174
Figure 4-28: GDP delay cost predictability Figure 4-29: Optimized arrival rates as a
as a function of robustness parameter ↵. function of robustness parameter ↵.

Figures 4-30 and 4-31 show the difference between the planned ground delays and the
expected airborne delays for the revised GDP plan and the initial GDP plan. When = 0,
the highest changes in ground delays were observed, revealing that updated information was
indeed leveraged to further optimize the plan during the revision. With the exception of test
case 2, more ground delays were assigned during the revision. Indeed, Figure 4-31 shows
that the updated information led to an increase in the expected airborne delays for all test
cases. When changes in ground delays were "allowed" during the revision (small values of ),
the model counterbalanced this increase in the expected airborne delays by assigning more
ground delays in order to minimize the expected overall delay costs. The only exception was
test case 2, as no changes in ground delays were observed despite the expected increase in
airborne delays. Further analysis revealed that, for test case 2, all flights eligible to have an
updated ground delay assignment had already been assigned the maximum possible ground
delay in the original plan, as they were scheduled to arrive at the hour after the end of the
program.
Figure 4-32 shows the planned delay costs associated with the revised and the initial GDP
plans, averaged over all test cases. Figure 4-33 shows the actual delay costs resulting from
the implementation of the revised and the initial GDP plans, averaged over all test cases. It
is observed that the revision was useful to increase the predictability of the expected GDP
delay costs, as the planned costs under the revised plan were closer to the actual delay costs.

175
Furthermore, when changes in ground delays were not penalized and were fully allowed
during the revision ( = 0), an average increase in efficiency of 2% was observed.

Figure 4-30: Difference between the revised Figure 4-31: Difference between the revised
and the initially planned ground delays as a and the initially expected airborne delays as
function of robustness parameter . a function of robustness parameter .

Figure 4-32: Planned delay costs for revised Figure 4-33: Actual delay costs for revised
and initial plans as a function of robustness and initial plans as a function of robustness
parameter (averaged over all test cases). parameter (averaged over all test cases).

Comparison between GDP costs resulting from data-driven and baseline AAR

The delay costs resulting from the use of the optimized AARs and the AARs actually im-
plemented by the FAA for designing the GDP plan were compared. Figure 4-34 shows that,

176
with the exception of test case 5, lower GDP costs were achieved with the data-driven capac-
ity predictions. When robustness was not accounted in the optimization of the GDP plan
(↵ = 0), a 2.4% reduction in the total GDP costs (sum of all test cases) was observed. When
robustness was accounted for in the optimization of the GDP plan, with ↵ = 0.6 (observed
value that contributed to increase efficiency on average), the reduction in total GDP costs
was 9.7%. Figure 4-35 shows the MAPE for the capacity predictions generated by the GP
model (averaged over the GDP planning horizon) for each test case. It is observed that,
although test case 5 was the exception, it showed similar level of error as the other cases,
suggesting that the AARs actually implemented on that date were closer to the true capacity
at the airport.

Figure 4-34: Comparison between GDP costs resulting from data-driven and baseline AAR.

177
Figure 4-35: MAPE of GP model capacity predictions (averaged over planning horizon) for
GDP test cases.

178
Chapter 5

Conclusions and Future Work

5.1 Summary
In this thesis, a flight trajectory data analytics framework was developed to provide a high-
fidelity characterization of air traffic flows from large-scale aircraft tracking data. Through
the application of machine learning methods to discover spatial and temporal trends in
aircraft movement, the framework allows to automatically learn the airspace structure, assess
the use of the airspace and identify patterns of usage of the airspace. For this, it includes three
modules: (1) clustering flight trajectories at spatial scale to identify trajectory patterns, (2)
trajectory classification to assess flight trajectory conformance against the learned airspace
structure and identify air traffic flows, (3) clustering air traffic flows at temporal scale to
identify traffic flow patterns. The framework was used to provide a detailed characterization
of air traffic flows in the terminal area of multi-airport systems.
We first leveraged the knowledge generated by the framework to develop a systematic
descriptive approach to analyze metroplex airspace design and use and to assess operational
performance. Novel quantitative metrics were created to summarize metroplex performance
in the areas of efficiency, capacity and predictability. The descriptive approach was demon-
strated with a comparative analysis of terminal area air traffic operations in three represen-
tative multi-airport systems: New York, Hong Kong and Sao Paulo. The results revealed
structural, operational and performance differences between these systems. We found that
the New York multi-airport system presents the most complex airspace design and the most

179
dynamic flow behavior. Interestingly, it exhibits the best levels of trajectory lateral and
temporal efficiency on average, yet the highest variability in operational performance. By
contrast, the Hong Kong and Sao Paulo multi-airport systems show a relative simple and
clean airspace design and lower variability in traffic flow behavior and performance. The re-
sults also highlighted different areas of action to be considered at each multi-airport system
towards improving their air traffic operations. The Hong Kong metroplex route structure
was found to be considerably less efficient, particularly for HKG, revealing opportunities
for improvement in airspace design in order to increase trajectory lateral performance. The
lower level of trajectory conformance and temporal efficiency observed for the Hong Kong
and Sao Paulo metroplexes revealed opportunities for improvement in tactical traffic flow
management towards better sequencing and scheduling of arriving flights. Sao Paulo was
found to be the lowest throughput performing system, emphasizing the importance of ac-
tions to expand capacity at the individual airports. Finally, the higher level of airport
interdependency and operational variability observed in the New York metroplex revealed
opportunities for improving the airspace design in order to reduce flow interdependencies and
provide more consistent all-weather operations. The particular influence of weather condi-
tions on airspace use and traffic performance in the New York metroplex also highlighted
the importance of integrating weather forecasting with decision-making in order to better
anticipate weather conditions, translate them into operational impact and plan traffic flow
management efficiently at this system.
Finally, in the last part of the thesis, we leveraged the knowledge generated by the flight
trajectory data analytics framework to develop a data-driven approach for metroplex config-
uration and airport capacity planning towards improved decision support. The approach has
two building blocks: predictive modeling for capacity estimation and prescriptive modeling
for capacity allocation.
In the capacity estimation block, machine learning methods are used to translate weather
forecasts into probabilistic forecasts of the metroplex configuration and airport capacity for
strategic planning horizons. First, a traffic flow pattern prediction model was developed
to deliver a probabilistic forecast of the traffic flow structure to be seen in the metroplex
terminal airspace on an hourly basis. At any given time, the metroplex flow pattern reflects

