You are on page 1of 15

November 27, 2000

Report on the European Workshop


“Aviation, Aerosols, Contrails and Cirrus Clouds” (A2C3),
Seeheim near Frankfurt/Main, July 10-12, 2000

Aircraft engine emissions cause the formation of small aerosol particles and contrails in the
upper troposphere and lower stratosphere, which - together with other aerosols - may
influence cirrus cloud formation (Figure 1, 2). The special report by the Intergovernmental
Panel on Climate Change (IPCC) on "Aviation and the Global Atmosphere" (IPCC, 1999)
concluded that the largest uncertainties limiting our ability to project aviation impacts on
climate and ozone comes from the influence of contrails and aerosols on cloud changes
(Figure 3). Under guidance of the World Meteorological Organisation (WMO), the ”Scientific
Assessment of Ozone Depletion: 1998” was prepared (WMO, 1999) which notes incomplete
understanding of the influence of aerosol and cloud particles on ozone in the lower
stratosphere and the upper troposphere. The Second (IPCC, 1996) and the forthcoming Third
Assessment Report of IPCC on ”Climate Change” note large uncertainties in quantifying the
direct and indirect effects of aerosols and changes in cloudiness on climate. At present, less is
known about the impact of aerosols and contrail-induced cirrus changes than about the impact
of gaseous emissions. Many of the emissions of aircraft engines may be reduced with more
advanced technology. But what can be done against contrails and cirrus changes? Therefore
this workshop concentrate on the aerosol and contrail issues.

The three-days European workshop “Aviation, Aerosols, Contrails and Cirrus Clouds” (A2C3)
took place in Seeheim, near Frankfurt/Main, Germany, July 10-12, 2000. The workshop
started with two days with four scientific sessions followed by one day with working group
meetings and discussions. 26 invited oral papers and 55 poster contributions were presented
and discussed by the 96 participants from Europe, Russia, and the USA, with representatives
from atmospheric sciences, combustion sciences, industry, and various agencies. The first
session provided the frame of reference for the following discussions by reviewing results of
previous assessments and by presenting results from scientific work in related fields. The
following three sessions were devoted to the specific workshop topics: 1) particle formation
and properties; 2) contrails and cloud changes; and 3) parameters effecting aviation impact.
The third day was dedicated to discussions at the posters, working groups and a closing panel
session. The proceedings (Schumann and Amanatidis, 2000) are in press and will appear in
the series of Air Pollution Research Reports of the European Commission. Main results of the
papers presented in the proceedings and in posters during the workshop are summarised
briefly below.

Black-Carbon and Nonvolatile Soot Particles

Aircraft soot particles (composed of black carbon and other non-volatile materilas) may act as
condensation and freezing nucleus for ice particle formation.

Soot emissions have been measured in altitude test facilities (A. Döpelheuer and C. Wahl, D.
E. Hagen et al.) and in exhaust plumes and contrails behind cruising aircraft (A. Petzold et
al.). Such measurements have been used to validate the soot calculation procedures which are
necessary to set up a 3d global aircraft-generated soot inventory (Figure 4). Aircraft emit on
average 0.04 g black carbon per kg of burnt fuel; 40% of all soot gets emitted above 10 km
altitude, mainly over North America and Europe (A. Döpelheuer and C. Wahl). The number
and mass of soot particles emitted per kg of burnt fuel by modern engines is lower (typically
0.01 g/kg, 3×1014 particles/kg fuel) than in older ones (typically 0.1 g/kg and 2×1015/kg). The

1
November 27, 2000

soot emission mass and number is independent of fuel sulphur content. Primary soot particles
are typically 25 to 60 nm large in diameter, soot agglomerates reach larger than 150 nm in
size (A. Petzold et al.).

Aircraft engine soot may not be as hydrophobic as assumed in the past. High initial water
uptake has been measured on aviation soot. Such soot particles acquire a substantial fraction
of a water monolayer already in the very young exhaust plume, which would explain contrail
formation independent of fuel sulphur content (O.Popovitcheva et al.).

The activation of soot particles from a graphite spark gap generator with mobility equivalent
diameters of 50 to 200 nm has been investigated in a new large aerosol chamber (AIDA in
Karlsruhe) at 255 and 239 K. For 255 K, the soot was observed to become activated to form
droplets when reaching liquid saturation; for 239 K, the droplets freeze and evaporate quickly
when the humidity drops below ice saturation (A. Nink, O. Möhler et al.).

The main processes leading to soot particles are now known and can be reasonably
approximated in models at least for laminar flame situations. However, a reliable technique
for quantitative calculation of soot emissions from engine combustors has still to be
developed. The most important contribution to soot comes from surface growth. Particle
interception and coagulation dominate the particle dynamics. Large amounts of soot forms in
the early combustion process but most of it gets oxidised before leaving the combustion
chamber. Soot prediction within aero-gas turbine combustion chambers require soot models
coupled to three-dimensional (3d) fluid dynamics and radiation transfer calculations (H.
Bockhorn, H. Brocklehurst, and D. Hu, P. Frank et al.).

Chemiions (CI) and Volatile particles

Large numbers (about 2.4×1017 per kg of burnt fuel) of volatile particles form behind aircraft
at cruise. The volatile particles may contribute to contrail and cirrus particle formation (Figure
5). The size spectrum of particles from 3 nm to 100 nm has been measured with a battery of
condensation nucleus counters of different size sensitivities; size spectra of larger particles are
measured with optical probes (A. Petzold et al.; F. Schröder). Smaller particles of different
masses are detectable as chemiions (CI).

Engines emit large numbers of positive and negative chemiions. Chemiions are produced by
chemistry in the combustors. They may enhance the growth of volatile particles, and hence
influence the formation of ice particles. Chemiions may have large sizes, possibly more than
8500 atomic mass units, and hence may reach up to the smallest condensation nuclei
measurable today (about 3 nm) (F. Arnold et al., K.-H. Wohlfrom et al.). Very massive CI
were observed even for nearly sulphur free fuels. However the number of negative CI grows
with fuel sulphur content (FSC). Negative CI are larger in mass than positive CI (K.-H.
Wohlfrom et al.). Negative CI contain HSO4-H2SO4 cluster ions. Positive CI have been
identified as at least partly oxygenated hydrocarbons (A. Kiendler et al.). Models explain the
measured number of chemiions in hot aircraft exhaust plumes by ion-ion-recombination and
initial ion concentrations of the order of 108 cm-3 (A. Sorokin et al.; F. Arnold et al.).

The concentration of volatile particles strongly increases as diameter decreases toward the
sizes of the large molecular clusters. The number of volatile particles larger than 3 nm
detectable in a young (order 0.5 to 10 s) plume increases with FSC for FSC in the range from
60 to 1000 µg/g (Figure 6). Particle formation does not change significantly when the FSC
gets reduced below 60 µg/g, apparently because of condensable hydrocarbons contributing to

2
November 27, 2000

the formation of volatile aerosols in the exhaust plume (Petzold et al.). The variation in the
number of measured volatile particles behind several aircraft can be explained with a model in
terms of variations in plume age, size detection limit of the particle counters, FSC, the amount
of condensable organic material, and, most important, with the number of CI emitted (B.
Kärcher et al.).

Sulphuric acid

Part of the sulphur contained in the aircraft fuel may get converted to sulphuric acid which
affects the formation of volatile particles and the hydration properties of soot. The conversion
fraction of fuel sulphur to S(VI) (SO3, H2SO4) depends on pressure and temperature at
combustor exit and on the residence times. The conversion is largest at a temperature of about
1000 K and high pressures. Cooling of turbine blades in the high pressure turbine enhances
the conversion (R. Miake-Lye et al.). In flight measurements indicate that around 3 % of FSC
gets converted to sulphuric acid, less than believed a few years ago (F. Arnold et al., K.-H.
Wohlfrom et al.).

