Filename
stringlengths 22
64
| Paragraph
stringlengths 8
5.57k
|
---|---|
Processed_The_driving_mode_of_shock-driven_turbulence.txt | these cases is highly compressive, which is what one might expect for shock-driven turbulence. We also find that 𝑏 is nearly independent of the size of the structures in the medium, which is desirable, as 𝑏 is supposed to be a measure of the turbulence driving mechanism (i.e., here the laser-driven shock), and therefore, 𝑏 should be independent of the medium that is driven. However, there is a weak tendency for 𝑏 to increase with decreasing foam void size, although 𝑏 overlaps on a 1-sigma level between small and medium, and medium and large, respectively. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | is effectively isothermal (Γ = 1), and use the simpler Eq. (1) to evaluate 𝑏. Making this simple (incorrect) assumption, we would find 𝑏 = 0.72, 0.63, 0.67, for the 12.5 𝜇m, 50 𝜇m, and 100 𝜇m void cases, respectively. Thus, the driving parameters would still be significantly larger than the natural mix (𝑏 ∼ 0.4), and would therefore still indicate strongly compressive driving of the turbulence. However, use of Eq. (1) is strongly discouraged in media with a polytropic index Γ ≠ 1, and Eq. (10) should be used instead, which indeed results in 𝑏 ∼ 1 for the shock-driven turbulence simulations analysed here. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Our numerical experiments suggest that shock-driven turbulence, such as supernova explosions or expanding radiation fronts driven by massive stars, may be considered compressive drivers, akin to the type of laser-induced, shock-driven turbulence in the numerical experiments presented here. Thus, we now want to place our simula- tion results into the bigger context of astrophysical observations, and compare them with observations in different environments, and with numerical studies related to astrophysical turbulence. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Fig. 11 shows the turbulent density fluctuations, 𝜎𝜌/(cid:104)𝜌(cid:105), as a func- tion of sonic Mach number3, in a wide range of turbulent environ- ments, from observations in the Warm Neutral Medium (WNM) over cold molecular clouds in the Milky Way, to the Large Magel- lanic Cloud, as well as previous idealised simulations of isothermal turbulence with different, controlled turbulence driving. The markers with black error-bars correspond to results from various astrophysi- cal observations, whereas the coloured markers are data points from several numerical simulations. To guide the eye, we show lines of constant 𝑏 for the isothermal case (Γ = 1), appropriate for the isother- mal simulations shown as gold, blue and magenta data points, as well as for the polytropic case with Γ = 0.1, appropriate for our shock- driven simulations presented above. We also show a shaded region, where 𝑏 = 1 and Γ = [0.0, 0.3], which is the range of Γ found in our study of shock-driven turbulence. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | The data points of the three different void size cases are represented by circles of different colours, with the 12.5, 50, and 100 𝜇m void cases shown in green, red, and blue, respectively. The red crosses are the data points for the 50 𝜇m void size case, but for different choices of the analysis volume (see Appendix F). The triangles, corresponding to each void size case (shown using the corresponding colour), represent the data points for different analysis times (see Appendix E). As expected from the discussion in Sec. 5, all our data points lie very close to or inside the region with 𝑏Γ = 1, indicating strongly compressive driving, irrespective of the size of the voids in the foam or the choice of analysis region or time. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Miville-Deschênes 2021), where the Mach numbers are of the order of unity or slightly subsonic. Depending on the appropriate value of Γ for the WNM, 𝑏 is always greater than 0.5, suggesting that the turbulence driving in the WNM is predominantly compressive in nature. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | It is worth noting that although some of the physical parameters, such as the absolute temperature, velocity, size and time scales in our numerical and laboratory experiments are starkly different from those observed in the ISM, the dimensionless parameters of the turbulence – that is, the 3D Mach number (M), the relative density fluctuations (𝜎𝜌/ (cid:104)𝜌(cid:105)), and the turbulence driving parameter (𝑏) – can be directly compared. In summary, we find that shocks are generally compressive drivers of turbulence (𝑏 ∼ 1). Since many drivers of turbulence in the ISM fall into the category of shock drivers (SN explosions, stellar winds and ionisation, gravitational collapse, spiral-arm compression, bow shocks from jets, etc.), we may expect compressive drivers of turbulence in many astrophysical environments in the ISM, with im- portant implications for the structure and star formation activity of interstellar clouds. It should be noted, however, that our simulations are highly simplified, with voids of constant radius in each simulation set. The ISM has a much more complex density structure. However, our finding that the value of the driving parameter is relatively insen- sitive to the size of the structures in the medium, lends support to the robustness of the main result of this study, namely that the driving mode of shock-driven turbulence is compressive, with a 𝑏-parameter of ∼ 1. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | We thank Kumar Raman, Mario Manuel, and Nino Landen for sup- porting this project. We acknowledge the NIF Discovery Science Pro- gram for allocating upcoming facility time on the NIF Laser to test aspects of the models and simulations discussed in this paper. C. F. ac- knowledges funding provided by the Australian Research Council (Future Fellowship FT180100495), and the Australia-Germany Joint Research Cooperation Scheme (UA-DAAD). The work of S. D. was supported in part by the LLNL-LDRD Program under Project No. 20- ERD-058. Work by LLNL was performed under the auspices of U.S. DOE under Award No. DE-AC52-07NA27344. We further ac- knowledge high-performance computing resources provided by the Australian National Computational Infrastructure (grant ek9) in the framework of the National Computational Merit Allocation Scheme and the ANU Merit Allocation Scheme, and by the Leibniz Rechen- zentrum and the Gauss Centre for Supercomputing (grant pr32lo). The simulation software FLASH was in part developed by the DOE- supported Flash Center for Computational Science at the University of Chicago. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | The simulation data in this article will be shared upon reasonable request to Christoph Federrath ([email protected]). |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | The 3D simulations were initialized with a density, velocity and pres- sure profile obtained with a 1D Lagrangian code, specifically devel- oped to prepare laser experiments by running radiative hydrodynam- ics multi-material simulations (Benuzzi-Mounaix et al. 2001). This CHARM code features a fully implicit Lagrangian solver with multi- group radiative transfer, different electron and ion temperature, laser energy deposition, thermal electronic conduction, ionic viscosity and SESAME equation of state for multiple materials. Opacities for each separate material are computed using a simple average atom model. This code has proven accurate enough to model laser experiments with direct and indirect drives on the Phebus laser (Benuzzi-Mounaix et al. 2001) and on the Omega laser (Kane et al. 2001). |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Figure A1. Space-time diagram of the material density. The colour map is logarithmically spaced between density of 0.01 g/cm3 and 10 g/cm3. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Figure A2. Density, velocity, and pressure profile of the 1D shock generated with the CHARM code, which we use to initialise the shock in the plastic ablator in our 3D simulations (c.f., Fig. 1). |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | For the purposes of initialising the shock in the 3D simulations, we take the 1D shock profile from here. Figure A2 shows the shock profile after the laser drive has been switched off, but before the shock has reached the Aluminum foil, which happens at 𝑡 = 5 ns in Fig. A1. We use this density, velocity, and pressure profile to initialise the shock inside the plastic ablator for the 3D simulations presented in the main part of the study (see Fig. 1). |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Fig. B1 shows the time evolution of the density structures as the shock passes through the foam, for all three simulations with different foam void sizes. We show 𝑥𝑦-slices of the density at three different time instances after the initiation of the shock. The bottom panels show the time at which the analysis is carried out (𝑡 = 75 ns), with the white rectangles representing the position of the analysis regions. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Figure B1. Same as the top panels of Fig. 2, but here we compare the time evolution of all three simulations with foam void sizes of 12.5 𝜇m, 50 𝜇m, and 100 𝜇m (from left to right), at times 𝑡1 = 25 ns (top panels), 𝑡2 = 50 ns (middle panels) and 𝑡3 = 75 ns (bottom panel). The turbulence analysis is carried out at 𝑡 = 75 ns in all three cases. For the analysis we selected a sufficiently large region of well-developed turbulence, indicated by the white rectangles. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Three constraints are applied to the fit: the normalisation condition and the continuity of the function (𝑝(𝑐𝑠/(cid:104)𝑐𝑠(cid:105)) and its derivative (𝑑𝑝(𝑐𝑠/(cid:104)𝑐𝑠(cid:105))/𝑑 (𝑐𝑠/(cid:104)𝑐𝑠(cid:105))) at the point of transition from the exponential to the Gaussian part, 𝑐𝑠𝑔 . Because of the constraints, there are only three free parameters, which are chosen to be the slope of the exponential part in the logarithmic PDF (𝜈), the width of the Gaussian (𝜎𝑐𝑠 ) and the transition point (𝑐𝑠𝑡 ). The mean of the Gaussian (𝑐𝑠0) and the normalisation constants, 𝐴 and 𝐵 can be derived from these parameters using the constraint conditions. As seen from Fig. 5, the exponential + Gaussian is a reasonably good fit for the sound speed PDF, with deviations only at very low sound speeds. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Table C1 presents the parameters of the density, sound speed and Mach number fits for all the three cases considered. Monte Carlo sampling was used to determine the errors in the fit parameters, where ∼ 500 PDFs were fitted, each with ∼ 10000 randomly sampled points from the analysis region. It is observed that the slopes of the over- density tail and the low sound speed tail decrease with increase in the void size, making the power law and exponential more dominant. The width of the LN part in the density PDF and Gaussian in the sound speed PDF increase with increase in the void size, though the difference between the 12.5 𝜇m and 50 𝜇m cases is not much. The values of M𝑥 and M𝑧 are almost identical for each of the three cases, which is a result of the 𝑥𝑧 symmetry. The 3D Mach numbers obtained from the data and from the Gaussian fits are in good agreement. Thus, non-Gaussian features of the Mach number PDFs can be neglected. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | As the shock moves through the foam, the foam and the walls are pushed outwards. Thus, the particles towards left move leftwards and the particles towards right move rightwards. Because of this, there are gradients in the x and z components of velocity (𝑣 𝑥 and 𝑣𝑧) along 𝑥 and 𝑦 directions respectively. In order to look to at the velocity gradients and to correct for them, we take 𝑦𝑧 slices of density for each value of 𝑥-position in the analysis volume. We calculate the mean 𝑣 𝑥 for each slice, and plot it as a function of position. The bottom left panel in Fig. D1 shows the 𝑣 𝑥 gradient which is fitted using a linear fit. It is observed that the gradient is negligible as compared to the overall turbulent 𝑣 𝑥 fluctuations in the analysis region. We subtracted the 𝑣 𝑥 gradient from the 𝑣 𝑥 data, and it was found that the Mach-number PDF almost remains unchanged. Similar results were obtained for the 𝑣𝑧 gradient along 𝑧-direction. Thus, the gradients in x and z components of the velocity have negligible effect on the turbulence properties and are neglected in the final analysis. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | the density gradient. The top right panel in Fig. D1 shows the gradient in density along 𝑦-direction, which is fitted using a linear fit. The density gradient is negligible as compared to the overall turbulent density fluctuations, and correcting for it keeps the density PDF almost unchanged. Therefore, the density gradient is also neglected in the final analysis. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Here, we investigate whether the properties of the turbulence depend on the choice of volume for the analysis region considered. We take different analysis volumes centred on the same point (𝑥 = 0.0 mm, 𝑦 = 3.47 mm, 𝑧 = 0.0 mm) for all the three void size cases. Five different cases are considered for each void size. Starting from the region for the main analysis, we first halve one side of the cuboidal region, then halve two sides, and finally halve all the sides, so that the final analysis region has one eighth of the original volume. Table E1 shows the results obtained. The polytropic Γ for the system does not change with changes in the analysis volume. It is seen that all the values of interest change only slightly with variations in the analysis volume. In fact, the variation in the driving parameter is less than 15% for all the three void size cases considered. Thus, we conclude that the turbulence properties measured remain largely independent of the choice of the analysis volume as long as the volume is not too small (due to statistical limitations) or too large (when the density and velocity gradients and/or boundary effects might become important). |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | To check whether the turbulence properties depend on the time cho- sen at which it is analysed, we consider three different time in- stances at which there is a sufficiently large region of turbulence: 50 ns, 65 ns, and 75 ns after the shock initiation. The dimensions of the analysis region are kept constant for all the three times (1.2 mm × 0.3 mm × 1.2 mm). However, its position is moved up- wards (along the positive 𝑦-direction) as time progresses, so as to select a region of well-developed turbulence at each time instance. Table F1 lists the turbulent density dispersion, Mach number, and the driving parameter calculated at each time. We find that the overall shapes of the PDFs remain largely independent of time (not shown here). Although the spread of the density PDFs, as well as the Mach number PDFs, decreases slightly as time progresses, the driving pa- rameter remains relatively unchanged. Thus, we conclude that the driving parameter in our case is largely independent of the choice of analysis time, as long as the analysis time and region are chosen such that the region contains largely fully-developed turbulent flow and is not directly affected by the shock (i.e., the shock has passed the analysis volume), and the analysis region is not affected by the boundaries of the setup. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Table C1. Fit parameters of the density PDF 𝑝 (𝑠) (Eq. 3), sound speed PDF 𝑝 (𝑐𝑠/(cid:104)𝑐𝑠 (cid:105)), and Mach number PDFs 𝑝 ( M𝑥,𝑦,𝑧 ) (Eq. 7), for the 3 foam void sizes. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Notes. The last three rows are derived quantities from the fitted parameters. For 𝜎𝜌/(cid:104)𝜌(cid:105) we integrate the numerical PDF via Eq. (5) up to 𝑠 = 1.5. This upper bound is only relevant for the 100 𝜇m case, because the PL section is very flat there, and integrating to infinity would yield unreasonably large 𝜎𝜌/(cid:104)𝜌(cid:105). |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | compute cells), half the standard resolution (384 × 512 × 384 com- pute cells), and double the standard resolution (1536 × 2048 × 1536 compute cells). Table G1 lists the quantities of interest. We see that the results are reasonably converged with the standard resolution that is used throughout this study. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | While the turbulence driving parameter is relatively insensitive to the numerical resolution, we note that mixing of material in the post-shock medium will likely depend on the numerical resolution (see e.g., Banda-Barragán et al. 2020; Banda-Barragán et al. 2021). Noting that numerical viscosity starts acting on scales of (cid:46) 30 grid cells (Kitsionas et al. 2009; Sur et al. 2010; Federrath et al. 2011), it is expected that mixing in the mid- and small-sized void cases is likely underestimated. |
