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

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

Galaxies, Volume 7, Issue 4 (December 2019) – 14 articles

Cover Story (view full-size image): The cover image shows the three-dimensional distribution of synchrotron intensity (in units of micro-Jansky pixel^-1) at 1.5 GHz of a transonic turbulent medium generated from magnetohydrodynamic simulations (Burkhart et al. 2009, ApJ, 693, 250). All intensity fluctuations are caused due to fluctuations of the magnetic field strength in the plane of the sky. The synchrotron intensity is computed using the COSMIC package presented in Basu et al., and performs polarization transfer calculations including the effects of Faraday rotation to generate realistic synthetic observations. 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:
27 pages, 1600 KiB  
Review
High Energy Radiation from Spider Pulsars
by Chung Yue Hui and Kwan Lok Li
Galaxies 2019, 7(4), 93; https://doi.org/10.3390/galaxies7040093 - 13 Dec 2019
Cited by 22 | Viewed by 4258
Abstract
The population of millisecond pulsars (MSPs) has been expanded considerably in the last decade. Not only is their number increasing, but also various classes of them have been revealed. Among different classes of MSPs, the behaviours of black widows and redbacks are particularly [...] Read more.
The population of millisecond pulsars (MSPs) has been expanded considerably in the last decade. Not only is their number increasing, but also various classes of them have been revealed. Among different classes of MSPs, the behaviours of black widows and redbacks are particularly interesting. These systems consist of an MSP and a low-mass companion star in compact binaries with an orbital period of less than a day. In this article, we give an overview of the high energy nature of these two classes of MSPs. Updated catalogues of black widows and redbacks are presented and their X-ray/ γ -ray properties are reviewed. Besides the overview, using the most updated eight-year Fermi Large Area Telescope point source catalog, we have compared the γ -ray properties of these two MSP classes. The results suggest that the X-rays and γ -rays observed from these MSPs originate from different mechanisms. Lastly, we will also mention the future prospects of studying these spider pulsars with the novel methodologies as well as upcoming observing facilities. Full article
(This article belongs to the Special Issue Observations of Gamma-Ray Pulsars)
Show Figures

Figure 1

Figure 1
<p>Period–period derivative (<span class="html-italic">P</span>–<math display="inline"><semantics> <mover accent="true"> <mi>P</mi> <mo>˙</mo> </mover> </semantics></math>) diagram of all currently known pulsars. The population can be divided into two groups (black and grey dots) based on <span class="html-italic">k</span>-means partitioning. The locations of black widow and redback millisecond pulsars (MSPs) in this parameter space are highlighted by the green and blue circles, respectively. Lines of constant dipolar surface magnetic field (Equation (<a href="#FD1-galaxies-07-00093" class="html-disp-formula">1</a>)) and characteristic age (Equation (<a href="#FD2-galaxies-07-00093" class="html-disp-formula">2</a>)) are shown. The dashed line illustrates the death line for radio pulsars by assuming a multipolar magnetic field configuration [<a href="#B2-galaxies-07-00093" class="html-bibr">2</a>].</p>
Full article ">Figure 2
<p>Galactic distributions of confirmed spider pulsars. The dark red points are the results of overlapping of redbacks and black widows in the same globular cluster.</p>
Full article ">Figure 3
<p>X-ray image of the bow-shock nebula associated with black widow MSP PSR B1957+20 obtained from the public data with an effective exposure of ∼165 ks as acquired by Chandra (Observation ID: 9088). This is the same data as used in the study by Huang et al. [<a href="#B63-galaxies-07-00093" class="html-bibr">63</a>].</p>
Full article ">Figure 4
<p>Orbital modulation of PSR B1957+20 in X-ray as observed by Chandra [<a href="#B63-galaxies-07-00093" class="html-bibr">63</a>]. The error bars correspond to <math display="inline"><semantics> <mrow> <mn>1</mn> <mi>σ</mi> </mrow> </semantics></math> uncertainties assuming Poisson noises. The eclipses of the radio pulses occur at the orbital phases of 0.2–0.3 and 1.2–1.3 which are highlighted by the blue regions. The grey regions represent the phases for extracting the X-ray spectrum of PSR B1957+20 in the eclipsing region in Huang et al. [<a href="#B63-galaxies-07-00093" class="html-bibr">63</a>]. Two orbital cycles are shown for clarity.</p>
Full article ">Figure 5
<p><math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray orbital modulation of PSR B1957+20 observed by Fermi large area telescope (LAT). This is discovered by Wu et al. [<a href="#B64-galaxies-07-00093" class="html-bibr">64</a>]. The error bars correspond to <math display="inline"><semantics> <mrow> <mn>1</mn> <mi>σ</mi> </mrow> </semantics></math> uncertainties assuming Poisson noises. The shaded regions correspond to the phase of radio eclipse (i.e., 0.2–0.3 and 1.2–1.3). Two orbital cycles are shown for clarity.</p>
Full article ">Figure 6
<p>X-ray orbital modulation of PSR J1023+0038 in the soft (0.3–2.0 keV) and hard (2.0–10.0 keV) bands, as obtained from the observation taken at 26 November 2008 with XMM-Newton [<a href="#B108-galaxies-07-00093" class="html-bibr">108</a>].</p>
Full article ">Figure 7
<p>UV, X-ray and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray lightcurves of PSR J1023+0038 from 1 June 2013 to 13 November 2013 are shown together in the main panel with different flux scales for each energy band (see upper left corner for details). On the other hand, the inset box shows the detailed evolution of the <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray emissions from 6 June to 24 July. Each data point of UV/X-ray represents an individual observation taken by Swift. Each <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray data points in the main panel and inset corresponds to two weeks and three days, respectively. In the cases where the detection significances is <math display="inline"><semantics> <mrow> <mo>≤</mo> <mn>3</mn> <mi>σ</mi> </mrow> </semantics></math>, upper limits at 95% confidence are given instead [<a href="#B28-galaxies-07-00093" class="html-bibr">28</a>].</p>
Full article ">Figure 8
<p>Schematic illustration for the emission nature of PSR J1023+0038 after 2013 late June in different wavelengths. The accretion disk extends beyond the light cylinder radius (<math display="inline"><semantics> <msub> <mi>R</mi> <mrow> <mi>l</mi> <mi>c</mi> </mrow> </msub> </semantics></math>). <math display="inline"><semantics> <msub> <mi>R</mi> <mi>s</mi> </msub> </semantics></math> is the distance to the intra-binary shock from the pulsar. <math display="inline"><semantics> <msub> <mi>R</mi> <mi>c</mi> </msub> </semantics></math> is the critical distance from the pulsar at which the <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays from its magnetosphere evaporate the disk matter at <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>&lt;</mo> <msub> <mi>R</mi> <mi>c</mi> </msub> <mo>∼</mo> <mn>3</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>9</mn> </msup> </mrow> </semantics></math> cm. UV/Optical photons mainly originate from the disk at <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>∼</mo> <msup> <mn>10</mn> <mrow> <mn>9</mn> <mo>−</mo> <mn>10</mn> </mrow> </msup> </mrow> </semantics></math> cm. Shock is formed through the interaction between the pulsar wind and the stellar wind. This produces the non-thermal X-ray emissions. The inverse-Compton process of the cold-relativistic pulsar wind off UV/Optical photons from the disk produces the additional <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays [<a href="#B28-galaxies-07-00093" class="html-bibr">28</a>].</p>
Full article ">Figure 9
<p>Multi-wavelength spectral energy distributions of a PSR J1023+0038 system before (<b>left</b>) and after (<b>right</b>) late June 2013. Calculations with a model consist of emission components from the pulsar magnetosphere (outer gap); shock and pulsar wind (PW) are compared with the observed data before and after the transition. For further details, please refer to [<a href="#B28-galaxies-07-00093" class="html-bibr">28</a>].</p>
Full article ">Figure 10
<p>Light curves of PSR J1048+2339 companion star with <math display="inline"><semantics> <msup> <mi>r</mi> <mo>′</mo> </msup> </semantics></math> and <math display="inline"><semantics> <msup> <mi>g</mi> <mo>′</mo> </msup> </semantics></math> band filter, as observed by a 1 m Lulin telescope and a 2 m Liverpool telescope between 11 March 2018 and 23 April, folded with an orbital period of 6 h [<a href="#B34-galaxies-07-00093" class="html-bibr">34</a>].</p>
Full article ">Figure 11
<p>Relation between <math display="inline"><semantics> <msub> <mi>L</mi> <mi>x</mi> </msub> </semantics></math> and <math display="inline"><semantics> <mover accent="true"> <mi>E</mi> <mo>˙</mo> </mover> </semantics></math> for 46 MSPs of different classes which are shown as different symbols in this plot [<a href="#B20-galaxies-07-00093" class="html-bibr">20</a>]. In addition, the upper-limits on <math display="inline"><semantics> <msub> <mi>L</mi> <mi>x</mi> </msub> </semantics></math> for 35 MSPs are included in the sample with which Lee et al. [<a href="#B20-galaxies-07-00093" class="html-bibr">20</a>] performed the survival analysis. The solid line illustrates the Akritas–Thiel–Sen (ATS) line inferred from this censored data. For comparison, the dashed line illustrates the result from the standard linear regression of X-ray detected MSPs. Moreover the relation reported by Possenti et al. [<a href="#B118-galaxies-07-00093" class="html-bibr">118</a>] based on a sample of 10 MSPs is displayed as the dotted line [<a href="#B20-galaxies-07-00093" class="html-bibr">20</a>].</p>
Full article ">Figure 12
<p>Comparison of the effective photon indices <math display="inline"><semantics> <mo>Γ</mo> </semantics></math> of black-widows (BWs) and redbacks (RBs) in X-ray (left panel). The comparison of the X-ray luminosities <math display="inline"><semantics> <msub> <mi>L</mi> <mi>x</mi> </msub> </semantics></math> of BWs and RBs (right panel). The <math display="inline"><semantics> <mrow> <mi>p</mi> <mo>-</mo> </mrow> </semantics></math>values resulting from the two-sample Kolmogorov–Smirnov (KS) test and Anderson–Darling (AD) test are given in each figure, and strongly indicate the differences between these two classes of MSPs [<a href="#B20-galaxies-07-00093" class="html-bibr">20</a>].</p>
Full article ">Figure 13
<p>Plot of <math display="inline"><semantics> <msub> <mi>L</mi> <mi>γ</mi> </msub> </semantics></math> vs. <math display="inline"><semantics> <mover accent="true"> <mi>E</mi> <mo>˙</mo> </mover> </semantics></math> for the <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray pulsars enlisted in the current version of the 4FGL catalog. The solid straight line illustrates the least square linear fit. The locations of black widows and redbacks in this parameter space are given by blue and green symbols, respectively.</p>
Full article ">Figure 14
<p>Comparisons of the step-wise empirical cumulative distributions of <math display="inline"><semantics> <msub> <mi>L</mi> <mi>γ</mi> </msub> </semantics></math> (upper-left panel), <math display="inline"><semantics> <mover accent="true"> <mi>E</mi> <mo>˙</mo> </mover> </semantics></math> (upper-right panel), <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray photon index (lower-left panel) and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray exponential factor <span class="html-italic">a</span> (lower-right panel) of black-widows and redbacks in <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-rays.</p>
Full article ">Figure 15
<p>A long-term <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray light curve of PSR J1023+0038 as observed by Fermi LAT at energies <math display="inline"><semantics> <mrow> <mo>&gt;</mo> <mn>100</mn> </mrow> </semantics></math> MeV from MJD 54697 (August 2008) to MJD 58580 (April 2019).</p>
Full article ">
22 pages, 6972 KiB  
Review
Red Supergiants, Yellow Hypergiants, and Post-RSG Evolution
by Michael S. Gordon and Roberta M. Humphreys
Galaxies 2019, 7(4), 92; https://doi.org/10.3390/galaxies7040092 - 3 Dec 2019
Cited by 13 | Viewed by 4786
Abstract
How massive stars end their lives remains an open question in the field of star evolution. While the majority of stars above ≳9 M will become red supergiants (RSGs), the terminal state of these massive stars can be heavily influenced by their [...] Read more.
How massive stars end their lives remains an open question in the field of star evolution. While the majority of stars above ≳9 M will become red supergiants (RSGs), the terminal state of these massive stars can be heavily influenced by their mass-loss histories. Periods of enhanced circumstellar wind activity can drive stars off the RSG branch of the HR Diagram. This phase, known as post-RSG evolution, may well be tied to high mass-loss events or eruptions as seen in the Luminous Blue Variables (LBVs) and other massive stars. This article highlights some of the recent observational and modeling studies that seek to characterize this unique class of stars, the post-RSGs and link them to other massive objects on the HR Diagram such as LBVs, Yellow Hypergiants and dusty RSGs. Full article
(This article belongs to the Special Issue Luminous Stars in Nearby Galaxies)
Show Figures

Figure 1

Figure 1
<p>Initial masses of observed supernova (SN) Type II-P progenitors in the Smartt et al. [<a href="#B6-galaxies-07-00092" class="html-bibr">6</a>] survey. Labels indicate theoretical limits for types of compact remnants. Darker shading is higher metallicity. The thick gray line represents a cumulative frequency distribution of a Salpeter IMF with Γ = −1.35. Figure reproduced from Smartt et al. [<a href="#B6-galaxies-07-00092" class="html-bibr">6</a>].</p>
Full article ">Figure 2
<p>Schematic HRD of Galactic warm hypergiants (and Var A in M33) illustrating the location of these massive stars relative to the Humphreys-Davidson limit [<a href="#B1-galaxies-07-00092" class="html-bibr">1</a>] and the “yellow void”—a temperature and luminosity band region for increased dynamical instability. The location of the LBV instability strip is also shown with the classical (LBV 1) and less-luminous (LBV 2) LBVs in their quiescent state.</p>
Full article ">Figure 3
<p><b>Left</b>: The combined color image of IRC +10420 from HST/WFPC2 [<a href="#B47-galaxies-07-00092" class="html-bibr">47</a>]. <b>Right</b>: Profile of the H<span class="html-italic">α</span> emission line showing the broad electron scattering wings and the split profile (adapted from Reference [<a href="#B48-galaxies-07-00092" class="html-bibr">48</a>]).</p>
Full article ">Figure 4
<p><b>Top</b>: Optical spectrum of Var A from 1985 [<a href="#B58-galaxies-07-00092" class="html-bibr">58</a>] showing H<span class="html-italic">α</span> emission and TiO absorption bands. <b>Bottom</b>: Optical spectrum of Var A from 2004 [<a href="#B59-galaxies-07-00092" class="html-bibr">59</a>] with the strongest emission lines marked.</p>
Full article ">Figure 5
<p><b>Left</b>: Light curve of Var A from 1950 to the present. The top panel shows the photographic and B-band magnitudes. The middle and bottom panels show the variability in the V-band and the B–V color. See Reference [<a href="#B59-galaxies-07-00092" class="html-bibr">59</a>]. <b>Right</b>: Spectral energy distribution of Var A from 1986 [<a href="#B58-galaxies-07-00092" class="html-bibr">58</a>,<a href="#B59-galaxies-07-00092" class="html-bibr">59</a>]. The plus signs show its apparent magnitudes at maximum light from Reference [<a href="#B57-galaxies-07-00092" class="html-bibr">57</a>].</p>
Full article ">Figure 6
<p>SEDs of warm hypergiant candidates in M31. The observed visual, 2MASS and IRAC magnitudes are shown as filled circles and WISE data as open circles. The extinction-corrected photometry is plotted as filled squares, with the measured line-of-sight <math display="inline"><semantics> <msub> <mi>A</mi> <mi>V</mi> </msub> </semantics></math> specified in each legend. The SED of J004621.05+421308.06 (<b>top</b>) reveals a prominent CS dust envelope in the IRAC and WISE bands. The WISE photometry of J004051.59+403303.00 (<b>bottom</b>) is suggestive of silicate dust emission but is most likely due to contamination from a nearby H II region and nebulosity. The dotted line is a curve of constant <math display="inline"><semantics> <msub> <mi>F</mi> <mi>ν</mi> </msub> </semantics></math>, which is evidence for free-free emission in wind. Figure adapted from Reference [<a href="#B13-galaxies-07-00092" class="html-bibr">13</a>].</p>
Full article ">Figure 7
<p>HR Diagrams of M31 (<b>top</b>) and M33 (<b>bottom</b>). Red circles represent the RSG sample from Gordon et al. [<a href="#B13-galaxies-07-00092" class="html-bibr">13</a>], black circles are the YSGs. Closed symbols are sources with evidence of mass loss, either in their spectra (for the YSGs) or their SEDs (for both the YSGs and RSGs). Non-rotating stellar evolution tracks for three mass bins from Ekström et al. [<a href="#B80-galaxies-07-00092" class="html-bibr">80</a>] are shown for comparison. The stars with mass loss, the post-RSG candidates, appear to dominate the upper portion of the HR diagram. Figures adapted from Reference [<a href="#B13-galaxies-07-00092" class="html-bibr">13</a>].</p>
Full article ">Figure 8
<p>Mass-loss rates vs. luminosity for Galactic RSGs. The solid line represents the de Jager et al. [<a href="#B21-galaxies-07-00092" class="html-bibr">21</a>] model for stellar <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>eff</mi> </msub> <mo>=</mo> <mn>4000</mn> </mrow> </semantics></math> K and the dotted line for <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>eff</mi> </msub> <mo>=</mo> <mn>3500</mn> </mrow> </semantics></math> K. Figure adapted from Reference [<a href="#B3-galaxies-07-00092" class="html-bibr">3</a>].</p>
Full article ">Figure 9
<p>Bolometric luminosity vs. total mass lost based on dust measurements for RSG candidates in M31 and M33. Closed circles are those with clear evidence for mass loss in their SEDs. Open circles are the less certain mass losers. We note that the RSGs with higher luminosity tend to have lost more mass, consistent with the prescription of de Jager et al. [<a href="#B21-galaxies-07-00092" class="html-bibr">21</a>] for mass loss in RSGs. Figure adapted from Reference [<a href="#B13-galaxies-07-00092" class="html-bibr">13</a>].</p>
Full article ">
14 pages, 606 KiB  
Article
The Origin of Large-Scale Magnetic Fields in Low-Mass Galaxies
by Prasanta Bera, Anvar Shukurov and Kandaswamy Subramanian
Galaxies 2019, 7(4), 91; https://doi.org/10.3390/galaxies7040091 - 29 Nov 2019
Cited by 3 | Viewed by 2616
Abstract
The origin of large-scale magnetic fields, detected in some low-mass (dwarf and irregular) galaxies via polarised synchrotron emission and Faraday rotation, has remained unexplained for a long time. We suggest that mean-field dynamos can be active in galaxies of this class despite their [...] Read more.
The origin of large-scale magnetic fields, detected in some low-mass (dwarf and irregular) galaxies via polarised synchrotron emission and Faraday rotation, has remained unexplained for a long time. We suggest that mean-field dynamos can be active in galaxies of this class despite their slow rotation because their discs are relatively thick. Earlier assessments of the possibility of the mean-field dynamo action in low-mass galaxies relied on estimates applicable to thin discs, such as those in massive spiral galaxies. Using both order-of-magnitude estimates and numerical solutions, we show that the strength of differential rotation required to amplify magnetic field reduces as the aspect ratio of the galactic gas layer increases. As in a thin disc, quadrupolar magnetic fields dominate in thick discs. Thus, the origin of large-scale magnetic fields in low-mass galaxies has been clarified. This class of galaxies provides a new ground for testing our understanding of galactic magnetism. Full article
(This article belongs to the Special Issue New Perspectives on Galactic Magnetism)
Show Figures

