ArticlePDF Available

Urban eddy covariance measurements reveal significant missing NOx emissions in Central Europe

Authors:

Abstract and Figures

Nitrogen oxide (NOx) pollution is emerging as a primary environmental concern across Europe. While some large European metropolitan areas are already in breach of EU safety limits for NO2, this phenomenon does not seem to be only restricted to large industrialized areas anymore. Many smaller scale populated agglomerations including their surrounding rural areas are seeing frequent NO2 concentration violations. The question of a quantitative understanding of different NOx emission sources is therefore of immanent relevance for climate and air chemistry models as well as air pollution management and health. Here we report simultaneous eddy covariance flux measurements of NOx, CO2, CO and non methane volatile organic compound tracers in a city that might be considered representative for Central Europe and the greater Alpine region. Our data show that NOx fluxes are largely at variance with modelled emission projections, suggesting an appreciable underestimation of the traffic related atmospheric NOx input in Europe, comparable to the weekend-weekday effect, which locally changes ozone production rates by 40%.
Content may be subject to copyright.
1
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
www.nature.com/scientificreports
Urban eddy covariance
measurements reveal signicant
missing NOx emissions in Central
Europe
T. Karl
1, M. Graus1, M. Striednig1, C. Lamprecht1, A. Hammerle2, G. Wohlfahrt
2, A. Held3,
L. von der Heyden3, M. J. Deventer4, A. Krismer5, C. Haun6, R. Feichter7 & J. Lee8
Nitrogen oxide (NOx) pollution is emerging as a primary environmental concern across Europe.
While some large European metropolitan areas are already in breach of EU safety limits for NO2,
this phenomenon does not seem to be only restricted to large industrialized areas anymore. Many
smaller scale populated agglomerations including their surrounding rural areas are seeing frequent
NO2 concentration violations. The question of a quantitative understanding of dierent NOx emission
sources is therefore of immanent relevance for climate and air chemistry models as well as air pollution
management and health. Here we report simultaneous eddy covariance ux measurements of NOx,
CO2, CO and non methane volatile organic compound tracers in a city that might be considered
representative for Central Europe and the greater Alpine region. Our data show that NOx uxes are
largely at variance with modelled emission projections, suggesting an appreciable underestimation of
the trac related atmospheric NOx input in Europe, comparable to the weekend-weekday eect, which
locally changes ozone production rates by 40%.
e nitrogen cycle1 is essential for maintaining the oxidizing capacity of the atmosphere and regulating ozone
in the lower atmosphere2. Perturbations due to rapid industrialization and agricultural activities have led to a
signicant increase of atmospheric nitrogen oxides (NOx) during the 20th century3. A regionally intense buildup
of photochemical smog due to the presence of nitrogen oxides, CO and non-methane volatile organic com-
pounds (NMVOC) was rst identied in the US and attributed as the main cause of severe ozone pollution in
many areas4. Decades of subsequent research activities ranging from detailed laboratory5, 6 and smog chamber79
studies to large scale eld campaigns1012 have led to a reasonably good mechanistic understanding of the for-
mation of tropospheric ozone, which is characterized by a complex nonlinear relationship between NOx and
reactive carbon species13. is interdependency gives regulators two key strategies to mitigate ozone pollution.
e eectiveness to control ozone thereby very much depends on the ratio between ambient OH reactivity and
NOx concentrations14, which can be described by relatively simple analytical relationships15. e development
of mechanistic regional16 and global air chemistry models17 has further given regulators and scientists powerful
tools to study tropospheric ozone formation16, where the mitigation of NOx emissions has emerged as one of the
key air pollution control strategies for ozone1820 and more recently also for particulate matter with a diameter
of 1 μm or less (PM1)21. Due to the toxicity, nitrogen dioxide (NO2) is also regulated as a hazardous air pollutant
itself22. For example, in Europe regulatory action under the EU ematic Strategy on Air Pollution is in place to
limit urban street canyon NO2 concentrations to 40 μg/m3 per year (or 200 μg/m3/h on less than 18 days/year)23.
Current trends across European air quality networks show that regulatory thresholds of NO2 are violated at many
stations, which does not seem to be limited to large population centers anymore (ref. 24, SI). In fact many rural
1Institute of Atmospheric and Cryospheric Sciences, University of Innsbruck, Innsbruck, Austria. 2Institute of Ecology,
University of Innsbruck, Innsbruck, Austria. 3Atmospheric Chemistry, University of Bayreuth, Innsbruck, Germany.
4Department of Geography, University of California, Berkeley, USA. 5Abteilung Waldschutz, Amt der Tiroler
Landesregierung, Innsbruck, Austria. 6Abteilung Geoinformation, Amt der Tiroler Landesregierung, Innsbruck,
Austria. 7Amt für Verkehrsplanung, Umwelt, Magistrat III Stadt Innsbruck, Innsbruck, Austria. 8National Centre for
Atmospheric Science and Department of Chemistry, University of York, York, UK. Correspondence and requests for
materials should be addressed to T.K. (email: thomas.karl@uibk.ac.at)
Received: 3 November 2016
Accepted: 19 April 2017
Published: xx xx xxxx
OPEN
www.nature.com/scientificreports/
2
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
areas and smaller towns see NO2 concentration levels rivaling those of large metropolitan areas. Owing to the
spatiotemporal variability and uncertainty of dierent anthropogenic NOx sources, it is dicult to attribute emis-
sion uncertainties to specic sectors in complex bottom-up emission inventories25 or top-down remote sensing
assessments26. Recently evidence has accumulated that rapid shis in transportation fuels can have signicant
impacts on air quality27, 28. In Europe for example the question about the increasing penetration of Diesel cars
raises concerns as to what extent such a technological change has been counterproductive to mitigating atmos-
pheric NO2 pollution under new emission regulation standards19, 29. e United States environmental protection
agency’s (US EPA) notice of violation of the Clean Air Act to a German automaker regarding Diesel engines has
sparked a number of new real world driving (RDE) emission tests across Europe, which show signicant man-
ufacturer and vehicle specic variability30, 31. ese new data suggest that the impact on up-scaled average eet
emissions needed for accurate air quality predictions remains unclear32.
A number of urban ux measurement sites for energy and CO2 have been established highlighting their
potential for surface-atmosphere exchange studies33, 34. In contrast, similar measurements for reactive gases are
oen still quite limited, owing to the complexity of the required measurement systems. A set of recently con-
ducted urban ground based and airborne NMVOC ux measurements revealed the usefulness to test bottom-up
emission inventories and revealed signicant discrepancies for some species3539. Urban ux measurements for
NOx are even more scarce32, 40, 41 indicating that constraints on emission sources in urban areas can be quite
uncertain. Here we improve upon existing work, by simultaneously measuring NOx, selected tracer NMVOCs,
CO and CO2 leading to a well constrained ux dataset, that allows testing our understanding of prominent NOx
emission sources.
Results
Obtaining ensemble average statistics on eet emissions by eddy covariance ux measure-
ments. A comprehensive set of eddy covariance measurements for NOx, marker NMVOC, CO and CO2 at
an urban location allows a direct comparison of relative ux ratios with bottom-up emission sources. e study
site located in Innsbruck (N 47°1551.50, E 11°236.77) at the center of the Inn valley, represents one of the
most strategically important Alpine crossing points for the transport of goods between Northern and Southern
Europe. Each year approximately 6 million vehicles42 pass through the east-west facing valley, which is about
10 km wide surrounded by mountain ridges about 2.5 km high. e valley topography leads to a very predictable
and pronounced wind system characterized by a topographic amplication factor (TAF) of about 343. Due to
the combination of signicant trac induced NOx emissions and increasingly stringent NO2 limit values, the
area is in non-attainment. Local authorities are facing legal proceedings by the European Commission for their
failure to control excessive levels of nitrogen oxides (Fig.S2), similar to many areas across Europe23. Tracer ux
relationships allow investigating to what extent urban emissions are caused by (a) trac, (b) urban residential
and (c) biomass burning/biofuel activities. Figure1 shows the diurnal evolution of Weekday (Tuesday-ursday)
and Sunday NOx uxes and concentrations along with mean trac count data at the site during the measurement
campaign (July – October, 2015). Median measured midday NOx mixing ratios in Innsbruck are comparable to
values reported for central London (10–14 ppbv), while corresponding observed uxes are about a factor of 3–4
lower (i.e. 3000–4000 ng/m2/s vs 700–1440 ng/m2/s)32. ese observations are consistent with the idea of an inten-
sication of air pollution proportional to TAF, and a corresponding eective lower air volume that pollutants are
being mixed into in steep valleys (Fig.S2, ref. 43). is comparison likely indicates that a much stronger reduction
of NOx emissions from the transport sector would be required in the Alps than for example in London in order
to achieve current air pollution standards along one of the busiest EU transport corridors across the Alps. Several
lines of evidence exclude signicant presence of biomass burning during the present study. e average ratio
between benzene and toluene uxes exhibited a typical value (3.3 ± 0.7; R2 = 0.85) characteristic for urban emis-
sion sources, dominated by fossil fuel combustion and evaporative/cold-start emissions. e correlation between
acetonitrile and benzene, toluene, NOx or CO2 uxes was low with an R2 of 0.07, 0.08, 0.02 and 0.06 respectively.
We also did not observe signicant excursions of other species recently suggested as additional biomass burning
markers44 such as furfural and furan showing a correlation coecient of R2 < 0.2 above their background uxes.
We observed an excellent correlation between CO2, benzene and NOx uxes (CO2/NOx: R2 = 0.86; benzene/NOx:
R2 = 0.75). e covariance between between NOx and CO2 (benzene) uxes yielded values of 0.91 (0.86). We
interpret these observations such that benzene, NOx and CO2 emissions are dominated by road trac with con-
tributions from residential combustion sources.
Benchmarking urban source emission ratios and inventories. To gain a more quantitative insight,
we investigated ux ratios between NOx and CO2 (FNOx/FCO2; Fig.2). e advantage of this approach is that it
allows determining the actual ensemble average of dierent emission sources based on measured ux ratios,
similar to an end-member un-mixing regression analysis. is allows us to compare our measurements to rela-
tive emission strengths reported in emission models and inventories. For large scale emission inventories (e.g.
grid cells > 1 km2) this approach also circumvents uncertainties related to assumptions of various downscaling
approaches. e observed FNOx/FCO2 ratios follow a diurnal cycle showing a ~40–50% variation throughout a day,
which reects the pronounced uctuation of trac activity across the city (e.g. ranging from about 97 vehicles/h
at night to 890 vehicles/h during daytime at a trac count station within the ux footprint). Since we can exclude
signicant industrial emissions within the ux footprint as well as biomass burning, the variation of FNOx/FCO2
should exhibit the characteristic behavior of city scale sources comprised of (1) a combination of vehicular emis-
sions and (2) residential/domestic combustion sources (e.g. oil and gas heating units). A minimization routine
(SI) allowed un-mixing these two end-members of the compositional data, reecting the actual emission ratios
for trac and urban residential combustion sources (Fig.2). e tted model (SI) can reproduce the diurnal cycle
and activity factors reasonably well, leading to a NOx/CO2 emission ratio for trac of 4.2(±0.3) × 103 ([mg/
www.nature.com/scientificreports/
3
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
m2/h] NOx/[mg/m2/h] CO2), and 0.20(±0.05) × 103 ([mg/m2/h] NOx/[mg/m2/h] CO2) for residential combus-
tion sources. ese calculated ratios are also depicted in Fig.2 by horizontal shaded blue and green lines. e
activity factors suggest that the ratio is dominated by trac, comprising about 85% of the activity averaged over
the entire day (and >95% during peak trac). We also investigated NOx/benzene ux source ratios revealing
Figure 1. Statistical plot of measured NOx uxes, mixing ratios and trac count data. e center dot shows
the ensemble median, where the box around it represents one standard deviation and whiskers the 25 and 75%
percentile. Individual extreme values are plotted as open circles. Panels A, B and C represent weekdays (i.e.
TUE-THU; composite of 609 individual data points) and panels D, E and F depict Sundays (193 individual data
points).
Figure 2. Diurnal cycle of median trac count data (black line - le axis), measured (blue circles) and model
tted (dashed blue line) mass ux ratio of NOx/CO2 (right axis). e corresponding calculated end members for
trac and residential combustion ratios are indicated by blue and green horizontal lines. e shading reects
one standard deviation. In addition colored dashed lines show predictions using xed emission ratios from
COPERT (magenta) and ACCMIP (red).
www.nature.com/scientificreports/
4
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
comparable dierences as observed for FNOx/FCO2. e CO2 ux weekend-weekday eect and CO/CO2 ratios close
to a recent road tunnel study45, all imply that a biogenic inuence on CO2 uxes due to photosynthetic uptake
or respiration can be considered negligible at this site. e obtained NOx/CO2 ux ratio for trac is signicantly
larger than predicted by a number of state of the art emission inventories and emission standards (Table1).
Generally, we observe 50–70% higher NOx emissions relative to CO2 from road trac than what is calculated
with the most recent trac emission models. In these detailed bottom – up models, mobile source emissions
are treated for dierent engine sizes and fuels, that, in the past, relied on standardized protocols obtained in test
facilities, but were recently updated based on a number of RDE tests30, 31. Exhaust from modern gasoline pow-
ered engines, despite higher ignition temperatures than those powered by Diesel, can be eectively treated for
NO along with NMVOC (and CO) using a three way catalytic (TWC) converter. TWC treatment can lead to a
10 fold reduction of NO. is has been hard to achieve for Diesel powered cars, which nowadays mostly rely on
selective catalytic reduction due to the high air to fuel ratio during combustion. Generally, dierent combustion
and exhaust treatment characteristics result in signicantly higher NOx/CO2 emission ratios for Diesel powered
cars than for gasoline. e modelled eet average contribution suggests that at least 90% of urban NOx emissions
should originate from Diesel driven vehicles at the present location. 85% is modelled to be emitted by the passen-
ger car eet based on the COPERT model and TRACCS database (ref. 46, SI). e current Austrian passenger car
eet comprises about 50% Diesel cars and the percentage across Europe grew at a substantially faster rate com-
pared to the US47. Based on our measurements the current average Austrian car eet emits about 36 times more
NOx per CO2 molecule compared to the US TIER II emission standard and a factor of 8–10 more than Euro 6
emission standards31. Factoring in dierences in fuel economy between the European and US car eet, this would
equate to about an order of magnitude more NOx emissions per travelled distance compared to newly introduced
emission standards. How comparable are these results to other European countries? Diesel engines dominate
the European passenger car market: 55% of all newly registered vehicles in the EU were powered by Diesel-fuel
in 201242; the penetration of Diesel cars of the two largest European economies bordering the Alps ranges from
30% (Germany) to 70% (France)42, more than an order of magnitude higher than in the US47. When comparing
measured NOx/CO2 ux ratios with current inventories used for IPCC atmospheric chemistry/climate48 and air
quality models49 we obtain a discrepancy up to about a factor of 3–4 (Table1). Incidentally, a recent comprehen-
sive model evaluation has suggested signicant discrepancies between regionally modelled and observed surface
NOx concentrations that seemed worse (e.g. by a factor of 2 during summers) over Europe than over the US25.
While a number of uncertainties (agricultural emissions, vertical mixing, biomass burning emissions, deposition)
can potentially result in modelled concentration biases in these models, our measurements suggest that road
transport related biases likely contribute signicantly to these discrepancies as they account for about half of the
total NOx emissions across Europe50.
What is the impact on atmospheric chemistry? e weekend – weekday eect (Fig.1) allows to gain
insight into changing NOx and NMVOC uxes on ozone production in more detail. In Austria, heavy duty vehi-
cle trac (trucks heavier than 7.5t and all road trains) is banned between Saturday 15:00 and Sunday 22:00 and
on public holidays between midnight and 22:00. Trac count data generally show a pronounced dierence in
Measurement/inventory ratio or
Measurement/emission standard ratio NOx/CO2 NOx/CO NOx/benzene
INNAQS/COPERT1,#
..
.
1 716
18
..
.
39
31
47
INNAQS/HBFA3.22
..
.
1 514
16
N/A
INNAQS/ACCMIP (trac)3
..
.
4 028
53
*
..
.
2 117
25
INNAQS/EMEP (trac)#
..
.
2 926
33
*N/A
INNAQS/US Tier II4
..
.
36 033 4
38 6
N/A
NOx/CO2 PC and LCV < 1305 kg NOx/CO2 LCV 1305 kg–3500 kg
INNAQS/Euro65Diesel
..
.
4 743
50
..
.
4 541
48
Petrol .
.
.
126
11 7
13 5
.
.
.
195
18 1
20 9
INNAQS/Euro55Diesel
..
.
2 119
23
..
.
2 018
21
Petrol .
.
.
126
11 7
13 5
.
.
.
195
18 1
20 9
INNAQS/Euro45Diesel
..
.
1 514
16
..
.
1 413
15
Petrol .
.
.
4 744
50
.
.
.
145
13 5
15 5
INNAQS/Euro35Diesel
..
.
0 9085
10
Petrol
..
.
2 523
27
..
.
180
16 7
19 3
Table 1. Measurement – inventory comparison (i.e. measured/modelled ux (emission) ratio). e uncertainty
range is given by sub- and superscripts. PC: passenger cars; LCV: light commercial vehicle. 1COPERT
emission model (SI). 2HBFA 3.2 (SI). Comparison with HBFA 3.3, that was published during the copy editing
phase,yielded an averagebias of 1.2. 3ACCMIP – (SI). 4US EPA22. 5EEA24. *For inventories that do not explicitly
report CO2, we converted data using the measured midday range of CO to CO2 ux ratios (3.6 to 4.6 ppbv/
ppmv), which fall close to a recent evaluation based on a road tunnel study45. #Data are trend adjusted for 2015
according to GAINS (http://gains.iiasa.ac.at/models/)23.
www.nature.com/scientificreports/
5
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
driving habits resulting in a factor of 1.9 ± 0.2 lower vehicle counts on Sunday than on weekdays. We calculated
typical ozone production rates for midday-aernoon conditions (11–16 h LT), when photochemistry peaks. e
corresponding NOx uxes are a factor of 2.1 ± 0.2 lower on Sundays, closely matching observations of vehi-
cle activity. Benzene and toluene uxes, representing the variation of anthropogenic NMVOC emissions, were
lower by a factor of 1.8 ± 0.3 and 2.0 ± 0.3 respectively. CO2 uxes changed by a similar factor of 2.3 ± 0.5. e
weekend-weekday comparison provides an independent conrmation that these pollutant emissions are domi-
nated by trac activity during the day. Average NOx concentrations are a factor of 2.5 ± 0.2 lower during Sundays.
Incidentally, the observed weekend – weekday reduction is comparable to the observed measurement-inventory
discrepancies or the eect if an entire car eet was converted from a Euro 5 (0.18 g/km) to a Euro 6 (0.08 g/km)
NOx emission standard. Our measurements therefore allows us to benchmark such a hypothetical regulatory
action in the real atmosphere. Changes in local ozone production are calculated following procedures outlined
before15, 51, where the sensitivity of local ozone production can be approximated by the ratio of radical termina-
tion (LN) processes (e.g. NO2 + OH) and photochemical radical production (Q):
δδ δ
δδ δ
−−
+−
L
Q
ONOJ
ONMVOC NO
22
3(1)
Nxx
xx
Here the δ symbol indicates the relative change between weekday and Sunday, J represents the photolysis rates,
and Ox = O3 + NO2. All terms on the right side can be inferred from measurements of the weekend eect, where
the anthropogenic change of NMVOCs is assumed to follow benzene. ere is evidence of a non-neglegible bio-
genic NMVOC (BVOC) presence at the site (e.g. 20–50% of the NMVOC reactivity) and these BVOCs do not
exhibit any anthropogenically related variation between weekdays and weekend. We apportioned the change of
NMVOC reactivity therefore into a biogenic and anthropogenic part using data from the PTR-QiTOF-MS instru-
ment. To achieve this, we estimated the total anthropogenic NMVOC reactivity from known urban concentration
ratios52, 53 scaled to benzene and compared this to the measured reactivity of BVOC, in particular the reactive
biogenic marker species isoprene and monoterpenes. is allowed to obtain upper (δNMVOC = δbenzene) and
lower (δNMVOC = δbenzene × [reactivity anthropogenic NMVOC]/[reactivity BVOC]) bounds for LN/Q. It can
further be shown15 that the sensitivity of local ozone production can be related to LN/Q according to:
=
dPO
dNO
ln ()
ln()
1
1
(2a)
x
L
Q
L
Q
3
3
2
1
2
N
N
=
dPO
dNMVOC
ln ()
ln()
1
(2b)
L
QL
Q
3
1
2
1
2
N
N
e calculated range for LN/Q delimits a ratio between 0.82 and 0.92 (theoretical maximum = 1) and places the
site in a NOx inhibited chemical regime for ozone production during weekdays15, 51. Incidentally these aernoon
values are systematically higher than reported for a rural site in Northern Italy with comparable NOx concentra-
tions, and more closely follow conditions reported for Milan in 199854. Our ux data clearly demonstrate that the
weekend eect of ozone production found in Innsbruck is largely a direct result of lower weekend NOx emissions.
As a consequence of the dominating nitrogen chemistry for radical loss calculated via eqs1 and 2, local ozone
production will therefore increase by 39% to 70% proportional to a reduction of NOx mixing ratios in Innsbruck.
It will decrease between 70% to 85% proportional to a reduction of NMVOC or CO. e impact on gross ozone
production (P(O3)) was also investigated independently using the Leeds Master Chemical Mechanism (MCM)
as a photochemical box model (refs 13 and 55, SI, Fig.3). e model simulates an increase of P(O3) by 40% (from
2.5 to 3.5 ppbv/h) on Sundays due to a 2 fold reduction of NOx, giving an explanation for the observed weekend
eect, which results in an increase of local aernoon ozone concentrations up to 24% (Fig.4). e model simula-
tion suggests a maximum increase of P(O3) up to 3 times (2.5 ppbv/h to 7.8 ppbv/h), if solely NOx concentrations
were decreased from current levels to about 2 ppbv (Fig.3). Below 2–4 ppbv of NOx, ozone production would
enter the NOx sensitive regime and gradually lead to decreasing P(O3). Next, we setup the MCM as a diurnally
constrained 0-dimensional diluting box model (SI) to study the sensitivity with respect to changes in modelled
ozone concentrations between weekdays and weekends. e model is thereby fully constrained by methane, CO,
measured NMVOC, NO, photolysis rates, and PBL height along with ancilliary meteorological variables (e.g. tem-
perature, humidity). NO2 and ozone are initialized by their measured initial concentrations, but are then allowed
to freely adjust during the model run. We chose a spin-up time of 3 days and used model results from the last day
of simulation to compare with measured ozone concentrations. Figure4 shows the diurnal patterns of modelled
and observed ozone concentrations. Observed (modelled) peak ozone concentrations are 49 ± 2 (52) ppbv and
43 ± 2 (44) ppbv on weekends and weekdays respectively. Modelled and observed weekend-weekday dierences
averaged over daytime hours correspond to 7.2 ± 1 and 7.6 ± 1.2. While the model reproduces the general diur-
nal ozone cycle and midday peak reasonably well, it underestimates absolute nighttime concentrations exhibiting
a stronger diurnal cycle. We attribute this shortcoming mainly to poor knowledge and constraints on entrainment
and advection processes that dominate observed nighttime distributions of ozone. While this approach has its
limitations, we believe it does a sucient job for investigating relative daytime changes of ozone concentrations
caused by the weekend eect. On average, daytime weekday concentrations are about 7–8 ppbv lower than on
weekends. ese results can be compared to two quite contrasting areas. In Mexico City51, ozone concentrations
change very little despite similar relative variations of NOx mixing ratios between weekday and weekend, which
www.nature.com/scientificreports/
6
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
can be attributed to a completely NOx saturated environment. On the other hand, a quite pronounced variation is
found in the southern California Air basin56, where average ozone concentrations are observed to change almost
twice as much compared to the present study. We interpret these observations such that they reect the change
in ozone production eciency (Fig.4), when going from environments with NOx saturated conditions to more
VOC limited conditions. In this context a number of explanations of the weekend eect on ozone mixing ratios
are oen discussed51: (1) dierent timing of NOx emissions on weekends and associated chemical repartitioning,
(2) carryover of previous day pollutants at the surface and alo, (3) higher weekend VOC emissions, (4) higher
weekend photolysis frequencies due to lower aerosol loadings, (5) changes in biogenic emissions due to a dierent
radiation regime caused by lower aerosol loadings and (6) lower overall weekend NOx emissions. We show for the
rst time, that, for Innsbruck, at least 85% of the NOx concentration change, leading to a 7–8 ppbv ozone increase
on weekends, can be directly associated with a change of the overall local atmospheric NOx ux.
Discussion
Continued urbanization in conjunction with rapid technological changes in the mobility sector poses a chal-
lenge for accurate up to date predictions of pollutant emissions. By constraining the actual uxes of nitrogen
oxides, NMVOC, CO2 and CO into the atmosphere our measurements provide an observationally based expla-
nation why NOx concentrations have hardly declined since the introduction of EURO 3 emission standards in
Central Europe. While the technological shi towards Diesel passenger cars might have helped curb CO2 emis-
sions through better fuel economy compared to gasoline powered cars in the past, it created a widespread prob-
lem of NO2 pollution across Europe23 that does not seem to be exclusively limited to the largest metropolitan
and industrialized areas. e presented ux measurements indicate that trac related NOx emissions in current
operational air quality models can be signicantly underestimated by up to a factor of 4 across countries exhib-
iting a sizeable fraction of Diesel powered cars in their eet. As Diesel fuels (including bio-diesel) could account
for 70% of the growth in transportation fuels by 2040, with signicant demand in Asian markets according to
industry projections57, a better understanding of the uncertainty in associated changes of NOx uxes and ozone
chemistry will therefore be important for future environmental impact studies. Our measurements show that pro-
jected signicant decreases in European NOx emissions from the mobile transport sector will lead to conditions
improving NO2 exposure limits, but could locally increase ozone levels on the short term. Concomitant reduction
measures for NOx and NMVOC (CO) might therefore still prove most eective to avoid parallel increases of
local ozone levels due to new NOx emission standards. is might be particularly important in areas where top-
ographic amplication can lead to a stronger accumulation of air pollutants than over at land and be a relevant
consideration for mountainous mega-cities58. Using the observed weekend eect as proxy for underestimated
NOx emissions (i.e. a factor 2–4 dierence), models would overestimate P(O3) by 30–40% under the observed
NOx inhibited – VOC limited regime and underestimate P(O3) downwind, once NOx concentrations fall below
2–4 ppbv. Here we demonstrate that parallel ux measurements of a wide range of chemical species can be used to
benchmark urban emission sources, complement traditional approaches and signicantly improve uncertainties
inherent to bottom-up scaling in atmospheric chemistry models.
Figure 3. Calculated ozone production rates (solid lines) and derivatives (dashed lines) as a function of NOx.
Blue and cyan represent base cases for the study site, taking into account changes in O3, NOx and NMVOCs
concentrations. e solid black line follows the observed weekend-weekday trajectory.
www.nature.com/scientificreports/
7
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
Methods
Instrumentation. NMVOC. A PTR-QiTOF instrument (Ionicon, Austria) was operated in hydronium
mode at standard conditions in the dri tube of 112 Townsend. e instrument was set up to sample ambient air
from a turbulently purged “3/8” Teon line. Every seven hours, zero calibrations were performed for 30 minutes
providing VOC free air from a continuously purged catalytical converter though a setup of soware controlled
solenoid valves. In addition, known quantities of a suite of VOC from a 1 ppm calibration gas standard (Apel
& Riemer, USA) were periodically added to the VOC free air and dynamically diluted into low ppbv mixing
ratios. NOx: A dual channel chemiluminescence instrument (CLD 899 Y; Ecophysics) was used for NO and NOx
measurements. e instrument was equipped with a metal oxide converter operated at 375C. e instrument
was operated in ux mode acquiring data at about 5 Hz, similar to measurements performed over a pasture59
and forest60. A NO standard was periodically introduced for calibration. Zeroing was performed once a day
close to midnight. CO2, H2O: A closed path eddy covariance system (CPEC 200; short inlet, enclosed IRGA
design; Campbell Scientic) measured three dimensional winds along with CO2 and H2O. An additional 3D sonic
anemometer (CSAT3; Campbell Scientic) was available for turbulence measurements at an alternative height
level (Fig.S3). Calibration for CO2 was performed once a day. O3: Ozone concentrations were obtained from a
closed-path UV photometric analyser (APOA-370, Horiba, Japan); CO: CO measurements were available for a
limited amount of time in August 2015. Ambient mole fractions of carbon monoxide (CO) were measured with
a quantum cascade laser spectrometer (CWQC-TILDAS-76-D, Aerodyne, USA) with a 76 m path length optical
cell at a wavenumber of ca. 2190 cm1. e QCL was operated at a pressure of ca. 4 kPa using a built-in pressure
controller and temperature of the optical bench and housing controlled to 35 °C. Fitting of absorption spectra
at 2 Hz, storing of calculated dry mole fractions, switching of zero/calibration valves, control of pressure lock,
correction for band broadening and other system controls were realized by the TDLWintel soware (Aerodyne,
USA) run on a PC synchronized in time with the system collecting the anemometer data using the NTP soware
(Meinberg, Germany).
Eddy covariance data analysis. e eddy covariance method is derived from the scalar budget equation aer
Reynolds decomposition, and in its simplest form for horizontally homogeneous ows normal to the surface,
where the mean vertical motion of wind (
w
) can be considered 0, relates the measured surface-atmosphere
exchange ux (F) to the covariance between vertical wind and concentration uctuation according to:
=′wcF,
where w represents the vertical uctuation of wind speed, and c the concentration uctuation. Brackets denote
the averaging interval. e ensemble average used here is 30 minutes. Fluxes were selected according to standard
quality control criteria, such as raw data despiking, correcting for high and low pass ltering biases, applying a
stationarity test and test on developed turbulent conditions (e.g. u* ltering)61. In addition we parsed the data to
make sure that the ux footprint would reect a representative urban area (SI).
Figure 4. Observed and modelled changes in ozone concentrations on weekdays (Tuesday – ursday) and
weekends (Sunday). Panel A depicts a statistical boxplot showing weekend-weekday dierences during daytime,
which are represented by the colored sections in panels B and C.
www.nature.com/scientificreports/
8
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
References
1. Fowler, D. et al. Eects of global change during the 21st century on the nitrogen cycle. Atmos. Chem. Phys. 15, 13849–13893 (2015).
2. Mons, P. S. et al. Tropospheric ozone and its precursors from the urban to the global scale from air quality to short-lived climate
forcer. Atmos. Chem. Phys. 15, 8889–8973 (2015).
3. Intergovernmental Panel on Climate Change (IPCC) “Climate Change 2014: Synthesis eport. Contribution of Woring Groups I,
II and III to the Fih Assessment eport of the Intergovernmental Panel on Climate Change, (IPCC, 2014, Geneva, https://www.
ipcc.ch/report/ar5/syr/).
4. Haagen-Smit, A. J., Bradley, C. E. & Fox, M. M. Ozone formation in photochemical oxidation of organic substances. Indust. and
Engineering Che mistr y 45, 2086–2089 (1953).
5. Calvert, J. G. Gaseous reactions. Annual Review of Physical Chemistry 11, 41–64 (1960).
6. Atinson, . inetics and mechanism of the gasphase reactions of the hydroxylradical with organic compounds under atmospheric
conditions. Chem. Rev. 86, 69–201 (1985).
7. Leone, J. A., Flagan, . C., Grosjean, D. & Seinfeld, J. H. An outdoor smog chamber and modeling study of toluene-NOx
photooxidation. Int. J. of Chem. Kinetics 17, 177–216 (1985).
8. Martin, M., Pons, M., Wirtz, ., Etzorn, T. & Senzig, J. An outdoor smog chamber study of the gas-phase chemistry of toluene and
xylenes in NOx/air systems, in e oxidizing capacity of the troposphere (Proceedings of the seventh European symposium on
physicochemical behavior of atmospheric pollutants, European Commission, Luxembourg, 1996) pp. 350–354.
9. arl, M. et al. inetic Study of the OH-isoprene and O3-isoprene reaction in the atmosphere simulation chamber SAPHI. GRL. 31,
doi:10.1029/2003GL019189 (2004).
10. Meagher, J. F., Cowling, E. B., Fehsenfeld, F. C. & Parhurst, W. J. Ozone formation and transport in the southeastern United States:
Overview of the SOS Nashville Middle Tennessee Ozone Study. J. Geophys. Res. 103, 22213–22223 (1998).
11. idley, B. A. & obinson, E. e Mauna-Loa-Observatory Photochmsitry Experiment. J. Geophys. Res. 97, 10285–10290 (1992).
12. Jacob, D. J. et al. Transport and Chemical Evolution over the Pacic (TACE-P) aircra mission: Design, execution, and rst results.
J. Geophys. Res. 108, 1–19 (2003).
13. Jenin, M. E., Saunders, S. M. & Pilling, M. J. The tropospheric degradation of volatile organic compounds: A protocol for
mechanism development. Atmos. Environ. 31, 81–104 (1997).
14. Brune, W. H. et al. Ozone production chemistry in the presence of urban plumes, Faraday Discussions, doi:10.1039/c5fd00204d
(2015).
15. leinman, L. I. et al. Dependence of ozone production on NO and hydrocarbons in the troposphere. GRL 24, 2299–2302 (1997).
16. Im, U. et al. Evaluation of operational on-line-coupled regional air quality models over Europe and North America in the context of
AQMEII phase 2. Part I: Ozone. Atmos. Environm. 115, 404–420 (2015).
17. Ehhalt, D. & Prather, M. “Atmospheric Chemistry and Greenhouse Gases” (Intergovernmetnal Panel on Climate Change, 3rd
assessment report, Chapter 4, 2001; http://www.grida.no/publications/other/ipcc_tar/).
18. de Gouw, J. A., Parrish, D. D., Frost, G. J. & Trainer, M. educed emissions of CO2, NOx, and SO2 from US power plants owing to
switch from coal to natural gas with combined cycle technology. Earths Future 2, 75–82 (2014).
19. Ehlers, C. et al. Twenty years of ambient observations of nitrogen oxides and specied hydrocarbons in air masses dominated by
trac emissions in Germany, Faraday Discussions, doi:10.1039/c5fd00180c (2015).
20. Carslaw, D. C. & Beevers, S. D. New Directions: Should road vehicle emissions legislation consider primary NO2? Atmos. Environ.
38, 1233–1234 (2004).
21. iendler-Scharr, A. et al. Ubiquity of organic nitrates from nighttime chemistry in the European submicron aerosol. GRL 43,
7735–7744 (2016).
22. U. S. Environmental Protection Agency, (EPA), “NAAQS Scientific and Technical Information”, https://www.epa.gov/naaqs/
nitrogen-dioxide-no2-primary-air-quality-standards retrieved on 07-01-2016, (2016).
23. iese wetter, G. et al. Modelling NO2 concentrations at the street level in the GAINS integrated assessment model: projections under
current legislation. Atmos. Chem. Phys. 14, 813–829 (2014).
24. European Environment Agency, “Air quality in Europe – 2015” (Tech. ep. 5/2015, ISBN 978-92-9213-702-1, doi:10.2800/62459,
2015).
25. Giordano, L. et al. Assessment of the MACC reanalysis and its inuence as chemical boundary conditions for regional air quality
modeling in AQMEII-2. Atmos. Environ. 115, 371–388 (2015).
26. Hecel, A. et al. Inuence of low spatial resolution a priori data on tropospheric NO2 satellite retrievals. Atmos. Meas. Technol. 4,
1805–1820 (2011).
27. Salvo, A. & Geiger, F. M. eduction in local ozone levels in urban Sao Paulo due to a shi from ethanol to gasoline use. Nature.
Geoscience. 7, 450–458 (2014).
28. Madronich, S. Atmospheric chemistry: Ethanol and ozone. Nature Geoscience 7, 395–397 (2014).
29. Carslaw, D. C. Evidence of an increasing NO2/NOx emissions ratio from road trac emissions. Atmos. Environm 39, 4793–4802
(2005).
30. Bundesministerium für Verehr und digitale Infrastrutur (BMVI), “Bericht der Untersuchungsommission “Volswagen” 2016”
(https://www.bmvi.de/SharedDocs/DE/Publiationen/LA/bericht-untersuchungsommission-volswagen.html, retrieved 2016)
[the easiest access to this source is via the UL].
31. Franco, V., Posada Sanches, F., German, J. & Moc, P. “eal-world exhaust emissions from modern diesel cars, Part 2” (Tech. eport:
e International Council on Clean Transportation, www.theicct.org, Washington, DC, 2014).
32. Lee, J. et al. Measurement of NOx uxes from a tall tower in central London, U and comparison with emission inventories. Environ.
Sci. & Tech 49, 1025–1034 (2015).
33. Christen, A. Atmospheric measurement techniques to quantify greenhouse gas emissions from cities. Urban Climate 10, 241–260
(2014).
34. Nemitz, E., Hargreaves, . J., McDonald, A. G., Dorsey, J. . & Fowler, D. Meteorological measurements of the urban heat budget
and CO2 emissions on a city scale. Environ. Sci. Tech. 36, 3139–3146 (2002).
35. Velasco, H. E. et al. Eddy covariance ux measurements of pollutant gases in urban Mexico City. Atmos. Chem. Phys. 9, 7325–7342
(2009).
36. Valach, A. C., Langford, B., Nemitz, E., Macenzie, A. . & Hewitt, C. N. Seasonal and diurnal trends in concentrations and uxes
of volatile organic compounds in central London. Atmos. Chem. Phys. 15, 7777–7796 (2015).
37. arl, T. et al. Emissions of volatile organic compounds inferred from airborne ux measurements over a megacity. Atmos. Chem.
Phys. 9, 271–285 (2009).
38. Langford, B., Davison, B., Nemitz, E. & Hewitt, C. N. Mixing ratios and eddy covariance ux measurements of volatile organic
compounds from an urban canopy (Manchester, U). Atm. Chem. Phys 9, 1971–1987 (2009).
39. antala, P. et al. Anthropogenic and biogenic inuence on VOC uex at an urban bachground site in Helsini, Finland. Atm os.
Chem. Phys. 16, 7981–8007 (2016).
40. Marr, L. C., Moore, T. O., lapmeyer, M. E. & illar, M. B. Comparison on NOx uxes measured by eddy covariance to emission
inventories and land use. Environ. Sci. & Techn 47, 1800–1808 (2013).
41. Vaughan, A. . et al. Spatially resolved ux measurements of NOx from London suggest signicantly hagher emissions than
predicted by inventories. Faraday Discu ssions 189, 455–472 (2016).
www.nature.com/scientificreports/
9
Scientific RepoRts | 7: 2536 | DOI:10.1038/s41598-017-02699-9
42. EUOSTAT, “Energy, transport and environment indicators” (Tech. eport 201, European Commission, ec.europa.eu, ISBN 978-92-
79-49471-0, doi:10.2785/547816, 2015).
43. Vergreiner, I. & Dreiseitl, E. Valley winds and slope winds – observations and elementary thoughts. Meteorol. Atmos. Phys. 36,
24–286 (1987).
44. Coggon, M. M. et al. Emissions of nitrogen-containing organic compounds from the burning of herbaceous and arboraceous
biomass: Fuel composition dependence and the variability of commonly used nitrile tracers. GRL, doi:10.1002/2016GL070562
(2016).
45. Popa, M. E. et al. Vehicle emissions of greenhouse gases and related tracers from a tunnel study: CO: CO2, N2O: CO2, CH4: CO2,
O2:CO2 ratios, and the stable isotopes 13C, 18O in CO2 and CO. Atmos. Chem. Phys. 14, 210–2123 (2014).
46. Papadimitriou, G. et al. “TACCS - Transport data collection supporting the quantitative analysis of measures relating to transport
and climate change” (Tech. ep. EMISIA No. 13.E.025.V1, 2013).
47. Harley, . A., Marr, L. C., Lehner, J. . & Giddings, S. N. Changes in motor vehicle emissions on diurnal to decadal time scales and
eects on atmospheric composition. Environ. Sci. Tech. 39, 5356–5362 (2005).
48. Lamarque, J. F. et al. Historical (1850–2000) gridded anthropogenic and biomass burning emissions of reactive gases and aerosols:
methodology and application. Atmos. Chem. Phys. 10, 7017–7039 (2010).
49. Vestreng, V. et al. Evolution of NOx emissions in Europe with focus on road transport control measures. Atmos. Chem. Phys 9,
1503–1520 (2009).
50. eis, S. et al. A. oad trac emissions –predictions of future contributions to regional ozone levels in Europe. Atmos. Environ. 34,
4701–4710 (2000).
51. Stephens, S. et al. Weely patterns of Mexico City’s surface concentrations of CO, NOx, PM10 and O3 during 1986–2007. Atmos.
Chem. Phys. 8, 5313–5325 (2008).
52. Hellen, H., Haola, H. & Laurila, T. Determination of source contributions of NMHCs in Helsini (60 degrees N, 25 degrees E)
using chemical mass balance and the Unmix multivariate receptor models. Atmos. Environm 37, 1413–1424 (2003).
53. B orbon, A. et al. Emission ratios of anthropogenic volatile organic compounds in northern mid-latitude megacities: Observations
versus emission inventories in Los Angeles and Paris. J. Geophys. Res. 118, 2041–2057 (2013).
54. Spirig, C., Neel, A., leinman, L. I. & Hjorth, J. NOx versus VOC limitation of O3 production in the Po valley : Local and integrated
vew based on observations, J. Geophys. Res. 107, doi:10.1029/2001JD000561 (2002)
55. Wolfe, G. & Thornton, J. The Chemistry of atmosphere-forest exchange (CAFÉ) model – Part 1: Model description and
characterization. Atmos. Chem. Phys. 11, 77–101 (2011).
56. im, S. W. et al. Modeling the weely cycle of NOx and CO emissions and their impacts on O3 in the Los Angeles – South Coast air
Basin during the CalNex 2010 eld campaign. J. Geophys. Res. Atmos. 121, 1340–1360 (2016).
57. ExxonMobil, “e Outloo for Energy: A view to 2040” (Techn. eport, http://corporate.exxonmobil.com/en/energy/energy-
outloo/introduction/a-view-to-2040 retrieved on 1st of August 2016).
58. Molina, L. T. et al. An overview of the MILAGO 2006 Campaign: Mexico City emissions and their transport and transformation.
Atmos. Chem. Phys. 10, 8697–8760 (2010).
59. Marx, O., Brümmer, C., Ammann, C., Wol, V. & Freibauer, A. TANC – a novel fast-response converter to measure total reactive
atmospheric nitrogen. Atmos. Meas. Tech. 5, 1045–1057 (2012).
60. ummel, U., Ammann, C., Gut, A., Meixner, F. X. & Andreae, M. O. Eddy covariance measruements of nitric oxide ux within an
Amazonian rainforest, J. Geophys. Res. 107, doi:10.1029/2001JD000520 (2002).
61. Foen, T., Leuning, ., Oncley, S. P., Mauder, M. & Aubinet, M. Corrections and data quality, Eddy Covariance: A Practical Guide to
Measurement and Data Analysis (luwer Academic Publishers, Dortrecht, ed. 1, 2010), pp. 85–131.
Acknowledgements
T.K. received support from the EC Seventh Framework Program (Marie Curie Reintegration Program, “ALP-
AIR”, grant no. 334084).
Author Contributions
Karl, T. and Graus, M. conceived the experiment. Karl, T., Graus, M., Striednig, M., Lamprecht C., Hammerle,
A., Wohlfahrt, G., Held, A., von der Heyden, L., and Deventer J. designed and performed the INNAQS eld
experiment, interpreted the data and wrote the manuscript with inputs from all co-authors; Lee, J. provided input
on the interpretation of NOx ux data; Krismer A. provided air quality data, Feichter R. performed and analyzed
trac ow data; Haun C. provided the bottom up stationary emission source data; Karl T. performed the model
simulations.
Additional Information
Supplementary information accompanies this paper at doi:10.1038/s41598-017-02699-9
Competing Interests: e authors declare that they have no competing interests.
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2017

Supplementary resource (1)

... World Health Organization (WHO) air quality (AQ) standards have been implemented in Europe as regulatory action under the EU Thematic Strategy on Air Pollution, which aims to limit urban street canyon NO 2 concentrations to 40 μg/m 3 per year (or 200 μg/m 3 per hour on less than 18 days/year) (16). Compared to the United States (3,17), where NO 2 and NO x have generally been declining steadily, current trends across urban European AQ networks show that NO 2 has not decreased as projected (14,18). Because of slower than expected decrease in NO 2 levels and stricter AQ standards, regulatory thresholds of NO 2 are now violated at many stations (19). ...
... The weekend-weekday effect for the summer of 2018 shows that peak O 3 was about 10 ppbv higher on weekends, reflecting lower NO x mixing ratios. This is comparable to what we observed in 2015, when the difference between weekend and weekday was on the order of 7 to 8 ppbv (18). Surface in situ observations of NO 2 toward the east agree to within the uncertainty of remote sensing observations obtained from a ground-based Pandora system, which has been used extensively to validate Tropospheric Monitoring Instrument (TROPOMI) observations ( fig. ...
... In urban areas, the establishment of the Leighton ratio is thought to be dominated by the O 3 -NO-NO 2 triad due to high-NO emissions. To estimate the influence of RO x chemistry in the present study, we use a chemical box model sensitivity analysis that previously calculated a range of RO x between 2 and 15 parts per trillion by volume ( pptv) (18). This is comparable to typically observed total peroxy radical densities (40). ...
Article
Full-text available
Nitrogen oxides (NO x ) play a central role in catalyzing tropospheric ozone formation. Nitrogen dioxide (NO 2 ) has recently reemerged as a key target for air pollution control measures, and observational evidence points toward a limited understanding of ozone in high-NO x environments. A complete understanding of the mechanisms controlling the rapid atmospheric cycling between ozone (O 3 )–nitric oxide (NO)–NO 2 in high-NO x regimes at the surface is therefore paramount but remains challenging because of competing dynamical and chemical effects. Here, we present long-term eddy covariance measurements of O 3 , NO, and NO 2 , over an urban area, that allow disentangling important physical and chemical processes. When generalized, our findings suggest that the depositional O 3 flux near the surface in urban environments is negligible compared to the flux caused by chemical conversion of O 3 . This leads to an underestimation of the Leighton ratio and is a key process for modulating urban NO 2 mixing ratios. As a consequence, primary NO 2 emissions have been significantly overestimated.
... The observed diurnal and seasonal variations agree with [48], who observed two diurnal peaks for the NO2 surface concentration at 07:00 UTC and 18:00 UTC, more evident during wintertime, associated with atmospheric dynamics and peaks in vehicular traffic. In fact, as reported by [53], the double peak daily cycle of NO2 is observable in the NO2 concentrations, but not in its vertical fluxes. It suggests that the amount of atmospheric NO2 is mainly driven by the dynamics of the atmospheric boundary layer, In fact, since the near-surface and the in situ concentration of NO 2 are closely linked to local emissions, in the urban area of Rome a morning peak between 07:00 and 09:00 UTC and a second peak between 20:00 and 22:00 UTC, higher than the morning one, are expected [51]. ...
... The observed diurnal and seasonal variations agree with [48], who observed two diurnal peaks for the NO 2 surface concentration at 07:00 UTC and 18:00 UTC, more evident during wintertime, associated with atmospheric dynamics and peaks in vehicular traffic. In fact, as reported by [53], the double peak daily cycle of NO 2 is observable in the NO 2 concentrations, but not in its vertical fluxes. It suggests that the amount of atmospheric NO 2 is mainly driven by the dynamics of the atmospheric boundary layer, and therefore to its development, also in relation to local weather conditions, and only secondarily to the volume of traffic. ...
Article
Full-text available
To assess the best measures for the improvement of air quality, it is crucial to investigate insitu and columnar pollution levels. In this study, ground-based measurements of nitrogen dioxide(NO2) and ozone (O3) collected in Rome (Italy) between 2017 and 2022 are analyzed. Pandora sun-spectrometers provided the time series of the NO2 vertical column density (VC-NO2), troposphericcolumn density (TC-NO2), near-surface concentration (SC-NO2), and the O3 vertical column density(VC-O3). In situ concentrations of NO2 and O3 are provided by an urban background air qualitystation. The results show a clear reduction of NO2 over the years, thanks to the recent ecologicaltransition policies, with marked seasonal variability, observable both by columnar and in situ data.Otherwise, O3 does not show inter-annual variations, although a clear seasonal cycle is detectable.The results suggest that the variation of in situ O3 is mainly imputable to photochemical reactionswhile, in the VC-O3, it is triggered by the predominant contribution of stratospheric O3. The outcomeshighlight the importance of co-located in situ and columnar measurements in urban environmentsto investigate physical and chemical processes driving air pollution and to design tailored climatechange adaptation strategies.
... The calculated flux coupled with a footprint model provides information on surface emissions, allowing for changes to be studied and for direct comparison to the emissions inventories used in policy development. Whilst most frequently used for measuring carbon dioxide exchange with ecosystems from stationary towers (Baldocchi et al., 2001;Griffis et al., 2008;, the technique has been extended to the urban canopy for both greenhouse gases and air pollutants (Langford et al., 2010a;Lee et al., 2014;Helfter et al., 2016;Karl et al., 2017), as well as to airborne measurements for the assessment of fluxes at a much greater spatial extent Metzger et al., 2013;Vaughan et al., 2017). Recently, Lamprecht et al. (2021) used long-term air pollutant emissions measurements to understand how COVID-19 restrictions impacted different sources of NO x in the small European city of Innsbruck, Austria. ...
... Highlighted in green are all the resulting possible solutions where, crucially, to achieve the observed reductions in NO x flux, there must have been a 73 %-100 % reduction in transport NO x emissions, with transport contributing > 70 % to total NO x emissions. This transport contribution percentage demonstrates the underestimation in the inventory of transport emissions, in agreement with Karl et al. (2017) and Drysdale et al. (2022). However, the most interesting observation is that a 73 %-100 % decrease in transport NO x emissions is seen for only a 32 % decrease in traffic load since 2017. ...
Article
Full-text available
Fluxes of nitrogen oxides (NOx=NO+NO2) and carbon dioxide (CO2) were measured using eddy covariance at the British Telecommunications (BT) Tower in central London during the coronavirus pandemic. Comparing fluxes to those measured in 2017 prior to the pandemic restrictions and the introduction of the Ultra-Low Emissions Zone (ULEZ) highlighted a 73 % reduction in NOx emissions between the two periods but only a 20 % reduction in CO2 emissions and a 32 % reduction in traffic load. Use of a footprint model and the London Atmospheric Emissions Inventory (LAEI) identified transport and heat and power generation to be the two dominant sources of NOx and CO2 but with significantly different relative contributions for each species. Application of external constraints on NOx and CO2 emissions allowed the reductions in the different sources to be untangled, identifying that transport NOx emissions had reduced by >73 % since 2017. This was attributed in part to the success of air quality policy in central London but crucially due to the substantial reduction in congestion that resulted from pandemic-reduced mobility. Spatial mapping of the fluxes suggests that central London was dominated by point source heat and power generation emissions during the period of reduced mobility. This will have important implications on future air quality policy for NO2 which, until now, has been primarily focused on the emissions from diesel exhausts.
... The calculated flux coupled with a footprint model provides information on surface emissions, allowing for changes to be studied and for direct comparison to the emissions inventories used in policy development. Whilst most frequently used for measuring carbon dioxide exchange with ecosystems from stationary towers (Baldocchi et al., 2001;Griffis et al., 2008;Butterbach-Bahl et al., 2013), the technique has been extended to the urban canopy for both greenhouse gases and air pollutants (Langford et al., 2010;Lee et al., 2014;Helfter 50 et al., 2016;Karl et al., 2017), as well as to airborne measurements for the assessment of fluxes at a much greater spatial extent (Vaughan et al., 2021;Metzger et al., 2013;Vaughan et al., 2017). ...
... Highlighted in green are all the resulting possible solutions where crucially, to achieve the observed reductions in NO x flux, there must have been a 75-100 % reduction in transport NO x emissions with transport contributing > 70 % to total NO x emissions. This transport contribution percentage demonstrates the underestimation in the inventory of transport emissions in agreement withKarl et al. (2017) andDrysdale et al. (2022). However, the most interesting observation 180 is that a 75-100 % decrease in transport NO x emissions is seen for only a 32 % decrease in traffic load since 2017.When compared to concentrations, we found that NO x concentrations at Marylebone Road, a kerbside monitoring site within the flux footprint, had declined by 62 % between the two periods (248 vs 95 µg m −3 ). ...
Preprint
Full-text available
Fluxes of nitrogen oxides (NOx = NO + NO2) and carbon dioxide (CO2) were measured using eddy covariance at the BT Tower in central London during the coronavirus pandemic. Comparing fluxes to those measured in 2017 prior to the pandemic restrictions and the introduction of the Ultra-Low Emissions Zone (ULEZ) highlighted a 75 % reduction in NOx emissions between the two periods but only a 20 % reduction in CO2 emissions and a 32 % reduction in traffic load. Use of a footprint model and the London Atmospheric Emissions Inventory (LAEI) identified transport and heat and power generation to be the two dominant sources of NOx and CO2 but with significantly different relative contributions for each species. Application of external constraints on NOx and CO2 emissions allowed the reductions in the different sources to be untangled identifying that transport NOx emissions had reduced by > 75 % since 2017. This was attributed in part to the success of air quality policy in central London, but crucially due to the substantial reduction in congestion that resulted from pandemic reduced mobility. Spatial mapping of the fluxes suggests that central London was dominated by point source heat and power generation emissions during the period of reduced mobility. This will have important implications on future air quality policy for NO2 which until now, has been primarily focused on the emissions from diesel exhausts.
... We should also mention that there might be large uncertainties for the calculations of emission factors, as discussed in Doumbia et al. (2021). Underestimates of traffic NO x emissions over Europe have been mentioned previously in several air quality modelling studies (von Schneidemesser et al., 2021;Karl et al., 2017). Although the model exhibits biases for NO 2 and O 3 , the modelled air quality changes presented in the next section are in line with other studies such as Querol et al. (2021) that present a comparison between data from years 2015-2019 and the lockdown and relaxation periods for the year 2020 in the city of Barcelona. ...
Article
Full-text available
Tropospheric ozone (O3) is an important surface pollutant in urban areas, and it has complex formation mechanisms that depend on the atmospheric chemistry and on meteorological factors. The severe reductions observed in anthropogenic emissions during the COVID-19 pandemic can further our understanding of the photochemical mechanisms leading to O3 formation and provide guidance for policies aimed at reducing air pollution. In this study, we use the Weather Research and Forecasting model with Chemistry (WRF-Chem) coupled with the urban canopy building effect parameterization and building energy model (BEP + BEM) to investigate changes in the ozone chemistry over the metropolitan area of Barcelona (AMB) and its atmospheric plume moving northwards, which is responsible for the highest number of hourly O3 exceedances in Spain. The trajectories of the air masses from the AMB to the Pyrenees are studied with the Lagrangian FLEXible PARTicle dispersion model with WRF (FLEXPART-WRF). The aim is to investigate the response of ozone chemistry to reduction in precursor emissions (NOx – nitrogen oxides; VOCs – volatile organic compounds). The results show that, with the reduction in emissions, (1) the ozone chemistry tends to enter the NOx-limited or transition regimes, but highly polluted urban areas are still in the VOC-limited regime; (2) the reduced O3 production is overwhelmed by reduced nitric oxide (NO) titration, resulting in a net increase in the O3 concentration (up to 20 %) in the evening; (3) the increase in the maximum O3 level (up to 6 %) during the highest emission-reduction period could be attributed to an enhancement in the atmospheric oxidants hydroxyl and nitrate radical (OH and NO3) given their strong link with O3 loss or production chemistry; (4) the daily maximum levels of ozone and odd oxygen species (Ox) generally decreased (4 %) in May – a period with intense radiation which favours ozone production – with the reduced atmospheric OH and NO3 oxidants, indicating an improvement in the air quality; and (5) ozone precursor concentration changes in the urban plume of Barcelona contribute significantly to the level of pollution along the 150 km south-to-north valley in the Pyrenees. Our results indicate that O3 abatement strategies cannot rely only on NOx emission control but must include a significant reduction in anthropogenic sources of VOCs. In addition, our results show that mitigation strategies intended to reduce O3 should be designed according to the local meteorology, air transport, particular ozone regimes, and oxidation capacity of the atmosphere of the urban area.
... The first general method is based on sets of equations referred to as bottom-up aggregation, in which small geographic and temporal scale measurements made in the laboratory or field are aggregated together with economic and technological data Olivier et al., 1994;van Amstel et al., 1999). The second approach uses directly measured point-source measurements such as those made via local flux towers (Geddes and Murphy, 2014;Haszpra et al., 2018;Karl et al., 2017;Lee et al., 2015). The third approach uses chemical transport models and either a Bayesian, 3D/4D variance, or Kalman Filter type of inversion (Cohen and Wang, 2014;Henderson et al., 2012;Hu et al., 2022;Napelenok et al., 2008). ...
Article
Current approaches to estimate NO x emissions fail to account for new and small sources, biomass burning, and sources which change rapidly in time, generally don't account for measurement error, and are either based on models, or do not consider wind, chemistry, and dynamical effects. This work introduces a new, model-free analytical environment that assimilates daily TROPOMI NO 2 measurements in a mass-conserving manner, to invert daily NO x emissions. This is applied over a rapidly developing and energy-consuming region of Northwest China, specifically chosen due to substantial economic and population changes, new environmental policies, large use of coal, and access to independent emissions measurements for validation, making this region representative of many rapidly developing regions found across the Global South. This technique computes a net NO x emissions gain of 118% distributed in a seesaw manner: a more than doubling of emissions in cleaner regions, chemical plants, and regions thought to be emissions-free, combined with a more than halving of emissions in city centers and at well-regulated steel and powerplants. The results allow attribution of sources, with major contributing factors computed to be increased combustion temperature, atmospheric transport, and in-situ chemical processing. Furthermore, sensitivity runs that represent the range of uncertainties in the TROPOMI NO 2 retrievals themselves are performed, and it is determined that the differences between the errors in TRO-POMI NO 2 lead to smaller errors in the emissions estimates, due to non-linear buffering built in from the physical bounds on the underlying physical and chemical driving factors. It is also observed that errors in the priori dataset contribute more than the uncertainties in the TROPOMI retrievals, opening a new door to recursive iteration between this approach and new emissions dataset development. It is hoped that these findings will drive a new look at emissions estimation and how it is related to remotely sensed measurements and associated uncertainties , especially applied to rapidly developing regions. This is especially important for understanding the loadings and impacts of short-lived climate forcers, and provides a bridge between remotely sensed data, measurement error, and models, while allowing for further improvement of identification of new, small, and rapidly changing sources.
... It was noted that the model underestimated NOX traffic emissions in urban areas on weekday between 6 AM and 5 PM. Similar underestimation of NOX was found in other studies as well (e.g., [19][20]). ...
Article
Full-text available
Many cities face low air quality. To better predict the exceedance of air quality limits, the traffic’s contribution to air pollution was analysed in this paper. Several studies used a twin site approach to determine the impact of urban traffic; however, it requires the deployment of stations at various locations. A time variant analysis to determine traffic’s contribution and regression analysis were applied to determine the weather’s impact. The results were validated using actual traffic data. It was found that the traffic’s contributions to CO and NO2 were 22 and 30%. It was noted that the seasonal fluctuation of NO2 is significantly influenced by precipitation. Long-term trends of pollutants require further research.
... last access: 30 January 2021; AmeriFlux, https://ameriflux.lbl.gov/, last access: 30 January 2021). A number of studies investigated eddy fluxes of chemical species and aerosols in urban settings (Nemitz et al., 2008;Velasco et al., 2009;Rantala et al., 2016;Lee et al., 2015;Karl et al., 2017;Vaughan et al., 2017;Striednig et al., 2020). Briefly, the method relies on the conservation equation of a scalar, which under homogenous conditions and can be simplified to ...
Article
Full-text available
Lockdown and the associated massive reduction in people's mobility imposed by SARS-CoV-2 (severe acute respiratory syndrome coronavirus 2) mitigation measures across the globe provide a unique sensitivity experiment to investigate impacts on carbon and air pollution emissions. We present an integrated observational analysis based on long-term in situ multispecies eddy flux measurements, allowing for quantifying near-real-time changes of urban surface emissions for key air quality and climate tracers. During the first European SARS-CoV-2 wave we find that the emission reduction of classic air pollutants decoupled from CO2 and was significantly larger. These differences can only be rationalized by the different nature of urban combustion sources and point towards a systematic bias of extrapolated urban NOx emissions in state-of-the-art emission models. The analysis suggests that European policies, shifting residential, public, and commercial energy demand towards cleaner combustion, have helped to improve air quality more than expected and that the urban NOx flux remains to be dominated (e.g., >90 %) by traffic.
Article
Full-text available
NO2 concentrations at the street level are a major concern for urban air quality in Europe and have been regulated under the EU Thematic Strategy on Air Pollution. Despite the legal requirements, limit values are exceeded at many monitoring stations with little or no improvement in recent years. In order to assess the effects of future emission control regulations on roadside NO2 concentrations, a downscaling module has been implemented in the GAINS integrated assessment model. The module follows a hybrid approach based on atmospheric dispersion calculations and observations from the AirBase European air quality database that are used to estimate site-specific parameters. Pollutant concentrations at every monitoring site with sufficient data coverage are disaggregated into contributions from regional background, urban increment, and local roadside increment. The future evolution of each contribution is assessed with a model of the appropriate scale: 28 x 28 km grid based on the EMEP Model for the regional background, 7 x 7 km urban increment based on the CHIMERE Chemistry Transport Model, and a chemical box model for the roadside increment. Thus, different emission scenarios and control options for long-range transport as well as regional and local emissions can be analysed. Observed concentrations and historical trends are well captured, in particular the differing NO2 and total NOx = NO + NO2 trends. Altogether, more than 1950 air quality monitoring stations in the EU are covered by the model, including more than 400 traffic stations and 70% of the critical stations. Together with its well-established bottom-up emission and dispersion calculation scheme, GAINS is thus able to bridge the scales from European-wide policies to impacts in street canyons. As an application of the model, we assess the evolution of attainment of NO2 limit values under current legislation until 2030. Strong improvements are expected with the introduction of the Euro 6 emission standard for light duty vehicles; however, for some major European cities, further measures may be required, in particular if aiming to achieve compliance at an earlier time.
Article
Full-text available
Ozone holds a certain fascination in atmospheric science. It is ubiquitous in the atmosphere, central to tropospheric oxidation chemistry, yet harmful to human and ecosystem health as well as being an important greenhouse gas. It is not emitted into the atmosphere but is a byproduct of the very oxidation chemistry it largely initiates. Much effort is focused on the reduction of surface levels of ozone owing to its health and vegetation impacts, but recent efforts to achieve reductions in exposure at a country scale have proved difficult to achieve owing to increases in background ozone at the zonal hemispheric scale. There is also a growing realisation that the role of ozone as a short-lived climate pollutant could be important in integrated air quality climate change mitigation. This review examines current understanding of the processes regulating tropospheric ozone at global to local scales from both measurements and models. It takes the view that knowledge across the scales is important for dealing with air quality and climate change in a synergistic manner. The review shows that there remain a number of clear challenges for ozone such as explaining surface trends, incorporating new chemical understanding, ozone–climate coupling, and a better assessment of impacts. There is a clear and present need to treat ozone across the range of scales, a transboundary issue, but with an emphasis on the hemispheric scales. New observational opportunities are offered both by satellites and small sensors that bridge the scales.
Article
Full-text available
The global nitrogen (N) cycle at the beginning of the 21st century has been shown to be strongly influenced by the inputs of reactive nitrogen (Nr) from human activities, including combustion-related NOx, industrial and agricultural N fixation, estimated to be 220 Tg N yr−1 in 2010, which is approximately equal to the sum of biological N fixation in unmanaged terrestrial and marine ecosystems. According to current projections, changes in climate and land use during the 21st century will increase both biological and anthropogenic fixation, bringing the total to approximately 600 Tg N yr−1 by around 2100. The fraction contributed directly by human activities is unlikely to increase substantially if increases in nitrogen use efficiency in agriculture are achieved and control measures on combustion-related emissions implemented. Some N-cycling processes emerge as particularly sensitive to climate change. One of the largest responses to climate in the processing of Nr is the emission to the atmosphere of NH3, which is estimated to increase from 65 Tg N yr−1 in 2008 to 93 Tg N yr−1 in 2100 assuming a change in global surface temperature of 5 °C in the absence of increased anthropogenic activity. With changes in emissions in response to increased demand for animal products the combined effect would be to increase NH3 emissions to 135 Tg N yr−1. Another major change is the effect of climate changes on aerosol composition and specifically the increased sublimation of NH4NO3 close to the ground to form HNO3 and NH3 in a warmer climate, which deposit more rapidly to terrestrial surfaces than aerosols. Inorganic aerosols over the polluted regions especially in Europe and North America were dominated by (NH4)2SO4 in the 1970s to 1980s, and large reductions in emissions of SO2 have removed most of the SO42− from the atmosphere in these regions. Inorganic aerosols from anthropogenic emissions are now dominated by NH4NO3, a volatile aerosol which contributes substantially to PM10 and human health effects globally as well as eutrophication and climate effects. The volatility of NH4NO3 and rapid dry deposition of the vapour phase dissociation products, HNO3 and NH3, is estimated to be reducing the transport distances, deposition footprints and inter-country exchange of Nr in these regions. There have been important policy initiatives on components of the global N cycle. These have been regional or country-based and have delivered substantial reductions of inputs of Nr to sensitive soils, waters and the atmosphere. To date there have been no attempts to develop a global strategy to regulate human inputs to the nitrogen cycle. However, considering the magnitude of global Nr use, potential future increases, and the very large leakage of Nr in many forms to soils, waters and the atmosphere, international action is required. Current legislation will not deliver the scale of reductions globally for recovery from the effects of Nr deposition on sensitive ecosystems, or a decline in N2O emissions to the global atmosphere. Such changes would require substantial improvements in nitrogen use efficiency across the global economy combined with optimization of transport and food consumption patterns. This would allow reductions in Nr use, inputs to the atmosphere and deposition to sensitive ecosystems. Such changes would offer substantial economic and environmental co-benefits which could help motivate the necessary actions.
Article
Full-text available
We present the Chemistry of Atmosphere-Forest Exchange (CAFE) model, a vertically-resolved 1-D chemical transport model designed to probe the details of near-surface reactive gas exchange. CAFE integrates all key processes, including turbulent diffusion, emission, deposition and chemistry, throughout the forest canopy and mixed layer. It is the first model of its kind to incorporate the Master Chemical Mechanism (MCM) and a suite of reactions for the oxidation of monoterpenes and sesquiterpenes, providing a more comprehensive description of the oxidative chemistry occurring within and above the forest. We use CAFE to simulate a young Ponderosa pine forest in the Sierra Nevada, CA. Utilizing meteorological constraints from the BEARPEX-2007 field campaign, we assess the sensitivity of modeled fluxes to parameterizations of diffusion, laminar sublayer resistance and radiation extinction. To characterize the general chemical environment of this forest, we also present modeled mixing ratio profiles of biogenic hydrocarbons, hydrogen oxides and reactive nitrogen. The vertical profiles of these species demonstrate a range of structures and gradients that reflect the interplay of physical and chemical processes within the forest canopy, which can influence net exchange.
Article
Full-text available
Eddy covariance (EC) flux measurements of the atmosphere/surface exchange of gases over an urban area are a direct way to improve and evaluate emissions inventories, and, in turn, to better understand urban atmospheric chemistry and the role that cities play in regional and global chemical cycles. As part of the MCMA-2003 study, we demonstrated the feasibility of using eddy covariance techniques to measure fluxes of selected volatile organic compounds (VOCs) and CO2 from a residential district of Mexico City (Velasco et al., 2005a, b). During the MILAGRO/MCMA-2006 field campaign, a second flux measurement study was conducted in a different district of Mexico City to corroborate the 2003 flux measurements, to expand the number of species measured, and to obtain additional data for evaluation of the local emissions inventory. Fluxes of CO2 and olefins were measured by the conventional EC technique using an open path CO2 sensor and a Fast Isoprene Sensor calibrated with a propylene standard. In addition, fluxes of toluene, benzene, methanol and C2-benzenes were measured using a virtual disjunct EC method with a Proton Transfer Reaction Mass Spectrometer. The flux measurements were analyzed in terms of diurnal patterns and vehicular activity and were compared with the most recent gridded local emissions inventory. In both studies, the results showed that the urban surface of Mexico City is a net source of CO2 and VOCs with significant contributions from vehicular traffic. Evaporative emissions from commercial and other anthropogenic activities were significant sources of toluene and methanol. The results show that the emissions inventory is in reasonable agreement with measured olefin and CO2 fluxes, while C2-benzenes and toluene emissions from evaporative sources are overestimated in the inventory. It appears that methanol emissions from mobile sources occur, but are not reported in the mobile emissions inventory.
Article
The formation of ozone through photochemical oxidation of alcohols, aldehydes, ketones, acids, and hydrocarbons, such as are present in gasoline, in the presence of small quantities of nitrogen oxides has been demonstrated. Ozone production without the addition of nitrogen oxides has been observed in the photochemical oxidation of biacetyl, bibutyryl, pyruvic acid, and butyl nitrite. The ozone produced in these reactions was identified by chemical and physical methods. The ozone formation is attributed to a peroxide radical chain reaction. The release of large quantities of hydrocarbons to the air and the simultaneous presence of nitrogen oxides from combustion processes explains the relatively high ozone content, and consequent severe rubber cracking, in the Los Angeles area. These findings should be considered in planning rubber storage facilities. In view of the irritating properties of the products formed by this photochemical oxidation, both hydrocarbons and nitrogen oxides should be considered as potential irritants when they occur simultaneously in the air at low concentrations.
Article
Volatile organic compounds (VOCs) emitted from residential wood and crop residue burning were measured in Colorado, US. When compared to the emissions from crop burning, residential wood burning exhibited markedly lower concentrations of acetonitrile, a commonly used biomass burning tracer. For both herbaceous and arboraceous fuels, the emissions of nitrogen-containing VOCs (NVOCs) strongly depend on the fuel nitrogen content; therefore, low NVOC emissions from residential wood burning result from the combustion of low-nitrogen fuel. Consequently, the emissions of compounds hazardous to human health, such as HNCO and HCN, and the formation of secondary pollutants, such as ozone generated by NOx, are likely to depend on fuel nitrogen. These results also demonstrate that acetonitrile may not be a suitable tracer for domestic burning in urban areas. Wood burning emissions may be best identified through analysis of the emissions profile rather than reliance on a single tracer species.
Article
In the atmosphere night time removal of volatile organic compounds (VOC) is initiated to a large extent by reaction with the nitrate radical (NO3) forming organic nitrates which partition between gas and particulate phase. Here we show based on particle phase measurements performed at a suburban site in the Netherlands that organic nitrates contribute substantially to particulate nitrate and organic mass. Comparisons with a chemistry transport model (CTM) indicate that most of the measured particulate organic nitrates are formed by NO3 oxidation. Using aerosol composition data from three intensive observation periods at numerous measurement sites across Europe, we conclude that organic nitrates are a considerable fraction of fine particulate matter (PM1) at the continental scale. Organic nitrates represent 34% to 44% of measured submicron aerosol nitrate and are found at all urban and rural sites, implying a substantial potential of PM reduction by NOx emission control.