Nothing Special   »   [go: up one dir, main page]

Next Issue
Volume 8, June
Previous Issue
Volume 7, December
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

Galaxies, Volume 8, Issue 1 (March 2020) – 27 articles

Cover Story (view full-size image): ΛCDM-predicted dark matter (DM) subhalos, not massive enough to retain baryons and become visible, could be detected in gamma rays from DM annihilation as unidentified sources (unIDs), provided DM is made of weakly interacting massive particles (WIMPs). Using the High Altitude Water Cherenkov (HAWC) observatory, we search for DM subhalo candidates within the detected unIDs. Only one, 2HWC J1040+308, is found to be a promising source. Lower-energy gamma-ray instruments such as Fermi-LAT or VERITAS do not detect the candidate, as one would expect from a DM perspective. This unID is also spatially extended, which is a “smoking gun” in DM subhalo searches. Finally, constraints on the annihilation cross-section are set by comparing this source to expectations based on state-of-the-art N-body cosmological simulations of the Galactic subhalo population. View this paper
  • 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:
28 pages, 1526 KiB  
Review
The UV Perspective of Low-Mass Star Formation
by P. Christian Schneider, H. Moritz Günther and Kevin France
Galaxies 2020, 8(1), 27; https://doi.org/10.3390/galaxies8010027 - 21 Mar 2020
Cited by 14 | Viewed by 3768
Abstract
The formation of low-mass ( M 2 M ) stars in molecular clouds involves accretion disks and jets, which are of broad astrophysical interest. Accreting stars represent the closest examples of these phenomena. Star and planet formation are also intimately [...] Read more.
The formation of low-mass ( M 2 M ) stars in molecular clouds involves accretion disks and jets, which are of broad astrophysical interest. Accreting stars represent the closest examples of these phenomena. Star and planet formation are also intimately connected, setting the starting point for planetary systems like our own. The ultraviolet (UV) spectral range is particularly suited for studying star formation, because virtually all relevant processes radiate at temperatures associated with UV emission processes or have strong observational signatures in the UV range. In this review, we describe how UV observations provide unique diagnostics for the accretion process, the physical properties of the protoplanetary disk, and jets and outflows. Full article
(This article belongs to the Special Issue Star Formation in the Ultraviolet)
Show Figures

Figure 1

Figure 1
<p>IUE spectrum of RU Lup (flux unit: <math display="inline"><semantics> <mrow> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>13</mn> </mrow> </msup> <mspace width="0.166667em"/> </mrow> </semantics></math>erg cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math> s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> Å<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>) with the main features labeled, one of the first high-fidelity FUV spectra of a classical T Tauri star (CTTS). From Gahm et al. [<a href="#B38-galaxies-08-00027" class="html-bibr">38</a>]. Reproduced with permission © ESO.</p>
Full article ">Figure 2
<p>Sketch of a CTTS system with regions that are relevant for FUV studies labeled. Relative positions are correct, but relative distances and sizes are not. The relevant components are (<b>a</b>) the pre-main-sequence star, (<b>b</b>) the accretion shock on the stellar surface, (<b>c</b>) the accretion funnel, (<b>d</b>) the inner, hot disk region (<math display="inline"><semantics> <mrow> <mi>T</mi> <mo>∼</mo> <mn>2</mn> <mo>,</mo> <mn>000</mn> </mrow> </semantics></math> K), (<b>e</b>) cooler regions of the inner disk surface (<math display="inline"><semantics> <mrow> <mi>T</mi> <mo>∼</mo> <mn>500</mn> </mrow> </semantics></math> K), (<b>f</b>) cool protoplanetary disk material, (<b>g</b>) a spherical (stellar) wind, which may be seen in emission or as an absorption component in strong emission lines (e.g., C <span class="html-small-caps">iv</span>), (<b>h</b>) an X-wind, and (<b>i</b>) a disk wind. <a href="https://www.overleaf.com/project/5e42c54e77c525000110cf18" target="_blank">https://www.overleaf.com/project/5e42c54e77c525000110cf18</a>.</p>
Full article ">Figure 3
<p>Comparison of a typical CTTS’s line profile (<b>top</b>), here C <span class="html-small-caps">iv</span>, and that of a weak-line T Tauri star (WTTS) of comparable mass (<b>bottom</b>). The dashed line marks the location of emission from fluorescent H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math>. From Ardila et al. [<a href="#B50-galaxies-08-00027" class="html-bibr">50</a>]. ©AAS. Reproduced with permission.</p>
Full article ">Figure 4
<p><b>Left</b>: Correlation between extinction-corrected C <span class="html-small-caps">iv</span> and accretion luminosity. This suggests that the C <span class="html-small-caps">iv</span> in CTTSs is mostly due to the accretion process. <b>Middle</b>: Correlation between C <span class="html-small-caps">iv</span> and total UV luminosity. <b>Right</b>: Relation between line fluxes. Data and fits from Yang et al. [<a href="#B62-galaxies-08-00027" class="html-bibr">62</a>].</p>
Full article ">Figure 5
<p>Schematic sketch of scenarios proposed for the structure of accretion columns. The stellar surface is at the bottom and material flows along a magnetic field line onto the stellar surface (indicated by the red arrow). The accretion shock (blue) forms on the stellar surface. Density is shown as gray scale. From left to right: (1) One homogeneous column with one density and a single infall velocity; (2) one column with a density stratification (high density in the center of the column indicated by the black region, lower density in the outer region indicated by the grayish color) and one infall velocity; (3) multiple columns that are individually homogeneous, i.e., have different densities but are otherwise equal; (4) one “column” is decomposed into individual blobs that are, individually, homogeneous in density.</p>
Full article ">Figure 6
<p><b>Left</b>: Spectral region of 1430–1470 Å for three prototypical gas-rich disks. All of the strong spectral features in this bandpass are emission lines from Ly<math display="inline"><semantics> <mi>α</mi> </semantics></math>-pumped fluorescent H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> (UV-H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math>; marked with blue diamonds and several bright features are labeled). From France et al. [<a href="#B79-galaxies-08-00027" class="html-bibr">79</a>]. <b>Right</b>: Kinematic model for the H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> emission including the effect of the COS line spread function (LSF). From Arulanantham et al. [<a href="#B84-galaxies-08-00027" class="html-bibr">84</a>]. Both figures: ©AAS. Reproduced with permission.</p>
Full article ">Figure 7
<p><b>Left</b>: Fourth positive band emission of CO from Schindhelm et al. [<a href="#B31-galaxies-08-00027" class="html-bibr">31</a>]. <b>Right</b>: CO absorption in the CTTS HN Tau. Data shown in black, fit of the CO transmission spectrum in red, and the green line indicates the bandheads (1 FEFU<math display="inline"><semantics> <mrow> <mspace width="3.33333pt"/> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </mrow> </semantics></math> ergs cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math> s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> Å<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>). From France et al. [<a href="#B32-galaxies-08-00027" class="html-bibr">32</a>]. Both figures: ©AAS. Reproduced with permission.</p>
Full article ">Figure 8
<p><b>Left</b>: FUV to optical spectrum of the nearest CTTS, TW Hya, after subtraction of the WTTS template spectrum of V819 Tau. The dashed line indicates the expected emission from the accretion model, which approximates the NUV continuum of TW Hya but significantly underestimates the FUV continuum, especially around 1600 Å. From Herczeg et al. [<a href="#B80-galaxies-08-00027" class="html-bibr">80</a>]. <b>Right</b>: Typical 1600 Å-Bump spectrum, here for CS Cha. The blue-dashed line indicates the polynomial model used to subtract the accretion continuum. From France et al. [<a href="#B36-galaxies-08-00027" class="html-bibr">36</a>]. Both figures: ©AAS. Reproduced with permission.</p>
Full article ">Figure 9
<p>Example of a Ly<math display="inline"><semantics> <mi>α</mi> </semantics></math> reconstruction based on H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> fluxes. Diamonds indicate the wavelengths at which measured H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> progressions are pumped, the dashed line indicates an H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> model for T = 2500 K and <math display="inline"><semantics> <mrow> <mo form="prefix">log</mo> <mi>N</mi> <mo>(</mo> <msub> <mi>H</mi> <mn>2</mn> </msub> <mo>)</mo> <mo>=</mo> <mn>18.5</mn> </mrow> </semantics></math> and a surface filling factor for the H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> material, as seen by the Ly<math display="inline"><semantics> <mi>α</mi> </semantics></math> emitting region. In this model, the intrinsic Ly<math display="inline"><semantics> <mi>α</mi> </semantics></math> line is Gaussian-shaped (dotted), the H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> gas sees the dashed profile, and the Ly<math display="inline"><semantics> <mi>α</mi> </semantics></math> profile observed by HST is subject to additional interstellar absorption (solid line). From Herczeg et al. [<a href="#B80-galaxies-08-00027" class="html-bibr">80</a>]. ©AAS. Reproduced with permission.</p>
Full article ">Figure 10
<p>Comparison of full and transitional disks in their reconstructed radial distributions of the observed H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> emission. From Hoadley et al. [<a href="#B81-galaxies-08-00027" class="html-bibr">81</a>]. ©AAS. Reproduced with permission.</p>
Full article ">Figure 11
<p><b>Left</b>: Position–velocity diagram for the C <span class="html-small-caps">iv</span> emission of the DG Tau jet. The red contours indicate emission components seen in traditional tracers like [O <span class="html-small-caps">i</span>]. Note that there is no emission at the stellar position, and all C <span class="html-small-caps">iv</span> emission originates within the jet. From Schneider et al. [<a href="#B33-galaxies-08-00027" class="html-bibr">33</a>]. <b>Right</b>: Image of UV-H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math> of DG Tau. Reproduced with permission © ESO.</p>
Full article ">Figure 12
<p><b>Top</b>: FUV continuum spectra of AA Tau during bright and dim states. <b>Bottom</b>: Ratio between bright and dim states with the evolution of emission lines shown in blue/red diamonds for the blue and red line wing, respectively. Notable is the bluer appearance of the continuum during the dim state, which is incompatible with just extra absorption and more compatible with scattering. From Schneider et al. [<a href="#B87-galaxies-08-00027" class="html-bibr">87</a>]. Reproduced with permission © ESO.</p>
Full article ">
11 pages, 1384 KiB  
Article
Shaping Planetary Nebulae with Jets and the Grazing Envelope Evolution
by Noam Soker
Galaxies 2020, 8(1), 26; https://doi.org/10.3390/galaxies8010026 - 18 Mar 2020
Cited by 14 | Viewed by 2452
Abstract
I argue that the high percentage of planetary nebulae (PNe) that are shaped by jets show that main sequence stars in binary systems can accrete mass at a high rate from an accretion disk and launch jets. Not only does this allow jets [...] Read more.
I argue that the high percentage of planetary nebulae (PNe) that are shaped by jets show that main sequence stars in binary systems can accrete mass at a high rate from an accretion disk and launch jets. Not only does this allow jets to shape PNe, but this also points to the importance of jets in other types of binary systems and in other processes. These processes include the grazing envelope evolution (GEE), the common envelope evolution (CEE), and the efficient conversion of kinetic energy to radiation in outflows. Additionally, the jets point to the possibility that many systems launch jets as they enter the CEE, possibly through a GEE phase. The other binary systems in which jets might play significant roles include intermediate-luminosity optical transients (ILOTs), supernova impostors (including pre-explosion outbursts), post-CEE binary systems, post-GEE binary systems, and progenitors of neutron star binary systems and black hole binary systems. One of the immediate consequences is that the outflow of these systems is highly-non-spherical, including bipolar lobes, jets, and rings. Full article
(This article belongs to the Special Issue Workplans II: Workshop for Planetary Nebula Observations)
27 pages, 3748 KiB  
Review
Gamma-Ray Dark Matter Searches in Milky Way Satellites—A Comparative Review of Data Analysis Methods and Current Results
by Javier Rico
Galaxies 2020, 8(1), 25; https://doi.org/10.3390/galaxies8010025 - 17 Mar 2020
Cited by 21 | Viewed by 3363
Abstract
If dark matter is composed of weakly interacting particles with mass in the GeV-TeV range, their annihilation or decay may produce gamma rays that could be detected by gamma-ray telescopes. Observations of dwarf spheroidal satellite galaxies of the Milky Way (dSphs) benefit from [...] Read more.
If dark matter is composed of weakly interacting particles with mass in the GeV-TeV range, their annihilation or decay may produce gamma rays that could be detected by gamma-ray telescopes. Observations of dwarf spheroidal satellite galaxies of the Milky Way (dSphs) benefit from the relatively accurate predictions of dSph dark matter content to produce robust constraints to the dark matter properties. The sensitivity of these observations for the search for dark matter signals can be optimized thanks to the use of advanced statistical techniques able to exploit the spectral and morphological peculiarities of the expected signal. In this paper, I review the status of the dark matter searches from observations of dSphs with the current generation of gamma-ray telescopes: Fermi-LAT, H.E.S.S, MAGIC, VERITAS and HAWC. I will describe in detail the general statistical analysis framework used by these instruments, putting in context the most recent experimental results and pointing out the most relevant differences among the different particular implementations. This will facilitate the comparison of the current and future results, as well as their eventual integration in a multi-instrument and multi-target dark matter search. Full article
(This article belongs to the Special Issue The Role of Halo Substructure in Gamma-Ray Dark Matter Searches)
Show Figures

Figure 1

Figure 1
<p>Expected gamma-ray spectral energy distribution for WIMPs of masses <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mi>χ</mi> </msub> <mo>=</mo> <mn>0</mn> <mo>.</mo> <mn>01</mn> <mo>,</mo> <mn>0</mn> <mo>.</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>10</mn> </mrow> </semantics></math> and 100 TeV annihilating with <math display="inline"><semantics> <mrow> <mrow> <mo>〈</mo> <mi>σ</mi> <mi mathvariant="italic">v</mi> <mo>〉</mo> </mrow> <mo>=</mo> <mn>3</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>26</mn> </mrow> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math> s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> into <math display="inline"><semantics> <mrow> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> (left) and <math display="inline"><semantics> <mrow> <msup> <mi>τ</mi> <mo>+</mo> </msup> <msup> <mi>τ</mi> <mo>−</mo> </msup> </mrow> </semantics></math> (right) pairs in a dSph with associated J-factor <math display="inline"><semantics> <mrow> <msub> <mi>J</mi> <mi>ann</mi> </msub> <mo>=</mo> <mn>5</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>21</mn> </msup> </mrow> </semantics></math> GeV<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> </semantics></math>; also shown are the sensitivity curves for the instruments considered in this paper. Fermi-LAT sensitivity curve [<a href="#B10-galaxies-08-00025" class="html-bibr">10</a>] corresponds to observations of a point-like source at Galactic coordinates <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mi>l</mi> <mo>,</mo> <mi>b</mi> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <msup> <mn>120</mn> <mo>∘</mo> </msup> <mo>,</mo> <msup> <mn>45</mn> <mo>∘</mo> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math> for 10 years, analyzed using the latest (Pass8) data reconstruction tools; HESS [<a href="#B11-galaxies-08-00025" class="html-bibr">11</a>], MAGIC [<a href="#B12-galaxies-08-00025" class="html-bibr">12</a>] and VERITAS [<a href="#B13-galaxies-08-00025" class="html-bibr">13</a>] curves correspond to 50 h of observations of a point-like source at low (Zd <math display="inline"><semantics> <mrow> <mo>≲</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>) zenith distance; HAWC curve [<a href="#B14-galaxies-08-00025" class="html-bibr">14</a>] is for five years of observations of a point-like source at a declination of +22<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>N. The flux sensitivity for 50 h observations with the future Cherenkov Telescope Array [<a href="#B15-galaxies-08-00025" class="html-bibr">15</a>] is shown for comparison.</p>
Full article ">Figure 2
<p>The 95% confidence-level upper limits to <math display="inline"><semantics> <mrow> <mo>〈</mo> <mi>σ</mi> <mi mathvariant="italic">v</mi> <mo>〉</mo> </mrow> </semantics></math> for the <math display="inline"><semantics> <mrow> <mi>χ</mi> <mi>χ</mi> <mo>→</mo> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <mrow> <mi>χ</mi> <mi>χ</mi> <mo>→</mo> <msup> <mi>τ</mi> <mo>+</mo> </msup> <msup> <mi>τ</mi> <mo>−</mo> </msup> </mrow> </semantics></math> (<b>right</b>) annihilation channels derived from 6-year observations of 15 dSphs with Fermi-LAT. The dashed black line shows the median of the distribution of limits obtained from 300 simulated realizations of the null hypothesis using LAT observations of high-Galactic-latitude empty fields, whereas green and yellow bands represent the symmetric 68% and 95% quantiles, respectively. The dashed gray curve corresponds to the thermal relic cross-section [<a href="#B54-galaxies-08-00025" class="html-bibr">54</a>]. Reprinted figure with permission from reference [<a href="#B37-galaxies-08-00025" class="html-bibr">37</a>]; copyright (2014) by the American Physical Society.</p>
Full article ">Figure 3
<p>The 95% confidence level upper limits to the cross section of dark matter annihilating into a combination of <math display="inline"><semantics> <mrow> <msup> <mi>W</mi> <mo>+</mo> </msup> <msup> <mi>W</mi> <mo>−</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>Z</mi> <mi>Z</mi> </mrow> </semantics></math> (<b>left</b>, reprinted figure with permission from reference [<a href="#B41-galaxies-08-00025" class="html-bibr">41</a>]; copyright (2014) by the American Physical Society) and <math display="inline"><semantics> <mrow> <mi>γ</mi> <mi>γ</mi> </mrow> </semantics></math> pairs (<b>right</b>, reprinted figure with permission from reference [<a href="#B44-galaxies-08-00025" class="html-bibr">44</a>], ©IOP Publishing Ltd. and Sissa Medialab; reproduced by permission of IOP Publishing; all rights reserved). Different lines show limits from individual dSphs and from their combination with and without Sagittarius. For the spectral line search, also the median of the distribution of limits obtained for simulated realizations of the null hypothesis is shown, together with the corresponding 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> and 2<math display="inline"><semantics> <mi>σ</mi> </semantics></math> symmetric quantiles.</p>
Full article ">Figure 4
<p>The 95% confidence level upper limits to <math display="inline"><semantics> <mrow> <mo>〈</mo> <mi>σ</mi> <mi mathvariant="italic">v</mi> <mo>〉</mo> </mrow> </semantics></math> (solid line) for the process <math display="inline"><semantics> <mrow> <mi>χ</mi> <mi>χ</mi> <mo>→</mo> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> from observations of the dSphs Segue 1 (<b>left</b>, reprinted figure under CC BY license from reference [<a href="#B45-galaxies-08-00025" class="html-bibr">45</a>]) and Ursa Major II (<b>right</b>, reprinted figure with permission from reference [<a href="#B46-galaxies-08-00025" class="html-bibr">46</a>], ©IOP Publishing Ltd. and Sissa Medialab; reproduced by permission of IOP Publishing; all rights reserved) with MAGIC; also shown are the median of the distribution of limits for the null-hypothesis, and the limits of the symmetric 68% and 95% quantiles. For Ursa Major II, both the results with and without considering <math display="inline"><semantics> <mover> <mi>J</mi> <mo>¯</mo> </mover> </semantics></math> statistical uncertainty are shown.</p>
Full article ">Figure 5
<p>The 95% confidence-level upper limits to the dark matter annihilation cross-section into <math display="inline"><semantics> <mrow> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <mrow> <msup> <mi>τ</mi> <mo>+</mo> </msup> <msup> <mi>τ</mi> <mo>−</mo> </msup> </mrow> </semantics></math> (<b>right</b>) pairs, obtained from dSph observations by VERITAS (black solid line), compared with results from other gamma-ray instruments (see legend for the details). Reprinted figure with permission from reference [<a href="#B48-galaxies-08-00025" class="html-bibr">48</a>]; copyright (2017) by the American Physical Society.</p>
Full article ">Figure 6
<p>The 95% confidence level upper limits to the annihilation cross-section of dark matter particles annihilating into <math display="inline"><semantics> <mrow> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <mrow> <msup> <mi>τ</mi> <mo>+</mo> </msup> <msup> <mi>τ</mi> <mo>−</mo> </msup> </mrow> </semantics></math> (<b>right</b>) pairs, from HAWC observations of dSphs (black solid line). Results from other gamma-ray instruments are also shown (see legend for details), as well as the median and 65% and 95% symmetric quantiles of the distribution of limits obtained under the null hypothesis. Figure reproduced with permission from reference [<a href="#B49-galaxies-08-00025" class="html-bibr">49</a>], ©AAS.</p>
Full article ">Figure 7
<p>The 95% confidence level upper limits to the cross-section for dark matter particles annihilating into <math display="inline"><semantics> <mrow> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <mrow> <msup> <mi>τ</mi> <mo>+</mo> </msup> <msup> <mi>τ</mi> <mo>−</mo> </msup> </mrow> </semantics></math> (<b>right</b>) pairs. Thick solid lines show the limits obtained by combining Fermi-LAT observations of 15 dSphs with MAGIC observations of Segue 1. Dashed lines show the limit obtained individually by MAGIC (short dashes) and Fermi-LAT (long dashes), respectively. The thin-dotted line, green and yellow bands show, respectively, the median and the two-sided 68% and 95% symmetric quantiles for the distribution of limits under the null hypothesis. Reprinted figure with permission from reference [<a href="#B22-galaxies-08-00025" class="html-bibr">22</a>], ©IOP Publishing Ltd. and Sissa Medialab; reproduced by permission of IOP Publishing; all rights reserved.</p>
Full article ">
6 pages, 258 KiB  
Article
X-ray Observations of Planetary Nebulae since WORKPLANS I and Beyond
by Martín A. Guerrero
Galaxies 2020, 8(1), 24; https://doi.org/10.3390/galaxies8010024 - 17 Mar 2020
Cited by 3 | Viewed by 2071
Abstract
Planetary nebulae (PNe) were expected to be filled with hot pressurized gas driving their expansion. ROSAT hinted at the presence of diffuse X-ray emission from these hot bubbles and detected the first sources of hard X-ray emission from their central stars, but it [...] Read more.
Planetary nebulae (PNe) were expected to be filled with hot pressurized gas driving their expansion. ROSAT hinted at the presence of diffuse X-ray emission from these hot bubbles and detected the first sources of hard X-ray emission from their central stars, but it was not until the advent of Chandra and XMM-Newton that we became able to study in detail their occurrence and physical properties. Here I review the progress in the X-ray observations of PNe since the first WORKshop for PLAnetary Nebulae observationS (WORKPLANS) and present the perspective for future X-ray missions with particular emphasis on eROSITA. Full article
(This article belongs to the Special Issue Workplans II: Workshop for Planetary Nebula Observations)
Show Figures

