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

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (17)

Search Parameters:
Keywords = supergiant stars

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 2282 KiB  
Article
Hybrid Isentropic Twin Stars
by Juan Pablo Carlomagno, Gustavo A. Contrera, Ana Gabriela Grunfeld and David Blaschke
Universe 2024, 10(9), 336; https://doi.org/10.3390/universe10090336 - 23 Aug 2024
Cited by 2 | Viewed by 479
Abstract
We present a study of hybrid neutron stars with color superconducting quark matter cores at a finite temperature that results in sequences of stars with constant entropy per baryon, s/nB=const. For the quark matter equation of state, [...] Read more.
We present a study of hybrid neutron stars with color superconducting quark matter cores at a finite temperature that results in sequences of stars with constant entropy per baryon, s/nB=const. For the quark matter equation of state, we employ a recently developed nonlocal chiral quark model, while nuclear matter is described with a relativistic density functional model of the DD2 class. The phase transition is obtained through a Maxwell construction under isothermal conditions. We find that traversing the mixed phase on a trajectory at low s/nB2 in the phase diagram shows a heating effect, while at larger s/nB the temperature drops. This behavior may be attributed to the presence of a color superconducting quark matter phase at low temperatures and the melting of the diquark condensate which restores the normal quark matter phase at higher temperatures. While the isentropic hybrid star branch at low s/nB2 is connected to the neutron star branch, it becomes disconnected at higher entropy per baryon so that the “thermal twin” phenomenon is observed. We find that the transition from connected to disconnected hybrid star sequences may be estimated with the Seidov criterion for the difference in energy densities. The radii and masses at the onset of deconfinement exhibit a linear relationship and thus define a critical compactness of the isentropic star configuration for which the transition occurs and which, for large enough s/nB2 values, is accompanied by instability. The results of this study may be of relevance for uncovering the conditions for the supernova explodability of massive blue supergiant stars using the quark deconfinement mechanism. The accretion-induced deconfinement transition with thermal twin formation may contribute to explaining the origin of eccentric orbits in some binary systems and the origin of isolated millisecond pulsars. Full article
(This article belongs to the Special Issue Studies in Neutron Stars)
Show Figures

Figure 1

Figure 1
<p>Phase diagram of the nonlocal chiral quark matter model in the <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>−</mo> <msub> <mi>μ</mi> <mi>B</mi> </msub> </mrow> </semantics></math> plane. The black dashed–double dotted line corresponds to the chiral crossover transition while the grey dashed–dotted line shows the second-order phase transition to 2SC. The solid-colored lines correspond to points with a constant entropy per baryon ratio <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Isotherms of pressure vs. the baryochemical potential for the hadronic DD2 model (dashed lines) and the nonlocal chiral quark model 3DNJL (solid lines). The highlighted crossing points indicate the Maxwell construction of first-order phase transitions.</p>
Full article ">Figure 3
<p>Maxwell transition points in the <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>−</mo> <msub> <mi>μ</mi> <mi>B</mi> </msub> <mo>−</mo> <mi>P</mi> </mrow> </semantics></math> space (shown as black dots) and the corresponding two-dimensional projections (shown as colored dots).</p>
Full article ">Figure 4
<p>Constant entropy per baryon trajectories for the DD2 hadronic EOS (dashed lines) and for the 3DNJL superconducting quark matter EOS (solid lines) in the QCD phase diagrams in the <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>−</mo> <msub> <mi>μ</mi> <mi>B</mi> </msub> </mrow> </semantics></math> plane. The black dashed–double dotted and the grey dashed–dotted lines indicate the corresponding QM phase transitions as in <a href="#universe-10-00336-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 5
<p>Constant entropy per baryon trajectories for the DD2 hadronic EOS (dashed lines) and for the 3DNJL superconducting quark matter EOS (solid lines) in the QCD phase diagrams in the <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>−</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> </mrow> </semantics></math> plane. The dotted lines indicate the proposed mixed-phase construction described in the text. The black dashed-double dotted and the grey dashed-dotted lines indicate the corresponding QM phase transitions as in <a href="#universe-10-00336-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 6
<p>Hybrid EOS (<b>a</b>) and squared speed of sound (<b>b</b>) as a function of the energy density for different values of entropy per baryon, <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> </mrow> </semantics></math>, indicated by different colors.</p>
Full article ">Figure 7
<p>Mass–radius diagram for hot hybrid neutron star sequences with different <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> </mrow> </semantics></math>. For <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> <mo>≥</mo> <mn>2.0</mn> </mrow> </semantics></math>, one recognizes the occurrence of thermal twin stars. The dotted straight line connects the onset masses for the deconfinement transition.</p>
Full article ">Figure 8
<p>Onset masses for the deconfinement transition as a function of the entropy per baryon (triangle up) <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> </mrow> </semantics></math> and a polynomial fit (solid line).</p>
Full article ">Figure 9
<p>Gravitational mass (solid lines) and baryon mass (dashed lines) versus central energy density for different values of entropy per baryon of the hot hybrid neutron star configurations.</p>
Full article ">Figure 10
<p>Gravitational mass defect construction for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>4.0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Same as <a href="#universe-10-00336-f010" class="html-fig">Figure 10</a> for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>3.0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Same as <a href="#universe-10-00336-f011" class="html-fig">Figure 11</a> for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>2.0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Mass defect vs. <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> </mrow> </semantics></math> (open squares) and exponential fit (red solid line).</p>
Full article ">Figure 14
<p>Dimensionless onset critical pressure <math display="inline"><semantics> <mrow> <msubsup> <mi>P</mi> <mi>c</mi> <mrow> <mi>H</mi> <mi>M</mi> </mrow> </msubsup> <mo>/</mo> <msubsup> <mi>ε</mi> <mi>c</mi> <mrow> <mi>H</mi> <mi>M</mi> </mrow> </msubsup> </mrow> </semantics></math> vs. <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> </mrow> </semantics></math> and a linear fit (red solid line).</p>
Full article ">Figure 15
<p>Fulfillment of the Seidov criterion <math display="inline"><semantics> <mrow> <mi>X</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math> indicates the gravitational instability that triggers the occurrence of thermal twin star branches (open square symbols). From the polynomial fit (red solid line), we obtain <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> <msub> <mrow> <mo>|</mo> </mrow> <mi>crit</mi> </msub> <mo>=</mo> <mn>1.81</mn> <mo>±</mo> <mn>0.04</mn> </mrow> </semantics></math>. Some additional intermediate QM onset points are included to improve the fit. The inset shows a zoom-in of the region with the onset of the deconfinement transition.</p>
Full article ">Figure 16
<p>In agreement with the result shown in <a href="#universe-10-00336-f015" class="html-fig">Figure 15</a> from the Seidov criterion <math display="inline"><semantics> <mrow> <mi>X</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>, this plot shows the onset of twin star configurations after the pointed value of <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>/</mo> <msub> <mi>n</mi> <mi>B</mi> </msub> <mo>=</mo> <mn>1.81</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 17
<p>The same as <a href="#universe-10-00336-f016" class="html-fig">Figure 16</a> but relating gravitational masses and central energy density.</p>
Full article ">
19 pages, 686 KiB  
Article
Probing the Propeller Regime with Symbiotic X-ray Binaries
by Marina D. Afonina and Sergei B. Popov
Universe 2024, 10(5), 205; https://doi.org/10.3390/universe10050205 - 3 May 2024
Cited by 2 | Viewed by 953
Abstract
At the moment, there are two neutron star X-ray binaries with massive red supergiants as donors. Recently, De et al. (2023) proposed that the system SWIFT J0850.8-4219 contains a neutron star at the propeller stage. We study this possibility by applying various models [...] Read more.
At the moment, there are two neutron star X-ray binaries with massive red supergiants as donors. Recently, De et al. (2023) proposed that the system SWIFT J0850.8-4219 contains a neutron star at the propeller stage. We study this possibility by applying various models of propeller spin-down. We demonstrate that the duration of the propeller stage is very sensitive to the regime of rotational losses. Only in the case of a relatively slow propeller model proposed by Davies and Pringle in 1981, the duration of the propeller is long enough to provide a significant probability to observe the system at this stage. Future determination of the system parameters (orbital and spin periods, magnetic field of the compact object, etc.) will allow putting strong constraints on the propeller behavior. Full article
(This article belongs to the Special Issue Universe: Feature Papers 2024 – Compact Objects)
Show Figures

