首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The Helios 1 spacecraft was launched in December 1974 into a heliocentric orbit of 0.3 AU perihelion distance. Helios 2 followed one year later on a similar orbit. Both spaceprobes carry on board micrometeoroid experiments each of which contains two sensors with a total sensitive area of 121 cm2. To date, only preliminary data are available from Helios 2. Therefore the results presented here mainly apply to data from Helios 1. The ecliptic sensor of Helios 1 measures dust particles which have trajectories with elevations from ?45° to + 55° with respect to the ecliptic plane. The south sensor detects dust particles with trajectory elevations from ?90° (ecliptic south-pole) to ?4°. The ecliptic sensor is covered by a thin film (3000 Å parylene coated with 750 Å aluminium) as protection against solar radiation. The other sensor is shielded by the spacecraft rim from direct sunlight and has an open aperture. Micrometeoroids are detected by the electric charge produced upon impact. During the first 6 orbits of Helios 1 around the sun the experiment registered a total of 168 meteoroids, 52 particles were detected by the ecliptic sensor and 116 particles by the south sensor. This excess of impacts on the south sensor with regard to the impacts on the ecliptic sensor is due predominantly to small impacts which are characterized by small pulse heights of the charge signals. But also large impacts were statistically significantly more abundant on the south sensor than on the ecliptic sensor. Most impacts on the ecliptic sensor were observed when it was pointing in the direction of motion of Helios (apex direction). In contrast to that the south sensor detected most impacts when it was facing in between the solar and antapex direction. Orbit analysis showed that the “apex” particles which are predominantly detected by the ecliptic sensor have eccentricities e < 0.4 or semi-major axes a ? 0.5 AU. From a comparison with corresponding data from the south sensor it is concluded that the average inclination f of “apex” particles is -i < 30°. The excess of impacts on the south sensor, called “eccentric” particles, have orbit eccentricities e > 0.4 and semimajor axes a > 0.5AU. β-meteoroids leaving the solar system on hyperbolic orbits are directly identified by the observed imbalance of outgoing (away from the sun) and ingoing particles. It is shown that “eccentric” particles, due to their orbital characteristics, should be observable also by the ecliptic sensor. Since they have not been detected by this sensor it is concluded that the only instrumental difference between both sensors, i.e. the entrance film in front of the ecliptic sensor, prevented them from entering it. A comparison with penetration studies proved that particles which do not penetrate the entrance film must have bulk densities ρ(g/cm3) below an upper density limit ρmax. It is shown that approximately 30% of the “eccentric” particles have densities below ρmax = 1 g/cm3.  相似文献   

2.
A model calculation to predict infrared Shuttle flow due to the radiative relaxation of vibration of the NO molecule is presented. Space Shuttles hit atmospheric NO molecules at a very high speed (≈ 8 km s?1) and excite vibrational and rotational motions up to the temperature of 54,000 K. With the electric dipole radiation of Δν = 1, 2, 3, and particularly 4 (ν is the vibrational quantum number), the excited NO molecules emit infrared radiation before they collide with other molecules. The total radiation power is estimated to be 170AμW, where A is the cross-section area of the Shuttle in m2 if no adsorption of the NO molecule takes place on the Shuttle surface. The intensity of each infrared line is calculated as a function of time, including all vibrational states up to ν = 35. For example, the 5039 cm?1 line (ν = 24 → 20) has a maximum intensity of about 2.3 × 10?21 W molecule at around 0.2 ms, which corresponds to 80 cm from the Shuttle surface if the recoil speed of the molecules is 4 km s?1.  相似文献   

3.
W.A. Traub  N.P. Carleton 《Icarus》1974,23(4):585-589
A spectroscopic search for H2O and CH4 in Comet Kohoutek (1973f) was made using a Pepsios interferometer. No evidence was found for either molecule, allowing us to set an upper limit on their production rates (on about 21 January 1974) of Q(H2O) < 6.2 × 1028 sec?1 and Q(CH4) < 2.0 × 1030 sec?1. If the cometary surface is water-ice, this production rate leads to a product (1 ? A)·(πR02) < 2.2 km2, where A is the Bond albedo, R0 is the nuclear radius, and we assume that all the absorbed solar energy is used to evaporate H2O.  相似文献   

