You are on page 1of 11

REVIEW OF SCIENTIFIC INSTRUMENTS 77, 034101 共2006兲

Aircraft instrument for simultaneous, in situ measurement of NO3 and N2O5


via pulsed cavity ring-down spectroscopy
William P. Dubé
NOAA Earth System Research Laboratory, Chemical Sciences Division, 325 Broadway, Boulder, Colorado
80305 and Cooperative Institute for Research in the Environmental Sciences, University of Colorado, Boulder,
Colorado 80309
Steven S. Brown
NOAA Earth System Research Laboratory, Chemical Sciences Division, 325 Broadway,
Boulder, Colorado 80305
Hans D. Osthoff
NOAA Earth System Research Laboratory, Chemical Sciences Division, 325 Broadway, Boulder, Colorado
80305 and Cooperative Institute for Research in the Environmental Sciences, University of Colorado, Boulder,
Colorado 80309
Maya R. Nunleya兲
NOAA Earth System Research Laboratory, R/CSD2, 235 Broadway, Boulder, CO 80305
Steven J. Ciciora, Mark W. Paris, and Richard J. McLaughlin
NOAA Earth System Research Laboratory, Chemical Sciences Division, 325 Broadway,
Boulder, Colorado 80305
A. R. Ravishankara
NOAA Earth System Research Laboratory, Chemical Sciences Division, 325 Broadway, Boulder, Colorado
80305 and Department of Chemistry and Biochemistry, University of Colorado, Boulder, Colorado 80309
共Received 17 November 2005; accepted 22 January 2006; published online 23 March 2006兲
This article describes a cavity ring-down spectrometer 共CaRDS兲 specifically designed and
constructed for installation on the NOAA WP-3D Orion 共P-3兲 aircraft for sensitive, rapid in situ
measurement of NO3 and N2O5. While similar to our previously described CaRDS instrument, this
instrument has significant improvements in the signal-to-noise ratio, the time resolution, and in
overall size and weight. Additionally, the instrument utilizes a custom-built, automated filter changer
that was designed and constructed to meet the requirement for removal of particulate matter in the
airflow while allowing fully autonomous instrument operation. The CaRDS instrument has a
laboratory detection sensitivity of 4 ⫻ 10−11 cm−1 in absorbance or 0.1 pptv 共pptv denotes parts per
trillion volume兲 of NO3 in a 1 s average, although the typical detection sensitivities encountered in
the field were 0.5 pptv for NO3 and 1 pptv for N2O5. The instrument accuracy is 25% for NO3 and
20%–40% for N2O5, limited mainly by the uncertainty in the inlet transmission. The instrument has
been deployed on the P-3 aircraft as part of a major field campaign in the summer of 2004 and
during several ground and tower deployments near Boulder, CO. © 2006 American Institute of
Physics. 关DOI: 10.1063/1.2176058兴

I. INTRODUCTION hydrolysis of N2O5 is rapid.3 Furthermore, NO3 is a strong


oxidant for a variety of hydrocarbons and sulfur
The nitrate radical 共NO3兲 and dinitrogen pentoxide
compounds.4 Thus, the fate of NO3 and N2O5 is important to
共N2O5兲 are important reactive intermediates in the atmo-
the cycling and removal of NOx and to the oxidation of vola-
spheric chemistry of nitrogen oxides.1 They form from the
tile organic compounds 共VOC兲 in the nighttime atmosphere,
oxidation of NO2 by O3 and the further association of NO2
and measurements of their concentrations are of substantial
with NO3.
interest to atmospheric chemistry.
NO2 + O3 → NO3 + O2 , 共1兲 The majority of the database for atmospheric measure-
ments of NO3 comes from open path differential optical ab-
NO3 + NO2 ↔ N2O5 . 共2兲 sorption spectroscopy 共DOAS兲, which was first applied to
the detection of NO3 over 25 years ago.5 These observations
Their formation is effectively reversed during the day by the
have been carried out both passively, using natural light
reaction of NO3 with NO 共present in sunlight from NO2
sources 共e.g., moonlight and starlight兲 for column density
photolysis兲 and by efficient NO3 photolysis 共lifetime of ⬃5 s
measurements,6 and actively, using artificial light sources
in direct sunlight兲.2 At night, reaction 共1兲 can consume a
aligned over a multikilometer path.7 Reported detection sen-
significant fraction 共50%–90%兲 of NO2, particularly if the
sitivities for active DOAS detection of NO3 range from
0.2 pptv8 共pptv denotes parts per trillion by volume兲 to sev-
a兲
Present address: Cargill, Inc., Wichita, Kansas 67214. eral pptv9 with integration times on the order of minutes.

0034-6748/2006/77共3兲/034101/11/$23.00 77, 034101-1 © 2006 American Institute of Physics

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-2 Dubé et al. Rev. Sci. Instrum. 77, 034101 共2006兲