Figure 1

Figure 1
<p>Donor mass loss rate <math display="inline"><semantics> <msub> <mover accent="true"> <mi>M</mi> <mo>˙</mo> </mover> <mi mathvariant="normal">w</mi> </msub> </semantics></math> (blue solid curve) and wind velocity <math display="inline"><semantics> <mrow> <msub> <mi>v</mi> <mi mathvariant="normal">w</mi> </msub> <mrow> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics></math> (black dashed curve) at <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>1280</mn> <mspace width="0.166667em"/> <msub> <mi>R</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math> over the time elapsed since the birth of the NS. Vertical lines indicate bi-stability jumps. The second bi-stability jump determines the change in donor evolution from the Main sequence to the red supergiant stage.</p>
Full article ">Figure 2
<p>Evolution of <math display="inline"><semantics> <mover accent="true"> <mi>M</mi> <mo>˙</mo> </mover> </semantics></math>, spin period <span class="html-italic">P</span>, and the absolute value of its time derivative <math display="inline"><semantics> <mover accent="true"> <mi>P</mi> <mo>˙</mo> </mover> </semantics></math> for the NS with a constant magnetic field <math display="inline"><semantics> <mrow> <mi>B</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>12</mn> </msup> </mrow> </semantics></math> G and initial spin period <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math> ms in a binary with the semi-major axis <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1280</mn> <mspace width="0.166667em"/> <msub> <mi>R</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math>. For each propeller model, the NS makes a transition from the ejector to the propeller and then to the accretor stage, except for the model D where the accretor stage is not reached. Black solid curves refer to all propeller models. Black dotted vertical lines and black-filled circles show the transition to the propeller stage, which is the same for all models. Panels on the right side, show the zoomed region near the ejector-propeller transition. Colored vertical dashed lines and colored-filled circles indicate the transition to the accretor stage for each model. The dashed lines are not shown on the left panels, since there they would overlap with the black dotted line. Colored curves for models A1, A, B, and C overlap at the accretor stage and are shown in green on the left panels. The black arrow in the left middle panel points to the sign change of <math display="inline"><semantics> <mover accent="true"> <mi>P</mi> <mo>˙</mo> </mover> </semantics></math> value at <math display="inline"><semantics> <mrow> <mn>6.7</mn> </mrow> </semantics></math> Myr.</p>
Full article ">Figure 3
<p>Evolution of <math display="inline"><semantics> <mover accent="true"> <mi>M</mi> <mo>˙</mo> </mover> </semantics></math>, spin period <span class="html-italic">P</span>, and the absolute value of its time derivative <math display="inline"><semantics> <mover accent="true"> <mi>P</mi> <mo>˙</mo> </mover> </semantics></math> for the NS with a constant magnetic field <math display="inline"><semantics> <mrow> <mi>B</mi> <mo>=</mo> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>12</mn> </msup> </mrow> </semantics></math> G. Line styles, points, and colors are the same as in <a href="#universe-10-00205-f002" class="html-fig">Figure 2</a>. The transition to the accretor stage is now shown in the left panels. Here the grey vertical dashed line shows the transition to the accretor stage for models A1, B, and C, while the red one is for propeller model A. The evolution of an NS within model D does not lead to the accretor stage. In the middle panels, the black arrow at <math display="inline"><semantics> <mrow> <mn>6.7</mn> </mrow> </semantics></math> Myr on the left and <math display="inline"><semantics> <mrow> <mn>6.679</mn> </mrow> </semantics></math> Myr on the right show the change in sign of <math display="inline"><semantics> <mover accent="true"> <mi>P</mi> <mo>˙</mo> </mover> </semantics></math>. After the NS period reaches <math display="inline"><semantics> <msub> <mi>P</mi> <mi>eq</mi> </msub> </semantics></math>, the value of <math display="inline"><semantics> <mover accent="true"> <mi>P</mi> <mo>˙</mo> </mover> </semantics></math> fluctuates visibly.</p>
Full article ">Figure 4
<p>Evolutionary stages of NSs with <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math> ms, <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1280</mn> <mspace width="0.166667em"/> <msub> <mi>R</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math> over the age <span class="html-italic">t</span> of the NS and its magnetic field <span class="html-italic">B</span>. Propeller model C is shown on the left panel and model D is shown on the right. Three evolutionary stages are shown in color: ejector (blue), propeller (light blue), and accretor (orange). White dashed lines correspond to the evolutionary phases of the donor star, as in <a href="#universe-10-00205-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 5
<p>The evolution of an NS with <math display="inline"><semantics> <mrow> <msub> <mi>P</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> s, <math display="inline"><semantics> <mrow> <mi>B</mi> <mo>=</mo> <mn>4</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>12</mn> </msup> </mrow> </semantics></math> G in a binary with <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1280</mn> <mspace width="0.166667em"/> <msub> <mi>R</mi> <mo>⊙</mo> </msub> </mrow> </semantics></math>. The first evolutionary stage of the NS is the propeller stage. Therefore there is no vertical dotted line for the ejector-propeller transition. Otherwise, all curves, lines, and dots have the same style as in <a href="#universe-10-00205-f003" class="html-fig">Figure 3</a>. Black arrows indicate changes in the sign of <math display="inline"><semantics> <mover accent="true"> <mi>P</mi> <mo>˙</mo> </mover> </semantics></math> as in <a href="#universe-10-00205-f002" class="html-fig">Figure 2</a> and <a href="#universe-10-00205-f003" class="html-fig">Figure 3</a>. In the middle right panel, the curve for model A1 has an additional sign change at <math display="inline"><semantics> <mrow> <mn>6.685</mn> </mrow> </semantics></math> Myr after the propeller-accretor transition, similar to the model A (<math display="inline"><semantics> <mrow> <mn>6.676</mn> </mrow> </semantics></math> Myr), which is also indicated by a black arrow.</p>
Full article ">
28 pages, 1902 KiB  
Article
Unveiling the Evolutionary State of Three B Supergiant Stars: PU Gem, ϵ CMa, and η CMa
by Julieta Paz Sánchez Arias, Péter Németh, Elisson Saldanha da Gama de Almeida, Matias Agustin Ruiz Diaz, Michaela Kraus and Maximiliano Haucke
Galaxies 2023, 11(5), 93; https://doi.org/10.3390/galaxies11050093 - 29 Aug 2023
Cited by 3 | Viewed by 2001
Abstract
We aim to combine asteroseismology, spectroscopy, and evolutionary models to establish a comprehensive picture of the evolution of Galactic blue supergiant stars (BSG). To start such an investigation, we selected three BSG candidates for our analysis: HD 42087 (PU Gem), HD 52089 ( [...] Read more.
We aim to combine asteroseismology, spectroscopy, and evolutionary models to establish a comprehensive picture of the evolution of Galactic blue supergiant stars (BSG). To start such an investigation, we selected three BSG candidates for our analysis: HD 42087 (PU Gem), HD 52089 (ϵ CMa), and HD 58350 (η CMa). These stars show pulsations and were suspected to be in an evolutionary stage either preceding or succeding the red supergiant (RSG) stage. For our analysis, we utilized the 2-min cadence TESS data to study the photometric variability, and we obtained new spectroscopic observations at the CASLEO observatory. We used non-LTE radiative transfer models calculated with CMFGEN to derive their stellar and wind parameters. For the fitting procedure, we included CMFGEN models in the iterative spectral analysis pipeline XTgrid to determine their CNO abundances. The spectral modeling was limited to changing only the effective temperature, surface gravity, CNO abundances, and mass-loss rates. Finally, we compared the derived metal abundances with prediction from Geneva stellar evolution models. The frequency spectra of all three stars show stochastic oscillations and indications of one nonradial strange mode, fr= 0.09321 d1 in HD 42087 and a rotational splitting centred in f2= 0.36366 d1 in HD 52089. We conclude that the rather short sectoral observing windows of TESS prevent establishing a reliable mode identification of low frequencies connected to mass-loss variabilities. The spectral analysis confirmed gradual changes in the mass-loss rates, and the derived CNO abundances comply with the values reported in the literature. We were able to achieve a quantitative match with stellar evolution models for the stellar masses and luminosities. However, the spectroscopic surface abundances turned out to be inconsistent with the theoretical predictions. The stars show N enrichment, typical for CNO cycle processed material, but the abundance ratios did not reflect the associated levels of C and O depletion. We found HD 42087 to be the most consistent with a pre-RSG evolutionary stage, HD 58350 is most likely in a post-RSG evolution and HD 52089 shows stellar parameters compatible with a star at the TAMS. Full article
(This article belongs to the Special Issue Theory and Observation of Active B-type Stars)
Show Figures