Nitric acid and other aerosol consitutents

H2O-H2SO4 aerosols take up HNO3 only at small mole fractions (<1%) and only for
temperatures below 222 K and high ambient humidity (G. Gleitsmann and R. Zellner). The
uptake coefficient of HNO3 (order 10-3 mg-1) and H2O (5×10-4 mg-1) on soot have been
measured. Uptake of HNO3 seems to be of little importance for volatile aerosols, soot
activation, and contrail formation in aircraft exhaust plumes (T. Keil et al.). The uptake of
NOy on cirrus cloud particles was inferred from measurements of gas- and particle phase NOy,
H2O, HNO3, and aerosols for temperatures near 200 K. About 2 to 10% of the NOy gets
removed by a coverage of ice particles with about 0.1 monolayers of NOy (H. Schlager et al.).
A thermodynamic model of hydrogen-ammonium-sulphate-nitrate-water aerosols has been
provided for upper tropospheric temperatures (S. Clegg and P. Brimblecombe). The non-
reactive uptake of HCl on frozen film ice surfaces was determined for temperatures from 205
– 235 K. The uptake coefficient decreases substantially when T increases from 205 to 215 K
(R. G. Hynes et al.).

Global aerosol distribution

A global stratospheric and tropospheric sulphate aerosol model has been developed, compared
to observations from the SAGE satellite instrument in terms of extinction, optical depth, mass
and surface area density, and applied to study future changes of stratospheric aerosols in
relation to projected changes of anthropogenic sulphur emissions. Subsonic aircraft were
found to contribute about 3% to the total aerosols mass in the stratosphere and 20 % of the
surface area density at 14 km altitude north of 45°N (G. Pitari and E. Mancini).

Contrails

Contrails trigger cirrus cloud formation. In air masses with ice-supersaturation of a few
percent no cirrus cloud would have formed otherwise, see Figure 2 (U. Schumann). Aerosols
from aircraft may influence cirrus formation, possibly long after the time of emission. The
crucial contrail parameters are their cover and optical properties. Contrails may warm the
Earth surface in particular over warm and bright surfaces. During day, the enhanced Earth’s
albedo caused by contrails may cause a cooling. Around 17% of the Earth is covered with air
masses which are ice-supersaturated and cold enough so that persistent contrails form when

3
November 27, 2000

aircraft fly in these region. The area coverage by such regions susceptible to contrails is 21%
when aircraft would burn liquid hydrogen instead of kerosene (K. Gierens and S. Marquart).
Several recent studies show that supersaturation with respect to ice is very common in the
upper troposphere (K. Gierens, E. Jensen, J. Ovarlez et al.), see Figure 7. In the Arctic winter,
supersaturation occurs even in the lowermost stratosphere due to the extremely low
temperatures in this region. Also the tropical upper troposphere is often ice supersaturated.
Cirrus clouds rarely extend more than 1 km above the tropopause (E. Jensen et al.). Large ice
supersaturation also occurs within cirrus clouds (J. Ovarlez et al.).

Short lived contrails of 2-engined aircraft were observed to evaporate earlier than contrails
from 4-engines aircraft, consistent with model calulations (K. Gierens and R. Sussmann).

The microphysical properties of contrails and cirrus clouds have been measured during more
than 20 airborne missions over Central Europe at temperatures below –50°C. Contrails
contain large concentrations of nearly spherical ice crystals, initially up to 105 cm-3, later
diluted to typically a few 100 cm-3, while the mean diameter range from 1-10 µm. Young
cirrus clouds contain mostly regularly shaped ice crystals in the size range of 10-20 µm at
typical concentrations of 2-5 cm-3. Contrail growth is affected only weakly by pre-existing
cirrus clouds (F. Schröder).

In contrails and cirrus clouds, the scattering phase function has been measured in situ with a
Polar Nephelometer. The scattering by cold ice clouds differs strongly from that of spherical
particles, implying 40% higher albedo, i.e. a stronger shortwave cooling by contrails than
computed for spherical particles (F. Auriol et al.).

Particle measurement systems may be strongly disturbed by electromagnetic interferences.


Older probes with slower electronics may have difficulties in registering crystal size and
number density accurately (E. Raschke et al.).

Contrails have been estimated to cover currently about 0.1% of the global sky (K. Gierens and
S. Marquart). Over the USA, the mean contrail coverage has been determined from 4 months
of satellite data to be 1.8%. For 55 contrails over different backgrounds the mean optical
depth and the effective particle diameter have been determined and found to be 0.46 and 36
µm, respectively, see Figure 10 (P. Minnis et al.). Over central Europe, the annual daytime
average contrail has been determined from 2 years of NOAA-14 satellite data with an
automated contrail detection algorithm. The annual day time average cover is 0.7% (Figure 9)
with a strong annual cycle (1% in winter, 0.4% in summer). The mean visible optical depth in
this region is determined from the same data to be 0.11. During night the cover is 3 time
smaller. The resultant radiative forcing is found to be one order of magnitude below the
results of previous assessments. (R. Meyer et al.). Data from traffic management systems
indicate strong diurnal variations in air traffic, which is important in assessing the mean
radiative forcing by contrails; contrails warm during night but may cool during day time (S.
Baughcum et al.).

A case study of a contrail has been reported which has been measured from ground with a
Lidar, and in situ with particle size spectrometers. Pyranometers and pyrgeometers were used
to measure the radiative shortwave and longwave vertical fluxes near the contrail. The solar
optical depth of the contrail varies from 0.05 to 0.23. The measured radiation fluxes indicate
smaller radiative forcing than what is computed for the measured size spectrum assuming
spherical ice particles (P. Wendling et al.).

4
November 27, 2000

Gas phase chemistry and heterogeneous reactions on contrail and cirrus particles

Aerosols and ice particles contribute to changes in ozone and other air constituents. NOx
emissions by subsonic aircraft lead to net ozone production in the tropopause region.
However, within contrails, halogen components may get activated and NOx and hydroxyl
radicals may get reduced, causing less ozone production, and ozone depletion is possible at
least locally within contrails. In the lowermost stratosphere, ozone destruction is caused by
heterogeneous chlorine activation on contrail ice particles, whereas in the upper troposphere
ozone production is compensated by heterogeneous reduction of OH radicals (S. Meilinger et
al.). A chemistry box model study shows a large sensitivity of upper tropospheric chemistry
to heterogeneous processes in cirrus clouds. Ice particle sedimentation may reduce upper
tropospheric HNO3, NOx, and OH, resulting in significant ozone loss (A. Meier and J.
Hendricks). A 3d chemical transport model (CTM) has been used to investigate the role of
subvisible ice clouds in chemical ozone loss near the tropopause. The calculated zonal mean
ozone loss reaches a maximum in the mid- and high latitude tropopause region of about 1% in
winter and 3% in summer (B. Bregman et al.).

Seven years of ozone, humidity and temperature profile observations at 21°S, 55°E reveal a
systematic diminution of ozone abundance in tropical cirrus clouds, possibly due to ice cloud
induced denoxification (S. Roumeau et al).

Only a few attempts have been made to model the formation and microphysics of cirrus
clouds (W. Wobrock and A. Flossmann; T. Chourlaton et al.). From measurements and
models, it was found that aerosols which passed through cirrus behave differently from
aerosol that was not yet contained in frozen particles (T. Choularton). Large aerosol may act
as ice nuclei by homogeneous freezing, small aerosols may enhance freezing by acting as
heterogeneous ice nucleus (W. Wobrock). Bulk and size-resolved cirrus process-models show
large differences even when applied to idealised cases with prescribed ambient conditions;
particle sedimentation rate is one of the crucial model parameters (GEWEX-cirrus modelling
initiative, poster of D. O’C. Starr).

Volatile aerosol freezes for temperatures, T, in the range 233.5 K < T < 273.15 K, when the
ambient air reaches liquid saturation. Below 233 K, homogeneous ice nucleation occurs in
particles of 0.2 µm radius within one minute when the relative humidity over ice (RHI)
exceeds RHI = 238.7 – 0.398 T (RHI in %, T in K; Koop et al.).

The ECHAM4 global circulation model (GCM) has been used to study the sensitivity of ice
clouds to different assumptions about ice nuclei. Ice nuclei are assumed to be either just a
function of temperature and altitude, insoluble carbonaceous aerosols or dust aerosols. The
different models cause variations in the longwave cloud forcing by 4 W m-2 (U. Lohmann). A
first GCM study with parameterised contrails was presented, showing variations in the cover
and optical depth of contrails with latitude, altitude, and season; the optical depth of contrails
appears to be larger over the USA than over Europe because of different temperature and
humidity conditions (M. Ponater et al.).

Lidar observations in Southern France (44°N, 6°E) have been used to determine the
occurrence frequency and altitude range of thin cirrus at various optical depths. Subvisible
cirrus (532 nm optical depth < 0.03) constitute a substantial portion (40%) of the calculated
mean optical depth. Fewer than 5% of the cirrus occurrences are (at most <1 km) above the
tropopause (L. Goldfarb et al.).

5
November 27, 2000

For the first time, in situ observations of cirrus clouds have been performed in the Southern
hemisphere during the Interhemispheric Differences in Cirrus Properties From Anthropogenic
Emissions (INCA) experiment in March/April 2000. An extensive scientific payload to
characterise aerosol-cloud interactions, cloud microphysical properties, and trace gases was
deployed on the DLR research aircraft Falcon. A second experiment in the Northern
Hemisphere midlatitudes has been performed in Sept./Oct. 2000 to compare clean and
polluted air masses in order to study whether anthropogenic emissions have a measurable
effect on cirrus cloud properties. First results were presented showing clear interhemispheric
differences in various aerosol components (volatile and non-volatile particles of various sizes;
crystal residuals and interstitial aerosols; see Figure 8) and trace gases (CO, O3, NO, and
NOy) (J. Ström et al.). The observations are accompanied by ground based Lidar (Immler et
al.) and satellite observations (P. Minnis et al.). At Punta Arenas the mean cirrus cover was
45%. On the way from Punta Arenas to Bremerhaven, the Lidar was mounted on the
POLARSTERN and observed extended cirrus clouds in the tropics (F. Immler et al.).

Climate Impact

Little progress has been made beyond that described in the IPCC report in understanding
which aviation parameter is most important in controlling the climate impact by contrails and
cirrus. The total aviation climate impact (in terms of radiative forcing, and without cirrus
changes) is about three times as large as the radiative forcing from aircraft-induced carbon
dioxide alone (R. Sausen; U. Schumann), see Figure 3 (IPCC,1999).

The climate response is more sensitive to aircraft induced ozone than to a CO2 perturbation
with the same radiative forcing. Global warming leads to a smaller increase in contrail
coverage than would be expected from growing air traffic in an unchanged atmosphere (R.
Sausen). Some observations indicate long-term cirrus cover trends (P. Minnis et al.). In
general, long term observations are needed to identify any systematic cloud or climate change
due to aviation, if present, because of the large variability of cloud systems (V. Ramaswamy).

Mitigation Potentials

Contrails form for thermodynamic reasons and, hence, can be avoided only by flying in
warmer air either below or above the altitude range (around the tropopause at midlatitudes)
which is cool enough to let contrails form. Recent experiments have provided evidence that
contrails form at lower altitude and hence more frequently when using more efficient engines.
However, more efficient engines also emit less carbon dioxides. The trade-off between both
effects depends on the relative sensitivity of climate to contrails and carbon dioxide changes
and on the scenario of future aviation (U. Schumann).

Fuel properties have some impact on the formation of soot, volatile particles, and the carbon
dioxide and water vapour emissions. The impact of variations within various jet fuels for soot
formation is small compared to the variability between various combustor concepts (poster by
P. Madden). Elimination of fuel sulphur would reduce the size, and to minor degree also the
number of volatile aerosol particles formed. However, volatile particles would form even for a
very low fuel sulphur content (B. Kärcher et al.).

Hydrogen fuelled aircraft have been proposed in order to reduce carbon dioxide emissions by
aircraft. Very low NOx emissions from such aircraft have been demonstrated. The radiative
forcing by the emitted water vapour is small, but the impact on contrails and on cirrus clouds

6
November 27, 2000

has still to be determined (H. Klug and M. Ponater). Replacing the fleet of kerosene driven
aircraft by a fleet of liquid hydrogen driven aircraft would pay off in terms of reduced
radiative forcing by carbon dioxides only in the long term, after several decades (S.Marquart
et al.). Contrails behind liquid hydrogen driven engines may have smaller climatic impact
because of different particle properties (larger in size but less in number) compared to
contrails behind kerosene burning engines.

It is not yet known whether the climatic impact depends just on fuel consumption of the
aircraft flying in sufficiently cold and humid air masses, or whether other parameters are more
or equally important, such as the number of aircraft in operation, the number and size of soot
particles emitted, the sulphur content in the fuel used, or just the flight route. In the absence of
a reliable estimate of the climatic impact of contrails, it is not yet possible to assess various
flight routing and operational options (U. Schumann).

Ongoing and future research programmes

Several research programmes were described: Atmospheric research in Germany supported


by the Bundesministerium für Bildung und Forschung (G. Hahn); laboratory studies of
heterogeneous processes on ice (CUT-ICE; J. N. Crowley et al.); aerosol/particulate research
activities supported by the NASA ultra-efficient engine technology (UEET) program (C. C.
Wey); measurements of aged aircraft exhaust in the ACCENT mission (R. Friedl et al.);
environmental considerations in aircraft design (C. Hume); the CRYOPLANE project
considering the potential of liquid hydrogen as aircraft fuel (H. Klug); perspective of the U.S.
federal aviation administration (H. Wesoky); overview on NASA’s plans to address the
concerns identified in the IPCC special report (D. Anderson and R. Friedl); programmatic
perspectives of the European research on aviation’s atmospheric effects (G. Amanatidis); and
industry and agencies perspectives were presented in opening remarks and panel discussions
(J. Raps, F. Beulé, P. Newton, F. Walle). A series of SULFUR experiments (1-7) was
performed by the Deutsches Zentrum für Luft- und Raumfahrt (DLR) in Germany; DLR and
partners of other major national research centres (within the Hermann von Helmholtz-
Gemeinschaft Deutscher Forschungszentren, HGF) perform a 3-years project Particles from
Aircraft Engines and Their Impact on Contrails, Cirrus Clouds, and Climate (PAZI).

Important open questions addressed in the working groups include: The prediction of the
number of soot particles emitted from the fleet of aircraft; the hydration and ice nucleation
properties of soot and other aerosol components; the global cover and optical properties of
contrails and aircraft induced cirrus, and their long-term trends; the heterogeneous chemistry
on cirrus and contrail particles and their effects on air composition; the statistics of vertical
wind, humidity, and cirrus properties in the tropopause region; the identification of the most
relevant aerosols component which should get reduced for an environmentally sustainable air
traffic development.

Aerosols are key issue for understanding of aircraft impact, ozone chemistry and climate
changes. Aviation emissions are contributing a small but growing fraction of all
anthropogenic disturbances to the climate system. In view of the rather well defined
disturbance, understanding of aviation effects helps in understanding the more general
aerosol-cloud-climate-chemistry interactions from all sources.

7
November 27, 2000

Figures Captions:

Figure 1. Mean altitude of the tropopause in June and December, distribution of NOx
emission source rate from aircraft versus altitude and latitude (the emission rate is the larger
the darker the shaded area), and indications of mean circulation and some relevant processes.

Figure 2. Spiral contrail observed in NOAA-14 AVHRR satellite data west of Denmark
induced by a military aircraft circulating over the North Sea at 1236 UT May 22, 1998,
indicating that long-lasting cirrus clouds are triggered by contrails which would not form
otherwise (U. Schumann)

Figure 3. Radiative forcing (W m-2) due to aircraft emissions in 1992 (IPCC, 1999). The
columns indicate the best estimate of forcing. The error bars denote the range within which
the best-estimate value is expected with a two-third probability. The available information on
cirrus clouds is insufficient to determine either a best estimate or an uncertainty range; the
dashed line indicates a range of possible best estimates. The estimate for total forcing does not
include the effect of changes in cirrus cloudiness. The evaluations below the graph (“good,”
“fair,” “poor,” “very poor”) are a relative appraisal associated with each component and
indicates the level of scientific understanding.

Figure 4. Global aircraft-generated soot inventory (vertical integral of emissions per month in
1992) (A. Döpelheuer et al.).

Figure 5. Formation processes of aerosols in an aircraft exhaust plume (K.-H. Wohlfrom).

Figure 6. Apparent particle emission indices of particles larger diameter d (PEId, d = 3, 5, and
14 nm), measured in aircraft exhaust plumes that do not cause a contrail as a function of fuel
sulphur content (FSC) behind the ATTAS jet aircraft (squares), a Boeing 737-300 (circles),
and an Airbus 310-300 (diamonds) (A. Petzold et al.).

Figure 7. Relative humidity with respect to liquid water versus ambient temperature. The
horizontal dashed line denotes liquid saturation, the full curve denotes the relative humidity
for ice saturation, and the thin curves with various line codings are lines of constant water
mixing ratio. The symbols denote the measured data as derived from frostpoint hygrometer
and temperature sensors during the POLINAT 2 project (J. Ovarlez and U. Schumann).

Figure 8. Number densities of particles versus latitude during the outbound transfer flight to
Punta Arenas, Chile. a) Particles larger than 6 nm diameter b) Particles larger than 120nm
diameter. (Note difference in scale between Figures a and b) (J. Ström et al.).

Figure 9. Contrail cover over Europe from NOAA-14 noon scenes for March 1995-February
1997. The lines indicate main traffic routes over Europe (R. Meyer et al.).

Figure 10. Contrails, 0.65 µm optical depth τ, effective diameter De, and longwave (LW) and
shortwave (SW) cloud radiative forcing (CRF) in a NOAA-12 infrared image, 1214 UTC 26
September 1996, with letter-identified contrails (left) and contrails selected by automated
algorithm (right, in green). Properties of identified contrails are given in the table (P. Minnis
et al.).

8
November 27, 2000

References

IPCC, Climate Change 1995 – The Science of Climate Change, edited by J. T. Houghton et
al., Intergovernmental Panel on Climate Change, Cambridge University Press, 878 pp.,
1996.

IPCC, Aviation and the Global Atmosphere, edited by J. E. Penner et al., Intergovernmental
Panel on Climate Change, Cambridge Univ. Press, New York, 373 pp., 1999.

Schumann, U. and G. T. Amanatidis (Editors), Aviation, Aerosols, Contrails and Cirrus


Clouds (A2C3), Proceedings of a European Workshop, Seeheim (near Frankfurt/Main),
Germany, July 10-12, 2000, Air Pollution Research Report 74, Report EUR 19428 EN,
European Commission, Brussels, pp. 314, in press, 2000.

WMO, Scientific Assessment of Ozone Depletion: 1998, World Meteorological Organisation,


Geneva, 1999.

Ulrich Schumann, Deutsches Zentrum für Luft- und Raumfahrt (DLR), Institut für Physik der
Atmosphäre, Oberpfaffenhofen, 82230 Wessling, Germany, (ulrich.schumann@dlr.de)

9
November 27, 2000

Figures:

Fig. 1. Mean altitude of the tropopause in June and December, distribution of NOx emission
source rate from aircraft versus altitude and latitude (the emission rate is the larger the darker
the shaded area), and indications of mean circulation and some relevant processes.

Fig. 2. Spiral contrail observed in NOAA-14 AVHRR satellite data west of Denmark induced
by a military aircraft circulating over the North Sea at 1236 UT May 22, 1998, indicating that
long-lasting cirrus clouds are triggered by contrails which would not form otherwise (U.
Schumann)

10
November 27, 2000

Fig. 3. Radiative forcing (W m-2) due to aircraft emissions in 1992 (IPCC, 1999). The
columns indicate the best estimate of forcing. The error bars denote the range within which
the best-estimate value is expected with a two-third probability. The available information on
cirrus clouds is insufficient to determine either a best estimate or an uncertainty range; the
dashed line indicates a range of possible best estimates. The estimate for total forcing does not
include the effect of changes in cirrus cloudiness. The evaluations below the graph (“good,”
“fair,” “poor,” “very poor”) are a relative appraisal associated with each component and
indicates the level of scientific understanding.

Fig: 4. Global aircraft-generated soot inventory (vertical integral of emissions per month in
1992) (A. Döpelheuer).

11
November 27, 2000

Fig. 5. Formation processes of aerosols in an aircraft exhaust plume. (K.-H. Wohlfrom).

PEI3
17
10
PEIs [kg ]
-1

16 PEI5
10

15
10 PEI14

1 10 100 1000
-1
FSC [mg kg ]

Fig. 6. Apparent particle emission indices of particles larger diameter d (PEId, d = 3, 5, and 14
nm), measured in aircraft exhaust plumes that do not cause a contrail as a function of fuel
sulphur content (FSC) behind the ATTAS jet aircraft (squares), a Boeing 737-300 (circles),
and an Airbus 310-300 (diamonds) (A. Petzold et al.).

12
November 27, 2000

Fig. 7. Relative humidity with respect to liquid water versus ambient temperature. The horizontal
dashed line denotes liquid saturation, the full curve denotes the relative humidity for ice saturation,
and the thin curves with various line codings are lines of constant water mixing ratio. The symbols
denote the measured data as derived from frostpoint hygrometer and temperature sensors during the
POLINAT 2 project (J. Ovarlez and U. Schumann).

Ferry to Punta Arenas: Latitudional distribution of aerosol > 6nm Ferry to Punta Arenas: Latitudional distribution of aerosol > 120 nm
2
Flights performed in upper troposphere (ozone < 100 ppbv) 60
Flights performed in upper troposphere (ozone < 100 ppbv)
and devided in bins by 1 deg of Latitude and devided in bins by 1 deg of Latitude
4
10
8 50
6
a) b)
4

40
2
N120 (cm )
-3
N6 (cm )
-3

3 30
10
8
6

4 20

10
2
10
8
6
0

-60 -50 -40 -30 -20 -10 0 10 20 30 40 50 60 70 -60 -50 -40 -30 -20 -10 0 10 20 30 40 50 60 70
Latitude Latitude

Fig. 8 Number densities of particles versus latitude during the outbound transfer flight to
Punta Arenas, Chile. a) Particles larger than 6 nm diameter b) Particles larger than 120nm
diameter. (Note difference in scale between Figures a and b) (J. Ström et al.).

13
November 27, 2000

Fig. 9. Contrail cover over Europe from NOAA-14 noon scenes for March 1995-February
1997. The lines indicate main traffic routes over Europe (R. Meyer et al.).

E
H
A
D
B
G

A B C D E F G H I
τ 0.37 0.25 0.42 0.28 0.22 0.49 0.14 0.21 0.20
De (µm) 18 19 25 14 30 21 45 30 21
-2
LWCRF (Wm ) 24.2 13.0 21.7 7.0 5.2 16.8 6.7 6.9 4.7
-2
SWCRF (Wm ) -10.1 -2.6 -6.8 -2.2 -1.1 -7.1 -0.8 -5.2 -2.2

Fig. 10. Contrails, 0.65 µm optical depth τ, effective diameter De, longwave (LW) and
shortwave (SW) cloud radiative forcing (CRF) in NOAA-12 IR images, 1214 UTC, 26
September 1996, with letter-identified contrails (left) and contrails selected by automated

14
November 27, 2000

algorithm (right, in green). Properties of identified contrails are given in the table. (P. Minnis
et al.).

15

You might also like