Public Comment - Exhibit G

From:            laura gibbons
To:                Commission-Public-Records
Subject:           [EXTERNAL] Written comment for the Feb 9, 2020 port meeting.
Date:              Saturday, February 6, 2021 9:09:31 AM

WARNING: External email. Links or attachments may be unsafe.
(This is written only - I won't be able to join the meeting. Thanks!)
Commissioners,
I would like to comment on item 11a, 2021 Committee Workplans. You need to add to the Aviation
Committee Workplan the task of figuring out how you are going to meet your 2030 target of a 50%
reduction in Scope 3 greenhouse gas emissions from 2007 levels. This is only 9 years out; you need to
get on it!
Please remember that AIRPLANE EMISSIONS HAVE 3 TIMES GREATER WARMING IMPACT ON
CLIMATE THAN ON-THE-GROUND EMISSIONS, making reducing aviation even more critical than its
carbon emissions might indicate. Here is the latest scientific study to quantify the factor of three:
(https://www.sciencedirect.com/science/article/pii/S1352231020305689?via%3Dihub)
Sincerely,
Laura Gibbons
Seattle








Atmospheric Environment 244 (2021) 117834

Contents lists available at ScienceDirect
Atmospheric Environment
journal homepage: http://www.elsevier.com/locate/atmosenv

The contribution of global aviation to anthropogenic climate forcing for
2000 to 2018
D.S. Lee a,*, D.W. Fahey b, A. Skowron a, M.R. Allen c,n, U. Burkhardt d, Q. Chen e, S.J. Doherty f,
S. Freeman a, P.M. Forster g, J. Fuglestvedt h, A. Gettelman i, R.R. De Leon a, L.L. Lim a, M.
T. Lund h, R.J. Millar c,o, B. Owen a, J.E. Penner j, G. Pitari l, M.J. Prather k, R. Sausen d, L.
J. Wilcox m
a Faculty of Science and Engineering, Manchester Metropolitan University, John Dalton Building, Chester Street, Manchester, M1 5GD, United Kingdom
b NOAA Chemical Sciences Laboratory (CSL), Boulder, CO, USA
c School of Geography and the Environment, University of Oxford, Oxford, UK
d Deutsches Zentrum fr Luft- und Raumfahrt (DLR), Institut fr Physik der Atmosphare, Oberpfaffenhofen, Germany
e State Key Joint Laboratory of Environmental Simulation and Pollution Control, College of Environmental Sciences and Engineering, Peking University, Beijing, 100871,
China
f Cooperative Institute for Research in Environmental Sciences (CIRES), University of Colorado, Boulder, CO, USA
g School of Earth and Environment, University of Leeds, Leeds, LS2 9JT, United Kingdom
h CICEROCenter for International Climate ResearchOslo, PO Box 1129, Blindern, 0318, Oslo, Norway
i National Center for Atmospheric Research, Boulder, CO, USA
j Department of Climate and Space Sciences and Engineering, University of Michigan, 2455 Hayward St., Ann Arbor, MI, 48109-2143, USA
k Department of Earth System Science, University of California, Irvine, 3329 Croul Hall, CA, 92697-3100, USA
l Department of Physical and Chemical Sciences, Universit`a dell'Aquila, Via Vetoio, 67100, L'Aquila, Italy
m National Centre for Atmospheric Science, Department of Meteorology, University of Reading, Earley Gate, Reading, RG6 6BB, UK
n Department of Physics, University of Oxford, Oxford, UK
o Committee on Climate Change, 151 Buckingham Palace Road, London, SW1W 9SZ, UK

HIGHLIGHTS            GRAPHICAL ABSTRACT
Global aviation warms Earth's surface
through both CO2 and net non-CO2
contributions.
Global aviation  contributes  a  few
percent  to  anthropogenic  radiative
forcing.
Non-CO2 impacts comprise about 2/3 of
the net radiative forcing.
Comprehensive and quantitative calcu-
lations of aviation effects are presented.
Data are made available to analyze past,
present  and  future  aviation  climate
forcing.

ARTICLE INFO            ABSTRACT
Global aviation operations contribute to anthropogenic climate change via a complex set of processes that lead to
Dedication: This paper is dedicated to the           a net surface warming. Of importance are aviation emissions of carbon dioxide (CO2), nitrogen oxides (NOx),
memory of Professor Ivar S. A. Isaksen of the
water vapor, soot and sulfate aerosols, and increased cloudiness due to contrail formation. Aviation grew

* Corresponding author.
E-mail address: d.s.lee@mmu.ac.uk (D.S. Lee).
https://doi.org/10.1016/j.atmosenv.2020.117834
Received 9 February 2020; Received in revised form 2 July 2020; Accepted 30 July 2020
Available online 3 September 2020
1352-2310/ 2020 Elsevier Ltd. All rights reserved.

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
University of Oslo, whose scientific excellence,        strongly over the past decades (19602018) in terms of activity, with revenue passenger kilometers increasing
friendship, and mentorship is sorely missed.          from 109 to 8269 billion km yr 1, and in terms of climate change impacts, with CO2 emissions increasing by a
factor of 6.8 to 1034 Tg CO2 yr 1. Over the period 20132018, the growth rates in both terms show a marked
Keywords:                                  increase. Here, we present a new comprehensive and quantitative approach for evaluating aviation climate
Aviation                                   forcing terms. Both radiative forcing (RF) and effective radiative forcing (ERF) terms and their sums are
Contrail cirrus
calculated for the years 20002018. Contrail cirrus, consisting of linear contrails and the cirrus cloudiness arising
Climate
from them, yields the largest positive net (warming) ERF term followed by CO2 and NOx emissions. The for-
Radiative forcing
CO2                                     mation and emission of sulfate aerosol yields a negative (cooling) term. The mean contrail cirrus ERF/RF ratio of
NOx                                     0.42 indicates that contrail cirrus is less effective in surface warming than other terms. For 2018 the net aviation
ERF is +100.9 milliwatts (mW) m 2 (595% likelihood range of (55, 145)) with major contributions from
contrail cirrus (57.4 mW m 2), CO2 (34.3 mW m 2), and NOx (17.5 mW m 2). Non-CO2 terms sum to yield a net
positive (warming) ERF that accounts for more than half (66%) of the aviation net ERF in 2018. Using
normalization to aviation fuel use, the contribution of global aviation in 2011 was calculated to be 3.5 (4.0, 3.4)
% of the net anthropogenic ERF of 2290 (1130, 3330) mW m 2. Uncertainty distributions (5%, 95%) show that
non-CO2 forcing terms contribute about 8 times more than CO2 to the uncertainty in the aviation net ERF in
2018. The best estimates of the ERFs from aviation aerosol-cloud interactions for soot and sulfate remain undetermined.
CO2-warming-equivalent emissions based on global warming potentials (GWP* method) indicate
that aviation emissions are currently warming the climate at approximately three times the rate of that associated
with aviation CO2 emissions alone. CO2 and NOx aviation emissions and cloud effects remain a continued focus of
anthropogenic climate change research and policy discussions.

1. Introduction                                                    time and combined to provide a net ERF for global aviation. Quantifying
the terms required new analyses of CO2 and NOx ERFs and recalibration
Aviation is one of the most important global economic activities in    of other individual ERFs accounting for factors not previously applied in
the modern world. Aviation emissions of CO2 and non-CO2 aviation ef-    a common framework.
fects result in changes to the climate system (Fig. 1). Both aviation CO2       In Lee et al. (2009), the net RF was calculated with and without the
and the sum of quantified non-CO2 contributions lead to surface    full contrail cirrus term but including an estimate for linear contrails.
warming. The largest contribution to anthropogenic climate change    The exclusion was based on the lack of a best estimate derived from
across all economic sectors comes from the increase in CO2 concentra-    existing studies. At that time radiative forcing estimates were limited to
tion, which is the primary cause of observed global warming in recent    linear or line-shaped contrails since the modelling approaches required
decades (IPCC, 2013, 2018). Aviation contributions involve a range of    scaling contrail formation frequency to observed coverage and only
atmospheric physical processes, including plume dynamics, chemical    satellite observations of linear contrails existed (Burkhardt et al., 2010).
transformations, microphysics, radiation, and transport. Aggregating    The contrail cirrus term requires the simulation of the whole contrail
these processes to calculate changes in a greenhouse gas component or a    cirrus life cycle, starting from persistent linear contrails which spread
cloud radiative effect is a complex challenge for contemporary atmo-    and often become later indistinguishable from natural cirrus. Persistent
spheric modeling systems. Given the dependence of aviation on burning    contrail formation requires ice-supersaturated conditions along a flight
fossil fuel, its significant CO2 and non-CO2 effects, and the projected    track, which are variable in space and time in the troposphere and
fleet growth, it is vital to understand the scale of aviation's impact on    tropopause region (Irvine et al., 2013). Estimating the RF from contrail
present-day climate forcing.                                           cirrus requires knowledge of complex microphysical processes, radiative
Historically, estimating aviation non-CO2 effects has been particu-    transfer, and the interaction with background cloudiness (Burkhardt
larly challenging. The primary (quantified) non-CO2 effects result from    et al., 2010). Contrail cirrus forcing dominates that of persistent linear
the emissions of NOx, along with water vapor and soot that can result in    contrails with the latter on the order of 10% of the combined forcing
contrail formation. Aviation aerosols are small particles composed of    (Burkhardt and Karcher, 2011). In the present study, we present a best
soot (black and organic carbon (BC/OC)) and sulfur (S) and nitrogen (N)    estimate and uncertainty based on the results from global climate
compounds. The largest positive (warming) climate forcings adding to    models employing process-based contrail cirrus parameterizations.
that of CO2 are those from contrail cirrus and from NOx-driven changes       Emissions of NOx from aviation lead to photochemical changes that
in the chemical composition of the atmosphere (Lee et al., 2009). Lee    increase global ozone (O3) formation while decreasing the lifetime and
et al. (2009) estimated that in 2005, aviation CO2 radiative forcing (RF    abundance of methane (CH4). The changes result in positive and nega-
(Wm 2)) was 1.59% of total anthropogenic CO2 RF and that the sum of    tive (cooling) RF contributions, respectively. Since Lee et al. (2009),
aviation CO2 and non-CO2 effects contributed about 5% of the overall    improved understanding and modeling capabilities have emerged, as
net anthropogenic forcing.                                            well as additional RF terms in response to NOx emissions, namely a
Understanding of aviation's impacts on the climate system has    longer-term decrease in background O3 and a reduction in H2O in the
improved over the decade since the last comprehensive evaluation (Lee    stratosphere in response to decreased CH4. Here, model results are used
et al., 2009), but remains incomplete. Published studies of aviation    to calculate the additional RF terms, and to incorporate the updated CH4
contributions to climate change generally focus on one or a few ERF    forcing  as  assessed  by  Etminan  et  al.  (2016)  and  the
terms. For example, about 20 studies are cited here that quantify the    equilibrium-to-transient corrections for the CH4 term (see Appendix D).
contribution from global NOx emissions. In contrast, only a few studies    Finally, aviation-specific efficacies (Appendix C) of the individual NOx
have addressed the net RF from global aviation (IPCC, 1999; Sausen    components are used to estimate a net NOx ERF for the first time.
et al., 2005; Lee et al., 2009). A more recent study updated some avia-       Lee et al. (2009) includes best estimates for the RFs resulting from
tion terms without providing a net RF (Brasseur et al., 2016). Here, a    the aerosol-radiation interactions (previously called direct effects) of
comprehensive analysis of individual aviation ERFs is undertaken in    soot and sulfate aerosols from aviation. However, no best estimates of
order to provide an overall ERF for global aviation, along with the    RFs from aerosol-cloud interactions (previously called indirect effects)
associated uncertainties, which is an analysis unavailable elsewhere.    were available in 2009. Subsequent studies discussed here have yet to
This step updates and improves the analysis of Lee et al. (2009). Best    provide a basis for best estimates of ERFs from aviation aerosol-cloud
estimates of individual aviation ERF terms are derived here for the first    interactions that may be significant.

2

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
The primary motivations for the present study are to provide an    scenario projections of aviation climate impacts to be assessed in context
updated, comprehensive evaluation of aviation climate forcings in terms    with other sectors, such as maritime shipping, ground transportation
of RF and ERF based on new calculations and the normalization of values    and energy generation. This updated understanding is especially
from published modeling studies, and to combine the resulting best es-    important given the potential role of international aviation in meeting
timates via a Monte-Carlo analysis to yield a best estimate for the net    the goals of the Paris Agreement (Section 2) on limiting future tem-
ERF for global aviation for the years 20002018. The three years 2018,    perature increases.
2011, and 2005 are notable because the year 2018 is the latest year for       The remaining sections address global aviation growth statistics
which air traffic and fuel use datasets are available, 2011 is the most    (Section 2); a brief summary of methods used in the analysis (Section 3);
recent year evaluated for net anthropogenic climate forcing by the IPCC    results for the ERF estimates of CO2, NOx, water vapor, contrail cirrus,
(IPCC, 2013), and 2005 is the year evaluated in the latest comprehensive    and aerosol-radiation and aerosol-cloud interactions with soot and sulaviation
and climate evaluation (Lee et al., 2009). By normalizing the    fate (Section 4); results for the net ERF of global aviation (Section 5);
calculations across these years, more specific and self-consistent com-    emission metrics (Section 6); and aviation CO2 vs non-CO2 forcings
parisons can be made of the changes in aviation contributions over time.    (Section 7). The appendices contain additional detailed information on
The normalization step requires addressing in each study, for example,    trends in aviation emissions (App. A); aviation CO2 radiative forcing
the choice of air traffic inventory, the integration of emissions along    calculations (App. B); radiative forcing, efficacy and ERF definitions
flight tracks, and the assumed jet-engine emission indices. The new best    (App. C); aviation NOx RF calculations (App. D); contrail cirrus RF
estimates of aviation ERF, for example, show that the 2018 value is    scaling factors and uncertainty (App. E); and emission equivalency
about 48% larger than the updated 2005 value.                         metric calculations (App. F). A Supplemental Data (SD) file is provided
In general, previous global aviation climate assessments have made    containing the interactive spreadsheet used to calculate RFs and ERFs
different assumptions concerning emissions, cloudiness effects, and    for each aviation term.
aviation operations (e.g., IPCC, 1999). Here, our self-consistent set of
component and net aviation ERFs for 2000 to 2018 allows historical and

















Fig. 1. Schematic overview of the processes by which aviation emissions and increased cirrus cloudiness affect the climate system. Net positive RF (warming)
contributions arise from CO2, water vapor, NOx, and soot emissions, and from contrail cirrus (consisting of linear contrails and the cirrus cloudiness arising from
them). Negative RF (cooling) contributions arise from sulfate aerosol production. Net warming from NOx emissions is a sum over warming (short-term ozone increase
) and cooling (decreases in methane and stratospheric water vapor, and a long-term decrease in ozone) terms. Net warming from contrail cirrus is a sum over
the day/night cycle. These contributions involve a large number of chemical, microphysical, transport and, radiative processes in the global atmosphere. The
quantitative ERF values associated with these processes are shown in Fig. 3 for 2018.
3





D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
2. Global aviation growth                                          growth is likely to be largely dependent upon the combustion of kerosene
fossil fuel (Jet A-1/A) (OECD, 2012), resulting in emission of CO2.
Global aviation fuel use and CO2 emissions have increased in the last    Renewable biofuels partially offset fossil fuel emissions but these have
four decades with large growth occurring in Asia and other developing    yet to be produced in sufficient quantities to offset growth of fossil fuel
regions due to the rapid expansion of civil aviation (Fig. 2 and Appendix    use. Furthermore, considerable uncertainties remain regarding the
A). Looking forward, this pattern of growth is expected to be main-    life-cycle emissions of biofuels, which determine the reductions in net
tainedfor example, of the 1229 orders of Airbus and 1031 orders of    CO2 emissions (e.g., Hari et al., 2015). There are current regulations
Boeing in 2017, 20.3% and 37.5%, respectively, are for airlines in the    regarding aviation emissions of CO2, NOx, and soot mass and number
Asia region (Airbus, 2017; Boeing, 2018). Airbus projects 41% of orders    based on decisions by the International Civil Aviation Organization
over the next two decades to be from the Asia-Pacific region (Airbus,    (ICAO). Under the 2016 Paris climate agreement, nations are commit-
2017). The uncertainty in this expectation has increased due to the    ting to limiting future increases in global temperatures with Nationally
slowdown in aviation operations in the early months of 2020 due to the    Determined Contributions (NDCs) (UNFCCC). Whereas domestic avia-
COVID-19 pandemic (Le Quere et al., 2020). Annual aviation emissions    tion CO2 emissions are included in the NDCs, CO2 emissions from inin
2020 are now expected to be below recent projections that are based    ternational aviation are not mentioned in the agreement. It remains
on historical growth.                                                 open as to whether emissions from international aviation or global
A striking feature of Fig. 2a is the sustained multi-decade growth in    emissions beyond greenhouse gases (e.g., short-lived (non-CO2) climate
CO2 emissions; the average rate for the period 19602018 is 15 Tg CO2    forcers) will be included in future international agreements.
yr 1. The growth rate for 2013 through 2018 is much larger (44 Tg CO2
yr 1). The annually averaged growth rate over the period 1970 to 2012    3. Methods
is 2.2% yr 1and for 2013 to 2018 is 5% yr 1(increase of 27%). In 2018,
global aviation CO2 emissions exceeded 1000 million tonnes per year for       The methodologies used to calculate ERF and RF for individual
the first time (see methodology for scaling 2016 IEA data in Appendix    aviation terms are described in this section, and results of these calcu-
A). The cumulative emissions of global aviation (19402018) are 32.6    lations are given in Section 4. Common to the methodologies is a
billion (109) tonnes of CO2, of which approximately 50% were emitted    comprehensive multi-page spreadsheet (see SD) that begins with a user's
in the last 20 years. Current (2018) CO2 emissions from aviation    guide. The spreadsheet pages include those for contrail cirrus, CO2, NOx,
represent approximately 2.4% of anthropogenic emissions of CO2    H2O, and sulfate and soot aerosol, along with CO2-equivalent metrics,
(including land use change) (Fig. 2c).                                  ERF probability distributions, ERF time series, and estimates of forcings
Aviation has grown strongly over time (Fig. 2b) in terms of available    from aerosol-cloud effects. The spreadsheet displays the results of
seat kilometers (ASK, a measure of capacity) and revenue passenger    aviation forcings provided by individual published studies. ERF and RF
kilometers (RPK, a measure of transport work). Fuel usage and hence    values were calculated for 2018 and other years based on the normalized
CO2 emissions have grown at a lesser rate than RPK, reflecting increases    values of ERF or RF per unit emission or distance, choice of appropriate
in aircraft efficiency derived from changes in technology, larger average    emission indices, and times series data on fuel use and distance travaircraft
sizes and increased passenger load factor. Aviation transport    elled. In the case of the contrail cirrus forcing, the flight-track distance
efficiency has improved by approximately eightfold since 1960, to 125    was chosen as the proxy over fuel usage. Annual global emissions are
gCO2 (RPK) 1.                                                      derived from fuel burn by multiplying by the average emission indices
At present and for some considerable time into the future, aviation    (Table 1). The combined and normalized results are used to create sets of

Fig. 2. Data related to the growth of aviation traffic
and CO2 emissions from 1940 to 2018. Panel (a):
Global aviation CO2 emissions. Underlying fuel usage
data for 1940 to 1970 are derived from Sausen and
Schumann (2000) and for 19702016 from Interna-
tional Energy Agency (UKDS, 2016) data, which
include international bunker fuels. For 2017/18, the
values are scaled from information from the International
Air Transport Association (see Appendix A).
The average annual increase of global emissions from
1960 to 2018 is 15 Tg CO2 yr 1 and the corresponding
decadal average growth rates are 8.0, 2.2, 3.0, 2.3
and 1.1% yr 1, yielding an overall average of 3.3%
yr 1. Panel (b): Global aviation traffic in RPK and
ASK  from  airlines.org  (http://airlines.org/datas
et/world-airlines-traffic-and-capacity/ ),   and   the
transport efficiency of global aviation in kg CO2 per
RPK. The passenger load factor defined as RPK/ASK
increased from about 60% in 1960 to 82% in 2018.
Panel (c): Total anthropogenic CO2 emissions and the
aviation fractions of this total with and without the
inclusion of CO2 emissions from land use change
(LUC) from the Global Carbon Budget 2018 (Le
Quere, 2018). Panel (d)(f): Additional aviation
emissions data by region and year. The yearly sums of
OECD and non-OECD values in (d) equal the respective
global total values. The regional values in (e) and
(f) also sum to equal the yearly global total values.
Note different vertical scales (http://www.oecd.org/
about/membersandpartners/) (UKDS, 2016) (Coun-
try listings in SD Spreadsheet).
4

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Table 1                                                             approach in Lee et al. (2009) to include ERF values in addition to the
Emission indices used in ERF and RF calculations.                           traditional RF values (Tables 2 and 3 and Fig. 3). The distinction be-
Emission    Emission index      Reference          Notes                  tween ERF and RF is presented in Appendix C. ERF is the preferred
metric for comparing the expected impacts of climate forcing terms
CO2        3.16 kg/kg fuel      ICAO (2018)
NOx       15.14 g/kg fuel     Fleming and Ziegler   2018, 2011             (Myhre et al., 2013). Its use derives from the stronger correlation be-
14.12 g/kg fuel      (2016)            2005                  tween ERF and the change in the equilibrium global-mean surface
Barrett et al. (2010)                        temperature for some forcing agents than for the corresponding RF. ERF
Water      1.231 kg/kg fuel     Barrett et al. (2010)                        is calculated as the change in net top-of-the-atmosphere (TOA) down-
vapor                                                            ward radiative flux after allowing for rapid adjustments in atmospheric
Soot       0.03 g/kg fuel       Barrett et al. (2010)
2  1014 particles/                                          temperatures, water vapor and clouds with globally-averaged sea sur-
kg fuela                                                  face and/or land surface temperatures unchanged. ERF is preferred over
Sulfur      1.2 g/kg fuel       Miller et al. (2010)    Assumed S content of      RF estimates because the imposed forcing and rapid responses to the
(SO2)                                       600 ppm               forcing cannot always be separately evaluated, especially for aerosols. In
a Assumes mean particle size in the range of 1179 nm diameter.              general, the largest differences between ERF and RF are expected for
aerosol-cloud interactions and contrail cirrus (Myhre et al., 2013;
RF and ERF aviation terms for the years 20002018. In addition to    Boucher et al., 2013). In calculating ERF values for 20002018, the
facilitating the present study, the spreadsheet also provides a quantita-    ERF/RF ratio is assumed to be constant with time.
tive framework for follow-on analyses.                                    Most of the results for the non-CO2 terms have associated statistics
Calculations of radiative forcing are expanded here beyond the    from which the median was chosen as the best estimate, including the
net aviation ERF and RF, and the net non-CO2 ERF and RF. For CO2 and
Table 2
Best estimates and high/low limits of the 90% likelihood ranges for aviation ERF components derived in this study.
ERF (mW m 2)                   2018a              2011a               2005a              Sensitivity to emissions                    ERF/RF
Contrail cirrus                   57.4 (17, 98)         44.1 (13, 75)         34.8 (10, 59)         9.36  10 10 mW m 2 km 1                0.42
CO2                          34.3 (28, 40)         29.0 (24, 34)         25.0 (21, 29)                                             1.0
Short-term O3 increase             49.3 (32, 76)         37.3 (24, 58)         33.0 (21, 51)         34.4  9.9 mW m 2 (Tg (N) yr 1) 1           1.37
Long-term O3 decrease              10.6 ( 20, 7.4)      7.9 ( 15, 5.5)       6.7 ( 13, 4.7)       9.3  3.4 mW m 2 (Tg (N) yr 1) 1           1.18
CH4 decrease                     21.2 ( 40, 15)      15.8 ( 30, 11)       13.4 ( 25, 9.4)      18.7  6.9 mW m 2 (Tg (N) yr 1) 1          1.18
Stratospheric water vapor decrease      3.2 ( 6.0, 2.2)       2.4 ( 4.4, 1.7)       2.0 ( 3.8, 1.4)      2.8  1.0 mW m 2 (Tg (N) yr 1) 1           1.18
Net NOx                        17.5 (0.6, 29)        13.6 (0.9, 22)         12.9 (1.9, 20)        5.5  8.1 mW m 2 (Tg (N) yr 1) 1
Stratospheric H2O increase           2.0 (0.8, 3.2)         1.5 (0.6, 2.4)         1.4 (0.6, 2.3)         0.0052  0.0026 mW m 2 (Tg (H2O) yr 1) 1     
Soot (aerosol-radiation)             0.94 (0.1, 4.0)        0.71 (0.1, 3.0)         0.67 (0.1, 2.8)        100.7  165.5 mW m 2 (Tg (BC) yr 1) 1        
Sulfate (aerosol-radiation)            7.4 ( 19, 2.6)       5.6 ( 14, 1.9)       5.3 ( 13, 1.8)       19.9  16.0 mW m 2 (Tg (SO2) yr 1) 1       
Sulfate and soot (aerosol-cloud)                                                                                               
Net ERF (only non-CO2 terms)        66.6 (21, 111)        51.4 (16, 85)         41.9 (14, 69)                                             
Net aviation ERF                  100.9 (55, 145)       80.4 (45, 114)         66.9 (38, 95)                                             
Net anthropogenic ERF in 2011                        2290 (1130, 3330)b                                                          
a The uncertainty distributions for all forcing terms are lognormal except for CO
2 and contrail cirrus (normal) and Net NOx (discrete pdf).
b Boucher et al., 2013. IPCC also separately estimated the contrail cirrus term for 2011 as 50 (20, 150) mW m 2.


