You are on page 1of 9

Journal of Atmospheric and Solar–Terrestrial Physics 212 (2021) 105488

Contents lists available at ScienceDirect

Journal of Atmospheric and Solar–Terrestrial Physics


journal homepage: www.elsevier.com/locate/jastp

Research paper

Dissipation of acoustic–gravity waves in the Earth’s thermosphere


A.K. Fedorenko, E.I. Kryuchkov, O.K. Cheremnykh, Y.A. Selivanov ∗
Space Research Institute of NANU and SSAU, 40, Acad. Glushkov ave., 03187, Kyiv, Ukraine

ARTICLE INFO ABSTRACT

Keywords: The dissipation of acoustic–gravity waves in the thermosphere due to molecular viscosity and thermal
Acoustic–gravity waves conductivity based on a modified system of hydrodynamic equations is studied. A modification of the system
Earth’s thermosphere of equations consists in additional taking into account in the linearized Navier–Stokes equations and heat
Molecular viscosity
transfer additional terms describing the transfer of momentum and energy due to the background density
Thermal conductivity
gradient, except for the usually taken into account velocity gradient. Using these modified equations, the local
Attenuation rate
dispersion equation of acoustic–gravity waves in an isothermal dissipative atmosphere is obtained. On this
basis, new general expressions for the attenuation rates in time and in space for both acoustic and gravity waves
in the entire spectral range are derived in the paper. The relationship between the rates of AGW temporal and
spatial attenuation is analyzed. The results obtained are consistent with satellite-based observations of AGW
at the heights of the thermosphere, indicating filtering of the wave spectrum due to viscosity and thermal
conductivity.

1. Introduction a limited range of heights of the thermosphere: 200 − 450 km. AGW
are recorded from low-orbit satellites in the form of fluctuations in the
The interest in studying acoustic–gravity waves (AGW) in the atmo- parameters of the atmosphere and ionosphere: density, temperature,
spheres of planets and the Sun has not been weakening for more than and velocity. By comparing synchronous measurements of different
half a century (Hines, 1960; Yeh and Liu, 1974; Beer, 1974; Roy et al., parameters with theory, one can reconstruct the spectral characteristics
2019). This is due to the important role played by AGW in the dynamics of waves and the directions of their motion (Fedorenko, 2009). Since
and energetics of planetary atmospheres. These waves can be gener- the temperature in the thermosphere almost does not change with
ated by sources of different nature and propagate over considerable height, this gives reason to apply the theoretical concepts developed for
distances, providing redistribution of the energy of disturbances in the an isothermal atmosphere when analyzing satellite data (Hines, 1960;
atmosphere, both in the horizontal plane and between different altitude Yeh and Liu, 1974).
levels. The earliest theoretical studies of AGW were carried out mainly However, when comparing the satellite observations of AGWs with
within the framework of the linear theory developed for an isothermal theory, a number of difficulties arise, primarily related to the strong
atmosphere (Hines, 1960; Yeh and Liu, 1974; Francis, 1975). Later, the
rarefaction of the neutral medium, as well as the presence of an
nonlinear theory of AGW was significantly developed (Belashov, 1990;
ionized component, the concentration of which increases with height.
Nekrasov et al., 1995; Kaladze et al., 2008; Stenflo and Shukla, 2009;
A clear feature of the observation of AGWs at satellite altitudes is the
Huang et al., 2014). The influence of changes in temperature, viscosity
amplitude–phase differences in wave fluctuations in different types of
and thermal conductivity with height, as well as of other features of the
atmospheric gases (Dudis and Reber, 1976; Fedorenko and Kryuchkov,
real atmosphere, on the propagation of AGWs are taken into account
2014). These differences are due to the fact that in the absence of strong
mainly by numerical simulation (Mayr et al., 1990; Piani et al., 2000;
turbulent mixing above the turbopause, different gases are distributed
Kshevetskii, 2001; Zhang and Yi, 2002; Cheremnykh et al., 2010).
To verify the theory and numerical models, experimental obser- in height with their individual altitude scales. Also, at the heights of
vations of the AGW in the atmosphere are necessary. To date, many the thermosphere, the propagation of AGW is strongly influenced by
different remote methods have been developed for ground-based and viscosity and thermal conductivity. In addition to attenuation of am-
satellite diagnostics of AGW in the atmosphere, a meaningful overview plitudes, viscosity can lead to selective filtering of the wave spectrum.
of these issues can be found in Meng et al. (2019). Direct satellite The features of viscous dissipation of gravity waves at different heights
measurements of the parameters of these waves are available only in depending on their spectral properties were studied in Vadas and Fritts

∗ Corresponding author.
E-mail address: yuraslv@gmail.com (Y. Selivanov).

https://doi.org/10.1016/j.jastp.2020.105488
Received 15 June 2020; Received in revised form 14 October 2020; Accepted 23 October 2020
Available online 16 November 2020
1364-6826/© 2020 Elsevier Ltd. All rights reserved.
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488

(2005), Fritts and Vadas (2008), Vadas and Nicolls (2012) and Lizunov 2. Basic equations
and Leontiev (2014).
Satellite observations of the AGW in the Earth’s polar thermosphere The general system of hydrodynamic equations describing an
indicate the predominance of waves with periods of the order of the isothermal compressible dissipative atmosphere without external
Brunt–Väisälä period and horizontal lengths in the interval of 400 − sources includes the equations of continuity, Navier–Stokes and heat
700 km (Johnson et al., 1995; Innis and Conde, 2002; Fedorenko and transfer (Landau and Lifshitz, 1987; Ladikov-Roev and Cheremnykh,
Kryuchkov, 2011). In addition, in the altitude range of 250 − 400 km 2010):
of the thermosphere, there is no pronounced dependence of the AGW 𝑑𝜌
amplitudes on altitude (Fedorenko et al., 2015). The AGW amplitudes + 𝜌∇ ⋅ 𝑉⃗ = 0, (1)
𝑑𝑡
in relative density fluctuations in the range of 1%–6% are equally often
𝑑 𝑉⃗
observed at different heights of the indicated interval. If we assume 𝜌 + ∇𝑃 − ∇ ⋅ 𝑆̂ − 𝜌𝑔⃗ = 0, (2)
𝑑𝑡( )
that the sources of these waves are of an auroral nature (for example,
𝑑 𝑐𝑉 𝑇
particles precipitations or ionospheric currents) and are located below 𝜌 + 𝑃 ∇ ⋅ 𝑉⃗ − ∇ ⋅ (𝜒∇𝑇 ) = 𝑄. (3)
the base of the thermosphere, then, according to the theory, in the 𝑑𝑡
𝑑 𝜕
absence of dissipation, their amplitudes should increase exponentially. Here 𝑑𝑡 = 𝜕𝑡 + 𝑉⃗ ⋅ ∇ is the material derivative, 𝑉⃗ is the particle velocity
The predominance, in satellite observations, of certain spectral scales of vector with the components 𝑉𝑖 and 𝑉𝑗 in the Cartesian coordinate
the AGW, as well as the absence of a noticeable increase in amplitudes system, 𝜌 is the density, 𝑇 is the temperature, 𝑃 is the pressure, 𝑔 is the
in the thermosphere, can be a consequence of viscosity and thermal gravity acceleration, 𝑆̂ is the tensor of viscous stresses with components
conductivity.
The main difficulty of the analytical approach when studying the ( ) ( )
2 𝜕𝑉𝑖 𝜕𝑉𝑗
atmosphere is that its properties (density, temperature, viscosity and 𝑆̂𝑖𝑗 = 𝛿𝑖𝑗 𝜀 − 𝜇 ∇ ⋅ 𝑉⃗ + 𝜇 + , (4)
3 𝜕𝑥𝑗 𝜕𝑥𝑖
thermal conductivity coefficients, etc.) vary with height. As a result,
the system of hydrodynamic equations describing wave perturbations where 𝜇 and 𝜀 are the dynamic and bulk viscosity coefficients, re-
in the atmosphere contains height-dependent coefficients. A barometric spectively, 𝜒 = 𝑐𝑃 𝜇∕𝑃 𝑟 is the thermal conductivity, 𝑃 𝑟 is the Prandtl
change in density with height is usually excluded in the AGW theory number, 𝑐𝑃 and 𝑐𝑉 are the specific heat at constant ( ) volume and
( )
for an isothermal atmosphere using the well-known substitution (Hines, pressure, respectively. The term 𝑄 = 𝑉⃗ ⋅ ∇ ⋅ 𝑆̂ −∇⋅ 𝑆̂ ⋅ 𝑉⃗ on the right-
1960; Yeh and Liu, 1974), which leads to a system of equations with hand side of (3) describes the heating of the medium due to viscous
constant coefficients. However, it is impossible to take into account dissipation. The axis 𝑧 of the Cartesian coordinate system is directed
in this way a change of temperature and transport coefficients with up. Since we assume that the atmosphere is uniform in the horizontal
height. Therefore, the analytical expression for the dispersion of waves plane, the dependence on the coordinate 𝑦 is excluded by turning the
in a viscous and non-isothermal atmosphere can be obtained only in coordinate system around the axis 𝑧 so that the axis 𝑥 is directed along
the local approximation, assuming that the parameters of the medium the horizontal component of the particle velocity.
change rather slowly at the scales of the vertical wavelength. The We will consider the Earth’s thermosphere at heights above 200 km,
local dispersion equation of the AGW with allowance for viscosity and where the temperature does not change much with height, and one type
thermal conductivity was obtained and studied, for example, in Vadas of atmospheric gas predominates — atomic oxygen. In the equilibrium
and Fritts (2005). state, the unperturbed pressure and density change in the thermosphere
Rates of AGW attenuation in time due to molecular viscosity and with height according to the barometric law are 𝑃𝑃 (𝑧) = 𝜌𝜌(𝑧) =
thermal conductivity were obtained in Vadas and Fritts (2005) and ( ) (𝑧0 ) (𝑧0 )
𝑧−𝑧0
Lizunov and Leontiev (2014). In Lizunov and Leontiev (2014), the exp − 𝐻 , where 𝑧0 is a certain initial altitude level, 𝐻 = 𝑅𝑇 ∕𝑔 is the
attenuation rate was calculated based on the general hydrodynamic atmosphere scale height, 𝑅 = 𝑅0 ∕𝑀, 𝑅0 is the universal gas constant,
approach, which describes the dissipation of the energy of acoustic 𝑀 is the molar mass of the gas.
waves. The expression for attenuation rate obtained in Vadas and Fritts As is the practice, in the linearized Navier–Stokes and heat transfer
(2005) from a linearized system of hydrodynamic equations contains a equations, the transfer of energy and momentum by the wave is taken
paradoxical result. At wavelengths greater than a certain critical value, into account through the gradient of the perturbed velocity. However,
the attenuation is replaced by an increase in amplitude, which is im- as shown in the paper Lebedev-Stepanov (2007), in a medium with
possible in a dissipative atmosphere. In our opinion, this paradox is due a background density gradient, the contribution of this latter factor
to incomplete consideration of transport processes in the hydrodynamic to transfer processes must also be taken into account. In this paper,
description of AGW dissipation using the Navier–Stokes and heat trans- following the approach developed in Lebedev-Stepanov (2007) for
fer equations. Wave perturbations propagate in the atmosphere against flows with a velocity gradient, we take into account in Eq. (2) the
the background of a barometric density distribution that is established additional momentum transfer due to the vertical density gradient
upon diffusion equilibrium of the atmospheric gas in the gravitational 𝜕𝑉 𝜕𝜌𝑉
using the following replacement 𝜇 𝜕𝑥 𝑖 → 𝜈 𝜕𝑥 𝑖 in all components of the
field. Therefore, an additional transfer of momentum and energy by the 𝑗 𝑗

wave due to the background density gradient must occur, in addition viscous stress tensor. It is indicated here that 𝜈 = 𝜇∕𝜌, 𝜈 is the kinematic
to the usually taken into account velocity gradient (Lebedev-Stepanov, viscosity coefficient. We will make a similar replacement in the heat
𝑐 𝜈
2007). In the Navier–Stokes and heat transfer equations, which are transfer equation (3): 𝜒 𝜕𝑥 𝜕𝑇
→ 𝑃𝑝 𝑟 𝜕𝜌𝑇
𝜕𝑥
. The indicated substitutions are
𝑖 𝑖
obtained for the case of a compressible medium with a uniform density,
actually equivalent to introducing 𝜌 in the Navier–Stokes and heat
this effect is not taken into account (Lebedev-Stepanov, 2007).
transfer equations under the sign of the derivative. As a result, we
In this paper, we obtain the local dispersion equation of the AGW
obtain from Eqs. (1)–(3) the linearized equations for the perturbed
based on a modified system of linearized hydrodynamic equations that
quantities 𝑉𝑥 , 𝑉𝑧 , 𝜌′ ∕𝜌, 𝑇 ′ ∕𝑇 in an isothermal dissipative atmosphere:
additionally take into account the transfer of energy and momentum ( )
due to the background density gradient. From these equations, we 𝜕 𝜌′ 1 𝜕𝜌
+ 𝑉 + ∇ ⋅ 𝑉⃗ = 0, (5)
calculate the expressions for the attenuation rates in time and in 𝜕𝑡 𝜌 𝜌 𝜕𝑧 𝑧
space, which more adequately describe the attenuation of acoustic–
( ′ )
gravity perturbations in the entire allowable spectral range. Using the 𝜕𝑉𝑥 𝜕 𝜌 𝑇′
+ 𝑅𝑇 +
expression for spatial attenuation rate, we analyze the nature of the 𝜕𝑡 𝜕𝑥 𝜌 𝑇
AGW attenuation with height depending on their spectral properties [
𝜕 ⃗
and temperature in the thermosphere. −𝜈 𝛥𝑉𝑥 + 𝑒 ∇ ⋅ 𝑉
𝜕𝑥

2
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488
( )]
1 𝜕𝜌 𝜕𝑉𝑥 𝜕𝑉 𝑔𝑘2𝑥 𝛼 𝑘2𝑥 𝜈 2 𝑏5
+ + (𝑒 − 1) 𝑧 = 0, (6) 𝐹2 = 𝑐𝑆2 𝑞 2 𝑏2 + + 𝑞 4 𝜈 2 𝑏4 +
,
𝜌 𝜕𝑧 𝜕𝑧 𝜕𝑥 𝐻𝑞 2 𝐻2
( ′ ) 𝛾
𝑏1 = + 𝑒 + 2,
𝜕𝑉𝑧 𝜕 𝜌 𝑇′ 𝑇′ 𝑃𝑟
+ 𝑅𝑇 + −𝑔 ( 𝛾 ) 𝛾
𝜕𝑡 𝜕𝑧 𝜌 𝑇 𝑇 1
[ 𝑏2 = 1 + , 𝑏3 = (𝑒 + 1) +1 + ,
𝜕 𝑃𝑟 𝑃𝑟 𝑃𝑟
−𝜈 𝛥𝑉𝑧 + 𝑒 ∇ ⋅ 𝑉⃗ 𝑏4 =
𝛾
(𝑒 + 1), 𝑏5 =
𝛾
(𝑒 − 1).
𝜕𝑧
( )] 𝑃𝑟 𝑃𝑟
1 𝜕𝜌 𝜕𝑉𝑥 𝜕𝑉
+ + (𝑒 + 1) 𝑧 = 0, (7) The presence of the imaginary part in the dispersion equation (11)
𝜌 𝜕𝑧 𝜕𝑥 𝜕𝑧 indicates a change in the amplitudes of the perturbations. The inclusion
( ) of the effects of viscosity and thermal conductivity in the original
𝜕 𝑇′
+ (𝛾 − 1) ∇ ⋅ 𝑉⃗ system of hydrodynamic equations means taking into account irre-
𝜕𝑡 𝑇
[ ( ′) ( ′ )] versible losses of wave energy. Therefore, we must obtain a decrease
𝛾𝜈 𝑇 1 𝜕𝜌 𝜕 𝑇 in the amplitudes of the waves, i.e. attenuation in the entire spectral
− 𝛥 + = 0. (8)
𝑃𝑟 𝑇 𝜌 𝜕𝑧 𝜕𝑧 𝑇 range of their existence. In the following sections, using the dispersion
It has been taken into account that for an ideal gas 𝑃 = 𝜌𝑅𝑇 , 𝑒 = equation (11), we consider the nature of the AGW attenuation in time
1∕3+𝜀∕𝜇, 𝑃 𝑟 = 0.7 (Kundu, 1990). In the isothermal case, we neglected and space.
the terms ∼ 𝑑𝜇∕𝑑𝑧 because 𝜇 ∼ 𝑇 0.7 (Dalgarno and Smith, 1962).
It can also be put in the thermosphere 𝑒 = 1∕3 since for monatomic 4. AGW attenuation in time
gases 𝜀 = 0 (Landau and Lifshitz, 1987). The contribution of the
density gradient to the transfer of energy and momentum by wave Let us calculate the attenuation rate in time for AGW in a dissipative
disturbances is taken into account in Eqs. (6)–(8) using additional terms atmosphere. To do this, we make the following change in Eq. (11):
proportional to 𝜌1 𝜕𝜌
𝜕𝑧
. In the system of equations commonly used to 𝜔 → 𝜔 + 𝑖𝛿𝑡 , where 𝛿𝑡 is the attenuation rate, and leave in it only terms
analyze linear disturbances in the atmosphere, these terms are absent linear in the small parameter 𝛿𝑡 . As a result, we obtain the following
(see, for example, Kundu, 1990; Vadas and Fritts, 2005). equation:
( )
𝜈 2 𝑐𝑆2 𝑞 2 𝑘2
3. Dispersion equation of AGW in a dissipative atmosphere 𝜔4 − 𝜔2 𝐹1 + 𝑘2𝑥 𝑁 2 𝑐𝑆2 + 𝑞4 + 𝑥
𝑃𝑟 𝐻2
( )
Solutions for small wave disturbances in the atmosphere will be +𝜈𝑞 2 3𝜔2 𝑏1 − 𝐹2 𝛿𝑡
sought in the following form (Hines, 1960): [ ]
[ ( )] +𝑖𝜔 4𝜔2 𝛿𝑡 − 2𝛿𝑡 𝐹1 − 𝜔2 𝜈𝑞 2 𝑏1 + 𝜈𝑞 2 𝐹2 = 0. (12)
𝑉𝑥 , 𝑉𝑧 , 𝜌′ ∕𝜌, 𝑇 ′ ∕𝑇 ∼ exp (𝑧∕2𝐻) exp 𝑖 𝜔𝑡 − 𝑘𝑥 𝑥 − 𝑘𝑧 𝑧 , (9)
Equating the imaginary part of (12) to zero, we determine the attenu-
where 𝜔 is the frequency, 𝑘𝑥 and 𝑘𝑧 are the horizontal and vertical ation rate in time:
components of the wave vector, respectively. The multiplier exp (𝑧∕2𝐻)
𝜈𝑞 2 𝜔2 𝑏1 − 𝐹2
takes into account the change in the amplitudes of disturbances with 𝛿𝑡 = ⋅ . (13)
2 2𝜔2 − 𝐹1
height in the atmosphere with barometric vertical density stratifica-
[ ]
tion (Hines, 1960). Substituting Eq. (9) into the system of Eqs. (5)–(8), For the accepted form of perturbations, ∼ exp 𝑖(𝜔 + 𝑖𝛿𝑡 )𝑡 − 𝑖(𝑘𝑥 𝑥 + 𝑘𝑧 𝑧) =
( ) [ ]
as a result, we obtain the dispersion equation exp −𝛿𝑡 𝑡 exp 𝑖(𝜔𝑡 − 𝑘𝑥 𝑥 − 𝑘𝑧 𝑧) , the attenuation in time is obtained at
( )( 𝜈 2
) 𝛿𝑡 > 0. For acoustic waves, the numerator and denominator in Eq. (13)
𝑘2𝑥 𝑁 2 𝑐𝑆2 + 𝑐𝑆2 𝑞 2 𝑖𝜔 + 𝜈𝑞 2 𝑖𝜔 + 𝑞
𝑃𝑟 are always positive, and for gravity waves they are always negative.
( 𝛾𝜈 2 ) 𝑔𝑘𝑥 𝜈
2
Therefore, in both of these cases 𝛿𝑡 > 0, and the AGW amplitudes decay
+ 𝑖𝜔𝛼 + 𝑞
𝑃𝑟 𝐻 in time in a dissipative atmosphere.
( 𝛾𝜈 2 ) 𝑘𝑥 𝜈
2 2 For high-frequency acoustic waves from (13) we obtain 𝛿 ∼ 𝜔2 ,
+𝑖𝜔 𝑖𝜔 + 𝑞 (𝑒 − 1) for low-frequency gravity waves 𝛿 ∼ 𝑘2𝑥 ∕𝜔2 . Whence one can see a
𝑃𝑟 2
( ) ( 𝐻 𝛾𝜈 ) [ ] significant difference in the character of attenuation of these types of
+𝑖𝜔 𝑖𝜔 + 𝜈𝑞 2 𝑖𝜔 + 𝑞 2 𝑖𝜔 + (𝑒 + 1) 𝜈𝑞 2 = 0. (10)
𝑃𝑟 waves at high and low frequencies. Next, we will focus on the analysis
√ √
Here 𝑐𝑆 = 𝛾𝑅𝑇 is the speed of sound, 𝑁 = 𝑔 𝛾 − 1∕𝑐𝑆 is the of the attenuation of internal gravity waves, since waves of this type
Brunt–Väisälä frequency, 𝛾 = 𝑐𝑃 ∕𝑐𝑉 is the specific heats ratio, 𝑞 2 = were observed from a satellites and therefore it is possible to compare
𝑘2𝑥 + 𝑘2𝑧 + 1∕4𝐻 2 , 𝛼 = 2 − 𝛾 + 𝑒(𝛾 − 1). Since the kinematic viscosity our theoretical results with experimental data.
coefficient in the atmosphere varies with height, dispersion equation Substituting Eq. (13) into Eq. (12), we obtain the dispersion equa-
(10) is valid only in the local approximation. It is performed within thin tion of the AGW in an isothermal atmosphere with dissipation:
[
layers, where the value of 𝜈 can be considered approximately constant. ( )
This is true for waves with 𝑘𝑧 > 𝜈1 𝑑𝜈 ≈ 𝐻1 . When 𝜈 = 0 in expression 2 2𝜔2 − 𝐹1 𝜔4 − 𝜔2 𝐹1 + 𝑘2𝑥 𝑁 2 𝑐𝑆2
𝑑𝑧
( )]
(10), it exactly coincides with the dispersion equation of the AGW in 𝜈 2 𝑐𝑆2 𝑞 2 𝑘2
an isothermal atmosphere without dissipation (see, for example, Hines + 𝑞4 + 𝑥
𝑃𝑟 𝐻2
(1960), Yeh and Liu (1974)). We represent Eq. (10) in the following [ ]
equivalent form, where the real and imaginary parts are separately +𝜈 2 𝑞 4 3𝜔4 𝑏21 − 4𝜔2 𝑏1 𝐹2 + 𝐹22 = 0. (14)
distinguished:
( ) The kinematic viscosity coefficient 𝜈 varies in the thermosphere in
4 2 2 2 2
𝜈 2 𝑐𝑆2 𝑞 2 4
𝑘2𝑥 the interval of 105 −106 m2 s−1 (Vadas and Nicolls, 2012), depending on
𝜔 − 𝜔 𝐹1 + 𝑘𝑥 𝑁 𝑐𝑆 + 𝑞 +
𝑃𝑟 𝐻2 the height and level of solar activity. The dispersion dependences (𝜔, 𝑘𝑥 )
( ) with 𝑘𝑧 = 0 and 𝑘𝑧 = 1∕𝐻 for both the acoustic and gravity branches of
+𝑖𝜔𝜈𝑞 2 𝐹2 − 𝜔2 𝑏1 = 0. (11)
the AGW taking into account dissipation (𝜈 = 1 ⋅ 106 m2 s−1 ) are shown
Here in Figs. 1a–1b by solid curves. For comparison, the same dependences
𝑘2𝑥 𝜈 2 (𝑒 − 1) are shown without taking into account dissipation (dashed curves). It
𝐹1 = 𝑐𝑆2 𝑞 2 + 𝑞 4 𝜈 2 𝑏3 + , is seen that due to dissipation, the frequency of acoustic and gravity
𝐻2

3
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488

Fig. 2. Attenuation rate in time for gravity (solid curves) and acoustic AGW branches
(dashed curves): 𝜈 = 2 ⋅ 105 m2 s−1 (curves below) and 𝜈 = 1 ⋅ 106 m2 s−1 (curves above),
𝑘𝑧 𝐻 = 1.

addition, as will be shown below, the independently calculated atten-


uations in time and space make it possible to calculate a reasonable
value for the vertical group velocity in a dissipative atmosphere. In
turn, this will make it possible to understand how the rate of vertical
energy transfer in a viscous isothermal atmosphere changes.
Suppose the AGW attenuates in space in the vertical direction. To
calculate the attenuation rate in space, we make a replacement 𝑘𝑧 →
𝑘𝑧 + 𝑖𝛿𝑧 in Eq. (11), where 𝛿𝑧 is the spatial attenuation rate. Similarly to
the procedure for the attenuation rate in time described in the previous
section, we obtain the expression for the AGW attenuation rate with
height:

𝜈𝑞 2 𝜔 𝜔2 𝑏1 − 𝐹2
𝛿𝑧 = − ⋅ , (16)
2 𝑘𝑧 𝑐 2 𝜔2 + 2𝑞 2 𝜈 2 𝑏3 𝜔2 − 𝐹3
𝑆
Fig. 1. Dispersion dependences of AGW in an isothermal atmosphere, taking into
( )
account viscosity and thermal conductivity at 𝜈 = 8 ⋅ 105 m2 s−1 (solid curves) and 𝜈2 𝑐2 𝑘2
without dissipation (dashed curves): (a) 𝑘𝑧 𝐻 = 0; (b) 𝑘𝑧 𝐻 = 1. where 𝐹3 = 𝑃 𝑟𝑆 3𝑞 4 + 𝐻𝑥2 . We note that expression (16) must be
used with caution when 𝑘𝑧 → 0, making sure that the assumption of
[
𝛿𝑧 smallness is satisfied. For solution of the form ∼ exp 𝑖(𝜔𝑡 − 𝑘𝑥 𝑥) − 𝑖
] ( ) [ ]
waves decreases. With a viscosity coefficient 𝜈 = 1 ⋅ 106 m2 s−1 , the (𝑘𝑧 + 𝑖𝛿𝑧 )𝑧 = exp 𝛿𝑧 𝑧 × exp 𝑖(𝜔𝑡 − 𝑘𝑥 𝑥 − 𝑘𝑧 𝑧) , the attenuation condi-
frequency shift becomes noticeable for the values 𝑘𝑥 𝐻 > 4 (see 1a–1b). tion in space is satisfied if 𝛿𝑧 < 0.
This corresponds to horizontal wavelengths 𝜆𝑥 ⩽ 75 km for values The dependences of the spatial attenuation rate, 𝛿𝑧 , on the hor-
𝐻 = 50 km characteristic of the thermosphere. According to satel- izontal scale 𝑘𝑥 are shown in Figs. 3a–3b for different values of 𝑘𝑧
lite data, most AGW observed at altitudes of the thermosphere have and viscosity coefficient, 𝜈. From these figures, the general tendency
horizontal scales of ∼ 400 km or more (Johnson et al., 1995; Innis of viscous filtration of internal gravity waves with height is clearly
and Conde, 2002; Fedorenko et al., 2015). Therefore, the influence visible. At a fixed value of 𝑘𝑧 , the rate of wave spacial attenuation
of molecular viscosity and thermal conductivity on the shift of the rapidly increases with increasing coefficient 𝜈 (Fig. 3a), and at a
real part of the frequency in the thermosphere can be neglected. In fixed value of 𝜈 the attenuation increases with increasing 𝑘𝑧 (Fig. 3b).
this case, the value 𝑞 2 needed to calculate the attenuation rate (13) There is a minimum of spatial attenuation in the range of horizontal
can be approximately found from the usual dispersion equation of the scales 𝑘𝑥 𝐻 = 0.4 − 0.8. With increase of 𝜈, the horizontal wave scale
AGW (Hines, 1960; Yeh and Liu, 1974): corresponding to the minimum attenuation manifests itself more dis-
1 𝑁 2 𝜔2 tinctly (see Fig. 3a). These results are in good agreement with satellite
𝑞 2 = 𝑘2𝑥 + 𝑘2𝑧 + = 𝑘2𝑥 + . (15)
4𝐻 2 𝜔2 𝑐𝑆2 observations, indicating the predominance at altitudes of 250 − 450 km
of the polar thermosphere of large-amplitude quasi-horizontal AGW
The attenuation rates of the acoustic and gravity AGW branches at
with wavelength of ∼ 400 − 700 km (Johnson et al., 1995; Innis and
𝑘𝑧 𝐻 = 1 and two values 𝜈 = 2⋅105 m2 s−1 and 1⋅106 m2 s−1 are presented
Conde, 2002; Fedorenko et al., 2015). At an atmosphere scale height
in Fig. 2. It can be seen that the quantity 𝛿𝑡 increases rapidly with
𝐻 = 50 km, these wavelengths correspond to 𝑘𝑥 𝐻 = 0.45 − 0.78,
increasing viscosity coefficient and decreasing horizontal wavelength.
i.e. to the interval of minimum AGW attenuation in the dissipative
thermosphere. Apparently, under conditions of a viscous atmosphere
5. AGW attenuation in space
far from sources, only waves with the spectral properties precisely
corresponding to weak attenuation can be observed, while other AGWs
Usually researchers limit themselves to calculation of the time atten-
are rapidly absorbed. Let us obtain expressions for spatial attenuation
uation rate (see, for example, Vadas and Fritts, 2005; Vadas and Nicolls,
rate in some special cases. For internal gravity waves in the Boussinesq
2012; Lizunov and Leontiev, 2014). However, in our opinion, the atten-
approximation 𝑞 2 = 𝑘2𝑥 𝑁 2 ∕𝜔2 , and from Eq. (16) it follows:
uation of waves with height is more demonstrative than attenuation in
time. The spatial attenuation rate shows how the exponential growth 𝜈𝑞 2 𝜔 𝐹2
𝛿𝑧 = ⋅ . (17)
of the wave’s amplitude slows down with height due to dissipation. In 2 𝑘𝑧 𝑐 2 𝜔2 − 𝐹3
𝑆

4
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488

expressions (13) and (16), the terms ∼ 𝜈 2 can be neglected. In this


approximation we get:

𝛿𝑡 𝑘 𝑐𝑆2 𝜔2
𝑉𝑔𝑟𝑧 = − = 𝑧 . (20)
𝛿𝑧 𝜔 2𝜔2 − 𝑐 2 𝑞 2
𝑆
Eq. (20) exactly coincides with the expression for the vertical group ve-
locity 𝑉𝑔𝑟𝑧 in an isothermal atmosphere without dissipation (Kryuchkov
and Fedorenko, 2012). For low-frequency internal gravity waves it
follows from Eq. (20) that 𝑉𝑔𝑟𝑧 = −𝑘𝑧 𝜔3 ∕𝑘2𝑥 𝑁 2 , and for high-frequency
acoustic waves, that 𝑉𝑔𝑟𝑧 = 𝑘𝑧 𝑐𝑆2 ∕𝜔. The vertical component of the
group velocity 𝑉𝑔𝑟𝑧 determines the rate of energy transfer by the wave
in the vertical direction. In an isothermal atmosphere without dissipa-
tion, 𝑉𝑔𝑟𝑧 reaches a maximum with spectral parameters 𝑘𝑥 𝐻 ≈ 0.5 and
𝑘𝑧 𝐻 ≈ −0.2 (Kryuchkov and Fedorenko, 2012). This means that waves
with such spectral properties provide the most efficient energy transfer
in the atmosphere from sources located ‘‘below’’.
The rate of vertical energy transfer in a dissipative atmosphere can
be determined using expressions (13) and (16):
2 2 2 2 2
𝛿𝑡 𝑘 𝑐 𝜔 + 2𝑞 𝜈 𝑏3 𝜔 − 𝐹3
𝑉𝑔𝑟𝑧 = − = 𝑧 𝑆 . (21)
𝛿𝑧 𝜔 2𝜔2 − 𝐹1
Dependencies of 𝑉𝑔𝑟𝑧 on 𝑘𝑥 𝐻 without regard to viscosity and thermal
conductivity are shown in Fig. 4a by dashed curves.
It can be seen that with an increase in the angle of inclination of the
wave vector to the horizontal plane, the value of 𝑉𝑔𝑟𝑧 decreases from
the maximum value with a simultaneous shift of the maximum position
to the side of increase of 𝑘𝑥 𝐻. Note that in Fig. 4a, we deliberately
left the cases of that go beyond the locality of the approximation
in order to show the tendency for a rapid increase in attenuation
with increasing 𝑘𝑧 𝐻 and the presence of scales corresponding to the
minimum attenuation of waves. The dependences of the vertical group
velocities, taking into account the viscosity of 𝜈 = 6 ⋅ 105 m2 s−1 ,
calculated by the formula (21), are shown by solid curves in the same
Fig. 3. The spatial attenuation of internal gravity waves −𝛿𝑧 𝐻: (a) at 𝑘𝑧 𝐻 = −1 and
figure.
different values of 𝜈: 2⋅105 m2 s−1 , 4⋅105 m2 s−1 , 6⋅105 m2 s−1 , 8⋅105 m2 s−1 , 1⋅106 m2 s−1 ; The effect of viscosity on the expected value 𝑉𝑔𝑟𝑧 increases with
(curves are arranged from bottom to top); (b) for 𝜈 = 4 ⋅ 105 m2 s−1 and different values increasing parameter 𝑞 2 . The vertical group velocities calculated in two
of 𝑘𝑧 𝐻 ∶ −1, −1.2, −1.4, −1.6 (curves are arranged from bottom to top) ways more and more differ with increasing 𝑘𝑥 and |𝑘𝑧 |. The rate of
vertical energy transfer in a dissipative atmosphere also decreases with
height as the viscosity coefficient, 𝜈, increases. This can be seen from
For high frequency acoustic waves 𝑞𝐴𝐶 = 𝜔2 ∕𝑐𝑆2 . Then the expression Fig. 4b, where the dependences of 𝑉𝑔𝑟𝑧 on 𝑘𝑥 𝐻 at 𝑘𝑧 𝐻 = −1 are shown
for spatial attenuation rate takes the form: for different values of 𝜈.
( )
𝜈 𝜔3 𝛾 −1
𝛿𝑧𝐴𝐶 ≈ − +𝑒+1 . (18) 7. Change of AGW amplitude with height
2 𝑘𝑧 𝑐 4 𝑃𝑟
𝑆

In the limit of low-frequency gravity waves, the attenuation rate is The spatial rate of the AGW attenuation has a simple physical
equal to: meaning, which can be understood from the following considerations.
4 4 ( ) In a stratified atmosphere without dissipation, the AGW amplitude in-
𝜈 𝑘𝑧 𝑁 1
𝛿𝑧𝐺𝑅 ≈ 1+ . (19) creases exponentially with increasing height as the background density
2 𝑘𝑧 𝜔5 𝑃𝑟
decreases. When dissipation is taken into account, with an increase in
Note that the different signs in the attenuation rates of acoustic and height, the loss of wave energy due to viscosity and thermal conduc-
gravity waves reflect the difference in their physical nature. For upward tivity also increases simultaneously, since 𝜈(𝑧) increases exponentially
propagating acoustic waves 𝑘𝑧 > 0, and for internal gravity waves with height. As a result, the change in wave amplitudes with height in
𝑘𝑧 < 0. Therefore, the attenuation condition 𝛿𝑧 < 0 is satisfied according an isothermal dissipative atmosphere can be written as:
to expressions (18) and (19) for both types of waves during their [ ( ) ]
1
propagation in the atmosphere from the bottom up. 𝑉𝑥 , 𝑉𝑧 ∼ exp + 𝛿𝑧 (𝑧) 𝑑𝑧 . (22)
∫ 2𝐻
Since for the adopted type of solution we have 𝛿𝑧 < 0, it follows
6. Relationship between AGW attenuation rates in time and in from Eq. (22) that at a certain height 𝑧 = 𝑧𝑚𝑎𝑥 the wave amplitude
space will reach its maximum value, and will decrease above 𝑧𝑚𝑎𝑥 . The
attenuation rate, 𝛿𝑧 < 0, depends on the wave spectral properties
The description of wave attenuation in time and in space are equiv- and parameters of the medium; therefore, the attenuation heights of
alent ways, and the rates 𝛿𝑡 and 𝛿𝑧 must be connected through the different spectral components must differ. Intuitively, the maximum
vertical group velocity. For the characteristic in the thermosphere heights in a dissipative atmosphere are reached by spectral harmonics
values of the parameters: 𝑐𝑆 ≈ 700 − 1000 ms−1 , 𝐻 ≈ 40 − 50 km, 𝜈 ∼ propagating with a maximum value of velocity 𝑉𝑔𝑟𝑧 . This is confirmed
105 − 106 m2 s−1 and typical spectral characteristics of the AGW in by the simple model below.

5
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488

Fig. 4. Dependences of the vertical group velocity 𝑉𝑔𝑟𝑧 (m s−1 ) on the horizontal
scale 𝑘𝑥 𝐻: (a) for 𝜈 = 6 ⋅ 105 m2 s−1 and different values of the vertical components
𝑘𝑧 𝐻 = −0.2, −0.5, −1; (b) for 𝑘𝑧 𝐻 = −1 and different values of the viscosity coefficient
𝜈 = 4 ⋅ 105 m2 s−1 , 8 ⋅ 105 m2 s−1 , 1 ⋅ 106 m2 s−1 (curves are arranged from top to bottom).
Dashed curves - excluding dissipation.

Fig. 5. The change in the AGW amplitude with height depending on the normalized
frequency 𝜔∕𝑁 at 𝑘𝑥 𝐻 = 0.5, 𝑇 = 1000 K: (a) 𝜔∕𝑁 = 0.1...0.64; (b) 𝜔∕𝑁 = 0.64...080.
Consider the interval of heights of the thermosphere in 200−400 km
with 𝑇 = 𝑐𝑜𝑛𝑠𝑡. Suppose that at the lower boundary of this interval,
the auroral source generates AGW of unit amplitude. In the absence
height, as a result of which the AGW penetrate higher and reach large
of viscosity, the amplitudes of such waves, regardless of their spectral
properties, would increase in this interval by a factor 2 amplitudes. Fig. 6b shows the change in the amplitude of the spectral
( of 𝑒 ≈ 7.3) when component 𝑘𝑥 𝐻 = 0.5, 𝑘𝑧 𝐻 = −1 with height, at temperatures of the
ℎ = 50 km. Let us set the dependence 𝜈(𝑧) = 𝜈0 exp (𝑧 − 𝑧0 )∕𝐻 so that
𝜈 = 1 ⋅ 106 m2 s−1 at an altitude of about 300 km, which approximately thermosphere: 𝑇 = 800 K, 1000 K and 1300 K, characteristic of low,
corresponds to the average level of solar activity. Let us calculate how medium, and high solar activity, respectively. It is seen that under
the AGW amplitudes change with height using Eqs. (16) and (22). conditions of high solar activity, AGW can propagate much higher
Dependences of the AGW amplitude on the normalized frequency 𝜔∕𝑁 in a dissipative atmosphere than with low activity. Also, with low
at 𝑘𝑥 𝐻 = 0.5 are shown in Fig. 5. With increasing frequency, the solar activity, the AGW amplitude in the thermosphere is, on average,
attenuation decreases, reaching a minimum at 𝜔 ≈ 0.64𝑁 (Fig. 5a). smaller due to greater attenuation.
With a further increase in frequency, the attenuation increases again
(Fig. 5b). Thus, a minimum of AGW attenuation with frequency is 8. Discussion
observed.
Dependences of amplitudes on height for 𝑘𝑥 𝐻 = 0.5 and sev-
For several years, the authors of this work had been analyzing the
eral values of 𝑘𝑧 𝐻 are shown in Fig. 5. It can be seen that in the
properties of AGWs at heights of the thermosphere from measurements
thermosphere it is waves with 𝑘𝑧 𝐻 ≈ −0.2, 𝑘𝑥 𝐻 = 0.5 reach the
on the Dynamics Explorer 2 satellite (Fedorenko and Kryuchkov, 2011;
maximum heights and amplitudes, and these conditions correspond to
Fedorenko et al., 2015). As a result, a number of observed features
the maxim of 𝑉𝑔𝑟𝑧 . The wave amplitudes with |𝑘𝑧 𝐻| > 1 increase in the
dissipative thermosphere with an average solar activity of no more than were revealed that did not fit into the framework of the standard AGW
2 times. This is consistent with satellite measurement data, indicating theory developed for an isothermal atmosphere without dissipation. We
the absence of a pronounced dependence of the AGW amplitudes on then considered that it would be more consistent to complicate the
altitude in the range of 250 − 450 km (Fedorenko et al., 2015). isothermal model by introducing viscosity than to abandon this model
In the viscous thermosphere, the AGW amplitudes and the maxi- and look for a suitable one among the abundant set of non-isothermal
mum heights of their penetration are strongly dependent on temper- models. In doing so, we proceeded from the following results. First,
ature. With increasing temperature, 𝜈(𝑧) changes more slowly with observations indicate the predominance of waves of certain spatial and

6
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488

associated only with the velocity gradient, both in analytical consid-


eration and in numerical simulation. For the first time, to take into
account the damping of small wave disturbances in an isothermal
atmosphere, we used a system of the form (5)–(8), where, in addition
to the usual terms, the terms responsible for the transfer of energy
and momentum due to the background density gradient are added.
We analytically considered AGW attenuation based on (5)–(8) and
obtained new expressions for the wave attenuation rates 𝛿𝑡 and 𝛿𝑧 .
The difference between these expressions and those obtained earlier by
other authors (Vadas and Fritts, 2005; Vadas and Nicolls, 2012; Lizunov
and Leontiev, 2014) is as follows: (1) the expressions are obtained
taking into account the additional transfer due to the background
density gradient; (2) they are valid for the entire frequency range of
AGW, including the gravity and acoustic regions. For example, in Vadas
and Fritts (2005) and Lizunov and Leontiev (2014), attenuation rates
in time were obtained only for low-frequency gravity waves, in the
Boussinesq approximation.
Taking into account that additional terms in the system of Eqs. (5)–
(8) led not only to an increase in the total attenuation, but, most
importantly, to the elimination of the paradoxical amplification of
amplitudes at large wavelengths, which follows from the system of hy-
drodynamic equations in its commonly used form. For a more accurate
study of the dissipation of AGWs in the atmosphere, it is necessary
to carry out numerical modeling. However, when modeling, it is also
necessary to take into account the additional terms given in (5)–(8),
proportional to (1∕𝜌)(𝑑𝜌∕𝑑𝑧).
We compared the observational data obtained earlier from satellite
measurements with the theoretical results obtained in this work on the
features of the attenuation of AGWs with different spectral properties.
AGW with horizontal scales of 400–700 km prevailing in satellite
measurements at thermosphere heights approximately correspond to
the minimum attenuation of waves in space (see Fig. 3). At the same
wave scales, the rate of vertical energy transfer is close to maximum
(Fig. 4), i.e. such waves most efficiently transfer energy along the
vertical and provide coupling between different altitude levels of the
atmosphere. With an increase in the viscosity coefficient, the rate of en-
ergy transfer along the vertical decreases (Fig. 4b). The observed AGW
frequencies near the BW frequency also correspond to the minimum of
wave attenuation (Fig. 5). In the future, it is of considerable interest to
Fig. 6. The change in the AGW amplitudes with height in an isothermal at- compare the statistical features of the AGW amplitudes during periods
mosphere with dissipation: (a) at 𝑘𝑥 𝐻 = 0.5 and different values of 𝑘𝑧 𝐻 = of high and low solar activity.
−0.2, −0.5, −0.8, −1, −1.2, −1.4; (b) at 𝑘𝑥 𝐻 = 0.5, 𝑘𝑧 𝐻 = −1 and different values of the
Let us compare the expression we obtained for the temporal rate of
temperature of the thermosphere: 𝑇 = 800 K, 1000 K and 1300 K.
AGW attenuation with the results of other authors (Vadas and Fritts,
2005; Lizunov and Leontiev, 2014). Since the AGW attenuation rates
in these works were calculated in the Boussinesq approximation, for
temporal scales in the thermosphere. Thus, the frequency estimates comparison we will also restrict ourselves to this approximation. Taking
for large-amplitude AGWs observed in the polar thermosphere accord- into account the AGW dissipation using the Navier–Stokes and heat
ing to the data of different authors (Johnson et al., 1995; Innis and
transfer equations leads to the expression for the attenuation rate in
Conde, 2002; Fedorenko and Kryuchkov, 2011) give values close to
time with the following form (Vadas and Fritts, 2005):
the boundary frequency of the Brunt–Väisälä for the gravity branch.
In this case, the horizontal wavelengths belong to the interval 400– 1 2𝑘𝑧 𝜈
( ) 1+ +
700 km (Johnson et al., 1995; Fedorenko and Kryuchkov, 2011). Note 𝜈 2 1 𝑃 𝑟 𝜔𝐻𝑃 𝑟 .
𝛿𝑡𝑉 = 𝑘 − ⋅ (23)
that in satellite measurements, the frequency is determined in the ref- 2 4𝐻 2 𝑘 𝜈 ( 1
)
1+ 𝑧 1+
erence frame of the medium, in contrast to ground-based observations, 2𝜔𝐻 𝑃𝑟
where the directly measured frequency is shifted due to the Doppler In a wide range of spectral parameters and characteristic conditions in
effect by the presence of background winds at the altitudes of AGW the thermosphere, one can approximately write:
propagation. The frequency values obtained in the reference frame of ( )( )
𝜈 2 1 1
the medium can be used to estimate other spectral characteristics using 𝛿𝑡𝑉 ≈ 𝑘 − 1+ . (24)
2 4𝐻 2 𝑃𝑟
dispersion relations. As follows from the dispersion equation, AGWs
with the above properties are characterized by a small inclination of the It follows from Eqs. (23) and (24) that at large values of 𝑘2 = 𝑘2𝑥 + 𝑘2𝑧
wave vector to the horizontal plane. Also, by the satellite observations (small wavelengths), the attenuation increases with decreasing their
in the thermosphere, there is no exponential dependence of AGW wavelength, which is consistent with the general hydrodynamic con-
amplitudes on altitude, as it should be in the atmosphere without cepts (Landau and Lifshitz, 1987). However, according to this expres-
attenuation (Fedorenko et al., 2015). The inability to explain these sion, the AGW amplitudes with 𝑘2 < 1∕4𝐻 2 should increase in a
observed features prompted us to consider how dissipation affects the dissipative atmosphere, which is impossible.
propagation and properties of AGWs in the thermosphere. In our work, the attenuation rate, 𝛿𝑡 , was calculated based on the
When studying the viscous damping of waves in the atmosphere, local dispersion equation (11), obtained from a linearized system of hy-
one usually takes into account the transfer of energy and momentum drodynamic equations, taking into account the additional contribution

7
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488

9. Conclusion

The dissipation of acoustic–gravity waves in the thermosphere due


to viscosity and thermal conductivity is studied based on a modified
system of hydrodynamic equations. A modification of the system of
equations consists in the inclusion in the linearized Navier–Stokes
and heat transfer equations of special terms describing the transfer
of momentum and energy due to the density gradient, in addition to
the usually taken into account the velocity gradient. (see Eqs. (5)–(8)).
Using these hydrodynamic equations, the local dispersion equation for
acoustic–gravity waves, Eq. (10), in an isothermal dissipative atmo-
sphere is obtained. An analysis of this dispersion equation showed that
viscosity and thermal conductivity practically do not affect the real part
of the AGW frequency under typical conditions in the thermosphere.
In this work, a general expression is obtained for the AGW attenua-
tion rate in time, Eq. (13), which describes the attenuation of acoustic
and gravity waves in the entire spectral range. The value 𝛿𝑡 increases
with increasing viscosity coefficient 𝜈 and with decreasing wavelength.
A general expression is also obtained for the AGW attenuation rate
Fig. 7. AGW attenuation rates in time at 𝜈 = 1 ⋅ 106 m2 s−1 according to different
in space, Eq. (16). From it follows the expression for the attenuation
authors: 1 - 𝛿𝑡 in this work; 2 - 𝛿𝑡𝑉 according to Vadas and Fritts (2005); 3 - 𝛿𝑡𝐿
according to Lizunov and Leontiev (2014). rates in particular cases of internal gravity waves in the Boussinesq ap-
proximation and sound waves. The rate of the AGW spatial attenuation,
like the time attenuation rate, rapidly increases with increasing coeffi-
of the density gradient to transport processes. From the general expres- cient 𝜈 and magnitude of 𝑘𝑧 . Waves with a small angle of inclination
sion for attenuation rate, Eq. (13), in the Boussinesq approximation, we of the wave vector to the horizontal plane attenuate much less. In the
obtain: interval of horizontal scales 𝑘𝑥 𝐻 = 0.3 − 0.8, a minimum of spatial
( ) attenuation of the AGW is observed.
1 𝑘2 𝑔𝛼 𝜈 2 𝑏5
1+ + 𝑥 + The relationship between the temporal and spatial attenuation rates
𝜈𝑞 2 𝑃 𝑟 𝑐 2 𝑞 2 𝐻𝑞 2 𝐻2
𝑆
𝛿𝑡 = ⋅ , (25) of AGW is analyzed. The rate of energy transfer in the vertical di-
2 𝑘2 𝜈 2 (𝑒 − 1)
1+ 𝑥 rection in a dissipative atmosphere is calculated using the quantities
𝐻 2 𝑐𝑆2 𝑞 2 𝛿𝑡 and 𝛿𝑧 . It is shown that the value of the vertical group velocity
where 𝑞 2 = 𝑘2 +1∕4𝐻 2 . For internal gravity waves in the thermosphere, 𝑉𝑔𝑟𝑧 decreases with height in a viscous atmosphere as the viscosity
the following approximate expression will be valid: coefficient 𝜈 increases. Using the expression for spatial attenuation
( )( ) rate (16), the character of AGW attenuation with height is analyzed
𝜈 2 1 1
𝛿𝑡 ≈ 𝑘 + 1+ . (26) depending on their spectral properties and at different temperatures in
2 4𝐻 2 𝑃𝑟
It is seen that an additional inclusion of the density gradient in the the thermosphere corresponding to different levels of solar activity.
transfer of energy and momentum by the wave eliminates the paradox The results obtained are consistent with satellite observations of
that arises in the attenuation rate 𝛿𝑡𝑉 . Instead of the factor 𝑘2 − 1∕4𝐻 2 the AGW at the heights of the thermosphere. The predominance of
present in the expressions for attenuation rates (23) and (24), we quasi-horizontal AGWs with horizontal wavelengths of ∼ 400 − 700 km
got a factor 𝑘2 + 1∕4𝐻 2 . As a result, Eqs. (25) and (26) describe the in satellite data at altitudes of 250 − 450 km can be explained by
attenuation of acoustic–gravity perturbations in the entire allowable filtering the wave spectrum due to viscosity and thermal conductivity.
spectral range. As a result, far from sources, only AGWs with spectral properties that
In the paper Lizunov and Leontiev (2014), the time attenuation rate,
correspond to minimal attenuation can be observed.
𝛿𝑡𝐿 , was calculated on the basis of a completely different approach,
namely, based on general hydrodynamic ideas about the dissipation
of the energy of acoustic waves under the influence of viscosity and CRediT authorship contribution statement
thermal conductivity. Additionally, polarization relations of the AGW
theory, Hines (1960), were also used. The attenuation rate obtained in
this way has the form (Lizunov and Leontiev, 2014): A.K. Fedorenko: Conceptualization, Methodology, Drafting the ar-
ticle. E.I. Kryuchkov: Methodology, Software. O.K. Cheremnykh:
𝛿𝑡𝐿 = 𝑎𝜈𝑘2 , (27)
Conceptualization, Methodology. Y.A. Selivanov: Conceptualization,
where 𝑎 = 0.9 − 0.95 at 𝑘 = (∞, 0) for a monatomic gas. Drafting the article, Manuscript preparation.
The results of comparing the temporal attenuation rates obtained by
different authors are shown in Fig. 7. The figure shows the dependences
of attenuation rates on the magnitude of the wave vector 𝑘 according Acknowledgments
to expressions (24), (26) and (27). Since the approximate expressions
(24) and (26) coincide up to a numerical factor, the curves 𝛿𝑡 and 𝛿𝑡𝑉
This research was supported by the National Research Foundation
go parallel to each other. The curve 𝛿𝑡𝐿 has a functionally different
form, but quantitatively well agrees with 𝛿𝑡 and 𝛿𝑡𝑉 at relatively small of Ukraine under grant 2020.02/0015 ‘‘Theoretical and experimental
values of 𝑘𝐻 < 2. A comparison with the data of the work (Lizunov studies of global perturbations of natural and artificial origin in the
and Leontiev, 2014) is also interesting from the point of view of the Earth–Atmosphere–Ionosphere system’’, and the Targeted integrated
feasibility of the local approximation used to obtain the attenuation program of the National Academy of Sciences of Ukraine on Space
rates 𝛿𝑡 and 𝛿𝑡𝑉 . Research.

8
A. Fedorenko et al. Journal of Atmospheric and Solar-Terrestrial Physics 212 (2021) 105488

References Kryuchkov, E., Fedorenko, A., 2012. Peculiarities of energy transport in the atmosphere
by acoustic gravity waves. Geomagn. Aeron. 52, 2. http://dx.doi.org/10.1134/
Beer, T., 1974. Atmospheric Waves. Adam Hilger, London, London, http://dx.doi.org/ s0016793212010057.
10.1063/1.3069210. Kshevetskii, S.P., 2001. Analytical and numerical investigation of nonlinear internal
Belashov, V.Y., 1990. Dynamics of nonlinear internal gravity waves at ionosphere gravity waves. Nonlinear Process. Geophys. 8 (1/2), 37–53. http://dx.doi.org/10.
F-region heights. Geomagn. Aeron. 30, 536–538. 5194/npg-8-37-2001.
Cheremnykh, O.K., Selivanov, Y.A., Zakharov, I.V., 2010. The influence of compressibil- Kundu, P., 1990. Fluid Dynamics. Elsevier, New York.
ity and non-isothermality of the atmosphere on the propagation of acoustic-gravity Ladikov-Roev, Y., Cheremnykh, O., 2010. Mathematical Models of Continuous Media,
waves. Kosm. Nauka Tehnol. 16 (1), 9–19. http://dx.doi.org/10.15407/knit2010. Vol. 552. Naukova Dumka, Kyiv.
01.009. Landau, L.D., Lifshitz, E.M., 1987. Fluid Mechanics. In: Theoretical Physics, vol. 6,
Dalgarno, A., Smith, F.J., 1962. The thermal conductivity and viscosity of atomic Pergamon, London.
oxygen. Planet. Space Sci. 9 (1–2), 1–2. http://dx.doi.org/10.1016/0032-0633(62) Lebedev-Stepanov, P., 2007. Flow of a viscous compressible medium: Invalidity of the
90064-8. Navier-Stokes equation. Dokl. Phys. Chem. 417 (2), 319–324. http://dx.doi.org/10.
Dudis, J.J., Reber, C.A., 1976. Composition effects in thermospheric gravity waves. 1134/S0012501607120019.
Geophys. Res. Lett. 3 (12), 727–730. http://dx.doi.org/10.1029/gl003i012p00727. Lizunov, G.V., Leontiev, A.Y., 2014. Height of the penetration into the ionospherte
Fedorenko, A., 2009. Determination characteristics of atmospheric gravity waves in the for internal atmosphere gravity waves. Kosm. Nauka Tehnol. 20 (4(89)), 31–41.
polar regions using mass-spectrometer satellite measurements. Radio Phys. Radio http://dx.doi.org/10.15407/knit2014.04.031.
Astron. V 14 (N3), 254–265. Mayr, H., Harris, I., Herrero, F., Spencer, N., Varosi, F., Pesnell, W., 1990. Thermo-
Fedorenko, A., Bespalova, A., Cheremnykh, O., Kryuchkov, E., 2015. A dominant spheric gravity waves: observations and interpretation using the transfer function
acoustic-gravity mode in the polar thermosphere. Ann. Geophys. 33, 101–108. model (TFM). Space Sci. Rev. 54 (3–4), 297–375. http://dx.doi.org/10.1007/
http://dx.doi.org/10.5194/angeo-33-101-2015. bf00177800.
Fedorenko, A., Kryuchkov, Y., 2011. Distribution of medium scale acoustic gravity Meng, X., Vergados, P., Komjathy, A., Verkhoglyadova, O., 2019. Upper atmospheric
waves in polar regions according to satellite measurement data. Geomagn. Aeron. responses to surface disturbances: An observational perspective. Radio Sci. 54 (11),
(Engl. Transl.) 51, 520–533. http://dx.doi.org/10.1134/s0016793211040128. 1076–1098. http://dx.doi.org/10.1029/2019rs006858.
Fedorenko, A., Kryuchkov, Y., 2014. Observed features of acoustic gravity waves in Nekrasov, A., Shalimov, S., Shukla, P., Stenflo, L., 1995. Nonlinear disturbances in the
the heterosphere. Geomagn. Aeron. 54 (1), 109–116. http://dx.doi.org/10.1134/ ionosphere due to acoustic gravity waves. J. Atmos. Terr. Phys. 57 (7), 732–742.
s0016793214010022. http://dx.doi.org/10.1016/0021-9169(94)00052-p.
Francis, S., 1975. Global propagation of atmospheric gravity waves: A review. J. Atmos. Piani, C., Durran, D., Alexander, M., Holton, J., 2000. A numerical study of three-
Terr. Phys. 37, 1011–1054. http://dx.doi.org/10.1016/0021-9169(75)90012-4. dimensional gravity waves triggered by deep tropical convection and their role in
Fritts, D., Vadas, S., 2008. Gravity wave penetration into the thermosphere: sensitivity the dynamics of the QBO. J. Atmos. Sci. 57 (22), 3689–3701. http://dx.doi.org/
to solar cycle variations and mean winds. Ann. Geophys. 26, 3841–3861. http: 10.1175/1520-0469(2000)057<3689:ansotd>2.0.co;2.
//dx.doi.org/10.5194/angeo-26-3841-2008. Roy, A., Roy, S., Misra, A., 2019. Dynamical properties of acoustic-gravity waves in
Hines, C., 1960. Internal gravity waves at ionospheric heights. Can. J. Phys. 38, the atmosphere. J. Atmos. Sol.-Terr. Phys. 186, 78–81. http://dx.doi.org/10.1016/
1441–1481. j.jastp.2019.02.009.
Huang, K., Zhang, S., Yi, F., Huang, C., Gan, Q., Gong, Y., Zhang, Y., 2014. Nonlinear Stenflo, L., Shukla, P., 2009. Nonlinear acoustic-gravity waves. J. Plasma Phys. 75 (6),
interaction of gravity waves in a nonisothermal and dissipative atmosphere. Ann. 841–847. http://dx.doi.org/10.1017/S0022377809007892.
Geophys. 32, 263–275. http://dx.doi.org/10.5194/angeo-32-263-2014. Vadas, S., Fritts, M., 2005. Thermospheric responses to gravity waves: Influences of
Innis, J., Conde, M., 2002. Characterization of acoustic-gravity waves in the upper ther- increasing viscosity and thermal diffusivity. J. Geophys. Res. 110 (D15), 15103.
mosphere using dynamics explorer 2 wind and temperature spectrometer (WATS) http://dx.doi.org/10.1029/2004JD005574.
and neutral atmosphere composition spectrometer (NACS) data. J. Geophys. Res. Vadas, S., Nicolls, D., 2012. The phases and amplitudes of gravity waves propagating
107, http://dx.doi.org/10.1029/2002JA009370. and dissipating in the thermosphere: Theory. J. Geophys. Res. 117 (A5), 30.
Johnson, F., Hanson, W., Hodges, R., Coley, W., Carignan, G., Spencer, N., 1995. Gravity http://dx.doi.org/10.1029/2011JA017426.
waves near 300 km over the polar caps. J. Geophys. Res. 100, 23993–24002. Yeh, K., Liu, C., 1974. Acoustic-gravity waves in the upper atmosphere. Rev. Geophys.
http://dx.doi.org/10.1029/95ja02858. Space Phys. 12 (2), 193–216. http://dx.doi.org/10.1029/rg012i002p00193.
Kaladze, T., Pokhotelov, O., Shah, H., Khan, M., Stenflo, L., 2008. Acoustic-gravity Zhang, S., Yi, F., 2002. A numerical study of propagation characteristics of gravity wave
waves in the Earth’s ionosphere. J. Atmos. Sol.-Terr. Phys. 1607–1616. http: packets propagating in a dissipative atmosphere. J. Geophys. Res. 107 (D14), 32.
//dx.doi.org/10.1016/j.jastp.2008.06.009. http://dx.doi.org/10.1029/2001jd000864.

You might also like