[go: up one dir, main page]

Next Issue
Volume 12, August
Previous Issue
Volume 12, April
 
 

Galaxies, Volume 12, Issue 3 (June 2024) – 11 articles

Cover Story (view full-size image): Galaxies that are edge-on to the line of sight offer a unique perspective for analyzing the dynamic relationship between the star forming disk and the surrounding gaseous halo. By focusing on radio observations of 35 edge-on galaxies, the CHANG-ES (Continuum Halos in Nearby Galaxies—an EVLA Survey) project has revealed stunning images of the dynamic disk–halo interface, as well as cosmic rays and magnetic fields that can extend over 10 kpc from the disk. A decade of scrutiny has shown that halo magnetic fields form stable patterns on large scales, likely generated by galactic dynamos. For a summary of the many diverse results from CHANG-ES, see https://www.mdpi.com/2075-4434/12/3/22
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Select all
Export citation of selected articles as:
18 pages, 3160 KiB  
Article
A Supermassive Binary Black Hole Candidate in Mrk 501
by Gustavo Magallanes-Guijón and Sergio Mendoza
Galaxies 2024, 12(3), 30; https://doi.org/10.3390/galaxies12030030 - 18 Jun 2024
Cited by 1 | Viewed by 645
Abstract
Using multifrequency observations, from radio to γ-rays of the blazar Mrk 501, we constructed their corresponding light curves and built periodograms using RobPer and Lomb–Scargle algorithms. Long-term variability was also studied using the power density spectrum and the detrended function analysis. Using [...] Read more.
Using multifrequency observations, from radio to γ-rays of the blazar Mrk 501, we constructed their corresponding light curves and built periodograms using RobPer and Lomb–Scargle algorithms. Long-term variability was also studied using the power density spectrum and the detrended function analysis. Using the software VARTOOLS Version 1.40, we also computed the analysis of variance, box-least squares and discrete fourier transform. The result of these techniques showed an achromatic periodicity ≲229d. This, combined with the result of pink-color noise in the spectra, led us to propose that the periodicity was produced via a secondary eclipsing supermassive binary black hole orbiting the primary one locked inside the central engine of Mrk 501. We built a relativistic eclipsing model of this phenomenon using Jacobi elliptical functions, finding a periodic relativistic eclipse occurring every ∼224d in all the studied wavebands. This implies that the frequency of the emitted gravitational waves falls slightly above 0.1 mHz, well within the operational range of the upcoming LISA space-based interferometer, and as such, these gravitational waves must be considered as a prime science target for future LISA observations. Full article
Show Figures

Figure 1

Figure 1
<p>Multifrequency light curves of the blazar Mrk 501. From left to right and top to bottom, the panels represent radio, optical, X- and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays at a <math display="inline"><semantics> <mrow> <mn>3</mn> <mi>σ</mi> </mrow> </semantics></math> confidence level for light curves, as described in the text.</p>
Full article ">Figure 2
<p>RobPer (magenta) and L-S (teal) periodograms (represented by their normalized coefficient of determination -NCoD), together with their corresponding window function (blue) for radio-, optical-, X- and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray observations of Mrk 501 are shown from left to right, top to bottom panels. The black dotted vertical line shows the mean periodicity between the RobPer and L-S peaks that are common in all frequencies (radio: <math display="inline"><semantics> <mrow> <mn>228.03</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>; optical: <math display="inline"><semantics> <mrow> <mn>226.77</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>; X-rays: <math display="inline"><semantics> <mrow> <mn>223.20</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>; and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays: <math display="inline"><semantics> <mrow> <mn>238.90</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>). <a href="#galaxies-12-00030-t002" class="html-table">Table 2</a> shows the periods obtained for the periodograms presented in the figure. The fact that these mean periodicities do not coincide with peaks in the window function reinforces their true periodic character.</p>
Full article ">Figure 3
<p>The figure shows the analysis of variance (AoV) for Mrk 501 using VARTOOLS. The left panel is the AoV for all frequencies: radio is in magenta with a periodicity of <math display="inline"><semantics> <mrow> <mn>227.2</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, optical is in blue with a periodicity of <math display="inline"><semantics> <mrow> <mn>226.73</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, X-rays, in teal, have a periodicity of <math display="inline"><semantics> <mrow> <mn>227.1</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays, in yellow with a periodicity of <math display="inline"><semantics> <mrow> <mn>229.2</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>. The mean value of <math display="inline"><semantics> <mrow> <mn>227.55</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math> is represented with a dashed vertical line. The gray vertical band zone represents the statistical range of these peaks. The right panel uses the same coloring scheme as the left one but for the harmonic analysis of variance (AoV-h) of the VARTOOLS software version 1.40. The periodicity of radio, optical, X- and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays are given via the following: <math display="inline"><semantics> <mrow> <mn>228.06</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>227.4</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>223.4</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>219.7</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, respectively, yielding an average value of <math display="inline"><semantics> <mrow> <mn>224.64</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, shown with the vertical dashed line. The vertical values on both panels were normalized to the maximum.</p>
Full article ">Figure 4
<p>The left panel shows the B-L Square algorithm of VARTOOLS used in all wavebands, applying the same coloring scheme of <a href="#galaxies-12-00030-f003" class="html-fig">Figure 3</a>, with the following periodicities in radio, optical, X- and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays: <math display="inline"><semantics> <mrow> <mn>220.96</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>224.141</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>220.96</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>240.404</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, with a mean value of <math display="inline"><semantics> <mrow> <mn>226.616</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math> represented using a dashed horizontal line. The right panel, with the same coloring scheme, uses the DFT VARTOOLS algorithm with resulting periodicities of <math display="inline"><semantics> <mrow> <mn>228.737</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>225.909</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>222.374</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>242.818</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math> and an average value of <math display="inline"><semantics> <mrow> <mn>229.959</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math>. The vertical values on both panels were normalized to the maximum. The gray vertical band zone represents the statistical range of these peaks.</p>
Full article ">Figure 5
<p>From left to right and top to bottom, the panels in the figure correspond to radio, optical, X- and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays’ power spectrum density (PSD) for the blazar Mrk 501. In all cases, the color of the noise of the signal is pink, according to the results in <a href="#galaxies-12-00030-t005" class="html-table">Table 5</a>.</p>
Full article ">Figure 6
<p>The illustration shows a black hole orbiting a central spherical source of light that eclipses the radiation detected by an observer. For simplicity, and in order to amplify the magnification effect detected by the observer, the example shown in the figure has the line of sight of the observer within the plane of orbit. The right plot shows the radiation flux detected by the observer. It consists of a numerical simulation of a Schwarzschild black hole orbiting a fixed, spherical source of light. Over an orbital period, the passage of the black hole through the line of sight of the observer magnifies the flux detected. The plotted flux and time are normalized to numerical units for a spherical source of radius five emitting isotropic radiation, a Schwarzschild radius of the black hole of one and an orbit of radius thirty. The ray-tracing technique used for this simulation was performed using a squared screen normal to the line of sight at a distance of one thousand. A video of this numerical simulation can be found at <a href="https://archive.org/details/blackhole_magnification" target="_blank">https://archive.org/details/blackhole_magnification</a>, accessed on 15 March 2024, and it was produced using a GNU General Public License (GPL) code named aztekas-shadows, which is under development and will eventually be available at <a href="https://aztekas.org" target="_blank">https://aztekas.org</a>, accessed on 15 March 2024, copyright ©2020 Gustavo Magallanes-Guijón, Sergio Mendoza and Milton Jair Santibañez-Armenta.</p>
Full article ">Figure 7
<p>The figure shows plots of an artificial eclipse that occurs two times. The software that produced them is described in <a href="#sec3dot4-galaxies-12-00030" class="html-sec">Section 3.4</a> and bears the copyright ©2022 Gustavo Magallanes-Guijón and Sergio Mendoza. From top to bottom, different values of the Jacobi elliptic function <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>0.9</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>0.999</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>0.99999</mn> </mrow> </semantics></math> were chosen and for all plots with the duration time of the eclipse, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi mathvariant="normal">e</mi> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, for a quiescent time, <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi mathvariant="normal">q</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>, and an amplitude, <math display="inline"><semantics> <mrow> <mi>A</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. The left column shows relativistic eclipses that produce magnification, which correspond to a plus sign in the simplified Equation (<a href="#FD6-galaxies-12-00030" class="html-disp-formula">6</a>), and the right column represents a standard, non-relativistic eclipse, showing the diminishing of the radiation represented by a minus sign in the same equation.</p>
Full article ">Figure 8
<p>From top to bottom, the figure shows <math display="inline"><semantics> <mrow> <mn>224</mn> <mspace width="0.166667em"/> <mi mathvariant="normal">d</mi> </mrow> </semantics></math> folded light curves for radio, optical, X- and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays. The solid curves are the best fit eclipse model described in <a href="#sec3dot4-galaxies-12-00030" class="html-sec">Section 3.4</a> constructed using the results of <a href="#galaxies-12-00030-t006" class="html-table">Table 6</a>. The shaded zone in each panel represents the duration of the eclipse. The dotted horizontal lines represent the <math display="inline"><semantics> <mrow> <mn>1</mn> <mi>σ</mi> </mrow> </semantics></math> significance level.</p>
Full article ">
14 pages, 7277 KiB  
Article
Planetary Nebula Morphologies Indicate a Jet-Driven Explosion of SN 1987A and Other Core-Collapse Supernovae
by Noam Soker
Galaxies 2024, 12(3), 29; https://doi.org/10.3390/galaxies12030029 - 6 Jun 2024
Cited by 3 | Viewed by 623
Abstract
I demonstrate the usage of planetary nebulae (PNe) to infer that a pair of jets shaped the ejecta of the core-collapse supernova (CCSN) SN 1987A. The main structure of the SN 1987A inner ejecta, the ‘keyhole’, comprises two low-intensity zones. The northern one [...] Read more.
I demonstrate the usage of planetary nebulae (PNe) to infer that a pair of jets shaped the ejecta of the core-collapse supernova (CCSN) SN 1987A. The main structure of the SN 1987A inner ejecta, the ‘keyhole’, comprises two low-intensity zones. The northern one has a bright rim on its front, while the southern one has an elongated nozzle. An earlier comparison of the SN 1987A ‘keyhole’ with bubbles in the galaxy group NGC 5813 led to its identification as a jet-shaped rim–nozzle structure. Here, I present rim–nozzle asymmetry in planetary nebulae (PNe), thought to be shaped by jets, which solidifies the claim that jets powered the ejecta of SN 1987A and other CCSNe. This finding for the iconic SN 1987A with its unique properties strengthens the jittering-jets explosion mechanism (JJEM) of CCSNe. In a few hundred years, the CCSN 1987A will have a complicated structure with two main symmetry axes, one along the axis of the three circumstellar rings that was shaped by two opposite 20,000-year pre-explosion jets, and the other along the long axis of the ‘keyhole’ that was shaped by the main (but not the only) jet pair of the exploding jets of SN 1987A in the frame of the JJEM. Full article
(This article belongs to the Special Issue Origins and Models of Planetary Nebulae)
Show Figures