Table 3
Best estimates and low/high limits of the 95% likelihood ranges for aviation RF components derived in this studya.
RF (mW m 2)                 2018b          2011b           2005b           Lee et al. (2009) 2005 values  Sensitivity to emissions (this work)
Contrail cirrus                111.4 (33, 189)    85.6 (25, 146)     67.5 (20, 115)     (11.8c)                 1.82  10 9 mW m 2 km 1
CO2                       34.3 (31, 38)     29.0 (26, 32)      25.0 (23, 27)      28.0
Short-term O3 increase          36.0 (23, 56)     27.3 (17, 42)      24.0 (15, 37)      26.3                   25.1  7.3 mW m 2 (Tg (N) yr 1) 1
Long-term O3 decrease           9.0 ( 17, 6.3)   6.7 ( 13, 4.7)    5.7 ( 11, 4.0)                          7.9  2.9 mW m 2 (Tg (N) yr 1) 1
CH4 decrease                  17.9 ( 34, 13) 13.4 ( 25, 9.3) 11.4 ( 21, 7.9) 12.5                   15.8  5.9 mW m 2 (Tg (N) yr 1) 1
Stratospheric water vapor decrease 2.7 ( 5.01.9)    2.0 ( 3.8, 1.4)   1.7 ( 3.2, 1.2)                         2.4  0.9 mW m 2 (Tg (N) yr 1) 1
Net NOx                     8.2 ( 4.8, 16)     6.5 ( 3.3, 12)     6.6 (1.9, 12)      13.8d                   1.0  6.6 mW m 2 (Tg (N) yr 1) 1
Stratospheric H2O increase       2.0 (0.8, 3.2)     1.5 (0.6, 2.4)      1.4 (0.6, 2.3)      2.8                    0.0052  0.0026 mW m 2 (Tg (H2O) yr 1) 1
Soot (aerosol-radiation)          0.94 (0.1, 4.0)    0.71 (0.1, 3.0)     0.67 (0.1, 2.8)     3.4                    100.7  165.5 mW m 2 (Tg (BC) yr 1) 1
Sulfate (aerosol-radiation)         7.4 ( 19, 2.6)   5.6 ( 14, 1.9)    5.3 ( 13, 1.8)    4.8                    19.9  16.0 mW m 2 (Tg (SO2) yr 1) 1
Sulfate and soot (aerosol-cloud)                                                                    
Net RF (only non-CO2 terms)      114.8 (35, 194)    88.4 (27, 149)     70.3 (22, 119)                           
Net aviation RF               149.1 (70, 229)    117.4 (56, 179)    95.2 (47, 144)     78.0                   
a ERF values are shown in Table 2.
b The uncertainty distributions for all forcing terms are lognormal except for CO
2 and contrail cirrus (normal) and Net NOx (discrete pdf).
c Linear contrails only; excludes the increase in cirrus cloudiness due to aged spreading contrails.
d Excludes updated CH
4 RF evaluation of Etminan et al. (2016) and equilibrium-to-transient correction.

5

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
















Fig. 3. Best-estimates for climate forcing terms from global aviation from 1940 to 2018. The bars and whiskers show ERF best estimates and the 595% confidence
intervals, respectively. Red bars indicate warming terms and blue bars indicate cooling terms. Numerical ERF and RF values are given in the columns with 595%
confidence intervals along with ERF/RF ratios and confidence levels. ERF and RF values are shown for other years in Tables 2 and 3, Fig. 6 and the SD spreadsheet. RF
values are multiplied by the respective ERF/RF ratio to yield ERF values. ERF/RF values designated as [1] indicate that no estimate is available yet. The basis for
confidence levels is presented in Table 4.
contrail cirrus, for which the sample sizes are small (3, in both cases), the    performance of LinClim and CICERO-SCM with respect to aviation
mean was used as the best estimate. The best estimates of the non-CO2    emissions is documented in the multi-model comparison of Khodayari
terms except contrail cirrus have associated uncertainties expressed as    et al. (2013). The CO2 concentrations attributable to aviation in 2018
5% and 95% confidence intervals calculated from 5, 95% percentile    based on LinClim, CICERO-SCM and FaIR are 2.9, 2.4 and 2.4 ppm,
statistics. The uncertainty distributions for all forcing terms other than    respectively, with concentrations nearly doubling in the last 20 years
CO2 and contrail cirrus are lognormal and that for net NOx has a discrete    (see SD spreadsheet). The ERF/RF ratio for CO2 is assumed to be unity.
probability distribution function (PDF). The uncertainties for the ERF    The resulting CO2 ERFs, as derived from global concentrations using
and RF of CO2 were taken from IPCC (2013) and fitted with a Monte    standard IPCC expressions (IPCC, 2001), are 38.6, 32.0 and 32.4 mW
Carlo analysis with a normal distribution (see Section 5). The un-    m 2, respectively. With only three model estimates, the average of 34.3
certainties for contrail cirrus were estimated partly from expert judge-    mWm 2 (5 and 95% percentiles of 29 and 40 mW m 2), is chosen be the
ment of the underlying processes, as described in Appendix E, again    CO2 RF best estimate.
fitted with a Monte Carlo analysis with a normal distribution.
4.2. NOx
4. Calculations of ERFs for aviation terms
The photochemical effects of aviation NOx emissions on the atmo-
4.1. CO2
spheric abundances of O3, CH4, carbon monoxide (CO) and reactive
hydrogen (HOx) are well established (Fuglestvedt et al., 1999). Earlier
The time series of aviation CO2 emissions is shown in Fig. 2 as
studies assessed the short-term increase of O3 and the longer-term
derived from combined kerosene and avgas usage (UKDS, 2016).
reduction in CH4 lifetime and abundance, which yield positive and
Calculating CO2 concentrations from emissions requires use of a global
negative RFs, respectively (IPCC, 1999; Sausen et al., 2005). Lee et al.
carbon-cycle model, which has a range of complexity from a compre-
(2009) introduced the concept of the 'net NOx' effect by combining the
hensive Earth system model (ESM) to a simple climate model (SCM),
two components, extending and updating the study of Sausen et al.
with the latter being based on a box model or impulse response function
(2005). Later studies expanded the analysis of NOx effects to include the
(IRF) model. Three SCMs were used here: LinClim, an IRF model based
long-term decreases in both O3 and stratospheric water vapor (SWV)
on Sausen and Schumann (2000) (Appendix B); the Finite-amplitude
resulting from the CH4 reduction. Both effects yield negative RFs
Impulse  Response  (FaIR)  model  (Millar  et  al.,  2017);  and  the
(Holmes et al., 2011; Myhre et al., 2011). In the present study, an
CICERO-SCM (Fuglestvedt and Berntsen, 1999; Skeie et al., 2017). The
ensemble of 20 NOx studies is assessed to provide NOx forcing best
6

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
estimates based on a wide range of global atmospheric chemistry
/climate models and a broad range of present-day aviation emission
inventories (details in Appendix D and SD spreadsheet). Results from 6
of the studies were adopted from Holmes et al. (2011).
The study ensemble represents various model methodologies in
calculating and treating both the short-term and the long-term NOx
components. In order to avoid gaps and additional uncertainties, standardized
ERFs were developed that estimated disparate elements (e.g.,
CH4 mediated decreases in SWV and long-term O3). Moreover, most of
the studies were based upon a parameterization of the CH4 response that
assumed a full equilibrium response. In order to calculate the transient
response for a specific year more accurately, a correction factor is
needed (Myhre et al., 2011). Here, the CH4 responses for individual
years were calculated (see Appendix D) using the difference between
two simulations with differing aviation NOx emissions. A number of
transient and equilibrium simulations were conducted with a 2D
chemical-transport model to find that the requirement for a correction
factor is well supported and that the 2018 value is 0.79 (see Transient vs.
equilibrium in Appendix D and Appendix Table D.2). In addition, a    Fig. 4. Results from an ensemble of 18 models from 20 studies for aviation NOx
scaling factor (1.23) is applied to derived CH4 ERF numbers to account    impacts: short-term O3 increases; CH4 reductions, CH4-induced long-term re-
for the effect of shortwave CH4 forcing, following Etminan et al. (2016)    ductions of O3, CH4-induced reductions of stratospheric water vapor (SWV) and
(see Appendix D). The existence and nature of correlations between the    Net NOx. Each data point represents a value of RF per unit emission (mW m 2
NOx RF components were also explored (see Correlations in Appendix D    (Tg N yr 1) 1) as normalized from a published study (see SD). CH4-induced O3
and Appendix Fig. D.1) since the degree of correlation between    and SWV are calculated using standardized methodology (see text for details).
short-term O3 and CH4 terms was a source of uncertainty in the calcu-    Note that the displayed values do not include correction factors to account for
the non-steady-state CH4 responses to NOx emissions and the new CH4 RF
lation of the net-NOx forcing in Lee et al. (2009). The work of Holmes
parameterization. These adjustments are applied in forming the best estimates
et al. (2011) supports the prior assumption of correlation, which is
as discussed in Appendix D.
greatly expanded here. Regardless of inter-model differences, significant
correlations are observed; for example, a significant negative correlation
4.3. Water vapor emissions
(p = 0.7) exists between the short-term and the long-term NOx RF
components.
A large fraction of annual aircraft emissions from the global fleet
The normalized sensitivity results for net NOx in units of mW m 2 (Tg
occurs in the stratosphere, primarily in the northern hemisphere (Forster
(N) yr 1) 1 for the individual modeling studies are shown in Fig. 4 along
et al., 2003). The accumulation of water vapor emissions perturbs the
with statistical parameters (see Ensemble values in Appendix D). Given
low background humidity in the lower stratosphere and changes the
the diversity of studies conducted over nearly two decades, the standard
water vapor radiative balance. Calculating the water vapor RF is
deviations of the distributions are reasonably small. In contrast, the sign
complicated by the sensitivity to the vertical and horizontal distribution
of the net-NOx RF obtained from summing over the 4 component values
of emissions, seasonal changes in tropopause heights, and short stratovaries
from positive to negative. The spread in NOx RF values is caused
spheric residence times. Some earlier studies do not include the water
by various factors (e.g., emissions inventories, experimental design or
vapor effect.
inter-model differences) and is particularly sensitive to the NOx distri-
The water vapor effects were explored in detail (see SD) using results
bution in the model background troposphere (Holmes et al., 2011). The
from nine studies: IPCC (1999), Marquart et al., (2001), Gauss et al.
NOx efficacies are 1.37 for the short-term ozone increases and 1.18 for
(2003), Ponater et al. (2006), Fromming et al. (2012), Wilcox et al.
methane decreases (Ponater et al., 2006). The efficacies do not equal the
(2012), Lim et al. (2015), Pitari et al. (2015) and Brasseur et al. (2016).
ERF/RF ratios, in general (Ponater et al., 2020; Appendix C); nonethe-
The reported RFs from these studies vary from 0.4 mW m 2 (Wilcox
less, in the present study, we assume the efficacies and the ERF/RF ratios
et al., 2012) through 1.5 mW m 2 (Fromming et al., 2012; Lim et al.,
are equal, in the absence of better information. The factor of 1.18 was
2015) to 3.0 mW m 2 (IPCC, 1999). The differences are attributed to the
similarly adopted for the CH4-mediated decreases in long-term ozone
different transport models used, with some contribution from the
and SWV. It is noted that these ratios are from one study and that, in
different meteorologies in different studies. Normalizing to the same
general, the ratio of ERF to RF for CH4 and tropospheric O3 are currently
emissions and averaging these reported estimates yields a water vapor
the subject of some debate (Smith et al., 2018; Xie et al., 2016;
sensitivity of 0.0052  0.0026 mW m 2 (Tg (H2O) yr 1) 1. Scaling this
Richardson et al., 2019). Given the strength of the net effect of the ERF
value linearly to emissions of 382 Tg H2O yields an ERF best estimate of
adjustment on the net NOx forcing (more than doubling over its
2.0 (0.8, 3.2) mW m 2 for 2018, which is well within the uncertainty
stratosphere-adjusted RF), these ratios warrant further study.
range of the 2005 Lee et al. (2009) value of 2.8 (0.39, 20.3) mW m 2.
The net-NOx ERF sensitivity of 5.5  8.1 mW m 2 (Tg (N) yr 1) 1
The ERF/RF ratio for stratospheric water increases is assumed to be
yields a 2018 best estimate of 17.5 (0.6, 28.5) mW m 2. This best estiunity.
We have greater confidence in the new estimate and its smaller
mate includes the correction factor for non-steady state conditions as
uncertainty since it is based on detailed physical studies, rather than a
well as the revised formulation of CH4 RF (Appendix D).
scaling of the earlier IPCC (1999) estimate. The new best estimate is also
Other potential short-term effects from NOx emissions involve the
in good agreement with the earlier results of Gauss et al. (2003) and
direct formation of nitrate aerosol and indirect enhancement of sulfate
Ponater et al. (2006), after scaling their results to account for emissions
aerosol. These effects, addressed in a few modelling studies, are assodifferences.
ciated with large uncertainties (Righi et al., 2013; Pitari et al., 2017;
Unger, 2011). The effects of NOx on aerosol abundances are not further
considered here owing to the limited number of studies and the large    4.4. Contrail cirrus
associated uncertainties.
The aviation fleet increases global cloudiness through the formation
of persistent contrails when the ambient atmosphere is supersaturated
7

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
with respect to ice (IPCC, 1999). Contrail cirrus, consisting of linear    groups: those connected with the upper tropospheric water budget and
contrails and the cirrus cloudiness arising from them, have cooling    the contrail cirrus scheme itself, and those associated with the change in
(short-wave) and warming (long-wave) effects, with the effect at night    radiative transfer due to the presence of contrail cirrus. We considered
being exclusively warming. In past assessments (e.g., IPCC, 1999; Lee    uncertainty in upper tropospheric ice-supersaturation frequencies and
et al., 2009), a best estimate was only available for the RF of linear    their simulation in global models and the uncertainty of ice-crystal
persistent contrails, in part because of the difficulty of quantifying the    numbers due to uncertainty in soot-number emissions, ice nucleation
cloudiness contribution of aging and spreading contrails (Minnis et al.,    within the plume, and loss processes in the contrail's vortex phase.
2013). The ERF of contrail cirrus was estimated for 2011 as 50 (20, 150)    Finally, an important uncertainty comes from the adjustment of natural
mW m 2 by Boucher et al. (2013). Results of a recent assessment of    clouds (Burkhardt and Karcher, 2011). There is also a small uncertainty
contrail cirrus and other aviation effects are included here, although the    associated with the contrail cirrus life cycle, which affects the difference
study did not propose new best estimates (Brasseur et al., 2016).          in nighttime and daytime contrail cirrus cover (Stuber et al., 2006)
A persistent contrail requires ice-supersaturated conditions along the    based on work analyzing the diurnal cycle (Chen and Gettelman, 2013;
flight track. Contrail cirrus life cycles are dependent on the temporal and    Newinger and Burkhardt, 2012).
spatial scales of the ice supersaturated areas, which are highly variable       Uncertainty connected with the radiative response to contrail cirrus
in the troposphere and tropopause region (e.g., Lamquin et al., 2012;    is largely due to the differences in the radiation schemes across climate
Irvine et al., 2013; Bier et al., 2017). Estimating the impact of contrail    models and the approximations made therein (Myhre et al., 2009;
cirrus on upper tropospheric cloudiness requires the simulation of    Gounou and Hogan, 2007); the background cloud field and its vertical
complex microphysical processes, contrail spreading, overlap with nat-    overlap with contrail cirrus; and assumptions about the homogeneity of
ural clouds, radiative transfer, and the interaction with background    the contrail cirrus field. Furthermore, the presence of very small ice
cloudiness (Burkhardt et al., 2010). We present new best estimates based    crystals (<5 m) (Bock and Burkhardt, 2016) and unknown ice-crystal
on the results of global climate models employing process-based contrail    habits (Markowicz and Witek, 2011) add to the uncertainty.
cirrus parameterizations (Appendix E). Due to the small number of in-       Our best estimate of the contrail cirrus uncertainty does not include
dependent estimates the uncertainty must be estimated from the sensi-    the impact of contrails forming within natural clouds, which was
tivities of the respective processes and the uncertainty in the underlying    recently shown to be observable from space (Tesche et al., 2016), or the
parameters and fields.                                                change in radiative transfer due to soot cores in contrail cirrus ice
Here, we consider RF and ERF estimates from global climate models    crystals (Liou et al., 2013), which decreases the albedo at solar wave-
(Burkhardt and Karcher, 2011; Bock and Burkhardt, 2016; Chen and    lengths and increases the top of atmosphere net RF. Both effects are very
Gettelman, 2013; Schumann et al., 2015; Bickel et al., 2020) to ulti-    likely to lead on average to an increase in contrail cirrus RF, causing our
mately produce an ERF best estimate. For the present study, the Chen    best estimate to be conservative. The estimated uncertainty relates to the
and Gettelman study was repeated with lower prescribed initial    average contrail cirrus RF. In specific synoptic situations, uncertainties
ice-crystal diameters, thereby bringing assumptions in line with mea-    may be much larger and correlated with each other.
surements e.g., Schumann et al. (2017). Since the RF estimates differ       In contrast to other aviation forcing terms, the average ERF/RF ratio
regarding the air traffic inventory, the measure of air traffic distance (i.    for contrail cirrus is estimated to be 0.42, much less than unity. The
e., taking only surface-projected or overall flight distances into account)    associated uncertainty is thought to be very large and dependent on
and the temporal resolution of the air traffic data, the estimates were    prevailing aviation traffic and its geographic distribution. The low ERF/
homogenized using known sensitivities (Bock and Burkhardt, 2016) (see    RF value is largely due to the reduction in natural cloudiness caused by
Appendix E). Furthermore, the estimates were corrected to account for    increased contrail cirrus similar to the reduction in natural cirrus
the underestimation of the contrail cirrus RF, as calculated by climate    cloudiness as reported by Burkhardt and Karcher (2011). The ERF/RF
models that use frequency bands, relative to more detailed line-by-line    value is the average of three global climate model studies: two that
radiative transfer calculations (Myhre et al., 2009). The Chen and Get-    estimated climate efficacies of 31% and 59% (Ponater et al., 2005; Rap
telman (2013) study is closer to a calculation of an ERF, since it accounts    et al., 2010) and a third that gave a direct estimate of the ERF of contrail
for fast feedbacks on natural clouds, which Bickel et al. (2020) show in    cirrus that is 35% of the corresponding RF (Bickel et al., 2020). These
their model explains most of the differences between an ERF and an RF    studies conclude that efficacies equal to that of CO2 overstate the role of
calculation. Bickel et al. (2020) presents an explicit calculation of the    cirrus changes due to aviation on global mean surface temperatures. The
contrail cirrus ERF and uses the same basic model formulation of Bock    average ERF/RF ratio was applied to the homogenized estimates of RF,
and Burkhardt, so the ERF calculation was not used here directly but    while the RF of Chen and Gettelman (2013) was interpreted as an ERF
rather the estimation of the ERF/RF ratio was used.                      (see above). Weighting each study equally, the resulting ERF for contrail
The RF best estimate for 2011 was calculated here for comparison to    cirrus is 57 (17, 98) mW m 2 for 2018. It is important to note that the
the most recent IPCC estimate (Boucher et al., 2013). With each study    uncertainty does not include any contribution coming from the ERF/RF
weighted equally, the resulting 2011 RF best estimate for contrail cirrus    estimate. Despite the large ERF/RF adjustment, this ERF term is the
(excluding any adjustments) is approximately 86 (25, 146) mW m 2 (see    largest for global aviation in 2018 and is comparable in magnitude to the
Table 3). The IPCC best estimate of 50 (20, 150) mW m 2 (including the    CO2 term in the normalized results for 2000 to 2018 (Fig. 6). While
natural cloud feedback) was derived from scaling and averaging two    comparable in magnitude, these ERFs have different implications for
studies. IPCC assigned a large uncertainty and low confidence to reflect    future climate change (Section 6).
important aspects with incomplete knowledge (e.g., spreading rate,
optical depth, and radiative transfer). The RF best estimate derived here
for 2018 is 111 (33, 189) mW m 2. The uncertainties in the present    4.5. Aerosol-radiation interaction
study are reduced due to the development of process-based approaches
simulating contrail cirrus in recent years. The uncertainty in the new RF       Aircraft engines directly emit soot, defined as mixture of BC and OC,
estimate, excluding the uncertainty in the ERF/RF scaling of individual    and precursors for sulfate (SO2 4 ) and nitrate (NO 3 ) aerosol along flight
RF values, is 70%, a value substantially lower than the factor of three    tracks. Soot aerosol is formed from the condensation of unburnt arostated
in IPCC.                                                      matic compounds in the combustor (e.g. Ebbinghaus and Wiesen, 2001)
The 70% uncertainty was derived differently than for the NOx    and sulfate aerosol from the oxidation of sulfur in the fuel (Dstan 91-91,
forcing due to the smaller number of available studies. Instead, the    2015). Most of the sulfur is emitted as SO2, whilst a small fraction (~3%)
uncertainty was derived from the combined uncertainties associated    is emitted as oxidized H2SO4 (Petzold et al., 2005). Most of the sulfate
with the processes involved (see Appendix E). The processes fall into two    aerosol is produced after emission from sulfur precursor compounds by
8

