Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-23T19:54:35.173Z Has data issue: false hasContentIssue false

15 - Space Weather Effects in the Ionosphere, in the Thermosphere and at Earth’s Surface

from Part IV - Space Weather

Published online by Cambridge University Press:  25 October 2019

Mioara Mandea
Affiliation:
Centre National d'études Spatiales, France
Monika Korte
Affiliation:
GeoforschungsZentrum, Helmholtz-Zentrum, Potsdam
Andrew Yau
Affiliation:
University of Calgary
Eduard Petrovsky
Affiliation:
Academy of Sciences of the Czech Republic, Prague
Get access

Summary

In the context of space weather effects, magnetosphere-ionosphere coupling is one of the fundamental processes controlling energy transfer and dissipation in geospace. Alfvén waves appear to play a key role in this coupling, specifically in coupling the dynamics of magnetospheric convection to the ionosphere and in generating the region 1 and region 2 global field-aligned current systems. The momentum transport from the magnetosphere to the ionosphere can be described as the result of the generation and propagation of Alfvén waves, for example as arising along newly reconnected magnetic field-lines, and in general in terms of their incidence on and reflection from the ionosphere. The thermosphere experiences dramatic changes in density and composition during magnetic storms. Intense Joule heating and particle precipitation at auroral latitudes cause intense thermal expansion, air upwelling and strong wind circulations. The Joule heating at E-layer altitudes can cause both density enhancements and depletions at higher altitudes, and complicate the interpretation of mass density anomalies at high latitudes. The thermospheric response to storms at middle and low latitudes is less complicated, where the averaged density enhancement is linearly proportional to the solar wind input. Magnetic substorms during active periods also cause mass density perturbations. Magnetic storms and substorms can cause disturbances up to thousands of nT at the Earth’s surface. The time derivative of the magnetic field provides a proxy for the associated geoelectric field, which can drive geomagnetically induced currents in Earthed conductors. The geoelectric field is thus a key quantity for space weather effects on technological systems such as power grids, and it can be obtained by modelling the magnetic field using ionospheric currents and model ground conductivity as inputs.

Type
Chapter
Information
Geomagnetism, Aeronomy and Space Weather
A Journey from the Earth's Core to the Sun
, pp. 229 - 250
Publisher: Cambridge University Press
Print publication year: 2019

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

Alekseev, D., Kuvshinov, A. and Palshin, N. (2015). Compilation of 3D global conductivity model of the Earth for space weather applications. Earth Planets Space, 67, 108, doi: 10.1186/s40623-015-0272-5.CrossRefGoogle Scholar
Alfvén, H. (1942). Existence of electromagnetic-hydromagnetic waves, Nature, 150, 405.Google Scholar
Allan, W. and Wright, A. N. (2000). Magnetotail waveguide: Fast and Alfvén waves in the plasma sheet boundary layer and lobe. J. Geophys. Res., 105(A1), 317–28.CrossRefGoogle Scholar
Anderson, B. J., Takahashi, K. and Toth, B. A. (2000). Sensing global Birkeland currents with Iridium engineering magnetometer data. Geophys Res Lett, 27(24), 4045–8.Google Scholar
Angelopoulos, V., Baumjohann, W., Kennel, C. F., et al. (1992). Bursty bulk flows in the inner central plasma sheet. J. Geophys. Res., 97(A4), 4027–39.Google Scholar
Angelopoulos, V., McFadden, J. P., Larson, D., et al. (2008). Tail reconnection triggering substorm onset. Science, 321(5891), 931–5.CrossRefGoogle ScholarPubMed
Bedrosian, P. A. and Love, J. J. (2015). Mapping geoelectric fields during magnetic storms: Synthetic analysis of empirical United States impedances. Geophys. Res. Lett., 42, doi: 10.1002/2015GL066636.Google Scholar
Beggan, C. D. (2015). Sensitivity of geomagnetically induced currents to varying auroral electrojet and conductivity models. Earth Planets Space, 67, 24, doi: 10.1186/s40623-014-0168-9.CrossRefGoogle Scholar
Bonner, L. R. and Schultz, A. (2017). Rapid prediction of electric fields associated with geomagnetically induced currents in the presence of three-dimensional ground structure: Projection of remote magnetic observatory data through magnetotelluric impedance tensors. Space Weather, 14, doi: 10.1002/2016SW001535.Google Scholar
Bruinsma, S. L. and Forbes, J. M. (2007). Global observation of traveling atmospheric disturbances (TADs) in the thermosphere. Geophys. Res. Lett., 34, L14103, doi: 10.1029/2007GL030243.CrossRefGoogle Scholar
Burke, W. J., Huang, C. Y., Marcos, F. A. and Wise, J. O. (2007). Interplanetary control of thermospheric densities during large magnetic storms. J. Atmos. Sol. Terr. Phys., 69, 279–87.Google Scholar
Burke, W. J., Lin, C. S., Hagan, M. P., Huang, C. Y., Weimer, D. R., Wise, J. O., Gentile, L. C. and Marcos, F. A. (2009). Storm time global thermosphere: A driven-dissipative thermodynamic system. J. Geophys. Res., 114, A06306, doi: 10.1029/2008JA013848.Google Scholar
Cagniard, L. (1953). Basic theory of the magneto-telluric method of geophysical prospecting. Geophysics, 18, 605–35, doi: 10.1190/1.1437915.CrossRefGoogle Scholar
Campbell, W. H. (1973). Spectral composition of geomagnetic field variations in the period range of 5 min to 2 hr as observed at the Earth’s surface. Radio Sci., 8, 929–32.Google Scholar
Chi, P. J., Russell, C. T. and Ohtani, S. (2009). Substorm onset timing via travel time magnetoseismology. Geophys. Res. Lett., 36(8).CrossRefGoogle Scholar
Cowley, S. W. H. (2000). Magnetosphere–ionosphere interactions: A tutorial review, in Magnetospheric Current Systems, ed. Ohtani, S.-I., Fujii, R., Hesse, M. and Lysak, R. L., American Geophysical Union, Washington, DC, doi: 10.1029/GM118p0091.Google Scholar
Coxon, J. C., Milan, S. E., Clausen, L. B. N., Anderson, B. J. and Korth, H. (2014), A superposed epoch analysis of the regions 1 and 2 Birkeland currents observed by AMPERE during substorms, J. Geophys. Res., 119(12), 9834–46.Google Scholar
Donovan, E., Mende, S., Jackel, B., et al. (2006). The azimuthal evolution of the substorm expansive phase onset aurora. Proc. ICS, 8, 5560.Google Scholar
Dungey, J. W. (1961). Interplanetary magnetic field and the auroral zones. Phys. Rev. Lett., 6(2), 47.Google Scholar
Emmert, J. T., Lean, J. L. and Picone, J. M. (2010). Record‐low thermospheric density during the 2008 solar minimum. Geophys. Res. Lett., 37, L12102, doi: 10.1029/2010GL043671.Google Scholar
Emmert, J. T. and Picone, J. M. (2010). Climatology of globally averaged thermospheric mass density. J. Geophys. Res., 115, A09326, doi: 10.1029/2010JA015298.Google Scholar
Engels, M., Korja, T. and the BEAR Working Group (2002). Multisheet modelling of the electrical conductivity structure in the Fennoscandian Shield. Earth Planets Space, 54, 559–73, doi: 10.1186/BF03353045.CrossRefGoogle Scholar
Forbes, J. M., Gonzalez, R., Marcos, F. A., Revelle, D. and Parish, H. (1996). Magnetic storm response of lower thermospheric density, J. Geophys. Res., 101, 2313–19.Google Scholar
Forbes, J. M., Lu, G., Bruinsma, S., Nerem, R. S. and Zhang, X. (2005). Thermosphere density variations due to the 15–24 April 2002 solar events from CHAMP/STAR accelerometer measurements, J. Geophys. Res., 110, A12S27, doi: 10.1029/2004JA010856.Google Scholar
Forsyth, C., Rae, I. J., Mann, I. R. and Pakhotin, I. P. (2017). Identifying intervals of temporally invariant field‐aligned currents from Swarm: Assessing the validity of single‐spacecraft methods. J. Geophys. Res., 122, 3411–19. doi: 10.1002/2016JA023708.Google Scholar
Fujiwara, H. and Miyoshi, Y. (2006). Characteristics of the large-scale traveling atmospheric disturbances during geomagnetically quiet and disturbed periods simulated by a whole atmosphere general circulation model, Geophys. Res. Lett., 33, L20108, doi: 10.1029/2006GL027103.Google Scholar
Friis‐Christensen, E., Lühr, H., Knudsen, D. and Haagmans, R. (2008). Swarm – an Earth observation mission investigating geospace. Adv. Space Res., 41, 210–16, doi: 10.1016/j.asr.2006.10.008.Google Scholar
Ganushkina, N. Y., Liemohn, M. W., Dubyagin, S., Daglis, I. A., Dandouras, I., De Zeeuw, D. L., Ebihara, Y., Ilie, R., Katus, R., Kubyshkina, M., Milan, S. E., Ohtani, S., Østgaard, N., Reistad, J. P., Tenfjord, P., Tofoletto, F., Zaharia, S. and Amariutei, O. (2015). Defining and resolving current systems in geospace. Ann. Geophys., 33, 13691402, doi: 10.5194/angeo-33-1369-2015.Google Scholar
Gjerloev, J. W. (2012). The SuperMAG data processing technique. J. Geophys. Res., 117, A09213, doi: 10.1029/2012JA017683.Google Scholar
Gjerloev, J. W., Ohtani, S., Iijima, T., Anderson, B., Slavin, J. and Le, G. (2011). Characteristics of the terrestrial field‐aligned current system. Ann. Geophys., 29, 1713–29, doi: 10.5194/angeo‐29‐1713‐2011.CrossRefGoogle Scholar
Grzesiak, M. (2000). Ionospheric Alfvén resonator as seen by Freja satellite. Geophys. Res. Lett., 27, 923–6, doi: 10.1029/1999GL010747.Google Scholar
Guo, J., Wan, W., Forbes, J. M., Sutton, E., Nerem, R. S., Woods, T. N., Bruinsma, S. and Liu, L. (2007). Effects of solar variability on thermosphere density from CHAMP accelerometer data. J. Geophys. Res., 112, A10308, doi: 10.1029/2007JA012409.Google Scholar
Iijima, T. and Potemra, T. A. (1976). The amplitude distribution of field‐aligned currents at northern high latitudes observed by Triad. J. Geophys. Res., 81(13), 2165–74.Google Scholar
Jacobs, J. A., Kato, Y., Matsushita, S. and Troitskaya, V. A. (1964). Classification of geomagnetic micropulsations. J. Geophys. Res., 69(1), 180–81.Google Scholar
Janhunen, P., Palmroth, M., Laitinen, T., Honkonen, I., Juusola, L., Facsko, G. and Pulkkinen, T. I. (2012). The GUMICS-4 global MHD magnetosphere–ionosphere coupling simulation. J. Atmos. Sol. Terr. Phys., 80, 4859, doi: 10.1016/j.jastp.2012.03.006.Google Scholar
Juusola, L., Kauristie, K., van de Kamp, M., Tanskanen, E. I., Mursula, K., Asikainen, T., Andreeova, K., Partamies, N., Vanhamäki, H. and Viljanen, A. (2015). Solar wind control of ionospheric equivalent currents and their time derivatives. J. Geophys. Res., 120, doi: 10.1002/2015JA021204.Google Scholar
Kalmoni, N. M. E., Rae, I. J., Murphy, K. R., et al. (2017). Statistical azimuthal structuring of the substorm onset arc: Implications for the onset mechanism. Geophys. Res. Lett., 44(5), 2078–87.CrossRefGoogle Scholar
Kan, J. K. and Lee, L. C. (1979). Energy coupling function and solar wind-magnetosphere dynamo. Geophys. Res. Lett., 6(7), 577–80.Google Scholar
Kaufman, A. A. and Keller, G. V. (1981). The Magnetotelluric Sounding Method. Elsevier, New York.Google Scholar
Keiling, A. and Takahashi, K. (2011). Review of Pi2 models. Space Sci. Rev., 161(1–4), 63148.CrossRefGoogle Scholar
Keiling, A., Wygant, J. R., Cattell, C. A., Mozer, F. S. and Russell, C. T. (2003). The global morphology of wave Poynting flux: Powering the aurora. Science, 299(5605), 383–6.CrossRefGoogle ScholarPubMed
Kelbert, A., Balch, C. C., Pulkkinen, A., Egbert, G. D., Love, J. J., Rigler, E. J. and Fujii, I. (2017). Methodology for time-domain estimation of storm-time geoelectric fields using the 3D magnetotelluric response tensors. Space Weather, 15, doi: 10.1002/2017SW001594.CrossRefGoogle Scholar