Figure 1

Figure 1
<p>HST/WFC3 images of SN 1987A in nine filters and at 35.5 years post-explosion, adapted from Rosu et al. (2024) [<a href="#B57-galaxies-12-00029" class="html-bibr">57</a>], with the marks identifying the bubbles (black arrows), nozzle, and rim (red arrows), from Soker (2024c [<a href="#B71-galaxies-12-00029" class="html-bibr">71</a>]). The bright structure with the low-intensity inner zones inside the purple line in the lower-right panel is the ‘keyhole’. The north bubble with its front rim (red arrow) and the south bubble with its nozzle (dashed light-blue arrow) form the rim–nozzle asymmetry.</p>
Full article ">Figure 2
<p>(<b>a</b>) Chandra 80 kpc by 80 kpc X-ray smoothed 0.3–3 keV image of the cooling-flow galaxy group NGC 5813 (adapted from the Chandra site and based on Randall et al. 2015; credit: NASA/CXC/SAO/S, Randall [<a href="#B52-galaxies-12-00029" class="html-bibr">52</a>]). (<b>b</b>): The inner 27 kpc by 27 kpc of the left panel. Marks on both panels are from Soker (2024c) [<a href="#B71-galaxies-12-00029" class="html-bibr">71</a>]. Note the <span class="html-italic">rim–nozzle asymmetry</span> of the bright inner structure: the northern bubble has rim 1E on its front, while the southern bubble has a nozzle (low-intensity zone) on its front.</p>
Full article ">Figure 3
<p>A collection of eight PNe that possess rim–nozzle asymmetry (panels (<b>a</b>–<b>h</b>)). The solid red arrows point at the rims, and the dashed light-blue arrows point at the nozzles. NGC 6881 (panel (<b>i</b>)) has two nozzles with the remnants of the jets (pointed at with thin yellow arrows but that can be seen only in the inset; see also Ramos-Larios et al. 2008 [<a href="#B51-galaxies-12-00029" class="html-bibr">51</a>]) that probably shaped the nozzles. The inset is an image from Kwok and Su (2005) [<a href="#B34-galaxies-12-00029" class="html-bibr">34</a>], in which low-ionization gas better shows the remnants of the jets. Sources and credits: NGC 2818: NASA, ESA, and Hubble Heritage Team; M 4-14 + M 3-28: Manchado et al. (1996) [<a href="#B37-galaxies-12-00029" class="html-bibr">37</a>]; NGC 6578: Palen et al. (2002) [<a href="#B46-galaxies-12-00029" class="html-bibr">46</a>]; Hubble 5: Bruce Balick, Vincent Icke, Garrelt Mellema, and NASA/ESA; MRSL 252: Corradi et al. (1997) [<a href="#B16-galaxies-12-00029" class="html-bibr">16</a>]; K1-10: Schwartz et al. (1992) [<a href="#B62-galaxies-12-00029" class="html-bibr">62</a>]; NGC 6790: Hubble site; NGC 6881: HST NASA/ESA, with inset from Kwok and Su (2005) [<a href="#B34-galaxies-12-00029" class="html-bibr">34</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) An image of new CCSNR candidate G107.7-5.1 adapted from Fesen et al. (2024) [<a href="#B20-galaxies-12-00029" class="html-bibr">20</a>]. Marks of the nozzle (dashed light-blue arrow), bubble (dashed green arrow), and rims are as in Soker (2024c) [<a href="#B71-galaxies-12-00029" class="html-bibr">71</a>]. (<b>b</b>) Image of the galaxy Hercules A in radio (blue), X-ray (purple), and optical (white). The radio shows the jets that inflate multiple rims; two are marked by the yellow arrows (credit: X-ray: NASA/CXC/SAO; optical: NASA/STScI; radio: NSF/NRAO/VLA). (<b>c</b>) Hα image of the PN KjPn 8 adapted from Lopez et al. (2000) [<a href="#B36-galaxies-12-00029" class="html-bibr">36</a>]. I added the marks of the nozzle (dashed light-blue arrow) and the three rims.</p>
Full article ">Figure 5
<p>(<b>a</b>) ROSAT X-ray image of Vela CCSNR based on Figure 1 from Sapienza et al. (2021) [<a href="#B61-galaxies-12-00029" class="html-bibr">61</a>], who marked the clumps and white line; the original X-ray image and identification of clumps A–F are from Aschenbach et al. (1995) [<a href="#B4-galaxies-12-00029" class="html-bibr">4</a>]. The thick yellow DEline, thick yellow FJ-line, and two dashed black lines are additions from Soker (2023b) [<a href="#B68-galaxies-12-00029" class="html-bibr">68</a>]. (<b>b</b>) Color-coded (inset), exposure-corrected eROSITA X-ray image of Vela adapted from Mayer et al. (2023) [<a href="#B39-galaxies-12-00029" class="html-bibr">39</a>]. The new eROSITA X-ray image allowed me (Soker 2024c [<a href="#B71-galaxies-12-00029" class="html-bibr">71</a>]) to identify clumps H and H2 at equal distances from the center (solid red line), solidifying the classification of a point-symmetric CCSNR. Here, I added the marks of three possible rim–nozzle asymmetry pairs.</p>
Full article ">Figure 6
<p>(<b>a</b>) A radio image of SNR G309.2-00.6 from Gaensler et al. (1998) [<a href="#B25-galaxies-12-00029" class="html-bibr">25</a>], with marks from the original image. Note that they identified a jet inside what I identify as a nozzle! (<b>b</b>) Focus on the inner area of the left panel with my identification of the rim–nozzle asymmetry, the two arcs (‘crescents’), and the symmetry axis of the arcs and the rim–nozzle structure (yellow line).</p>
Full article ">Figure 7
<p>A JWST image of SN 1987a with marks from Matsuura et al. (2024) [<a href="#B38-galaxies-12-00029" class="html-bibr">38</a>]. I added the marks of arcs, as I identified their termed ‘crescents’ with arcs, as used in earlier studies of CCSNRs.</p>
Full article ">
11 pages, 3803 KiB  
Article
Wave-Particle Interactions in Astrophysical Plasmas
by Héctor Pérez-De-Tejada
Galaxies 2024, 12(3), 28; https://doi.org/10.3390/galaxies12030028 - 6 Jun 2024
Viewed by 466
Abstract
Dissipation processes derived from the kinetic theory of gases (shear viscosity and heat conduction) are employed to examine the solar wind that interacts with planetary ionospheres. The purpose of this study is to estimate the mean free path of wave-particle interactions that produce [...] Read more.
Dissipation processes derived from the kinetic theory of gases (shear viscosity and heat conduction) are employed to examine the solar wind that interacts with planetary ionospheres. The purpose of this study is to estimate the mean free path of wave-particle interactions that produce a continuum response in the plasma behavior. Wave-particle interactions are necessary to support the fluid dynamic interpretation that accounts for the interpretation of various features measured in a solar wind–planet ionosphere region; namely, (i) the transport of solar wind momentum to an upper ionosphere in the presence of a velocity shear, and (ii) plasma heating produced by momentum transport. From measurements conducted in the solar wind interaction with the Venus ionosphere, it is possible to estimate that in general terms, the mean free path of wave-particle interactions reaches λH ≥ 1000 km values that are comparable to the gyration radius of the solar wind particles in their Larmor motion within the local solar wind magnetic field. Similar values are also applicable to conditions measured by the Mars ionosphere and in cometary plasma wakes. Considerations are made in regard to the stochastic trajectories of the plasma particles that have been implied from the measurements made in planetary environments. At the same time, it is as possible that the same phenomenon is applicable to the interaction of stellar winds with the ionosphere of exoplanets, and also in regions where streaming ionized gases reach objects that are subject to rotational motion in other astrophysical problems (galactic flow–plasma interactions, black holes, etc.). Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Lower panel</b>) Trajectory of the Mariner 5 spacecraft projected in cylindrical coordinates in its flyby past Venus. Labels 1 to 5 along the trajectory mark important events in the plasma properties (a bow shock is identified at features 1 and 5), and the intermediate plasma transition occurs at features 2 and 4). (<b>Upper panel</b>) Magnetic field intensity and its latitudinal and azimuthal orientation, together with the plasma properties (thermal speed, density, and bulk speed) measured around Venus [<a href="#B1-galaxies-12-00028" class="html-bibr">1</a>].</p>
Full article ">Figure 2
<p>Ion speed and temperature measured along the orbit of Venera 10 on 19 April 1976. The Venera orbit in cylindrical coordinates is shown at the top. The temperature burst at position 1 was recorded during a flank crossing of a bow shock. A boundary layer is apparent by the increase in temperature and decrease in speed, and is initiated by the intermediate transition at the position labeled 2. A latter discontinuity in the boundary layer temperature profile corresponds to the boundary of the magneto-tail (from [<a href="#B6-galaxies-12-00028" class="html-bibr">6</a>]).</p>
Full article ">Figure 3
<p>Vector velocity speeds of the trans-terminator flow in the Venus upper ionosphere measured with instruments onboard the Pioneer Venus Orbiter spacecraft [<a href="#B9-galaxies-12-00028" class="html-bibr">9</a>].</p>
Full article ">Figure 4
<p>Measured flow velocities versus VEX altitude for solar wind H<sup>+</sup> ions, and ionospheric H<sup>+</sup> and O<sup>+</sup> ions. The curve marked v<sub>esc</sub> illustrates escape velocity versus altitude above Venus. The data points represent average values in 50 km altitude intervals sampled within Y = +0.5 of the dawn-dusk Meridian (<b>left panel</b>) and of the noon–midnight Meridian (<b>right panel</b>). Regions and boundaries are marked on the right-hand side as the I-sphere (the ionopause (IP), and the ionosheath (IMB) (from Lundin et al., (2011) [<a href="#B17-galaxies-12-00028" class="html-bibr">17</a>]).</p>
Full article ">Figure 5
<p>Mean free path values λ<sub>H</sub> of the solar wind obtained in wave-particle interactions and also in magnetic field fluctuations using the solar wind thermal speed V<sub>T</sub> and its kinematic viscosity coefficient ﬠ during the Mariner 5 trajectory in <a href="#galaxies-12-00028-f001" class="html-fig">Figure 1</a>. The connecting line labeled “W” refers to a value in the Venus inner and in the outer ionosheath that is implied by wave-particle interactions. The connecting line labeled “F” is implied by the magnetic field fluctuations.</p>
Full article ">Figure 6
<p>(<b>Upper panel</b>) Trajectory of the PVO in orbit 87 projected on one quadrant in cylindrical coordinates. The bow shock, the intermediate transition, and the ionopause are indicated. (<b>Lower panel</b>) Ion flux values measured as a function of energy in cycles I, II, III, and IV state their start time (their position is noted in the upper panel along the PVO trajectory). Positions A, B, and C in spectrum III mark the time when the ion fluxes were obtained ([<a href="#B24-galaxies-12-00028" class="html-bibr">24</a>]).</p>
Full article ">Figure 7
<p>View of a corkscrew vortex flow in fluid dynamics. Its geometry is equivalent to that of a vortex flow in the Venus wake, with its width and position varying during the solar cycle. Near the solar cycle minimum, the vortex is located closer to Venus (located by the right side) and there are also indications that its width becomes smaller with increasing distance downstream from the planet [<a href="#B13-galaxies-12-00028" class="html-bibr">13</a>].</p>
Full article ">
17 pages, 1175 KiB  
Article
The Response of the Inner Dark Matter Halo to Stellar Bars
by Daniel A. Marostica, Rubens E. G. Machado, E. Athanassoula and T. Manos
Galaxies 2024, 12(3), 27; https://doi.org/10.3390/galaxies12030027 - 28 May 2024
Cited by 2 | Viewed by 559
Abstract
Barred galaxies constitute about two-thirds of observed disc galaxies. Bars affect not only the mass distribution of gas and stars but also that of the dark matter. An elongation of the inner dark matter halo is known as the halo bar. We aim [...] Read more.
Barred galaxies constitute about two-thirds of observed disc galaxies. Bars affect not only the mass distribution of gas and stars but also that of the dark matter. An elongation of the inner dark matter halo is known as the halo bar. We aim to characterize the structure of the halo bars, with the goal of correlating them with the properties of the stellar bars. We use a suite of simulated galaxies with various bar strengths, including gas and star formation. We quantify the strengths, shapes, and densities of these simulated stellar bars. We carry out numerical experiments with frozen and analytic potentials in order to understand the role played by a live responsive stellar bar. We find that the halo bar generally follows the trends of the disc bar. The strengths of the halo and stellar bars are tightly correlated. Stronger bars induce a slight increase in dark matter density within the inner halo. Numerical experiments show that a non-responsive frozen stellar bar would be capable of inducing a dark matter bar, but it would be weaker than the live case by a factor of roughly two. Full article
Show Figures