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Fig. 5. Summary of RF estimates for aerosol-cloud
interactions for aviation aerosol as calculated in the
SD spreadsheet for a variety of published results
normalized to 2018 air traffic and 600 ppm fuel sulfur.
The results are shown for soot; total particulate
organic matter (POM), sulfate and ammonia (NH3);
and sulfate aerosol from the indicated studies. The
color shading gradient in the symbols indicates
increasing positive or negative magnitudes. No best
estimate was derived in the present study for any
aerosol-cloud effect due to the large uncertainties. In
previous studies, the estimates for the soot aerosolcloud
effect are associated with particularly large
uncertainty in magnitude and uncertainty in the sign
of the effect (Penner et al., 2009, 2018; Zhou and
Penner, 2014). As part of the present study, an author
(JEP) re-evaluated these earlier studies and concluded
that the Penner et al. (2018) results supersede the
earlier Penner et al. (2009) and Zhou and Penner
(2014) results because of assumptions regarding up-
draft velocities during cloud formation. In addition, a
bounding sensitivity case in which all aviation soot
acts as an IN in Penner et al. (2018) is not included
here.








oxidation in the ambient atmosphere. Both aerosol types create RFs from    nucleate on aerosol particles. Thus, aerosol-cloud interactions involving
aerosol-radiation interactions: soot absorbs short-wave radiation lead-    aviation aerosol potentially result in an ERF. Aviation soot and sulfate
ing to net warming and sulfate aerosol scatters incoming short-wave    particles are the predominant primary and secondary aerosol from
radiation leading to net cooling (IPCC, 1999). As figures of merit, year    aircraft. The uncertainties in evaluating the aerosol-cloud interactions of
2000 global aviation emissions increase aerosol mass for both soot and    aviation soot and sulfate preclude best estimates of ERF contributions.
sulfate by a few percent and aerosol number by 1030% near air traffic    Given the potential importance of these ERF terms, placeholders are
flight corridors in the northern extratropics (Righi et al., 2013).           included in Fig. 3. Furthermore, to promote progress towards future best
Past calculations of aerosol-radiation RF values using a variety of    estimates, the results of relevant modeling studies were compiled and
global aerosol models have yielded values of a few mW m 2 and with    normalized to global aviation fuel usages in 2005, 2011, 2018, to a soot
large uncertainties (e.g., Righi et al., 2013; Gettelman and Chen, 2013;    emission index, and to a fuel S content of 600 pm (except in the cases of
Lee et al., 2009). In the present study, 10 estimates across 8 models were    low fuel-S content tests) (see Fig. 5 and spreadsheet). As noted in the
used to evaluate soot and sulfate aerosol normalized RFs (IPCC, 1999;    caption of Fig. 5, some earlier wide-ranging values for the soot aerosol-
Sausen et al., 2005; Fuglestvedt et al., 2008; Balkanski et al., 2010;    cloud interaction have been superseded by a more recent study (Penner
Gettelmann and Chen, 2013; Unger et al., 2013; Pitari et al., 2015;    et al., 2018).
Brasseur et al., 2016) (see SD spreadsheet). Averaging the normalized
values yields a 2018 best estimate of the soot aerosol-radiation RF of 0.9    4.6.1. Sulfate aerosol
(0.1, 4.0) mW m 2 for 0.0093 Tg soot emitted. The corresponding best       Aviation sulfate aerosol primarily affects liquid clouds in the backestimate
for sulfate aerosol is 7.4 ( 19, 3) mW m 2 for 0.37 Tg SO2    ground atmosphere. Sulfate aerosol is very efficient as a cloud condenemitted.
The uncertainties are derived from the standard deviation of    sation nuclei (CCN) for liquid clouds, and for promoting homogeneous
the model values. The ERF/RF ratios for soot and sulfate are assumed to    freezing of solution particles at cold temperatures, thus nucleating ice
be unity in the absence of any estimates of this ratio.                    clouds. Two integrated model simulations (Kapadia et al., 2016; Get-
telman and Chen, 2013) found large impacts on liquid clouds from
4.6. Aerosol-cloud interaction                                          aviation sulfate aerosol that is transported to liquid clouds at lower altitudes
over oceans, which have low albedo. The reported RF values in
Aerosol-cloud interactions are those processes by which aerosols    these studies, when scaled appropriately, are 37 to 76 mW m 2 in
influence cloud formation. For example, cloud droplets and ice crystals    2018, excluding a low fuel-sulfur case. Note that the study of Righi et al.
9

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Fig. 6. Timeseries of calculated ERF values and con-
fidence intervals for annual aviation forcing terms
from 2000 to 2018. The top panel shows all ERF terms
and the bottom panel shows only the NOx terms and
net NOx ERF. All values are available in the SD
spreadsheet, in Tables 2 and 3, and in Fig. 3 for 2018
values. The net values are not arithmetic sums of the
annual values because the net ERF, as shown in Fig. 3
for 2018, requires a Monte Carlo analysis that properly
includes uncertainty distributions and correlations
(see text).







(2013) that yields an RF of 213 mW m2 in 2018 includes sulfate    calculated in some studies for moderate ice-nucleating efficiencies
aerosol-cloud interactions but cannot be directly compared with Kapa-    (Pitari et al., 2015; Gettelman and Chen, 2013).
dia et al. (2016) and Gettelman and Chen (2013), since the former treats       In sensitivity tests, if soot processed within contrails is assumed to be
the combined effects of sulfate, nitrate and particulate organic matter    an efficient IN particle, then the RF may be negative by up to 330 mW
(POM) rather than isolating the effects of sulfate as done in the latter    m 2 due to reductions in ice crystal number in regions dominated by
studies. While these RF estimates do not support a best estimate at    homogeneous freezing (Penner et al., 2018; see Fig. 5). The RF could be
present, they do suggest that the sign of the sulfate aerosol-cloud effect    significantly smaller (less negative) if additional ice-forming particles,
on low-level clouds is likely to be negative (i.e., a cooling), similar to the    such as secondary organic aerosol (SOA), are already present in the
ERF for the aerosol-cloud interactions of other anthropogenic sources of    background atmosphere (Penner et al., 2018; Gettelman and Chen,
sulfate aerosol (IPCC, 2013).                                          2013). In addition, increases in ice crystal numbers occur when the
Sulfate  aerosol-cloud  interaction  forcing  estimates  are  highly    background atmosphere has much lower sulfate or haze-forming aerosol
dependent on the sensitivity (or susceptibility) of the cloud radiative    number concentrations and is dominated by heterogeneous freezing,
field to aerosol perturbations, which is dependent on uncertain model    causing forcings near zero or even positive (Zhou and Penner, 2014).
processes and the model background aerosol state. Clouds that form    Other studies predict decreases in cirrus number for smaller numbers of
with small CCN number concentrations in the background atmosphere    larger soot particles (Hendricks et al., 2011), resulting in a slight
are more sensitive to CCN perturbations. Forcing by these cloud effects    warming (Gettelman and Chen, 2013).
are largely concentrated near flight corridors over oceans because the       A dominant uncertainty for the aerosol-cloud effect from soot is the
high albedo contrast between the ocean surface and clouds increases    IN properties of aviation soot aerosol. Some laboratory studies indicate
forcing sensitivity to CCN perturbations.                                soot particles are not efficient ice nuclei (DeMott et al., 1999), while
A large uncertainty was also reported for the magnitude of the    other studies indicate higher efficiencies (Mohler et al., 2005; Hoose and
aerosol-cloud ERF from all anthropogenic activities, estimated for 2011    Mohler, 2012). The possibility that contrail-processed soot particles
to be 450 ( 1200, 0.0) mW m 2 (Myhre et al., 2013). A more recent    would show enhanced IN activity after sublimation in the background
estimate of the aerosol-cloud RF from all anthropogenic activities has a    atmosphere was addressed in the laboratory (Mahrt et al., 2020). The
68% confidence interval of 650 to 1600 mW m 2 (Bellouin et al.,    effect was limited to large soot particles, suggesting that the impact of
2019). In general, aerosol-cloud interactions contribute the largest un-    aviation soot on cloudiness may be overestimated in previous studies
certainty in calculations of anthropogenic ERF (IPCC, 2013).              that assume soot processed through contrails and not covered by a sulfate
coating is an efficient IN (Penner et al., 2018).
4.6.2. Soot                                                            Another source of uncertainty is soot number concentrations. For
The magnitude and the sign of the global RF from aviation soot ef-    individual engines, the soot number can vary by two orders of magnifects
on background cloudiness remain highly uncertain. The un-    tude (Agarwal et al., 2019). Soot number concentrations from aviation
certainties  center  on  the  difficulties  in  accurately  simulating    vary with the assumed size of the particles emitted as well as the mass
homogeneous and heterogeneous ice nucleation in the background at-    emissions. Soot emissions from aircraft are set as a regulatory parameter
mosphere, variations in the treatment of updraft velocities during cirrus    for the landing/take-off (LTO) cycle by ICAO and are measured in terms
formation, and the lack of knowledge of the ice nucleating (IN) ability of    of mass. Robust conversion factors from mass to number have recently
aviation soot particles during their atmospheric lifetime (Zhou and    been developed for the ICAO-LTO cycle (Agarwal et al., 2019) but have
Penner, 2014; Penner et al., 2018).                                    not yet been made for cruise, although other methodologies exist (Teoh
Two studies find moderate effects of soot aerosol on ice clouds,    et al., 2019).
depending on the ice nucleating efficiency and the size distribution. RF
values of about 1113 mW m 2 (normalized to 2018 emissions) are
10

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
5. Calculated net aviation ERF and RF values                       contributions of aviation to the net ERF in 2011 are 3.5% (4.0, 3.4%)
and 1.59% (1.65, 1.56%) for the sum of all terms and the CO2 term
ERF and RF values for the terms associated with global aviation    alone, respectively. The 2005 and 2018 percentages are likely the same
emissions and cloudiness are given in Tables 2 and 3, respectively, for    because the fraction of aviation CO2 emissions of total anthropogenic
the years 2018, 2011, and 2005, along with uncertainties, sensitivities to    CO2 emissions has averaged 2.1% (0.15) for the last two decades (see
emissions and the ERF/RF ratio for selected terms. ERF values are shown    Fig. 2). Normalized relative probabilities of CO2 and non-CO2 ERFs for
for all years in Fig. 6. All ERF and RF values are available in the analysis    2018 as derived from the Monte Carlo simulations show that non-CO2
spreadsheet (SD). Through normalization and scaling, all 2000 to 2018    uncertainties are the predominant contribution to the uncertainty in the
values are self-consistent. The sensitivity of each term to emission    aviation net ERF (Fig. 7). IPCC also separately estimated the contrail
magnitudes or flight track distances is derived in the normalization    cirrus term for 2011 as 50 (20, 150) mW m 2 as discussed above, which
process. ERF best estimates and uncertainties (95% confidence limits)    compares well with the updated value of 44.1 (13, 75) mW m 2.
are highlighted for year 2018 in Fig. 3 along with their assessed confi -       The determination of net aviation ERFs and their uncertainties
dence levels. No best estimates are included for sulfate and soot aerosol-    shown in Fig. 3 and accompanying tables required a Monte Carlo
cloud interactions because of the substantial uncertainties noted above.    approach to summing over terms with discrete probability distributions.
However, placeholder spaces are included in both Tables 2 and 3 and    A similar method was employed in Lee et al. (2009). PDFs of each term
Fig. 3 to indicate the potential importance of these terms and to flag the    were constructed from the respective individual studies as normal,
associated knowledge gaps for consideration in future research and    lognormal or discrete distributions (see SD spreadsheet). Monte Carlo
assessment activities. The confidence levels and their justifications    samplings (one million random points) of the individual forcing PDFs
shown in Fig. 3 are obtained by employing the methodology of Mas-    were then used to combine terms to yield net ERFs and the uncertainties
trandrea et al. (2011), which is based on evidence and agreement in    (95% likelihood range) for the sum of all terms and for only non-CO2
accordance with IPCC guidance (Table 4).                              terms (Fig. 7). The forcing terms are generally assumed to be indepen-
In Fig. 3, contrail cirrus formation yields the largest positive    dent (uncorrelated) with the notable exception of the NOx component
(warming) ERF term, followed by CO2 and NOx emissions. For the 1940    terms which have strong paired correlations as shown in Appendix
to 2018 period, the net aviation ERF is +100.9 mW m 2 (595% like-    Fig. D.1. Only the short-term O3 and CH4 terms were included in Lee
lihood range of (55, 145)) with major contributions from contrail cirrus    et al. (2009) and a 100% correlation was assumed, in part, because the
(57.4 mW m 2), CO2 (34.3 mW m 2), and NOx (17.5 mW m 2). The    assumption of uncorrelated effects was deemed less acceptable. A subaerosol
and water vapor terms represent minor contributions. The for-    sequent study showed that these terms are indeed strongly correlated
mation and emission of sulfate aerosol yields the only significant    (R2 = 0.79) (Holmes et al., 2011), similar to the present results in Apnegative
(cooling) term. Non-CO2 terms sum to yield a positive    pendix Fig. D.1. The Holmes et al. (2011) study further concluded that
(warming) ERF that accounts for 66% of the aviation net ERF in 2018    the assumption of 100% correlation in this case would lead to an un-
(66.6 (21, 111) mW m 2). The application of ERF/RF ratios more than    derestimate of uncertainty in the NOx RF. Another correlation of forcing
halves the RF value of contrail cirrus while approximately doubling the    terms not considered here may be the dependence of the soot direct
NOx value. ERF/RF ratios were not included in the Lee et al. (2009)    effect and contrail properties on the soot number index since ice
analysis. Uncertainty distributions (5%, 95%) show that non-CO2 forc-    nucleation at the time of contrail formation depends on the soot number
ing terms contribute about 8 times more than CO2 to the uncertainty in    index (e.g., Karcher, 2018).
the aviation net ERF in 2018. The best estimates of the ERFs from
aviation aerosol-cloud interactions remain undetermined.                6. Emission equivalency metrics
The time series of ERF values for individual terms is shown in Fig. 6
for the 20002018 period. Through normalization and scaling the terms       Using the best estimate ERFs, we calculate updated aviation-specific
are self-consistent over this period. The increase in all of the terms with    Global Warming Potential (GWP) and Global Temperature change Potime
is consistent with the growth of aviation fuel burn and CO2 emis-    tential (GTP) values, presented for 20-, 50-, and 100-year time horizons
sions over the same period (Fig. 2). Note that net ERF values shown for    in Table 5. These metrics assign so-called 'CO2-emission equivalences'
each year are not linear sums over the component terms due to the    for non-CO2 emissions via ratios of time-integrated ERF and changes in
separate probability distributions associated with each component term    future temperatures, respectively. The choice of metric depends upon
in the sum, and instead are calculated with a Monte Carlo sampling    the particular underlying application (Fuglestvedt et al., 2010) such that
method described below.                                             there is no uniquely 'correct' metric or time horizon, and alternative
A comparison of updated RF estimates with Lee et al. (2009) values    metrics are available. GWP and GTP are the most commonly applied
for 2005 is given in Table 3. The large increase in the contrail cirrus RF    metrics and the values calculated here allow a comparison with previous
between 2005 and 2018 results in part because the 2005 value only    estimations (e.g., Lee et al., 2010; Lund et al., 2017). In calculating the
includes linear contrails. In Lee et al. (2009), only an estimate of 2005    GWPs and GTPs, the CO2 IRF from Joos et al. (2013) is used and the
contrail cirrus was provided rather than a best estimate. The present    climate response IRF from Boucher and Reddy (2008) for the GTPs (see
study now includes a process-based model estimate of the contrail cirrus    Appendix F for futher details about the metrics calculations).
term (Section 4.4). The NOx treatment in Lee et al. (2009) did not       GWPs and GTPs for contrail cirrus and for water vapor reported here
include the negative forcing contributions of the long-term O3 decrease    are similar to, albeit slightly smaller than, corresponding results previor
the SWV decrease, the updated treatment of CH4 of Etminan et al.    ously reported, while soot and sulfate numbers are larger in magnitude
(2016), nor an equilibrium-to-transient correction. As a result, the    (positive and negative) than previous estimates (Fuglestvedt et al., 2010;
updated RF values for NOx are approximately a factor of 2 smaller.    Lund et al., 2017). The Fuglestvedt et al. (2010) estimates for soot are
Incorporating all the updated information in the RF calculations of the    based on RF due to soot emissions from all sources, not just aviation,
NOx and contrail cirrus terms yields an approximately 30% increase in    which yields a lower radiative efficiency (i.e., forcing per unit emission)
the net aviation RF for 2005, from 78.0 to 95.2 mW m 2. In the ERF    than in the present study. Also given in Table 5 are CO2-equivalent
evaluation for 2005 the net aviation forcing is reduced from 95.2 to 66.9    aviation emissions, along with ratios of total CO2-equivalent emissions
mW m 2 because the ERF/RF ratios for NOx and contrail cirrus are    to CO2 emissions. Such ratios are sometimes used as 'multipliers' to
different than unity.                                                 illustrate the additional climate impact from aviation non-CO2 terms
In seeking comparison of net aviation ERF with net anthropogenic    over those from CO2 emissions alone. Here, estimated multipliers for
ERF, we note that IPCC (Myhre et al., 2013) provides a value for    2018 range from 1.0 to 4.0 depending on the choice of time horizon and
17502011  of  2290  (1130,  3330)  mW  m 2.  The  percentage    emission metric. This is broadly consistent with what has been reported
11

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Table 4a
Confidence levels for the ERF estimates in Fig. 3.

























*This term has the additional uncertainty of the derivation of an effective radiative forcing from a radiative forcing.
**This term differs from 'Very High' level in IPCC (2013) because additional uncertainties are introduced by the assessment of marginal aviation CO2 emissions and
their resultant concentrations in the atmosphere from simplified carbon cycle models.