180
a particular utilization pattern of runways and airspace, providing a complete picture of
the multi-airport system configuration state. Three candidate traffic flow pattern prediction
models were created using different supervised learning methods for multi-way classification.
After performance evaluation, a Random Forests model was selected for traffic flow pattern
prediction. For the New York multi-airport system, the model showed an average prediction
accuracy of 83% for a short-term 1-hour forecast, 64% for a 3-hour forecast, and 52% for a
6-hour forecast. Despite the diminishing accuracy for long-term forecasts, it showed higher
performance when compared with a static model that only considers the current flow pattern
to predict the next hours. For a 6-hour forecast horizon, the performance gap was 13%. The
probabilistic forecast was found to be useful, as the results indicated that, for many instances
in which the model incorrectly predicted the traffic flow pattern, it was right on the second
best option. Indeed, the prediction accuracy was 93% for a 1-hour forecast and stabilized
at about 77% for longer look-ahead times, if the top two most likely flow patterns in the
probabilistic forecast were considered instead of just the most likely one.
Second, an airport capacity prediction model was developed to deliver probabilistic arrival
capacity forecasts on an hourly basis. The model takes as input the metroplex flow pattern
forecast and weather features representing the forecast environmental conditions in the ter-
minal area. Three different methods for machine learning were evaluated in a supervised
learning regression scheme for arrival rate prediction: Bayesian Linear Regression, Random
Forests, and Gaussian Process. The Gaussian Process model was chosen based on accuracy of
point prediction and uncertainty quantification. An analysis of feature importance revealed
that the metroplex configuration is the second most important factor influencing the arrival
rates. Indeed, it was found that the two-layer prediction model outperformed a simplified
model that does not account for metroplex system state in the capacity estimation process,
especially for JFK and for longer forecast horizons. We also found that the data-driven
capacity predictions obtained with the Gaussian Process model reduced the prediction error
by 5.4% at JFK, 3.0% at EWR and 1.5% at LGA when compared with the baseline capacity
estimates reported by the FAA.
Finally, in the capacity allocation block, optimization methods are used for AAR pre-
scription and ground delay planning towards demonstrating the usability of the data-driven

181
capacity predictions and evaluating their impacts in the design of strategic TMIs such as
GDPs. An AAR planning model was developed to prescribe optimal AARs that minimize
the overall expected delay costs. The model has the following key features: (1) it is stochas-
tic, i.e., capacity uncertainty is accounted for in the form of capacity profiles drawn from
actual predictive distributions; (2) it enables revision of the plan in order to take advantage
of updated information and more accurate predictions throughout the planning horizon; (3)
it incorporates robustness goals, i.e., the decision-maker has the ability to select a desired
level robustness for the TFM plan based on its relative valuation between efficiency and pre-
dictability. The model was used to optimize the traffic flow management plan for five test
cases corresponding to historical GDP events at JFK. The results revealed that incorporating
robustness in the design of the TFM plan not only helps increasing delay predictability, but
can also lead to increased efficiency by protecting against errors in capacity predictions. The
optimized AARs generated an overall reduction of 2.4% in the total GDP costs, if compared
with the AARs actually implemented by the FAA. When robustness was accounted for in
the optimization of the GDP plan, the observed reduction in total GDP costs was 9.7%.

5.2 Future Research Directions


The flight trajectory data analytics framework was demonstrated at the terminal area scale
in this thesis. One clear direction for future research involves the use of the framework for
characterization of air traffic flows at other scales. For instance, characterizing national-level
air traffic flows can generate useful insights about en route traffic behavior for supporting
en route air traffic management. One might be interested in exploiting routing patterns
between origin and destination to diagnose en route flight inefficiencies, to better understand
the factors that drive the selection of a particular route by flight operators, to discover
feasible rerouting options or to assess the impacts of adverse weather conditions (such as
convective weather) on route acceptability and capacity. At the sector level, characterization
of traffic flow patterns may be useful for complexity management applications. One might be
interested in exploiting patterns of air traffic control complexity in order to predict periods
of low/high complexity and better inform staffing or sector reconfiguration decisions. While

182
the methods incorporated in the framework have general applicability, adaptations should
be needed (e.g., data representation for clustering) to tailor the analysis for the intended
application.
The predictive models developed in this thesis could benefit from higher quality infor-
mation describing the input variables. For instance, traffic flow pattern predictions could
be improved if actual demand information is used instead of flight schedules. In an actual
implementation, this information could be retrieved from the Flight Scheduled Monitor of
the TFMS. In the same way, capacity predictions could be improved with the use of different
weather forecasts. Indeed, one particular limitation of the predictive models developed in
this work is the use of short-term forecasts of convective weather impacts (ARSI). Ideally,
the models should also incorporate features associated with long-term forecasts of convec-
tive weather impacts to improve the quality of strategic airport capacity predictions. In
this direction, existing weather translation tools could be leveraged, such as the Traffic Flow
Impact [148], which was designed to translate long-term convective weather forecasts into a
measure of airspace permeability for strategic time horizons.
Within the airport capacity planning framework, one potential research direction is ex-
tending the capacity allocation model to account for differences in robustness goals between
different flight operators, along the lines of decentralized approaches for traffic flow manage-
ment, and investigating whether it is possible to prescribe an optimal time for revision of
the traffic flow management plan based on the uncertainty in capacity forecasts.
Finally, this thesis was focused on evaluating the potential impacts of data-driven models
towards supporting ATM decision-making using historical data, but a lot of work is required
into actual decision-support tool development. In this direction, human-in-the-loop simu-
lations could be performed in order to test the models for a wider range of scenarios and
obtain user feedback. As mentioned previously, the predictive models could be enhanced
with a richer set of features. Research efforts should also focus on how these tools can be
designed to best leverage the strengths of analytics and computation to enable fast and
reliable real-time implementation as well as to foster human engagement.