Figure 1

Figure 1
<p>The variation of the critical rotation speed <math display="inline"><semantics> <msub> <mi>V</mi> <mi mathvariant="normal">c</mi> </msub> </semantics></math> required for the dynamo action (for the quadrupolar magnetic field geometry) with the disc aspect ratio <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math>: numerical results are shown with crosses and the solid line represents the approximate value from Equation (<a href="#FD16-galaxies-07-00091" class="html-disp-formula">16</a>) with <math display="inline"><semantics> <mrow> <msup> <mi>k</mi> <mn>2</mn> </msup> <mo>=</mo> <mn>8</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>As in <a href="#galaxies-07-00091-f001" class="html-fig">Figure 1</a> but for the critical dynamo number, Equation (<a href="#FD15-galaxies-07-00091" class="html-disp-formula">15</a>).</p>
Full article ">Figure 3
<p>The cylindrical components (<b>a</b>) <math display="inline"><semantics> <msub> <mi>B</mi> <mi>r</mi> </msub> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <msub> <mi>B</mi> <mi>ϕ</mi> </msub> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <msub> <mi>B</mi> <mi>z</mi> </msub> </semantics></math> of the marginally stable quadrupolar eigenfunction of the <math display="inline"><semantics> <mrow> <mi>α</mi> <mi>ω</mi> </mrow> </semantics></math>-dynamo equations for <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>1</mn> <mspace width="0.166667em"/> <mi>kpc</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mn>0</mn> </msub> <mo>/</mo> <msub> <mi>v</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>2.96</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math> normalised to <math display="inline"><semantics> <mrow> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <msubsup> <mi>B</mi> <mi>r</mi> <mn>2</mn> </msubsup> <mo>+</mo> <msubsup> <mi>B</mi> <mi>z</mi> <mn>2</mn> </msubsup> <mo>+</mo> <msubsup> <mi>B</mi> <mi>ϕ</mi> <mn>2</mn> </msubsup> <mo>)</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. Panel (<b>d</b>) shows poloidal magnetic lines (white), providing an indication of the relationship between <math display="inline"><semantics> <msub> <mi>B</mi> <mi>z</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>B</mi> <mi>r</mi> </msub> </semantics></math>, and the pitch angle of magnetic lines (in degrees, colour). Only the region where <math display="inline"><semantics> <mrow> <msubsup> <mi>B</mi> <mi>r</mi> <mn>2</mn> </msubsup> <mo>+</mo> <msubsup> <mi>B</mi> <mi>z</mi> <mn>2</mn> </msubsup> <mo>+</mo> <msubsup> <mi>B</mi> <mi>ϕ</mi> <mn>2</mn> </msubsup> <mo>≥</mo> <mn>0.01</mn> <mo movablelimits="true" form="prefix">max</mo> <mrow> <mo>(</mo> <msubsup> <mi>B</mi> <mi>r</mi> <mn>2</mn> </msubsup> <mo>+</mo> <msubsup> <mi>B</mi> <mi>z</mi> <mn>2</mn> </msubsup> <mo>+</mo> <msubsup> <mi>B</mi> <mi>ϕ</mi> <mn>2</mn> </msubsup> <mo>)</mo> </mrow> </mrow> </semantics></math> is shown.</p>
Full article ">Figure 4
<p>As in <a href="#galaxies-07-00091-f003" class="html-fig">Figure 3</a> but for the dipolar parity. The critical circular rotation speed <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mi mathvariant="normal">c</mi> </msub> <mo>/</mo> <msub> <mi>v</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>17</mn> </mrow> </semantics></math> is higher than for the marginal quadrupolar mode. Unlike the quadrupolar field of <a href="#galaxies-07-00091-f003" class="html-fig">Figure 3</a>, the dipolar magnetic field varies periodically in time (see <a href="#galaxies-07-00091-f005" class="html-fig">Figure 5</a>).</p>
Full article ">Figure 5
<p>The form of the marginally stable dipolar mode of <a href="#galaxies-07-00091-f004" class="html-fig">Figure 4</a> in various phases of its oscillation: magnetic lines of the poloidal field are shown white and the toroidal component is colour-coded (arbitrary units) for three different times within the oscillation period (<math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>&lt;</mo> <msub> <mi>t</mi> <mn>2</mn> </msub> <mo>&lt;</mo> <msub> <mi>t</mi> <mn>3</mn> </msub> </mrow> </semantics></math>).</p>
Full article ">
14 pages, 2468 KiB  
Article
Gamma-Ray Sensitivity to Dark Matter Subhalo Modelling at High Latitudes
by Francesca Calore, Moritz Hütten and Martin Stref
Galaxies 2019, 7(4), 90; https://doi.org/10.3390/galaxies7040090 - 26 Nov 2019
Cited by 13 | Viewed by 2518
Abstract
Searches for “dark” subhaloes in gamma-ray point-like source catalogues are among promising strategies for indirect dark matter detection. Such a search is nevertheless affected by uncertainties related, on the one hand, to the modelling of the dark matter subhalo distribution in Milky-Way-like galaxies, [...] Read more.
Searches for “dark” subhaloes in gamma-ray point-like source catalogues are among promising strategies for indirect dark matter detection. Such a search is nevertheless affected by uncertainties related, on the one hand, to the modelling of the dark matter subhalo distribution in Milky-Way-like galaxies, and, on the other hand, to the sensitivity of gamma-ray instruments to the dark matter subhalo signals. In the present work, we assess the detectability of dark matter subhaloes in Fermi-LAT catalogues, taking into accounts uncertainties associated with the modelling of the galactic subhalo population. We use four different halo models bracketing a large set of uncertainties. For each model, adopting an accurate detection threshold of the LAT to dark matter subhalo signals and comparing model predictions with the number of unassociated point-sources in Fermi-LAT catalogues, we derive upper limits on the annihilation cross section as a function of dark matter mass. Our results show that, even in the best-case scenario (i.e., DMonly subhalo model), which does not include tidal disruption from baryons, the limits on the dark matter parameter space are less stringent than current gamma-ray limits from dwarf spheroidal galaxies. Comparing the results obtained with the different subhalo models, we find that baryonic effects on the subhalo population are significant and lead to dark matter constraints that are less stringent by a factor of ∼2 to ∼5. This uncertainty comes from the unknown resilience of dark matter subhaloes to tidal disruption. 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>Scatter plot of <math display="inline"><semantics> <mi mathvariant="script">J</mi> </semantics></math>-factor values, <math display="inline"><semantics> <mi mathvariant="script">J</mi> </semantics></math> within 0.1<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>, as a function of subhalo mass, <math display="inline"><semantics> <msub> <mi>M</mi> <mi>SH</mi> </msub> </semantics></math>, in one realisation of the Monte Carlo simulations for each subhalo model—<b>top left</b>: DMonly, <b>top right</b>: Phat-ELVIS, <b>bottom left</b>: <tt>SL17-fragile</tt>, <b>bottom right</b>: SL17-resilient. The colour-bar represents the subhalo distance from Earth, hereafter <math display="inline"><semantics> <msub> <mi>d</mi> <mi>SH</mi> </msub> </semantics></math>. The realisation shown is the one containing the lowest mass subhalo. We remind that we have applied a cut of <math display="inline"><semantics> <mi mathvariant="script">J</mi> </semantics></math> (&lt;0.1<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>) &gt; 10<math display="inline"><semantics> <msup> <mrow/> <mn>17</mn> </msup> </semantics></math><math display="inline"><semantics> <mrow> <msup> <mi>GeV</mi> <mn>2</mn> </msup> <msup> <mi>cm</mi> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Same as in <a href="#galaxies-07-00090-f001" class="html-fig">Figure 1</a> displaying all subhaloes selected as grey dots and those which would be detectable in the 2FHL catalogue as coloured points. The results are shown for a DM mass of 100 GeV, <span class="html-italic">b</span>-quark annihilation, and a cross section of <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>24</mn> </mrow> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>/s.</p>
Full article ">Figure 3
<p>Same as in <a href="#galaxies-07-00090-f001" class="html-fig">Figure 1</a> displaying all subhaloes selected as grey dots and those which would be detectable in the 2FHL catalogue as coloured points. The results are shown for a DM mass of 1.5 TeV, <span class="html-italic">b</span>-quark annihilation, and a cross section of <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>22</mn> </mrow> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mn>3</mn> </msup> </semantics></math>/s.</p>
Full article ">Figure 4
<p>For the same realisations as in <a href="#galaxies-07-00090-f002" class="html-fig">Figure 2</a>, we display the corresponding all-sky gamma-ray maps of the selected haloes. Fluxes are computed assuming a DM mass of 100 GeV and an annihilation cross section into <math display="inline"><semantics> <mrow> <mi>b</mi> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> </mrow> </semantics></math> of <math display="inline"><semantics> <mrow> <mn>5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>24</mn> </mrow> </msup> <mspace width="0.166667em"/> <msup> <mi>cm</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>, for the 3FGL catalogue set-up. We overlay the LAT sensitivity threshold curves at fixed latitude values. The orange circles indicate the subhaloes that are above threshold, and that would therefore be detectable in the 3FGL catalogue. <b>Top left</b>: DMonly; <b>top right</b>: Phat-ELVIS; <b>bottom left</b>: SL17-fragile; <b>bottom right</b>: SL17-resilient. The orange circles indicating the detectable subhaloes have a diameter of <math display="inline"><semantics> <msup> <mn>7</mn> <mo>∘</mo> </msup> </semantics></math>.</p>
Full article ">Figure 5
<p>Upper limits on the dark matter (DM) annihilation cross-section, <math display="inline"><semantics> <mrow> <mo>〈</mo> <mi>σ</mi> <mspace width="3.33333pt"/> <mi>v</mi> <mo>〉</mo> </mrow> </semantics></math>, from the observation of 16 (4) DM subhalo candidates, <math display="inline"><semantics> <msub> <mi>N</mi> <mi>cand</mi> </msub> </semantics></math>, in the 3FGL (2FHL) catalogue (the number of candidates is taken from [<a href="#B16-galaxies-07-00090" class="html-bibr">16</a>]). We show the limit for DMonly (purple curve), Phat-ELVIS (red dashed curve), SL17-resilient (blue dotted-dashed curve) and SL17-fragile (green dotted curve). The same colour-code applies to uncertainty bands which represent the spread due to the 1000 Monte Carlo realisations for each subhalo model. <span class="html-italic">Top left (right) panel</span>: Annihilation into <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>) for the 3FGL catalogue. <span class="html-italic">Bottom left (right) panel</span>: Annihilation into <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>) for the 2FHL catalogue.</p>
Full article ">Figure 6
<p>Upper limits on the DM annihilation cross section, <math display="inline"><semantics> <mrow> <mo>〈</mo> <mi>σ</mi> <mspace width="3.33333pt"/> <mi>v</mi> <mo>〉</mo> </mrow> </semantics></math>, from the observation of 16 (4) DM subhalo candidates, <math display="inline"><semantics> <msub> <mi>N</mi> <mi>cand</mi> </msub> </semantics></math>, in the 3FGL—purple solid—(2FHL—red solid) catalogue, for the DMonly. The sensitivity reach (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>cand</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>) of the 3FGL (2FHL) is also shown by the turquoise dashed-dotted line (blue dashed-dotted line). The same colour-code applies to uncertainty bands which represent the spread due to the 1000 Monte Carlo realisations of the subhalo model. Left (Right) panel: Annihilation into <span class="html-italic">b</span>-quark (<math display="inline"><semantics> <mi>τ</mi> </semantics></math> lepton) finale states. Overlaid, the limits from gamma-ray observations towards dwarf spheroidal galaxies from Albert et al. 2016 [<a href="#B38-galaxies-07-00090" class="html-bibr">38</a>] (black dotted), and Calore et al. 2018 [<a href="#B9-galaxies-07-00090" class="html-bibr">9</a>] (red dotted).</p>
Full article ">
28 pages, 5305 KiB  
Article
An In-Depth Investigation of Faraday Depth Spectrum Using Synthetic Observations of Turbulent MHD Simulations
by Aritra Basu, Andrew Fletcher, Sui Ann Mao, Blakesley Burkhart, Rainer Beck and Dominic Schnitzeler
Galaxies 2019, 7(4), 89; https://doi.org/10.3390/galaxies7040089 - 23 Nov 2019
Cited by 16 | Viewed by 3611
Abstract
In this paper, we present a detailed analysis of the Faraday depth (FD) spectrum and its clean components obtained through the application of the commonly used technique of Faraday rotation measure synthesis to analyze spectro-polarimetric data. To directly compare the Faraday depth spectrum [...] Read more.
In this paper, we present a detailed analysis of the Faraday depth (FD) spectrum and its clean components obtained through the application of the commonly used technique of Faraday rotation measure synthesis to analyze spectro-polarimetric data. To directly compare the Faraday depth spectrum with physical properties of a magneto-ionic medium, we generated synthetic broad-bandwidth spectro-polarimetric observations from magnetohydrodynamic (MHD) simulations of a transonic, isothermal, compressible turbulent medium. We find that correlated magnetic field structures give rise to a combination of spiky, localized peaks at certain FD values, and broad structures in the FD spectrum. Although most of these spiky FD structures appear narrow, giving an impression of a Faraday thin medium, we show that they arise from strong synchrotron emissivity at that FD. Strong emissivity at a FD can arise because of both strong spatially local polarized synchrotron emissivity at a FD or accumulation of weaker emissions along the distance through a medium that have Faraday depths within half the width of the rotation measure spread function. Such a complex Faraday depth spectrum is a natural consequence of MHD turbulence when the lines of sight pass through a few turbulent cells. This therefore complicates the convention of attributing narrow FD peaks to the presence of a Faraday-rotating medium along the line of sight. Our work shows that it is difficult to extract the FD along a line of sight from the Faraday depth spectrum using standard methods for a turbulent medium in which synchrotron emission and Faraday rotation occur simultaneously. Full article
(This article belongs to the Special Issue New Perspectives on Galactic Magnetism)
Show Figures