Figure 1

Figure 1
<p>(bottom) Time evolution of the number of PNe detected by the <span class="html-italic">Chandra</span> (red) and <span class="html-italic">XMM-Newton</span> (blue) X-ray observatories. (top) Time evolution of the number of papers (black) and citations to these papers (blue) based on those X-ray observations (source Astronomycal Database System, ADS, at <a href="https://ui.adsabs.harvard.edu" target="_blank">https://ui.adsabs.harvard.edu</a>). Note that the left and right axes in this panel refer to numbers of papers and citations, respectively, as denoted by their labels and colors. The red-dashed line is a linear fit to the number of papers, implying a publication rate ≃3 papers per year on the X-ray emission from PNe.</p>
Full article ">Figure 2
<p>Time evolution of the exposure time required for a <span class="html-italic">Chandra</span> ACIS-S observation to obtain a given number of counts for a plasma model typical of the diffuse X-ray emission in PNe. Simulations were obtained using the <span class="html-italic">Chandra</span> PIMMS v4.9 simulator, which includes the most up-to-date calibrations. A plasma with a typical 1.6 × 10<math display="inline"><semantics> <msup> <mrow/> <mn>6</mn> </msup> </semantics></math> K temperature, low hydrogen column density, and slightly subsolar chemical composition has been assumed.</p>
Full article ">Figure 3
<p>Distribution of exposure time (left) and sensitivity (rigth) for PNe expected in the final eRASS map of the whole sky. Most PNe will be located in areas of the eRASS with total exposure times shorter than 20 ks, but a significant number of them, ≃200, will have longer exposure times. Accordingly, a significant number of PNe, those with long exposure times in the final eRASS, will have observations with sensitivities adequate to detect X-ray fluxes smaller than ≃2×10<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>15</mn> </mrow> </msup> </semantics></math> erg cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math> s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>.</p>
Full article ">
8 pages, 1771 KiB  
Editorial
WORKPLANS: Workshop on Planetary Nebula Observations
by Isabel Aleman, Jeronimo Bernard-Salas, Joel H. Kastner, Toshiya Ueta and Eva Villaver
Galaxies 2020, 8(1), 23; https://doi.org/10.3390/galaxies8010023 - 16 Mar 2020
Cited by 1 | Viewed by 2890
Abstract
This workshop is the second of the WORKPLANS series, which we started in 2016. The main goal of WORKPLANS is to build up a network of planetary nebulae (PNe) experts to address the main open questions in the field of PNe research. The [...] Read more.
This workshop is the second of the WORKPLANS series, which we started in 2016. The main goal of WORKPLANS is to build up a network of planetary nebulae (PNe) experts to address the main open questions in the field of PNe research. The specific aims of the WORKPLANS workshop series are (i) to discuss and prioritize the most important topics to be investigated by the PN community in the following years; (ii) to establish a network of excellent researchers with complementary expertise; (iii) to formulate ambitious observing proposals for the most advanced telescopes and instrumentation presently available (ALMA, SOFIA, VLT, GTC, HST, etc.), addressing those topics; and (iv) to develop strategies for major proposals to future observatories (JWST, ELT, SPICA, Athena, etc.). To achieve these goals, WORKPLANS II brought together experts in all key sub-areas of the PNe research field, namely: analysis and interpretation of PNe observational data; theoretical modeling of gas and dust emission; evolution from Asymptotic Giant Branch stars (PNe progenitors) to PNe; and the instrumentation and technical characteristics of the relevant observatories. Full article
(This article belongs to the Special Issue Workplans II: Workshop for Planetary Nebula Observations)
Show Figures

Figure 1

Figure 1
<p>Participants of WORKPLANS II.</p>
Full article ">
19 pages, 1131 KiB  
Article
Diagnosing Magnetic Field Geometry in Blazar Jets Using Multi-Frequency, Centimeter-Band Polarimetry and Radiative Transfer Modeling
by Margo Aller, Philip Hughes, Hugh Aller and Talvikki Hovatta
Galaxies 2020, 8(1), 22; https://doi.org/10.3390/galaxies8010022 - 10 Mar 2020
Cited by 1 | Viewed by 2393
Abstract
We use multi-frequency linear polarization observations from the University of Michigan blazar program (UMRAO), in combination with radiative transfer simulations of emission from a relativistic jet, to investigate the time-dependent flow conditions, including magnetic field geometry, in an example blazar OT 081. We [...] Read more.
We use multi-frequency linear polarization observations from the University of Michigan blazar program (UMRAO), in combination with radiative transfer simulations of emission from a relativistic jet, to investigate the time-dependent flow conditions, including magnetic field geometry, in an example blazar OT 081. We adopt a scenario incorporating relativistic shocks during flaring, and both ordered axial and helical magnetic field components and magnetic turbulence in the underlying flow; these constituents are consistent with the observed periods of ordered behavior in the polarization intermixed with stochastic variations. The simulations are able to reproduce the global features of the observed light curves, including amplitude and spectral evolution of the linear polarization, during four time periods spanning 25 years. From the simulations, we identify the signature of a weak-to-strong helical magnetic field on the polarization, but conclude that a dominant helical magnetic field is not consistent with the UMRAO polarization data. The modeling identifies time-dependent changes in the ratio of the ordered-to-turbulent magnetic field, and changes in the flow direction and Lorentz factor. These suggest the presence of jet-like structures within a broad envelope seen at different orientations. Full article
(This article belongs to the Special Issue Polarimetry as a Probe of Magnetic Fields in AGN Jets)
Show Figures

Figure 1

Figure 1
<p>Longterm UMRAO light curves illustrate the range of variability in consecutive outbursts in the emission from the example program source OT 081 (1749 + 096). From bottom to top, 30-day averages of the total flux density (S), polarized flux (P), and electric vector position angle (EVPA or X) are shown. Green crosses, blue circles, and red triangles denote the data at 14.5, 8.0, and 4.8 GHz respectively. Modeled epochs are indicated with time labels in the lower panel.</p>
Full article ">Figure 2
<p>Comparison of the observed light curves in the form of daily-averaged data (bottom plots) and simulations (top plots) for epochs T1985 (left) and T1996 (right) showing time in years. The data at the three frequencies are symbol and color coded to match in the data and simulations. The simulated light curves were computed for 20 time steps using the adopted flow parameters listed in <a href="#galaxies-08-00022-t001" class="html-table">Table 1</a> and the values of the shock attributes listed in <a href="#app1-galaxies-08-00022" class="html-app">Appendix A</a> <a href="#galaxies-08-00022-t0A1" class="html-table">Table A1</a> and <a href="#galaxies-08-00022-t0A2" class="html-table">Table A2</a>. Upward black arrows mark the adopted shock onset times.</p>
Full article ">Figure 3
<p>Comparison of the the data (bottom plots) and simulations (top plots) for T2008 (<b>left</b>) and T2010 (<b>right</b>). Symbols, number of time steps in the simulation, and arrows are as in <a href="#galaxies-08-00022-f002" class="html-fig">Figure 2</a>. The shock parameters are given in <a href="#app1-galaxies-08-00022" class="html-app">Appendix A</a> <a href="#galaxies-08-00022-t0A3" class="html-table">Table A3</a> and <a href="#galaxies-08-00022-t0A4" class="html-table">Table A4</a>.</p>
Full article ">Figure 4
<p>(<b>a</b>) The simulation for the model for T2010 with <math display="inline"><semantics> <mi>θ</mi> </semantics></math> = 1.1<math display="inline"><semantics> <mo>°</mo> </semantics></math>. (<b>b</b>) The simulation with the same parameters with the exception of the viewing angle, which has been changed to 1.4<math display="inline"><semantics> <mo>°</mo> </semantics></math>.</p>
Full article ">Figure 5
<p>Simulated T1985 light curves with the addition of a weak helical magnetic field component. Panel (<b>a</b>) shows the simulated light curves for T1985 assuming a negligible helical magnetic field (an order multiple of 0.05; see text). As would be expected, this simulation is nearly identical to our adopted model (<a href="#galaxies-08-00022-t001" class="html-table">Table 1</a>) which did not include a helical magnetic field. Panel (<b>b</b>) shows the simulated light curves for a helical field with strength approaching that of the random component (an order multiple of 0.5). Time is shown in arbitrary units.</p>
Full article ">Figure 6
<p>Simulated T1985 light curves with the addition of a strong helical magnetic field component. Panel (<b>a</b>) shows the result for a helical field of strength comparable to that of the random component (an order multiple of 1.0, see text) and panel (<b>b</b>) shows the result for a helical field with strength that dominates over the random component (an order multiple of 10). Time is shown in arbitrary units.</p>
Full article ">
10 pages, 471 KiB  
Review
(Sub)mm-Wavelength Observations of Pre-Planetary Nebulae and Young Planetary Nebulae
by Carmen Sánchez Contreras
Galaxies 2020, 8(1), 21; https://doi.org/10.3390/galaxies8010021 - 10 Mar 2020
Viewed by 3731
Abstract
This is a non-comprehensive review of observations of pre-Planetary Nebulae (pPNe) and young Planetary Nebulae (yPNe) at (sub)mm-wavelengths, a valuable window for probing multi-phased gas and dust in these objects. This contribution focuses on observations of molecular lines (from carbon monoxide—CO—and other species), [...] Read more.
This is a non-comprehensive review of observations of pre-Planetary Nebulae (pPNe) and young Planetary Nebulae (yPNe) at (sub)mm-wavelengths, a valuable window for probing multi-phased gas and dust in these objects. This contribution focuses on observations of molecular lines (from carbon monoxide—CO—and other species), and briefly at the end, on hydrogen radio recombination lines from the emerging H ii regions at the center of yPNe. The main goal of this contribution is to show the potential of (sub)mm-wavelength observations of pPNe/yPNe to help the community to devise and develop new observational projects that will bring us closer to a better understanding of these latest stages of the evolution of low-to-intermediate (∼0.8–8 M ) mass stars. Full article
(This article belongs to the Special Issue Workplans II: Workshop for Planetary Nebula Observations)
Show Figures

Figure 1

Figure 1
<p>Adapted from Figure 8 of Sánchez Contreras et al. [<a href="#B58-galaxies-08-00021" class="html-bibr">58</a>]. Post-asymptotic giant branch (post-AGB) mass-loss vs. effective temperatures adopted by post-AGB models (solid black lines) and empirically deduced (symbols). The thick dashed and dot-dashed lines indicate the mass-loss rates typically assumed by evolutionary models for the three objects in with mmRRL detections. See Sánchez Contreras et al. [<a href="#B58-galaxies-08-00021" class="html-bibr">58</a>] for details.</p>
Full article ">
27 pages, 5016 KiB  
Review
Luminous Blue Variables
by Kerstin Weis and Dominik J. Bomans
Galaxies 2020, 8(1), 20; https://doi.org/10.3390/galaxies8010020 - 29 Feb 2020
Cited by 44 | Viewed by 5421
Abstract
Luminous Blue Variables are massive evolved stars, here we introduce this outstanding class of objects. Described are the specific characteristics, the evolutionary state and what they are connected to other phases and types of massive stars. Our current knowledge of LBVs is limited [...] Read more.
Luminous Blue Variables are massive evolved stars, here we introduce this outstanding class of objects. Described are the specific characteristics, the evolutionary state and what they are connected to other phases and types of massive stars. Our current knowledge of LBVs is limited by the fact that in comparison to other stellar classes and phases only a few “true” LBVs are known. This results from the lack of a unique, fast and always reliable identification scheme for LBVs. It literally takes time to get a true classification of a LBV. In addition the short duration of the LBV phase makes it even harder to catch and identify a star as LBV. We summarize here what is known so far, give an overview of the LBV population and the list of LBV host galaxies. LBV are clearly an important and still not fully understood phase in the live of (very) massive stars, especially due to the large and time variable mass loss during the LBV phase. We like to emphasize again the problem how to clearly identify LBV and that there are more than just one type of LBVs: The giant eruption LBVs or η Car analogs and the S Dor cycle LBVs. Full article
(This article belongs to the Special Issue Luminous Stars in Nearby Galaxies)
Show Figures