183
THIS PAGE INTENTIONALLY LEFT BLANK

184
Appendix A

Trajectory Clustering Analysis Results -


New York Metroplex

Table A.1: DBSCAN parameter settings used for trajectory clustering.

Airport flow MinPts " Number of clusters identified


JFK Arrivals 6 1.4 20
EWR Arrivals 6 1.4 14
LGA Arrivals 6 1.2 16
JFK Departures 6 1.2 24
EWR Departures 6 1.3 13
LGA Departures 6 1.3 18

185
Figure A-1: JFK arrival trajectory clusters; Figure A-2: JFK arrival trajectories labeled
each color represents one cluster. as noise.

Figure A-3: Centroids of JFK arrival trajec- Figure A-4: Distribution of JFK arrival tra-
tory clusters. jectories by cluster; grey bar represents the
percentage of noise.

186
Figure A-5: JFK departure trajectory clus- Figure A-6: JFK departure trajectories la-
ters; each color represents one cluster. beled as noise.

Figure A-7: Centroids of JFK departure tra- Figure A-8: Distribution of JFK departure
jectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

187
Figure A-9: EWR arrival trajectory clus- Figure A-10: EWR arrival trajectories la-
ters; each color represents one cluster. beled as noise.

Figure A-11: Centroids of EWR arrival tra- Figure A-12: Distribution of EWR arrival
jectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

188
Figure A-13: EWR departure trajectory Figure A-14: EWR departure trajectories
clusters; each color represents one cluster. labeled as noise.

Figure A-15: Centroids of EWR departure Figure A-16: Distribution of EWR depar-
trajectory clusters. ture trajectories by cluster; grey bar repre-
sents the percentage of noise.

189
Figure A-17: LGA arrival trajectory clus- Figure A-18: LGA arrival trajectories la-
ters; each color represents one cluster. beled as noise.

Figure A-19: Centroids of LGA arrival tra- Figure A-20: Distribution of LGA arrival
jectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

190
Figure A-21: LGA departure trajectory Figure A-22: LGA departure trajectories la-
clusters; each color represents one cluster. beled as noise.

Figure A-23: Centroids of LGA departure Figure A-24: Distribution of LGA departure
trajectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

191
THIS PAGE INTENTIONALLY LEFT BLANK

192
Appendix B

Trajectory Clustering Analysis Results -


Hong Kong Metroplex

Table B.1: DBSCAN parameter settings used for trajectory clustering.

Airport flow MinPts " Number of clusters identified


HKG Arrivals 6 1.4 15
SZX Arrivals 6 1.4 8
MFM Arrivals 6 1.6 6
HKG Departures 6 1.4 10
SZX Departures 6 1.4 12
MFM Departures 6 1.4 6

193
Figure B-1: HKG arrival trajectory clusters; Figure B-2: HKG arrival trajectories la-
each color represents one cluster. beled as noise.

Figure B-3: Centroids of HKG arrival tra- Figure B-4: Distribution of HKG arrival tra-
jectory clusters. jectories by cluster; grey bar represents the
percentage of noise.

194
Figure B-5: HKG departure trajectory clus- Figure B-6: HKG departure trajectories la-
ters; each color represents one cluster. beled as noise.

Figure B-7: Centroids of HKG departure Figure B-8: Distribution of HKG departure
trajectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

195
Figure B-9: SZX arrival trajectory clusters; Figure B-10: SZX arrival trajectories la-
each color represents one cluster. beled as noise.

Figure B-11: Centroids of SZX arrival tra- Figure B-12: Distribution of SZX arrival
jectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

196
Figure B-13: SZX departure trajectory clus- Figure B-14: SZX departure trajectories la-
ters; each color represents one cluster. beled as noise.

Figure B-15: Centroids of SZX departure Figure B-16: Distribution of SZX departure
trajectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

197
Figure B-17: MFM arrival trajectory clus- Figure B-18: MFM arrival trajectories la-
ters; each color represents one cluster. beled as noise.

Figure B-19: Centroids of MFM arrival tra- Figure B-20: Distribution of MFM arrival
jectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

198
Figure B-21: MFM departure trajectory Figure B-22: MFM departure trajectories
clusters; each color represents one cluster. labeled as noise.

Figure B-23: Centroids of MFM departure Figure B-24: Distribution of MFM depar-
trajectory clusters. ture trajectories by cluster; grey bar repre-
sents the percentage of noise.

199
THIS PAGE INTENTIONALLY LEFT BLANK

200
Appendix C

Trajectory Clustering Analysis Results -


Sao Paulo Metroplex

Table C.1: DBSCAN parameter settings used for trajectory clustering.

Airport flow MinPts " Number of clusters identified


GRU Arrivals 6 1.4 10
CGH Arrivals 6 1.4 6
VCP Arrivals 6 1.4 8
GRU Departures 6 1 14
CGH Departures 6 1 10
VCP Departures 6 1.4 12

201
Figure C-1: GRU arrival trajectory clusters; Figure C-2: GRU arrival trajectories labeled
each color represents one cluster. as noise.

Figure C-3: Centroids of GRU arrival tra- Figure C-4: Distribution of GRU arrival tra-
jectory clusters. jectories by cluster; grey bar represents the
percentage of noise.

202
Figure C-5: GRU departure trajectory clus- Figure C-6: GRU departure trajectories la-
ters; each color represents one cluster. beled as noise.

Figure C-7: Centroids of GRU departure Figure C-8: Distribution of GRU departure
trajectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

203
Figure C-9: CGH arrival trajectory clusters; Figure C-10: CGH arrival trajectories la-
each color represents one cluster. beled as noise.