Figure 1

Figure 1
<p>Synthetic spectra of Stokes parameters of synchrotron emission generated by <tt>COSMIC</tt> for a uniform slab. Numerically computed quantities are shown as the data points and the analytical functions for the linearly polarized quantities are shown as dashed lines. <span class="html-italic"><b>Top left</b></span>: Spectrum of the total synchrotron flux density (grey points) and the linearly polarized flux density (blue points). <span class="html-italic"><b>Top right</b></span>: Variation of the factional polarization with <math display="inline"><semantics> <msup> <mi>λ</mi> <mn>2</mn> </msup> </semantics></math>. <span class="html-italic"><b>Bottom left</b></span>: Variation of fractional Stokes <span class="html-italic">Q</span> and <span class="html-italic">U</span> parameters as a function of <math display="inline"><semantics> <msup> <mi>λ</mi> <mn>2</mn> </msup> </semantics></math>. <span class="html-italic"><b>Bottom right</b></span>: Variation of the angle of the plane of linear polarization (<math display="inline"><semantics> <mi>θ</mi> </semantics></math>) with <math display="inline"><semantics> <msup> <mi>λ</mi> <mn>2</mn> </msup> </semantics></math>.</p>
Full article ">Figure 2
<p>Synthetic spectra of Stokes parameters of synchrotron emission generated by <tt>COSMIC</tt> for an internal Faraday dispersion model. Different shades of data points in the inset of the top right panel show the fractional polarization computed by averaging over different areas. All other panels, data points and lines have the same meaning as in <a href="#galaxies-07-00089-f001" class="html-fig">Figure 1</a>. In the main plots only the results for averaging over 50 × 50 pixel<sup>2</sup> are shown.</p>
Full article ">Figure 3
<p>Distribution of strengths of the three components of the magnetic field from the isothermal, compressible turbulence simulations used for our analysis are shown as the histograms. The dashed lines show Gaussian distributions with mean and standard deviation computed from the corresponding field component.</p>
Full article ">Figure 4
<p><span class="html-italic"><b>Left</b></span>: Variation of emission measure (EM) as a function of dispersion measure (DM) for an ionization fraction of 0.5 of the simulation volume. The different shades for the symbols represent different averaging scales. The black line shows the best-fit, and the grey dashed and dotted lines are from Berkhuijsen and Müller [<a href="#B56-galaxies-07-00089" class="html-bibr">56</a>] and Pynzar’ [<a href="#B57-galaxies-07-00089" class="html-bibr">57</a>], respectively. <span class="html-italic"><b>Right</b></span>: Distribution of free electron density <math display="inline"><semantics> <msub> <mi>n</mi> <mi mathvariant="normal">e</mi> </msub> </semantics></math> in the simulation volume. The solid black line shows the best-fit gamma function representation of the distribution.</p>
Full article ">Figure 5
<p><span class="html-italic"><b>Left</b></span>: 3-D synchrotron emissivity per pixel at 1 GHz. Synchrotron emissivity was computed assuming a constant density of CREs throughout the 3-D volume. <span class="html-italic"><b>Right</b></span>: 2-D synchrotron intensity at 1 GHz integrated along <span class="html-italic">z</span>-axis.</p>
Full article ">Figure 6
<p>Linearly polarized intensities at 0.5 GHz (<b>left</b>), 1.5 GHz (<b>middle</b>) and 5 GHz (<b>right</b>) including the effects of Faraday depolarization.</p>
Full article ">Figure 7
<p>Synthetic spectra of Stokes parameters of synchrotron emission generated by <tt>COSMIC</tt> from MHD simulations. Here we have used <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mo>−</mo> <mn>0.8</mn> </mrow> </semantics></math> and assumed optically thin synchrotron emission. Quantities computed from <tt>COSMIC</tt> are shown as the data points. <span class="html-italic"><b>Top left</b></span>: Spectrum of the total synchrotron intensity is shown as the grey points and the linearly polarized intensity is shown as the blue points. <span class="html-italic"><b>Top right</b></span>: Variation of the factional polarization with <math display="inline"><semantics> <msup> <mi>λ</mi> <mn>2</mn> </msup> </semantics></math>. <span class="html-italic"><b>Bottom left</b></span>: Variation of Stokes <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>/</mo> <msub> <mi>I</mi> <mi>sync</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>U</mi> <mo>/</mo> <msub> <mi>I</mi> <mi>sync</mi> </msub> </mrow> </semantics></math> parameters as a function of <math display="inline"><semantics> <msup> <mi>λ</mi> <mn>2</mn> </msup> </semantics></math>. <span class="html-italic"><b>Bottom right</b></span>: Variation of the angle of the plane of linear polarization (<math display="inline"><semantics> <mi>θ</mi> </semantics></math>) with <math display="inline"><semantics> <msup> <mi>λ</mi> <mn>2</mn> </msup> </semantics></math>. All the dashed lines represents the best-fit obtained from Stokes <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>,</mo> <mi>U</mi> </mrow> </semantics></math> fitting. The insets shows the wavelength regime below 700 cm<sup>2</sup> within which successful fits were obtained (see text for details).</p>
Full article ">Figure 8
<p>Comparison of the Faraday depth map computed from the MHD simulations (<b>left</b>) with the Faraday depth map reconstructed from RM synthesis applied to synthetic observations in the frequency range 0.5 to 6 GHz (<b>right</b>).</p>
Full article ">Figure 9
<p><span class="html-italic"><b>Left</b></span>: Distribution of the difference between FD estimated by applying RM synthesis (<math display="inline"><semantics> <msub> <mi>FD</mi> <mrow> <mi>RM</mi> <mspace width="0.166667em"/> <mi>synthesis</mi> </mrow> </msub> </semantics></math>) and the FD obtained from the MHD simulation (<math display="inline"><semantics> <msub> <mi>FD</mi> <mi>MHD</mi> </msub> </semantics></math>) computed using Equation (<a href="#FD9-galaxies-07-00089" class="html-disp-formula">A4</a>). The different colors for the histograms are for different frequency coverages. <span class="html-italic"><b>Right</b></span>: RMSF for the corresponding frequency coverages shown in the left-hand panel.</p>
Full article ">Figure 10
<p>Faraday depth spectra for the uniform slab (<b>left</b>) and internal Faraday dispersion (IFD) models (<b>right</b>). The red lines are the clean components. The blue dots show the relative intrinsic synchrotron emissivities at respective FD computed from the simulated volume described in <a href="#sec4-galaxies-07-00089" class="html-sec">Section 4</a> and represents a model of the intrinsic Faraday depth spectrum. For the IFD model, we have averaged 10 × 10 pixel<sup>2</sup> pixels.</p>
Full article ">Figure 11
<p><span class="html-italic"><b>Left</b></span>: Variation of <math display="inline"><semantics> <msub> <mi>B</mi> <mo>⊥</mo> </msub> </semantics></math> (circles) and cumulative FD (blue line) with distance along three random LOS each of which is one pixel wide. <math display="inline"><semantics> <msub> <mi>B</mi> <mo>⊥</mo> </msub> </semantics></math> is colored based on the angle of the polarized synchrotron emission. <span class="html-italic"><b>Middle</b></span>: Faraday depth spectrum for the corresponding LOS. The blue points show the local polarized emissivity at that FD. The red lines are the clean FD components and the green dot-dashed line is located at the FD along that LOS. The black points show the summed linearly polarized synchrotron emissivity per FD bin of size half the RMSF. The amplitudes of the polarized emissivities are normalized to the peak FD clean component. <span class="html-italic"><b>Right</b></span>: Same as the middle-panel zoomed around the peak of the Faraday depth spectrum for a clearer visualization.</p>
Full article ">
23 pages, 4750 KiB  
Review
Massive Stars in the Tarantula Nebula: A Rosetta Stone for Extragalactic Supergiant HII Regions
by Paul A. Crowther
Galaxies 2019, 7(4), 88; https://doi.org/10.3390/galaxies7040088 - 8 Nov 2019
Cited by 40 | Viewed by 4839
Abstract
A review of the properties of the Tarantula Nebula (30 Doradus) in the Large Magellanic Cloud is presented, primarily from the perspective of its massive star content. The proximity of the Tarantula and its accessibility to X-ray through radio observations permit it to [...] Read more.
A review of the properties of the Tarantula Nebula (30 Doradus) in the Large Magellanic Cloud is presented, primarily from the perspective of its massive star content. The proximity of the Tarantula and its accessibility to X-ray through radio observations permit it to serve as a Rosetta Stone amongst extragalactic supergiant HII regions since one can consider both its integrated characteristics and the individual properties of individual massive stars. Recent surveys of its high mass stellar content, notably the VLT FLAMES Tarantula Survey (VFTS), are reviewed, together with VLT/MUSE observations of the central ionizing region NGC 2070 and HST/STIS spectroscopy of the young dense cluster R136, provide a near complete Hertzsprung-Russell diagram of the region, and cumulative ionizing output. Several high mass binaries are highlighted, some of which have been identified from a recent X-ray survey. Brief comparisons with the stellar content of giant HII regions in the Milky Way (NGC 3372) and Small Magellanic Cloud (NGC 346) are also made, together with Green Pea galaxies and star forming knots in high-z galaxies. Finally, the prospect of studying massive stars in metal poor galaxies is evaluated. Full article
(This article belongs to the Special Issue Luminous Stars in Nearby Galaxies)
Show Figures