Figure 1

Figure 1
<p>This figure taken from Burggraf (2015) [<a href="#B18-galaxies-08-00020" class="html-bibr">18</a>] shows a lightcurve spanning more than 100 years of the LBV and original Hubble Sandage Variable Var B in M 33. In addition to the B magnitudes upper section the spectral type if know for the same date is plotted in the lower section. Note the for S Dor cycles typical changes in the spectral type.</p>
Full article ">Figure 2
<p>This figure shows the classical plot by Wolf [<a href="#B17-galaxies-08-00020" class="html-bibr">17</a>] (<b>top</b>) and in a new version (<b>bottom</b>) we plotted the luminosity L and change in Temperature <math display="inline"><semantics> <mo>Δ</mo> </semantics></math>T<math display="inline"><semantics> <msub> <mrow/> <mi>eff</mi> </msub> </semantics></math> for a new way to visualize the amplitude- luminosity-relation.</p>
Full article ">Figure 3
<p>HRD with Galactic and LMC LBVs and LBV candidates. Circles are used for LBVs with an emission line (optical/NIR) nebulae, squares for all others. If an S Dor cycle has been observed both the cool (open symbol) and the hot phase (filled symbol) are marked. Otherwise an open grayish symbol is used. In color evolutionary tracks for different masses are added. The tracks are based on the data from the Geneva code for Z = 0.02 and v<math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>r</mi> <mi>o</mi> <mi>t</mi> </mrow> </msub> </semantics></math> = 300 km/s, colors code the generally three different evolutionary scenarios, see text for details.</p>
Full article ">Figure 4
<p>HST images of LBV nebulae sorted by morphology: hourglass AG Car [<a href="#B54-galaxies-08-00020" class="html-bibr">54</a>], R 127 with bipolar attachments, weakly bipolar He 3-519 [<a href="#B55-galaxies-08-00020" class="html-bibr">55</a>], spherical S 61 [<a href="#B52-galaxies-08-00020" class="html-bibr">52</a>] and last in row irregular R 143 [<a href="#B52-galaxies-08-00020" class="html-bibr">52</a>].</p>
Full article ">Figure 5
<p>An LBT LUCI AO [FeII] image of the inner nebula (or inner shell) of P Cygni (Weis et al., in prep). The images has pixel scale of 0.015”/pixel and resolve scales down to 85 AU.</p>
Full article ">Figure 6
<p>The nebula around <math display="inline"><semantics> <mi>η</mi> </semantics></math> Carinae in the optical and X-ray. <b>Left</b>: An optical F658N HST image in greyscale, the Homunculus nebula additionally marked in contour to distinguish it from the outer ejecta, shown only in grey scale [<a href="#B58-galaxies-08-00020" class="html-bibr">58</a>]. <b>Right</b>: A CHANDRA Xray image with color coded energy regimes, green:0.2-0.6 keV, red: 0.6–1.2 keV and blue 1.2–12 keV color version of Figure 1 in [<a href="#B95-galaxies-08-00020" class="html-bibr">95</a>].</p>
Full article ">Figure 7
<p>Plot of the gas-phase metallicity of nearby galaxies versus their distance and spatial resolution. Only a selection of the galaxies in the Local volume are plotted, but the sample is complete for the significantly starforming galaxies in the Local Group. Metalicities of the inner disk are chosen for the spiral galaxies with metallicity gradients, the metalicities of stars in the outer disk of these galaxies can be a a few faction of tens solar lower. Galaxies with LBVs and/or LBV candidates are plotted as red dots, the other galaxies are plotted as blue dots. Plot was adapted and updated from [<a href="#B163-galaxies-08-00020" class="html-bibr">163</a>].</p>
Full article ">
33 pages, 15972 KiB  
Article
Density Profiles of 51 Galaxies from Parameter-Free Inverse Models of Their Measured Rotation Curves
by Robert E. Criss and Anne M. Hofmeister
Galaxies 2020, 8(1), 19; https://doi.org/10.3390/galaxies8010019 - 26 Feb 2020
Cited by 12 | Viewed by 5594
Abstract
Spiral galaxies and their rotation curves have key characteristics of differentially spinning objects. Oblate spheroid shapes are a consequence of spin and reasonably describe galaxies, indicating that their matter is distributed in gravitationally interacting homeoidal shells. Here, previously published equations describing differentially spinning [...] Read more.
Spiral galaxies and their rotation curves have key characteristics of differentially spinning objects. Oblate spheroid shapes are a consequence of spin and reasonably describe galaxies, indicating that their matter is distributed in gravitationally interacting homeoidal shells. Here, previously published equations describing differentially spinning oblate spheroids with radially varying density are applied to 51 galaxies, mostly spirals. A constant volumetric density (ρ, kg m−3) is assumed for each thin homeoid in these formulae, after Newton, which is consistent with RCs being reported simply as a function of equatorial radius r. We construct parameter-free inverse models that uniquely specify mass inside any given r, and thus directly constrain ρ vs. r solely from velocity v (r) and galactic aspect ratios (assumed as 1:10 for spirals when data are unavailable). Except for their innermost zones, ρ is proven to be closely proportional to rn, where the statistical average of n for all 36 spirals studied is −1.80 ± 0.40. Our values for interior densities compare closely with independently measured baryon density in appropriate astronomical environments: for example, calculated ρ at galactic edges agrees with independently estimated ρ of intergalactic media (IGM). Our finding that central densities increase with galaxy size is consistent with behavior exhibited by diverse self-gravitating entities. Our calculated mass distributions are consistent with visible luminosity and require no non-baryonic component. Full article
Show Figures