Because these techniques rely on the strong visible absorp-


tion bands of the nitrate radical, they do not detect N2O5 ␣= 冉 冊
1 1 1

c ␶ ␶0
, 共3a兲
directly, although average N2O5 concentrations over the
DOAS path can be calculated from simultaneous measure- RL
ments of NO3, NO2, and temperature, as long as the variabil- 共A兲 = ␣ . 共3b兲
␴TE
ity in NO2 and NO3 is small over the path.10 Until recently,
the only in situ technique for NO3 detection was cryogenic Here c is the speed of light, and RL is the ratio of the cavity
collection followed by electron spin resonance 共ESR兲 length to the length over which the sample is present, as
spectroscopy.11 In the last few years, several sensitive, fast described below.
time response, real-time in situ techniques have been devel- There are now several varieties of cavity ring-down
oped for NO3 detection. One of the principal advantages of spectroscopy and related techniques such as integrated cavity
these instruments is the capability to detect N2O5 directly via output spectroscopy 共ICOS兲, using both pulsed and cw laser
its thermal decomposition to NO3 in a heated inlet. One re- light sources and broadband sources.22 Pulsed CaRDS takes
cent approach has been cavity ring-down spectroscopy advantage of the passive coupling that occurs between nano-
共CaRDS兲, a high sensitivity direct absorption technique. This second time scale pulsed lasers and high finesse optical cavi-
method relies on the strong visible absorption bands of NO3, ties as a result of the overlap of a relatively wide-bandwidth
similar to DOAS. Our group,12,13 Simpson and coworkers,14 laser source 共e.g., Nd:YAG 共YAG denotes yttrium aluminum
and Ball et al.15 have all reported CaRDS based instruments, garnet兲 laser pumped dye lasers, ⌬␯ ⬃ 3 GHz兲 with multiple
with detection sensitivities for NO3 or for the sum of NO3 cavity resonances 共free spectral range ⬃0.2 GHz兲 and the
+ N2O5 of 0.3– 2 pptv with integration times from seconds to broadening of cavity resonances in the presence of laser
minutes. Recently Ball and coworkers16 have also demon- pulses comparable to the cavity round trip time.23 The pas-
strated laboratory detection of NO3 by a related technique sive coupling allows optimization of the instrument sensitiv-
using a broadband light source. Laser induced fluorescence ity through the use of mirrors of arbitrarily high reflectivity.
共LIF兲 has been recently applied to NO3 and N2O5 detection, The disadvantages to pulsed CaRDS include the larger size
as demonstrated by both Wood et al.17,18 and Matsumoto et and power requirements of pulsed laser systems in compari-
al.,19 albeit with somewhat larger detection limits and inte- son to cw diode lasers, and the laser bandwidth effects23,24
gration times 共4 – 80 pptv in 10– 1 min兲 in the instruments that makes the detection of compounds on discrete mid or
reported to date. Finally, Slusher et al.20 have applied chemi- near infrared rovibrational features more difficult. Therefore,
cal ionization mass spectrometry 共CIMS兲 to detect the sum for atmospheric applications, pulsed CaRDS is most appli-
of NO3 and N2O5 by their reactions with I−, both of which cable to broad band absorption and/or scattering measure-
yield NO−3 . ments, such as the quantification of aerosol extinction25 or
This article describes a CaRDS instrument designed and the detection of visible absorbers with wide bandwidths such
tested for aircraft deployment. Included in the description are as NO3 or NO2.26
improvements in the sensitivity and the inlet design for NO3
and N2O5 sampling implemented since the previous descrip- A. Optics, electronics, and vibration isolation
tion of our ground and ship based instrument.13 Also in- Figure 1 shows the layout of the optical and inlet system
cluded are the data from a test flight on the NOAA WP-3D for the aircraft NO3 and N2O5 CaRDS instrument. Optically,
Orion aircraft in Tampa, FL in March of 2004, the first 共to the instrument is similar to the previously described ground-
our knowledge兲 example of NO3 or N2O5 measurements based, two channel NO3 and N2O5 instrument.13 The light
from an aircraft. To date, the instrument has been deployed at source is a small footprint, pulsed, Nd:YAG laser 共Big Sky
ground and tower sites near our laboratories in Boulder, CO, Laser, ultra-CFR兲 pumped dye laser 共Dakota Technologies,
and on the NOAA P-3 during the New England Air Quality Northern Lights兲 that produces ⬃1 mJ, 6 ns pulses at a rep-
Study in Portsmouth, NH 共http://www.al.noaa.gov/2004/兲 for etition rate of 33 Hz. The dye laser wavelength is tuned to a
six weeks in July and August, 2004. point on the broad maximum of the NO3 absorption spec-
trum near 662 nm 共Ref. 27兲 that is not resonant with any of
the weak, discrete water vapor absorptions in this region.
II. INSTRUMENT DESCRIPTION
Since the 662 nm NO3 absorption band varies only slowly
Cavity ring-down spectroscopy 共CaRDS兲 is a high sen- with wavelength 共bandwidth= 80 cm−1兲, it is possible to
sitivity, direct absorption technique based on the measure- choose an arbitrary point near the peak of the NO3 absorp-
ment of the time constant for single exponential decay of tion band that minimizes the optical absorption due to water
light intensity from an optical cavity.21 Measurement of this vapor. Periodic scans of the water vapor absorption spectrum
time constant in the presence 共␶兲 and absence 共␶0兲 of an in ambient air between 661– 663 nm provided a convenient
absorbing species gives an absolute measurement of the ab- means of calibrating the dye laser wavelength, which was
sorbance 共␣, cm−1, also referred to as the absorption coeffi- stable to ±0.02 nm. The breadth of the NO3 absorption band
cient or the extinction兲, and, therefore, the absorber’s con- relative to the linewidth of the dye laser 共0.5 cm−1兲 ensures
centration 关共A兲, molecules cm−3兴 as long as the absorption that there are no laser linewidth artifacts in the exponential
cross section 共␴, cm2 molecule−1兲 and the transmission effi- ring-down traces.23 The beam from the dye laser propagates
ciency 共TE兲 for the target compound through the inlet system through an optical isolator and a series of irises and lenses
are known: 共50– 100 cm focal length兲 to minimize the spot size at the far

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-3 Aircraft measurement of NO3 and N2O5 Rev. Sci. Instrum. 77, 034101 共2006兲

FIG. 3. 共a兲 Schematic of the instrument frame and optical table mounted in
the P-3 fuselage. 共b兲 Inlet fast flow system and winglet.

tivity rather than Rayleigh scattering.28 The mirror reflectiv-


ity, corrected for Rayleigh losses, is 99.9987% 共13 ppm total
FIG. 1. 共Color online兲 Layout of the optical and inlet systems.
loss兲. The effective path length, shown on the top axis, is
55 km for one e-folding time and several hundred kilometers
end of the optical cavities. A 50/ 50 beamsplitter separates over the five lifetime range over which we typically fit each
the laser output into two beams of approximately equal in- decay.
tensity, each of which is aligned on the axis of two separate The optical system is mounted on an aluminum bread-
optical cavities. board attached to a custom-made steel frame that houses the
Each cavity is a stable resonator of near-confocal geom- computer, electronics, photomultiplier tubes, flow and tem-
etry, consisting of two 1 in. diameter, 1 m radius of curva- perature controllers, laser power supply, dye laser circulators,
ture dielectric coated, high reflectivity mirrors, separated by gas cylinders 共NO, zero air, and He兲, and pumps. Figure 3共a兲
93 cm. Each mirror is mounted in a purge volume separated shows the arrangement and its mount inside the cabin of the
from the air sample by an orifice 共⬃1 / 4 in. diameter兲 and P-3. The overall size and weight of the instrument are 110
purged with dry zero air to maintain mirror cleanliness. The ⫻ 52.5⫻ 88 cm3 and 135 kg, respectively. Although this in-
purge volumes have proven effective in preventing degrada- strument is still relatively large, it is considerably smaller
tion in mirror reflectivity, which has been observed to remain than our prototype instrument and it meets the requirements
constant or improve slightly, even during extended field de- for installation on the P-3. We anticipate further size and
ployments of six weeks or longer. The purge volumes are weight reductions in the future.
mounted in commercial, 2 in. optical mounts. Light transmit- Vibration is always a concern on a mobile platform, par-
ted through the rear cavity mirror is collected using optical ticularly a propeller-driven aircraft. Previous experience with
fiber and imaged on to a small, red-sensitive photomultiplier field deployments of the pulsed CaRDS system on a ship29
tube through a bandpass filter centered at 660 nm to reject has shown, however, that it is rather robust with respect to
the stray light at other wavelengths. Ring-down traces are vibration. Thus, our approach to vibration isolation was a bit
digitized on a 16 bit oscilloscope card mounted in an exter- different than typical. We chose not to vibration isolate the
nal peripheral component interconnect 共PCI兲 bus and ana- optical table from the frame and the support systems. In-
lyzed on a laptop computer. Ring-down time constants are stead, we chose to purposely couple the optical table to the
measured at 1 s time resolution from a linear fit to the natural frame and the nonvibrating support equipments 共such as the
logarithm of a coadded trace of 33 individual decay profiles. computer, gas bottles, flow controllers, etc.兲 We vibration
Figure 2 shows an example ring-down trace and exponential isolated this much larger mass from the aircraft and the vi-
fit at 662 nm. The time constant in 830 mbar of laboratory brating support equipment. This additional “complicated”
air 共Boulder, CO兲 is 183 ␮s, limited mainly by mirror reflec- mass enhanced the vibration isolation between the optical
table and the aircraft. Vibration that leaked by the wire-rope
vibration isolators was attenuated, to a degree, by the pur-
posefully supple support frame. Moreover, the instruments
and support equipment help “soak up” and damp the leakage
vibration before it reaches the optical table. This approach
proved quite successful as there were no vibration-related
problems during the deployment of the instrument. The in-
strument sensitivity, as measured by the precision of the ring-
down time constant, was identical when the aircraft was on
the ground without engines running and in flight.