Figure C-11: Centroids of CGH arrival tra- Figure C-12: Distribution of CGH arrival
jectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

204
Figure C-13: CGH departure trajectory Figure C-14: CGH departure trajectories la-
clusters; each color represents one cluster. beled as noise.

Figure C-15: Centroids of CGH departure Figure C-16: Distribution of CGH depar-
trajectory clusters. ture trajectories by cluster; grey bar repre-
sents the percentage of noise.

205
Figure C-17: VCP arrival trajectory clus- Figure C-18: VCP arrival trajectories la-
ters; each color represents one cluster. beled as noise.

Figure C-19: Centroids of VCP arrival tra- Figure C-20: Distribution of VCP arrival
jectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

206
Figure C-21: VCP departure trajectory Figure C-22: VCP departure trajectories la-
clusters; each color represents one cluster. beled as noise.

Figure C-23: Centroids of VCP departure Figure C-24: Distribution of VCP departure
trajectory clusters. trajectories by cluster; grey bar represents
the percentage of noise.

207
THIS PAGE INTENTIONALLY LEFT BLANK

208
Bibliography

[1] J. Gudmundsson, P. Laube, and T. Wolle. Encyclopedia of GIS, chapter Movement


patterns in spatio-temporal data, pages 726–732. Springer, 1st edition, 2008.

[2] New York Air Route Traffic Control Center. TRACON letters of agreement. http:
//nyartcc.org/wiki/index.php/LGA_SOP, (accessed 2017-06).

[3] J. P. B. Clarke, L. Ren, E. McClain, D. Schleicher, S. Timar, A. Saraf, D. Crisp,


R. Gutterud, R. Laroza, T. Thompson, C. Cross, and T. Lewis. Evaluating concepts
for operations in metroplex terminal area airspace. Journal of Aircraft, 49(3):758–773,
2012.

[4] International Air Transport Association. Fact Sheet: Industry Statistics. http://www.
iata.org/pressroom/facts_figures/Pages/index.aspx, (accessed 2017-06).

[5] Air Transport Action Group. Aviation: Benefits Beyond Borders. https://
www.aviationbenefits.org/media/149668/abbb2016_full_a4_web.pdf, (accessed
2017-06).

[6] International Air Transport Association. IATA forecasts passenger demand to double
over 20 years. http://www.iata.org/pressroom/pr/Pages/2016-10-18-02.aspx,
(accessed 2017-06).

[7] Federal Aviation Administration. FAA Aerospace Forecast - Fiscal Years 2016-
2036. https://www.faa.gov/data_research/aviation/aerospace_forecasts/
media/FY2016-36_FAA_Aerospace_Forecast.pdf, (accessed 2017-06).

[8] International Civil Aviation Organization. Global Air Navigation Plan for CNS/ATM
Systems. http://www.icao.int/publications/Documents/9750_2ed_en.pdf, (ac-
cessed 2017-06).

[9] Bureau of Transportation Statistic. On-time performance - flight delays at a


glance. http://www.transtats.bts.gov/HomeDrillChart.asp?URL_SelectMonth=
11&URL_SelectYear=2016, (accessed 2017-06).

[10] Eurocontrol. Annual Network Operations Report 2015. https://www.eurocontrol.


int/publications/annual-network-operations-report-2015, (accessed 2017-06).

209
[11] National Center of Excellence for Aviation Operations Research. Total De-
lay Impact Study. http://its.berkeley.edu/sites/default/files/NEXTOR_TDI_
Report_Final_October_2010.pdf, (accessed 2017-06).

[12] Federal Aviation Administration. NextGen. https://www.faa.gov/nextgen/, (ac-


cessed 2017-06).

[13] Single European Sky ATM Research. Discover SESAR. http://www.sesarju.eu/


index.php/discover-sesar, (accessed 2017-06).

[14] Transport Ministry of Land, Infrastructure and Tourism. CARATS: Collaborative


Actions for Renovation of Air Traffic Systems. http://www.mlit.go.jp/common/
000128185.pdf, (accessed 2017-06).

[15] Brazilian Department of Airspace Control. SIRIUS Program. https://www.decea.


gov.br/sirius/?lang=en, (accessed 2017-06).

[16] International Civil Aviation Organization. A39-WP/304: China’s Strategy for Modern-
izing Air Traffic Management. https://www.icao.int/Meetings/a39/Documents/
WP/wp_304_en.pdf, (accessed 2017-06).

[17] Federal Aviation Administration. The Future of the NAS. https://www.faa.gov/


nextgen/media/futureofthenas.pdf, (accessed 2017-06).

[18] Eurocontrol. European ATM Master Plan 2015: The Roadmap for Delivering High Per-
forming Aviation for Europe. https://ec.europa.eu/transport/sites/transport/
files/modes/air/sesar/doc/eu-atm-master-plan-2015.pdf, (accessed 2017-06).

[19] H. Hu, Y. Wen, T.-S. Chua, and X. Li. Toward scalable systems for big data analytics:
A technology tutorial. IEEE Access, 2:652–687, 2014.

[20] G. Blackett. Analytics Network - OR and Analytics. https://www.theorsociety.


com/Pages/SpecialInterest/AnalyticsNetwork_analytics.aspx, (accessed 2017-
06).

[21] D. Bertsimas, A. O’Hair, and D. Pulleyblank. The Analytics Edge. Dynamic Ideas
LLC, 1st edition, 2016.

[22] Commission of the European Communities. Air Traffic Management. http://aei.


pitt.edu/1134/1/air_traffic_white_paper_COM_96_57.pdf, (accessed 2017-06).

[23] P. Belobaba, A. Odoni, and C. Barnhart. The Global Airline Industry. Wiley, 2nd
edition, 2015.

[24] International Civil Aviation Organization. Rules of the Air, Annex 2 of the Convention
on International Civil Aviation. ICAO, 10th edition, 2005.

210
[25] Federal Aviation Administration. AC 90-115 - Collaborative Trajectory Op-
tions Program (CTOP). https://www.faa.gov/regulations_policies/advisory_
circulars/index.cfm/go/document.information/documentID/1024987, (accessed
2017-06).