Figure 1

Figure 1
<p>(<b>left</b>) Optical image of the Tarantula Nebula from the MPG/ESO 2.2m WFI, with NGC 2060 and SN1987A indicated; (<b>centre</b>) Optical VLT/FORS2 image centred on NGC 2070, with Hodge 301 to the upper right; (<b>right</b>) an infrared VLT/MAD image of the central R136 region, with the massive colliding wind binary Mk 34 indicated. Credit: ESO/P. Crowther/C.J. Evans.</p>
Full article ">Figure 2
<p>(<b>left</b>) Chandra ACIS X-ray logarithmic intensity image of the core of NGC 2070 from T-ReX, centred on R136c, adapted from [<a href="#B38-galaxies-07-00088" class="html-bibr">38</a>], showing the relative brightness of the colliding wind binary Melnick 34 (WNh5 + WN5h) [<a href="#B37-galaxies-07-00088" class="html-bibr">37</a>] to the R136a star cluster (hosting multiple WN5h stars) and R136c (WN5h+?); (<b>right</b>) HST WFC3/F555W logarithmic intensity image of the same 19 × 19 arcsec region, highlighting the rich stellar population of R136a with respect to R136c and Melnick 34.</p>
Full article ">Figure 3
<p>Hertzsprung-Russell diagram of the Tarantula Nebula, based on results from VLT FLAMES Tarantula Survey (VFTS) [<a href="#B50-galaxies-07-00088" class="html-bibr">50</a>,<a href="#B51-galaxies-07-00088" class="html-bibr">51</a>,<a href="#B52-galaxies-07-00088" class="html-bibr">52</a>,<a href="#B53-galaxies-07-00088" class="html-bibr">53</a>,<a href="#B54-galaxies-07-00088" class="html-bibr">54</a>], MUSE [<a href="#B11-galaxies-07-00088" class="html-bibr">11</a>], Hubble Space Telescope (HST)/STIS [<a href="#B55-galaxies-07-00088" class="html-bibr">55</a>] and other literature results, with typical uncertainties from each survey indicated. Filled symbols are within NGC 2070, open symbols elsewhere in the Tarantula. Non-rotating tracks for 10, 15, 25, 40, 60, 100 and 200 <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math> LMC metallicity stars have been included from [<a href="#B60-galaxies-07-00088" class="html-bibr">60</a>,<a href="#B61-galaxies-07-00088" class="html-bibr">61</a>] which terminate at the onset of He-burning.</p>
Full article ">Figure 4
<p>Pie charts, courtesy Hugues Sana and Selma de Mink, illustrating the fraction of O stars undergoing single stellar evolution versus mergers, primaries being stripped of their envelopes, and secondaries being spun up, for Milky Way young clusters (left) and VFTS O stars (right), adapted from [<a href="#B63-galaxies-07-00088" class="html-bibr">63</a>,<a href="#B64-galaxies-07-00088" class="html-bibr">64</a>].</p>
Full article ">Figure 5
<p>Cumulative distribution of rotational rates for single VFTS O (red) and B stars (blue) [<a href="#B69-galaxies-07-00088" class="html-bibr">69</a>,<a href="#B71-galaxies-07-00088" class="html-bibr">71</a>].</p>
Full article ">Figure 6
<p>Unclumped mass-loss rates of O-type, Of/WN and Wolf-Rayet stars in the Tarantula Nebula (based on results from VFTS [<a href="#B50-galaxies-07-00088" class="html-bibr">50</a>,<a href="#B51-galaxies-07-00088" class="html-bibr">51</a>,<a href="#B53-galaxies-07-00088" class="html-bibr">53</a>], HST/STIS [<a href="#B55-galaxies-07-00088" class="html-bibr">55</a>] and other surveys [<a href="#B37-galaxies-07-00088" class="html-bibr">37</a>,<a href="#B56-galaxies-07-00088" class="html-bibr">56</a>,<a href="#B59-galaxies-07-00088" class="html-bibr">59</a>] for WR stars). Filled symbols are within NGC 2070, open symbols elsewhere in the Tarantula. Theoretical mass-loss rates for zero age main sequence massive stars at the Large Magellanic Cloud (LMC) metallicity [<a href="#B83-galaxies-07-00088" class="html-bibr">83</a>] are included (solid line), based on LMC metallicity evolutionary models [<a href="#B60-galaxies-07-00088" class="html-bibr">60</a>,<a href="#B61-galaxies-07-00088" class="html-bibr">61</a>].</p>
Full article ">Figure 7
<p>Comparison between observed He <span class="html-small-caps">ii</span> <math display="inline"><semantics> <mi>λ</mi> </semantics></math>1640 emission equivalent widths in R136 [<a href="#B5-galaxies-07-00088" class="html-bibr">5</a>] versus predicted emission from BPASS (v.2.2.1, red) and Starburst99 (blue) population synthesis models (absorption lines are shown as negative values).</p>
Full article ">Figure 8
<p>Cumulative ionizing output (10<math display="inline"><semantics> <msup> <mrow/> <mn>50</mn> </msup> </semantics></math> ph/s) from spectroscopically classified early-type stars in the Tarantula, obtained from VFTS [<a href="#B50-galaxies-07-00088" class="html-bibr">50</a>,<a href="#B51-galaxies-07-00088" class="html-bibr">51</a>,<a href="#B53-galaxies-07-00088" class="html-bibr">53</a>], VLT/MUSE [<a href="#B11-galaxies-07-00088" class="html-bibr">11</a>], HST/STIS [<a href="#B55-galaxies-07-00088" class="html-bibr">55</a>] and literature results [<a href="#B56-galaxies-07-00088" class="html-bibr">56</a>], updated from [<a href="#B12-galaxies-07-00088" class="html-bibr">12</a>]. Specific regions within 30 Dor are indicated from <a href="#galaxies-07-00088-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 9
<p>Comparison between the integrated star-formation rate versus size of the Tarantula (filled red square) and star-forming knots from galaxies spanning a range of redshifts, adapted from [<a href="#B106-galaxies-07-00088" class="html-bibr">106</a>].</p>
Full article ">Figure 10
<p>BPT diagram illustrating the similarity in integrated strengths between the Tarantula Nebula (red square), Green Pea (green circles), extreme Green Peas (blue diamonds), Lyman-continuum leaking Green Peak (pink triangles), updated from [<a href="#B108-galaxies-07-00088" class="html-bibr">108</a>], plus SDSS star-forming galaxies.</p>
Full article ">Figure 11
<p>Comparison between present-day star formation rates, as measured by H<math display="inline"><semantics> <mi>α</mi> </semantics></math> luminosity [<a href="#B113-galaxies-07-00088" class="html-bibr">113</a>], distance modulus (mag), and oxygen metal content (squares: ≥20% of solar value, triangles: &lt;20% of solar value, for Local Group dwarf galaxies. Metal-poor galaxies possess low star-formation rates, so host small numbers of OB stars, and these are ≥6 magnitudes fainter than Magellanic Cloud counterparts.</p>
Full article ">Figure 12
<p>Comparison between far UV spectroscopy of mid O giants in metal-rich [<a href="#B117-galaxies-07-00088" class="html-bibr">117</a>] and metal-deficient [<a href="#B116-galaxies-07-00088" class="html-bibr">116</a>] environments, illustrating the extreme differences in wind features (e.g., N<span class="html-small-caps">v</span> <math display="inline"><semantics> <mi>λ</mi> </semantics></math>1240, Si <span class="html-small-caps">iv</span><math display="inline"><semantics> <mi>λ</mi> </semantics></math>1400, C <span class="html-small-caps">iv</span><math display="inline"><semantics> <mi>λ</mi> </semantics></math>1550) and the iron forest (Fe <span class="html-small-caps">iv-v</span>).</p>
Full article ">
30 pages, 5281 KiB  
Review
Relativistic Jets in Gamma-Ray-Emitting Narrow-Line Seyfert 1 Galaxies
by Filippo D’Ammando
Galaxies 2019, 7(4), 87; https://doi.org/10.3390/galaxies7040087 - 7 Nov 2019
Cited by 22 | Viewed by 4129
Abstract
Before the launch of the Fermi Gamma-ray Space Telescope satellite only two classes of active galactic nuclei (AGN) were known to generate relativistic jets and thus to emit up to the γ -ray energy range: blazars and radio galaxies, both hosted in giant [...] Read more.
Before the launch of the Fermi Gamma-ray Space Telescope satellite only two classes of active galactic nuclei (AGN) were known to generate relativistic jets and thus to emit up to the γ -ray energy range: blazars and radio galaxies, both hosted in giant elliptical galaxies. The discovery by the Large Area Telescope (LAT) on-board the Fermi satellite of variable γ -ray emission from a few radio-loud narrow-line Seyfert 1 galaxies (NLSy1) revealed the presence of an emerging third class of AGN with powerful relativistic jets. Considering that NLSy1 are usually hosted in late-type galaxies with relatively small black hole masses, this finding opened new challenging questions about the nature of these objects, the disc/jet connection, the emission mechanisms at high energies, and the formation of relativistic jets. In this review, I will discuss the broad-band properties of the γ -ray-emitting NLSy1 included in the Fourth Fermi LAT source catalog, highlighting major findings and open questions regarding jet physics, black hole mass estimation, host galaxy and accretion process of these sources in the Fermi era. Full article
(This article belongs to the Special Issue Particle Acceleration Processes in Astrophysical Jets)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Left panel</span>: LAT light curve of SBS 0846+513 in the 0.1–100 GeV energy range during 1 April–28 August 2012 with 7 days or 1 day (shown as red squares) time bins. Adapted from [<a href="#B38-galaxies-07-00087" class="html-bibr">38</a>]. <span class="html-italic">Right panel</span>: LAT light curve of PKS 1502+036 in the 0.1–300 GeV energy range during 11 December 2015–9 January 2016, with 1-d time bins (top panel), 12-h time bins (middle panel), and 6-h time bins (bottom panel). Adapted from [<a href="#B62-galaxies-07-00087" class="html-bibr">62</a>]. In both panels arrow refers to 2-<math display="inline"><semantics> <mi>σ</mi> </semantics></math> upper limits. Upper limits are computed when <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>S</mi> </mrow> </semantics></math> &lt; 10.</p>
Full article ">Figure 2
<p><span class="html-italic">Left panel</span>: LAT light curve of PMN J0948+0022 in the 0.1–100 GeV energy range during 1 December 2012–31 January 2013, with 7-day time bins. Red solid squares represent daily fluxes reported for the period of high activity. The horizontal line indicates the period of the VERITAS observation. Adapted from [<a href="#B61-galaxies-07-00087" class="html-bibr">61</a>]. <span class="html-italic">Right panel</span>: SED of PMN J0948+0022 in the MeV-to-TeV energy range. The LAT spectrum was extrapolated to the TeV energies and corrected for EBL absorption using the model of [<a href="#B93-galaxies-07-00087" class="html-bibr">93</a>]. <span class="html-italic">Fermi</span>-LAT and VERITAS data points and upper limits are shown. Adapted from [<a href="#B61-galaxies-07-00087" class="html-bibr">61</a>]. In both panels, arrow refers to 2-<math display="inline"><semantics> <mi>σ</mi> </semantics></math> upper limits. Upper limits are computed when <math display="inline"><semantics> <mrow> <mi>T</mi> <mi>S</mi> </mrow> </semantics></math> &lt; 10.</p>
Full article ">Figure 3
<p><span class="html-italic">Left panel</span>: <span class="html-italic">XMM-Newton</span> EPIC pn (black), MOS1 (red), and MOS2 (green) data of PMN J0948+0022 collected on 28–29 May 2011 shown as a ratio to a power law with photon index <math display="inline"><semantics> <mo>Γ</mo> </semantics></math> = 1.48. Adapted from [<a href="#B103-galaxies-07-00087" class="html-bibr">103</a>]. <span class="html-italic">Right panel</span>: data/model ratio for the <span class="html-italic">XMM-Newton</span> EPIC pn X-ray spectrum of 1H 0323+342 collected on 23–24 August 2015. A model including two power-laws and two Gaussian Iron emission lines with the Galactic value in the direction of the source fixed to N<math display="inline"><semantics> <msub> <mrow/> <mrow> <mi mathvariant="normal">H</mi> <mo>,</mo> <mspace width="0.166667em"/> <mi>Gal</mi> </mrow> </msub> </semantics></math> = 2.33 × 10<math display="inline"><semantics> <msup> <mrow/> <mn>21</mn> </msup> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math> is applied to the data. Adapted from [<a href="#B98-galaxies-07-00087" class="html-bibr">98</a>].</p>
Full article ">Figure 4
<p>Photon indices obtained from broken power-law fits to <span class="html-italic">XMM-Newton</span> spectra of <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray-emitting NLSy1. <math display="inline"><semantics> <msub> <mo>Γ</mo> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mo>Γ</mo> <mn>2</mn> </msub> </semantics></math> are the photon indices below and above the break, respectively. One of the sources is best described by a simple power-law. In case of 1H 0323+342 the broken power-law is just an approximation, as the spectrum show additional complexity. Adapted from [<a href="#B99-galaxies-07-00087" class="html-bibr">99</a>].</p>
Full article ">Figure 5
<p><span class="html-italic">Left panel</span>: VLBA image at 15 GHz of SBS 0846+513. Component W1 corresponds to the core region, component W2 to a knot in the jet, component E is an extended low-surface brightness structure with a steep spectrum. Adapted from [<a href="#B56-galaxies-07-00087" class="html-bibr">56</a>]. <span class="html-italic">Right panel</span>: VLBA image at 15 GHz of PMN J0948+0022 collected on 2011 May 26. The vectors superimposed on the total intensity contours show the percentage polarization and the position angle of the electric vector. Adapted from [<a href="#B103-galaxies-07-00087" class="html-bibr">103</a>]. In both images, it is provided the restoring beam, plotted in the bottom-left corner, the peak flux density in mJy/beam and the first contour (f.c.) intensity in mJy/beam, which is three times the off-source noise level. Contour levels increase by a factor of 2.</p>
Full article ">Figure 6
<p><span class="html-italic">Left panel</span>: The separation between the core component of SBS 0846+513 and the knot ejected in 2009 as a function of time. The solid line represents the regression fit to the 15 GHz VLBA MOJAVE data, while the dotted lines represent the uncertainties from the fit parameters. Dashed lines indicate the beginning and the peak of the radio flare observed by OVRO in 2009. Adapted from [<a href="#B38-galaxies-07-00087" class="html-bibr">38</a>]. <span class="html-italic">Right panel</span>: Jet width profile of 1H 0323+342 as a function of (projected) distance <span class="html-italic">z</span> from the 43 GHz core. The dashed line (A) represents a best-fit model for the inner jet (<span class="html-italic">z</span> &lt; 6.5 mas), indicating a parabolic collimating jet. The dashed line (B) represents a best-fit model for the outer (<span class="html-italic">z</span> &gt; 7.5 mas) jet, indicating a hyperbolic expanding jet. The vertical dashed line (C) indicates the location of the quasi-stationary feature S. Adapted from [<a href="#B143-galaxies-07-00087" class="html-bibr">143</a>].</p>
Full article ">Figure 7
<p>SED of the four NLSy1 detected by <span class="html-italic">Fermi</span>-LAT during the first year of operation. The synchrotron self-absorption is clearly visible around 10<math display="inline"><semantics> <msup> <mrow/> <mrow> <mn>11</mn> <mo>−</mo> <mn>12</mn> </mrow> </msup> </semantics></math> Hz. The short dashed light blue line indicates the synchrotron component, while the long dashed orange line is the synchrotron-self Compton emission. The dot-dashed line refers to external Compton emission and the dotted black line represents the contribution of the accretion disc, X-ray corona and the infrared torus. The continuous blue line is the sum of all the contributions. Adapted from [<a href="#B8-galaxies-07-00087" class="html-bibr">8</a>].</p>
Full article ">Figure 8
<p>SED and models for the 2013 and 2011 activity states of PMN J0948+0022. The filled circles are the data from the 2013 flaring state, and the open squares are the data from the 2011 intermediate state taken from [<a href="#B103-galaxies-07-00087" class="html-bibr">103</a>]. The dashed curve indicates the disc and coronal emission, and the dotted line indicates the thermal dust emission. Solid lines represent models consistent with scattering dust torus radiation, while the dashed-dotted curve represents a model consistent with the scattering of BLR radiation. Arrows refer to 2<math display="inline"><semantics> <mi>σ</mi> </semantics></math> upper limits on the source flux. The VERITAS upper limits are corrected for EBL absorption using the model of [<a href="#B93-galaxies-07-00087" class="html-bibr">93</a>]. Adapted from [<a href="#B61-galaxies-07-00087" class="html-bibr">61</a>].</p>
Full article ">Figure 9
<p>SED data (squares) and model fit (solid curve) of SBS 0846+513 (<span class="html-italic">left panel</span>) and PKS 1502+036 (<span class="html-italic">right panel</span>) in flaring activity with the thermal emission components shown as dashed curves. The data points were collected by OVRO at 15 GHz, <span class="html-italic">Swift</span> (UVOT and XRT) and <span class="html-italic">Fermi</span>-LAT. The SED in the quiescent state (taken from [<a href="#B56-galaxies-07-00087" class="html-bibr">56</a>]) and average state, respectively, ares shown as circles. Adapted from [<a href="#B147-galaxies-07-00087" class="html-bibr">147</a>] (<span class="html-italic">left panel</span>) and [<a href="#B62-galaxies-07-00087" class="html-bibr">62</a>] (<span class="html-italic">right panel</span>).</p>
Full article ">Figure 10
<p>Two-dimensional surface-brightness profile decomposition of 1H 0323+342 in <span class="html-italic">J</span>-band for Model A (PSF+Bulge; <span class="html-italic">left panel</span>) and Model B (PSF+Bulge+Disc; <span class="html-italic">right panel</span>). <span class="html-italic">Top left subpanel</span>: the observed image in a field of view of 20 arcsec × 20 arcsec. <span class="html-italic">Middle left subpanel</span>: model used to describe the surface brightness distribution. <span class="html-italic">Bottom left subpanel</span>: the residual image. <span class="html-italic">Top right subpanel</span>: radial profile of the surface brightness distribution. The filled circles show the observations, and the solid, pointed, and dashed lines represent the model, PSF, and host galaxy, respectively. The exponential disk component is shown in orange. <span class="html-italic">Bottom right subpanel</span>: residuals. Adapted from [<a href="#B181-galaxies-07-00087" class="html-bibr">181</a>].</p>
Full article ">Figure 11
<p><span class="html-italic">Left panels</span>: central 13 × 13 arcsec<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math> of the images in the <span class="html-italic">J</span> and <math display="inline"><semantics> <msub> <mi>K</mi> <mi>s</mi> </msub> </semantics></math> band of PKS 1502+036, top and bottom, respectively. <span class="html-italic">Center panels</span>: GALFIT models using a Sérsic profile combined with a nuclear PSF. <span class="html-italic">Right panels</span>: residual images after subtracting the model. Colour bars are in mag arcsec<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math> (left-hand and centre panels). Adapted from [<a href="#B179-galaxies-07-00087" class="html-bibr">179</a>].</p>
Full article ">Figure 12
<p>Surface brightness decomposition in the <span class="html-italic">J</span>-band of PKS 1502+036. The observed profiles are the black dots, the nuclear PSF is the dot–dashed light blue curve, the bulge component is reproduced with a Sérsic model (green dot line). In the bottom panels the residuals are shown. In the residual panel crosses represent bulge + disc component. Adapted from [<a href="#B179-galaxies-07-00087" class="html-bibr">179</a>].</p>
Full article ">
13 pages, 1229 KiB  
Article
Features of Structure and Absorption in the Jet-Launching Region of M87
by Wei Zhao, Xiaoyu Hong, Tao An, Xiaofeng Li, Xiaopeng Cheng and Fang Wu
Galaxies 2019, 7(4), 86; https://doi.org/10.3390/galaxies7040086 - 31 Oct 2019
Cited by 3 | Viewed by 2823
Abstract
M87 is one of the best available source for studying the AGN jet-launching region. To enrich our knowledge of this region, with quasi-simultaneous observations using VLBA at 22, 43 and 86 GHz, we capture the images of the radio jet in M87 on [...] Read more.
M87 is one of the best available source for studying the AGN jet-launching region. To enrich our knowledge of this region, with quasi-simultaneous observations using VLBA at 22, 43 and 86 GHz, we capture the images of the radio jet in M87 on a scale within several thousand R s . Based on the images, we analyze the transverse jet structure and obtain the most accurate spectral-index maps of the jet in M87 so far, then for the first time, we compare the results of the two analyses and find a spatial association between the jet collimations and the local enhancement of the density of external medium in the jet-launching region. We also find the external medium is not uniform, and greatly contributes to the free-free absorption in this region. In addition, we find for the jet in M87, its temporal morphology in the launching region may be largely affected by the local, short-lived kink instability growing in itself. Full article
Show Figures

