Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-06-06T22:00:31.659Z Has data issue: false hasContentIssue false

7 - Lidars for Profiling Aerosol Optical Properties, Atmospheric Temperature and Wind

Published online by Cambridge University Press:  24 February 2022

Chiao-Yao She
Affiliation:
Colorado State University
Jonathan S. Friedman
Affiliation:
Universidad Ana G. Mendez
Get access

Summary

In Chapter 7, we present lidars for profiling atmospheric parameters such as aerosol optical properties, temperatures, and winds. We start with a description of methods for profiling the lidar aerosol-molecular ratio and determining aerosol optical properties. We present and compare techniques for these measurements, including Rayleigh and vibrational Raman integration, rotational Raman technique, and the use of multiple receiver channels with custom-built interference filters. We follow with a description of high-spectral resolution lidar (HSRL), including a detailed discussion of notch (atomic or molecular vapor) filters in the Cabannes scattering detection channels and what is required to make an “ideal” or near-ideal filter. From there, we describe wind profiling methods using Cabannes–Mie scattering, comparing coherent versus incoherent lidars. For HSRL, we describe a near-ideal filter based on absorption in potassium vapor at 770 nm. Following parameter profiling with Cabannes–Mie lidar, we close by describing temperature and wind profiling with laser-induced fluorescence. Here, we focus on Na and Fe lidars. We give summaries of daylight measurements and data processing algorithms, including uncertainty in measurements. We close by discussing scientific contributions by and challenges for these lidars.

Type
Chapter
Information
Atmospheric Lidar Fundamentals
Laser Light Scattering from Atoms and Linear Molecules
, pp. 138 - 243
Publisher: Cambridge University Press
Print publication year: 2022

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

Hair, J. W., Hostetler, C. A., Cook, A. L. et al. (2008). Airborne high spectral resolution lidar for profiling aerosol optical properties. Appl. Opt. 47(36), 67346753.CrossRefGoogle ScholarPubMed
Krueger, D. A., She, C.-Y., and Yuan, T.. (2015). Retrieving mesopause temperature and line-of-sight wind from full-diurnal-cycle Na lidar observations. Appl. Opt. 54(32), 94699489.Google Scholar
Hansch, T. W., Shahin, I. S., and Schawlow, A. L.. (1971). High resolution saturation spectroscopy of the sodium D lines with a pulsed tunable dye laser. Phys. Rev. Lett. 27(11), 707710.CrossRefGoogle Scholar
Kuchta, E., Alvarez, II R. J., Li, Y. H., Krueger, D. A., and She, C. Y.. (1990). Collisional broadening of Bal line (553.5 nm) by He or Ar. Appl. Phys. B, 50(2), 129132.CrossRefGoogle Scholar
Arie, A., Schiller, S., Gustafson, E. K., and Byer, R. L.. (1992). Absolute frequency stabilization of diode-laser-pumped Nd:YAG lasers to hyperfine transitions in molecular iodine. Opt. Lett. 17(17), 12041206.Google Scholar
Arie, A. and Byer, R. L.. (1993). Frequency stabilization of the 1064-nm Nd: YAG lasers to Doppler-broadened lines of iodine. Appl. Opt. 32(36), 73827386.Google Scholar
Fiedler, J., and von Cossart, G.. (1999). Automated lidar transmitter for multiparameter investigations within the Arctic atmosphere. IEEE Trans. Geosci. Remote Sensing, 37(2), 748755.Google Scholar
She, C. Y. and Yu, J. R.. (1995). Doppler-free saturation fluorescence spectroscopy of Na atoms for atmospheric applications. Appl. Opt., 34(6), 10631075.CrossRefGoogle Scholar
Fiocco, G., Beneditti-Michelangeli, G., Maischberger, K., and Madonna, E.. (1971). Measurement of temperature and aerosol to molecule ratio in the troposphere by optical radar. Nature, Phys. Sci., 229(3), 7879.CrossRefGoogle Scholar
