Sie sind auf Seite 1von 44

THE NATIONAL ACADEMIES PRESS

This PDF is available at http://nap.edu/24881 SHARE


Dispersion Modeling Guidance for Airports Addressing Local


Air Quality Health Concerns

DETAILS

44 pages | 8.5 x 11 | PAPERBACK


ISBN 978-0-309-44654-9 | DOI 10.17226/24881

CONTRIBUTORS

GET THIS BOOK Saravanan Arunachalam, Alejandro Valencia, Matthew C. Woody, Michelle G. Snyder,
Jiaoyan Huang, Jeffrey Weil, Philip Soucacos, and Sandy Webb; Airport
Cooperative Research Program; Transportation Research Board; National Academies
FIND RELATED TITLES of Sciences, Engineering, and Medicine


Visit the National Academies Press at NAP.edu and login or register to get:

Access to free PDF downloads of thousands of scientic reports


10% off the price of print titles
Email or social media notications of new titles related to your interests
Special offers and discounts

Distribution, posting, or copying of this PDF is strictly prohibited without written permission of the National Academies Press.
(Request Permission) Unless otherwise indicated, all materials in this PDF are copyrighted by the National Academy of Sciences.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

AIRPORT COOPERATIVE RESEARCH PROGRAM

ACRP RESEARCH REPORT 179


Dispersion Modeling Guidance
for Airports Addressing Local
Air Quality Health Concerns

Saravanan Arunachalam
Alejandro Valencia
Matthew C. Woody
Michelle G. Snyder
Jiaoyan Huang
Institute for the Environment
The University of North Carolina at Chapel Hill
Chapel Hill, NC

in association with

Jeffrey Weil
Cooperative Institute for Research in Environmental Sciences
University of Colorado at Boulder
Boulder, CO

Philip Soucacos
Booz allen hamIlton, InC.
Herndon, VA
and

Sandy Webb
The Environmental Consulting Group, LLC
Crownsville, MD

Subscriber Categories
Aviation Environment

Research sponsored by the Federal Aviation Administration

2017

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

AIRPORT COOPERATIVE RESEARCH PROGRAM ACRP RESEARCH REPORT 179

Airports are vital national resources. They serve a key role in trans- Project 02-58
portation of people and goods and in regional, national, and interna- ISSN 2572-3731 (Print)
tional commerce. They are where the nations aviation system connects ISSN 2572-374X (Online)
with other modes of transportation and where federal responsibility for ISBN 978-0-309-44654-9
managing and regulating air traffic operations intersects with the role of Library of Congress Control Number 2017951641
state and local governments that own and operate most airports. Research
2017 National Academy of Sciences. All rights reserved.
is necessary to solve common operating problems, to adapt appropriate
new technologies from other industries, and to introduce innovations into
the airport industry. The Airport Cooperative Research Program (ACRP)
serves as one of the principal means by which the airport industry can COPYRIGHT INFORMATION
develop innovative near-term solutions to meet demands placed on it. Authors herein are responsible for the authenticity of their materials and for obtaining
The need for ACRP was identified in TRB Special Report 272: Airport written permissions from publishers or persons who own the copyright to any previously
Research Needs: Cooperative Solutions in 2003, based on a study spon- published or copyrighted material used herein.
sored by the Federal Aviation Administration (FAA). ACRP carries out Cooperative Research Programs (CRP) grants permission to reproduce material in this
applied research on problems that are shared by airport operating agen- publication for classroom and not-for-profit purposes. Permission is given with the
cies and not being adequately addressed by existing federal research understanding that none of the material will be used to imply TRB, AASHTO, FAA, FHWA,
programs. ACRP is modeled after the successful National Cooperative FMCSA, FRA, FTA, Office of the Assistant Secretary for Research and Technology, PHMSA,
or TDC endorsement of a particular product, method, or practice. It is expected that those
Highway Research Program (NCHRP) and Transit Cooperative Research reproducing the material in this document for educational and not-for-profit uses will give
Program (TCRP). ACRP undertakes research and other technical activi- appropriate acknowledgment of the source of any reprinted or reproduced material. For
ties in various airport subject areas, including design, construction, legal, other uses of the material, request permission from CRP.
maintenance, operations, safety, policy, planning, human resources, and
administration. ACRP provides a forum where airport operators can
cooperatively address common operational problems.
NOTICE
ACRP was authorized in December 2003 as part of the Vision 100
Century of Aviation Reauthorization Act. The primary participants in The research report was reviewed by the technical panel and accepted for publication
according to procedures established and overseen by the Transportation Research Board
the ACRP are (1) an independent governing board, the ACRP Oversight
and approved by the National Academies of Sciences, Engineering, and Medicine.
Committee (AOC), appointed by the Secretary of the U.S. Department of
Transportation with representation from airport operating agencies, other The opinions and conclusions expressed or implied in this report are those of the
researchers who performed the research and are not necessarily those of the Transportation
stakeholders, and relevant industry organizations such as the Airports Research Board; the National Academies of Sciences, Engineering, and Medicine; or the
Council International-North America (ACI-NA), the American Associa- program sponsors.
tion of Airport Executives (AAAE), the National Association of State
The Transportation Research Board; the National Academies of Sciences, Engineering, and
Aviation Officials (NASAO), Airlines for America (A4A), and the Airport Medicine; and the sponsors of the Airport Cooperative Research Program do not endorse
Consultants Council (ACC) as vital links to the airport community; (2) TRB products or manufacturers. Trade or manufacturers names appear herein solely because
as program manager and secretariat for the governing board; and (3) the they are considered essential to the object of the report.
FAA as program sponsor. In October 2005, the FAA executed a contract
with the National Academy of Sciences formally initiating the program.
ACRP benefits from the cooperation and participation of airport
professionals, air carriers, shippers, state and local government officials,
equipment and service suppliers, other airport users, and research organi-
zations. Each of these participants has different interests and responsibili-
ties, and each is an integral part of this cooperative research effort.
Research problem statements for ACRP are solicited periodically but
may be submitted to TRB by anyone at any time. It is the responsibility
of the AOC to formulate the research program by identifying the highest
priority projects and defining funding levels and expected products.
Once selected, each ACRP project is assigned to an expert panel
appointed by TRB. Panels include experienced practitioners and
research specialists; heavy emphasis is placed on including airport
professionals, the intended users of the research products. The panels
prepare project statements (requests for proposals), select contractors,
and provide technical guidance and counsel throughout the life of the Published research reports of the
project. The process for developing research problem statements and
AIRPORT COOPERATIVE RESEARCH PROGRAM
selecting research agencies has been used by TRB in managing coop-
erative research programs since 1962. As in other TRB activities, ACRP are available from
project panels serve voluntarily without compensation. Transportation Research Board
Primary emphasis is placed on disseminating ACRP results to the Business Office
500 Fifth Street, NW
intended users of the research: airport operating agencies, service pro- Washington, DC 20001
viders, and academic institutions. ACRP produces a series of research
reports for use by airport operators, local agencies, the FAA, and other and can be ordered through the Internet by going to
interested parties; industry associations may arrange for workshops, http://www.national-academies.org
training aids, field visits, webinars, and other activities to ensure that and then searching for TRB
results are implemented by airport industry practitioners. Printed in the United States of America

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

The National Academy of Sciences was established in 1863 by an Act of Congress, signed by President Lincoln, as a private, non-
governmental institution to advise the nation on issues related to science and technology. Members are elected by their peers for
outstanding contributions to research. Dr. Marcia McNutt is president.

The National Academy of Engineering was established in 1964 under the charter of the National Academy of Sciences to bring the
practices of engineering to advising the nation. Members are elected by their peers for extraordinary contributions to engineering.
Dr. C. D. Mote, Jr., is president.

The National Academy of Medicine (formerly the Institute of Medicine) was established in 1970 under the charter of the National
Academy of Sciences to advise the nation on medical and health issues. Members are elected by their peers for distinguished contributions
to medicine and health. Dr. Victor J. Dzau is president.

The three Academies work together as the National Academies of Sciences, Engineering, and Medicine to provide independent,
objective analysis and advice to the nation and conduct other activities to solve complex problems and inform public policy decisions.
The National Academies also encourage education and research, recognize outstanding contributions to knowledge, and increase
public understanding in matters of science, engineering, and medicine.

Learn more about the National Academies of Sciences, Engineering, and Medicine at www.national-academies.org.

The Transportation Research Board is one of seven major programs of the National Academies of Sciences, Engineering, and Medicine.
The mission of the Transportation Research Board is to increase the benefits that transportation contributes to society by providing
leadership in transportation innovation and progress through research and information exchange, conducted within a setting that
is objective, interdisciplinary, and multimodal. The Boards varied committees, task forces, and panels annually engage about 7,000
engineers, scientists, and other transportation researchers and practitioners from the public and private sectors and academia, all
of whom contribute their expertise in the public interest. The program is supported by state transportation departments, federal
agencies including the component administrations of the U.S. Department of Transportation, and other organizations and individuals
interested in the development of transportation.

Learn more about the Transportation Research Board at www.TRB.org.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

COOPERATIVE RESEARCH PROGRAMS

CRP STAFF FOR ACRP RESEARCH REPORT 179


Christopher J. Hedges, Director, Cooperative Research Programs
Lori L. Sundstrom, Deputy Director, Cooperative Research Programs
Michael R. Salamone, Manager, Airport Cooperative Research Program
Joseph D. Navarrete, Senior Program Officer
Hana Vagnerova, Senior Program Assistant
Eileen P. Delaney, Director of Publications
Sharon Lamberton, Editor

ACRP PROJECT 02-58 PANEL


Field of Environment
Kristoffer Russell, Dallas/Fort Worth International Airport, DFW Airport, TX (Chair)
Edward L. Carr, ICF, San Francisco, CA
Renee L. Dowlin, Jviation, Portland, OR
Prem Lobo, Center of Excellence for Aerospace Particulate Emissions Reduction Research, Rolla, MO
Randy J. McGill, Greater Toronto Airports Authority, Georgetown, ON, Canada
Barbara Morin, Rhode Island Department of Health, Providence, RI
Mohammed Majeed, FAA Liaison
Christine Gerencher, TRB Liaison

AUTHOR ACKNOWLEDGMENTS
The Atmospheric Dispersion Modeling System at Airports (ADMS-Airport) is a proprietary disper-
sion model for aircraft-related sources developed and maintained by Cambridge Environmental Research
Consultants (CERC). The authors of this report would like to thank CERC for allowing the use of the
model in this comparative study, and for guidance in its application. The authors also gratefully acknowl-
edge the discussions with Christopher DesAutels of Exponent in the use of the CALPUFF model, the
Electric Power Research Institute (EPRI) for providing an advanced version of the SCIPUFF model before
public release, and Biswanath Chowdhury of Sage Management for assistance with troubleshooting and
applying the SCIPUFF model for this study.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

FOREWORD

By Joseph D. Navarrete
Staff Officer
Transportation Research Board

ACRP Research Report 179 provides guidance for selecting and applying dispersion mod-
els to study local air quality health impacts resulting from airport-related emissions. The
report should be of particular interest to airport environmental practitioners and regulators
who wish to learn about the unique challenges associated with modeling emissions in an
airport setting for the purposes of understanding their potential impacts on human health.

The Aviation Environmental Design Tool (AEDT) is the required regulatory emissions
and dispersion model for US airports; it employs the EPAs AERMOD dispersion model,
a Gaussian plume model. AERMOD is typically used to model dispersion from point and
area sources (e.g., power plants, industrial activities) and to assess local air quality impacts.
In recent years, however, airports have been asked to address more complex public health
issues associated with airport activity. These public health studies often call for the use
of high-fidelity, time-varying dispersion models such as CALPUFF, SCIPUFF/SCICHEM,
ADMS-Airport, and LASPORT, some of which, in addition to their higher resolution, provide
additional chemical transformation mechanisms and processes not included in AERMOD.
No established process exists for modeling airport sources with these models, however,
which has led to inconsistent practices. Research was needed to provide guidance for select-
ing and using dispersion models to address local air quality health concerns.
The research team, led by the University of North Carolina at Chapel Hill, began with a
review of recent relevant literature. This was followed by a review of model input require-
ments (aircraft activity, ground support equipment, on-road vehicles, stationary sources,
and weather). The research team then conducted an intercomparison of four dispersion
models, AERMOD, SCICHEM, CALPUFF, and ADMS-Airport, using actual airport air
quality monitoring data. The results of this analysis were used to develop the guidance.
This guidance document includes a primer on airport air quality and dispersion modeling,
a decision matrix to help practitioners select the most appropriate modeling approach based
on research needs, and guidance on how to approach an airport air quality study that requires
dispersion modeling. Recognizing that the state of the practice would benefit from additional
investigation, the guidance document also identifies areas of further research to improve dis-
persion models, with specific focus on the unique characteristics of aircraft sources compared
to other source types. The contractors final report, which provides detail on their research
approach and findings, is available at www.trb.org by searching ACRP Project 02-58.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

CONTENTS

1 Summary
4 Chapter 1 Introduction
6 Chapter 2Primer on Airport Air Quality
and Dispersion Modeling
6 2.1 Overview of Airport Air Quality Modeling
11 2.2 Selecting a Dispersion Model for Airport Air Quality Analysis
13 Chapter 3 Airport Modeling Studies
16 Chapter 4 Models versus Data Inputs
16 4.1 Input Data Requirements
16 4.1.1 AERMOD
16 4.1.2 CALPUFF
17 4.1.3 SCICHEM
17 4.1.4 ADMS-Airport
18 4.2 Modeling Systems
20 Chapter 5 Dispersion Model Intercomparison
25 Chapter 6 Future Research Needs
25 6.1 Incorporation of Background Pollutant Concentrations
26 6.2 Source Characterization
26 6.3 Inventory of UFPs
27 6.4 Plume Rise from Aircraft Emissions
27 6.5 Aircraft Downwash Effects on Plume Rise and Dispersion
28 6.6 Aircraft Dispersion Based on Instantaneous Line Puffs
29 6.7 Effects of Light Winds and Atmospheric Stability on Dispersion
29 6.8 Other Limitations
30 6.9 Interim Guidance
31 Glossary
34 Bibliography and References