Figure 1

Figure 1
<p>22 and 43 GHz contour plots overlapping with pseudo-color images of the same frequency respectively, with the synthesized beam shown at the top-left corner of each panel. Panel (<b>a</b>) shows the image of M87 at 22 GHz with a synthesized beam of 0.93 × 0.45 mas at a position angle of 12.6<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>; contours in this image are from (−1, 1, 2, 4, 8…)<math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <mn>0.86</mn> <mspace width="3.33333pt"/> <mi>mJy</mi> <mspace width="3.33333pt"/> <msup> <mi>beam</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>. Panel (<b>b</b>) shows the image of M87 at 43 GHz with a synthesized beam of 0.56 × 0.23 mas at a position angle of 16.7<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>; contours in this image are from (−1, 1, 2, 4, 8…) <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <mn>0.54</mn> <mspace width="3.33333pt"/> <mi>mJy</mi> <mspace width="3.33333pt"/> <msup> <mi>beam</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>86 GHz contour plot overlapping with pseudo-color image of the same frequency of M87; contours are from (−1, 1, 2, 4, 8…) <math display="inline"><semantics> <mrow> <mo>×</mo> <mspace width="3.33333pt"/> <mn>0.60</mn> <mspace width="3.33333pt"/> <mi>mJy</mi> <mspace width="3.33333pt"/> <msup> <mi>beam</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>. The synthesized beam of 0.24 × 0.12 mas at a position angle of <math display="inline"><semantics> <mrow> <mo>−</mo> <msup> <mn>17.7</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> is shown at the top-left corner.</p>
Full article ">Figure 3
<p>The left panels shows the images restored with circular beams as contours. The restored beams are shown at the top-left corner of each panel. Locations of sample slices are presented as lines superimposed on the contours. The ridgeline is presented as red curves superimposed on the contour plots. The right panels present sample slices, showing the intensity as functions of position angles PAs referring to the jet axis. The northern limb is at positive values, while its southern counterpart is at negative values. The blue horizontal lines represent 5<math display="inline"><semantics> <mi>σ</mi> </semantics></math> of the restored images.</p>
Full article ">Figure 4
<p>(<b>a</b>) Position angles of the northern limb in M87 as functions of core distance <span class="html-italic">d</span>. (<b>b</b>) Position angles of the southern limb in M87 as functions of core distance <span class="html-italic">d</span>. The colors represent the measured frequencies, magenta for 86 GHz, blue for 43 GHz and green for 22 GHz. The uncertainty of <math display="inline"><semantics> <mi>PA</mi> </semantics></math> is estimated as <math display="inline"><semantics> <mfrac> <msup> <mn>20</mn> <mo>∘</mo> </msup> <mi>SNR</mi> </mfrac> </semantics></math>, in which <math display="inline"><semantics> <msup> <mn>20</mn> <mo>∘</mo> </msup> </semantics></math> is a typical width of the jet limb, and SNR is the ratio between the intensity peak and the <math display="inline"><semantics> <mi>σ</mi> </semantics></math> of images. only peaks with intensity above 5<math display="inline"><semantics> <mi>σ</mi> </semantics></math> are presented to make sure the result is reliable.</p>
Full article ">Figure 5
<p>(<b>a</b>) The apparent opening angle <math display="inline"><semantics> <msub> <mi>ψ</mi> <mrow> <mi>a</mi> <mi>p</mi> <mi>p</mi> </mrow> </msub> </semantics></math> as functions of core distance <span class="html-italic">d</span>. Parabolic collimation profile determined by Asada &amp; Nakamura is shown for reference. (<b>b</b>) The jet center offset <math display="inline"><semantics> <msub> <mi>δ</mi> <mrow> <mi>a</mi> <mi>p</mi> <mi>p</mi> </mrow> </msub> </semantics></math> as functions of core distance <span class="html-italic">d</span>. The jet axis is are shown for reference. The colors represent the measured frequencies: magenta for 86 GHz, blue for 43 GHz and green for 22 GHz. The uncertainties of <math display="inline"><semantics> <msub> <mi>ψ</mi> <mrow> <mi>a</mi> <mi>p</mi> <mi>p</mi> </mrow> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>δ</mi> <mrow> <mi>a</mi> <mi>p</mi> <mi>p</mi> </mrow> </msub> </semantics></math> are estimated as <math display="inline"><semantics> <msqrt> <mrow> <mo>Δ</mo> <msubsup> <mi>PA</mi> <mi>North</mi> <mn>2</mn> </msubsup> <mo>+</mo> <mo>Δ</mo> <msubsup> <mi>PA</mi> <mi>South</mi> <mn>2</mn> </msubsup> </mrow> </msqrt> </semantics></math>. Only peaks with intensity above 5<math display="inline"><semantics> <mi>σ</mi> </semantics></math> are presented to make sure the result is reliable.</p>
Full article ">Figure 6
<p>Panel (<b>a</b>) shows 2D cross-correlation of the optically thin regions of 22 and 43 GHz images. Panel (<b>b</b>) 2D cross-correlation of the optically thin regions of 43 and 86 GHz images.</p>
Full article ">Figure 7
<p>The pseudo-color maps of spectral-index distribution with color bars superimposed on the total intensity contour plots. Panel (<b>a</b>) shows 22–43 GHz spectral-index map superimposed on the contour plot of 22 GHz. Panel (<b>c</b>) shows the 43–86 GHz spectral-index map superimposed on the contour plot of 43 GHz. Their noise maps showing distribution of uncertainty are also presented in panel (<b>b</b>) and (<b>d</b>) respectively.</p>
Full article ">
19 pages, 5355 KiB  
Review
Blazar Optical Polarimetry: Current Progress in Observations and Theories
by Haocheng Zhang
Galaxies 2019, 7(4), 85; https://doi.org/10.3390/galaxies7040085 - 27 Oct 2019
Cited by 21 | Viewed by 4139
Abstract
Polarimetry has been a standard tool to probe the active galactic nucleus (AGN) jet magnetic field. In recent years, several optical polarization monitoring programs have been carried out, bringing in many exciting new results and insights into jet dynamics and emission. This article [...] Read more.
Polarimetry has been a standard tool to probe the active galactic nucleus (AGN) jet magnetic field. In recent years, several optical polarization monitoring programs have been carried out, bringing in many exciting new results and insights into jet dynamics and emission. This article discusses current progress in blazar optical polarimetry. The main focus is the variability of polarization signatures, which has spurred a lot of theoretical studies. These novel developments have provided unique constraints on the blazar flares and emphasized the role of the magnetic field in jet evolution. Optical polarimetry will continue to act as an essential component in the multi-messenger study of AGN jets, in particular with the upcoming high-energy polarimetry. Comparing to first-principle numerical simulations, future multi-wavelength polarimetry can shed light on jet dynamics, particle acceleration, and radiation processes. Full article
(This article belongs to the Special Issue Jet Physics of Accreting Super Massive Black Holes)
Show Figures

Figure 1

Figure 1
<p>Two examples of long-term optical and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray monitoring with optical polarization signatures. (<b>Left</b>): Observations of the blazar S5 0716+71 from 2008–2014. The top panel shows the gamma-ray light curve, the second panel shows variability of the gamma-ray spectral index, the third panel shows the optical V-band light curve, the fourth panel shows the ratio of gamma-ray to optical flux, the fifth panel shows variability of the polarization degree, and the bottom panel shows variability of the polarization angle. The figure is reproduced from Figure 1 in [<a href="#B58-galaxies-07-00085" class="html-bibr">58</a>] with the permission of the American Astronomical Society (AAS). (<b>Right</b>): Observations of the blazar 3C 279 from 2008–2014. The temporal behaviors of the polarimetric observations (P, PA, polarized flux) compared with the gamma-rays and the UV-continuum for 3C 279. The figure is reproduced from Figure 4 in [<a href="#B64-galaxies-07-00085" class="html-bibr">64</a>].</p>
Full article ">Figure 2
<p>Statistical studies of the RoboPol data. (<b>Upper left</b>): gamma-ray-loud (GL) blazars show a systematically higher polarization degree than the gamma-ray-quiet (GQ) ones. The figure is reproduced from Figure 1 in [<a href="#B74-galaxies-07-00085" class="html-bibr">74</a>]. (<b>Upper right</b>): observed polarization degree distribution (black line) versus random walk simulations (blue lines). Figure reproduced from Figure 15 in [<a href="#B78-galaxies-07-00085" class="html-bibr">78</a>]. (<b>Lower left</b>): normalized gamma-ray flare level versus time lag between gamma-ray flare and polarization angle rotation. Figure reproduced from Figure 5 in [<a href="#B73-galaxies-07-00085" class="html-bibr">73</a>]. (<b>Lower right</b>): the distribution of the ratio of polarization degree during angle rotation over that outside the angle rotation. Figure reproduced from Figure 9 in [<a href="#B80-galaxies-07-00085" class="html-bibr">80</a>].</p>
Full article ">Figure 3
<p>Observations of BL Lac showing multi-wavelength flares with an optical polarization angle swing. (<b>a</b>–<b>d</b>), dependence on time of the flux of radiation from BL Lac over a two-year interval at the indicated wavebands. (<b>e</b>–<b>h</b>), enlargements of the 0.25 yr time interval marked by vertical dotted lines in panels (<b>a</b>–<b>d</b>) but with optical R-band polarization angle (<b>g</b>) and degree of polarization P (<b>h</b>) respectively replacing the X-ray spectral index (<b>b</b>) and radio flux density (<b>d</b>) (whereas e and f, respectively, show the magnified intervals in (<b>a</b>,<b>c</b>)). The figure is reproduced from Figure 2 in [<a href="#B44-galaxies-07-00085" class="html-bibr">44</a>] with permission from Nature.</p>
Full article ">Figure 4
<p>Examples of optical polarization angle rotations. All four observations, from top to bottom, show the optical flux, polarization degree, and polarization angle. Upper left: PKS 1510+089. This figure is reproduced from Figure 4 in [<a href="#B57-galaxies-07-00085" class="html-bibr">57</a>], with permission from the American Astronomical Society (AAS). Upper right: B2 2308+34. This figure is reproduced from Figure 2 in [<a href="#B46-galaxies-07-00085" class="html-bibr">46</a>]. Lower left: S5 0716+71. This figure is reproduced from Figure 1 in [<a href="#B72-galaxies-07-00085" class="html-bibr">72</a>], with permission from the American Astronomical Society (AAS). Lower right: S4 0954+658. This figure is reproduced from Figure 2 in [<a href="#B53-galaxies-07-00085" class="html-bibr">53</a>], with permission from the American Astronomical Society (AAS).</p>
Full article ">Figure 5
<p>A sketch of the spiral motion model. The figure is reproduced from Figure 3 in [<a href="#B44-galaxies-07-00085" class="html-bibr">44</a>] with permission from Nature.</p>
Full article ">Figure 6
<p>Simulated linear polarization behavior over 5000 time-steps generated by the turbulent extreme multi-zone (TEMZ) model for the case of a 100% turbulent magnetic field, no shock, and acceleration of electrons in each of 18,816 cells. Upper left: polarization vs. time; upper right: Stokes Q vs. U scatter plot; middle: color-dependence (B-band minus R-band) of the optical polarization (dashed lines in upper-right and middle plots indicate the mean values); bottom: sample simulated maps of 230 GHz polarized intensity with polarization vectors superposed, convolved with circular Gaussian restoring beams of (<b>left</b>) 2 and (<b>right</b>) 20 microarcseconds FWHM, the former to show the underlying structure, and the latter to approximate the resolution of VLBI at 230 GHz. In the images, the polarization vectors are color-coded according to the 45-degree-wide range in which they fall, for ease of visualization. In this and subsequent figures, <math display="inline"><semantics> <msub> <mi>β</mi> <mrow> <mi>t</mi> <mi>u</mi> <mi>r</mi> <mi>b</mi> </mrow> </msub> </semantics></math> is the randomly directed velocity (in units of the speed of light) of the turbulent cells. This figure is reproduced from Figure 2 in [<a href="#B89-galaxies-07-00085" class="html-bibr">89</a>].</p>
Full article ">Figure 7
<p>(<b>Left</b>): a sketch of the magnetic field geometry and the internal shock model (panel <b>a</b>) and the illustration of light-crossing delay effects (panel <b>b</b>). The different colors in panel <b>a</b> represent the location of the shock at different epochs of the blazar flare. Panel <b>b</b> then shows the apparent location of the shock at equal photon arrival time at the observer. (<b>Right</b>): fittings of spectra, light curves, and polarization signatures of the 3C 279 flaring event [<a href="#B81-galaxies-07-00085" class="html-bibr">81</a>] based on the model. The figure is reproduced from Figures 4 and 8 in [<a href="#B96-galaxies-07-00085" class="html-bibr">96</a>] with the permission of the American Astronomical Society (AAS).</p>
Full article ">Figure 8
<p>(<b>Left</b>): Snapshots of the magnetic field strength (upper row), particle number density (middle row), and the polarized emission map (lower row) of the simulation region. In the lower row, the color indicates the total flux at each zone, while the segments represent the relative polarized flux. (<b>Right</b>): From top to bottom are the relative flux, polarization degree, and angle in the optical band. The figure is reproduced from Figures 2 and 5 in [<a href="#B106-galaxies-07-00085" class="html-bibr">106</a>] with the permission of the American Astronomical Society (AAS).</p>
Full article ">Figure 9
<p>Predictions of the spectral and temporal high-energy polarization signatures. (<b>Left</b>): The X-ray to <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray spectral polarization degree based on the spectral fitting of the TXS 0506+056 observation. The optical polarization degree is taken as 10%. The spectral models 1–4 are the pure leptonic model, leptonic model with a hadronic cascade component, proton synchrotron model with a considerable hadronic cascade component, and proton-synchrotron-dominated model. The shaded regions correspond to the X-ray and MeV <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray bands. The figure is reproduced from Figure 2 in [<a href="#B107-galaxies-07-00085" class="html-bibr">107</a>] with the permission of the American Astronomical Society (AAS). (<b>Right</b>): The spectra, light curves, and polarization signatures of the proton-synchrotron-dominated model. The A–D epochs in the upper panels correspond to the spectra and spectral polarization degree before the <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-ray flare, during the rising phase, at the flare peak, and after the flare. This figure is reproduced from Figure 4 in [<a href="#B109-galaxies-07-00085" class="html-bibr">109</a>] with permission from the American Astronomical Society (AAS).</p>
Full article ">
25 pages, 6326 KiB  
Article
Planets in Binaries: Formation and Dynamical Evolution
by Francesco Marzari and Philippe Thebault
Galaxies 2019, 7(4), 84; https://doi.org/10.3390/galaxies7040084 - 16 Oct 2019
Cited by 33 | Viewed by 4343
Abstract
Binary systems are very common among field stars, yet the vast majority of known exoplanets have been detected around single stars. While this relatively small number of planets in binaries is probably partly due to strong observational biases, there is, however, statistical evidence [...] Read more.
Binary systems are very common among field stars, yet the vast majority of known exoplanets have been detected around single stars. While this relatively small number of planets in binaries is probably partly due to strong observational biases, there is, however, statistical evidence that planets are indeed less frequent in binaries with separations smaller than 100 au, strongly suggesting that the presence of a close-in companion star has an adverse effect on planet formation. It is indeed possible for the gravitational pull of the second star to affect all the different stages of planet formation, from proto-planetary disk formation to dust accumulation into planetesimals, to the accretion of these planetesimals into large planetary embryos and, eventually, the final growth of these embryos into planets. For the crucial planetesimal-accretion phase, the complex coupling between dynamical perturbations from the binary and friction due to gas in the proto-planetary disk suggests that planetesimal accretion might be hampered due to increased, accretion-hostile impact velocities. Likewise, the interplay between the binary’s secular perturbations and mean motion resonances lead to unstable regions, where not only planet formation is inhibited, but where a massive body would be ejected from the system on a hyperbolic orbit. The amplitude of these two main effects is different for S- and P-type planets, so that a comparison between the two populations might outline the influence of the companion star on the planet formation process. Unfortunately, at present the two populations (circumstellar or circumbinary) are not known equally well and different biases and uncertainties prevent a quantitative comparison. We also highlight the long-term dynamical evolution of both S and P-type systems and focus on how these different evolutions influence the final architecture of planetary systems in binaries. Full article
(This article belongs to the Special Issue Habitability of Planets in Stellar Binary Systems)
Show Figures