Figure 1

Figure 1
<p>Our three BSG stars in the HR diagram according to the values derived in [<a href="#B11-galaxies-11-00093" class="html-bibr">11</a>]. The evolutionary tracks are taken from [<a href="#B22-galaxies-11-00093" class="html-bibr">22</a>].</p>
Full article ">Figure 2
<p>Evolution of the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line profile of HD 42087 between 2006 and 2020. The emission component weakened by 2020.</p>
Full article ">Figure 3
<p>Evolution of the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line profile of HD 52089 between 2013 and 2015.</p>
Full article ">Figure 4
<p>Evolution of the H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line profile of HD 58350 between 2006 and 2020. The emission component has strengthened between 2006 and 2013 and became noticeably weaker by 2020.</p>
Full article ">Figure 5
<p><b>Top panel</b>: Light curve of HD 42087 corresponding to Sectors 43, 44, and 45. <b>Lower panels</b>: Amplitude spectra for each sector and all sectors combined.</p>
Full article ">Figure 6
<p><b>Top panel</b>: Light curve of HD 52089 corresponding to Sectors 6 and 7. <b>Lower panels</b>: Amplitude spectra for each sector and all sectors combined.</p>
Full article ">Figure 7
<p><b>Top panel</b>: Light curve of HD 52089 corresponding to Sectors 33 and 34. <b>Lower panels</b>: Amplitude spectra for each sector and all sectors combined.</p>
Full article ">Figure 8
<p><b>Top panel</b>: Light curve of HD 58350 acquired during Sector 34. <b>Lower panel</b>: Amplitude spectrum for the same sector.</p>
Full article ">Figure 9
<p>Best-fit <span class="html-small-caps">XTgrid/CMFGEN</span> model for HD 42087. In each panel, the CASLEO observation is in grey, the CMFGEN model in black, and the residuals, shifted by +1.1 for clarity, are in red.</p>
Full article ">Figure 10
<p>Best-fit <span class="html-small-caps">XTgrid/CMFGEN</span> model for HD 52089. In each panel, the CASLEO observation is in grey, the CMFGEN model in black, and the residuals, shifted by +1.1 for clarity, are in red.</p>
Full article ">Figure 11
<p>Best-fit <span class="html-small-caps">XTgrid/CMFGEN</span> model for HD 58350. In each panel, the CASLEO observation is in grey, the CMFGEN model in black, and the residuals, shifted by +1.1 for clarity, are in red.</p>
Full article ">Figure 12
<p>Left panels: HR diagram (<b>top</b>) and mass evolution (<b>bottom</b>) showing the position of HD 42087 and evolutionary tracks with initial masses of 24, 26, and 28 <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math>. Middle panels: The same as in the left panels for HD 52087 and initial masses of 10, 12, and 14 <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math>. Right panels: the same as before for HD 58350 and initial masses of 20 and 22 <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math>. Solid and dashed lines represent models with <math display="inline"><semantics> <mrow> <mo>Ω</mo> <mo>/</mo> <msub> <mo>Ω</mo> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> 0.568 from [<a href="#B22-galaxies-11-00093" class="html-bibr">22</a>] and the interpolated ones for <math display="inline"><semantics> <mrow> <mo>Ω</mo> <mo>/</mo> <msub> <mo>Ω</mo> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> 0.4, respectively.</p>
Full article ">Figure 13
<p>SED of HD 42087. All data points were taken from the VizieR Photometry Viewer service. The photometric data were de-reddened using <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>(</mo> <mi>B</mi> <mo>−</mo> <mi>V</mi> <mo>)</mo> </mrow> </semantics></math> = 0.4 mag. The green points were used to match the slope of the passband convolved CMFGEN fluxes to the observations, and the model was normalized to the observed SED in the 2MASS/J band. The binned IUE spectrum is included with black dots.</p>
Full article ">Figure 14
<p>SED of HD 52089. Same as <a href="#galaxies-11-00093-f013" class="html-fig">Figure 13</a>, but de-reddening was conducted using <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>(</mo> <mi>B</mi> <mo>−</mo> <mi>V</mi> <mo>)</mo> </mrow> </semantics></math>= 0.005 mag.</p>
Full article ">Figure 15
<p>SED of HD 58350. Same as <a href="#galaxies-11-00093-f013" class="html-fig">Figure 13</a>, but de-reddening was conducted using <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>(</mo> <mi>B</mi> <mo>−</mo> <mi>V</mi> <mo>)</mo> </mrow> </semantics></math> = 0.18 mag (see main text).</p>
Full article ">Figure 16
<p>The measured CNO abundance patterns in all three stars compared to the solar pattern from Asplund et al. [<a href="#B52-galaxies-11-00093" class="html-bibr">52</a>]. In all three stars, C and O are depleted and N is overabundant compared to the solar mixture. The mean CNO abundances found by Searle et al. [<a href="#B17-galaxies-11-00093" class="html-bibr">17</a>] for Galactic BSGs are also shown for reference.</p>
Full article ">Figure 17
<p>N/C (<b>bottom</b>) and N/O (<b>top</b>) abundance ratios by number from Searle et al. [<a href="#B17-galaxies-11-00093" class="html-bibr">17</a>] and Georgy et al. [<a href="#B67-galaxies-11-00093" class="html-bibr">67</a>] compared to our measurements, as well as Geneva solar metallicity (Z = 0.014), <math display="inline"><semantics> <mrow> <mo>Ω</mo> <mo>/</mo> <msub> <mo>Ω</mo> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> </mrow> </semantics></math> 0.4 interpolated evolutionary tracks for 5, 10, 20, 23, and 26 <math display="inline"><semantics> <msub> <mi>M</mi> <mo>⊙</mo> </msub> </semantics></math>. For reference, we also show B-type main sequence stars from Lyubimkov et al. [<a href="#B68-galaxies-11-00093" class="html-bibr">68</a>].</p>
Full article ">Figure 18
<p>N/C and N/O abundance ratio correlations for our sample and the same tracks as in <a href="#galaxies-11-00093-f017" class="html-fig">Figure 17</a>. We included Searle et al. [<a href="#B17-galaxies-11-00093" class="html-bibr">17</a>] and Georgy et al. [<a href="#B67-galaxies-11-00093" class="html-bibr">67</a>] samples and the B-type main sequence stars from Lyubimkov et al. [<a href="#B68-galaxies-11-00093" class="html-bibr">68</a>].</p>
Full article ">
21 pages, 1509 KiB  
Article
Discovering New B[e] Supergiants and Candidate Luminous Blue Variables in Nearby Galaxies
by Grigoris Maravelias, Stephan de Wit, Alceste Z. Bonanos, Frank Tramper, Gonzalo Munoz-Sanchez and Evangelia Christodoulou
Galaxies 2023, 11(3), 79; https://doi.org/10.3390/galaxies11030079 - 19 Jun 2023
Cited by 2 | Viewed by 1551
Abstract
Mass loss is one of the key parameters that determine stellar evolution. Despite the progress we have achieved over the last decades we still cannot match the observational derived values with theoretical predictions. Even worse, there are certain phases, such as the B[e] [...] Read more.
Mass loss is one of the key parameters that determine stellar evolution. Despite the progress we have achieved over the last decades we still cannot match the observational derived values with theoretical predictions. Even worse, there are certain phases, such as the B[e] supergiants (B[e]SGs) and the Luminous Blue Variables (LBVs), where significant mass is lost through episodic or outburst activity. This leads to various structures forming around them that permit dust formation, making these objects bright IR sources. The ASSESS project aims to determine the role of episodic mass in the evolution of massive stars, by examining large numbers of cool and hot objects (such as B[e]SGs/LBVs). For this purpose, we initiated a large observation campaign to obtain spectroscopic data for ∼1000 IR-selected sources in 27 nearby galaxies. Within this project we successfully identified seven B[e] supergiants (one candidate) and four Luminous Blue Variables of which six and two, respectively, are new discoveries. We used spectroscopic, photometric, and light curve information to better constrain the nature of the reported objects. We particularly noted the presence of B[e]SGs at metallicity environments as low as 0.14 Z. Full article
(This article belongs to the Special Issue Theory and Observation of Active B-type Stars)
Show Figures