12

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Table 4b                                                           that when the C-cycle feedback is consistently accounted for, the
Basis for confidence levels in Table 4a.a                                   non-CO2 emission metrics increase, but less so than initially suggested
by Myhre et al. (2013). They also find that removing the C-cycle feedback
from both numerator and denominator give similar metric values
as including it in both places. Using the CO2 IRF without the C-cycle
feedback provided by Gasser et al., 2017, we calculate a second set of
aviation emission metrics (Table F.1a and Table F.1b), showing that the
changes to the GWP100 and GTP100 values from those given in Table 5
are rather small.
In response to the challenges related to comparing short-lived and
long-lived forcing components, a number of new 'flow-based' methods
have been introduced representing both short-lived and long-lived
climate forcers explicitly as 'warming-equivalent' emissions that have
approximately the same impact on the global average surface temperature
over multi-decade to century timescales (Lauder et al., 2012; Allen
a The basis for the confidence level is given as a combination of evidence    et al., 2016, 2018; Cain et al., 2019; Collins et al., 2019). A simple
(limited, medium, robust) and agreement (low, medium and high) based on    version of these methods, known as GWP*, defines the average annual
guidance given by Mastrandrea et al. (2011).                               rate of CO2-warming-equivalent emissions (E*CO2e) over a period of t
years arising from a particular component of RF or ERF by (Cain et al.,
2019):
/
E*CO2e = [(1  )H / AGWPH] F   t + [ / AGWPH] F,            (1)
where F is the ERF change and F the average ERF arising from that
component over that period, AGWPH is the Absolute GWP of CO2 (Wm 2
kg 1 year) over time-horizon H and  is a small coefficient depending on
the previous history of that RF component. Eqn (1) gives the rate of CO2
emission that would, alone, create the same rate of global temperature
increase as the combined effect of aviation climate forcings. For historically
small and/or rapidly changing RF components,  may be
neglected, and hence to a good approximation, total CO2-warmingequivalent
emissions over this period (tE*CO2e) are approximated by an
increase in forcing, F, multiplied by H / AGWPH (see Appendix F ),
which is about 1000 GtCO2 per W/m2 for H in the range 20100 years
(Myhre et al., 2013; IPCC, 2018, Figure SPM.1, caption). This result
follows from the definition of AGWP: since all GWP calculations assume
a linearization, the AGWPH is equivalent to the forcing change resulting
from the emission of H tonnes of CO2 spread over H years (Shine et al.,
2005), so AGWPH/H is the forcing change per tonne of CO2. Under the
historical profile of increasing global annual aviation-related emissions
and associated ERFs, CO2-warming-equivalent emissions based on
GWP* indicate that aviation emissions are currently warming the
climate around three times faster than that associated with aviation CO2
Fig. 7. Probability distribution functions (PDFs) for aviation ERFs in 2018    emissions alone (Table 5).
based on the results in Fig. 3 and Table 2. PDFs are shown for separately for       It is important to note that, unlike the conventional GWP and GTP
CO2, the sum of non-CO2 terms, and the net aviation ERF. Since the area of each    metrics given in Table 5, the ratio between total CO2-warming-equiva-
distribution is normalized to the same value, relative probabilities can be    lent emissions from all forcing agents and those from CO2 alone will
intercompared. Uncertainties are expressed by a distribution about the best-    change substantially if future aviation emissions deviate from their
estimate value that is normal for CO2 and contrail cirrus, and lognormal for
current growth trajectory (calculated here over the period 20002018).
all other components. A one-million-point Monte Carlo simulation run was used
If annual global aviation emissions were to stabilize, this ratio declines
to calculate all PDFs.
towards unity, as F/t would decline to zero. This does not indicate,
however, that the non-CO2 effects do not have a warming affect. This
and used previously (Lee et al., 2010). The broad range emphasizes the
human-induced  warming  still  represents  a  mitigation  potential.
challenges associated with developing comparisons of emission equiv-
Warming-equivalent emissions capture the fact that constant emission of
alences for short- and long-lived climate forcers within a common
short-lived climate forcers maintain an approximately constant level of
framework and how such considerations strongly depend on the chosen
warming, whilst constant emissions of long-lived climate forcers, such as
perspective.
CO2, continue to accumulate in the atmosphere resulting in a constantly
One of the significant uncertainties in calculating GWPs and GTPs is
increasing level of associated warming. Hence warming-equivalent
the treatment of climate-carbon (C-cycle) feedbacks in the modeling
emissions show that the widely-used assumption of a constant 'multiframework.
The efficiency of carbon sinks reduces with increasing
plier', assuming that net warming due to aviation is a constant ratio of
warming (Ciais et al., 2013) and this climate feedback is implicitly
warming due to aviation CO2 emissions alone, only applies in a situation
included in the Absolute GWP of CO2 through the IRF used (Joos et al.,
in which aviation emissions are rising exponentially such that the rate of
2013). However, Myhre et al. (2013) highlighted that this introduces an
change of non-CO2 RF is approximately proportional to the rate of CO2
inconsistency since the numerators for the GWP and GTP do not include
emissions (assuming non-CO2 RF is proportional to CO2 emissions, and
such a climate carbon feedback. One of the studies that have proposed
noting that the rate of change any quantity is proportional to that
ways of addressing this inconsistency is Gasser et al. (2017). They show

13

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Table 5
Emission metrics and corresponding CO2-equivalent emissions for the ERF components of 2018 aviation emissions and cloudiness.
Metrics
ERF term                           GWP20             GWP50             GWP100             GTP20             GTP50             GTP100
CO2                              1                1                1                 1                1                1
Contrail cirrus (Tg CO2 basis)             2.32               1.09               0.63               0.67              0.11              0.09
Contrail cirrus (km basis)                39                18                11                 11               1.8               1.5
Net NOx                            619               205               114                 222              69              13
Aerosol-radiation
Soot emissions                      4288              2018              1166               1245              195              161
SO2 emissions                       832               392               226               241              38               31
Water vapor emissions                  0.22               0.10               0.06               0.07              0.01              0.008
CO2-eq emissions (Tg CO2 yr 1) for 2018
ERF term                         GWP20         GWP50         GWP100         GTP20         GTP50         GTP100         GWP*100 (E*CO2e)
CO2                            1034          1034          1034           1034          1034          1034          1034
Contrail cirrus (Tg CO2 basis)           2399          1129          652            695           109           90            1834
Contrail cirrus (km basis)              2395          1127          651            694           109           90            1834
Net NOx                          887           293           163             318          99          19            339
Aerosol-radiation
Soot emissions                    40            19            11             12           2            2             20
SO2 emissions                     310           146           84            90           14           12            158
Water vapor emissions              83            39            23             27           4            3             42
Total CO2-eq (using km basis)         4128          2366          1797           1358          1035          1135          3111
Total CO2-eq/CO2                 4.0            2.3            1.7            1.3           1.0           1.1            3.0

quantity only when both are growing exponentially). In contrast, under    halt anthropogenic global warming on multi-decadal time scales." (IPCC,
a future hypothetical trajectory of decreasing aviation emissions, this    2018, bullet A2.2, SPM). Crucially, both conditions would need to be
GWP* based multiplier could fall below unity, as a steadily falling rate of    met to halt global warming. Hence, to halt aviation's contribution to
emission of (positive) short-lived climate forcers has the same effect on    global warming, the aviation sector would need to achieve net-zero CO2
global temperature as active removal of CO2 from the atmosphere. The    emissions and declining non-CO2 radiative forcing (unless balanced by
GWP* based 'multiplier' calculated here (which depends on the ratio of    net negative emissions from another sector): neither condition is suffi -
the increase in net aviation warming to the increase in warming due to    cient alone. Some combination of reductions in CO2 emissions and
aviation CO2 emissions alone over the recent past), should not be    non-CO2 forcings might halt further warming temporarily, but only for a
applied to future scenarios that deviate substantially from the current    few years: it would not be possible to offset continued warming from
trend of increasing aviation-related emissions. The broad range of values    CO2  by  varying  non-CO2  radiative  forcing,  or vice versa,  over
for a 'multiplier' presented here is an illustration of the limitations of    multi-decade timescales.
using a constant multiplier in the assessment of climate impacts of       That aviation's non-CO2 forcings are not included in global climate
aviation, and a reminder that the choice of metric for such a multipler    policy has resulted in studies as to whether they could be incorporated
involves subjective choices.                                           into existing policies, such as the European Emissions Trading Scheme,
using an appropriate overall emissions 'multiplier'; however, scientific
7. Aviation CO2 vs non-CO2 forcings                                uncertainty has so far precluded this (Faber et al., 2008). In addition, as
noted above, the multiplier is highly dependent on the future emissions
Since IPCC (1999), the comparison of aviation CO2 RF with the    scenario (Section 6). Alternatively, proposals have been made to reduce
non-CO2 RFs has been a major scientific topic, as well as a discussion    aviation's non-CO2 forcings by, for example, avoiding contrail formation
point amongst policy makers and civil society (ICAO, 2019). Aviation as    by re-routing aircraft (Matthes et al., 2017), or optimizing flight times to
a sector is not unique in having significant non-CO2 forcings; the same is    avoid the more positive (warming) fractional forcings (e.g., by avoiding
true of agriculture with significant CH4 and N2O emissions, or maritime    night flights, Stuber et al., 2006). There is a developing body of literashipping
with net-negative current-day RF despite CO2 emissions of a    ture on this topic (e.g., Newinger and Burkhardt, 2012; Yin et al., 2018).
similar magnitude to those from aviation (Fuglesvedt et al., 2009).    Similarly, studies have assessed whether changes in cruise altitudes
However, unlike direct emissions of the greenhouse gases N2O and CH4    could mitigate NOx impacts (e.g. Fromming et al., 2012). The potential
from the agricultural sector, aviation non-CO2 forcings are not covered    impacts of changes in technology have also been examined to reduce the
by the former Kyoto Protocol. It is unclear whether future developments    non-CO2 forcings such as lowering the emission index for NOx (Freeman
of the Paris Agreement or ICAO negotiations to mitigate climate change,    et al., 2018) or soot particle number emissions (Moore et al., 2017) to
in general, will include short-lived indirect greenhouse gases like NOx    reduce net NOx and contrail cirrus forcings, respectively (Burkhardt
and CO, aerosol-cloud effects, or other aviation non-CO2 effects. Avia-    et al., 2018).
tion is not mentioned explicitly in the text of the Paris Agreement, but       Avoidance of contrail formation through re-routing can incur a fuel
according to its Article 4, total global greenhouse-gas emissions need to    penalty and therefore additional CO2 emissions during a flight, and
be reduced rapidly to achieve a balance between anthropogenic emis-    changes in combustor technology to minimize NOx generally increases
sions by sources and removals by sinks of greenhouse gases in the second    marginal fuel burn and CO2 emission. Both methods invoke the usage of
half of this century.                                                  climate metrics such as those calculated and presented in Section 6 to
The IPCC concludes: "Reaching and sustaining net-zero global anthro-    evaluate whether there is a net climate benefit or disbenefit over a
pogenic CO2 emissions and declining net non-CO2 radiative forcing would    defined period. In examining such mitigation scenarios involving

14




D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
tradeoffs (e.g. Teoh et al., 2020), the perceived success or otherwise of    Skowron: Investigation, Methodology, Writing - review & editing, Data
the outcome will be a function of the user's choice of metric and time    curation, Formal analysis, Software. M.R. Allen: Writing - review &
horizon. A limitation noted for the GWP is that it has an 'artificial    editing, Investigation, Methodology, Writing - original draft, Data
memory' over longer time horizons, since the integrated-RF nature of    curation, Formal analysis. U. Burkhardt: Writing - review & editing,
the metric accumulates 'signal' over time that the climate system has    Investigation, Methodology, Writing - original draft, Data curation,
'forgotten' (Fuglestvedt et al., 2010). The GTP, being an 'end point'    Formal analysis. Q. Chen: Writing - review & editing, Investigation,
metric that captures the temperature response, overcomes this limita-    Methodology, Writing - original draft, Data curation, Formal analysis. S.
tion of the GWP but is not yet in usage within current climate policy.      J. Doherty: Writing - review & editing, Investigation, Methodology,
Changes to aviation operations or technology that result in a    Writing - original draft, Data curation, Formal analysis. S. Freeman:
reduction of a non-CO2 forcing with the added consequence of increased    Writing - review & editing, Investigation, Methodology, Writing - orig-
CO2 emissions can result in net reductions of forcing on short timescales    inal draft, Data curation, Formal analysis. P.M. Forster: Writing - rewhile
increasing the net forcing on longer timescales (e.g., Freeman    view & editing, Investigation, Methodology, Writing - original draft,
et al., 2018). In a case study of contrail avoidance through routing    Data curation, Formal analysis. J. Fuglestvedt: Writing - review &
changes, Teoh et al. (2019) found that the resultant small increase in    editing, Investigation, Methodology, Writing - original draft, Data
CO2 emissions still reduces the net forcing over a timescale of 100 years.    curation, Formal analysis. A. Gettelman: Writing - review & editing,
In such 'tradeoff cases' the balance between non-CO2 and CO2 forcings    Investigation, Methodology, Writing - original draft, Data curation,
have to be weighted carefully, since CO2 accumulates in the atmosphere    Formal analysis. R.R. De Leon: Writing - review & editing, Investigaand
a fraction has millennial timescales (Archer and Brovkin, 2008;    tion, Methodology, Writing - original draft, Data curation, Formal
IPCC, 2007). Prior to the COVID-19 pandemic, global aviation traffic    analysis. L.L. Lim: Writing - review & editing, Investigation, Methodand
emissions were projected to grow to 2050 (Fleming and de Lepinay,    ology, Writing - original draft, Data curation, Formal analysis. M.T.
2019). As the COVID-19 pandemic diminishes, aviation traffic is likely    Lund: Writing - review & editing, Investigation, Methodology, Writing -
to recover to meet projected rates on varying timescales (IATA, 2020),    original draft, Data curation, Formal analysis. R.J. Millar: Writing -
with continued growth further increasing CO2 emissions. Thus, reducing    review & editing, Investigation, Methodology, Writing - original draft,
CO2 aviation emissions will remain a continued focus in reducing future    Data curation, Formal analysis. B. Owen: Writing - review & editing,
anthropogenic climate change, along with aviation non-CO2 forcings.    Investigation, Methodology, Writing - original draft, Data curation,
The latter increase the current-day impact on global average tempera-    Formal analysis. J.E. Penner: Writing - review & editing, Investigation,
tures by a factor of around 3 (using GWP*) above that due to CO2 alone.    Methodology, Writing - original draft, Data curation, Formal analysis. G.
Pitari: Writing - review & editing, Investigation, Methodology, Writing -
Funding                                                     original draft, Data curation, Formal analysis. M.J. Prather: Writing -
review & editing, Investigation, Methodology, Writing - original draft,
DSL, AS, RRdL, LL, BO acknowledge support from the UK Depart-    Data curation, Formal analysis. R. Sausen: Writing - review & editing,
ment for Transport. PMF acknowledges support of the European Union's    Investigation, Methodology, Writing - original draft, Data curation,
Horizon 2020 Research and Innovation Programme under grant agree-    Formal analysis. L.J. Wilcox: Writing - review & editing, Investigation,
ment number 820829 (CONSTRAIN) by the UK National Environment    Methodology, Writing - original draft, Data curation, Formal analysis.
Research Council (NERC) SMURPHS project (NE/N006038/1). MRA
acknowledges support from the EU H2020 grant agreement number
821205 (FORCeS) and the Oxford Martin Programme on Climate Pol-    Declaration of competing interest
lutants. MTL and JSF acknowledges support from the Norwegian
Research Council (RCN) grant number 300718 (AVIATE), for which DSL       The authors declare that they have no known competing financial
and RS have a collaboration agreement. JEP acknowledges support from    interests or personal relationships that could have appeared to influence
the National Science Foundation (NSF 1540954).                        the work reported in this paper.
CRediT authorship contribution statement                          Acknowledgements
D.S. Lee: Investigation, Methodology, Writing - review & editing,       We gratefully acknowledge discussions with many colleagues during
Data curation, Formal analysis, Project administration, Supervision. D.    the preparation of this paper, in particular Andreas Bier and Bernd
W. Fahey: Investigation, Methodology, Writing - review & editing, Data    Karcher. We acknowledge help with graphical displays from Beth Tully
curation, Formal analysis, Project administration, Supervision. A.    (Fig. 1) and Chelsea R. Thompson (Figs. 57).

Supplementary data
Supplementary data to this article is a spreadsheet that can be found online at: https://doi.org/10.1016/j.atmosenv.2020.117834.
Appendices.
A. Trends in aviation CO2 emissions
Global aviation CO2 emissions for 19401970 were taken from Sausen and Schumann (2000) and for the years 19712016 were calculated from
International Energy Agency (IEA) data on usage of JET-A and aviation gasoline, largely from annual 'Oil Information' digests (e.g., https://webstore.
iea.org/oil-information-2019). The regional data are from the same source but accessed online from the IEA Oil Information (19602017) held at the
UK Data Service (IEA, 2019). Note that these data are proprietary and must be purchased from IEA. Data were unavailable for 2017 and 2018, so
incremental annual percentage increases in global aviation fuel usage and, therefore CO2 emissions, for those years were taken from reports of the
International Air Transport Association (IATA, 2019). Some uncertainties exist from the annual fuel estimations and to a much smaller extent, the
emission factors. The IEA does not give uncertainties for annual kerosene fuel sales or usage. Sausen and Schumann (2000), from which the 1940 to

15

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
1970 data are based here, estimated that the uncertainty in cumulative fuel consumption from 1940 to 1995 (their dataset) is 20%. There is a known
discrepancy of IEA estimates of aviation fuel usage being greater by about 10% than that derived from bottom-up global civil aviation inventories.
Actual fuel usage is likely to be somewhere between the two estimates: aviation emissions inventories are known to be incomplete, with only
scheduled traffic being available from some air traffic regions, and fuel usage potentially being underestimated from flight routing and cruise altitudes;
IEA data on the other hand includes military aviation fuel (not included in civil aviation inventories) and a small fraction of kerosene not used in
aviation, but sold for that purpose (Lee et al., 2009). The CO2 emission factors for aviation fuel on the other hand are well determined, and the
uncertainty is likely within 1%.
B. Aviation CO2 radiative forcings

Calculation of CO2 concentrations from emissionsLinClim SCM
The response of CO2 concentrations, C(t), to a CO2 aviation emissions rate, E(t), is modelled using the method described in Hasselmann et al.
(1997) and is expressed as:
t
C(t)=   GC(t  t )E(t )dt                                                                                                (B.1)
t0

where in Eqn (B.1)
5
GC(t)=    je  t/j                                                                                                       (B.2)
j=0

and j in Eqn( B.2) is the e-folding time of mode j and the equilibrium response of mode j to a unit emissions of jj.
The mode parameters used in this study are presented in Sausen and Schumann (2000) and approximate the carbon-cycle model in Maier-Reimer
and Hasselmann (1987). The applicability of these parameters in the context of aviation response was tested in a model intercomparison exercise
(Khodayari et al., 2013). For the time horizon of 5060 years into the future, these were found to compare well with other more sophisticated
carbon-cycle models such as MAGICC 6.0, which is widely used in the IPCC Fourth Assessment Report (IPCC, 2007). Beyond this horizon, aviation CO2
concentrations begin to have an impact on the ocean and biosphere uptake of CO2 and the non-linearities of the system must be accounted for.
Calculation of CO2 concentrations from emissionsCICERO-2 SCM
The CICERO-2 SCM (Fuglestvedt and Berntsen, 1999; Skeie et al., 2017) uses interconnected process-specific IRFs with explicit treatment of air-sea
and air-biosphere exchange of CO2 (Joos et al., 1996; Alfsen and Berntsen, 1999) that forms a nonlinear carbon cycle. The ocean and biosphere IRFs in
CICERO-2 express how the CO2 impulse decays within each reservoir. The CO2 partial pressure in each reservoir is calculated as a function of the
carbon in that reservoir, and the CO2 partial pressure in each reservoir is related to the CO2 partial pressure in the atmosphere by explicitly solving for
the atmosphere/ocean/biosphere CO2 mass transfer. Therefore, the CICERO-2 carbon cycle takes into account the nonlinearity in ocean chemistry and
biosphere uptake at high CO2 partial pressures since it represents the atmospheric change in CO2 as a function of total background.
Calculation of CO2 concentrations from emissionsFaIR SCM
The FaIR SCM is described by Millar et al. (2017) and summarized as follows. FaIR is a modified version of the IPCC AR5 four time-constant
impulse response function (IRF) model, which represents the evolution of atmospheric CO2 by partitioning emissions of anthropogenic CO2 between
four reservoirs of an atmospheric CO2 concentrations change, following a pulse emission (see Myhre et al., 2013 for more details). In more
comprehensive models, ocean uptake efficiency declines with accumulated CO2 in ocean sinks (Revelle and Suess, 1957) and uptake of carbon into
both terrestrial and marine sinks are reduced by warming (Friedlingstein et al., 2006). FAIR captures some of these dynamics within the simple IRF
structure, mimicking the behaviour of Earth System Models/Earth System Models of Intermediate Complexity in response to finite-amplitude CO2
injections; this is achieved by introducing a state-dependent carbon uptake with a single scaling factor, , to all four of the time constants in the carbon
cycle of the IPCC AR5 impulse response model used for the calculation of CO2-equivalence metrics. This approach is described in more detail by Millar
et al. (2017).
C. Radiative forcing, efficacy and effective radiative forcing (ERF)
Radiative forcing (RF) has been introduced as a predictor for the expected equilibrium global mean of the (near) surface temperature change Ts
that results from the introduction of climate forcers, such as additional atmospheric CO2 or a change in the solar irradiation (e.g., IPCC, 2007):
Ts =  RF                                                                                                                  (C.1)
where  is the climate sensitivity parameter (K (W m 2) 1). Several definitions of RF exist. According to the simplest one, the instantaneous RF is the
change in the total irradiation (incoming short-wave solar radiation minus the outgoing long-wave terrestrial radiation) at the top of the atmosphere
over the industrial era. However, for most of the climate forcers a better definition (with respect to the linearity of Eq. (C.1)) is the stratosphereadjusted
RF at the tropopause. Here, after the introduction of the new climate forcer, the temperature of the stratosphere is allowed to reach a
new radiative equilibrium, while all other atmospheric state variables are kept constant. The stratosphere-adjusted RF at the tropopause was used in
many of the earlier IPCC reports (IPCC, 1999) and in earlier assessments of aviation climate impacts (Sausen et al., 2005; Lee et al., 2009).