Figure 1

Figure 1
<p>Schematics. <b>(a)</b> Graph comparing spin of a rigid top (with null central velocities) to Keplerian orbits (which have high central velocities) and to galactic RC, which show a gradation between these limiting cases. Mass components cannot be summed to produce RCs [<a href="#B9-galaxies-08-00019" class="html-bibr">9</a>], as has been pursued in the Newtonian orbital model approach (e.g., [<a href="#B12-galaxies-08-00019" class="html-bibr">12</a>,<a href="#B13-galaxies-08-00019" class="html-bibr">13</a>,<a href="#B14-galaxies-08-00019" class="html-bibr">14</a>]. <b>(b)</b> Nesting of similar, homogeneous homeoids to form a stratified oblate body, with axes labeled.</p>
Full article ">Figure 2
<p>Isophotes of edge-on galaxies at various wavelengths: (left) New General Catalog (NGG) 3109 in the B-band. Modified after figure 6 in Carignan, C. <b>1985</b>, Light and mass distribution of the magellanic-type spiral NGC 3109. <span class="html-italic">Astrophys. J. 299</span>, 59–73 [<a href="#B24-galaxies-08-00019" class="html-bibr">24</a>] with permissions; (middle) NCG 4594 (Sombrero) isophotes at 3.6 µm (circles are foreground stars). Modified after figure 1 in Gadotti, D.A. and Sánchez-Janssen, R. <b>2012</b>, Surprises in image decomposition of edge-on galaxies: Does Sombrero have a (classical) bulge? <span class="html-italic">Mon. Not. R. Astron. Soc. 423</span>, 877–888 [<a href="#B26-galaxies-08-00019" class="html-bibr">26</a>], with permissions from Oxford University Press on behalf of the R. Astron. Soc.; (right) Grey shades = the L-band from averaging 30 spirals, using the 22 µm contour to represent the disk. Modified after figure 6 in [<a href="#B21-galaxies-08-00019" class="html-bibr">21</a>] (Wiegert, T. and 9 others, <b>2015</b>. CHANG-ES. IV. Radio continuum emission of 35 edge-on galaxies observed with the Karl G. Jansky very large array in D configuration—Data release 1. Astron. J., 150:81, doi:10.1088/0004-6256/150/3/81), with permissions.</p>
Full article ">Figure 3
<p>Inverse models showing the effect of varying <span class="html-italic">c</span>/<span class="html-italic">a</span> over a large range for an internally probed galaxy, the Milky Way. Compiled RCs are used to calculate (<b>a</b>) mass inside and (<b>b</b>) density at any given radius. In part (<b>a</b>), upper box describes line patterns associated with RCs [<a href="#B58-galaxies-08-00019" class="html-bibr">58</a>,<a href="#B59-galaxies-08-00019" class="html-bibr">59</a>,<a href="#B66-galaxies-08-00019" class="html-bibr">66</a>] using the left axis and two additional constraints on <span class="html-italic">v</span> near the Sun [<a href="#B81-galaxies-08-00019" class="html-bibr">81</a>,<a href="#B82-galaxies-08-00019" class="html-bibr">82</a>] and from globular cluster data [<a href="#B60-galaxies-08-00019" class="html-bibr">60</a>]. Lower box lists aspect ratios assumed in analyzing each dataset for mass, shown on the right axis. Error bars on <span class="html-italic">v</span> from [<a href="#B58-galaxies-08-00019" class="html-bibr">58</a>] are shown. Downturns in <span class="html-italic">M</span><sub>in</sub> and ρ indicate that the “edge” of the Galaxy is gradual and between 18 and 30 kpc. In (<b>b</b>), density above <span class="html-italic">r</span> = 0.1 kpc was fit to a power law, as shown.</p>
Full article ">Figure 4
<p>Expanded view of the Milky Way center. Mass and density are shown on left and right axes, respectively. The fit to ρ is for <span class="html-italic">r</span> &lt; 0.2 kpc. Central mass is estimated by extrapolating the mass curve (dotted line) and by assuming a high density (dashed line). See <a href="#galaxies-08-00019-f003" class="html-fig">Figure 3</a> for data sources.</p>
Full article ">Figure 5
<p>Comparison of three datasets on Triangulum to calculate mass (<b>a</b>) and density (<b>b</b>) vs. radius. RCs from Kam et al. [<a href="#B14-galaxies-08-00019" class="html-bibr">14</a>] (grey crosses) were used to calculate mass (grey diamonds). RCs of Sofue [<a href="#B61-galaxies-08-00019" class="html-bibr">61</a>] (black dashed line) were used to calculate mass (heavy black dotted line) and density (black dotted line). RCs of Corbelli and Salucci [<a href="#B70-galaxies-08-00019" class="html-bibr">70</a>] (black points with error bars) were used to calculate mass for <span class="html-italic">c</span>/<span class="html-italic">a</span> = 1 (solid line) and <span class="html-italic">c</span>/<span class="html-italic">a</span> = 0.1 (medium dashed line) and density for <span class="html-italic">c</span>/<span class="html-italic">a</span> = 0.1 (circles and solid line). Agreement is good and the masses extracted are similar. The fit to ρ is over all <span class="html-italic">r</span> of [<a href="#B70-galaxies-08-00019" class="html-bibr">70</a>]. An abrupt drop off in density is not observed. RC data are limited to the visible disk.</p>
Full article ">Figure 6
<p>Inverse models of interesting spirals. Velocity and <span class="html-italic">M</span><sub>in</sub>/(10<sup>9</sup><span class="html-italic">M</span><sub>Sun</sub>) are both plotted on the left axis. (<b>left</b>) Sombrero galaxy, showing that disparities between RC measurements have little effect. Heavy solid line = RC of [<a href="#B71-galaxies-08-00019" class="html-bibr">71</a>] was used to calculate <span class="html-italic">M</span><sub>in</sub> (light solid) and ρ (open diamonds). Heavy broken lines = RC of [<a href="#B72-galaxies-08-00019" class="html-bibr">72</a>] used to calculate mass (light broken line) and density (dots and fit). The visual edge is beyond the measurements; (<b>right</b>) Counter-rotating NGC-4826. RCs from deBlok et al. [<a href="#B54-galaxies-08-00019" class="html-bibr">54</a>]. Results for mass and density are close to similar sized normally rotating NGC 7793 (<a href="#galaxies-08-00019-f0A3" class="html-fig">Figure A3</a>). We assumed that <span class="html-italic">v</span> = 0 at 4 kpc, midway between the nearest counter-rotating data points. The breadth of the <span class="html-italic">v</span> = 0 region affects the mass calculation, but little alters the density profile.</p>
Full article ">Figure 7
<p>Small galaxies: (<b>a</b>) Low surface brightness, dwarf irregular galaxy NGC 4789a. RCs from Bottema and Pestaña [<a href="#B64-galaxies-08-00019" class="html-bibr">64</a>]. <span class="html-italic">M</span><sub>in</sub> ratioed to 10<sup>6</sup> <span class="html-italic">M</span><sub>Sun</sub> Two fits to density are shown: solid = power law and dots = exponential; (<b>b</b>) Compact dwarf elliptical M32. RCs from Howley et al. [<a href="#B55-galaxies-08-00019" class="html-bibr">55</a>]. <span class="html-italic">M</span><sub>in</sub> ratioed to 10<sup>6</sup> <span class="html-italic">M</span><sub>Sun</sub>. Density above 0.6 kpc is described by a power law, but not an exponential.</p>
Full article ">Figure 8
<p>The prototype polar ring galaxy NGC 4650a. RC raw data from Iodice et al. [<a href="#B80-galaxies-08-00019" class="html-bibr">80</a>] are similar to previous reports. (<b>a</b>) Comparison of RC data from different sectors and fits; see text. (<b>b</b>) The fit that excluded the uncertain dataset to provide <span class="html-italic">M</span><sub>in</sub>/(10<sup>8</sup><span class="html-italic">M</span><sub>Sun</sub>) and ρ.</p>
Full article ">Figure 9
<p>Various measures for the density of galaxies compared to their size, set to the visual edge at 25 B-magnitude arcsec<sup>−2</sup>. (<b>a</b>) Spiral and similar galaxies. M82 is included since near-IR images [<a href="#B23-galaxies-08-00019" class="html-bibr">23</a>] suggest it is a spiral. Highest and lowest densities are depicted, but these values depend somewhat on smallest and largest radii explored in each RC study. For reference, we include ρ measured at <span class="html-italic">r</span> = 0.1 kpc and ρ measured at the visual edge. The polar ring galaxy has lower ρ than ordinary spirals of similar size. (<b>b</b>) Morphologies other than spiral. Fits to ellipticals (lines) roughly also describe the lenticular and spheroidal classes and are similar to fits for spirals.</p>
Full article ">Figure 10
<p>Density trends in spiral galaxies: (<b>a</b>) Radius of the turndown in density on the position of 0.1 kpc. Density at 0.1 kpc strongly increases with the size of both spiral and elliptical galaxies (<a href="#galaxies-08-00019-f009" class="html-fig">Figure 9</a>a,b). The slopes of the regression lines for these types are similar, despite the disparity in the number of samples (36 spirals vs. 4 ellipticals), and disparities in central density between ellipticals and spirals. The three lenticular and eight dwarf galaxies measured also fall on or near these trends. The polar-ring type is less dense, due to its shape, with considerable matter out of the plane. Notably, the smallest galaxies, the dwarf spheroidals, have densities near the crossing of the trends and show little variation in ρ with <span class="html-italic">r</span>. Interestingly, increasing tightness of the spiral arms is associated with higher ρ at 0.1 kpc (<a href="#galaxies-08-00019-f010" class="html-fig">Figure 10</a>b). This finding is consistent with spiral arms being concentrations of stars.</p>
Full article ">Figure 11
<p>Galaxy mass–size relationships. (<b>a</b>) Dependence of galaxy mass on the position of the visible edge. (<b>b</b>) Dependence of the largest mass calculated for each galaxy on the radius where the largest mass was calculated. In most cases, this radius is the outer cutoff of available data, but some exceptions exist. Fits are to certain types, as listed in the inset.</p>
Full article ">Figure 12
<p>Dependence of density at various galactic positions on the visible luminosity of the galaxy: (<b>a</b>) Data for spirals and the polar-ring galaxy; (<b>b</b>) Other galaxy types, with a wider range of luminosity. The dwarf spheroidals show a flat dependence (dash-dotted curve). For the smallest galaxies, the average density is ~10<sup>−21</sup> kg m<sup>−3</sup>, i.e., where trends for the large spirals converge.</p>
Full article ">Figure 13
<p>Dependence of calculated parameters on measured visible luminosity of our galaxy sample. Least squares fits are shown: (<b>a</b>) Mass of all galaxy types. The maximum mass is affected by the termination of velocity measurements with distance. The visual edge provides a more consistent measure of galaxy size in relationship to luminosity. Power law fits give exponents very close to unity: the linear fits shown are similar. Grey line indicates a 1:1 correspondence; (<b>b</b>) Power (−<span class="html-italic">n</span>) for fits to density vs. radius. The least squares fit for spirals (solid curve) shows that density is more concentrated in the centers of larger galaxies. The polar-ring galaxy is less concentrated than normal spirals of similar size.</p>
Full article ">Figure 14
<p>Dependence of luminosity in the visible and at 21 cm wavelengths on galaxy size, as determined from the visual edge at 25 B-magnitude arcsec<sup>−2</sup>. Data compiled in <a href="#galaxies-08-00019-t004" class="html-table">Table 4</a>. Fits exclude the compact ellipsoid, M32.</p>
Full article ">Figure A1
<p>Andromeda. Inverse models showing the effect of varying <span class="html-italic">c</span>/<span class="html-italic">a</span> for an externally probed galaxy, with the most extensive RCs available. Compiled RCs of Sofue [<a href="#B59-galaxies-08-00019" class="html-bibr">59</a>] are used to calculate: (<b>a</b>) <span class="html-italic">M</span><sub>in</sub> and; (<b>b</b>) density vs. radius. Labels on curves list the aspect ratios considered.</p>
Full article ">Figure A2
<p>Large spirals: NGC 253 RCs from Sofue et al. [<a href="#B66-galaxies-08-00019" class="html-bibr">66</a>] and derived <span class="html-italic">M</span> and ρ are colored grey. RCs from Hlavacek-Larrondo et al. [<a href="#B69-galaxies-08-00019" class="html-bibr">69</a>] are black curves. Although velocity data differ substantially, similar mass and density are obtained near the visible edge. Differences at small <span class="html-italic">r</span> affect the fits to ρ; NCG 925 has low velocities which produce low density; NGC 2599 RCs from Noordemeer et al. [<a href="#B61-galaxies-08-00019" class="html-bibr">61</a>]. This spiral has both high <span class="html-italic">v</span> and a strong decline in <span class="html-italic">v</span> with radius. The increase in <span class="html-italic">v</span> at low <span class="html-italic">r</span> was assumed; UCG 2855 RCs below 13.7 kpc from Sofue et al. [<a href="#B66-galaxies-08-00019" class="html-bibr">66</a>], and RCs above 13.7 kpc from Roelfsema and Allen [<a href="#B62-galaxies-08-00019" class="html-bibr">62</a>]. The merge produces unrealistically high density over a short interval but does not seem to affect the fit. A drop off in density exists at high <span class="html-italic">r</span>, whereas the trend is flatter than the fit near the galaxy center.</p>
Full article ">Figure A2 Cont.
<p>Large spirals: NGC 253 RCs from Sofue et al. [<a href="#B66-galaxies-08-00019" class="html-bibr">66</a>] and derived <span class="html-italic">M</span> and ρ are colored grey. RCs from Hlavacek-Larrondo et al. [<a href="#B69-galaxies-08-00019" class="html-bibr">69</a>] are black curves. Although velocity data differ substantially, similar mass and density are obtained near the visible edge. Differences at small <span class="html-italic">r</span> affect the fits to ρ; NCG 925 has low velocities which produce low density; NGC 2599 RCs from Noordemeer et al. [<a href="#B61-galaxies-08-00019" class="html-bibr">61</a>]. This spiral has both high <span class="html-italic">v</span> and a strong decline in <span class="html-italic">v</span> with radius. The increase in <span class="html-italic">v</span> at low <span class="html-italic">r</span> was assumed; UCG 2855 RCs below 13.7 kpc from Sofue et al. [<a href="#B66-galaxies-08-00019" class="html-bibr">66</a>], and RCs above 13.7 kpc from Roelfsema and Allen [<a href="#B62-galaxies-08-00019" class="html-bibr">62</a>]. The merge produces unrealistically high density over a short interval but does not seem to affect the fit. A drop off in density exists at high <span class="html-italic">r</span>, whereas the trend is flatter than the fit near the galaxy center.</p>
Full article ">Figure A3
<p>Comparison of moderate-sized galaxies: NGC 3034 is the irregular M82 galaxy. RCs from Greco et al. [<a href="#B63-galaxies-08-00019" class="html-bibr">63</a>]. An abrupt drop off in density is not observed, because RC data are limited to the visible galaxy. NGC 7793 RCs from deBlok et al. [<a href="#B54-galaxies-08-00019" class="html-bibr">54</a>] have fairly low velocity, and lower density.</p>
Full article ">Figure A4
<p>High-velocity large spirals from the compilation of Bottema and Pestaña [<a href="#B64-galaxies-08-00019" class="html-bibr">64</a>]. Inner densities for similar visible edges depend on velocities.</p>
Full article ">Figure A4 Cont.
<p>High-velocity large spirals from the compilation of Bottema and Pestaña [<a href="#B64-galaxies-08-00019" class="html-bibr">64</a>]. Inner densities for similar visible edges depend on velocities.</p>
Full article ">Figure A5
<p>Three small spirals: RC data from Bottema and Pestaña [<a href="#B64-galaxies-08-00019" class="html-bibr">64</a>].</p>
Full article ">Figure A6
<p>Spirals in Messier’s Catalog studied by Sofue et al. [<a href="#B53-galaxies-08-00019" class="html-bibr">53</a>,<a href="#B66-galaxies-08-00019" class="html-bibr">66</a>]. These large and luminous galaxies have a variety of RC patterns. Tabular data were downloaded from [<a href="#B65-galaxies-08-00019" class="html-bibr">65</a>].</p>
Full article ">Figure A6 Cont.
<p>Spirals in Messier’s Catalog studied by Sofue et al. [<a href="#B53-galaxies-08-00019" class="html-bibr">53</a>,<a href="#B66-galaxies-08-00019" class="html-bibr">66</a>]. These large and luminous galaxies have a variety of RC patterns. Tabular data were downloaded from [<a href="#B65-galaxies-08-00019" class="html-bibr">65</a>].</p>
Full article ">Figure A7
<p>Irregular galaxies: For Holmberg II, squares and dashed line = RCs from Oh et al. [<a href="#B74-galaxies-08-00019" class="html-bibr">74</a>]. Grey diamonds = RCs from Bureau and Carnigan [<a href="#B85-galaxies-08-00019" class="html-bibr">85</a>], who state their velocities are not accurate beyond 10 kpc, and so this dataset was not used in the calculations. Dotted line = calculated mass. Circles = calculated density. Two different fits are shown as solid line and widely spaced dotted curve; WLM is unusual due to its isolation RCs from Leaman et al. [<a href="#B73-galaxies-08-00019" class="html-bibr">73</a>]; M81dWb RCs from Oh et al. [<a href="#B74-galaxies-08-00019" class="html-bibr">74</a>].</p>
Full article ">Figure A8
<p>Dwarf spheroidal galaxies. RCs presented by Salucci et al. [<a href="#B75-galaxies-08-00019" class="html-bibr">75</a>] were digitized. Original data from Walker et al. [76.77] and Mateo et al. [<a href="#B78-galaxies-08-00019" class="html-bibr">78</a>]. Fornax is larger, more like the dwarf irregulars in <a href="#galaxies-08-00019-f0A7" class="html-fig">Figure A7</a>. For Draco, we compare extractions from the published data with those from the same data, but smoothed, and find little effect on the results. For Leo I, we compare extractions from the published data with those for an interpolated data set. Interpolating at <span class="html-italic">r</span> below the lowest measurement is equivocal and the choice of v affects interior density.</p>
Full article ">Figure A8 Cont.
<p>Dwarf spheroidal galaxies. RCs presented by Salucci et al. [<a href="#B75-galaxies-08-00019" class="html-bibr">75</a>] were digitized. Original data from Walker et al. [76.77] and Mateo et al. [<a href="#B78-galaxies-08-00019" class="html-bibr">78</a>]. Fornax is larger, more like the dwarf irregulars in <a href="#galaxies-08-00019-f0A7" class="html-fig">Figure A7</a>. For Draco, we compare extractions from the published data with those from the same data, but smoothed, and find little effect on the results. For Leo I, we compare extractions from the published data with those for an interpolated data set. Interpolating at <span class="html-italic">r</span> below the lowest measurement is equivocal and the choice of v affects interior density.</p>
Full article ">Figure A9
<p>Lenticular galaxies. RCs of UGC 3993 and NGC 7786 from Noordemeer et al. [<a href="#B61-galaxies-08-00019" class="html-bibr">61</a>]. RCs of NGC 2768 from Forbes et al. [<a href="#B42-galaxies-08-00019" class="html-bibr">42</a>].</p>
Full article ">Figure A10
<p>Elliptical galaxies. RCs of nearly spherical NGC 2434 from Rix et al. [<a href="#B68-galaxies-08-00019" class="html-bibr">68</a>]. RCs of dwarf NGC 4431 from Toluba et al. [<a href="#B79-galaxies-08-00019" class="html-bibr">79</a>]. RCs of NCG 3379 from Romanowsky et al. [<a href="#B31-galaxies-08-00019" class="html-bibr">31</a>].</p>
Full article ">
3 pages, 665 KiB  
Editorial
Radio Galaxies at TeV Energies: Preface
by Dorit Glawion
Galaxies 2020, 8(1), 18; https://doi.org/10.3390/galaxies8010018 - 22 Feb 2020
Cited by 1 | Viewed by 1999
Abstract
The majority of the known extragalactic sky from TeV gamma-ray energies consists of blazars having plasma jets pointing in the direction of the line-of-sight, which results in a large Doppler boosting of their emission. Up to now, only six galaxies with a larger [...] Read more.
The majority of the known extragalactic sky from TeV gamma-ray energies consists of blazars having plasma jets pointing in the direction of the line-of-sight, which results in a large Doppler boosting of their emission. Up to now, only six galaxies with a larger viewing angle have been detected in the TeV range. These objects also show fascinating properties, such as fast variability or spectral features and are called “radio galaxies”. The TeV radio galaxies provide a unique laboratory for studying key aspects of active galactic nuclei. This Special Issue of Galaxies targets these exciting objects. Full article
(This article belongs to the Special Issue Radio Galaxies at TeV Energies)
16 pages, 7299 KiB  
Review
From SN 2010da to NGC 300 ULX-1: Ten Years of Observations of an Unusual High Mass X-Ray Binary in NGC 300
by Breanna A. Binder, Stefania Carpano, Marianne Heida and Ryan Lau
Galaxies 2020, 8(1), 17; https://doi.org/10.3390/galaxies8010017 - 18 Feb 2020
Cited by 5 | Viewed by 5975
Abstract
In May 2010, an intermediate luminosity optical transient was discovered in the nearby galaxy NGC 300 by a South African amateur astronomer. In the decade since its discovery, multi-wavelength observations of the misnamed “SN 2010da” have continually reshaped our understanding of this high [...] Read more.
In May 2010, an intermediate luminosity optical transient was discovered in the nearby galaxy NGC 300 by a South African amateur astronomer. In the decade since its discovery, multi-wavelength observations of the misnamed “SN 2010da” have continually reshaped our understanding of this high mass X-ray binary system. In this review, we present an overview of the multi-wavelength observations and attempt to understand the 2010 transient event, and later, the reclassification of this system as NGC 300 ULX-1: a red supergiant + neutron star ultraluminous X-ray source. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Left</b>): The initial discovery image [<a href="#B8-galaxies-08-00017" class="html-bibr">8</a>] of SN 2010da by L.A.G. Monard [<a href="#B1-galaxies-08-00017" class="html-bibr">1</a>]. (<b>Right</b>): An RGB-rendered IR composite image of NGC 300, with insets showing the 3.6 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m image of the progenitor (∼6.5 years pre-outburst), ∼4.5 years post-outburst, and the difference subtracted image. Reproduced with permission from Lau et al. [<a href="#B9-galaxies-08-00017" class="html-bibr">9</a>] (their Figure 1).</p>
Full article ">Figure 2
<p>The mid-IR color–magnitude diagram (CMD) showing the SN 2010da progenitor (red square) and local luminous blue variables (LBVs) and LBV candidates (filled circles and triangles, respectively). Open circles and filled triangles show the CMD locations of LBV-like supernova (SN) “impostors.” Reproduced with permission from Khan et al. [<a href="#B3-galaxies-08-00017" class="html-bibr">3</a>].</p>
Full article ">Figure 3
<p>The IR spectral energy density (SED) of SN 2010da progenitor (black squares show measurements, black downward facing triangles show upper limits). The SEDs of two other optical progenitors, NGC 300 OT2008-1 and SN 2008S, are shown (in blue and brown, respectively) for comparison. The dotted red line shows the SED of the known LBV AG Car; the addition of 12 mag extinction is required for this SED to fit the observed fluxes of SN 2010da (dashed red line). Reproduced with permission from Berger and Chornock [<a href="#B19-galaxies-08-00017" class="html-bibr">19</a>].</p>
Full article ">Figure 4
<p><b>Left</b>: Adapted from Figure 1 in Binder et al. [<a href="#B23-galaxies-08-00017" class="html-bibr">23</a>], showing the <span class="html-italic">Chandra</span> image of SN 2010da. The yellow circle indicates the location of the X-ray source coincident with SN 2010da (the three fainter sources to the upper-left are unrelated). <b>Right</b>: Figure 2 from Binder et al. [<a href="#B23-galaxies-08-00017" class="html-bibr">23</a>] showing the <span class="html-italic">Hubble</span>/Advanced Camera for Surveys (ACS) image of the same region. The white cross and circle indicate the location of the X-ray source; the green cross and circle show the location of the likely massive donor star.</p>
Full article ">Figure 5
<p>Figure 1 from Carpano et al. [<a href="#B28-galaxies-08-00017" class="html-bibr">28</a>], showing the period evolution during the deep <span class="html-italic">XMM-Newton</span>/<span class="html-italic">NuSTAR</span> observations.</p>
Full article ">Figure 6
<p>The <span class="html-italic">XMM-Newton</span> and <span class="html-italic">NuSTAR</span> spectra of NGC 300 ULX-1 (top panel of Figure 3 in Carpano et al. [<a href="#B28-galaxies-08-00017" class="html-bibr">28</a>]). <span class="html-italic">XMM-Newton</span>/EPIC pn spectra are shown in black and blue, while the MOS spectra are shown in red, green, magenta and cyan. <span class="html-italic">NuSTAR</span> FPMA and FPMB spectra are shown in yellow and orange, respectively.</p>
Full article ">Figure 7
<p><b>Left</b>: <span class="html-italic">Swift</span>/XRT long term light curve from 2018 January to 2019 May. The corresponding hardness ratios (<span class="html-italic">S</span>: 0.2–1.5 keV, <span class="html-italic">H</span>: 1.5–10 keV) are shown in the bottom panel. <b>Right</b>: Evolution of the pulse period from <span class="html-italic">NICER</span> (pink) and <span class="html-italic">Swift</span> (blue) data covering the period from January 2018 to January 2019, when pulsations could be detected.</p>
Full article ">Figure 8
<p>(<b>Left</b>) <span class="html-italic">Spitzer</span>/IRAC mid-IR light curve of SN 2010da/NGC 300 ULX-1 from 2004–2019. The red and blue points correspond to photometry measured at 3.6 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and 4.5 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m, respectively. Only 3.6 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m observations were serendipitously obtained near the 2010 outburst. (<b>Right</b>) Spectral energy distribution (SED) of SN 2010da/NGC 300 ULX-1 during pre-outburst (December 2007) and post-outburst phases (late 2014/early 2015) overlaid on SED templates of 122 RSG and 11 sgB[e] stars in the Large Magellanic Cloud, cataloged by Bonanos et al. [<a href="#B55-galaxies-08-00017" class="html-bibr">55</a>]. Solid lines correspond to the median VIJHKs and <span class="html-italic">Spitzer</span>/IRAC magnitudes of the SED template stars, and the surrounding shaded regions indicate the 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> spread in the magnitudes of the distribution. The wavelength on the <span class="html-italic">x</span>-axis is shown in units of <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m. Both figures are reproduced and modified from Lau et al. [<a href="#B53-galaxies-08-00017" class="html-bibr">53</a>].</p>
Full article ">Figure 9
<p>Deep Xshooter spectroscopy of NGC 300 ULX-1 in October 2018 (black). The composite model (blue) requires a RSG atmosphere (red), excess dust emission (red short-dashed line) and a power-law blue excess (red long-dashed line) attributed to an irradiated accretion disk. Reproduced with permission from Heida et al. [<a href="#B57-galaxies-08-00017" class="html-bibr">57</a>].</p>
Full article ">Figure 10
<p>Mid-IR and X-ray light curve of SN 2010da/NGC 300 ULX-1 taken between MJD 57300 and MJD 58770 by <span class="html-italic">Spitzer</span>/IRAC and <span class="html-italic">Swift</span>/XRT, respectively. This figure is modified from Lau et al. [<a href="#B53-galaxies-08-00017" class="html-bibr">53</a>].</p>
Full article ">
25 pages, 466 KiB  
Article
Frequency of Planets in Binaries
by Mariangela Bonavita and Silvano Desidera
Galaxies 2020, 8(1), 16; https://doi.org/10.3390/galaxies8010016 - 18 Feb 2020
Cited by 27 | Viewed by 3709
Abstract
The frequency of planets in binaries is an important issue in the field of extrasolar planet studies because of its relevance in the estimation of the global planet population of our galaxy and the clues it can give to our understanding of planet [...] Read more.
The frequency of planets in binaries is an important issue in the field of extrasolar planet studies because of its relevance in the estimation of the global planet population of our galaxy and the clues it can give to our understanding of planet formation and evolution. Multiple stars have often been excluded from exoplanet searches, especially those performed using the radial velocity technique, due to the technical challenges posed by such targets. As a consequence and despite recent efforts, our knowledge of the frequency of planets in multiple stellar systems is still rather incomplete. On the other hand, the lack of knowledge about the binarity at the time of the compilation of the target samples means that our estimate of the planet frequency around single stars could be tainted by the presence of unknown binaries, especially if these objects have a different behavior in terms of planet occurrence. In a previous work we investigated the binarity of the objects included in the Uniform Detectability sample defined by Fisher and Valenti (2005), showing how more than 20% of their targets were, in fact, not single stars. Here, we present an update of this census, made possible mainly by the information now available thanks to the second Gaia Data Release. The new binary sample includes a total of 313 systems, of which 114 were added through this work. We were also able to significantly improve the estimates of masses and orbital parameters for most of the pairs in the original list, especially those at close separations. A few new systems with white dwarf companions were also identified. The results of the new analysis are in good agreement with the findings of our previous work, confirming the lack of difference in the overall planet frequency between binaries and single stars but suggesting a decrease in the planet frequency for very close pairs. Full article
(This article belongs to the Special Issue Habitability of Planets in Stellar Binary Systems)
Show Figures

Figure 1

Figure 1
<p>Mass vs. semi-major axes of companions compatible with the observed <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>μ</mi> </mrow> </semantics></math> for HD 67458 (<b>left panel</b>) and HD 80913 (<b>right panel</b>), calculated using the COPAINS (Code for Orbital Characterization of Astrometrically Inferred New Systems; see [<a href="#B47-galaxies-08-00016" class="html-bibr">47</a>] for details) and assuming a uniform distribution for the eccentricity. The red curve shows Gaia’s sensitivity limits [<a href="#B27-galaxies-08-00016" class="html-bibr">27</a>,<a href="#B50-galaxies-08-00016" class="html-bibr">50</a>]. The blue dot in the left panel shows the position of the companion to HD 67458 retrieved in Gaia DR2. The blue solid line in the right panel shows the position of the companions compatible with the RV trend observed by [<a href="#B38-galaxies-08-00016" class="html-bibr">38</a>] for HD 80913.</p>
Full article ">Figure 2
<p>Critical semi-major axis vs. mass ratio (<math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mi>A</mi> </msub> <mo>/</mo> <msub> <mi>M</mi> <mi>B</mi> </msub> </mrow> </semantics></math>) for the pairs in the Uniform Detectability (UD) sample with (stars) and without (filled circles) planetary companions. The new pairs not in BD07 are shown in light blue and orange, respectively.</p>
Full article ">Figure 3
<p>Histogram of the projected separation for the pairs in the updated (dark blue) and the original (light blue) UD binary sample. The number of stars with planets (P-Stars) in the final sample is shown in light orange. The solid line corresponds to <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>=</mo> <msup> <mn>2</mn> <mrow> <mo>″</mo> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Histogram of the critical semi-major axis (<math display="inline"><semantics> <msub> <mi>a</mi> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> </mrow> </msub> </semantics></math>, <b>left panel</b>) and mass ratio (<math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mi>B</mi> </msub> <mo>/</mo> <msub> <mi>M</mi> <mi>A</mi> </msub> </mrow> </semantics></math>, <b>right panel</b>) for the pairs in the updated (dark blue) and the original (light blue) UD binary sample. The number of stars with planets (P-Stars) in the final sample is shown in light orange.</p>
Full article ">Figure 5
<p><b>Left:</b> Fraction of planet-host binaries as a function of <math display="inline"><semantics> <msub> <mi>a</mi> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> </mrow> </msub> </semantics></math>. <b>Right:</b> Cumulative distribution of the <math display="inline"><semantics> <msub> <mi>a</mi> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> </mrow> </msub> </semantics></math> for UD pairs with (red) and without (blue) planetary companions.</p>
Full article ">
17 pages, 316 KiB  
Article
A Multi-Wavelength View of OJ 287 Activity in 2015–2017: Implications of Spectral Changes on Central-Engine Models and MeV-GeV Emission Mechanism
by Pankaj Kushwaha
Galaxies 2020, 8(1), 15; https://doi.org/10.3390/galaxies8010015 - 14 Feb 2020
Cited by 17 | Viewed by 3103
Abstract
A diverse range of observational results and peculiar properties across the domains of observation have made OJ 287 one of the best-explored BL Lac objects on the issues of relativistic jets and accretion physics as well as the strong theory of gravity. We [...] Read more.
A diverse range of observational results and peculiar properties across the domains of observation have made OJ 287 one of the best-explored BL Lac objects on the issues of relativistic jets and accretion physics as well as the strong theory of gravity. We here present a brief compilation of observational results from the literature and inferences/insights from the extensive studies but focus on the interpretation of its ∼12-yr quasi-periodic optical outbursts (QPOOs) and high energy emission mechanisms. The QPOOs in one model are attributed to the disk-impact related to dynamics of the binary SMBHs while alternative models attribute it to the geometrical effect related to the precession of a single jet or double jets. We discuss implications of the new spectral features reported during the 2015–2017 multi-wavelength high activity of the source—a break in the NIR-optical spectrum and hardening of the MeV-GeV emission accompanied by a shift in the location of its peak, in the context of the two. The reported NIR-optical break nicely fits the description of a standard accretion disk emission from an SMBH of mass 10 10 M while the time of its first appearance at the end of May, 2013 (MJD 56439) is in close coincidence with the time of impact predicted by the disk-impact binary SMBH model. This spectral and temporal coincidence with the model parameters of the disk-impact binary SMBH model provides independent evidence in favor of the model over the geometrical models which argue for a total central-engine mass in the range of 10 7 - 9 M . On the other hand, the MeV-GeV spectral change is naturally reproduced by the inverse Compton scattering of photons from the broad-line region and is consistent with the detection of broad emission lines during the previous cycles of quasi-periodic outbursts. Combining this with previous SED studies suggests that in, OJ 287, the MeV-GeV emission results from external Comptonization. Full article
(This article belongs to the Special Issue Monitoring the Non-Thermal Universe)
Show Figures

Figure 1

Figure 1
<p>Broadband and optical SEDs showing the spectral phases exhibited by OJ 287 to date. (<b>a</b>) Typical broadband SEDs of OJ 287 from a 2009 observation [<a href="#B28-galaxies-08-00015" class="html-bibr">28</a>] showing three different flux states: Flare (magenta, MJD: 55124-55131), moderate (green, MJD: 55131-55152) and quiescent (cyan, MJD: 55152-55184) MW emission. The solid curve is the total emission with synchrotron, SSC and EC-IR component shown respectively in dashed, dotted, and dot-dashed curves (see [<a href="#B28-galaxies-08-00015" class="html-bibr">28</a>], for modeling details). (<b>b</b>) Broadband SEDs of the source in its new spectral phase during 2016–2017 with magenta (MJD: 57359-57363; [<a href="#B60-galaxies-08-00015" class="html-bibr">60</a>]) showing a flare SED, black (MJD: 57786; [<a href="#B27-galaxies-08-00015" class="html-bibr">27</a>]) showing a typical SED during the VHE phase and red (MJD: 57871; [<a href="#B27-galaxies-08-00015" class="html-bibr">27</a>]) showing the quiescent SED state after the VHE activity with the lowest flux across EM spectrum. For reference/comparison, the quiescent SED (cyan) from panel (<b>a</b>) is also shown. The solid curves are the model produced spectrum (see [<a href="#B27-galaxies-08-00015" class="html-bibr">27</a>,<a href="#B60-galaxies-08-00015" class="html-bibr">60</a>], for details), while the dotted curve is the standard accretion-disk spectrum of a <math display="inline"><semantics> <mrow> <mo>∼</mo> <msup> <mn>10</mn> <mn>10</mn> </msup> <mspace width="3.33333pt"/> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math> SMBH; (<b>c</b>) NIR-optical spectrum highlighting the timing of tbe appearance of spectral break [<a href="#B60-galaxies-08-00015" class="html-bibr">60</a>]. The dashed curve is the accretion-disk spectrum drawn again for clarity. (<b>d</b>) NIR-optical SEDs before (MJD 57455.5), during and after the VHE activity of OJ 287, showing return to the typical power-law NIR-optical spectrum. For comparison, one of the SED (black) from (<b>c</b>) is also shown.</p>
Full article ">
8 pages, 711 KiB  
Article
Circular Geodesics Stability in a Static Black Hole in New Massive Gravity
by Andrés Aceña, Ericson López and Franklin Aldás
Galaxies 2020, 8(1), 14; https://doi.org/10.3390/galaxies8010014 - 14 Feb 2020
Cited by 3 | Viewed by 2294
Abstract
We study the existence and stability of circular geodesics in a family of asymptotically AdS static black holes in New Massive Gravity theory. We show that the mathematical sign of the hair parameter determines the existence of such geodesics. For a positive hair [...] Read more.
We study the existence and stability of circular geodesics in a family of asymptotically AdS static black holes in New Massive Gravity theory. We show that the mathematical sign of the hair parameter determines the existence of such geodesics. For a positive hair parameter, the stability regions follow the usual pattern, with the innermost geodesic being null, unstable, and separated from the horizon, followed by a region of unstable timelike geodesics and then a region of stable timelike geodesics, which extends in the asymptotic region. Full article
29 pages, 11021 KiB  
Review
Massive Star Formation in the Ultraviolet Observed with the Hubble Space Telescope
by Claus Leitherer
Galaxies 2020, 8(1), 13; https://doi.org/10.3390/galaxies8010013 - 9 Feb 2020
Cited by 12 | Viewed by 4817
Abstract
Spectroscopic observations of a massive star formation in the ultraviolet and their interpretation are reviewed. After a brief historical retrospective, two well-studied resolved star clusters and the surrounding H II regions are introduced: NGC 2070 in the Large Magellanic Cloud and NGC 604 [...] Read more.
Spectroscopic observations of a massive star formation in the ultraviolet and their interpretation are reviewed. After a brief historical retrospective, two well-studied resolved star clusters and the surrounding H II regions are introduced: NGC 2070 in the Large Magellanic Cloud and NGC 604 in M33. These regions serve as a training set for studies of more distant clusters, which can no longer be resolved into individual stars. Observations of recently formed star clusters and extended regions in star-forming galaxies in the nearby universe beyond the Local Group are presented. Their interpretation relies on spectral synthesis models. The successes and failures of such models are discussed, and future directions are highlighted. I present a case study of the extraordinary star cluster and giant H II region in the blue compact galaxy II Zw 40. The review concludes with a preview of two upcoming Hubble Space Telescope programs: ULLYSES, a survey of massive stars in nearby galaxies, and CLASSY, a study of massive star clusters in star-forming galaxies. Full article
(This article belongs to the Special Issue Star Formation in the Ultraviolet)
Show Figures

Figure 1

Figure 1
<p>International Ultraviolet Explorer (<span class="html-italic">IUE</span>) spectra of NGC 2070 (<b>left</b>) and NGC 604 (<b>right</b>), two extragalactic massive star clusters and the Large Magellanic Cloud (LMC) and M33, respectively. The spectra show the typical UV spectral features of hot, massive stars. From [<a href="#B4-galaxies-08-00013" class="html-bibr">4</a>,<a href="#B5-galaxies-08-00013" class="html-bibr">5</a>].</p>
Full article ">Figure 2
<p><span class="html-italic">IUE</span> low-dispersion spectrum of the nearby starburst galaxy NGC 1705, whose UV light is dominated by a bright super star cluster. The main observed and expected spectral features are identified at the top. The Milky Way foreground lines are labeled at the bottom. From [<a href="#B2-galaxies-08-00013" class="html-bibr">2</a>].</p>
Full article ">Figure 3
<p>Hubble Space Telescope/Space Telescope Imaging Spectrograph (<span class="html-italic">HST/STIS)</span> spectrum of the central super star cluster in NGC 1705. The spectra were taken with the echelle gratings E140M and E230M, which have resolving powers of 46,000 and 30,000, respectively. Compare the quality of the spectrum to that shown in <a href="#galaxies-08-00013-f002" class="html-fig">Figure 2</a>. From [<a href="#B6-galaxies-08-00013" class="html-bibr">6</a>].</p>
Full article ">Figure 4
<p><span class="html-italic">HST/STIS</span> 52″ × 0.2″ apertures overlaid on an <span class="html-italic">HST</span> image of R136. The central region is shown in the insert at the top right, which has a radius of 2.05″, corresponding to 0.5 parsec at the distance of the LMC. From [<a href="#B15-galaxies-08-00013" class="html-bibr">15</a>].</p>
Full article ">Figure 5
<p>Left: the 2″- (corresponding to 8 pc) wide aperture of <span class="html-italic">STIS</span> superposed on a U image of NGC 604. Right: the <span class="html-italic">STIS</span> spectral image of OB stars in NGC 604, showing the individually resolved spectra. The vertical stripe to the left is geocoronal Ly-α. From [<a href="#B20-galaxies-08-00013" class="html-bibr">20</a>].</p>
Full article ">Figure 6
<p>Left: examples of <span class="html-italic">STIS</span>/G140L UV spectra of the most luminous stars in NGC 604. Right: a Hertzsprung-Russell diagram of the most massive stars in NGC 604. The upper and lower figure are for two different calibrations of spectral type versus stellar parameters. From [<a href="#B22-galaxies-08-00013" class="html-bibr">22</a>].</p>
Full article ">Figure 7
<p>Left: an <span class="html-italic">STIS</span> UV image of the star-forming region near the center of M83. Many UV-bright star <span class="html-italic">clusters</span> can be seen. Compare with <a href="#galaxies-08-00013-f005" class="html-fig">Figure 5</a>, which shows individual <span class="html-italic">stars</span>. The field size is approximately 280 pc × 280 pc. Right: the corresponding spectra of the clusters obtained by using slitless spectroscopy in the region with <span class="html-italic">STIS</span>. From [<a href="#B24-galaxies-08-00013" class="html-bibr">24</a>].</p>
Full article ">Figure 8
<p>Composite spectrum constructed from a sample of star-cluster spectra obtained with the Faint Object Spectrograph (<span class="html-italic">FOS</span>) and the Goddard High Resolution Spectrograph (<span class="html-italic">GHRS</span>) onboard <span class="html-italic">HST</span>. Strong interstellar lines are identified below the spectrum; stellar-wind and photospheric lines are labeled above. The wind lines are marked with a ″W″. Nebular emission lines are identified with an ″E″. From [<a href="#B31-galaxies-08-00013" class="html-bibr">31</a>].</p>
Full article ">Figure 9
<p><span class="html-italic">HST</span> UV images of the star-forming galaxies hosting many massive young clusters. The overlaid circles indicate the 2.5″ aperture of the <span class="html-italic">HST Cosmic Origins Spectrograph</span> (<span class="html-italic">COS</span>). The entrance aperture encompasses a single (NG 4214) or multiple clusters (e.g., I Zw 18). The linear diameter of the field enclosed by the aperture is (from left to right, top to bottom): 220 pc, 650 pc, 165 pc, 37 pc, 46 pc, 280 pc, 46 pc, 587 pc, 58 pc. From [<a href="#B32-galaxies-08-00013" class="html-bibr">32</a>].</p>
Full article ">Figure 10
<p><span class="html-italic">COS</span> G130M spectra of the star clusters seen in the images in <a href="#galaxies-08-00013-f009" class="html-fig">Figure 9</a>. The spectra are arranged in order of increasing oxygen abundance from top to bottom. The major absorption lines in each galaxy are labeled in the uppermost panel. From [<a href="#B32-galaxies-08-00013" class="html-bibr">32</a>].</p>
Full article ">Figure 11
<p>Schematic showing the principal components of population synthesis models. (<b>a</b>): input IMF, isochrones, and spectral library; (<b>b</b>): adopted star-formation history, single stellar population (SSP) calculated from the top row ingredients, and dust attenuation; (<b>c</b>): final spectrum for a continuously forming population (CSP) by combining the components in the middle row. From [<a href="#B43-galaxies-08-00013" class="html-bibr">43</a>].</p>
Full article ">Figure 12
<p>Comparison of the average spectrum of the 46 local star-forming regions shown in <a href="#galaxies-08-00013-f008" class="html-fig">Figure 8</a> (dark spectrum) and a simulated spectrum following a Salpeter IMF with stars up to 100 M<sub>⊙</sub> (light spectrum). Models and data of the wind features of N V 1240, Si IV 1400, and C IV 1550 are in excellent agreement. The narrow absorption features (e.g., at 1260 Å, 1335 Å, and blended with the Si IV 1400 P Cygni profile) are interstellar and are not modeled in the synthetic spectrum. From [<a href="#B55-galaxies-08-00013" class="html-bibr">55</a>].</p>
Full article ">Figure 13
<p>Comparison of the simulated spectra of a single stellar population with a solar chemical composition (thick lines) and an observational <span class="html-italic">IUE</span> library of Galactic stars (thin). The age steps from 1 to 20 Myr are labeled on the right. From [<a href="#B59-galaxies-08-00013" class="html-bibr">59</a>].</p>
Full article ">Figure 14
<p>Comparison of evolution models at solar metallicity with rotation velocities on the zero-age main-sequence of 0 and 550 km s<sup>−1</sup>. The velocities may be higher than actually observed in stars but were chosen to illustrate the effect of rotation on the evolutionary tracks. From [<a href="#B62-galaxies-08-00013" class="html-bibr">62</a>].</p>
Full article ">Figure 15
<p>Comparison of the UV spectra of R136a, the central region of 30 Doradus, and the massive star cluster #5 in NGC 5253. Note the strong, broad He II 1640 line, which indicates He-enriched stars with dense stellar winds. From [<a href="#B70-galaxies-08-00013" class="html-bibr">70</a>].</p>
Full article ">Figure 16
<p><span class="html-italic">STIS</span> UV (left) and Advanced Camera for Surveys (<span class="html-italic">ACS</span>) Hα (right) images of the star cluster and H II region in II Zw 40. Table 1600. Å. The circle denotes the position of the <span class="html-italic">COS</span> aperture, which encompasses a physical area of 135 pc in diameter. From [<a href="#B76-galaxies-08-00013" class="html-bibr">76</a>].</p>
Full article ">Figure 17
<p>Observed UV spectrum of SSC-N. No correction for reddening or redshift was applied. Line identifications are above each spectrum. From [<a href="#B76-galaxies-08-00013" class="html-bibr">76</a>].</p>
Full article ">Figure 18
<p>Comparison of the observed spectrum of SSC-N (black) and the best-fit simulation (blue). The data are in restframe wavelengths and have been corrected for the Milky Way and internal reddening. The model assumes a star cluster of mass 9.1 × 10<sup>5</sup> M<sub>⊙</sub>, age 2.8 × 10<sup>6</sup> yr and chemical composition of 1/7<sup>th</sup> solar. For comparison, the best-fit model with solar chemical composition is shown as well (red, dashed). From [<a href="#B76-galaxies-08-00013" class="html-bibr">76</a>].</p>
Full article ">Figure 19
<p>Equivalent width of stellar He II 1640 versus time. The value for SSC-N is the open symbol with error bars. The lines show the predictions from four different simulations: solar and subsolar compositions with and without rotation. The predictions for the two subsolar models are very small at any time and are therefore not visible in the graph. From [<a href="#B76-galaxies-08-00013" class="html-bibr">76</a>].</p>
Full article ">Figure 20
<p>Left: Baldwin Phillips &amp; Terlevich (BPT) diagram comparing II Zw 40 to other samples: the BCD sample of [<a href="#B30-galaxies-08-00013" class="html-bibr">30</a>]; extreme Green Peas of [<a href="#B92-galaxies-08-00013" class="html-bibr">92</a>]; Mrk 71 [<a href="#B93-galaxies-08-00013" class="html-bibr">93</a>]; <span class="html-italic">z </span>≈ 2–3 galaxies of [<a href="#B94-galaxies-08-00013" class="html-bibr">94</a>]. Solid black line: an extreme starburst classification line from [<a href="#B95-galaxies-08-00013" class="html-bibr">95</a>]. The grayscale 2D histogram indicates the density of star-forming galaxies in SDSS [<a href="#B91-galaxies-08-00013" class="html-bibr">91</a>]. Right: the computed line ratios of Si III]/C III] versus O III]/C III], using the models of [<a href="#B96-galaxies-08-00013" class="html-bibr">96</a>]. Blue lines: calculations with a constant ionization parameter, from log <span class="html-italic">U</span><sub>0</sub> = −1 (dark blue) to log <span class="html-italic">U</span><sub>0</sub> = −4 (light blue). Calculations with constant oxygen abundance are connected with lines having different colors, from dark purple to orange. II Zw 40: purple star; BCD sample of [<a href="#B30-galaxies-08-00013" class="html-bibr">30</a>]: yellow circles; <span class="html-italic">z</span> ≈ 2 galaxies of [<a href="#B97-galaxies-08-00013" class="html-bibr">97</a>]: grey squares. From [<a href="#B76-galaxies-08-00013" class="html-bibr">76</a>].</p>
Full article ">
16 pages, 556 KiB  
Article
Entropy and Mass Distribution in Disc Galaxies
by John Herbert Marr
Galaxies 2020, 8(1), 12; https://doi.org/10.3390/galaxies8010012 - 8 Feb 2020
Cited by 6 | Viewed by 3722
Abstract
The relaxed motion of stars and gas in galactic discs is well approximated by a rotational velocity that is a function of radial position only, implying that individual components have lost any information about their prior states. Thermodynamically, such an equilibrium state is [...] Read more.
The relaxed motion of stars and gas in galactic discs is well approximated by a rotational velocity that is a function of radial position only, implying that individual components have lost any information about their prior states. Thermodynamically, such an equilibrium state is a microcanonical ensemble with maximum entropy, characterised by a lognormal probability distribution. Assuming this for the surface density distribution yields rotation curves that closely match observational data across a wide range of disc masses and galaxy types and provides a useful tool for modelling the theoretical density distribution in the disc. A universal disc spin parameter emerges from the model, giving a tight virial mass estimator with strong correlation between angular momentum and disc mass, suggesting a mechanism by which the proto-disc developed by dumping excess mass to the core or excess angular momentum to a satellite galaxy. The baryonic-to-dynamic mass ratio for the model approaches unity for high mass galaxies, but is generally <1 for low mass discs, and this discrepancy appears to follow a similar relationship to that shown in recent work on the Radial Acceleration Relation (RAR). Although this may support Modified Newtonian Dynamics (MOND) in preference to a Dark Matter (DM) halo, it does not exclude undetected baryonic mass or a gravitational DM component in the disc. Full article
Show Figures