Schwiesow, R. L. and Lading, L.. (1981). Temperature profiling by Rayleigh-scattering lidar. Appl. Opt., 20(11), 19721979.CrossRefGoogle ScholarPubMed
Shimizu, H., Lee, S. A., and She, C. Y.. (1983). High spectral resolution lidar system with atomic blocking filters for measuring atmospheric parameters. Appl. Opt., 22(9), 13731381.CrossRefGoogle ScholarPubMed
She, C. Y., Alvarez II, R. J., Caldwell, L. M, and Krueger, D. A. (1992). High-spectral-resolution RayleighMie lidar measurement of aerosol and atmospheric profiles. Opt. Lett., 17(7), 541543.CrossRefGoogle ScholarPubMed
Hair, J. W., Caldwell, L. M., Krueger, D. A., and She, C.-Y.. (2001). High-spectral-resolution lidar with iodine-vapor filters: measurement of atmospheric-state and aerosol profiles. Appl. Opt., 40(30), 52805294.CrossRefGoogle ScholarPubMed
Ansmann, A., Riebesell, M., and Weitkamp, C.. (1990). Measurement of atmospheric aerosol extinction profiles with a Raman lidar. Opt. Lett., 15(13), 746748.Google Scholar
Eloranta, E. (2005). High spectral resolution lidar. Chapter 5 in Lidar Range-Resolved Optical Remote Sensing of the Atmosphere, Weitkamp, C., ed., Springer.Google Scholar
Ansmann, A., Wandinger, U., Riebesell, M., Weitkamp, C., and Michaelis, W.. (1992). Independent measurement of extinction and backscatter profiles in cirrus clouds by using a combined Raman elastic-backscatter lidar. Appl. Opt. 31(33), 71137131.Google Scholar
Ansmann, A. and Müller, D. (2005). Lidar and atmospheric aerosol particles. Chapter 4 in Lidar Range-Resolved Optical Remote Sensing of the Atmosphere, Weitkamp, C., ed., Springer.Google Scholar
Shipley, S. T., Tracy, D. H., Eloranta, E. W. et al. (1983). High spectral resolution lidar to measure optical scattering properties of atmospheric aerosols. 1: Theory and instrumentation. Appl. Opt., 22(23), 37163724.Google Scholar
Alvarez II, R. J., Caldwell, L. M., Li, Y. H., Krueger, D. A., and She, C. Y.. (1990). High spectral resolution lidar measurement of tropospheric backscatter-ratio with barium atomic blocking filters. J. Atmos. Oceanic Technol., 7(6), 876881.Google Scholar
Piironen, P. and Eloranta, E. W.. (1994). Demonstration of a high-spectral-resolution lidar based on an iodine absorption filter. Opt. Lett., 19(3), 234236.Google Scholar
Gerstenkorn, S. and Luc, P.. (1978). Atlas du Spectre d’Absorption de la Molecule d’Iode 14800–20000 cm−1. Paris: Editions du Centre National de la Recherche Scientifique (CNRS).Google Scholar
Forkey, J. N., Lempert, W. R., and Miles, R. B. (1997). Corrected and calibrated I2 absorption model at frequency-doubled Nd:YAG laser wavelengths, Appl. Opt., 36(27), 67296738.CrossRefGoogle ScholarPubMed
Zhang, Y.-P., Liu, D., Shen, X. et al. (2017). Design of iodine absorption cell for high-spectral-resolution lidar. Opt. Ex., 25(14), 1591315926.Google Scholar
Eloranta, E., Razenkov, I., and Garcia., J. (2018). HSRL measurements of Lidar ratios in the presence of oriented ice crystals. Abstract. American Meteorological Society website. Accessed at https://ams.confex.com/ams/2019Annual/webprogram/Paper351863.htmlGoogle Scholar
Eloranta, E. (1998). Practical model for the calculation of multiply scattered lidar returns. Appl. Opt. 37(12), 24642472.Google Scholar
Elterman, L. (1954). Seasonal trends of temperature, density, and pressure to 66 km obtained with the searchlight probing technique. Jour. Geophys. Res. 59(3), 351358.CrossRefGoogle Scholar
Hauchecorne, A., and Chanin, M. L.. (1980). Density and temperature profiles obtained by lidar between 35 and 70 km, Geophys. Res. Letters 7(8), 565568.Google Scholar