[26] N. M. Smith, C. Brasil, P. U. Lee, N. Buckley, C. Gabriel, C. Mohlenbrink, F. Omar,


B. Parke, C. Speridakos, and H. S. Yoo. Integrated demand management: Coordinating
strategic and tactical flow scheduling operations. In 16th AIAA Aviation Technology,
Integration, and Operations Conference, 2016.

[27] Federal Aviation Administration. Air Traffic by the Numbers. https://www.faa.gov/


air_traffic/by_the_numbers/, (accessed 2017-06).

[28] Federal Aviation Administration. Passenger boarding (enplanement) and all-cargo data
for U.S. airports. https://www.faa.gov/airports/planning_capacity/passenger_
allcargo_stats/passenger/, (accessed 2017-06).

[29] R. De Neufville and A. Odoni. Airport Systems: Planning, Design, and Management.
McGraw-Hill, 2nd edition, 2013.

[30] P. A. Bonnefoy. Scalability of the Air Transportation System and Development of


Multi-Airport Systems: A Worldwide Perspective. Ph.D. dissertation, Massachusetts
Institute of Technology, 2008.

[31] L. Ren, J. P. B. Clarke, D. Schleicher, S. Timar, A. Saraf, D. Crisp, R. Gutterud,


T. Lewis, and T. Thompson. Contrast and comparison of metroplex operations (an
air traffic management study of Atlanta, Los Angeles, New York, and Miami. In 9th
AIAA Aviation Technology, Integration, and Operations Conference, 2009.

[32] A. D. Donaldson and R. J. Hansman. Capacity improvement potential for the new
york metroplex system. In 10th AIAA Aviation Technology, Integration and Operations
Conference, 2010.

[33] A. R. Odoni. Flow Control of Congested Networks, chapter The Flow Management
Problem in Air Traffic Control, pages 269–288. Springer-Verlag, Berlin, 1987.

[34] M. Terrab and A. R. Odoni. Strategic flow management for air traffic control. Opera-
tions Research, 41(1):138–152, 1993.

[35] O. Richetta and A. R. Odoni. Solving optimally the static ground-holding policy
problem in air traffic control. Transportation Science, 27(3):228–238, 1993.

[36] M. Ball, A. R. Odoni, and R. Rifkin. Stochastic integer program with dual network
structure and its application to the ground-holding problem. Operations Research,
51(1):167–171, 2003.

[37] O. Richetta and A. R. Odoni. Dynamic solution to the ground-holding problem in air
traffic control. Transportation Research Part A: Policy and Practice, 28(3):167–185,
1994.

211
[38] A. Mukherjee and M. Hansen. Dynamic stochastic model for the single airport ground
holding problem. Transportation Science, 41(4):444–456, 2007.

[39] Y. Liu and M. Hansen. Incorporating predictability into cost optimization for ground
delay programs. Transportation Science, 50(1):132–149, 2015.

[40] P. Swaroop and M. Ball. Consensus-building mechanism for setting service expecta-
tions in air traffic flow management. Transportation Research Record: Journal of the
Transportation Research Board, 2325:87–96, 2013.

[41] A. Evans, V. Vaze, and C. Barnhart. Airline-driven performance-based air traffic


management: Game theoretic models and multicriteria evaluation. Transportation
Science, 50(1):180–203, 2016.

[42] M. Ball, P. Swaroop, C. Barnhart, C. Yan, M. Hansen, L. Kang, Y. Liu, and V. Vaze.
Service level expectation setting for air traffic flow management: Practical challenges
and benefits assessment. In 12th USA/Europe Air Traffic Management Research and
Development Seminar, 2017.

[43] C. Yan, V. Vaze, and C. Barnhart. Airline-driven ground delay programs: a benefits
assessment. Transportation Research Part C: Emerging Technologies, 89:268–288, 2018.

[44] D. Bertsimas and S. S. Patterson. The air traffic flow management problem with
enroute capacities. Operations Research, 46(3):406–422, 1998.

[45] D. Bertsimas and S. S. Patterson. The traffic flow management rerouting problem in air
traffic control: a dynamic network flow approach. Transportation Science, 34:239–255,
2000.

[46] D. Bertsimas, G. Lulli, and A. R. Odoni. An integer optimization approach to large-


scale air traffic flow management. Operations Research, 59(1):211–227, 2011.

[47] A. Agustín, Alonso-Ayuso A., L. F. Escudero, and C. Pizarro. On air traffic flow man-
agement with rerouting. Part i: Deterministic case. European Journal of Operational
Research, 219:156–166, 2012.

[48] A. Agustín, Alonso-Ayuso A., L. F. Escudero, and C. Pizarro. On air traffic flow
management with rerouting. Part ii: Stochastic case. European Journal of Operational
Research, 219:167–177, 2012.

[49] H. Balakrishnan and B. G. Chandran. Optimal large-scale air traffic flow management.
http://web.mit.edu/Hamsa/www/pubs/BalakrishnanChandran_ATFM.pdf, 2014.

[50] P. K. Menon, G. D. Sweriduk, T. Lam, G. M. Diaz, and K. Bilimoria. Computer-aided


eulerian air traffic flow modeling and predictive control. AIAA Journal of Guidance,
Control and Dynamics, 29:12–19, 2006.

[51] B. Sridhar, T. Soni, K. Sheth, and G. B. Chatterji. Aggregate flow model for air-traffic
management. Journal of Guidance, Control, and Dynamics, 26:992–997, 2006.

212
[52] D. Sun and A. M. Bayen. Multicommodity eulerian-lagrangian large-capacity cell
transmission model for en route traffic. Journal of Guidance, Control, and Dynamics,
31(3):616–628, 2008.

[53] A. Marzuoli, M. Gariel, A. Vela, and E. Feron. Data-based modeling and optimization
of en route traffic. Journal of Guidance, Control, and Dynamics, 37(6):1930–1945,
2014.