Figure 1

Figure 1
<p>Spectra of objects classified as B[e]SGs (including the B[e]SG candidate NGC7793-1). (Left) The full spectra for all stars with small offsets for better illustration purposes. The most prominent emission features are indicated. (Right) The region around H<math display="inline"><semantics> <mi>α</mi> </semantics></math> is highlighted to emphasize the relative strength of the emission compared to the continuum.</p>
Full article ">Figure 2
<p>Similar to <a href="#galaxies-11-00079-f001" class="html-fig">Figure 1</a>, but for LBVc. We note the lack of forbidden emission lines.</p>
Full article ">Figure 3
<p>The region between the [O <span class="html-small-caps">i</span>] and H<math display="inline"><semantics> <mi>α</mi> </semantics></math> line, that showcases multiple Fe <span class="html-small-caps">ii</span> emission lines. We note the clear presence of [O <span class="html-small-caps">i</span>] <math display="inline"><semantics> <mi>λ</mi> </semantics></math>6300 line for the B[e]SGs (left and middle panels, with the exception of the candidate NGC7793-1, due to the problematic spectrum; see text for more) and its absence from the LBVc spectra (right panel).</p>
Full article ">Figure 4
<p>The region around the [Ca <span class="html-small-caps">ii</span>] emission doublet. Its presence is evident in some B[e]SGs (left and middle panels, including NGC7793-1 candidate source, that suffers from data reduction artifacts due to slit overlaps), while LBVc (right panel) do not typically exhibit these lines (except for NGC55-3).</p>
Full article ">Figure 5
<p>The light curves from the Pan-STARRS survey for the B[e]SGs WLM-1 and NGC247-1. Each panel (per filter) shows the difference of each epoch from the mean value (noted on the y-axis label). See text for more.</p>
Full article ">Figure 6
<p>Same as <a href="#galaxies-11-00079-f005" class="html-fig">Figure 5</a>, but for the candidate LBVs NGC3109-1 and NGC247-2.</p>
Full article ">Figure 7
<p>(Left) The mid-IR <span class="html-italic">WISE</span> CCD for B[e]SGs and LBVs, including sources from the MCs (after [<a href="#B7-galaxies-11-00079" class="html-bibr">7</a>]) and our sample (for 5 out of 11 sources with <span class="html-italic">WISE</span> data). In general, the separation also holds for the new sources, with the exception of NGC55-1 (see text for more). (Right) IR CMD combining near-IR <span class="html-italic">J</span>-band (available for only five of our sources) with <span class="html-italic">Spitzer</span> [3.6]. We notice that, in this case, the newly found sources are consistent with the positions of the MC sources.</p>
Full article ">Figure 8
<p>(Left) The optical (<span class="html-italic">Gaia</span>) CMD, plotting BP–RP vs. M<math display="inline"><semantics> <msub> <mrow/> <mi>G</mi> </msub> </semantics></math> magnitude. We included all our sample and the MC sources from [<a href="#B7-galaxies-11-00079" class="html-bibr">7</a>] (except for two sources without a complete dataset in both <span class="html-italic">Gaia</span> and <span class="html-italic">Spitzer</span> surveys). (Right) The mid-IR (<span class="html-italic">Spitzer</span>) CMD using the IR color [3.6]–[4.5] vs. M<math display="inline"><semantics> <msub> <mrow/> <mrow> <mo>[</mo> <mn>4.5</mn> <mo>]</mo> </mrow> </msub> </semantics></math>. In this case, there is a significant improvement in the separation between the two classes. The position of NGC7793-1 favors a B[e]SG nature (see text for more).</p>
Full article ">Figure 9
<p>Similar to <a href="#galaxies-11-00079-f008" class="html-fig">Figure 8</a> but plotting the IR color [3.6]–[4.5] vs. the optical M<math display="inline"><semantics> <msub> <mrow/> <mi>G</mi> </msub> </semantics></math> magnitude. Similar to the IR CMD we saw relatively good separation between the two classes, with LBVs being brighter in the optical and less dusty compared to the B[e]SGs.</p>
Full article ">Figure 10
<p>The cumulative distribution function of the B[e]SGs and LBVs (including candidates) from this work and the literature. We notice (for the first time) the presence of B[e]SGs in lower metallicity environments and the fact that the two populations are not totally different (see text for more).</p>
Full article ">
23 pages, 968 KiB  
Article
Dense Molecular Environments of B[e] Supergiants and Yellow Hypergiants
by Michaela Kraus, Michalis Kourniotis, María Laura Arias, Andrea F. Torres and Dieter H. Nickeler
Galaxies 2023, 11(3), 76; https://doi.org/10.3390/galaxies11030076 - 16 Jun 2023
Cited by 3 | Viewed by 1422
Abstract
Massive stars expel large amounts of mass during their late evolutionary phases. We aim to unveil the physical conditions within the warm molecular environments of B[e] supergiants (B[e]SGs) and yellow hypergiants (YHGs), which are known to be embedded in circumstellar shells and disks. [...] Read more.
Massive stars expel large amounts of mass during their late evolutionary phases. We aim to unveil the physical conditions within the warm molecular environments of B[e] supergiants (B[e]SGs) and yellow hypergiants (YHGs), which are known to be embedded in circumstellar shells and disks. We present K-band spectra of two B[e]SGs from the Large Magellanic Cloud and four Galactic YHGs. The CO band emission detected from the B[e]SGs LHA 120-S 12 and LHA 120-S 134 suggests that these stars are surrounded by stable rotating molecular rings. The spectra of the YHGs display a rather diverse appearance. The objects 6 Cas and V509 Cas lack any molecular features. The star [FMR2006] 15 displays blue-shifted CO bands in emission, which might be explained by a possible close to pole-on oriented bipolar outflow. In contrast, HD 179821 shows blue-shifted CO bands in absorption. While the star itself is too hot to form molecules in its outer atmosphere, we propose that it might have experienced a recent outburst. We speculate that we currently can only see the approaching part of the expelled matter because the star itself might still block the receding parts of a (possibly) expanding gas shell. Full article
(This article belongs to the Special Issue Theory and Observation of Active B-type Stars)
Show Figures

Figure 1

Figure 1
<p>HR diagram with evolutionary tracks for LMC (<math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <mn>0.006</mn> </mrow> </semantics></math>, [<a href="#B76-galaxies-11-00076" class="html-bibr">76</a>], left panel) and solar (<math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <mn>0.014</mn> </mrow> </semantics></math>, [<a href="#B42-galaxies-11-00076" class="html-bibr">42</a>], right panel) metallicity for models of rotating stars (<math display="inline"><semantics> <mrow> <mi>v</mi> <mo>/</mo> <msub> <mi>v</mi> <mi>crit</mi> </msub> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math>) with initial masses from <math display="inline"><semantics> <mrow> <mn>20</mn> <mrow> <mo>−</mo> </mrow> <mn>40</mn> </mrow> </semantics></math> M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math>. Shown are the positions of LMC (blue symbols) and Galactic (red symbols) objects. Only B[e]SGs with hot circumstellar molecular CO gas are shown. These populate similar evolutionary tracks as the YHGs. The minimum and maximum temperature values (where known) of the YHGs are connected by dashed lines. YHGs with reported (at least once) CO band emission are shown with triangles. The stars of the current study are labeled.</p>
Full article ">Figure 2
<p>Best fitting model (red) to the normalized Phoenix spectra (black) of LHA 120-S 12 (<b>top</b>) and LHA 120-S 134 (<b>bottom</b>). For each star we display the entire fit to the total spectrum (top panels) and the zoom to the band heads (bottom panels). Emission from the Pfund series (blue dashed line), detected in the spectrum of LHA 120-S 134, is included in the total fit.</p>
Full article ">Figure 3
<p>Normalized medium-resolution <span class="html-italic">K</span>-band spectra of the YHGs taken with GNIRS. For better visualization, the spectra have been offset along the flux axis. Positions of the CO band heads and of the lines from the Pfund series are marked by ticks. The lines of Br<math display="inline"><semantics> <mi>γ</mi> </semantics></math> and of the Na <span class="html-small-caps">i</span> doublet are labeled as well.</p>
Full article ">Figure 4
<p>Best fitting model (red) to the observed (black) <span class="html-italic">K</span>-band spectrum of [FMR2006] 15 for the model including a rotational component (<b>top</b>) and a pure Gaussian broadening (<b>bottom</b>).</p>
Full article ">Figure 5
<p>Residual spectrum of [FMR2006] 15 after subtraction of the blue-shifted CO band emission. For illustration purposes, a synthetic model spectrum of a cool supergiant is shown as well (shifted down along the flux axis for better visualization) with parameters similar to those determined for [FMR2006] 15.</p>
Full article ">Figure 6
<p>Comparison of the <span class="html-italic">K</span>-band spectrum of HD 179821 (<b>top</b>) with synthetic spectra for effective temperatures of 5400 K (<b>middle</b>) and 6600 K (<b>bottom</b>). For illustration purposes, the synthetic model spectra are included in this figure and shifted down along the flux axis for better visualization.</p>
Full article ">
40 pages, 90706 KiB  
Article
Synthetic Light Curve Design for Pulsating Binary Stars to Compare the Efficiency in the Detection of Periodicities
by Aldana Alberici Adam, Gunther F. Avila Marín, Alejandra Christen and Lydia Sonia Cidale
Galaxies 2023, 11(3), 69; https://doi.org/10.3390/galaxies11030069 - 31 May 2023
Cited by 1 | Viewed by 1275
Abstract
B supergiant stars pulsate in regular and quasi-regular oscillations resulting in intricate light variations that might conceal their binary nature. To discuss possible observational bias in a light curve, we performed a simulation design of a binary star affected by sinusoidal functions emulating [...] Read more.
B supergiant stars pulsate in regular and quasi-regular oscillations resulting in intricate light variations that might conceal their binary nature. To discuss possible observational bias in a light curve, we performed a simulation design of a binary star affected by sinusoidal functions emulating pulsation phenomena. The Period04 tool and the WaveletComp package of R were used for this purpose. Thirty-two models were analysed based on a combination of two values on each of the k = 6 variables, such as multiple pulsations, the amplitude of the pulsation, the pulsation frequency, the beating phenomenon, the light-time effect, and regular or quasi-regular periods. These synthetic models, unlike others, consider an ARMA (1, 1) statistical noise, irregular sampling, and a gap of about 4 days. Comparing Morlet wavelet with Fourier methods, we observed that the orbital period and its harmonics were well detected in most cases. Although the Fourier method provided more accurate period detection, the wavelet analysis found it more times. Periods seen with the wavelet method have a shift due to the slightly irregular time scale used. The pulsation period hitting rate depends on the wave amplitude and frequency with respect to eclipse depth and orbital period. None of the methods was able to distinguish accurate periods leading to a beating phenomenon when they were longer than the orbital period, resulting, in both cases, in an intermediate value. When the beating period was shorter, the Fourier analysis found it in all cases except for unsolved quasi-regular periods. Overall, the Morlet wavelet analysis performance was lower than the Fourier analysis. Considering the strengths and disadvantages found in these methods, we recommend using at least two diagnosis tools for a detailed time series data analysis to obtain confident results. Moreover, a fine-tuning of trial periods by applying phase diagrams would be helpful for recovering accurate values. The combined analysis could reduce observational bias in searching binaries using photometric techniques. Full article
(This article belongs to the Special Issue Theory and Observation of Active B-type Stars)
Show Figures