Kelbert, A., Kuvshinov, A., Velimsky, J., Koyama, T., Ribaudo, J., Sun, J., Martinec, Z. and Weiss, C. J. (2014). Global 3-D electromagnetic forward modelling: A benchmark study. Geophys. J. Int., 197, 785814, doi: 10.1093/gji/ggu028.Google Scholar
Kepko, L., Kivelson, M. G. and Yumoto, K. (2001). Flow bursts, braking, and Pi2 pulsations. J. Geophys. Res., 106(A2), 1903–15, doi: 10.1029/2000JA000158.Google Scholar
Kepko, L., McPherron, R. L., Amm, O., Apatenkov, S., Baumjohann, W., et al. (2015). Substorm current wedge revisited. Space Sci. Rev., 190(1–4), 146.Google Scholar
Knudsen, D. J., Kelley, M. C. and Vickrey, J. F. (1992).Alfvén waves in the auroral ionosphere: A numerical model compared with measurements. J. Geophys. Res., 97(A1), 7790, doi: 10.1029/91JA02300.CrossRefGoogle Scholar
Knudsen, D. J., Burchill, J. K., Buchert, S. C., et al. (2017). Thermal ion imagers and Langmuir probes in the Swarm electric field instruments. J. Geophys. Res., 122, 2655–73, doi: 10.1002/2016JA022571.Google Scholar
Kwak, Y.-S., Richmond, A. D., Deng, Y., Forbes, J. M. and Kim, K.-H. (2009). Dependence of the high-latitude thermospheric densities on the interplanetary magnetic field. J. Geophys. Res., 114, A05304, doi: 10.1029/2008JA013882.Google Scholar
Lei, J., Thayer, J. P., Burns, A. G., Lu, G. and Deng, Y. (2010). Wind and temperature effects on thermosphere mass density response to the November 2002 Geomagnetic Storm. J. Geophys. Res., 115, A05303, doi: 10.1029/2009JA014754.Google Scholar
Lehtinen, M. and Pirjola, R. (1985). Currents produced in earthed conductor networks by geomagnetically-induced electric fields. Ann. Geophys., 3, 479–84.Google Scholar
Liu, H. and Lühr, H. (2005). Strong disturbance of the upper thermospheric density due to magnetic storms: CHAMP observations. J. Geophys. Res., 110, A09S29, doi: 10.1029/2004JA010908.Google Scholar
Liu, H., Lühr, H. and Watanabe, S. (2007). Climatology of the Equatorial Thermospheric Mass Density Anomaly. J. Geophys. Res., 112, A05305, doi: 10.1029/2006JA012199.Google Scholar
Liu, R., Lühr, H. and Ma, S. Y. (2010a). Storm-time related mass density anomalies in the polar cap as observed by CHAMP. Ann. Geophys., 28(1), 165–80.Google Scholar
Liu, R., Lühr, H., Doornbos, E. and Ma, S. Y. (2010b). Thermospheric mass density variations during geomagnetic storms and a prediction model based on the merging electric field. Ann. Geophys., 28, 1633–45, doi: 10.5194/angeo-28-1633-2010.Google Scholar
Liu, R., Ma, S.-Y. and Lühr, H. (2011). Predicting storm-time thermospheric mass density variations at CHAMP and GRACE altitudes. Ann. Geophys., 29, 443–53, doi: 10.5194/angeo-29-443-2011.CrossRefGoogle Scholar
Lockwood, M., Cowley, S. W. H. and Freeman, M. P. (1990). The excitation of plasma convection in the high‐latitude ionosphere. J. Geophys. Res., 95(A6), 7961–72.CrossRefGoogle Scholar
Love, J. J., Pulkkinen, A., Bedrosian, P. A., Jonas, S., Kelbert, A., Rigler, E. J., Finn, C. A., Balch, C. C., Rutledge, R., Wagge, R. M., Sabata, A. T., Kozyra, J. U. and Black, C. E. (2016). Geoelectric hazard maps for the continental United States. Geophys. Res. Lett., 43, 9415–24, doi: 10.1002/2016GL070469.Google Scholar
Lu, G., Richmond, A. D., Lühr, H. and Paxton, L. (2016). High-latitude energy input and its impact on the thermosphere. J. Geophys. Res., 121, 7108–24, doi: 10.1002/2015JA022294.Google Scholar
Lühr, H., Park, J., Gjerloev, J. W., Rauberg, J., Michaelis, I., Merayo, J. M. G. and Brauer, P. (2015). Field‐aligned currents’ scale analysis performed with the Swarm constellation. Geophys. Res. Lett., 42, 18, doi: 10.1002/2014GL062453.Google Scholar
Lühr, H., Park, J., Ritter, P. and Liu, H. (2012). In-situ CHAMP observation of ionosphere-thermosphere coupling. Space Sci. Rev., 168, 237–60, doi: 10.1007/s11214-011-9798-4.Google Scholar
Lühr, H., Rother, M., Köhler, W., Ritter, P. and Grunwaldt, L. (2004). Thermospheric up-welling in the cusp region, evidence from CHAMP observations, Geophys. Res. Lett., 31, L06805, doi: 10.1029/2003GL019314.CrossRefGoogle Scholar
Lyon, J. G., Fedder, J. A. and Mobarry, C. M. (2004). The Lyon-Fedder-Mobarry (LFM) global MHD magnetospheric simulation code. J. Atmos. Sol. Terr. Phys., 66, 1333–50, doi: 10.1016/j.jastp.2004.03.020.Google Scholar
Lysak, R. L. (1991). Feedback instability of the ionospheric resonant cavity. J. Geophys. Res., 96, 1553–68, doi: 10.1029/90JA02154.Google Scholar
Mann, I. R., Milling, D. K., Rae, I. J., et al. (2008). The upgraded CARISMA magnetometer array in the THEMIS era. Space Sci. Rev., 141(1–4), 413–51.Google Scholar
Mann, I. R., Wright, A. N., Mills, K. J. and Nakariakov, V. M. (1999). Excitation of magnetospheric waveguide modes by magnetosheath flows. J. Geophys. Res., 104(A1), 333–53.CrossRefGoogle Scholar
Marti, L., Yiu, C., Rezaei-Zare, A. and Boteler, D. (2014). Simulation of geomagnetically induced currents with piecewise layered-Earth models. IEEE Trans. Power Delivery, 29, 1886–93, doi: 10.1109/TPWRD.2014.2317851.CrossRefGoogle Scholar
McPherron, R. L., Russell, C. T. and Aubry, M. P. (1973). Satellite studies of magnetospheric substorms on August 15, 1968: 9. Phenomenological model for substorms. J. Geophys. Res., 78(16), 3131–49.Google Scholar
Milan, S. E. (2013). Modeling Birkeland currents in the expanding/contracting polar cap paradigm, J. Geophys. Res., 118(9), 5532–42.Google Scholar
Milan, S. E., Clausen, L. B. N., Coxon, J. C., Carter, J. A., Walach, M.-T., Laundal, K., Østgaard, N., Tenfjord, P., Reistad, J., Snekvik, K., Korth, H. and Anderson, B. J. (2017). Overview of solar wind-magnetosphere–ionosphere-atmosphere coupling and the generation of magnetospheric currents. Space Sci. Rev., doi: 10.1007/s11214-017-0333-0.Google Scholar
Miles, D. M., Mann, I. R., Pakhotin, I. P., et al. (2018). Alfvénic dynamics and fine structuring of discrete auroral arcs: Swarm and e‐POP observations. Geophys. Res. Lett., 45(2), 545–55.Google Scholar
Milling, D. K., Rae, I. J., Mann, I. R., et al. (2008). Ionospheric localisation and expansion of long‐period Pi1 pulsations at substorm onset. Geophys. Res. Lett., 35(17).Google Scholar
Müller, S., Lühr, H. and Rentz, S. (2009). Solar and magnetospheric forcing of the low latitude thermospheric mass density, as observed by CHAMP. Ann. Geophys., 27, 2087–99.Google Scholar
Murphy, K. R., Rae, I. J., Mann, I. R., et al. (2009). Wavelet‐based ULF wave diagnosis of substorm expansion phase onset. J. Geophys. Res., 114(A1).Google Scholar
Nakamura, R., Baumjohann, W., Schödel, R., Brittnacher, M., Sergeev, V. A., Kubyshkina, M., Mukai, T. and Liou, K. (2001). Earthward flow bursts, auroral streamers, and small expansions. J. Geophys. Res., 106(A6), 10791–802, doi: 10.1029/2000JA000306.Google Scholar
Newell, P. T., Sotirelis, T., Liou, K., Meng, C. I. and Rich, F. J. (2007). A nearly universal solar wind-magnetosphere coupling function inferred from magnetospheric state variables. J. Geophys. Res., 112, A01206, doi: 10.1029/2006JA012015.Google Scholar
Ngwira, C. M., Pulkkinen, A. A., Bernabeu, E., Eichner, J., Viljanen, A. and Crowley, G. (2015). Characteristics of extreme geoelectric fields and their possible causes: Localized peak enhancements. Geophys. Res. Lett., 42, 6916–21, doi: 10.1002/2015GL065061.Google Scholar
Ngwira, C. M., Pulkkinen, A., Kuznetsova, M. M. and Glocer, A. (2014). Modeling extreme ‘Carrington-type’ space weather events using three-dimensional global MHD simulations. J. Geophys. Res., 119, 4456–74, doi: 10.1002/2013JA019661.Google Scholar
Nikitina, L., Trichtchenko, L. and Boteler, D. H. (2016). Assessment of extreme values in geomagnetic and geoelectric field variations for Canada. Space Weather, 14, doi: 10.1002/2016SW001386.Google Scholar
Ogino, T., Walker, R. J. and Ashour-Abdalla, M. (1994). A global magnetohydrodynamic simulation of the response of the magnetosphere to a northward turning of the interplanetary magnetic field. J. Geophys. Res., 99, 11027–42, doi: 10.1029/93JA03313.Google Scholar
Ohtani, S. I. (2004). Flow bursts in the plasma sheet and auroral substorm onset: Observational constraints on connection between midtail and near-Earth substorm processes. Space Sci. Rev., 113(1–2), 7796.CrossRefGoogle Scholar
Pakhotin, I. P., Mann, I. R., Lysak, R. L., et al. (2018). Diagnosing the role of Alfvén waves in magnetosphere‐ionosphere coupling: Swarm observations of large amplitude nonstationary magnetic perturbations during an interval of northward IMF. J. Geophys. Res., 123(1), 326–40.Google Scholar
Peticolas, L. M., Craig, N., Odenwald, S. F., et al. (2009). The Time History of Events and Macroscale Interactions during Substorms (THEMIS) education and outreach (E/PO) program, in The THEMIS Mission, pp. 557–83, Springer, New York.Google Scholar
Pirjola, R. (2010). Derivation of characteristics of the relation between geomagnetic and geoelectric variation fields from the surface impedance for a two-layer Earth. Earth Planets Space, 62, 287–95.Google Scholar
Pirjola, R. and Lehtinen, M. (1985). Currents produced in the Finnish 400 kV power transmission grid and in the Finnish natural gas pipeline by geomagnetically-induced electric fields. Ann. Geophys., 3, 485–91.Google Scholar
Prölss, G. W. (1987). Storm-induced changes in the thermospheric composition at middle latitudes, Planet. Space Sci., 35, 807–11.Google Scholar
Prölss, G. W. (1997). Magnetic storm associated perturbations of the upper atmosphere, in Magnetic Storms, ed. Tsurutani, B. T., Gonzalez, W. D., Kamide, Y. and Arballo, J. K., pp. 227–41, Geophys. Monogr. 98, AGU, Washington, DC.Google Scholar
Prölss, G. W. (2005). Physics of the Earth’s Space Environment, Springer, Berlin.Google Scholar
Prölss, G. W. (2011). Density perturbations in the upper atmosphere caused by the dissipation of solar wind energy, Surv. Geophys., 32, 101–95, doi: 10.1007/s10712-010-9104-0.Google Scholar
Pulkkinen, A., Bernabeu, E., Eichner, J., Viljanen, A. and Ngwira, C. (2015). Regional-scale high-latitude extreme geoelectric fields pertaining to geomagnetically induced currents. Earth Planets Space, 67, 93, doi: 10.1186/s40623-015-0255-6.Google Scholar
Pulkkinen, A., Bernabeu, E., Thomson, A., Viljanen, A., Pirjola, R., Boteler, D., Eichner, J., Cilliers, P. J., Welling, D., Savani, N. P., Weigel, R. S., Love, J. J., Balch, C., Ngwira, C. M., Crowley, G., Schultz, A., Kataoka, R., Anderson, B., Fugate, D., Simpson, J. J. and MacAlester, M. (2017). Geomagnetically induced currents: Science, engineering and applications readiness. Space Weather, 15, doi: 10.1002/2016SW001501.Google Scholar
Pulkkinen, A., Kataoka, R., Watari, S. and Ichiki, M. (2010). Modeling geomagnetically induced currents in Hokkaido, Japan. Adv. Space Res., 46, 1087–93, doi: 10.1016/j.asr.2010.05.024.Google Scholar
Pulkkinen, A., Klimas, A., Vassiliadis, D., Uritsky, V. and Tanskanen, E. (2006). Spatiotemporal scaling properties of the ground geomagnetic field variations. J. Geophys. Res., 111(A3), A03305, doi: 10.1029/2005JA011294.CrossRefGoogle Scholar
Püthe, C. and Kuvshinov, A. (2013). Towards quantitative assessment of the hazard from space weather: Global 3-D modellings of the electric field induced by a realistic geomagnetic storm. Earth Planets Space, 65, 1017–25, doi: 10.5047/eps.2013.03.003.Google Scholar
Püthe, C., Manoj, C. and Kuvshinov, A. (2014). Reproducing electric field observations during magnetic storms by means of rigorous 3-D modelling and distortion matrix co-estimation. Earth Planets Space, 66, 162, doi: 10.1186/s40623-014-0162-2.Google Scholar
Rae, I. J., Mann, I. R., Murphy, K. R., et al. (2009a). Timing and localization of ionospheric signatures associated with substorm expansion phase onset. J. Geophys. Res., 114(A1).Google Scholar
Rae, I. J., Mann, I. R., Angelopoulos, V., et al. (2009b). Near‐Earth initiation of a terrestrial substorm. J. Geophys. Res., 114(A7).Google Scholar
Rae, I. J., Murphy, K. R., Watt, C. E., et al. (2017). Using ultra-low frequency waves and their characteristics to diagnose key physics of substorm onset. Geosci. Lett., 4(1), 23.Google Scholar
Raeder, J., Larson, D., Li, W., Kepko, E. L. and Fuller-Rowell, T. (2008). OpenGGCM simulations for the THEMIS mission. Space Sci. Rev., 141, 535, doi: 10.1007/s11214-008-9421-5.Google Scholar
Ritter, P., Lühr, H. and Doornbos, E. (2010). Substorm-related thermospheric density and wind disturbances derived from CHAMP observations. Ann. Geophys., 28, 1207–20, doi: 10.5194/angeo-28-1207-2010.Google Scholar
Ritter, P., Lühr, H. and Rauberg, J. (2013). Determining field‐aligned currents with the Swarm constellation mission. Earth Planet Space, 65(11), 1285–94, doi: 10.5047/eps.2013.09.006.Google Scholar
Russell, C. T., Chi, P. J., Dearborn, D. J., Ge, Y. S., Kuo-Tiong, B., Means, J. D., Pierce, D. R., Rowe, K. M. and Snare, R. C. (2008). THEMIS ground-based magnetometers. Space Sci. Rev., 141(1–4), 389412.Google Scholar
Scholer, M. (1970). On the motion of artificial ion clouds in the magnetosphere. Planet. Space Sci., 18, 977.Google Scholar
Song, Y. and Lysak, R. L. (2001). Towards a new paradigm: From a quasi-steady description to a dynamical description of the magnetosphere. Space Sci. Rev., 95(1–2), 273–92.Google Scholar
Strangeway, R. J. (2012). The relationship between magnetospheric processes and auroral field-aligned current morphology. Auror. Phenomenol. Magnetos. Process. Earth Planets, 197, 355–64.Google Scholar
Takahashi, K., Lee, D.-H., Nosé, M., Anderson, R. R. and Hughes, W. J. (2003). CRRES electric field study of the radial mode structure of Pi2 pulsations. J. Geophys. Res., 108, 1210, doi: 10.1029/2002JA009761, A5.Google Scholar
Tamao, T. (1964). The structure of three-dimensional hydromagnetic waves in a uniform cold plasma. J. Geomagn. Geoelectr., 18, 89114.CrossRefGoogle Scholar
Tanskanen, E. I., Viljanen, A., Pulkkinen, T. I., Pirjola, R., Häkkinen, L., Pulkkinen, A. and Amm, O. (2001). At substorm onset, 40 % of AL comes from underground. J. Geophys. Res., 106, 13119–34.Google Scholar
Thomson, A., Dawson, E. and Reay, S. (2011). Quantifying extreme behaviour in geomagnetic activity. Space Weather, 9, S10001, doi: 10.1029/2011SW000696.Google Scholar
Toth, G., Sokolov, I. V., Gombosi, T. I., Chesney, D. R., Clauer, C. R., DeZeeuw, C. D. L., Hansen, K. C., Kane, K. J., Manchester, W. B., Oehmke, R. C., Powell, K. G., Ridley, A. R., Roussev, I. I., Stout, Q. F., Volberg, O., Wolf, R. A., Sazykin, S., Chan, A., Yu, B. and Kota, J. (2005). Space weather modeling framework: A new tool for the space science community. J. Geophys. Res., 110. doi: 10.1029/2005JA011126.Google Scholar
Untiedt, J. and Baumjohann, W. (1993). Studies of polar current systems using the IMS Scandinavian magnetometer array. Space Sci. Rev., 63, 245390, doi: 10.1007/BF00750770.Google Scholar
Vasseur, G. and Weidelt, P. (1977). Bimodal electromagnetic induction in non-uniform thin sheets with an application to the northern Pyrenean induction anomaly. Geophys. J. R. Astron. Soc., 51, 669–90, doi: 10.1111/j.1365-246X.1977.tb04213.x.Google Scholar
Viljanen, A., Amm, O. and Pirjola, R. (1999). Modelling geomagnetically induced currents during different ionospheric situations. J. Geophys. Res., 104, 28059–72, doi: 10.1029/1999JA900337.Google Scholar
Viljanen, A., Pirjola, R., Wik, M., Adam, A., Pracser, E., Sakharov, Ya. and Katkalov, Yu. (2012). Continental scale modelling of geomagnetically induced currents. J. Space Weather Space Clim., 2, A17, doi: 10.1051/swsc/2012017.Google Scholar
Viljanen, A., Pulkkinen, A., Pirjola, R., Pajunpää, K., Posio, P. and Koistinen, A. (2006). Recordings of geomagnetically induced currents and a nowcasting service of the Finnish natural gas pipeline system. Space Weather, 4, S10004, doi: 10.1029/2006SW000234.Google Scholar
Viljanen, A., Wintoft, P. and Wik, M. (2015). Regional estimation of geomagnetically induced currents based on the local magnetic or electric field. J. Space Weather Space Clim., 5, A24, doi: 10.1051/swsc/2015022.Google Scholar
Watari, S., Kunitake, M., Kitamura, K., Hori, T., Kikuchi, T., Shiokawa, K., Nishitani, N., Kataoka, R., Kamide, Y., Aso, T., Watanabe, Y. and Tsuneta, Y. (2009). Measurements of geomagnetically induced current in a power grid in Hokkaido, Japan. Space Weather, 7, S03002, doi: 10.1029/2008SW000417.Google Scholar
Weaver, J. T. (1964). On the separation of local geomagnetic fields into external and internal parts. Z. Geophys., 30, 2936.Google Scholar
Wei, L. H., Homeier, N. and Gannon, J. L. (2013). Surface electric fields for North America during historical geomagnetic storms. Space Weather, 11, 451–62, doi: 10.1002/swe.20073.Google Scholar
Weigel, R. S. (2017). A comparison of methods for estimating the geoelectric field. Space Weather, 15, 430–40, doi: 10.1002/2016SW001504.Google Scholar
Weigel, R. S., Klimas, A. J. and Vassiliadis, D. (2003). Solar wind coupling to and predictability of ground magnetic fields and their time derivatives, J. Geophys. Res., 108(A7), 1298, doi: 10.1029/2002JA009627.Google Scholar
Weimer, D. R. (2001). Maps of ionospheric field‐aligned currents as a function of the interplanetary magnetic field derived from Dynamics Explorer 2 data. J. Geophys. Res., 106(A7), 12889–902.Google Scholar
Weimer, D. R. (2013). An empirical model of ground-level geomagnetic perturbations. Space Weather, 11, 107–20, doi: 10.1002/swe.20030.Google Scholar
Wintoft, P., Wik, M. and Viljanen, A. (2015). Solar wind driven empirical forecast models of the time derivative of the ground magnetic field. J. Space Weather Space Clim., 5, A7, doi: 10.1051/swsc/2015008.Google Scholar
Wright, A. N. (1996). Transfer of magnetosheath momentum and energy to the ionosphere along open field lines. J. Geophys. Res., 101(A6), 13169–78.Google Scholar
Wright, A. N. and Mann, I. R. (2006). Global MHD eigenmodes of the outer magnetosphere, in Magnetospheric ULF Waves: Synthesis and New Directions, ed. Takahashi, K., Chi, P. J., Denton, R. E. and Lysak, R. L., pp. 5172, American Geophysical Union, Washington, DC.Google Scholar
Wygant, J. R., Keiling, A., Cattell, C. A., et al. (2002). Evidence for kinetic Alfvén waves and parallel electron energization at 4–6 RE altitudes in the plasma sheet boundary layer. J. Geophys. Res., 105(A8), doi: 10.1029/2001JA900113.Google Scholar
Wygant, J. R., Keiling, A., Cattell, C. A., et al. (2000). Polar spacecraft based comparisons of intense electric fields and Poynting flux near and within the plasma sheet-tail lobe boundary to UVI images: An energy source for the aurora. J. Geophys. Res., 105(A8), 18675–92.Google Scholar
Zhang, J. J., Wang, C. and Tang, B. B. (2012). Modeling geomagnetically induced electric field and currents by combining a global MHD model with a local one-dimensional method. Space Weather, 10, S05005, doi: 10.1029/2012SW000772.Google Scholar
Zhou, Y. L., Ma, S. Y., Lühr, H., Xiong, C. and Reigber, C. (2009). An empirical relation to correct storm-time thermospheric mass density modeled by NRLMSISE-00 with CHAMP satellite air drag data, Adv. Space Res., 43, 819–28.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
×