[54] A. Bombelli, L. Soler, E. Trumbauer, and K. D. Mease. Strategic air traffic planning
with fréchet distance aggregation and rerouting. Journal of Guidance, Control, and
Dynamics, 40(5):1117–1129, 2017.

[55] J. E. Beasley, M. Krishnamoorthy, Y.M. Sharaiha, and D. Abramson. Scheduling


aircraft landings - the static case. Transportation Science, 34(2):180–197, 2000.

[56] H. Balakrishnan and B. Chandran. Algorithms for scheduling runway operations under
constrained position shifting. Operations Research, 58(6):1650–1665, 2010.

[57] G. Solveling, S. Solak, J. P. Clarke, and E. Johnson. Runway operations optimiza-


tion in the presence of uncertainties. Journal of Guidance, Control, and Dynamics,
34(5):1373–1382, 2011.

[58] M. C. R. Murça and C. Müller. Control-based optimization approach for aircraft


scheduling in a terminal area with alternative arrival routes. Transportation Research
Part E, 73:96–113, 2015.

[59] D. Bertsimas, M. Frankovich, and A. Odoni. Optimal selection of airport runway


configurations. Operations research, 59(6):1407–1419, 2011.

[60] S. Rathinam, J. Montoya, and Y. Jung. An optimization model for reducing aircraft
taxi times at the dallas fort worth international airport. In 26th International Congress
of the Aeronautical Sciences, 2008.

[61] W. Malik, G. Gupta, and Y. Jung. Managing departure aircraft release for efficient
airport surface operations. In AIAA Guidance, Navigation, and Control Conference,
2010.

[62] I. Simaiakis, M. Sandberg, and H. Balakrishnan. Dynamic control of airport depar-


tures: algorithm development and field evaluation. IEEE Transactions on Intelligent
Transportation Systems, 15(1):285–295, 2013.

[63] C. Verlhac and S. Manchon. Optimization of opening scheme. In 4th USA/Europe Air
Traffic Management Research and Development Seminar, 2001.

[64] M. Bloem, P. Kopardekar, and P. Gupta. Algorithms for combining airspace sectors.
Air Traffic Control Quarterly, 17(3):245–268, 2009.

[65] D. Gianazza. Forecasting workload and airspace configuration with neural networks
and tree search methods. Artificial Intelligence, 174(7):530–549, 2010.

213
[66] H. Trandac, P. Baptiste, and V. Duong. A constraint-programming formulation for
dynamic airspace sectorization. In AIAA/IEEE Digital Avionics Systems Conference,
2002.

[67] M. Drew. Analysis of an optimal sector design method. In 27th Digital Avionics System
Conference, 2008.

[68] A. Basu, J. S. B. Mitchell, and G. Sabhnani. Geometric algorithms for optimal airspace
design and air traffic controller workload balancing. In 10th Workshop on Algorithm
Engineering and Experiments and 5th Workshop on Analytic Algorithmics and Combi-
natorics, 2008.

[69] M. Xue. Airspace sector redesign based on voronoi diagrams. Journal of Aerospace
Computing, 6(12):605–615, 2009.

[70] C. R. Brinton, K. Leiden, and J. Hinkey. Airspace sectorization by dynamic density.


In 9th AIAA Aviation Technology, Integration and Operations Conference, 2009.

[71] J. Li, T. Wang, M. Savai, and I. Hwang. Graph-based algorithm for dynamic airspace
configuration. Journal of Guidance, Control, and Dynamics, 33(4):1082–1094, 2010.

[72] G. R. Sabhnani, A. Yousefi, and J. S. B. Mitchell. Flow conforming operational airspace


sector design. In 10th AIAA Aviation Technology, Integration and Operations Confer-
ence, 2010.

[73] M. Gariel, A. N. Srivastava, and E. Feron. Trajectory clustering and an application


to airspace monitoring. IEEE Transactions on Intelligent Transportation Systems,
12(4):1511–1524, 2011.

[74] G. R. Sabhnani, A. Yousefi, I. Kostitsyna, J. S. B. Mitchell, and V. Polishchuk. Al-


gorithmic traffic abstraction and its application to nextgen generic airspace. In 10th
AIAA Aviation Technology, Integration, and Operations Conference, 2010.

[75] C. A. Provan, L. Cook, and J. Cunningham. A probabilistic airport capacity model for
improved ground delay program planning. In AIAA/IEEE Digital Avionics Systems
Conference, 2011.

[76] D. M. Pfeil and H. Balakrishnan. Identification of robust terminal-area routes in


convective weather. Transportation Science, 46(1):56–73, 2012.

[77] J. J. Rebollo and H. Balakrishnan. Characterization and prediction of air traffic delays.
Transportation Research Part C: Emerging Technologies, 44:231–241, 2014.

[78] L. Li, M. Gariel, R. J. Hansman, and R. Palacios. Anomaly detection in onboard-


recorded flight data using cluster analysis. In 30th Digital Avionics Systems Conference,
2011.

[79] Y. S. Chati. Statistical Modeling of Aircraft Engine Fuel Burn. Ph.D. thesis, Mas-
sachusetts Institute of Technology, 2018.

214
[80] G. Antonini and J. P. Thiran. Counting pedestrians in video sequences using trajec-
tory clustering. IEEE Transactions on Circuits and Systems for Video Technology,
16(8):1008–1020, 2016.

[81] Z. Fu, W. Hu, and T. Tan. Similarity based vehicle trajectory clustering and anomaly
detection. In 12th IEEE International Conference on Image Processing, 2005.

[82] S. J. Gaffney and P. Smyth. Trajectory clustering with mixtures of regression models.
In 5th ACM SIGKDD International Conference on Knowledge Discovery and Data
Mining, 1999.

[83] S. J. Gaffney, A. W. Robertson, P. Smyth, S. J. Camargo, and M. Ghil. Probabilistic


clustering of extratropical cyclones using regression mixture models. Climate Dynam-
ics, 29(4):423–440, 2007.