Note: Photographs, figures, and tables in this report may have been converted from color to grayscale for printing.
The electronic version of the report (posted on the web at www.trb.org) retains the color versions.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Summary

Dispersion Modeling Guidance


for Airports Addressing Local
Air Quality Health Concerns
The motivation for this research was to develop guidance for analysts in applying disper-
sion models to study local air quality health impacts from airport emissions. (Terms shown in
italics appear with brief definitions in the glossary.) Airport analysts have used the Emissions
and Dispersion Modeling System (EDMS), since replaced by the Aviation Environmental
Design Tool (AEDT), for modeling local air quality. AEDT uses EPAs AERMOD disper-
sion model, a Gaussian plume model in which concentrations are most commonly 1-hour
averages. Typically, AERMOD is used to model dispersion from point and area sources (e.g.,
power plants, industrial activities) and is used to assess local air quality impacts. A range of
other models is available, however, and these models are time-varying and take into account
additional chemical transformation mechanisms, which may expand analysts capabilities
and provide added insights. To date, these models have had limited use in the United States
in studies of airport-related local air quality and health impacts; therefore, proper guidance
on their use for modeling airport-related emissions sources has been lacking.
ACRP Project 02-58 conducted a direct intercomparison of four dispersion models using
a common set of input data for a single airport. This guidance document begins with a
primer on dispersion modeling and then presents the results of the model intercomparison.
It also provides guidance on selecting an appropriate model for future studies of the health
impacts of airport emissions.
Based upon an extensive review of the literature, four models were chosen for the inter-
comparison: AERMOD, SCICHEM, CALPUFF and ADMS-Airport. These models were
chosen based upon several criteria and based upon an extensive literature review. AERMOD
is the de facto model used for regulatory dispersion modeling in the United States. In addi-
tion, AERMOD is coupled with EDMS/AEDT for airport-related modeling in the United
States. The first three models are US-based dispersion models available at no cost, while
the fourth is a UK-based proprietary model that has been used to study airport-related air
quality, mostly in Europe. Comparing the four models, AERMOD and ADMS-Airport have
a limited treatment of chemical transformation (converting NO to NO2, which are likely
important for capturing localized effects), whereas SCICHEM and CALPUFF have addi-
tional processes for gas-phase and aerosol chemistry, with SCICHEM more detailed than
CALPUFF. Also, AERMOD and ADMS-Airport are Gaussian plume models, and SCICHEM
and CALPUFF are puff models that were originally designed to model large elevated sources
such as power plants but nevertheless provide a viable alternate choice given their detailed
treatment of chemical processes.
To support a model intercomparison study, a detailed suite of measurements at a large
airport is ideal. The researchers reviewed the literature and chose the Los Angeles International
Airport (LAX) in Southern California as the candidate airport. In 2012, LAX was one of the

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

2 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

top five airports for commercial air traffic in the world, with more than 700,000 landing
and take-off operations per year. LAX also is situated close to a very large metropolitan area.
During 20112012, Los Angeles World Airports (LAWA) conducted a detailed Air Quality
Source Apportionment Study (LAX AQSAS) in which detailed measurements of more than
400 chemical compounds were made for two 6-week periods during both summer and
winter at 17 locations in and around the airport. Multiple source-based and receptor-based
models were applied to perform source apportionment. Of the 17 data-collection sites, four
sites had detailed measurements of various pollutants on an hourly resolution, and the other
sites had 7-day average measurements.
The research team used the detailed emissions inventories that were created in support of
the LAX AQSAS using EDMS as input data to AERMOD; however, these emissions inven-
tories could not be directly used in the other three models. Therefore, the researchers devel-
oped converters to adapt EDMS-based inputs to the other three models. This guidance
document highlights key issues faced by the project team during this conversion process
and provides suggestions for addressing them in future development of the AEDT model.
A key aspect of the LAX EDMS inventories is that there are more than 5,000 aircraft
sources at LAX, and when combined with the other airport-related and background sources
within the study domain, there are nearly 6,000 sources to be modeled. Although the con-
version from EDMS-AERMOD to SCICHEM and CALPUFF was relatively straightforward,
the conversion of EDMS-based outputs for use in ADMS-Airport was challenging. For
future applications of ADMS-Airport, the project team recommends that airport practitio-
ners start directly from EDMS inputs.
Minor issues also arose because of differences between the version of EDMS used by
LAWA in 20112012 and the more recent version used for this project. Improvements and
bug fixes to the more recent EDMS caused differences in the modeled emissions for some
sources at LAX, compared to the modeled emissions from the LAX AQSAS.
Another key challenge faced during the model intercomparison was missing meteorologi-
cal data for a few specific time periods. While this is usually a non-issue with steady-state
models such as AERMOD, non-steady-state models such as CALPUFF and SCICHEM need
continuous (hourly) valid meteorological data. Thus, airport practitioners should ensure
there are no gaps in input meteorological fields, should they choose one of these alternate
models. To fill the gaps for the few hours of missing data, the researchers used data from
one of the four nearby meteorological stations near LAX.
The general objective of ACRP Project 02-58 was to use equivalent input data in all four
models to keep the inputs consistent so that any differences in output would be the result
of differences in the dispersion models. In some cases, the researchers performed additional
sensitivity simulations in a given model to take advantage of any specific enhancements that
the model offered to provide improved characterization of local-scale air quality at the air-
port. To support the model evaluation and intercomparison, the research team programmed
all models to predict concentrations at the LAWA AQSAS measurement locations. The team
also set the models to predict pollutant concentrations for a uniform Cartesian grid centered
on the airport for a 5 5 km region (with receptors every 500 m), a Polar grid centered on
the airport for a 50 50 km region (with receptors every 5 km), and flagpole receptors aloft
to capture vertical gradients. Based on computational demands of some models, however,
this expanded set of receptors was used only in AERMOD and CALPUFF, and not in the
other two models.
All models were configured to predict seven pollutantsCO, NOx, SO2, VOCs, TOG,
PM2.5 and PM10on an hourly basis for each of the two 6-week periods.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Summary3

The researchers were able to assess the computational demands of each model and con-
cluded that, in its current form, the SCICHEM model has computational demands that may
not be practical. Improvements to source characterization (e.g., reducing the number of
sources or reducing the number of elongated sources) or to simplify the second-order clo-
sure treatment (which provides a direct relationship between the predicted dispersion rates
and the measurable turbulent velocity statistics of the wind field at a high computational
cost) could make SCICHEM more practical for use by airports. All models but one (ADMS-
Airport) can be run on Linux Operating Systems, and thus parallel processing on multiple
CPUs is possible. ADMS-Airport runs only through a Windows-based GUI. However, note
that the Run Manager system available to ADMS-Airport users allows parallel processing
on multiple Windows CPUs and/or PCs. With the exception of AERMOD, the models could
not directly support the nearly 5,000 area sources to represent aircraft activity and needed
to be customized in order to use the models with this many sources.
The model outputs were compared against observations, and against each other, using
an extensive set of graphical and statistical measures of model performance. To provide a
summary assessment of model performance, the research team developed a scoring scheme
using select performance criteria. Based on this scoring scheme, AERMOD and ADMS-
Airport results seemed to match more closely with observations. All models underestimated
the observed PM2.5, which was due to the lack of data on background concentrations in the
local-scale modeling. A need exists for providing information on background concentra-
tions using a regional-scale model like the Community Multi-scale Air Quality (CMAQ)
model, or using geostatistical techniques such as Space-Time Ordinary Kriging (STOK) of
observed concentrations from remote background locations.
The researchers concluded the study by identifying several areas of potential future
research:
1. Incorporation of background concentrations,
2. Representation of aircraft sources at the airport,
3. Inventory of ultra-fine particles,
4. Plume rise from aircraft emissions,
5. Aircraft downwash effects on plume rise and dispersion,
6. Aircraft dispersion based on instantaneous line puffs,
7. Effects of light winds and atmospheric stability on dispersion, and
8. Other limitations such as lack of chemical treatment in some models.
With regard to plume rise from aircraft emissions, in this guide the researchers propose
and discuss three general plume rise models:
1. The existing model with a zero plume rise,
2. An empirical model for plume rise and initial spread based on the LIDAR measurements,
and
3. A fluid-mechanical entrainment model (FEM).

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Chapter 1

Introduction

FAAs Aviation Emissions and Air Quality Handbook (which is available for download on the
FAA website) has three primary objectives:
1. To provide guidance, procedures, and methodologies appropriate for use in carrying out air
quality assessments prepared in association with FAA-supported projects/actions;
2. To help ensure that these air quality assessments meet the requirements of the National
Environmental Policy Act (NEPA), the federal Clean Air Act (CAA), and other relevant laws
and regulations; and
3. To provide a process for users to determine when an air quality assessment is considered
necessary, the type of analysis that is appropriate, and the level of effort that is warranted.
The approach taken by the FAA handbook is broad in scope for air quality assessments,
whereas ACRP Research Report 179 provides detailed information about airport dispersion mod-
eling for airport staff with an interest in or responsibility for the impacts of airport emissions
on air quality in the airport vicinity. More specifically, this guidebook presents information for
selecting and using specific dispersion models to address local air quality health concerns.
The AEDT is FAAs required model for airport air quality analysis. Two key components of
AEDT are the EDMS, which converts airport activity into an emissions inventory, and the Amer-
ican Meteorological Society/Environmental Protection Agency (EPA) Regulatory Model, called
AERMOD, which takes the emissions inventory and computes pollutant concentrations based
on the dispersion of the pollutants from emission sources to receptor locations. AERMOD is a
general-purpose dispersion model developed and maintained by EPA and commonly used for
regulatory purposes. Other dispersion models have been developed, often for specific emission
source types or special applications. Some dispersion models include atmospheric chemistry to
track the changes to pollutant species over time. Some models use different approaches to rep-
resent how pollutants are emitted from different sources and the movement of those pollutants
through the environment. Some models focus on changes within a small area, such as an airport
boundary, while others compute changes over a much larger region. No single approach is best
for all applications. ACRP Research Report 179 describes four dispersion models and provides
guidance for selecting the most appropriate dispersion model for a particular study based on
model capabilities and limitations, data requirements unique to airports, pollutant(s) of con-
cern, resource availability, and output requirements.
The guidance presented in this document is based on a comparison of the performance of four
dispersion models applied under the same conditions at the same airport. A common EDMS
emissions inventory was used to provide the necessary inputs for each dispersion model. The
basis for the model inputs came from a detailed emissions analysis of LAX that was conducted by
the Los Angeles World Airports (LAWA) in 20112012. Table 1 summarizes the models evalu-
ated for ACRP Project 02-58 and highlights important differences among the models.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Introduction5

Table 1. Summary of important aspects of four dispersion models evaluated in this study.

Aircraft
Model Platform Sources Model Species Chemistry Met Inputs Output
Inputs
CO2, H2O, CO,
Windows, Aircraft, GSE, VOC, NO2, NOx, 3 options for AERMET using
AERMOD Point
Linux and stationary Activity-based NO -> NO2 NWS, or user-
(EDMS) SOx, PM10, PM2.5, based
GUI sources conversion dened
air toxics
CO, VOC, NOx, Simplied
SOx, PM10, PM2.5, chemistry for
Windows, CALMET, MMIF Point
CALPUFF User-dened User-dened air toxics secondary
Linux or user-dened based
particle
formation
Windows, CO, VOC, NO2, Detailed gas-
NOx, SOx, O3, MMIF or user- Point
Linux and User-dened phase and
SCICHEM User-dened dened based
GUI PM10, PM2.5, aerosol
air toxics chemistry
Aircraft (full LTO General
cycle), APU, GPU, trac CO, VOC, NO2, NO -> NO2
NOx, SOx, O3, AERMET using
ADMS- Windows GSE, engine information conversion; Point
NWS, or user-
Airport GUI startup, motor or individual PM10, PM2.5, Limited O3 based
dened
trac, other trac air toxics chemistry
(user-dened) movements
Abbreviations: GUI = graphical user interface; GSE = ground support equipment; LTO = landing and take-o ; APU = auxiliary power unit;
GPU = ground power unit; AERMET = a meteorological data preprocessor used in AERMOD; CALMET = a diagnostic meteorological model;
MMIF = Mesoscale Model Interface Program; NWS = National Weather Service. Model species listed: CO = carbon monoxide; CO2 = carbon
dioxide; H2O = water; NOx = nitrogen oxides, including NO (nitric oxide) and NO2 (nitrogen dioxide); O3 = ozone; PM10 = particulate matter
of size 10 microns and below; PM2.5 = particulate matter of size 2.5 microns and below; SOx = sulfur oxides; VOCs = volatile organic
compounds; and air toxics (pollutants deemed hazardous because they cause or may cause serious health eects or adverse
environmental and ecological eects).

Chapter 2 provides the following information:


A primer on dispersion modeling, describing basic concepts of dispersion modeling, the phys-
ical and meteorological effects that cause pollutants to disperse, and the regulatory context
that motivates the need for dispersion modeling;
Summaries of the four dispersion models compared in this study; and
A decision tree for selecting a specific dispersion model for a particular study or application.