Figure 1

Figure 1
<p>All circumstellar (S-type orbits, as opposed to circumbinary P-type orbits) planet-hosting binaries with separation ≤500 au (as of July 2019). Companion stars are displayed as yellow circles, whose radius is proportional to <math display="inline"><semantics> <msup> <mrow> <mo>(</mo> <msub> <mi>M</mi> <mn>2</mn> </msub> <mo>/</mo> <msub> <mi>M</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>3</mn> </mrow> </msup> </semantics></math>. Planets are marked as blue circles whose radius is proportional to <math display="inline"><semantics> <msup> <mrow> <mo>(</mo> <msub> <mi>m</mi> <mrow> <mi>p</mi> <mi>l</mi> </mrow> </msub> <mo>)</mo> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>3</mn> </mrow> </msup> </semantics></math>. The horizontal lines represent the radial excursion of the planet and star orbits (when they are known). When the binary orbit is known, the displayed distance is the semi-major axis, otherwise it is the projected current separation. The short vertical lines correspond to the outer limit of the orbital stability region around the primary estimated by Holman and Wiegert [<a href="#B7-galaxies-07-00084" class="html-bibr">7</a>] assuming prograde orbits and coplanarity (for the seemingly “unstable” planet HD59686Ab, a retrograde orbit was suggested as a possible stable solution [<a href="#B8-galaxies-07-00084" class="html-bibr">8</a>]. Planets detected by the radial velocity method are written in black, planets detected by transit are in blue, and planets detected by other methods are in red.</p>
Full article ">Figure 2
<p>Red histogram: Marginalized distribution <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </semantics></math> of projected separations for all stellar companions to S-type exoplanet-host stars found by [<a href="#B18-galaxies-07-00084" class="html-bibr">18</a>] for a large sample of Kepler Objects of Interest (KOIs). The blue line represents the predicted distribution if binaries were drawn from the distribution reported by [<a href="#B2-galaxies-07-00084" class="html-bibr">2</a>]. Taken from [<a href="#B18-galaxies-07-00084" class="html-bibr">18</a>], courtesy of the Astrophysical Journal.</p>
Full article ">Figure 3
<p>Distribution of the semi-major axis of all S-type exoplanet-hosting binaries with a projected separation up to 500 au. When the semi-major axis <math display="inline"><semantics> <msub> <mi>a</mi> <mi>b</mi> </msub> </semantics></math> is not known and only the projected separation <math display="inline"><semantics> <mi>ρ</mi> </semantics></math> is available, <math display="inline"><semantics> <mrow> <msub> <mrow/> <mi>b</mi> </msub> <mi>a</mi> </mrow> </semantics></math> is estimated through the statistical relation <math display="inline"><semantics> <mrow> <mi>l</mi> <mi>o</mi> <mi>g</mi> <mrow> <mo>(</mo> <msub> <mi>a</mi> <mi>b</mi> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mi>l</mi> <mi>o</mi> <mi>g</mi> <mrow> <mo>(</mo> <mi>ρ</mi> <mo>)</mo> </mrow> <mo>+</mo> <mn>0.13</mn> </mrow> </semantics></math> [<a href="#B1-galaxies-07-00084" class="html-bibr">1</a>,<a href="#B2-galaxies-07-00084" class="html-bibr">2</a>]. The solid-line curve represents the normalized <math display="inline"><semantics> <msub> <mi>a</mi> <mi>b</mi> </msub> </semantics></math> distribution for field stars derived by [<a href="#B2-galaxies-07-00084" class="html-bibr">2</a>].</p>
Full article ">Figure 4
<p>Frequency of disk–bearing tight binaries (with projected separation <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>≤</mo> <mn>40</mn> <mspace width="0.166667em"/> </mrow> </semantics></math> au) as a function of age in several young stellar associations (from [<a href="#B23-galaxies-07-00084" class="html-bibr">23</a>], courtesy of the Astrophysical Journal).</p>
Full article ">Figure 5
<p>The accretion behavior in a planetesimal disk around <math display="inline"><semantics> <mi>α</mi> </semantics></math> Cen B, estimated with numerical simulations taking into account gas drag. The relative importance of different types of collisional outcomes is displayed as a function of radial distance. <span class="html-italic">Red</span>: impacts for which the impact velocity <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>v</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>≥</mo> <msub> <mi>v</mi> <mrow> <mi>e</mi> <mi>r</mi> <mi>o</mi> <mo>−</mo> <mi>M</mi> </mrow> </msub> </mrow> </semantics></math>, where <math display="inline"><semantics> <msub> <mi>v</mi> <mrow> <mi>e</mi> <mi>r</mi> <mi>o</mi> <mo>−</mo> <mi>M</mi> </mrow> </msub> </semantics></math> is the threshold velocity beyond which an impact between two objects of sizes <math display="inline"><semantics> <msub> <mi>s</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>s</mi> <mn>2</mn> </msub> </semantics></math> always results in net mass loss. <span class="html-italic">Yellow</span>: Uncertain outcome. The erosion vs. accretion net balance depends on the physical composition of the planetesimals. <span class="html-italic">Green</span>: <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>v</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>≤</mo> <msub> <mi>v</mi> <mrow> <mi>e</mi> <mi>s</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>, where <math display="inline"><semantics> <msub> <mi>v</mi> <mrow> <mi>e</mi> <mi>s</mi> <mi>c</mi> </mrow> </msub> </semantics></math> is the escape velocity of the (<math display="inline"><semantics> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </semantics></math>) pair. Accretion can here proceed unimpeded, in a “runaway growth” way, as around a single star. <span class="html-italic">Light blue</span>: <math display="inline"><semantics> <mrow> <msub> <mi>v</mi> <mrow> <mi>e</mi> <mi>s</mi> <mi>c</mi> </mrow> </msub> <mo>≤</mo> <mo>Δ</mo> <msub> <mi>v</mi> <mrow> <msub> <mi>s</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>2</mn> </msub> </mrow> </msub> <mo>≤</mo> <msub> <mi>v</mi> <mrow> <mi>e</mi> <mi>r</mi> <mi>o</mi> <mo>−</mo> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math>. Collisions result in net accretion, but <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>v</mi> </mrow> </semantics></math> are high enough to switch off the fast-runaway growth mode. The two thick blue lines denote the location of the inner limit of the “optimistic” and “conservative” habitable zone around the star. The planetesimal size distribution is assumed to be a Maxwellian centered on 5 km (modified from [<a href="#B36-galaxies-07-00084" class="html-bibr">36</a>]).</p>
Full article ">Figure 6
<p>Numerical simulation of the mutual accretion within an initial disk of lunar-sized embryos around <math display="inline"><semantics> <mi>α</mi> </semantics></math> Centauri B. At the end of the simulation, four terrestrial planets were formed. (Taken from [<a href="#B47-galaxies-07-00084" class="html-bibr">47</a>], courtesy of the Astrophysical Journal)</p>
Full article ">Figure 7
<p>Accretion vs. Erosion behavior of a population of kilometer-sized planetesimals in the habitable zone around <math display="inline"><semantics> <mi>α</mi> </semantics></math> Centauri B when varying the binary’s separation and eccentricity. The white circle marks the current orbit of the binary. This shows that planetesimal accretion is possible if the binary has suffered a stellar encounter that shrank its initial wider orbit to its current value (taken from [<a href="#B36-galaxies-07-00084" class="html-bibr">36</a>], courtesy of the MNRAS).</p>
Full article ">Figure 8
<p>Figure from [<a href="#B74-galaxies-07-00084" class="html-bibr">74</a>] showing at different evolutionary times the relative impact velocities for a putative planetesimal population in Kepler-16. In the upper panel the impact velocities between equal size R = 5 km planetesimals, computed at different radial distances from the baricenter of the binary, are compared with the critical erosion velocity (black dashed line). The continuous black line shows the average impact velocity for the population. In the lower panel the impact speeds are computed for larger planetesimals R = 25 km in size.</p>
Full article ">Figure 9
<p>Figure taken from [<a href="#B88-galaxies-07-00084" class="html-bibr">88</a>]. The values of <math display="inline"><semantics> <msub> <mi>e</mi> <mi>F</mi> </msub> </semantics></math> (top panel) and <math display="inline"><semantics> <msub> <mi>g</mi> <mi>s</mi> </msub> </semantics></math> (bottom panel) are computed for the planet in HD 196885 (the vertical dashed line shows its location). The red dots are reference outcomes of numerical simulations [<a href="#B7-galaxies-07-00084" class="html-bibr">7</a>], the magenta curve is derived from [<a href="#B87-galaxies-07-00084" class="html-bibr">87</a>], the blue curve from the first order model of [<a href="#B88-galaxies-07-00084" class="html-bibr">88</a>] while the green curve is the second order model of [<a href="#B88-galaxies-07-00084" class="html-bibr">88</a>]. The black curve shows the amplitude of the short period variations of the eccentricity.</p>
Full article ">Figure 10
<p>Figure taken from [<a href="#B84-galaxies-07-00084" class="html-bibr">84</a>]. The FMA analysis is performed on a Jupiter-size single planet orbiting in a binary system with <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>25</mn> </mrow> </semantics></math> AU and <math display="inline"><semantics> <mrow> <msub> <mi>e</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (top panel) and <math display="inline"><semantics> <mrow> <msub> <mi>e</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math> (bottom panel). The diffusion index <math display="inline"><semantics> <msub> <mi>c</mi> <mi>s</mi> </msub> </semantics></math> of the main secular frequency of the planet is drawn vs. its semi-major axis. Small values of <math display="inline"><semantics> <msub> <mi>c</mi> <mi>s</mi> </msub> </semantics></math> means low diffusion, while large values connote chaotic orbits. The black dash-dotted line marks the critical semi-major axis computed from the empirical formula of [<a href="#B7-galaxies-07-00084" class="html-bibr">7</a>]. The green dashed lines show the location of mean motion resonances between the planet and the companion star. The blue squares show stable cases whose orbits were numerically integrated over 4 Gyr, the yellow circles unstable ones.</p>
Full article ">
36 pages, 1317 KiB  
Review
A Census of B[e] Supergiants
by Michaela Kraus
Galaxies 2019, 7(4), 83; https://doi.org/10.3390/galaxies7040083 - 29 Sep 2019
Cited by 45 | Viewed by 4816
Abstract
Stellar evolution theory is most uncertain for massive stars. For reliable predictions of the evolution of massive stars and their final fate, solid constraints on the physical parameters, and their changes along the evolution and in different environments, are required. Massive stars evolve [...] Read more.
Stellar evolution theory is most uncertain for massive stars. For reliable predictions of the evolution of massive stars and their final fate, solid constraints on the physical parameters, and their changes along the evolution and in different environments, are required. Massive stars evolve through a variety of short transition phases, in which they can experience large mass-loss either in the form of dense winds or via sudden eruptions. The B[e] supergiants comprise one such group of massive transition objects. They are characterized by dense, dusty disks of yet unknown origin. In the Milky Way, identification and classification of B[e] supergiants is usually hampered by their uncertain distances, hence luminosities, and by the confusion of low-luminosity candidates with massive pre-main sequence objects. The extragalactic objects are often mistaken as quiescent or candidate luminous blue variables, with whom B[e] supergiants share a number of spectroscopic characteristics. In this review, proper criteria are provided, based on which B[e] supergiants can be unambiguously classified and separated from other high luminosity post-main sequence stars and pre-main sequence stars. Using these criteria, the B[e] supergiant samples in diverse galaxies are critically inspected, to achieve a reliable census of the current population. Full article
(This article belongs to the Special Issue Luminous Stars in Nearby Galaxies)
Show Figures

Figure 1

Figure 1
<p>HR diagram showing the positions of the classical MC B[e]SG sample [<a href="#B18-galaxies-07-00083" class="html-bibr">18</a>]. The stellar evolutionary tracks at SMC metallicity for stars rotating initially with 40% of their critical velocity are also included (from [<a href="#B19-galaxies-07-00083" class="html-bibr">19</a>]). The dotted square contains objects that display CO band emission (except for S 89, see <a href="#sec2dot3-galaxies-07-00083" class="html-sec">Section 2.3</a>). For brevity and readability, the identifiers LHA 120 and LHA 115 for objects within the LMC and SMC, respectively, have been omitted.</p>
Full article ">Figure 2
<p>Sketch of the generation of the typical CO band head profile. (<b>a</b>) Spectrum around the (2-0) band head of the CO first-overtone bands for a hot gas with velocity dispersion of a few km s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. (<b>b</b>) Profile of a single line from a rotating gas ring with a velocity, projected to the line of sight, of 66 km s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> as seen with a spectral resolution of 6 km s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. (<b>c</b>) Total synthetic CO band head spectrum resulting from the convolution of the band transitions in (<b>a</b>) with the profile of the ring in (<b>b</b>). (<b>d</b>) CO band head observations of the Galactic B[e]SG CPD-57 2874 [<a href="#B58-galaxies-07-00083" class="html-bibr">58</a>].</p>
Full article ">Figure 3
<p>Synthetic spectra of the combined emission from <math display="inline"><semantics> <msup> <mrow/> <mn>12</mn> </msup> </semantics></math>CO and <math display="inline"><semantics> <msup> <mrow/> <mn>13</mn> </msup> </semantics></math>CO for different values of the <math display="inline"><semantics> <msup> <mrow/> <mn>12</mn> </msup> </semantics></math>C/<math display="inline"><semantics> <msup> <mrow/> <mn>13</mn> </msup> </semantics></math>C ratio. The computations have been performed for the following physical parameters: a <math display="inline"><semantics> <msup> <mrow/> <mn>12</mn> </msup> </semantics></math>CO column density of <math display="inline"><semantics> <mrow> <mn>2</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>21</mn> </msup> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </msup> </semantics></math>, a gas temperature of 3000 K, a line-of-sight rotational velocity of 66 km s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, and a spectral resolution of 50 km s<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>.</p>
Full article ">Figure 4
<p>Evolution of the <math display="inline"><semantics> <msup> <mrow/> <mn>12</mn> </msup> </semantics></math>C/<math display="inline"><semantics> <msup> <mrow/> <mn>13</mn> </msup> </semantics></math>C isotope ratio along the solar metallicity tracks of a star with initial mass of 32 M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math> and initial rotation speeds <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>/</mo> <msub> <mi>v</mi> <mi>crit</mi> </msub> </mrow> </semantics></math> ranging from 0 to 0.4 (corresponding to <math display="inline"><semantics> <mrow> <mo>Ω</mo> <mo>/</mo> <msub> <mo>Ω</mo> <mi>crit</mi> </msub> <mo>=</mo> <mn>0.0</mn> <mo>;</mo> <mn>0.1</mn> <mo>;</mo> <mn>0.2</mn> <mo>;</mo> <mn>0.3</mn> <mo>;</mo> <mn>0.4</mn> <mo>;</mo> <mn>0.5</mn> <mo>;</mo> <mn>0.568</mn> </mrow> </semantics></math>). The individual tracks have been obtained from the interpolation tool SCYCLIST provided by the Geneva group. For clarity of the plot, we truncated the evolutionary tracks within the red supergiant regions. Included are the positions of the MC B[e]SG sample from <a href="#galaxies-07-00083-t002" class="html-table">Table 2</a> with known <math display="inline"><semantics> <msup> <mrow/> <mn>12</mn> </msup> </semantics></math>C/<math display="inline"><semantics> <msup> <mrow/> <mn>13</mn> </msup> </semantics></math>C ratio, following the same color coding as for the tracks. The Galactic objects are excluded due to their highly uncertain luminosities. Depending on the initial rotation speed of the star, the observed ratio can be reached either in the pre-RSG (moderate rotator) or post-RSG (slow rotator) phase.</p>
Full article ">Figure 5
<p>Demonstration of the separation of the B[e]SGs from the quiescent LBVs within the near-IR (J–H versus H–K diagram (<b>left</b>)) and the WISE diagram (W1–W2 versus W2–W4 (<b>right</b>)). Shown are the positions of the classical MC B[e]SG sample and of the MC LBV sample. IR colors of the objects are provided in <a href="#galaxies-07-00083-t003" class="html-table">Table 3</a>. The solid line represents the positions of regular supergiants with empirical colors taken from [<a href="#B106-galaxies-07-00083" class="html-bibr">106</a>] for solar metallicity stars.</p>
Full article ">Figure 6
<p>Location of the new LMC (light blue) and SMC (purple) samples with respect to those MC B[e]SGs that meet all the required classification criteria (dark blue) and LBVs (red triangles) in the near-IR (<b>top</b>) and the WISE diagram (<b>bottom</b>). Filled circles are used for confirmed B[e]SGs, filled stars for candidates, and empty circles for misclassified objects. A sample of late-type stars and supergiants in the MCs (small gray crosses, from [<a href="#B115-galaxies-07-00083" class="html-bibr">115</a>]) and a sample of (dereddened, see text) Galactic HAeBe stars (black plus signs, from [<a href="#B116-galaxies-07-00083" class="html-bibr">116</a>]) are also included. The arrow in each panel indicates the direction of the reddening, and their length complies with a value of <math display="inline"><semantics> <mrow> <msub> <mi>A</mi> <mi mathvariant="normal">V</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>. As the color excess is not known for all MC objects, no extinction correction has been applied. However, the MC stars have in general relatively small color excess values (see <a href="#galaxies-07-00083-t004" class="html-table">Table 4</a> and <a href="#galaxies-07-00083-t005" class="html-table">Table 5</a>), which would shift them only marginally in the diagrams.</p>
Full article ">Figure 7
<p>As <a href="#galaxies-07-00083-f006" class="html-fig">Figure 6</a> but for the location of the M31 (light blue) and M33 (purple) samples in the near-IR (<b>top</b>) and the WISE diagram (<b>bottom</b>) based on their observed colors. Typical values for the objects’ reddening are <math display="inline"><semantics> <mrow> <msub> <mi>A</mi> <mi mathvariant="normal">V</mi> </msub> <mo>≤</mo> <mn>1.5</mn> </mrow> </semantics></math> mag [<a href="#B142-galaxies-07-00083" class="html-bibr">142</a>].</p>
Full article ">Figure 8
<p>Near-IR diagrams as in <a href="#galaxies-07-00083-f006" class="html-fig">Figure 6</a>, showing the locations of the Galactic confirmed B[e]SGs (<b>top</b>) and B[e]SG candidates (<b>bottom</b>). The colors of the Galactic objects have been corrected for interstellar extinction (<a href="#galaxies-07-00083-t009" class="html-table">Table 9</a>).</p>
Full article ">
34 pages, 1850 KiB  
Review
Biosignatures Search in Habitable Planets
by Riccardo Claudi and Eleonora Alei
Galaxies 2019, 7(4), 82; https://doi.org/10.3390/galaxies7040082 - 29 Sep 2019
Cited by 7 | Viewed by 7005
Abstract
The search for life has had a new enthusiastic restart in the last two decades thanks to the large number of new worlds discovered. The about 4100 exoplanets found so far, show a large diversity of planets, from hot giants to rocky planets [...] Read more.
The search for life has had a new enthusiastic restart in the last two decades thanks to the large number of new worlds discovered. The about 4100 exoplanets found so far, show a large diversity of planets, from hot giants to rocky planets orbiting small and cold stars. Most of them are very different from those of the Solar System and one of the striking case is that of the super-Earths, rocky planets with masses ranging between 1 and 10 M with dimensions up to twice those of Earth. In the right environment, these planets could be the cradle of alien life that could modify the chemical composition of their atmospheres. So, the search for life signatures requires as the first step the knowledge of planet atmospheres, the main objective of future exoplanetary space explorations. Indeed, the quest for the determination of the chemical composition of those planetary atmospheres rises also more general interest than that given by the mere directory of the atmospheric compounds. It opens out to the more general speculation on what such detection might tell us about the presence of life on those planets. As, for now, we have only one example of life in the universe, we are bound to study terrestrial organisms to assess possibilities of life on other planets and guide our search for possible extinct or extant life on other planetary bodies. In this review, we try to answer the three questions that also in this special search, mark the beginning of every research: what? where? how? Full article
(This article belongs to the Special Issue Habitability of Planets in Stellar Binary Systems)
Show Figures

