Hostname: page-component-76fb5796d-vvkck Total loading time: 0 Render date: 2024-04-26T19:31:15.344Z Has data issue: false hasContentIssue false

The Radiocarbon Intracavity Optogalvanic Spectroscopy Setup at Uppsala

Published online by Cambridge University Press:  09 February 2016

Gerriet Eilers
Affiliation:
Dept. of Physics and Astronomy, Ion Physics, Box 516, SE-751 20, Uppsala, Sweden
Anders Persson
Affiliation:
Dept. of Physics and Astronomy, Ion Physics, Box 516, SE-751 20, Uppsala, Sweden
Cecilia Gustavsson
Affiliation:
Dept. of Physics and Astronomy, Ion Physics, Box 516, SE-751 20, Uppsala, Sweden
Linus Ryderfors
Affiliation:
Dept. of Physics and Astronomy, Ion Physics, Box 516, SE-751 20, Uppsala, Sweden
Emad Mukhtar
Affiliation:
Dept. of Chemistry, ångström Laboratory, Box 523, SE-751 20, Uppsala, Sweden
Göran Possnert
Affiliation:
Dept. of Physics and Astronomy, Ion Physics, Box 516, SE-751 20, Uppsala, Sweden
Mehran Salehpour*
Affiliation:
Dept. of Physics and Astronomy, Ion Physics, Box 516, SE-751 20, Uppsala, Sweden
*
3Corresponding author: mehran.salehpour@physics.uu.se.

Abstract

Accelerator mass spectrometry (AMS) is by far the predominant technology deployed for radiocarbon tracer studies. Applications are widespread from archaeology to biological, environmental, and pharmaceutical sciences. In spite of its excellent performance, AMS is expensive and complicated to operate. Consequently, alternative detection techniques for 14C are of great interest, with the vision of a compact, user-friendly, and inexpensive analytical method. Here, we report on the use of intracavity optogalvanic spectroscopy (ICOGS) for measurements of the 14C/12C ratio. This new detection technique was developed by Murnick et al. (2008). In the infrared (IR) region, CO2 molecules have strong absorption coefficients. The IR-absorption lines are narrow in line width and shifted for different carbon isotopes. These properties can potentially be exploited to detect 14CO2, 13CO2, or 12CO2 molecules unambiguously. In ICOGS, the sample is in the form of CO2 gas, eliminating the graphitization step that h is required in most AMS labs. The status of the ICOGS setup in Uppsala is presented. The system is operational but not yet fully developed. Data are presented for initial results that illustrate the dependence of the optogalvanic signal on various parameters, such as background and plasma-induced changes in the sample gas composition.

Type
Articles
Copyright
Copyright © 2013 by the Arizona Board of Regents on behalf of the University of Arizona 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Bradley, LC, Soohoo, KL, Freed, C. 1986. Absolute frequencies of lasing transitions in nine CO2 isotopic species. IEEE Journal of Quantum Electronics 22(2):234–67.Google Scholar
Chedin, A. 1975. The carbon dioxide molecule: potential, spectroscopic, and molecular constants from its infrared spectrum. Journal of Molecular Spectroscopy 76(1–3):430–91.Google Scholar
Deju, N, Kan, T, Wolga, GJ. 1968. Gain distribution, population densities, and rotational temperature for the (00°1)-(10°0) rotation-vibration transitions in a flowing CO,-N,-He laser. IEEE Journal of Quantum Electronics 4(5):256–60.Google Scholar
Demtröder, W. 1981. Laser Spectroscopy. Berlin: Springer-Verlag.Google Scholar
Duarte, FJ. 1995. Tunable Lasers Handbook. San Diego: Academic Press. 63 p.Google Scholar
Galli, I, Bartalini, S, Borri, S, Cancio, P, Mazzotti, D, De Natale, P, Giusfredi, G. 2011. Molecular gas sensing below parts per trillion: radiocarbon-dioxide optical detection. Physical Review Letters 107:270802, doi:10.1103/PhysRevLett.107.270802.Google Scholar
Hartmann, B, Kleman, B. 1966. Laser lines from CO2 in the 11 to 18 micron region. Canadian Journal of Physics 44:1609–12.Google Scholar
Hendriksen, BLM, Ackermann, MD, van Rijn, R, Stoltz, D, Popa, I, Balmes, O, Resta, A, Wermeille, D, Felici, R, Ferrer, S, Frenken, JWM. 2010. The role of steps in surface catalysis and reaction oscillations. Nature Chemistry 2:730–4.Google Scholar
Huang, X, Yung, YL. 2004. A common misunderstanding about the Voigt line profile. Journal of the Atmospheric Sciences 61(13):1630–2.Google Scholar
Ilkmen, E. 2009. Intracavity optogalvanic spectroscopy for radiocarbon analysis with attomole sensitivity , New Brunswick: Rutgers University.Google Scholar
Labrie, D, Reid, J. 1981. Radiocarbon dating by infrared laser spectroscopy. Applied Physics A 24(4):381–6.Google Scholar
Lappin, G, Garner, CR. 2003. Big physics, small doses: the use of AMS and PET in human microdosing of development drugs. Nature Reviews 2:233–40.Google Scholar
May, RD, May, PH. 1986. Solid-state radio frequency oscillator for optogalvanic spectroscopy: detection of nitric oxide using the 2–0 overtone transition. Review of Scientific Instruments 57(9):2242–5.Google Scholar
Meyer, TW, Rhodes, CK, Haus, HA. 1993. High-resolution line broadening and collisional studies in CO2 using nonlinear spectroscopic techniques. Physical Review A 12(5):1993–2008.Google Scholar
Murnick, DE, Peer, BJ. 1994. Laser-based analysis of carbon isotope ratios. Science 263(5149):945–7.Google Scholar
Murnick, DE, Dogru, O, Ilkmen, E. 2008. Intracavity optogalvanic spectroscopy. An analytical technique for 14C analysis with subattomole sensitivity. Analytical Chemistry 80(13):4820–4.Google Scholar
Salehpour, M, Possnert, G, Bryhni, H. 2008. Subattomole sensitivity in biological accelerator mass spectrometry. Analytical Chemistry 80(10):3515–21.Google Scholar
Silfvast, WT. 1996. Laser Fundamentals. Cambridge: Cambridge University Press.Google Scholar