Chapter 3 provides summaries of prior published studies that compared different dispersion
models in airport air quality applications. These case studies illustrate some of the benefits,
limitations, and qualitative differences among the dispersion models.
Chapter 4 addresses differences in input data requirements for each model. EDMS and AERMOD
have been developed cooperatively so that EDMS outputs are compatible with AERMOD input
requirements. For the other dispersion models, some amount of conversion or transformation of
the input data was required.
Chapter 5 presents the model intercomparison. This chapter describes how the different mod-
els compared on a range of factors such as high and low pollutant concentrations, model sensi-
tivity, run time, and ease of use.
Chapter 6 describes potential improvements to the models examined. Modeling results could
be improved by modifying how sources are represented, use of meteorological data, incorpora-
tion of atmospheric chemistry, and other factors.
A glossary also is provided. A combined Bibliography and References list follows the glossary.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Chapter 2

Primer on Airport Air Quality


and Dispersion Modeling

This chapter describes how airport emissions disperse, mix with background pollutants and
react and/or transform in the atmosphere. It also describes the regulatory framework important
to understanding the relative significance of different sources and pollutants. It also discusses
the reasons for conducting dispersion modeling, alternative modeling approaches, and when,
where, and why different approaches are appropriate.

2.1 Overview of Airport Air Quality Modeling


Airport air quality modeling is used to determine the impact of airport emissions on people
and the environment. Models are numerical approximations of physical phenomena, and spe-
cific air quality models approximate the physical and chemical processes that occur in the atmo-
sphere. Given that these are approximations, uncertainties exist both in the inputs and outputs
of a model. Accordingly, airport operators should exercise caution in the application and use of
models. However, air quality models have played a significant role in understanding the sources
of air pollution, and in developing alternate emissions scenarios to reduce air pollution.
The air quality modeling process begins with quantifying the mass (e.g., lbs. or kg) of pol-
lutant emissions from all sources at the airport to produce an airport emissions inventory. An
inventory is useful to compare the emissions from these sources to better understand the relative
contributions of each source, and may be an essential element of regulatory reports, planning
studies, or sustainability programs.
Major emissions sources at airports are aircraft engines, auxiliary power units, ground sup-
port equipment, and ground access vehicles. These sources are generally described as mobile
sources (e.g., aircraft and passenger automobiles, which involve discrete vehicles emitting pol-
lutants); area sources (e.g., painting booth or parking garages, in which many sources emit pol-
lutants over a wide area); or point sources (e.g., emergency generators, boilers used to heat water
for terminal heating, and other stationary equipment). Quantifying emissions from these many
sources over a common time period is the first step in creating an inventory.
To understand the actual impact of these emissions, it is necessary to determine the pollutant
concentration (mass per unit volume, which may be measured in ppm or mg/m3) at the point
where exposure takes place.
EPA has set limits on exposure to six common pollutants to protect public health and envi-
ronmental welfare against the effects of outdoor air pollution, and these limits are referred to as
the National Ambient Air Quality Standards (NAAQS). These pollutants, generally referred to
as criteria pollutants, are carbon monoxide (CO), lead (Pb), nitrogen dioxide (NO2), ozone (O3),
particulate matter of a size less than 2.5 microns (PM2.5) and of a size less than 10 microns (PM10),

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Primer on Airport Air Quality and Dispersion Modeling 7

and sulfur dioxide (SO2). It is also important to know the average concentration of criteria pol-
lutants over a specified averaging time period, because that is how US air quality regulations are
defined. The primary limits are health-based standards geared toward protecting people who
are sensitive or at-risk, including asthmatics, children, and elderly people. The secondary limits
are designed to prevent impacts to animals, vegetation, and physical structures, and to prevent
reduced visibility. Table 2 summarizes the limits and averaging times for the six common pollut-
ants. The researchers also refer to EPAs Green Book, which has extensive information (including
tables and maps) on individual areas of the country that are in non-attainment areas for each
criteria pollutant (see https://www.epa.gov/green-book). As of September 2016, there were 130
US airports located within areas designated as being in non-attainment or maintenance of the
NAAQS for one or more criteria pollutants (see https://www.faa.gov/airports/environmental/
vale/ for current and historical attainment data).
Calculating concentrations is very complex and requires computer models to quantify and
track the movement of pollutants from the emission source as they spread out due to weather,
local terrain, and mixing. Atmospheric motion determines the overall speed and direction with
which emissions travel. Atmospheric motion is primarily responsible for the mixing, or disper-
sion, that takes place within the ambient atmosphere, creating a plume of pollution.
An important factor in how emissions disperse has to do with plume dynamics, or the physi-
cal condition of the emissions plume. The temperature of the plume affects its buoyancy. High
temperature emissions will rise once released into the surrounding air as a result of the tempera-
ture difference. The greater the temperature difference, the greater the buoyancy. Another factor
in how emissions disperse is the plume velocity. High velocity emissions lead to shear with the
local wind and to turbulence, which causes entrainment of and mixing with the surrounding air.
Low velocity emissions will have much less shear and lower entrainment. The direction of the
plume is determined by the local wind direction, while mixing is related to small-scale effects
like turbulence. Likewise, terrain characteristics and local building structures can affect local
pollutant concentrations by affecting wind patterns and generating turbulence. Other emission-
specific processes also may have an effect, such as dry and wet deposition. All of these physical
processes affect atmospheric dispersion and lead to a three-dimensional, time-dependent con-
centration distribution of the pollutants.
Adding further complexity, many pollutants undergo chemical reaction and transform in the
atmosphere. Understanding the transformation of pollutants by chemical reaction is essential
for determining the health impacts of emissions. Also, the pollutants released at the airport (i.e.,

Table 2. Pollutants, averaging times, and levels for primary and secondary standards.

Pollutant Primary/Secondary Limit Averaging Time Level


Primary 8 hours 9 ppm
Carbon Monoxide (CO)
Primary 1 hour 35 ppm
Lead (Pb) Primary & Secondary Rolling 3-month average 0.15 g/m3
Primary 1 hour 100 ppb
Nitrogen Dioxide (NO2)
Primary & Secondary Annual 53 ppb
Ozone (O3) Primary & Secondary 8 hours 0.070 ppm
Particulate Matter (PM2.5) Primary Annual 12.0 g/m3
Secondary Annual 15.0 g/m3
Primary & Secondary 24 hours 35.0 g/m3
(PM10) Primary & Secondary 24 hours 150 g/m3
Primary 1 hour 75 ppb
Sulfur Dioxide (SO2)
Secondary 3 hours 0.5 ppm
Source: https://www.epa.gov/criteria-air-pollutants/naaqs-table (accessed December 15, 2016)

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

8 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

source emissions) mix with the pollutants that are already in the atmosphere (i.e., background
emissions) forming a more complex mixture. To truly understand the health and environmental
impacts of airport emissions, it is necessary to understand and track all of these factors.
Air dispersion models are used to track the movement and transformation of pollutants
over time in the atmosphere. They are composed of a sequence of mathematical equations that
require information about the physical setting, emissions sources and their temperature, veloc-
ity, direction and spatial location, background pollutant concentrations, weather conditions,
surface characteristics of the airport, and locations of receptors (specified points at which mea-
surements are taken to gauge the exposure to pollutants of the people and environments at those
locations). Even though air dispersion models are simplified representations of reality, they are very
complex and are based on the current best scientific understanding of the factors that influence the
movement and transformation of pollutants in the atmosphere. The various air dispersion mod-
els differ in the basic dispersion assumptions they make, as well as how they represent emission
sources. Rather than representing an aircraft taking off as continuously rising, for example, an
air dispersion model might represent the aircraft as a series of area sources at increasing height,
which allows for simplified computation. Another model might represent the aircraft as a series
of emissions puffs (non-continuous distributions of concentrations) occurring at increasing
height along the take-off path.
After quantifying emissions from the specified sources over a common time period, the mod-
els compute the impact of local meteorology, or how weather and other atmospheric conditions
cause the emissions to migrate or disperse into the environment. Wind speed and direction,
ambient temperature stratification, and surface heating and cooling are the most significant
meteorological factors that cause emissions to disperse.
Local geographical features can disrupt the effects of meteorology, and these are accounted
for in dispersion models. These features could include airport buildings and other structures
on the airfield, as well as nearby hills, which make up complex terrain. Downwash (the effect
of the turbulent wake in the lee of a building) is a term used to represent the potential effects of
a building on the dispersion of emissions from a source. Downwash is considered for sources
characterized as point, line, or area sources. The height and proximity of a point source to a
structure can be used to determine the significance of downwash.
It bears repeating that to evaluate health-related impacts, it is important to consider both the
pollutants in the source emissions and emissions chemistry; that is, how the pollutants chemi-
cally transform once they have been emitted. Current research indicates that, with regard to
airport emissions, the human health effects of PM and NOx are generally the most significant,
given that high NOx concentrations can lead to high O3 concentrations, and O3 is an important
secondary pollutant that affects human health. At most airports, aircraft are the largest source
of NOx emissions.
Aircraft-generated pollutants generally transform in three different zones:
1. Immediately after exiting the combustor within the engine,
2. Downstream from the engine in the hot exhaust plume, and
3. After emissions have cooled and mixed with the ambient atmosphere.
At the aircraft engine exit, hot combustion gases mix with ambient air to quickly cool the
gas stream. Some gases, like heavy hydrocarbons, can condense under these conditions to form
aerosol particles. In the exhaust plume, as emissions continue to cool, some molecules undergo
chemical reactions and produce other molecules that can also condense into particles.
Similarly, gaseous and particle emissions from cars, trucks and ground vehicles that have
exhaust pipes, catalytic converters or particle traps, and mufflers will transform in the exhaust

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Primer on Airport Air Quality and Dispersion Modeling 9

plume after mixing with the ambient atmosphere. Most of the aviation-related PM that reaches
airport communities comprises particles released during ground operations, landings, and take-
offs. Models vary in the level of detail with which they treat atmospheric chemistry.
Once the air quality models have computed an emissions inventory and evaluated the effects
of meteorology, plume dynamics, and emissions chemistry, they determine the pollutant con-
centration at defined receptor sites, including locations of expected maximum concentration,
locations where employees are present, and locations where the general public is commonly
present. The results show the degree to which airport employees, passengers, citizens living
nearby, and the local community are subject to airport emissions impacts.
EDMS, developed by FAA and required for air quality analyses of aviation sources, is EPAs
preferred airport air quality model. EDMS was recently incorporated into FAAs AEDT, which
integrates noise and emissions models and helps assess their interdependencies. The core com-
ponents of EDMS prepare airport emissions inventories. Dispersion capabilities are added to
process the EDMS inventory and determine pollutant concentrations at specified receptor sites.
The standard dispersion model preferred by EPA for use with EDMS is AERMOD. Other
dispersion models also can be used with EDMS, depending on the needs of the particular study.
The four dispersion models emphasized in this guidebook are:
AERMOD: AERMOD is a steady-state Gaussian plume dispersion model that was developed
and is maintained by EPA. The term steady state means that the local meteorological con-
ditions are not changing with time and approximate the flow field. Gaussian refers to the
shape of the concentration profile within the plume (specifically that it reflects a Gaussian
or normal distribution). This model incorporates air dispersion based on the boundary
layer turbulence structure and scaling concepts and includes treatment of both surface and
elevated sources, as well as both simple and complex terrain. It is non-proprietary and is EPAs
preferred regulatory dispersion model for near field (< 50 km) applications. It predicts the
dispersion of both primary gas and aerosol emissions and includes chemistry for the conver-
sion of NOx to NO2 and decay of SO2, dry and wet deposition, plume buoyancy, and complex
terrain. EDMS accounts for emissions from aircraft, auxiliary power units (APU), ground
support equipment (GSE), and stationary sources, which are dispersed as predicted by AERMOD
to produce pollutant concentrations. The EDMS/AERMOD combination is used for the vast
majority of airport air quality analyses performed in the United States.
CALPUFF: The California Puff (CALPUFF) model is a non-proprietary, non-steady-state
Lagrangian Gaussian puff model maintained and distributed for no cost by Exponent. The
term non-steady state means that the local meteorological conditions can change with time,
and Lagrangian refers to following or tracking a puff or parcel of contaminants in space and
time. EPA has identified CALPUFF as a preferred model for assessing the impacts of long-
range transport of pollutants (greater than 50 km). Long-range transport is usually assessed
when primary pollutants from an elevated source are transported to downwind distances and,
when they chemically react with other pollutants, form secondary pollutants that affect human
health. CALPUFF uses overlapping puffs to represent a continuous plume (see Figure 1).
Along with the dispersion of primary gas-phase and aerosol species, plume dynamics, and wet
and dry deposition, CALPUFF includes particle formation of nitrates and sulfates (from NOx
and SO2, respectively) and anthropogenic secondary organic aerosols. CALPUFF applications
typically assess long-range transport to distances as far as 300 km from large point sources
such as power plants.
SCICHEM: The Second-Order Integrated Puff Model with Chemistry (SCICHEM) is a
non-steady-state Lagrangian puff dispersion model based on the Second-Order Integrated
Puff Model (SCIPUFF). The term second-order refers to the models turbulence closure

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

10 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

scheme and to assumptions/hypotheses used in the solution to the equations governing the
wind, turbulence, and concentration of the puff transport and dispersion. SCICHEM uses
a collection of Gaussian puffs to represent concentration fields. The model is non-proprietary
and is designated by EPA as an alternative dispersion model to be used on a case-by-
case basis for both short- and long-range (> 50 km) regulatory applications. SCICHEM
accounts for puff dynamics, complex terrain, wet and dry deposition, and secondary par-
ticle formation. Plumes are represented in three dimensions by numerous puffs that are
advected and dispersed independently, reflecting the local meteorology. Puff merging and
splitting occurs to reflect the variability inherent in weather. SCICHEM simulates chemi-
cal processes in the gas, aerosol and aqueous phases, with chemical transformation taking
place in the plumes. Traditionally, SCICHEM has been used to model large point sources,
such as power plants.
ADMS-Airport: The Atmospheric Dispersion Modeling System at Airports (ADMS-Airport)
is a Gaussian plume dispersion model for aircraft-related sources developed and maintained
by Cambridge Environmental Research Consultants (CERC). ADMS-Airport is a propri-
etary model, which means users must obtain a license from the developer for its use. ADMS-
Airport accounts for chemical reactions for NO, NO2, and O3, as well as the production of
sulfate particles from SO2. It can accommodate complex terrain and can account for puffs or
plumes, wet and dry deposition, and plume dynamics. It includes emissions sources found
at airports, including aircraft, APU, GSE, on-road mobile sources, and airport stationary
sources and uses algorithms designed specifically to model dispersion from aircraft engines.
The aircraft jet model within ADMS-Airport includes equations for conservation of mass,
momentum, heat, and pollutant species. It computes the effect that movement of the aircraft
engine has on reducing the effective buoyancy of the exhaust. This calculation is particularly
important for evaluating dispersion from the high-momentum, buoyant take-off ground
roll from aircraft. ADMS-Airports ability to model atmospheric chemistry and its aircraft jet
model are important reasons for considering its use. ADMS-Airport has been used to model
air quality at Londons Heathrow Airport as part of the UK Department for Transports Proj-
ect for Sustainable Development of Heathrow and it is one of the models used by the Inter-
national Civil Aviation Organization Committee on Aviation Environmental Protection
(ICAO CAEP).