4.
《Icarus》1987,71(3):337-349
This paper represents a final report on the gravity analysis of radio Doppler and range data generated by the Deep Space Network (DSN) with Mariner 10 during two of its encounters with Mercury in March 1974 and March 1975. A combined least-squares fit to Doppler data from both encounters has resulted in a determination of two second degree gravity harmonics, J2 = (6.0 ± 2.0) × 10−5 and C22 = (1.0 ± 0.5) × 10−5, referred to an equatorial radius of 2439 km, plus an indication of a gravity anomaly in the region of closest approach of Mariner 10 to Mercury in March 1975 amounting to a mass deficiency of about GM = −0.1 km3sec−2. An analysis is included that defends the integrity of previously published values for the mass of Mercury (H. T. Howard et al. 1974, Science 185, 179–180; P. B. Esposito, J. D. Anderson, and A. T. Y. Ng 1978, COSPAR: Space Res. 17, 639–644). This is in response to a published suggestion by R. A. Lyttleton (1980, Q. J. R. Astron. Soc. 21, 400–413; 1981, Q. J. R. Astron. Soc. 22, 322–323) that the accepted values may be in error by more than 30%. We conclude that there is no basis for being suspicious of the earlier determinations and obtain a mass GM = 22,032.09 ± 0.91 km3sec−2 or a Sun to Mercury mass ratio of 6,023,600 ± 250. The corresponding mean density of Mercury is 5.43 ± 0.01 g cm−3. The one-sigma error limits on the gravity results include an assessment of systematic error, including the possibility that harmonics other than J2and C22 are significantly different from zero. A discussion of the utility of the DSN radio range data obtained with Mariner 10 is included. These data are most applicable to the improvement of the ephemeris of Mercury, in particular the determination of the precession of the perihelion.  相似文献   

5.
To redetermine the Galactic spiral density wave parameters, we have performed a spectral (Fourier) analysis of the radial velocities for 44 masers with known trigonometric parallaxes, proper motions, and line-of-sight velocities. The masers are distributed in awide range of Galactocentric distances (3.5 kpc < R < 13.2 kpc) and are characterized by a wide scatter of position angles ?? in the Galactic XY plane. This has required an accurate allowance for the dependence of the perturbation phase both on the logarithm of the Galactocentric distances and on the position angles of the objects. To increase the significance of the extraction of periodicities from data series with large gaps, we have proposed and implemented a spectrum reconstruction method based on a generalized maximum entropy method. As a result, we have extracted a periodicity describing a spiral density wave with the following parameters from the maser radial velocities: the perturbation amplitude f R = 7.7 ?1.5 +1.7 km s?1, the perturbation wavelength ?? = 2.2 ?0.1 +0.4 kpc, the pitch angle of the spiral density wave i = ?5 ?0.9° +0.2° , and the phase of the Sun in the spiral density wave ?? ?? = ?147 ?17° +3° .  相似文献   

6.
A highly sensitive device has been developed for the detection of micrometeoroids with masses extending down to and below the expected classical solar radiation pressure limit. This limit occurs at a mass of about 10−16 kg for a particle density of 3 × 103 kg m−3. The detector operates on the principle that charge (Q) is released when a hypervelocity microparticle impacts on a solid surface. This process has previously been investigated at particle velocities (ν) greater than 1 km sec−1 and the empirical relationship Q = Kmανβ obtained where m is the particle mass and K, α and β are constants. In the present study the validity of this relationship has been demonstrated for iron particles with masses from 10−16 to 10−13 kg in the velocity range 0.05–1.4 km sec−1 impacting on a molybdenum target. A value of α ≈ 1 is indicated and a value of β of 3.2 ± 0.4 has been obtained.  相似文献   

7.
We present results of the dual-frequency radio sounding of the Venusian ionosphere carried out by the Venera 9 and 10 satellites in 1975. Thirteen height profiles of electron density for different solar zenith angles varying from 10 to 87° have been obtained by analyzing the refraction bending of radiorays in the sounded ionssphere. The main maximum of electron density at a height of 140–150 km depends on the solar zenith angle and is 1.4 to 5 × 105 cm?3. The lower maximum is determined definitely to be at ~130 km high. In the main and lower maxima the electron density variations with solar zenith angle are in good agreement with the Chapman layer theory. For the first time it is found that the height of the upper boundary for the daytime ionosphere (hi) depends regularly on the solar zenith angle. At Z < 60°, hi does not exceed 300 km while at Z > 60°, it increases with Z and comes up to ~ 600 km at Z ~ 80°.  相似文献   

8.
F.A. Gifford  R.P. Hosker  K.S. Rao 《Icarus》1978,36(1):133-146
The spreading angle of a number of light and dark Martian streaks is determined from selected Mariner 9 images. The resulting frequency distributions of spreading half-angles have maxima at ~5° for light, and ~7° for dark streaks; however the dark streaks have a secondary maximum spreading angle at ~14°. The smaller values, which include most streaks, are interpreted as crater-wake spreading phenomena. The larger value, found in only a few dark streaks or “tails,” may result from atmospheric diffusion and subsequent deposition of material from isolated sources such as vents or blowouts. An atmospheric diffusion-deposition analysis is presented, assuming this streak origin, from which it is possible to deduce the eddy diffusivity, K, in Mars' boudary layer. Calculated K values are found to agree with various theoretical estimates. They lie in the range 107 and 109 cm2 sec?1 and exhibit the proper scale dependence. Thus it appears that, in addition to streak-derived wind direction patterns and speed information, it is possible in a few cases to derive information on Mars' boundary-layer turbulence from streak-spreading measurements.  相似文献   

9.
The spectrum of Saturn was measured from 80 to 350 cm?1 (29 to 125 μm) with ≈6-cm?1 resolution using a Michelson interferometer aboard NASA's Kuiper Airborne Observatory. These observations are of the full disk, with little contribution from the rings. For frequencies below 300 cm?1, Saturn's brightness temperature rises slowly, reaching ≈111°K at 100 cm?1. The effective temperature is 96.8 ± 2.5°K, implying that Saturn emits 3.0 ± 0.5 times as much energy as it receives from the Sun. The rotation-inversion manifolds of NH3 that are prominent in the far-infrared spectrum of Jupiter are not observed on Saturn. Our models predict the strengths to be only ≈2 to 5°K in brightness temperature because most of the NH3 is frozen out; this is comparable to the noise in our data. By combining our data with those of an earlier investigation when the Saturnicentric latitude of the Sun was B′ = 21.2°, we obtain the spectrum of the rings. The high-frequency end of the ring spectrum (ν > 230 cm?1) has nearly constant brightness temperature of 85°K. At lower frequencies, the brightness temperature decreases roughly as predicted by a simple absorption model with an optical depth proportional to ν1.5. This behavior could be due to mu-structure on the surface of the ring particles with a scale size of 10 to 100 μm and/or to impurities in their composition.  相似文献   

10.
We present equivalent widths of Venus CO2 scans of the P branch (P8–P32) of the 5ν3 band at 8689 Å, the P16 line of the 5ν3 band, and the P14 line of the ν1 + 5ν3 band at 7820 Å covering phase angles between 5°.1 and 170°. The equivalent widths reach a minimum at 10°, in agreement with a phase function with a backward lobe at 160° which is caused by a single internal reflection within the cloud particles. This is evidence that Venus cloud particles are composed of liquid droplets. Maximum equivalent widths are observed at ~60°, a value which is closer to the maximum of single-layer Mie scattering models than to that of two-layer models. At high phase angles we observe equivalent widths greater than those computed from homogeneous scattering models, indicating that at high altitudes the mixing ratio of scattering particles to CO2 increases with depth. At all phase angles, particularly at large phase angles, the temporal and spatial variations in the observed equivalent widths confuse the phase variation.  相似文献   

11.
《Icarus》1987,70(2):354-365
Liquid solutions of N2 containing up to one-third CH4 can exist on Triton's surface in regions T > 62.5°K. More generally, subsurface oceans of N2 solution are expected to be stable beneath overlying, thermally insulating, less dense layers of the abundant light hydrocarbon products of radiochemical synthesis: C2H6, C3H8, and C4H10. Cosmic rays are the main source of energy, capable of producing synthesis of organic compounds from N2CH4 solutions on the surface. For baseline Triton models with R = 2500 km, ϱ = 2.1 g cm−3, and Ts = 65 or 55°K, respectively, 4 × 10−3 or 7 × 10−3 erg cm−2 sec−1 (49 or 87% of the total incident flux) is deposited within a few meters below the surface. Using yields from laboratory experiments, we estimate the quantities of products produced: over 4.5 billion years, the cosmic ray flux alone produces 2 to 4 m of organic product, about half of which is C2H6. For ocean depths <250 m, C2H6 will reach its saturation limit and form a surface “slick.” For ocean depths <10 km, all of the other products also oversaturate and exsolve, adding to the surface slick and/or to a denser bottom sediment. Products produced from solid N2CH4 mixtures will accumulate as evaporite deposits because of the rapid volatile transport (of N2 and CH4) over Triton's surface. The complex, reddish organic solid found in laboratory simulations is probably the source of Triton's reddish color. Estimated yields over 4.5 billion years (for 7 × 10−3 erg cm−2 sec−1) are 190 (C2H6), 58 (NH3), 17 (HCN), 3.5 (HN3), 2.5 (C4H10), 0.35 (CH3CN), and 0.14 (C2H5N3) g cm−2. More basic laboratory work on the low-temperature, low-pressure solvent properties and phase equilibria of N2-hydrocarbon systems is clearly needed.  相似文献   

12.
Based on kinematic data on masers with known trigonometric parallaxes and measurements of the velocities of HI clouds at tangential points in the inner Galaxy, we have refined the parameters of the Allen-Santillan model Galactic potential and constructed the Galactic rotation curve in a wide range of Galactocentric distances, from 0 to 20 kpc. The circular rotation velocity of the Sun for the adopted Galactocentric distance R 0 = 8 kpc is V 0 = 239 ± 16 km s?1. We have obtained the series of residual tangential, ΔV θ , and radial, V R , velocities for 73 masers. Based on these series, we have determined the parameters of the Galactic spiral density wave satisfying the linear Lin-Shu model using the method of periodogram analysis that we proposed previously. The tangential and radial perturbation amplitudes are f θ = 7.0±1.2 km s?1 and f R = 7.8±0.7 km s?1, respectively, the perturbation wave length is λ = 2.3±0.4 kpc, and the pitch angle of the spiral pattern in a two-armed model is i = ?5.2° ±0.7°. The phase of the Sun ζ in the spiral density wave is ?50° ± 15° and ?160° ± 15° from the residual tangential and radial velocities, respectively.  相似文献   

13.
At the distancer?1015 cm from NP 0532 the plasma concentration decreases so that the intense low-frequency wave (ν=30 Hz) can propagate. The interaction of this wave with the electrons ejected from the pulsar should result in the IR radiation withF ν~102 fu at λ~10 μ. This flux is the order of the excess IR radiation from the Crab Nebula.  相似文献   

14.
Both hypervelocity impact and dynamic spall experiments were carried out on a series of well-indurated samples of gabbro to examine the relation between spall strength and maximum spall ejecta thickness. The impact experiments carried out with 0.04- to 0.2-g, 5- to 6-km/sec projectiles produced decimeter- to centimeter-sized craters and demonstrated crater efficiencies of 6 × 10?9 g/erg, an order of magnitude greater than in metal and some two to three times that of previous experiments on less strong igneous rocks. Most of the crater volume (some 60 to 80%) is due to spall failure. Distribution of cumulative fragment number, as a function of mass of fragments with masses greater than 0.1 g yield values of b = d(log Nf)/d log(m) ?0.5 ?0.6, where N is the cumulative number of fragments and m is the mass of fragments. These values are in agreement or slightly higher than those obtained for less strong rocks and indicate that a large fraction of the ejecta resides in a few large fragments. The large fragments are plate-like with mean values of B/A and C/A 0.8 0.2, respectively (A = long, B = termediate, and C = short fragment axes). The small equant-dimensioned fragments (with mass < 0.1 g and B ~ 0.1 mm) represent material which has been subjected to shear failure. The dynamic tensile strenght of San Marcos gabbro was determined at strain rates of 104 to 105 sec?1 to be 147 ± 9 MPa. This is 3 to 10 times greater than inferred from quasi-static (strain rate 100 sec?1) loading experiments. Utilizing these parameters in a continuum fracture model predicts a tensile strenght of σmε?[0.25–0.3], where ε is strain rate. It is suggested that the high spall strenght of basic igneous rocks gives rise to enhanced cratering efficiencies due to spall in the <102-m crater diamter strength-dominated regime. Although the impact spall mechanism can enhance cratering efficiencies it is unclear that resulting spall fragments achieve sufficient velocities such that fragments of basic rocks can escape from the surfaces of planets such as the Moon or Mars.  相似文献   

15.
Absolute spectrophotometry of the coma of Comet Kohoutek 1973f is discussed for the nights of January 24 and 26, 1974. Specific intensities are measured for spectral features and a continuum band in the wavelength region λλ3460–6062Å. The (0, 0) band of the Δν = 0 sequence of the violet system of the cyanogen molecule is analyzed and column densities of 1.7 × 1015m?2 and 3.4 × 1014m?2 are found for January 24 and 26, 1974, respectively. The analysis of the bands of C2 will be reported in a second paper of this series.  相似文献   

16.
A spectrum of Jupiter between 6000 and 12 000 cm? at high resolution (0.05 cm?) was recorded with a Michelson interferometer at Palomar Mountain in October 1974. An analysis of the R branch of the 3ν3CH4 band with the reflecting-layer model, taking into account the H2 absorption which occurs in the same spectral range, leads to a Lorentzian half-width of 0.09 ± 0.02 cm?1, a rotational temperature of 175 ± 10° K, and a CH4 abundance of order 52m atm. Five lines of the 13CH43ν3 band have been identified; a comparison with new laboratory spectra indicates that the 13CH4/12CH4 ratio in the Jupiter atmosphere is close to the terrestrial ratio.  相似文献   

17.
《Icarus》1987,71(1):30-45
This is the second paper devoted to the numerical study of planetary collisions as a possible scenario for forming the Moon. We present a series of nine simulations of a collision between the protoearth and an impactor of various sizes. The mass ratio between the protoearth and the impactor ranged from 0.1 to 0.25. We were able to model both planets with iron cores, having modified our smoothed particle hydrodynamics code to allow the inclusion of up to 10 different material types. Two different relative velocities at infinity for the impactor were considered: ν = 0 km/sec and ν = 10 km/sec. We show that for a low-velocity collision and an impactor in the mass range 6.5 × 1026Mimpactor ≤ 8.2 × 1026 g, more than a lunar mass of iron-poor material is thrown into orbit. For an impactor with a mass within this range, the ejected mass that goes into orbit is for the most part divided comparably into material orbiting inside the Roche limit and into material orbiting outside the Roche limit. This material is either spread out in the form of a disk, or, for a relatively narrow range of masses toward the lower end of the range, clumped into an object of about lunar mass beyond the Roche limit. For impactors more massive than about 8.2 × 1026 g we found that there is too little mass thrown into orbit. For very small mass impactors well over a lunar mass is placed in orbit, but a large amount of it is iron. In the high-velocity range we did not find a possible mass range for the impactor that would lead to the formation of an iron-poor disk massive enough to form the Moon.  相似文献   

18.
Altitude dependences of [CO2] and [CO2+] are deduced from Mariner 6 and 7 CO2+ airglow measurements. CO2 densities are also obtained from ne radio occultation measurements. Both [CO2] profiles are similar and correspond to the model atmosphere of Barth et al. (1972) at 120 km, but at higher altitudes they diverge and at 200–220 km the obtained [CO2] values are three times less the model. Both the airglow and radio occultation observations show that a correction factor of 2.5 should be included into the values for solar ionization flux given by Hinteregger (1970). The ratio of [CO2+]/ne is 0.15–0.2 and, hence, [O]/[CO2] is ~3% at 135 km. An atmospheric and ionospheric model is developed for 120–220 km. The calculated temperature profile is characterized by a value of T ≈ 370°K at h ? 220 km, a steep gradient (~2°/km) at 200-160 km, a bend in the profile at 160 km, a small gradient (~0.7°/km) below and a value of T ≈ 250°K at 120 km. The upper point agrees well with the results of the Lyman-α measurements; the steep gradient may be explained by molecular viscosity dissipation of gravity and acoustical waves (the corresponding energy flux is 4 × 10?2 erg cm?2sec?1 at 180 km). The bend at 160 km may be caused by a sharp decrease of the eddy diffusion coefficient and defines K ≈ 2 × 108cm2sec?1; and the low gradient gives an estimate of the efficiency of the atmosphere heating by the solar radiation as ? ≈ 0.1.  相似文献   

19.
To study the peculiarities of the Galactic spiral density wave, we have analyzed the space velocities of Galactic Cepheids with propermotions from the Hipparcos catalog and line-of-sight velocities from various sources. First, based on the entire sample of 185 stars and taking R 0 = 8 kpc, we have found the components of the peculiar solar velocity (u , v ) = (7.6, 11.6) ± (0.8, 1.1) km s?1, the angular velocity of Galactic rotation Ω0 = 27.5 ± 0.5 km s?1 kpc?1 and its derivatives Ω′0 = ?4.12 ± 0.10 km s?1 kpc?2 and Ω″0 = 0.85 ± 0.07 km s?1 kpc?3, the amplitudes of the velocity perturbations in the spiral density wave f R = ?6.8 ± 0.7 and f θ = 3.3 ± 0.5 km s?1, the pitch angle of a two-armed spiral pattern (m = 2) i = ?4.6° ± 0.1° (which corresponds to a wavelength λ = 2.0 ± 0.1 kpc), and the phase of the Sun in the spiral density wave χ = ?193° ± 5°. The phase χ has been found to change noticeably with the mean age of the sample. Having analyzed these phase shifts, we have determined the mean value of the angular velocity difference Ω p ? Ω, which depends significantly on the calibrations used to estimate the individual ages of Cepheids. When estimating the ages of Cepheids based on Efremov’s calibration, we have found |Ω p ? Ω0| = 10 ± 1stat ± 3syst km s?1 kpc?1. The ratio of the radial component of the gravitational force produced by the spiral arms to the total gravitational force of the Galaxy has been estimated to be f r0 = 0.04 ± 0.01.  相似文献   

20.
We performed for the first time stereoscopic triangulation of coronal loops in active regions over the entire range of spacecraft separation angles (?? sep??6°,43°,89°,127°,and 170°). The accuracy of stereoscopic correlation depends mostly on the viewing angle with respect to the solar surface for each spacecraft, which affects the stereoscopic correspondence identification of loops in image pairs. From a simple theoretical model we predict an optimum range of ?? sep??22°??C?125°, which is also experimentally confirmed. The best accuracy is generally obtained when an active region passes the central meridian (viewed from Earth), which yields a symmetric view for both STEREO spacecraft and causes minimum horizontal foreshortening. For the extended angular range of ?? sep??6°??C?127° we find a mean 3D misalignment angle of ?? PF??21°??C?39° of stereoscopically triangulated loops with magnetic potential-field models, and ?? FFF??15°??C?21° for a force-free field model, which is partly caused by stereoscopic uncertainties ?? SE??9°. We predict optimum conditions for solar stereoscopy during the time intervals of 2012??C?2014, 2016??C?2017, and 2021??C?2023.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号