Figure 1

Figure 1
<p>Projected dark matter mass of halo particles within <math display="inline"><semantics> <mrow> <mo>|</mo> <mi>z</mi> <mo>|</mo> <mo>&lt;</mo> <mn>1</mn> </mrow> </semantics></math> kpc at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> Gyr. The contours highlight the shape of the inner halo. As in <a href="#galaxies-12-00027-t001" class="html-table">Table 1</a>: columns are initial halo shapes; rows are initial gas fractions.</p>
Full article ">Figure 2
<p>Strength of the disc and halo bars as a function of time. The three columns correspond to the three different initial halo triaxialities, respectively from left to right. The five rows, from top to bottom, correspond to increasing initial gas fraction (see <a href="#galaxies-12-00027-t001" class="html-table">Table 1</a>). The orange curves represent the disc bar and their scale is shown in the left ticks. For better comparison, we use a different scale for the halo bar (measured within <math display="inline"><semantics> <mrow> <mo>|</mo> <mi>z</mi> <mo>|</mo> <mo>&lt;</mo> <mn>1</mn> </mrow> </semantics></math> kpc), depicted in blue, to the right.</p>
Full article ">Figure 3
<p>Halo bar strength as a function of disc bar strength at <span class="html-italic">t</span> = 10 Gyr. Circles, squares and triangles stand for haloes 1, 2 and 3, respectively. Color is according to the initial gas fraction. Scales of halo <math display="inline"><semantics> <msub> <mi>A</mi> <mn>2</mn> </msub> </semantics></math> and disc <math display="inline"><semantics> <msub> <mi>A</mi> <mn>2</mn> </msub> </semantics></math> differ by approximately an order of magnitude.</p>
Full article ">Figure 4
<p>From top to bottom, dark matter halo axis ratios (<math display="inline"><semantics> <mrow> <mi>b</mi> <mo>/</mo> <mi>a</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>/</mo> <mi>a</mi> </mrow> </semantics></math>) and triaxiality parameter <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>B</mi> <mi>A</mi> </mrow> </msub> </semantics></math>, respectively, as a function of disc bar strength at <span class="html-italic">t</span> = 10 Gyr. Circles, squares and triangles stand for haloes 1, 2 and 3, respectively. Color is according to initial gas fraction.</p>
Full article ">Figure 5
<p>Dark matter halo <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>/</mo> <mi>a</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>c</mi> <mo>/</mo> <mi>a</mi> </mrow> </semantics></math> axis ratios as a function of one another at <span class="html-italic">t</span> = 10 Gyr. Circles, squares and triangles stand for haloes 1, 2 and 3, respectively. Color is according to initial gas fraction. Data above close to the diagonal line indicate prolateness of the halo.</p>
Full article ">Figure 6
<p>Central dark matter density (inside 5 kpc) as a function of disc bar strength at <span class="html-italic">t</span> = 10 Gyr. Circles, squares and triangles stand for haloes 1, 2 and 3, respectively. Color is according to initial gas fraction. For better visualization, grey lines connect galaxies of the same initial halo (halo 1, halo 2 or halo 3).</p>
Full article ">Figure 7
<p>Face-on and side-on projections of the stellar disc from model 101 at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> Gyr. Contours represent the potential and colors represent the density. These frozen particles were used in run 101-F.</p>
Full article ">Figure 8
<p>Strength of the halo bar of galaxies 101, 101-F and 101-A as a function of time. All galaxies evolve inside live haloes. Galaxy 101 (black line) is the default model and it features a live <span class="html-italic">N</span>-body disc. Galaxy 101-F (orange) evolves in the presence of a frozen barred disc potential. Galaxy 101-A (green) has a time-dependent analytic potential emulating the growth of a barred disc. Galaxy 101-R (blue) has a rotating rigid disc.</p>
Full article ">Figure 9
<p>(<b>Left</b>): in the rotating rigid disc 101-R, the stellar mass is grown smoothly from 0 to <math display="inline"><semantics> <msub> <mi>M</mi> <mi mathvariant="normal">d</mi> </msub> </semantics></math> during the first 1 Gyr. (<b>Right</b>): the measured pattern speed from the bar in model 101 (symbols) was used to impose a rotation to model 101-R (line).</p>
Full article ">
22 pages, 2559 KiB  
Article
ALMA Band 3 Source Counts: A Machine Learning Approach to Contamination Mitigation below 5 Sigma
by Ivano Baronchelli, Matteo Bonato, Gianfranco De Zotti, Viviana Casasola, Michele Delli Veneri, Fabrizia Guglielmetti, Elisabetta Liuzzo, Rosita Paladino, Leonardo Trobbiani and Martin Zwaan
Galaxies 2024, 12(3), 26; https://doi.org/10.3390/galaxies12030026 - 20 May 2024
Viewed by 632
Abstract
We performed differential number counts down to 4.25 sigma using ALMA Band 3 calibrator images, which are known for their high dynamic range and susceptibility to various types of contamination. Estimating the fraction of contaminants is an intricate process due to correlated non-Gaussian [...] Read more.
We performed differential number counts down to 4.25 sigma using ALMA Band 3 calibrator images, which are known for their high dynamic range and susceptibility to various types of contamination. Estimating the fraction of contaminants is an intricate process due to correlated non-Gaussian noise, and it is often compounded by the presence of false positives generated during the cleaning phase. In addition, calibrator extensions further complicate the counting of background sources. In order to address these challenges, our strategy employs a machine learning-based approach utilizing the UMLAUT algorithm. UMLAUT assigns a value to each detection, and it considers how likely it is for there to be a genuine background source or a contaminant. With respect to this goal, we provide UMLAUT with eight observational input parameters, each automatically weighted using a gradient descent method. Our methodology significantly improves the precision of differential number counts, thus surpassing conventional techniques, including visual inspection. This study contributes to a better understanding of radio sources, particularly in the challenging sub-5 sigma regime, within the complex context of a high dynamic range of ALMA calibrator images. Full article
Show Figures