B. Inlet system
The inlet system consists of three parts: an initial, fast
FIG. 2. Ring-down trace with single exponential fit. flow system that brings the air sample from outside the

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-4 Dubé et al. Rev. Sci. Instrum. 77, 034101 共2006兲

boundary layer of the aircraft into the cabin; a filter that inlet and exhaust. The inside of the manifold is machined to
removes ambient aerosol from the sampled air flow; and a the shape of the tubing and the fittings at the ends, and is
slower flow system that brings the air sample on to the axes split into a top and bottom half that bolt together. This design
of the two ring-down cells. Aside from the filter housings, all allows periodic removal and exchange of the Teflon tubing to
components of the cells and all inlet plumbing are con- help suppress wall losses due to aging of the tubing between
structed of commercial 1 / 2 in. outer diameter 共OD兲 PFA fit- flights. It also maintains the alignment of the PFA tubing on
tings and tubings. the center of the optical detection axis. The temperature of
The sampling position of the NO3 and N2O5 instrument the manifold is actively set by the data acquisition software
on its first deployment on the NOAA P-3 was somewhat to match that of the air flow immediately upstream of the
rearward in the aircraft, where the estimated boundary layer manifold. This active temperature control is necessary to
thickness is ⬃30 cm. To ensure sampling of air outside this suppress turbulent flow noise 共see below兲 arising from tem-
boundary layer, the fast flow portion of the inlet was housed perature gradients within the sampled air flow. The air
in a winglet 关Fig. 3共b兲兴 that protruded 46 cm from the outer sample entering the heated channel passes through two dif-
fuselage skin. The sampling tube sheath is a thin-wall copper ferent heating stages of 45 cm each. The first is a precon-
pipe with an ID slightly larger than 1 / 2 in. 共⬃13 mm兲. This verter stage that rapidly heats the incoming airflow and in-
copper sheath protects and guides an inner liner made of duces rapid thermal decomposition of N2O5. The
1 / 2 in. 共⬃13 mm兲 PFA tubing. The inner liner of the sam- preconverter stage is controlled to a temperature of 120 ° C
pling inlet can be easily removed and replaced from inside using a copper sheath wrapped in heat tape and silicone rub-
the aircraft cabin. The flow rate through the fast flow system ber insulation. A second, similarly constructed heating stage
is 40– 50 LPM 共LPM denotes liter per minute兲, which is the with a set temperature of 70 or 75 ° C allows the air sample
sum of the full capacity of an oil-free diaphragm pump, mea- to cool prior to entering the axis of the ring-down cell. Here,
sured at 25 LPM, and the controlled flows in the slow flow a single aluminum manifold, similar to that on the ambient
system, which totaled approximately 20 LPM. The flow rate channel but controlled at 70 or 75 ° C, maintains a constant
in this section of the inlet is not critical so long as it is fast temperature in the flow along the entire optical detection
enough to suppress wall loss of NO3 and N2O5 共see below兲. region.
The slow flow system samples air from the fast flow mani- In spite of our attempts to maintain constant temperature
fold at a branch point just inside the aircraft fuselage. The air throughout the flow along the optical detection axis, we
sample is filtered in the slow flow part of the inlet, immedi- found that there is considerable noise in the heated channel.
ately downstream of the branch point with the fast flow sys- The noise is the result, presumably, of the temperature gra-
tem. Filters are housed in and changed by an automated de- dient between the ambient temperature purge volumes and
vice, described in the next section. As shown in Fig. 1, the the heated air sample. This gradient may induce some vari-
slow flow system branches into two air samples immediately able thermal lensing at the interface that slightly alters the
below the filter, each of which is controlled with electronic alignment in the optical cavity. Introduction of a mixture of
mass flow controllers at the exhaust and a second diaphragm 20% He in zero air to the purge volumes to approximately
pump common to both flows. The pressure in the ring-down match the refractive index of the heated air sample to the
cells is not controlled and varies with the ambient pressure cooler purge volume partially offsets this noise 共⬃2⫻ reduc-
on the exterior of the aircraft. To maintain constant residence tion兲. Similar effects have been observed at surface pressure
time within the slow flow systems, the data acquisition soft- in the ambient channel when the air sample was much colder
ware sets the two flow rates to a constant volumetric flow as than the purge volumes. In summertime aircraft sampling,
the cell pressure changes with aircraft altitude. The flow rate thermally induced noise was not important on the ambient
on the ambient temperature side 共NO3 detection兲 is 16 LPM side since the coldest air was sampled at the highest altitude,
volumetric, while on the heated side 共N2O5 + NO3 detection兲, and the effect was smaller at lower pressures.
it is 8 LPM. These flow rates allow adequate residence time
for NO titration of NO3 to zero the absorption signal and C. Automated filter changer
thermal conversion of N2O5 to NO3 while minimizing wall The largest component to the visible extinction signal in
loss 共see below兲. ambient air is typically due to scattering and absorption by
Insulation along the common section of the slow flow aerosol. Not only is the aerosol extinction signal large, but it
system serves to minimize the temperature rise between the is also gives rise to a statistically noisy CaRDS signal due to
共typically兲 colder air, sampled outside of the aircraft and the small numbers of large, optically active particles in the de-
cabin temperature. This is important to minimize shifts in the tection volume.25 Consequently, high sensitivity gas-phase
NO3 – N2O5 equilibrium, which can give an artifact in the pulsed CaRDS measurements require filtering of the aerosol.
NO3 signal when the ratio of N2O5 to NO3 is large.30 We This requirement poses a significant challenge for the sam-
measure this temperature shift and model the required cor- pling of reactive trace gases and leads to by far the largest
rection to the NO3 signal. During the NEAQS 2004 deploy- single source of uncertainty in sampling NO3 and N2O5 with
ment, this correction was generally less than 2% but occa- this instrument. To minimize loss of NO3 and N2O5, we use
sionally as large as 15% in large, cold NO2 plumes. As 47 mm diameter, 2 ␮m pore size, 25 ␮m thick poly tet-
shown in Fig. 1, an aluminum manifold with Peltier elements rafluoroethylene 共PTFE兲 filters 共Pall-Gelman兲. Because of
maintains a constant temperature in the ambient channel aging effects in the filters 共see below兲, we typically change
along the ring-down axis, including the elbow fittings at the them at 1 h intervals, although in some cases the filter

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-5 Aircraft measurement of NO3 and N2O5 Rev. Sci. Instrum. 77, 034101 共2006兲

Nitric oxide reacts rapidly with NO3 to form NO2,31 whose


absorption cross section at 662 nm is nearly four orders of
magnitude smaller than that of NO3.32