Figure 1

Figure 1
<p>The original Visible, NIR and IR Earth spectrum taken by NIMS instrument during the Galileo fly–by on Earth [<a href="#B34-galaxies-07-00082" class="html-bibr">34</a>].</p>
Full article ">Figure 2
<p>Observed Vis-NIR spectrum of the Earthshine obtained by [<a href="#B38-galaxies-07-00082" class="html-bibr">38</a>]. The reflectivity of vegetation is dominated by a sharp rise in reflectivity for wavelengths longer than 0.70 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m, plus some smaller bumps at shorter and longer wavelengths [<a href="#B36-galaxies-07-00082" class="html-bibr">36</a>,<a href="#B38-galaxies-07-00082" class="html-bibr">38</a>]). In the spectrum, the rise of the flux due to plants reflection is only about 6% of the nearby continuum. This is because at the time of observation only about 17% of the projected area was land (for details on the observing method see [<a href="#B36-galaxies-07-00082" class="html-bibr">36</a>]. In the spectrum are also indicated most of the molecules that are addressed in the text.</p>
Full article ">Figure 3
<p>Cross section for photodissociation of H<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math>O, CO<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math>, O<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math>. Taken by Selsis et al. [<a href="#B81-galaxies-07-00082" class="html-bibr">81</a>].</p>
Full article ">Figure 4
<p>The planet-star contrast as a function of the planetary system separation. In the plot are shown the performances of current high contrast imagers together with future ground and space-based instrumentation (see <a href="#sec8-galaxies-07-00082" class="html-sec">Section 8</a>). This plot, updated to 2018, is taken by the NASA Exoplanet Exploration Program Office page (<a href="https://exoplanets.nasa.gov/exep/technology/technology-overview/" target="_blank">https://exoplanets.nasa.gov/exep/technology/technology-overview/</a>). The orange points are the young self-luminous planets imaged by ground-based telescope in the NIR. The black points are estimates of contrast for planets detected with the radial velocity method. Solid lines are the performances of ground-based coronagraph (orange lines) and space-based (black lines). For the space-based coronagraph, both HST/ACS (in black) and JWST/NIR Cam (in orange) are reported. SPHERE contrast curve represents the best contrast achieved on Sirius, while the GPI labeled line represents its typical performance. The solid black line shows the predicted contrast curve for WFIRST-CGI (see <a href="#sec8-galaxies-07-00082" class="html-sec">Section 8</a>). The names of SS planets show the contrast star-planet of the analogous planets placed at 10 pc of distance.</p>
Full article ">Figure 5
<p>The effective wavelength coverage of each instrument indicates what species of features can be potentially detected by what instrument. The colored filled circles identify the central wavelength of the absorption band of the moleculas named on the vertical axis. On the top of the plot, the wavelength range of the considered instruments is indicated by the corresponding horizontal colored solid line. OST and MIRI@JWST the final wavelength of their used spectral range is also indicated. The use of colors has the unique aim to make the plot clearer.</p>
Full article ">
54 pages, 9674 KiB  
Review
Dark Matter Haloes and Subhaloes
by Jesús Zavala and Carlos S. Frenk
Galaxies 2019, 7(4), 81; https://doi.org/10.3390/galaxies7040081 - 25 Sep 2019
Cited by 102 | Viewed by 10540
Abstract
The development of methods and algorithms to solve the N-body problem for classical, collisionless, non-relativistic particles has made it possible to follow the growth and evolution of cosmic dark matter structures over most of the universe’s history. In the best-studied case—the cold [...] Read more.
The development of methods and algorithms to solve the N-body problem for classical, collisionless, non-relativistic particles has made it possible to follow the growth and evolution of cosmic dark matter structures over most of the universe’s history. In the best-studied case—the cold dark matter or CDM model—the dark matter is assumed to consist of elementary particles that had negligible thermal velocities at early times. Progress over the past three decades has led to a nearly complete description of the assembly, structure, and spatial distribution of dark matter haloes, and their substructure in this model, over almost the entire mass range of astronomical objects. On scales of galaxies and above, predictions from this standard CDM model have been shown to provide a remarkably good match to a wide variety of astronomical data over a large range of epochs, from the temperature structure of the cosmic background radiation to the large-scale distribution of galaxies. The frontier in this field has shifted to the relatively unexplored subgalactic scales, the domain of the central regions of massive haloes, and that of low-mass haloes and subhaloes, where potentially fundamental questions remain. Answering them may require: (i) the effect of known but uncertain baryonic processes (involving gas and stars), and/or (ii) alternative models with new dark matter physics. Here we present a review of the field, focusing on our current understanding of dark matter structure from N-body simulations and on the challenges ahead. 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>Dimensionless linear dark matter power spectrum in different dark matter models. In the current paradigm, cold dark matter (CDM), the power spectrum keeps on rising to well below subgalactic scales. Alternative models such as warm dark matter (WDM) or interacting dark matter (DAOs) have a cutoff at or slightly below galactic scales, which determines the abundance and structure of small-mass dark matter haloes and subhaloes and the galaxies within. In the black hashed area, the dark matter is constrained by the observed large-scale distribution of galaxies (e.g., [<a href="#B30-galaxies-07-00081" class="html-bibr">30</a>,<a href="#B31-galaxies-07-00081" class="html-bibr">31</a>]) and the Ly-<math display="inline"><semantics> <mi>α</mi> </semantics></math> forest constraints on WDM [<a href="#B28-galaxies-07-00081" class="html-bibr">28</a>] to behave as CDM. Figure adapted from [<a href="#B32-galaxies-07-00081" class="html-bibr">32</a>].</p>
Full article ">Figure 2
<p>Illustration of the initial conditions for an <span class="html-italic">N</span>-body simulation. <span class="html-italic">Left:</span> the dimensionless linear CDM power spectrum. The vertical dashed lines mark the modes corresponding to the maximum and minimum scales that can be represented in the initial conditions: the fundamental mode, <math display="inline"><semantics> <mrow> <mn>2</mn> <mi>π</mi> <mo>/</mo> <mi>L</mi> </mrow> </semantics></math>, and the Nyquist mode, <math display="inline"><semantics> <mrow> <mi>π</mi> <mo>/</mo> <mi>d</mi> </mrow> </semantics></math>, where <span class="html-italic">L</span> and <span class="html-italic">d</span> are the cube length and interparticle separation, respectively. <span class="html-italic">Right</span>: a realization of the dark matter density field generated from the power spectrum on the left at redshift <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>127</mn> </mrow> </semantics></math> using <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msup> <mn>1024</mn> <mn>3</mn> </msup> </mrow> </semantics></math> particles in a cosmological cube of co-moving side, <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>40</mn> </mrow> </semantics></math> Mpc/h. The code MUSIC [<a href="#B65-galaxies-07-00081" class="html-bibr">65</a>] was used to generate the particle distribution and the Pynbody package [<a href="#B72-galaxies-07-00081" class="html-bibr">72</a>] to create the image.</p>
Full article ">Figure 3
<p>Emergence of the cosmic web. <span class="html-italic">Left:</span> evolution of the (projected) dark matter density field in a slab of length <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math> Mpc/h and thickness 15 Mpc/h from the Millennium-II simulation [<a href="#B77-galaxies-07-00081" class="html-bibr">77</a>]. The redshift corresponding to each snapshot is shown on the top right. <span class="html-italic">Right:</span> The dimensionless dark matter power spectrum (solid lines) at the redshifts shown on the left. For comparison, also shown are: the linear power spectrum (thin grey lines) and the non-linear power spectrum for the lower resolution but larger scale (500 Mpc/h) Millennium I simulation (in dotted lines; [<a href="#B4-galaxies-07-00081" class="html-bibr">4</a>]). The dashed lines show the Poisson noise limit for the Millennium I (left) and Millennium-II (right) simulations. Figure adapted from [<a href="#B77-galaxies-07-00081" class="html-bibr">77</a>]<a href="#fn018-galaxies-07-00081" class="html-fn">18</a>.</p>
Full article ">Figure 4
<p>The galaxy distribution in various redshift surveys and in mock catalogues constructed from the Millennium simulation [<a href="#B4-galaxies-07-00081" class="html-bibr">4</a>]. The small slice at the top shows the CfA2 “Great Wall” [<a href="#B81-galaxies-07-00081" class="html-bibr">81</a>], with the Coma cluster at the center. Just above is a section of the Sloan Digital Sky Survey in which the “Sloan Great Wall” [<a href="#B82-galaxies-07-00081" class="html-bibr">82</a>] is visible. The wedge on the left shows one half of the 2-degree-field galaxy redshift survey [<a href="#B83-galaxies-07-00081" class="html-bibr">83</a>]. At the bottom and on the right, mock galaxy surveys constructed using a semi-analytic model applied to the simulation [<a href="#B84-galaxies-07-00081" class="html-bibr">84</a>] are shown, selected to have geometry and magnitude limits matching the corresponding real surveys. Adapted from [<a href="#B85-galaxies-07-00081" class="html-bibr">85</a>].</p>
Full article ">Figure 5
<p>Halo mass function for different dark matter models (adapted from [<a href="#B20-galaxies-07-00081" class="html-bibr">20</a>]). <span class="html-italic">Left:</span> The large-scale dark matter distribution in a slab of a 64 Mpc/h cube in different cosmologies: CDM and WDM in the top left and bottom right, respectively; two interacting dark matter models in the other two panels. <span class="html-italic">Right:</span> The halo mass function at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> for the models on the left. The transparent light blue region marks the resolution limit of the simulations. The cutoff in the primordial linear power spectrum of the non-CDM models results in a lower abundance of low-mass haloes, visible in the panels on the left and quantified in the halo mass function on the right.</p>
Full article ">Figure 6
<p>The structure of CDM haloes. The different panels show several characteristics of the spatial (left) and dynamical (right) structure of a Milky Way-size CDM halo (<math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mn>200</mn> </msub> <mo>∼</mo> <mn>1</mn> <mo>.</mo> <mn>8</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>12</mn> </msup> </mrow> </semantics></math> M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math>; <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>200</mn> </msub> <mo>∼</mo> <mn>250</mn> </mrow> </semantics></math> kpc) from the Aquarius project [<a href="#B138-galaxies-07-00081" class="html-bibr">138</a>]. The top panels show the spherically averaged radial density (left; [<a href="#B138-galaxies-07-00081" class="html-bibr">138</a>]<a href="#fn021-galaxies-07-00081" class="html-fn">21</a>) and velocity dispersion (right; [<a href="#B121-galaxies-07-00081" class="html-bibr">121</a>]<a href="#fn022-galaxies-07-00081" class="html-fn">22</a>) profiles, which are nearly universal for haloes in dynamical equilibrium. The bottom panels show the halo shape (left: moment of inertia axis ratios, and triaxiality: <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mrow> <mo>(</mo> <msup> <mi>a</mi> <mn>2</mn> </msup> <mo>−</mo> <msup> <mi>b</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> <mo>/</mo> <mrow> <mo>(</mo> <msup> <mi>a</mi> <mn>2</mn> </msup> <mo>−</mo> <msup> <mi>c</mi> <mn>2</mn> </msup> <mo>)</mo> </mrow> </mrow> </semantics></math>; [<a href="#B139-galaxies-07-00081" class="html-bibr">139</a>]<a href="#fn023-galaxies-07-00081" class="html-fn">23</a>) and local dark matter velocity distribution near the solar circle: <math display="inline"><semantics> <mrow> <mn>2</mn> <mspace width="3.33333pt"/> <mi>kpc</mi> <mo>&lt;</mo> <mi>r</mi> <mo>&lt;</mo> <mn>9</mn> <mspace width="3.33333pt"/> <mrow/> <mspace width="3.33333pt"/> <mi>kpc</mi> </mrow> </semantics></math> (right; [<a href="#B140-galaxies-07-00081" class="html-bibr">140</a>]<a href="#fn024-galaxies-07-00081" class="html-fn">24</a>).</p>
Full article ">Figure 7
<p>Structure of haloes in models with different types of dark matter: collisional (SIDM; <math display="inline"><semantics> <mrow> <msub> <mi>σ</mi> <mi>T</mi> </msub> <mo>/</mo> <msub> <mi>m</mi> <mi>χ</mi> </msub> <mo>≳</mo> <mn>1</mn> </mrow> </semantics></math> cm<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math>/g) and with a galactic-scale free-streaming cutoff (WDM; <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mi>χ</mi> </msub> <mo>∼</mo> <mn>2</mn> <mo>.</mo> <mn>3</mn> </mrow> </semantics></math> keV). <span class="html-italic">Upper panels:</span> projected dark matter distribution of a Milky Way-size halo in the SIDM model (left panel; [<a href="#B57-galaxies-07-00081" class="html-bibr">57</a>]<a href="#fn026-galaxies-07-00081" class="html-fn">26</a>) and in the WDM model (right panel; [<a href="#B153-galaxies-07-00081" class="html-bibr">153</a>]<a href="#fn027-galaxies-07-00081" class="html-fn">27</a>). <span class="html-italic">Bottom left:</span> spherically averaged density profiles. WDM haloes are well described by an NFW profile, but have lower concentrations than their CDM counterparts of the same mass; SIDM haloes develop flat density cores during a transient stage that inevitably ends with the collapse of the core once the gravothermal catastrophe is triggered. <span class="html-italic">Bottom right:</span> spherically averaged velocity dispersion profiles. WDM haloes still obey the universal scaling for the pseudo-phase-space density, <math display="inline"><semantics> <mrow> <mi>ρ</mi> <mo>/</mo> <msup> <mi>σ</mi> <mn>3</mn> </msup> <mo>∝</mo> <msup> <mi>r</mi> <mrow> <mo>−</mo> <mn>1</mn> <mo>.</mo> <mn>875</mn> </mrow> </msup> </mrow> </semantics></math>, at most radii, except in the very center, which results from a similar velocity dispersion profile to that in CDM but shifted downwards and to the right as a result of the lower concentration. SIDM haloes develop isothermal density cores of size of the order of the scale radius.</p>
Full article ">Figure 8
<p>Dark matter subhaloes. <span class="html-italic">Left:</span> schematic representation of a dark matter halo <span class="html-italic">merger tree</span> (taken from [<a href="#B176-galaxies-07-00081" class="html-bibr">176</a>]<a href="#fn030-galaxies-07-00081" class="html-fn">30</a>) at discrete redshifts. In a hierarchical model, haloes grow by the accretion of smaller neighboring haloes (A,B,C,D), which become subhaloes at the time when they first cross the virial radius of the host halo. The main branch of the tree represents the evolution of the main progenitor (shown in blue). Since this process occurs across the entire hierarchy of structures, there are subhaloes within subhaloes (sub-subhaloes; like a, b, c in system D) and so on. <span class="html-italic">Right:</span> a simulated Milky Way-size CDM halo from the Aquarius project (figure taken from [<a href="#B138-galaxies-07-00081" class="html-bibr">138</a>]<a href="#fn031-galaxies-07-00081" class="html-fn">31</a>; this is the same halo illustrated in <a href="#galaxies-07-00081-f006" class="html-fig">Figure 6</a>). The circles in the main image mark six subhaloes that are shown enlarged in the surrounding panels, as indicated by the labels. Sub-subhaloes are clearly visible (corresponding to the configuration illustrated in the last step, <math display="inline"><semantics> <msub> <mi>z</mi> <mn>0</mn> </msub> </semantics></math>, in the left panel). The bottom row shows several generations of sub-subhaloes contained within subhalo f.</p>
Full article ">Figure 9