16

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
While Eq. (C.1) is a fairly good approximation for many nearly spatially homogeneously distributed climate forcers, such as global increases of CO2
or CH4, Eq. (C.1) fails to some extent for many forcers that are heterogeneously distributed either horizontally or vertically; such is the case for
aviation-induced ozone perturbations and contrail cirrus (e.g., Hansen et al., 1997, 2005; Forster and Shine, 1997; Stuber et al., 2005). To overcome
this problem Hansen and Nazarenko (2004) introduced the efficacy, ri, into Eqn (C.1):
Ts = ri CO2 RF = i RF with i = ri CO2                                                                                          (C.2)
Here CO2 is the climate sensitivity parameter for a CO2 perturbation. While  in (C.1) is considered a universal constant, which can only be determined
by climate models and hence is model dependent, i depends on the type of forcing, as does ri. (While rCO2 is 1 by definition, rlinear contrails is < 1
(Ponater et al., 2005; Rap et al., 2010)). Eqn (C.2) can also be expressed differently:
Ts = CO2 RFi* with RFi* = ri RF                                                                                                 (C.3)
In Eqn (C.3) RFi* is the forcing modified by the efficacy, which yields a better approximation for the surface temperature change than RF. However,
the calculation of the RFi* is computationally much more expensive than the calculation of RF, as it requires the determination of the equilibrium
temperature change, Ts, with a comprehensive climate model.
As an alternative, the effective radiative forcing (ERF) has been introduced as a more practical indicator of the eventual global mean temperature
response (IPCC, 2013). While RFi* assumes equilibrium climate change, ERF only includes all 'fast' atmospheric responses to a given climate forcer.
For example, rapid adjustments in cloud cover, such as from aerosols, or in properties that respond to changes in water vapor, can either increase or
decrease the initial RF. In contrast, the instantaneous, stratosphere-adjusted, and effective RFs for well-mixed greenhouse gases are nearly equal. In
practice, ERF is determined with a comprehensive climate model, which calculates a new equilibrium radiative imbalance, while the sea surface
temperature and/or the global surface temperature is kept constant. As a consequence, an ERF value is expected to be somewhere between RF and RFi*
values and closer to RFi* values.
D. Aviation NOx radiative forcings

Impacts of NOx emissions on ozone, methane and stratospheric water vapor
Model studies. In this ensemble analysis of the climate forcing from aviation NOx emissions, the results of 20 studies published since the IPCC
(1999) aviation report were considered: IPCC (1999), Sausen et al. (2005), Stordal et al. (2006), Kohler et al. (2008), Hoor et al. (2009), Myhre et al.
(2011), Fromming et al. (2012), Olivie et al. (2012), Gottschaldt et al. (2013), Kohler et al. (2013), Olsen et al. (2013), Skowron et al. (2013),
Khodayari et al. (2014a, 2014b), Khodayari et al., 2014, Svde et al. (2014), Skowron et al. (2015), Pitari et al. (2015), Kapadia et al. (2016), Pitari
et al. (2017), Lund et al. (2017). Three studies that reported results from a 100-year integration of a pulse NOx emission (Wild et al., 2001; Derwent
et al., 2001; Stevenson et al., 2004) were not included in this analysis, nor has as Unger et al. (2010) which uses a different methodology to the
aforementioned.
This model ensemble represents various methodologies in calculating and treating the long-term effects; in order to avoid gaps and additional
uncertainties, standardized RFs for reductions in CH4-induced O3 and SWV were adopted, except for one study that calculates the 'real' long-term
effects from their 50-yr integrations (Pitari et al., 2017):
All analyzed short-term O3 RFs account for a stratospheric adjustment: Assuming that it reduces the instantaneous RF by ~20% (Myhre et al., 2013;
Stevenson et al., 1998), a factor of 0.8 was applied to any O3 RF that is an instantaneous RF (e.g., in the cases of Khodayari et al. (2014a, 2014b) and
Olsen et al. (2013)).
Reductions in CH4-induced O3 and SWV are defined as 50% (Myhre et al., 2013) and 15% (Myhre et al., 2007) of reported CH4 RFs, respectively.
This is applicable for studies that either originally did not provide CH4-induced O3 and SWV estimates (e.g., IPCC, 1999; Sausen et al., 2005; Olsen
et al., 2013) or derived these RFs using another assumptions (e.g., Stordal et al., 2006; Kohler et al., 2008; Hoor et al., 2009; Gottschaldt et al.,
2013; Kohler et al., 2013; Skowron et al., 2013; Khodayari et al., 2014a).
Further assumptions regarding data treatment are:
Fromming et al. (2012), Olivie et al. (2012), Khodayari et al. (2014b) and Kapadia et al. (2016) provide the short-term O3 RFs only and p-TOMCAT
in Stordal et al. (2006) calculates just the long-term effects; thus, these numbers are included in the respective NOx variable analysis but do not
contribute to the net NOx estimate.
Whenever the same estimate appears repetitively in subsequent studies, it is treated as a single entry: this is the case for CAM4 short-term O3 RF
that appears in Khodayari et al. (2014a, 2014b) and Olsen et al. (2013), CAM5 short-term O3 RF that can be found in Khodayari et al. (2014a,
2014b) and NASA ModelE2 short-term O3 and CH4 RFs presented by Unger et al. (2013) and Olsen et al. (2013).
In addition, the ERF estimates for the CH4 term include shortwave RF (Etminan et al., 2016). The inclusion of shortwave forcing in the simplified
expression increases CH4 RF from aviation NOx emissions by 23% (based on MOZART-3 CTM runs driven for all the aircraft emission inventories
represented in the model ensemble) (Table D.1).
Ensemble values. This ensemble analysis covers a period of almost two decades; however, none of the RF per unit of emitted N estimates show any
trends over time of publication and the spread in RF per unit of emitted N values has not changed. The short-term O3 RF varies from 6.2 to 45.1 mW
m 2 (Tg (N) yr 1) 1, where these values come from the NASA ModelE2 (Olsen et al., 2013) and p-TOMCAT (Hoor et al., 2009) models, respectively.
The long-term CH4 RF varies from 27.9 to 8.1 mW m 2 (Tg (N) yr 1) 1, from the p-TOMCAT (Kohler et al., 2008) and MOZART3 (Skowron et al.,


17

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
2015) models, respectively. The spread of other CH4-induced long-term effects follows that of CH4. The net-NOx RF varies from 17.5 to 11.9 mW m 2
(Tg (N) yr 1) 1 from ECHAM/MESSy (Gottschaldt et al., 2013) and CAM4 (Khodayari et al., 2014a), respectively. The Results from the mid-1990s
CTMs are within the envelope of RFs generated more recently (Fig. 3). The numbers from IPCC (1999) and related studies, Sausen et al. (2005)
and Lee et al. (2009), where the non-CO2 effects were originally calibrated to the results from IPCC (1999), do not alter the best NOx RF values and
their uncertainties (Table D.2).
Correlations. The correlations between the NOx RF components are shown in Fig. D.1. In addition to the significant negative correlations between
the short-term and the long-term aviation RF components, correlations between the net-NOx effect and its components are also apparent, especially for
the short-term O3 and net-NOx components; however, their strength is around half. The high correlations (p = 1, R2 = 1) across the long-term effects is
expected since CH4-induced O3 and SWV are all derived based on CH4 RFs. In units of mW m 2 (Tg(N yr 1) 1, 49% of this ensemble short-term O3 RF is
concentrated between 20 and 35, 43% of CH4 RFs is found between 14 and 10, 41% of CH4-induced O3 RFs is between 7 and 5 and 45% of SWV
RFs vary from 2.5 to 1.5. Of the normalized net-NOx RFs resulting from this ensemble, 44% are observed between 5 and 10 mW m 2 (Tg(N)
yr 1) 1.
Transient vs. equilibrium. In calculating the CH4 RF response to aviation NOx emissions, the lack of steady-state conditions is an important
consideration. Since methane (CH4) has a lifetime of the order 812 years (largely model-dependent) any NOx perturbation takes on the order ~40
years to come within 2% of the steady state solution. Moreover, the timescale of removal of CH4 from the atmosphere is made longer through a positive
chemical feedback (Prather, 1994). In order to overcome the necessity to run a global chemical transport model (CTM) with full chemistry for such
long integrations, a parameterization to account for this perturbation was originally developed by Fuglestvedt et al. (1999) and has been widely
adopted since then. However, with the significant annual increases in aviation NOx emissions over the last several decades (Fig. D.2a) the CH4
response does not reach its steady-state value in any given year of emissions, so the steady-state solution generally overstates the CH4 response in a
particular year from historical time-evolving emissions. Similar considerations apply to other sectors with substantial NOx emissions such as shipping
(Myhre et al., 2011). If steady-state conditions are utilized, there is a conceptual and quantitative mismatch when comparing the NOx RF from aviation
with other RF terms, since RF represents a particular condition at a point in time, not the steady-state conditions. To remedy this mismatch, Myhre
et al. (2011) suggested that a factor accounting for the non-steady-state condition of CH4 be introduced, thereby modifying the CH4 impact for a given
year of interest, and further suggested that for the aviation RF in the year 2000 the CH4 term be reduced by approximately 35% for aircraft emissions
using a simplified estimation derived from Grewe and Stenke (2008).
Here, we present an updated methodology to calculate the non-steady-state aviation-NOx-induced CH4 perturbation for the specific year of 2018.
The method relies on transient and steady-state runs of the TROPOS 2D CTM. The results of the steady-state runs using constant emissions for a given
year are compared with those of transient runs using background historical surface emissions from anthropogenic activities and the corresponding
aviation NOx emissions. The latter requires full implementation of time-varying CH4 emissions into the model simulation, a requirement that is not a
standard set-up for many of the CTM/GCMs currently in use where CH4 conditions are defined from observations as fixed concentrations with
relaxation terms introduced to accommodate perturbations to these concentrations. The use of CTM runs explicitly accounts for changing background
atmospheric conditions over the integration period as well as the change in emission rate dependence of the O3 and CH4 responses.
Method. In order to compare these two methods, two types of experiments were performed:
Transient experiment: a long-term simulation with anthropogenic (surface and aviation) emissions evolving over time covering the period
19502050, using historical data up to 2000 and the RCP-4.5 scenario after 2000 (Fig. D.2a),
Steady-state experiment: a 100-year simulation with constant anthropogenic (surface and aviation) emissions representing the year 2000, 2018 or
2050 (Fig. D.2a); the steady-state CH4 response starts to be observed 6070 years into the run.
Each of these experiments was run twice, with and without aviation emissions, and the difference between these two Results defined as the aircraft
response (e.g., Fig. D.2d-f). The initial concentrations of CH4 were set using the observations from NOAA surface stations (Montzka et al., 2000) for
1950 and 2000; for the year 2050 the CH4 concentrations are taken from projections of the MAGICC model (Meinshausen et al., 2011). The background
anthropogenic emissions of CO, CH4, NOx, N2O, and non-methane volatile organic carbon (NMVOC) compounds, as well as aircraft NOx
emissions, evolve during the period 19502050 (Lamarque et al., 2010; Clarke et al., 2007) (Fig. D.2a). The natural emissions from soils and oceans
were kept constant and represent the year 2000 (Prather et al., 2001).
The TROPOS CTM is a latitudinally-averaged, two-dimensional Eulerian global tropospheric chemistry model extensively evaluated by Hough
(1989, 1991). The model's domain extends from pole-to-pole (24 latitudinal grid cells) and from the surface to an altitude of 24 km (12 vertical layers).
TROPOS is driven by chemistry, emissions, transport, removal processes and upper boundary conditions. There are 56 chemical species in the
chemical mechanism of the model, which consists of 91 thermal reactions, 27 photolytic reactions and 7 more reactions, which include night-time NO3
chemistry. The reaction rates and cross sections were updated to the evaluation of Sander et al. (2006) (see Skowron et al., 2009). There are no fixed
concentrations within the model domain other than the upper boundary conditions, which are specified for long-lived species and for gases that have
stratospheric sources. This 2D CTM has the disadvantage of zonal symmetry but has the advantage of an adequate chemical scheme and computational
efficiency, such that long-term integrations can be reasonably performed. Owing to the aforementioned reasons, the O3 response in TROPOS is
overestimated by a factor of ~2 by comparison with a range of up-to-date 3D models. As a consequence, the CH4 results in Fig. D.2d-f were reduced
accordingly. This modification of the original TROPOS responses does not affect the core result of this study, which is the relative difference of CH4
responses between transient and equilibrium methods.
Results. Fig. D.2b shows the evolution of the global CH4 burden over the period 19502050 in the transient TROPOS simulation. There is a steady
growth in the atmospheric CH4 burden, with a small decline over the period 19972007 in response to the decrease in CH4 emissions over the period
19902000. The steady-state simulations for the year 2000 and 2050 agree well (within 1%) with transient CH4 responses for the respective years. A
similar agreement is observed for modelled transient and steady-state CH4 lifetimes in Fig. D.2c. Most of the CH4 loss in the atmosphere is driven by
OH and the oxidative capacity of the atmosphere changes over time (thus CH4 lifetime as well), influenced by emissions of CO, NOx, NMVOC or CH4.
Fig. D.2c shows the evolution of global CH4 lifetime (LT) over the period 19502050: there is a decrease in the CH4 lifetime between 1950 and 2000


18

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
(until around 2007), whilst under the RCP-4.5 scenario the opposite is observed, with the CH4 lifetime increasing by 3.5% by the end of 2050
compared with 2000. The TROPOS CH4 lifetimes agree relatively well with other studies (e.g., Holmes et al., 2013; Voulgarakis et al., 2013; Dalsren
et al., 2016) not only in terms of absolute numbers but also the rate of changes; a detailed comparison is presented in Table D.3. The perturbation
lifetime of CH4 in TROPOS is 37% longer than its global lifetime and the sensitivity coefficient s = ln(LT)/ln(CH4) is 0.27, placing these estimates in
the middle of model ranges (e.g., Prather et al., 2001; Holmes et al., 2011). These terms were calculated using a 5% increase of CH4 global levels for the
year 2000. There is no need to apply the feedback factor (1.37) to the TROPOS CH4 estimates as it is already included in the observed responses;
TROPOS does not have a fixed boundary conditions, so CH4 and OH can freely interact.
Aircraft NOx emissions, via the chemical coupling to OH and HO2, enhance OH, which reduces the global CH4 lifetime. Fig. D.2d shows the
evolution of the CH4 lifetime reduction in the transient 19502050 simulation and in steady-state runs for conditions representing the years 2000 and
2050. In the transient run, there is a steady decrease of global CH4 lifetime as a consequence of a constant increase of aviation NOx emissions during
the period 19502050. The agreement in 2000 and 2050 between the transient and steady-state CH4 lifetime reductions is within 6% (on a global
scale) (see Table D.3). These relatively small differences in CH4 lifetime lead to much more pronounced differences in the associated global CH4
burdens as shown in Fig. D.2e. In contrast to the lifetime results, the CH4 burden response in the transient run lags behind the steady-state CH4
response with differences of 27% in the year 2000 and 20% in the year 2050. Similarly, the calculations for 2018 emissions yield a multiplicative
correction factor of 0.79 (Fig. D.2f), which has been incorporated into the ERF values of CH4, long-term O3 and SWV shown in Fig. 5.
The CH4 results contrast with O3 changes from aircraft NOx emissions, which agree within 3% between transient and steady-state experiments with
aircraft O3 burdens of 10.3 and 10.6 Tg (O3), respectively, in the year 2000. These TROPOS O3 magnitudes are at the upper limit of model ranges, as
present-day aircraft O3 perturbations found in the literature vary from 3 to 11 Tg (O3) (e.g., Hoor et al., 2009; Holmes et al., 2011; Khodayari et al.,
2014a). The aircraft O3 burden increases by 41% in 2050, reaching 17.2 and 18.0 Tg(O3) for transient and steady-state experiments, respectively. This
agrees with other studies (e.g., Olsen et al., 2013) that report a multi-model average increase of 44% in O3 burden from future aircraft NOx emissions
under the RCP-4.5 scenario.
The present approach is in general agreement with that presented by Grewe and Stenke (2008), which accounts for CH4 concentrations not being in
steady-state with OH changes in the year of simulation. The present CTM Results further demonstrate the importance of explicitly calculating CH4
changes in response to time-dependent aviation NOx emissions rather than assuming constant emissions. The difference between transient and
steady-state CH4 for the year 2000 found with TROPOS is smaller than that resulting from the Grewe and Stenke (2008) approach (Myhre et al., 2011)
(27% and 35%, respectively). Table D.4presents a further comparison of CH4 correction factors derived in this study. The systematic differences are
likely due to the Grewe and Stenke (2008) values being based on a simplified chemistry/climate model (AirClim) and the present TROPOS simulations
having a different experimental setup (all our emissions (surface + aircraft) are time-varying) and a full chemical reaction scheme with explicit
calculations performed on time-varying emissions. Indeed, if TROPOS is run with constant background emissions representing the year 2000 in a
similar manner using Grewe and Stenke (2008) methodology, the difference between transient and steady-state CH4 for the year 2000 increases from
27% to 31%. This change shows that background emissions modify the CH4 correction factor and further emphasizes the need to have surface and
aircraft emissions that simultaneously follow historical pathways. In other studies using the Grewe and Stenke (2008) methodology, CH4 correction
factors vary from 0.74 to 1.15 depending on the investigated year (2025 or 2050) and aircraft emission scenario (SRES A1B, B1 and B1 ACARE) (the
factor can be larger than 1 if the aircraft emissions are assumed to decrease in the preceding years) (Hodnebrog et al., 2011, 2012).
Uncertainties in the CH4 correction factor are associated mainly with inter-model differences and the applied emission scenarios; the correction
factor is sensitive, within ~10%, to inter-model differences (based on two models, TROPOS and AirClim) and it can vary by another  10% depending
on emission scenario (based on a range of RCP projections up to 2050). Given that the uncertainties of the CH4 correction factor on the net-NOx RF are
rather small, especially when compared with overall uncertainties, we do not include in the estimated uncertainty of the net-NOx RF value a separate
uncertainty due to the correction factor.
E. Contrail cirrus
The global contrail cirrus RF is calculated by homogenizing existing estimates through the use of specific scaling factors. The factors relate to the
choice of air traffic inventory and its basis year; the use of the full 3D flight distance; the use of hourly air traffic data; the feedback of natural clouds;
and correcting for weaknesses in the radiative transfer calculations. The corrections and scaling actions are:
The estimate of Chen and Gettelman (2013) was corrected by redoing the CAM simulation using a lower ice crystal radius of 7 m and a larger
contrail cross-sectional area of 0.09 km2 for the initialization of contrails at an age of about 1520 min, in agreement with observations (Schumann
et al., 2017). The resulting change in cirrus cloudiness including the adjustment in cloudiness due to the presence of contrail cirrus leads to a
radiative forcing of 57 mW m 2.
A scaling S1 of 1.4 is applied for estimates based on the AERO2k inventory for the year 2002 instead of the AEDT inventory for the year 2006 (Bock
and Burkhardt, 2016);
A scaling S2 of 1.14 is applied to estimates that are based on track distance instead of slant distance (Bock and Burkhardt, 2016). The 'slant' air
traffic distance is the full flight distance and not the ground projected 'track' distance.
A scaling S3 of 0.87 is applied to estimates that used monthly instead of hourly resolved air traffic data. This scaling is based on an estimate for the
impact of the temporal resolution of the air traffic data of 25% to 30% within CAM (Chen et al., 2012) and one of no significant change in
ECHAM4-CCMod.
A scaling S4 of 1.15 is applied to account for the underestimation of RF in radiative transfer calculations that use frequency bands instead of line by
line calculations (Myhre et al., 2009).



19

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
The study details and scaling results are shown in Table E.1. Weighting each estimate equally, the best estimate of global contrail cirrus RF is
approximately 66 mW m 2. As noted in the main text, the Chen and Gettelman (2013) calculation is interpreted as being closer to an ERF than an RF,
so was excluded from this averaging. This mean RF estimate does not include the RF due to contrails forming within natural cirrus. Uncertainty due to
scalings S3S4 is included in the uncertainty discussion below, whereas uncertainty in scalings S1S2, namely updating the ECHAM4-CCMod estimates
using sensitivities from ECHAM5-CCMod, is neglected.
The statistical uncertainty of global contrail cirrus RF cannot be estimated from the small number of available studies. Uncertainties affecting our
contrail cirrus estimates are, on the one hand, due to (A) uncertainties in the radiative response to the presence of contrail cirrus and, on the other
hand, (B) uncertainties in the upper tropospheric water budget and the contrail cirrus scheme. In most cases, we can only infer very rough estimates for
the uncertainties related to specific processes.
(A) Uncertainties associated with the radiative response to contrail cirrus are:
A1 Uncertainty related to the model's radiative transfer scheme of approximately 35% (Myhre et al., 2009).
A2 Uncertainty in the inhomogeneity of ice clouds within a grid box of a climate model (Carlin et al., 2002; Pomroy and Illingworth, 2000), the
vertical cloud overlap, and the use of plane parallel geometry as compared to full 3D radiative transfer (Gounou and Hogan, 2007), which
together amount to approximately 35%.
A3 Uncertainty estimating radiative transfer in a global climate model in the presence of very small ice crystals within young contrails, which may
amount to about 10% (Bock and Burkhardt, 2016). The uncertainty is dependent on the contrail cirrus ice water content.
A4 Uncertainty due to the ice crystal habit is approximately 20% according to Markowicz and Witek (2011).
A5 Uncertainty in the radiative transfer due to soot cores within the contrail cirrus ice crystals is thought to be large, as the change in the shortwave
(SW) albedo is large (Liou et al., 2013). The soot impact on contrail cirrus RF has not yet been quantified.
Overall, uncertainty in the radiative response to contrail cirrus (excluding A3) is estimated to be about 55%, assuming independence of different
uncertainties and excluding the impact of ice crystal soot cores. The uncertainty A3 is included in the uncertainty estimate under (B) because A3 and
B2 are dependent uncertainties.
(B) Uncertainty in contrail cirrus RF associated with the upper-tropospheric water budget and the contrail cirrus scheme are:
B1 Uncertainty in contrail cirrus RF associated with the uncertainty in upper-tropospheric ice supersaturation. This results from a lack of
knowledge in ambient conditions due to the low vertical resolution of satellite instruments (Lamquin et al., 2012) and to the ability of models to
reproduce the observed statistics of ice supersaturation. This contributes about 20% to uncertainty.
B2 There is uncertainty related to ice crystal number densities within young contrails. Ice nucleation within the plume can vary drastically
depending on the water supersaturation reached within the plume and on the soot emissions (Karcher et al., 2015, 2018). This dependency on
the atmospheric state leads to a reduction in the number of nucleated ice crystals in particular in the tropics and at lower flight levels (Bier and
Burkhardt, 2019) leading to a large uncertainty in the impact of tropical and subtropical air traffic. Depending on the atmospheric state and ice
crystal numbers, a varying fraction of ice crystals can be lost in the contrail vortex phase (Unterstrasser, 2014). We assume an uncertainty in
average contrail ice crystal numbers after the vortex phase of about 50% leading to an uncertainty in contrail cirrus RF of about 20%. This
estimate of the sensitivity of contrail cirrus RF to ice crystal numbers in newly formed contrails is based on simulations with ECHAM5-CCMod
(Burkhardt et al., 2018).
B3 The uncertainty in the lifetime of contrail cirrus, affecting the day-/night-time contrail cover, has only a small impact on the estimated contrail
cirrus RF (Chen and Gettelman, 2013; Newinger and Burkhardt, 2012). We estimate the associated uncertainty to be 510%.
B4 From the sensitivity of the contrail cirrus RF to the temporal resolution in the air traffic dataset in ECHAM5 and CAM, we deduce an uncertainty
of about 10%.
B5 The estimate of the feedback of natural clouds, due to contrail cirrus changing the water and heat budget of the upper troposphere, is very
uncertain and has not been properly quantified yet (Burkhardt and Karcher, 2011; Schumann et al., 2015). We assume here the uncertainty
related to this estimate to be only slightly smaller than the estimate itself, or about 15%.
B6 Uncertainty in the RF estimate of Chen and Gettelman (2013) to assumptions in the initial ice-crystal radii and contrail cross-sectional areas is
about 33%.
We assume independence of the uncertainties except for the dependence of A3 and B3 on the uncertainty in B2. The overall uncertainty due to the
water budget and the contrail cirrus scheme (including uncertainty A3) is about 40% and more than 50% in the case of the Chen and Gettelman
(2013). From the two different sources of uncertainty (list A, radiative, and list B, contrail cirrus properties, above) we calculate an overall contrail
cirrus RF uncertainty of about 70%, assuming independence of the overall uncertainties described in A and B.
Note that we do not attempt to infer an estimate for the uncertainty of the factor ERF/RF. When calculating the contrail cirrus ERF, the error range
given refers to the error range of contrail cirrus RF and not ERF.
F. Emission metrics calculations
We calculate the AGWP and AGTP, and corresponding GWPs and GTPs, for aviation CO2, NOx (which encompasses the ERF of short-term O3, CH4,
CH4-induced O3 and SWV), soot, SO2, and contrail cirrus. The methodology and analytical expressions for the emissions metrics are described in detail
in previous literature (e.g., Fuglestvedt et al., 2010; Myhre et al., 2013). The impulse response function (IRF) that describes the atmospheric decay of
CO2 upon emission is taken from Joos et al. (2013). For the other species, the atmospheric decay is given by a constant e-folding time taken as the



20



D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
'perturbation lifetime'. The lifetimes used here are broadly consistent with Fuglestvedt et al. (2010). The radiative efficiency (RE) for CO2 is calculated
using year 2018 background concentrations of 407 ppm (annual mean, from monthly mean observed concentrations from NOAA GMD - ftp://aftp.cmd
l.noaa.gov/products/trends/co2/co2_mm_gl.txt). This yields a RE of 1.68  10 15 W m 2 kg 1), 4% lower than used in the IPCC Fifth Assessment
report (AR5) (Myhre et al., 2013). The climate response IRF is taken from Boucher and Reddy (2008). The latter has an inherent equilibrium climate
sensitivity (ECS) of 1.06K (W m 2) 1, equivalent to a 3.9K equilibrium response to a doubling of CO2.
For the calculation of the average rate of CO2-warming-equivalent emissions for aviation non-CO2 forcings (ECO2e*) under the GWP* metric in
Table 5, we use the relationship between recent changes in effective RF and CO2-equivalent emissions from Allen et al. (2018) (or Equation (1) with
= 0),
ECO2e* = [F / t]  [H / AGWPH(CO2)]                                                                                            (F.1)
where F in Eqn (F.1) is the change in ERF over the recent period, t, and AGWPH(CO2) is the absolute global warming potential of CO2 at time horizon
H. We use updated AGWPH(CO2) values incorporating the updated radiative efficiency of CO2 as described in the previous paragraph. Allen et al. (2018)
used a backward-looking period of 20 years as t, whereas here we use a backward-looking 18-yr period as our time series of ERF components only
extends back to 2000.
G. List of Acronyms and abbreviations used in tables and figures of the Appendices
ACARE  Advisory Council for Aeronautical Research in Europe
ACCMIP Atmospheric Chemistry and Climate Model Intercomparison Project
AEDT   Aviation Environmental Design Tool
AEM   Advanced Emission Model
AERO2K Global aircraft emissions data project for climate impacts evaluation
AGAGE  Advanced Global Atmospheric Gases Experiment
CAM   Community Atmosphere Model
CCMod  Contrail Cirrus Module
CH3CCl3 Methyl chloroform
COCIP   Contrail Cirrus Prediction Tool
CTM   Chemical Transport Model
ECHAM European Centre/Hamburg Model
IPCC    Intergovernmental Panel on Climate Change
MAGICC Model for the Assessment of Greenhouse Gas Induced Climate Change
MOZART Model for OZone And Related chemical Tracers
NOAA  National Oceanic and Atmospheric Administration
QUANTIFY Quantifying the Climate Impact of Global and European Transport System
REACT4C Reducing Emissions from Aviation by Changing Trajectories for the benefit of Climate
RCP    Representative Concentration Pathway
SRES    Special Report on Emission Scenarios
TAR    Third Assessment Report
TRADEOFF Aircraft emissions: contribution of different climate components to changes in radiative forcingtradeoff to reduce atmospheric impact
TROPOS 2D global TROPOSpheric model
WDCGG World Data Centre for Greenhouse Gases
Table D.1
The CH4 RFs derived for all the aircraft emission inventories
that are present in the model ensemble.a
Inventories              CH4 RF, mW m 2
Old                New
AEDT                   6.67               8.22
AEM                  6.82              8.41
AERO2K                 7.09               8.74
REACT4C                6.97               8.59
QUANTIFY               6.96               8.58
TRADEOFF               7.11               8.76
a Values are those represented in the model ensemble based on
MOZART-3 CTM simulations (Old) and recalculated values
using a revised simplified expression for the CH4 RF (New) as
presented by Etminan et al. (2016). The NOx emissions of each
inventory are normalized so that all RFs are scaled to the same
global total emissions (0.71 Tg(N) yr 1) as in the REACT4C
model.



21

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834



Table D.2
The best NOx RFs per unit emission derived for datasets that include and exclude late 1990s numbers and related estimates, see
text for details.
Components                 Value               Uncertainty*              Value               Uncertainty*
(mW m 2 (Tg (N) yr 1) 1
with IPCC (1999)                              without IPCC (1999)
Short-term O3                25.6                7.3                    25.1                7.2
CH4                       13.8              4.7                     13.4              4.5
CH4-induced O3               6.9               2.3                     6.7               2.3
SWV                     2.1              0.7                   2.0              0.7
Net NOx                    3.9                 5.7                    4.0                 5.8
*Stated uncertainties are one standard deviation (68% confidence interval).

Table D.3
Methane response in TROPOS and other studies
Variable                      Year    2D CTM, TROPOS          Literature
Transient   Steady-statea   Study                  Ref   Model/Years                Variable estimate/change
CH4 burden, Tg                2000   4770.8     4785.1
IPCC TAR                    1998                     4850 Tg
Voulgarakis et al. (2013)          ACCMIP                   4750d Tg
Dalsren et al. (2016)            Oslo CTM3                 4560d Tg
Dalsren et al. (2016)            19702012                 +15%
This studyc                                            +13%
2050   5051.6     5081.4       Voulgarakis et al. (2013)          ACCMIP                   5000d Tg
Voulgarakis et al. (2013)                                   +5.3d %
This studyc                   20002050                 +5.9%
CH4 abundance, ppb             2000   1784.2     1787.5       Observations                  NOAA                    1773 ppb
AGAGE                  1774 ppb
WDCGG                1783 ppb
2050   1886.2     1897.6       Meinshausen et al. (2011)         MAGICC                   1833 ppb
CH4 lifetime (CH4+OH)b, yr        2000   10.6       10.5         Prather et al. (2012)             CH3CCl3-based              11.2  1.3 yr
Voulgarakis et al. (2013)          ACCMIP                   9.8  1.6 yr
Holmes et al. (2013)             1980/852000/05             2.2  1.8%
This studyc                                             2.06%
Voulgarakis et al. (2013)          19802000                  4%
This studyc                                             2%
2050   11.0       11.0         Voulgarakis et al. (2013)          20002050                 +1.0d %
This studyc                                            +3.5%
aircraft CH4 lifetime (CH4+OH), yr   2000    0.137      0.145       Hoor et al. (2009)               AERO2K                    1.55% Tg(N) 1
Myhre et al. (2011)              QUANTIFY                  1.46% Tg(N) 1
Holmes et al. (2011)             Model ensemble              1.77% Tg(N) 1
Svde et al. (2014)              REACT4C dENOx = QUANTIFY    1.36% Tg(N) 1
This studyc                                             1.48% Tg(N) 1
2050    0.293      0.311       Hodnebrog et al. (2011)          SRES B1                    1.61% Tg(N) 1
B1 ACARE                  1.48% Tg(N) 1
Hodnebrog et al. (2012)          SRES A1B                   1.22% Tg(N) 1
Khodayari et al. (2014a)          AEDT Scenario1              1.88% Tg(N) 1
AEDT Baseline               1.59% Tg(N) 1
This studyc                   RCP45                     1.36% Tg(N) 1
a this is an average of the last 10 years of simulations
b the chemcial (
CH4+OH) lifetime is around 7% greater than the total CH4 lifetime, as modelled by TROPOS
c numbers are based on transient simulation
d numbers might not be very accurate as they are read directly from the graphs found in the respecitve papers





22

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834


Table D.4
Calculated CH4 correction factors
Aviation emissions year              CH4 correction factors
This study              Grewe and Stenke (2008) methodology
2000                           0.73                  0.65
2005                           0.75                  0.73
2011                           0.78                  0.81
2018                           0.79                  0.86

Table E.1
Scaling of contrail cirrus RF and ERF Results a
Model             Inventory        Representation of flight distance     RF (mW/m2)     Scalings      Scaled RF (mW/m2)b     Reference
ECHAM4-CCMod     AERO2K 2002     track                        38            S1, S2, S4     70                  Burkhardt and Karcher (2011)
ECHAM5-CCMod     AEDT 2006       slant                        56            S3, S4       56                  Bock and Burkhardt (2016)
COCIP            AEDT 2006       flight vectors                   63            S4          72                  Schumann et al. (2015)
CAM5            AEDT 2006      slant                      13 [57]c        S3, S4       57                 Chen and Gettelman (2013)
Best estimate                                                                            66d
a Adapted from Table 1 of Bock and Burkhardt (2016).
b RF that would be expected in 2006 when using slant distance from the AEDT inventory with hourly resolution.
c An updated simulation (see text) yielded 57 mW m 2.
d The best estimate is of RFs, and excludes the Chen and Gettelman (2013) results since this is closer to an ERF (see main text).

Table F.1a
Emission metrics and corresponding CO2-equivalent emissions for the ERF components of 2018 aviation emissions and cloudiness using CO2 IRF without C-cycle
feedbacks from Gasser et al., 2017, and climate IRF from Boucher and Reddy (2008).
Metrics
ERF term                           GWP20             GWP50             GWP100             GTP20             GTP50             GTP100
CO2                              1                1                1                 1                1                1
Contrail cirrus (Tg CO2 basis)             2.39               1.15               0.68               0.70              0.11              0.10
Contrail cirrus (km basis)                40                19                11                 12               1.9               1.6
Net NOx                            637               216               122                 231              75              14
Aerosol-radiation
Soot emissions                      4409              2125              1252               1295              210              177
SO2 emissions                       856               412               243               251              41               34
Water vapor emissions                  0.22               0.11               0.06               0.07              0.01              0.009

Table F.1b
Emission metrics and corresponding CO2-equivalent emissions for the ERF components of 2018 aviation emissions and cloudiness using CO2 IRF without C-cycle
feedbacks, and climate IRF from Gasser et al. (2017).
Metrics
ERF term                           GWP20             GWP50             GWP100             GTP20             GTP50             GTP100
CO2                              1                1                1                 1                1                1
Contrail cirrus (Tg CO2 basis)             2.39               1.15               0.68               0.3               0.19              0.15
Contrail cirrus (km basis)                40                19                11                 4                3.3               2.6
Net NOx                            637               216               122                 420              18              22
Aerosol-radiation
Soot emissions                      4409              2125              1252               466               360              284
SO2 emissions                       856               412               243               90               70               55
Water vapor emissions                  0.22               0.11               0.06               0.03              0.018             0.014






23

D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834

























Fig. D.1. Matrix of pair-wise scatter plots of RF values from NOx terms: short-term O3, CH4, CH4-induced O3, SWV and net NOx (i.e., the sum of all 4 components), all
represented as normalized RFs (mW m 2 (Tg(N)yr 1) 1) from the ensemble studies (see details in text). The red line is the linear fit, the ellipse shows the 95%
confidence level and histograms present frequencies.






24



D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834





























Fig. D.2. (a) Past and future anthropogenic emissions of CO, CH4, NOx, NMVOC, N2O and aircraft NOx (IIASA RCP Database: http://www.iiasa.ac.at/web-apps/tnt/
RcpDb/). Dots represent conditions for 'constant 2000' and 'constant 2050' simulations. (b) Evolution of the global CH4 burden in TROPOS for transient aircraft NOx
emissions combining historical emissions (19502000) and RCP-4.5 emissions (20002050); and constant emissions for the years 2000 and 2050. (c) Global CH4
lifetime due to aircraft NOx emissions in TROPOS for transient emissions combining historical emissions (19502000) and RCP-4.5 emissions (20002050); and
constant emissions for the years 2000 and 2050. (d) Global CH4 lifetime reduction due to aircraft NOx emissions in TROPOS for transient emissions combining
historical emissions (19502000) and RCP-4.5 emissions (20002050); and constant emissions for the years 2000 and 2050. The dashed lines represent 2000 and
2050 equilibrium values (light and dark blue) and 2000 and 2050 transient values (red). (e) Global CH4 burden reduction due to aircraft NOx emissions in TROPOS

25















































































D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
for transient emissions combining historical emissions (19502000) and RCP-4.5 emissions (20002050); and constant emissions for the years 2000 and 2050. The
dashed lines represent 2000 and 2050 equilibrium values (light and dark blue) and 2000 and 2050 transient values (red). (f) Global CH4 burden reduction due to
aircraft NOx emissions in TROPOS for transient emissions combining historical emissions (19502000) and RCP-4.5 emissions (20002050); and constant emissions
for the year 2018. The dashed lines represent 2018 equilibrium (green) and transient values (red).

References                                                       Chen, C.-C., Gettelman, A., Craig, C., Minnis, P., Duda, D.P., 2012. Global contrail
coverage simulated by CAM5 with the inventory of 2006 global aircraft emissions.
J. Adv. Model. Earth Syst. 4 https://doi.org/10.1029/2011MS000105, 04003.
Airbus, 2017. Global Market Forecast 20172036. Airbus, France.
Chen, C.-C., Gettelman, A., 2013. Simulated radiative forcing from contrails and contrail
Allen, M.R., Fuglestvedt, J.S., Shine, K.P., Reisinger, A., Pierrehumbert, R.T., Forster, P.
cirrus. Atmos. Chem. Phys. 13, 1252512536. https://doi.org/10.5194/acp-13-
M., 2016. New use of global warming potentials to compare cumulative and
12525-2013.
shortlived climate pollutants. Nat. Clim. Change 6 (8), 773776. https://doi.org/
Ciais, P., Sabine, C., Bala, G., Bopp, L., Brovkin, V., Canadell, J., Chhabra, A., DeFries, R.,
10.1038/nclimate2998.
Galloway, J., Heimann, M., Jones, C., Le Quere, C., Myneni, R.B., Piao, S.,
Allen, M.R., Shine, K.P., Fuglestvedt, J.S., Millar, R.J., Cain, M., Frame, D.J., Macey, A.
Thornton, P., 2013. Carbon and other biogeochemical cycles. In: Stocker, T.F.,
H., 2018. A solution to the misrepresentations of CO2-equivalent emissions of short-
Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y.,
lived climate pollutants under ambitious mitigation. npj Clim. Atmos. Sci. 1 https://
Bex, V., Midgley, P.M. (Eds.), Climate Change 2013: the Physical Science Basis.
doi.org/10.1038/s41612-018-0026-8, 16.
Contribution of Working Group I to the Fifth Assessment Report of the
Agarwal, A., Speth, R.L., Fritz, T.M., Jacob, S.D., Rindlisbacher, T., Iovinelli, R.,
Intergovernmental Panel on Climate Change. Cambridge University Press,
Owen, B., Miake-Lye, R.C., Sabnis, J.S., Barrett, S.R.H., 2019. SCOPE11 method for
Cambridge, UK and New York, NY, USA.
estimating aircraft black carbon mass and particle number emissions. Environ. Sci.
Clarke, L., Edmonds, J., Jacoby, H., Pitcher, H., Reilly, J., Richels, R., 2007. "Scenarios of
Technol. 53, 13641373. https://doi.org/10.1021/acs.est.8b04060.
Greenhouse Gas Emissions and Atmospheric Concentrations." Sub-report 2.1A of
Alfsen, K.H., Berntsen, T.. An efficient and accurate carbon cycle model for use in simple
Synthesis and Assessment Product 2.1 by the U.S. Climate Change Science Program
climate mdels. CICERO, Oslo, Norway. https://core.ac.uk/reader/52082516.
and the Subcommittee on Global Change Research. Department of Energy, Office of
Archer, D., Brovkin, V., 2008. The millennial atmospheric lifetime of anthropogenic CO2.
Biological & Environmental Research, Washington, 7 DC, p. 54. https://globalcha
Climatic Change 90, 283297. https://doi.org/10.1007/s10584-008-9413-1.
nge.mit.edu/sites/default/files/CCSP_SAP2-1a-FullReport.pdf .
Balkanski, Y., Myhre, G., Gauss, M., Radel, G., Highwood, E.J., Shine, K.P., 2010. Direct
Collins, W.J., Frame, D.J., Fuglestvedt, J.S., Shine, K.P., 2019. Stable climate metrics for
radiative effect of aerosols emitted by transport: from road, shipping and aviation.
emissions of short and long-lived species  combining steps and pulses. Environ. Res.
Atmos. Chem. Phys. 10 (10), 44774489. https://doi.org/10.5194/acp-10-4477-
Lett. 15 (2) https://doi.org/10.1088/1748-9326/ab6039, 024018.
2010.
Dalsren, S.B., Myhre, C.L., Myhre, G., Gomez-Pelaez, A.J., Svde, O.A., Isaksen, I.S.A.,
Barrett, S., Prather, M., Penner, J., Selkirk, H., Balasubramanian, S., Dopelheuer, A.,
Weiss, R.F., Harth, C.M., 2016. Atmospheric methane evolution the last 40 years.
Fleming, G., Gupta, M., Halthore, R., Hileman, J., Jacobson, M., Kuhn, S.,
Atmos. Chem. Phys. 16, 30993126. https://doi.org/10.5194/acp-16-3099-2016.
Lukachko, S., Miake-Lye, R., Petzold, A., Roof, C., Schaefer, M., Schumann, U.,
DeMott, P.J., Chen, Y., Kreidenweis, S.M., Rogers, D.C., Sherman, D.E., 1999. Ice
Waitz, I., Wayson, R., 2010. Guidance on the Use of AEDT Gridded Aircraft
formation by black carbon particles. Geophys. Res. Lett. 26, 24292432. https://doi.
Emissions in Atmospheric Models. Massachusetts Institute for Technology,
org/10.1029/1999GL900580.
Laboratory for Aviation and the Environment. LAE-2010-008-N. http://citeseerx.ist.
Derwent, R.G., Collins, W.J., Johnson, C.E., Stevenson, D.S., 2001. Transient behaviour
psu.edu/viewdoc/download?doi=10.1.1.719.2090&rep=rep1&type=pdf.
of tropospheric ozone precursors in a global 3-D CTM and their indirect greenhouse
Bellouin, N., Quaas, J., Gryspeerdt, E., Kinne, S., Stier, P., Watson-Parris, D., Boucher, O.,
effects. Climatic Change 49, 463487. https://doi.org/10.1023/A:1010648913655.
Carslaw, K.S., Christensen, M., Daniau, A.-L., Dufresne, J.-L., Feingold, G., 2019.
Dstan 91-91, 2015. Turbine Fuel, Kerosene Type, Jat A-1. Ministry of Defence, Defence
Bounding global aerosol radiative forcing of climate change. Rev. Geophys. 58
Standard 91-91", Issue 7, Amendment 3. Defence Equipment and Support. UK
https://doi.org/10.1029/2019RG000660 e2019RG000660.
Defence Standardization, Glasgow, UK.
Bickel, M., Ponater, M., Bock, L., Burkhardt, U., Reineke, S., 2020. Estimating the
Ebbinghaus, A., Wiesen, P., 2001. Aircraft fuels and their effects upon engine emissions.
effective radiative forcing of contrail cirrus. J. Clim. 33, 19912005. https://doi.org/
Air Space Eur. 3, 101103. https://doi.org/10.1016/S1290-0958(01)90026-7.
10.1175/JCLI-D-19-0467.1.
Etminan, M., Myhre, G., Highwood, E.J., Shine, K.P., 2016. Radiative forcing of carbon
Bier, A., Burkhardt, U., Bock, L., 2017. Synoptic control of contrail cirrus life cycles and
dioxide, methane, and nitrous oxide: a significant revision of the methane radiative
their modification due to reduced soot number emissions. J. Geophys. Res. Atmos.
forcing. Geophys. Res. Lett. 43, 12,61412,623. https://doi.org/10.1002/
122 (21), 11,58411,603. https://doi.org/10.1002/2017JD027011.
2016GL071930.
Bier, A., Burkhardt, A.U., 2019. Variability in contrail ice nucleation and its dependence
Faber, J., Greenwood, D., Lee, D.S., Mann, M., de Leon, P.M., Nelissen, D., Owen, B.,
on soot number emissions. J. Geophys. Res. Atmos. 124, 33843400. https://doi.
Ralph, M., Tilston, J., van Velzen, A., van de Vreede, G., 2008. Lower NOx at Higher
org/10.1029/2018JD029155.
Altitudes: Policies to Reduce the Climate Impact of Aviation NOx Emissions. CE-
Bock, L., Burkhardt, U., 2016. Reassessing properties and radiative forcing of contrail
Delft, 08.7536.32, Delft, The Netherlands.
cirrus using a climate model. J. Geophys. Res. Atmos. 121, 97179736. https://doi.
Fleming, G., Ziegler, U., 2016. Environmental Trends in Aviation to 2050. In 'ICAO
org/10.1002/2016JD025112.
Environmental Report, 2016'. International Civil Aviation Organization, Montreal.
Boeing, 2018. Orders and Deliveries for January 2018. http://www.boeing.
https://www.icao.int/environmental-protection/Documents/EnvironmentalRepo
com/commercial/#/orders-deliveries.
rts/2019/ENVReport2019_pg17-23.pdf.
Boucher, O., Randall, D., Artaxo, P., Bretherton, C., Feingold, G., Forster, P.,
Fleming, G., de Lepinay, I., 2019. Environmental Trends in Aviation to 2050", in ICAO
Kerminen, V.-M., Kondo, Y., Liao, H., Lohmann, U., Rasch, P., Satheesh, S.K.,
Environmental Report, 2019 Destination Green the Next Chapter. ICAO, Montreal,
Sherwood, S., Stevens, B., Zhang, X.Y., 2013. Clouds and aerosols. In: Sotcker, T.F.,
2019. https://www.icao.int/environmental-protection/Documents/Environmenta
Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Doschung, J., Nauels, A., Xia, Y.,
lReports/2019/ENVReport2019_pg17-23.pdf.
Bex, V., Midgley, P.M. (Eds.), Climate Change 2013: The Physical Science Basis.
Forster, P.M.d.F., Shine, K.P., 1997. Radiative forcing and temperature trends from
Contribution of Working Group I to the Fifth Assessment Report of the
stratospheric ozone changes. J. Geophys. Res. 102, 1084110855. https://doi.org/
Intergovernmental Panel on Climate Change. Cambridge University Press,
10.1029/96JD03510.
pp. 571657. https://doi.org/10.1017/CBO9781107415324.016.
Forster, C., Stohl, A., James, P., Thouret, V., 2003. The residence times of aircraft
Boucher, O., Reddy, M.S., 2008. Climate trade-off between black carbon and carbon
emissions in the stratosphere using a mean emission inventory and emissions along
dioxide emissions. Energy Pol. 36, 193200. https://doi.org/10.1016/j.
actual flight tracks. J. Geophys. Res. Atmos. 108, 8524. https://doi.org/10.1029/
enpol.2007.08.039.
2002JD002515.
Brasseur, G.P., Gupta, M., Anderson, B.E., Balasubramanian, S., Barrett, S., Duda, D.,
Freeman, S., Lee, D.S., Lim, L.L., Skowron, A., De Leon, R.R., 2018. Trading off aircraft
Fleming, G., Forster, P.M., Fuglestvedt, J., et al., 2016. Impact of aviation on climate:
fuel burn and NOx emissions for optimal climate policy. Environ. Sci. Technol. 52,
FAA's aviation climate change research initiative (ACCRI) phase II. Bull. Am.
24982505. https://doi.org/10.1021/acs.est.7b05719.
Meteorol. Soc. 97, 561583. https://doi.org/10.1175/BAMS-D-13-00089.1.
Friedlingstein, P., Cox, P., Betts, R., Bopp, L., von Bloh, W., Brovkin, V., Cadule, P.,
Burkhardt, U., Karcher, B., Schumann, U., 2010. Global Modelling of the contrail and
Doney, S., Eby, M., Fung, I., Bala, G., John, J., Jones, C., Joos, F., Kato, T.,
contrail cirrus climate impact. Bull. Am. Meteorol. Soc. 91, 479484. https://doi.
Kawamiya, M., Knorr, W., Lindsay, K., Matthews, H.D., Raddatz, T., Rayner, P.,
org/10.1175/2009BAMS2656.1.
Reick, C., Roeckner, E., Schnitzler, K.-G., Schnur, R., Strassmann, K., Weaver, A.J.,
Burkhardt, U., Karcher, B., 2011. Global radiative forcing from contrail cirrus. Nat. Clim.
Yoshikawa, C., Zeng, N., 2006. Climate-carbon cycle feedback analysis: Results from
Change 1, 5458. https://doi.org/10.1038/nclimate1068.
the C4MIP model intercomparison. J. Clim. 19, 33373353. https://doi.org/
Burkhardt, U., Bock, L., Bier, A., 2018. Mitigating the contrail cirrus climate impact by
10.1175/JCLI3800.1.
reducing aircraft soot number emissions. npj Clim. Atmos. Sci. 1, 37. https://doi.
Fromming, C., Ponater, M., Dahlmann, K., Grewe, V., Lee, D.S., Sausen, R., 2012.
org/10.1038/s41612-018-0046-4.
Aviation-induced radiative forcing and surface temperature change in dependency of
Cain, M., Lynch, J., Allen, M.R., Fuglestvedt, J.S., Frame, D.J., Macey, A.H., 2019.
the emission altitude. J. Geophys. Res. Atmos. 117, 97179736. https://doi.org/
Improved calculation of warming-equivalent emissions for short-lived climate
10.1029/2012JD018204.
pollutants. npj Clim. Atmos. Sci. 2, 29. https://doi.org/10.1038/s41612-019-0086-
Fuglestvedt, J.S., Berntsen, T., 1999. A Simple Model for Scenario Studies of Changes in
4.
Climate, Version 1.0. CICERO, Oslo, Norway, p. 59. https://cicero.oslo.no/no/public
Carlin, B., Fu, Q., Lohmann, U., Mace, G., Sassen, K., Comstock, J., 2002. High-cloud
ations/internal/326.
horizontal inhomogeneity and solar albedo bias. J. Clim. 15, 23212339. https://
doi.org/10.1175/1520-0442(2002)015<2321:HCHIAS>2.0.CO;2.