NO + NO3 → 2NO2 , 共4兲

k共298 K兲 = 2.6 ⫻ 10−11 cm3 molecule−1 s−1

Addition of ⬃30 ppbv 共ppbv denotes parts per billion by


volume兲 of NO just above the filter is just sufficient to titrate
NO3 to ⬎99% completion prior to entry into the ambient
channel
共⬃0.25 s兲. There is a brief 共0.2 s兲 purge of the NO addition
line with zero air subsequent to the zero acquisition to pre-
vent leakage of residual NO from this line into the sample
stream during signal acquisition. Contamination of NO2 in
the NO mixture is small 共⬍10−3 of total NOx present as
NO2兲 and is undetectable as an absorption signal in either
channel. Reaction of NO with ambient O3 to produce NO2 is
too slow to produce a measurable signal on the ambient
FIG. 4. Schematic of the sealed, automated filter changer. channel. However, on the heated channel, it does generate a
small negative offset 共equivalent to 艋0.3 pptv NO3兲 because
of the longer residence time and the increase in the rate co-
changes occur more frequently. Thus, for the instrument to
efficient for this reaction with temperature. We model this
be fully autonomous, an automated filter changer had to be
offset from the residence time in the heated inlet and the
devised and constructed. It is shown in Fig. 4.
measured concentration of ambient O3 and subtract it from
The filter changer was designed with minimal gas con-
the data.
tact surface in the flow path to minimize wall losses. Two
The combination of 662 nm optical detection and titra-
funnel-like PFA pistons are sealed to either side of the PTFE
tion with NO provides a highly specific detection scheme for
filter. The pistons are withdrawn by applying a vacuum to the
NO3 and in the case of thermal conversion to NO3, N2O5 as
back side of the pistons to allow the filter to be exchanged.
well.13 Prior to the 2004 deployments of the NO3 CaRDS
They are then moved back into place and held firmly against
instruments, we had observed no evidence for any interfering
the filter by applying compressed air to the back sides
absorber. During the New England Air Quality Study
of the pistons. A rotating disk transports filters from a
共NEAQS兲, we observed a small interference in the heated
stack contained in the fresh filter reservoir. The fresh filter is
channel that correlated with the observed structure in per-
then moved into place between the two pistons by the
oxyacetylnitrate 共PAN兲. This interference was typically
rotating disk. Simultaneously, the used filter is moved away
⬍1 pptv, with maximum values of 2 – 4 pptv. It does not
from between the pistons and transported to the used filter
affect the interpretation of nocturnal results from this instru-
reservoir.
ment, although it is of approximately the same order as N2O5
The interior of the device is sealed from the exterior to
observed during the day. Its analysis has been described in a
allow the filter changer to operate in the pressurized cabin of
recent publication.33
the aircraft. The interior of the filter changer case is vented to
The NO titration allows periodic zero determination 共ev-
the aircraft exterior 共ambient兲 pressure. Thus, when the pis-
ery 0.5– 5 min兲 without the need to scan the wavelength of
tons retract to change a filter, there is no significant air move-
the dye laser away from the peak of the NO3 absorption
ment that might buffet the filter or contaminate the inlet.
spectrum. Variability in the background absorption of NO2
Also, cabin pressure integrity is maintained during a filter
and O3, and to a much lesser extent water vapor, can produce
change. Since the pistons are actuated pneumatically,
small, spurious structure in the 662 nm absorption signal.
changes in the ambient pressure on the piston faces alter the
During surface based sampling, the interpolation of the base-
net force acting on the pistons. Springs are needed to assist
line variation between subsequent zero measurements is suf-
withdrawal of the pistons at high altitude because the pres-
ficient to track these changes; however, large instantaneous
sure on the piston faces is not adequate to reliably move the
changes in O3 and NO2 are commonplace during aircraft
pistons. The springs are chosen carefully so as to provide
sampling. Each 1 ppbv of O3 or NO2 produces a baseline
adequate force to withdraw the pistons, but not so much
shift equivalent to 0.1 and 0.15 pptv of NO3, respectively.
force that they prevent the pistons from sealing reliably to
We corrected the instrument baseline for the measured varia-
the filter.
tion in the concentrations of NO2 and O3 prior to analysis of
the raw signals; the interference from water vapor was ex-
D. Zero acquisition ceedingly small. By far the largest changes in the instrument
The instrument zero is determined by addition of a small baseline were due to Rayleigh scattering as the inlet pressure
quantity of NO to the air sample, as described previously.13 changed with aircraft altitude. Fortunately, these changes

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-6 Dubé et al. Rev. Sci. Instrum. 77, 034101 共2006兲

Repeated measurements gave a first order loss rate coeffi-


cient and associated uncertainty of kloss = 0.2± 0.05 s−1. When
the instrument is deployed in the aircraft, the calculated resi-
dence time from the tip of the inlet to the midpoint of the
ambient ring-down cell is 艋0.5 s 共艋0.1 s in the fast flow
section and ⬃0.4 s in the slow flow leading to the ambient
side兲, giving a total loss of 共10± 3兲% for NO3 to the walls of
the tubing.
The lower trace in Fig. 5 shows the measurement of the
transmission through the tubing plus the custom-made pis-
tons that hold the membrane filter in place 共there is no filter
present for the data in the figure兲. The pistons have a ma-
chined PFA surface rather than the molded surface present in
FIG. 5. Measurement of the wall loss rate 共upper trace, solid points兲 through the commercial PFA fittings and tubing in the rest of the inlet
1 / 2 in. PFA tubing from exponential fit to the ratio of the NO3 concentration
measured in two channels against the mean residence time between them. system. Although the loss rate coefficient is similar with the
Repeated measurements gave a net loss rate of 0.2± 0.05 s−1. The lower pistons in place, the intercept at zero residence time is
trace 共open circles兲 shows the additional loss due to insertion of the ma- smaller, indicating that the pistons act as a point source loss
chined, PFA pistons that hold the filter in place. The loss rate is similar in
for NO3. Several measurements using different sets of pis-
this trace, but there is a constant offset indicating a point source loss for
NO3 in the flow system. tons gave a net loss of 共10± 5兲%. Of the materials that we
tested 共including PFA, PTFE, Kel-f, etc.兲, machined PFA had
the lowest surface losses, although the machining process
tended to be smooth and continuous, so that interpolation of
clearly degrades the NO3 transmission efficiency compared
the baseline between successive zero acquisitions produced
to commercial, molded components.
no large, spurious signal offsets.
The other important loss for NO3 is the Teflon mem-
brane filters themselves. This loss has been measured by re-
E. Sampling efficiency and measurement accuracy
peated insertion and removal of a filter into a flow of NO3 in
There are a number of factors that govern the accuracy zero air and is 共7 ± 2兲% for clean filters. 共Loss of N2O5 on
of the NO3 and N2O5 measurements, but the most important clean filters is negligible.兲 Our previous measurement
is the inlet transmission efficiency, which is denoted by TE in showed a decrease in the NO3 transmission efficiency
Eq. 共3b兲. Therefore, this section mainly outlines improve- through the filters of up to 10% per hour due to filter aging
ments to the inlet system for NO3 and N2O5 detection since while sampling in ambient air. Thus, filter aging represents a
our previously described prototype instrument13,29 had an in- potentially large uncertainty for the NO3 concentration mea-
let constructed of 1 in. diameter glass cells coated with ha- surement because the aging rate may vary with the compo-
locarbon wax. The two channel 662 nm absorption measure- sition of the air mass. There are several diagnostics to check
ment capability provides a method of accurately determining for the effects of filter aging while in the field, including
the wall loss rates for NO3 in different types of inlet tubing discontinuities upon change of a filter in the measured con-
and the transmission efficiency of various inlet components, centrations of NO3 and N2O5: the steady state lifetimes of
such as the filter and its housing, from a measurement of the these compounds 共i.e., their concentrations divided by their
same NO3 and/or N2O5 sample flowed sequentially through production rate from the reaction of NO2 with O3兲 and the
the two ring-down cells. Synthetic NO3 or N2O5 samples ratio of N2O5 to NO3. For the majority 共⬎90% 兲 of all filter
were prepared from a flow of zero air over crystalline N2O5 changes on the aircraft, there was no evidence for filter aging
共synthesized by a standard method34兲 in a trap at −78 ° C. effects, although in a few cases there were clear discontinui-
The sample was twice diluted in zero air and heated, when ties across filter changes. Data showing such discontinuities
desired, to convert N2O5 to NO3, giving a sample with a have been either corrected for the estimated, time-dependent
concentration range of 109 – 1010 molecules cm−3 or roughly change in the transmission efficiency or else discarded from
40– 400 pptv at atmospheric pressure. the data set. Based on the previously reported aging rate for
the NO3 transmission efficiency of 10% per hour and the 7%
1. NO3 accuracy transmission efficiency for clean filters, the NO3 filter trans-
Figure 5 shows a measurement of the wall loss rate for mission efficiency used for reduction of field data is
NO3 measured in 3 / 8 in. ID PFA tubing. The transmission 共88± 7兲%. The net transmission efficiency for NO3 is the
between the two channels was fitted to an exponential decay product of that for the inlet tubing, the filter housings and the
as a function of the residence time, calculated as the plug filter, or 共71± 13兲%. The nearly 30% loss of NO3 radicals
flow time between the midpoints of the two ring-down cells. through the current inlet system illustrates the difficulties of
Plug flow is appropriate for a system in which the calculated sampling this reactive compound.
rate constant for diffusion of NO3 to the wall of the cylin- We verified the inlet transmission efficiency by a second
drical tubing35 共kd ⬃ 2 s−1兲 is an order of magnitude faster method, conversion of NO3 to NO2 via reaction 共4兲 followed
than the observed NO3 loss rate coefficient, and in which the by pulsed CaRDS measurement of the resulting NO2 concen-
90° bends in the path between the two ring-down cells tend tration at 532 nm.26 We are currently developing this method
to mix the gas flow at the center with that at the walls. into an in-field calibration for NO3 and N2O5. The net trans-