[84] J. G. Lee, J. Han, and K.Y. Whang. Trajectory clustering: A partition-and-group


framework. In Association for Computing Machinery’s Special Interest Group on Man-
agement of Data Conference, 2007.

[85] J. Kim and H. S. Mahmassani. Spatial and temporal characterization of travel pat-
terns in a traffic network using vehicle trajectories. Transportation Research Part C:
Emerging Technologies, 59:375–390, 2015.

[86] R. Wen, W. Yan, A.N. Zhang, N.Q. Chinh, and O. Akcan. Spatio-temporal route
mining and visualization for busy waterways. In IEEE International Conference on
Systems, Man, and Cybernetics, 2016.

[87] A. Eckstein. Automated flight track taxonomy for measuring benefits from perfor-
mance based navigation. In Integrated Communications, Navigation and Surveillance
Conference, 2009.

[88] F. Rehm. Clustering of flight tracks. In AIAA Infotech@Aerospace Conferece, 2010.

[89] M. Enriquez. Identifying temporally persistent flows in the terminal airspace via spec-
tral clustering. In 10th USA/Europe Air Traffic Management Research and Develop-
ment Seminar, 2013.

[90] H. Arneson, A. Bombelli, A. S. Torne, and E. Tse. Analysis of convective weather


impact on pre-departure routing of fights from fort worth center to new york center.
In 17th AIAA Aviation Technology, Integration, and Operations Conference, 2017.

[91] A. Bombelli, A. S. Torne, E. Trumbauer, and K. D. Mease. Automated route clustering


for air traffic modeling. In AIAA Modeling and Simulation Technologies Conference,
2017.

[92] L. Song, C. Wanke, and D. P. Greenbaum. Predicting sector capacity for tfm decision
support. In 6th AIAA Aviation Technology, Integration, and Operations Conference,
2006.

215
[93] S. Sidiropoulos, K. Han, A. Majumdar, and W. Y. Ochieng. Robust identification
of air traffic flow patterns in metroplex terminal areas under demand uncertainty.
Transportation Research Part C: Emerging Technologies, 75:212–227, 2017.

[94] E. P. Gilbo. Airport capacity: Representation, estimation, optimization. IEEE Trans-


actions on Control Systems Technology, 1(3):144–154, 1993.

[95] G. F. Newell. Airport capacity and delays. Transportation Science, 13(3):201–241,


1979.

[96] M. Ignaccolo. A simulation model for airport capacity and delay analysis. Transporta-
tion Planning and Technology, 26(2):135–170, 2003.

[97] P. B. Liu, M. Hansen, and A. Mukherjee. Scenario-based air traffic flow management:
From theory to practice. Transportation Research Part B: Methodological, 42(7):685–
702, 2008.

[98] G. Buxi and M. Hansen. Generating probabilistic capacity profiles from weather fore-
cast: A design-of-experiment approach. In 9th USA/Europe Air Traffic Management
Research and Development Seminar, 2011.

[99] J. Cox and M. J. Kochenderfery. Probabilistic airport acceptance rate prediction. In


AIAA Modeling and Simulation Technologies Conference, 2016.

[100] J. C. Jones, R. DeLaura, M. Pawlak, S. Troxel, and N. Underhill. Predicting and


quantifying risk in airport capacity profile selection for air traffic management. In 14th
USA/Europe Air Traffic Management Research and Development Seminar, 2017.

[101] J. C. Jones and R. DeLaura. Predicting airport capacity in the presence of winds. In
17th AIAA Aviation Technology, Integration, and Operations Conference, 2017.

[102] S. Houston and D. Murphy. Predicting runway configurations at airports. In Annual


Meeting of Transportation Research Board, 2012.

[103] V. Ramanujam and H. Balakrishnan. Data-driven modeling of the airport configuration


selection process. IEEE Transactions on Human-Machine Systems, 45(4):490–499,
2015.

[104] J. Avery and H. Balakrishnan. Predicting airport runway configuration: A discrete-


choice modeling approach. In 11th USA/Europe Air Traffic Management Research and
Development Seminar, 2015.

[105] J. Avery. Data-driven modeling of the airport runway configuration selection process
using maximum likelihood discrete-choice models. S.M. thesis, Massachusetts Institute
of Technology, 2016.

[106] M. Vlachos, G. Kollios, and D. Gunopulos. Discovering similar multidimensional tra-


jectories. In 18th International Conference on Data Engineering, 2002.

216
[107] H. Zhu, J. Luo, H. Yin, X. Zhou, J. Z. Huang, and F. B. Zhan. Mining trajectory
corridors using fréchet distance and meshing grids. In 14th Pacific-Asia Conference on
Knowledge Discovery and Data Mining, 2010.

[108] A. K. Jain, M. N. Murty, and P. J. Flynn. Data clustering: a review. ACM Computing
Surveys, 31:264–323, 1999.

[109] A. K. Jain. Data clustering: 50 years beyond k-means. Pattern Recognition Letters,
31:651–666, 2010.

[110] M. Ester, H. P. Kriegel, J. Sander, and X. Xu. A density-based algorithm for discover-
ing clusters in large spatial databases with noise. In 2nd International Conference on
Knowledge Discovery and Data Mining, 1996.

[111] O. Arbelaitz, I. Gurrutxaga, J. Muguerza, J. M. Pérez, and I. Perona. An extensive


comparative study of cluster validity indices. Pattern Recognition, 46:243–256, 2013.

[112] P. J. Rousseeuw. Silhouettes: A graphical aid to the interpretation and validation of


cluster analysis. Journal of Computational and Applied Mathematics, 20:53–65, 1987.

[113] E. Alpaydin. Introduction to Machine Learning. MIT Press, 2nd edition, 2010.

[114] L. Breiman. Random forests. Machine Learning, 45:5–32, 2001.

[115] V. Chandola, A. Banerjee, and V. Kumar. Anomaly detection: A survey. ACM


Computing Surveys, 41(3):1–58, 2009.

[116] G. Shaffer and V. Vovk. A tutorial on conformal prediction. Journal of Machine


Learning Research, 9:371–421, 2008.

