• No results found

Expected performance of the superconducting submillimeter-wave limb emission sounder compared with aircraft data

N/A
N/A
Protected

Academic year: 2021

Share "Expected performance of the superconducting submillimeter-wave limb emission sounder compared with aircraft data"

Copied!
13
0
0

Loading.... (view fulltext now)

Full text

(1)

Expected performance of the Superconducting Submillimeter-Wave Limb Emission Sounder compared with aircraft data

S. A. Buehler, 1 C. L. Verdes, 1 S. Tsujimaru, 2 A. Kleinbo¨hl, 1 H. Bremer, 1 M. Sinnhuber, 1 and P. Eriksson 3

Received 5 May 2004; revised 26 January 2005; accepted 18 March 2005; published 25 June 2005.

[ 1 ] The simulated retrieval performance of a submillimeter wave limb sounder was compared with that of an up-looking instrument with identical observation frequency bands and comparable noise temperature. The frequency bands were 624.32 – 626.32 and 649.12 – 650.32 GHz, and the retrieval simulations focused on the key trace gas species O 3 , HCl, and ClO. As expected, the limb geometry leads to a better altitude resolution and larger measurement altitude range. The same retrieval setup was applied to measured spectra, taken by the up-looking Airborne Submillimeter Radiometer (ASUR) instrument on 4 September 2002 at 19.11E, 71.90N and on 19 September 2002 at 44.10E, 4.10S. The observed structures in the fit residual near the HCl spectral lines at 625.9 GHz lead to the conclusion that the pressure shift parameter of HCl is likely to be higher than the value in the HITRAN spectroscopic database. Depending on the assumed

temperature dependence of the shift, the HCl pressure shift value consistent with the ASUR data is 0.090 – 0.117 MHz/hPa instead of the 0.030 MHz/hPa reported in HITRAN.

This result is in good agreement with very recent independent laboratory work which suggests a value of 0.110 MHz/hPa for the shift.

Citation: Buehler, S. A., C. L. Verdes, S. Tsujimaru, A. Kleinbo¨hl, H. Bremer, M. Sinnhuber, and P. Eriksson (2005), Expected performance of the Superconducting Submillimeter-Wave Limb Emission Sounder compared with aircraft data, Radio Sci., 40, RS3016, doi:10.1029/2004RS003089.

1. Introduction

[ 2 ] Satellite-based millimeter wave limb sounding is a well established technique for the observation of atmo- spheric trace gases such as ozone, water vapor, chlorine compounds, and many others. Two successful instru- ments of this type were the Microwave Limb Sounder (MLS) [Barath et al., 1993] and the Millimeter-wave Atmospheric Sounder (MAS) [Hartmann et al., 1996].

Recently, instruments have moved toward higher fre- quencies into the submillimeter wave spectral range, an example of this type of instrument is Odin-SMR [Murtagh et al., 2002].

[ 3 ] The Superconducting Submillimeter-Wave Limb Emission Sounder (SMILES) [Masuko et al., 2002] is an instrument concept designed to measure with excep- tionally low noise stratospheric trace gases that have only weak spectroscopic signatures. Its unique feature is a superconductor-insulator-superconductor (SIS) mixer, a device that has so far not been operated in space since it works only if cooled down to 4.5 K. The first SMILES- type instrument will be JEM/SMILES. It is planned to operate on the exposed facility of the Japanese Experi- ment Module (JEM) of the International Space Station (ISS) from the year 2008. Its bands have been selected to allow the measurement of various trace gases that are important for the understanding of stratospheric ozone chemistry, notably ozone itself including its isotopes, and several chlorine compounds. This paper focuses on the species O 3 , HCl, and ClO, since they are regarded as standard products of JEM/SMILES, and they can be easily detected by an airborne instrument.

[ 4 ] The Airborne Submillimeter Radiometer (ASUR) is an aircraftborne submillimeter sensor used for study- ing stratospheric trace gases. Since its development in

1

Institute of Environmental Physics, University of Bremen, Bremen, Germany.

2

National Institute of Information and Communications Technology, Tokyo, Japan.

3

Department of Radio and Space Science, Chalmers University of Technology, Gothenburg, Sweden.

Copyright 2005 by the American Geophysical Union.

0048-6604/05/2004RS003089$11.00

RS3016 1 of 13

(2)

1991 by the University of Bremen, ASUR has contrib- uted to all major European Arctic ozone campaigns (EASOE, SESAME I+III, THESEO, THESEO-2000/

SOLVE, EUPLEX) as well as to several satellite valida- tion experiments (MLS, GOME, ILAS, SAGE III, SCIAMACHY). In 2002 ASUR participated in the LEONID-MAC campaign on board the NASA DC-8 to study changes of mesospheric composition during the passage of the Earth through comet dust trails. For detailed information see Crewell et al. [1994], de Valk et al. [1997], Sugita et al. [2002], Fahey et al. [2001], Bremer et al. [2002], von Ko¨nig et al. [2002], and Kleinbo¨hl et al. [2003]. The discussed species O 3 , HCl, and ClO are standard products of ASUR.

[ 5 ] The ASUR instrument looks upward, under a constant zenith angle of 78 and can be tuned over a wide frequency range from 604 GHz to 662 GHz. For the present study it was tuned to observe the planned JEM/

SMILES bands. Some technical details of both instru- ments are summarized in Table 1. Like the planned SMILES, ASUR also features a state of the art SIS mixer. The cooling for the ASUR mixer is achieved by a cryostat with liquid helium, whereas for the space instrument SMILES it is achieved by a three-stage mechanical cooler, using helium in a closed cycle.

[ 6 ] The aim of this study is to confirm that the spectroscopic properties of the atmosphere in the desig- nated SMILES bands are correctly represented by state of the art radiative transfer models, and to confirm the predicted SMILES performance by comparing it to the observed ASUR performance. The aim is not to present an exhaustive discussion of JEM/SMILES retrieval ca- pabilities, for such a discussion see Masuko et al. [2002].

[ 7 ] In order to do this investigation one needs a forward model, in this case a radiative transfer model,

and a retrieval model. The forward model used is the Atmospheric Radiative Transfer Simulator ARTS [Buehler et al., 2005]. Spectroscopic data are taken from three different sources: The current versions of HITRAN [Rothman et al., 2003], JPL [Pickett et al., 1998], and MYTRAN [Perrin et al., 2005; Verdes et al., 2005;

Demaison et al., 2004]. The latter database has been generated within a study funded by the European Space Agency (ESA), focusing on the planned MASTER instrument. Figure 1 shows simulated spectra for both instruments, the limb-looking satellite sensor and the up- looking aircraft sensor. The inversion software used is Qpack [Eriksson et al., 2005]. It uses the Optimal Estimation Method (OEM), also often called ‘1-D var,’

which has been described for example by Rodgers [1976, 2000]. The nonlinear version of OEM, with Levenberg- Marquardt iteration, was used.

[ 8 ] This paper is organized as follows: Section 2 presents some very brief retrieval theory, mainly intended to introduce the necessary terminology. Sec- tions 3.1 and 3.2 present the simulated SMILES and ASUR performances, respectively. Section 4 presents and discusses the observed ASUR performance and the consistency of the observed spectroscopy in the SMILES bands with model predictions. Finally, section 5 presents summary and conclusions.

2. Retrieval Theory

[ 9 ] A retrieval method must be used to extract the desired species concentration profiles from measured or simulated submillimeter wave spectra. For this study the optimal estimation method was chosen, which is de- scribed in detail in the well known book of Rodgers Table 1. Some Key Instrumental Parameters of the Planned Satellite Instrument JEM/SMILES Compared With the Existing Aircraft Instrument ASUR a

Parameters SMILES ASUR

Observation method limb looking up looking

Platform International Space Station aircraft

Platform altitude, km 400 (nominal) 10 – 13

Frequency range, GHz

Band A 624.32 – 625.52 604 – 662 (tunable)

Band B 625.12 – 626.32

Band C 649.12 – 650.32

Receiver type SIS mixer SIS mixer

Sideband treatment single sideband (SSB) SSB

System noise temperature, K <700 450 – 1000

Spectrometer Acousto Optical Spectrometer (AOS) AOS

Spectral resolution (FWHM), MHz 1.4 1.26

Effective antenna beam width, deg 0.096 (HPBW, elevation) 1.3

Observation range 10 – 60 km in tangent height 78 in zenith angle

Pointing uncertainty 0.34 km, random 0.2

a The acronyms FWHM and HPBW mean full width at half maximum and half power beam width, respectively. The ASUR system noise

temperature value comes from the calibration procedure and depends on the observation frequency.

(3)

[2000]. The purpose of this section is not to derive or even describe the optimal estimation method, but just to define the terms used in the subsequent discussion.

[ 10 ] In the retrieval context one assumes that the atmospheric state x, in our case the species concen- trations, and the measurement y, in our case the spectra, are linked by a function called forward model according to

y ¼ F x; b ð Þ þ ; ð1Þ where b represents input parameters to F that are not retrieved but assumed to be well known, and  represents measurement errors. Which of the input parameters to F belong to x and which to b is a matter of choice.

[ 11 ] In order to perform a retrieval and a basic error analysis, one needs the Jacobian of F with respect to the state vector x

K x ¼ @F x; b ð Þ

@x ð2Þ

and the Jacobian of the inverse (retrieval) model I with respect to the measurement y

D ¼ @I y ð Þ

@y : ð3Þ

Another important quantity in the context of retrieval is the averaging kernel matrix A which basically describes how the retrieved state x ret is smoothed relative to the true state x true :

x ret  x ref ¼ A x ð true  x ref Þ; ð4Þ

where x ref is some reference state. The rows of A are called averaging kernels. For an ideal retrieval they should be delta functions with value unity where retrieval altitude matches true altitude and value zero everywhere else. For a real retrieval they will have a peak value less than unity and a finite width. The averaging kernel matrix can be used to define the measurement response R s (i) of species s as the row sum of the averaging kernel matrix, where the sum goes only over those columns that belong to species s:

R s ð Þ ¼ i X j2

j¼j1

A ij : ð5Þ

Here j1 and j2 are start and stop indices of species s in the state vector. The measurement response is useful, because it can be used to define the ‘good’ retrieval altitude range as the region where the response is close to unity.

[ 12 ] Other important retrieval quantities are various covariance matrices, in particular the measurement error covariance matrix S y , the a priori covariance matrix S a , the retrieval measurement error covariance matrix

S m ¼ DS y D T ; ð6Þ

the retrieval smoothing error covariance matrix

S s ¼ A  I ð ÞS a ð A  I Þ T ; ð7Þ and the total retrieval error covariance matrix

S ¼ S m þ S s : ð8Þ Figure 1. Simulated submillimeter wave brightness temperatures as a function of frequency: limb

sensor for a tangent altitude of 25 km, antenna beam width 0.096 (half power beam width elevation) (solid curve) and up-looking sensor, flight altitude 10 km, zenith-looking angle 78

(dashed curve). The spectral lines belonging to the target species are labeled in the plots. See color

version of this figure in the HTML.

(4)

In relations (6) and (7) I is the identity matrix, and the superscript T indicates the matrix transpose.

[ 13 ] The terms used to assess the retrieval performance are retrieval error, defined as the square root of the diagonal elements of S, retrieval measurement error, defined as the square root of the diagonal elements of S m , and smoothing error, defined as the square root of the diagonal elements of S s . More details on retrieval theory and explanations on these quantities can be found in the work by Rodgers [2000]. A discussion of the application of the optimal estimation method specifically for the case of a submillimeter wave limb sounder can be found in work by Verdes [2002] or Buehler [1999]. Note that S y and S a are input quantities for the retrieval, whereas A, S, S m , and S s are output quantities.

3. Retrieval Simulations

3.1. Simulated SMILES Performance

[ 14 ] To make a realistic estimate of the limb sounder performance, several quantities have to be retrieved simultaneously from a simulated limb scan cycle. The retrieval setup is summarized in Table 2, and Figure 2 shows the assumed atmospheric state, that is, profiles of temperature, O 3 , HCl, and ClO. Other trace gases with weaker spectroscopic signatures, HNO 3 , HO 2 , H 2 O 2 , and BrO were also included in the simulation, but are not shown.

[ 15 ] Retrieved was the volume mixing ratio (VMR) in relative units, that is, x = VMR true /VMR reference , and the temperature in absolute units (Kelvin), as described in detail by Verdes et al. [2002]. The reference VMR profiles are the ones shown in Figure 2. These

retrievals were using different bands for different key species: band B for O 3 and HCl, band C for ClO. The retrievals included altitude profiles of two absorption offsets, which are necessary to remove uncertainties due to absorption continua [Kuhn, 2004]. The absorp- tion offset was assumed to vary linearly across the band between these two values, which are at the band edges.

[ 16 ] The measurement error covariance matrix S y was assumed to be diagonal, with the square root of the Table 2. Retrieval Setup Summary for the SMILES and ASUR Simulation a

Retrieval Condition SMILES ASUR

Platform altitude, km 400 10

Tangent height or zenith angle 10 – 100 km (2 km step) 78

Spectroscopic database MYTRAN MYTRAN

Retrieval altitude range, km 0 – 90 0 – 90

Retrieval grid spacing 2 km for ClO and 3 km for T 2 km for ClO and 3 km for T

A priori profile subarctic winter subarctic winter

A priori error covariance

b

50% error 50% error

T retrieval 5 K a priori error no

Continuum retrieval 2 absorption offset profiles 2 absorption offset profiles

Pointing offset retrieval 0.2 a priori error no

Antenna integration yes no

Frequency grid spacing, MHz 2.7 2.7

Channel response convolution 2.7 MHz boxcar 2.7 MHz boxcar

System temperature, K 500 500

Integration time 0.5 s per spectrum, 45 s per scan 60 s per spectrum

Measurement error, K 0.86 0.08

a The retrieval grid spacing was chosen to be identical to the tangent altitude spacing of the limb instrument, except for ClO and temperature retrieval.

b The a priori error covariance matrix had an exponentially decaying correlation with 5 km correlation length.

Figure 2. Assumed subarctic winter atmospheric state:

temperature (solid curve), O 3 (short-dashed curve), HCl

(long-dashed curve), and ClO (dash-dotted curve).

(5)

diagonal elements set to 0.86 K, which was calculated with the radiometer noise formula

DT ¼ aT sys

ffiffiffiffiffi

p Bt ð9Þ

from the assumed system noise temperature T sys of 500 K, the assumed channel bandwidth B of 2.7 MHz, and the assumed integration time t of 0.5 s. The receiver-type dependent factor a was assumed to be 2. This noise figure can be regarded as typical for a single SMILES spectrum.

No off-diagonal elements of S y were introduced in the simulation, which means that it was assumed that the measurement noise is uncorrelated between the different channels. This assumption is not generally true for acousto-optical spectrometers, since they normally do some frequency oversampling. To account for this, the channel bandwidth was taken to be two times the nominal bandwidth, assuming binning of pairs of neighboring channels. For these two-channel bins the noise is uncorrelated. Further assumed was a nondiagonal a priori covariance matrix S a with the square root of the diagonal elements corresponding to an error of 50% and an exponentially decaying correlation between the retrieval layers, with a correlation length of 5 km.

[ 17 ] Figure 3 summarizes the result of the retrieval simulation for the three key species O 3 , HCl, and ClO.

The left plots of Figure 3 show retrieval error, retrieval measurement error, and smoothing error. The middle plots show averaging kernels and measurement response, and the right plots show the averaging kernel full width at half maximum. One can see from Figure 3 that O 3

retrieval works well at altitudes roughly between 15 and 50 km. The retrieval measurement error is mostly around 5% for the range where the altitude resolution evaluated from the full width at half maximum of the averaging kernels is below approximately 2.5 km. In the same altitude range the smoothing error is around 3%. HCl retrieval also works well roughly between 15 and 50 km.

The retrieval measurement error is around 5%, while the smoothing error is larger, varying between 5% and 15%.

ClO retrieval works well between 25 and 45 km, but with somewhat higher errors. The retrieval measurement error is between 8% and 15% in the quoted altitude range, while the smoothing error is generally higher. Only in the altitude range of 30 – 40 km the retrieval measurement error dominates the smoothing error. In the same altitude range the full width at half maximum of the averaging kernels is about 5 km. Note that ClO retrieval perfor- mance will be significantly better at low altitudes for an enhanced ClO situation inside the polar vortex.

3.2. Simulated ASUR Performance

[ 18 ] The performance of the up-looking ASUR instru- ment for the same spectral bands was simulated with the

same setup (see Table 2). Only the geometry and the assumed measurement noise were different (see Table 1).

The viewing angle was 78, and the measurement noise was set to be 0.08 K, corresponding to an integration time of 60 s. The antenna response is not an issue for up- looking geometry, hence a simple pencil beam was assumed. The continuum retrieval was formally treated in the same way as for the limb sensor. However, for the continuum parameters there is no altitude information in the measurement, so that the retrieval effectively just scales the total absorption of the troposphere. This is reflected in the averaging kernels for the absorption offsets (not shown), which are very wide and flat for ASUR. Temperature retrieval was not considered as little information can be obtained for up-looking geometry.

Figure 4 shows the simulated ASUR retrieval perfor- mance for O 3 , HCl, and ClO.

[ 19 ] A good measurement response for O 3 is achieved at altitudes between 15 km and 45 km. The retrieval measurement error at these altitudes is about 5% or slightly larger, while the smoothing error is much larger, reaching a value of 20% at the edges of the altitude range.

[ 20 ] The retrieval performance of HCl is slightly worse. The best performance is achieved between 15 and 40 km. The retrieval measurement error is about 10%. However, the error due to the smoothing effects is larger (25%, in the altitude range where good measure- ment response is achieved). The averaging kernels are between 5 and 10 km wide at altitudes where a good retrieval performance is achieved (15 – 40 km). The ClO profile can be retrieved with an acceptable performance only at altitudes between 30 and 40 km. The retrieval measurement error is about 10% while the smoothing error is larger than 25%. The vertical resolution in the same altitude range, given by the width of the averaging kernels, is about 10 km. Note that, as in the case of SMILES, ClO retrieval performance is significantly better at low altitudes if ClO is enhanced. From other ASUR campaigns ClO profiles have been retrieved down to approximately 18 km.

[ 21 ] The main difference between the SMILES and ASUR instrument is that ASUR has a coarser vertical resolution due to the up-looking geometry. This leads to significant smoothing effects, as can be seen from the smoothing error curves in Figure 4.

4. ASUR Data

[ 22 ] The data used for this study were taken during the

SCIAMACHY validation campaign from 4 to 26 Sep-

tember 2002. For comparison with the SMILES satellite

sensor, dedicated observations were made in the

SMILES spectral bands. We discuss two measurements,

one for band B recorded on 4 September 2002 over the

(6)

arctic region (longitude 19.11, latitude 71.90), and one for band C on 19 September 2002 over the tropical region (longitude 44.10, latitude 4.10).

[ 23 ] Individual spectra were integrated to reach a total integration time of 102 s for band B and 490 s for band C. A species retrieval was carried out with these data, using the same setup as for the model simulations. The temperature data were taken from the reanalysis of the

National Centers for Environmental Prediction (NCEP), which was obtained from the Distributed Active Archive Center (DAAC) at NASA Goddard Space Flight Center (http://daac.gsfc.nasa.gov/).

[ 24 ] Figure 5 shows a measured spectrum for band B (top plot) as well as the fit residual (bottom plot). This can be used to estimate the instrument noise. To do this, several small frequency regions away from spectral lines Figure 3. Simulated JEM/SMILES retrieval performance for a measurement noise of 0.86 K. The

left plots show retrieval error (solid curve), retrieval measurement error (dashed curve), and

smoothing error (dash-dotted curve) for (top) O 3 , (middle) HCl, and (bottom) ClO. The middle

plots show averaging kernels for (top) O 3 , (middle) HCl, and (bottom) ClO. For the sake of clarity,

only kernels for every third retrieval altitude are plotted. The heavy line shows the measurement

response. The right plots show full width at half maximum (FWHM) of averaging kernels

presented in the middle plots. See color version of this figure in the HTML.

(7)

were identified, and the mean DT b of the residual and its standard deviation s DT

b

were calculated. The s DT

b

is the instrument noise. The results are summarized in Table 3. They show that the noise is significantly higher than the noise estimated from the Y factor measurement that is part of the calibration procedure.

This is believed to be due to the very long integration times used, which emphasize noise sources other than pure radiometric noise. One can also observe that the noise increases toward the band edges, which is believed to be due to the poorer performance of the mixer toward the band edges and the limited pass- bands of the intermediate frequency chain. Despite

these observations, the ASUR measurement is still exceptionally low noise for measurements at these frequencies, and could not be achieved with a conven- tional mixer. Note that it was the actual noise level observed in the fit residual that was used to derive the measurement error covariance matrix for the retrieval, not the lower Y factor noise level.

[ 25 ] A further striking feature of Figure 5 is that the residual shows some structures near the spectral lines, indicating a poor fit. Two possible causes of these structures were investigated, (1) a frequency offset in the measurement, and (2) incorrect spectroscopic data for individual spectral lines.

Figure 4. Same as for Figure 3 but for simulated ASUR retrieval performance for a measurement

noise of 0.08 K. See color version of this figure in the HTML.

(8)

[ 26 ] The frequency calibration of the acousto-optical spectrometer (AOS) was done a few times during the flight by using the output of a comb generator as a reference. During the calibration, the relation between the frequency grid and the AOS channel number was determined by a third-order polynomial fit. However, the data may be distorted by a residual frequency calibration error, due to changes in the cabin temperature between the calibrations. To check this, a second-order polyno- mial frequency offset was added to the list of retrieval parameters, yielding retrieved frequency offset values of up to one AOS channel. As shown by Figure 6, the residual with retrieved frequency offset is improved, but still shows suspicious features around the HCl lines in band B.

[ 27 ] These features cannot be explained by a frequency calibration error of the instrument and must therefore be due to errors in the spectroscopic parameters used.

Table 4 shows spectroscopic parameters for the transi- tions of concern. The center frequencies of the HCl lines are in close agreement in all three databases studied.

Also, test retrievals with slightly modified line center frequencies did not improve the residual. However, a

substantial improvement was achieved by including the retrieval of the pressure shift parameter for the HCl lines, as demonstrated by Figure 7. The spectral signature of an error in the pressure shift parameter is quite different from the signature of a general frequency shift, since the former leads to asymmetric line wings far away from the

Table 3. Mean Brightness Temperature Residuals and Their Standard Deviations for the Integrated ASUR Measurement a

Frequency Range, GHz  T b , K s

DTb

, K T

sys

, K Band B

625.12 – 625.22 0.0207 0.2173 1735

625.42 – 625.82 0.0004 0.1340 1070

626.20 – 626.32 0.1184 0.2912 2325

Band C

649.12 – 649.40 0.0170 0.1453 2545

649.50 – 650.00 0.0258 0.0790 1383

650.20 – 650.32 0.0248 0.1814 3175

a The integration time was 102 and 490 s for band B and band C, respectively. The last column gives the system temperature consistent with the observed noise level and a receiver a factor of 2.

Figure 5. (top) Measured ASUR spectrum in band B for 19.11E, 71.90N on 4 September 2002,

1614 UT. (bottom) Residual (data-model) from a retrieval using the same setup as described in

section 3.2. See color version of this figure in the HTML.

(9)

line center while leaving the center in place, whereas the latter shifts the center frequencies but maintains the symmetry of the line wings.

[ 28 ] The pressure shift d depends on temperature according to

d T ð Þ ¼ d 0

T 0

T

  s

0

; ð10Þ

where d 0 is the pressure shift at the reference temperature T 0 (T 0 = 296 K for HITRAN), T is temperature, and s 0 is the temperature exponent of the shift. Pickett [1980]

gives a relationship between s 0 and the temperature exponent n 0 of the air-broadening parameter

s 0 ¼ 1 4 þ 3

2 n 0 ; ð11Þ

Figure 6. Same as residual plot in Figure 5 but with frequency offset retrieval. See color version of this figure in the HTML.

Table 4. Comparison of Some Spectroscopic Parameters of Major Species in the SMILES Bands a Frequency,

GHz

g

air

,

MHz/hPa n

air

d

0

,

MHz/hPa Reference

O

3

625.3700653 2.09 0.76 - HITRAN

625.3714686(833) 2.09 (2) 0.76 (2) - MYTRAN

625.3714686(833) - - - JPL

H

35

Cl

625.9016573(300000) 2.64 0.69 0.03 HITRAN

625.9185356(300000) 2.64 0.69 0.03 HITRAN

625.9320862(300000) 2.64 0.69 0.03 HITRAN

625.9016584(1000) 2.63 (13) 0.69 (7) - MYTRAN

625.9187272(1000) 2.63 (13) 0.69 (7) - MYTRAN

625.9320081(1000) 2.63 (13) 0.69 (7) - MYTRAN

625.9016030(1000) - - - JPL

625.9187560(1000) - - - JPL

625.9320070(1000) - - - JPL

2.566 (5) 0.73 (3) 0.110 (2) Drouin [2003], s

0

= 0.405

- - - 0.068 (11) von Ko¨nig [2001], s

0

= 1.285

- - - 0.094 (3) H. Ozeki, personal communication (2004)

- - - 0.090 (11) This work, s

0

= 1.285

- - - 0.117 (11) This work, s

0

= 0.405

ClO

649.4512 2.10 (6) 0.85 (9) - Oh and Cohen [1994]

649.4512 2.52 0.50 - HITRAN

a Here g

air

is the air-broadening parameter, n

air

is its temperature dependence coefficient, and d

0

is the pressure shift parameter as defined in (10).

The reference temperature T

0

is 296 K. The numbers in parentheses represent the 1 sigma error for the last digits. Both the line catalogues and our

analysis assume that pressure-broadening parameters for all three HCl lines are the same.

(10)

according to this the s 0 value consistent with our MYTRAN catalogue value of n 0 = 0.69 would be 1.285. However, recent laboratory measurements by Drouin [2003] suggest a quite different value of s 0 = 0.405. Since the pressure shift becomes important below approximately 25 km altitude, where the atmospheric temperature is much colder than the reference tempera- ture, the retrieved value of d 0 depends significantly on the assumed s 0 . For s 0 = 1.285 the retrieved value is d 0 = 0.090 ± 0.011 MHz/hPa, for s 0 = 0.405 the retrieved value is d 0 = 0.117 ± 0.011 MHz/hPa. Thus, if we use the same temperature exponent as Drouin [2003], our result is very close to his laboratory value of 0.110 ± 0.002 MHz/hPa.

[ 29 ] An independent earlier investigation for ASUR by von Ko¨nig [2001] yielded a value of 0.068 ± 0.011 MHz/

hPa retrieved from 44 measured spectra for the same line during the HIMSPEC campaign 1999. That value is somewhat lower than our value with the s 0 = 1.285 temperature exponent, which was also used by von Ko¨nig [2001]. The difference can be explained by the colder temperatures in the lower stratosphere, which were approximately 200 K during the HIMSPEC campaign versus approximately 220 K for the data used in this study.

[ 30 ] Furthermore, a prototype of the SMILES receiver was used to measure the HCl pressure shift at room temperature in the laboratory (H. Ozeki, personal com- munication, 2004). The result of that measurement, d 0 = 0.0939 ± 0.0025 MHz/hPa, also is in better agreement with our value and the value of Drouin than with the HITRAN value. Since this measurement was taken at room temperature it should be rather independent of the assumed temperature exponent.

[ 31 ] Figure 8 shows retrieved O 3 and HCl profiles for the data (19.11E, 71.90N, 4 September 2002,

1614 UT), and ClO profile for the data (44.10E, 4.10S, 19 September 2002, 0515 UT). O 3 and HCl were retrieved simultaneously from band B, while ClO was retrieved from band C. Because of lack of indepen- dent data it was not attempted to validate these profiles;

however, ASUR profiles from other bands have been validated [de Valk et al., 1997; Crewell et al., 1995].

5. Summary and Conclusions

[ 32 ] The simulated retrieval performance of a submil- limeter wave limb sounder was compared with that of an up-looking instrument with identical observation fre- quency bands and comparable noise temperature. The frequency bands were 624.32 – 626.32 GHz and 649.12 – 650.32 GHz, the retrieval simulations focused on the key trace gas species O 3 , HCl, and ClO. As expected, the limb geometry leads to a better altitude resolution and larger measurement altitude range.

[ 33 ] In a second step, the same retrieval setup was applied to measured spectra, taken by the ASUR instrument on 4 September 2002, at 19.11E 71.90N and on 19 September 2002, at 44.10E 4.10S. The observed structures in the fit residual near the HCl spectral lines at 625.9 GHz lead to the conclusion that the pressure shift parameter of HCl is likely to be higher than the HITRAN value and closer to the recent measurements reported by Drouin [2003] and von Ko¨nig [2001]. Depending on the assumed tempe- rature dependence of the shift, the HCl pressure shift value consistent with the ASUR data is 0.090 – 0.117 MHz/hPa, instead of the 0.030 MHz/hPa reported in HITRAN.

[ 34 ] Catalogue values of other spectroscopic parame- ters in the observed frequency region appear to be consistent with the ASUR spectra. However, it should Figure 7. Same as residual plot in Figure 5 but with pressure shift retrieval for the HCl lines. See

color version of this figure in the HTML.

(11)

be noted that not all spectroscopic parameters can be tested with the methodology of this study: Line positions are normally known more precisely than they can be measured with a spectrometer, so that only gross errors in the database, which are sometimes present, would have been detected. Furthermore, biases in line intensities and

pressure-broadening parameters of all observed lines of a given species would not show up in the fit residual.

[ 35 ] Acknowledgments. We thank Y. Kasai and H. Ozeki for discussions and also the SMILES team in general. Further thanks are due to the ARTS radiative transfer community, many Figure 8. Retrieved O 3 and HCl profiles for the measurement 19.11E, 71.90N on 4 September

2002, 1614 UT, and ClO profile for the measurement 44.10E, 4.10S on 19 September 2002,

0515 UT: (top) O 3 , (bottom left) HCl, and (bottom right) ClO. The error bars indicate the retrieval

error (effects of noise plus smoothing), and the thick part of the error bars indicates the retrieval

measurement error (effect of noise only). The HCl pressure shift parameter was retrieved

simultaneously. See color version of this figure in the HTML.

(12)

of whom have indirectly contributed by implementing features to the ARTS model, and to two anonymous reviewers for their astute and constructive comments on the draft manuscript. This study was funded by the German Federal Ministry of Education and Research (BMBF), within the DLR project SMILES, grant 50 EE 9815.

References

Barath, F. T., et al. (1993), The Upper Atmosphere Research Satellite microwave limb sounder instrument, J. Geophys.

Res., 98, 10,751 – 10,762.

Bremer, H., M. von Ko¨nig, A. Kleinbo¨hl, H. Ku¨llmann, K. Ku¨nzi, K. Bramstedt, J. P. Burrows, K.-U. Eichmann, M. Weber, and A. P. H. Goede (2002), Ozone depletion observed by the Airborne Submillimeter Radiometer (ASUR) during the Arctic winter 1999/2000, J. Geophys.

Res., 107(D20), 8277, doi:10.1029/2001JD000546.

Buehler, S. A. (1999), Microwave limb sounding of the strato- sphere and upper troposphere, Ph.D. thesis, 262 pp., Univ.

of Bremen, Bremen, Germany.

Buehler, S. A., P. Eriksson, T. Kuhn, A. von Engeln, and C. Verdes (2005), ARTS, the atmospheric radiative transfer simulator, J. Quant. Spectrosc. Radiat. Transfer, 91, 65 – 93, doi:10.1016/j.jqsrt.2004.05.051.

Crewell, S., K. Ku¨nzi, H. Nett, T. Wehr, and P. Hartogh (1994), Aircraft measurements of ClO and HCl during EASOE 1991/

92, Geophys. Res. Lett., 21, 1267 – 1270.

Crewell, S., R. Fabian, K. Ku¨nzi, H. Nett, W. Read, J. Waters, and T. Wehr (1995), Comparison of ClO measurements by airborne and spaceborne microwave radiometers in the Arc- tic winter stratosphere 1993, Geophys. Res. Lett., 22, 1489 – 1492.

Demaison, J., et al. (2004), Characterization of millimetre- wave spectroscopic signatures, Final Rep. 16377/02/NL/

FF, Eur. Space Res. and Technol. Cent., Noordwijk, Netherlands.

de Valk, J. P. J. M. M., et al. (1997), Airborne heterodyne measurements of stratospheric ClO, HCl, O 3 and N 2 O dur- ing SESAME-I over northern Europe, J. Geophys. Res., 102, 1391 – 1398.

Drouin, B. J. (2003), Temperature dependent pressure-induced lineshape of the HCl J = 1 0 rotational transition in nitrogen and oxygen, J. Quant. Spectrosc. Radiat. Transfer, 83, 321 – 331.

Eriksson, P., C. Jimenez, and S. A. Buehler (2005), Qpack, a general tool for instrument simulation and retrieval work, J. Quant. Spectrosc. Radiat. Transfer, 91, 47 – 64, doi:10.1016/j.jqsrt.2004.05.050.

Fahey, D. W., et al. (2001), The detection of large HNO 3 - containing particles in the winter Arctic stratosphere, Science, 291, 1026 – 1031.

Hartmann, G. K., et al. (1996), Measurements of O 3 , H 2 O, and ClO in the middle atmosphere using the millimeter-wave atmospheric sounder (MAS), Geophys. Res. Lett., 23, 2313 – 2316.

Kleinbo¨hl, A., et al. (2003), Vortexwide denitrification of the Arctic polar stratosphere in winter 1999/2000 determined by remote observations, J. Geophys. Res., 108(D5), 8305, doi:10.1029/2001JD001042.

Kuhn, T. (2004), Atmospheric absorption models for the milli- meter wave range, Ph.D. thesis, Univ. of Bremen, Bremen, Germany.

Masuko, H., et al. (2002), JEM/SMILES mission plan, ver- sion 2.1, technical report, Natl. Space Dev. Agency of Jpn., Tokyo.

Murtagh, D., et al. (2002), An overview of the Odin atmo- spheric mission, Can. J. Phys., 80, 309 – 319.

Oh, J. J., and E. A. Cohen (1994), Pressure broadening of ClO by N 2 and O 2 near 204 and 649 GHz and new frequency measurements between 632 and 725 GHz, J. Quant. Spec- trosc. Radiat. Transfer, 52, 151 – 156.

Perrin, A., C. Puzzarini, J.-M. Colmont, C. Verdes, G. Wlodarczak, G. Cazzoli, S. Buehler, J.-M. Flaud, and J. Demaison (2005), Molecular line parameters for the MASTER (millimeter wave acquisitions for stratosphere/tro- posphere exchange research) database, J. Atmos. Chem., in press.

Pickett, H. M. (1980), Effects of velocity averaging on the shapes of absorption lines, J. Chem. Phys., 73(12), 6090 – 6094.

Pickett, H. M., R. L. Poynter, E. A. Cohen, M. L. Delitsky, J. C.

Pearson, and H. S. P. Mueller (1998), Submillimeter, milli- meter, and microwave spectral line catalog, J. Quant. Spec- trosc. Radiat. Transfer, 60, 883 – 890.

Rodgers, C. (1976), Retrieval of atmospheric temperature and composition from remote measurements of thermal radia- tion, Rev. Geophys., 14, 609 – 624.

Rodgers, C. (2000), Inverse Methods for Atmospheric Sound- ing: Theory and Practice, World Sci., Hackensack, N. J.

Rothman, L. S., et al. (2003), The HITRAN molecular spectro- scopic database: Edition of 2000 including updates through 2001, J. Quant. Spectrosc. Radiat. Transfer, 82, 5 – 44.

Sugita, T., et al. (2002), Validation of ozone measurements from the Improved Limb Atmospheric Spectrometer, J. Geo- phys. Res., 107(D24), 8212, doi:10.1029/2001JD000602.

Verdes, C. (2002), Deriving atmospheric temperature and instru- mental pointing from millimeter/sub-millimeter limb sound- ing measurements, Univ. of Bremen, Bremen, Germany.

Verdes, C., S. A. Buehler, A. von Engeln, T. Kuhn, K. Kuenzi, P. Eriksson, and B. M. Sinnhuber (2002), Pointing and tem- perature retrieval from millimeter/sub-millimeter limb soundings, J. Geophys. Res., 107(D16), 4299, doi:10.1029/

2001JD000777.

Verdes, C. L., S. A. Buehler, A. Perrin, J.-M. Flaud, J. Demaison, G. Wlodarczak, J.-M. Colmont, G. Cazzoli, and C. Puzzarini (2005), A sensitivity study on spectroscopic parameter accuracies for a mm/sub-mm limb sounder instrument, J. Mol. Spectrosc., 229(2), 266 – 275.

von Ko¨nig, M. (2001), Chloraktivierung und PSC-Bildung in

der arktischen Stratosphaere, Ph.D. thesis, Univ. of Bremen,

Bremen, Germany.

(13)

von Ko¨nig, M., et al. (2002), Using gas-phase nitric acid as an indicator of PSC composition, J. Geophys. Res., 107(D20), 8265, doi:10.1029/2001JD001041.



H. Bremer, S. A. Buehler, A. Kleinbo¨hl, M. Sinnhuber, and C. L. Verdes, Institute of Environmental Physics, University of

Bremen, Otto-Hahn-Allee 1, D-28259 Bremen, Germany.

(sbuehler@uni-bremen.de)

P. Eriksson, Department of Radio and Space Science, Chalmers University of Technology, Gothenburg SE-41296, Sweden.

S. Tsujimaru, National Institute of Information and Com-

munications Technology, 4-2-1 Nukui-Kitamachi, Koganei,

Tokyo 184-8795, Japan.

References

Related documents

The trend was the same for all classes; the outer beams had a higher mean backscatter value and lower standard deviation compared with values from midrange and center beams, which

A linear regression analysis was performed with the result that there is a linear relationship between the performance of the Industrial Transportation Companies and the variables

The main motive of this research is to extract maximum electrical power from the WEC by active rectification and smoothing the power fluctuation of the wave energy converter through

The properties of the spin waves are qualitatively affected by the magnetic field landscape surrounding the nanocontact, caused by the vectorial superposition of applied, dipolar

These simulations, performed at fixed bias current I DC = 8 mA, suggests that the reason for the s− or p−like symmetry of the excitation is due to the interplay between the

Ziel des Projekts war die Entwicklung des Instrumentes, zusammen mit dem SMILES Team in Japan, und die Entwicklung und Validation von Algorithmen zur SMILES Datenauswertung..

Comparison of Simulated Radiances from Radiosonde Data and Mi- crowave Sounding Data for the Validation of the Forward Model for Water Vapor Retrieval. Master thesis, Institute

– Same order of magnitude as measurement noise (for integration time 10  atmospheric) – Statistical error, i.e., goes down when data.