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-7 Aircraft measurement of NO3 and N2O5 Rev. Sci. Instrum. 77, 034101 共2006兲

mission measured from conversion to NO2 was 共71± 12兲%


and was constant for an inlet that had sampled ambient air
over the course of 8 h. The error estimate on the NO2 con-
version test comes from the uncertainties in the cross sec-
tions for NO2 at 532 nm 共6%兲,26 the NO3 absorption cross
section at 662 nm 共10%兲,27 and experimental scatter 共3%兲.
Since the error in the two measurements of the transmission
efficiency are comparable, we take the net transmission effi-
ciency as 共70± 13兲%. The resulting uncertainty due to inlet
transmission in the measured NO3 concentrations in Eq. 共3兲
is 20%.
The remaining uncertainty in the NO3 measurement
comes from the cross section for absorption at 662 nm and
the ratio of the length over which the air sample is present to
the total length of the optical cavity. We have used values for
the temperature dependent cross sections from Yokelson
et al.,27
␴共NO3兲 = 关4.56 − 0.007 87 ⫻ T共K兲兴
⫻ 10−17 cm2 molecule−1 共5兲
The stated uncertainty in the peak cross section for NO3 is
10%. The ratio of distance between the mirrors to the length
over which the absorber is present in the cavity, or RL in Eq.
共3b兲, is defined by the lengths of the purge and sample vol-
umes shown in Fig. 1. We have measured this value both
from the physical dimensions of the cells and from the mea-
sured visible absorption in the Chappius bands due to known
quantities of O3, as described previously.13,36 The RL deter-
mined from physical cell dimensions was 1.22, while that
determined from O3 absorption was 1.16. The 5% difference
was presumably due to diffusion of O3 into the orifices sepa-
rating the purge and sample volumes. 共Independent tests de-
termined that it was not due to a limitation in the accuracy of FIG. 6. 共a兲 Measurement of the NO3 wall loss rate in the heated system at
the ring-down absorption measurement.兲 For NO3 detection, typical flows, converter, and measurement cell temperatures. 共b, c兲 Measure-
we assume that the RL value of 1.22 is correct since the more ment of the N2O5 conversion efficiency as a function of heated channel flow
rate and preconverter temperature in the field using two different CaRDS
reactive NO3 will be lost to the walls upon diffusion outside instruments sampling from the same manifold. The reference instrument had
of the sample volume. However, we also assume an uncer- a flow of 6 SLPM and a preconverter temperature of 120 ° C. Variation of
tainty of 5% in this quantity. Combining the cross section the flow and temperature on the second instrument showed that the conver-
sion is not a strong function of the conditions over a flow range of
uncertainty 共10%兲, the uncertainty in RL 共5%兲 and the trans-
4 – 8 SLPM and preconverter temperature of 90– 140 ° C.
mission efficiency 共20%兲 yields a net uncertainty in NO3 of
25%. The accuracy of the measured ambient NO3 concentra-
tions is dominated by the uncertainty in the inlet transmis- converter and reaction heater shown in Fig. 1, and the second
sion efficiency. ring-down cell was heated. Figure 6共a兲 shows that the first
order loss rate coefficient was the same as that measured at
2. N2O5 accuracy ambient temperature or 0.2± 0.05 s−1. In the field instrument,
The sampling efficiency for N2O5 depends on its trans- the residence time through the heated sections of the flow
mission through the ambient temperature sections of the inlet system to the midpoint of the ring-down cell is 0.7 s, so that
prior to the split that divides the flow to the ambient and the net transmission efficiency for NO3 in the heated portion
heated channels, the conversion efficiency for thermal disso- of the inlet is 共87± 3兲%. A verification of this result via con-
ciation of N2O5 to NO3 and the loss rate for NO3 in the version of NO3 to NO2 as described above gave a transmis-
heated sections of the inlet. The N2O5 wall loss rate is slow, sion for NO3 through this system of 共93± 12兲%.
and the conversion efficiency for thermal dissociation of Measurement of NO3 or NO3 + N2O5 at 75 ° C depends
N2O5 to NO3 is essentially quantitative as long as the con- upon the parameterization of the temperature dependence of
ditions of inlet temperature and flow have been optimized. the NO3 absorption cross sections given above. This cross
Therefore, the wall loss rate for NO3 in the heated inlet is the section is extrapolated from the temperature dependence
most important contribution to the N2O5 sampling efficiency. measured below room temperature, and decreases by ap-
Measurement of the heated wall loss rate coefficient for NO3 proximately 20% between room temperature and 75 ° C. We
was similar to that described above, except that the test sec- have independently verified the high temperature cross sec-
tion between the two ring-down cells consisted of the pre- tion in our laboratory from conversion of NO3 to NO2, find-

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-8 Dubé et al. Rev. Sci. Instrum. 77, 034101 共2006兲

ing a value of 共1.85± 0.20兲 ⫻ 10−17 cm2 molecule−1, in good tainty in the NO3 cross section and a 5% uncertainty in RL,
agreement with the extrapolation. A recent calculation of the as described above. The fractional uncertainty in the quantity
temperature dependence of the NO3 absorption cross section T2NO3 can be taken only as the uncertainty in the transmis-
above room temperature37 is also in agreement with the ex- sion through the heated inlet since the uncertainties in this
trapolated of the Yokelson et al. data and with the new mea- transmission efficiency are the same as those used to derive
surements. the concentration in the ambient channel. This quantity is
Optimization of the conditions of temperature and flow ␦共T2兲 = 33%. The fractional uncertainty in T1 is ␦共T1兲 = 13%.
for the conversion of N2O5 to NO3 came from tests similar to Propagating the errors in these quantities gives an expression
the wall loss rate measurement in the laboratory and from for the uncertainty in the N2O5 measurement,
in-field comparison of the aircraft NO3 and N2O5 CaRDS
instrument to our second, ground-based instrument. Figures
6共b兲 and 6共c兲 show the result of this optimization from the
␦共N2O5兲 = 再 关␦共SUM兲SUM兴2 + 关␦共T2兲T2NO3兴2
共SUM − T2NO3兲2
+ ␦共T1兲2 冎 1/2

field data under cold conditions 共T = 264– 266 K兲 and large 共7兲
NO2 mixing ratio 共⬃20 ppbv兲 where the predicted ratio of Under conditions where NO3 ⬇ N2O5, the uncertainty due to
N2O5 to NO3 was ⬎1000: 1. To test the efficiency of the the NO3 subtraction is the most important, while for N2O5
N2O5 conversion, the conditions in the heated channel on Ⰷ NO3, the uncertainty depends mainly on the efficiency of
one of the instruments were held constant at a flow of N2O5 conversion and sampling as described above. The
6 SLPM 共SLPM denotes standard liter per minute兲 and a range of values for ␦共N2O5兲 from field data lie between
preconverter temperature of 120 ° C, while the heated chan- 20%–40% for N2O5 ⬎ 5 pptv, but increase to unity for
nel flow and temperature were varied in the other instrument. N2O5 ⬍ 1 pptv.
The figure shows that conversion of N2O5 is not a strong
function of either temperature or flow under these conditions.
There is wide latitude to set the flow between 4 – 8 SLPM F. Detection sensitivities
and the preconverter temperature between 100 and 140 ° C The detection sensitivity of the CaRDS instrument can
while still maintaining an N2O5 signal within 90% of its be calculated from the limit of Eqs. 共3兲 and 共3a兲 in which ␶
optimum value. Actual flows of 5 – 6 SLPM, preconverter approaches ␶0,38
temperature of 110– 120 ° C, and reaction heater and mea-
冑2 ⌬␶
surement cell temperature of 70– 75 ° C were used for the ␣min = , 共8兲
aircraft measurements during the NEAQS 2004 campaign. c ␶20
Although these tests show that conversion of N2O5 to NO3
goes fully to its equilibrium value at the 70– 75 ° C tempera- RL
关NO3兴min = ␣min . 共9兲
ture of the heated ring-down cell, there is still a small frac- ␴TE
tion of N2O5 that remains undissociated at equilibrium Here ⌬␶ is the smallest measurable change in ␶0, normally
at this temperature. This equilibrium necessitates an taken as the standard deviation in ␶0 from repeated measure-
NO2-dependent correction ranging from 0%–5% for NO2 ments. Under laboratory conditions, ␶0 ⬇ 180 ␮s and ␴共␶兲
from 0 – 20 ppbv to the measured sum of NO3 + N2O5 in the ⬇ 0.03 ␮s 关where ␴共␶兲 has been taken as the standard devia-
heated channel 共equilibrium constants from the NASA/JPL tion from repeated 1 s measurements兴, giving ␣min = 4
recommendation31兲. Typical in-flight corrections were less ⫻ 10−11 cm−1 Hz−1/2. For a 662 nm NO3 absorption cross
than 2%. section of 2.2⫻ 10−17 cm2 molecule−1, RL value of 1.22, and
Finally, in calculating the concentration of N2O5, it is TE value of 0.7, the corresponding detection sensitivity for
important to accurately subtract the contribution from NO3 NO3 is 0.1 pptv at 298 K and 1 atm. In the field, the sensi-
to the signal in the heated channel, which can be comparable tivity is nearly always larger because there is a variable
to or even larger than that due to N2O5 for summer condi- amount of noise on both channels that depends on the flow
tions. Based on the losses, loss rates, and residence times through the ring-down cells, possibly as a result of turbu-
outlined above, the net transmission of the NO3 signal lence in the flow under conditions with modest temperature
through the entire flow system leading to the heated ring- gradients not present in the laboratory. During summer con-
down cell, including the tubing, filter, and filter housings, is ditions, actual detection sensitivities for NO3 varied between
共60± 15兲%. 0.2– 0.5 pptv 共1␴兲, while under cold conditions at low alti-
The measured N2O5 mixing ratio depends on the sum of tude 共i.e., higher ambient pressure兲, the sensitivity for NO3
NO3 and N2O5, which is measured inside the heated channel degrades to as much as 3 pptv. On the heated channel for
共denoted SUM兲, the NO3 mixing ratio derived from the am- detection of the sum of NO3 and N2O5, the turbulent noise is
bient channel, the transmission efficiency for N2O5 共T1 be- always large at the optimal flow rate for N2O5 conversion.
low兲, and the transmission efficiency for NO3 in the heated Although matching the index of refraction of the purge vol-
inlet 共T2 below兲, ume to that of the heated air sample by addition of He to the
SUM − T2NO3 purge flow helps to suppress this noise, the actual detection
N 2O 5 = . 共6兲 sensitivity for the sum of NO3 and N2O5 is a variable
T1
0.5– 2 pptv, with a typical value of approximately 1 pptv.
The estimated fractional uncertainty in the quantity of SUM Figure 7 shows the Allan variance plots39 for 800 s seg-
is ␦共SUM兲 = 15%, taken as the linear sum of a 10% uncer- ments of NO3 data from the laboratory and from the P-3

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-9 Aircraft measurement of NO3 and N2O5 Rev. Sci. Instrum. 77, 034101 共2006兲

FIG. 7. Allan variance plots of the instrument baseline in the laboratory


共plot A兲 and from the aircraft during a daylight flight on July 27, 2004 共plot
FIG. 8. Results from the test flight of the CaRDS instrument on the NOAA
B兲. The laboratory data have a sensitivity of 0.13 pptv 共1␴兲 at 1 s but
P-3 in March, 2004. Upper left: flight track. Lower: time series of NO3,
average to a detection limit below 0.03 pptv at 30– 40 s. The aircraft data
N2O5, and altitude. Upper right: variation of the instrument detection sensi-
are considerably noisier, with a 1 s sensitivity of 0.3 pptv and a detection
tivity 共lower axis兲 and ring-down time constant 共top axis兲 with altitude. The
limit at 10 s of averaging of 0.15 pptv.
increase in sensitivity due to the increasing ring-down time constant and
reduced turbulent flow noise approximately cancels the decrease in ambient
aircraft during a daytime flight 共July 27, 2004兲. The labora- number density, leading to a nearly altitude independent sensitivity.
tory data show a detection sensitivity of 0.13 pptv 共one stan-
dard deviation, 1␴, from repeated 1 s measurements兲 at 1 s, 2004. The results illustrate several interesting features of the
but average to an optimal detection limit of 0.03 pptv for a instrument performance. Figure 8 shows a flight track for the
30 s average. The data shown in the figure were recorded at test flight, a time series of the data and the altitude, and a plot
a total flow of 1 – 2 SLPM. Similar detection sensitivities of the detection sensitivity versus altitude. The flight took off
have been observed under laboratory conditions at flow rates just after the local sunset from MacDill Air Force Base in
up to 16 LPM, although flow-dependent noise often degrades Tampa, FL, where the NOAA Aircraft Operation Center is
the detection sensitivity by a factor of two at this flow rate. located, and ran several level legs at increasing altitudes to
The data from the aircraft are consistently noisier than the approximately 7 km over the eastern Gulf of Mexico, return-
best case laboratory data and are also less amenable to signal ing at low altitude 共⬃1 km兲 over populated areas in west
averaging, with a 1 s detection sensitivity for this trace of central Florida. There was very little NO3 or N2O5 at the
0.3 pptv and an optimal detection sensitivity of 0.15 pptv for higher altitude legs on this flight, although there was clear
a 10 s average. The Allan plots illustrate the degree to which signal on the lower altitude legs over the Florida peninsula,
averaging increases the instrument sensitivity under low sig- where the aircraft presumably sampled anthropogenic NOx
nal conditions. However, the observed concentrations of sources. To our knowledge, this is the first example of a
NO3 and N2O5 in the atmosphere, particularly from the air- measurement of NO3 or N2O5 from an aircraft platform. The
craft, are so highly variable that signal averaging can result lack of signal on the higher altitude legs allowed for an as-
in a significant loss of information. Therefore, the 1 s sensi- sessment of the instrument sensitivity as a function of alti-
tivities are the most appropriate measure of instrument tude from the statistics on the baseline, as shown in the upper
performance. right hand graph of Fig. 8. Because the instrument samples at
a modest 共⬍10 mbars兲 pressure drop from the ambient, ex-
ternal pressure, the background ring-down time constant, ␶0,
III. RESULTS AND SAMPLE DATA
increases monotonically with altitude. 共Because of slightly
The NO3 and N2O5 CaRDS instrument was deployed on dirty mirrors on the test flight, the ␶0 value at 1 atm was
the NOAA P-3 aircraft on March 30, 2004 for a flight to test ⬃160 ␮s, rather than the more typical 180 ␮s.兲 Also, the
instrument performance in advance of the planned deploy- turbulent flow noise described above tends to decrease with
ment during the NEAQS field campaign in the summer of increasing altitude, presumably because the Reynolds num-

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-10 Dubé et al. Rev. Sci. Instrum. 77, 034101 共2006兲

ber in the inlet decreases at constant volumetric flow. Both of ACKNOWLEDGMENTS


these effects tend to increase the sensitivity of the instrument
with respect to the absolute number density of absorbers. This work was supported by the NOAA Climate Goal’s
However, the decrease in the ambient number density with Climate Forcing Program and NOAA’s Air Quality Program.
altitude approximately cancels these increases, and the sen- One of the authors 共M. N.兲 thanks the NOAA Educational
sitivity for NO3 in mixing ratio units is approximately con- Partnership Program with Minority Serving Institutions for
stant with altitude over a range of 0 – 7 km. This invariance an internship.
with altitude suggests that the instrument would work 1
U. Platt and F. Heintz, Isr. J. Chem. 34, 289 共1994兲.
equally well with an inlet controlled at a constant subambient 2
R. P. Wayne et al., Atmos. Environ., Part A 25A, 1 共1991兲.
temperature down to roughly 0.5 atm. It further suggests that 3
U. F. Platt, A. M. Winer, H. W. Bierman, R. Atkinson, and J. N. Pitts, Jr.,
with higher reflectivity mirrors, where the major limitation to Environ. Sci. Technol. 18, 365 共1984兲; S. S. Brown et al., Geophys. Res.
the time constant is due to Rayleigh scattering rather than Lett. 31 共2004兲.
4
A. M. Winer, R. Atkinson, and J. N. Pitts, Jr., Science 224, 156 共1984兲; U.
mirror reflectivity, a pressure controlled inlet 共or high alti- Platt, G. LeBras, G. Poulet, J. P. Burrows, and G. Moortgat, Nature 共Lon-
tude sampling兲 could yield a substantial increase in detection don兲 348, 147 共1990兲; S. A. Penkett, J. N. Blake, P. Lightman, A. R. W.
sensitivity. Marsh, P. Anwyl, and G. Butcher, J. Geophys. Res. 98, 2865 共1993兲.
5
J. F. Noxon, R. B. Norton, and W. R. Henderson, Geophys. Res. Lett. 5,
675 共1978兲; U. Platt, D. Perner, A. M. Winer, G. W. Harris, and J. N. Pitts,
Jr., ibid. 7, 89 共1980兲; J. F. Noxon, R. B. Norton, and E. Marovich, ibid.
IV. SUMMARY AND FUTURE WORK 7, 125 共1980兲.
6
S. Solomon, H. L. Miller, J. P. Smith, R. W. Sanders, G. H. Mount, A. L.
This article has described a pulsed, visible cavity ring-
Schmeltekopf, and J. F. Noxon, J. Geophys. Res. 94, 11041 共1989兲; S. R.
down spectrometer for in situ detection of NO3 and N2O5 Aliwell and R. L. Jones, Geophys. Res. Lett. 23, 2585 共1996兲.
7
from an aircraft. The instrument detects NO3 in an ambient J. M. C. Plane and N. Smith, in Spectroscopy in Environmental Science,
channel on its strong visible absorption band at 662 nm, and edited by R. J. H. Clark and R. E. Hester 共Wiley, Chichester, 1995兲, p.
223; U. Platt, in Air Monitoring by Spectroscopic Techniques, edited by
the sum of NO3 and N2O5 in a separate, heated channel via M. W. Sigrist 共Wiley, New York, 1994兲, p. 27.
the thermal conversion of N2O5 to NO3. It is small enough 8
B. J. Allan, G. McFiggans, J. M. C. Plane, H. Coe, and G. G. McFadyen,
for deployment on the NOAA WP-3D aircraft and has per- J. Geophys. Res. 105, 24 共2000兲.
9
formed reliably on that platform for one test flight in March A. Geyer, R. Ackermann, R. Dubois, B. Lohrmann, R. Müller, and U.
Platt, Atmos. Environ. 35, 3619 共2001兲; F. Heintz et al., J. Geophys. Res.
of 2004 and for the duration of a six week field deployment 101, 22891 共1996兲.
in July-August 2004. At a time resolution of 1 s, the best 10
R. Atkinson, A. M. Winer, and J. N. Pitts, Jr., Atmos. Environ. 20, 331
case detection limits 共1␴兲 for NO3 and N2O5 are approxi- 11
共1986兲.
mately 0.1 pptv, although actual detection limits encountered D. Mihelcic, D. Klemp Müsgen, H. W. Pätz, and A. Volz-Thomas, J.
Atmos. Chem. 16, 313 共1993兲; A. Geyer et al., J. Geophys. Res. 104, 26
in the field were 0.2– 0.5 pptv for NO3 and 0.5– 2 pptv for 共1999兲.
N2O5. Although cavity ring-down spectroscopy is an abso- 12
S. S. Brown, H. Stark, S. J. Ciciora, and A. R. Ravishankara, Geophys.
lute, direct absorption method, the inlet transmission for NO3 Res. Lett. 28, 3227 共2001兲.
13
and the conversion efficiency of N2O5 to NO3 are less than S. S. Brown, H. Stark, S. J. Ciciora, R. J. McLaughlin, and A. R. Ravis-
hankara, Rev. Sci. Instrum. 73, 3291 共2002兲.
unity and add considerable uncertainty to the determination 14
M. D. King, E. M. Dick, and W. R. Simpson, Atmos. Environ. 34, 685
of their ambient concentrations. The estimated accuracy of 共2000兲; W. R. Simpson, Rev. Sci. Instrum. 74, 3442 共2003兲; J. D. Ayers,
the NO3 measurement is 25%, while the estimated N2O5 L. Apodaca, W. R. Simpson, and D. S. Baer, Appl. Phys. B 共in press兲.
15
S. M. Ball, I. M. Povey, E. G. Norton, and R. L. Jones, Chem. Phys. Lett.
accuracy depends on the ratio of N2O5 to NO3 but is gener-
342, 113 共2001兲.
ally in the range of 20%–40%. 16
S. M. Ball, J. M. Langridge, and R. L. Jones, Chem. Phys. Lett. 398, 68
There are a number of improvements to the instrument 共2004兲; M. Bitter et al., Atmos. Chem. Phys. Discuss. 5, 3491 共2005兲.
17
currently underway that should improve its sensitivity, accu- E. C. Wood, P. J. Wooldridge, J. H. Freese, T. Albrecht, and R. C. Cohen,
Environ. Sci. Technol. 37, 5732 共2003兲.
racy, and versatility. The first is the incorporation of an ad- 18
E. C. Wood, T. H. Bertram, P. J. Wooldridge, and R. C. Cohen, Atmos.
ditional 532 nm CaRDS measurement for NO2 using a frac- Chem. Phys. 5, 483 共2005兲.
19
tion of the 532 nm light used to pump the dye laser, as J. Matsumoto, N. Ksougi, H. Imai, and Y. Kajii, Rev. Sci. Instrum. 76,
has been recently demonstrated for our ground-based 共2005兲.
20
D. L. Slusher, L. G. Huey, D. J. Tanner, F. Flocke, and J. M. Roberts, J.
instrument.26 Detection of NO2 not only provides measure- Geophys. Res. 109, 共2004兲.
ment of an additional, related nitrogen oxide, but also pro- 21
A. O’Keefe, and D. A. G. Deacon, Rev. Sci. Instrum. 59, 2544 共1988兲; K.
vides a mean to calibrate the inlet transmission directly in the W. Busch and M. A. Busch, ACS Symposium Series 共American Chemical
field from the conversion of NO3 to NO2 via reaction with Society, Washington, DC, 1999兲.
22
S. M. Ball and R. L. Jones, Chem. Rev. 共Washington, D.C.兲 103, 5239
NO.17 Further work to better quantify the absorption cross 共2003兲; G. Berden, R. Peeters, and G. Meijer, Int. Rev. Phys. Chem. 19,
section of NO3 and its temperature dependence, currently 565 共2000兲; S. S. Brown, Chem. Rev. 共Washington, D.C.兲 103, 5219
underway in our laboratory, will also improve the instrument 共2003兲.
23
accuracy. We are also working on reduction of the flow- P. Zalicki and R. N. Zare, J. Chem. Phys. 102, 2708 共1995兲.
24
S. M. Newman, I. C. Lane, A. J. Orr-Ewing, D. A. Newnham, and J.
related noise, especially on the heated channel, to bring the Ballard, J. Chem. Phys. 110, 10749 共1999兲.
25
field detection sensitivity closer to that achievable in the A. Pettersson, E. R. Lovejoy, C. A. Brock, S. S. Brown, and A. R. Rav-
laboratory. Future planned deployments include experiments ishankara, J. Aerosol Sci. 35, 995 共2004兲.
26
H. D. Osthoff, S. S. Brown, T. B. Ryerson, T. J. Fortin, B. M. Lerner, E.
at a tall tower near our laboratories in Boulder, CO for ver-
J. Williams, A. Pettersson, T. Baynard et al., J. Geophys. Res. 共in press兲.
tical profiling of NO3 and N2O5 and further air quality stud- 27
R. J. Yokelson, J. B. Burkholder, R. W. Fox, R. K. Talukdar, and A. R.
ies with the NOAA P-3. Ravishankara, J. Phys. Chem. 98, 13144 共1994兲.

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp
034101-11 Aircraft measurement of NO3 and N2O5 Rev. Sci. Instrum. 77, 034101 共2006兲

28
R. Penndorf, J. Opt. Soc. Am. 47, 176 共1957兲. D. L. Albritton, and E. E. Ferguson, J. Chem. Phys. 68, 2085 共1978兲.
29
M. Aldener et al., J. Geophys. Res. 共submitted兲. 35
D. M. Murphy and D. W. Fahey, Anal. Chem. 59, 2753 共1987兲; Y. Ru-
30
S. S. Brown, H. Stark, T. B. Ryerson, E. J. Williams, D. K. Nicks, Jr., M. dich, R. K. Talukdar, T. Imamura, R. W. Fox, and A. R. Ravishankara,
Trainer, F. C. Fehsenfeld, and A. R. Ravishankara, J. Geophys. Res. 108, Chem. Phys. Lett. 261, 467 共1996兲.
共2003兲. 36
J. B. Burkholder and R. K. Talukdar, Geophys. Res. Lett. 21, 581 共1994兲.
31 37
S. P. Sander et al., Chemical Kinetics and Photochemical Data for Use in J. Orphal, C. E. Fellows, and P. M. Flaud, Sci. News 共Washington, D. C.兲
Atmospheric Studies 共JPL, Pasadena, CA, 2003兲. 108, 共2003兲.
32
A. C. Vandaele et al., J. Quant. Spectrosc. Radiat. Transf. 59, 171 共1998兲. 38
S. S. Brown, H. Stark, and A. R. Ravishankara, Appl. Phys. B 75, 173
33
S. S. Brown et al., J. Photochem. Photobiol., A 共in press兲. 共2002兲.
34 39
J. A. Davidson, A. A. Viggiano, C. J. Howard, I. Dotan, F. C. Fehsenfeld, P. Werle, R. Mücke, and F. Slemr, Appl. Phys. B 57, 131 共1993兲.

Downloaded 27 Apr 2006 to 140.172.241.113. Redistribution subject to AIP license or copyright, see http://rsi.aip.org/rsi/copyright.jsp

You might also like