3. Phenomenological description of a meteor.

advertisement
1
Доклад на Ассамблее : COSPAR 04-A-00016: B1.4/C1.4-0012-04 ORAL
Опубликовано: Advances in Space Research Vol 39, Number 4, 2007
«Meteors: Quantitative Spectroscopy»
Guest Editor: Peter Jenniskens
PECULARITIES OF METEOR RADIATION
Vladimir A. Smirnov
Academy of Communication of Ukraine
smirnovw@tm.odessa.ua
ABSTRACT
In this article the characteristics of meteor’s radiation are investigated.
The character of a radiating spectrum depends first of all on the speed of a
meteor, and also on the chemical and mechanical structure of the meteoroid.
The spectrum changes along the path of a meteor in the atmosphere
according to the law of Wien. For fast meteors, terminal flares are often
observed, which are formed due to the interaction of meteoric plasma with
the particles in the atmosphere. Such features present an opportunity for the
meteor plasma to radiate even when the speed is more than 60 km/s by the
mechanism of cascade radiation, similar to the radiation of gas-dynamic
lasers.
Keywords: Spectra of a meteors; Spectrophotometrical analysis;
Physical theory; Intensity; Radiation efficiency; Kinetic equation; Flares;
Laser.
1.History of meteor spectroscopy
The second son of J. Herschel and the grandson of W. Herschel with a Russian name
Aleksander Stevat Herschel (1836 - 1907) in 1863 with the help of binoculars, supplied by
prisms designed by D. Brauning, first constructed a meteor spectroscope.The field of vision of
his spectroscope was equal to 20 degrees. On, January 18, 1864, A. S. Herschel for the first
time observed a meteor spectrum (Herschel, 1866).
The initial research of meteor’s spectra was a period of visual observations by A.S.
Herschel, followed by the first photographic spectrograms obtained by S. N. Blajko in 1904
and 1907 (Smirnov, 1994).
The results of subsequent spectroscopic studies were reviewed
in the early ‘60s by Peter Millman (1906 - 1990) in his work "The General Survey of Meteor
Spectra".
2
Millman, originally from a missionary family, took after his parents the "mission" – to
organize the development of spectral meteor research in various countries. Besides his
regularly publishing of lists of meteor spectrograms made all over the world, he also dispatched
his works to many young researchers and was their teacher. Millman, for example, inspired my
own 1962 publication on the photometry of two meteor spectra (Smirnov 1962).
Millman’s classification of spectral photographs in classes “W”, “X”, “Y”, and “Z” had a
certain physical
sense: the spectra of a Y-type differed by including the radiation of ionized calcium in H and
K-lines. In spectra of Z-type were intensive lines of the sodium D-line and of neutral
magnesium. In spectra of Z-type, the lines of neutral iron and chromium were the most
intensive. The spectra of Y and X-types by their structure corresponded to that expected for the
stony meteorites called chondrites. The Z-type and W-type corresponded to spectra expected
from iron meteorites having an asteroid’s origin {Millman & Cook, 1955, Millman, 1963). Of
all known spectra, 68% were from meteor showers: among them 31 % were Perseids, 14 % Leonids, and 7 % - on Geminids.
The meteors of the same shower have the same velocity and, probably, a similar chemical
structure, determined by a common cosmogenic origin. As a result, the spectra of showers are
like a "family", having a similar character of radiation.
The spectrum of a meteor is not only a function of a chemical structure, but
also of velocity, or rather, the meteoroids kinetic energy. The spectrum of high-speed streams
of Perseids, Orionids, and Leonids all belonged to Y-type. The meteors of low-speed streams Geminids and Giacobinids – all belonged to X-type. Lines of more than 120 multiplet's of iron
were identified, as well as 10 multiplet’s of ionized iron, magnesium, calcium, and oxygen. In
addition, bands of molecular nitrogen and lines of atomic nitrogen and ionized nitrogen were
identified. Meteoroids contained a significant abundance of neutral sodium, chromium,
manganese, nickel, cobalt, carbon, lithium, barium, strontium, and hydrogen, as well as ions of
magnesium, oxygen, silicon, and aluminum. More than 10 band heads were recognized from
bands of FeO, MgO, CaO, CN, C2, CO, CH, CO  etc.
The application of methods of star photometry and spectrophotometry to the meteor
Spectra, and especially to photographic traces of integrated light, was burdened by plenty of
measurement and observational errors. This was considered in detail in Smirnov (1994).
In one of the first works of the author, devoted to the photometry of meteor traces "About
the photometric effect of a meteor being more dispersed than a point source" the analysis of
photometric profiles of the photographic traces of moving stars and meteors was carried out. It
was shown that the photometric profile of a meteor image is wider and fainter than that of the
stars. This made it necessary in the reduction of photographed meteor traces, to not consider the
meteor a point-image, but to consider the real trace caused by a certain volume and geometrical
size. In that respect, a meteor trace differs from that of a star of the same brightness. The same
concerns also the reduction of meteor spectra, where each "spectral line" represents a
monochromatic image of a meteor.
In the method of Spectro-photometrical analysis of a spectrogram by A. Cook and P.
Millman (1955)
, the densities of radiating atoms and ions were correctly determined from absolute