26











































































D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Fuglestvedt, J.S., Berntsen, T.K., Isaksen, I.S.A., Mao, H.T., Liang, X.Z., Wang, W.C.,       Irvine, E.A., Hoskins, B.J., Shine, K.P., 2013. A Lagrangian analysis of ice-supersaturated
1999. Climatic forcing of nitrogen oxides through changes in tropospheric ozone and        air over the North Atlantic. J. Geophys. Res. Atmos. 119, 90100. https://doi.org/
methane; global 3D model studies. Atmos. Environ. 33, 961977.                     10.1002/2013JD020251.
Fuglestvedt, J., Berntsen, T., Myhre, G., Rypdal, K., Skeie, R.B., 2008. Climate forcing      IATA, 2019. Economic performance of the airline industry. https://www.iata.org/cont
from the transport sectors. Proc. Natl. Acad. Sci. U.S.A 105 (2), 454458. https://         entassets/f88f0ceb28b64b7e9b46de44b917b98f/iata-economic-performance-of-th
doi.org/10.1073/pnas.0702958104.                                          e-industry-end-year-2018-report.pdf.
Fuglestvedt, J.S., Berntsen, T., Eyring, V., Isaksen, I., Lee, D.S., Sausen, R., 2009.          IATA, 2020. Outlook for air travel in the next 5 years. https://www.iata.org/en/iata-r
Shipping emissions: from cooling to warming of climateand reducing impacts on         epository/publications/economic-reports/covid-19-outlook-for-air-travel-in-the-n
health. Environ. Sci. Technol. 43, 90579062. https://doi.org/10.1021/es901944r.        ext-5-years/.
Fuglestvedt, J.S., Shine, K.P., Berntsen, T., Cook, J., Lee, D.S., Stenke, A., Skeie, R.B.,      ICAO, 2018. ICAO Carbon Emissions Calculator Methodology version 11, June 2018.
Velders, G.J.M., Waitz, I.A., 2010. Transport impacts on atmosphere and climate:         https://www.icao.int/environmental-protection/CarbonOffset/Documents/Metho
Metrics. Atmos. Environ. 44, 46484677. https://doi.org/10.1016/j.                   dology%20ICAO%20Carbon%20Calculator_v11-2018.pdf. (Accessed 19 May 2020).
atmosenv.2009.04.044.                                                ICAO, 2019. Destination Green the Next Chapter. ICAO Environmental Report, Montreal.
Gasser, T., Peters, G.P., Fuglestvedt, J.S., Collins, W.J., Shindell, D.T., Ciais, P., 2017.         https://www.icao.
Accounting for the climate-carbon feedback in emission metrics. Earth Syst. Dynam.        int/environmental-protection/Documents/ICAO-ENV-Report2019-F1-WEB%20(1).
8, 235253. https://doi.org/10.5194/esd-8-235-2017.                             pdf.
Gauss, M., Isaksen, I.S.A., Wong, S., Wang, W.C., 2003. Impact of H2O emissions from      IEA, 2019. International Energy Agency. International Energy Agency Oil Information,
cryoplanes and kerosene aircraft on the atmosphere. J. Geophys. Res. Atmos. 108         1960-2017 [data collection], twelfth ed. https://doi.org/10.5257/iea/oil/2019-1
(D10), 4304. https://doi.org/10.1029/2002JD002623.                            UK Data Service. SN: 5187.
Gettelman, A., Chen, C., 2013. The climate impact of aviation aerosols. Geophys. Res. Lett.     IPCC, 1999. Aviation and the global atmosphere. In: Penner, J.E., Lister, D.H., Griggs, D.
40, 27852789. https://doi.org/10.1002/grl.50520.                              J., Dokken, D.J., McFarland, M. (Eds.), Intergovernmental Panel on Climate Change
Gounou, A., Hogan, R.J., 2007. A sensitivity study of the effect of horizontal photon          Special Report. Cambridge University Press, Cambridge, UK, 1999. https://www.
transport on the radiative forcing of contrails. J. Atmos. Sci. 64, 17061716. https://        ipcc.ch/report/aviation-and-the-global-atmosphere-2/.
doi.org/10.1175/JAS3915.1.                                            IPCC, 2001. UK. In: Houghton, J.T., Ding, Y., Griggs, D.J., Noguer, M., van der Linden, P.
Gottschaldt, K., Voigt, C., Jockel, P., Righi, M., Deckert, R., Dietmller, S., 2013. Global        J., Dai, X., Maskell, K., Johnson, C.A. (Eds.), Climate Change 2001: the Scientific
sensitivity of aviation NOx effects to the HNO3-forming channel of the HO2 + NO         Basis. Contribution of Working Group I to the Third Assessment Report of the
reaction. Atmos. Chem. Phys. 13, 30033025. https://doi.org/10.5194/acp-13-           Intergovernmental Panel on Climate Change. Cambridge University Press. htt
3003-2013.                                                           ps://www.ipcc.ch/site/assets/uploads/2018/07/WG1_TAR_FM.pdf.
Grewe, V., Stenke, A., 2008. AirClim: an efficient tool for climate evaluation of aircraft     IPCC, 2007. Climate change 2007. "Mitigation of climate change. In: Metz, B.,
technology. Atmos. Chem. Phys. 8, 46214639. https://doi.org/10.5194/acp-8-           Davidson, O.R., Bosch, P.R., Dave, R., Meyer, L.A. (Eds.), Contribution of Working
4621-2008.                                                           Group III to the Fourth Assessment Report of the Intergovernmental Panel on
Hansen, J., Sato, M., Ruedy, R., 1997. Radiative forcing and climate response.              Climate Change. Cambridge University Press, UK. https://www.ipcc.ch/report
J. Geophys. Res. Atmos. 102 (D6), 68316864. https://doi.org/10.1029/               /ar4/wg3/.
96JD03436.                                                        IPCC, 2013. In: Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K.,
Hansen, J., Nazarenko, L., 2004. Soot climate forcing via snow and ice albedos. Proc.         Boschung, J., Nauels, A., Xia, Y., Bex, V., Midgley, P.M. (Eds.), Climate Change 2013:
Natl. Acad. Sci. U.S.A. 101, 423428. https://doi.org/10.1073/pnas.2237157100.         the Physical Science Basis, Contribution of Working Group I to the Fifth Assessment
Hansen, J., Sato, M., Ruedy, R., Nazarenko, L., Lacis, A., Schmidt, G.A., Russell, G.,           Report of the Intergovernmental Panel on Climate Change. Cambridge University
Aleinov, I., Bauer, M., Bauer, S., Bell, N., Cairns, B., Canuto, V., Chandler, M.,           Press, Cambridge, United Kingdom and New York, NY, USA, 2013. https://www.ipcc
Cheng, Y., Del Genio, A., Faluvegi, G., Fleming, E., Friend, A., Hall, T., Jackman, C.,        .ch/report/ar5/wg1/.
Kelley, M., Kiang, N., Koch, D., Lean, J., Lerner, J., Lo, K., Menon, S., Miller, R.,       IPCC, 2018. In: Masson-Delmotte, V., Zhai, P., Portner, H.-O., Roberts, D., Skea, J.,
Minnis, P., Novakov, T., Oinas, V., Perlwitz, Ja, Perlwitz, Ju, Rind, D., Romanou, A.,        Shukla, P.R., Pirani, A., Moufouma-Okia, W., Pean, C., Pidcock, R., Connors, S.,
Shindell, D., Stone, P., Sun, S., Tausnev, N., Thresher, D., Wielicki, B., Wong, T.,          Matthews, J.B.R., Chen, Y., Zhou, X., Gomis, M.I., Lonnoy, E., Maycock, T.,
Yao, M., Zhang, S., 2005. Efficacy of climate forcings. J. Geophys. Res. Atmos. 110,        Tignor, M., Waterfield, T. (Eds.), Global Warming of 1.5C. An IPCC Special Report
D18104. https://doi.org/10.1029/2005JD005776.                               on the Impacts of Global Warming of 1.5C above Pre-industrial Levels and Related
Hari, T.K., Yaakob, Z., Binitha, N., 2015. Aviation biofuel from renewable resources:          Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the
routes, opportunities and challenges. Renew. Sustain. Energy Rev. 42, 12341244.        Global Response to the Threat of Climate Change, Sustainable Development, and
https://doi.org/10.1016/j.rser.2014.10.095.                                    Efforts to Eradicate Poverty, 2018. https://www.ipcc.ch/sr15/download/.
Hasselmann, K., Hasselmann, S., Giering, R., Ocana, V., von Storch, H., 1997. Sensitivity     Joos, F., Bruno, M., Fink, R., Stocker, T.F., Siegenthaler, U., LeQuere, C., Sarmiento, J.L.,
study of optimal CO2 emission paths using a simplified structural integrated             1996. An efficient and accurate representation of complex oceanic and biospheric
assessment model (SIAM). Climatic Change 37, 345386. https://doi.org/10.1023/        models for anthropogenic carbon uptake. Tellus 48B, 397e417. https://doi.org/
A:1005339625015.                                                      10.1034/j.1600-0889.1996.t01-2-00006.x.
Hendricks, J., Karcher, B., Lohmann, U., 2011. Effects of ice nuclei on cirrus clouds in a     Joos, F., Roth, R., Fuglestvedt, J.S., Peters, G.P., Enting, I.G., von Bloh, W., Brovkin, V.,
global climate model. J. Geophys. Res. Atmos. 116, 21562202. https://doi.org/          Burke, E.J., Eby, M., Edwards, N.R., Friedrich, T., Frolicher, T.L., Halloran, P.R.,
10.1029/2010JD015302.                                                 Holden, P.B., Jones, C., Kleinen, T., Mackenzie, F.T., Matsumoto, K.,
Hodnebrog, ., Berntsen, T.K., Dessens, O., Gauss, M., Grewe, V., Isaksen, I.S.A.,             Meinshausen, M., Plattner, G.-K., Reisinger, A., Segschneider, J., Shaffer, G.,
Koffi, B., Myhre, G., Olivie, D., Prather, M.J., Pyle, J.A., Stordal, F., Szopa, S.,            Steinacher, M., Strassmann, K., Tanaka, K., Timmermann, A., Weaver, A.J., 2013.
Tang, Q., van Velthoven, P., Williams, J.E., demark, K., 2011. Future impact of         Carbon dioxide and climate impulse response functions for the computation of
non-land based traffic emissions on atmospheric ozone and OH  an optimistic           greenhouse gas metrics: a multi-model analysis. Atmos. Chem. Phys. 13, 27932825.
scenario and a possible mitigation strategy. Atmos. Chem. Phys. 11, 11,29311,317.        https://doi.org/10.5194/acp-13-2793-2013.
https://doi.org/10.5194/acp-11-11293-2011.                                Kapadia, Z.Z., Spracklen, D.V., Arnold, S.R., Borman, D.J., Mann, G.W., Pringle, K.J.,
Hodnebrog, ., Berntsen, T.K., Dessens, O., Gauss, M., Grewe, V., Isaksen, I.S.A.,             Monks, S.A., Reddington, C.L., Benduhn, F., Rap, A., Scott, C.E., Butt, E.W.,
Koffi, B., Myhre, G., Olivie, D., Prather, M.J., Stordal, F., Szopa, S., Tang, Q., van         Yoshioka, M., 2016. Impacts of aviation fuel sulfur content on climate and human
Velthoven, P., Williams, J.E., 2012. Future impact of traffic emissions on               health. Atmos. Chem. Phys. 16, 1052110541. https://doi.org/10.5194/acp-16-
atmospheric ozone and OH based on two scenarios. Atmos. Chem. Phys. 12,             10521-2016.
12,21112,225. https://doi.org/10.5194/acp-12-12211-2012.                    Karcher, B., Burkhardt, U., Bier, A., Bock, L., Ford, I.J., 2015. The microphysical pathway
Holmes, C.D., Tang, Q., Prather, M.J., 2011. Uncertainties in climate assessment for the        to contrail formation. J. Geophys. Res. Atmos. 120, 78937927. https://doi.org/
case of aviation NO. Proc. Natl. Acad. Sci. U.S.A. 108 (27), 1099711002. https://        10.1002/2015JD023491.
doi.org/10.1073/pnas.1101458108.                                       Karcher, B., 2018. Formation and radiative forcing of contrail cirrus. Nat. Commun. 9,
Holmes, C.D., Prather, M.J., Svde, O.A., Myhre, G., 2013. Future methane, hydroxyl,         1824. https://doi.org/10.1038/s41467-018-04068-0.
and their uncertainties: key climate and emission parameters for future predictions.     Khodayari, A., Tilmes, S., Olsen, S.C., Phoenix, D.B., Wuebbles, D.J., Lamarque, J.-F.,
Atmos. Chem. Phys. 13, 285302. https://doi.org/10.5194/acp-13-285-2013.            Chen, C.-C., 2014b. Aviation 2006 NOx-induced effects on atmospheric ozone and
Hoor, P., Borken-Kleefeld, J., Caro, D., Dessens, O., Endresen, O., Gauss, M., Grewe, V.,        HOx in community Earth system model (CESM). Atmos. Chem. Phys. 14, 99259939.
Hauglustaine, D., Isaksen, I.S.A., Jockel, P., Lelieveld, J., Myhre, G., Meijer, E.,           https://doi.org/10.5194/acp-14-9925-2014.
Olivie, D., Prather, M., Schnadt-Poberaj, C., Shine, K.P., Staehelin, J., Tang, Q., van     Khodayari, A., Wuebbles, D.J., Olsen, S., Fuglestvedt, J.S., Berntsen, T., Lund, M.T.,
Aardenne, J., van Velthoven, P., Sausen, R., 2009. The impact of traffic emissions on        Waitz, I., Wolfe, P., Forster, P.M., Meinshausen, M., Lee, D.S., Lim, L.L., 2013.
atmospheric ozone and OH: Results from QUANTIFY. Atmos. Chem. Phys. 9,             Intercomparison of the capabilities of simplified climate models to project the effects
31133136. https://doi.org/10.5194/acp-9-3113-2009.                            of aviation CO2 on climate. Atmos. Environ. 75, 321328. https://doi.org/10.1016/
Hoose, C., Mohler, O., 2012. Heterogeneous ice nucleation on atmospheric aerosols: a         j.atmosenv.2013.03.055.
review of results from laboratory experiments. Atmos. Chem. Phys. 12, 98179854.     Khodayari, A., Olsen, S.C., Wuebbles, D.J., 2014a. Evaluation of aviation NOx-induced
https://doi.org/10.5194/acp-12-9817-2012.                                    radiative forcings for 2005 and 2050. Atmos. Environ. 91, 95103. https://doi.org/
Hough, A.M., 1989. The development of a two-dimensional global tropospheric model         10.1016/j.atmosenv.2014.03.044.
1. The model transport. Atmos. Environ. 23, 12351261. https://doi.org/10.1016/     Kohler, M.O., Radel, G., Dessens, O., Shine, K.P., Rogers, H.L., Wild, O., Pyle, J.A., 2008.
0004-6981(89)90150-9.                                                  Impact of perturbation of nitrogen oxide emissions from global aviation. J. Geophys.
Hough, A.M., 1991. Development of a two-dimensional global tropospheric model:           Res. Atmos. 113, D11305. https://doi.org/10.1029/2007JD009140.
model chemistry. J. Geophys. Res. Atmos. 96, 73257362. https://doi.org/10.1029/     Kohler, M.O., Radel, G., Shine, K.P., Rogers, H.L., Pyle, J.A., 2013. Latitudinal variation
90JD01327.                                                           of the effect of aviation NOx emissions on atmospheric ozone and methane and