Chanin, M.-L., and Hauchecorne, A.. (1981). Lidar observation of gravity and tidal waves in the stratosphere and mesosphere, J. Geophys. Res. 86(C10), 97159721.CrossRefGoogle Scholar
Nedeljkovic, D., Hauchecorne, A., and Chanin, M. L.. (1993). Rotational Raman lidar to measure the atmospheric temperature from the ground to 30 km, IEEE Trans. Geosci. Remote Sens. 31(1), 90101.Google Scholar
Behrendt, A. and Reichardt, J.. (2000). Atmospheric temperature profiling in the presence of clouds with a pure rotational Raman lidar by use of an interference-filter-based polychromator. Appl. Opt. 39(9), 13721378.CrossRefGoogle ScholarPubMed
Behrendt, A., Nakamura, T., and Tsuda, T.. (2004). Combined temperature lidar for measurements in the troposphere, stratosphere, and mesosphere. Appl. Opt. 43(14), 29302939.Google Scholar
Arshinov, Yu. F., Bobrovnikov, S. M, Zuev, V. E, and Mitev, V. M. (1983). Atmospheric temperature measurement using a pure rotational Raman lidar. Appl. Opt. 22(19), 29842990.Google Scholar
Cooney, J. A. (1984). Atmospheric temperature measurement using a pure rotational Raman lidar: comment. Appl. Opt. 23(5), 653654.Google Scholar
Alvarez, R. J., II. (1991). Measurement of tropospheric temperature and aerosol extinction using high-spectral-resolution lidar. Ph.D. thesis, Colorado State University. University Microfilms International, Ann Arbor, MI.Google Scholar
Krueger, D. A., Caldwell, L. M., Alvarez II, R. J., and She, C. Y. (1993). Self-consistent method for determining vertical profiles of aerosol and atmospheric properties using high-spectral-resolution Rayleigh–Mie lidar. J. Atm. Oceanic Tech. 10(4), 533545.Google Scholar
Alvarez II, R. J., Caldwell, L. M, Wolyn, P. G et al. (1993). Profiling temperature, pressure, and aerosol properties using a high spectral resolution lidar employing atomic blocking filters. J. Atm. Oceanic Tech. 10(4), 546556.Google Scholar
She, C.-Y., Yue, J., Yan, Z.-A. et al. (2007). Direct-detection Doppler wind measurements with a CabannesMie lidar: A. Comparison between iodine vapor filter and FabryPerot interferometer methods. Appl. Opt., 46(20), 44344443.CrossRefGoogle ScholarPubMed
Tellinghuisen, J. (1973). Resolution of the visible-infrared absorption spectrum of I2 into three contributing transitions. J. Chem. Phys. 58(7), 28212834.CrossRefGoogle Scholar
Forkey, J. N. (1996). Development and demonstration of filtered Rayleigh scattering – a laser based flow diagnostic for planar measurement of velocity, temperature and pressure. Ph.D. dissertation, Department of Mechanical and Aerospace Engineering, Princeton University.Google Scholar
Hair, J. W. (1998). A high spectral resolution lidar at 532 nm for simultaneous measurement of atmospheric state and aerosol profiles using iodine vapor filters. Ph.D. dissertation, Department of Physics, Colorado State University.Google Scholar
Hua, D.-X., Uchida, M., and Kobayashi, T.. (2004). Ultraviolet high-spectral-resolution Rayleigh–Mie lidar with a dual-pass Fabry–Perot etalon for measuring atmospheric temperature profiles of the troposphere. Opt. Lett. 29(10), 10631065.Google Scholar
Kawahara, T. D., Nozawa, S., Saito, N. et al. (2017). Sodium temperature/wind lidar based on laser-diode-pumped Nd:YAG lasers deployed at Tromsø, Norway (69.6°N, 19.2°E). Opt. Express 25(12), A491A501.Google Scholar
Munk, A., Jungbluth, B., Strotkamp, M. et al. (2018). Diode-pumped alexandrite ring laser in single-longitudinal mode operation for atmospheric lidar measurements. Opt. Express 26(12), 14,928–14,935.Google Scholar
Harrell, S. D., She, C.-Y., Yuan, T. et al. (2009). Sodium and potassium vapor Faraday filters revisited: Theory and applications. J. Opt. Soc. Amer. B, 26(4), 659670.Google Scholar
Voss, E., and Weitkamp, C.. (1992). Investigations on atomic-vapor filter high-spectral-resolution lidar for temperature measurements. Proc. 16th. Int. Laser Radar Conf., Cambridge, MA, NASA Conf. Pub. 3158, Part 2, 699702.Google Scholar
She, C.-Y., Krueger, D. A., Yan, Z.-A., and Hu, X.. (2021). Atomic vapor filter revisited: a Cabannes scattering temperature/wind lidar at 770 nm. Opt. Express 29(3), 43384362.Google Scholar
Werner, C. (2005). Doppler wind lidar. Chapter 12 in Lidar Range-Resolved Optical Remote Sensing of the Atmosphere, Weitkamp, C., ed., Springer.Google Scholar
Post, M. J. and Cupp, R. E.. (1990). Optimizing a pulsed Doppler lidar. Appl. Opt. 29(28), 41454158.Google Scholar
Kavaya, M. J., Beyon, J. Y., Koch, G. J. et al. (2014). The Doppler Aerosol Wind (DAWN) airborne, wind-profiling coherent-detection lidar system: overview and preliminary flight results. J. Atm. Oceanic Tech. 31(4), 826842. doi: https://doi.org/10.1175/JTECH-D-12-00274.1.Google Scholar
Yu, J., Singh, U., Barnes, N., and Petros, M.. (1998). 125-mJ diode-pumped injection-seeded Ho:Tm:YLF laser. Opt. Lett., 23(10), 780782. doi: https://doi.org/10.1364/OL.23.000780.Google Scholar
Henderson, S. W., Gatt, P., Rees, D., and Huffaker, R. M.. (2005). Wind lidar. Chapter 7 in Laser Remote Sensing, Fujii, T., & Fukuchi, T, eds., CRC Press, Taylor and Francis Group.Google Scholar
Kingston, R. H. (1995). Optical Sources, Detectors, and Systems. Academic Press. doi: https://doi.org/10.1016/B978-0-12-408655-5.X5000-8.Google Scholar
Goodman, J. W. (1985). Statistical Optics. John Wiley and Sons, New York.Google Scholar
Henderson, S. W. (2013). Review of Fundamental Characteristics of Coherent and Direct Detection Doppler Receivers and Implications to Wind Lidar System Design, Proc. 17th Coherent Laser Radar Conference (CLRC 2013), 4549, ISBN: 9781629931494.Google Scholar
She, C. Y., Yu, J. R., Latifi, H., and Bills, R. E.. (1992). High-spectral-resolution fluorescence light detection and ranging for mesospheric sodium temperature measurements. Appl. Opt., 31(12), 20952106.Google Scholar
Chanin, M. L., Garnier, A., Hauchecorne, A., and Porteneuve, J.. (1989). A Doppler lidar for measuring winds in the middle atmosphere. Geophys. Res. Lett. 16(11), 12731276, doi: https://doi.org/10.1029/GL016i011p01273.Google Scholar
Korb, C. L., Gentry, B., and Weng, C.. (1992). Edge technique: Theory and application to the lidar measurement of atmospheric wind. Appl. Opt. 31(21), 42024213.Google Scholar
Korb, C. L., Gentry, B. M., Li, S. X., and Flesia., C. (1998). Theory of the double-edge technique for Doppler lidar wind measurement. Appl. Opt. 37(15), 30973104.Google Scholar
Flesia, C., and Korb., C. L. (1999). Theory of the double-edge molecular technique for Doppler lidar wind measurement. Appl. Opt. 38(3), 432440.Google Scholar
Korb, C. L., Gentry, B., and Li, X.. (1997). Edge technique Doppler lidar wind measurements with high vertical resolution. Appl. Opt. 36(24), 59765983.Google Scholar
Gentry, B.M., Chen, H., and Li, S. X.. (2000). Wind measurements with 355 nm molecular Doppler lidar. Opt. Lett. 25(17), 12311233.Google Scholar
Liu, Z. S., Chen, W. B., Zhang, T. L., Hair, J. W., and She, C. Y.. (1996). An incoherent Doppler lidar for ground-based atmospheric wind profiling. Appl. Phys. B 64(5), 561566.Google Scholar
Friedman, J. S., Tepley, C. A., Castleberg, P. A., and Roe, H.. (1997). Middle-atmospheric Doppler lidar using an iodine-vapor edge filter. Opt. Lett. 22(21), 16481650.Google Scholar
Liu, Z.-S., Wu, D., Liu, J.-T. et al. (2002). Low-altitude atmospheric wind measurement from the combined Mie and Rayleigh backscattering by Doppler lidar with an iodine filter. Appl. Opt. 41(33), 70797086.Google Scholar
Wang, Z., Liu, Z., Liu, L. et al. (2010). Iodine-filter-based mobile Doppler lidar to make continuous and full-azimuth-scanned measurements: Data acquisition and analysis system, data retrieval methods, and error analysis. Appl. Opt., 49(36), 69606978.Google Scholar
Yan, Z., Hu, X., Guo, W. et al. (2017). Development of a mobile Doppler lidar system for wind and temperature measurements at 30–70 km. J. Quant. Spectrosc. Radiat. Transf., 188, 5259.CrossRefGoogle Scholar
Baumgarten, G. (2010). Doppler Rayleigh/Mie/Raman lidar for wind and temperature measurements in the middle atmosphere up to 80 km. Atmos. Meas. Tech., 3(6), 15091518, doi: https://doi.org/10.5194/amt-3-1509-2010.Google Scholar
von Zahn, U., von Cossart, G., Fiedler, J. et al. (2000). The ALOMAR Rayleigh/Mie/Raman lidar: Objectives, configuration, and performance. Ann. Geophys. 18(7), 815833.CrossRefGoogle Scholar
Hecht, E. (1998). Optics. 3rd ed., Addison-Wesley, pp. 413417.Google Scholar
Huang, W., Chu, X., Williams, B. P. et al. (2009). Na double-edge magneto-optic filter for Na lidar profiling of wind and temperature in the lower atmosphere. Opt. Lett., 34(2), 199201.Google Scholar
Huang, W., Chu, X., Wiig, J. et al. (2009). Field demonstration of simultaneous wind and temperature measurements from 5 to 50 km with a Na double-edge magneto-optic filter in a multi-frequency Doppler lidar. Opt. Lett., 34(10), 15521554.Google Scholar
Witschas, B., Lemmerz, C., and Reitebuch, O.. (2014). Daytime measurements of atmospheric temperature profiles (2–15 km) by lidar utilizing Rayleigh–Brillouin scattering. Opt. Lett., 39(7), 19721975.CrossRefGoogle ScholarPubMed
Xia, H., Dou, X., Sun, D. et al. (2012). Mid-altitude wind measurements with mobile Rayleigh Doppler lidar incorporating system-level optical frequency control method. Opt. Express, 20(14), 1528615300.Google Scholar
Xia, H., Dou, X., Shangguan, M. et al. (2014). Stratospheric temperature measurement with scanning Fabry–Perot interferometer for wind retrieval from mobile Rayleigh Doppler lidar. Opt. Express, 22(18), 2177521789. doi: https://doi.org/10.1364/OE.22.021775.Google Scholar
She, C.-Y. (2005). On atmospheric lidar performance comparison: from power aperture product to power aperture mixing ratio scattering cross-section product. Modern Optics, 52(18), 27232729, doi: https://doi.org/10.1080/09500340500352618.Google Scholar
Gardner, C. S. (2004). Performance capabilities of middle-atmosphere temperature lidars: comparison of Na, Fe, K, Ca, Ca+, and Rayleigh systems. Appl. Opt., 43(25), 49414956.Google Scholar
Gibson, A. J., Thomas, L., and Bhattachacharyya, S. K.. (1979). Laser observation of the ground-state hyperfine structure of sodium and of temperature in the upper atmosphere. Nature 281(5727), 131132.Google Scholar
Fricke, K. H. and von Zahn, U.. (1985). Mesopause temperature derived from probing the hyperfine structure of the D2 resonance line of sodium by lidar. J. Atmos. Terr. Phys. 47(5), 499512.Google Scholar
She, C. Y., Latifi, H., Yu, J. R. et al. (1990). Two frequency lidar techniques for mesospheric Na temperature measurements. Geophys. Res. Lett. 17(7), 929932.Google Scholar
She, C. Y. and Yu, J. R.. (1994). Simultaneous three-frequency Na lidar measurements of radial wind and temperature in the mesopause region. Geophys. Res. Lett. 21(17), 17711774.Google Scholar
Chu, X., Pan, W., Papen, G. C., Gardner, C. S., and Gelbwachs, J. A.. (2002). Fe Boltzmann temperature lidar: design, error analysis and initial results at the North and South Poles. Appl. Opt. 41(21), 44004410.Google Scholar
Höffner, J., and Lautenbach, J.. (2009). Daylight measurements of mesopause temperature and vertical wind with the mobile scanning iron lidar. Opt. Lett. 34(9), 13511353.Google Scholar
Chu, X., Huang, W., Thayer, J. P., Wang, Z., and Smith, J. A.. (2010). Progress in MRI Fe-Resonance/Rayleigh/Mie Doppler Lidar. In Proceedings of the 25th International Laser Radar Conference, Saint Petersburg, Russia, 947950.Google Scholar
Papen, G. C., Pfenninger, W. M., and Simonich, D. M.. (1995). Sensitivity analysis of Na narrowband wind-temperature lidar systems. Appl. Opt. 34(3), 480498.Google Scholar
Chen, H., White, M. A., Krueger, D. A., and She, C. Y.. (1996). Daytime mesopause temperature measurements using a sodium-vapor dispersive Faraday filter in lidar receiver. Opt. Lett. 21(15), 10931095.CrossRefGoogle ScholarPubMed
Chen, H., She, C. Y., and Korevaar, E.. (1993). Sodium vapor dispersive Faraday filter. Opt. Lett. 18(12), 10191021.Google Scholar
Fricke-Begemann, C., Alpers, M., and Höffner, J.. (2002). Daylight rejection with a new receiver for potassium resonance temperature lidars. Opt. Lett. 27(21), 19321934.Google Scholar
Sox, L., Wickwar, V. B., Yuan, T., and Criddle, N. R.. (2018). Simultaneous Rayleigh-scatter and sodium resonance lidar temperature comparisons in the mesosphere-lower thermosphere. J. Geophys. Res. Atmos., 123(18), 10,688–10,706, doi: https://doi.org/10.1029/2018jd029438.CrossRefGoogle Scholar
Clemesha, B. R., Simonich, D. M., Batista, P. P., Vondrak, T., and Plane, J. M. C. (2004). Negligible long‐term temperature trend in the upper atmosphere at 23°S. J. Geophys. Res., 109(D5), D05302, https://doi.org/10.1029/2003JD004243.Google Scholar
She, C.-Y., Berger, U, Yan, Z.-A et al. (2019). Solar response and long‐term trend of midlatitude mesopause region temperature based on 28 years (1990–2017) of Na lidar observations. J. Geophys. Res., 124(8), 71407156, https://doi.org/10.1029/2019JA026759.Google Scholar
Li, T., Fang, X., Liu, W., Gu, S.-Y., and Dou, X.. (2012). Narrowband sodium lidar for the measurements of mesopause region temperature and wind. Appl. Opt., 51(22), 54015411, doi: https://doi.org/10.1364/AO.51.005401.Google Scholar
Xia, Y., Du, L.-F., Cheng, X.-W. et al. (2017). Development of a solid-state sodium Doppler lidar using an all-fiber-coupled injection seeding unit for simultaneous temperature and wind measurements in the mesopause region. Opt. Express, 25(5), 52645278, doi: https://doi.org/10.1364/OE.25.005264.Google Scholar
von Zahn, U., and Hőffner, J.. (1996). Mesopause temperature profiling by potassium lidar. Geophys. Res. Lett., 23(2), 141144, doi: https://doi.org/10.1029/95GL03688.Google Scholar
Lübken, F.-J., Lautenbach, J., Hőffner, J., Rapp, M., and Zecha, M.. (2009). First continuous temperature measurements within polar mesosphere summer echoes. J. Atmos. Sol.-Terr. Phys., 71(3–4), 453463, doi: https://doi.org/10.1016/j.jastp.2008.06.001.Google Scholar
Friedman, J. S., and Chu, X.. (2007). Nocturnal temperature structure in the mesopause region over the Arecibo Observatory (18.35°N, 66.75°W): Seasonal variations. J. Geophys. Res., 112(D14), D14107, doi: https://doi.org/10.1029/2006JD008220.Google Scholar
She, C. Y. (1994). New lidar reveals seasonal temperature variations in a midlatitude mesopause region. Optics & Photonics News, December issue, pp. 2324.Google Scholar
She, C. Y., Chen, S. S., Hu, Z. L. et al. (2000). Eight-year climatology of nocturnal temperature and sodium density in the mesopause region (80 to 105 km) over Fort Collins, CO (41°N, 105°W). Geophys. Res. Lett., 27(20), 32893292.CrossRefGoogle Scholar
Yuan, T., She, C.-Y., Krueger, D. A. et al. (2008). Climatology of mesopause region temperature, zonal wind, and meridional wind over Fort Collins, Colorado (41°N, 105°W), and comparison with model simulations. J. Geophys. Res., 113(D3), D03105, doi: https://doi.org/0.1029/2007JD008697.Google Scholar
She, C.-Y. and Krueger, D. A.. (2007). Laser-induced fluorescence: Spectroscopy in the sky. Optics & Photonic News, 18(9), 3541.Google Scholar
She, C. Y., Liu, A. Z., Yuan, T. et al. (2021). MLT science enabled by atmospheric lidars. In Upper Atmospheric Dynamics and Energetics, Space Physics and Aeronomy Collection, vol. 4, ed. Wang, Wenbin, Zhang, Yongliang, Paxton, Larry J., chap. 20. John Wiley and Sons. doi: https://doi.org/10.1002.9781119507512.ch20.Google Scholar
Lübken, F.-J. and Höffner, J. (2021). VAHCOLI, a new concept for lidars: technical setup, science applications, and first measurements. Atmos. Meas. Tech., 14(5), 38153836. https://doi.org/10.5194/amt-14-3815-2021.Google Scholar
Vance, J. D., She, C. Y., and Moosmüller, H.. (1998). Continuous-wave, all-solid-state, single-frequency 400-mW source at 589 nm based on doubly resonant sum-frequency mixing in a monolithic lithium niobate resonator. Appl. Opt. 37(21), 48914896.Google Scholar
Kaifler, B., Büdenbender, C., Mahnke, P. et al. (2017). Demonstration of an iron fluorescence lidar operating at 372 nm wavelength using a newly developed Nd: YAG laser. Opt. Lett., 42(15), 28582861. doi: https://doi.org/10.1364/OL.42.002858.Google Scholar
Fu, S., Shi, W., Feng, Y., et al. (2017). Review of recent progress on single‐frequency fiber lasers [Invited]. J. Opt. Soc. Am. B, 34(3), A49A62. doi: https://doi.org/10.1364/JOSAB.34.000A49.Google Scholar
Strelnikov, B., Szewczyk, A., Strelnikova, I. et al. (2017). Spatial and temporal variability in MLT turbulence inferred from in situ and ground-based observations during the WADIS-1 sounding rocket campaign. Ann. Geophys., 35(3), 547565. doi: https://doi.org/10.5194/angeo-35-547-2017.Google Scholar
Lübken, F.-J. (1997). Seasonal variation of turbulent energy dissipation rates at high latitudes as determined by in situ measurements of neutral density fluctuations. J. Geophys. Res., 102(D12), 13,441–13,456. doi: https://doi.org/10.1029/97JD00853.Google Scholar
Guo, Y., Liu, A. Z., and Gardner, C. S.. (2017). First Na lidar measurements of turbulence heat flux, thermal diffusivity, and energy dissipation rate in the mesopause region. Geophys. Res. Lett., 44(11), 57825790. doi: https://doi.org/10.1002/2017GL073807.CrossRefGoogle Scholar
Pfrommer, T. and Hickson, P.. (2010). High-resolution lidar observations of mesospheric sodium and implications for adaptive optics. J. Opt. Soc. Am. A, 27(11), A97A105.Google Scholar
Pfrommer, T., Hickson, P., and She, C.-Y.. (2009). A large-aperture sodium fluorescence lidar with very high resolution for mesopause dynamics and adaptive optics studies. Geophys. Res. Lett., 36(15), L15831. doi: https://doi.org/10.1029/2009GL038802.Google Scholar
Taylor, L. R., Feng, Y., and Bonaccini Calia, D.. (2010). 50 W CW visible laser source at 589 nm obtained via frequency doubling of three coherently combined narrow-band Raman fibre amplifiers. Opt. Express 18(8), 85408555.CrossRefGoogle Scholar
She, C.-Y., Abo, M., Yue, J. et al. (2011). Mesopause-region temperature and wind measurements with pseudorandom modulation continuous-wave (PMCW) lidar at 589 nm. Appl. Opt. 50(18), 29162926.Google Scholar
Butler, D. J., Davies, R. I., Redfern, R. M. et al. (2003). Measuring the absolute height and profile of the mesospheric sodium layer using a continuous wave laser. Astron. Astrophys. 403(2), 775785. doi: https://doi.org/10.1051/0004-6361:20030379.Google Scholar
Hellemeier, J. A., Bonaccini Calia, D., Hickson, P., Otarola, A., and Pfrommer, T.. (2020). Measuring line-of-sight sodium density structure using laser guide stars. Monthly Notice of the Royal Astronomical Society MNRAS 494(2), 27982808.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×