Figure 1

Figure 1
<p>Band 3 images of calibrator J1550 + 0527. In the (<b>left panel</b>) is an example of an image excluded from our sample due to surpassing certain thresholds specified for image rejection (the specific values of Galactic latitude, <math display="inline"><semantics> <mrow> <msub> <mi>σ</mi> <mi>pix</mi> </msub> <mo>,</mo> <mi>H</mi> <mi>D</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>G</mi> <mi>D</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>D</mi> </mrow> </semantics></math>, and FWHM are reported in the text for both the images). In the (<b>right panel</b>) is the image included in our sample for the same calibrator. The general shape of the noise patterns was not directly related to any of the selection parameters in particular. However, the same parameters were related with the probability to observe the non-random patterns. The scientific notation “<span class="html-italic">a</span>E<span class="html-italic">b</span>” indicates <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>×</mo> <msup> <mn>10</mn> <mi>b</mi> </msup> </mrow> </semantics></math>. For example, 1.5E8 stands for <math display="inline"><semantics> <mrow> <mn>1.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>The completeness measurements (the black-filled circles and error bars) and best-fitting curve (the black line and red-filled area). Both the measurements and the fitting curve were corrected for the bias due to the fraction of the spurious sources misidentified as injected sources.</p>
Full article ">Figure 3
<p>Contamination fraction as a function of the signal-to-noise ratio (black-filled circles and error bars). The best-fitting logistic curve and its associated uncertainty are shown with a black line and a red-filled area, respectively. This type of contamination accounts only for the fraction of false detections due to the stochastic distribution of the pixels fluxes. Other sources of contamination are taken into account separately.</p>
Full article ">Figure 4
<p>Ratio between the recovered total flux (<math display="inline"><semantics> <msubsup> <mi>F</mi> <mrow> <mi>AUTO</mi> </mrow> <mi>output</mi> </msubsup> </semantics></math>) and flux originally injected for the simulated sources (<math display="inline"><semantics> <msubsup> <mi>F</mi> <mrow> <mi>tot</mi> </mrow> <mi>input</mi> </msubsup> </semantics></math>) as a function of the intrinsic SNR. Within each SNR bin, the data dispersion are indicated by the black error bars (where the uncertainty in the average measurements was found to be negligible and smaller than the circles used in the plot). Three different SNR thresholds, <math display="inline"><semantics> <mrow> <mn>3</mn> <mi>σ</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>4</mn> <mi>σ</mi> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mn>5</mn> <mi>σ</mi> </mrow> </semantics></math>, are indicated using dashed vertical lines. It is possible to note that only sources injected below <math display="inline"><semantics> <mrow> <mn>4</mn> <mi>σ</mi> </mrow> </semantics></math> or <math display="inline"><semantics> <mrow> <mn>5</mn> <mi>σ</mi> </mrow> </semantics></math> were substantially affected by flux boosting. This, however, does not indicate that the flux boosting effect can be ignored, even when using these safer thresholds. The continuous curves (with red denoting a <math display="inline"><semantics> <mrow> <mn>3</mn> <mi>σ</mi> </mrow> </semantics></math> threshold, green for <math display="inline"><semantics> <mrow> <mn>4</mn> <mi>σ</mi> </mrow> </semantics></math>, and blue for <math display="inline"><semantics> <mrow> <mn>5</mn> <mi>σ</mi> </mrow> </semantics></math>) visually indicate the fraction of sources, which were injected at various levels of SNR, entering the sample. The fraction can be estimated comparing the part of dispersion bars above and below the curve considered. For example, when considering a <math display="inline"><semantics> <mrow> <mn>5</mn> <mi>σ</mi> </mrow> </semantics></math> threshold (blue curve), half of the sources <span class="html-italic">injected</span> at <math display="inline"><semantics> <mrow> <mn>5</mn> <mi>σ</mi> </mrow> </semantics></math> will be excluded from the sample, while this fraction decreases to ∼16% when the <span class="html-italic">injected</span> flux is <math display="inline"><semantics> <mrow> <mn>7</mn> <mi>σ</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>Example of the flux-boosting correction curve estimation for a source in our sample. In the (<b>upper panel</b>), we segmented the theoretical counts curve (black line) and the sum of the predicted number counts from the “C2Ex” model by [<a href="#B25-galaxies-12-00026" class="html-bibr">25</a>] for RL-AGNs with the number counts of SF galaxies, as predicted by [<a href="#B26-galaxies-12-00026" class="html-bibr">26</a>], into logarithmic flux bins. The centers of the bins are denoted by vertical, dashed lines. Given the SNR of each bin, we obtained the distribution of recovered SNRs from our simulations. All these curves are integrated, with results displayed from the lowest yellow curve (representing the first bin) to the highest one (representing the sum over all bins). The ratio between the final integrated curve and the theoretical number counts curve yields the correction curve displayed in the (<b>bottom panel</b>). The red-shaded area in the (<b>bottom panel</b>) represents the associated uncertainty (which is directly derived from the uncertainty on the integrated curve and is depicted as two dotted lines in the (<b>upper panel</b>)). The horizontal blue line represents the correction required for this specific source (which is detected at approximately <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>σ</mi> </mrow> </semantics></math>). This correction curve comprehensively addresses both the completeness and flux-boosting effects.</p>
Full article ">Figure 6
<p>(<b>Left</b>) image of the ALMA calibrator J0215-0222. We measured the average flux along a vertical slit (shown in red), which is positioned to include the positions of both the calibrator and the main lobe. The fluxes, averaged along the x axis, are shown in the plot on the (<b>right</b>). The scientific notation “<span class="html-italic">a</span>E<span class="html-italic">b</span>” indicates <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>×</mo> <msup> <mn>10</mn> <mi>b</mi> </msup> </mrow> </semantics></math>. For example, 1.5E8 stands for <math display="inline"><semantics> <mrow> <mn>1.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Example of an image affected by multiple false detections. We display all the contaminants detected above <math display="inline"><semantics> <mrow> <mn>3</mn> <mi>σ</mi> </mrow> </semantics></math>. For each source, we indicate the ID, which is followed by the SNR in parentheses. It is possible to note that the distribution of contaminants exhibited a distinct spatial pattern characterized by an elliptical shape. The scientific notation “<span class="html-italic">a</span>E<span class="html-italic">b</span>” indicates <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>×</mo> <msup> <mn>10</mn> <mi>b</mi> </msup> </mrow> </semantics></math>. For example, 1.5E8 stands for <math display="inline"><semantics> <mrow> <mn>1.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Differential number counts obtained by correcting for completeness, flux boosting, and for the contamination by noise fluctuation. Conversely, the contamination by calibrator extensions and false positives, which substantially affected our number counts above ∼0.3 mJy, was not corrected. The model predictions by [<a href="#B26-galaxies-12-00026" class="html-bibr">26</a>] for dusty, star-forming galaxies and by [<a href="#B25-galaxies-12-00026" class="html-bibr">25</a>] for radio sources are shown by the green and red dashed lines, respectively.</p>
Full article ">Figure 9
<p>Differential number counts obtained by eliminating the contaminants that were caused due to the calibrator extensions and cleaning residuals identified through visual inspection. The theoretical models of [<a href="#B25-galaxies-12-00026" class="html-bibr">25</a>,<a href="#B26-galaxies-12-00026" class="html-bibr">26</a>] are shown by green and red dashed lines, respectively.</p>
Full article ">Figure 10
<p>The differential number counts derived by excluding the detections classified (with a probability exceeding 50%) as calibrator extensions or cleaning residuals by <span class="html-italic">UMLAUT</span>. For cases below ∼0.3 mJy (vertical dotted line), we relied on our own visual inspection. The theoretical models of [<a href="#B25-galaxies-12-00026" class="html-bibr">25</a>,<a href="#B26-galaxies-12-00026" class="html-bibr">26</a>] are shown by the green and red dashed lines, respectively.</p>
Full article ">Figure 11
<p>The differential number counts derived by weighting the detections using the probabilities, which were computed by <span class="html-italic">UMLAUT</span>, of being calibrator extensions or cleaning residuals. The weights were computed only for sources above ∼0.3 mJy (vertical dotted line), and we relied on our visual classification for the few detections that occurred below this threshold. The theoretical models of [<a href="#B25-galaxies-12-00026" class="html-bibr">25</a>,<a href="#B26-galaxies-12-00026" class="html-bibr">26</a>] are shown by the green and red dashed lines, respectively.</p>
Full article ">
49 pages, 1670 KiB  
Article
High-Redshift Quasars at z ≥ 3: Radio Variability and MPS/GPS Candidates
by Yulia Sotnikova, Alexander Mikhailov, Timur Mufakharov, Tao An, Dmitry Kudryavtsev, Marat Mingaliev, Roman Udovitskiy, Anastasia Kudryashova, Vlad Stolyarov and Tamara Semenova
Galaxies 2024, 12(3), 25; https://doi.org/10.3390/galaxies12030025 - 15 May 2024
Cited by 1 | Viewed by 784
Abstract
We present a study of the radio variability of bright, S1.4100 mJy, high-redshift quasars at z3 on timescales of up to 30–40 yrs. The study involved simultaneous RATAN-600 measurements at the frequencies of 2.3, 4.7, 8.2, 11.2, and [...] Read more.
We present a study of the radio variability of bright, S1.4100 mJy, high-redshift quasars at z3 on timescales of up to 30–40 yrs. The study involved simultaneous RATAN-600 measurements at the frequencies of 2.3, 4.7, 8.2, 11.2, and 22.3 GHz in 2017–2020. In addition, data from the literature were used. We have found that the variability index, VS, which quantifies the normalized difference between the maximum and minimum flux density while accounting for measurement uncertainties, ranges from 0.02 to 0.96 for the quasars. Approximately half of the objects in the sample exhibit a variability index within the range from 0.25 to 0.50, which is comparable to that observed in blazars at lower redshifts. The distribution of VS at 22.3 GHz is significantly different from that at 2.3–11.2 GHz, which may be attributed to the fact that a compact AGN core dominates at the source’s rest frame frequencies greater than 45 GHz, leading to higher variability indices obtained at 22.3 GHz (the VS distribution peaks around 0.4) compared to the lower frequencies (the VS distribution at 2.3 and 4.7 GHz peaks around 0.1–0.2). Several source groups with distinctive variability characteristics were found using the cluster analysis of quasars. We propose seven new candidates for gigahertz-peaked spectrum (GPS) sources and five new megahertz-peaked spectrum (MPS) sources based on their spectrum shape and variability features. Only 6 out of the 23 sources previously reported as GPS demonstrate a low variability level typical of classical GPS sources (VS<0.25) at 4.7–22.3 GHz. When excluding the highly variable peaked-spectrum blazars, we expect no more than 20% of the sources in the sample to be GPS candidates and no more than 10% to be MPS candidates. Full article
Show Figures