Figure 1

Figure 1
<p>NGC 2366’s (<b>a</b>) Rotational velocity for the LN model (solid line) and observations ([<a href="#B38-galaxies-08-00012" class="html-bibr">38</a>]). (<b>b</b>) Radial distribution of density (<math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math>/kpc<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math>) for the LN model (solid line) and Freeman exponential disc model (dashed line).</p>
Full article ">Figure 2
<p>Rotational Curve (RC) for the massive galaxy M31, with rotational velocity for the LN model (solid line) and observations with error bars ([<a href="#B39-galaxies-08-00012" class="html-bibr">39</a>,<a href="#B40-galaxies-08-00012" class="html-bibr">40</a>,<a href="#B41-galaxies-08-00012" class="html-bibr">41</a>]).</p>
Full article ">Figure 3
<p>Log-log plot of angular momentum vs. disc mass for the LN model galaxies (squares) and the theoretical slope 5/3 (purple solid line). Idealised evolutionary changes are shown for a disc with excess mass dumping mass into the core while conserving angular momentum (R–S ); and a disc with excess angular momentum shedding mass and angular momentum to a satellite galaxy, leaving a less massive disc (P–Q ).</p>
Full article ">Figure 4
<p>Log-log plot of the modelled lognormal dynamic disc mass vs. the estimated observational baryonic masses (<math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mi>g</mi> </msub> <mo>+</mo> <msup> <mi>M</mi> <mo>*</mo> </msup> </mrow> </semantics></math>) for a wide range of galaxy masses and types. The solid line is the line of mass equality; the dashed line is for <math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mrow> <mi>d</mi> <mi>y</mi> <mi>n</mi> <mi>a</mi> <mi>m</mi> <mi>i</mi> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mi>f</mi> <mrow> <mo>(</mo> <msub> <mi>M</mi> <mrow> <mi>b</mi> <mi>a</mi> <mi>r</mi> <mi>y</mi> <mi>o</mi> <mi>n</mi> <mi>i</mi> <mi>c</mi> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math>, with <math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mo>†</mo> </msub> <mo>=</mo> <mn>3</mn> <mo>.</mo> <mn>98</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>10</mn> </msup> <mspace width="3.33333pt"/> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math> (see the text).</p>
Full article ">
7 pages, 1113 KiB  
Article
The Microvariable Activity of BL Lacertae
by Alberto C. Sadun, Masoud Asadi-Zeydabadi, Lauren Hindman and J. Ward Moody
Galaxies 2020, 8(1), 11; https://doi.org/10.3390/galaxies8010011 - 7 Feb 2020
Cited by 2 | Viewed by 2610
Abstract
We report on seven nights of optical observation taken over a two-week period, and the resultant analysis of the intermediate-frequency peaked BL Lac object (IBL), BL Lac itself, at redshift z = 0.069. The microvariable behavior can be confirmed over the course of [...] Read more.
We report on seven nights of optical observation taken over a two-week period, and the resultant analysis of the intermediate-frequency peaked BL Lac object (IBL), BL Lac itself, at redshift z = 0.069. The microvariable behavior can be confirmed over the course of minutes for each night. A relativistic beaming model was used in our analysis, to infer changes to the line of sight angles for the motion of the different relativistic components. This model has very few free parameters. The light curves we generated show both high and moderate frequency cadence to the variable behavior of BL Lac itself, in addition to the well documented long-term variability. Full article
Show Figures

Figure 1

Figure 1
<p>Light curves (smoothed) of BL Lac over seven nights in October 2017.</p>
Full article ">Figure 2
<p>Light curves of BL Lac for the entire night of Oct 27, (<b>a</b>) BL Lac variability, and (<b>b</b>) variability of the comparison stars.</p>
Full article ">Figure 3
<p>Geometric model showing the source emitting independent optical components which result in relativistic beaming. The Lorentz factor γ is taken to be 10 [<a href="#B12-galaxies-08-00011" class="html-bibr">12</a>], and the angle to the line of sight is designated θ, as shown.</p>
Full article ">Figure 4
<p>Histogram of the luminosity magnification of the microvariable outbursts of BL Lac.</p>
Full article ">Figure 5
<p>Histogram of the angle deviations of the microvariable outbursts of BL Lac.</p>
Full article ">
29 pages, 1376 KiB  
Review
Radiation-Driven Stellar Eruptions
by Kris Davidson
Galaxies 2020, 8(1), 10; https://doi.org/10.3390/galaxies8010010 - 5 Feb 2020
Cited by 13 | Viewed by 3243
Abstract
Very massive stars occasionally expel material in colossal eruptions, driven by continuum radiation pressure rather than blast waves. Some of them rival supernovae in total radiative output, and the mass loss is crucial for subsequent evolution. Some are supernova impostors, including SN precursor [...] Read more.
Very massive stars occasionally expel material in colossal eruptions, driven by continuum radiation pressure rather than blast waves. Some of them rival supernovae in total radiative output, and the mass loss is crucial for subsequent evolution. Some are supernova impostors, including SN precursor outbursts, while others are true SN events shrouded by material that was ejected earlier. Luminous Blue Variable stars (LBV’s) are traditionally cited in relation with giant eruptions, though this connection is not well established. After four decades of research, the fundamental causes of giant eruptions and LBV events remain elusive. This review outlines the basic relevant physics, with a brief summary of essential observational facts. Reasons are described for the spectrum and emergent radiation temperature of an opaque outflow. Proposed mechanisms are noted for instabilities in the star’s photosphere, in its iron opacity peak zones, and in its central region. Various remarks and conjectures are mentioned, some of them relatively unfamiliar in the published literature. Full article
(This article belongs to the Special Issue Luminous Stars in Nearby Galaxies)
Show Figures

Figure 1

Figure 1
<p>The empirical upper boundary and LBV instability strip in the Hertzsprung-Russell Diagram. In reality they are ill-defined and may depend on rotation and chemical composition. The interval between the LBV strip and the boundary is very uncertain. The zero-age main sequence on the left side shows initial masses, and most of a very massive star’s evolution occurs at roughly twice the initial luminosity.</p>
Full article ">Figure 2
<p>Luminosity record of SN2011ht based on visual-wavelength brightness [<a href="#B42-galaxies-08-00010" class="html-bibr">42</a>]. The vertical scale, expressed in solar units, neglects variations in the bolometric correction but is adequate for conceptual purposes. The relative faintness after <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>∼</mo> <mn>130</mn> </mrow> </semantics></math> d is highly abnormal if this object was a true supernova.</p>
Full article ">Figure 3
<p>Spectrum of SN2011ht at three different times [<a href="#B42-galaxies-08-00010" class="html-bibr">42</a>], cf. <a href="#galaxies-08-00010-f002" class="html-fig">Figure 2</a>. Both scales are logarithmic, the three tracings have differing vertical offsets, and the marks near 6100 Å indicate <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mi>λ</mi> </msub> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>14</mn> </mrow> </msup> </mrow> </semantics></math> erg cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math> s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> Å<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. Gaps at <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>&gt;</mo> <mn>7000</mn> </mrow> </semantics></math> Å are obscured by terrestrial atmospheric features. Concerning the line profiles, see <a href="#sec4dot2-galaxies-08-00010" class="html-sec">Section 4.2</a> and Figure 6.</p>
Full article ">Figure 4
<p>Photosphere temperatures in simplified opaque outflow models with <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mn>10</mn> <mn>8</mn> </msup> <mspace width="0.166667em"/> <msub> <mi>L</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>V</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math> km s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. The curves show temperatures corresponding to thermalization depths of 1 and 2. <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>(</mo> <msub> <mi>τ</mi> <mi>th</mi> </msub> <mo>)</mo> </mrow> </semantics></math> depends approximately on the quantity <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>M</mi> <mo>˙</mo> </mover> <msup> <mi>V</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <msup> <mi>L</mi> <mrow> <mo>−</mo> <mn>0.67</mn> </mrow> </msup> </mrow> </semantics></math>. Temperature values here are imprecise and very likely overestimated, because the models are highly idealized; see text.</p>
Full article ">Figure 5
<p>Radii of the photospheric locations shown in <a href="#galaxies-08-00010-f004" class="html-fig">Figure 4</a>. For large <math display="inline"><semantics> <mover accent="true"> <mi>M</mi> <mo>˙</mo> </mover> </semantics></math> the photosphere becomes geometrically thin (small <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>r</mi> <mo>/</mo> <mi>r</mi> </mrow> </semantics></math>) and thus resembles a plane-parallel model.</p>
Full article ">Figure 6
<p>Emission line profile with moderate Thomson scattering. The dashed curve on the left side is a mirror image of the right side. Note that the line wings extend far beyond the velocity indicated by P Cyg absorption. This example is H<math display="inline"><semantics> <mi>α</mi> </semantics></math> in the radiation-driven outflow of SN 2011ht [<a href="#B42-galaxies-08-00010" class="html-bibr">42</a>], but other giant eruptions produce similar line shapes.</p>
Full article ">Figure 7
<p>Sketch of the mass column density at temperatures below <span class="html-italic">T</span>, in a stellar atmosphere and in a dense outflow. These curves are based merely on idealized textbook-style models of <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>(</mo> <mi>r</mi> <mo>)</mo> </mrow> </semantics></math>, but the difference between them is qualitatively valid.</p>
Full article ">
15 pages, 1195 KiB  
Review
Rotating Disk Galaxies without Dark Matter Based on Scientific Reasoning
by James Q. Feng
Galaxies 2020, 8(1), 9; https://doi.org/10.3390/galaxies8010009 - 1 Feb 2020
Cited by 6 | Viewed by 4983
Abstract
The most cited evidence for (non-baryonic) dark matter has been an apparent lack of visible mass to gravitationally support the observed orbital velocity of matter in rotating disk galaxies, yet measurement of the mass of celestial objects cannot be straightforward, requiring theories derived [...] Read more.
The most cited evidence for (non-baryonic) dark matter has been an apparent lack of visible mass to gravitationally support the observed orbital velocity of matter in rotating disk galaxies, yet measurement of the mass of celestial objects cannot be straightforward, requiring theories derived from the known physical laws along with some empirically established semi-quantitative relationship. The most reliable means for determining the mass distribution in rotating disk galaxies is to solve a force balance equation according to Newton’s laws from measured rotation curves, similar to calculating the Sun’s mass from the Earth’s orbital velocity. Another common method to estimate galactic mass distribution is to convert measured brightness from surface photometry based on empirically established mass-to-light ratio. For convenience, most astronomers commonly assumed a constant mass-to-light ratio for estimation of the so-called “luminous” or “visible” mass, which would not likely be accurate. The mass determined from a rotation curve typically exhibits an exponential-like decline with galactrocentric distance, qualitatively consistent with observed surface brightness but often with a larger disk radial scale length. This fact scientifically suggests variable mass-to-light ratio of baryonic matter in galaxies without the need for dark matter. Full article
Show Figures

Figure 1

Figure 1
<p>Photographic images of circular thin-disk galaxies with small, amorphous, centrally located bulge.</p>
Full article ">Figure 2
<p>Definition sketch of the rotating thin-disk galaxy model, where the mass is assumed to distribute axisymmetrically in the circular disk of uniform thickness <span class="html-italic">h</span> with a variable density <span class="html-italic">ρ</span> and rotation velocity <span class="html-italic">V</span> as functions of the radial distance from galactic center <span class="html-italic">r</span> (but independent of the polar angle <span class="html-italic">θ</span>).</p>
Full article ">Figure 3
<p>Profiles of the Milky Way rotation velocity <span class="html-italic">V</span>(r) and mass density <span class="html-italic">ρ</span>(r) for the disk portion and bulge portion as noted with the thick line as a reference from the pure disk model without a bulge (taken from Figure 7 of Ref. [<a href="#B21-galaxies-08-00009" class="html-bibr">21</a>]). Noteworthy here is that the portion of mass density profile (shown with the thick line, as roughly a combined mass density profile from both disk and bulge) for <span class="html-italic">r</span> in the interval [0.1, 0.9] appears nearly linear in the semi-log plot (when when the abruptly varying ends around <span class="html-italic">r</span> = 0 and 1 are trimmed out), indicating an approximately exponential decline of mass density with galactocentric distance.</p>
Full article ">
3 pages, 170 KiB  
Editorial
Acknowledgement to Reviewers of Galaxies in 2019
by Galaxies Editorial Office
Galaxies 2020, 8(1), 8; https://doi.org/10.3390/galaxies8010008 - 27 Jan 2020
Viewed by 1438
Abstract
The editorial team greatly appreciates the reviewers who have dedicated their considerable time and expertise to the journal’s rigorous editorial process over the past 12 months, regardless of whether the papers are finally published or not [...] Full article
5 pages, 974 KiB  
Article
Investigating Multiwavelength Lognormality with Simulations—Case of Mrk 421
by Nachiketa Chakraborty
Galaxies 2020, 8(1), 7; https://doi.org/10.3390/galaxies8010007 - 16 Jan 2020
Cited by 5 | Viewed by 2238
Abstract
Blazars are highly variable and display complex characteristics. A key characteristic is the flux probability distribution function or flux PDF whose shape depends upon the form of the underlying physical process driving variability. The BL Lacertae Mrk 421 is one of the brightest [...] Read more.
Blazars are highly variable and display complex characteristics. A key characteristic is the flux probability distribution function or flux PDF whose shape depends upon the form of the underlying physical process driving variability. The BL Lacertae Mrk 421 is one of the brightest and most variable blazars across the electromagnetic spectrum. It has been reported to show hints of lognormality across the spectrum from radio to gamma-ray histograms of observed fluxes. This would imply that the underlying mechanisms may not conform to the “standard” additive, multi-zone picture, but could potentially have multiplicative processes. This is investigated by testing the observed lightcurves at different wavelengths with time-series simulations. We find that the simulations reveal a more complex scenario, than a single lognormal distribution explaining the multiwavelength lightcurves of Mrk 421. Full article
Show Figures

Figure 1

Figure 1
<p>Figure shows the histogram of the observed Fermi-LAT lightcurves compared to those for simulations. Using simulations confidence intervals (dashed red) are derived at the 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> level for each bin. (<b>a</b>) Simulations of the fluxes that are normally distributed. (<b>b</b>) Simulations of the logarithm of fluxes that are normally distributed or lognormal simulations. The lognormal simulations fit the lightcurves significantly better with SW <span class="html-italic">p</span>-values of <math display="inline"><semantics> <mrow> <mn>0.013</mn> </mrow> </semantics></math> relative to <math display="inline"><semantics> <mrow> <mn>1.78</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>14</mn> </mrow> </msup> </mrow> </semantics></math> or <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>≪</mo> <mn>0.001</mn> </mrow> </semantics></math> for the normal case.</p>
Full article ">Figure 2
<p>Similar to <a href="#galaxies-08-00007-f001" class="html-fig">Figure 1</a> for BAT lightcurves. Both (<b>a</b>) normal simulations and (<b>b</b>) lognormal ones are not great fits with SW <span class="html-italic">p</span>-values of <math display="inline"><semantics> <mrow> <mn>8.29</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>16</mn> </mrow> </msup> </mrow> </semantics></math> (i.e., <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>≪</mo> <mn>0.001</mn> </mrow> </semantics></math>) and <math display="inline"><semantics> <mrow> <mn>1.30</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math> (also, <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>≪</mo> <mn>0.001</mn> </mrow> </semantics></math>), even though the latter appears to be better. This mismatch is potentially due to the multi-modal, bursty structure.</p>
Full article ">Figure 3
<p>Similar to <a href="#galaxies-08-00007-f001" class="html-fig">Figure 1</a> for OVRO lightcurves. Between (<b>a</b>) normal simulations and (<b>b</b>) lognormal ones, the normal distribution is a somewhat better fit with SW <span class="html-italic">p</span>-values of <math display="inline"><semantics> <mrow> <mn>2.31</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math> relative to <math display="inline"><semantics> <mrow> <mn>4.80</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math>. Once again the multi-modal, bursty structure makes it complex to explain with a single model for PDF.</p>
Full article ">
28 pages, 737 KiB  
Review
Applications of Stellar Population Synthesis in the Distant Universe
by Elizabeth R. Stanway
Galaxies 2020, 8(1), 6; https://doi.org/10.3390/galaxies8010006 - 8 Jan 2020
Cited by 6 | Viewed by 4793
Abstract
Comparison with artificial galaxy models is essential for translating the incomplete and low signal-to-noise data we can obtain on astrophysical stellar populations to physical interpretations which describe their composition, physical properties, histories and internal conditions. In particular, this is true for distant galaxies, [...] Read more.
Comparison with artificial galaxy models is essential for translating the incomplete and low signal-to-noise data we can obtain on astrophysical stellar populations to physical interpretations which describe their composition, physical properties, histories and internal conditions. In particular, this is true for distant galaxies, whose unresolved light embeds clues to their formations and evolutions, and their impacts on their wider environs. Stellar population synthesis models are now used as the foundation of analysis at all redshifts, but are not without their problems. Here we review the use of stellar population synthesis models, with a focus on applications in the distant Universe. Full article
(This article belongs to the Special Issue Star Formation in the Ultraviolet)
Show Figures