Where to Obtain Models


The four primary models discussed in ACRP Research Report 179 can be obtained
using links at the following websites:

AERMOD: https://www3.epa.gov/ttn/scram/dispersion_prefrec.htm#aermod
CALPUFF: https://www3.epa.gov/ttn/scram/dispersion_prefrec.htm#calpuff
SCICHEM: https://sourceforge.net/projects/epri-dispersion/files/SCICHEM/
ADMS-Airport: http://www.cerc.co.uk/environmental-software/ADMS-Airport-
model.html

Figure 1 shows a schematic of plume versus puff models. The left panel shows the instanta-
neous plume as it is realistically observed, versus the average plume that is modeled by AERMOD
and ADMS-Airport. The right panel shows how the plume can be modeled as a sequence of
puffs, a method that is especially useful with time-varying winds, as in the case of SCICHEM
and CALPUFF.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Primer on Airport Air Quality and Dispersion Modeling 11

Puffs
Plume

Figure 1. Schematic showing plumes versus puffs.

2.2Selecting a Dispersion Model for


Airport Air Quality Analysis
Selecting the right model to use for a particular study at an airport is important because the
data and manpower requirements can vary widely, affecting the cost of conducting the study.
Questions a researcher may ask when choosing a model are:
Am I creating an emissions inventory to meet a regulatory requirement for determining pol-
lutant concentrations?
Will the requirements for computing pollutant concentrations be driven by regulatory needs
or evaluation of pollutant health impacts?
Does the model have the scientific rigor for the pollutants of interest, and is it accepted for
practice in the scientific/regulatory modeling community?
A decision tree can assist with choosing an air quality model (see Figure 2). In this figure,
although AERMOD and ADMS-Airport are the preferred models for primary pollutants, they
do include treatment of the conversion from NO to NO2, which could be important at localized
scales. CALPUFF has a relatively simpler parameterized scheme for PM2.5, but SCICHEM has a
more complex treatment for other secondary pollutants like O3 and PM2.5.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

12 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Figure 2. Decision tree for selection of an air quality model.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Chapter 3

Airport Modeling Studies

Some of the dispersion models described in this chapter have been used in modeling studies
of local airport air quality, particularly for NOx, NO2 and CO, in addition to the intercompari-
son study for this project. This chapter summarizes key findings from individual studies that
highlight some differences in model performance. More information about these studies can
be found in the contractors final report for ACRP Project 02-58, which can be accessed using
a link on the project webpage. ADMS-Airport, EDMS-AERMOD, and LASPORT (Lagrangian
Simulation of Aerosol-Transport for Airports), which were not specifically used in ACRP Proj-
ect02-58, were all designed to be used for airports (Janicke, Fleuti, and Fuller 2007). The Inter-
national Civil Aviation Organization Committee on Aviation Environmental Protection (ICAO
CAEP) approved all three models for assessing air quality impacts of airport emission sources
(International Civil Aviation Organization 2011). Generally, applications of these three models
are country-specific: ADMS-Airport is the preferred model at London Heathrow Airport and
London Gatwick Airport in the United Kingdom; LASPORT is generally used at airports in
Switzerland and Germany; and AEDT is designated as the required model by the FAA for assessing
air quality impacts of airports within the United States.
With the exception of some analyses at Heathrow Airport (Carruthers et al. 2007), no com-
prehensive model evaluations have used the same airport and sources, meteorology, and moni-
toring data until this ACRP-sponsored model intercomparison at LAX. Most evaluations have
been of individual models at a single airport.
The nature or quality of a model evaluation depends strongly on the concentration averaging
time, given that the effects of turbulence and wind direction variability can fluctuate significantly
across time periods of different lengths. Large turbulence intensities and variability lead to large
concentration fluctuations and statistical variability for short (e.g., 1-hour) averaging periods
in contrast to the smaller fluctuations associated with longer (e.g., annual) averaging periods.
Evaluations for short averaging times are negatively affected by the large variability. Finally, the
airport location, surrounding terrain, and local meteorology and climatology also affect the
quality of a model evaluation. More complex terrain and meteorology lead to more uncertainty
in the meteorological inputs to models and a larger variance between model predictions and
observations.
In a model intercomparison study at Heathrow Airport that included ADMS-Airport, EDMS
and LASPORT, ADMS-Airports annual average NOx predictions were closest to actual mea-
surements (Carruthers et al. 2007). The good performance for ADMS-Airport was attributed to
the models more robust treatment of plume buoyancy and dispersion for jet aircraft exhaust
plumes compared to that used by the other models. Also, the annual mean NOx concentra-
tions for these three models varied by a factor of 2.1 despite their use of the same emissions and
meteorological data. These results were based on model comparisons with observations at nine

13

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

14 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

NOx monitoring stations around Heathrow over a year. In addition, modeled airport NOx con-
centration contributions were approximately 1.7 times too high at a single monitoring station
north of the northern runway. It is unclear how the apportionment analysis was conducted and
if it applied to the annual average concentrations. In a more recent study, EDMS-AERMOD
performed equally as well as ADMS-Airport in matching the observed annual-averaged NOx and
NO2 concentrations around Heathrow Airport (Sabatino et al. 2011).
Other studies have applied ADMS-Airport further at Heathrow Airport. One study examined
ADMS-Airport predictions of NOx, NO2, and PM10 around Heathrow and found good model
performance for each of the three pollutants (Carruthers et al. 2008). Another study found that
ADMS-Airport accurately represented the overall pattern of NOx concentrations around Heath-
row with the treatment of aircraft exhaust as buoyant jets (as opposed to volume sources), which
resulted in significantly different results at receptors nearest the airport (Carruthers et al. 2011).
At low wind speeds, however, predicted concentrations were too high, suggesting that the model
overestimated the impact of nearby passive sources, such as those along the airport perimeter road.
Such overpredictions also may have been caused by underestimates of turbulence fluctuations and
wind variability under low wind conditions. In a fourth study using ADMS-Airport over several
days at a busy regional airport, predicted NOx concentrations for complete aircraft traffic including
aircraft trajectories were found to be satisfactory (Sarrat, Aubry, and Chaboud 2012).
Prior to the development of ADMS-Airport, ADMS-Urban was applied at the Manchester
and Heathrow airports. One study found that aircraft flying at altitudes between 2001,000 m
had minimum impact on ground-level concentrations (Peace et al. 2006). This finding high-
lighted the importance of investigating the total contribution from many distributed sources to
local air quality at an airport versus considering the airport as just one source. In the Heathrow
Airport study, ADMS-Urban tended to overestimate the concentrations of NOx at the one moni-
tor site considered (Farias and ApSimon 2006).
Another study using LASPORT reported airport-attributable NO2 concentrations were typi-
cally below 1 g/m3 at locations 3 km or more away from the airport, and that major highways
dominated the regional air pollution (Fleuti and Hofmann 2005). However, comparisons of mea-
sured and modeled NO2 concentrations were mixed with model predictions at some locations
that were well correlated with measurements, while others were not. For example, LASPORT
underpredicted NOx concentrations at monitors near roadways, which were dominated by road
traffic, but overestimated NOx concentrations near runways, which were dominated by aircraft
activities. The overestimation may have been due to insufficient plume rise or issues with low
winds and wind and turbulence variability. The study by Fleuti and Hofmann suggested that the
emission factors for aircraft in actual operation were lower than the results of ICAO certification
tests, a result that other studies have corroborated. Because it is a proprietary model, however,
LASPORT was not used in the ACRP Project 02-58 study.
Other work with EDMS-AERMOD has found a range of performance results, usually for
short-term (1-hour) averaged concentrations. For example, one study examined lead concen-
trations from aircraft piston engines at a general aviation airport near Santa Monica, California
(Carr et al. 2011). This study showed that the model-to-monitor performance at two sites was
good to excellent (within a factor of 2), particularly on 4 of the 6 modeled days.
EDMS-AERMOD also was used for the LAX AQSAS. In an extensive evaluation of AERMOD
using four measurement sites around LAX, model performance was reported as generally fair
to poor (Tetra Tech 2013). In particular, AERMOD showed that greater than 50% of modeled
values of NOx differed from observations by at least a factor of 2. This poorer performance rela-
tive to the results at Heathrow was likely attributed to: (1) the shorter averaging time used at LAX
(1 hour); (2) the more complex terrain at this coastline site, with generally more complicated

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Airport Modeling Studies 15

meteorology including land- and sea-breezes; and (3) a greater frequency of unstable conditions,
with light and variable winds that would lead to higher concentration variability. Moreover, the
neglect of plume rise for the aircraft sources probably led to high overpredictions of concentra-
tions (> factor of 2).
In a 3-day study comprising only 18 hours, a comparison of EDMS-AERMOD predictions
of 1-hour. CO concentrations with observations around Washington Dulles International Air-
port showed that the model frequently captured the hourly trend in the data but, overall, it
underpredicted the measurements (Martin 2006). There was evidence that mobile sources (i.e.,
automobiles) were the largest CO contributor, but low measurements reported for many hours
suggested either an underestimate of airport traffic emissions or possibly background and traffic
sources were not included in the study.
Another study found that EDMS-AERMOD performed quite well for annual average pre-
dictions around Heathrow Airport (Sabatino et al. 2011). The EDMS-AERMOD results were
determined to be within 20% for NOx and NO2 concentrations. Trends of hourly EDMS NOx
predictions at Budapest Ferenc Liszt International Airport (Ferihegy Airport) also agreed well
with these measurements (Steib et al. 2008). However, CO predictions in the 2008 Ferihegy
Airport study were mixed and generally underestimated the peak CO observations, which was
possibly due to CO transport from the nearby Budapest urban area, not included in the emis-
sions data.
Recent work reported in ACRP Report 135: Understanding Airport Air Quality and Public
Health Studies Related to Airports found that PM2.5 dominated the overall health risk posed by
airport emissions (Kim et al. 2012). Especially for aviation-attributable PM, considerations of
chemistry could have significant implications on both the total PM mass and composition. The
three models evaluated in these studiesLASPORT, EDMS, and ADMS-Airportare limited in
their consideration of atmospheric chemistry, as well as their treatment of fine particulate matter
(PM2.5) by either including only primary PM2.5 (LASPORT, EDMS) or by including only some
secondary PM2.5 formation pathways (ADMS-Airport).
Recent efforts to quantify secondary PM formation from aircraft found that secondary organic
aerosols (SOA) make up a significant amount of aircraft PM after a few hours of chemical pro-
cessing (Miracolo et al. 2011; Woody et al. 2015). Furthermore, an additional study found that
aircraft-attributable PM (which already accounted for secondary inorganic PM) was enhanced
by up to 10% near the airport and 20% downwind (Riley et al. 2016).
Recent studies, including one by Peters et al. (2016), have measured ultra-fine PM emissions
on and near airports; however, these data have not been incorporated into or evaluated using
airport air quality dispersion models. Findings from these studies indicate that aircraft produce
particlespredominantly in the 1020 nm size rangethat are smaller than the particles pro-
duced by other sources commonly found at airports. On-airport measurements found peak
concentrations of these particles under arriving aircraft and behind aircraft taking off. Peak
concentrations also were found off-airport downwind of the runways. A clear relationship exists
between LTO operations, wind direction, and distance to the airport and the ultra-fine particle
(UFP) concentration that is observed at monitoring sites around the airport. The contribution
decreases with increasing distances, but effects were measurable at a distance of 7 km from the
airport.
In other recent studies, a 4- to 5-fold increase in particle number concentrations (PNC) was
observed 810 km downwind of LAX, and a doubling of PNC was observed at a site 4 km down-
wind of Bostons Logan International Airport (Hudda et al. 2014; Hudda and Fruin 2016). In
contrast, airport activity does not contribute more to black carbon, NOx and PM10 concentra-
tions at monitoring sites than does traffic at nearby roadways.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Chapter 4

Models versus Data Inputs

4.1 Input Data Requirements


On May 29, 2015, FAA revised its policy on air quality modeling procedures to identify the
Aviation Environmental Design Tool (AEDT) Version 2b as the required model to perform air
quality analyses for aviation, replacing EDMS. At the time ACRP Project 02-58 was being com-
pleted, AEDT 2b had not been publicly released; therefore, after coordination with the ACRP
project panel, the researchers decided to use the latest version of EDMS (5.1.4.1) to generate the
airport emissions inventory.
AEDT uses the same algorithms for estimating emissions as EDMS, and has the same cou-
pling of dispersion to the AERMOD model. AEDT 2b implements the key functionalities from
Integrated Noise Model (INM), EDMS, and AEDT 2a applications. Although mostly legacy
functionalities were incorporated, differences in the models resulted in new overall capabilities,
such as emissions dispersion from curved flight tracks (resulting from INM legacy capability)
and noise using real weather (resulting from EDMS legacy capability). The ability to estimate
emissions from curved flight tracks and then disperse them is a significant improvement over
EDMS which assumes straight in and straight out tracks.
Our general objective was to use equivalent input data in all four models to keep the inputs
consistent so that any differences in output are the result of differences in the dispersion models.
However, the researchers identified a dispersion modeling limitation where some of the disper-
sion models (e.g., CALPUFF and SCIPUFF) are able to only model area sources with four edges.
In the EDMS study used for the LAWA report, some EDMS sources such as parking structures
and gates were modeled with more than four edges. To ensure consistency among all dispersion
software models, the researchers converted the parking structures to have four edges.
AEDT also has the ability to estimate full-flight emissions of multiple pollutants for global
aircraft activity, as described in Wilkerson et al. (2010). Ongoing enhancements include the abil-
ity to estimate both non-volatile PM mass and number, and will be available in a future release.

4.1.1AERMOD
AERMOD is included in EDMS/AEDT; hence, dispersion modeling for local-scale air quality
and health using AERMOD can be made straight from the models and is relatively straight-
forward, as long as the focus is solely on primary pollutants. For AERMOD to be used with one of
the other models (e.g., CALPUFF, SCICHEM or ADMS-Airport), additional work needs to be done.

4.1.2CALPUFF
CALPUFF uses the same area-source treatment for aircraft sources as AERMOD, and it retains
the same aspect ratio. To elaborate, EDMS represents most aircraft sources during landing and

16

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Models versus Data Inputs 17

take-off with a dimension of 200 m 20 m, thus yielding an aspect ratio of 10.0, which is retained
in both AERMOD and CALPUFF modeling. Ideally, all area sources in CALPUFF should have
an aspect ratio of less than 2.0. Sources with an aspect ratio greater than 2.0 will lead to some
performance problems in the near field (at distances of 15 km), but not in the far field (at dis-
tances greater than 5 km). In the research for ACRP Project 02-58, non-aircraft sources that are
usually estimated on an annual basis with temporal profiles were converted to hourly emissions
to be compatible with the CALPUFF model.

4.1.3SCICHEM
The version of SCICHEM used in this study also models aircraft sources as area sources.
However, SCICHEMs algorithm for modeling area sources involves splitting the area sources
into smaller point sources, and then modeling them as individual Gaussian points. Thus, area
sources that were elongated in one direction [or with aspect ratio (length/width) > 1.0] were split
into individual point/stack sources. Even though AERMOD and CALPUFF maintain the same
dimension when modeling aircraft sources, SCICHEM splits the source into multiple smaller
Gaussian points. By following this procedure for the LAX study, the researchers found that the
total number of sources increases by about a factor of 10. This is the key reason behind the exces-
sive run times associated when using SCICHEM to model aircraft emissions as area sources.
Like CALPUFF, SCICHEM does not support hour-of-day or static emission sources; hence,
pollutant-specific hourly emissions for a single day in each season were generated for these
sources and combined with the AERMODS hourly emissions files (labeled .hre files).

4.1.4ADMS-Airport
ADMS-Airport allows two methods for inputting aircraft emissions data to compute dispersion.
1. ADMS-VOL: The Volume option uses volume sources from aircraft sources. EDMS/AERMOD
outputs the aircraft airborne emissions in area sources 20 m apart (vertically). For ACRP
Project 02-58, these area sources were converted to volume sources in ADMS-Airport with a
height of 20 m, thus making a very straightforward conversion process.
2. ADMS-AIR: The Air File option uses the models performance and chemistry capabilities
to model aircraft sources as jets. This method uses an aircraft performance engine that
uses specific aircraft and track positions to disperse the emissions. Unfortunately, specific
aircraft and track position data are not available in the EDMS outputs, because EDMS out-
puts aircraft emissions in area sources without any information on which aircraft contributed
to these emissions. Therefore, the research team made several approximations when using
the Air File (ADMS-AIR) option. For each hour, EDMS provides the area sources that had
emissions (i.e., an aircraft passed through them) and the sources that did not (i.e., no aircraft
passed through them). The active area sources were used to determine an average hourly air-
craft track for a specific runway, and the researchers used that track for all aircraft in ADMS-
Airport. The complexity of this ADMS-Airport modeling method required a reduction in
the different aircraft types used; the researchers therefore mapped all the EDMS aircraft to
11 ADMS aircraft that could be used in the ADMS-Airport modeling. The emissions along
each hourly track were distributed to the aircraft based on the number of flights by each air-
craft and the aircraft emission indexes (EIs).
The research team encountered no problems converting non-aircraft sources from EDMS
to ADMS-Airport. EDMS gate and parking sources are already area sources and were eas-
ily converted to volume sources. EDMS roadways are line sources with an EDMS-specified
dispersion width, which was used to construct the volume sources. In EDMS, stationary
sources can be area, point, or volume. Point sources in EDMS are specified with one set of
coordinates (for the point); the coordinates are used, along with the source diameter, and

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

18 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

height (provided in EDMS), to construct volume sources. However, ADMS-Airport has an


option to model roadways as road sources that include the treatment of street canyons, tunnels,
and traffic-induced turbulence. ADMS-Airport also has a treatment for point sources that incor-
porates buoyancy and momentum, and effects of nearby buildings. The ACRP Project 02-58
researchers did not use either of these options in this study. Rather, the team converted
non-aircraft sources that were estimated on an annual basis with temporal profiles to hourly
emissions to be compatible with the ADMS-Airport model.
This is the first attempt to adapt the EDMS emissions for ADMS-Airport dispersion in the
United States. The ADMS-Airport inputs are not directly compatible with EDMS, and the
researchers had to make several assumptions and create conversion algorithms. EDMS and
AEDT both utilize AERMOD and have the same dispersion output format. However, AEDT
has additional output persist options, allowing for more details in its performance output
files (e.g., results by flight rather than aggregate emissions results). In addition, AEDT has
the ability to specify a detailed track for each aircraft in AEDT (including curved tracks)
using aircraft radar data or other similar information. These AEDT features can be used
in future studies to minimize the conversion assumptions, which in turn will increase the
fidelity in creating ADMS-Airport inputs using the Air File (ADMS-AIR) option.

4.2 Modeling Systems


AERMOD, CALPUFF, and SCICHEM models can be run on the Windows or Linux operating
platforms. Commercial versions of AERMOD are available with GUI front-ends on Windows
for ease of use, but the researchers compiled and installed all models on a Linux server. Doing
this enabled the research team to run multiple instances for the two seasons and for various pol-
lutants and receptor combinations. ADMS-Airport only runs through a Windows-based GUI,
so for this model the researchers were limited to that environment.
The LAX AQSAS provided access to an unprecedented dataset of ambient measurements at a
large airport, but the data are only useful for validating the models at discrete locations. To develop
dispersion model applications for local-scale health assessments, the researchers recommend that a
gridded set of receptors be created in and around the airport, similar to what the researchers did for
this study. This approach requires a Cartesian grid with receptors every 500 meters going up to 5 km
from the airport, and a Polar grid with receptors every 5 km extending to 50 km from the airport.
For airports near densely populated urban areas, the number of receptors may be increased (e.g.,
spacing one receptor every 100 meters), and near roadways increased even more (e.g., spacing
one receptor every 10 meters) to capture the spatial gradients of traffic-related air pollutants in
the near-road environment. The model then predicts concentrations at all these receptors at the
ground level. This level of detail will be helpful to understand the spatial field of concentrations
from the airport that will be used to study health impacts.
To obtain seasonality, it is suggested that a future modeling study look at capturing estimates
from at least two different seasons (e.g., summer versus winter) to examine the effects of meteo-
rological conditions on the model predictions.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Models versus Data Inputs 19

Table 3. CPU-hours used by the four models.

LAWA Receptors, 2m Heights, 7 Pollutants*


Winter 2012 Summer 2012
# CPUs Airc Airp Bg All # CPUs Airc Airp Bg All
AERMOD 7 n/a n/a n/a 3.4 7 n/a n/a n/a 3.4
CALPUFF 1 1.4 1.7 0.3 2.0 1 1.0 1.1 0.2 1.3
SCICHEM 7 1167.9 1742.1 1439.1 3181.1 7 1652.0 2084.1 1459.6 3543.7
ADMS-AIR 7 27.3 27.7 0.5 28.3 7 22.0 22.4 0.5 22.9
ADMS-VOL 7 1.6 2.0 0.5 2.6 7 1.2 1.6 0.5 2.1
# CPUs = The number of CPUs used when running the model; Airc = aircraft sources; Airp = airport sources, including Airc; Bg = background
sources (sources outside the airport but within the study region); All = the sum of Airp+Bg. AERMOD was run for all sources in one
execution; for this model no data are broken out under Airc, Airp, and Bg. All models except ADMS-Airport (ADMS-AIR and ADMS-VOL)
were run on Redhat Linux OS with Intel5866 x86_64 processor with 48 GB memory; ADMS-AIR and ADMS-VOL were run on Windows 7
OS using the same processor.
* Particulate matter (PM) counted separately for PM2.5 and PM10.

Table 3 presents computational times in CPU-hours taken by the four modeling systems to
perform the modeling for the LAX AQSAS for the winter and summer seasons.
As seen in Table 3, SCICHEM modeling run times are prohibitively long. Unless additional
work is done to improve source characterization, SCICHEM is not likely to be a viable option to
perform local air quality modeling of airport sources.
Another key distinction between steady-state models (such as AERMOD and ADMS-Airport)
and non-steady-state models (such as CALPUFF and SCICHEM) is the generation of input
meteorological data. When processing input meteorological data from National Weather Service
(NWS) sites for AERMOD, some data gaps are normal and do not create an issue for steady-state
models, aside from missing concentration fields. In the case of non-steady-state models, how-
ever, missing hours of meteorological data need to be addressed before performing the model
simulations. For this study the research team created a complete dataset by filling missing hours
of data with observations from nearby NWS sites and by computing certain meteorological
variables from first principles. (A description of the procedures used appears in the contractors
final report, which can be obtained using a link from the ACRP Project 02-58 webpage.)

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Chapter 5

Dispersion Model Intercomparison

The general objective of the intercomparison was to use equivalent input data in all four
models, keeping the inputs consistent so that any differences in output would be the result of dif-
ferences in the dispersion models. Of the four models, only AERMOD could directly support the
nearly 5,000 area sources to represent aircraft activity; the other models needed to be custom-
ized for this application. For example, CALPUFF had to be modified to increase the number of
area-source puffs from 200 to 6,000, and subsequently the total number of puffs from 100,000 to
500,000. Similarly, SCICHEM hit CPU-memory limits with the number of puffs that were gen-
erated with approximately 5,000 area sources. To compensate for this, the aircraft sources were
split into 4 subsets, and thus required four separate executions each time. Although AERMOD
and CALPUFF continue to keep the same dimension for the area source during their modeling
of aircraft sources, SCICHEM splits this source into 10 smaller Gaussian points. By doing this for
the LAX study, the researchers found that the total number of sources increases by about a factor
of 10. This increase is the key reason behind the excessive run times associated with SCICHEM
modeling of aircraft emissions as area sources. There were 4,179 aircraft sources whose aspect
ratio was 10.0; hence, SCICHEM split each of the 4,179 sources into 10 Gaussian point sources,
resulting in a total of 41,790 sources. Thus, what started as 6,170 sources in EDMS/AERMOD
were modeled as 52,035 sources in SCICHEM.
To support the model evaluation and intercomparison, the researchers configured all the
models to predict concentrations at the LAWA AQSAS measurement locations. The research
team also set the AERMOD and CALPUFF models to predict pollutant concentrations for a uni-
form Cartesian grid centered on the airport for a 5 5 km region (with receptors every 500 m), a
Polar grid centered on the airport for a 50 50 km region (with receptors every 5 km), and flag-
pole receptors aloft to capture vertical gradients. Based upon computational demands of some
models, however, this expanded set of receptors was used only with AERMOD and CALPUFF,
and not with the other two models. All models were configured to predict seven pollutants: CO,
NOx, SO2, VOCs, TOG, PM2.5 and PM10 on an hourly basis for each of the two 6-week peri-
ods. AERMOD, SCICHEM, and CALPUFF were configured with most of their default options;
ADMS-Airport was used with the Air File (ADMS-AIR) input option, and all results have been
presented for this combination of four models, unless otherwise stated.
In the cases of CALPUFF, SCICHEM, and ADMS-Airport, the researchers performed addi-
tional sensitivity simulations to take advantage of specific enhancements that each model offered
to provide improved characterization of local-scale air quality at the airport. The simulation
scenarios were:
1. CALPUFF Sensitivity #1 with the Slug option for aircraft sources,
2. CALPUFF Sensitivity #2 with CALMET-based meteorological inputs,
3. SCICHEM Sensitivity #1 with reduced number of sources,

20

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Dispersion Model Intercomparison 21

4. SCICHEM Sensitivity #2 with full chemistry, and


5. ADMS-Airport Sensitivity #1 with Volume source option for aircraft sources (ADMS-VOL).
The model outputs were compared against observations at the 17 LAWA sites, as well as
against each other, using an extensive set of graphical and statistical measures of model perfor-
mance (including mean bias, mean error, fractional bias, factor of 2, normalized mean square
error [NMSE] and geometric mean bias).
The next sections in this chapter provide a brief summary of the main results, followed by
a discussion of their meaning with respect to problems with the dispersion models. Figure 3
presents a summary of model performance for NOx at the four LAWA core sites compared to
observations, as a function of wind speed and wind direction. All models overpredict observed
concentrations at CE, the site located east of the South runway and the site with the highest
observed concentrations of NOx. The overprediction in CALPUFF was the highest, with AERMOD
closest to the observed concentrations. Similar comparisons are presented for PM2.5 in Figure 4.