Figure 1

Figure 1
<p>The variability (<b>top row</b>), modulation (<b>middle row</b>), and fractional variability (<b>bottom row</b>) index distributions for the whole sample, peaked-spectrum (PS) sources, and blazars.</p>
Full article ">Figure 2
<p>The variability index <math display="inline"><semantics> <msub> <mi>V</mi> <msub> <mi>S</mi> <mrow> <mn>4.7</mn> </mrow> </msub> </msub> </semantics></math> versus the number of observations N<sub>obs</sub> for the quasars. The size of the symbols is proportional to the years of monitoring in the rest frame, we define three scales: <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>rest</mi> </msub> <mo>≤</mo> <mspace width="3.33333pt"/> <mn>5</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>&lt;</mo> <msub> <mi>t</mi> <mi>rest</mi> </msub> <mo>≤</mo> <mn>10</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>rest</mi> </msub> <mo>&gt;</mo> <mn>10</mn> </mrow> </semantics></math> yrs. Bright quasars have been observed more frequently and for a longer period.</p>
Full article ">Figure 3
<p>The distribution of the observational frequencies converted to the rest frame (<b>top</b>), and the distribution of the light curve durations in the rest frame, <math display="inline"><semantics> <msub> <mi>t</mi> <mi>rest</mi> </msub> </semantics></math> in yrs (<b>bottom</b>).</p>
Full article ">Figure 4
<p>The variability indices plotted relatively to the rest frame frequencies. The outlier source is J0941+1145 with <math display="inline"><semantics> <msub> <mi>V</mi> <mi>S</mi> </msub> </semantics></math> equal to 0.86 and 0.96.</p>
Full article ">Figure 5
<p>The variability index <math display="inline"><semantics> <msub> <mi>V</mi> <mi>S</mi> </msub> </semantics></math> at 4.7 (<b>top</b>) and 22.3 GHz (<b>bottom</b>) versus redshift <span class="html-italic">z</span> for blazars (orange circles) and non-blazars (blue triangles). The size of the symbols is defined as in <a href="#galaxies-12-00025-f002" class="html-fig">Figure 2</a>. The five well-known blazars are tagged: J1026+2542 at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>5.25</mn> </mrow> </semantics></math> [<a href="#B63-galaxies-12-00025" class="html-bibr">63</a>,<a href="#B64-galaxies-12-00025" class="html-bibr">64</a>]; J1430+4204 at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>4.71</mn> </mrow> </semantics></math> [<a href="#B65-galaxies-12-00025" class="html-bibr">65</a>]; J0324−2918 at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>4.63</mn> </mrow> </semantics></math> [<a href="#B66-galaxies-12-00025" class="html-bibr">66</a>]; J0525−3343 at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>4.41</mn> </mrow> </semantics></math> [<a href="#B67-galaxies-12-00025" class="html-bibr">67</a>,<a href="#B68-galaxies-12-00025" class="html-bibr">68</a>]; J1028−0844 at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>4.27</mn> </mrow> </semantics></math> [<a href="#B69-galaxies-12-00025" class="html-bibr">69</a>].</p>
Full article ">Figure 6
<p>The <math display="inline"><semantics> <msub> <mi>V</mi> <msub> <mi>S</mi> <mrow> <mn>4.7</mn> </mrow> </msub> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>V</mi> <msub> <mi>S</mi> <mrow> <mn>22.3</mn> </mrow> </msub> </msub> </semantics></math> versus the rest frame timescale <math display="inline"><semantics> <msub> <mi>t</mi> <mi>rest</mi> </msub> </semantics></math> for the PS and non-PS quasars (orange and blue triangles). The symbol sizes are proportional to N<sub>obs</sub>, they correspond to 10, 20, 50, 100, and more number of observations. The dashed lines mark the 0.25 and 0.50 variability index levels.</p>
Full article ">Figure 7
<p>The variability indices <math display="inline"><semantics> <msub> <mi>V</mi> <msub> <mi>S</mi> <mrow> <mn>22.3</mn> </mrow> </msub> </msub> </semantics></math> (<b>top</b>) and <math display="inline"><semantics> <msub> <mi>V</mi> <msub> <mi>S</mi> <mrow> <mn>8.2</mn> </mrow> </msub> </msub> </semantics></math> (<b>bottom</b>) versus the high-frequency spectral index <math display="inline"><semantics> <msub> <mi>α</mi> <mi>high</mi> </msub> </semantics></math>, measured from the averaged historical spectra of the PS blazars and PS non-blazars candidates (orange and blue stars) and the blazars/non-blazars with another spectral types (orange and blue circles). The vertical dotted line corresponds to the spectral index <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mo>−</mo> <mn>0.5</mn> </mrow> </semantics></math>, the horizontal line corresponds to <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mi>S</mi> </msub> <mo>=</mo> <mn>0.25</mn> </mrow> </semantics></math>. Some variable quasars with steep-averaged radio spectra are tagged.</p>
Full article ">Figure 8
<p>The relation <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mi>high</mi> </msub> <mo>−</mo> <msub> <mi>S</mi> <mrow> <mn>11.2</mn> </mrow> </msub> </mrow> </semantics></math> for the blazar J0646+4451 is shown in the top panel. The spectral index <math display="inline"><semantics> <msub> <mi>α</mi> <mi>high</mi> </msub> </semantics></math> evolution is presented in the middle panel. Corresponding RATAN-600 quasi-simultaneous spectra are shown in the bottom panel (green color) together with the literature data (grey color). The orange line corresponds to the ultra-steep spectrum measured in 2011.</p>
Full article ">Figure 9
<p>The distributions of the parameters for PS sources. The observed and rest frame peak frequencies are on top, and the flux densities at the peak frequency and FWHMs are at the bottom.</p>
Full article ">Figure 10
<p>Redshift versus rest frame peak frequency for 47 PS quasars from the sample under investigation. The PS sources from [<a href="#B105-galaxies-12-00025" class="html-bibr">105</a>] are marked by grey triangles, from [<a href="#B106-galaxies-12-00025" class="html-bibr">106</a>] by star symbols, and from [<a href="#B37-galaxies-12-00025" class="html-bibr">37</a>] by grey squares.</p>
Full article ">Figure A1
<p>Correlation matrix with Kendall correlation coefficients.</p>
Full article ">Figure A2
<p><b>Left</b>: clusters in the primary component (PC) coordinates; <b>right</b>: 2D t-SNE representation of the left panel.</p>
Full article ">Figure A3
<p>Distributions of the considered characteristics for different clusters. Each panel corresponds to a certain characteristic, the <span class="html-italic">y</span>-axes show the cluster ordinal numbers. The distributions are shown as box plots, see description in the text.</p>
Full article ">Figure A4
<p>Comparison of the PCA+k-means clustering (<b>upper left panel</b>) with 5 runs of the SOM training (<b>other panels</b>).</p>
Full article ">Figure A5
<p>The RATAN-600 light curves, the measurements at 22.3 GHz are colored by orange, at 11.2 GHz by magenta, at 8.2 GHz by green, at 4.7 GHz by blue.</p>
Full article ">Figure A6
<p>The RATAN-600 light curves, the measurements at 22.3 GHz are colored by orange, at 11.2 GHz by magenta, at 8.2 GHz by green, at 4.7 GHz by blue.</p>
Full article ">Figure A7
<p>The RATAN-600 light curves, the measurements at 22.3 GHz are colored by orange, at 11.2 GHz by magenta, at 8.2 GHz by green, at 4.7 GHz by blue.</p>
Full article ">Figure A8
<p>The RATAN-600 light curves, the measurements at 22.3 GHz are colored by orange, at 11.2 GHz by magenta, at 8.2 GHz by green, at 4.7 GHz by blue.</p>
Full article ">Figure A9
<p>The RATAN-600 light curves, the measurements at 22.3 GHz are colored by orange, at 11.2 GHz by magenta, at 8.2 GHz by green, at 4.7 GHz by blue.</p>
Full article ">Figure A10
<p>The RATAN-600 light curves, the measurements at 22.3 GHz are colored by orange, at 11.2 GHz by magenta, at 8.2 GHz by green, at 4.7 GHz by blue.</p>
Full article ">Figure A11
<p>The radio spectra of PS quasars constructed using the RATAN-600 (green) and the literature data from CATS (grey).</p>
Full article ">Figure A12
<p>The radio spectra of PS quasars constructed using the RATAN-600 (green) and the literature data from CATS (grey).</p>
Full article ">
16 pages, 14681 KiB  
Article
Galaxy Groups as the Ultimate Probe of AGN Feedback
by Dominique Eckert, Fabio Gastaldello, Ewan O’Sullivan, Alexis Finoguenov, Marisa Brienza and the X-GAP Collaboration
Galaxies 2024, 12(3), 24; https://doi.org/10.3390/galaxies12030024 - 13 May 2024
Cited by 3 | Viewed by 887
Abstract
The co-evolution between supermassive black holes and their environment is most directly traced by the hot atmospheres of dark matter halos. The cooling of the hot atmosphere supplies the central regions with fresh gas, igniting active galactic nuclei (AGN) with long duty cycles. [...] Read more.
The co-evolution between supermassive black holes and their environment is most directly traced by the hot atmospheres of dark matter halos. The cooling of the hot atmosphere supplies the central regions with fresh gas, igniting active galactic nuclei (AGN) with long duty cycles. The outflows from the central engine tightly couple with the surrounding gaseous medium and provide the dominant heating source, preventing runaway cooling. Every major modern hydrodynamical simulation suite now includes a prescription for AGN feedback to reproduce the realistic populations of galaxies. However, the mechanisms governing the feeding/feedback cycle between the central black holes and their surrounding galaxies and halos are still poorly understood. Galaxy groups are uniquely suited to constrain the mechanisms governing the cooling–heating balance, as the energy supplied by the central AGN can exceed the gravitational binding energy of halo gas particles. Here, we provide a brief overview of our knowledge of the impact of AGN on the hot atmospheres of galaxy groups, with a specific focus on the thermodynamic profiles of the groups. We then present our on-going efforts to improve on the implementation of AGN feedback in galaxy evolution models by providing precise measurements of the properties of galaxy groups. We introduce the XMM-Newton Group AGN Project (X-GAP), a large program on XMM-Newton targeting a sample of 49 galaxy groups out to R500c. Full article
(This article belongs to the Special Issue Multi-Phase Fueling and Feedback Processes in Jetted AGN)
Show Figures