[117] D. Devetyarov and I. Nouretdinov. Prediction with confidence based on a random


forest classifier. In 6th IFIP WG 12.5 International Conference, 2010.

[118] S. Bhattacharyya. Confidence in predictions from random tree ensembles. Knowledge


Information Systems, 35:391–410, 2013.

[119] U. Johansson, H. Boström, T. Löfström, and H. Linusson. Regression conformal pre-


diction with random forests. Machine Learning, 97:155–176, 2014.

[120] R. Laxhammar and G. Falkman. Conformal prediction for distribution-independent


anomaly detection in streaming vessel data. In 1st International Workshop on Novel
Data Stream Pattern Mining Techniques, ACM, 2010.

[121] Port Authority of New York and New Jersey. Airport Traffic Report. https://www.
panynj.gov/airports/pdf-traffic/ATR2016.pdf, (accessed 2018-01).

[122] Hong Kong International Airport. Air Traffic Statistics. http://www.


hongkongairport.com/eng/pdf/business/statistics/2016e.pdf, (accessed 2018-
01).

217
[123] Civil Aviation Resource Network. Transportation Statistics. http://i.carnoc.com/
detail/385649, (accessed 2018-01).

[124] Macau International Airport. Passenger Figure and Movement Statistics. http://www.
macau-airport.com/en/media-centre/facts-figures/statistics-passengers,
(accessed 2018-01).

[125] GRU Airport. Operational Information. https://www.gru.com.br/en/


institutional/sobre-gru-airport/operational-information, (accessed 2018-
01).

[126] Aeroportos Brasil Viracopos. Statistics and Publications. http://www.viracopos.


com/institucional/estatisticas-e-publicacoes/, (accessed 2018-01).

[127] Infraero Aeroportos. Statistics. http://www4.infraero.gov.br/


acesso-a-informacao/institucional/estatisticas/, (accessed 2018-01).

[128] Airports Council International. Annual Traffic Data. http://www.aci.aero/


Data-Centre/Annual-Traffic-Data, (accessed 2017-11).

[129] Federal Aviation Administration. Faa JO Order 7210.3Z, Facility Operation and Ad-
ministration. http://www.faa.gov/documentLibrary/media/Order/7210.3Z_FAC.
pdf, (accessed 2018-01).

[130] P. Glaab, R. Tamburro, and P. Lee. Analysis of the capacity potential of current day
and novel configurations for new york’s john f. kennedy airport. In 16th AIAA Aviation
Technology, Integration, and Operations Conference, 2016.

[131] Civil Aviation Department of the Hong Kong Special Administrative Region. Air
Traffic Control Facilities. https://www.cad.gov.hk/english/cad_division.html,
(accessed 2018-01).

[132] FlightRadar24. Live Air Traffic. https://www.flightradar24.com/, (accessed 2018-


01).

[133] International Civil Aviation Organization. Doc9854-AN/458: Global Air Traffic Man-
agement Operational Concept. ICAO, 1st edition, 2005.

[134] R. Koenker and K. Hallock. Quantile regression. Journal of Economic Perspectives,


15:143–156, 2001.

[135] Federal Aviation Administration. JFK, EWR and LGA traffic management tips.
https://www.fly.faa.gov, (accessed 2017-06).

[136] Civil Air Navigation Services Organization. Recommended Key Performance Indi-
cators for Measuring ANSP Operational Performance. https://www.canso.org/
recommended-key-performance-indicators-measuring-ansp-operational-performance,
(accessed 2018-01).

218
[137] Federal Aviation Administration. 14 CFR Part 1 - Definitions and Abbreviations.
https://www.ecfr.gov/cgi-bin/retrieveECFR?n=14y1.0.1.1.1, (accessed 2018-
04).

[138] M. Robinson, R. DeLaura, and N. Underhill. The Route Availability Planning Tool
(RAPT): evaluation of departure management decision support in New York during the
2008 convective weather season. In 8th USA/Europe Air Traffic Management Research
and Development Seminar, 2009.

[139] C. M. Bishop. Neural Networks for Pattern Recognition. Oxford University Press, 1st
edition, 1995.

[140] J. Platt. Probabilistic outputs for support vector machines and comparisons to regu-
larized likelihood methods. Advances in large margin classifiers, 10(3):61–74, 1999.

[141] T.-F. Wu, C.-J. Lin, and R. C. Weng. Probability estimates for multi-class classification
by pairwise coupling. Journal of Machine Learning Research, 5:975–1005, 2004.

[142] N. Meinshausen. Quantile regression forests. Journal of Machine Learning Research,


7:983–999, 2006.

[143] C. E. Rasmussen and C. K. I. Williams. Gaussian Processes for Machine Learning.


MIT Press, 1st edition, 2006.

[144] J. Vanhatalo, J. Riihimäki, J. Hartikainen, P. Jylänki, V. Tolvanen, and A. Vehtari.


Gpstuff: Bayesian modeling with gaussian processes. Journal of Machine Learning
Research, 14:1175–1179, 2013.

[145] International Civil Aviation Organization. Doc 9750-AN/963: 2016-2030


global Air Navigation Plan. https://www.icao.int/airnavigation/Documents/
GANP-2016-interactive.pdf, (accessed 2018-04).

[146] J. Ferguson, A. Q. Kara, K. Hoffman, and L. Sherry. Estimating domestic US airline


cost of delay based on European model. Transportation Research Part C: Emerging
Technologies, 33:311–323, 2013.

[147] Eurocontrol. Standard inputs for Eurocontrol cost benefit analyses.


https://www.eurocontrol.int/sites/default/files/publication/files/
standard-input-for-eurocontrol-cost-benefit-analyses-2015.pdf, (accessed
2017-12).

[148] M. P. Matthews, M. S. Veillette, J. C. Venuti, R. A. DeLaura, and J. K. Kuchar.


Heterogeneous convective weather forecast translation into airspace permeability with
prediction intervals. Journal of Air Transportation, 14(2):41–54, 2016.

219

You might also like