Figure 1

Figure 1
<p>Synthetic light curve of an eclipsing binary affected by ARMA (1, 1) noise.</p>
Full article ">Figure 2
<p>Example of applying wavelet analysis for a multi-periodic time series. Left panel: sinusoidal series with periods 30, 38, and 80. Central panel: Scalogram or wavelet power spectrum of the series. The top in the color code indicates a higher significance for periods. Periods with a higher power are detected with solid black lines. Right panel: average wavelet power of the series in logarithm scale. In this diagram, the beating periods 30 and 38 are more difficult to identify individually, unlike period 80, but they show a striking feature in the scalogram. Although all the periods represented in red have a significance level less than or equal to 0.05, the most relevant ones, about 30, 38, and 80, are the three highest average wavelet power. They were obtained using the WaveletComp Monte Carlo sampling-based test.</p>
Full article ">Figure 3
<p>Synthetic light curves obtained from the simulation design. The model number is placed according to the list in <a href="#galaxies-11-00069-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 4
<p>Results of the wavelet analysis for Model No. 0, 1, 5, 17, and 21. From left to right, the wavelet power spectrum and the average wavelet power for the second part of each synthetic light curve.</p>
Full article ">Figure 5
<p>Periodograms obtained for the models No. 0, 1, 5, 17, and 21.</p>
Full article ">Figure 6
<p>Comparison between the proportion of correctness of the original period obtained by wavelet (cyan) and Fourier (grey) analysis. The thick solid black line represents the median of the data, and the red symbol stands for an outlier or extreme value.</p>
Full article ">Figure 7
<p>Boxplots of hit ratio of the original periods with the wavelet (cyan) and Fourier (grey) analysis for each variable involved. The thick solid black line represents the median of the data, and the red symbol stands for an outlier or extreme value. The hit rate is on the <span class="html-italic">y</span>-axis and the two categories (indicated by <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> and 1) for each variable are shown on <span class="html-italic">x</span>-axis: beating phenomenon (absence/presence), period of the pulsation (minor/major), amplitude of the pulsation (<math display="inline"><semantics> <mrow> <mn>4</mn> <mo>%</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>20</mn> <mo>%</mo> </mrow> </semantics></math>), number of pulsations periods (1/2), and quasi-regular pulsations (absence/presence).</p>
Full article ">Figure 8
<p>Light curves obtained by TESS of pulsating binary stars.</p>
Full article ">Figure A1
<p>Periodograms obtained from the Fourier analysis for models No. 0 to No. 23.</p>
Full article ">Figure A1 Cont.
<p>Periodograms obtained from the Fourier analysis for models No. 0 to No. 23.</p>
Full article ">Figure A2
<p>Periodograms obtained from the Fourier analysis for models No. 24 to No. 32.</p>
Full article ">Figure A3
<p>Results of the wavelet analysis for models No. 0 to No. 5. From left to right the wavelet power spectrum and the average wavelet power for the first and second part of each synthetic curve.</p>
Full article ">Figure A4
<p>Results of the wavelet analysis for models No. 6 to No. 11. From left to right the wavelet power spectrum and the average wavelet power for the first and second part of each synthetic curve.</p>
Full article ">Figure A5
<p>Results of the wavelet analysis for models No. 12 to No. 17. From left to right the wavelet power spectrum and the average wavelet power for the first and second part of each synthetic curve.</p>
Full article ">Figure A6
<p>Results of the wavelet analysis for models No. 18 to No. 23. From left to right the wavelet power spectrum and the average wavelet power for the first and second part of each synthetic curve.</p>
Full article ">Figure A7
<p>Results of the wavelet analysis for models No. 24 to No. 29. From left to right the wavelet power spectrum and the average wavelet power for the first and second part of each synthetic curve.</p>
Full article ">Figure A8
<p>Results of the wavelet analysis for models No. 30 to No. 32. From left to right the wavelet power spectrum and the average wavelet power for the first and second part of each synthetic curve.</p>
Full article ">
25 pages, 3424 KiB  
Article
The C/M Ratio of AGB Stars in the Local Group Galaxies
by Tongtian Ren, Biwei Jiang, Yi Ren and Ming Yang
Universe 2022, 8(9), 465; https://doi.org/10.3390/universe8090465 - 7 Sep 2022
Cited by 5 | Viewed by 1638
Abstract
The number ratio of carbon-rich to oxygen-rich asymptotic giant branch (AGB) stars (the so-called C/M ratio) is closely related to the evolution environment of the host galaxy. This work studies the C/M ratio in 14 galaxies within the Local Group with the most [...] Read more.
The number ratio of carbon-rich to oxygen-rich asymptotic giant branch (AGB) stars (the so-called C/M ratio) is closely related to the evolution environment of the host galaxy. This work studies the C/M ratio in 14 galaxies within the Local Group with the most complete and clean sample of member stars identified in our previous works. The borderlines between carbon-rich AGB and oxygen-rich AGB stars as well as red supergiants are defined by Gaussian mixture model fitting to the number density in the (JK)/K diagram for the member stars of the LMC and M33, and then applied to the other galaxies by shifting the difference in the position of tip red giant branch (TRGB). The C/M ratios are obtained after precise and consistent categorization. Although for galaxies with larger distance modulo there is greater uncertainty, the C/M ratio is clearly found to decrease with the color index (JK)0 of TRGB as the indicator of metallicity, which agrees with previous studies and can be explained by the fact that carbon stars are more easily formed in a metal-poor environment. Furthermore, the C/M ratio within M33 is found to increase with galactocentric distance, which coincides with this scenario and the galactic chemical evolution model. On the other hand, the C/M ratio within M31 is found to decrease with galactocentric radius, which deserves further study. Full article
(This article belongs to the Section Solar and Stellar Physics)
Show Figures

Figure 1

Figure 1
<p>Determination of the borderlines between different types of stars in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>(</mo> <mi>J</mi> <mo>−</mo> <mi>K</mi> <mo>)</mo> </mrow> <mn>0</mn> </msub> <mo>/</mo> <msub> <mi>K</mi> <mn>0</mn> </msub> </mrow> </semantics></math> diagram in the LMC and M33. The boundaries (red dots) of each magnitude cut (red masks) calculated from GMM fitting (right panels), are used to fit the borderlines (dashed line). The dashed lines on either side of Line 3 represent the location of its 50% confidence interval. The insert panels show the reference for detecting the <span class="html-italic">K</span> magnitudes borderlines (Line 5) between O-AGBs and TP-AGBs with the Sobel filter.</p>
Full article ">Figure 1 Cont.
<p>Determination of the borderlines between different types of stars in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>(</mo> <mi>J</mi> <mo>−</mo> <mi>K</mi> <mo>)</mo> </mrow> <mn>0</mn> </msub> <mo>/</mo> <msub> <mi>K</mi> <mn>0</mn> </msub> </mrow> </semantics></math> diagram in the LMC and M33. The boundaries (red dots) of each magnitude cut (red masks) calculated from GMM fitting (right panels), are used to fit the borderlines (dashed line). The dashed lines on either side of Line 3 represent the location of its 50% confidence interval. The insert panels show the reference for detecting the <span class="html-italic">K</span> magnitudes borderlines (Line 5) between O-AGBs and TP-AGBs with the Sobel filter.</p>
Full article ">Figure 2
<p>The classification by shifting the boundary lines of the LMC and M33, the benchmark galaxies, to each galaxy according to their positions of TRGB. The C-AGBs and O-AGs are represented by red and blue dots, respectively. Note that 12 galaxies, except M31 and M33, have been extinction corrected.</p>
Full article ">Figure 3
<p>Dependence of <math display="inline"><semantics> <mrow> <mo form="prefix">lg</mo> <mo>(</mo> <mi mathvariant="normal">C</mi> <mo>/</mo> <mi mathvariant="normal">M</mi> <mo>)</mo> </mrow> </semantics></math> with <math display="inline"><semantics> <msubsup> <mrow> <mo>(</mo> <mi>J</mi> <mo>−</mo> <mi>K</mi> <mo>)</mo> </mrow> <mn>0</mn> <mi>TRGB</mi> </msubsup> </semantics></math> for the 12 galaxies. The red dashed line represents the linear fitting whose result with the Pearson’s correlation coefficient is displayed in the lower-left corner.</p>
Full article ">Figure 4
<p>Comparison of the borderlines with previous studies, where the red solid lines represent this work and the grey dashed lines represent the work by Cioni et al. [<a href="#B67-universe-08-00465" class="html-bibr">67</a>] for the LMC and by Paper I for M31 and M33.</p>
Full article ">Figure 5
<p>Verification of the classification areas in <a href="#universe-08-00465-f001" class="html-fig">Figure 1</a> with Padova stellar evolutionary models. The red and blue dots represent C-AGBs and O-AGBs at an age of 1.00 Gyr, respectively, and the yellow and green dots represent C-AGBs and O-AGBs at an age of 0.63 Gyr, respectively. The background is the sample stars in the LMC from Paper II.</p>
Full article ">Figure 6
<p>The previous spectroscopic classifications of AGB stars in the LMC (left) and NGC6822 (right) are used for the validation of our <math display="inline"><semantics> <mrow> <mi>J</mi> <mo>/</mo> <mi>J</mi> <mo>−</mo> <mi>K</mi> </mrow> </semantics></math> CMD photometric method. The C-AGBs stars and O-AGBs identified from previous spectral observations are indicated by red and blue dots, respectively. The black dashed line represents the borderlines in our work, and the green dashed line in the right panel represents the borderlines suggested by Kacharov et al. [<a href="#B78-universe-08-00465" class="html-bibr">78</a>] based on the result of the NGC 6822 AGB stars’ spectroscopic analysis.</p>
Full article ">Figure 7
<p>Comparison of the results of the LMC taking sloped TRGB and horizontal TRGB. The red line represents the sloped TRGB, and the white dots represent the saddle points we detected in different bins by the ridge method of number density analysis. The O-AGBs suggested by the sloped TRGB are represented by blue dots.</p>
Full article ">Figure 8
<p>The radial distribution of the C/M ratio, C-AGB stars and O-AGB stars in M31 (panel a,b,c), and in M33 (panel d,e,f). An exponential function is used to fit the radial distribution of C-AGBs and O-AGBs with the results displayed, where the region with uncertain photometry and locate in the ring area in M31 are not used and labelled by the dashed line. Besides, the 50% confidence limits of the C/M ratio are represented with dashed-dot lines on both sides of the C/M ratio median lines.</p>
Full article ">Figure 9
<p>The radial distribution of the mean K magnitude of the sources labelled as C-AGBs or O-AGBs in M31 (<b>upper</b>) and M33 (<b>bottom</b>).</p>
Full article ">
10 pages, 8583 KiB  
Article
Follow-Up of Extended Shells around B[e] Stars
by Tiina Liimets, Michaela Kraus, Alexei Moiseev, Nicolas Duronea, Lydia Sonia Cidale and Cecilia Fariña
Galaxies 2022, 10(2), 41; https://doi.org/10.3390/galaxies10020041 - 1 Mar 2022
Cited by 9 | Viewed by 2741
Abstract
B[e] stars are massive B type emission line stars in different evolutionary stages ranging from pre-main sequence to post-main sequence. Due to their mass loss and ejection events these objects deposit huge amounts of mass and energy into their environment and enrich it [...] Read more.
B[e] stars are massive B type emission line stars in different evolutionary stages ranging from pre-main sequence to post-main sequence. Due to their mass loss and ejection events these objects deposit huge amounts of mass and energy into their environment and enrich it with chemically processed material, contributing significantly to the chemical and dynamical evolution of their host galaxies. However, the large-scale environments of these enigmatic objects have not attracted much attention. The first and so far only catalog reporting the detection of extended shells around a sample of B[e] stars was an Hα imaging survey carried out in the year 2001, and was limited to bright targets in the northern hemisphere. We have recently started a follow-up of those targets to detect possible evolution of their nebulae in the plane of the sky over a baseline of two decades. Furthermore, we extend our survey to southern targets and fainter northern ones to complement and complete our knowledge on large-scale ejecta surrounding B[e] stars. Besides imaging in Hα and selected nebular lines, we utilize long-slit and 3D spectral observations across the nebulae to derive their physical properties. We discovered pronounced nebula structures around 15 more objects, resulting in a total of 27 B[e] stars with a large-scale nebula. Here we present our (preliminary) results for three selected objects: the two massive supergiants MWC137 and MWC 314, and the unclassified B[e] star MWC 819. Full article
(This article belongs to the Special Issue Asymmetric Planetary Nebulae 8e)
Show Figures

Figure 1

Figure 1
<p>Large-scale nebula of MWC 137. ALFOSC H<span class="html-italic">α</span> iamge with blue and the <span class="html-italic">Sptizer</span> 3.6 μm with red. FOV <math display="inline"><semantics> <mrow> <msup> <mn>2</mn> <mo>′</mo> </msup> <mo>×</mo> <msup> <mn>2</mn> <mo>′</mo> </msup> </mrow> </semantics></math>. Figure taken from [<a href="#B13-galaxies-10-00041" class="html-bibr">13</a>] (their Figure 10, <span class="html-italic">©</span> AAS, reproduced with permission).</p>
Full article ">Figure 2
<p>Radial velocity measurements of the MWC 137 from the FPI observations (<b>left</b>) and electron density distribution across the nebula obtained from the ratio of the [SII] <math display="inline"><semantics> <mrow> <mi>λ</mi> <mi>λ</mi> </mrow> </semantics></math> 6716,6731 lines measured in the NOT+ALFOSC long-slit spectra (<b>right</b>). On both panels north is up and east to the left. Figures taken from [<a href="#B14-galaxies-10-00041" class="html-bibr">14</a>] (their Figure 4a, left panel, and Figure 9, right panel, <span class="html-italic">©</span> AAS, reproduced with permission).</p>
Full article ">Figure 3
<p>H<span class="html-italic">α</span> image (INT) of the bipolar nebula around MWC 314. The pointings for the spectroscopic FPI observations are shown in red. See text for more details.</p>
Full article ">Figure 4
<p>Velocity field of the bipolar nebula around MWC 314 measured from the [SII] lines observed with the FPI.</p>
Full article ">Figure 5
<p>Nebular features of the B[e] star MWC 819 seen in 2001 from [<a href="#B10-galaxies-10-00041" class="html-bibr">10</a>] (<b>left</b>) and in our new, deeper image taken in 2020 (<b>right</b>). North is up, east is left.</p>
Full article ">
43 pages, 8180 KiB  
Review
Space-Based Photometry of Binary Stars: From Voyager to TESS
by John Southworth
Universe 2021, 7(10), 369; https://doi.org/10.3390/universe7100369 - 30 Sep 2021
Cited by 34 | Viewed by 5053
Abstract
Binary stars are crucial laboratories for stellar physics, so have been photometric targets for space missions beginning with the very first orbiting telescope (OAO-2) launched in 1968. This review traces the binary stars observed and the scientific results obtained from the early days [...] Read more.
Binary stars are crucial laboratories for stellar physics, so have been photometric targets for space missions beginning with the very first orbiting telescope (OAO-2) launched in 1968. This review traces the binary stars observed and the scientific results obtained from the early days of ultraviolet missions (OAO-2, Voyager, ANS, IUE), through a period of diversification (Hipparcos, WIRE, MOST, BRITE), to the current era of large planetary transit surveys (CoRoT, Kepler, TESS). In this time observations have been obtained of detached, semi-detached and contact binaries containing dwarfs, sub-giants, giants, supergiants, white dwarfs, planets, neutron stars and accretion discs. Recent missions have found a huge variety of objects such as pulsating stars in eclipsing binaries, multi-eclipsers, heartbeat stars and binaries hosting transiting planets. Particular attention is paid to eclipsing binaries, because they are staggeringly useful, and to the NASA Transiting Exoplanet Survey Satellite (TESS) because its huge sky coverage enables a wide range of scientific investigations with unprecedented ease. These results are placed into context, future missions are discussed, and a list of important science goals is presented. Full article
(This article belongs to the Section Solar and Stellar Physics)
Show Figures

Figure 1

Figure 1
<p>Comparison of the brightness variation of a transiting planet (<b>left</b>) and an EB (<b>right</b>), both observed by TESS in sector 2. WASP-4 was chosen as it shows one of the deepest planetary transits known. <math display="inline"><semantics> <mi>ζ</mi> </semantics></math> Phe has a typical brightness variation for an EB and was observed by TESS at the same time as WASP-4. The light curve of <math display="inline"><semantics> <mi>ζ</mi> </semantics></math> Phe (brightness <math display="inline"><semantics> <mrow> <mi>V</mi> <mo>=</mo> <mn>4.01</mn> </mrow> </semantics></math>) has been artificially degraded so its scatter matches that of the much fainter WASP-4 system (<math display="inline"><semantics> <mrow> <mi>V</mi> <mo>=</mo> <mn>12.46</mn> </mrow> </semantics></math>). The axes are the same on both plots. Only a small fraction of the light curves are shown.</p>
Full article ">Figure 2
<p>The OAO-2 satellite light curve of the EB VV Ori (<math display="inline"><semantics> <mrow> <mi>P</mi> <mo>=</mo> <mn>1.485</mn> </mrow> </semantics></math> d) in seven UV passbands (labelled), published by Eaton [<a href="#B42-universe-07-00369" class="html-bibr">42</a>].</p>
Full article ">Figure 3
<p>The <span class="html-italic">Hipparcos</span> satellite light curves of three example EBs. V455 Aur (<b>top</b>) was chosen as it was these data that enabled its discovery. <math display="inline"><semantics> <mi>ζ</mi> </semantics></math> Phe (<b>middle</b>) was selected for comparison with the TESS data in <a href="#universe-07-00369-f001" class="html-fig">Figure 1</a>. W UMa (<b>bottom</b>) was picked as it is the prototype eclipsing contact binary; the TESS data of this object are shown in Figure 9 for comparison.</p>
Full article ">Figure 4
<p>The WIRE satellite light curves of four EBs (labelled) plotted according to orbital phase.</p>
Full article ">Figure 5
<p>Parts of the <span class="html-italic">CoRoT</span> satellite light curves of four EBs (labelled).</p>
Full article ">Figure 6
<p>Example <span class="html-italic">Kepler</span> satellite light curves of a set of binary systems. The names and the <span class="html-italic">Kepler</span> quarters are labelled.</p>
Full article ">Figure 7
<p>Example TESS light curves of binary systems. The names and sectors are labelled. The data for <math display="inline"><semantics> <mi>α</mi> </semantics></math> Dra have been binned to decrease the size of the image file.</p>
Full article ">Figure 8
<p>Example TESS light curves of binary stars containing pulsating stars. The names and sectors are labelled.</p>
Full article ">Figure 9
<p>Example TESS light curves of semi-detached and contact systems. The <span class="html-italic">y</span>-axis for <math display="inline"><semantics> <mi>β</mi> </semantics></math> Lyr is not labelled as the eclipse depths in the TESS data are incorrect.</p>
Full article ">Figure 10
<p>Example TESS light curves of CVs. The magnitude scales are not reliable due to the faintness of these objects for TESS.</p>
Full article ">Figure 11
<p>Mass-radius diagram for EBs in the DEBCat catalogue [<a href="#B264-universe-07-00369" class="html-bibr">264</a>]. Results from space-based telescopes are colour-coded according to source. Errorbars are not plotted as they are almost all smaller than the point size.</p>
Full article ">Figure 12
<p>Hertzsprung-Russell diagram for EBs in the DEBCat catalogue [<a href="#B264-universe-07-00369" class="html-bibr">264</a>]. Results from space-based telescopes are colour-coded according to source.</p>
Full article ">
25 pages, 24668 KiB  
Article
Approximate Analytical Periodic Solutions to the Restricted Three-Body Problem with Perturbation, Oblateness, Radiation and Varying Mass
by Fabao Gao and Yongqing Wang
Universe 2020, 6(8), 110; https://doi.org/10.3390/universe6080110 - 4 Aug 2020
Cited by 13 | Viewed by 2630
Abstract
Against the background of a restricted three-body problem consisting of a supergiant eclipsing binary system, the two primaries are composed of a pair of bright oblate stars whose mass changes with time. The zero-velocity surface and curve of the problem are numerically studied [...] Read more.
Against the background of a restricted three-body problem consisting of a supergiant eclipsing binary system, the two primaries are composed of a pair of bright oblate stars whose mass changes with time. The zero-velocity surface and curve of the problem are numerically studied to describe the third body’s motion area, and the corresponding five libration points are obtained. Moreover, the effect of small perturbations, Coriolis and centrifugal forces, radiative pressure, and the oblateness and mass parameters of the two primaries on the third body’s dynamic behavior is discussed through the bifurcation diagram. Furthermore, the second- and third-order approximate analytical periodic solutions around the collinear solution point L3 in two-dimensional plane and three-dimensional spaces are presented by using the Lindstedt-Poincaré perturbation method. Full article
(This article belongs to the Section Planetary Sciences)
Show Figures

Figure 1

Figure 1
<p>Zero velocity surfaces for <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0.5</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> </mrow> </semantics></math> and 4, respectively.</p>
Full article ">Figure 2
<p>Zero-velocity surfaces of the third body for <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>4.16993</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>3.59308</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>3.57940</mn> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>2.82017</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Zero-velocity curves of the third body: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>4.16993</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>3.59308</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>3.57940</mn> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>2.82017</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) Bifurcation diagrams of Coriolis force <math display="inline"><semantics> <mi>α</mi> </semantics></math> and centrifugal force <math display="inline"><semantics> <mi>β</mi> </semantics></math> in the frames <math display="inline"><semantics> <mrow> <mi>α</mi> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>β</mi> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>α</mi> <mi>β</mi> <mi>u</mi> </mrow> </semantics></math>, respectively, when <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>μ</mi> <mi>m</mi> </msub> <mo>=</mo> <mn>0.48785</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mspace width="4pt"/> <msub> <mi>q</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.9988</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>q</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.9985</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.024</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>α</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.02</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) Bifurcation diagrams of oblateness coefficient of the <span class="html-italic">i</span>th (<math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> </mrow> </semantics></math>) primary in the frames <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>2</mn> </msub> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <msub> <mi>α</mi> <mn>2</mn> </msub> <mi>u</mi> </mrow> </semantics></math>, respectively, when <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi>m</mi> </msub> <mo>=</mo> <mn>0.48785</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.9988</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.9985</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1.001</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1.002</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) Bifurcation diagrams of the arbitrary sum <math display="inline"><semantics> <mi>κ</mi> </semantics></math> of the masses of the primaries in the frames <math display="inline"><semantics> <mrow> <mi>κ</mi> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>κ</mi> <mi>u</mi> <mi>w</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>κ</mi> <mi>v</mi> <mi>w</mi> </mrow> </semantics></math>, respectively, when <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi>m</mi> </msub> <mo>=</mo> <mn>0.48785</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <mspace width="4pt"/> <msub> <mi>q</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.9988</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>q</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.9985</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.024</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>α</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.02</mn> <mo>,</mo> <mspace width="4pt"/> <mi>α</mi> <mo>=</mo> <mn>1.001</mn> <mo>,</mo> <mspace width="4pt"/> <mi>β</mi> <mo>=</mo> <mn>1.002</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>c</b>) Bifurcation diagrams of the mass parameter <math display="inline"><semantics> <msub> <mi>μ</mi> <mi>m</mi> </msub> </semantics></math> in the frames <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi>m</mi> </msub> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi>m</mi> </msub> <mi>u</mi> <mi>w</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mi>m</mi> </msub> <mi>v</mi> <mi>w</mi> </mrow> </semantics></math>, respectively, when <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>q</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.9988</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>q</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.9985</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.024</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.02</mn> <mo>,</mo> <mspace width="4pt"/> <mi>α</mi> <mo>=</mo> <mn>1.001</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1.002</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>c</b>) Bifurcation diagrams of radiation factor of the <span class="html-italic">i</span>th (<math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mn>2</mn> </mrow> </semantics></math>) primary in the frames <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mn>1</mn> </msub> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mn>2</mn> </msub> <mi>u</mi> <mi>v</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>q</mi> <mn>1</mn> </msub> <msub> <mi>q</mi> <mn>2</mn> </msub> <mi>u</mi> </mrow> </semantics></math>, respectively, when <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> <mo>,</mo> <mspace width="4pt"/> <msub> <mi>μ</mi> <mi>m</mi> </msub> <mo>=</mo> <mn>0.48785</mn> <mo>,</mo> </mrow> </semantics></math><math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.024</mn> <mo>,</mo> <msub> <mi>α</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.02</mn> <mo>,</mo> <mspace width="4pt"/> <mi>α</mi> <mo>=</mo> <mn>1.001</mn> <mo>,</mo> <mspace width="4pt"/> <mi>β</mi> <mo>=</mo> <mn>1.002</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>The approximate periodic solutions in two-dimensional plane for (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, respectively.</p>
Full article ">Figure 10
<p>The approximate periodic solutions in three-dimensional space for (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math>, and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>, respectively.</p>
Full article ">
16 pages, 7299 KiB  
Review
From SN 2010da to NGC 300 ULX-1: Ten Years of Observations of an Unusual High Mass X-Ray Binary in NGC 300
by Breanna A. Binder, Stefania Carpano, Marianne Heida and Ryan Lau
Galaxies 2020, 8(1), 17; https://doi.org/10.3390/galaxies8010017 - 18 Feb 2020
Cited by 5 | Viewed by 5964
Abstract
In May 2010, an intermediate luminosity optical transient was discovered in the nearby galaxy NGC 300 by a South African amateur astronomer. In the decade since its discovery, multi-wavelength observations of the misnamed “SN 2010da” have continually reshaped our understanding of this high [...] Read more.
In May 2010, an intermediate luminosity optical transient was discovered in the nearby galaxy NGC 300 by a South African amateur astronomer. In the decade since its discovery, multi-wavelength observations of the misnamed “SN 2010da” have continually reshaped our understanding of this high mass X-ray binary system. In this review, we present an overview of the multi-wavelength observations and attempt to understand the 2010 transient event, and later, the reclassification of this system as NGC 300 ULX-1: a red supergiant + neutron star ultraluminous X-ray source. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Left</b>): The initial discovery image [<a href="#B8-galaxies-08-00017" class="html-bibr">8</a>] of SN 2010da by L.A.G. Monard [<a href="#B1-galaxies-08-00017" class="html-bibr">1</a>]. (<b>Right</b>): An RGB-rendered IR composite image of NGC 300, with insets showing the 3.6 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m image of the progenitor (∼6.5 years pre-outburst), ∼4.5 years post-outburst, and the difference subtracted image. Reproduced with permission from Lau et al. [<a href="#B9-galaxies-08-00017" class="html-bibr">9</a>] (their Figure 1).</p>
Full article ">Figure 2
<p>The mid-IR color–magnitude diagram (CMD) showing the SN 2010da progenitor (red square) and local luminous blue variables (LBVs) and LBV candidates (filled circles and triangles, respectively). Open circles and filled triangles show the CMD locations of LBV-like supernova (SN) “impostors.” Reproduced with permission from Khan et al. [<a href="#B3-galaxies-08-00017" class="html-bibr">3</a>].</p>
Full article ">Figure 3
<p>The IR spectral energy density (SED) of SN 2010da progenitor (black squares show measurements, black downward facing triangles show upper limits). The SEDs of two other optical progenitors, NGC 300 OT2008-1 and SN 2008S, are shown (in blue and brown, respectively) for comparison. The dotted red line shows the SED of the known LBV AG Car; the addition of 12 mag extinction is required for this SED to fit the observed fluxes of SN 2010da (dashed red line). Reproduced with permission from Berger and Chornock [<a href="#B19-galaxies-08-00017" class="html-bibr">19</a>].</p>
Full article ">Figure 4
<p><b>Left</b>: Adapted from Figure 1 in Binder et al. [<a href="#B23-galaxies-08-00017" class="html-bibr">23</a>], showing the <span class="html-italic">Chandra</span> image of SN 2010da. The yellow circle indicates the location of the X-ray source coincident with SN 2010da (the three fainter sources to the upper-left are unrelated). <b>Right</b>: Figure 2 from Binder et al. [<a href="#B23-galaxies-08-00017" class="html-bibr">23</a>] showing the <span class="html-italic">Hubble</span>/Advanced Camera for Surveys (ACS) image of the same region. The white cross and circle indicate the location of the X-ray source; the green cross and circle show the location of the likely massive donor star.</p>
Full article ">Figure 5
<p>Figure 1 from Carpano et al. [<a href="#B28-galaxies-08-00017" class="html-bibr">28</a>], showing the period evolution during the deep <span class="html-italic">XMM-Newton</span>/<span class="html-italic">NuSTAR</span> observations.</p>
Full article ">Figure 6
<p>The <span class="html-italic">XMM-Newton</span> and <span class="html-italic">NuSTAR</span> spectra of NGC 300 ULX-1 (top panel of Figure 3 in Carpano et al. [<a href="#B28-galaxies-08-00017" class="html-bibr">28</a>]). <span class="html-italic">XMM-Newton</span>/EPIC pn spectra are shown in black and blue, while the MOS spectra are shown in red, green, magenta and cyan. <span class="html-italic">NuSTAR</span> FPMA and FPMB spectra are shown in yellow and orange, respectively.</p>
Full article ">Figure 7
<p><b>Left</b>: <span class="html-italic">Swift</span>/XRT long term light curve from 2018 January to 2019 May. The corresponding hardness ratios (<span class="html-italic">S</span>: 0.2–1.5 keV, <span class="html-italic">H</span>: 1.5–10 keV) are shown in the bottom panel. <b>Right</b>: Evolution of the pulse period from <span class="html-italic">NICER</span> (pink) and <span class="html-italic">Swift</span> (blue) data covering the period from January 2018 to January 2019, when pulsations could be detected.</p>
Full article ">Figure 8
<p>(<b>Left</b>) <span class="html-italic">Spitzer</span>/IRAC mid-IR light curve of SN 2010da/NGC 300 ULX-1 from 2004–2019. The red and blue points correspond to photometry measured at 3.6 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and 4.5 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m, respectively. Only 3.6 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m observations were serendipitously obtained near the 2010 outburst. (<b>Right</b>) Spectral energy distribution (SED) of SN 2010da/NGC 300 ULX-1 during pre-outburst (December 2007) and post-outburst phases (late 2014/early 2015) overlaid on SED templates of 122 RSG and 11 sgB[e] stars in the Large Magellanic Cloud, cataloged by Bonanos et al. [<a href="#B55-galaxies-08-00017" class="html-bibr">55</a>]. Solid lines correspond to the median VIJHKs and <span class="html-italic">Spitzer</span>/IRAC magnitudes of the SED template stars, and the surrounding shaded regions indicate the 1<math display="inline"><semantics> <mi>σ</mi> </semantics></math> spread in the magnitudes of the distribution. The wavelength on the <span class="html-italic">x</span>-axis is shown in units of <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m. Both figures are reproduced and modified from Lau et al. [<a href="#B53-galaxies-08-00017" class="html-bibr">53</a>].</p>
Full article ">Figure 9
<p>Deep Xshooter spectroscopy of NGC 300 ULX-1 in October 2018 (black). The composite model (blue) requires a RSG atmosphere (red), excess dust emission (red short-dashed line) and a power-law blue excess (red long-dashed line) attributed to an irradiated accretion disk. Reproduced with permission from Heida et al. [<a href="#B57-galaxies-08-00017" class="html-bibr">57</a>].</p>
Full article ">Figure 10
<p>Mid-IR and X-ray light curve of SN 2010da/NGC 300 ULX-1 taken between MJD 57300 and MJD 58770 by <span class="html-italic">Spitzer</span>/IRAC and <span class="html-italic">Swift</span>/XRT, respectively. This figure is modified from Lau et al. [<a href="#B53-galaxies-08-00017" class="html-bibr">53</a>].</p>
Full article ">
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 4774
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 ">
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 4832
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 ">
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 4800
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 ">
14 pages, 1067 KiB  
Review
The Complex Upper HR Diagram
by Roberta M. Humphreys
Galaxies 2019, 7(3), 75; https://doi.org/10.3390/galaxies7030075 - 23 Aug 2019
Cited by 3 | Viewed by 3747
Abstract
Several decades of observations of the most massive and most luminous stars have revealed a complex upper HR Diagram, shaped by mass loss, and inhabited by a variety of evolved stars exhibiting the consequences of their mass loss histories. This introductory review presents [...] Read more.
Several decades of observations of the most massive and most luminous stars have revealed a complex upper HR Diagram, shaped by mass loss, and inhabited by a variety of evolved stars exhibiting the consequences of their mass loss histories. This introductory review presents a brief historical overview of the HR Diagram for massive stars, highlighting some of the primary discoveries and results from their observation in nearby galaxies. The sections in this volume include reviews of our current understanding of different groups of evolved massive stars, all losing mass and in different stages of their evolution: the Luminous Blue Variables (LBVs), B[e] supergiants, the warm hypergiants, Wolf–Rayet stars, and the population of OB stars and supergiants in the Magellanic Clouds. Full article
(This article belongs to the Special Issue Luminous Stars in Nearby Galaxies)
Show Figures

Figure 1

Figure 1
<p>The HR Diagram for the brightest stars in the LMC and SMC from [<a href="#B5-galaxies-07-00075" class="html-bibr">5</a>]. LMC stars are closed symbols and those in the SMC are shown as open symbols.</p>
Full article ">Figure 2
<p>The HR Diagram, M<math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>B</mi> <mi>o</mi> <mi>l</mi> </mrow> </msub> </semantics></math> vs. log T, for the luminous stars in the Milky Way from [<a href="#B25-galaxies-07-00075" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>The HR Diagram, M<math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>B</mi> <mi>o</mi> <mi>l</mi> </mrow> </msub> </semantics></math> vs. log T, for the luminous stars in the LMC from [<a href="#B25-galaxies-07-00075" class="html-bibr">25</a>].</p>
Full article ">Figure 4
<p>The schematic HR Diagrams for M31 and M33 showing the positions of the confirmed LBVs and candidate LBVs shown respectively, as filled and open blue circles, the warm hypergiants as green circles and the B[e] supergiants as orange circles. The LBV transits during their high mass loss state are shown as dashed blue lines. The LBV/S Dor instability strip is outlined in blue. The 15 and 20 M<math display="inline"><semantics> <msub> <mrow/> <mo>⊙</mo> </msub> </semantics></math> tracks are from [<a href="#B87-galaxies-07-00075" class="html-bibr">87</a>] with rotation are shown to provide a reference for the lower mass B[e]sgs, which are probable rotators. (Higher mass tracks are not shown due to crowding.) The supergiant population is shown in the background in light gray. Reproduced from [<a href="#B61-galaxies-07-00075" class="html-bibr">61</a>].</p>
Full article ">
Back to TopTop