Figure 3. Comparisons among models (model-to-model) and with observed NOX


concentrations for summer 2012.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

22 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Figure 4. Comparisons among models and with observed PM2.5 concentrations for
summer 2012.

All models underpredict observed PM2.5 at all four sites, which highlights the lack of background
concentrations in the local-scale modeling, thus pointing to the need to incorporate background
concentrations using another regional-scale model like the CMAQ model, or to use geostatisti-
cal techniques such as Space-Time Ordinary Kriging (STOK) of observed concentrations from
remote background locations.
The performance results show that the highest observed and predicted concentrations of
NOx usually occur at night, typically after 8:00 p.m. (stable conditions) with the minimum
during daytime, from 10:00 a.m. to 6:00 p.m. Additional maxima cluster around night-to-day
and day-to-night transition periods (i.e., from 6:008:00 a.m. and from 7:008:00 p.m.). With
respect to atmospheric stability, as defined by the inverse Monin-Obukhov length (1/L), the
highest concentrations usually occur under near-neutral conditions (1/L ~ 0), which happen
at or near the transition periods, an observation that is consistent with the time-of-day results.
High concentrations occur for both unstable conditions (daytime, 1/L < 0) and stable condi-
tions (1/L > 0); that is, they exist on both sides of the neutral stability limit. Furthermore, the

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Dispersion Model Intercomparison 23

Table 4. Model performance


objective scores, based upon observed
and modeled distributions of NOX,
CO, and SOX.

High Low Good


AERMOD 3 6 15
CALPUFF 12 3 9
SCICHEM 3.5 3 17.5
ADMS-Airport 7 4 13

highest concentrations occur under light winds (13 m/s typically) for both stable and unstable
conditions. Based upon the specific pollutant that is being considered, and how it is evaluated
according to the NAAQS (i.e., whether the focus is on a short-term maximum concentration or
on a longer-term average concentration), these patterns with model performance have implica-
tions for airport-related local air quality health impacts.

The models generally overestimate the highest concentrations, as shown by quantile-quantile


(Q-Q) and box-and-whisker plots (shown in the contractors final report, available online),
and the largest statistical variability or highest concentration fluctuations are associated with
the maximum concentrations. This pattern points to problems of plume transport and disper-
sion under light wind and stable atmospheric conditions, which have been found previously for
AERMOD and other models (Cirillo and Poli 1992; Sharan and Yadav 1998). Similar light wind
problems occur under convective conditions with these models (Weil, Corio, and Brower 1997).

Given the number of pollutants that were modeled at different sites, during two different seasons,
the researchers developed a simple objective scoring scheme to group all model results into one
of three bins: Good (modeled mean between the 25th and 75th percentile of observed means),
High (modeled mean above the 75th percentile of observed means) and Low (modeled mean
less than the 25th percentile of observed means). The resulting predictions for NOx, CO and SOX
were combined. One point was awarded for each combination of model/site/pollutant if the result
fell in the appropriate category of Good, High, or Low. Half a point was awarded if the model result
fell in the borderline region between any two categories. The researchers excluded PM2.5 from this
analysis because all models underestimated PM2.5 levels.

The resulting scores appear in Table 4 and Table 5 in relation to two specific metrics of model
performance. To compute the scores in Table 4, the researchers looked at the ranges of hourly
concentrations predicted by each model and compared them to the observed ranges. Each model
was eligible to score a maximum of 24 points (6 pollutants at the 4 LAWA core sites). Ideally, the
model providing results that most closely matched observed conditions would have 24 points
under the Good category.

Table 5. Model performance


objective scores, based upon NMSE
vs. FB for NOX, CO and SOX.

High Low Good


AERMOD 1 5
CALPUFF 3 3
SCICHEM 2 4
ADMS-Airport 1 5

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

24 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Based upon this scoring scheme, in the Good category SCICHEMwith a score of 17.5
seems to slightly outperform AERMOD with a score of 15 and ADMS-Airport with a score of 13.
CALPUFF has 12 points under the High category, with only 9 under the Good category.
Table 5 shows a scatter of normalized mean square error (NMSE) versus fractional bias (FB) in
making the same determination. Each model was eligible to score a total of 6 points. Based upon
this metric, both AERMOD and ADMS-Airporteach with a score of 5 in the Good category
seem to slightly outperform the other two models. Using this methodology, CALPUFF scored 3
in both the High and Good categories.
The following conclusions were drawn from the model intercomparisons for the LAX AQSAS:
Modeling systems and input datasets
EDMS-based emissions inventories are not directly usable by dispersion models other than
AERMOD, and significant efforts need to be devoted to this task.
AERMET-based meteorology for non-steady-state models such as CALPUFF and SCIPUFF
needs to be reviewed carefully to ensure continuous hours with valid meteorological data.
The EDMS approach to modeling area sources does not translate well to SCICHEM, caus-
ing significantly longer runtimes, for approximately 5,000 sources. Additional work is
needed to aggregate sources.
Key air pollutants of interest from a health risk point of view are fine particulate matter (i.e.,
PM2.5), followed by O3, and then air toxics to a relatively smaller degree.
The researchers focused the model intercomparison on NOx, PM2.5, CO, and SO2 as key pol-
lutants of interest. Given that O3 is a secondary pollutant, not all of the four chosen models
could predict it. Nevertheless, O3 is not formed appreciably in the immediate vicinity of the
airport, due to rapid titration by high levels of NOx by aircraft.
Model predictions
Models tend to overpredict summer NOx but underpredict winter NOx levels. NOx measured
during winter (55 - 100 g/m3) was higher than NOx measured during summer (7.5 - 35 g/m3).
The AERMOD- and SCICHEM-predicted means are closer to observations, whereas
ADMS-Airport and CALPUFF tend to overpredict.
Compared to the other models, CALPUFF predicts the highest contributions from aircraft
sources.
AERMOD- and SCICHEM-predicted distributions are closer to observed than CALPUFF
and ADMS-Airport.
PM2.5 is a criteria pollutant of special interest in relation to airports local air quality and
health impacts. Predictions of PM2.5 are poor across-the-board, pointing to lack of back-
ground concentrations, and hence secondary components of PM2.5.
A need exists to incorporate regional background concentrations using hybrid techniques.
Future research could focus on local-scale models that can incorporate this process with-
out substantially affecting model runtimes.
Maximum concentrations are overpredicted by AERMOD, possibly because of missing plume
rise, but means are reasonably predicted. This result shows the potential for AERMOD and
other models to be conservative in application to short-term maximum concentrations (such
as the 1-hour form of the NO2 NAAQS, which requires computation of the 98th percentile of
the 1-hour daily maximum averaged over 3 years), but reasonable for predicting long-term
concentrations (such as the annual average form of the NO2 NAAQS).
Both CALPUFF and ADMS-Airport show much larger contributions from non-aircraft sources,
highlighting potential differences in treatment of aircraft sources.
UFP was not modeled, because the EDMS inventories did not support them. Given evolving
literature on airport-based UFP studies, EDMS or AEDT could be enhanced to generate emis-
sions of UFP for activities at airports.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Chapter 6

Future Research Needs

Our model evaluation and other analyses have identified several areas of improvement for
dispersion models applied to aircraft and airport sources. They include:

1. Incorporation of background concentrations,


2. Representation of aircraft sources at the airport,
3. Inventory of ultra-fine particles (UFPs),
4. Plume rise from aircraft emissions,
5. Aircraft downwash effects on plume rise and dispersion,
6. Aircraft dispersion based on instantaneous line puffs,
7. Effects of light winds and atmospheric stability on dispersion, and
8. Other limitations, such as lack of chemical treatment in some models.

This chapter presents a short description of each of these problem areas and offers some
recommendations. Given that AEDT (coupled with AERMOD) is the current regulatory model
for modeling aircraft emissions at airports in the United States, all the recommendations in this
guidebook focus on this model.

6.1Incorporation of Background
Pollutant Concentrations
The models used in this study capture the contributions to local air quality from airport sources;
however, the regional background also contributes to air quality. To account for background
contributions, estimates can be incorporated from a regional-scale model like the Community
Multi-scale Air Quality (CMAQ) model, or ambient monitoring data from background sites can
be used through a geostatistical technique, such as Space-Time Ordinary Kriging (STOK), which
was developed to support exposure assessment studies (Arunachalam et al. 2014).
The CMAQ model is a three-dimensional Eulerian (i.e., gridded) atmospheric chemistry and
transport modeling system developed and maintained by EPA. (The term Eulerian refers to a
fixed spatial coordinate system, which means the model is not Lagrangian and does not fol-
low the plume or flow; the term is also used to refer to the solution scheme, which is based
on solving equations that govern variables such as the wind, temperature, and concentration,
on a fixed-grid system.) CMAQ was designed as a modular system, able to incorporate data
from other models that have alternate mathematical processes, so it works well in concert with
EDMS. CMAQ has multi-pollutant capabilities to address air quality issues such as O3, PM,
and air toxics, all of which are known to have adverse health impacts. It is non-proprietary and
includes emission, meteorology, and chemical modeling components. The chemistry-transport
component can simulate chemical transformation and fate. CMAQ is a multi-scale system that

25

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

26 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

can be applied over local, regional and continental areas with progressively finer resolution in
a series of nested grids. With the temporal flexibility of the model, simulations can evaluate
short-term (weeks to months) transport from local sources as well as longer-term (annual to
multi-year) pollutant evolution. CMAQs atmospheric chemistry modeling capability and its
ability to model large geographical areas are important aspects for using the model with EDMS
to analyze airport air quality. NWS uses CMAQ to produce daily US forecasts for O3 air quality,
and it is also used by states to assess implementation actions needed to attain EPAs air qual-
ity regulations. CMAQ can be obtained at no cost from the Center for Community Modeling
and Analyses System (CMAS), which is hosted at the University of North Carolina at Chapel
Hill. CMAQ has a large user community across the globe, and user support is offered through
the CMAS Center. Although the research team did not use CMAQ in ACRP Project 02-58, it
is discussed here as an alternate regional-scale model that can be used to estimate background
concentrations. CMAQ was used in conjunction with AERMOD using a hybrid approach to
predict total concentrations at an airport in a previous ACRP-funded study published as ACRP
Report 71, and explained later in this chapter (Kim et al. 2012).

CMAQ
Details on CMAQ resources and products can be found at the following website:

https://www.cmascenter.org/cmaq/

6.2 Source Characterization


The evaluation of AERMOD and other models at LAX required the use of about 5,000 area
sources to describe the aircraft plumes and trajectories. This was by far the largest source cat-
egory used in the airport modeling, and it required the most computational resources and time.
Each area source was modeled as a continuous release. The SCICHEM model had the second-
order closure treatment and offered the most detailed chemical treatment for gas-phase and aerosol
species that are of interest, but even without turning on the chemical mechanism and running
SCICHEM in tracer mode, the computational burden was prohibitively large. Additional work is
suggested to perform a range of sensitivity simulations to determine the level of detail adequate
for accurately characterizing local-scale air quality and health. The researchers suggest exploration
of these options:
1. Revisit the current approach for EDMS/AERMOD to create a uniform set of rectangular area
sources to represent aircraft activity during landing and take-offs;
2. Consider aggregating these sources to have fewer numbers, and evaluate potential loss of
information in characterizing local air quality and health; and
3. Consider alternate treatments. (Additional discussion of alternate treatments is provided in
this chapter in the section on line puffs.)

6.3 Inventory of UFPs


Concern about contributions to adverse health effects from UFPs (< 100 nm in size) from
airport emissions is increasing. Currently, EDMS (and AEDT) inventories do not report UFP as
a separate category. As described in the methodology for estimating PM emissions from aircraft,
almost all particles emitted by aircraft are < 100 nanometers, with a small fraction ranging in size

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Future Research Needs 27

between 100250 nanometers (Petzold et al. 1999; Wayson et al. 2008; Lobo et al. 2015; Brem
et al. 2016; Ortega et al. 2016). However, current official guidance for the AEDT model refers to
these emissions as PM2.5. It is suggested that future updates to AEDT explicitly include an inven-
tory of ultra-fine particles from aircraft engines as a primary pollutant.

6.4 Plume Rise from Aircraft Emissions


A major limitation in applying AERMOD to aircraft emissions is not accounting for plume
rise from the engine exhaust. This leads to overestimation of ground-level concentrations
(GLCs) for on-airport and near-airport measurements, as found in the June 2013 LAX AQSAS.
The results showed that the modeled NOx GLCs at four monitor sites around LAX were roughly
an order of magnitude higher than observations at the high end of the quantile-quantile (Q-Q)
plots. The researchers found similar overpredictions of 1-hour NOx concentrations in ACRP
Project 02-58, not only for AERMOD but also for the other models. A correction for plume rise
would be needed to improve the GLC predictions.
In AERMOD, the aircraft plume is modeled using a series of area sources that are treated as
neutrally buoyant releases from a fixed height (the engine height zs above the surface). No option
exists to treat distance-dependent plume rise, but the source height could be greater than zs. In view
of this limitation, the researchers suggest that the area-source height be equal to the engine height
plus the final rise of the aircraft plume or the effective source height as used in the Gaussian
plume model. The final rise typically depends on the wind speed, stability, and engine heat or
buoyancy flux, and separate final rise models can be used to address neutral, stable, and unstable
atmospheric conditions.
The researchers suggest consideration of three general plume rise models for aircraft emissions:
1. The existing model with a zero plume rise;
2. An empirical model for plume rise and initial spread based on the LIDAR measurements of
Wayson et al. (2004 and 2008); and
3. A fluid-mechanical entrainment model (FEM).
The FEM suggested in this guidebook is a more general and fundamental model for aircraft
plume rise. From the FEM (presented in greater detail in the Contractors Final Report), sev-
eral choices or options are available for using the plume rise in AERMOD and other dispersion
models. The choices include:
1. An average rise (over time) at the aircraft starting position akin to the average rise obtained
from LIDAR by Wayson et al. (2008);
2. An average rise over the ground roll;
3. The final rise based on atmospheric conditions that limit plume risestable stratification
and atmospheric turbulence; and
4. A local or separate plume rise for each elemental area source along the aircraft trajectory.
Further research and comparisons between model-predicted concentrations and observations
are required to help determine the most appropriate plume rise prediction.

6.5Aircraft Downwash Effects on Plume Rise


and Dispersion
Wing-tip (WT) vortices and downwash are generated by the pressure difference across the
bottom and top wing surfaces, which leads to the rotational flow about the wings. (For more
information, see Appendix G in the contractors final report). Once generated, the WT vortices

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

28 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

leave the wing as trailing vortices, which are long lived and can cause safety issues for nearby aircraft.
Extensive literature examines WT vortices, or wakes, and their effects on engine exhaust dispersion
and concentrations (Garnier, Brunet, and Jacquin 1997; Schumann et al. 1995; Unterstrasser et al.
2014). However, most studies address dispersion at high altitude, where the vortices are far removed
from the effects of solid boundaries.
Based on their review, the research team suggests further work/research be performed on the
effects of aircraft wake downwash, specifically to:
1. Develop a new, coupled plume rise-wake model for predicting and assessing the effects of wake
vortices on plume rise, dispersion, and GLCs. This model would be directed mostly at the take-
off phase, but also should be made applicable to plume behavior during the ground roll.
2. Conduct further analysis of observed near-runway surface concentrations of aircraft pol-
lutants (at any airport) to determine if high concentrations occur near the runway take-off
end and how they compare with those near the runway starting position. Results from ACRP
Project 02-58 show that high concentrations do indeed occur near the starting end.
3. Consider a potential field experiment deploying a dense array of near-runway surface moni-
tors of an aircraft pollutant that can be measured in real time (e.g., NOx). The experiment
could again determine if high concentrations occur near the take-off end of the runway and
how these concentrations compare with concentrations near the aircraft starting position.
Such an experiment could be considered if:
i. Any existing near-runway monitors suggest that concentrations near the take-off end are
significant, and/or
ii. Sufficiently high surface concentrations are found using the new coupled plume rise-wake
model.
4. Analyze existing large eddy simulation results of wake vortices for aircraft on or near the ground
and use these to guide the development and testing of the coupled plume rise-wake model.

6.6Aircraft Dispersion Based on Instantaneous


Line Puffs
The evaluation of AERMOD and other models at LAX required the use of approximately
5,000 area sources to describe the aircraft plumes and trajectories. This was by far the largest
source category in the airport modeling and required the most computational resources and
time; each area source was modeled as a continuous release. To improve the calculation effi-
ciency, the researchers believe that it would be useful to consider instantaneous line puffs or
segmented line puffs as the basic source configuration for the aircraft plume. Essentially, one
would model the ground-roll plume as one line puff and the initial climb-out plume as a second
line puffalbeit, as a tilted or slanted line puff.
Two key reasons for considering a line puff as the main source geometry for the aircraft
plume are:
1. The line puff is more physically representative of the plume than a number of continuous area
sources, each with its own plume, and
2. The concentration and dose at some location and time due to a single aircraft (or line puff)
can be determined and then the concentration and dose from many aircraft can be super-
imposed to obtain appropriate values over an averaged time (e.g., 1 hour).
An example of this approach is the dose or exposure due to an instantaneous crosswind line
source (in the y direction, normal to the mean wind) that is transported by the local wind; the
dose or exposure is the integral of the concentration over time. Near the ground, a crosswind
line source could be created by a mobile vehicle releasing tracer along a road normal to the mean

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Future Research Needs 29

wind; similarly, an elevated crosswind line source could be generated by an aircraft releasing
tracer along a path normal to the mean wind. The contractors final report includes an analyti-
cal dose expression for a crosswind line source from which the concentration can be found by
dividing the dose by the average time.
The researchers suggest that it would be useful to conduct a preliminary feasibility study to
develop and implement the line-puff model for aircraft plumes and compare its concentra-
tion predictions and performance, especially run time, with those of the existing area-source
approach. The results of such a study would provide an assessment of the potential benefits of
the approach for further consideration.

6.7Effects of Light Winds and Atmospheric


Stability on Dispersion
Analysis of the model performance results shows clear trends of predicted and observed con-
centrations with time of day, wind speed, and atmospheric stability, especially at the four LAWA
core sites.
Several processes or causes are either not included or incompletely addressed in the models
evaluated, leading to the overestimation of maximum concentrations in light winds. They are:
1. Along-wind dispersion (Sharan and Yadav 1998), which is typically not considered in a
Gaussian plume model and leads to a reduction in concentration, but which can be accom-
modated more easily in a puff model;
2. Plume or puff meandering in light winds, which leads to enhanced lateral dispersion and
lower concentrations;
3. Insufficient dispersion and turbulence parameterization (e.g., of the lateral root mean square
[RMS] turbulence velocity, sv, or the RMS angular wind deviation, sq) when using the fric-
tion velocity (up) to parameterize sv and the dispersion (sy) due to the difficulty in uniquely
characterizing up in very low winds, creating a very large scatter and uncertainty in up(Qian
and Venkatram 2011); and
4. Adequate accounting for the increase of wind speed with height near the ground using the
log law or similarity theory for the wind speed increase, which is important for near-surface
sources (Eckman 1994; Qian and Venkatram 2011).
As noted by Qian and Venkatram, use of direct measurements of the lateral turbulence, sv or
sq, leads to improved estimates of lateral dispersion and concentration in light winds (Qian and
Venkatram 2011).
One suggestion would be to take one model (e.g., AERMOD) or a surrogate (AERMOD simu-
lator) and address/improve all or as many of these four processes as possible and re-evaluate the
model with the LAWA data set. This approach would permit an assessment of proper treatment
of the processes and the resulting change in the model performance.

6.8 Other Limitations


The range of local-scale models considered in this study had a wide range in their treatment of
chemical processes. From a health impact point of view, the six criteria pollutants of interest in
the NAAQS are CO, Pb, NO2, SO2, O3, and PM2.5. Given the large amount of NOx emissions from
aircraft, any O3 that is formed is immediately titrated; hence, O3 concentrations in the vicinity
of the airport are usually of less interest. At a minimum, the local-scale models used for airport
dispersion modeling should have the ability to predict CO, NO2, SO2, and directly emitted PM2.5.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

30 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

AERMOD has a detailed three-tiered approach to predict NO2 given NOx concentrations:
(1) the Plume Volume Molar Ratio Method (PVMRM), (2) the Ozone Limiting Method (OLM),
and (3) the Ambient Ratio Method (ARM). These approaches are designed primarily for emis-
sions from tall stacks such as power plants and other large stationary sources. ACRP has spon-
sored a separate study to develop a NOx-to-NO2 chemistry module specifically designed for
aircraft sources in EDMS-AEDT. At publication of ACRP Research Report 179, the final report
from that study was not yet available. It is recommended that future dispersion studies consider
use of the new module and the associated guidance, when they become available, to predict NO2.
Given that secondary PM2.5 has a more homogenous characteristic spatially, estimates for
secondary PM2.5 can be obtained from a more comprehensive model like CMAQ. Hybrid
approaches described in other studies can be used to obtain total PM2.5 concentrations (Isakov
et al. 2009; Davis and Arunachalam 2009; Kim et al. 2012; Yim et al. 2015; Chang et al. 2017).
In these hybrid approaches, the local-scale contribution for PM2.5 is assumed to be primary
(directly emitted), and the secondary component is assumed to be formed from atmospheric
chemical reactions by interactions with other emitted species from non-aircraft sources. The
secondary component is estimated from the regional-scale CMAQ model, and then the two
estimates (primary and secondary) are simply added to obtain the hybrid estimate. From these
prior studies, the hybrid approach has been shown to give much better model performance
than single models when compared to observations at local scales. The advantages of the hybrid
approach are that it incorporates complex chemistry as well as accounts for regional-scale trans-
port of the PM2.5 component from upwind regions of the airport region. Given that this hybrid
approach requires additional expertise in a more complex model such as CMAQ, however, one
can also obtain a spatial field of these concentrations using statistical approaches such as STOK,
as was illustrated by Arunachalam et al. in support of another environmental exposure study
(Arunachalam et al. 2014). Given that aircraft-related air quality impacts are primarily due to
PM2.5 (primary in the near field and secondary in the far field), use of a hybrid approach that
estimates both near- and far-field concentrations will enhance this characterization to a large
degree and improve model performance.

6.9 Interim Guidance


Given that several of the recommendations put forth in this report are for the longer-term,
until dispersion models are improved, the interim recommendation is for airport air quality
practitioners to adopt the following approaches:
1. Assume that most or all of the PM2.5 emitted from aircraft during LTO cycles is of a size less
than 100 nanometers, and hence the AEDT-based PM2.5 inventory for aircraft emissions is
essentially the same as aircraft inventories of UFP.
2. When it becomes available, use the new NOx-to-NO2 module being developed as part of
ACRP Project 02-43 for estimating NO2 concentrations. An alternative is to use the three-
tiered approach from AERMOD in sensitivity mode and assess NO2 model performance.
3. Use a hybrid modeling approach to estimate PM2.5 (using CMAQ or STOK, as described in
this chapter) for improved model performance.
4. Explore the treatment of aircraft sources as volume sources instead of the current default
approach, which treats them as area sources in AERMOD, and then assess model performance.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Glossary

ADMS-Airport Atmospheric Dispersion Modeling System developed specifically


for airport assessments.
AEDT Aviation Environmental Design Tool.
AERMET Meteorological data preprocessor for AERMOD.
AERMOD American Meteorological Society (AMS)/EPA Regulatory Model.
Advection Transport mechanism of a substance (or pollutant) due to the fluids
bulk motion.
Air toxics Airborne pollutants deemed hazardous because they cause or may
cause serious health effects or adverse environmental and ecologi-
cal effects; also called hazardous air pollutants.
APU Auxiliary power unit.
Area source Agglomeration of many sources that have low emission rates spread
over a large area (e.g., a parking lot) and that are too numerous to
treat individually.
Buoyancy (or upthrust) Upward force exerted by a fluid.
CALPUFF California Puff model.
Chemical reaction Process of rearrangement of an element or a compound, thus
forming a new chemical species, as opposed to a physical or nuclear
change.
CMAQ Community Multi-Scale Air Quality model.
Complex terrain A region having complex topography, with variations in land use,
likely to create unique local weather characteristics.
Dispersion Process by which atmospheric pollutants disseminate due to wind
and vertical stability.
Dispersion models Mathematical simulation of air pollutants dispersing in the ambi-
ent atmosphere.
Downwash Change in direction of air deflected by the aerodynamic action of
aircraft (or other vehicle movement).
Dry and wet deposition Atmospheric process of removal of air pollutants from the atmo-
sphere onto the ground, without (dry) or with (wet) precipitation.
EDMS Emissions and Dispersion Modeling System.

31

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

32 Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Emissions chemistry Chemical transformation of pollutants within a plume or the


atmosphere.
Emissions inventory Quantity or mass (e.g., lbs. or kg) of pollutant emissions from all
sources on the airport or those within the study scope.
Eulerian A fixed spatial coordinate system (i.e., not Lagrangian or following
the plume or flow) and the solution scheme for the wind, tempera-
ture, and concentration, which is based on solving the equations
governing these variables (wind, concentration, etc.) on an Eulerian
or fixed-grid system.
Gaussian Concentration profile shape (in y and z) within the plume (i.e., a
Gaussian or normal distribution).
GSE Ground support equipment.
GLC Ground-level concentration.
Lagrangian Moving puffs (as plumes) as a modeling frame of reference.
Local meteorology Weather-related parameters for the immediate vicinity of the
source.
Maintenance area An area that was designated non-attainment for one of the National
Ambient Air Quality Standards (NAAQS) but later met the stan-
dard and was re-designated to attainment. To ensure that the
air quality in this area continues to meet the NAAQS, states are
required to develop and implement Maintenance State Implemen-
tation Plans.
Mobile sources Moving vehicles that emit pollutants (e.g., aircraft, automobiles,
and ground support equipment).
Momentum Force of movement.
Monin-Obukhov (M-O) Height at which turbulence is generated more by buoyancy than by
wind shear; often used as indicator of atmospheric stability. In the
daytime over land, M-O length is typically between 1 to 50 meters.
MOVES Motor Vehicle Emission Simulator; EPAs emission modeling sys-
tem that estimates emissions for mobile sources.
NAAQS National Ambient Air Quality Standards.
Non-attainment A non-attainment area is considered to have air quality worse than
the NAAQS. An area may be a non-attainment area for one pollut-
ant and an attainment area for others.
Non-steady state Local meteorological conditions that change with time.
Plume A vertically or longitudinally moving, rising, or expanding mass
of gas.
Plume dynamics Of or related to the motion of a plume under the action of various
atmospheric conditions.
Point sources Emission sources that come from a specific, stationary location
(e.g., boilers, turbines, generators, cooling towers).
Pollutant Abundance of a pollutant within a reference volume (e.g., ppm or
concentration mg/m3).

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Glossary33

Puff Spatial distribution of concentrations created by an instantaneous


source (i.e., not continuous).
QA/QC Quality assurance/quality control; process used to measure the per-
formance of a dispersion model relative to measured values.
Receptor A specified point in space or on the ground at which pollutant
concentrations can be measured or computed. Sensitive receptors
include, but are not limited to, hospitals, schools, daycare facilities,
housing for elderly people, and convalescent facilities. Occupants
of these areas are more susceptible to the adverse effects of expo-
sure to pollutants. Off-site receptors include those located beyond
the property boundary of the source region, and areas within the
property boundary to which the public has routine and unre-
stricted access during or outside business hours.
SCICHEM Second-Order Integrated Puff Model with Chemistry.
Second-order Refers to the turbulence closure scheme and assumptions/
hypotheses used in the solution to the equations governing the wind,
turbulence, and concentration of the puff transport and dispersion.
Steady state Meteorological conditions that do not change with time and are an
approximation to the flow field.
Shear Movement of a gas mass (e.g., a plume) across the ambient or pre-
vailing atmosphere.
Stationary source A source of emissions that is immobile (e.g., a fuel tank, degreaser,
or power plant).
UFP Ultra-fine particles, or particles of diameter less than 100 nanometers.
Velocity Speed at which a gas moves with respect to a stationary reference,
such as the airport.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Bibliography and References

Arunachalam, S., et al. (2014). Estimating Regional Background Air Quality using Space/Time Ordinary
Kriging to Support Exposure Studies. Int. J. Environ. Res. Public Health, 11(10), 1051810536; doi:10.3390/
ijerph111010518.
Brem, B., et al. (2016). Particulate Matter and Gas Phase Emission Measurement of Aircraft Engine Exhaust. Final
Report prepared by EMPA Advanced Analytical Technologies/ETH Zurich for Swiss Federal Office of Civil
Aviation FOCA (Bundesamt fr Zivilluftfahrt, BAZL), June 2016.
Carr, E., et al. (2011). Development and Evaluation of an Air Quality Modeling Approach to Assess Near-Field
Impacts of Lead Emissions from Piston-Engine Aircraft Operating on Leaded Aviation Gasoline. Atmo-
spheric Environment, 45(32), 57955804.
Carruthers, D., S. Gray, K. Johnson, and C. McHugh (2007). Intercomparison of Five Modelling Approaches
Including ADMS-Airport and EDMS/AERMOD for Predicting Air Quality in the Vicinity of London
Heathrow Airport. In: Air and Waste Management AssociationGuideline on Air Quality Models: Applica-
tions and FLAG Developments 2006. A&WMA, 247258.
Carruthers, D., et al. (2008). ADMS-Airport: Model Inter-Comparisons and Model Validation. Presented at
12th International Conference of Harmonisation within Atmospheric Dispersion Modelling for Regulatory
Purposes, October 69, 2008, Cavtat, Croatia.
Carruthers, D., C. McHugh, M. Jackson, and K. Johnson (2011). Developments in ADMS-Airport to Take
Account of Near Field Dispersion and Applications to Heathrow Airport. International Journal of Environ-
ment and Pollution, 44(14), 332341.
Chang, S.-Y., et al. (2017). Finely Resolved On-Road PM2.5 and Estimated Premature Mortality in Central North
Carolina. Risk Analysis. John Wiley & Sons; doi:10.1111/risa.12775.
Cirillo, M., and A. Poli (1992). An Intercomparison of Semiempirical Diffusion Models Under Low Wind Speed,
Stable Conditions. Atmos. Environ., 26A, 765774.
Davis, N., and S. Arunachalam (2009). A Hybrid CMAQ and AERMOD Approach to Investigate the Impact of
Airports on Local Air Quality. In Proceedings of the 8th Annual Models-3/CMAS Users Conference, Chapel
Hill, NC, Oct 2009.
Eckman, R.M. (1994). Re-Examination of Empirically Derived Formulas for Horizontal Diffusion from Surface
Sources. Atmos. Environ., 28, 265272.
Farias, F., and H. ApSimon (2006). Relative Contributions from Traffic and Aircraft NOx Emissions to Exposure
in West London. Environmental Modelling & Software, 21(4), 477485.
Fleuti, E., and P. Hofmann (2005). Zurich Airport 2004: A Comparison of Modelled and Measured Air Quality.
European Organisation for the Safety of Air Navigation, EUROCONTROL.
Garnier, F., S. Brunet, and L. Jacquin (1997). Modelling Exhaust Plume Mixing in the Near Field of an Aircraft.
Ann. Geophysicae, 15, 14681477.
Hudda, N., and S.A. Fruin (2016). International Airport Impacts to Air Quality: Size and Related Properties of
Large Increases in Ultrafine Particle Number Concentrations. Environ. Sci. Technol. 50, 33623370.
Hudda, N., T. Gould, K. Hartin, T.V. Larson, and S.A. Fruin (2014). Emissions from an International Airport
Increase Particle Number Concentrations 4-fold at 10 km Downwind. Environ. Sci. Technol. 48, 66286635.
International Civil Aviation Organization (2011). Airport Air Quality Manual, Doc 9889, 1st Ed. ISBN 978-92-
9231-862-8, Available at: http://www.icao.int/publications/Documents/9889_cons_en.pdf.
Isakov, V., et al. (2009). Combining Regional- and Local-Scale Air Quality Models with Exposure Models for Use
in Environmental Health Studies. J. Air & Waste Manage. Assoc. 59, 461472.
Janicke, U., E. Fleuti, and I. Fuller (2007). LASPORTA Model System for Airport-Related Source Systems
Based on a Lagrangian Particle Model. Proceedings of the 11th International Conference on Harmonisation

34

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Bibliography and References 35

within Atmospheric Dispersion Modelling for Regulatory Purposes, p. 352. Available at: http://www.harmo.org/
conferences/Proceedings/_Cambridge/publishedSections/Op352-356.pdf.
Kim. B., et al. (2012). ACRP Report 71: Guidance for Quantifying the Contribution of Airport Emissions to Local Air
Quality. Transportation Research Board of the National Academies, Washington, D.C. Available at: http://
onlinepubs.trb.org/onlinepubs/acrp/acrp_rpt_071.pdf.
Lobo, P., et al. (2015). Measurement of Aircraft Engine Non-Volatile PM Emissions: Results of the Aviation-
Particle Regulatory Instrumentation Demonstration Experiment (A-PRIDE) 4 Campaign, Aerosol Sci. Tech.,
49:7,472484, DOI: 10.1080/02786826.2015.1047012.
Martin, A. (2006). Verification of FAAs Emissions and Dispersion Modeling System (EDMS). Electronic Theses
and Dissertations. 1044. University of Central Florida, Orlando, FL. Available at: http://stars.library.ucf.
edu/etd/1044.
Miracolo, M.A., et al. (2011). Secondary Aerosol Formation from Photochemical Aging of Aircraft Exhaust in a
Smog Chamber. Atmos. Chem. Phys., 11, 41354147; doi:10.5194/acp-11-4135-2011.
Ortega, I.K., et al. (2016). Measuring Non-Volatile Particle Properties in the Exhaust of an Aircraft Engine.
J. Aerospace Lab, AL 11-08, June 2016. Available at: http://www.aerospacelab-journal.org/sites/www.
aerospacelab-journal.org/files/AL11-08.pdf.
Peace, H., J. Maughan, B. Owen, and D. Raper (2006). Identifying the Contribution of Different Airport Related
Sources to Local Urban Air Quality. Environmental Modelling & Software, (21)4, 532538.
Peters, J., P. Berghmans, J. Van Laer, and E. Frijns (2016). UFPen BC-Metingen Rondom de Luchthaven van
Zaventem. Studieuitgevoerd in opdracht van de VMM en van BIM, 2016/MRG/R/0493.
Petzold, A., A. Dopelheuer, C.A. Brock, and R. Schroder (1999). In Situ Observations and Model Calculations of
Black Carbon Emission by Aircraft at Cruise Altitude. J. Geophys. Res. 104(D18), 2217122181.
Qian, W., and A. Venkatram (2011). Performance of Steady-State Dispersion Models Under Low Wind-Speed
Conditions. Bound.-Layer Meteor., 138, 475491.
Riley, E., et al. (2016). Ultrafine Particle Size as a Tracer for Aircraft Turbine Emissions. Atmos. Environ. 139
(2016) 20e29.
Sabatino, S.D., E. Solazzo, and R. Britter (2011). The Sustainable Development of Heathrow Airport: Model
Inter-Comparison Study. International Journal of Environment and Pollution, (44)1/2/3/4, 351358.
Sarrat, C., S. Aubry, and T. Chaboud (2012). Modelling Air Traffic Impact on Local Air Quality With IESTA and
ADMS-Airport: Validation Using Field Measurements on a Regional Airport. Presentation at 28th Congress
of International Council of the Aeronautical Sciences, September 2328, 2012, Brisbane, Australia.
Schumann, U., et al. (1995). Estimate of Diffusion Parameters of Aircraft Exhaust Plumes Near the Tropopause
from Nitric Oxide and Turbulence Measurements. J. Geophy. Res., 100, 1414714162.
Sharan, M., and A.K. Yadav (1998). Simulation of Experiments Under Light Wind, Stable Conditions by a
Variable K-Theory Model. Atmos. Environ., 32, 34813492.
Steib, R., K. Labancz, Z. Ferenczi, and B. Alfoldy (2008). Airport (Budapest Feribegy-Hungary) Air Quality Analysis
Using the EDMS Modeling System, Part I: Model Development and Testing. IDOJARAS, (112)2, 99112.
Tetra Tech, Inc. (2013). LAX Air Quality and Source Apportionment Study. Los Angeles World Airports. Available
at: http://www.lawa.org/airQualityStudy.aspx?id=7716.
Unterstrasser, S., R. Paoli, I. Solch, C. Kuhnlein, and T. Gerz (2014). Dimension of Aircraft Exhaust Plumes at
Cruise Conditions: Effect of Wake Vortices. Atmos. Chem. Phys., 14, 27132733.
Wayson, R.L., G.G. Fleming, B. Kim, W.L. Eberhard, and W.A. Brewer (2004). Final Report: The Use of Lidar to
Characterize Aircraft Initial Plume Characteristics. Letter Report, Federal Aviation Administration, Wash-
ington, D.C.
Wayson, R.L., et al. (2008). Lidar Measurement of Exhaust Plume Characteristics from Commercial Jet Turbine
Aircraft at the Denver International Airport. Tech. Rep. FAA-AEE-08-02, Federal Aviation Administration,
Washington, D.C.
Weil, J.C., L.A. Corio, and R.P. Brower (1997). A PDF Dispersion Model for Buoyant Plumes in the Convective
Boundary Layer. J. Appl. Meteor., 36, 9821003.
Wilkerson, J.T., et al. (2010). Analysis of Emission Data from Global Commercial Aviation: 2004 and 2006.
Atmos. Chem. Phys. Atmos. Chem. Phys. 10, 63916408.
Woody, M.C., et al. (2015). Estimates of Non-traditional Secondary Organic Aerosols from Aircraft SVOC and
IVOC Emissions Using CMAQ. Atmos. Chem. Phys., 15, 69296942; doi:10.5194/acp-15-69292015.
Yim, S.H.L., et al. (2015). Global, Regional and Local Health Impacts of Civil Aviation Emissions. Environ. Res.
Lett. 10(3), 034001. IOP Science.

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

Abbreviations and acronyms used without definitions in TRB publications:


A4A Airlines for America
AAAE American Association of Airport Executives
AASHO American Association of State Highway Officials
AASHTO American Association of State Highway and Transportation Officials
ACINA Airports Council InternationalNorth America
ACRP Airport Cooperative Research Program
ADA Americans with Disabilities Act
APTA American Public Transportation Association
ASCE American Society of Civil Engineers
ASME American Society of Mechanical Engineers
ASTM American Society for Testing and Materials
ATA American Trucking Associations
CTAA Community Transportation Association of America
CTBSSP Commercial Truck and Bus Safety Synthesis Program
DHS Department of Homeland Security
DOE Department of Energy
EPA Environmental Protection Agency
FAA Federal Aviation Administration
FAST Fixing Americas Surface Transportation Act (2015)
FHWA Federal Highway Administration
FMCSA Federal Motor Carrier Safety Administration
FRA Federal Railroad Administration
FTA Federal Transit Administration
HMCRP Hazardous Materials Cooperative Research Program
IEEE Institute of Electrical and Electronics Engineers
ISTEA Intermodal Surface Transportation Efficiency Act of 1991
ITE Institute of Transportation Engineers
MAP-21 Moving Ahead for Progress in the 21st Century Act (2012)
NASA National Aeronautics and Space Administration
NASAO National Association of State Aviation Officials
NCFRP National Cooperative Freight Research Program
NCHRP National Cooperative Highway Research Program
NHTSA National Highway Traffic Safety Administration
NTSB National Transportation Safety Board
PHMSA Pipeline and Hazardous Materials Safety Administration
RITA Research and Innovative Technology Administration
SAE Society of Automotive Engineers
SAFETEA-LU Safe, Accountable, Flexible, Efficient Transportation Equity Act:
A Legacy for Users (2005)
TCRP Transit Cooperative Research Program
TDC Transit Development Corporation
TEA-21 Transportation Equity Act for the 21st Century (1998)
TRB Transportation Research Board
TSA Transportation Security Administration
U.S.DOT United States Department of Transportation

Copyright National Academy of Sciences. All rights reserved.


Dispersion Modeling Guidance for Airports Addressing Local Air Quality Health Concerns

ADDRESSSERVICEREQUESTED

Washington, DC 20001
500 Fifth Street, NW
TRANSPORTATIONRESEARCHBOARD
NON-PROFIT ORG.

ISBN 978-0-309-44654-9
COLUMBIA, MD
PERMIT NO. 88

90000
U.S. POSTAGE
PAID

9 780309 446549

Copyright National Academy of Sciences. All rights reserved.

Das könnte Ihnen auch gefallen