Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Determination of N2O isotopomers with quantum cascade laser based absorption spectroscopy

Open Access Open Access

Abstract

We present an analytical technique based on direct absorption laser spectroscopy for high precision and simultaneous determination of the mixing ratios of the most abundant nitrous oxide isotopic species: 14N15N16O, 15N14N16O and 14N2 16O. A precision of 0.5‰ was achieved for the site specific isotope ratios of N2O at 90 ppm using an averaging time of 300 s.

©2008 Optical Society of America

1. Introduction

Nitrous oxide (N2O) is a stratospheric ozone depleting substance and exerts the fourth largest radiative forcing. The predominant sources of N2O on a global scale are microbial production in fertilized agricultural lands as well as biomass and fossil fuel burning. While the major sink, stratospheric destruction, is well quantified, the strength of N2O sources remains largely uncertain, due to the complexity of pathways involved [1].

Important information about the geochemical cycle of N2O can be obtained by measuring the intramolecular distribution of 15N within atmospheric N2O [2]. Being a linear, nonsymmetric molecule (N-N-O), with one nitrogen atom at the center (α site) and one at the end (β site), there are two isotopomers containing one heavy isotope of nitrogen, namely 14N15N16O and 15N14N16O, referred to as 15Nα and 15Nβ, respectively. The corresponding isotope ratios are usually reported in the δ-notation, where δ15N denotes the relative difference in per mil (‰) of the 15N/14N ratio of the sample versus the reference material (atmospheric N2), thus δ15Nβ for example denotes the site specific ratio of 15N14N16O vs. 14N14N16O.

Currently, the only analytical technique for N2O isotope measurements at ambient concentrations is the laboratory-based isotope-ratio mass-spectrometry (IRMS). Unfortunately, isotopomers such as 14N15N16O and 15N14N16O, cannot directly be distinguished by IRMS, as they have identical molecular mass. The determination of site-selective isotopic composition involves the complex analysis of the NO+ and N2O+ ions [3, 4] and the adopted methods lead to an unresolved discrepancy of almost 30 ‰ for the absolute intramolecular distribution of 15N in N2O. The difference seems partly be due to differences in calibration strategies [5]. These limitations might be avoided by the development of more direct and independent analytical approaches. Moreover, continuous measurements under field conditions would open a completely new field of applications for N2O isotopomer studies.

This motivates the development of an alternative analytical technique based on direct absorption laser spectroscopy. In contrast to IRMS, laser spectroscopy offers the inherent advantage of site selectivity combined with high sensitivity and time resolution. So far, the published spectroscopic measurements of N2O isotope ratios were mostly done in pure N2O and suffered from limited precision [6–8]. In this paper, we demonstrate the precise and simultaneous determination of the site specific isotope abundance ratios of N2O at concentrations as low as 9 ppm with a pulsed quantum cascade laser based spectrometer. This instrument is adequate to determine the isotopic signature of major biogenic and technical processes where the concentrations of the emitted N2O can increase up to several tens of ppm [9–10]. Furthermore, isotopic analysis at ambient concentrations would be accessible by coupling the laser spectrometer to a preconcentration unit.

2. Experimental

2.1 Instrumental setup and data analysis

Figure 1(a) shows the experimental setup employing a single-mode, pulsed quantum cascade laser (QCL) emitting at 4.6 µm, a multipass absorption cell and a detection system. The spectrometer is an improved version of an instrument built by Aerodyne Research Inc. [11–12]. Briefly, the laser (Alpes Lasers SA, CH) operates at quasi-room temperature with 10 ns pulses at a repetition rate of 1 MHz and it is frequency tuned with a sweep rate of 8 kHz by a sub-threshold current ramp that rapidly modulates its temperature. The tuning rate of the laser is determined using the fringe spectrum of a solid state Ge-etalon with a free spectral range of 0.0485 cm-1. The laser beam passes through a wedged BaF2 beamsplitter and is then coupled into the multipass cell with an optical path length of 56 m in a volume of 0.5 L (Aerodyne Research Inc., USA). The outcoming beam is focused on a thermoelectrically cooled detector (PDI-2TE-5, Vigo System SA, PL). The fraction of light that has been reflected by the beamsplitter is directly coupled into the same detector. Due to the large path length difference between reference and sample pulse, the former reaches the detector 150 ns earlier than the pulse exiting the multipass cell. Since the detector is fast enough (τ=60 ns) it can temporally resolve these two pulses. Normalizing the sample pulse with the reference pulse reduces the effect of pulse-to-pulse fluctuations and, therefore, significantly improves the signal-to-noise ratio [11].

The mixing ratios are obtained by simultaneously fitting a low-order polynomial to the spectral baseline and a Gaussian-convoluted Voigt profile to the observed absorption lines, taking into account the measured path length, gas temperature (~306 K), pressure (8 kPa) and laser line width (0.014 cm-1).

 figure: Fig. 1.

Fig. 1. (a). Schematic of the optical layout (QCL: quantum cascade laser, D: detector). (b) Measured (dots) and fitted spectrum (solid line) of the N2O isotopes (top) and the corresponding line strengths (bottom). The tail of a strong absorption line of 14N14N16O (2188.19 cm-1) is clearly visible and was considered in the fit up to 2188.1 cm-1.

Download Full Size | PDF

The selected absorption lines for this study are shown in Fig. 1(b). They are located within the scanning range of the laser and have sufficiently strong line strength with similar absorbance for each isotopic species. This, however, also implies considerable differences in the lower state energies (E”) of the selected absorption lines and results in a pronounced temperature sensitivity. With E”=1205.9 cm-1 the absorption line of the main isotope has a significantly different E” than the two other isotopomers 14N15N16O (65.4 cm-1) and 15N14N16O (110.1 cm-1). These differences translate into a temperature sensitivity of 17.3 ‰ K-1 and 16.6 ‰ K-1 for δ15Nα and δ15Nβ, respectively. To account for this, the gas temperature is controlled to 0.1 K and measured with a precision of one mK using a calibrated 10 kΩ thermistor in the absorption cell. As discussed in [8], these absorption lines are essentially interference-free at ambient concentrations. For field applications with large changes in H2O, CO or CO2 we would, nevertheless, consider using specific adsorbing or drying agents.

2.2 Preparation of isotopic standards

A set of gas mixtures with distinct N2O isotopomer compositions was prepared in two steps. First, isotopically pure (>98 %) 15N14N16O and 14N15N16O (Cambridge Isotope Laboratories, USA) were diluted with high purity synthetic air (>99.999 %, 20.5 % O2, Messer, CH) and their exact mixing ratio determined gravimetrically. The gas cylinders contained 359 ± 3 ppm (15Nα) and 353 ± 3 ppm (15Nβ) of N2O, respectively. Additionally, the isotopic purity of both tanks was determined by the quantification of 14N2 16O (<2 ppm) using laser spectroscopy. Secondly, in order to span the δ-scale, five calibration gases were produced in 50 L stainless steel cylinders. These were first filled with known volumes of medical N2O (>98 %, Messer, CH), supplemented with exact amounts of isotopic pure N2O (15Nα and/or 15Nβ) and diluted with a defined volume of high purity synthetic air (Table 1). The exact amounts of added N2O were determined using a high precision flow measurement device (molbox1, DH Instruments Inc., USA), and the dilution with synthetic air was controlled gravimetrically. Calibration gas concentrations were assigned using laser spectroscopy, referenced to a certified standard (92.0 ± 0.2 ppm N2O in synthetic air, Messer, CH). Similarly to the use of atmospheric N2 for IRMS, isotopomer specific ratios here are reported relative to medical N2O in synthetic air which is used as reference to calculate the δ-values from the isotope ratios.

Tables Icon

Table 1. N2O concentrations and δ-values of calibration gases with ±σ (δ reference point: medical N2O)

3. Results and discussion

3.1 Precision and long-term stability

The spectrometer’s performance regarding precision and stability was characterized using the Allan variance technique [13]. At a concentration of 90 ppm N2O, a short term precision (1 Hz) of 6.1 ‰ for the 15Nβ/14N ratio was obtained, while the Allan variance reached its minimum at an integration time of 300 s, corresponding to a precision of 0.46‰ (see Fig. 2). Similar results were obtained for the 15Nα/14N ratio (data not shown). The isotopomer concentrations reported here are weighted by the isotopic abundance as given in the HITRAN database [14]. The 15Nβ/15Nα ratio measurement had a comparable precision (0.67‰ @ 300 s), which indicates that temperature fluctuations do not limit instrumental precision. At 9 ppm, the precision was 5 to 10 times lower than at 90 ppm (see Fig. 2).

 figure: Fig. 2.

Fig. 2. Allan variance of the normalized isotope ratio 15Nβ/14N (see text) at 9 ppm (a) and 90 ppm (b) of N2O.

Download Full Size | PDF

The long-term stability of the instrument was assessed by performing an unattended measurement over several days while alternating every 10 minutes between sample gas (α50) and reference N2O (Ref). A small vacuum pump continuously drew the gas through the absorption cell, while a manually adjusted metering valve maintained a gas flow of 0.15 L/min. During these measurements, a slight drift in gas pressure (ΔP=0.3 kPa) and temperature (ΔT=0.1°C) was observed. The pressure dependence of the analytical procedure had previously been determined experimentally as 4‰/kPa. Hence, the changes in these parameters can not explain the shift of 80‰ observed in the derived isotope ratio illustrated in Fig 3(a). The drift was due to an unusually large change in laser intensity of 11 % during the measurement period. Moreover, the overall spectral range (0.43 cm-1) shifted by 4% towards higher wavenumbers. About 90 % of this took place in the first 15 hours after turning on the instrument, i.e. in a situation when thermal equilibrium of the whole system was not achieved yet. However, even in such extreme situations the average δ15Nβ of the sample gas could be determined with a standard deviation of 0.7‰ (see Fig. 3(b)). For δ15Nα a standard deviation of 0.5‰ was achieved, which is similar to the precision obtained by the Allan variance technique. This clearly demonstrates the necessity of a frequent calibration procedure.

 figure: Fig. 3.

Fig. 3. Repeated measurements of a gas cylinder (α50). (a) Time series of the 15Nβ/14N ratio and the drift corrected averaged δ15Nβ values (10 min) during 2.5 days. (b) After drift correction averaged δ15Nβ values (10 min) were plotted as a histogram with a 0.2‰ bin width. The distribution is nearly Gaussian with a standard deviation of 0.7‰ and a shift of 0.3‰ with respect to the expected δ-value.

Download Full Size | PDF

3.2 Calibration procedure

Direct absorption spectroscopy is an absolute method that, in principle, allows for a straightforward determination of the concentration, and hence the calculation of isotope ratios from the measured signals. In practice, however, the accuracy of this calculation depends on many parameters, such as the laser line width and shape, and the quality of the molecular line parameters. Additionally, instrumental nonlinearity and drifts will bias the retrieved isotope ratios. To account for this, the instrument needs to be equipped with a calibration system.

The retrieved isotope ratios showed an accentuated dependence on the N2O concentration, which is consistent with observations from measurements of CO2 isotopologues [15]. This dependence on total N2O was characterized by dynamically diluting various calibration gases with N2O-free synthetic air. The experimental results for δ15Nβ are well described by a linear interpolation. The very slight deviations in the slope values for individual dilution curves are probably due to instrumental drifts. Their average value of 0.28‰ ppm-1 is an excellent approximation to account for the dependence on total N2O for all δ-values, as shown in Fig. 4(a). Based on this linear interpolation, the expected δ-values of the calibration gases were plotted against the spectroscopically measured isotope ratios which were normalized to an arbitrarily selected N2O concentration (60 ppm in this case). This yielded a calibration function that links the measured isotope ratio values to the δ-scale (Fig. 4(b)). The linearity and potential accuracy of the instrument can be assessed from the residual of the linear interpolation, which has a 1σ standard deviation of 0.34‰. Given the linear behavior depicted in Fig. 4, the entire calibration of the spectrometer can in principle be performed by single-point measurements of three adequate reference gases. Two are needed to determine the dependency of isotope ratios on the N2O concentration and a third one to define the calibration function that links the normalized isotope ratio values to the δ-scale. The same procedure can be applied to δ15Nα.

 figure: Fig. 4.

Fig. 4. (a). The isotopomer specific 15Nβ/14N ratio measured for different N2O concentrations. (b) The linear relationship between δ15Nβ of the calibration gases and the interpolated 15Nβ/14N ratios determined by laser spectroscopy.

Download Full Size | PDF

4. Conclusions and outlook

A versatile analytical method for high precision, continuous and simultaneous determination of the site specific isotope abundance ratios of N2O at trace level is presented. In contrast to IRMS, this optical technique offers a direct way to investigate the intramolecular isotope distribution of nitrogen without any sample pretreatment. The instrument achieves a precision of 0.5‰ with 300 s averaging time for 90 ppm of N2O which is more than one order of magnitude better than previously reported values. Measurements of various, isotopically enriched gas samples showed a linear instrumental response to changing δ-values. This allows for a straightforward calibration of the system.

The precision and sensitivity of the current spectrometer is adequate for example to differentiate between the most important biogenic N2O emitting processes, nitrification and denitrification, where differences in the site preference (δ15Nα - δ15Nβ) exceed 30‰ and concentrations are in the ppm-range [16]. Other field applications, such as source characterization of anthropogenic N2O emissions, are also directly accessible. For atmospheric applications, further improvements are needed. A significant step forward would be the use of a high power, continuous wave QCL that should soon be available. Further enhancements may be obtained by the latest generation of TE-cooled detectors or the implementation of a reference cell as demonstrated for CO2 isotope measurements [16]. At the current state of laser spectroscopy, the most promising way to reach excellent precision at ambient N2O concentration may be the addition of an automated preconcentration unit.

Acknowledgments

The authors thank Alpes Lasers for providing the laser source at the required wavelength. M. Zahniser, D. Nelson and B. McManus are acknowledged for their continuous support. We acknowledge the criticisms and helpful suggestions provided by the anonymous reviewers.

References and links

1. IPCC, Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the IPCC (Cambridge University Press, Cambridge and New York, 2007).

2. L.Y. Stein and Y. L. Yung, “Production, isotopic composition, and atmospheric fate of biologically produced nitrous oxide,” Annu. Rev. Earth Planet Sci. 31, 329–356 (2003). [CrossRef]  

3. T. Röckmann, J. Kaiser, C. A. M. Brenninkmeijer, and W. A. Brand, “Gas chromatography/isotope-ratio mass-spectrometry method for high-precision position-dependent 15N and 18O measurement of atmospheric nitrous oxide,” Rapid Commun. Mass Spectrom. 17, 1897–1908 (2003). [CrossRef]   [PubMed]  

4. S. Toyoda and N. Yoshida, “Determination of nitrogen isotopomers of nitrous oxide on a modified isotope ratio mass spectrometer,” Anal. Chem. 71, 4711–4718 (1999). [CrossRef]  

5. M. B. Westley, B. N. Popp, and T. M. Rust, “The calibration of the intramolecular nitrogen isotope distribution in N2O measured by isotope ratio mass spectrometry,” Rapid Commun. Mass Spectrom. 21, 391–405 (2007). [CrossRef]   [PubMed]  

6. K. Uehara, K. Yamamoto, T. Kikugawa, and N. Yoshida, “Site-selective isotopic ratio measurement of nitrous oxide using 2 µm diode lasers,” Spectrochim. Acta A 59, 957–962 (2003). [CrossRef]  

7. G. Gagliardi, S. Borri, F. Tamassia, F. Capasso, C. Gmachl, D. L. Sivco, J. N. Baillargeon, A. L. Hutchinson, and A. Y. Cho, “A frequency-modulated quantum cascade laser for spectroscopy of CH4 and N2O isotopomers,” Isotopes Environ. Health Stud. 41, 313–321 (2005). [CrossRef]  

8. H. Waechter and M. W. Sigrist, “Mid-infrared laser spectroscopic determination of isotope ratios of N2O at trace levels using wavelength modulation and balanced path length detection,” Appl. Phys. B 87, 539–546 (2007). [CrossRef]  

9. G. Drewitt and J. S. Warland, “Continuous measurements of belowground nitrous oxide concentrations,” Soil Sci. Soc. Am. J. 71, 1–7 (2007). [CrossRef]  

10. J. A. Martín, M. Yates, P. Ávila, S. Suárez, and J. Blanco, “N2O formation in low temperature selective catalytic reduction of nitrogen oxides with V2O5/TiO2 catalysts,” Appl. Catal. B Environ. 70, 330–334 (2007). [CrossRef]  

11. D. D. Nelson, J. H. Shorter, J. B. McManus, and M. S. Zahniser, “Sub-part-per-billion detection of nitric oxide in air using a thermoelectrically cooled mid-infrared quantum cascade laser spectrometer,” Appl. Phys. B 75, 343–350 (2002). [CrossRef]  

12. B. Tuzson, M. J. Zeeman, M. S. Zahniser, and L. Emmenegger, “Quantum cascade laser based spectrometer for in situ stable carbon dioxide isotope measurements,” Infra. Phys. Technol. 51, 198–206 (2008). [CrossRef]  

13. P. Werle, R. Mucke, and F. Slemr, “The limits of signal averaging in atmospheric trace-gas monitoring by tunable diode-laser absorption spectroscopy (TDLAS),” Appl. Phys. B 57, 131–139 (1993). [CrossRef]  

14. L. S. Rothman, D. Jacquemarta, and A. Barbe, et al. “The HITRAN 2004 molecular spectroscopic database,” J. Quantum Spectrosc. Radiat. Transfer96, 139–204 (2005).

15. B. Tuzson, J. Mohn, M. J. Zeeman, R. A. Werner, W. Eugster, M. S. Zahniser, D. D. Nelson, J. B. McManus, and L. Emmenegger, “High precision and continuous field measurements of δ13C and δ18O in carbon dioxide with a cryogenic free QCLAS,” Appl. Phys. B (to be published).

16. R. L. Sutka, N. E. Ostrom, P. H. Ostrom, J. A. Breznak, H. Gandhi, A. J. Pitt, and F. Li, “Distinguishing nitrous oxide production from nitrification and denitrification on the basis of isotopomer abundances,” Appl. Environ. Microbiol. 72, 638–644 (2006). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1.
Fig. 1. (a). Schematic of the optical layout (QCL: quantum cascade laser, D: detector). (b) Measured (dots) and fitted spectrum (solid line) of the N2O isotopes (top) and the corresponding line strengths (bottom). The tail of a strong absorption line of 14N14N16O (2188.19 cm-1) is clearly visible and was considered in the fit up to 2188.1 cm-1.
Fig. 2.
Fig. 2. Allan variance of the normalized isotope ratio 15Nβ/14N (see text) at 9 ppm (a) and 90 ppm (b) of N2O.
Fig. 3.
Fig. 3. Repeated measurements of a gas cylinder (α50). (a) Time series of the 15Nβ/14N ratio and the drift corrected averaged δ15Nβ values (10 min) during 2.5 days. (b) After drift correction averaged δ15Nβ values (10 min) were plotted as a histogram with a 0.2‰ bin width. The distribution is nearly Gaussian with a standard deviation of 0.7‰ and a shift of 0.3‰ with respect to the expected δ-value.
Fig. 4.
Fig. 4. (a). The isotopomer specific 15Nβ/14N ratio measured for different N2O concentrations. (b) The linear relationship between δ15Nβ of the calibration gases and the interpolated 15Nβ/14N ratios determined by laser spectroscopy.

Tables (1)

Tables Icon

Table 1. N2O concentrations and δ-values of calibration gases with ±σ (δ reference point: medical N2O)

Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.