27




































































D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
related climate metrics. Atmos. Environ. 64, 19. https://doi.org/10.1016/j.          Montzka, S.A., Spivakovsky, C.M., Butler, J.H., Elkins, J.W., Lock, L.T., Mondeel, D.J.,
atmosenv.2012.09.013.                                                   2000. New observational constraints for atmospheric hydroxyl on global and
Lamarque, J.-F., Bond, T.C., Eyring, V., Granier, C., Heil, A., Klimont, Z., Lee, D.,            hemispherical scales. Science 288, 500503. https://doi.org/10.1126/
Liousse, C., Mieville, A., Owen, B., Schultz, M.G., Shindell, D., Smith, S.J.,              science.288.5465.500.
Stehfest, E., van Aardenne, J., Cooper, O.R., Kainuma, M., Mahowald, N.,            Moore, R.H., Thornhill, K.L., Weinzierl, B., Sauer, D., D'Ascoli, E., Kim, J., et al., . Biofuel
McConnell, J.R., Naik, V., Riahi, K., van Vuuren, D.P., 2010. Historical (18502000)        blending reduces particle emissions from aircraft engines at cruise conditions.
gridded anthropogenic and biomass burning emissions of reactive gases and             Nature 543, 411415. https://doi.org/10.1038/nature21420.
aerosols: methodology and application. Atmos. Chem. Phys. 10, 70177039. https://     Myhre, G., Nilsen, J.S., Gulstad, L., Shine, K.P., Rognerud, B., Isaksen, I.S.A., 2007.
doi.org/10.5194/acp-10-7017-2010.                                         Radiative forcing due to stratospheric water vapor from CH4 oxidation. Geophys.
Lamquin, N., Stubenrauch, C.J., Gierens, K., Burkhardt, U., Smit, H., 2012. A global          Res. Lett. 34, L01807. https://doi.org/10.1029/2006GL027472.
climatology of upper tropospheric ice supersaturation occurrence inferred from the     Myhre, G., Kvalevg, M., Radel, G., Cook, J., Shine, K.P., Clark, H., Karcher, F.,
Atmospheric Infrared Sounder calibrated by MOZAIC. Atmos. Chem. Phys. 12,           Markowicz, K., Kardas, A., Wolkenberg, P., Balkanski, Y., Ponater, M., Forster, P.,
381405. https://doi.org/10.5194/acp-12-381-2012.                              Rap, A., de Leon, R.R., 2009. Intercomparison of radiative forcing calculations of
Lauder, A.R., Enting, I.G., Carter, J.O., Clisby, N., Cowie, A.L., Henry, B.K., Raupach, M.        stratospheric water vapour and contrails. Meteorol. Z. 18, 585596. https://doi.org/
R., 2012. Offsetting methane emissions  an alternative to emission equivalence         10.1127/0941-2948/2009/0411.
metrics. Int. J. Greenh. Gas Contr. 12, 419429. https://doi.org/10.1016/j.           Myhre, G., Shine, K.P., Radel, G., Gauss, M., Isaksen, I.S.A., Tang, Q., Prather, M.J.,
ijggc.2012.11.028.                                                      Williams, J.E., van Velthoven, P., Dessens, O., Koffi, B., Szopa, S., Hoor, P.,
Lee, D.S., Fahey, D., Forster, P.M., Newton, P.J., Wit, R.C.N., Lim, L.L., Owen, B.,            Grewe, V., Borken-Kleefeld, J., Berntsen, T.K., Fuglestvedt, J.S., 2011. Radiative
Sausen, R., 2009. Aviation and global climate change in the 21st century. Atmos.         forcing due to changes in ozone and methane caused by the transport sector. Atmos.
Environ. 43, 35203537. https://doi.org/10.1016/j.atmosenv.2009.04.024.              Environ. 45, 387394. https://doi.org/10.1016/j.atmosenv.2010.10.001.
Lee, D.S., Pitari, G., Grewe, V., Gierens, K., Penner, J.E., Petzold, A., Prather, M.,         Myhre, G., Shindell, D., Breon, F.-M., Collins, W., Fuglestvedt, J., Huang, J., Koch, D.,
Schumann, U., Bais, A., Berntsen, T., Iachetti, D., Lim, L.L., Sausen, R., 2010.            Lamarque, J.-F., Lee, D., Mendoza, B., Nakajima, T., Robock, A., Stephens, G.,
Transport impacts on atmosphere and climate: Aviation. Atmos. Environ. 44,            Takemura, T., Zhang, H., 2013. Anthropogenic and natural radiative forcing. In:
46784734. https://doi.org/10.1016/j.atmosenv.2009.06.005.                       Climate Change 2013: the Physical Science Basis, Contribution of Working Group I to
Le Quere, C., 76 others, 2018. Global carbon budget 2018. Earth Syst. Sci. Data 10,          the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.
21412194. https://doi.org/10.5194/essd-10-2141-2018.                          Cambridge University Press. https://www.ipcc.ch/report/ar5/wg1/.
Le Quere, C., Jackson, R.B., Jones, M.W., Smith, A.J.P., Abernethy, S., Andrew, R.M., De-     Newinger, C., Burkhardt, U., 2012. Sensitivity of contrail cirrus radiative forcing to air
Gol1, A.J., Willis, D.R., Shan, Y., Canadell, J.G., Friedlingstein, P., Creutzig, F.,           traffic scheduling. J. Geophys. Res. Atmos. 117, D10205. https://doi.org/10.1029/
Peters, G.P., 2020. Temporary reduction in daily global CO2 emissions during the         2011JD016736.
COVID-19 forced confinement. Nat. Clim. Change. https://doi.org/10.1038/s41558-     OECD, 2012. Green growth and the future of aviation. In: Paper Prepared for the 27th
020-0797-x.                                                           Round Table on Sustainable Development Held at OECD Headquarters 2324
Lim, L.L., Lee, D.S., Owen, B., Skowron, A., Matthes, S., Burkhardt, U., Dietmuller, S.,         January 2012. OECD. https://www.oecd.org/sd-roundtable/papersandpublicat
Pitari, G., Di Genova, G., Iachetti, D., Isaksen, I., Svde, O.A., 2015. REACT4C:           ions/49482790.pdf.
simplified mitigation study. In: TAC-4 Proceedings, June 22nd to 25th, 2015,         Olivie, D.J.L., Cariolle, D., Teyss`edre, H., Salas, D., Voldoire, A., Clark, H., Saintpp.
181185. Bad Kohlgrub. https://www.pa.op.dlr.de/tac/2015/Proceedings_of_         Martin, D., Michou, M., Karcher, F., Balkanski, Y., Gauss, M., Dessens, O., Koffi, B.,
TAC4_conference_final.pdf .                                                Sausen, R., 2012. Modeling the climate impact of road transport, maritime shipping
Liou, K.N., Takano, Y., Yue, Q., Yang, P., 2013. On the radiative forcing of contrail cirrus        and aviation over the period 18602100 with an AOGCM. Atmos. Chem. Phys. 12,
contaminated by black carbon. Geophys. Res. Lett. 40, 778784. https://doi.org/         14491480. https://doi.org/10.5194/acp-12-1449-2012.
10.1002/GRL.50110.                                                 Olsen, S.C., Brasseur, G.P., Wuebbles, D.J., Barrett, S.R.H., Dang, H., Eastham, S.D.,
Lund, M.T., Aamaas, B., Berntsen, T., Bock, L., Burkhardt, U., Fuglestvedt, J.S., Shine, K.        Jacobson, M.Z., Khodayari, A., Selkirk, H., Sokolov, A., Unger, N., 2013. Comparison
P., 2017. Emission metrics for quantifying regional climate impacts of aviation.           of model estimates of the effects of aviation emissions on atmospheric ozone and
Earth Syst. Dynam. 8, 547563. https://doi.org/10.5194/esd-8-547-2017.               methane. Geophys. Res. Lett. 40, 60046009. https://doi.org/10.1002/
Mahrt, F., Kilchhofer, K., Marcolli, C., Gronquist, P., David, R.O., Rosch, M.,                2013GL057660.
Lohmann, U., Kanji, Z.A., 2020. The impact of cloud processing on the ice nucleation     Penner, J.E., Chen, Y., Wang, M., Liu, X., 2009. Possible influence of anthropogenic
abilities of soot particles at cirrus temperatures. J. Geophys. Res. 125 https://doi.         aerosols on cirrus clouds and anthropogenic forcing. Atmos. Chem. Phys. 9,
org/10.1029/2019JD030922 e2019JD030922.                                  879896. https://doi.org/10.5194/acp-9-879-2009.
Maier-Reimer, E., Hasselmann, K., 1987. Transport and storage of CO2 in the oceanan     Penner, J.E., Zhou, C., Garnier, A., Mitchell, D.L., 2018. Anthropogenic aerosol indirect
inorganic ocean-circulation carbon cycle model. Clim. Dynam. 2, 6390. https://doi.        effects in cirrus clouds. J. Geophys. Res. Atmos. 123, 11,65211,677. https://doi.
org/10.1007/BF01054491.                                                org/10.1029/2018JD029204.
Matthes, M., Grewe, V., Dahlmann, K., Fromming, C., Irvine, E., Lim, L., Linke, F.,        Petzold, A., Gysel, M., Vancassel, X., Hitzenberger, R., Puxbaum, H., Vrochticky, S.,
Lhrs, B., Owen, B., Shine, K., Stromatas, S., Yamashita, H., Yin, F., 2017. A concept        Weingartner, E., Baltensperger, U., Mirabel, P., 2005. On the effects of organic
for multi-criteria environmental assessment of aircraft trajectories. Aerospace 4 (42).        matter and sulphur-containing compounds on the CCN activation of combustion
https://doi.org/10.3390/aerospace4030042.                                    particles. Atmos. Chem. Phys. 5, 31873203. https://doi.org/10.5194/acp-5-3187-
Markowicz, K.M., Witek, M.L., 2011. Simulations of contrail optical properties and           2005.
radiative forcing for various crystal shapes. J. Appl. Meteorol. Climatol. 50,          Pitari, G., Iachetti, D., Genova, G., De Luca, N., Svde, O.A., Hodnebrog, ., Lee, D.S.,
17401755. https://doi.org/10.1175/2011JAMC2618.1.                           Lim, L.L., 2015. Impact of coupled NOx/aerosol aircraft emissions on ozone
Marquart, S., Sausen, R., Ponater, M., Grewe, V., 2001. Estimate of the climate impact of        photochemistry and radiative forcing. Atmosphere 6, 751782. https://doi.org/
the cryoplanes. Aero. Sci. Technol. 5, 7384. https://doi.org/10.1016/S1270-9638        10.3390/atmos6060751.
(00)01084-1.                                                       Pitari, G., Cionni, I., Di Genova, G., Svde, O.A., Lim, L., 2017. Radiative forcing from
Mastrandrea, M.D., Mach, K.J., Plattner, G.K., Edenhofer, O., Stocker, T.F., Field, C.B.,         aircraft emissions of NOx: model calculations with CH4 surface flux boundary
Ebi, K.L., Matschoss, P.R., 2011. The IPCC AR5 guidance note on consistent             condition. Meteorol. Z. 26 (6), 663687. https://doi.org/10.1127/metz/2016/0776.
treatment of uncertainties: a common approach across the working groups. Climatic     Pomroy, H.R., Illingworth, J.A., 2000. Ice cloud inhomogeneity: quantifying bias in
Change 108, 675691. https://doi.org/10.1007/s10584-011-0178-6.                  emissivity from radar observations. Geophys. Res. Lett. 27, 21012104. https://doi.
Meinshausen, M., Smith, S.J., Calvin, K., Daniel, J.S., Kainuma, M.L.T., Lamarque, J.-F.,        org/10.1029/1999GL011149.
Matsumoto, K., Montzka, S.A., Raper, S.C.B., Riahi, K., Thomson, A., Velders, G.J.M.,     Ponater, M., Marquart, S., Sausen, R., Schumann, U., 2005. On contrail climate
van Vuuren, D.P.P., 2011. The RCP greenhouse gas concentrations and their             sensitivity. Geophys. Res. Lett. 32, L10706. https://doi.org/10.1029/
extensions from 1765 to 2300. Climatic Change 109, 213241. https://doi.org/          2005GL022580.
10.1007/s10584-011-0156-z.                                            Ponater, M., Pechtl, S., Sausen, R., Schumann, U., Httig, G., 2006. Potential of the
Millar, R.J., Nicholls, Z.R., Friedlingstein, P., Allen, M.R., 2017. A modified impulse-          cryoplane technology to reduce aircraft climate impact: a state-of-the-art assessment.
response representation of the global near-surface air temperature and atmospheric        Atmos. Environ. 40, 69286944. https://doi.org/10.1016/j.atmosenv.2006.06.036.
concentration response to carbon dioxide emissions. Atmos. Chem. Phys. 17,          Ponater, M., Bickel, M., Bock, L., Burkhardt, U., 2020. Towards determining the efficacy
72137228. https://doi.org/10.5194/acp-17-7213-2017.                           of contrail cirrus. In: Matthes, S., Blum, A. (Eds.), Making Aviation Environmentally
Miller, M., Brook, P., Eyers, C., 2010. Reduction of Sulphur Limits in Aviation Fuel           Sustainable, 3rd ECATS Conference, Book of Abstracts, vol. 1, ISBN 978-1-910029-
Standards (SULPHUR). EASA Research Project EASA.2008/C11. European Aviation        58-9, 51-44. http://www.ecats-network.eu/uploads/2020/06/ECATS_Main_B
Safety Agency. https://www.easa.europa.eu/sites/default/files/dfu/2009-SULPHU         ookOfAbstracts_Vol1_final.pdf .
R-Reduction%20of%20sulphur%20limits%20in%20aviation%20fuel%20standards      Prather, M.J., 1994. Lifetimes and eigenstates in atmospheric chemistry. Geophys. Res.
-Final%20Report.pdf.                                                    Lett. 21, 801804. https://doi.org/10.1029/94GL00840.
Minnis, P., Bedka, S.T., Duda, D.P., Bedka, K.M., Chee, T., Ayers, J.K., Palikonda, R.,       Prather, M., Ehhalt, D., Dentener, F., Derwent, R., Dlugokencky, E., 2001. Atmospheric
Spangenberg, D.A., Khlopenkov, K.V., Boeke, R., 2013. Linear contrail and contrail        chemistry and greenhouse gases. In: Houghton, J.T. (Ed.), Climate Change 2001: the
cirrus properties determined from satellite data. Geophys. Res. Lett. 40, 32203226.        Scientific Basis, Contribution of Working Group I to the Third Assessment Report of
https://doi.org/10.1002/grl.50569.                                          the Intergovernmental Panel on Climate Change. Cambridge University Press,
Mohler, O., Bttner, S., Linke, C., Schnaiter, M., Saathof, H., Stetzer, O., Wagner, R.,          Cambridge, United Kingdom and New York, NY, USA, pp. 239287. https://www.
Kramer, M., Mangold, A., Ebert, V., Schurath, U., 2005. Effect of sulfuric acid coating        ipcc.ch/site/assets/uploads/2018/03/TAR-04.pdf.
on heterogenous ice nucleation by soot aerosol particles. J. Geophys. Res. 110 (D11)     Prather, M.J., Holmes, C.D., Hsu, J., 2012. Reactive greenhouse gas scenarios: systematic
https://doi.org/10.1029/2004JD005169.                                      exploration of uncertainties and the role of atmospheric chemistry. Geophys. Res.
Lett. 39, L09803. https://doi.org/10.1029/2012GL051440.

28






















































D.S. Lee et al.                                                                                         Atmospheric Environment 244 (2021) 117834
Rap, A., Forster, P.M., Haywood, J.M., Jones, A., Boucher, O., 2010. Estimating the        Stevenson, D.S., Johnson, C.E., Collins, W.J., Derwent, R.G., Shine, K.P., Edwards, J.M.,
climate impact of linear contrails using the UK Met Office climate model. Geophys.        1998. Evolution of tropospheric ozone radiative forcing. Geophys. Res. Lett. 25,
Res. Lett. 37, L20703. https://doi.org/10.1029/2010GL045161.                      38193822. https://doi.org/10.1029/1998GL900037.
Revelle, R., Suess, H.E., 1957. Carbon dioxide exchange between atmosphere and ocean     Stevenson, D.S., Doherty, R.M., Sanderson, M.G., Collins, W.J., Johnson, C.E.,
and the question of an increase of atmospheric CO2 during the past decades. Tellus 9,        Derwent, R.G., 2004. Radiative forcing from aircraft NOx emissions: mechanisms and
1827. https://doi.org/10.1111/j.2153-3490.1957.tb01849.x.                       seasonal dependence. J. Geophys. Res. Atmos. 109, D17307. https://doi.org/
Richardson, T.B., Forster, P.M., Smith, C.J., Maycock, A.C., Wood, T., Andrews, T.,           10.1029/2004JD004759.
Boucher, O., Faluvegi, G., Flaschner, D., Hodnegrog, ., Kasoar, M., Kirkevg, A.,      Stordal, F., Gauss, M., Myhre, G., Mancini, E., Hauglustaine, D.A., Kohler, M.O.,
Lamarque, J.-F., Mlmenstadt, J., Myhre, G., Olivie, D., Portmann, R.W., Samset, B.        Berntsen, T., Stordal, E.J.G., Iachetti, D., Pitari, G., Isaksen, I.S.A., 2006. TRADEOFFs
H., Shawki, D., Shindell, D., Stier, P., Takemura, T., Voulgarakis, A., Watson-            in climate effects through aircraft routing: forcing due to radiatively active gases.
Parris, D., 2019. Efficacy of climate forcings in PDRMIP models. J. Geophys. Res.:         Atmos. Chem. Phys. Discuss. 6, 1073310771.
Atmospheres 124. https://doi.org/10.1029/2019JD030581.                      Stuber, N., Ponater, M., Sausen, R., 2005. Why radiative forcing might fail as a predictor
Righi, M., Hendricks, J., Sausen, R., 2013. The global impact of the transport sectors on        of climate change. Clim. Dynam. 24, 497510. https://doi.org/10.1007/s00382-
atmospheric aerosol: simulations for year 2000 emissions. Atmos. Chem. Phys. 13,        004-0497-7.
99399970. https://doi.org/10.5194/acp-13-9939-2013.                        Stuber, N., Forster, P., Radel, G., Shine, K., 2006. The importance of the diurnal and
Sander, S.P., Friedl, R.R., Ravishankara, A.R., Golden, D.M., Kolb, C.E., Kurylo, M.J.,          annual cycle of air traffic for contrail radiative forcing. Nature 441, 864867.
Molina, M.J., Moortgat, G.K., Finlayson-Pitts, B.J., 2006. Chemical Kinetics and          https://doi.org/10.1038/nature04877.
Photochemical Data for Use in Atmospheric Studies. JPL Publ, 062, No. 15. https     Teoh, R., Schumann, U., Majumdat, A., Stettler, M.E.J., 2020. Mitigating the climate
://jpldataeval.jpl.nasa.gov/pdf/JPL_02-25_rev02.pdf.                              forcing of aircraft contrails by small-scale diversions and technology adoption.
Sausen, R., Schumann, U., 2000. Estimates of the climate response to aircraft CO2 and         Environ. Sci. Technol. 54, 29412950. https://doi.org/10.1021/acs.est.9b05608.
NOx emissions scenarios. Climatic Change 44, 2758.                          Teoh, R., Stettler, M.E.J., Majumdar, A., Schumann, U., Graves, B., Boies, A.M., 2019.
Sausen, R., Isaksen, I., Grewe, V., Hauglustaine, D., Lee, D.S., Myhre, G., Kohler, M.O.,        A methodology to relate black carbon particle number and mass emissions.
Pitari, G., Schumann, U., Stordal, F., Zerefos, C., 2005. Aviation radiative forcing in        J. Aerosol Sci. 132, 4459. https://doi.org/10.1016/j.jaerosci.2019.03.006.
2000: an update on IPCC (1999). Meteorol. Z. 14, 555561. https://doi.org/          Tesche, M., Achtert, P., Glantz, P., Noone, K.J., 2016. Aviation effects on already-existing
10.1127/0941-2948/2005/0049.                                            cirrus clouds. Nat. Commun. 7, 12016. https://doi.org/10.1038/ncomms12016.
Schumann, U., Penner, J.E., Chen, Y., Zhou, C., Graf, K., 2015. Dehydration effects from     UKDS, 2016. http://stats.ukdataservice.ac.uk/index.aspx?r=349678&DataSetCode=IE
contrails in a coupled contrail-climate model. Atmos. Chem. Phys. 15, 1117911199.        A_COAL_BA, 2016.
https://doi.org/10.5194/acp-15-11179-2015.                                UNFCCC. https://unfccc.int/nationally-determined-contributions-ndcs.
Schumann, U., Baumann, R., Baumgardner, D., Bedka, S.T., Duda, D.P.,                Unger, N., Bond, T.C., Wang, J.S., Koch, D.M., Menon, S., Shindell, D.T., Bauer, S., 2010.
Freudenthaler, V., Gayet, J.-F., Heymsfield, A.J., Minnis, P., Quante, M., Raschke, E.,        Attribution of climate forcing to economic sectors. Proc. Natl. Acad. Sci. U.S.A. 107,
Schlager, H., Vazquez-Navarro, M., Voigt, C., Wang, Z., 2017. Properties of              33823387. https://doi.org/10.1073/pnas.0906548107.
individual contrails: a compilation of observations and some comparisons. Atmos.      Unger, N., 2011. Global climate impact of civil aviation for standard and desulfurized jet
Chem. Phys. 17, 403438. https://doi.org/10.5194/acp-17-403-2017.                  fuel. Geophys. Res. Lett. 38, 16. https://doi.org/10.1029/2011GL049289.
Shine, K.P., Fuglestvedt, J.S., Hailemariam, K., Stuber, N., 2005. Alternatives to the        Unger, N., Zhao, Y., Dang, H., 2013. Mid-21st century chemical forcing of climate by the
global warming potential for comparing climate impacts of emissions of greenhouse        civil aviation sector. Geophys. Res. Lett. 40, 641645. https://doi.org/10.1002/
gases. Climatic Change 68, 281302. https://doi.org/10.1007/s10584-005-1146-9.        grl.50161.
Skeie, R.B., Fuglestvedt, J., Berntsen, T., Peters, G.P., Andrew, R., Allen, M.,             Unterstrasser, S., 2014. Large-eddy simulation study of contrail microphysics and
Kallbekken, S., 2017. Perspective has a strong effect on the calculation of historical        geometry during the vortex phase and consequences on contrail-to-cirrus transition.
contributions to global warming. Environ. Res. Lett. 12 https://doi.org/10.1088/         J. Geophys. Res. Atmos. 119, 75377555. https://doi.org/10.1002/2013JD021418.
1748-9326/aa5b0a, 024022.                                            Voulgarakis, A., Naik, V., Lamarque, J.-F., Shindell, D.T., Young, P.J., Prather, M.J.,
Skowron, A., Lee, D.S., Hurley, J., 2009. "Aviation NOx global warming potential. In: 2nd        Wild, O., Field, R.D., Bergmann, D., Cameron-Smith, P., Cionni, I., Collins, W.J.,
International Conference on Transport, Atmosphere and Climate, 2528 June 2009.        Dalsren, S.B., Doherty, R.M., Eyring, V., Faluvegi, G., Folberth, G.A., Horowitz, L.
Aachen/Maastricht, Germany/Netherlands. https://www.pa.op.dlr.de/tac/200            W., Josse, B., McKenzie, I.A., Nagashima, T., Plummer, D.A., Righi, M., Rumbold, S.
9/proceedings/FB2010-10.pdf.                                             T., Stevenson, D.S., Strode, S.A., Sudo, K., Szopa, S., Zeng, G., 2013. Analysis of
Skowron, A., Lee, D.S., de Leon, R.R., 2013. The assessment of the impact of aviation NOx        present-day and future OH and methane lifetime in the ACCMIP simulations. Atmos.
on ozone and other radiative forcing responsesThe importance of representing           Chem. Phys. 13, 25632587. https://doi.org/10.5194/acp-13-2563-2013.
cruise altitudes accurately. Atmos. Environ. 74, 159168. https://doi.org/10.1016/j.     Wilcox, L., Shine, K.P., Hoskins, B.J., 2012. Radiative forcing due to aviation water
atmosenv.2013.03.034.                                                   vapour emissions. Atmos. Environ. 63, 113. https://doi.org/10.1016/j.
Skowron, A., Lee, D.S., de Leon, R.R., 2015. Variation of radiative forcings and global         atmosenv.2012.08.072.
warming potentials from regional aviation NOx emissions. Atmos. Environ. 104,       Wild, O., Prather, M.J., Akimoto, H., 2001. Indirect long-term global radiative cooling
6978. https://doi.org/10.1016/j.atmosenv.2014.12.043.                          from NOx emissions. Geophys. Res. Lett. 28, 17191722. https://doi.org/10.1029/
Svde, O.A., Matthes, S., Skowron, A., Iachetti, D., Lim, L., Owen, B., Hodnebrog, ., Di        2000GL012573.
Genova, G., Pitari, G., Lee, D.S., Myhre, G., Isaksen, I.S.A., 2014. Aircraft emission     Xie, B., Zhang, H., Wang, Z., Zhao, S., Fu, Q., 2016. A modelling study of effective
mitigation by changing route altitude: a multi-model estimate of aircraft NOx            radiative forcing and climate response due to tropospheric ozone. Adv. Atmos. Sci.
emission impact on O3 photochemistry. Atmos. Environ. 95, 468479. https://doi.        33, 819828. https://doi.org/10.1007/s00376-016-5193-0.
org/10.1016/j.atmosenv.2014.06.049.                                     Yin, F., Grewe, V., Fromming, C., Yamashita, H., 2018. Impact on flight trajectory
Smith, C.J., Kramer, R.J., Myhre, G., Forster, P.M., Soden, B.J., Andrews, T., Boucher, O.,        characteristics when avoiding the formation of persistent contrails for transatlantic
Faluvegi, G., Flaschner, D., Hodnebrog, ., Kasoar, M., Kharin, V., Kirkevg, A.,          routes. Transport. Res. Transport Environ. 65, 466484. https://doi.org/10.1016/j.
Lamarque, J.-F., Mlmenstadt, J., Olivie, D., Richardson, T., Samset, B.H.,              trd.2018.09.017.
Shindell, D., Stier, P., Takemura, T., Voulgarakis, A., Watson-Parris, D., 2018.         Zhou, C., Penner, J.E., 2014. Aircraft soot indirect effect on large-scale cirrus clouds: is
Understanding rapid adjustments to diverse forcing agents. Geophys. Res. Lett. 45         the indirect forcing by aircraft soot positive or negative? J. Geophys. Res. Atmos.
https://doi.org/10.1029/2018GL079826.                                      119, 11,30311,320. https://doi.org/10.1002/2014JD021914.











29

Limitations of Translatable Documents

PDF files are created with text and images are placed at an exact position on a page of a fixed size.
Web pages are fluid in nature, and the exact positioning of PDF text creates presentation problems.
PDFs that are full page graphics, or scanned pages are generally unable to be made accessible, In these cases, viewing whatever plain text could be extracted is the only alternative.