Figure 1

Figure 1
<p>The cosmic, volume-averaged star formation rate density history as a function of redshift. Data points show the compilation of (Madau and Dickinson [<a href="#B10-galaxies-08-00006" class="html-bibr">10</a>]; see references therein as well), with infrared-derived and dust-corrected ultraviolet-derived star formation rates distinguished by colour. The volume-averaged flux measured by each data set has been calibrated by Madau and Dickinson [<a href="#B10-galaxies-08-00006" class="html-bibr">10</a>] against stellar population synthesis models. UV-derived rates are also subject to large corrections for assumed dust extinction, which remains uncertain at the highest redshifts. The rise in volumetric star formation rate from Cosmic Dawn to the peak at Cosmic Noon, followed by a decline as star formation downsizes, is clearly visible.</p>
Full article ">Figure 2
<p>The impact of the assumed initial mass function (IMF), age, metallicity and binary fraction on the far-ultraviolet (ionising) emission from a stellar population produced by BPASS v2.2 [<a href="#B59-galaxies-08-00006" class="html-bibr">59</a>]. <b>Top Left</b>: the effect of age for a population incorporating binary models at Z = 0.1 Z<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math>, which has been forming stars continuously at the rate of 1 M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math> per year for different lengths of star bursts. The assumed initial mass function is a broken power law with a slope of −1.35 between 0.5 and 300 M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math> and −0.3 between 0.1 and 0.5 M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math>. Top right: the effect on a 10 million year old binary population with the same IMF of altering the stellar metallicity. Bottom left: the effect on a 10 million year old binary population at Z = 0.1 Z<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math> of altering the upper slope of the IMF, or replacing the low mass break with an exponential (Chabrier [<a href="#B60-galaxies-08-00006" class="html-bibr">60</a>]) cut-off. <b>Bottom right</b>: The effect on the same population of altering the upper mass cut-off and of removing the lower break (i.e., a constant IMF slope of −1.35 between 0.1 and 100 M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math>) on populations incorporating only single or single and binary stellar evolution pathways.</p>
Full article ">Figure 3
<p>The power of ultraviolet line diagnostics to diagnose the physical conditions in ionised gas. In each case the same ionising spectrum (a continuously star forming stellar population at 100 million years old) is processed through a nebular gas cloud with the same gas density (log(<math display="inline"><semantics> <msub> <mi>n</mi> <mi>e</mi> </msub> </semantics></math>/cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>) = 2.3) and a range of ionisation parameters as labelled on the figures (decreasing this parameter is broadly equivalent to increasing the mean distance of the cloud from the source or decreasing the star formation rate). In the first three panels we show strong ultraviolet and optical emission line ratios which may be used as ionisation and metallicity diagnostics. On the lower right panel we demonstrate the importance of including dust grain surface physics (colour) or omitting it (greyscale) in an otherwise identical radiative transfer model (a subset of ionisation parameters and metallicities is shown for clarity). While the effect at low ionisation parameters is small, the higher ionisation parameters typical of the high star formation densities and binary populations seen in the distant Universe are very sensitive to the assumed physics. All models are from BPASS v2.2 [<a href="#B59-galaxies-08-00006" class="html-bibr">59</a>], processed with CLOUDY. The metallicity of each line is labelled with “zem4”, “zem5”, “z001”, etc., indicating metallicity mass fractions of <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> <mo>,</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> <mo>,</mo> <mn>0.001</mn> </mrow> </semantics></math> etc.</p>
Full article ">Figure 4
<p>The effects of these stellar population choices on diagnostic diagrams. Models compared were the BPASS models as before (orange) and the nebular synthesis models of Gutkin et al. [<a href="#B142-galaxies-08-00006" class="html-bibr">142</a>] (blue, based on the 2016 version of the GalaxEv models of BC03). BPASS models assume a broken power law IMF, with an upper slope of −1.35 and an upper mass cut-off at 300 M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math>, while the nebular gas electron density is 300 cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>. The Gutkin et al. [<a href="#B142-galaxies-08-00006" class="html-bibr">142</a>] model assumes an IMF with the same slope and upper mass limit, but a Chabrier [<a href="#B60-galaxies-08-00006" class="html-bibr">60</a>] lower mass cut-off for <math display="inline"><semantics> <mrow> <msub> <mi>n</mi> <mi>e</mi> </msub> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math>, [C/O] = 1 and a dust-to-metal ratio of 0.5. Both models assume constant star formation and are shown at matched metallicities. The flux in the ultraviolet doublets was summed before ratios were taken. The primary difference was in the handling of stellar evolution and atmosphere.</p>
Full article ">
14 pages, 901 KiB  
Article
Constraints to Dark Matter Annihilation from High-Latitude HAWC Unidentified Sources
by Javier Coronado-Blázquez and Miguel A. Sánchez-Conde
Galaxies 2020, 8(1), 5; https://doi.org/10.3390/galaxies8010005 - 30 Dec 2019
Cited by 9 | Viewed by 3502
Abstract
The Λ CDM cosmological framework predicts the existence of thousands of subhalos in our own Galaxy not massive enough to retain baryons and become visible. Yet, some of them may outshine in gamma rays provided that the dark matter is made of weakly [...] Read more.
The Λ CDM cosmological framework predicts the existence of thousands of subhalos in our own Galaxy not massive enough to retain baryons and become visible. Yet, some of them may outshine in gamma rays provided that the dark matter is made of weakly interacting massive particles (WIMPs), which would self-annihilate and would appear as unidentified gamma-ray sources (unIDs) in gamma-ray catalogs. Indeed, unIDs have proven to be competitive targets for dark matter searches with gamma rays. In this work, we focus on the three high-latitude ( | b | 10 ) sources present in the 2HWC catalog of the High Altitude Water Cherenkov (HAWC) observatory with no clear associations at other wavelengths. Indeed, only one of these sources, 2HWC J1040+308, is found to be above the HAWC detection threshold when considering 760 days of data, i.e., a factor 1.5 more exposure time than in the original 2HWC catalog. Other gamma-ray instruments, such as Fermi-LAT or VERITAS at lower energies, do not detect the source. Also, this unID is reported as spatially extended, making it even more interesting in a dark matter search context. While waiting for more data that may shed further light on the nature of this source, we set competitive upper limits on the annihilation cross section by comparing this HAWC unID to expectations based on state-of-the-art N-body cosmological simulations of the Galactic subhalo population. We find these constraints to be particularly competitive for heavy WIMPs, i.e., masses above ∼25 (40) TeV in the case of the b b ¯ ( τ + τ ) annihilation channel, reaching velocity-averaged cross section values of 2 × 10 25 ( 5 × 10 25 ) cm 3 ·s 1 . Although far from testing the thermal relic cross section value, the obtained limits are independent and nicely complementary to those from radically different DM analyses and targets, demonstrating once again the high potential of this DM search approach. Full article
(This article belongs to the Special Issue The Role of Halo Substructure in Gamma-Ray Dark Matter Searches)
Show Figures

Figure 1

Figure 1
<p>VHE sources listed in the TeVCat online catalog as of November 2019, here shown in Hammer-Aitoff projection and Galactic coordinates. Background is the gamma-ray sky as seen by Fermi-LAT in 4 years of operation. Note the higher density of sources, in particular of unIDs, along the Galactic plane. The figure was generated with the TeVCat online tool [<a href="#B44-galaxies-08-00005" class="html-bibr">44</a>].</p>
Full article ">Figure 2
<p>HAWC minimum flux, <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math>, needed to reach source detection (<math display="inline"><semantics> <mrow> <mi>T</mi> <mi>S</mi> <mo>=</mo> <mn>25</mn> </mrow> </semantics></math>) at a sky declination equal to <math display="inline"><semantics> <mrow> <msup> <mn>30.87</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. In blue, the <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math> as originally reported in [<a href="#B62-galaxies-08-00005" class="html-bibr">62</a>] for the case of a point source described by a power law spectrum with index <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Γ</mi> <mo>=</mo> <mn>2.63</mn> </mrow> </semantics></math> and 507 days of exposure. The red curve shows the <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math> for our DM subhalo candidate instead (with a spatial extension of <math display="inline"><semantics> <mrow> <msup> <mn>0.5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Γ</mi> <mo>=</mo> <mn>2.08</mn> </mrow> </semantics></math> and 760 days of exposure). This one is computed taking into account (i) the improvement factor of ∼5 derived for this unID at a pivot energy of 7 TeV [<a href="#B46-galaxies-08-00005" class="html-bibr">46</a>], marked in the figure as a dashed grey vertical line; (ii) the difference in exposure times, which makes the <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math> improve by a factor <math display="inline"><semantics> <mrow> <msqrt> <mrow> <mn>507</mn> <mo>/</mo> <mn>760</mn> </mrow> </msqrt> <mo>=</mo> <mn>0.82</mn> </mrow> </semantics></math>; and (iii) a factor ∼2 worsening due to the extension of the source, according to [<a href="#B34-galaxies-08-00005" class="html-bibr">34</a>]. See text for further details on these computations. The overall <math display="inline"><semantics> <msub> <mi>F</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math> improvement is roughly an order of magnitude at low energies, while at the highest energies there is no improvement at all.</p>
Full article ">Figure 3
<p>J-factor distributions for the brightest subhalo across 1000 realizations of the DM subhalo population as derived in [<a href="#B23-galaxies-08-00005" class="html-bibr">23</a>] from the Via Lactea II N-body simulation. Red and blue histograms show, respectively, the values before and after applying our selection cuts on the simulated DM subhalos, namely a mass cut (<math display="inline"><semantics> <mrow> <mi>M</mi> <mo>≤</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> <msub> <mi mathvariant="normal">M</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math>) and a coordinate cut (<math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mi>b</mi> <mo>|</mo> <mo>&gt;</mo> </mrow> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>). The dashed, purple vertical line marks the value of the J-factor that will be used to set the 95% C.L. DM limits in <a href="#sec3-galaxies-08-00005" class="html-sec">Section 3</a>, and it is defined as the one above which 95% of the J-factor distribution is contained.</p>
Full article ">Figure 4
<p>95% upper limits on the DM annihilation cross section as derived from HAWC unIDs and predictions from N-body cosmological simulations. Indeed, these constraints use 2HWC J1040+308 alone, i.e., the only HAWC unID located at <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> <mi>b</mi> <mo>|</mo> </mrow> <mo>≥</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> and surviving our selection criteria in <a href="#sec2dot2-galaxies-08-00005" class="html-sec">Section 2.2</a>. Left (right) panel shows the 95% C.L. upper limits for the <math display="inline"><semantics> <mrow> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <msup> <mi>τ</mi> <mo>+</mo> </msup> <msup> <mi>τ</mi> <mo>−</mo> </msup> </mrow> </semantics></math>) annihilation channel. For comparison, we also show as a blue, dot-dashed line the DM constraints obtained from Fermi-LAT unIDs using the same methodology [<a href="#B47-galaxies-08-00005" class="html-bibr">47</a>]. Limits from the observation of the Galactic center region by H.E.S.S., i.e., the best DM constraints achieved from IACT observations at present, are included in both panels as well as a green, dashed line [<a href="#B65-galaxies-08-00005" class="html-bibr">65</a>]. Finally, a dotted, orange line shows the constraints from <span class="html-italic">Fermi</span>-LAT observations of dSphs [<a href="#B59-galaxies-08-00005" class="html-bibr">59</a>]. Please note that we did not include in this Figure the results in [<a href="#B35-galaxies-08-00005" class="html-bibr">35</a>], as a one-to-one comparison would be misleading given the fact that the methodology and underlying assumptions in both works are significantly different; see discussion in <a href="#sec3-galaxies-08-00005" class="html-sec">Section 3</a> for details.</p>
Full article ">
45 pages, 1834 KiB  
Review
Synthesizing Observations and Theory to Understand Galactic Magnetic Fields: Progress and Challenges
by Rainer Beck, Luke Chamandy, Ed Elson and Eric G. Blackman
Galaxies 2020, 8(1), 4; https://doi.org/10.3390/galaxies8010004 - 21 Dec 2019
Cited by 74 | Viewed by 4981
Abstract
Constraining dynamo theories of magnetic field origin by observation is indispensable but challenging, in part because the basic quantities measured by observers and predicted by modelers are different. We clarify these differences and sketch out ways to bridge the divide. Based on archival [...] Read more.
Constraining dynamo theories of magnetic field origin by observation is indispensable but challenging, in part because the basic quantities measured by observers and predicted by modelers are different. We clarify these differences and sketch out ways to bridge the divide. Based on archival and previously unpublished data, we then compile various important properties of galactic magnetic fields for nearby spiral galaxies. We consistently compute strengths of total, ordered, and regular fields, pitch angles of ordered and regular fields, and we summarize the present knowledge on azimuthal modes, field parities, and the properties of non-axisymmetric spiral features called magnetic arms. We review related aspects of dynamo theory, with a focus on mean-field models and their predictions for large-scale magnetic fields in galactic discs and halos. Furthermore, we measure the velocity dispersion of H i gas in arm and inter-arm regions in three galaxies, M 51, M 74, and NGC 6946, since spiral modulation of the root-mean-square turbulent speed has been proposed as a driver of non-axisymmetry in large-scale dynamos. We find no evidence for such a modulation and place upper limits on its strength, helping to narrow down the list of mechanisms to explain magnetic arms. Successes and remaining challenges of dynamo models with respect to explaining observations are briefly summarized, and possible strategies are suggested. With new instruments like the Square Kilometre Array (SKA), large data sets of magnetic and non-magnetic properties from thousands of galaxies will become available, to be compared with theory. Full article
(This article belongs to the Special Issue New Perspectives on Galactic Magnetism)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of magnetic field components: mostly isotropic turbulent field <math display="inline"><semantics> <msub> <mi mathvariant="bold-italic">B</mi> <mi>iso</mi> </msub> </semantics></math> (<b>left</b>), mostly anisotropic turbulent field <math display="inline"><semantics> <msub> <mi mathvariant="bold-italic">B</mi> <mi>an</mi> </msub> </semantics></math> (<b>middle</b>), and mostly regular field <math display="inline"><semantics> <msub> <mi mathvariant="bold-italic">B</mi> <mi>reg</mi> </msub> </semantics></math> (<b>right</b>). The red circle represents the telescope beam (courtesy: Andrew Fletcher).</p>
Full article ">Figure 2
<p>Radial variation of the ratio of ordered to unresolved (isotropic turbulent) field strengths <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>≡</mo> <msub> <mi>B</mi> <mi>ord</mi> </msub> <mo>/</mo> <msub> <mi>B</mi> <mi>iso</mi> </msub> </mrow> </semantics></math>, derived from the observed degree of polarization of the synchrotron emission at 4.86 GHz (6.2 cm wavelength), averaged over all azimuthal angles of each radial ring in the galaxy’s plane (i.e., corrected for inclination). The largest radius of each plot is limited by the extent of the map of thermal emission that needs to be subtracted from the map of total emission to derive the map of total synchrotron emission.</p>
Full article ">Figure 3
<p>Radial variation of the pitch angle <math display="inline"><semantics> <msub> <mi>p</mi> <mi mathvariant="normal">o</mi> </msub> </semantics></math> (absolute values) of the ordered field, computed from the original maps in Stokes <span class="html-italic">Q</span> and <span class="html-italic">U</span>, averaged over all azimuthal angles of each radial ring in the galaxy’s plane. For NGC 6946 [<a href="#B55-galaxies-08-00004" class="html-bibr">55</a>] data at 8.46 GHz and 4.86 GHz (3.6 cm and 6.2 cm wavelengths) allowed us to correct the pitch angles for Faraday rotation, while those for IC 342 [<a href="#B184-galaxies-08-00004" class="html-bibr">184</a>], M 83 [<a href="#B173-galaxies-08-00004" class="html-bibr">173</a>], and M 101 [<a href="#B174-galaxies-08-00004" class="html-bibr">174</a>] are based on apparent polarization angles at 4.86 GHz. The low angular resolution of the M 101 observations corresponds to a spatial resolution of about 5 kpc at the adopted distance, so that the radial variation is smeared out. The spatial resolution for the other three galaxies is 0.4–0.6 kpc.</p>
Full article ">Figure 4
<p>Maps and radial profiles for M 51 (NGC 5194). Top left: H <span class="html-small-caps">i</span> total intensity map in units of <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math> pc<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math>, top middle: H <span class="html-small-caps">i</span> velocity dispersion map in units of <math display="inline"><semantics> <mrow> <mspace width="0.166667em"/> <mrow> <mi>km</mi> <mspace width="0.166667em"/> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, top right: stellar mass surface density map in units of <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math> pc<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math>, bottom left: azimuthally-averaged stellar mass surface density (black) with the lower limits of the error bars (blue), bottom middle: thresholded stellar surface density map showing arm (colour scale, in units of <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math> pc<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math>) and inter-arm (grey scale) portions of galaxy, bottom right: azimuthally-averaged H <span class="html-small-caps">i</span> velocity dispersions for arm (red) and inter-arm (green) portions of galaxy. Error bars represent the interquartile range of surface densities in each ring.</p>
Full article ">Figure 5
<p>Maps and radial profiles for M 74 (NGC 628); see <a href="#galaxies-08-00004-f004" class="html-fig">Figure 4</a> caption for full details.</p>
Full article ">Figure 6
<p>Maps and radial profiles for NGC 6946; see <a href="#galaxies-08-00004-f004" class="html-fig">Figure 4</a> caption for full details.</p>
Full article ">Figure 7
<p>Each row is similar to the bottom row of <a href="#galaxies-08-00004-f004" class="html-fig">Figure 4</a>, <a href="#galaxies-08-00004-f005" class="html-fig">Figure 5</a> and <a href="#galaxies-08-00004-f006" class="html-fig">Figure 6</a>, but now WISE 12 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m data, which traces star formation rate, is used to define arm and inter-arm regions. Rows from top to bottom show NGC 5194, NGC 628, and NGC 6946.</p>
Full article ">Figure A1
<p>Variation with azimuthal angle of Faraday rotation (<math display="inline"><semantics> <mrow> <mi>R</mi> <mi>M</mi> </mrow> </semantics></math>) of the barred spiral galaxy M 83, measured between 2.8 cm wavelength [<a href="#B237-galaxies-08-00004" class="html-bibr">237</a>] and 6.3 cm wavelength [<a href="#B173-galaxies-08-00004" class="html-bibr">173</a>], both at <math display="inline"><semantics> <msup> <mn>75</mn> <mrow> <mo>′</mo> <mo>′</mo> </mrow> </msup> </semantics></math> beam width, in sectors of two radial ranges 4–8 kpc and 8–12 kpc in the galaxy plane (inclined by <math display="inline"><semantics> <msup> <mn>24</mn> <mo>∘</mo> </msup> </semantics></math>). The azimuthal angle <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> is counted counter-clockwise from the north-eastern major axis (position angle <math display="inline"><semantics> <msup> <mn>45</mn> <mo>∘</mo> </msup> </semantics></math>). The error bars show the <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>M</mi> </mrow> </semantics></math> dispersion in each sector, including systematic variations, and hence are upper limits of the statistical uncertainties.</p>
Full article ">
29 pages, 1228 KiB  
Article
ConvoSource: Radio-Astronomical Source-Finding with Convolutional Neural Networks
by Vesna Lukic, Francesco de Gasperin and Marcus Brüggen
Galaxies 2020, 8(1), 3; https://doi.org/10.3390/galaxies8010003 - 20 Dec 2019
Cited by 25 | Viewed by 3232
Abstract
Finding and classifying astronomical sources is key in the scientific exploitation of radio surveys. Source-finding usually involves identifying the parts of an image belonging to an astronomical source, against some estimated background. This can be problematic in the radio regime, owing to the [...] Read more.
Finding and classifying astronomical sources is key in the scientific exploitation of radio surveys. Source-finding usually involves identifying the parts of an image belonging to an astronomical source, against some estimated background. This can be problematic in the radio regime, owing to the presence of correlated noise, which can interfere with the signal from the source. In the current work, we present ConvoSource, a novel method based on a deep learning technique, to identify the positions of radio sources, and compare the results to a Gaussian-fitting method. Since the deep learning approach allows the generation of more training images, it should perform well in the source-finding task. We test the source-finding methods on artificial data created for the data challenge of the Square Kilometer Array (SKA). We investigate sources that are divided into three classes: star forming galaxies (SFGs) and two classes of active galactic nuclei (AGN). The artificial data are given at two different frequencies (560 MHz and 1400 MHz), three total integration times (8 h, 100 h, 1000 h), and three signal-to-noise ratios (SNRs) of 1, 2, and 5. At lower SNRs, ConvoSource tends to outperform a Gaussian-fitting approach in the recovery of SFGs and all sources, although at the lowest SNR of one, the better performance is likely due to chance matches. The Gaussian-fitting method performs better in the recovery of the AGN-type sources at lower SNRs. At a higher SNR, ConvoSource performs better on average in the recovery of AGN sources, whereas the Gaussian-fitting method performs better in the recovery of SFGs and all sources. ConvoSource usually performs better at shorter total integration times and detects more true positives and misses fewer sources compared to the Gaussian-fitting method; however, it detects more false positives. Full article
Show Figures

