10
Atmos. Meas. Tech., 6, 927–936, 2013 www.atmos-meas-tech.net/6/927/2013/ doi:10.5194/amt-6-927-2013 © Author(s) 2013. CC Attribution 3.0 License. Atmospheric Measurement Techniques Open Access Selective measurements of NO, NO 2 and NO y in the free troposphere using quantum cascade laser spectroscopy B. Tuzson 1 , K. Zeyer 1 , M. Steinbacher 1 , J. B. McManus 2 , D. D. Nelson 2 , M. S. Zahniser 2 , and L. Emmenegger 1 1 Empa, Swiss Federal Laboratories for Materials Science and Technology, Laboratory for Air Pollution and Environmental Technology, ¨ Uberlandstr. 129, 8600 D¨ ubendorf, Switzerland 2 Aerodyne Research, Inc., Center for Atmospheric and Environmental Chemistry, Billerica, MA 01821, USA Correspondence to: B. Tuzson ([email protected]) Received: 11 December 2012 – Published in Atmos. Meas. Tech. Discuss.: 20 December 2012 Revised: 8 March 2013 – Accepted: 19 March 2013 – Published: 9 April 2013 Abstract. A quantum cascade laser based absorption spec- trometer for continuous and direct measurements of NO and NO 2 was employed at the high-altitude monitoring site Jungfraujoch (3580 m a.s.l., Switzerland) during a three- month campaign in spring/summer 2012. The total reactive nitrogen, NO y , was also measured in the form of NO after conversion on a gold catalyst. The aim was to assess the suitability of the instrument for long-term monitoring of the main reactive nitrogen species under predominantly free tro- pospheric air conditions. A precision (1σ ) of 10 and 3 ppt for NO and NO 2 was achieved under field conditions after 180 s averaging time. The linear dynamic range of the instrument has been verified for both species from the detection limit up to 45 ppb. The spectrometer shared a common sampling inlet with a chemiluminescence-based analyzer. The comparison of the time series shows excellent agreement between the two techniques and demonstrates the adequacy of the laser spec- troscopic approach for this kind of demanding environmental applications. 1 Introduction In the troposphere, the abundance of reactive nitrogen species is primarily influenced by emissions from anthro- pogenic sources in the form of nitrogen oxides (NO x = NO + NO 2 ). Once in the atmosphere, these species are involved in a number of different chemical pathways and oxidized to a large variety of other reactive nitrogen species denoted as NO y (Crutzen, 1979; Logan, 1983). Despite their low atmo- spheric concentrations, nitrogen oxides represent an impor- tant factor regarding air quality given their contribution to local ozone production and secondary organic aerosol forma- tion. Furthermore, the adverse effect of NO 2 on the human respiratory system and the formation of nitric acid (HNO 3 ), which is the dominant sink of NO x (via dry and wet depo- sition), makes it a harmful air pollutant (Wesely and Hicks, 2000; Jacob, 2000). The significance of NO x in tropospheric chemistry triggered the interest in instrumental development. This is a challenging task given the substantial spatial and temporal variability of NO x , which ranges from a few ppt at remote locations to > 100 ppb in urban regions. Never- theless, it has been addressed over the last two decades by a wide variety of measurement techniques and by even larger number of instrumentations (e.g., see review by Clemitshaw, 2004 and references therein). Although recent intercompari- son studies demonstrate that several techniques perform well (Fehsenfeld et al., 1990; Fuchs et al., 2010), they still in- volve research grade instruments that are not yet established as routine monitoring tools. At present, chemiluminescence detection (CLD) can be considered as the standard technique, which is routinely used in field missions and air quality monitoring (Demerjian, 2000; GAW Report #195, 2011). Although the method has very good precision for NO, it is unable to directly measure NO 2 . Nevertheless, this has successfully been overcome by including the catalytic or photolytic conversion of NO 2 to NO prior to analysis. Besides the necessity of systematic de- termination of the conversion efficiency, the conversion may also lead to interferences of other nitrogen compounds that can readily be converted to NO (Dunlea et al., 2007; Stein- bacher et al., 2007), possibly resulting in overestimated NO 2 Published by Copernicus Publications on behalf of the European Geosciences Union.

AMT 6-927-2013 - Copernicus.org

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: AMT 6-927-2013 - Copernicus.org

Atmos. Meas. Tech., 6, 927–936, 2013www.atmos-meas-tech.net/6/927/2013/doi:10.5194/amt-6-927-2013© Author(s) 2013. CC Attribution 3.0 License.

EGU Journal Logos (RGB)

Advances in Geosciences

Open A

ccess

Natural Hazards and Earth System

Sciences

Open A

ccess

Annales Geophysicae

Open A

ccess

Nonlinear Processes in Geophysics

Open A

ccess

Atmospheric Chemistry

and Physics

Open A

ccess

Atmospheric Chemistry

and Physics

Open A

ccess

Discussions

Atmospheric Measurement

TechniquesO

pen Access

Atmospheric Measurement

Techniques

Open A

ccess

Discussions

Biogeosciences

Open A

ccess

Open A

ccess

BiogeosciencesDiscussions

Climate of the Past

Open A

ccess

Open A

ccess

Climate of the Past

Discussions

Earth System Dynamics

Open A

ccess

Open A

ccess

Earth System Dynamics

Discussions

GeoscientificInstrumentation

Methods andData Systems

Open A

ccess

GeoscientificInstrumentation

Methods andData Systems

Open A

ccess

Discussions

GeoscientificModel Development

Open A

ccess

Open A

ccess

GeoscientificModel Development

Discussions

Hydrology and Earth System

Sciences

Open A

ccess

Hydrology and Earth System

Sciences

Open A

ccess

Discussions

Ocean Science

Open A

ccess

Open A

ccess

Ocean ScienceDiscussions

Solid Earth

Open A

ccess

Open A

ccess

Solid EarthDiscussions

The Cryosphere

Open A

ccess

Open A

ccess

The CryosphereDiscussions

Natural Hazards and Earth System

Sciences

Open A

ccess

Discussions

Selective measurements of NO, NO2 and NOy in thefree troposphere using quantum cascade laser spectroscopy

B. Tuzson1, K. Zeyer1, M. Steinbacher1, J. B. McManus2, D. D. Nelson2, M. S. Zahniser2, and L. Emmenegger1

1Empa, Swiss Federal Laboratories for Materials Science and Technology, Laboratory for Air Pollution and EnvironmentalTechnology,Uberlandstr. 129, 8600 Dubendorf, Switzerland2Aerodyne Research, Inc., Center for Atmospheric and Environmental Chemistry, Billerica, MA 01821, USA

Correspondence to:B. Tuzson ([email protected])

Received: 11 December 2012 – Published in Atmos. Meas. Tech. Discuss.: 20 December 2012Revised: 8 March 2013 – Accepted: 19 March 2013 – Published: 9 April 2013

Abstract. A quantum cascade laser based absorption spec-trometer for continuous and direct measurements of NOand NO2 was employed at the high-altitude monitoringsite Jungfraujoch (3580 m a.s.l., Switzerland) during a three-month campaign in spring/summer 2012. The total reactivenitrogen, NOy, was also measured in the form of NO afterconversion on a gold catalyst. The aim was to assess thesuitability of the instrument for long-term monitoring of themain reactive nitrogen species under predominantly free tro-pospheric air conditions. A precision (1σ ) of 10 and 3 ppt forNO and NO2 was achieved under field conditions after 180 saveraging time. The linear dynamic range of the instrumenthas been verified for both species from the detection limit upto 45 ppb. The spectrometer shared a common sampling inletwith a chemiluminescence-based analyzer. The comparisonof the time series shows excellent agreement between the twotechniques and demonstrates the adequacy of the laser spec-troscopic approach for this kind of demanding environmentalapplications.

