ArticlePDF Available

A high precision pulsed quantum cascade laser spectrometer for measurements of stable isotopes of carbon dioxide

Authors:

Abstract and Figures

We describe a prototype instrument using a Peltier cooled quantum cascade laser for precise measurement of stable carbon (13 C/ 12 C) isotopologue ratios in atmospheric CO 2 . Using novel optics and signal processing techniques in a compact instrument, we are able to detect the difference between sample and reference with a precision of 0.1ø (2 standard error of mean of 11 samples) in 10 min of analysis time. The standard deviation of 0.18ø for individual 30 s measurements shows that this prototype instrument already approaches the best reported literature values using continuous wave lead alloy tunable diode lasers. The application of pulsed near room-temperature quantum cascade lasers to this demanding problem opens the possibility of field worthy rapid response isotopic instrumentation and attests to the maturity of these lasers as spectroscopic sources.
Content may be subject to copyright.
A high precision pulsed quantum cascade laser spectrometer for
measurements of stable isotopes of carbon dioxide
J. B. MCMANUS*y, D. D. NELSONy, J. H. SHORTERy,
R. JIMENEZz, S. HERNDONy, S. SALESKA§ and M. ZAHNISERy
yAerodyne Research, Inc., 45 Manning Road, Billerica, MA US 01821
zHarvard University, Department of Earth and Planetary Sciences,
20 Oxford Street, Cambridge, MA, US 02138
§University of Arizona, Department of Ecology and
Evolutionary Biology, Tucson, AZ 85721, USA
(Received 15 February 2005; in final form 2 July 2005)
We describe a prototype instrument using a Peltier cooled quantum cascade
laser for precise measurement of stable carbon (13C/12 C) isotopologue ratios in
atmospheric CO
2
. Using novel optics and signal processing techniques in a
compact instrument, we are able to detect the difference between sample and
reference with a precision of 0.1ø(2standard error of mean of 11 samples)
in 10 min of analysis time. The standard deviation of 0.18øfor individual 30 s
measurements shows that this prototype instrument already approaches the best
reported literature values using continuous wave lead alloy tunable diode lasers.
The application of pulsed near room-temperature quantum cascade lasers to this
demanding problem opens the possibility of field worthy rapid response isotopic
instrumentation and attests to the maturity of these lasers as spectroscopic
sources.
1. Introduction
The measurement of isotopic ratios in the natural environment is a challenging
problem that can be addressed with high resolution laser spectroscopic techniques
[1–9]. Most of the prior work in laser spectroscopic measurement of isotopic
ratios has been with continuous wave (CW) cryogenically cooled lasers (especially
mid-infrared lead alloy semiconductor lasers) [1–7]. The advent of pulsed near room-
temperature quantum cascade lasers (QCLs) [10, 11] opens the possibility of field
worthy rapid response instrumentation for measurement of stable isotopologues.
New isotopic measurement modes may be implemented, such as eddy covariance
flux determination. The successful application of QCLs to isotopic measurements
also attests to their maturity as spectroscopic sources.
*Corresponding author. Email: mcmanus@aerodyne.com
Journal of Modern Optics
Vol. 52, No. 16, 10 November 2005, 23092321
Journal of Modern Optics
ISSN 0950–0340 print/ISSN 1362–3044 online #2005 Taylor & Francis
http://www.tandf.co.uk/journals
DOI: 10.1080/09500340500303710
Measurement of the ratios of atomic isotopes in the molecules that participate
in environmental exchange processes is an important tool for understanding those
processes and the natural environment [12–15]. This is because there are charac-
teristic isotopic enrichments or depletions generated by the chemical reactions or
biochemical processes leading to sources and sinks of atmospheric trace gases. Stable
isotope abundances are typically expressed in terms of delta units (,ø), which are a
comparison of the ratio of two isotopes in a sample (e.g. Rx¼13 C=12 C), to the same
ratio in a standard, R
std
. The delta unit is defined as: x¼½ðRxRstdÞ=Rstd1000.
Isotopic deltas of interest in the natural environment may be 1øor less, and the
desired measurement precision is 0.1ø.
Isotopic ratio measurements are difficult because the ratio variations are small,
the minor isotopes are relatively rare, and the molecules of interest may be quite
dilute in the natural environment. As an example, consider one of the stronger
isotopic enhancements due to biological processes, the fixing of carbon in photo-
synthesis [12, 13]. Different classes of plants discriminate differently against uptake
of 13CO2relative to 12 CO2, so that woody plants (‘C3’ photosynthesis) have
13C16øand grasses (‘C4’ photosynthesis) have 13 C4ø, relative to
atmospheric CO
2
. One can measure the ratio 13 C/12C to identify the class of plants
involved in atmospheric sources and sinks of CO
2
[14, 15]. In atmospheric sampling,
the average concentration of CO
2
is 380 ppm and the concentration of 13CO2
is 3.8 ppm. The difference in concentration of 13CO2produced by burning C3
versus C4 plants (with the resulting CO
2
diluted to near background levels) would be
0.05 ppm. The standard method for determining trace gas stable isotope ratios
is isotope ratio mass spectrometry (IRMS) [12, 13], which provides 13 C with
precision of 0.01–0.05ø. However, IRMS has a number of experimental draw-
backs, especially in the great care needed in sample preparation (often involving
chemical processing or purification). Commercial IRMS units require permanent
installation for reliable operation, so that field samples are normally returned to
the laboratory and analysed long after they are taken.
Real time continuous isotopic measurements would greatly facilitate the
investigation of environmental exchange processes, allowing guidance of sampling
by real time analysis and enabling new techniques. If isotopic measurements can be
made with sufficient accuracy at a rate of 10 Hz, then the correlation of upward
wind velocity with concentration yields area flux, a technique called eddy covariance
flux measurement. Recently, lead alloy tunable diode lasers (TDLs) have been
applied to continuous in situ atmospheric isotope ratio measurements [1–7].
Several of these systems have been reported with precisions on the order of 0.1ø
[1, 4–6]. Lead alloy TDLs represent an advance in isotope measurement methods
but they are far from ideal, due to the need for liquid nitrogen (LN2) cooling, and
the lasers often change characteristics after cooling cycles.
Quantum cascade lasers (QCLs) provide alternative sources of high resolution
mid-infrared radiation that have overcome some of the difficulties associated with
lead alloy TDLs. QCLs can operate in the mid-infrared, where many molecules
display their strongest fundamental vibrational bands, for the highest sensitivity.
QCLs can operate near room temperature, and offer single-mode continuous tuning
and good power output. Since QC lasers were first demonstrated [16], they have
2310 J. B. McManus et al.
undergone rapid development [10, 11] and been used for atmospheric trace gas
measurements, in CW operation with LN2 cooling [17] and pulsed with thermo-
electric cooling [8, 9, 18–22].
In this paper we describe an instrument based on a pulsed quantum cascade laser
for precise measurement of carbon isotope ratios (13C/12 C) and present laboratory
results. The new QC laser spectrometer employs a dual path length optical absorp-
tion cell [1] and improved signal processing techniques. The precision and minimal
detectable absorbance we obtain with pulsed QC lasers is comparable to that
achieved with LN2 cooled CW TDLs, despite the broader linewidths inherent to
pulsed operation.
2. Instrument design
In the design of this instrument we have addressed problems that are specific to
isotopes and general problems for precision measurements with pulsed QC lasers.
At the needed level of precision (of 104, or 0.1ø) a number of subtle effects can
limit the ultimate performance, such as the method of operation of the laser, or
choices of spectral features and the method of deriving concentrations. The isotopic
specific measurement problems include the large difference in abundances of
the minor and major isotopes, and the temperature dependence of absorption
strengths. Our group at Aerodyne Research, Inc. has worked over the past several
years improving the precision of pulsed QCL spectrometers for several applications,
resulting in many of the general approaches described here [20–22]. The QCL
spectrometer combines commercially available QC lasers, an optical system, and
a computer-controlled electronic system for driving the laser and data processing.
The spectrometer is designed for simultaneous measurement of sample, pulse
normalization and frequency-lock spectra. We have used commercially available
distributed feedback InGaAs–AlInAs/InP QC lasers designed for pulsed operation
(Alpes Lasers, Neuchaˆ tel, Switzerland). An essential aspect of this work is that
it has been possible to select a QC laser that covers the frequency range from 2310
to 2314 cm
1
, which meets the line selection criteria described below.
Pulsed laser operation presents two main problems for high precision mea-
surements, noise due to pulse-to-pulse energy variations and the increase in laser
linewidth. We have greatly reduced the problem of pulse-to-pulse variations by
normalizing [20–22]. By taking advantage of the time delay of light propagation
through the multipass cell, we can use a single infrared detector for both the
absorption signal and pulse normalization. The problem of laser linewidth and
the associated problem of lineshape stability remain precision limiting issues. Even
though we minimize the linewidth by using short pulses (10 ns) and operate the
laser close to threshold, the laser width is greater than the molecular (Doppler)
absorption linewidth.
Stability against temperature changes is a key factor in precision measurements
and the basic thermal stability may be improved by careful selection of the two
absorption lines. Infrared linestrengths vary with temperature, so drift in the instru-
ment temperature can mask small isotopic deltas [5–7]. However, if the transitions
A high precision pulsed quantum cascade laser spectrometer 2311
for both isotopes have similar lower state energy levels, the temperature variation
of the two line strengths are closely matched. If the instrument temperature is to
within 0.1 K, the lower state energy level separated can be only 100 to 300 cm
1
to
maintain a precision of 0.2 to 0.5ø[6]. We have previously reported surveys of
thermal sensitivity for lines for the isotopes of CO
2
and CH
4
[1]. The line pairs
around 2314 cm
1
require a temperature stability of only 0.2 K to obtain a precision
of 0.1ø[1]. The temperature stability requirement is negligible using the line pairs
at 2311 cm
1
suggested by Weidmann et al. [8], which we have used extensively in
this study. The choice of spectral features to reduce thermal sensitivity may be less
optimal for gases other than CO
2
, since the following factors also need considera-
tion.
1. When using a single laser, lines from the major and minor isotopes must be
within the frequency scanning range of the laser.
2. The lines should have large transition moments, for maximum sensitivity.
3. The lines should be free of interference from other atmospheric gases.
A key design feature of our laser spectrometer is a dual-path multipass cell used
to compensate for the large difference in concentration between major and minor
isotopologues of CO
2
, an effect limiting precision. When long path lengths are
used to increase the absorption depth due to the minor constituent, then the major
constituent absorption depth may be excessive, giving essentially no transmission at
line centre. Our approach uses two different paths through the same measurement
volume for the major and minor isotope, where the ratio of path lengths is similar to
the abundance ratio, yielding a similar absorption depth for both isotopologues
[1]. The dual path-length cell is based on an astigmatic Herriott cell [23], with
the two paths corresponding to either 174 or 2 traverses of the 0.32 m base length [23].
A diagram of the cell coupling arrangement is shown in figure 1. The beam for the
174 pass pattern (for 13CO2) is injected at an angle to the centre axis and it exits
on the opposite side of the axis. The 2 pass pattern is generated by placing a
beamsplitter (BaF
2
) on the centre axis, reflecting 8% of the light straight into
the cell. The 2-pass beam then returns on the same axis after reflecting from the
back mirror. The resulting optical path lengths of 56.08 and 0.753 m have a ratio
of 74:1, nearly balancing the 12CO2=13CO2isotopomer ratio of 91:1. The dual
pathlength concept also greatly lessens the effect of temperature and pressure
variations since both isotopomers are measured in the same volume.
The instrument optical system, shown in figure 2, collects light from the QC laser
and directs it along four different paths: two paths through the absorption cell,
plus the pulse normalization and frequency-locking reference paths. The highly
divergent QC laser beam is collected by a reflecting microscope objective and
imaged at a pinhole (used only for initial alignment). After the first focus,
the beam passes through a wedged BaF
2
beamsplitter. The front reflection beam is
used for the pulse normalization spectrum, which defines the wavelength dependence
of the laser power, along with its temporal variability. The optical path length
of the pulse normalization beam is matched to the sample path length outside the
absorption cell, insuring that residual atmospheric absorption in that path is
cancelled. The second reflection from the beamsplitter is directed through a short
2312 J. B. McManus et al.
absorption cell with low pressure CO
2
, and its spectrum is used to actively control
the average frequency of the laser by changing the Peltier temperature. The beam
transmitted through the beam splitter is re-imaged into the astigmatic Herriott
multipass cell [23]. In this study, we sample at a relatively low pressure (7 torr),
which serves to reduce consumption of sample gas, minimize overlap of adjacent
lines and reduce optical depth at ambient concentrations. The optical board is
thermally controlled to better than 1 K and the instrument enclosure is purged with
CO
2
free dry air.
The spectral measurement is based on rapidly sweeping the laser across the
selected absorption lines, digitizing the direct-absorption signals and quantitatively
174 Pass Outpu
t
2-Pass Output
Laser Input
BaF2 Beamsplitter
Astigmatic
Herriott Cell
56m @ 174 passes
0.75m @ 2 passes
Figure 1. Optical arrangement to produce dual pathlength optical absorption cell.
A high precision pulsed quantum cascade laser spectrometer 2313
fitting the spectra to derive absolute concentrations. A unified computer system,
using ‘TDL Wintel’ software developed at Aerodyne Research, Inc., drives the
laser and synchronously analyses data in real time. A diagram of the signal
processing system is shown in figure 3. The laser is excited just above threshold
with 10–20 ns electrical pulses at 1 MHz. Spectral scans (over 0.5 cm
1
) are
Dual
Detector
QCL in housing
Trace Laser
Beamsplitter
Absorption Cell
15x
Main Path
Norm. Path
174 2
0.33m
0.64m
Figure 2. Optical layout of single quantum cascade laser instrument. The optical base
measures 0:33 0:64 m.
~100
Pulse
s
~0.5 cm1
Ramp
Peltier
100 µs
1µs
Current
QCL
i(t)
Absorption
Detector
Frequency
t
time
V(t)
S
PN
PN S PN S
…..
1µs
100 µs
100 µs
VN(t)
VS(t)
Average
Divide
Fit spectra
Calc. Conc.
Adj. Peltier
~10 ns
Figure 3. Laser control and signal processing schematic.
2314 J. B. McManus et al.
obtained by simultaneously applying a sub-threshold current ramp, which modulates
the laser temperature and thus its spectral frequency. Fast (5 MHz) DAC/ADC
boards synchronously trigger the pulse electronics and integrate the resulting signals
from LN2-cooled InSb photodiodes. The DAC board also generates a TTL gate that
defines the ramp duration, including a short laser-off period at the end of each sweep
to measure the detector offset. Since this instrument employs separate beam paths
and detectors to monitor each isotope simultaneously, we employ dual data
acquisition processes operating in parallel, to collect data through separate
acquisition boards. We run two instances of TDL Wintel with processes that are
precisely synchronized by operating the instances as ‘master and slave’. The master
instance controls the laser and sends timing signals to the slave, which then
constructs appropriate timing gates for its detector. A typical spectral sweep consists
of 100–400 pulses, giving a spectral sweep rate of 2.5–10 kHz. Concentrations and
laser linewidths are determined in real time from the spectra through a nonlinear
least-squares fitting algorithm (Levenberg-Marquardt) that uses fitting functions
composed from Voigt [24] molecular profiles (based on the HITRAN [25] database
plus measured pressure and temperature), convolved with a (Gaussian) laser line-
shape and added to a polynomial baseline. Derived mixing ratios are typically
accurate to 5% without calibration.
3. Laboratory data
The basic data produced by the instrument are the two absorption spectra, for 13 CO2
using the long path (56 m) and 12CO2using the short path (0.74 m). The spectral pair
near 2311 cm
1
is shown in figure 4, for a sample with CO
2
concentration 350 ppm
at a pressure of 7.5 torr. The data here are reported at 1 Hz, where the concentration
noise is 150 ppb. Each panel in figure 4 displays three curves, (1) the experimental
absorption spectrum, (2) a spectral fit to the data, a convolution of the Voigt
molecular lineshape with the laser lineshape, and (3) a spectral simulation of only
the molecular lineshape. We employ a simple Gaussian instrument profile in the
convolution, which does a good job of capturing the measured lineshape and area,
even though some details of the observed shape are missed by using a symmetrical
profile. Comparing the molecular line to the data shows the degradation in
resolution due to the significant laser linewidth. For the 13 CO2absorption line
the apparent peak absorbance is 0.1, but the actual peak absorbance is three times
greater. The 12CO2line at 2311.105 cm
1
(figure 4(a)) is optically black at centre
even though that is not apparent from the experimental spectrum. Nevertheless,
the analysis does a good job of reproducing the experimental 12 CO2spectrum. These
lines are strong enough that the absorbance area is not conserved and accurate
results can only be obtained by doing a convolution of the molecular spectrum with
the instrumental width.
In figure 5 we show in more detail the contributions to the spectrum in the
long absorption path for 13CO2. The underlying 12CO2lines contribute significantly
to the absorption area in the wings of the 13 CO2absorption region, despite the low
pressure and good separation. To compensate for this effect, the fitting procedure is
A high precision pulsed quantum cascade laser spectrometer 2315
done for the two isotopomers simultaneously. Although it would be possible to vary
both species in the fit, there is insufficient 12 CO2absorption in the 13 CO2region to
accurately retrieve its concentration. Instead, the 12 CO2concentration is set to the
more accurate value determined from the short path absorption spectrum. This
greatly reduces the dependence of the 13 CO2=12CO
2
ratio determination on
variations in the 12CO2concentration.
The long term stability of the instrument is evaluated using the Allan variance
technique, which distinguishes high frequency random noise from drifts at
longer time scales [26]. The contributions of random noise to the variance (2)
decrease as t
1
, so a plot of variance with increasing integration time is expected to
have a (log-log) slope of 1 at short times, while at longer times the variance curve
levels off and rises due to drifts. In figure 6 we show time series data and an Allan
1.0
(a)
(b)
0.8
0.6
0.4
0.2
0.0
TRANSMISSION
2311.42311.32311.22311.12311.0
FREQUENCY (cm1)
1.0
0.8
0.6
0.4
0.2
0.0
TRANSMISSION
174 PASS - 56m
2 PASS - 0.74 m
SPECTRA
CONVOLVED FIT
VOIGT SHAPE
7.5 Torr
12CO2 350 ppm
Laser Line Width 0.0105 cm1
13CO2
2311.399
12CO2
2311.106
Figure 4. CO
2
isotopologue spectra with pulsed QC laser. The cell contains 350 ppm
CO
2
in air at 7.5 torr. The laser linewidth is 0.0105 cm
1
hwhm. The graphs show recorded
spectra (large dots), fit functions (solid lines) and simulated Voigt absorption lines (dashed
lines), for the long path (upper panel, (a)) and short path (lower panel, (b)).
2316 J. B. McManus et al.
plot of both isotopologues and their ratio for a 10 h period. The 13 CO2=12CO
2
ratio
has a 1 s standard deviation of 0.5ø. However, the variance does not decrease with
t1 as would be expected for random (white) noise, but rather behaves as f0:5
‘pink’ noise, indicating that that there are correlations in noise sources at time scales
of 1–100 s. The variance minimum at 200 s corresponds to a minimum
Allan
of 0.1ø.
The precision of an isotopic ratio measurement can be improved by alternating
between a sample and reference, with the time for each measurement less than
the averaging at which the Allan variance begins to increase. We present an
alternating measurement in figure 7, sampling gases from two tanks: the ‘sample’
containing 350 ppm total 12 CO2and unknown 13 C, was compared with a ‘reference’
mixture of CO
2
with 13CPDB ¼49:4ø(determined by mass spectrometry
courtesy of Professor Dan Schrag at Harvard University). For the data in figure 7,
the averaging time for each sample was 25 s, with 5 s between samples for gas
flushing. For this experiment we observe that the sample air has 13CCO2¼14:9ø,
relative to PDB. The difference between sample and reference is obtained with a
precision of 0.1ø(2standard error of mean of 11 samples) in 10 min of analysis
time. The standard deviation of 0.18øin individual 30 s intervals shows that this
prototype instrument already approaches the best reported literature values using
1.00
0.98
0.96
0.94
0.92
TRANSMISSION
2311.452311.402311.352311.302311.252311.20
FREQUENCY
(
cm
1
)
SPECTRUM
12
CO
2
350 ppm (FIXED)
13
CO
2
336 ppm (FIT)
Pressure 7.5 Torr
Laser Line Width 0.0105 cm
1
Figure 5. Detail of the contributions to the fit spectrum in the long absorption path.
There are small contributions from a weak 12CO2line that is nearly coincident with the
13CO2line.
A high precision pulsed quantum cascade laser spectrometer 2317
continuous lead-alloy TDLs [4]. While the averaged data yields a best precision
in delta of 0.1ø, the large-signal accuracy of the spectroscopically derived delta is
substantially less (10%). The spectroscopic accuracy is limited by uncertainty
in the tabulated HITRAN line strengths, and by uncertainty in the spectral baseline,
which relates to uncertainty in the laser lineshape. To achieve higher accuracy it is
necessary to calibrate with standards.
4. Discussion
In the prototype instrument described above, we achieve a precision in isotopic ratio
measurements of 0.18øin 30 s (1) and 0.1øin 600 s (2). This result is achieved
by carefully controlling numerous system parameters, especially laser base tempera-
ture and average centre wavelength, as well as the temperature of the optical
108
107
Allan Var (σ2)
Allan Var (σ2)
100101102103104
INTEGRATION TIME (s)
0.988
0.987
0.986
0.985
0.984
0.983
RATIO
1/20/2005
9:00AM 12:00 PM 3:00 PM
364×103
362
360
358
CO2 (ppb)
103
2
4
104
2
4
σAllan=45 ppb
τAllan=200 s
σAllan=0.1 ‰
τAllan=200 s
12CO2
13CO2 / 0.01124
1 s rms 0.5 ‰
Figure 6. Noise in isotope ratio and Allan plot showing long term drift behaviour.
2318 J. B. McManus et al.
table and the pulse electronics. We believe the largest source of noise and drift in the
1 to 100 s time scale is residual laser frequency fluctuations, affecting both the peak
position and the laser linewidth. Another significant drift effect is variation in the
residual CO
2
in the optical enclosure.
The success to date of our QC laser spectrometer for isotopic measurements
provides the opportunity to comment on the utility of these lasers as spectroscopic
sources. The measurement method presented is possible because a high-quality QCL
was available at the desired wavelength, a significant factor given the current scarcity
of QC laser manufacturers. The laser is very well behaved in tuning with temperature
and current. The reproducibility of the output as we ramp and pulse the current
contributes to the final precision performance of the instrument. Pulse energy noise
can be divided out with our ‘pulse normalization’ approach, greatly reducing
multiplicative noise. With careful system control, we can achieve noise performance
approaching the best reported literature values using CW TDLs [4]. The laser used
in this study has operated continuously for more than three months with negligible
change in characteristics.
There are several directions available to improve the performance of this
instrument. We continue to improve stabilization and control of the laser and
345×103
340
335
330
325
CO2 (ppb)
8:52 AM
1/21/2005
8:54 AM 8:56 AM 8:58 AM 9:00 AM 9:02 AM
0.97
0.96
0.95
0.94
RATIO / 0.01124
35.5×103
35.0
34.5
34.0
33.5
DIFFERENCE
STD DEV = 0.18‰
SAMPLE-REFERENCE
mean = 34.5 ± 0.1 ‰ (2σ,n=11)
12CO213CO2 / 0.01124
RMS = 0.46 ‰ (1 Hz)
SAMPLE
REFERENCE
Figure 7. Alternating measurements of 13CO2=12CO2in sample and reference air from
tanks. Reference was made by diluting pure CO
2
(13CPDB ¼49:4ø) with dry N
2
to match
sample gas ½12CO2of 350 ppm. Reference gas ½12CO2of 345 ppm (bottom panel) and
isotope ratio of 0:940 ¼60ø(middle panel) are directly retrieved (uncalibrated) values.
The flow rate of 0.5 SLPM in a cell volume of 0.5 l at pressure of 7 torr corresponds to a cell
flushing (1/e) time of 0.6 s. The first 5 s of each 30 s interval were discarded.
A high precision pulsed quantum cascade laser spectrometer 2319
instrument overall, which has been a key element of the performance achieved to
date. While we have focused to date on 13 CO2and 12CO2, we could also apply these
techniques to, e.g. 13CH4=12 CH4or 12 C18O16 O=12 C16O2. If we can achieve sufficient
precision with fast response time then we could measure isotopic fluxes by eddy
covariance. We have previously presented a dual QCL instrument [22] and we plan
to modify that design to include the isotope dual-path optics. The instrument
presented here has a LN2 cooled detector, but with continued improvement
in thermoelectrically cooled infrared detectors a fully non-cryogenic instrument
is possible. This would allow completely unattended long term operation and greatly
advance the application of infrared spectroscopy to isotope analysis.
5. Summary and conclusions
A high precision pulsed-quantum cascade laser spectrometer operating in the
mid-infrared (2311 cm
1
) has been demonstrated to determine isotopic abundances
by measuring 13CO2=12 CO
2
in atmospheric air. The instrument uses a dual path-
length absorption cell with a longer path (56 m) for the minor isotope and a
shorter path for the major isotope (0.74 m), with absorption lines selected for
similar lower state energies to negate the effects of sample temperature variation.
By stabilizing the pulse electronics and normalizing pulse-to-pulse intensity varia-
tions, measurement precision for each isotopologue of 5 104(1 s rms) relative to
their ambient mixing ratios may be achieved. The difference in the isotopologue
ratio between two samples of ambient air can be determined by alternating between
sample and reference with a replicate precision of 0.18ø(1) standard deviation
for 30 s averaging intervals. This corresponds to a standard error of mean over
11 cycles of 0.1ø(2) in 10 min of measurement time alternating between sample
and reference in a flowing system.
Acknowledgments
We gratefully acknowledge the contributions of Patrick Kirwin and Jeff Mulholland
(Aerodyne Research, Inc.) to the engineering and construction of the QCL spectro-
meter. We also thank Steve Wofsy and Bruce Daube (Harvard University) and
Frank Tittel (Rice University) for valuable advice and discussions. We thank
Dan Schrag at Harvard University for isotopic analysis. We thank Antoine Muller,
Yargo Bonneti, Guillaume Vanderputte and Stephan Blaser of AlpesLasers for
providing the QC laser and other high quality components used in this system.
Funding for instrument development has been provided by the US Department of
Energy STTR and SBIR programs.
References
[1] J.B. McManus, M.S. Zahniser, D.D. Nelson, et al., Spectrochim. Acta A 58 2465 (2002).
[2] K. Uehara, K. Yamamoto, K. Kikugawa, et al., Sens. Actuators B 74 173 (2001).
2320 J. B. McManus et al.
[3] T.J. Griffis, J.M. Baker, S.D. Sargent, et al., Agric. Forest Meteorol. 124 15 (2004).
[4] D.R. Bowling, S.D. Sargent, B.D. Tanner, et al., Agric. Forest Meteorol. 118 1 (2003).
[5] P. Bergamaschi, A.M. Brenninkmeijer, M. Hahn, et al., J. Geophys. Res. 103 8227 (1998).
[6] P. Bergamaschi, M. Schupp and G.W. Harris, Appl. Opt. 33 7704 (1994).
[7] J.F. Becker, T.B. Sauke and M. Loewenstein, Appl. Opt. 31 1921 (1992).
[8] D. Weidmann, G. Wysocki, C. Oppenheimer, et al., Appl. Phys. B Lasers Opt. 80
255 (2005).
[9] A.A. Kosterev, R.F. Curl, F.K. Tittel, et al., Opt. Lett. 24 1764 (1999).
[10] J. Faist, F. Capasso, C. Sirtori, et al., Electron. Lett. 32 560 (1996)
[11] F. Capasso, R. Paiella, R. Martini, et al., IEEE J. Quantum Electron. 38 511 (2002).
[12] P. Ghosh and W.A. Brand, Int. J. Mass Spectrom. 228 1 (2003).
[13] S.E. Trumbore, in Biogenic Trace Gases: Measuring Emissions from Soil and Water,
edited by P.A. Matson and R. C. Harriss (Blackwell Science Ltd, Oxford, 1995).
[14] C.-T. Lai, A.J. Schauer, C. Owensby, et al., J. Geophys. Res. 108(D18) 4566 (2003).
[15] S.R. Saleska, S.D. Miller, D.M. Matross, et al., Science 302 1554 (2003).
[16] J. Faist, F. Capasso, D.L. Sivco, et al., Science 264 553 (1994).
[17] C.R. Webster, G.J. Flesch, D.C. Scott, et al., Appl. Opt. 40 321 (2001).
[18] A.A. Kosterev and F.K. Tittel, IEEE J. Quantum Electron. 38 582 (2002).
[19] A.A. Kosterev, F.K. Tittel, R. Kohler, et al., Appl. Opt. 41 1169 (2002).
[20] D.D. Nelson, J.H. Shorter, J.B. McManus, et al., Appl. Phys. B Lasers Opt. 75 343 (2002).
[21] D.D. Nelson, J.B. McManus, S. Urbanski, et al., Spectrochim. Acta A 60 3325 (2004).
[22] R. Jimenez, S. Herndon, J.H. Shorter, et al., SPIE Proc. 5738 37 (2005).
[23] J.B. McManus, P.L. Kebabian and M.S. Zahniser, Appl. Opt. 34 3336 (1995).
[24] J. Humlı
´cek, JQSRT 27 437 (1982).
[25] L.S. Rothman, A. Barbe, D.C. Benner, et al., JQSRT 82 5 (2003).
[26] D.W. Allan, Proc. IEEE 54 221 (1966).
A high precision pulsed quantum cascade laser spectrometer 2321
... Laser absorption spectroscopy determines isotopologue abundances by detecting the absorbance of laser light by the analyte gas (McManus et al., 2006). Laser spectroscopy provides a quick and cost-effective approach for quantifying isotopologue ratios directly in CO 2 gas and has the potential to replace IRMS for many typical applications (Nelson et al., 2008;Prokhorov et al., 2019;Stoltmann et al., 2017;Tuzson et al., 2008;Wang et al., 2020;Yanay et al., 2022). ...
... During an approximately 4-min-long measurement cycle, the oxygen isotope ratios of the analyte are calculated every second. Specifically, the TDLWintel software calculates mole fractions (X; also referred to as mixing ratios) of the three investigated molecular species ("626," "627," and "628") from the absorption spectra through spectral fitting (McManus et al., 2006). For the spectral fitting, the concentration of the "free-path CO 2 " species in the TDLWintel was set to zero. ...
Article
Full-text available
Analyzing the triple oxygen isotope (Δ′ ¹⁷ O) composition of carbonates and air CO 2 can provide valuable information about Earth system processes. However, accurately measuring the abundance of the rare ¹⁷ O‐bearing CO 2 isotopologue using isotope ratio mass spectrometry presents significant challenges. Consequently, alternative approaches, such as laser spectroscopy, have been developed. Here, we describe an adaptable dual inlet system for a tunable infrared laser direct absorption spectrometer (TILDAS) that maintains stable instrumental conditions for subsequent sample and reference measurements. We report ∆′ ¹⁷ O measurements on three types of samples: reference CO 2 , CO 2 derived from the acid digestion of carbonates, and air CO 2 . The external repeatability (±1 σ ) for reference‐sample‐reference bracketing measurements is generally better than ±10 ppm, close to the average internal error of ±6 ppm. Our results demonstrate that laser spectroscopy is a capable technique for measuring triple oxygen isotopes with the precision required to resolve variations in the ∆′ ¹⁷ O values of air CO 2 and to use the ∆′ ¹⁷ O of carbonates for paleothermometry.
... ATom flew south along the center of the Pacific basin, following 180°W within ~15° longitude, and north along the center of the Atlantic basin, generally between 15° and 30°W (SI Appendix, Fig. S16). SI Appendix, Table S5 describes the measurements used in our analysis (72)(73)(74)(75)(76)(77)(78)(79)(80)(81)(82)(83)(84)(85)(86)(87)(88), which spans ATom-1, 2, 3, and 4, corresponding to Northern Hemisphere summer (28 July to 22 August, 2016), winter (26 January to 22 February, 2017), fall (28 September to 26 October, 2017), and spring (24 April to 21 May, 2018 (30). Box model simulations of HO x (OH plus hydroperoxyl, HO 2 ) forced by measurements (excluding HO x ) along the ATom-1 and 2 flight tracks show consistency with ATom-1 and 2 observed HO x (24). ...
Article
Full-text available
The hydroxyl radical (OH) fuels atmospheric chemical cycling as the main sink for methane and a driver of the formation and loss of many air pollutants, but direct OH observations are sparse. We develop and evaluate an observation-based proxy for short-term, spatial variations in OH (ProxyOH) in the remote marine troposphere using comprehensive measurements from the NASA Atmospheric Tomography (ATom) airborne campaign. ProxyOH is a reduced form of the OH steady-state equation representing the dominant OH production and loss pathways in the remote marine troposphere, according to box model simulations of OH constrained with ATom observations. ProxyOH comprises only eight variables that are generally observed by routine ground- or satellite-based instruments. ProxyOH scales linearly with in situ [OH] spatial variations along the ATom flight tracks (median r2 = 0.90, interquartile range = 0.80 to 0.94 across 2-km altitude by 20° latitudinal regions). We deconstruct spatial variations in ProxyOH as a first-order approximation of the sensitivity of OH variations to individual terms. Two terms modulate within-region ProxyOH variations-water vapor (H2O) and, to a lesser extent, nitric oxide (NO). This implies that a limited set of observations could offer an avenue for observation-based mapping of OH spatial variations over much of the remote marine troposphere. Both H2O and NO are expected to change with climate, while NO also varies strongly with human activities. We also illustrate the utility of ProxyOH as a process-based approach for evaluating intermodel differences in remote marine tropospheric OH.
... In order to minimise such phenomena, it is necessary to limit the width of the pulses to a maximum of tens of nm and to maintain their amplitude close to the laser excitation threshold. Thus, the QCL can be tuned in the range of 1-2 cm −1 with the pulse repetition period ranging from several dozen to several thousand hertz [15]. ...
Article
Full-text available
Numerical research into the QCL tunability aspects in respect to being applied in chemical substance detection systems is covered in this paper. The QCL tuning opportunities by varying power supply conditions and geometric dimensions of the active area have been considered. Two models for superlattice finite (FSML) and infinite (RSM) size were assumed for simulations. The results obtained have been correlated with the absorption map for selected chemical substances in order to identify the potential detection possibilities.
Chapter
Stable carbon isotopes are a powerful tool to study photosynthesis. Initial applications consisted of determining isotope ratios of plant biomass using mass spectrometry. Subsequently, theoretical models relating C isotope values to gas exchange characteristics were introduced and tested against instantaneous online measurements of 13C photosynthetic discrimination. Beginning in the twenty-first century, laser absorption spectroscopes with sufficient precision for determining isotope mixing ratios became commercially available. This has allowed collection of large data sets at lower cost and with unprecedented temporal resolution. More data and accompanying knowledge have permitted refinement of 13C discrimination model equations, but often at the expense of increased model complexity and difficult parametrization. This chapter describes instantaneous online measurements of 13C photosynthetic discrimination, provides recommendations for experimental setup, and presents a thorough compilation of equations available to researchers. We update our previous 2018 version of this chapter by including recently improved descriptions of (photo)respiratory processes and associated fractionations. We discuss the capabilities and limitations of the diverse 13C discrimination model equations and provide guidance for selecting the model complexity needed for different applications.
Article
Full-text available
Mesophyll conductance to CO2 from the intercellular airspace to the CO2-H2O exchange site has been estimated using δ18O measurements (gm18). However, the gm18 estimates are affected by the uncertainties in the δ18O of leaf water where CO2-H2O exchange takes place and the degree of equilibration between CO2 and H2O. We show that measurements of Δ17O (i.e. Δ17O=δ17O-0.528×δ18O) can provide independent constraints on gm (gmΔ17) and that these gm estimates are less affected by fractionation processes during gas exchange. The gm calculations are applied to combined measurements of δ18O and Δ17O, and gas exchange in two C3 species, sunflower (Helianthus annuus L. cv "sunny") and ivy (Hedera hybernica L.), and the C4 species maize (Zea mays). The gm18 and gmΔ17--- estimates agree within the combined errors (p value 0.876). Both approaches are associated with large errors when the isotopic composition in the intercellular air space becomes close to the CO2-H2O exchange site. Although variations in Δ17O are low, it can be measured with much higher precision compared to δ18O. Measuring gmΔ17 has a few advantages compared to gm18: i) it is less sensitive to uncertainty in the isotopic composition of leaf water at the isotope exchange site and ii) the relative change in the gm due to an assumed error in the equilibration fraction θeq is lower for gmΔ17 compared to gm18. Thus, using Δ17O can complement and improve the gm estimates in settings where the δ18O of leaf water varies strongly, affecting the δ18O (CO2) difference between the intercellular air space and the CO2-H2O exchange site.
Article
Oxygen and carbon stable isotope ratios (18O/16O, 13C/12C, and 17O/16O) of CO2 have been crucial in helping us understand Earth and planetary systems. These ratios have also been used in medicine for the noninvasive diagnosis of diseases from exhaled breath and for quantifying biochemical or metabolic reactions and in determining the production area of agricultural products. The current method for measuring the stable isotope ratios of CO2 is primarily gas-source isotope ratio mass spectroscopy (IRMS). Due to the recent demand for isotopic microanalysis of carbonates and organic compounds, the sample size required for isotopic measurements has been reduced to approximately 2 nmol CO2 (equivalent to 0.2 μg CaCO3 and 24 ng carbon) by using high-precision IRMS. We report a novel method using tunable mid-infrared laser direct absorption spectroscopy (TILDAS) for sensitive measurements of 18O/16O and 13C/12C in subnanomolar CO2. This method can accurately measure 18O/16O and 13C/12C in CO2 with a repeatability of less than 0.03‰ (n = 28) in a range of 0.3 nmol (equivalent to 0.03 μg CaCO3 and 3.8 ng carbon) to 30 nmol. This is a sample size 1 order of magnitude smaller than currently available sensitive analytical techniques. In addition, the TILDAS system measures 17O/16O simultaneously with a repeatability of less than 0.06‰ (n = 28). Our method is a major advance in supersensitive CO2 stable isotopic analyses for various fields.
Article
Full-text available
In situ measurements of aerosol microphysical, chemical, and optical properties were made during global-scale flights from 2016–2018 as part of the Atmospheric Tomography Mission (ATom). The NASA DC-8 aircraft flew from ∼ 84∘ N to ∼ 86∘ S latitude over the Pacific, Atlantic, Arctic, and Southern oceans while profiling nearly continuously between altitudes of ∼ 160 m and ∼ 12 km. These global circuits were made once each season. Particle size distributions measured in the aircraft cabin at dry conditions and with an underwing probe at ambient conditions were combined with bulk and single-particle composition observations and measurements of water vapor, pressure, and temperature to estimate aerosol hygroscopicity and hygroscopic growth factors and calculate size distributions at ambient relative humidity. These reconstructed, composition-resolved ambient size distributions were used to estimate intensive and extensive aerosol properties, including single-scatter albedo, the asymmetry parameter, extinction, absorption, Ångström exponents, and aerosol optical depth (AOD) at several wavelengths, as well as cloud condensation nuclei (CCN) concentrations at fixed supersaturations and lognormal fits to four modes. Dry extinction and absorption were compared with direct in situ measurements, and AOD derived from the extinction profiles was compared with remotely sensed AOD measurements from the ground-based Aerosol Robotic Network (AERONET); this comparison showed no substantial bias. The purpose of this work is to describe the methodology by which ambient aerosol properties are estimated from the in situ measurements, provide statistical descriptions of the aerosol characteristics of different remote air mass types, examine the contributions to AOD from different aerosol types in different air masses, and provide an entry point to the ATom aerosol database. The contributions of different aerosol types (dust, sea salt, biomass burning, etc.) to AOD generally align with expectations based on location of the profiles relative to continental sources of aerosols, with sea salt and aerosol water dominating the column extinction in most remote environments and dust and biomass burning (BB) particles contributing substantially to AOD, especially downwind of the African continent. Contributions of dust and BB aerosols to AOD were also significant in the free troposphere over the North Pacific. Comparisons of lognormally fitted size distribution parameters to values in the Optical Properties of Aerosols and Clouds (OPAC) database commonly used in global models show significant differences in the mean diameters and standard deviations for accumulation-mode particles and coarse-mode dust. In contrast, comparisons of lognormal parameters derived from the ATom data with previously published shipborne measurements in the remote marine boundary layer show general agreement. The dataset resulting from this work can be used to improve global-scale representation of climate-relevant aerosol properties in remote air masses through comparison with output from global models and assumptions used in retrievals of aerosol properties from both ground-based and satellite remote sensing.
Article
Full-text available
Using laser absorption spectrometry for the measurement of stable isotopes of atmospheric CO2 instead of the traditional isotope ratio mass spectrometry method decreases sample preparation time significantly, and uncertainties in the measurement accuracy due to CO2 extraction and isobaric interferences are avoided. In this study we present the measurement performance of a new dual-laser instrument developed for the simultaneous measurement of the δ13C, δ18O and δ17O of atmospheric CO2 in discrete air samples, referred to as the Stable Isotopes of CO2 Absorption Spectrometer (SICAS). We compare two different calibration methods: the ratio method, based on the measured isotope ratio and a CO2 mole fraction dependency correction, and the isotopologue method, based on measured isotopologue abundances. Calibration with the ratio method and isotopologue method is based on three different assigned whole-air references calibrated on the VPDB (Vienna Pee Dee Belemnite) and the WMO 2007 (World Meteorological Organization) scale for their stable isotope compositions and their CO2 mole fractions, respectively. An additional quality control tank is included in both methods to follow long-term instrument performance. Measurements of the quality control tank show that the measurement precision and accuracy of both calibration methods is of similar quality for δ13C and δ18O measurements. During one specific measurement period the precision and accuracy of the quality control tank reach WMO compatibility requirements, being 0.01 ‰ for δ13C and 0.05 ‰ for δ18O. Uncertainty contributions of the scale uncertainties of the reference gases add another 0.03 ‰ and 0.05 ‰ to the combined uncertainty of the sample measurements. Hence, reaching WMO compatibility for sample measurements on the SICAS requires reduction of the scale uncertainty of the reference gases used for calibration. An intercomparison of flask samples over a wide range of CO2 mole fractions has been conducted with the Max Planck Institute for Biogeochemistry, resulting in a mean residual of 0.01 ‰ and −0.01 ‰ and a standard deviation of 0.05 ‰ and 0.07 ‰ for the δ13C measurements calibrated using the ratio method and the isotopologue method, respectively. The δ18O could not be compared due to depletion of the δ18O signal in our sample flasks because of storage times being too long. Finally, we evaluate the potential of our Δ17O measurements as a tracer for gross primary production by vegetation through photosynthesis. Here, a measurement precision of
Article
Full-text available
A semiconductor injection laser that differs in a fundamental way from diode lasers has been demonstrated. It is built out of quantum semiconductor structures that were grown by molecular beam epitaxy and designed by band structure engineering. Electrons streaming down a potential staircase sequentially emit photons at the steps. The steps consist of coupled quantum wells in which population inversion between discrete conduction band excited states is achieved by control of tunneling. A strong narrowing of the emission spectrum, above threshold, provides direct evidence of laser action at a wavelength of 4.2 micrometers with peak powers in excess of 8 milliwatts in pulsed operation. In quantum cascade lasers, the wavelength, entirely determined by quantum confinement, can be tailored from the mid-infrared to the submillimeter wave region in the same heterostructure material.
Article
Full-text available
We report the development of a field-deployable, pulsed quantum cascade laser spectrometer. The instrument is designed to measure the 13C/12C isotopic ratio in the CO2 released from volcanic vents. Specific 12CO2 and 13CO2 absorption lines were selected around 4.3m, where the P-branch of 12CO2 overlaps the R-branch of 13CO2 of the 0001–0000 vibrational transition. This particular selection makes the instrument insensitive to temperature variations. A dual-channel cell balances the two absorption signals. We provide details of the instrument design and a preliminary demonstration of its performance based on laboratory measurements of 16O12C16O and 16O12C18O.
Article
The quantum cascade (QC) laser1 is a new optical source in which one type of carrier, typically electrons, cascading down an electronic staircase, make transitions between energy levels created by quantum confinement. This source has the unique feature of having its emission wavelength entirely controlled by its geometry, i.e., the layer thickness, and not by the material's bandgap. Compared to lasers based on interband transitions the QC laser has the advantage of a large T0 and high temperature operation since the main non-radiative channel is optical phonon emission rather than the Auger effect.
Article
This book covers information on the study of biosphere-atmosphere trace gas exchange including the following subject areas: currently available methods and approaches used to measure trace gases exchanges; the methods that will be used in the future; and the studies, mathematical models, and extrapolation approaches used to describe the fluxes over a range of spatial and temporal scales.
Article
Stable isotope ratios of various ecosystem components and net ecosystem exchange (NEE) CO2 fluxes were measured in a C3-C4 mixture tallgrass prairie near Manhattan, Kansas. The July 2002 study period was chosen because of contrasting soil moisture contents, which allowed us to address the effects of drought on photosynthetic CO2 uptake and isotopic discrimination. Significantly higher NEE fluxes were observed for both daytime uptake and nighttime respiration during well-watered conditions when compared to a drought period. Given these differences, we investigated two carbon-flux partitioning questions: (1) What proportions of NEE were contributed by C3 versus C4 species? (2) What proportions of NEE fluxes resulted from canopy assimilation versus ecosystem respiration? To evaluate these questions, air samples were collected every 2 hours during daytime for 3 consecutive days at the same height as the eddy covariance system. These air samples were analyzed for both carbon isotope ratios and CO2 concentrations to establish an empirical relationship for isoflux calculations. An automated air sampling system was used to collect nighttime air samples to estimate the carbon isotope ratios of ecosystem respiration (δR) at weekly intervals for the entire growing season. Models of C3 and C4 photosynthesis were employed to estimate bulk canopy intercellular CO2 concentration in order to calculate photosynthetic discrimination against 13C. Our isotope/NEE results showed that for this grassland, C4 vegetation contributed ˜80% of the NEE fluxes during the drought period and later ˜100% of the NEE fluxes in response to an impulse of intense precipitation. For the entire growing season, the C4 contribution ranged from ˜68% early in the spring to nearly 100% in the late summer. Using an isotopic approach, the calculated partitioned respiratory fluxes were slightly greater than chamber-measured estimates during midday under well-watered conditions. In addition, time series analyses of our δR measurements revealed that occasionally during periods of high wind speed (increasing the sampling footprint) the C3 cropland and forests surrounding the C4 prairie could be detected and had an impact on the carbon isotopic signal. The implication is that isotopic air sampling of CO2 can be useful as a tracer for evaluating the fetch of upwind airflow in a heterogeneous ecosystem.
Article
Non-cryogenic, laser-absorption spectroscopy in the mid-infrared has wide applications for practical detection of trace gases in the atmosphere. We report measurements of nitric oxide in air with a detection limit less than 1nmole/mole (<1ppbv) using a thermoelectrically cooled quantum cascade laser operated in pulsed mode at 5.26μm and coupled to a 210-m path length multiple-pass absorption cell at reduced pressure (50Torr). The sensitivity of the system is enhanced by operating under pulsing conditions which reduce the laser line width to 0.010cm-1 (300MHz) HWHM, and by normalizing pulse-to-pulse intensity variations with temporal gating on a single HgCdTe detector. The system is demonstrated by detecting nitric oxide in outside air and comparing results to a conventional tunable diode laser spectrometer sampling from a common inlet. A detection precision of 0.12ppb Hz-1/2 is achieved with a liquid-nitrogen-cooled detector. This detection precision corresponds to an absorbance precision of 1×10-5Hz-1/2 or an absorbance precision per unit path length of 5×10-10cm-1 Hz-1/2. A precision of 0.3ppb Hz-1/2 is obtained using a thermoelectrically cooled detector, which allows continuous unattended operation over extended time periods with a totally cryogen-free instrument.
Article
We demonstrate a new method of absorption spectroscopy for precise isotopic composition measurements of environmental molecules such as (∼1/100) and (∼0.4/0.4/100). We combined wavelength-stabilized near-infrared semiconductor lasers with a multipass absorption cell in which laser beams passed different times to compensate the large abundance difference and balance the signal levels. Then we could choose strong absorption lines whose absorption coefficients have almost the same temperature dependencies so as to minimize the effect of the changes in gas temperature on the ratio measurements. Normalization of the second-harmonic absorption signal by the dc signal restored a good proportionality between the signal amplitude and the absorbance. An intercomparison between absorption spectroscopy and mass spectrometry for pure natural-abundance methane samples proved that the precision of the present method is ±3×10−4 (±0.3‰) and the relative accuracy is about the same. The demonstrated high-precision indicates that the laser-spectroscopic method can complement mass spectrometry, especially for isotopic molecular species of a small mass difference, and is potentially a useful means for precise analysis of the biogeochemical cycles of the substances.
Article
This paper describes the status circa 2001, of the HITRAN compilation that comprises the public edition available through 2001. The HITRAN compilation consists of several components useful for radiative transfer calculation codes: high-resolution spectroscopic parameters of molecules in the gas phase, absorption cross-sections for molecules with very dense spectral features, aerosol refractive indices, ultraviolet line-by-line parameters and absorption cross-sections, and associated database management software. The line-by-line portion of the database contains spectroscopic parameters for 38 molecules and their isotopologues and isotopomers suitable for calculating atmospheric transmission and radiance properties. Many more molecular species are presented in the infrared cross-section data than in the previous edition, especially the chlorofluorocarbons and their replacement gases. There is now sufficient representation so that quasi-quantitative simulations can be obtained with the standard radiance codes.In addition to the description and justification of new or modified data that have been incorporated since the last edition of HITRAN (1996), future modifications are indicated for cases considered to have a significant impact on remote-sensing experiments.
Article
Methods for computing the complex probability function w(z) = e−z2 erfc (−iz), which is related to the Voigt spectrum line profiles, are developed. The basic method is a rational approximation, minimizing the relative error of the imaginary part on the real axis. It is complemented by other methods in order to increase efficiency and to overcome the inevitable failure of any rational approximation near the real axis. The procedures enable one to evaluate both real and imaginary parts of w(z) with high relative accuracy. The methods are simple, as demonstrated by a sample FORTRAN program.
Article
Stable isotope ratios of the life science elements carbon, hydrogen, oxygen and nitrogen vary slightly, but significantly in major compartments of the earth. Owing mainly to antropogenic activities including land use change and fossil fuel burning, the ratio of CO2 in the atmosphere has changed over the last 200 years by 1.5 parts per thousand (from about 0.0111073 to 0.0110906). In between interglacial warm periods and glacial maxima, the ratio of precipitation in Greenland has changed by as much as 5 parts per thousand (0.001935–0.001925). While seeming small, such changes are detectable reliably with specialised mass spectrometric techniques. The small changes reflect natural fractionation processes that have left their signature in natural archives. These enable us to investigate the climate of past times in order to understand how the Earth’s climatic system works and how it can react to external forcing. In addition, studying contemporary isotopic change of natural compartments can help to identify sources and sinks for atmospheric trace gases provided the respective isotopic signatures are large enough for measurement and have not been obscured by unknown processes. This information is vital within the framework of the Kyoto process for controlling CO2 emissions.