spectrophotometry of the monochrome images of "spectral bands" and "lines". However, the
use of a curve of growth construction, and defining the self-absorption in a volume of meteoric
3
radiation by means of photometric profiles of the same imaginary (infinitely narrow) "lines", is
not correct.
Application to the photometrically integrated
meteor images is even worse. Here the photometric mistakes are caused not only by that
mistake, but alsow because for the curve of growth construction it is necessary to enter the
photometric corrections caused by deviations from the law of Bunser-Roscoe.
The regular mistake within the construction of an integrated photometric curve of growth
turns out to be caused by
the radiation in the violet side of the spectrum to be underestimated because of the drop in film
sensitivity of the photographic panchromatic emulsion as a function of exposure. When this
was not correctly taken into account, photographic experimental data were in doubt.
The use of electronic CCD-devices do not have this problem and will allow measuring the
photometric brilliance of fixed objects
at once in absolute power units.
The detection of meteors and meteor spectra with orthicons and superorthicons was
started in the ‘60s. More recent use of supersensitive all-sky video camera systems has allowed
to record thousands of spectra and meteor traces.
During the 1998-2001 Leonid meteor storm campaigns, the high-sensitivity video-devices
were widely applied. , That included meteor observations from space, among others from the
destroyed Soviet station "Mir" (World), albeit that no meteor spectra were measured. .
2. The physical theory of meteors in historic context
A theory of meteor radiation was developed that could not explain all peculiarities of meteor
luminescence without contradictions. The equations of the the physical theory of meteors were
formulated by J.Hoppe, and in the former Soviet Union by B. Levin (1958).
A correct solution of the equations of meteor physics clearly requires the the considertation
of both spatial and temporal changes (Smirnov, 1995). Otherwise conservation laws are no
longer obeyed.
As is well known, the the physical theory of meteors is described starting from the
following considerations. The meteoroid (with unknown shape and chemical composition)
meets a certain amount of air mass
dma  S (t )  ( H )V (t )dt ,
(1)
during the time interval dt, where S(t) - cross-section of meteoroid, (H) is the atmospheric
density at altitude H, and v(t) is the velocity of the meteoroid. Multiplying the equation with V,
we obtain an expression for the air mass stream passing over the meteoroid:
vdma  Sv 2 dt .
(2)
If we neglect the additional moment imposed by the gravity of the Earth and the back-wards
directed moment due to the mass losses of the meteoroid, the motion of the meteoroid with
mass m  is described as
m dv   ГSv 2 dt ,
(3)
4
where  is the coefficient of resistance, characterizing the momentum transfer to the
atmospheric particles passing the meteoroid.
Then, the “equation of braking “ of the physical theory of meteors is represented by the
following formula:
dv
m
  ГSv 2 .
(4)
dt
In the last equation all the parameters except the empirically determined velocity and
accelerations, are unknown. Also the laws of evolution of these parameters during a gradual
penetration of the meteor into denser layers of the atmosphere is unknown. Apart from this, the
meteoroid mass m↓ is supposed to be independent from the time t in disagreement with any
model of the evolution of the meteor phenomenon.
Equation (4) was used again for the determination of the “dynamic masses" of meteors
(Bronshten, 1981). However, V.A.Bronshten (1918 – 2004) noted in his studies that the timedependent value of the velocity can be used to determine the instantaneous meteoroid mass. In
other words, it will permit the application of equation (4) to the case of a stationary
(“instantaneous") meteor mass. Obviously, the application of the “equation of braking” of the
meteor stipulates a stationary “dynamical” mass.
In the same manner, the application of the equation of mass loss below (7) stipulates the
condition of a stationary velocity. Indeed, according to the law of energy conservation, the
kinetic energy of the air mass encountered by the meteoroid is
1 2
1
v dm a  Sv 3 dt
2
2
(5)
Introducing a heat transfer coefficient  characterizing the conversion of kinetic energy into
heat, and the specific heat of the meteoroid Q, we obtain
1
Qdm    Sv 3 dt .
(6)
2
The mass-loss equation then turns into
dm
1
  Sv 3
(7)
dt
2
Evidently, the last equation is written down for the condition of constant velocity V, which
disagrees with the actual flight of the meteoroid. Here, the mass loss in turn defines the
instantaneous value of the velocity.
In the third equation of physical theory – the "equation of emission", - the number of
unknown quantities is smaller than in the previous ones, (4) and (7). This is why its application
is less complicated. Photometric data deliver the observed intensity of meteor radiation being
related with the loss of kinetic energy through
dW
mv 2
I  
, where
W
, i.e. assuming v as independent from t, we
dt
2
obtain
Q
5
I  
dm v 2
,
dt 2
(8)
where coefficient  characterizes the fraction of kinetic energy of the meteoroid converted into
radiation upon disintegration of the body by some m (t). Equation (8) should only be correct if
the deceleration of the meteoroid is neglected and only for short sections of the luminous path
where changes of the parameters characterizing the phenomenon’s evolution can be neglected.
Meteor radiation emerges from the plasma generated as a consequence of the meteoroid’s
interaction with the atmosphere along its trajectory. The theory of meteor radiation is a physical
theory of non-equilibrium from a chemically active plasma.
One cannot solve equation (8) for the intensity I of integral light uniquely. For every
meteor, the coefficient  as well as its mass m are unknown and depending on the disintegration
of each individual meteoroid. Therefore, the “photometric mass” obtained from integrating (8)
generally differs from the “dynamical mass” obtained from (7). The calculations of such
“masses”, just like other parameters, contain systematic errors because they are derived with
certain assumptions about the numerical values of the averaged parameters of the meteor, as
shown by I. N. Kovshun (1937 – 2004) (Kovshun 1971).
The absolute spectrophotometry of meteors permits measuring the intensity of such
monochromatic images in absolute units for every moment during the flight of the meteor. If
we neglect induced emission and self-absorption in these lines, the radiation intensity emitted
in a certain moment and in a unit volume of particles is determined by the total exposure of the
film
I ik  N i Aik h ik
(9)
where N i is a concentration of radiating particles, Aik is the Einstein probability coefficient of
the transition i  k, and h ik is the energy of the quantum. [The astronomical application of
(9) requires an additional factor of 1/(4  )].
Cook and Millman (1955) used spectral line intensities determined in absolute units to measure
the number of radiating atoms in the meteor spectrum.
Equation (9) allows the “photometric mass” of the meteoroid during the flight, which is
practically derived from the data of absolute spectrophotometry, such as the concentration of
radiating particles. Nevertheless, a knowledge of the geometrical shape of the meteoric plasma
is necessary for the determination of N.
If we neglect the inductive transitions in atoms and self-absorption, the intensity of
monochrome meteor radiation is defined by equation (9) for the stationary case.
The intensity of monochrome radiation of the meteor image can be described by the sum
I ik = I shock  I casc ,
(10)
where I shock is defined by the direct impacts of meteor plasma particles with electrons, while
I casc is defined by cascaded transitions following that initial impact. The value if ik will vary
6
through the entire path of the meteor, because the excitation condition change continuously
during the flight of the meteor in the Earth’s atmosphere (Smirnov, 1994).
e 2
A
lf ik N i ,
(11)
m0 c
Studies of meteor radiation usually neglect the self-absorption, which influences the
coefficient of radiation efficiency. If the stream of light enters an absorbing layer of vapor with
thickness l, classical physics delivers us with the absorption in this layer,
where N i is the concentration of particles in the lower atomic level, e is the electron charge,
m 0 is the electron mass, c is the speed of light, and f ik is the oscillator power corresponding to
a spectral line with frequency  ik .
The absorption A can be directly measured by the spectral line profile, because it relates
to the line widths. The dependence A ( f ik , N i ) is obtained experimentally as a “curve of
growth function”. As it was pointed out above, the spectrograph method does not allow the
determination of spectral line profiles because the meteor images have both longitudinal and
radial extend. Moreover, the value of l in equation (11) being defined by the geometrical size of
the plasma “cloud” remains unknown. For its determination, it is necessary to construct a
model of the meteor phenomenon.
For calculating the total power being radiated by a meteor, it is necessary to sum up the
intensities at various wavelengths, in various time intervals, and of various parts through which
the radiation passes, in absolute units. In other words, the determination of the power P implies
the calculation of the triple integral:
P
2 t 2
   Iddtds
(12)
1 t1 S
In practice, when averaging the meteor’s monochromic radiation power, the obtained
intensities refer to wavelength intervals, time intervals, and an area chosen for the radiation
model, for example, a sphere with 100 km radius. In this approach, we also have to take into
account that the radiation of a spherically symmetric meteor emission takes place at 100 km
distance from the camera lens.
As it was explained above, the coefficient must depend not only on the velocity, but also
on the altitude of flight H, the interaction energy between atmospheric particles and the meteor
plasma that is connected with the mass and volume occupied by the radiating particle. Thus, the
luminous efficiency is dependent on the kinetic energy of the meteoroid..
In the range of kinetic energies typical for meteors, the luminous efficiency may be
obtained on the basis of spectrophotometric data if a certain value for the meteoroid mass is
assumed. Knowing the total power of radiation from meteor spectral images, we can compare
it with the kinetic energy of the meteoroid if we assume a certain value for the mass, for
example, 10 3 kg. For instance, a meteor photographed on August 12/13, 1965, whose data are
given in the Table 1 of the supplement of the book by Smirnov (1994), had a kinetic energy of
1.8. 10 6 J. The observed power of radiation is P = Pi, where Pi are the powers of monochrome
meteor radiation. For a given meteor
7

P
i
i
mv 2 / 2
 5,9  10  3
(13)
Of course, the determination of such coefficient requires the presumption of the meteoroid
mass. Nowadays, the accuracy of the determination of the meteor radiation efficiency is
defined by the method of averaging the various parameters, which characterize the radiation of
the meteor body.
3. Phenomenological description of a meteor.
According to I. S. Astapovich (1908 –1976) (Astapovich, 1958) the radiation of a meteor
peaks in the middle of a meteor’s path. There, the meteor enters rather dense atmosphere layers,
and the length of the molecular free path becomes comparable with the meteor body dimension.
As it follows from the reference data, at the height of 90 km the average free path of air
particles is 2.4 cm, while at 80 km – 0.4 cm. It is possible to assume that these geometric
values are related to the dimensions of the radiating meteoroids.
The observed phenomenon – the meteor luminescence – is determined in the atmosphere by
a number of facts defining the necessary conditions of meteor emission. For visual registration
of a meteor, a certain power of radiation is needed.
If a star with magnitude m  6 m is observed visually it means that this star created
illumination E determined from the equality m = -13.89 - 2.5lg E* . Hence, E*  1,107.10 8 lux. .
The luminous flux Ф  Es , where s – is the illuminated area. As 1 lumen in the area of
maximum eye sensitivity is equal to 1.47.10-3 Wt , the flux density is E  1,627.10 11Wt / m 2 .
However, the difference between the apparent speed of a star image movement along the sky
and that of meteor movement is 10 3  10 4 times. That’s why the flux density of radiation of a
weak meteor Ф  for keeping the same influence on the eye retina as from a star with magnitude
6 must be consequently higher and should be equal to not less than E   2.10 7 Wt / m 2 at
every point of luminescence.
The same calculation may be carried out also for the glow of separate spectral lines in the
case of meteor spectra. If the distance to a meteor is R=100 km, we can observe a radiating
power at every point of meteor’s path, being equal to about 4R 2 Ф  2,5.10 3 Wt .
The interaction of a meteoroid with atmospheric particles, first of all, leads to warming up of
its surface due to the energy of impacts. For time interval t  0,1 sec that is equal to the
average time of meteor movement, the maximum heat conductivity of soil with superficial
sand (Chudnovsky, 1976) is:
q
Ql
J
,
 1,72.10  4
StT
mm. K
(14)
8
T
- temperature gradient, S  1m 2 . Such experimental definition of soil heat
l
conductivity corresponded to a soil density 1.5 g / cm 3 and porosity – 42%. With soil porosity
where
90% and density 0.2 g / cm 3 heat conductivity reduces and becomes equal to J
.
0,48.10  4
mm. K
With such comparatively low heat conductivity values of substances including those of a
meteoroid, it is possible to explain the “long-life” of its solid phase during its movement in the
atmosphere and conservation of its sluggishness.
As we can see, while determining the heat conductivity of a meteoroid, it is necessary to
consider the temporal factor t. While the meteor penetrates deeper into the atmosphere, the
process of heat conductivity accelerates.
Further, while the meteor is penetrating into the denser layers of the atmosphere, the boiling
saturated vapor forms a a cloud surrounding the solid phase of the meteoroid. It is not difficult
to find its density and concentration, e.g. for nitrogen, from the correlation:

p
 1,5kg / m 3 ,
RT
(15)
where p - is the pressure of saturated vapor of nitrogen at boiling temperature 77ºK, µ - mass
of nitrogen kilomole, R - universal gaseous constant. Hence, the molecular concentration of
saturated vapor of nitrogen is 3,2.10 25 particles / m 3 . It may be noticed, that the energy lost for
melting of the meteoroid is derived from the energy of collisions and can not be used for
forming the cascading meteor plasma. (specific melting heat of iron is 270 kJ/kg , for nitrogen
- 25,7 kJ/kg).
The meteor color is changing depending on its brightness. In the brightest parts of its path,
as well as during flares, radiation in the “blue” part of spectrum becomes equal to or dominant
over the “red” part. At the beginning of the path, in higher layers of the atmosphere, the “blue”
radiation is absent from the spectrum, as a rule. At the end of radiation both “red” and “blue”
lines disappear simultaneously. If we assume that radiation of the nitrogen bands is determined
by atmospheric gas, it means, that at the beginning of the meteor path the radiation is defined
by the atmospheric luminescence.
Let’s consider the average velocity of a meteor equal to velocity of the Perseid shower V
= 6.10 4 m / s. Then at the height of 100 km the meteor encounters a molecular nitrogen flow
equal to 3.10 4 g / s through cm 2 , and at the height of 80 km every square centimeter of
meteoroid cross-section encounters a flux of 1,2.10 2 g / s .
In other words, every square centimeter of meteoroid cross-section encounters
18
10 particles/s at the height of 100 km and 5.10 20 particles/s at the height of 80 km. It is the
energy released in collisions of atmospheric particles with the meteoroid’s surface that leads to
its warning up, boiling, and luminescence of evaporated meteor plasma. For all matter to
evaporate, the temperature of the heated meteoroid must be not less than the boiling
temperature of iron 3023K, as iron is a constituent element almost of all meteoroids.
Thus, at the height of 80 km at meteor velocity v, the energy of interaction reaches
9
J
(16)
,
cm 2 s
where W is the kinetic energy of impacting nitrogen molecules, N a  encountering flow of
particles.
Photometric measurements of spectrograms along the meteor path confirm that the
maximum energy distribution along the meteor spectrum shifts from the long-wave part around
6000Å to the short-wave part around 4000Å in accordance with energetic interactions due to
collision with atmosphere particles so far as energetic interactions are increased while a meteor
penetrates into dense layers of the atmosphere. Moreover, the maximum radiation is often
formed at the final flare of a meteor.
According to Wien’s Law, the brightness peaks at 4800Å for blackbody (from saturated gas)
temperatures at 7100ºK [Smirnov, 1994]. Wien’s Law : max T  b, allows us to determine
from experimental data the wavelength, corresponding to the maximum of the energy
distribution along the spectrum  max and moreover it gives us values (in high approximation)
of the corresponding excitation temperature of the meteor radiation. If we assume that the
velocity of movement of a meteoroid’s vapor together with its parental body equals
v=6. 10 4 m / s, then the corresponding temperature derived from the kinetic energy of
impacting nitrogen molecules is :

mv 2
T
 404.10 4 K ,
(17)
3k
W  wN a  42.10 3
where m - the mass of molecule, k – Boltzsmann’s constant. For such a high “temperature”,
not only the solid and liquid meteoroid phase cannot exist, but also the maximum of energy
distribution along the spectrum shifts to the X-ray part of the spectrum with  max  7 Å.
As such a phenomenon during meteor glow really isn’t observed, we have to believe that
by the time evaporated meteoroid particles emit light, they are slowed down to relative speeds
with the ambient environment of not higher than 2,5.10 3 m / s. Just these velocities of
movement of meteor plasma particles create radiation at the observed temperature

4.10 3  7.10 3 K . It is such a radiation that corresponds to the visible spectrum of a meteor.
The inertia of the meteoroid is high compared to that of the evaporated meteoroid
particles. That’s why the velocity of a meteoroid during its movement decreases much slower
than that of the evaporated meteoroid particles when slowed down by encountering gas flow
while colliding. Within some determined limits velocity v of the meteoroid may be assumed
constant. Being slowed down the particles of the meteor plasma, however, lag the meteor
body, forming its coma and wake.
During a certain period of time t, the concentration of particles in every point of the meteor
path varies according to the law, derived from the kinetic equation [Biberman et al., 1982]:
2

 t  0, 4  
 23 N 0
N (t )  N e (t )27.10
t    1 
9/2 0 
T
 t 0 
 
0

1/ 2
(18)
10
Here N e (t )  is the law of electron concentration depending on time. For a spherical
plasma cloud it is
3
 t0 
(19)
N e (t )  N e 0   .
t 
N e0 , N 0 , t 0 , T0  initial values of electron concentration, of plasma particles
concentration, of time and temperature. In the given case we have
N 0  3,2.10 25 particles / m 3  concentration of saturated vapor when nitrogen is evaporating,
T0  3026  K  boiling temperature of iron, t 0  9.10 5 s and N e0  can be used according to
Smirnov (1995). Such scattering of meteoroid substance, with its subsequent radiation of
meteor plasma, takes place until all mass is lost. . A meteor’s mass dissipates fully at the end of
the path, m (t )  0 .
4. Possible laser amplification in the flares of meteors
A Leonid meteor spectrum photographed by the author on November 17, 1965,
showed an amplification of radiation in the ionized Calcium lines so rapidly
that the rate of brightness changes surpassed that of any possible mechanical
changes of the structure of the meteoroid (Figure 1).
So, the existence of double final flares of fast shower meteors cannot be explained by the
mechanical cause of breaking of surface layers of a meteoroid body
The brightness changes occurred within 0,05 seconds. However, the
transmission of energy for evaporation of 1 gram of meteoroid requires much
more time. Hence, the light emanating from the meteor during flares is
independent from the mechanical changes in the meteoroid’s structure.
The plasmoid density decreases within the time limit according to the law
m
t 

 0  0 
3
4 / 3R
t 
 
m
4 / 3R 3
t0 
 0 
 t 
 
3
3
(20)
( 22)
Hence, the time scale (Zeldovich et al., 1966; Biberman et al, 1982) is approximately
equal to
1.2
t 
T  T0  0  ,
t 


m

t 0  
3 

4
/
3

V
 0

1/ 3
(21)
11
If the behavior of system is supposed to be similar to gas but with some effectively
adiabatic constants present, one can write the law of the plasma cooling in the form
T  A  1  t 3( 1) ;
(22)
where A is the constant entropy, γ = Cp/Cv - the ratio of heat capacities at constant pressure
and constant volume.
Smirnov (1994) proposed that the flare emission is enhanced by amplification as in a laser. At
the early stage of expanding gas, all the relaxation process in plasma occur swiftly, and thegas
is considered to be in thermodynamic equilibrium. Further on at lower temperature, expanding
ionization and dissociation process are slowed down, and more energy is being consumed.
However, the speed of inverse processes, recombination in particular, depend on
density, temperature, and time according to the law of power (20, 22). The events of ionization
can be neglected in this case. The emission is only due to the recombination and subsequent
cascade transition.
When the meteor penetrates deeper into the atmosphere, the density of kinetic energy flux of
interaction between it and atmospheric particles (they may be considered immovable) increases
according to Equation (16). While cooling, the expanding meteor plasma into the ambient
‘vacuum’ characteristic radiation in the form of double flares of meteor ( v  6.10 4 m / s ) can
arise, in conditions similar to that of a gas-dynamical laser.
Actually, as it follows from “Laser handbook” [M. 1978, table 6.1] the expenditure of
radiating gas at emission laser is 1,8.10 23 perticles / s, the intensification coefficient achieving
10 1 m 1 .Smirnov (1994) derived the same intensification coefficient due to laser amplification
during the final flare of a  7 m Perseid in the lines of ionized calcium. For doing this, it is
necessary to assume the value of iniital meteor mass m (t 0 ). The laser amplification is caused
by a population of atoms in excited states. During the early stages of meteoric plasma
scattering, all the relaxation processes are fast, and the gas is in a condition of thermodynamic
equilibrium. However, owing to ambipolar diffusion the degree of excitation of particles will
increase at the expense of a recombination. In plasma atoms, there are populated levels lying
near to the edge of the continuous spectrum. A particular inversion can be formed, followed
by an amplification of radiation.
While the meteoric plasma fades, the number of atoms in excited states will decrease due to
spontaneous transitions.The speed of disintegration of a plasma V is connected to her
temperature T and pressure p by parities:
T0
 1  V

 1
T
  V saund
2

 ;

12
p 0    1  V

 1
p 
2  V saund





2
(23)
  1
 ,


were Vsaund - sound speed in the environment, p0, T0 - pressure and temperature in an initial
stage of disintegration of the plasma.
Only fast meteors can emit sufficient amplified photons before depopulation prevents any
enhanced emission.
5. Conclusions
In each point along its trajectory, a meteor appears distinctive from other points along the path,
due to changing interaction conditions with the atmosphere. The kinetic energy of impacts is
carried out by radiation, except for the energy used to evaporate the superficial layers of the
meteoroid and heating of the atmosphere.
The meteoric radiation derives from a plasma formed as a result of the interaction of the
meteoroid with the atmosphere along the trajectory of flight. The theory of radiation of a
meteor is a physical theory of a non-equilibrium chemically active plasma.
The application of the equations of physics for a non-equilibrium low-temperature plasma
need to be combined with a modern physical approach to the changing conditions in the meteor
to move forward the field of physics of the meteoric phenomenon. That could include such
novel phenomena as laser amplified radiation in meteors.
REFERENCES
Astapovich I.S., Meteor Phenomena in the Land Atmosphere, "Fizmatgiz",
Moscow , M, 1958.
Biberman L.M., Vorobjev V.S., Yakubov I.T. The Kinetics of Non-equilibrium
Low-temperature Plasma , "Science", Moscow , M , 1982.
Bronshten V.A.,1981. The physics of meteor phenomena. "Nauka", Moscow, 416 p.
Cook A.F., Millman P. Photometrik analysis of a spectrogram of a Perseid meteor. Ap. J. 1955.
Vol. 121, No 1. P.250 - 270.
Chudnovsky A.F., Heat-physics of Soils, "Science", Moscow , M, 1976.
Herschel A.S., 1866. Prismatic spectra of the August meteor, 1866 // The Intellectual
Observer
Review of Natural History, Microscopy Research and Recreative Science, October, August,
1866; The coming meteor shower. The spectra of meteor // The Intellectual Observer, August,
1866 .
Kovshun I.N., 1971, Photometric meteor mass scale for meteor photographic observations.
13
Astronomical messenger, v.V, No.3, p.185-189
Levin B.J. (1958) Physics of the theory of meteors and meteor substance in the Solar System.
Academy of Science Publishers, Moscow.
Ljubarsky, K. A., 1961. Spectrographical of supervision of meteors in Turkmenistan.
The Bulletin of a Commission on comets and meteors of Astronomical Advice of an
Academy of Sciences SSSR, 5, 37 - 43.
Millman P.M. (1963) . A general survey of meteor spectra // Smithsonian Contribs. Astrophys.
V.7. - P. 119 - 127.
Millman P., McKinley D.W.R. The Solar System: Vol. 1Y. - The University of Chicago Press,
1963. - P. 674 - 773.
Öpic E.K.(1958). Physics of meteor flight in the atmosphere. N.-Y.174p.
Smirnov V. A., 1962. Photometry of two spectra of meteors using a method of A. Cook and
P. Millman. Publications of the Odessa University 152, 55 – 60.
Smirnov V., 1967. The identification of spectral lines of meteors with the help
laboratory dispersive standards. Astronomical magazine, 14, 1316 – 1323.
Smirnov V. A. The Spectra of Transient Atmospheric Luminous Phenomena: Meteors,
"Physico-Mathematical Literature", Moscow , M, 1994.
Smirnov V.A.,(1997) Methods of Non-radiating Meteor Mass Calculation, Planetary and
Space Science, 45 , 847 – 851.
Smirnov V.A. (2001). Spatio-temporal correlations characterizing the Meteor Phenomenon.
Proceedings of the International Meteor Conference, Pucioasa, Romania 21-24 September
2000. Published by IMO 2001 edited by Rainer Arlt, Mihaela Triglav, and Chris Trainer.
P.88 - 95.
Smirnov V.A Coefficient of Meteor Plasma Radiation. // Proceeding IMC Frasso Sabino.
Published by the International Meteor Organization 2000 edited by Rainer Arlt. P. 47 - 58.
Zeldovich Ya.B., Raiser Yu.P., (1966) The physics of shock waves and high-temperature
hydrodynamic phenomena. Nauka, Moscow.
Figure 1. Spectrum of a Leonid, photographed by the author in the settlement
Kryzhanovka on November 17/18, 1965. A band in the red part of the spectrum due to
molecular nitrogen begins emitting before other spectral lines from meteoric metal
atoms. In a final flare, there are strong H and K lines of ionized calcium. This spectrum
is of type Y by P.Millman.
14
Figure 2. Two spectra of fireballs that show a change in brilliance typical for revolving
meteoroids.
The top spectrum was obtained by Kronid Ljubarsky (1961) on September 7, 1958.
He identified 30 spectral lines. This spectrum is of Type Z.
The lower spectrum of a fireball was obtained by the author at a station of the
Odessa Astronomical observatory in Kryzhanovka on August 12/13, 1965.
Type of a spectrum is Type Z.
Identification of spectral lines were made with the help of the nitric standard (Smirnov 1967).
Download