<p>Initial conditions for the orbits of subhaloes infalling into haloes of present-day mass <math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mn>200</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>12</mn> </msup> </mrow> </semantics></math> M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math> (figures taken from [<a href="#B186-galaxies-07-00081" class="html-bibr">186</a>]<a href="#fn033-galaxies-07-00081" class="html-fn">33</a>; see that paper for similar plots for other host masses). <span class="html-italic">Upper left:</span> the distribution of <span class="html-italic">formation redshifts</span> (defined as the redshift at which the mass of the main progenitor of the halo was half its present value). These and the other histograms in this figure are normalized such that the integral over the distribution is unity. <span class="html-italic">Lower left</span>: distribution of infall (accretion) redshifts of subhaloes of different mass ratios, <math display="inline"><semantics> <mi>ξ</mi> </semantics></math> (relative to the host halo at the time of accretion; see legend). <span class="html-italic">Middle:</span> distributions of radial (upper panel) and tangential (lower panel) subhalo orbital velocities at infall, relative to the virial velocity of the host, for the same halo mass and subhalo-to-halo mass ratios as in the lower-left panel. <span class="html-italic">Right:</span> bivariate distribution of orbital parameters for infalling haloes into hosts of mass <math display="inline"><semantics> <mrow> <msub> <mi>M</mi> <mn>200</mn> </msub> <mrow> <mo>(</mo> <mi>z</mi> <mo>=</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mn>10</mn> <mn>13</mn> </msup> </mrow> </semantics></math> M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math>.</p>
Full article ">Figure 10
<p><b>Left:</b> Distribution in the 2D radial phase-space plane of subhaloes identified in a Milky Way-size halo simulation (Via Lactea II [<a href="#B188-galaxies-07-00081" class="html-bibr">188</a>]; figure adapted from [<a href="#B189-galaxies-07-00081" class="html-bibr">189</a>]<a href="#fn035-galaxies-07-00081" class="html-fn">35</a>). Subhaloes are color-coded according to their infall time (measured from <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>). Subhaloes that are just being accreted are radially infalling, while those that were accreted earlier and have completed many orbits lose energy through dynamical friction and sink towards the halo center. <b>Right:</b> the 2D radial phase-space structure of simulation particles in a different Milky Way-size halo simulation (Aquarius [<a href="#B138-galaxies-07-00081" class="html-bibr">138</a>]; figure adapted from [<a href="#B190-galaxies-07-00081" class="html-bibr">190</a>]<a href="#fn036-galaxies-07-00081" class="html-fn">36</a>). Each particle is color-coded according to the number of caustics it passes (roughly proportional to the number of orbits executed by a given particle). The top panel includes bound subhaloes, while the bottom one does not. In the latter, tidal streams from disrupted subhaloes are more clearly visible.</p>
Full article ">Figure 11
<p>Tidal effects in subhaloes. <b>Left:</b> evolution of a subhalo in a circular orbit in a static host halo potential. Since the tidal field strength is constant, the subhalo gradually loses mass (red particles are bound to the subhalo, black particles are unbound) as it orbits in the host halo creating characteristic tidal streams (figure adapted from [<a href="#B208-galaxies-07-00081" class="html-bibr">208</a>]<a href="#fn040-galaxies-07-00081" class="html-fn">40</a>). <b>Right:</b> the effect of tidal shocks. For nearly radial orbits, the variations in the potential near pericentre are rapid (relative to the internal dynamical timescale of the subhalo) and this leads to an impulsive <span class="html-italic">tidal shock</span>, which causes a drastic removal of mass (upper right) and a change in the dark matter distribution (bottom right). In the upper panel the fraction of stripped mass as a function of circularity (see <a href="#sec3dot2-galaxies-07-00081" class="html-sec">Section 3.2</a>), given by the impulsive approximation, is compared with that in a controlled simulation (figure adapted from [<a href="#B209-galaxies-07-00081" class="html-bibr">209</a>]<a href="#fn041-galaxies-07-00081" class="html-fn">41</a>). The model works quite well for radial orbits but it fails for circular orbits (as in the left panel), for which an adiabatic model is more appropriate (Equation (<a href="#FD14-galaxies-07-00081" class="html-disp-formula">14</a>)). In the lower panel, tidal shocking is seen to reduce the mass in the central regions but preserves the asymptotic NFW behavior, while the outer regions become considerably steeper than NFW (figure adapted from [<a href="#B215-galaxies-07-00081" class="html-bibr">215</a>]<a href="#fn042-galaxies-07-00081" class="html-fn">42</a>).</p>
Full article ">Figure 12
<p>Dynamical friction experienced by subhaloes. <b>Left:</b> simulation of a subhalo orbiting a Milky Way-size halo; the initial mass ratio and circularity of the orbit are 0.1 and 0.5, respectively. The images show the projected over- (or under-) density relative to the initial value at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, at different times during the evolution. The thick solid line marks the subhalo orbit, which decays over time due to dynamical friction. This gravitational process induces a wake in the host halo trailing behind the satellite (most clearly visible in the top left panel). The dipole feature at the center of the host is caused by the tidal effect of the subhalo, which perturbs the position of the halo potential minimum. This effect diminishes with time as the satellite is stripped of mass (figure adapted from [<a href="#B217-galaxies-07-00081" class="html-bibr">217</a>]<a href="#fn043-galaxies-07-00081" class="html-fn">43</a>). <b>Right:</b> evolution of the radial distance of a simulated subhalo orbiting a Milky Way-size halo (figure taken from [<a href="#B218-galaxies-07-00081" class="html-bibr">218</a>]<a href="#fn044-galaxies-07-00081" class="html-fn">44</a>). The orbit decays by dynamical friction on a timescale that strongly depends on the initial mass ratio (different colors) and circularity of the orbit (dashed and solid lines). The timescales are well approximated by the fitting formula (Equation (<a href="#FD18-galaxies-07-00081" class="html-disp-formula">18</a>)), which is an improvement over the classical Chandrasekhar formula (Equation (<a href="#FD17-galaxies-07-00081" class="html-disp-formula">17</a>)).</p>
Full article ">Figure 13
<p>Subhalo abundance. <b>Left:</b> the subhalo velocity function at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> for haloes of different maximum circular velocity, from ∼150 km/s to ∼1000 km/s (bottom to top). In terms of the velocity ratio, <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <msub> <mi>V</mi> <mi>sub</mi> </msub> <mo>/</mo> <msub> <mi>V</mi> <mi>h</mi> </msub> </mrow> </semantics></math>, the velocity function is nearly universal, scaling as <math display="inline"><semantics> <msup> <mi>x</mi> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics></math> (dashed line) with a scale-dependent normalization (see Equation (<a href="#FD20-galaxies-07-00081" class="html-disp-formula">20</a>); figure adapted from [<a href="#B223-galaxies-07-00081" class="html-bibr">223</a>]<a href="#fn048-galaxies-07-00081" class="html-fn">48</a>). <b>Right:</b> the number density of subhaloes as a function of halocentric distance in units of the virial radius for Milky Way-size haloes (triangles) and cluster-size haloes (circles). All subhaloes with <math display="inline"><semantics> <mrow> <msub> <mi>m</mi> <mi>sub</mi> </msub> <mo>/</mo> <msub> <mi>M</mi> <mi>h</mi> </msub> <mo>&gt;</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> </mrow> </semantics></math> have been included. The dashed lines are the average NFW fits to the density profiles of the hosts. These functions have been normalized to unity at the virial radius. Figure adapted from [<a href="#B224-galaxies-07-00081" class="html-bibr">224</a>].</p>
Full article ">Figure 14
<p>The inner structure of subhaloes. <span class="html-italic">Left:</span> spherically averaged density profile of subhaloes (which is remarkably similar to that of isolated haloes). The plot shows the density profile of a Milky Way-size halo (solid black line) and eight of its largest subhaloes (color lines). The vertical dotted line marks the radius beyond which the simulation results are converged. The self-similarity in the central region is better appreciated in the inset where the density and radius are scaled to their values at the scale radius, <math display="inline"><semantics> <msub> <mi>r</mi> <mi>s</mi> </msub> </semantics></math>. The figure is for the Via Lactea II simulation and is adapted from [<a href="#B188-galaxies-07-00081" class="html-bibr">188</a>]. <span class="html-italic">Upper right:</span> mean relation between the maximum circular velocity, <math display="inline"><semantics> <msub> <mi>V</mi> <mi>max</mi> </msub> </semantics></math>, and the radius at which it is achieved, <math display="inline"><semantics> <msub> <mi>r</mi> <mi>max</mi> </msub> </semantics></math>, for subhaloes within <math display="inline"><semantics> <msub> <mi>r</mi> <mn>50</mn> </msub> </semantics></math> (the radius within which the mean enclosed density is 50 times the critical density) of one the Milky Way-size halo simulations in the Aquarius project, at different resolution levels (color lines). The red dashed lines show the scatter (<math display="inline"><semantics> <mrow> <mn>68</mn> <mo>%</mo> </mrow> </semantics></math> of the distribution) for the highest resolution level. The dotted line is a fit to the mean relation for subhaloes and lies systematically below the equivalent line for isolated haloes (solid line). <span class="html-italic">Lower right</span>: a measure of concentration for subhaloes (see Equation (<a href="#FD24-galaxies-07-00081" class="html-disp-formula">24</a>)) within different radial ranges, as given in the legend. The solid line corresponds to isolated haloes. Figures adapted from [<a href="#B138-galaxies-07-00081" class="html-bibr">138</a>]<a href="#fn051-galaxies-07-00081" class="html-fn">51</a>.</p>
Full article ">
16 pages, 658 KiB  
Article
Properties of Subhalos in the Interacting Dark Matter Scenario
by Ángeles Moliné, Jascha A. Schewtschenko, Miguel A. Sánchez-Conde, Alejandra Aguirre-Santaella, Sofía A. Cora and Mario G. Abadi
Galaxies 2019, 7(4), 80; https://doi.org/10.3390/galaxies7040080 - 21 Sep 2019
Cited by 2 | Viewed by 2717
Abstract
One possible and natural derivation from the collisionless cold dark matter (CDM) standard cosmological framework is the assumption of the existence of interactions between dark matter (DM) and photons or neutrinos. Such a possible interacting dark matter (IDM) model would imply a suppression [...] Read more.
One possible and natural derivation from the collisionless cold dark matter (CDM) standard cosmological framework is the assumption of the existence of interactions between dark matter (DM) and photons or neutrinos. Such a possible interacting dark matter (IDM) model would imply a suppression of small-scale structures due to a large collisional damping effect, even though the weakly-interacting massive particle (WIMP) can still be the DM candidate. Because of this, IDM models can help alleviate alleged tensions between standard CDM predictions and observations at small mass scales. In this work, we investigate the properties of the DM halo substructure or subhalos formed in a high-resolution cosmological N-body simulation specifically run within these alternative models. We also run its CDM counterpart, which allowed us to compare subhalo properties in both cosmologies. We show that, in the lower mass range covered by our simulation runs, both subhalo concentrations and abundances are systematically lower in IDM compared to the CDM scenario. Yet, as in CDM, we find that median IDM subhalo concentration values increase towards the innermost regions of their hosts for the same mass subhalos. Similarly to CDM, we find IDM subhalos to be more concentrated than field halos of the same mass. Our work has a direct application to studies aimed at the indirect detection of DM where subhalos are expected to boost the DM signal of their host halos significantly. From our results, we conclude that the role of the halo substructure in DM searches will be less important in interacting scenarios than in CDM, but is nevertheless far from being negligible. 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>Median halo concentrations and <math display="inline"> <semantics> <mrow> <mn>1</mn> <mi>σ</mi> </mrow> </semantics> </math> errors as found in our set of simulations, Box (blue) and LGs (red), at <math display="inline"> <semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>. (<b>a</b>) Left panel: median <math display="inline"> <semantics> <msub> <mi>c</mi> <mi mathvariant="normal">V</mi> </msub> </semantics> </math> values as a function of <math display="inline"> <semantics> <msub> <mi>V</mi> <mi>max</mi> </msub> </semantics> </math>. (<b>b</b>) Right panel: <math display="inline"> <semantics> <msub> <mi>c</mi> <mn>200</mn> </msub> </semantics> </math> as a function of <math display="inline"> <semantics> <msub> <mi>M</mi> <mn>200</mn> </msub> </semantics> </math>. In both panels, the circle symbols refer to the IDM simulations, whereas the triangles to CDM.</p>
Full article ">Figure 2
<p>Median subhalo concentrations and <math display="inline"> <semantics> <mrow> <mn>1</mn> <mi>σ</mi> </mrow> </semantics> </math> errors as found in our set of simulations, Box (blue) and LGs (red), at <math display="inline"> <semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>. The circle symbols represent the results from the IDM simulations, whereas the triangle symbols correspond to the CDM results. (<b>a</b>) Left panel: the median <math display="inline"> <semantics> <msub> <mi>c</mi> <mi mathvariant="normal">V</mi> </msub> </semantics> </math> as a function of <math display="inline"> <semantics> <msub> <mi>V</mi> <mi>max</mi> </msub> </semantics> </math>. (<b>b</b>) Right panel: <math display="inline"> <semantics> <msub> <mi>c</mi> <mn>200</mn> </msub> </semantics> </math> as a function of <math display="inline"> <semantics> <msub> <mi>m</mi> <mn>200</mn> </msub> </semantics> </math> as obtained using Equations (<a href="#FD6-galaxies-07-00080" class="html-disp-formula">6</a>) and (<a href="#FD7-galaxies-07-00080" class="html-disp-formula">7</a>) for every subhalo in the simulations.</p>
Full article ">Figure 3
<p>Median subhalo concentrations and <math display="inline"> <semantics> <mrow> <mn>1</mn> <mi>σ</mi> </mrow> </semantics> </math> errors as a function of <math display="inline"> <semantics> <msub> <mi>x</mi> <mi>sub</mi> </msub> </semantics> </math>, i.e., the distance to the centre of the host halo normalized to <math display="inline"> <semantics> <msub> <mi>R</mi> <mn>200</mn> </msub> </semantics> </math>. We show results for <math display="inline"> <semantics> <msub> <mi>c</mi> <mi mathvariant="normal">V</mi> </msub> </semantics> </math> (left) and <math display="inline"> <semantics> <msub> <mi>c</mi> <mn>200</mn> </msub> </semantics> </math> (right) as derived from our set of LG simulations.</p>
Full article ">Figure 4
<p><b>Left panel</b>: Median halo (open circles) and subhalo (filled circles) <math display="inline"> <semantics> <msub> <mi>c</mi> <mi mathvariant="normal">V</mi> </msub> </semantics> </math> concentration values and corresponding <math display="inline"> <semantics> <mrow> <mn>1</mn> <mi>σ</mi> </mrow> </semantics> </math> errors, as a function of <math display="inline"> <semantics> <msub> <mi>V</mi> <mi>max</mi> </msub> </semantics> </math>, as found in our set of simulations for interacting dark matter (IDM) at <math display="inline"> <semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>: Box (blue) and LGs (red). <b>Right panel</b>: the same, but for <math display="inline"> <semantics> <msub> <mi>c</mi> <mn>200</mn> </msub> </semantics> </math> as a function of <math display="inline"> <semantics> <msub> <mi>m</mi> <mn>200</mn> </msub> </semantics> </math>.</p>
Full article ">Figure 5
<p>Cumulative number of subhalos, <math display="inline"> <semantics> <mrow> <mi>N</mi> <mo>(</mo> <mo>&gt;</mo> <msub> <mi mathvariant="normal">m</mi> <mn>200</mn> </msub> <mo>)</mo> </mrow> </semantics> </math>, as a function of subhalo mass, <math display="inline"> <semantics> <msub> <mi>m</mi> <mn>200</mn> </msub> </semantics> </math>, in the case of IDM (circle symbols) and CDM (triangles) as obtained from Box (blue) and LG (red) simulations at <math display="inline"> <semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics> </math>. We also show the corresponding fits using Equation (<a href="#FD8-galaxies-07-00080" class="html-disp-formula">8</a>) with the best-fit parameters reported in <a href="#galaxies-07-00080-t002" class="html-table">Table 2</a> (solid coloured lines).</p>
Full article ">Figure 6
<p>Number density of subhalos as a function of distance to the host halo centre, <math display="inline"> <semantics> <mrow> <msub> <mi>x</mi> <mi>sub</mi> </msub> <mo>=</mo> <msub> <mi>r</mi> <mi>sub</mi> </msub> <mo>/</mo> <msub> <mi>R</mi> <mn>200</mn> </msub> </mrow> </semantics> </math>. We show results for both IDM (circle symbols) and CDM (triangles). Both cases refer to the LG simulation set; see the text for details.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop