Filename
stringlengths
22
64
Paragraph
stringlengths
8
5.57k
Processed_He_and_Ne_ages_of_large_presolar_silicon_carbide_g.txt
Zinner, E. 2007, in Treatise on Geochemistry Update, ed. H. D. Holland, K. K. Turekian & A.
Processed_He_and_Ne_ages_of_large_presolar_silicon_carbide_g.txt
which after decay contributes ~50% of the final 3He yield, is similar to that for directly produced 3He.
Processed_He_and_Ne_ages_of_large_presolar_silicon_carbide_g.txt
spectra based on the momentum distribution given by Greiner et al. (1975) (see text and Figure 2).
Processed_He_and_Ne_ages_of_large_presolar_silicon_carbide_g.txt
while open circles (Ne not detected) show upper limits based on upper limits to total 21Ne.
Processed_Growth_of_human_population_in_Australia,_1000-10,0.txt
Nielsen, R. W. aka Nurzynski, J. (2013). Growth of human population in Australia, 1000-10,000 years BP.
Processed_Growth_of_human_population_in_Australia,_1000-10,0.txt
None of the two equations [eqns (1) or (2)] fit the data (Johnson & Brook, 2013).
Processed_The_"universal"_radio_X-ray_flux_correlation_:_the.txt
The “universal” radio/X-ray flux correlation : the case study of the black hole GX 339−4.
Processed_The_"universal"_radio_X-ray_flux_correlation_:_the.txt
S. Corbel1,2⋆ , M. Coriat3,1, C. Brocksopp4, A.K. Tzioumis5, R. P. Fender3, J.A. Tomsick6, M. M. Buxton7, C. D. Bailyn7 1Laboratoire AIM (CEA/IRFU - CNRS/INSU - Universit´e Paris Diderot), CEA DSM/IRFU/SAp, F-91191 Gif-sur-Yvette, France 2Institut Universitaire de France, 75005 Paris, France. 3School of Physics and Astronomy, University of Southampton, Highfield, Southampton, SO17 1BJ, UK. 4Mullard Space Science Laboratory, University College London, Holmbury St Mary, Dorking, Surrey RH5 6NT, UK. 5Australia Telescope National Facility, CSIRO, P.O. Box 76, Epping NSW 1710, Australia. 6Space Sciences Laboratory, 7 Gauss Way, University of California, Berkeley, CA 94720-7450, USA. 7Astronomy Department, Yale University, P.O. Box 208101, New Haven, CT 06520-8101, USA.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
The Euclid mission of the European Space Agency will provide weak gravitational lensing and galaxy clustering surveys that can be used to constrain the standard cosmological model and its extensions, with an opportunity to test the properties of dark matter beyond the minimal cold dark matter paradigm. We present forecasts from the combination of the Euclid weak lensing and photometric galaxy clustering data on the parameters describing four interesting and representative non-minimal dark matter models: a mixture of cold and warm dark matter relics; unstable dark matter decaying either into massless or massive relics; and dark matter experiencing feeble interactions with relativistic relics. We model these scenarios at the level of the non-linear matter power spectrum using emulators trained on dedicated N -body simulations. We use a mock Euclid likelihood and Monte Carlo Markov Chains to fit mock data and infer error bars on dark matter parameters marginalised over other parameters. We find that the Euclid photometric probe (alone or in combination with cosmic microwave background data from the Planck satellite) will be sensitive to the effect of each of the four dark matter models considered here. The improvement will be particularly spectacular for decaying and interacting dark matter models. With Euclid , the bounds on some dark matter parameters can improve by up to two orders of magnitude compared to current limits. We discuss the dependence of predicted uncertainties on different assumptions: inclusion of photometric galaxy clustering data, minimum angular scale taken into account, modelling of baryonic feedback effects. We conclude that the Euclid mission will be able to measure quantities related to the dark sector of particle physics with unprecedented sensitivity. This will provide important information for model building in high-energy physics. Any hint of a deviation from the minimal cold dark matter paradigm would have profound implications for cosmology and particle physics.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
Understanding the nature of dark matter (DM) is one of the priority targets within the communities of cosmology, as- troparticle, and high-energy physics. Over the past decade, the Large Hadron Collider (LHC) results and the absence of direct or indirect DM detection have shown that the sit- uation concerning the nature of DM is wide open. Weakly Interacting Massive Particles (WIMPs) are only one candi- date among many possibilities (Bertone et al. 2005; Feng 2010). Particle-like DM could have a large range of plausi- ble mass, lifetime, annihilation cross-section, and scattering cross-sections.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
et al. 2004; Audren et al. 2014), and cross-sections describ- ing either its self-interaction (Spergel & Steinhardt 2000; Feng et al. 2009) or its feeble interaction with other species (Boehm et al. 2001; Cyr-Racine et al. 2016).
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
From a high-energy physics point of view, non-standard DM models are easy to motivate. The logic pursued success- fully by high-energy physicists for almost a century consists of postulating additional symmetries (rather than adding individual particles) in order to explain unaccounted ex- perimental results. The current standard model of particle physics is known to be incomplete (Workman et al. 2022) and the assumption of new symmetries usually comes to- gether with a rich dark sector, i.e., several new particles with new interactions, with potentially more than one pop- ulation surviving until today and contributing to DM or dark radiation. From this point of view, having just one decoupled, stable, and cold relic in our Universe does not sound much more natural than being surrounded by one or more dark species with potentially non-trivial properties. High-energy physicists often suggest that, given the rich- ness of the visible sector, there is no obvious reason for the dark sector to reduce to a single CDM relic particle.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
The astrophysics community is sometimes reluctant to investigate the possible consequences of non-minimal particle-physics assumptions in cosmology as long as the minimal ΛCDM model has not been ruled out. The situa- tion is however evolving given the accumulation of tensions or unresolved questions in cosmological observations (like the small-scale CDM crisis, Hubble tension or S8 tension), see for instance Verde et al. (2019), Abdalla et al. (2022). In this context, it sounds at least reasonable to investigate the possibility of detecting some effects induced by non- minimal DM models. Of course, it is still possible that fu- ture observations only provide bounds on these models and leave us with plain CDM as a preferred case. Even in this case, it would be extremely interesting for particle physics model-builders to have such bounds, since constraints from accelerators or astroparticle experiments usually probe a different regime or different model assumptions than cos- mological data.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
ies and the spectra of tens of millions of galaxies out to redshift of about 2. The combination of spectroscopy and photometry will allow us to reconstruct the matter power spectrum up to 1% accuracy. Since the matter power spectrum could be strongly affected by the nature of DM, Euclid is a perfect tool for testing non- minimal DM properties. It may either confirm the standard CDM paradigm or discover some new DM features. The goal of this work is precisely to estimate the sensitivity of Euclid to different DM parameters beyond its mere relic density. Given the wide range of possible alternatives to standard CDM, we cannot explore all possibilities. We will concentrate on four examples of non-minimal scenarios that are still compatible with current data and could be either constrained or detected by Euclid . Our choice of models is dictated by simple considerations. First, we should se- lect some representative cases. Since in non-minimal mod- els, DM particles are expected to either free-stream (with some velocity dispersion) and/or decay (with some rate) and or scatter (with some cross-sections), we pick up ex- amples in each of these three categories. A well-motivated example of DM with a velocity dispersion is warm DM; some simple examples of decaying DM consist of particles with a constant decay rate, decaying either into relativistic or non-relativistic daughter particles; and a representative case of scattering DM is that of DM interacting with dark radiation. Second, we are interested in models such that galaxy redshift surveys could provide stronger bounds than other observables, and in particular, than cosmic microwave background (CMB) and/or Lyman-α forest data. For rea- sons detailed in the next sections, this would not be the case for pure warm DM or pure decaying DM. Thus, going to the next level of complexity, we will assume a mixture of either cold and warm dark matter, or of stable and unsta- ble particles. In conclusion, we will focus on four interesting and representative models: a mixture of cold and warm dark matter, a mixture of stable and unstable particles decaying either into relativistic or non-relativistic particles, and dark matter interacting with dark relativistic relics.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
Non-minimal DM properties may affect the growth of structures in the Universe in different ways, at different times, and on different scales. Thus, they can leave sev- eral types of signatures in the 2-point correlation function of matter fluctuations in Fourier space, called the mat- ter power spectrum. This spectrum can be reconstructed from several types of cosmological observations at differ- ent redshifts. The modified growth of structure induced by non-minimal DM models could also affect other statisti- cal probes of structure formation (higher-order correlation functions, halo mass function, peak and void statistics), but in this work we only consider its impact on the matter power spectrum.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
Euclid will deliver several types of observations. Among these, the weak lensing (WL) survey and the galaxy clus- tering (GC) photometric survey will be ideal to constrain DM properties, since they will both provide a measurement of the matter power spectrum down to small scales and up to high redshift. These two surveys will return maps in tomographic bins that can be analysed all together (tak- ing into account cross-correlations between WL and galaxy density maps). In addition to this joint data set, called the photometric probe, Euclid will provide other observations. The Euclid spectroscopic galaxy redshift survey will play an essential role for constraining several cosmological mod- els and parameters. Cluster number counts will also convey very useful information. However, these surveys will not provide information on such small scales as WL, and their implementation in sensitivity forecasts relies on a different methodology than for the photometric probe. In particular, they require a different approach to model non-linear effects for each non-minimal DM model. Thus, for simplicity, we choose to concentrate only on the Euclid photometric probe in this work.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
tions, their free parameters and their effects on the linear matter power spectrum. In Sect. 3, we explain how to model the effect of these scenarios at the level of the non-linear power spectrum, using emulators trained on dedicated N - body simulations. In Sect. 4, we summarise the assumptions and numerical pipelines used in our parameter sensitivity forecasts for the Euclid photometric probe. We present our results for each model in Sect. 5 and provide final conclu- sions in Sect. 6.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
Many particle DM properties can be tested with cosmol- ogy (Gluscevic et al. 2019). As mentioned in the introduc- tion, we only focus here on four particular models. On the one hand, these models constitute representative samples of the three most plausible properties of non-minimal par- ticle DM: a non-negligible velocity dispersion, some decay rate, or a non-negligible scattering rate. On the other hand, within their respective category, they account for the sim- plest scenarios that can be constrained better by weak lens- ing and galaxy surveys than CMB and Lyman-α data.
Processed_Euclid_preparation._Sensitivity_to_non-standard_pa.txt
In this model, a fraction fwdm of the total DM fractional density Ωdm is assumed to be warm, so that Ωdm = Ωcdm + fwdm) Ωdm+fwdm Ωdm. The warm dark matter Ωwdm = (1 (WDM) component possesses a thermal (root mean square) velocity vrms that depends on the temperature-to-mass ra- tio Twdm/mwdm. WDM would revert to CDM in the limit vrms This mixed cold plus warm dark matter (CWDM) model has been studied previously, for instance, in Boyarsky et al. (2009a), Schneider (2015), Murgia et al. (2017), Murgia et al. (2018), Parimbelli et al. (2021), or Hooper et al. (2022). It may account either for cosmologies with two dis- tinct DM components, or also, effectively, for cosmologies with a single DM component with a non-thermal distri- bution, such as resonantly-produced sterile neutrinos (Bo- yarsky et al. 2009b). This model has been often invoked as a possible solution to the small-scale CDM crisis (Ander- halden et al. 2013; Maccio et al. 2013). Current best con- straints come from Lyman-α forest surveys (Hooper et al. 2022), Milky Way satellites (Diamanti et al. 2017), and WL surveys (Hervas-Peters et al. 2023), see Sect. 5.1.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
We have performed monitoring observations of the 3-mm flux density to- ward the Galactic Center compact radio source Sgr A* with the Australia Tele- scope Compact Array since 2005 October. Careful calibrations of both elevation- dependent and time-dependent gains have enabled us to establish the variability behavior of Sgr A*. Sgr A* appeared to undergo a high and stable state in 2006 June session, and a low and variable state in 2006 August session. We report the results, with emphasis on two detected intra-day variation events during its low states. One is on 2006 August 12 when Sgr A* exhibited a 33% fractional varia- tion in about 2.5 hr. The other is on 2006 August 13 when two peaks separated by about 4 hr, with a maximum variation of 21% within 2 hr, were seen. The observed short timescale variations are discussed in light of two possible scenar- ios, i.e., the expanding plasmon model and the sub-Keplerian orbiting hot spot model. The fitting results indicate that for the adiabatically expanding plasmon model, the synchrotron cooling can not be ignored, and a minimum mass-loss rate of 9.7 × 10−10M⊙ yr−1 is obtained based on parameters derived for this mod- ified expanding plasmon model. Simultaneous multi-wavelength observation is crucial to our understanding the physical origin of rapid radio variability in Sgr A*.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
There is compelling evidence that Sagittarius A∗ (Sgr A*), the extremely compact radio source at the dynamical center of the Galaxy, is associated with a 4 × 106M⊙ black hole (Eckart & Genzel 1996; Ghez et al. 2000; Sch¨odel et al. 2002; Eisenhauer 2003). Since its discovery in 1974 (Balick & Brown 1974), Sgr A* has been observed extensively with radio telescopes in the northern hemisphere, and temporal flux variations at millimeter wavelengths were reported. With VLA observations, Yusef-Zadeh et al. (2006b) detected an increase of flux density at a fractional level of 7% and 4.5% at 7- and 13-mm, respectively, with a duration of about 2 hr. The peak flare emission at 7-mm led the 13-mm peak flare by 20-40 minutes. Mauerhan et al. (2005) detected intra-day variations (IDVs) of about 20% and in some cases up to 40% at 3-mm using the Owens Valley Radio Observatory (OVRO). The rise and decay occurred on a timescale of 1-2 hr. At 2-mm, Miyazaki et al. (2004) reported a 30% flux increase in 30 minutes from the monitoring of the Nobeyama Millimeter Array (NMA). On the other hand, flares with violent intensity increases in very short timescales have also been detected at infrared and X-ray bands (Genzel et al. 2003; Baganoff et al. 2001; Eckart et al. 2006b), inferring that these emissions from Sgr A* originate within very vicinity of the central massive black hole. This is further strengthened by the simultaneous detection of X-ray, infrared and sub-mm flares (Eckart et al. 2004, 2006a, 2008a, 2008b; Yusef-Zadeh et al. 2006a, 2008; Marrone et al. 2008).
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Since Sgr A* is embedded in thick thermal material, it is particularly difficult to observe its intrinsic structure. But observations of IDV can give indirect constraints on the source emission geometry and emission mechanisms. However, previous monitoring observations of Sgr A* from the northern hemisphere have been strictly limited to a short observing window (< 7 hr/day) for the Galactic Center region. We have performed monitoring observations of flux density toward Sgr A* at 3-mm since 2005 October when for the first time the Australia Telescope Compact Array (ATCA) of the Australia Telescope National Facility (ATNF) was available at 3-mm. The ATCA is an interferometer consisting of five 22-m radio telescopes at Narrabri, Australia where Sgr A* passes almost overhead, allowing a much longer observing window (> 8 hr at elevation angles above 40◦). As such, the ATCA calibrations and flux density measurements of Sgr A* are expected to be more accurate.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Smax−Smin 2 (Smax+Smin), here Smax and Smin refer to the maximum and minimum value of flux density, respectively.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
In 2005 and 2006, we performed 3-mm ATCA flux density monitoring of Sgr A* over 50 hr in the following 3 sessions: 2005 October 18, 2006 June 9 and August 9-13. Dual (linear) polarization double sideband (DSB) HEMT receivers were used. The first ever 3-mm ATCA monitoring of Sgr A* was performed on 2005 October 18 when the data were simultaneously recorded, in both the lower (93.504 GHz) and upper (95.552 GHz) sidebands, in 32 channels of a total bandwidth of 128 MHz. For the observations in 2006, the data were recorded in two slightly different 3-mm bands: the lower sideband (86.243 GHz) was set to the transition frequency of the SiO J=2-1 v=1 line with 256 channels of a total bandwidth of 16 MHz, the upper sideband (88.896 GHz) was a wideband with 32 channels of a total bandwidth of 128 MHz. Since the continuum data of the lower sideband with narrow bandwidth have relatively low signal to noise ratio, only the upper sideband data were used for Sgr A* and other continuum sources.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
On 2005 October 18, we observed Sgr A* in the H168C configuration of the ATCA, with a maximum baseline of 192 m, uv range of 13 - 61kλ and a synthesized beam of 2.′′9×1.′′7. On 2006 June 9, the observations were performed in the 1.5D configuration with a maximum baseline of 1439 m, covering uv range of 20 - 430kλ and yielding a synthesized beam of 2.′′1 × 0.′′3. In 2006 August, the observations were performed in SPLIT5 configuration with a maximum baseline of 1929 m, covering uv range of 3 - 570kλ and yielding a synthesized beam of 1.′′3 × 0.′′2. In this array, the spacing between antennas 2 and 3 and antennas 3 and 4 are only 31 m, causing severe shadow effect, especially for antenna 3.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
measurements on June 9, August 12 and 13. The pointing accuracy was checked every half an hour by observing VX Sgr, a known strong SiO maser source. The instrumental gain and phase were calibrated by alternating observations of Sgr A* and all secondary calibrators. In 2006 June, the observations were performed using the following sequence: OH2.6-0.4 (2 min), Sgr A* (5 min), PKS 1730-130 (2 min), and PKS 1921-293 (1 min). In 2006 August, the observing sequence was PKS 1710-269 (2 min), OH2.6-0.4 (1 min), Sgr A* (10 min), PKS 1710-269 (2 min), OH2.6-0.4 (1 min), PKS 1730-130 (1 min), and PKS 1921-293 (1 min). The observing details have been summarized in Table 1.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
All the data processing was conducted using the ATNF MIRIAD package (Sault et al. 1995). At millimeter wavelengths, the atmosphere can no longer be approximately transparent. The opacity effect is included in an effective system temperature - the so-called “above atmosphere” system temperature (Ulich 1980) for the ATCA measurements at 3- mm. The bandpass corrections were made using the strong ATCA calibrator 3C279. For amplitude calibration, we first applied a nominal elevation-dependent gains of the antennas and then used calibrators to further determine the additional corrections. On 2005 October 18, the flux scale was based on observation of Uranus. On 2006 June 9, the flux density scale was determined using PKS 1730-130, assuming its flux density of 2.27 Jy at 3-mm. In 2006 August, we derived the flux density scale with another brighter radio source PKS 1921-293, which is reported to be 8.44 Jy during our observations from the ATCA calibrator list on web. From the ATCA calibrator flux density monitoring data during 2003 to 2006, we estimated its mean flux density of 8.66 Jy with a standard deviation of 1.04, implying a dispersion of about 12%. PKS 1921-293 is probably better than that of other calibrators simply because PKS 1921-293 data usually have very high signal-to-noise ratio. So, we expect an accuracy ≤ 20% for the absolute amplitude calibration in these observations. After the phase self- calibration, the data were averaged in 5 minutes bin to search for shorter timescale variability. The flux density of Sgr A* was estimated by fitting a point source model to visibilities on the projected baselines longer than 25kλ (about 85 m at 3-mm) to suppress the contamination from the surrounding extended components (Miyazaki et al. 2004; Mauerhan et al. 2005). Both the fitting error reported by MIRIAD and the rms of the residual visibilities were used to get the final error estimate.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
in a large fluctuation in its gain correction and thus a bigger modulation index. During observations in 2006 August, the 3-mm flux density of PKS 1710-269 is around 0.5 Jy, only one fiftieth of that of OH2.6-0.4, therefore the signal-to-noise ratio of PKS 1710-269 data is much lower than that of OH2.6-0.4. This explains why the modulation indices of gain corrections derived from PKS 1710-269 are relatively high.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
During the first 3-mm ATCA observation of Sgr A* on 2005 October 18, the flux density of Sgr A* ran up to 3.5 Jy, much brighter than the normally expected 1.5 Jy in the quiescent phase, and thus Sgr A* was very likely to be in an active phase during our observation. Although Sgr A* seems to vary in its total flux density as a function of time, we can not rule out the possibility of the elevation-dependent gain effect as use of the only secondary calibrator PKS 1730-130 (16.2◦ from Sgr A*) severely limited the amplitude calibration. Because of this, starting from observations in 2006 June and August (see § 2), we paid a particular attention to the strategy of calibration. As a result, the light-curves of Sgr A* at 3-mm in 2006 are reliably obtained (Figure 2). All the data were calibrated using OH2.6-0.4. The flux densities were estimated by fitting a point source model to visibility data on the projected baselines longer than 25kλ. Following is a detailed description of the results from each observation.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
The flux density of Sgr A* was relatively high (around 3 Jy) but stable on 2006 June 9. As shown in Figure 1, the modulation index of Sgr A* is small and comparable to that of antenna gains. Therefore, no IDV was detected.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
During the first three days in the 2006 August session (August 9, 10 and 11), very limited data were available. The flux density of Sgr A* was decreased from 2.52 to 2.25 Jy in 1 hr on August 9, stayed around 1.9 Jy quite stably during the 4 hr run on August 10 and around 2.0 Jy over the 2 hr observation on August 11. So, we conclude that no ascertained IDV was detected.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Two clear IDV events were seen in the last two days of the 2006 August session. As shown in the light-curves of Sgr A* and other sources on 2006 August 12 (Figure 3 left), first the flux density of Sgr A* decreased from 1.65 to 1.50 Jy, and then increased to 2.11 Jy in 2.5 hr before decreasing again to 1.90 Jy. The fractional flux density variation is estimated to be 33%. On 2006 August 13 (Figure 3 right), the flux density of Sgr A* first increased from 1.95 to 2.14 Jy, reached its first peak before decreasing to 1.80 Jy in 1.7 hr. Then it reached the second peak 2.22 Jy in 1.9 hr, and declined to 1.98 Jy in 1.2 hr. The maximum fractional flux density variation is 21% with a timescale of about 2 hr. As shown in Figure 1, the modulation indices of Sgr A* on both August 12 and 13 are much greater than that of gain corrections, supporting that the observed flux density variations are most likely to be real.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
al. 2003), which are in accord with our ATCA observations. As is shown in Figure 2, the mean flux density of Sgr A* dropped from 2.97 to 2.16 Jy from 2006 June to August session. The day-to-day fractional variation of Sgr A* appeared to be low from 2006 August 10 to 13. Comparison of flux densities in two observing sessions in 2006 indicates that Sgr A* appeared to undergo a high state in 2006 June session, and a low state in 2006 August session. Such different states were also noted by Herrnstein et al. (2004). They found a bimodal distribution of flux densities at centimeter wavelength and thought that it might indicate the existence of two distinct states of accretion onto the supermassive black hole. Different radiation states usually have connections to certain physical parameters or radiation model. For example, a unified inner advection dominated accretion flow (ADAF) model with different accretion rates and consequently different geometries of accretion flow has been proposed to explain five distinct spectral states that have been identified in black hole X-ray binaries (BHXBs), namely the quiescent, low, intermediate, high and very high states (Esin et al. 1997). Supposing that the accretion model of Sgr A* has something in common with that of BHXBs, the accretion rate in 2006 August session should be smaller than that in June session, but the accretion rate may not change much over days in August.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Expanding plasmon model of van der Laan (1966) was invoked to explain observed time delay in variation of Sgr A* at 7- and 13-mm (Yusef-Zadeh et al. 2006a, 2006b, 2008). In this model, rather than the synchrotron cooling, the adiabatic cooling associated with expansion of the emitting plasma is responsible for the decline of flare. Flaring at a given frequency is produced through the adiabatic expansion of an initially optically thick blob of synchrotron-emitting relativistic electrons. The initial rise of the flux density is produced by the increase in the surface area of blob while it still remains optically thick; the curve turns over once the blob becomes optically thin because of the reduction in the magnetic field, the adiabatic cooling of electrons, and the reduced column density as the blob expands. Such kind of blob ejected from an ADAF is also thought to be a possible explanation for nonthermal flares and recombination X-ray lines in low-luminosity active galactic nuclei and radio-loud quasars (Wang et al. 2000).
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Our observed IDVs with different amplitudes and timescales seem consistent with the expanding plasmon model in the context of jet or outflow. The amplitudes and timescales vary with the relativistic particle energy distribution, expanding velocity and size of the blob. To apply the model to the light-curves on 2006 August 12 and 13, we first assumed a power-law spectrum of the relativistic particle energy (n(E) ∝ E−p). Hornstein et al. (2007) reported a constant spectral index of 0.6 using multi-band IR observations of several flares. Here we adopt a spectral index of 0.6, corresponding to the particle spectral index of 2.2, the energy of the particles was assumed to range from 10 MeV to 3 GeV. The expanding velocity was supposed to be constant. As is stated by Yusef-Zadeh (2008), the relationship between the quiescent and flaring states of Sgr A* is not fully understood. Their results indicate that the quiescent emission at 7- and 13-mm varies on different days. The minimum flux density was 1.5 Jy during our 2006 August observing session, the quiescent flux density, if it does exist, should not be more than this value. We then assume a quiescent flux density of 1.4 Jy, while the flare is produced by the blob. Other parameters were derived by means of the weighted least square method. We adopt exponentially increasing step length for number density during the fitting in order to improve efficiency. The uncertainties of the parameters were assessed by scaling up the 68.3 % confidence region of parameter space, as an increase of χ2 from χ2 ν with the reduced chi squares, χ2 min/Ndof , where Ndof is the difference between the number of data and the number of fitting parameters (c.f. Shen et al. 2003).
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
which implies that τcrit(R0) mainly depends on p. Since c1γ0R−3 0, B0 is the most sensitive parameter for the evolution of flux S(ν, R). To illustrate this, we choose typical values of p = 2.2, γ0 = 20, R0 = 4rg and v = 0.004c and show the resulting model light curves at 90 GHz while B0 is 10, 20, 30, 40 and 50 Gauss in Figure 4. Results from expanding plasmon model of van der Laan (1966) are shown in dotted lines, and results from expanding plasmon model with synchrotron cooling are shown in solid lines. The differences between two results are significant above 30 Gauss, implying that the synchrotron cooling of electrons should not be ignored for strong magnetic field.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
The best-fit model for the light-curve of 2006 August 12 is plotted as a solid line in Figure 5 left. Two blobs were required to fit the data, which were assumed to appear at 7.0 and 10.3 UT. We attribute the turnover in light curve to the birth of a new blob, so the second blob was assumed to appear before the flux increases. The corresponding initial −1 × 10−3c, electron blob radius 1.8+1.6 number density of 2.56−2.52 × 107cm−3 and 5.12+5.12 −5 and 7+4 −2 Gauss were derived from the fit. The uncertainty that was failed to be assessed was left blank, if not such a sensitive parameter. The peak flux densities of two blobs are estimated to be 0.26 and 0.59 Jy, respectively. The half-power durations are 1.7 and 5.2 hr, respectively. Blob mass of 2.3 × 1020 g and 1.4 × 1021 g were estimated.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
In principle, the expanding plasmon model can also be used to interpret the 2000 March 7 NMA short millimeter flare reported by Miyazaki et al. (2004). In their observation, the peak flux density at the 140 GHz band is apparently larger than that at the 100 GHz band. The spectral variation suggests that the energy injection to photons occurred in the higher frequency regime first and the emitting frequency was shifted to the millimeter-wavelength regime with time, which is well consistent with the scenario predicted by expanding plasmon model. A time delay of 1.5 hr was observed for NIR and sub-mm flare on 2008 June 3 (Eckart et al. 2008b), which has been explained with a similar model with adiabatically expanding source components. There, the spectral index (0.9 to 1.8), expansion velocity (0.005c) and source size (∼2 rg) are fairly consistent with the parameters derived here. In order to compare with their modeling results, we calculate the optical depth at sub-mm based on parameters derived here. Take the first blob of 2006 Aug 13 for example, according to Pacholczyk (1970), the optical depth at 90 GHz is calculated to be 8.23 at t0. The critical optical depth at which the flux density for any particular frequency peaks at t0, is calculated to be 1.82 based on Eq.(12). According to Eq.(7), the optical depth at 345 GHz is 0.14, which is smaller than the critical value. Therefore, at both NIR and sub-mm wavelengths, the emission is optically thin at the beginning. Since we attribute the turnover in light curve to the emergence of a new blob, so the new blob is assumed to appear right before the observed flux density increases. Extending the model in time might help give time delay between NIR and sub-mm, however, the birth time of blob is difficult to determine. Future simultaneous multi-wavelength observation, especially the correlation between optically thin (such as NIR/X-ray) and 3-mm flaring emission is expected to help improve the model fitting.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
explaining the NIR 17 minutes quasi-periodic oscillation (Genzel et al. 2003). The hot spot model is applied to radio band by including the effects of disk opacity for a typical RIAF model (Broderick & Loeb 2006). In these studies, the hot spot is always close to the innermost stable circular orbit (ISCO), thus the NIR 17 minutes quasi-periodic oscillation can be produced. Since the creation of such a hot spot is still under discussion, it is also possible that such kind of spot may appear somewhere away from the ISCO and thus produce quasi-periodic oscillation with a longer timescale.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
In the accretion disk, neighboring annuli of differentially rotating matter experience a viscous shear that transports angular momentum outwards and allows matter to slowly spiral in towards the center of the potential (Merloni 2002). As a result, the gas rotates with a sub-Keplerian angular velocity (Narayan et al. 1997). In the following we assume that the rotation of hot spot is also sub-Keplerian and fit our detected IDV events using a sub-Keplerian rotating hot spot model. To simplify the calculation, the angular velocity is assumed to be 0.4 times of Keplerian angular velocity of a Schwarzschild black hole.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
We assumed the values of most physical parameters of the hot spot model the same as those in the expanding plasmon model when starting fitting the hot spot model to light- curves. These include: the energy range of relativistic particles, the particle spectral index and the quiescent flux density. Then we estimated other parameters by means of weighted least square method. The magnetic field was assumed to range from 1 to 100 Gauss. In RIAF model, the electron number density of accretion disk is about 2 × 106 cm−3 at a distance 20rg from the central black hole (Yuan et al. 2003). Since the hot spot is modeled by an overdensity of non-thermal electrons, it is safe to assume the electron number density ranges from 4 × 106 to 1 × 108 cm−3. In addition, the accretion disk is assumed to be edge-on to maximize the boosting effect (Huang et al. 2007; 2008). The final result is the combination of the quiescent flux density and flux density of hot spot. The derived parameters are summarized in Table 3.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
The hot spot model for the light-curve of 2006 August 13 is plotted as a solid line in Figure 6 right. The quiescent flux density was 1.4 Jy too. One hot spot is required to fit the data. Radius of 4.1 ± 0.3rg, magnetic intensity of 6+2 −5 Gauss and electron number density of 6+44 × 106 cm−3 were derived. The hot spot is at 11.4+0.4 −0.2rg from the central black hole.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
The electron cooling timescales due to synchrotron losses are calculated to be greater than 2 days at 90 GHz, much longer than the observed variation timescale, thus the syn- chrotron energy loss can be ignored in the fitting. Since the synchrotron cooling time is long, the life time of hot spot should mainly depend on the dynamical timescale. Reid et al. (2008) analyzed the limits on the position wander of Sgr A*, ruling out the possibility of hot spots with orbital radius above 15rg that contribute more than 30% of the total 7-mm flux. All the orbital radius listed in Table 3 are smaller than 15rg. Hence, the presented hotspot model is not in contradiction with their result.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
The discussion above shows that both the expanding plasmon model and the orbiting hot spot model can be used to interpret the detected two IDV events. Given that the former model predicts a time delay in flare emission, while the latter does not, the time delay between different frequencies in flare emission is believed to be critical to distinguish between them (Yusef-zadeh et al. 2006b). Recently, Yusef-zadeh et al. (2008) detected time lags of 20.4±6.8, 30±12 and 20±6 minutes between the flare peaks observed at 13-mm and 7-mm. At shorter wavelength, a possible time delay of 110±17 minutes between X-rays and 850-µm was observed (Marrone et al. 2008). Though these observations seem to support the expanding plasmon model, the hot spot model is still a possible explanation, especially for the observed nearly symmetrical light-curves.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
We presented the results of the ATCA flux density monitoring of Sgr A* at 3-mm, with emphasis on the detected two IDV events. Comparison of flux densities in two observing sessions in 2006 indicates that Sgr A* appeared to undergo a high state in June session, and a low state in August session. On 2006 August 12, Sgr A* exhibits a 33% fractional variation in about 2.5 hr. Two peaks with a separation of 4 hr are seen on 2006 August 13 flare which exhibits a maximum variation of 21% within 2 hr.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
these derived parameters. We assume that the rotation of hot spot is sub-Keplerian while applying the hotspot model. It seems that both models can reasonably fit the detected IDV events. Future simultaneous multi-wavelength monitoring is expected to discriminate them and tell us where such kind of IDV events come from.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
The Australia Telescope Compact Array is part of the Australia Telescope which is founded by the Commonwealth of Australia for operation as a National Facility managed by the CSIRO.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
This work was supported in part by the National Natural Science Foundation of China (grants 10573029, 10625314, 10633010 and 10821302) and the Knowledge Innovation Pro- gram of the Chinese Academy of Sciences (Grant No. KJCX2-YW-T03), and sponsored by the Program of Shanghai Subject Chief Scientist (06XD14024) and the National Key Basic Research Development Program of China (No. 2007CB815405).
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Trippe, S., Paumard, T., Ott T., Gillessen, S., Eisenhauer, F., Martins, F., & Genzel, R.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Fig. 1.— Modulation index of flux density of Sgr A*, and gain corrections of five antennas (labeled 1, 2, 3, 4 and 5). They are derived from calibrators OH2.6-0.4 and PKS 1710-269 (from left to right) for three observations on 2006 June 9, August 12 and 13 (from top to bottom). Many data obtained from antenna 3 in August session were shadowed and have been flagged, so the uncertainty of this antenna is big compared with other antennas. During observation in 2006 August, the 3-mm flux density of PKS 1710-269 was around 0.5 Jy, only one fiftieth of that of OH2.6-0.4, therefore the modulation index of gain corrections derived from this source are particularly high.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Fig. 2.— ATCA 3-mm light-curves of Sgr A* in 2006. Two detected IDV events are indicated (arrows).
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Fig. 3.— ATCA 3-mm light-curves on 2006 August 12 (Left) and 13 (Right) of Sgr A* (middle panel), secondary calibrators OH 2.6-0.4 (top panel) and PKS 1710-269 (bottom panel).
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Fig. 4.— Two kind of theoretical model light curves as a function of expanding blob radius at 90 GHz with different B0. Results from expanding plasmon model of van der Laan (1966) are shown in dotted lines, and results from expanding plasmon model with synchrotron cooling are shown in solid lines.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Fig. 5.— The solid line represents the expanding plasmon model fitting to the observed 3- mm light-curves on 2006 August 12 (left) and 13 (right) with synchrotron radiation cooling taken into account. An assumed quiescent flux density of 1.4 Jy is indicated by the straight dotted line. The blobs used to fit the data are indicated by dashed curves.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Fig. 6.— The light-curves produced by the sub-Keplerian orbiting hot spot model on 2006 August 12 (left) and 13 (right). An assumed quiescent flux density of 1.4 Jy is indicated by the straight dotted line. Two hot spots used to fit the data on August 12 (left) are indicated by dashed curves.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Table 1: ATCA Observations of Sgr A* in 2006. Length is duration of the observation. IF1&IF2 are intermediate frequencies for the lower and upper sidebands, respectively, with the corresponding bandwidth of BW1&BW2. Range of baselines are indicated by uv range. Beam is the ATCA synthesized beam.
Processed_The_variability_of_Sagittarius_A*_at_3_millimeter.txt
Table 2: Parameters of the expanding plasmon model. We took into account the synchrotron cooling of electrons while applying the adiabatically expanding plasmon model to the light- curves. Here, t0 is the time at which the blob was assumed to be generated, R0 is the initial radius, v is the expanding velocity, N0 is the initial electron number density, B0 is the initial magnetic field strength, Sp is the peak flux density of the blob, and χ2 ν is the reduced chi squares.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
Abstract— This paper presents a robust distributed coor- that achieves generation of collision-free dination protocol trajectories for multiple unicycle agents in the presence of stochastic uncertainties. We build upon our earlier work on semi-cooperative coordination and we redesign the coordination controllers so that the agents counteract a class of state (wind) disturbances and measurement noise. Safety and convergence is proved analytically, while simulation results demonstrate the efficacy of the proposed solution.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
Coordination in multi-agent systems has attracted much attention over the last decade with a plethora of theoretical and practical problems that this paper can not cite in their entirety; for recent overviews the reader is referred to [1]–[4]. A fundamental problem of interest in the area of dis- tributed coordination and control is the decentralized multi- agent motion planning, which mainly focuses on generating collision-free trajectories for multiple agents (e.g., unmanned vehicles, robots) so that they reach preassigned goal loca- tions under limited sensing, communication, and interaction capabilities. Numerous elegant methodologies on planning the motion for a single agent (robot) have appeared in recent years, with the most popular being (i) sampling- based methods, including probabilistic roadmaps [5], and rapidly-exploring random trees [6], [7], (ii) Lyapunov-based methods, including either the definition of closed-form feed- back motion plans via potential functions or vector fields, or computation of Lyapunov-based feedback motion plans via sum-of-squares programming [8], [9], and (iii) graph search and decision-theoretic methods, see also [10], [11] for a detailed presentation. Although each method has its own merits and caveats, arguably Lyapunov-based methods (often termed reactive) are particularly popular for multi- agent motion planning problems, as they offer scalability with the number of agents, and the merits of Lyapunov-based control design and analysis.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
considering them as bounded disturbances. In [12], a stable uncertainty is assumed to be bounded in H∞-norm by some prior given desired tolerance, and a state space observer along with a robust controller are designed by using the algebraic Riccati equations. In [13], an additive l2-norm bounded disturbance is considered, and a robust controller is proposed for the distributed cooperative tracking problem in a leader-follower network by using Lyapunov stability theorems. In [14], both bounded disturbances and unmod- eled dynamics are assumed in the dynamics of agents; an identifier for each agent is designed to estimate the unknown disturbances and unmodeled dynamics. Related work consid- ering bounded deterministic disturbances can be found in the design of finite-time consensus algorithms with mismatched disturbances [15], and the rotating consensus control with mixed model uncertainties and external disturbances [16].
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
Another way of modeling uncertainty is by Gaussian random processes. In [17], communication noises are de- scribed by a standard Brownian motion, and the mean square consensus in the multi-agent system is achieved by proposing a stochastic approximation-type gain vector, which attenuates the effect of noises. This work is extended to networks with Markovian switching topologies [18], and leader-follower networks [19] with a similar noise-attenuation controller. In [20], it is assumed that the measurements for each agent are disturbed by white noises. Robust consensus can be achieved by applying stochastic Lyapunov analysis. Based on this idea, in [21], the average-consensus problem of first- order multi-agent systems is considered, and a necessary and sufficient condition is proposed for robust consensus by using probability limit theory. The aforementioned methods are efficient in solving the coordination problems with stochastic uncertainties in measurement or system dynamics; however, safety (i.e., the generation of collision-free trajectories) is not considered.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
The authors are with the Department of Aerospace Engineering, Uni- versity of Michigan, Ann Arbor, MI, USA; {kgarg, dongkunh, dpanagou}@umich.edu.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
The authors would like to acknowledge the support of the Automotive Re- search Center (ARC) in accordance with Cooperative Agreement W56HZV- 14-2-0001 U.S. Army TARDEC in Warren, MI and the support of the NASA Grant NNX16AH81A.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
hence offers scalability with the number of agents along with provable guarantees. In summary, the contributions of this paper are: (i) a robust, semi-cooperative coordination protocol that accommodates for a class of stochastic dis- turbances in the agents’ dynamics and measurements, and (ii) the derivation of analytical bounds on the navigation (estimation) and final state errors of the agents in terms of the considered uncertainties.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
The paper is organized as follows: Section II includes an overview of the modeling of the system under the effect of disturbances. Section III presents the robust coordination protocol along with the safety and convergence analysis. Section IV evaluates the performance of the proposed method via two simulation scenarios. Our conclusions and thoughts on future work are summarized in Section V.
Processed_Robust_Semi-Cooperative_Multi-Agent_Coordination_i.txt
radius Rc centered at ri = (cid:2)xi , denoted as Ci : {r ∈ R2 | (cid:107)ri − r(cid:107) ≤ Rc}. We denote Ni the set of neighboring agents k ∈ Ci of agent i. We assume that each agent i can measure the position rk, orientation θk and receive the linear velocity uk of any agent k lying in Ci.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
ATCA search for 21 cm emission from a candidate damped Ly-α absorber at z = 0.101.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Nissim Kanekar1 ⋆, Jayaram N Chengalur1,4⋆⋆, Ravi Subrahmanyan2 ⋆⋆⋆ & Patrick Petitjean3† 1 National Centre for Radio Astrophysics, Post Bag 3, Ganeshkhind, Pune 411 007, India 2 Australia Telescope National Facility, CSIRO, Locked bag 194, Narrabri, NSW 2390, Australia 3 Institut d’Astrophysique de Paris - CNRS, 98bis Boulevard Arago, F-75014 Paris, France 4 Visiting Scientist, NFRA, P O Bus 2, 7990 AA Dwingeloo, The Netherlands.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Abstract. We report a deep search for 21 cm emis- sion/absorption from the z ∼ 0.101 candidate damped Lyman-α system towards PKS 0439−433, using the Aus- tralia Telescope Compact Array (ATCA). The spectrum shows a weak absorption feature — at the 3.3σ level — which yields a lower limit of 730 K on the spin temperature of the system. No H i emission was detected: the 3σ upper limit on the H i mass of the absorber is 2.25 × 109M⊙, for a velocity spread of ∼ 70 km s−1. The low H i mass and the high spin temperature seem to rule out the possibility that the absorber is a large gas-rich spiral galaxy.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
At low redshifts too, DLA systems are expected to be mainly associated with spiral galaxies. Rao & Briggs (1993) used the optical luminosity function and the aver- age H i content of a given galactic morphological type to conclude that the cross section for damped absorption at z = 0 is dominated by large spiral galaxies; ∼ 90% of the H i at z = 0 resides in large spirals. An alternative to this hybrid approach to determining the local H i mass den- sity is to directly use the observed H i mass function. The latter, as determined by blind H i surveys is, however, cur- rently controversial. Zwaan et al. (1997) find that the mass function is well fit by a Schecter function with a fairly flat α = −1.2 slope at the faint end; this implies that the ma- jor contributors to the H i mass density at z = 0 are L⋆ galaxies. On the other hand, Schneider et al. (1998, see also Rosenberg et al. 2000) suggest that the space den- sity of low mass (MH i< 108 M⊙) galaxies is considerably larger than that predicted by the Schecter function fit of Zwaan et al. (1997); a substantial fraction of the H i at z = 0 may then lie in smaller systems.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Damped Lyman-α (DLA) systems are objects with such high neutral hydrogen column densities (NH i >∼ 1020 cm−2), that their optical depth in the damping wings of the Lyman-α line is appreciable. Given a background quasar, this results in a broad absorption feature, eas- ily detectable in even moderate resolution optical spec- troscopy. DLA systems are the main repository of neutral gas at high redshift (z ∼ 3) and have, therefore, tra- ditionally been assumed to be the progenitors of large spiral galaxies (Wolfe 1988, Prochaska & Wolfe 1997, Prochaska & Wolfe 1998). In support of this hypothesis, the comoving mass density of neutral gas in these ob- jects at z ∼ 3 is comparable to the stellar mass density in bright galaxies today (Storrie-Lombardi & Wolfe 2000, Storrie-Lombardi et al. 1996, Lanzetta et al. 1991), con- sistent with the gas having been converted into stars in the intervening period.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
spin temperatures arise naturally in dwarfs because these systems have low metallicities and pressures and, conse- quently, a larger fraction of the warm phase of H i as com- pared to large spiral disks (Chengalur & Kanekar 2000). It is possible, of course, that the absorbing galaxies, though faint in the optical, are nonetheless exceedingly gas-rich, and their large H i envelopes cause them to be preferentially detected in absorption surveys. At the very lowest redshifts, the latter hypothesis can be tested through deep searches for H i 21 cm emission from the absorbers. Such searches yield direct estimates of the H i mass and can thus be used to check whether or not the optically faint galaxies which give rise to DLA absorption have anomalously large H i content. We describe, in this letter, a deep search for 21 cm emission/absorption from a candidate damped absorber at z ∼ 0.101 towards the quasar PKS 0439−433 (Petitjean et al. 1996). The ob- servations were carried out with the Australia Telescope Compact Array (ATCA). No emission was detected, re- sulting in strong constraints on the H i mass of the ab- sorber.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Heliocentric velocity, centred at z = 0.101 Fig. 1. ATCA 9 km s−1 resolution spectrum towards PKS 0439−433 . The x-axis is heliocentric velocity in km s−1, centred at z = 0.101. Weak absorption can be seen, close to v = 0.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
continuum subtraction might result in spectral artefacts (Cornwell et al. 1992). We hence also tried an alternate analysis procedure to check for spectral errors that might arise due to the confusing sources in the field. This in- volved the use of the task UVSUB, to subtract out the individual point sources; UVLIN was then run to remove any residual continuum emission and the resulting data set was again mapped to obtain the spectral data cube. The spectra obtained from the two procedures were found to be identical within the noise for each data set.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
An RMS noise of 1.56 mJy was obtained per chan- nel at the original velocity resolution of 1.8 km s−1; this is close to the theoretical sensitivity of the ATCA, for our observing parameters. Figure 1 shows the spectrum smoothed to a resolution of 9 km s−1: weak absorption can be seen close to z = 0.101. This occurs at a helio- centric frequency of 1290.144 MHz, corresponding to a redshift of 0.10097 ± 0.00003; this is in good agreement with the redshift z = 0.101 obtained from metal lines (Petitjean et al. 1996). The RMS noise on the spectrum is ∼ 0.76 mJy while the feature is ∼ 2.5 mJy deep (and only one channel wide), i.e. a 3.3σ result. The absorption was seen in both the XX and YY polarizations separately, although, of course, at even lower significance levels. Fig- ure 2 shows a zoomed-in version of the spectrum at the original resolution of 1.8 km s−1; the feature of figure 1 can be seen to be a few channels wide here (although within the noise). We note that the narrowness of the feature makes it very unlikely that it arises as an artefact from the continuum subtraction procedure.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Heliocentric velocity, centred at z = 0.101 Fig. 2. Zoomed-in version of the ATCA 1.8 km s−1 reso- lution spectrum towards PKS 0439−433 . The x-axis is heliocentric velocity in km s−1, centred at z = 0.101.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Heliocentric velocity, centred at z = 0.101 Fig. 4. ATCA 70 km s−1 resolution spectrum towards PKS 0439−433 . The x-axis is heliocentric velocity in km s−1, centred at z = 0.101. This spectrum was also ob- tained after dropping antenna 6 and has a noise of 0.29 mJy.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Heliocentric velocity, centred at z = 0.101 Fig. 3. ATCA 30 km s−1 resolution spectrum towards PKS 0439−433 . The x-axis is heliocentric velocity in km s−1, centred at z = 0.101. This spectrum was obtained after dropping antenna 6 and has a noise of 0.48 mJy.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
absorber1. Since typical sizes of spirals tend to be larger than this (∼ 30 − 40 kpc), there was a possibility that we might be resolving out some of the emission. We therefore re-analyzed the data after flagging out antenna 6, which gives rise to the long ATCA baselines: this resulted in a synthesized beam of ∼ 40′′ on the sky, corresponding to a length scale of 50 kpc at z = 0.101. This spectrum was also searched for emission after smoothing to a va- riety of velocity resolutions between 30 km s−1 and 70 km s−1: no statistically significant emission was detected. The 30 km s−1 and 70 km s−1 resolution spectra (after dropping antenna 6) are shown respectively in figures 3 and 4; the RMS noise levels are 0.48 mJy and 0.29 mJy. Finally, the spectrum was also smoothed to a resolution of 200 km s−1 to search for wide emission (note that this spectrum, which is not shown here, has only seven inde- pendent points). Again, no emission was seen; the RMS noise per 200 km s−1 channel is 0.15 mJy.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
where Ts is in K, NH i in cm−2 and dV in km s−1. The covering factor f can usually be estimated from VLBI observations; unfortunately, no such observations exist in the literature for PKS 0439−433 . The source, however, has an exceedingly flat spectrum, with flux values of 330 mJy at 1.29 GHz (our observations), 320 mJy at 2.7 GHz and 300 mJy at 5 GHz (Quiniento et al. 1988), and is thus likely to be very compact. We hence assume a cov- ering factor of unity; this should, of course, be verified by VLBI observations. Next, the z = 0.101 absorber towards PKS 0439−433 is a candidate damped system, with strong Mg II, Al II, Fe II, Si II and C IV absorption lines seen in the HST spectrum (Petitjean et al. 1996). The equiva- lent width ratios of these low-ionization lines indicate that the system is probably damped (see also Rao & Turnshek 2000), with NH i ∼ 1020 cm−2 (although the Lyman-α line has itself not so far been observed for this absorber). This estimate of NH i agrees well with the X-ray spectrum of PKS 0439−433 which shows absorption corresponding to NH i = 2.3 ± 0.8 × 1020 per cm2 (Wilkes et al. 1992), with a Galactic contribution of 1.3 ± 0.1 × 1020 per cm2 (Lockman & Savage 1995). The system is thus quite likely to be a moderate column density damped absorber. Our 9 km s−1 resolution ATCA spectrum has a peak optical depth τmax ∼ 0.0076; equation (1) then yields a column density of NH i = 1.4×1017 Ts cm−2. Combining this with the column density of Petitjean et al. (1996) gives a spin temperature of ∼ 730 K. Note that this is a lower limit to the spin temperature, since the upper limit to the optical depth is ∼ 0.0076. If the 3.3σ feature of figure 1 is not real, it would imply that the system has an even higher spin temperature. As discussed in Chengalur & Kanekar (2000), spiral galaxies tend to have low spin temperatures (Ts <∼ 300 K); it is thus unlikely that the absorber is a large spiral galaxy.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
where ∆V is in km s−1, S is in Jy and MH i is in solar masses. DL is the luminosity distance of the cloud in Mpc (DL = 413.72 Mpc, for a system at z = 0.101).
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Velocity widths in dwarf galaxies are of the order of a few tens of km s−1. On the other hand, in the case of spirals, the kinematics are dominated by rotation, and the velocity spreads of H i emission profiles depend crucially on the inclination of the system to the line of sight. Face- on spirals tend to have velocity widths similar to those of dwarfs, ∆V ∼ 20 − 30 km s−1; however, disks with large inclinations can have emission spread over a few hundred km s−1. Even in the latter cases, however, velocity crowd- ing in the tangent points results in a characteristic ‘double- horned’ profile, where each horn is ∼ 30 − 40 km s−1 wide, with a wide plateau in between. In the intermediate case, of inclination angles of the order of 45◦, velocity widths of ∼ 70 − 100 km s−1 are common. We will hence use two velocity widths, ∆V = 30 km s−1 and ∆V = 70 km s−1, to obtain limits on the H i mass of the absorber.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
The 30 km s−1 resolution spectrum, with angular res- olution ∼ 40′′ (i.e. without antenna 6), yields a 3σ up- per limit of 1.44 mJy on the line flux. If we assume the emission profile has ∆V ∼ 30 km s−1, equation (2) gives MH i(3σ) < 1.6 × 109M⊙. The 70 km s−1 resolution spec- trum has a 3σ upper limit of 0.87 mJy on the emission; for ∆V = 70 km s−1, this gives MH i(3σ) < 2.25 × 109M⊙. Both the above estimates are smaller than the H i mass of the Milky Way, for which MH i ≈ 5 × 109 M⊙. Fi- nally, the 200 km s−1 resolution spectrum also rules out the possibility that the emission is spread over a wide velocity range; this spectrum yields an upper limit of MH i(3σ) < 3.3 × 109 M⊙, again smaller than the H i mass of the Milky Way.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Petitjean et al. (1996) identified the absorber as an L ∼ 0.45L⋆ spiral galaxy (impact parameter ∼ 7 kpc), from its colours in ground based images; they also estimated an inclination angle of 51◦, assuming a disk morphology. The spatial resolution of the imaging was, however, not high enough to confirm the morphology. Our observations, on the other hand, show that the H i content of the absorber is less than that of normal spirals like the Milky Way, and seem to rule out the possibility that it is a large, gas-rich spiral galaxy. It is, of course, possible that the system is an early-type, gas-poor spiral. High resolution HST imaging of the absorber is necessary to resolve this issue.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
of Chengalur & Kanekar (2000) that the high Ts damped absorbers are likely to be low-mass galaxies. Acknowledgements It is a pleasure to thank Bob Sault for help in installing MIRIAD as well as with analysis pro- cedures. NK thanks the ATNF for hospitality during and after the observations. The Australia Telescope is funded by the Commonwealth of Australia for operation as a Na- tional Facility managed by CSIRO. This work was partly funded by a bezoekersbeurs from NWO to JNC. JNC is grateful for the hospitality offered by NFRA for the period during which part of this work was done.
Processed_ATCA_search_for_21_cm_emission_from_a_candidate_da.txt
Turnshek D. A., Rao S. M., Lane W., Monier E., Nestor D., 2000. To appear in: Hibbard J. E., Rupen M. P., van Gorkom J. H. (eds.), Proceedings of the “Gas and Galaxy Evolution” conference (astro-ph:0009096).
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
The role of the charm quark in the dynamics underlying the ∆I = 1/2 rule for kaon decays can be understood by studying the dependence of kaon decay amplitudes on the charm quark mass using an effective ∆S = 1 weak Hamiltonian in which the charm is kept as an active degree of freedom. Overlap fermions are employed in order to avoid renormalization problems, as well as to allow access to the deep chiral regime. Quenched results in the GIM limit have shown that a significant part of the enhancement is purely due to low-energy QCD effects; variance reduction techniques based on low-mode averaging were instrumental in determining the relevant weak effective low- energy couplings in this case. Moving away from the GIM limit requires the computation of diagrams containing closed quark loops. We report on our progress to employ a combination of low-mode averaging and stochastic volume sources in order to control these contributions. Results showing a significant improvement in the statistical signal are presented.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
c(cid:13) Copyright owned by the author(s) under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike Licence.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
where Hw is the ∆S = 1 effective weak Hamiltonian and δI the ππ-scattering phase shift. In exper- iments it is observed that the kaon decay amplitude into two pions with total isospin I = 0 is about twenty times larger than the amplitude into a state with I = 2, i.e.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
This enhancement, referred to as ∆I = 1/2 rule, remains one of the long-standing problems in hadron physics. Within the Standard Model, short-distance QCD and electroweak effects yield only a moderate enhancement. Therefore, the main contribution is expected to come from long- distance, i.e. non-perturbative, QCD effects or, if this is not the case, from new physics. Lattice QCD is the only known technique that allows to attack the problem from first principles, and pos- sibly to reveal the origin of the ∆I = 1/2 rule. In the low-energy regime of QCD various sources for the enhancement are possible. These include pionic final state interactions at around 100 MeV; physics at an intrinsic QCD scale of ΛQCD ≈ 250 MeV; or physics at the scale of the charm quark, i.e. around 1.3 GeV. It remains unclear whether the experimental observation is the result of an accumulation of several effects, or mainly due to a single cause or mechanism. A theoretically well-defined strategy to disentangle non-perturbative QCD contributions from var- ious sources was proposed in Ref. [1], with the specific aim to reveal the role of the charm quark in the explanation of the ∆I = 1/2 rule. The possibility that the enhancement is mainly due to its mass being decoupled from the light quark mass scale was pointed out a long time ago [2].
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
g− 1 g+ 1 Here, the LECs g± 1 are the couplings multiplying the counterparts of the four-quark operators Q± 1 in the effective Hamiltonian of the low-energy theory [3]. They can be determined by computing suitable correlation functions of Q± 2 in LQCD and matching them to the corresponding expressions in ChPT. The matching can be performed either in the standard p-regime of ChPT, or in the ε-regime. The advantage of the latter is that no new LECs appear at next-to-leading order, which a priori may allow for a better control of systematic uncertainties. The complicated renormalization and mixing patterns of four-fermion operators usually encoun- tered in lattice formulations can be avoided through the use of the Neuberger-Dirac (overlap) oper- ator [4], i.e.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
This is not large enough to explain the experimental ratio but is already significant and cannot be attributed to penguin diagrams.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
The challenge consists in the closed quark loop Su/c(z, z) with either a charm quark (c) or an up quark (u) running through it. By means of conventional techniques of computing quark propagators involving point sources, it is not possible to sample over the coordinate z and noise dominates over the statistical signal. More sophisticated “all-to-all” propagators have to be developed to compute these contractions.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
An essential step towards reducing the intrinsic stochastic noise is the application of “dilution”. In this work dilution is applied in spin, color and time, i.e. each of the Nr noise vectors of the ensemble has random entries only on one spin-color component of a single timeslice with all other entries set to zero.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
where C2 is the two-point function of the left-handed current J0 = ( ¯Ψγ0P− ˜Ψ). The overall improvement of LMA+SVS compared to LMA is illustrated in Figure 2. It reveals that LMA+SVS is effective for the terms consisting of 2 or 3 low-mode propagators. The variance is reduced significantly, most notably when the charm quark in the closed loop is heavy. In the latter case the absolute error of the sum of the 5 terms is roughly halved. For the term with a single low- mode propagator the technique of LMA+SVS shows no significant improvement; presumably in this case the use of multiple independent stochastic estimates for several high-mode parts increases the intrinsic stochastic noise and the technique deteriorates its performance.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
We have reported on the progress of our ongoing project to understand the role of the charm quark and its associated mass scale in non-leptonic decays of kaons into two pions. When the charm quark mass is decoupled from the light quark masses, it is hard to obtain statistical signals for “eye”-diagrams. A combination of low-mode averaging and stochastic volume sources is applied to cure this. We observe a significant variance reduction for several contributions, even though the overall error remains sizable. In the next step of this project the results for the bare ratios will be renormalized. To this purpose the contributions of the operators Q± 2 have to be taken into account.
Processed_Variance_reduction_techniques_for_a_quantitative_u.txt
Figure 2: Individual contributions to the ratios R± classified by their number of low-mode propagators for a light (top) and heavy (bottom) charm quark. Blue data points (displaced to the right for better visibility) result from LMA only, whereas LMA + SVS is used for the red data points (the stochastic estimate is only considered for a diagram if its noise is reduced, otherwise the LMA data is kept). The contribution with 0 low-mode propagators, i.e. Chhhh, is not computed stochastically and Call refers to the sum of all 5 terms.
Processed_States_of_Toeplitz-Cuntz_algebras.txt
Abstract. We characterize the state space of a Toeplitz-Cuntz algebra T On in terms of positive operator matrices Ω on Fock space which satisfy sl Ω ≤ Ω, where sl Ω is the operator matrix obtained from Ω by taking the trace in the last variable. Essential states correspond to those matrices Ω which are slice-invariant. As an application we show that a pure essential product state of the fixed-point algebra for the action of the gauge group has precisely a circle of pure extensions to T On.
Processed_States_of_Toeplitz-Cuntz_algebras.txt
and every endomorphism is of this form for some n and π; see [2, 13, 7].
Processed_States_of_Toeplitz-Cuntz_algebras.txt
Arveson has generalized these ideas to the continuous case through the use of product systems. Every representation φ of a continuous product system E on Hilbert space gives rise to a semigroup α = { αt : t > 0 } of endomorphisms of B(H), and φ is said to be essential if each αt is unital; such semigroups are called E0-semigroups, and are the primary objects of study in Arveson’s series [2, 3, 4, 5].