Processed_The_driving_mode_of_shock-driven_turbulence.txt | Figure D1. Plots showing the gradients in the density (top panels) and the velocity (bottom panels) in the analysis region along 𝑥-direction (left panels) and 𝑦-direction (right panels) respectively. It is seen that there is a net density gradient along the direction of the shock propagation, and a gradient in the 𝑥-component of the velocity along the 𝑥-direction. These two gradients are fitted with a linear fit, represented by red dashed lines. Overall, the contribution of the gradients is negligible as compared to the turbulent density and velocity fluctuations. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | GPS and CSS radio sources are the objects of choice to investigate the evolu- tion of young radio-loud AGN. Previous investigations, mainly based on number counts and source size distributions, indicate that GPS/CSS sources decrease significantly in radio power when evolving into old, extended objects. We sug- gest this is preceded by a period of increase in radio luminosity, which lasts as long as the radio source is confined within the core-radius of its host galaxy. We have selected a sample of nearby compact radio sources, unbiased by radio spec- trum, to determine their luminosity function, size distribution, dynamical ages, and emission line properties in a complete and homogeneous way. First results indicate that the large majority of objects (>80%) exhibit classical GPS/CSS radio spectra, and show structures consistent with them being compact double or compact symmetric objects. This sample provides an ideal basis to further test and constrain possible evolution scenarios, and to investigate the relation between radio spectra and morphologies, orientation and Doppler boosting in samples of young radio-loud AGN, in an unbiased way. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | It has been shown beyond reasonable doubt that the archetypical Gigahertz Peaked Spectrum (GPS) radio galaxies are young objects, with typical dynamical ages of 102−3 years (Polatidis & Conway 2002, and references therein). These sources are therefore the prime candidates to be the young progenitors of classical large-size radio sources, which subsequently evolve into Compact Steep Spectrum (CSS), and FRI and/or FRII radio sources. Although it is not clear whether all GPS sources are young objects (certainly the GPS quasars seem to be an unrelated distinct class of objects), GPS galaxies are the key objects to study the early evolution of radio-loud AGN. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Current views on radio source evolution are mainly based on the radio power - linear size (P − D) diagram, and source size distributions. These suggest that if GPS and CSS sources evolve into large size radio sources, they should decrease in radio power by at least an order of magnitude, to account for their relatively high space density (Fanti et al. 1995; Readhead et al. 1996, O’Dea & Baum 1997). This luminosity evolution scenario is backed up by theory, ie. it is expected that a radio source with a constant jet power, expanding in a medium with a declining density profile, will decrease in radio power (eg. Baldwin 1982; Kaiser & Alexander 1997). |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | We have used the existing samples of Stanghellini et al. (1998; bright GPS galaxies), Fanti et al. 1990 (bright CSS galaxies), and Snellen et al. (1998a; faint GPS galaxies) to study early radio source evolution. This has been described in detail in Snellen et al. (2000). We noticed a distinct difference between the redshift distribution of GPS galaxies from the Stanghellini et al. sample and 3C galaxies, with the GPS galaxies biased towards higher redshifts (95% confidence in the KS test). First of all, this makes the number count analyses mentioned above, which are averaged over a large redshift range, less straightforward. Furthermore, it is important to understand how two populations, which evolve into each other on timescales several orders of magnitude less than the Hubble time, can have different redshift distributions. It is clear that to explain this redshift bias in a flux density limited sample, the luminosity function of GPS galaxies has to be different from that of old, extended objects, if the former are to develop into the latter. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | We advocate that it is the qualitatively different individual luminosity development of young radio sources compared to that of old objects which causes a flatter overall luminosity function. If radio sources increase in radio power until they grow beyond a certain size and thereafter decrease in radio power, a randomly aged population of compact radio sources will be biased towards higher luminosities than a population of large-size objects, producing a flatter luminosity function. This behaviour in luminosity evolution is expected from theory, in which the density profile of the core of the host galaxy is flat (see Fig. 1, and the density only starts to decrease outside this core radius (Snellen et al. 2000; Alexander 2000; O’Dea 2002). Some new evidence for this scenario is also presented by Perucho & Marti 2002a,b). |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Figure 1: Typical density profile of a galaxy (dashed line), and the expected luminosity evolution as function of size for a radio source (solid line), according to Snellen et al. (2000). |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | where B is the magnetic field strength. If the weak dependence on B is taken out by assuming an equipartition magnetic field, then the SSA components sizes are found to be linearly proportional to the overall size of the GPS/CSS sources as measured from VLBI, over a luminosity and linear size range of several orders of magnitude. This indicates that the ratio of overall-size to lobe-size stays constant while the sources grow, and that therefore young radio sources evolve in a self-similar way. Furthermore, it provides strong evidence that the turnovers in the radio spectra of GPS/CSS sources are indeed caused by SSA, and not free-free absorption (FFA). Further evidence in favour of this scheme is presented by Tschager et al. (2002), who find that a new sample of very faint CSS galaxies also follow the relation given above. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Although it is shown that FFA does occur in some compact GPS sources (eg. Kameno et al. 2002; Inoue et al. 2002; Vermeulen et al., 2002), it seems that even in these objects FFA only generally affects faint components close to the nucleus, and not the mini-lobes which form the dominant components in the overall radio spectrum. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Although it is possible to fit the spectra of individual sources with fine-tuned FFA models, it is not clear to us why the conditions of FFA would change in such way over a large range of radio source luminosities and linear sizes, that they exactly mimic the frequency and strength of the turnovers as expected for SSA. This makes a very strong case for SSA being the dominant absorption process in GPS and CSS galaxies. Highly accurate flux density monitoring can now be used to further test self-similar evolution by comparing the long-term flux density variability with what is expected for this scenario (eg. Tingay et al. 2002). |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Several other matters have to be settled. The link between the radio spectra and VLBI morphologies have to be further investigated. Do all GPS galaxies have compact symmetric (CSO) morphologies, such as the bright archetypes? What effect could orientation have on the classification of an object as either a GPS source or a CSO (eg. Snellen et al. 1998b)? Is Doppler boosting really not important in these objects, as is always assumed, or does it significantly affect our statistical analyses? And even more general, but not too obvious, how do you exactly decide what is and what is not a GPS source? To address many of these issues, we are studying a new sample of nearby, young radio-loud AGN. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Figure 2: The redshift distribution of the sources in the new sample of nearby young radio-loud AGN. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | will therefore be the main goal of our investigation, which will form a powerful tool to constrain evolution models. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | In addition, as a spin-off, a complete sample of nearby young radio-loud AGN allows us to undertake unique studies of their HI absorption distribution (using VLBI in L- band), X-ray, infrared and optical cluster properties. This will undoubtedly give new insights in the intra- and intergalactic environments of young radio-loud AGN. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Selection of a comprehensive nearby sample has only become possible since the con- struction of the new generation of radio surveys. Especially the 1.4-GHz FIRST survey (White et al. 1997) with its high resolution is ideal for selecting compact radio sources, and therefore forms the basis of our sample. Only the region north of declination +30◦ was used, so it overlaps with the WENSS at 325 MHz (Rengelink et al. 1997). |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | The first step was the selection of all sources in FIRST with an angular size of < 2′′ and a flux density S1.4GHz > 100 mJy. It is important to note that no selection was made on radio spectra, but only on angular size, allowing in a later stage an unbiased investigation of possible orientation and/or beaming effects on the radio spectra and morphologies of young radio-loud AGN. Unfortunately, by selecting at 1.4 GHz, sources with spectral turnovers at higher frequencies (very compact objects) may have been missed. To catch these, sources from the flat spectrum CLASS survey (with S5GHz>30 mJy; Myers et al. 1995) were added to the sample, with an 8.4-5 GHz spectral index such that their 1.4-GHz flux density would have been brighter than 100 mJy, if they were not synchrotron self absorbed. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | Figure 3: Example radio spectra and morphologies of sources in the new sample of nearby young radio-loud AGN. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | The positions of the remaining objects were cross-correlated with the optical APM/ POSS-I catalogue, and only those objects which coincided within 10′′ with a galaxy with a red magnitude <16.m5 were selected. This yielded a total of 55 objects, which were subsequently correlated with other surveys such as 4C, NVSS (Condon et al. 1998), GB6 (Gregory et al. 1996), and CLASS, and images were obtained from the NVSS and the Digitized Sky Survey (DSS, Lasker et al. 1990) to check whether the optical ID was genuine, and to see whether any large scale radio emission may have been missed (on which basis they would be rejected). The literature was searched for any available redshifts, and optical spectra were taken for the remaining objects with the Calar-Alto 2.2-m and the 2.5-m Isaac Newton telescopes. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | of the angular sizes expected for these objects (see section 1.2). This information was used to divide the sample into small-size, intermediate-size, and large-size sub-samples. These sub-samples were then observed with the VLBA+EVN at 5 GHz (∼1.5 mas resolution), the EVN at 1.6 GHz (∼15 mas resolution), and MERLIN at 1.6 GHz (∼150 mas resolution) respectively. So far, all but one source have been observed in this way. In addition, optical imaging in B, V, R, and I band filter, and HI absorption observations with WSRT are underway. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | The first results look very promising. More than 80% of the objects have radio spectra which can be classified as classical GPS or CSS sources (see Fig. 3), with the remainder also showing a spectral peak but clearly exhibiting variability. Most of the objects have morphologies consistent with them being compact double or compact symmetric objects, as expected for this class of objects, but some seem to show core-jet or more complex structures. This sample is ideal to further study the relation between radio spectrum, morphology, orientation, and Doppler boosting. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | At this stage, the foundation has been laid to determine directly the local luminosity function, source size distribution, and the dynamical age distribution for young radio- loud AGN, without cosmological bias. |
Processed_Constraining_the_evolution_of_young_radio-loud_AGN.txt | We wish to thank Tasso Tzioumis for the excellent organisation of the workshop, and the villagers of Kerastari for their warm hospitality. This paper is partly based on observations collected at the German-Spanish Astronomical Center, Calar Alto, Spain, operated by the Max-Planck-Institute f¨ur Astronomie, Heidelberg, jointly with the Spanish National Commission for Astronomy; the Isaac Newton Telescope (through the service programme), which is operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Insti- tuto de Astrofisica de Canarias; The European VLBI Network (EVN, making use of EC’s Access to Research Infrastructures Programme, under Contract No. HPRI-CT- 1999-00045), which is a joint facility of European, Chinese and other radio astronomy institutes funded by their national research councils; MERLIN, a National Facility operated by the University of Manchester at Jodrell Bank Observatory on behalf of PPARC. The National Radio Astronomy Observatory is a facility of the National Sci- ence Foundation operated under cooperative agreement by Associated Universities, Inc. KHM was supported by a Marie-Curie fellowship of the European Commission. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | ABSTRACT The Australia Telescope Compact Array and the Very Large Array are currently being upgraded to operate with wide bandwidths; interferometers dedicated to the measurement of cosmic microwave background anisotropies are being designed with large instantaneous bandwidths for high sensitivity. Interferometers with wide instan- taneous bandwidths that do not operate with correlators capable of decomposing the bands into narrow channels suffer from ‘bandwidth smearing’ effects in wide- field imaging. The formalism of mosaic imaging is extended here to interferometers with finite bandwidths to examine the consequences for the imaging of wide fields if very wide instantaneous bandwidths are used. The formalism presented here provides an understanding of the signal processing associated with wide-band interferometers: mosaicing may be viewed as decomposing visibilities over wide observing bands and, thereby, avoiding ‘bandwidth smearing’ effects in wide-field imaging. In particular, the formalism has implications for interferometer measurements of the angular power spectrum of cosmic microwave background anisotropies: mosaic-mode observing with wide-band radio interferometers decompose wide-band data and synthesize narrow filters in multipole space. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | When observing continuum radio sources, the sensitivity of a radio interferometer depends on the square root of the bandwidth, apart from other factors. For this reason, arrays are usually designed to have the maximum continuum bandwidths permitted by the state of the art and, additionally, the instantaneous observing bandwidths of existing interferometer arrays are often upgraded over time as the technology associated with sampling rates, transmission bandwidths and correlator speeds improve. Interferometer arrays that Fourier synthesize images of relatively bright celestial sources are often dynamic range limited owing to artefacts resulting from calibration errors and other systematics; it is in the imaging of the relatively weak sources that thermal noise proves to be the limiting factor. The radio interferometer imaging of cosmic microwave background (CMB) anisotropies, in total intensity and in polarization, is an example of an astrophysical problem where the thermal noise limitation, arising partly from the limited bandwidths in radio interferometers, provides a strong argument in favour of using bolometers – which are intrinsically wide bandwidth devices – for the detection instead of radio interferometers. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Interferometer bandwidths are limited by the state of the technology. If the visibility correlations are to be measured by digital correlators, the wide bandwidth analog signals from the antenna elements would have to be sampled at appropriately high speeds. Interferometers in use today that have the widest bandwidths do not digitise the antenna signals; their correlators are analog multipliers. And an interesting proposal for a dedicated CMB interferometer uses bolometer detectors to measure the interferometer fringe (Ali et al. 2002). |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | When the celestial source that is being imaged has an intensity distribution over a wide field, or sources distributed over a wide field are being imaged together, wide bandwidth interferometers have special problems and the visibilities suffer from an effect known as ‘bandwidth smearing’ (see, for example, Cotton (1999) and Bridle (1999)). An alternate description of this effect is in Section 2 for the case where a simple two-element interferometer observes sources distributed over a wide field. In Section 3, the formalism of wide-field mosaicing is extended to a wide bandwidth interferometer. The analysis shows that just as mosaic imaging techniques may be adopted for overcoming the limitations posed by the size of the antenna elements, mosaic imaging may also be adopted for overcoming the limitations posed by the size of the bandwidth. Later sections discuss the use of wide bandwidth interferometers in wide field surveys of the sky and measuring the angular power spectrum of CMB temperature anisotropies. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | only avoid the problem but also improve the imaging fidelity. However, these bandwidth synthesis methods require that the interferometer correlations be measured separately in multiple frequency channels covering the observing band; in other words, a spectral-line correlator is required for the continuum imaging. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | where ~s is the baseline vector: ~s = ~b/λ, and ~r is a unit vector towards the sky solid angle element dΩ (Clark 1999). It is assumed here that the radiation is spatially incoherent across celestial sources and that the sources are extremely distant as compared to the spacing b. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | If the effective apertures of the antennas cover finite areas on the ground, the interferometer response will be an integral of the spatial coherence function over the spatial frequencies sampled by the baseline vectors between elements of one antenna aperture and elements of the other aperture. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | To take a specific example, assume that the antennas forming the interferometer have identical circular apertures, uni- formly illuminated, with diameter d and with a spacing b between their centres. The antenna apertures are assumed to be in the same plane as the interferometer baseline vector and, therefore, the antenna pointing coincides with the interferometer phase centre. The antenna configuration on the ground (real space) is shown in panel a of Fig. 1. The interferometer response (measured visibility) is an integral of the spatial coherence function over one of two circular regions, with radius d/λ, of the spatial frequency domain; these circular regions are located symmetrically in the 2-D spatial frequency space with their centres b/λ from the origin. These sampling regions are shown in panel b of Fig. 1. The values of the coherence function at points in the spatial frequency plane that are inversion symmetric about the origin are complex conjugates of each other and the implementation of the complex correlator determines the choice of the sampling region. The spatial frequencies are sampled over the range (b d)/λ to (b + d)/λ with a weighting that linearly decreases from the centres of the circular sampled regions to zero at the edges. For the special case where the antenna apertures are adjacent and b = d, the coherence function is sampled over the range 0–2d/λ. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Interferometers with finite antenna apertures provide visibilities that are integrals over spatial frequency space: the weighted averaging effectively results in a loss of information on the detailed variations of the coherence function over distances in spatial frequency that are smaller than 2d/λ. The weighting function in spatial frequency space has a Fourier transform that is the antenna far-field radiation power pattern and, therefore, in the sky domain, the averaging (in spatial frequency space) results in that the interferometer response to brightness away from the antenna pointing centre is attenuated by the primary beam pattern of the antennas. The interferometer does not respond to brightness outside the primary beam and this is a problem for wide-field imaging. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | winds (2dξ/λ) times. If the source is offset by a distance ξ = λ/(2d) = c/(2dν), the phase of the coherence function winds through 2π radians across the sampled circle in spatial frequency domain and the response is severely attenuated. It is this averaging of a rotating coherence function vector, across spatial frequency space, that results in the loss of information on the sky brightness distribution and is usually called the ‘primary-beam attenuation’. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | In the example we have been considering of an interferometer formed between a pair of circular and uniformly illuminated apertures, the finite bandwidth results in that the circular regions sampled in spatial frequency space scale with increasing ∆ν)/c at the bottom of the frequency range to 2d(ν◦+∆ν)/c frequency. The diameter of the sampled regions scales from 2d(ν◦ at the highest observing frequency. The centre of the circular region also shifts outwards from b(ν◦ ∆ν)/c to b(ν◦ + ∆ν)/c. The movement of the circular sampling region in spatial frequency space, from one end of the observing band to the other, is depicted in panel c of Fig. 1. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | where ~s◦ is the baseline vector between the centres of the apertures, in wavelengths, at the centre frequency ν◦. Ws(~s, ν) is the auto-correlation of the aperture illumination at frequency ν and is, in general, a frequency dependent sampling function in spatial frequency space. Wν(ν) is the weighting function corresponding to the bandpass shape. The weighting functions are assumed to be normalised so that their integrals are unity. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Consider the coherence function at a baseline ~b◦ owing to a celestial source that is offset from the interferometer phase centre and at a location defined by the vector ~ξ. Additionally, let the source offset be along the baseline; i.e., let ~s◦, which is equal to ~b◦/λ, and ~ξ be vectors with the same direction cosines in their respective planes. The coherence function at this spatial frequency will have a phase ∆ν. A source offset by c/(2b◦∆ν) will have a coherence function that winds through 2π radians across the observing band. It is the loss of information arising from this averaging across the observing band that is called ‘bandwidth smearing’. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | The number of phase turns across the observing bandwidth will exceed the number of turns across the visibility space covered by finite apertures if (∆ν/ν◦) > (d/b◦). The averaging length, in spatial frequency space, owing to the finite aperture is proportional to d/b◦ whereas that owing to the finite bandwidth depends on ∆ν/ν◦. The relative magnitudes of these two quantities decides which effect dominates in any interferometer. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Scanning the sky with an interferometer, and using the scan data to reconstruct the distribution in the coherence function across the spatial frequencies sampled by the apertures forming the interferometer, was proposed by Ekers & Rots (1979). As far as I know, Rao & Velusamy (1984) were the first to explicitely implement this observing scheme and decompose interferometer visibility data in the spatial frequency domain and then use the finely sampled visibility data to reconstruct wide field images. I extend this formalism below to the case where an interferometer mosaic observes a wide field using a very wide bandwidth. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | The δ-function above represents the Dirac δ-function. The Fourier transformation of the visibility data acquired in different pointings is seen to yield a weighted average of the coherence function. The averaging in equation (8d) is over frequency and not over spatial frequency and, therefore, this method of deriving the visibilities in spatial frequency domain is capable of avoiding the problems associated with band-width smearing related effects. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | In this case, the Fourier transformation of the visibilities that were acquired with the interferometer pointed at different offsets yields weighted samples of the coherence function; the relative weighting depends on the aperture illuminations of the antennas forming the interferometer pair and the bandpass shape. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | It may be noted here that if the mosaic observations are made using an interferometer array of antennas, all with diameters d and observing over a fixed band covering a range ∆ν around a centre frequency ν◦, avoiding aliasing in the spatial frequency domain requires that equation (11) be satisfied for the longest baseline. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Explicit transformation of the visibilities V ( ~s◦, ~ξ) into visibilities V ( ~s◦, ~χ) distributed over the spatial frequency range given by equation (10) is possible using equation (9) provided that the Nyquist sampling criterion is satisfied. The discrete Fourier transformation (DFT) of the visibilities V ( ~s◦, ~ξ), which are discrete samples in offset ~ξ space, would yield samples of the visibility V ( ~s◦, ~χ) in spatial frequency. If the mosaic grid measures 2ξm/∆ξ complex visibilities along any sky dimension, the DFT would yield 2ξm/∆ξ independent measures of the complex visibility in the conjugate space which is the spatial frequency ~χ domain. Samples spaced 1/(2ξm) wavelengths apart would be independent. These visibility samples are weighted averages of the coherence function in spatial frequency domain with a point spread function (PSF) that is determined by the sampling in ~ξ space and is independent of the bandwidth; the samples V ( ~s◦, ~χ) are essentially averages of the coherence function over spatial frequency ranges that are 1/(2ξm) wavelengths wide. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | The implication here is that even if the observing bandwidth is large, and in the extreme case if (∆ν/ν◦) exceeds (d/b) so that the attenuation owing to bandwidth smearing exceeds the primary beam attenuation in the individual fields, mosaicing observations with a sampling that satisfies equation (11) can image large fields of view. The DFT to spatial frequencies would decompose the observed visibilities – that are integrals of the coherence function over wide spatial frequency ranges corresponding to the wide bandwidths used – into samples that correspond to regions that are effectively 1/(2ξm) wavelengths wide. Just as mosaicing has the ability to decompose the observations made with wide-aperture antennas into visibilities corresponding to those obtained with small aperture arrays, mosaic mode observing has the ability to decompose the visibilities made with wide bandwidths into those corresponding to narrow bands. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Samples of these visibility estimates, which are separated by 1/(2ξm) wavelengths, are independent and may be used to reconstruct images of the sky. These images are that of the true sky intensity distribution convolved by a PSF that is the Fourier transform of the weights Wsν(~χ). The image would be tapered by the DFT of Ξ′(~χ), which is simply the top-hat function Ξ(~ξ) that encompasses the entire sky area covered by the mosaic pattern. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Assuming that the sky temperature anisotropies in the CMB are Gaussian random fluctuations with random phase, they may be completely described by the Cl coefficients of spherical harmonic decompositions. Observations of CMB anisotropies attempt to measure the angular power spectrum, which is the distribution of l(l + 1)Cl/(2π) over multipole l space. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | frequency s◦ corresponds to a multipole mode l◦ = 2πs◦ and the variance in CMB visibilities that are measured over a band ∆s is a measure of the CMB anisotropy over a multipole range ∆l = 2π∆s. Interferometer measurements of CMB anisotropy hence provide estimates of Cl power in which the telescope filter functions are defined by the aperture illuminations of the antennas and the projected baseline length. If the interferometer mosaic images a wide field, these visibility data at multiple pointings might be used to derive visibilities corresponding to narrow filters in spatial frequency space. Therefore, mosaic mode observations are a method for decomposing the interferometer Cl measurements into those corresponding to narrow telescope filter functions in l space: mosaic observing improves l-space resolution (White et al. 1999; Subrahmanyan 2002). Moreover, it has been argued that drift scanning the sky, using a co-mounted antenna array operated as interferometers, is a useful technique for rejecting systematics (Subrahmanyan 2002); drift scanning is an implementation of the mosaicing technique. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Consider the case where a two-element interferometer, consisting of apertures of diameter d and baseline b, observes an n n mosaic of pointings for a total time t (the time spent at each pointing is t/n2). Assume first that the observations are made (d/nb). Assume further that the pointings are separated by the required Nyquist rate using a narrow bandwidth (∆ν/ν◦) of (c/2dν◦). The signal to noise ratio (SNR) in the measurement of the flux density of a point source that appears in any one of the pointings will be proportional to d2√t/n. If, instead, this interferometer that is formed of apertures of diameter d observes the source position for the entire time t, the SNR would be proportional to d2√t. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | The analysis suggests that when interferometers with wide instantaneous bandwidths are used for the mosaic imaging of wide fields, including the particular case of mosaic-mode CMB observations, the optimum bandwidth corresponds to the case where (∆ν/ν◦) = (d/b); the SNR degrades if the instantaneous bandwidth used exceeds (2d/b)ν◦. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Interferometers with wide fractional bandwidths are of particular interest at high frequencies where errors due to telescope pointing and interferometer phase may be significant. therefore, the limitations arising from such errors are considered below. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | where θν◦ is the FWHM of the primary beam at frequency ν◦, NA is the number of antenna elements in the array and ǫp is the systematic pointing error associated with any antenna; it is assumed that the errors are uncorrelated between antennas. Mosaic scanning is also limited by time varying phase errors that introduce a varying phase error over the sky scans. If ǫφ is the R.M.S. phase error (in radians) during the scanning then the visibility amplitudes following the decomposition in spatial frequency space would have fractional errors of order ǫφ. In the case of interferometer mosaic imaging, assuming that these errors are antenna based (as would be expected for atmospheric phase errors), the image plane fidelity would be approximately limited by these errors to NA/ǫφ. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Wide band interferometers – in which the band is not finely sub-divided in multi-channel receivers – instantaneously sample a wide domain in spatial frequency space. Mosaic imaging of wide fields, which are made by covering the wide fields with a grid of pointings and subsequently transforming the visibilities that are measured in the multiple pointings to visibility space, yields samples of the visibilities distributed in the spatial frequency domain. The formalism of mosaic imaging of wide fields has been extended here to the case where the interferometers operate with wide instantaneous bandwidths. The mosaicing technique may be viewed as effectively decomposing wide-band data into spatial frequency bins. The formalism presented here develops the understanding of the mosaicing technique that reconstructs wide-field images without ‘bandwidth smearing’ effects. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | The signal-to-noise ratio in a mosaicing interferometer improves as the bandwidth is increased; however, the usefulness of the wide band for wide-field imaging diminishes once the bandwidth exceeds (2d/b)ν◦. Admittedly, a multi-channel receiver improves upon the sensitivity of the mosaicing interferometer; however, the gain is marginal if the total band is less than (2d/b)ν◦. The considerations discussed here are relevant both to the case where the mosaicing interferometer attempts to reconstruct sky images and to the case where the mosaicing interferometer attempts a measurement in spatial frequency space of the angular power spectrum of the sky brightness fluctuations. |
Processed_Radio_Interferometers_with_wide_bandwidths.txt | Figure 1. Panel a shows the configuration and dimensions of the aperture elements forming the interferometer in real space. Panel b shows the visibility sampling in spatial frequency space. Panel c shows the change in the sampling of the spatial frequency domain across the observing band. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | highest redshift bin. As a result, lensing magnification cannot be neglected in the spectroscopic survey, especially if we want to determine the growth factor, one of the most promising ways to test general relativity with Euclid. We also find that, by including lensing magnification with a simple template, this shift can be almost entirely eliminated with minimal computational overhead. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | we show that this is enough to de-bias the analysis, reducing the shifts to less than 0.1σ. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | The European Space Agency’s Euclid satellite mission (Laureijs et al., 2011, Amendola et al., 2018) aims at shedding new light on the so-called dark components of the Universe, namely dark matter and dark energy. Dark matter, a mysterious form of mat- ter that does not seem to emit light, yet it accounts for more than 80% of the total matter content of the Universe, forms the bulk of the large-scale cosmic structure, upon which galaxies form and evolve. The latter is even more elusive, and it is what drives the current accelerated expansion of the Universe, contributing to about 70% of the total cosmic energy budget (see, e.g., Bull et al., 2016, for a review of the current concordance cosmolog- ical model and the main theoretical challenges it faces). In fact, there is another possibility to explain the effects we ascribe to (either or both) dark matter and dark energy: that the theory we use to analyse the data is incorrect. This approach goes under the name of ‘modified gravity’ (e.g., Clifton et al., 2012). Thanks to the extent and exquisite precision of Euclid’s data, we shall soon be able to further test general relativity on scales far from the strong-gravity regime where it has been tested to supreme preci- sion (see, e.g., Cardoso & Pani, 2019, for a review of the current status). |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | Euclid will consist of two primary probes: a catalogue of about 30 million galaxies with spectroscopic redshift informa- tion, spanning a redshift range between z = 0.8 and z = 1.8, and a catalogue of 1.5 billion galaxy images with photometric redshifts down to z = 2 (see Laureijs et al., 2011, Amendola et al., 2018, for further details on the specifics of Euclid surveys). One of the main goals of the spectroscopic survey of Euclid is to measure the so-called growth rate which is very sensitive to the theory of gravity, see for example Alam et al. (2017). How- ever, in order to robustly test alternatives to general relativity, it is crucial to take into account in the analysis all of the relevant effects. One effect that has been overlooked in previous forecasts regarding the performance of the Euclid spectroscopic survey is lensing magnification (Matsubara, 2004). |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | The paper is structured as follows. In the next section, we present the fluctuations of galaxy number counts within linear perturbation theory, concentrating on the redshift-space 2-point correlation function. In Sect. 3 we present the relevant quanti- ties from the Euclid Flagship simulations used in this work. In Sect. 4 we explain the methods used in our analysis, which are the Fisher matrix and Markov chain Monte Carlo (MCMC). In Sect. 5 we discuss our results and present the method to de-bias the analysis, and in Sect. 6 we conclude. Some details and com- plementary results are delegated to several appendices. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | where we use the Einstein summation convention over repeated indices. Here a(η) is the cosmic scale factor evaluated at con- formal time η, and c is the speed of light. In the above, as well as subsequent equations, a prime denotes the derivative with re- spect to conformal time, and H = a′/a = Ha denotes the confor- mal Hubble parameter. We normalize the scale factor to 1 today, a0 = 1 such that H0 = H0. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | reasons that, for spectroscopic surveys, the correlation function is a more promising summary statistics than the angular power spectrum and we therefore need to determine the impact of lens- ing magnification on this statistic. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | Here r = r(z) is the comoving distance evaluated at redshift z, rs is the comoving distance between the observer and the source, b = b(z, m∗) is the linear bias of galaxies with magnitude below m∗, the magnitude limit of the survey, δ is the gauge-invariant density fluctuation representing the density in comoving gauge, ∂r is the derivative w.r.t. the comoving distance r, V is the ve- locity of sources in the longitudinal gauge, and ∆Ω denotes the angular part of the Laplacian. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | In spectroscopic surveys, there are two standard estimators used to extract information from galaxy number counts: the 2-point correlation function (2PCF), and its Fourier transform, the power spectrum. In this paper, we concentrate on the correlation func- tion, since lensing magnification can be included in this estima- tor in a straightforward way. This is not the case for the power spectrum, which requires non-trivial extensions to account for magnification (see, e.g., Grimm et al., 2020, Castorina & di Dio, 2022). |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | The 2-point correlation function can be calculated in the curved-sky, i.e., without assuming that the two directions n and n′ are parallel ξ(n, z, n′, z′) = ⟨∆(n, z)∆(n′, z′)⟩ . |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | Here ¯N is the cumulative number of objects brighter than the magnitude cut m∗ (for a magnitude-limited sample). The first line of Eq. (4) corresponds to the standard terms of density fluc- tuations and redshift space distortions while the second line is the lensing magnification which is the subject of the present pa- per, see Bonvin & Durrer (2011), Challinor & Lewis (2011), Jeong et al. (2012), Di Dio et al. (2013) for details. 1 The dots at the end of Eq. (4) stand for the “large-scale relativistic terms” which we omit in our analysis. These terms are suppressed by factors λH c−1, where λ is the comoving wavelength of the per- turbations, see, e.g., Di Dio et al. (2013), Jelic-Cizmek et al. (2021), Euclid Collaboration: Lepori et al. (2022). It is well known that the large-scale relativistic terms are relevant only at very large scales and do not significantly impact the even mul- tipoles of the correlation function on sub-Hubble scales (Lorenz et al., 2018, Yoo et al., 2009, Bonvin & Durrer, 2011). Some of these relativistic terms will, however, be detectable by measur- ing odd multipoles in the correlation of two different tracers (see Bonvin et al., 2014, Gaztanaga et al., 2017, Lepori et al., 2020, Beutler & Di Dio, 2020, Saga et al., 2022, Bonvin et al., 2023). A detailed study of the signal-to-noise ratio of all relativistic effects in simulated mock catalogues adapted to Euclid’s spectroscopic survey will be presented in Euclid collaboration: Elkhashab et al. (in preparation). |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | where jℓ denotes the spherical Bessel function of order ℓ, and Pδδ(k, ¯z) is the linear matter power spectrum in the comoving gauge. We see that, in the flat-sky approximation, density and RSD are fully encoded in the first three even multipoles. The ex- pressions for computing higher-order multipoles without using the flat-sky approximation can be found in Appendix E.1. |
Processed_Euclid_Preparation._TBD._Impact_of_magnification_o.txt | Here H0 = 100 h km s−1 Mpc−1, k⊥ = k − kµn is the projec- tion of the Fourier space wavevector k to the direction normal to n, and J0 denotes the Bessel function of zeroth order. The first two lines of Eq. (12) contain the density-magnification correla- tion: when computing the multipoles, one averages over all ori- entations of the pair of voxels. For each orientation, the galaxy which is further away is lensed by the one in the foreground. This effect clearly has a non-trivial dependence on the orienta- tion angle µ which enters in the argument of the Bessel function J0. The last two lines contain the magnification-magnification correlation, due to the fact that both galaxies are lensed by the same foreground inhomogeneities. In all the terms, the functions are evaluated at the mean redshift of the bin ¯z, ¯r denotes the comoving distance at that redshift, and a(r) denotes the scale factor evaluated at comoving distance r. For completeness, in Appendix E.2 we list the semi-analytic expressions which allow us to efficiently evaluate the flat-sky magnification terms. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | Among the 1000 H i-brightest HIPASS sources (Koribalski et al. 2004) HIPASS J0731–69 stands out as the only extragalactic source without a clear optical counterpart. HIPASS J0731–69, however, has a clear link with NGC 2442, a spiral galaxy with asymetric spi- ral arms and disturbed morphology located in the NGC 2434 group. The H i reservoir that constitutes HIPASS J0731–69 was most likely dislodged from NGC 2442 dur- ing a recent tidal encounter (Bekki et al. 2005). |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | survey-grade optical counterparts (Haynes et al. 2011; Cannon et al. 2015). Out of the 15,855 sources of the 40% ALFALFA catalog 15,041 of them are extragalactic. Recent observing campaigns targeting H i sources with- out optical counterparts in the ALFALFA survey have revealed ultra-low surface brightness structures, tidal dwarf galaxies and objects with exceptionally high mass- to-light ratio (Lee-Waddell et al. 2014; Janowiecki et al. 2015; Cannon et al. 2015). |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | Other intergalactic H i clouds without clear optical counterparts have also been discovered outside the main all sky surveys. For instance, Osterloo & van Gorkom (2005) report a large H i tail (3.4 × 108M⊙) near the cen- ter of the Virgo cluster. The origin of this cloud is likely ram pressure stripping of NGC 4388 by M 86. Another example of intergalactic H i, also in Virgo, is H i 1225+01 (Giovanelli & Haynes 1989; Chengalur et al. 1995). H i 1225+01 has two main components separated by 100 kpc with H i masses of 109M⊙. One of the two components is associated with a low surface brightness galaxy while the second component does not have an identified optical counterpart (Matsuoka et al. 2012). See also the exam- ple of the Vela cloud, a large and massive intergalactic H i cloud residing in the NGC 3256 group (English et al. 2010). |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | The Leo Triplet (NGC 3623, NGC 3627, and NGC 3628) is the classical example of a large tail of neutral hydrogen (Haynes et al. 1979) where a faint tidal tail of stars, coincidental with the H i peaks, was discovered with deep optical follow-up (Chromey et al. 1998). |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | Similarly, in this paper, we use new deep Gemini im- ages to search for the first signs of star formation within two massive H i clouds recently discovered near the spiral galaxy IC 5270. The two H i clouds that are the target of this study have neutral hydrogen masses comparable to the Milky Way’s. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | Note. — Column (1): Target cloud; Column (2) and (3): R.A. and Dec of peak H i intensity; Column (4): H i mass in solar masses; Column (5) Observing Date; Column (6): Image quality in percentile; Column (7): Cloud cover in percentile; Column (8): Total exposure time in seconds. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | The recent analysis of early commissioning data ob- tained with the Australian Square Kilometer Array Pathfinder (ASKAP) revealed the existence of three galaxy-size clouds of neutral hydrogen in the galaxy group IC 1459 without any documented optical emission (Serra et al. 2015). One of these clouds is linked to NGC 7418 while the two other clouds, which are the target of this study, are located near IC 5270. Each of these two large H i accumulations have the characteristic H i mass of galaxies (∼ 109M⊙) but, intriguingly, have no optical counterpart in the Digital Sky Survey – see Figure 1, top panel. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | est sky conditions (Surface Brightness 20th percentile). These observing conditions are optimal to detect ex- tended low-surface brightness objects in g-band. The seeing ranged between 0.9′′ and 1.5′′, resulting in an effec- tive image seeing in the coadded image of 1.1′′. Fourteen exposures of 200 s were obtained for the North cloud, and eleven exposures of the North-Eastern cloud. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | A 20′′ wide dither pattern was chosen to cover the gaps between the three GMOS detectors, and the two GMOS pointings overlap so that we could image a continuous field over the area of the H i detections. A summary of the observations is given in Table 1. |
Processed_Gemini_Follow-up_of_two_massive_HI_clouds_discover.txt | The galaxy group IC 1459 is located at a distance of 29 Mpc (Blakeslee et al. 2001) and is named after its brightest (mB = 11.0 mag), early type member. |
Subsets and Splits