1 Introduction

In the troposphere, the abundance of reactive nitrogenspecies is primarily influenced by emissions from anthro-pogenic sources in the form of nitrogen oxides (NOx = NO+ NO2). Once in the atmosphere, these species are involvedin a number of different chemical pathways and oxidized toa large variety of other reactive nitrogen species denoted asNOy (Crutzen, 1979; Logan, 1983). Despite their low atmo-spheric concentrations, nitrogen oxides represent an impor-

tant factor regarding air quality given their contribution tolocal ozone production and secondary organic aerosol forma-tion. Furthermore, the adverse effect of NO2 on the humanrespiratory system and the formation of nitric acid (HNO3),which is the dominant sink of NOx (via dry and wet depo-sition), makes it a harmful air pollutant (Wesely and Hicks,2000; Jacob, 2000). The significance of NOx in troposphericchemistry triggered the interest in instrumental development.This is a challenging task given the substantial spatial andtemporal variability of NOx, which ranges from a few pptat remote locations to> 100 ppb in urban regions. Never-theless, it has been addressed over the last two decades bya wide variety of measurement techniques and by even largernumber of instrumentations (e.g., see review byClemitshaw,2004and references therein). Although recent intercompari-son studies demonstrate that several techniques perform well(Fehsenfeld et al., 1990; Fuchs et al., 2010), they still in-volve research grade instruments that are not yet establishedas routine monitoring tools.

At present, chemiluminescence detection (CLD) can beconsidered as the standard technique, which is routinelyused in field missions and air quality monitoring (Demerjian,2000; GAW Report #195, 2011). Although the method hasvery good precision for NO, it is unable to directly measureNO2. Nevertheless, this has successfully been overcome byincluding the catalytic or photolytic conversion of NO2 toNO prior to analysis. Besides the necessity of systematic de-termination of the conversion efficiency, the conversion mayalso lead to interferences of other nitrogen compounds thatcan readily be converted to NO (Dunlea et al., 2007; Stein-bacher et al., 2007), possibly resulting in overestimated NO2

Published by Copernicus Publications on behalf of the European Geosciences Union.

Page 2: AMT 6-927-2013 - Copernicus.org

928 B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy

readings. The alternating measurement of NO and NOx canalso lead to artifacts in the NO2 determination when the am-bient NOx levels experience rapid changes. Thus, a setupwith two individual CLDs (three in case of NO, NOx andNOy observations) is the preferred – but not inexpensive –configuration.

For the present study, we adopted the dual laser de-sign approach recently reported byMcManus et al.(2010).Our choice for quantum cascade laser based direct absorp-tion spectroscopy (QCLAS) was driven by two major fac-tors: (i) our previous campaigns performed in a large vari-ety of environments including grassland, forest and mountainsites (Tuzson et al., 2008, 2011; Neftel et al., 2010; Kam-mer et al., 2011) showed that QCLAS is a robust and reliabletechnique to be employed under various field and measure-ment conditions, and (ii) we obtained up to date the longest(more than four years of continuous operation) time series ofCO2 isotope ratio measurements at a mountain site, whichreflects the sturdy operation capability of the QCLs (Sturmet al., 2013). In addition to these considerations, the directabsorption technique offers a simple and straightforward so-lution to perform absolute spectroscopic quantification of themolecular species of interest.

In the following sections we describe the laser instrument,the measurement strategy and the experimental setup usedfor the campaign to perform direct and simultaneous mea-surements of NO and NO2. The validation of the time se-ries recorded during three months of continuous operation isdone against the standard CLD method. A short discussion isgiven on the discrepancy seen when employing the catalyticgold converter for NOy measurements.

2 Experimental

2.1 Instrumentation

2.1.1 NOx QCLAS

The spectroscopic instrument (Aerodyne Research, Inc.,USA) employed in this study is based on continuous wavequantum cascade laser (cw-QCL) technology. A recent re-view on optical design, performance characteristics and ap-plication fields of this kind of analyzer has been presentedby McManus et al.(2010). Briefly, the spectrometer embod-ies two room-temperature cw-QCLs (Alpes Lasers, Switzer-land) emitting in the spectral range of 1600 and 1900 cm−1,respectively. The individual laser beams are first collected bya pair of reflective objectives then combined and co-orientedby a dichroic mirror (LohnStar Optics, Inc., USA) and sentthrough a wedged beam splitter that results in three differ-ent beams. The main beam is further shaped by a series ofreflective optics and coupled into an astigmatic Herriott mul-tipass absorption cell (AMAC-200), which folds the laser ra-diation for a total number of 434 reflections. This results in

an effective path length of 204 m (McManus et al., 2011).When the exit condition of the cavity is fulfilled, the laserbeam leaves the cell through the same entrance hole andis then focused on a thermoelectrically cooled IR detector(Vigo Systems, Poland). The fraction of the laser beam re-flected from the first surface of the beam splitter is directedthrough a short (5 cm) cell filled with a 1: 1 mixture of NOand NO2 and subsequently aimed onto a second detector.This reference path assures the locking of the laser frequencyto well-defined absorption lines even in situations when theambient NOx concentrations drop below the limit of detec-tion or when the multipass cell is filled with NOx-scrubbedair for spectral background measurements. The third beamreflected from the back surface of the wedge is used for ac-curate frequency tuning rate determination of the lasers byplacing a germanium etalon in its path. This beam is alsofocused onto the reference detector. A flipping mechanismselects between the two beams (reference cell or etalon).

Beside the optical assembly design, the electronics havealso experienced some major improvements. A significant re-duction of the hardware size has been achieved by using com-pact USB-based data acquisition cards (NI USB-6251 OEM,National Instruments, Inc., USA) and embedded mini-PCs.Low noise laser drivers (Wavelength Electronics, Inc., USA)combined with dedicated power-supply design and thermalmanagement led to unprecedented stability. Details regard-ing the laser control and spectral fitting software (TDLWin-tel, Aerodyne Research, Inc., USA) have been discussedpreviously (Nelson et al., 2004).

Overall, the spectrometer consisted of two full-size 19-inch modules (6U) incorporating optics and electronics, re-spectively. These units were built into a rack together withthe custom-made gas handling and calibration module (3U),a UPS (2U, APC Smart series) to bridge short power out-ages, and a rack-mount thermoelectric liquid chiller (4U,Thermorack 300, Solid State Cooling Systems) used forcooling the lasers, IR-detectors and temperature stabilizingthe optical module. The total weight, including the vacuumpump, of the setup was 175 kg and its power consumptionapproximately 0.8 kW.

The major advantage of a direct spectroscopic detec-tion of NO2 is the immunity from interferences associatedwith photochemical conversion of NO2 to NO prior to de-tection. However, this technique may suffer from interfer-ences of other nature, such as overlapping absorption fea-tures from other molecular species or pressure broadeningeffects mainly due to water (Tuzson et al., 2010). Therefore,we performed a detailed investigation of the simulated ab-sorption spectra, generated based on typical experimental pa-rameters and the HITRAN database (Rothman et al., 2009),in the range of 1600 and 1900 cm−1, where NO2 and NOhave their fundamental ro-vibrational modes. The aim was toselect the strongest absorption lines possible with the strin-gent condition that the spectral interferences from water va-por or other trace gases are minimal. The best compromises

Atmos. Meas. Tech., 6, 927–936, 2013 www.atmos-meas-tech.net/6/927/2013/

Page 3: AMT 6-927-2013 - Copernicus.org

B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy 929

1.000

0.998

0.996

0.994

Tra

nsm

ittan

ce

1600.101600.001599.90

Wavenumber (cm-1

)

-1.0-0.50.00.51.0

Res

x10

-4

1.0000

0.9995

0.9990

0.9985

0.9980

Transm

ittance

1900.41900.31900.21900.1

Laser #1 Laser #2

NO CO2

H2O NO2

Fig. 1. The selected spectral windows with the absorption lines of NO and NO2 as measured in ambient air with the dual-laser spectrometerwith a multi-pass cell of 204 m path-length at 50 hPa sample pressure. The gray circles indicate measurement points, while the black line isa least-square fit to the data. The residual is shown in the top panel. The colored sticks indicate individual absorption lines of molecules thatabsorb within the scanning range of the lasers.

we could identify were the doublet lines located at 1900.0706and 1900.0816 cm−1 for NO and the group of lines in thevicinity of 1599.9 cm−1 for NO2 (see Fig.1). Simulationsof typical real-time scenarios involving low NOx concentra-tions and various humidity levels indicated low interferenceeffects (< 70 ppt NOx per volume % H2O), which was laterconfirmed experimentally. Based on these examinations, twocw-QCLs were selected correspondingly with single modeemission powers higher than 10 mW at the specified wave-lengths. The tuning range of the lasers was chosen such thatit also included a water vapor absorption line (Fig.1). Thus,the air sample humidity could be measured continuously to-gether with the NOx species, which allowed for analysis andcorrection of water-related effects.

The instrumental precision has been enhanced by address-ing the issue of interference fringes. These are mainly pro-duced by scattered light within the multipass cell and maysignificantly limit the achievable SNR improvement avail-able from multipass cells. They typically generate an ab-sorption equivalent noise level of 10−4 in the backgroundsignal. A significant improvement was obtained by carefulinvestigation and selection of circulation patterns that min-imize fringes near the absorption line-width and by usinghigh-reflectivity mirrors that lead to lower fringes by reduc-ing the ratio of scattered light to transmitted light (McManuset al., 2011). In addition, a mechanical modulation (dither-ing) of the cell mirrors by a piezo-crystal helps to average outfringes. As a result, a precision ofσmin = 1ppt for NO2 after12 min of averaging was achieved in the laboratory. Underfield conditions, however, this performance may suffer fromenvironmental influences. Therefore, the instrument’s stabil-ity was investigated on-site by measuring NOx-free air fora time period of 1 h with a detection bandwidth of 1 Hz andcalculating the associated Allan variance as a function of av-

eraging time (Werle et al., 1993). This two-sample variance-plot is shown in Fig.2 and allows defining the optimum cal-ibration time scale. The variance curve reveals flicker noisecontributions to the signal, manifested as deviations from theexpected 1/τ slope, at around 10 and 220 s. Similar behav-ior has been described byWerle (2011) and associated tooptical fringes. Following his theoretical framework, it waspossible to closely reproduce the observed deviations by as-suming two periodic fringe components with frequencies of4.3×10−2 Hz and 2.2×10−3 Hz, respectively. In fact, the 23 speriodicity correlates well with the PID-cycle of the temper-ature controller employed for stabilizing the optical module.The slow fringe is probably due to opto-mechanical drifts orvariations in the driver electronics, reflecting the temperaturefluctuations (±2◦C) experienced at the monitoring station.Nevertheless, the precision achieved at the monitoring site af-ter 180 s averaging time was 3 and 10 ppt, which correspondsto an absorbance noise level of 5.9×10−6 and 8.8×10−6 forNO2 and NO, respectively. Normalizing to the 204 m opticalpath-length, this corresponds to an absorbance per unit pathlength sensitivity of 5× 10−10 cm−1.

The strategy to cope with the drift issue was the fre-quent and periodic background (zero air) spectrum subtrac-tion from the measured spectrum. Therefore, a cost-effectiveNOx filter that consists of a stainless-steel container (1 L)filled with grained activated charcoal (Merck KGaA, Ger-many) was used to generate NOx-scrubbed air for spectralbackground measurements. The NOx filter efficiency andpotential systematic errors, when using the NOx scrubberfor background definition, have been verified by compar-ing the values measured for ambient air through the filterwith data obtained when directly measuring high purity (5.5)He and N2. As minor differences (< 30 ppt) were observed,we decided to check the NOx scrubber efficiency during the

www.atmos-meas-tech.net/6/927/2013/ Atmos. Meas. Tech., 6, 927–936, 2013

Page 4: AMT 6-927-2013 - Copernicus.org

930 B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy

10-6

10-5

10-4

10-3

Alla

n V

aria

nce

(ppb

2 )

1 10 100 1000Integration Time (s)

-50

0

50

NO

2 (p

pt)

3000200010000

collected data time (s)

-100

0

100

NO

(pp

t)

ffringe = 2.2·10-3

Hz

ffringe = 4.3·10-2

Hz

ffringe = 5·10-4

Hz

Fig. 2. Allan variance plots for NO and NO2 measured with thedual-laser instrument at the Jungfraujoch using NOx-free air as gassample. The gray lines correspond to calculated white noise plusfringes (ffringe) contributing as flicker noise (details in text).

campaign at every 2 h against N2. The filter solution was thepreferred method over the use of dry N2 as zero air, becausethe NOx filter does not significantly alter the water vapormixing ratio of the air flowing through it, and thus the zero airbetter reconciles the sampled air humidity and contributes inreducing possible detection artifacts that could be induced byspectral interferences due to water absorption lines. The mostefficient method to maintain the best stability and accuracy ofthe spectrometer has been achieved by automatically record-ing a background spectrum every 10 min and subtracting itfrom each subsequent sample spectrum before fitting. Un-certainties due to background drifts, mainly caused by tem-perature fluctuations, can be significant (> 40 ppt), if no ac-tion is taken. However, the configuration of the measurementcycle, as we have used in the field mission, allowed in thepost-processing step to minimize these uncertainties. Theircontributions to the 10 min averaged values were estimatedto 5 ± 4 ppt for NO2 and 11± 7 ppt for NO. More detailsabout the measurement cycle and data analysis are presentedin the Sect. 3.

2.1.2 Chemiluminescence NOx detection

The routine, long-term monitoring of NO, NOx and NOyat the Jungfraujoch is performed with commercially avail-

Fig. 3. Experimental setup at Jungfraujoch. The equipment withinthe dashed boxes was operated for a three-month period, while theother instrumentation is part of the permanent monitoring program.

able chemiluminescence analyzers (CraNOx, Eco Physics,Switzerland). The chemiluminescence detection (CLD) tech-nique, developed in the 1970s, is the reference method rec-ommended by the US EPA (Demerjian, 2000) and by Eu-ropean legislation (European Standard, EN 14211, 2012)for the measurement of NOx in monitoring networks. Themethod relies on the light emission from the electronicallyexcited NO2, a product generated in the gas-phase reactionof ambient NO with O3 added in excess to the air sample.The emission intensity is directly proportional to the concen-tration of NO in the air, and the emitted photons are detectedby a photomultiplier tube. NOx is measured as NO after pho-tolytic conversion of NO2. The atmospheric NO2 is then de-rived by subtraction of NO from the NOx signal (Ridley andHowlett, 1974; Ryerson et al., 2000; Sadanaga et al., 2010).NOy species are converted to NO on a heated gold catalyst(300◦C) in the presence of CO as a reducing agent (Faheyet al., 1985).

The measurement procedure as well as the calibrationstrategy has repeatedly been described in detail (Zellwegeret al., 2000; Pandey Deolal et al., 2012). Here we just re-call that, for the sensitive measurement of the NO–NOx–NOyspecies, three dedicated CLDs are used. The detection lim-its of the CLDs for NO and NOy after 2 min averaging are20 ppt, whereas it is 50 ppt for the NO2 channel due to theincomplete conversion and the determination of NO2 by thedifference of two measured (NO and NO+ converted NO2)signals. The overall measurement uncertainty (1σ ) is 9 % forNOy and 5 % for NOx, including CLD precision, the NOstandard uncertainty, and the conversion efficiencies of thephotolytic convertor (PLC). Data are processed and reportedas 10 min averages.

Since the PLC–CLD setup is used as the standard method,these measurements are taken as reference for validating theQCLAS data.

Atmos. Meas. Tech., 6, 927–936, 2013 www.atmos-meas-tech.net/6/927/2013/

Page 5: AMT 6-927-2013 - Copernicus.org

B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy 931

2.2 Site description and sampling setup

Field measurements took place between 22 March and25 June 2012 at the high alpine research station Jungfraujoch(JFJ, 46.55◦ N, 7.98◦ E , elevation 3580 m). The monitoringstation is one of the 28 global sites of the Global AtmosphereWatch (GAW) program of the World Meteorological Organi-zation (WMO). Given its location, the air sampled here isexpected to be representative for the lower free troposphereover Europe, which is mainly the case in the autumn and win-ter season, while in late spring and summer it can be influ-enced by the planetary boundary layer (Baltensperger et al.,1997; Lugauer et al., 1998; Zellweger et al., 2003). Long-term continuous in situ monitoring of trace gases such asNO, NO2, NOy, CO, and O3 is routinely performed by Empawithin the activities of the Swiss National Air Pollution Mon-itoring Network (NABEL).

The ambient air was sampled using the monitoring site’sheated stainless steel main inlet (inner diameter 9 cm) sit-uated about 3 m above ground, in which the ambient air isdrawn at a high flow rate of> 800 Lmin−1 (Fig. 3). Themain manifold is temperature controlled to keep the gas tem-perature at its lower end at constant 10◦C. Potential lossesof NOy in the sampling lines were minimized by placingthe gold converter close to the main inlet, connected withabout 1 m long PFA tubing. The time required for the airto reach the converter was about 2 s. The photolytic con-verter (PLC 762, Eco Physics, Switzerland) for the NO2measurement was placed next to the NOy converter. Down-stream of the converters, the air is drawn through separatePFA tubing to the chemiluminescence analyzers. The air forthe NO analysis is drawn directly from the main manifoldwith PFA tubing.

The NO analyzers are automatically calibrated every 39 hby measuring zero air and NO standard gas (5 ppm NO in N2,referenced against NIST Standard Reference Material (SRM2629a) and NMI Primary Reference Material (PRM BD11))diluted with zero air to approximately 48 ppb. The conver-sion efficiency of the converter is determined by gas phasetitration of NO with ozone every 78 h. The conversion effi-ciency of the PLC usually ranged from 74 to 58 %.

The QCLAS was operated alongside the CLDs for theduration of the study. An additional gold converter (CON765, Eco Physics, Switzerland) was mounted next to theNABEL NOy converter and shared the same CO source(purity 99.997 %, Messer-Griesheim GmbH), whereas thesampling line was independently connected to the maininlet. The selection between ambient air and air passingthrough the gold converter was accomplished by a 3-way all-Teflon solenoid valve (Teqcom Ind., USA). The flow rate of1 Lmin−1 was manually adjusted by a PTFE needle valveVM (PKM, Switzerland) mounted upstream of the cell (seeFig. 3), while the cell pressure (' 50 hPa) was establishedby selecting the appropriate pumping rate with an integratedpotentiometer on the oil-free diaphragm pump (N920 series,

KNF, Germany). Measurements were done continuously at1 s time resolution and post-processed to meet the 10 minaveraging time scale of the CLD.

3 Results and discussion

For free tropospheric air composition monitoring, high in-strumental precision alone is not sufficient to obtain the nec-essary accuracy. In fact, an optimal calibration and samplingstrategy is equally important and has a decisive influence onthe data quality. The strategy employed in this study exploitsthe very high precision and the fast response of the instru-ment when measuring zero air, and it accounts for drifts thatmay dominate the system at longer time scales. It consists ofvarious steps repeated in half-hour sequence. First, a back-ground spectrum is recorded using NOx-scrubbed air as de-scribed in the previous section. This is achieved by directingthe sampled ambient air through the NOx filter by switchingthe solenoid valve V3 (see Fig.3) and purging the absorp-tion cell for 120 s. The last 10 s are then averaged and usedto determine the spectral baseline and perform a backgroundcorrection by subtracting it from all the subsequent spectra.The second step involves measuring NO and NO2 simulta-neously for 10 min by drawing sample air directly throughthe absorption cell. In the next step, the solenoid valve V1is switched together with V2 to divert the air flow across thegold converter, and thus the NOy mixing ratio is measured asNO during another 10 min. The NO2 signal can later be usedto verify the converter efficiency for NO2 or to monitor base-line changes due to humidity variations or potential instru-mental drifts. After a 20 min measurement period, the sys-tem is switched back to measure zero air again, followed byanother background definition for the next cycle. The Allanvariance plot in Fig.2 suggests that at least two slow fringecomponents prevent long averaging times by introducing aninstrumental drift. This shows up in the time series of thebackground signal as a linear trend, and averaging over morethan 10 min would obviously reduce the measurement accu-racy. However, re-measuring the “zero-level” after the ambi-ent air measurements provides a convenient way to monitorthe amount of drift. The difference between the latter and ini-tial “zero-level” measurements was considered as real drift inthe system, and a linear interpolation between the two valueswas used to correct the 1 s measurement values within thisinterval prior to averaging. For more confidence in the “ab-solute” zero baseline determination, dry N2 from a gas cylin-der (purity 5.5) was measured instead of NOx-scrubbed airevery 2 h. This acts as an additional parameter for monitor-ing the performance of the NOx filter and to correct the NOxdata for systematic offsets. Over the whole campaign, theN2 measurements showed slightly lower NO (−32± 15 ppt)and slightly higher NO2 (15 ± 9 ppt) values than the NOx-scrubbed air. Although nearly negligible, the values were in-cluded as offset correction for the QCLAS data. Moreover,

www.atmos-meas-tech.net/6/927/2013/ Atmos. Meas. Tech., 6, 927–936, 2013

Page 6: AMT 6-927-2013 - Copernicus.org

932 B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy

-50

0

50

Res

(pp

t)

50403020100NONabel (ppb)

50

40

30

20

10

0

NO

QC

LAS (

ppb)

slope: 0.9727 ± 0.0004

R2 = 0.999

-0.20.00.2

Res

(pp

b)

50403020100NO2 Nabel (ppb)

50

40

30

20

10

0

NO

2 Q

CLA

S (

ppb)

slope: 0.987 ± 0.002

R2 = 0.994

Fig. 4. Calibration plots of the QCLAS data for NO and NO2. Theabscissa represents the values determined from standard gas dilu-tion (details in text). The fit (black line) is determined by weightedorthogonal distance regression and its residual shown in the upperpart of each plot.

it also aided in establishing the water-vapor-induced interfer-ences, which were applied to correct the measurement val-ues. The amount of this correction was empirically deter-mined as being−16.8 ± 3.7 ppt NO and−57.3 ± 3.5 pptNO2 for 1 % volume H2O. The relative humidity (RH) at JFJvaried between 20 and 100 %, depending on meteorologicalconditions and wind directions, but predominantly (' 87 %of all values) it was above 75 % RH. During the campaign,the temperature gradually increased from−11◦C (in March)to around 0◦C (in June), while the air pressure remainedrather constant at 654± 8 hPa. The water vapor was below0.3 % during the whole campaign, and thus its effect wasrather marginal, leading to a minor correction only (i.e., 5 to20 ppt in the measured NO and NO2 values). Since the watervapor corrections of the NOx data are based on the spectro-scopically measured H2O values, potential lag times have noinfluence on the data quality.

Dynamic dilutions of standard NO2 (96.3 ± 3.7 ppb) andNO (5 ± 0.1 ppm) gases were used to establish the span ofthe laser spectrometer for the NOx measurements from thedetection limit up to 50 ppb, and to calibrate the raw spectro-scopic data against the standard scale used by NABEL (seeFig.4 and previous section). The accuracy of this calibration

is dominated by the accuracy of the mass flow controllers(MFCs) used for the dilution. These MFCs (red-y smart se-ries, Voegtlin, Switzerland) were referenced against a high-accuracy venturi flowmeter (molbloc, DH Instruments/Fluke,USA) at Empa. The estimated uncertainties resulted from thedilution process were 0.1 % and 0.4 % for NO and NO2, re-spectively. Since NO2 gas mixtures in cylinders are known tobe unstable, the NO2 concentration was first determined witha calibrated CLD followed by QCLAS measurements anddynamic dilution (Fig.4). The sampling line was taken thesame, and thus we assumed that losses of NO2 on the metal-lic surfaces of the pressure regulator and MFCs are similar.Obviously, the total uncertainty in this case is mainly givenby systematic errors due to losses. To summarize, the totaluncertainty of the QCLAS at large NOx values is determinedby the accuracy of the calibration gases, which lies in thepercent range (±2–4 %), while at sub-ppb levels it is mainlygiven by the uncertainty of the “zero-level” measurements,which is below 10 ppt.

Finally, it is noted that all parameters discussed abovewere determined on the monitoring station at JFJ. This as-sured a direct link to the retrieved tropospheric NOx data.Figure5 shows an exemplary window of 21 days from thetime series recorded by both techniques during the campaign.The challenge of such tropospheric air measurements is wellillustrated by the NO data, where∼ 78% of the values arebelow 0.1 ppb. Despite of this, the variations in the mixingratio profiles of NO and NO2 compare very well between theCLD and the QCLAS. The quantitative agreement is illus-trated in Fig.6 in the form of correlation plots and the as-sociated regression results for NOx and NOy measured overthe three-month mission. For the analysis, we employed theorthogonal distance regression method, which includes theerrors in both dependent and independent variables (X- andY-values) as well as the weighting with the expected magni-tude of errors for each individual data point. The calibratedQCLAS data were averaged to 10 min (600 points) to matchthe time resolution of the CLD, and the error of the mean ofeach individual averaged value was used to filter out thosedata points that exhibited too large (3σ of the CLD detec-tion limit) variations. This minimized the scatter due to syn-chronization and time response differences between instru-ments in situations where the NOx concentrations changedabruptly during the averaging time interval. Statistical anal-ysis on the differences between all NOx data reported by thetwo analyzers showed that more than 94 % of individual datapoints agreed to within±50 ppt and were included for theregression analysis (Fig.6).

In the case of the NOy time series, we observed a system-atic discrepancy of about 0.26 ppb, which, over the measure-ment period, displayed short-term (< 12 h) variations of upto 1.2 ppb. Additional tests, involving different gases (am-bient air, dry N2 and diluted NO standard) as common in-put sample to the converters and alternating measurementswith QCLAS and CLD, revealed that the two converters

Atmos. Meas. Tech., 6, 927–936, 2013 www.atmos-meas-tech.net/6/927/2013/

Page 7: AMT 6-927-2013 - Copernicus.org

B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy 933

6

4

2

0

NO

y (p

pb)

31.03.2012 05.04.2012 10.04.2012 15.04.2012 20.04.2012Date

1.2

0.8

0.4

0.0

NO

(pp

b)

4

3

2

1

0

NO

2 (p

pb)

CraNOx QCLAS

Fig. 5. An exemplary sequence of NO, NO2 and NOy time series measured at the Jungfraujoch with the two different methods. Note thedifferent scaling of the y-axis. The gaps in the QCLAS data were caused by additional on-site works, such as calibration and converterinvestigation (see details in text).

1.0

0.5

0.0

NO

QC

LAS (

ppb)

1.00.50.0NO NABEL (ppb)

offset: -7.2 ± 0.4 pptslope: 1.002 ± 0.006

3

2

1

0

NO

2 Q

CLA

S (

ppb)

3210NO2 NABEL (ppb)

offset: -10.6 ± 0.8 pptslope: 0.989 ± 0.004

4

2

0

NO

y Q

CLA

S (

ppb)

420NOy NABEL (ppb)

offset: -95.7 ± 2.5 pptslope: 1.155 ± 0.002

3

2

1

0

NO

x Q

CLA

S (

ppb)

3210NOx NABEL (ppb)

offset: -17.3 ± 1 pptslope: 0.988 ± 0.004

Fig. 6. Correlation plots of the QCLAS and CraNOx data.The results of the regression analysis are given as slope andintercept values.

generate indeed different amounts of NOy in good agree-ment with the discrepancy seen for the time series. Figure7shows the clear bias seen by the QCLAS when switchingbetween the two Au converters sharing the same inlet withdiluted NO standard as sample gas. The difference disap-peared only when dry N2 was flowing through the convert-ers. This analysis eliminates suspected issues related to CLDtechnique (i.e., quenching, O3 and H2O effects), as it shows

3.8

3.6

3.4

3.2

3.0

2.8

NO

[ppb

]

22:45 23:00 23:15 23:30 23:45 00:00 Time

-0.2

-0.1

0.0

0.1

0.2

NO

2 [p

pb]

Fig. 7.NO and NO2 mixing ratios measured with the QCLAS whenswitching between the two gold converters measuring NO standardgas diluted with “zero” air. The gray areas correspond to the mea-surements of the converter used by the NABEL monitoring instru-mentation.

that the difference still exists when alternating the convert-ers and analyzing the NOy directly with the QCLAS. There-fore, the mismatch between the two gold converters has fullybeen attributed to their slightly different conversion efficien-cies. Even though, for NO2, the converters’ efficiencies werefound comparable (within 98.5 %), Au catalyst may have dif-ferent efficiencies for the various reactive nitrogen speciesor may suffer from contributions of other interfering com-pounds. According to systematic investigations, species suchas nitric acid (HNO3) and peroxyacyl nitrates (PANs) canhave conversion efficiencies up to 95 %, whereas ammonia

www.atmos-meas-tech.net/6/927/2013/ Atmos. Meas. Tech., 6, 927–936, 2013

Page 8: AMT 6-927-2013 - Copernicus.org

934 B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy

(NH3) and hydrogen cyanide (HCN) in the presence of watervapor show little but variable contribution (1–5 %), and theirmagnitude depends on the surface history and temperatureof the converter (Kliner et al., 1997; Fahey et al., 1985; Volz-Thomas et al., 2005). Although we had no speciation mea-surements for these compounds, literature data allow for anestimate of their contribution. Partitioning measurements ofNOy in the lower free troposphere (Zellweger et al., 2003) re-vealed that PAN is the dominant NOy species (∼ 38 %) dur-ing spring, followed by inorganic nitrate (HNO3 (7 %) andparticulate nitrate (17 %)). Mixing ratio values of∼ 200 pptfor HCN at the JFJ were reported byRinsland et al.(2000).Considering the above values and assuming that the conver-sion efficiency of the thermally unstable PAN in the gold con-verter should be very similar to the conversion efficiency forNO2, the observed mismatch between the converters remainsunclear. Interestingly, we also observed a significant differ-ence in the CO2 content of the gas samples after the con-verters, despite the similar CO flow rates (see Fig.3). Whilethe CO2 level remained unchanged after passing through theconverter with lower NOy values, the other converter almostdoubled the amount of CO2. Further investigation of the NOydata revealed that the temporal variability of its relative dif-ference closely follows the humidity fluctuations in the am-bient air. The observed discrepancies in NOy were system-atically becoming larger when the relative humidity droppedbelow 70 %. This may well be caused by issues related toHNO3 such as conversion efficiency of the gold converter,memory effects or inlet transmission efficiencies, but at thepresent we have no plausible explanation for this behavior.Finally, it should be noted that slightly lower NOy values(6–10 %) in the NABEL system have already been observedduring intercomparison campaigns (Zellweger et al., 2003;Pandey Deolal et al., 2012), and were attributed to poten-tial losses of HNO3 in the inlet system. This probably has tobe reconsidered, and further investigations of Au convertersas well as their regular comparisons are recommended forassessing the data quality.

4 Conclusions

The validation and long-term performance assessment ofa quantum cascade laser spectrometer was successfully ac-complished at the high-altitude research station JFJ by mea-suring the tropospheric NOx and NOy mixing ratios ac-companied by standard monitoring CLD technique. To ourknowledge, this is the first intercomparison study of this kind,involving a laser absorption technique for direct and simul-taneous measurement of both NO and NO2, under field con-ditions. The time series recorded with QCLAS show verygood correlation with the CLD data even at sub-ppb mix-ing ratio levels. A dedicated gas sampling/handling methodcombined with the high instrumental precision (< 10 ppt)delivered the accuracy and drift suppression needed for re-

liable background tropospheric air NOx monitoring. Sinceboth techniques give similar results, there is no fundamentalreason to replace the well-established and reliable CLD tech-nique by QCLAS. Nevertheless, the possibility to detect NO2selectively that is, without chemical conversion, and withhigh time resolution (1 Hz) is very attractive. Furthermore,significant advances in quantum cascade laser technology letus anticipate the appearance of compact multi-color sources,which would lead to substantial simplification of the opticaldesign resulting in more robust, compact, and cost-effectiveanalyzers.

The intercomparison of the NOy data revealed differencesthat could not be attributed to any of the measurement tech-niques, but rather were associated with the conversion effi-ciency of the gold catalyst. This was confirmed in a dedi-cated set of experiments, but the exact reason for the differ-ent behavior of the two seemingly identical catalysts remainsunclear. Obviously, there are limitations of the existing NOymethod, which may require the development of more robustconversion schemes or, preferably, measurement techniquesfor compound-specific determination of the NOy species.

Acknowledgements.This work was supported by the FederalOffice for the Environment (FOEN), Switzerland and the SwissNational Science Foundation (SNSF/R’Equip). We thank the Inter-national Foundation High Altitude Research Stations Jungfraujochand Gornergrat (HFSJG) for access to the facilities at the ResearchStation Jungfraujoch and the custodians for on-site support. BeatSchwarzenbach (Empa) is acknowledged for calibrating our NO2working standard. We also acknowledge the valuable contributionsof Ryan McGovern, Stanley Huang and Michael Agnese (ARI).

Edited by: P. Werle

References

Baltensperger, U., Gaggeler, H. W., Jost, D. T., Lugauer, M.,Schwikowski, M., Weingartner, E., and Seibert, P.:Aerosol climatology at the high-alpine site Jungfraujoch,Switzerland, J. Geophys. Res.-Atmos., 102, 19707–19715,doi:10.1029/97JD00928, 1997.

Clemitshaw, K. C.: A review of instrumentation and measurementtechniques for ground-based and airborne field studies of gas-phase tropospheric chemistry, Crit. Rev. Env. Sci. Tec., 34, 1–108, doi:10.1080/10643380490265117, 2004.

Crutzen, P. J.: The role of NO and NO2 in the chemistry of thetroposphere and stratosphere, Annu. Rev. Earth Pl. Sc., 7, 443–472, doi:10.1146/annurev.ea.07.050179.002303, 1979.

Demerjian, K. L.: A review of national monitoring net-works in North America, Atmos. Environ., 34, 1861–1884,doi:10.1016/S1352-2310(99)00452-5, 2000.

Dunlea, E. J., Herndon, S. C., Nelson, D. D., Volkamer, R. M.,San Martini, F., Sheehy, P. M., Zahniser, M. S., Shorter, J. H.,Wormhoudt, J. C., Lamb, B. K., Allwine, E. J., Gaffney, J. S.,Marley, N. A., Grutter, M., Marquez, C., Blanco, S., Car-denas, B., Retama, A., Ramos Villegas, C. R., Kolb, C. E.,

Atmos. Meas. Tech., 6, 927–936, 2013 www.atmos-meas-tech.net/6/927/2013/

Page 9: AMT 6-927-2013 - Copernicus.org

B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy 935

Molina, L. T., and Molina, M. J.: Evaluation of nitrogen diox-ide chemiluminescence monitors in a polluted urban environ-ment, Atmos. Chem. Phys., 7, 2691–2704, doi:10.5194/acp-7-2691-2007, 2007.

Fahey, D. W., Eubank, C. S., Hubler, G., and Fehsenfeld, F. C.:Evaluation of a catalytic reduction technique for the measure-ment of total reactive odd-nitrogen NOy in the atmosphere, J.Atmos. Chem., 3, 435–468, doi:10.1007/BF00053871, 1985.

Fehsenfeld, F. C., Williams, E. J., Buhr, M. P., Hubler, G., Lang-ford, A. O., Murphy, P. C., Parrish, D. D., Norton, R. B.,Fahey, D. W., Drummond, J. W., Mackay, G. I., Roychowd-hury, U. K., Hovermale, C., Mohnen, V. A., Demerjian, K. L.,Galvin, P. J., Calvert, J. G., Ridley, B. A., Grahek, F.,Heikes, B. G., Kok, G. L., Shetter, J. D., Walega, J. G.,Elsworth, C. M., and Schiff, H. I.: Intercomparison of NO2measurement techniques, J. Geophys. Res., 95, 3579–3597,doi:10.1029/JD095iD04p03579, 1990.

Fuchs, H., Ball, S. M., Bohn, B., Brauers, T., Cohen, R. C.,Dorn, H.-P., Dube, W. P., Fry, J. L., Haseler, R., Heitmann, U.,Jones, R. L., Kleffmann, J., Mentel, T. F., Musgen, P., Rohrer, F.,Rollins, A. W., Ruth, A. A., Kiendler-Scharr, A., Schlosser, E.,Shillings, A. J. L., Tillmann, R., Varma, R. M., Venables, D. S.,Villena Tapia, G., Wahner, A., Wegener, R., Wooldridge, P. J.,and Brown, S. S.: Intercomparison of measurements of NO2concentrations in the atmosphere simulation chamber SAPHIRduring the NO3Comp campaign, Atmos. Meas. Tech., 3, 21–37,doi:10.5194/amt-3-21-2010, 2010.

GAW Report #195. A WMO/GAW Expert Workshop on GlobalLong-term Measurements of Nitrogen Oxides and Recom-mendations for GAW Nitrogen Oxides Network (WMO TDNo. 1570), available at:www.wmo.int/pages/prog/arep/gaw/documents/FinalGAW 195 TD No 1570web.pdf, 2011.

Jacob, D. J.: Heterogeneous chemistry and tropospheric ozone,Atmos. Environ., 34, 2131–2159, doi:10.1016/S1352-2310(99)00462-8, 2000.

Kammer, A., Tuzson, B., Emmenegger, L., Knohl, A., Mohn, J.,and Hagedorn, F.: Application of a quantum cascade laser-basedspectrometer in a closed chamber system for real-timeδ13C andδ18O measurements of soil-respired CO2, Agr. Forest Meteorol.,151, 39–48, doi:10.1016/j.agrformet.2010.09.001, 2011.

Kliner, D. A. V., Daube, B. C., Burley, J. D., and Wofsy, S. C.: Labo-ratory investigation of the catalytic reduction technique for mea-surement of atmospheric NOy, J. Geophys. Res.-Atmos., 102,10759–10776,doi:10.1029/96JD03816, 1997.

Logan, J. A.: Nitrogen oxides in the troposphere: globaland regional budgets, J. Geophys. Res., 88, 10785–10807,doi:10.1029/JC088iC15p10785, 1983.

Lugauer, M., Baltensperger, U., Furger, M., Gaggeler, H. W.,Jost, D. T., Schwikowski, M., and Wanner, H.: Aerosol transportto the high Alpine sites Jungfraujoch (3454 m a.s.l.) and ColleGnifetti (4452 m a.s.l.), Tellus B, 50, 76–92,doi:10.1034/j.1600-0889.1998.00006.x, 1998.

McManus, J. B., Zahniser, M. S., Nelson, D. D., Shorter, J. H.,Herndon, S., Wood, E., and Wehr, R.: Application of quantumcascade lasers to high-precision atmospheric trace gas measure-ments, Opt. Eng., 49, 111124–111135, doi:10.1117/1.3498782,2010.

McManus, J. B., Zahniser, M. S., and Nelson, D. D.: Dual quan-tum cascade laser trace gas instrument with astigmatic Her-

riott cell at high pass number, Appl. Opt., 50, A74–A85,doi:10.1364/AO.50.000A74, 2011.

Neftel, A., Ammann, C., Fischer, C., Spirig, C., Conen, F.,Emmenegger, L., Tuzson, B., and Wahlen, S.: N2O ex-change over managed grassland: Application of a quan-tum cascade laser spectrometer for micrometeorologicalflux measurements, Agr. Forest Meteorol., 150, 775–785,doi:10.1016/j.agrformet.2009.07.013, 2010.

Nelson, D. D., McManus, B., Urbanski, S., Herndon, S., and Zah-niser, M. S.: High precision measurements of atmospheric ni-trous oxide and methane using thermoelectrically cooled mid-infrared quantum cascade lasers and detectors, Spectrochim.Acta A, 60, 3325–3335, doi:10.1016/j.saa.2004.01.033, 2004.

Pandey Deolal, S., Brunner, D., Steinbacher, M., Weers, U., andStaehelin, J.: Long-term in situ measurements of NOx and NOyat Jungfraujoch 1998–2009: time series analysis and evaluation,Atmos. Chem. Phys., 12, 2551–2566, doi:10.5194/acp-12-2551-2012, 2012.

Ridley, B. A. and Howlett, L. C.: An instrument for nitric oxidemeasurements in the stratosphere, Rev. Sci. Instrum., 45, 742–748, doi:10.1063/1.1686726, 1974.

Rinsland, C. P., Mahieu, E., Zander, R., Demoulin, P., For-rer, J., and Buchmann, B.: Free tropospheric CO, C2H6, andHCN above central Europe: Recent measurements from theJungfraujoch station including the detection of elevated columnsduring 1998, J. Geophys. Res.-Atmos., 105, 24235–24249,doi:10.1029/2000JD900371, 2000.

Rothman, L. S., Gordon, I. E., Barbe, A., Chris Benner, D.,Bernath, P. F., Birk, M., Boudon, V., Brown, L. R., Campar-gue, A., Champion, J. -P., Chance, K., Coudert, L. H., Dana, V.,Devi, V. M., Fally, S., Flaud, J. -M., Gamache, R. R., Gold-man, A., Jacquemart, D., Kleiner, I., Lacome, N., Lafferty, W. J.,Mandin, J. -Y., Massie, S. T., Mikhailenko, S. N., Miller, C. E.,Moazzen-Ahmadi, N., Naumenko, O. V., Nikitin, A. V., Or-phal, J., Perevalov, V. I., Perrin, A., Predoi-Cross, A., Rins-land, C. P., Rotger, M.,Simeckova, M., Smith, M. A. H.,Sung, K., Tashkun, S. A., Tennyson, J., Toth, R. A., Van-daele, A. C. and Vander Auwera, J.: The HITRAN 2008 molec-ular spectroscopic database, J. Quant. Spectrosc. Ra., 110, 533–572, doi:10.1016/j.jqsrt.2009.02.013, 2009.

Ryerson, T. B., Williams, E. J., and Fehsenfeld, F. C.:An efficient photolysis system for fast-response NO2 mea-surements, J. Geophys. Res.-Atmos., 105, 26447–26461,doi:10.1029/2000JD900389, 2000.

Sadanaga, Y., Fukumori, Y., Kobashi, T., Nagata, M., Take-naka, N., and Bandow, H.: Development of a selective light-emitting diode photolytic NO2 converter for continuously mea-suring NO2 in the atmosphere, Anal. Chem., 82, 9234–9239,doi:10.1021/ac101703z, 2010.

Steinbacher, M., Zellweger, C., Schwarzenbach, B., Bugmann, S.,Buchmann, B., Ordonez, C., Prevot, A. S. H., and Hueglin, C.:Nitrogen oxide measurements at rural sites in Switzerland: biasof conventional measurement techniques, J. Geophys. Res.-Atmos., 112, D11307,doi:10.1029/2006JD007971, 2007.

Sturm, P., Tuzson, B., Henne, S., and Emmenegger, L.: Track-ing isotopic signatures of CO2 at Jungfraujoch with laser spec-troscopy: analytical improvements and exemplary results, At-mos. Meas. Tech. Discuss., 6, 423–459, doi:10.5194/amtd-6-423-2013, 2013.

www.atmos-meas-tech.net/6/927/2013/ Atmos. Meas. Tech., 6, 927–936, 2013

Page 10: AMT 6-927-2013 - Copernicus.org

936 B. Tuzson et al.: Selective measurements of tropospheric NOx and NOy

Tuzson, B., Mohn, J., Zeeman, M. J., Werner, R. A., Eugster, W.,Zahniser, M. S., Nelson, D. D., McManus, J. B., and Emmeneg-ger, L.: High precision and continuous field measurements ofδ13C andδ18O in carbon dioxide with a cryogen-free QCLAS,Appl. Phys. B, 92, 451–458, doi:10.1007/s00340-008-3085-4,2008.

Tuzson, B., Hiller, R. V., Zeyer, K., Eugster, W., Neftel, A., Am-mann, C., and Emmenegger, L.: Field intercomparison of two op-tical analyzers for CH4 eddy covariance flux measurements, At-mos. Meas. Tech., 3, 1519–1531, doi:10.5194/amt-3-1519-2010,2010.

Tuzson, B., Henne, S., Brunner, D., Steinbacher, M., Mohn, J.,Buchmann, B., and Emmenegger, L.: Continuous isotopiccomposition measurements of tropospheric CO2 at Jungfrau-joch (3580 m a.s.l.), Switzerland: real-time observation of re-gional pollution events, Atmos. Chem. Phys., 11, 1685–1696,doi:10.5194/acp-11-1685-2011, 2011.

Volz-Thomas, A., Berg, M., Heil, T., Houben, N., Lerner, A., Pet-rick, W., Raak, D., and Patz, H.-W.: Measurements of total oddnitrogen (NOy) aboard MOZAIC in-service aircraft: instrumentdesign, operation and performance, Atmos. Chem. Phys., 5, 583–595, doi:10.5194/acp-5-583-2005, 2005.

Werle, P.: Accuracy and precision of laser spectrometers for tracegas sensing in the presence of optical fringes and atmosphericturbulence, Appl. Phys. B, 102, 313–329, doi:10.1007/s00340-010-4165-9, 2011.

Werle, P., Mucke, R., and Slemr, F.: The limits of signal averag-ing in atmospheric trace-gas monitoring by tunable diode-laserabsorption spectroscopy (TDLAS), Appl. Phys. B, 57, 131–139,doi:10.1007/BF00425997, 1993.

Wesely, M. L. and Hicks, B. B.: A review of the current status ofknowledge on dry deposition, Atmos. Environ., 34, 2261–2282,doi:10.1016/S1352-2310(99)00467-7, 2000.

Zellweger, C., Ammann, M., Buchmann, B., Hofer, P., Lugauer, M.,Ruttimann, R., Streit, N., Weingartner, E., and Baltensperger, U.:Summertime NOy speciation at the Jungfraujoch, 3580 m abovesea level, Switzerland, J. Geophys. Res.-Atmos., 105, 6655–6667,doi:10.1029/1999JD901126, 2000.

Zellweger, C., Forrer, J., Hofer, P., Nyeki, S., Schwarzenbach, B.,Weingartner, E., Ammann, M., and Baltensperger, U.: Partition-ing of reactive nitrogen (NOy) and dependence on meteorolog-ical conditions in the lower free troposphere, Atmos. Chem.Phys., 3, 779–796, doi:10.5194/acp-3-779-2003, 2003.

Atmos. Meas. Tech., 6, 927–936, 2013 www.atmos-meas-tech.net/6/927/2013/