Figure 1

Figure 1
<p>Impact of AGN feedback on galaxy group atmospheres. The left-panel shows an SDSS <span class="html-italic">i</span>-band image of the brightest group galaxy IC 2476 with X-ray observations of the IGrM superimposed in red. The white contours shows radio emission from the associated radio galaxy B2 0924+30 from LoTSS DR2 data [<a href="#B20-galaxies-12-00024" class="html-bibr">20</a>]. The right-hand panel shows the relation between the binding energy of IGrM gas particles and the energy injection from the central BH (figure reproduced from [<a href="#B17-galaxies-12-00024" class="html-bibr">17</a>]). The dashed line indicates equality between BH and binding energy.</p>
Full article ">Figure 2
<p>Impact of AGN feedback on the thermodynamic profiles of galaxy groups. The figure shows the profiles of IGrM gas density (<b>a</b>) and entropy (<b>b</b>) in four state-of-the-art simulations (EAGLE, Illustris-TNG100, SIMBA, and ROMULUS; [<a href="#B40-galaxies-12-00024" class="html-bibr">40</a>]). Increasing the feedback energy leads to lower gas densities (<b>left</b>) and higher gas entropies (<b>right</b>), highlighting the sensitivity of the IGrM to the feedback scheme. For comparison, the dashed curves show observed galaxy group thermodynamic profiles from <span class="html-italic">Chandra</span> [<a href="#B45-galaxies-12-00024" class="html-bibr">45</a>] and XMM-<span class="html-italic">Newton</span> [<a href="#B46-galaxies-12-00024" class="html-bibr">46</a>].</p>
Full article ">Figure 3
<p>X-GAP group selection cross-matching extended sources detected in RASS with SDSS spectroscopic galaxy groups [<a href="#B56-galaxies-12-00024" class="html-bibr">56</a>]. The plot shows the RASS X-ray luminosity and the corresponding mass estimated through a luminosity–mass relation as a function of the source redshift. The points are color-coded by the apparent size <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>500</mn> </msub> </semantics></math> of <math display="inline"><semantics> <msub> <mi>R</mi> <mrow> <mn>500</mn> <mi>c</mi> </mrow> </msub> </semantics></math>. The selected sample is highlighted by the red stars.</p>
Full article ">Figure 4
<p>Analysis results for four groups matching the X-GAP selection criteria with available observations in the XMM-<span class="html-italic">Newton</span> archive. The selected groups exhibit a flat brightness distribution, as highlighted in the (<b>a</b>), where we show the combined XMM-<span class="html-italic">Newton</span>/EPIC image of one of the systems (SDSSTG 8050 = A1213) with RASS contours overlaid in green. The low surface brightness implies a strong entropy excess (<b>b</b>) extending all the way out to <math display="inline"><semantics> <msub> <mi>R</mi> <mrow> <mn>500</mn> <mi>c</mi> </mrow> </msub> </semantics></math>. The gas is evacuated towards the outskirts, where the gas fraction rises sharply (<b>c</b>). Given our selection process, we are able to determine the gas fraction at <math display="inline"><semantics> <msub> <mi>R</mi> <mrow> <mn>500</mn> <mi>c</mi> </mrow> </msub> </semantics></math> without requiring any extrapolation.</p>
Full article ">Figure A1
<p>Background subtracted, vignetting corrected, and adaptively smoothed <span class="html-italic">XMM-Newton</span>/EPIC maps of X-GAP groups in the [0.7–1.2] keV band. Shown are SDSSTG 828, SDSSTG 885, SDSSTG 1011 (top row), SDSSTG 1162, SDSSTG 1398, SDSSTG 1601 (second row), SDSSTG 1695, SDSSTG 2424, SDSSTG 2620 (third row), SDSSTG 3128, SDSSTG 3460, SDSSTG 3513 (bottom row). The magenta squares show the position of SDSS member galaxies selected using the FoF algorithm [<a href="#B19-galaxies-12-00024" class="html-bibr">19</a>]. In all cases, the axes are right ascension and declination in unit of degrees.</p>
Full article ">Figure A2
<p>Same as <a href="#galaxies-12-00024-f0A1" class="html-fig">Figure A1</a> for SDSSTG 3669, SDSSTG 4047, SDSSTG 4436 (top row), SDSSTG 4654, SDSSTG 4936, SDSSTG 5742 (second row), SDSSTG 6058, SDSSTG 6159, SDSSTG 8050 (third row), SDSSTG 8102, SDSSTG 9178, SDSSTG 9370 (bottom row).</p>
Full article ">Figure A3
<p>Same as <a href="#galaxies-12-00024-f0A1" class="html-fig">Figure A1</a> for SDSSTG 9399, SDSSTG 9647, SDSSTG 9695 (top row), SDSSTG 9771, SDSSTG 10094, SDSSTG 10159 (second row), SDSSTG 10842, SDSSTG 11320, SDSSTG 11631 (third row), SDSSTG 11844, SDSSTG 12349, SDSSTG 15354 (bottom row).</p>
Full article ">Figure A4
<p>Same as <a href="#galaxies-12-00024-f0A1" class="html-fig">Figure A1</a> for SDSSTG 15641, SDSSTG 15776, SDSSTG 16150 (top row), SDSSTG 16393, SDSSTG 21128, SDSSTG 22635 (second row), SDSSTG 24595, SDSSTG 28674, SDSSTG 35976 (third row), SDSSTG 39344, SDSSTG 40241, SDSSTG 46701 (bottom row).</p>
Full article ">
9 pages, 10540 KiB  
Article
Doppler Tomography of the Circumstellar Disk of the Be Star κ Draconis
by Ilfa A. Gabitova, Anatoly S. Miroshnichenko, Sergey V. Zharikov, Ainash Amantayeva and Serik A. Khokhlov
Galaxies 2024, 12(3), 23; https://doi.org/10.3390/galaxies12030023 - 7 May 2024
Viewed by 781
Abstract
κ Draconis is a binary system with a classical Be star as the primary component. Its emission-line spectrum consists of hydrogen lines, notably the Hα line with peak intensity ratio (V/R) variations phase-locked with the orbital period P = 61.55 days. Among [...] Read more.
κ Draconis is a binary system with a classical Be star as the primary component. Its emission-line spectrum consists of hydrogen lines, notably the Hα line with peak intensity ratio (V/R) variations phase-locked with the orbital period P = 61.55 days. Among binaries demonstrating the Be phenomenon, κ Dra stands out as one of a few systems with a discernible mass of its secondary component. Based on more than 200 spectra obtained in 2014–2023, we verified the physical parameters and constructed the mass function. We used part of these data obtained in 2014–2021 to investigate regions in the circumstellar disk of the primary component that emit the Hα line using the Doppler tomography method. The results show that the disk has a non-uniform density distribution with a prominent enhancement at Vy ≈ 99 km s1 and Vx6 km s1 that corresponds to a cloud-like source of the double-peaked Hα line profile. We argue that this enhancement’s motion is responsible for the periodic variations in the Hα V/R ratio, which is synchronised in orbital phase with the radial velocity (RV) of absorption lines from the atmosphere of the primary component. Full article
Show Figures

Figure 1

Figure 1
<p>Examples of the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line profile in the spectra of <math display="inline"><semantics> <mi>κ</mi> </semantics></math> Dra. The observational dates are shown in each panel. For more on the choice of the zero-phase epoch, see <a href="#sec3-galaxies-12-00023" class="html-sec">Section 3</a>.</p>
Full article ">Figure 2
<p>(<b>Top</b>): V/R variations folded with the orbital period 61.55 days. (<b>Bottom</b>): Temporal variations in the equivalent width (EW) of the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line. The colors correspond to observation dates, as shown on the bar on the right.</p>
Full article ">Figure 3
<p>(<b>Left panels</b>): Average H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line profiles. Black solid line shows the original profile, dashed line is a model spectrum of the primary component’s atmosphere, and solid red line is the result of subtraction of the model profile from the original one. (<b>Middle panels</b>): Reconstructed trailed spectra of the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line folded with the orbital period P = 61.55 days. (<b>Right panels</b>): Doppler maps of the system. The map is centered at the system’s center of mass, which is marked by the cross. The plus sign marks the center of mass of the primary component at <math display="inline"><semantics> <msub> <mi mathvariant="normal">V</mi> <mi>y</mi> </msub> </semantics></math><math display="inline"><semantics> <mrow> <mo>≈</mo> <mo>−</mo> <mn>7</mn> </mrow> </semantics></math> km <math display="inline"><semantics> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="normal">V</mi> <mi>x</mi> </msub> </semantics></math> = 0 km <math display="inline"><semantics> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. The center of mass of the secondary component is marked with another plus sign at <math display="inline"><semantics> <msub> <mi mathvariant="normal">V</mi> <mi>y</mi> </msub> </semantics></math> ≈ 50 km <math display="inline"><semantics> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="normal">V</mi> <mi>x</mi> </msub> </semantics></math> = 0 km <math display="inline"><semantics> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> with the Roche lobe plotted around it. The dashed line marks <math display="inline"><semantics> <mrow> <mi>v</mi> <mspace width="0.166667em"/> <mo form="prefix">sin</mo> <mi>i</mi> </mrow> </semantics></math> = 85 km <math display="inline"><semantics> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> that corresponds to the tidal truncation disk radius. The color of the Doppler maps corresponds to arbitrary units of emission intensity (the yellow–blue–red–black palette corresponds to a change from low to high intensity).</p>
Full article ">Figure 4
<p>The brightness distribution in the disk, transformed from the Doppler map of the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> emission line to the XY plane of the system. The colorbar shows normalized relative intensity. The Be star is located at the origin.</p>
Full article ">
27 pages, 21113 KiB  
Article
CHANG-ES XXXI—A Decade of CHANG-ES: What We Have Learned from Radio Observations of Edge-on Galaxies
by Judith Irwin, Rainer Beck, Tanden Cook, Ralf-Jürgen Dettmar, Jayanne English, Volker Heesen, Richard Henriksen, Yan Jiang, Jiang-Tao Li, Li-Yuan Lu, Crystal Mele, Ancla Müller, Eric Murphy, Troy Porter, Richard Rand, Nathan Skeggs, Michael Stein, Yelena Stein, Jeroen Stil, Andrew Strong, Rene Walterbos, Q. Daniel Wang, Theresa Wiegert and Yang Yangadd Show full author list remove Hide full author list
Galaxies 2024, 12(3), 22; https://doi.org/10.3390/galaxies12030022 - 6 May 2024
Viewed by 1211
Abstract
CHANG-ES (Continuum Halos in Nearby Galaxies—an EVLA Survey) is an ambitious project to target 35 nearby disk galaxies that are edge-on to the line of sight. The orientation permits both the disk and halo regions to be studied. The observations were initially at [...] Read more.
CHANG-ES (Continuum Halos in Nearby Galaxies—an EVLA Survey) is an ambitious project to target 35 nearby disk galaxies that are edge-on to the line of sight. The orientation permits both the disk and halo regions to be studied. The observations were initially at 1.5 GHz (L-band) and 6.0 GHz (C-band) in a variety of VLA array configurations, and in all four Stokes parameters, which allowed for spatially resolved images in total intensity plus polarization. The inclusion of polarization is unique to an edge-on galaxy survey and reveals the galaxies’ halo magnetic fields. This paper will summarize the results to date, some of which are new phenomena, never seen prior to CHANG-ES. For example, we see that ‘X-type’ fields, as well as rotation measure reversals, are common features of spiral galaxies. Further observations at 3.0 GHz (S-band) as well as future scientific opportunities will also be described. Full article
(This article belongs to the Special Issue The 10th Anniversary of Galaxies: New Perspectives on Radio Surveys)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>See <a href="https://public.nrao.edu/news/2020-image-contest-winners" target="_blank">https://public.nrao.edu/news/2020-image-contest-winners</a> accessed on 4 May 2024 and [<a href="#B8-galaxies-12-00022" class="html-bibr">8</a>] for these images which won Second Prize (left) and Honorable Mention (right) in the NRAO image contest held as part of celebrating the 40th anniversary of the VLA. On the left, the magnetic field of NGC 5775 is highlighted in superb detail using the CHANG-ES D-array and C-band linear polarization data. Displayed in fuchsia near the disk and in blue farther out, the curvy field lines are highlighted via an LIC algorithm (<a href="#sec2dot2dot3-galaxies-12-00022" class="html-sec">Section 2.2.3</a>) and extend as high as 8 kpc from the plane. Superimposed is an optical image constructed from the Hubble Space Telescope data of the galaxy, showing hot ionized gas in rose and broad-band optical in pale yellow. On the right, the magnetic field of NGC 4666 is illustrated using the same LIC algorithm as on the left. Image by Y. Stein with support from J. English. See also [<a href="#B9-galaxies-12-00022" class="html-bibr">9</a>]. See Figure 8 for versions with coordinates.</p>
Full article ">Figure 2
<p>Visual representation of the CHANG-ES frequency coverage. The red and blue bands represent the L-band and C-band coverage of the original observations (see [<a href="#B1-galaxies-12-00022" class="html-bibr">1</a>] for a tabular version), while green represents newer S-band data discussed in <a href="#sec4dot1-galaxies-12-00022" class="html-sec">Section 4.1</a>. The gaps in L-band and S-band are because of RFI. The gap from 4 to 5 GHz is because the lowest available C-band frequency was 5 GHz during CHANG-ES observations. Shading is for aesthetics and does not indicate any technical differences in the band.</p>
Full article ">Figure 3
<p>Spectral index color map: this triple divergent color scheme has 3 boundaries. Boundary 1 is assigned cyan and the increasingly positive values of <math display="inline"><semantics> <mi>α</mi> </semantics></math> have sequential colors on a standard color wheel. Cyan’s complementary color, orange, marks divergent point 2, the upper boundary for “steep” spectral values, and colors of more negative <math display="inline"><semantics> <mi>α</mi> </semantics></math> are also sequential. The end points of the color bar are yellow and purple which are also complementary colors, which work with human color perception. The gray-ish divergent point 3 is created by mixing the orange and cyan in steps to create tones of intermediate color. This example demonstrates that a flat spectrum central AGN can be discerned almost at a glance with this color scheme. Tutorials for this, along with a single divergent point error map, are found at <a href="https://github.com/mlarichardson/CosmosCanvas" target="_blank">https://github.com/mlarichardson/CosmosCanvas</a> (accessed on 4 May 2024).</p>
Full article ">Figure 4
<p>On the left, polarization contours plus B-field orientations over a grayscale optical SDSS image of NGC 5792, from [<a href="#B39-galaxies-12-00022" class="html-bibr">39</a>]. On the right, L-band total intensity contours over the B/L to C/C spectral index map of NGC 3628 [<a href="#B24-galaxies-12-00022" class="html-bibr">24</a>].</p>
Full article ">Figure 5
<p>Sample images from [<a href="#B31-galaxies-12-00022" class="html-bibr">31</a>]. On the left, an optical image of NGC 4438 is shown in colour. Two compact sources are specified in fuchsia. The central source is the AGN, and the southern source is a new cross-matched radio/X-ray source. On the right, an optical Milky Way binary star system is shown in colour. These stars are near the Sombrero galaxy and have Gaia designations 530134264552397440 (right star) and 3530134264552424960. The blue and red dots and error ellipses show the positions of the L-band and C-band sources, respectively. The black dots with error circles denote the two X-ray sources. The scale showing 1 arcsec is displayed on the lower left.</p>
Full article ">Figure 6
<p>On the left, the weighted mean linear polarization (colors) and matching magnetic field orientations from the stacking of 28 CHANG-ES galaxies, showing an underlying X-shaped field, from [<a href="#B29-galaxies-12-00022" class="html-bibr">29</a>]. The RA and Decl. axes have arbitrary centering. On the right, Comparison between C-band (red) and L-band (white) magnetic field (B) orientations at the same resolution in the galaxy NGC 2683, superimposed on an S-band (<a href="#sec4-galaxies-12-00022" class="html-sec">Section 4</a>) polarization image. The galaxy’s major axis extends from the north-east to the south-west. It should be noted how much the B field orientation has rotated between C-band and L-band. From [<a href="#B56-galaxies-12-00022" class="html-bibr">56</a>].</p>
Full article ">Figure 7
<p>NGC 3044 C-band polarized emission, RM-corrected, from [<a href="#B29-galaxies-12-00022" class="html-bibr">29</a>]. Contours denote the total intensity and show that the major axis extends from the north-west (advancing side) to the south-east (receding side). On the left, the polarized intensity (color) with magnetic field orientations superimposed. On the right, the rotation measure map (color), with magnetic orientations superimposed (see <a href="#sec3dot4dot3-galaxies-12-00022" class="html-sec">Section 3.4.3</a>).</p>
Full article ">Figure 8
<p>Percentage of polarization in NGC 5775 (<b>left</b>) and NGC 4666 (<b>right</b>) in color. Contours represent the total intensity emission and the beam size is displayed as a filled black circle in a lower corner. From [<a href="#B29-galaxies-12-00022" class="html-bibr">29</a>]. Attractive images of these galaxies are shown in <a href="#galaxies-12-00022-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 9
<p>Magnetic field orientations of NGC 4631 Faraday rotations measures (from [<a href="#B65-galaxies-12-00022" class="html-bibr">65</a>]) with RM signs plotted in color (blue: negative, green: positive). Position angles are traced using an LIC algorithm as in <a href="#galaxies-12-00022-f001" class="html-fig">Figure 1</a>. The background optical image uses data from the Mayall 4-m telescope, collected by Maria Patterson and Rene Walterbos of New Mexico State University. Image composition by Jayanne English, University of Manitoba.</p>
Full article ">Figure 10
<p>In the left image, we see RM patterns of NGC 4217. The RM image (upper right) shows two boxes which highlight a feature extending almost 7 kpc into the halo (left box) and two bubble-like features (right box). Models showing a helical field (top left, labelled <b>a</b>) and circumferential field (bottom right, labelled <b>c</b> and <b>d</b>) correspond to the two boxes, respectively. A possible orientation of the in-disk field is shown on the lower left (labelled <b>b</b>). From [<a href="#B19-galaxies-12-00022" class="html-bibr">19</a>]. In the right image, the galaxy NGC 4388 shows complex structure in polarization intensity (PI). Lower inset: Total intensity emission; the marked difference compared to PI should be noted. From [<a href="#B66-galaxies-12-00022" class="html-bibr">66</a>].</p>
Full article ">Figure 11
<p>One model from [<a href="#B74-galaxies-12-00022" class="html-bibr">74</a>] from a fit to NGC 4631, which includes outflow. On the left, reversing RMs are seen from an edge-on perspective. On the right, magnetic spirals continue from the disk to the halo. One slice at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> is shown from a face-on perspective.</p>
Full article ">Figure 12
<p>NGC 3044 images that include S-band data. On the left, <span class="html-italic">XMM-Newton</span> 0.3–1.2 keV intensity contours superimposed on the VLA S-band total intensity image in color. On the right, magnetic field orientations from combined S-band and C-band data superimposed on an H<math display="inline"><semantics> <mi>α</mi> </semantics></math> image of NGC 3044, from [<a href="#B56-galaxies-12-00022" class="html-bibr">56</a>].</p>
Full article ">Figure 13
<p>On the left, (single frame) molecular gas mass measured with the IRAM 30m observations from the CO-CHANGES project vs. atomic gas mass collected from the archive [<a href="#B81-galaxies-12-00022" class="html-bibr">81</a>]. Best-fit relations from different literature sources are plotted in colored lines for comparison, with the black solid line the best-fit of the CO-CHANGES sample. On the right (three frames labelled (<b>a</b>–<b>c</b>)), a pilot study of NGC 3556 from the eDIG-CHANGES project [<a href="#B83-galaxies-12-00022" class="html-bibr">83</a>]: (<b>a</b>) The multi-slit mask overlaid on a snapshot H<math display="inline"><semantics> <mi>α</mi> </semantics></math> image of NGC 3556. Each slit produces a narrow-band spectrum covering the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line and [N II]<math display="inline"><semantics> <mrow> <mi>λ</mi> <mi>λ</mi> <mn>6548</mn> <mo>,</mo> <mn>6583</mn> </mrow> </semantics></math> doublet. (<b>b</b>) The line centroid velocity map showing clear rotating disk structures. (<b>c</b>) PV diagram extracted from the midplane of the galaxy. The three parallel bright bands separated in the velocity space are the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line (middle) and the two [N II] lines. The contours are the CO <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>=</mo> <mn>1</mn> <mo>−</mo> <mn>0</mn> </mrow> </semantics></math> PV diagram constructed with our IRAM 30 m data [<a href="#B81-galaxies-12-00022" class="html-bibr">81</a>]. PV diagrams of the two gas phases have comparable slopes, indicating that they follow similar dynamics.</p>
Full article ">
18 pages, 2825 KiB  
Article
Flattened Galaxy Rotation Curves in the Exochronous Metric
by Robin Booth
Galaxies 2024, 12(3), 21; https://doi.org/10.3390/galaxies12030021 - 24 Apr 2024
Viewed by 968
Abstract
We examine some of the consequences of the Exochronous (timeless) metric and the associated ΣGR cosmological model for the formation of galaxies, and, in particular, their characteristic rotation curves. We show how the cumulative curvature from the multiple spatial hypersurfaces in this [...] Read more.
We examine some of the consequences of the Exochronous (timeless) metric and the associated ΣGR cosmological model for the formation of galaxies, and, in particular, their characteristic rotation curves. We show how the cumulative curvature from the multiple spatial hypersurfaces in this model leads to a modified version of the Poisson equation, in which the gravitational potential is computed over 4D space. Using this new form of the Poisson equation, we derive an analytic expression for gravitational potential as a function of radial distance for a uniform gas cloud undergoing gravitational collapse. We show that this results in a radial velocity profile that provides an excellent fit with commonly observed galaxy rotation curves, and hence fully accounts for the effects previously ascribed to dark matter. An expression can be derived for the equivalent matter density profile corresponding to the ΣGR gravitational potential, from which it is evident that this is very similar in form to the well-known Navarro–Frenk–White profile. As a further illustration of the consequences of adopting the Exochronous metric, we show how the principle can readily be incorporated into particle-mesh N-body simulations of large-scale structure evolution, using a relaxation solver for the solution to the Poisson equation and the evolution of the gravitational potential. Examples of the use of this simulation model are shown for the following cases: (a) the initial evolution of a large-scale structure, and (b) galaxy formation from a gravitationally collapsing gas cloud. In both cases, it is possible to directly visualise the build-up of the gravitational potential in 3D space as the simulation evolves and note how this corresponds to what is currently assumed to be dark matter. Full article
Show Figures

Figure 1

Figure 1
<p>Galaxy rotation curves from observations of 20 galaxies reported in [<a href="#B7-galaxies-12-00021" class="html-bibr">7</a>].</p>
Full article ">Figure 2
<p>Galaxy rotation curves. The solid blue line shows the flat galaxy rotation curve resulting from the potentials contributed by baryonic matter in the galaxy core and disk and the dark matter in the galaxy’s halo. The dashed orange line is the rotation curve that would be observed in the absence of dark matter.</p>
Full article ">Figure 3
<p>NFW density profile.</p>
Full article ">Figure 4
<p>Metric foliations.</p>
Full article ">Figure 5
<p>The Exochronous metric. (<b>a</b>) Standard spacetime. (<b>b</b>) Exochronous superspace. (<b>c</b>) Superspace projection onto 3D space.</p>
Full article ">Figure 6
<p>Collapsing gas cloud galaxy model. The galaxy starts with an initial radius <math display="inline"><semantics> <msub> <mi>R</mi> <mn>0</mn> </msub> </semantics></math> and density <math display="inline"><semantics> <msub> <mi>ρ</mi> <mn>0</mn> </msub> </semantics></math>, and collapses down under gravity to a final core radius, <math display="inline"><semantics> <msub> <mi>R</mi> <mi>c</mi> </msub> </semantics></math>, depicted by the yellow circle. The dashed blue circles depict a spherical shell of gas at radius <span class="html-italic">r</span> from the galactic centre, with a thickness <math display="inline"><semantics> <mrow> <mi mathvariant="normal">d</mi> <mi>r</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Evolution of the <math display="inline"><semantics> <mo>Σ</mo> </semantics></math>GR potential over a range of galaxy core radii (<math display="inline"><semantics> <msub> <mi>R</mi> <mi>c</mi> </msub> </semantics></math>) as a function of the fraction of initial radius (<math display="inline"><semantics> <mrow> <mi>r</mi> <mo>/</mo> <msub> <mi>R</mi> <mn>0</mn> </msub> </mrow> </semantics></math>). The dashed lines plot the gravitational potential due solely to the galaxy core. The solid curves show the total potential, including the galaxy halo.</p>
Full article ">Figure 8
<p>Comparison of the <math display="inline"><semantics> <mo>Σ</mo> </semantics></math>GR effective density profile with the NFW profile.</p>
Full article ">Figure 9
<p>Cumulative effective halo mass as a function of radial distance from centre of galaxy.</p>
Full article ">Figure 10
<p>Galaxy rotation curves.</p>
Full article ">Figure 11
<p>Rotation curves for selected galaxies in the Ursa Major cluster. The figures for <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>R</mi> <mn>0</mn> </msub> <mo>,</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mi>c</mi> </msub> </semantics></math> in the boxes are the best fit parameters for the <math display="inline"><semantics> <mo>Σ</mo> </semantics></math>GR model, and are also summarised in <a href="#galaxies-12-00021-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 12
<p>Snapshot from N-body simulation, showing typical “Cosmic web” structure, including filaments, clusters, and voids.</p>
Full article ">Figure 13
<p>A contour plot of the gravitational potential for a 2D slice though the simulation volume.</p>
Full article ">Figure 14
<p>Galaxy formation. (<b>a</b>) Elliptical galaxy from collapsing gas cloud. (<b>b</b>) Gravitational ‘halo’ surrounding galaxy. (<b>c</b>) Three-dimensional gravitational potential. (<b>d</b>) Halo plus rotation curve overlay.</p>
Full article ">
16 pages, 1306 KiB  
Review
Investigating the Properties of the Relativistic Jet and Hot Corona in AGN with X-ray Polarimetry
by Dawoon E. Kim, Laura Di Gesu, Frédéric Marin, Alan P. Marscher, Giorgio Matt, Paolo Soffitta, Francesco Tombesi, Enrico Costa and Immacolata Donnarumma
Galaxies 2024, 12(3), 20; https://doi.org/10.3390/galaxies12030020 - 23 Apr 2024
Viewed by 1016
Abstract
X-ray polarimetry has been suggested as a prominent tool for investigating the geometrical and physical properties of the emissions from active galactic nuclei (AGN). The successful launch of the Imaging X-ray Polarimetry Explorer (IXPE) on 9 December 2021 has expanded the previously restricted [...] Read more.
X-ray polarimetry has been suggested as a prominent tool for investigating the geometrical and physical properties of the emissions from active galactic nuclei (AGN). The successful launch of the Imaging X-ray Polarimetry Explorer (IXPE) on 9 December 2021 has expanded the previously restricted scope of polarimetry into the X-ray domain, enabling X-ray polarimetric studies of AGN. Over a span of two years, IXPE has observed various AGN populations, including blazars and radio-quiet AGN. In this paper, we summarize the remarkable discoveries achieved thanks to the opening of the new window of X-ray polarimetry of AGN through IXPE observations. We will delve into two primary areas of interest: first, the magnetic field geometry and particle acceleration mechanisms in the jets of radio-loud AGN, such as blazars, where the relativistic acceleration process dominates the spectral energy distribution; and second, the geometry of the hot corona in radio-quiet AGN. Thus far, the IXPE results from blazars favor the energy-stratified shock acceleration model, and they provide evidence of helical magnetic fields inside the jet. Concerning the corona geometry, the IXPE results are consistent with a disk-originated slab-like or wedge-like shape, as could result from Comptonization around the accretion disk. Full article
(This article belongs to the Special Issue Multi-Phase Fueling and Feedback Processes in Jetted AGN)
Show Figures

Figure 1

Figure 1
<p>A schematic representation of SED for RLAGN (HSP: red solid line, and LSP: gray dotted line) and RQAGN (black solid line with labeled colored lines for different emission components). The yellow shaded area indicates the X-ray regime. The figure is reproduced from <a href="#galaxies-12-00020-f001" class="html-fig">Figure 1</a> in [<a href="#B8-galaxies-12-00020" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>IXPE observation results of Mrk 501 and Mrk 421. (<b>Left</b>) Multiwavelength polarization degree of Mrk 501 from radio rays to X-rays. The black symbols represent observations conducted between March 8th and 10th, while the red symbols represent observations from March 26th to 28th. The open symbols indicate the intrinsic optical polarization degree corrected for the host galaxy. The figure was reproduced from <a href="#galaxies-12-00020-f003" class="html-fig">Figure 3</a> in [<a href="#B57-galaxies-12-00020" class="html-bibr">57</a>]. (<b>Right</b>) X-ray polarization angle rotation in Mrk 421. The symbols identify multiwavelength polarimetry measurements obtained from telescopes as labeled. The figure was reproduced from <a href="#galaxies-12-00020-f002" class="html-fig">Figure 2</a> in [<a href="#B59-galaxies-12-00020" class="html-bibr">59</a>].</p>
Full article ">Figure 3
<p>Physical properties of energy-stratified shock emissions from HSP blazars. (<b>Left</b>) X-ray polarization vs. the <math display="inline"><semantics> <msub> <mi mathvariant="normal">Π</mi> <mi mathvariant="normal">X</mi> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi mathvariant="normal">Π</mi> <mi mathvariant="normal">O</mi> </msub> </semantics></math> ratio; (<b>Right</b>) X-ray power law photon index vs. the <math display="inline"><semantics> <msub> <mi mathvariant="normal">Π</mi> <mi mathvariant="normal">X</mi> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi mathvariant="normal">Π</mi> <mi mathvariant="normal">O</mi> </msub> </semantics></math> ratio of HSP blazars observed by IXPE. Each colored point indicates different IXPE observations labeled in the legend.</p>
Full article ">Figure 4
<p>Schematic view of different X-ray corona geometries: slab, conical, and lamppost. The bluish and reddish areas represent the corona and disk emission regions, respectively. Blue and orange arrows indicate upscattered X-ray and UV/optical disk photons, respectively. The black arrow represents the predicted polarization properties (degree in number and angle in direction of the arrow), estimated from [<a href="#B77-galaxies-12-00020" class="html-bibr">77</a>].</p>
Full article ">Figure 5
<p>IXPE observation result of NGC 4151. (<b>Left</b>) Polarization contours (<math display="inline"><semantics> <mrow> <mn>68</mn> <mo>%</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>90</mn> <mo>%</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mn>99</mn> <mo>%</mo> </mrow> </semantics></math> detection significance levels); (<b>middle</b>) X-ray spectrum analysis result; (<b>right</b>) comparison of IXPE measurements and expected X-ray polarization in slab and wedge geometry coronae calculated from MONK simulations. The figure was reproduced with kind permission from Oxford University Press and the Royal Astronomical Society from [<a href="#B78-galaxies-12-00020" class="html-bibr">78</a>].</p>
Full article ">Figure 6
<p>X-ray polarization contours of the integrated analysis of MCG-05-23-16 I and II (red; <math display="inline"><semantics> <mrow> <mn>68</mn> <mo>%</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>90</mn> <mo>%</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mn>99</mn> <mo>%</mo> </mrow> </semantics></math> detection significance levels). The black dashed line indicates the direction of the NLR, while the dotted line denotes the perpendicular direction, implying an angle parallel to the disk. Each colored area represents the expected polarization properties calculated by MONK, with relatively saturated regions indicating the expected degree of polarizaton for inclinations in the range of approximately 30–50°. The figure was reproduced with kind permission from Oxford University Press and the Royal Astronomical Society from Figure 8 in [<a href="#B80-galaxies-12-00020" class="html-bibr">80</a>].</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop