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 46 毫秒
1.
The results of a search for maser emission in the methanol lines 8?1-70 E at 229.8 GHz, 3?2-4?1 E at 230.0 GHz, 00-1?1 E at 108.9 GHz, and in the J 1-J 0 E series near 165 GHz in star-forming regions are reported. At least two masers and two candidates have been detected at 229.8 GHz. Thus, methanol masers have been detected in the 1-mm band for the first time. At 108.9 GHz, masers have been detected toward G345.01+1.79 and possibly toward M8E as well. Thermal emission was found toward 28 objects. The 229.8-GHz sources are class I masers, whereas the 108.9-GHz sources are class II masers. An analysis using a large velocity-gradient method shows that the 229.8-GHz masers can appear at densities of about 3×104 cm?3. The ratios of the flux densities in different class I lines toward DR 21(OH) and DR 21 West can be approximated in models with gas kinetic temperatures of about 50 K. Detection of the 108.9 GHz masers toward G345.01+1.79 and M8E may provide information about the geometry of these objects.  相似文献   

2.
The excitation of methanol in the absence of external radiation is analyzed, and LTE methods for probing interstellar gas considered. It is shown that rotation diagrams correctly estimate the gas kinetic temperature only if they are constructed using lines whose upper levels are located in the same K-ladders, such as the J0?J?1E lines at 157 GHz, the J1?J0E lines at 165 GHz, and the J2?J1E lines at 25 GHz. The gas density must be no less than 107 cm?3. Rotation diagrams constructed from lines with different K values for their upper levels (e.g., 2K?1K at 96 GHz, 3K?2K at 145 GHz, 5K?4K at 241 GHz) significantly underestimate the temperature, but enable estimation of the density. In addition, diagrams based on the 2K?1K lines can be used to estimate the methanol column density within a factor of about two to five. It is suggested that rotation diagrams should be used in the following manner. First, two rotation diagrams should be constructed, one from the lines at 96, 145, or 241 GHz, and another from the lines at 157, 165, or 25 GHz. The former diagram is used to estimate the gas density. If the density is about 107 cm?3 or higher, the latter diagram reproduces the temperature fairly well. If the density is around 106 cm?3, the temperature obtained from the latter diagram should be multiplied by a factor of 1.5–2. If the density is about 105 cm?3 or lower, then the latter diagram yields a temperature that is lower than the kinetic temperature by a factor of three or more, and should be used only as a lower limit for the kinetic temperature. The errors in the methanol column density determined from the integrated intensity of a single line can be more than an order of magnitude, even when the gas temperature is well known. However, if the J0?(J ? 1)0E lines, as well as the J1?(J ? 1)1A+ or A? lines are used, the relative error in the column density is no more than a factor of a few.  相似文献   

3.
The methanol-line spectra in two maser condensations at velocities ~41 and ~45 km/s in the star-forming region W48 have been studied. The intensity of the 20-3?1 E (12.2 GHz) line is anticorrelated with that of the 51-60 A + (6.7 GHz) line: the intensity of the 51-60 A + (6.7 GHz) line is greater at ~41 km/s than at ~45 km/s, while the opposite is true of the 20-3?1 E (12 GHz) line. The remaining class II methanol lines in this source demonstrate the same behavior as the 20-3?1 E (12 GHz) line. This contradicts current concepts about the maser line intensities in various methanol transitions: according to model calculations, the intensities of all class II lines should vary in phase. This effect is confirmed for a large homogeneous sample of 67 sources. Possible explanations of the observed effect are proposed; one suggests the possible role of “transpumping” of the methanol-level populations in the maser condensations. The relationships between the variations of the 20-3?1 E (12 GHz) and 51-60 A + (6.7 GHz) line intensities, which are present for all 67 sources considered, may indicate that the condensations are at different distances from the pumping source. The presence of condensations at various distances from the pumping source in all 67 sources can be understood if they are ice planets revolving in different orbits around massive stars or protostars.  相似文献   

4.
An analysis of the flux densities of the 51-60 A + (6.7 GHz) and 20-3?1 E (12.2 GHz) class II methanol maser lines in a large and homogeneous sample of maser sources has been carried out. For convenience, the maser lines were divided into three groups: group I contains spectral features for the lines most prominent in the 51-60 A + (6.7 GHz) transition, group II contains spectral features for the lines strongest in the 20-3?1 E (12.2 GHz) transition, group III contains spectral features for which the velocities of the emission maxima of the two lines coincide. The same dependence was found for group II and group III: log S 6.7=(0.79±0.05)×log S 12.2+(0.79±0.05). The spectral features in group I do not obey this relation, and deviations from a linear dependence are considerably greater. It is suggested that methanol class II masers be divided into a subclass IIa, which has special conditions favoring 6.7 GHz masers, and a subclass IIb, which is comprised of the 12.2 GHz masers and those 6.7 GHz masers that necessarily accompany them under the same conditions.  相似文献   

5.
The results of spectral observations of the region of massive star formation L379IRS1 (IRAS18265–1517) are presented. The observations were carried out with the 30-m Pico Veleta radio telescope (Spain) at seven frequencies in the 1-mm, 2-mm, and 3-mm wavelength bands. Lines of 24 molecules were detected, from simple diatomic or triatomic species to complex eight- or nine-atom compounds such as CH3OCHO or CH3OCH3. Rotation diagrams constructed from methanol andmethyl cyanide lines were used to determine the temperature of the quiescent gas in this region, which is about 40–50 K. In addition to this warm gas, there is a hot component that is revealed through high-energy lines of methanol and methyl cyanide, molecular lines arising in hot regions, and the presence of H2O masers and Class II methanol masers at 6.7 GHz, which are also related to hot gas. One of the hot regions is probably a compact hot core, which is located near the southern submillimeter peak and is related to a group of methanol masers at 6.7 GHz. High-excitation lines at other positions may be associated with other hot cores or hot post-shock gas in the lobes of bipolar outflows. The rotation diagrams can be use to determine the column densities and abundances of methanol (10?9) and methyl cyanide (about 10?11) in the quiescent gas. The column densities of A- and E-methanol in L379IRS1 are essentually the same. The column densities of other observedmolecules were calculated assuming that the ratios of the molecular level abundances correspond to a temperature of 40 K. The molecular composition of the quiescent gas is close to that in another region of massive star formation, DR21(OH). The only appreciable difference is that the column density of SO2 in L379IRS1 is at least a factor of 20 lower than the value in DR21(OH). The SO2/CS and SO2/OCS abundance ratios, which can be used as chemical clocks, are lower in L379IRS1 than in DR21(OH), suggesting that L379IRS1 is probably younger than DR21(OH).  相似文献   

6.
Forty-eight objects were detected in the 5?1–40 E methanol line at 84.5 GHz during a survey of Class I maser sources. Narrow maser features were found in 14 of these. Broad quasi-thermal lines were detected toward other sources. One of the objects with narrow features at 84.5 GHz, the young bipolar outflow L1157, was also observed in the 80–71 A + line at 95.2 GHz; a narrow line was detected at this frequency. Analysis showed that the broad lines are usually inverted. The quasi-thermal profiles imply that there are no more than a few line opacities. These results confirm the plausibility of models in which compact Class I masers appear in extended sources as a result of a preferential velocity field.  相似文献   

7.
Observations of the molecular cloud G1.6-0.025 in the 2K-1K and J0-J?1E series and 5?1-40E line of CH3OH, the (2-1) and (3-2) lines of SiO, and the 7?7-6?6 line of HNCO are described. Maps of the previously observed extended cloud with Vlsr~50 km/s and high-velocity clump with Vlsr~160 km/s, as well as a newly detected clump with Vlsr~0 km/s, have been obtained. The extended cloud and high-velocity clump have a nonuniform structure. The linewidths associated with all the objects are between 20 and 35 km/s, as is typical of clouds of the Galactic center. In some directions, emission at velocities from 40 to 160 km/s and from ?10 to +75 km/s is observed at the clump boundaries, testifying to a connection between the extended cloud and the high-velocity clump and clump at Vlsr~0 km/s. Compact maser sources are probaby contributing appreciably to the emission of the extended cloud in the 5?1-40E CH3OH line. Non-LTE modeling of the methanol emission shows that the extended cloud and high-velocity clump have a relatively low hydrogen density (<104 cm?3). The specific column density of methanol in the extended cloud exceeds 6×108 cm?3s, and is 4×108?6×109 cm?3s in the high-velocity clump. The kinetic temperatures of the extended cloud and high-velocity clump are estimated to be <80 K and 150–200 K, respectively. Possible mechanisms that can explain the link between the extended cloud with Vlsr~50 km/s and the clumps with Vlsr~0 km/s and ~160 km/s are briefly discussed.  相似文献   

8.
A molecular cloud and high-velocity outflow associated with the star-forming region L379 IRS3 have been mapped in the 6?1-50E methanol and CS (3-2) lines using the 12-meter Kitt Peak telescope. The estimated CS column density and abundance in the molecular cloud are 8×1014 cm?2 and 4×10?9, respectively. LVG modeling of the methanol emission constrains the gas density in the cloud to (1–4)×105 cm?3 and the gas kinetic temperature to 20–45 K. The upper limit on the density of the high-velocity gas is 105 cm?3.  相似文献   

9.
A survey has been made of 27 Galactic star-forming regions in the (CH3CN) 6K–5K, 5K–4K, and 8K–7K lines of methyl cyanide (CH3CN) at 110, 92, and 147 GHz. Twenty-five sources were detected at 110 GHz, nineteen at 92 GHz, and three at 147 GHz. The strongest CH3CN emission arises in hot cores in regions of massive star formation. The abundance of CH3CN in these objects exceeds 10?9 as a consequence of grain mantle evaporation. Weaker CH3CN lines were found in a number of sources. These can arise in either warm (30–50 K), dense (>104 cm?3) clouds or in hot regions with cooler gas.  相似文献   

10.
The results of a survey of 63 Galactic star-forming regions in the 6K–5K and 5 K –4K methyl acetylene lines at 102.5 and 85.5 GHz are presented. Fourty-three sources were detected at 102.5 GHz, and twenty-five at 85.5 GHz. Emission was detected toward molecular clouds with kinetic temperatures of 20–60 K (so-called “warm clouds”). The CH3CCH abundances in these clouds are about several ×10?9. Five sources (NGC 2264, G30.8-0.1, G34.26+0.15, DR 21(OH), S140) were mapped using the maximum-entropy method. The sizes of the mapped clouds fall in the range 0.1–1.7 pc, and the clouds have virial masses of 90–6200 M and densities between 6×104 and 6×105 cm?3. The CH3CCH sources coincide spatially with the CO and CS sources. Chemical-evolution simulations show that the typical methyl acetylene abundances in the observed clouds correspond to ages of ≈6×104 years.  相似文献   

11.
Thirty-four Galactic star-forming regions have been observed in the J=4?3 line of cyanoacetylene (HC3N) at 36.4 GHz. Emission was detected toward 17 sources. The HC3N column density was determined in the detected sources, and in eight of them (NGC 2264, L379, W51E1/E2, DR 21 West, DR 21 (OH), DR 21, S140, and Cep A), the relative cyanoacetylene abundance was estimated. The HC3N abundance in these objects is about 1–70×10?10.  相似文献   

12.
Room temperature X-irradiation of some natural beryls produced several new absorption lines in the electron paramagnetic resonance (EPR) spectrum, a known series of optical absorption lines in the 500–700 nm range, and a shift of the absorption edge to lower energies. Several of the new EPR lines and part of the irradiation-induced shift of the absorption edge disappeared after a few days at room temperature, and were not examined in detail. However, three of the paramagnetic centres responsible for the new EPR lines were stable at room temperature and two of these have previously been identified as atomic hydrogen and the methyl radical, CH3. These species were stable to ~150 and ~450°C respectively. The third stable species, hitherto unreported, showed a single-line EPR spectrum of axial symmetry, with g∥=2.0051 and g⊥=2.0152. This spectrum was found to be intensity-correlated with the series of optical bands in the 500–700 nm range, after thermal bleaching at 175°C. The EPR and optical spectra are therefore assigned to the same species. It is argued that this species is the CO 3 ? molecular ion, located in the widest part of the structural channel and aligned with the plane of the molecule perpendicular to the c axis. The EPR spectrum is consistent with a 2 A2 ground state of a CO 3 ? molecule with trigonal symmetry, and this requires that the optical transition has a 2 A22 E′ character. Most of the features in the optical spectrum can be assigned to coupling of a totally symmetric mode of frequency ~1020 cm?1 onto a zero-phonon line at 14,490 cm?1 and a second weaker line at 16,020 cm?1. However, both of these two fundamental lines are structured, and the two components show strong temperature-dependent derivative-shaped magnetic circular dichroism (MCD). Furthermore, the overall sign of the MCD for the line at 16,020 cm?1 is opposite to that at 14,490 cm?1. The separation (~120 cm?1) of the two components of the 14,490 cm?1 line is much larger than that expected from spin-orbit interaction, and the origin of this splitting is not yet understood.  相似文献   

13.
Observations of various types of objects in the northern sky were obtained at 44 GHz in the 70-61 A + methanol line on the 20-m Onsala radio telescope (Sweden), in order to search for Class I methanol maser emission in the interstellar medium: regions of formation of high-mass stars, dust rings around HII regions, and protostellar candidates associated with powerful molecular outflows and Galactic HII regions. Seven new Class Imethanolmasers have been discovered toward regions of formation of highmass stars, and the existence of two previously observed masers confirmed. The following conclusions are drawn: (1) neither the association of a bipolar outflow manifest in the wings of CO lines with a highmass protostellar object (HMPO) nor the presence of thermal emission in lines of complex molecules are sufficient conditions for the detection of Class I methanol emission; no association with HMPOs radiating at 44 GHz was found for EGOs (a new class of object tracing bipolar outflows); (2) the existence of H2O masers and Class II methanol masers in the region of aHMPOenhances the probability of detecting Class I methanol emission toward the HMPO; Class II methanol masers with stronger line fluxes are associated with Class I methanol masers.  相似文献   

14.
Elastic behavior and pressure-induced structural evolution of synthetic boron-mullite “Al5BO9” (a = 5.678(2) Å, b = 15.015(4) Å and c = 7.700(3) Å, space group Cmc21, Z = 4) were investigated up to 7.4 GPa by in situ single-crystal X-ray diffraction with a diamond anvil cell under hydrostatic conditions. No phase transition or anomalous compressional behavior occurred within the investigated P range. Fitting the P–V data with a truncated second-order (in energy) Birch-Murnaghan Equation-of-State (BM-EoS), using the data weighted by the uncertainties in P and V, we obtained: V 0 = 656.4(3) Å3 and K T0 = 165(7) GPa (β V0 = 0.0061(3) GPa?1). The evolution of the Eulerian finite strain versus normalized stress (f EF E plot) leads to an almost horizontal trend, showing that a truncated second-order BM-EoS is appropriate to describe the elastic behavior of “Al5BO9” within the investigated P range. The weighted linear regression through the data points gives: F E(0) = 159(11) GPa. Axial compressibility coefficients yielded: β a  = 1.4(2) × 10?3 GPa?1, β b  = 3.4(4) × 10?3 GPa?1, and β c  = 1.7(3) × 10?3 GPa?1 (β a :β b :β c  = 1:2.43:1.21). The highest compressibilities observed in this study within (100) can be ascribed to the presence of voids represented by five-membered rings of polyhedra: Al1–Al3–Al4–Al1–Al3, which allow accommodating the effect of pressure by polyhedral tilting. Polyhedral tilting around the voids also explains the higher compressibility along [010] than along [001]. The stiffer crystallographic direction observed here might be controlled by the infinite chains of edge-sharing octahedra running along [100], which act as “pillars”, making the structure less compressible along the a-axis than along the b- and c-axis. Along [100], compression can only be accommodated by deformation of the edge-sharing octahedra (and/or by compression of the Al–O bond lengths), as no polyhedral tilting can occur. In addition, a comparative elastic analysis among the mullite-type materials is carried out.  相似文献   

15.
Our analysis of many years of infrared photometry of the unique object FG Sge indicates that the dust envelope formed around the supergiant in August 1992 is spherically symmetrical and contains compact, dense dust clouds. The emission from the spherically symmetrical dust envelope is consistent with the observed radiation from the star at 3.5–5 µm, and the presence of the dust clouds can explain the radiation observed at 1.25–2.2 µm. The mean integrated flux from the dust envelope in 1992–2001 was ~(1.0±0.2)×10?8 erg s?1cm?2. The variations of its optical depth in 1992–2001 were within 0.5–1.0. The maximum density of the dust envelope was recorded in the second half of 1993 and corresponded to mean optical depths as high as unity. Several times in the interval from 1992 to 2001, the dusty material of the envelope partially dissipated and was then replenished. For example, the optical depth of the dust cloud at λ=1.25 µm during the last brigthness minimum in the J band was τ1.25≈4.3, which is much higher than the optical depth of the dust envelope of FG Sge. During maxima of the J brightness, the mean spectral energy distribution at 0.36–5 µm can be represented as a combination of radiation from a G0 supergiant that is attenuated by a dust envelope with a mean optical depth of 0.65±0.15 and emission from the spherically symmetrical dust envelope itself, with the temperature of the graphite grains being 750±150 K. At minima of the J brightness, only radiation from the dust envelope is observed at 1.65–5 µm, with the radiation from the supergiant barely detectable at 1.25 µm. As a result, the integrated flux during J minima is almost half that during J maxima. The mean mass of the spherically symmetrical dust envelope of FG Sge in 1992–2001 was (3 ± 1) × 10?7M. This envelope’s mass varied by nearly a factor of two during 1992–2001, in the range (2 – 4) × 10?7M. In Autumn 1992, the mass-loss rate from the supergiant exceeded 2 × 10?7M/yr. The average rate at which matter was injected into the envelope during 1993–2001 was 10?8M/yr. The mean rate of dissipation of the dust envelope was about 1 × 10?8M/yr. During 1992–2001, the supergiant lost about 8.7 × 10?7M. The parameters of the dust envelope were relatively constant from 1999 until the middle of 2001.  相似文献   

16.
The temperature dependence of diffusion is usually found to follow the Arrhenius law: D = D0e?E/RT Winchell (1969) showed that there is commonly an inter-dependence between D0 and E (for diffusion in silicate glasses), such that diffusion of different species show a positive correlation on a log D0 vs E plot. A similar effect was noted by Hofmann (1980) for cation diffusion in basalt. This implies that diffusion rates of different species tend to converge at a particular temperature; this effect is known as the ‘compensation effect’. I will show that this effect is also present for diffusion in feldspars and olivines. The equations for the compensation lines (with E given in kcal/mol) are: basalt—E = 50 + 7.5 log D0 feldspar—E = 50.7 + 3.4 log D0 olivine—E = 78.0 + 7.5 log D0 The convergence, or crossover, temperatures for diffusion in various materials are: obsidian—3400°C basalt—1370°C olivine—1360°C feldspar—460°C Compensation plots are useful for evaluating and comparing experimental diffusion data (though of limited usefulness in a predictive sense) and for understanding ‘closure temperatures’ for diffusion in petrogenetic processes (since closure temperature, the temperature at which natural diffusion processes are frozen in, is dependent on E, log d0, and cooling rate). I show that most diffusing species in feldspar have a closure-temperature close to the crossover or convergence temperature, implying that all species in feldspars can be expected to ‘freeze-in’ simultaneously at temperatures in the range 400–600°C (for cooling rates in the range 101–105°C/myr). Closure temperatures of various species in olivine, on the other hand, span a much larger range (800°C) for a similar range in cooling rates, implying that different elements in olivine will record different time-temperature stages in petrogenetic processes.  相似文献   

17.
Seventy-eight molecules have been detected as a result of a spectral survey of the star-forming region DR21(OH) at 84–115 GHz. The abundances of most molecules are typical of those in the dense cores of molecular clouds. The rotational temperatures derived using the lines of most molecules fall in the range 9–56 K, which is also typical for dense cores. However, emission from high-lying levels of methanol and sulfur dioxide was detected; since the rotational temperatures for methanol and sulfur dioxide are 252 and 186 K, this indicates the presence of hot regions. Another fact indicating the existence of hot regions is the detection of CH3OCHO, CH3CH2OH, and CH3OCH3, which have thus far been observed only in hot cores and shock-heated regions. An interesting result is the tentative detection of the J = 2 − 1, v = 1 SiO line, with the upper level energy of 1775 K. This is probably a maser line, similar to but weaker than the well-known SiO masers in the star-forming regions Orion-KL,W51(N), and Sgr B2(N).  相似文献   

18.
Observations of 26 regions of low-mass star formation and 17 regions of massive star formation in the 5−1-40 E, 70-61 A +, 80-71 A +, and 2K-1K methanol lines at 44.1, 84.5, 95.2 GHz, and 96.7 GHz yielded detections of methanol emission in 11 low-mass and 12 high-mass regions. The strongest lines in the low-mass regions were found towards bipolar outflows driven by Class 0 protostars with luminosities higher than or of the order of 10 L . These lines usually consist of cores 1–2 km s−1 in width, which are emitted by quiescent gas, and broader wings, emitted by gas accelerated by high-velocity jets. The temperature of the accelerated gas derived from rotational diagrams and statistical equilibrium calculations is roughly 20–50 K. This means that a significant fraction of the accelerated gas cools to such temperatures. The widths of the lines detected in the massive star-forming regions are 2–3 km s−1 or higher. Weak, broad wings were found towards only two sources: L1287 and AFGL5142. For most sources, the statistical-equilibrium calculations yielded gas temperatures of about 20–30 K and densities of about 104–106 cm−3, which are typical for warm clouds. However, different transitions emit in regions with different physical conditions located within the main beam of the telescope. Most of the 96.7 GHz emission arises in warm gas with kinetic temperatures of about 30 K, while most of the 95.2 GHz emission may arise in hot regions around Young Stellar Objects and/or be related to the wings of bipolar outflows. Published in Russian in Astronomicheskiĭ Zhurnal, 2007, Vol. 84, No. 1, pp. 48–59. The article was translated by the author.  相似文献   

19.
Results of interferometric observations of the class I methanol masers OMC-2 and NGC 2264 in the 70-61 A + and 80-71 A + lines at 44 and 95 GHz, respectively, are presented. The maser spots are distributed along the arcs bent toward infrared sources, which are young stellar objects. The distributions of the maser spots at 44 and 95 GHz are virtually identical, and the fluxes from the brightest spots are similar. The measured sizes of the maser spots at 44 GHz are, on average, about 50 AU. The brightness temperature of the strongest components at 44 GHz is 1.7 × 107 K and 3.9 × 107 K for OMC-2 and NGC 2264, respectively. A simple model for the excitation of Class I methanol masers is proposed; it yields an estimate of the limiting brightness temperature of the emission. The model is based solely on the properties of the methanol molecule without invoking the physical parameters of the medium. Using it, we showed that the emission opening angles for NGC 2264 and OMC-2 do not exceed 3° and 4.5°, respectively. The depth of the masing region is about 1000 AU. The emission directivity is naturally realized in the model of of maser consisting of a thermalized core and a thin inverted envelope, probably, with an enhanced methanol abundance. The maser emission has the greatest intensity in the direction tangential to the envelope. The size of the masing envelope estimated from the measured depth and spot extens is ~2 × 104 AU, or 0.15 pc. This size is close to the sizes of the dense molecular cores surrounding the young stellar objects IRS 4 in OMC-2 and IRS 1 in NGC 2264.  相似文献   

20.
Six young bipolar outflows in regions of low-intermediate-mass star formation were observed in the 70-61 A +, 80-71 A +, and 5−1-40 E methanol lines at 44, 95, and 84 GHz, respectively. Narrow features were detected towards NGC 1333-IRS4A, HH 25MMS, and L1157-B1. The flux densities of the detected lines are not higher than 11 Jy, which is much lower than the flux densities of strong maser lines in regions of high-mass star formation. Analysis shows that the narrow features are most likely masers. Published in Russian in Astronomicheskiĭ Zhurnal, 2006, Vol. 83, No. 4, pp. 327–336. This text was submitted by the authors in English.  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