검색 전체 메뉴
PDF
맨 위로
OA 학술지
On the Electric Fields Produced by Dipolar Coulomb Charges of an Individual Thundercloud in the Ionosphere
  • 비영리 CC BY-NC
  • 비영리 CC BY-NC
ABSTRACT
On the Electric Fields Produced by Dipolar Coulomb Charges of an Individual Thundercloud in the Ionosphere
KEYWORD
thundercloud , ionosphere , electrostatic field
  • 1. INTRODUCTION

    Thunderclouds are tropospheric sources of intense electrostatic fields and electromagnetic radiation. It is known that lightning-associated electric fields penetrate into the ionosphere; they have been observed in the E and F regions as transient electric fields with a typical duration of 10-20 ms and a magnitude of 1-50 mV/m (e.g., Kelley et al. 1985, 1990; Vlasov & Kelley 2009). According to the theoretical model of global atmospheric electricity developed by Hays & Roble (1979), the African array of multiple thunderclouds is responsible for the steady state electrostatic field of ~300 μV/m at ionospheric altitudes for nighttime conditions. The calculations by Park & Dejnakarintra (1973) showed that an isolated giant thundercloud could produce electrostatic fields of ~700 μV/m in the nighttime midlatitude ionosphere. However, Park & Dejnakarintra (1973) neglected the ionospheric Pedersen conductivity above 150 km. The purpose of this study is to theoretically examine the mapping of electrostatic fields of coulomb charges of an individual thundercloud into the midlatitude ionosphere, taking into account the height-integrated Pedersen conductivities of both hemispheres.

    2. BASIC EQUATIONS

    In the simplest thundercloud model, the electrical structure of a thundercloud is represented by two volume Coulomb charges of the same absolute value Q but opposite signs, with a positive charge in the upper part of the thundercloud and a negative charge in the lower part of the thundercloud (e.g., Chalmers 1967). Typical thunderclouds extend from 2-3 km to 8-12 km in altitude, and so-called giant thunderclouds extend above an altitude of 20 km (e.g., Uman 1969; Weisberg 1976). The magnitude of Q is estimated to range from 5 to 25 coulombs for the typical thunderclouds, whereas in giant thunderclouds, Q may exceed 50 coulombs (e.g., Malan 1963; Kasemir 1965).

    We use a cylindrical coordinate system (r, φ, z), in which the origin is placed at the earth’s surface and the z axis points upward and passes through the centers of thundercloud positive and negative volume charges. The mapping of thundercloud electrostatic field into the ionosphere is studied following a similar formalism to that used by Park & Dejnakarintra (1973). In the steady state case, the electrostatic field distribution above the thundercloud is described by the following equations:

    image
    image
    image

    where J is the electric current density, σ is the electrical conductivity tensor, and E and Φ are the electrostatic field and potential, respectively. If we assume that the geomagnetic field B is vertical and the electrical conductivity tensor depends only on z, the following equation for the electrostatic potential Φ cna be obtained from (1), (2), and (3) :

    image

    where σp is the Pedersen conductivity and σ0 is the specific conductivity. The atmospheric conductivity below 70 km is isotropic since drifts of charged particles are not affected by the geomagnetic field. Equation (4) can be solved analytically if the conductivities σ0 and σp are exponential functions of z. In the case of isotropic conductivity (setting σ0=σp=b exp(z/h), where b and h are constants), we obtain

    image

    where J0 is the zero-order Bessel function of the first kind, A1 and B1 are coefficients, and c1= -l/(2h) -[l/(4h2)+k2]1/2, c2= -l/(2h) + [l/(4h2)+k2]1/2. For the anisotropic region, where we let σ0 = b0 exp(z/h0) and σp = bp exp(z/hp), the solution to Equation (4) is

    image

    where Jν and Kν are the ν-order modified Bessel functions of the first and the second kind, respectively, and A2 and B2 are coefficients, ν=hp/(hp-h0), f=2νh0(bp/b0)1/2 exp[-z(2 νh0)]. The coefficients A1, B1, A2, and B2 are determined from the boundary conditions.

    The electrostatic field components are given by

    image
    image

    Since the geomagnetic field B is assumed to be vertical, Er is perpendicular to B, while Ez is parallel to B.

    Above 90 km, the geomagnetic field lines are practically equipotential because the geomagnetic field aligned conductivity σ0 is much higher than the transverse conductivity σp. It allows us to consider the ionoshperic region above 90 km as a thin conducting layer with a geomagnetic field line integrated Pedersen conductivity Σp, and the continuity equation of electric current at z=90 km takes the following form:

    image

    where ▽ denotes the gradient operator in the two dimensions transverse to B, and the factor 2 before Σp accounts for a contribution of the Pedersen conductivity of the magnetically conjugate ionosphere. Equation (9) is explicitly expressed as

    image

    We use the conductivity model as shown in Fig. 1. Below 70 km, the conductivity is isotropic and varies exponentially with z as σ01=σp1=b1 exp(z/h1) from 0 to 40 km, and as σ02=σp2=b2[exp(z-z1)/h2] from 40 to 70 km (where z1 = 40 km) with the values of b1,2 and h1,2 to approximately fit the atmospheric conductivity models by Cold & Pierce (1965) below 40 km and by Swider (1988) from 40 to 70 km. In the anisotropic region between 70 and 90 km, σ0, and σp are exponentially extrapolated from 70 km to their equinoctial midday and midnight values at z=90 km. At z ≥ 90 km, the conductivities are found from

    image
    image

    where subscripts e and i denote electrons and the ith ion species, Ne and Ni are the electron and ion densities, e is the electron charge, me and mi are the electron and ion masses, νe and νi are the electron and ion momentum transfer collision frequencies, and ωe and ωi are the electron and ion gyrofrequencies, respectively. The frequencies νe and νi are from Schunk (1988). The required input parameters are taken from the empirical ionospheric model IRI-2012 (http://omniweb.gsfc.nasa.gov/vitmo/iri2012_vitmo.html) and the neutral atmosphere model NRLMSIS-00 (http://ccmc.gsfc.nasa.gov/modelweb/models/nrlmsise00.php).

    Our calculations show that during solar minimum, in Equinox, the magnitude of Ʃp at middle latitudes is commonly in the ranges of 5.0-8.0 S and 0.1-0.2 S for day and night, respectively. However, the nighttime Ʃp can be as low as 0.05 S. Under solar maximum conditions, Ʃp is several times larger than in solar minimum.

    3. RESULTS AND DISCUSSION

    To compute the electrostatic potential above the thundercloud from (5) and (6), we impose the following boundary conditions:

    1. Φ=(Q/4πε0)[(r2+(zb-hp )2)-1/2-(r2+(zb-hn )2)-1/2 ] at z=zb 2. Φ is continuous at z=40 km 3. σ0 ∂Φ/∂z=2Ʃp (∂2Φ/∂r2+1/r ∂Φ/∂r) at z=90 km

    where ε0 is the vacuum permittivity, zb is the altitude of the plane setting directly above the thundercloud top, and hp and hn are the altitudes of positive and negative charge centers of the thundercloud, respectively. The first boundary condition follows from the accepted electrical model of the thundercloud. We assume that the thundercloud does not affect the atmospheric conductivity at zzb.

    Fig. 2 shows the computed electrostatic field component Er normalized to Q as a function of r in the nighttime and daytime midlatitude ionosphere at z≥90 km for the typical thundercloud (zb=10 km, hn=3 km, hp=8 km) and for the giant thundercloud (zb=20 km, hn=5 km, hp=17 km). Solar minimum conditions are considered with Ʃ p=0.05 S at night and Ʃ p=5.0 S by day. All curves show similar behavior, attaining first a maximum and then revealing a gradual lowering. At night, the thundercloud electrostatic field is transmitted into the ionosphere much better than during the daytime. For a typical thundercloud, Er reaches its nighttime maximum value of ~2.6 μV/m (for Q=25 coulombs) at r~35 km. In the case of the giant thundercloud, the nighttime maximum magnitude of Er is ~270 μV/m (for Q=50 coulombs) and rmax~40 km. The daytime maximum values of Er are one order of magnitude less than their nighttime values. Thus, the steady state electrostatic fields associated with the individual typical thunderclouds have very small magnitudes at ionospheric altitudes. In the case of a giant thundercloud, Er is two orders of magnitude larger. Note that Park & Dejnakarintra (1973) discovered that the maximum magnitude of the transverse electrostatic field produced in the nighttime midlatitude ionosphere by a giant thundercloud with Q=50 coulombs can be as large as ~700 μV/m, which is about 2.6 times more than in our estimate. This difference can mainly be attributed to the fact that Park & Dejnakarintra (1973) ignored the Pedersen conductivity above 150 km.

    4. CONCLUSION

    Our computations show that the geomagnetic field line integrated Pedersen conductivity of the ionosphere plays an important role in troposphere-ionosphere electrostatic coupling. Even for nighttime conditions in solar minimum, when the values of Ʃ p are minimal, the electrostatic charges of the individual thundercloud can drive only small electrostatic fields at ionospheric altitudes.

참고문헌
  • 1. Chalmers JA 1967 Atmospheric electricity P.309-342 google
  • 2. Cole Jr. RK, Pierce ET 1965 Electrification in the Earth's atmosphere for altitudes between 0 and 100 kilometers [J. Geophys. Res.] Vol.70 P.2735-2749 google cross ref
  • 3. Hays PB, Roble RG 1979 A quasi-static model of global atmospheric electricity, 1. The lower atmosphere [J. Geophys. Res.] Vol.84 P.3291-3305 google cross ref
  • 4. Kasemir HW, Coroniti SC 1965 The thundercloud, in Problems of Atmospheric and Space Electricity P.215-235 google
  • 5. Kelley MC, Siefring CL, Pfaff RF, Kintner PM, Larsen M 1985 Electrical measurements in the atmosphere and the ionosphere over an active thunderstorm: 1. Campaign overview and initial ionospheric results [J. Geophys. Res.] Vol.90 P.9815-9823 google cross ref
  • 6. Kelley MC, Ding JG, Holzworth RH 1990 Intense ionospheric electric and magnetic field pulses generated by lightning [Geophys. Res. Lett.] Vol.17 P.2221-2224 google cross ref
  • 7. Malan DJ 1963 Physics of Lightning P.23-37 google
  • 8. Park CG, Dejnakarintra M 1973 Penetration of thundercloud electric fields into the ionosphere and magnetosphere: 1. Middle and subauroral latitudes [J. Geophys. Res.] Vol.78 P.6623-6633 google cross ref
  • 9. Schunk RW 1988 A mathematical model of the middle and high latitude ionosphere [Pure Appl. Geophys.] Vol.127 P.255-303 google cross ref
  • 10. Swider W 1988 Ionic mobility, mean mass, and conductivity in the middle atmosphere from near ground level to 70 km [Radio Sci.] Vol.23 P.389-399 google cross ref
  • 11. Uman MA 1969 Lightning P.10-35 google
  • 12. Vlasov MN, Kelley MC 2009 Electron heating and airglow emission due to lightning effects on the ionosphere [J. Geophys. Res.] Vol.114 P.A00E06 google cross ref
  • 13. Weisberg JS 1976 Meteorology: The Earth and Its Weather P.107-122 google
OAK XML 통계
이미지 / 테이블
  • [ ] 
  • [ ] 
  • [ ] 
  • [ ] 
  • [ ] 
  • [ ] 
  • [ ] 
  • [ ] 
  • [ ] 
  • [ ] 
  • [ Fig. 1. ]  Model altitude profiles of specific (σ0) and Pedersen (σp) conductivities. The numbers next to the curves indicate conductivity scale heights in kilometers within each altitude section.
    Model altitude profiles of specific (σ0) and Pedersen (σp) conductivities. The numbers next to the curves indicate conductivity scale heights in kilometers within each altitude section.
  • [ ] 
  • [ ] 
  • [ Fig. 2. ]  Calculated magnitude of the thundercloud electrostatic field strength Er normalized to Q, as a function of r, at ionospheric altitudes z≥90 km for the typical and giant thundercloud at night and by day.
    Calculated magnitude of the thundercloud electrostatic field strength Er normalized to Q, as a function of r, at ionospheric altitudes z≥90 km for the typical and giant thundercloud at night and by day.
(우)06579 서울시 서초구 반포대로 201(반포동)
Tel. 02-537-6389 | Fax. 02-590-0571 | 문의 : oak2014@korea.kr
Copyright(c) National Library of Korea. All rights reserved.