Figure 1

Figure 1
<p>ConvoSource architecture and examples of inputs when training (real maps and solution maps), features detected at the output of the first and third convolutional layer, as well as the resulting reconstructed image of the solution map. During testing, only the real maps are input into the network, and the predictions are given using the weights from the trained network.</p>
Full article ">Figure 2
<p>(Left panel) Real map of a panel containing a combination of SFGs, SS, and FS sources at B1 at 1000 h. (Middle panel) True source locations at SNR = 2. (Right panel) True source locations at SNR = 5. The yellow, blue, and green pixels indicate SFGs, SS, and FS sources, respectively. In this particular case, both the SS and FS sources are very close together and very faint, which presents a challenge for both source-finders. The panels have a side length of 50 × 50 pixels.</p>
Full article ">Figure 3
<p>(Left panel) Real map of a panel containing a combination of SFGs, SS, and FS sources at B2 at 8 h. (Middle panel) True source locations at SNR = 2. There are two SFGs and one each of SS and FS galaxies. (Right panel) True source locations at SNR = 5. At this SNR, only one SFG and one SS source remain. The other SFG and FS sources had a total flux that was lower than the cut-off threshold at that SNR. The yellow, blue, and green pixels indicate SFGs, SS, and FS sources, respectively. The panels have a side-length of 50 × 50 pixels.</p>
Full article ">Figure 4
<p>(<b>Left</b>) Segmentation of a portion of the primary beam corrected images in the training set area. (<b>Right</b>) Segmentation of the solution map in the same area. These images are generated from the B1 1000 h dataset, using an SNR = 5 to determine the threshold of flux for injecting the solutions. Each block formed a single 50 × 50 pixel image that was input into the ConvoSource algorithm. The blocks on the left make up the training set images (train_X), and the blocks on the right make up the solution set images (train_Y).</p>
Full article ">Figure 5
<p>F1 scores at SNR = 1, across the two frequencies B1 (560 MHz) and B2 (1400 MHz) and the three integration times. There are three results given from ConvoSource, depending on the augmentation used when training. The blue bar represents no augmentation; orange represents augmenting the SS and FS sources; and the green bar represents augmenting all sources. The graphs show that PyBDSF usually performed better compared to ConvoSource at this SNR. Although it appeared that ConvoSource performed better across the SFGs and all sources in the B1 dataset, for all integration times, the better performance appeared to be explained by the increased proportion of chance matches at this SNR, as shown in <a href="#galaxies-08-00003-f006" class="html-fig">Figure 6</a>. However, it should be noted that these sources had very low significance given the SNR.</p>
Full article ">Figure 6
<p>Showing the effect of randomly rotating the reconstructed matrix of source locations to investigate the proportion of chance findings.</p>
Full article ">Figure 7
<p>The top and bottom rows show a couple of examples of a real map for B1 at 8 h (first column), the solutions when injected into the map given the SNR = 1 threshold (second column), the predicted locations by ConvoSource after training on the original images only (third column), and the predicted locations by PyBDSF (fourth column).</p>
Full article ">Figure 8
<p>F1 scores at SNR = 2, across the two frequencies B1 (560 MHz) and B2 (1400 MHz) and the three integration times. There are three results given from ConvoSource, depending on the augmentation used when training. The blue bar represents no augmentation; orange represents augmenting the SS and FS sources; and the green bar represents augmenting all sources.</p>
Full article ">Figure 9
<p>The top and bottom rows show a couple of examples of a real map for B2 at 8 h (first column), the solutions when injected into the map given the SNR = 2 threshold (second column), the predicted locations by ConvoSource after training on the original images only (third column), and the predicted locations by PyBDSF (fourth column).</p>
Full article ">Figure 10
<p>F1 scores at SNR = 5, across the two frequencies B1 (560 MHz) and B2 (1400 MHz) and the three integration times. There are three results given from ConvoSource, depending on the augmentation used when training. The blue bar represents no augmentation; orange represents augmenting the SS and FS sources; and the green bar represents augmenting all sources.</p>
Full article ">Figure 11
<p>The top and bottom rows show a couple of examples of a real map for B2 at 8 h (first column), the solutions when injected into the map given the SNR = 5 threshold (second column), the predicted locations by ConvoSource after training on the original images only (third column), and the predicted locations by PyBDSF (fourth column).</p>
Full article ">Figure 12
<p>Training and validation losses across the three integration times at SNR = 5 across B1 and B2 datasets in the left and rights panels, respectively.</p>
Full article ">Figure A1
<p>Precision values at SNR = 1.</p>
Full article ">Figure A2
<p>Recall values at SNR = 1.</p>
Full article ">Figure A3
<p>Precision values at SNR = 2.</p>
Full article ">Figure A4
<p>Recall values at SNR = 2.</p>
Full article ">Figure A5
<p>Precision scores at SNR = 5.</p>
Full article ">Figure A6
<p>Recall scores at SNR = 5.</p>
Full article ">
15 pages, 1363 KiB  
Article
ASASSN-13db 2014–2017 Eruption as an Intermediate Luminosity Optical Transient
by Amit Kashi, Amir M. Michaelis and Leon Feigin
Galaxies 2020, 8(1), 2; https://doi.org/10.3390/galaxies8010002 - 19 Dec 2019
Cited by 9 | Viewed by 2826
Abstract
The low mass star ASASSN-13db experienced an EXor outburst in 2013, which identified it as a Young Stellar Object (YSO). Then, from 2014 to 2017 it had another outburst, longer and more luminous than the earlier. We analyze the observations of the second [...] Read more.
The low mass star ASASSN-13db experienced an EXor outburst in 2013, which identified it as a Young Stellar Object (YSO). Then, from 2014 to 2017 it had another outburst, longer and more luminous than the earlier. We analyze the observations of the second outburst, and compare it to eruptions of Intermediate Luminosity Optical Transients (ILOTs). We show that the decline of the light curve is almost identical to that of the V838 Mon, a prototype of a type of ILOT known as Luminous Red Nova (LRN). This similarity becomes conspicuous when oscillations that are associated with rotation are filtered out from the light curve of ASASSN-13db. We suggest that the eruption was the result of accretion of a proto-planet of a few Earth masses. The proto-planet was shredded by tidal forces before it was accreted onto the YSO, releasing gravitational energy that powered the outburst for 800 days , and ended in a 55 days decline phase. When the accretion material started depleting the accretion rate lowered and the eruption light curve declined for almost two months. Then it exhausted completely, creating a sharp break in the light curve. Another possibility is that the mass was a result of an instability in the proto-planetary disk that lead to a large episode of accretion from an inner viscous disk. We find that the variation of the temperature of the outburst is consistent with the surface temperature expected from a depleted viscous accretion disk. The 2014–2017 outburst of ASASSN-13db may be the least energetic ILOT to have been discovered to date, with an energy budget of only 10 42 erg . Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Comparing the <span class="html-italic">V</span>-band light curves of A13db1417 [<a href="#B47-galaxies-08-00002" class="html-bibr">47</a>] and V838 Mon (Bond et al. [<a href="#B2-galaxies-08-00002" class="html-bibr">2</a>], Starrfield et al. [<a href="#B39-galaxies-08-00002" class="html-bibr">39</a>], Sparks et al. [<a href="#B63-galaxies-08-00002" class="html-bibr">63</a>]). The magnitude scale is the apparent magnitude for A13db1417. The light curve of V838 Mon was shifted by <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>V</mi> <mrow> <mi mathvariant="normal">V</mi> <mn>838</mn> <mspace width="3.33333pt"/> <mi>Mon</mi> </mrow> </msub> <mo>=</mo> <mn>6.9</mn> </mrow> </semantics></math> mag to match the second peak before decline. The time axis focuses on the end of the <math display="inline"><semantics> <mrow> <mo>≈</mo> <mn>800</mn> <mspace width="3.33333pt"/> <mi>days</mi> </mrow> </semantics></math> duration of A13db1417 (see Sicilia Aguilar et al. [<a href="#B47-galaxies-08-00002" class="html-bibr">47</a>]) which is the <math display="inline"><semantics> <mrow> <mo>≃</mo> <mn>55</mn> <mspace width="3.33333pt"/> <mi>days</mi> </mrow> </semantics></math> decline phase. The peak at <math display="inline"><semantics> <mrow> <mi>JD</mi> <mo>≃</mo> </mrow> </semantics></math> 2,457,728 marks <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. (<b>b</b>) Same us the upper panel, but the light curves were shifted to match the peak just before decline. In addition, the time axis of the light curve V838 Mon is scaled by a factor of 1.3 relative to the matched peak. This results in that the two light curves match for about 4 mag.</p>
Full article ">Figure 2
<p>The light curve of A13db1417, with the variability resulted by rotation of stellar spot filtered out, compared to the light curve of V838 Mon. This variability caused oscillations of <math display="inline"><semantics> <mrow> <mi>δ</mi> <mi>V</mi> <mo>≃</mo> <mn>1</mn> </mrow> </semantics></math> mag. By filtering out the effect of rotation, we isolate the component resulted from accretion. The filtered signal matches better the scaled light curve of V838 Mon.</p>
Full article ">Figure 3
<p>The light curve of A13db1417 compared to that of the FU Ori eruption of V1647 Ori [<a href="#B66-galaxies-08-00002" class="html-bibr">66</a>]. It can be seen that the two curves are very different and the decline does not follow the same slope. This suggest that this objects are different. Note that the LCOGT observation at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mo>−</mo> <mn>2</mn> <mspace width="3.33333pt"/> <mi>days</mi> </mrow> </semantics></math> may be an outlier.</p>
Full article ">Figure 4
<p>The effective temperature of A13db1417, obtained from different filters as indicated in the legend. The calculation was performed assuming black-body emission, which is apparently not the emitting spectrum, hence the differences between the estimated in different filter pairs. Nevertheless we can see that the effective temperature is declining from <math display="inline"><semantics> <mrow> <mo>≃</mo> <mn>4500</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">K</mi> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mo>≃</mo> <mn>2000</mn> <mspace width="3.33333pt"/> <mi mathvariant="normal">K</mi> </mrow> </semantics></math> during the eruption. Over-plotted is the effective temperature from of V838 Mon, adopted from [<a href="#B67-galaxies-08-00002" class="html-bibr">67</a>]. It is evident that both objects have a similar decline.</p>
Full article ">Figure 5
<p>A focus on the fast increase in luminosity at the beginning of A13db1417. Observations are taken from [<a href="#B47-galaxies-08-00002" class="html-bibr">47</a>].</p>
Full article ">Figure 6
<p>An estimate to a classical accretion disk surface temperature, according to Equation (<a href="#FD7-galaxies-08-00002" class="html-disp-formula">7</a>). When the disk is depleted the mass accretion rate decreases and so does the surface temperature of the disk.</p>
Full article ">
28 pages, 3043 KiB  
Review
Relativistic Jets from AGN Viewed at Highest Angular Resolution
by Kazuhiro Hada
Galaxies 2020, 8(1), 1; https://doi.org/10.3390/galaxies8010001 - 18 Dec 2019
Cited by 19 | Viewed by 8715
Abstract
Accreting supermassive black holes in active galactic nuclei (AGN) produce powerful relativistic jets that shine from radio to GeV/TeV γ-rays. Over the past decade, AGN jets have extensively been studied in various energy bands and our knowledge about the broadband emission and rapid [...] Read more.
Accreting supermassive black holes in active galactic nuclei (AGN) produce powerful relativistic jets that shine from radio to GeV/TeV γ-rays. Over the past decade, AGN jets have extensively been studied in various energy bands and our knowledge about the broadband emission and rapid flares are now significantly updated. Meanwhile, the progress of magnetohydrodynamic simulations with a rotating black hole have greatly improved our theoretical understanding of powerful jet production. Nevertheless, it is still challenging to observationally resolve such flaring sites or jet formation regions since the relevant spatial scales are tiny. Observations with very long baseline interferometry (VLBI) are currently the only way to directly access such compact scales. Here we overview some recent progress of VLBI studies of AGN jets. As represented by the successful black hole shadow imaging with the Event Horizon Telescope, the recent rapid expansion of VLBI capability is remarkable. The last decade has also seen a variety of advances thanks to the advent of RadioAstron, GMVA, new VLBI facilities in East Asia as well as to the continued upgrade of VLBA. These instruments have resolved the innermost regions of relativistic jets for a number of objects covering a variety of jetted AGN classes (radio galaxies, blazars, and narrow-line Seyfert 1 galaxies), and the accumulated results start to establish some concrete (and likely universal) picture on the collimation, acceleration, recollimation shocks, magnetic field topology, and the connection to high-energy flares in the innermost part of AGN jets. Full article
(This article belongs to the Special Issue Jet Physics of Accreting Super Massive Black Holes)
Show Figures

Figure 1

Figure 1
<p>M87 images from galactic to event horizon scales. (Top left) HST image of M87 (credit: NASA, ESA and the Hubble Heritage Team (STScI/AURA)). (Top right) EAVN 43 GHz image of M87 (credit: EAVN Collaboration). (Bottom) EHT 230 GHz image of M87 (credit: EHT Collaboration).</p>
Full article ">Figure 2
<p>Collimation and acceleration of the M87 jet. (<b>Left</b>) Jet collimation profile of M87 (adapted from work in [<a href="#B47-galaxies-08-00001" class="html-bibr">47</a>]). (<b>Right</b>) Jet velocity profile of M87 as a function distance (adapted from work in [<a href="#B58-galaxies-08-00001" class="html-bibr">58</a>]).</p>
Full article ">Figure 3
<p>VLBA RM profile of the M87 jet as a function of distance from the BH (adapted from work in [<a href="#B69-galaxies-08-00001" class="html-bibr">69</a>]).</p>
Full article ">Figure 4
<p>(<b>Left</b>) RadioAstron 22 GHz image of 3C 84 observed in 2013 (adapted from work in [<a href="#B101-galaxies-08-00001" class="html-bibr">101</a>]). (<b>Right</b>) Trajectory of C3 from August 2015 to June 2016, obtained by KaVA and VLBA at 43 GHz (adapted from work in [<a href="#B102-galaxies-08-00001" class="html-bibr">102</a>]). A red rectangular area overlaid on the RadioAstron image corresponds to the plotting area of the C3 trajectory.</p>
Full article ">Figure 5
<p>Magnetic flux measured in pc-scale jets via core-shift (<math display="inline"><semantics> <msub> <mo>Φ</mo> <mi>jet</mi> </msub> </semantics></math>) for a sample of 76 MOJAVE sources (68 blazars (filled circles) and 8 radio galaxies (open circles)), plotted as a function of <math display="inline"><semantics> <mrow> <msubsup> <mi>L</mi> <mrow> <mi>acc</mi> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msubsup> <mi>M</mi> </mrow> </semantics></math> that is proportional to the theoretical MAD magnetic flux threading BH (<math display="inline"><semantics> <msub> <mo>Φ</mo> <mi>BH</mi> </msub> </semantics></math>) and is estimated by the accretion disk luminosity (<math display="inline"><semantics> <msub> <mi>L</mi> <mi>acc</mi> </msub> </semantics></math>) and BH mass (<span class="html-italic">M</span>) (adapted from the work in [<a href="#B139-galaxies-08-00001" class="html-bibr">139</a>]).</p>
Full article ">Figure 6
<p>(<b>Left</b>) RM distributions (color) of BL Lacertae made by 15/22/43 GHz data, overlaid on a RadioAstron 22 GHz total intensity contour map (adapted from [<a href="#B146-galaxies-08-00001" class="html-bibr">146</a>]). (<b>Right</b>) RM map of CTA 102 made by 43/86 GHz data, overlaid on a 86 GHz GMVA total intensity contour map (adapted from work in [<a href="#B150-galaxies-08-00001" class="html-bibr">150</a>]). Both images show significant RM gradients around the core.</p>
Full article ">Figure 7
<p>VLBA observations of 1H 0323+342 by [<a href="#B218-galaxies-08-00001" class="html-bibr">218</a>]. (<b>Left</b>) Contours are 15 GHz total intensity distributions while color scales are polarized fluxes. (<b>Right</b>) Jet width profile as a function of distance from core, showing a parabolic to conical transition at 7 mas (8 pc projected) where the stationary, compressed and polarized feature S is located. Figures adapted from work in [<a href="#B218-galaxies-08-00001" class="html-bibr">218</a>].</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop