generated from alexlyttle/uob-thesis-template
-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathlyttle21.bib
3251 lines (3015 loc) · 489 KB
/
lyttle21.bib
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
@article{Abadi.Barham.ea2016,
title = {{{TensorFlow}}: {{A}} System for Large-Scale Machine Learning},
shorttitle = {{{TensorFlow}}},
author = {Abadi, Mart{\'i}n and Barham, Paul and Chen, Jianmin and Chen, Zhifeng and Davis, Andy and Dean, Jeffrey and Devin, Matthieu and Ghemawat, Sanjay and Irving, Geoffrey and Isard, Michael and Kudlur, Manjunath and Levenberg, Josh and Monga, Rajat and Moore, Sherry and Murray, Derek G. and Steiner, Benoit and Tucker, Paul and Vasudevan, Vijay and Warden, Pete and Wicke, Martin and Yu, Yuan and Zheng, Xiaoqiang},
year = {2016},
month = may,
abstract = {TensorFlow is a machine learning system that operates at large scale and in heterogeneous environments. TensorFlow uses dataflow graphs to represent computation, shared state, and the operations that mutate that state. It maps the nodes of a dataflow graph across many machines in a cluster, and within a machine across multiple computational devices, including multicore CPUs, general-purpose GPUs, and custom designed ASICs known as Tensor Processing Units (TPUs). This architecture gives flexibility to the application developer: whereas in previous "parameter server" designs the management of shared state is built into the system, TensorFlow enables developers to experiment with novel optimizations and training algorithms. TensorFlow supports a variety of applications, with particularly strong support for training and inference on deep neural networks. Several Google services use TensorFlow in production, we have released it as an open-source project, and it has become widely used for machine learning research. In this paper, we describe the TensorFlow dataflow model in contrast to existing systems, and demonstrate the compelling performance that TensorFlow achieves for several real-world applications.},
archiveprefix = {arXiv},
eprint = {1605.08695},
journal = {ArXiv e-prints},
keywords = {and Cluster Computing,Computer Science - Artificial Intelligence,Computer Science - Distributed,Parallel},
primaryclass = {cs.DC}
}
@article{Addison.Wright.ea2021,
title = {{{TOI}}-257b ({{HD}} 19916b): A Warm Sub-Saturn Orbiting an Evolved {{F}}-Type Star},
shorttitle = {{{TOI}}-257b ({{HD}} 19916b)},
author = {Addison, Brett C. and Wright, Duncan J. and Nicholson, Belinda A. and Cale, Bryson and Mocnik, Teo and Huber, Daniel and Plavchan, Peter and Wittenmyer, Robert A. and Vanderburg, Andrew and Chaplin, William J. and Chontos, Ashley and Clark, Jake T. and Eastman, Jason D. and Ziegler, Carl and Brahm, Rafael and Carter, Bradley D. and Clerte, Mathieu and Espinoza, N{\'e}stor and Horner, Jonathan and Bentley, John and Jord{\'a}n, Andr{\'e}s and Kane, Stephen R. and Kielkopf, John F. and Laychock, Emilie and Mengel, Matthew W. and Okumura, Jack and Stassun, Keivan G. and Bedding, Timothy R. and Bowler, Brendan P. and Burnelis, Andrius and {Blanco-Cuaresma}, Sergi and Collins, Michaela and Crossfield, Ian and Davis, Allen B. and Evensberget, Dag and Heitzmann, Alexis and Howell, Steve B. and Law, Nicholas and Mann, Andrew W. and Marsden, Stephen C. and Matson, Rachel A. and O'Connor, James H. and Shporer, Avi and Stevens, Catherine and Tinney, C. G. and Tylor, Christopher and Wang, Songhu and Zhang, Hui and Henning, Thomas and Kossakowski, Diana and Ricker, George and Sarkis, Paula and Schlecker, Martin and Torres, Pascal and Vanderspek, Roland and Latham, David W. and Seager, Sara and Winn, Joshua N. and Jenkins, Jon M. and Mireles, Ismael and Rowden, Pam and Pepper, Joshua and Daylan, Tansu and Schlieder, Joshua E. and Collins, Karen A. and Collins, Kevin I. and Tan, Thiam-Guan and Ball, Warrick H. and Basu, Sarbani and Buzasi, Derek L. and Campante, Tiago L. and Corsaro, Enrico and {Gonz{\'a}lez-Cuesta}, L. and Davies, Guy R. and {de Almeida}, Leandro and {do Nascimento}, Jr., Jose-Dias and Garc{\'i}a, Rafael A. and Guo, Zhao and Handberg, Rasmus and Hekker, Saskia and Hey, Daniel R. and Kallinger, Thomas and Kawaler, Steven D. and Kayhan, Cenk and Kuszlewicz, James S. and Lund, Mikkel N. and Lyttle, Alexander and Mathur, Savita and Miglio, Andrea and Mosser, Benoit and Nielsen, Martin B. and Serenelli, Aldo M. and Aguirre, Victor Silva and Theme{\ss}l, Nathalie},
year = {2021},
month = apr,
volume = {502},
pages = {3704--3722},
issn = {0035-8711},
doi = {10.1093/mnras/staa3960},
abstract = {We report the discovery of a warm sub-Saturn, TOI-257b (HD 19916b), based on data from NASA's Transiting Exoplanet Survey Satellite (TESS). The transit signal was detected by TESS and confirmed to be of planetary origin based on radial velocity observations. An analysis of the TESS photometry, the MINERVA-Australis, FEROS, and HARPS radial velocities, and the asteroseismic data of the stellar oscillations reveals that TOI-257b has a mass of MP = 0.138 {$\pm$} 0.023 \$\textbackslash rm \{M\_J\}\$ (43.9 {$\pm$} 7.3 \$\textbackslash, M\_\{\textbackslash rm \textbackslash oplus\}\$ ), a radius of RP = 0.639 {$\pm$} 0.013 \$\textbackslash rm \{R\_J\}\$ (7.16 {$\pm$} 0.15 \$\textbackslash, \textbackslash mathrm\{ R\}\_\{\textbackslash rm \textbackslash oplus\}\$ ), bulk density of \$0.65\^\{+0.12\}\_\{-0.11\}\$ (cgs), and period \$18.38818\^\{+0.00085\}\_\{-0.00084\}\$ \$\textbackslash rm \{days\}\$ . TOI-257b orbits a bright (V = 7.612 mag) somewhat evolved late F-type star with M* = 1.390 {$\pm$} 0.046 \$\textbackslash rm \{M\_\{sun\}\}\$ , R* = 1.888 {$\pm$} 0.033 \$\textbackslash rm \{R\_\{sun\}\}\$ , Teff = 6075 {$\pm$} 90 \$\textbackslash rm \{K\}\$ , and vsin i = 11.3 {$\pm$} 0.5 km s-1. Additionally, we find hints for a second non-transiting sub-Saturn mass planet on a {$\sim$}71 day orbit using the radial velocity data. This system joins the ranks of a small number of exoplanet host stars ({$\sim$}100) that have been characterized with asteroseismology. Warm sub-Saturns are rare in the known sample of exoplanets, and thus the discovery of TOI-257b is important in the context of future work studying the formation and migration history of similar planetary systems.},
author+an = {87=highlight},
journal = {Monthly Notices of the Royal Astronomical Society},
keywords = {asteroseismology,planetary systems,stars: individual (TIC 200723869/TOI-257),techniques: photometric,techniques: radial velocities,techniques: spectroscopic}
}
@article{Ahumada.Prieto.ea2020,
title = {The 16th Data Release of the Sloan Digital Sky Surveys: {{First}} Release from the {{APOGEE}}-2 Southern Survey and Full Release of {{eBOSS}} Spectra},
author = {Ahumada, Romina and Prieto, Carlos Allende and Almeida, Andr{\'e}s and Anders, Friedrich and Anderson, Scott F. and Andrews, Brett H. and Anguiano, Borja and Arcodia, Riccardo and Armengaud, Eric and Aubert, Marie and Avila, Santiago and {Avila-Reese}, Vladimir and Badenes, Carles and Balland, Christophe and Barger, Kat and {Barrera-Ballesteros}, Jorge K. and Basu, Sarbani and Bautista, Julian and Beaton, Rachael L. and Beers, Timothy C. and Benavides, B. Izamar T. and Bender, Chad F. and Bernardi, Mariangela and Bershady, Matthew and Beutler, Florian and Bidin, Christian Moni and Bird, Jonathan and Bizyaev, Dmitry and Blanc, Guillermo A. and Blanton, Michael R. and Boquien, M{\'e}d{\'e}ric and Borissova, Jura and Bovy, Jo and Brandt, W. N. and Brinkmann, Jonathan and Brownstein, Joel R. and Bundy, Kevin and Bureau, Martin and Burgasser, Adam and Burtin, Etienne and {Cano-D{\'i}az}, Mariana and Capasso, Raffaella and Cappellari, Michele and Carrera, Ricardo and Chabanier, Sol{\`e}ne and Chaplin, William and Chapman, Michael and Cherinka, Brian and Chiappini, Cristina and Doohyun Choi, Peter and Chojnowski, S. Drew and Chung, Haeun and Clerc, Nicolas and Coffey, Damien and Comerford, Julia M. and Comparat, Johan and {da Costa}, Luiz and Cousinou, Marie-Claude and Covey, Kevin and Crane, Jeffrey D. and Cunha, Katia and Ilha, Gabriele da Silva and Dai, Yu Sophia and Damsted, Sanna B. and Darling, Jeremy and Davidson, James W., Jr. and Davies, Roger and Dawson, Kyle and De, Nikhil and {de la Macorra}, Axel and De Lee, Nathan and Queiroz, Anna B{\'a}rbara de Andrade and Deconto Machado, Alice and {de la Torre}, Sylvain and Dell'Agli, Flavia and {du Mas des Bourboux}, H{\'e}lion and {Diamond-Stanic}, Aleksandar M. and Dillon, Sean and Donor, John and Drory, Niv and Duckworth, Chris and Dwelly, Tom and Ebelke, Garrett and Eftekharzadeh, Sarah and Davis Eigenbrot, Arthur and Elsworth, Yvonne P. and Eracleous, Mike and Erfanianfar, Ghazaleh and Escoffier, Stephanie and Fan, Xiaohui and Farr, Emily and {Fern{\'a}ndez-Trincado}, Jos{\'e} G. and Feuillet, Diane and Finoguenov, Alexis and Fofie, Patricia and {Fraser-McKelvie}, Amelia and Frinchaboy, Peter M. and Fromenteau, Sebastien and Fu, Hai and Galbany, Llu{\'i}s and Garcia, Rafael A. and {Garc{\'i}a-Hern{\'a}ndez}, D. A. and Oehmichen, Luis Alberto Garma and Ge, Junqiang and Maia, Marcio Antonio Geimba and Geisler, Doug and Gelfand, Joseph and Goddy, Julian and {Gonzalez-Perez}, Violeta and Grabowski, Kathleen and Green, Paul and Grier, Catherine J. and Guo, Hong and Guy, Julien and Harding, Paul and Hasselquist, Sten and Hawken, Adam James and Hayes, Christian R. and Hearty, Fred and Hekker, S. and Hogg, David W. and Holtzman, Jon A. and Horta, Danny and Hou, Jiamin and Hsieh, Bau-Ching and Huber, Daniel and Hunt, Jason A. S. and Chitham, J. Ider and Imig, Julie and Jaber, Mariana and Angel, Camilo Eduardo Jimenez and Johnson, Jennifer A. and Jones, Amy M. and J{\"o}nsson, Henrik and Jullo, Eric and Kim, Yerim and Kinemuchi, Karen and Kirkpatrick, Charles C., IV and Kite, George W. and Klaene, Mark and Kneib, Jean-Paul and Kollmeier, Juna A. and Kong, Hui and Kounkel, Marina and Krishnarao, Dhanesh and Lacerna, Ivan and Lan, Ting-Wen and Lane, Richard R. and Law, David R. and Le Goff, Jean-Marc and Leung, Henry W. and Lewis, Hannah and Li, Cheng and Lian, Jianhui and Lin, Lihwai and Long, Dan and {Longa-Pe{\~n}a}, Pen{\'e}lope and Lundgren, Britt and Lyke, Brad W. and Ted Mackereth, J. and MacLeod, Chelsea L. and Majewski, Steven R. and Manchado, Arturo and Maraston, Claudia and Martini, Paul and Masseron, Thomas and Masters, Karen L. and Mathur, Savita and McDermid, Richard M. and Merloni, Andrea and Merrifield, Michael and M{\'e}sz{\'a}ros, Szabolcs and Miglio, Andrea and Minniti, Dante and Minsley, Rebecca and Miyaji, Takamitsu and Mohammad, Faizan Gohar and Mosser, Benoit and Mueller, Eva-Maria and Muna, Demitri and {Mu{\~n}oz-Guti{\'e}rrez}, Andrea and Myers, Adam D. and Nadathur, Seshadri and Nair, Preethi and Nandra, Kirpal and {do Nascimento}, Janaina Correa and Nevin, Rebecca Jean and Newman, Jeffrey A. and Nidever, David L. and Nitschelm, Christian and Noterdaeme, Pasquier and O'Connell, Julia E. and Olmstead, Matthew D. and Oravetz, Daniel and Oravetz, Audrey and Osorio, Yeisson and Pace, Zachary J. and Padilla, Nelson and {Palanque-Delabrouille}, Nathalie and Palicio, Pedro A. and Pan, Hsi-An and Pan, Kaike and Parker, James and Paviot, Romain and Peirani, Sebastien and Ram{\'r}ez, Karla Pe{\~n}a and Penny, Samantha and Percival, Will J. and {Perez-Fournon}, Ismael and {P{\'e}rez-R{\`a}fols}, Ignasi and Petitjean, Patrick and Pieri, Matthew M. and Pinsonneault, Marc and Poovelil, Vijith Jacob and Povick, Joshua Tyler and Prakash, Abhishek and {Price-Whelan}, Adrian M. and Raddick, M. Jordan and Raichoor, Anand and Ray, Amy and Rembold, Sandro Barboza and Rezaie, Mehdi and Riffel, Rogemar A. and Riffel, Rog{\'e}rio and Rix, Hans-Walter and Robin, Annie C. and {Roman-Lopes}, A. and {Rom{\'a}n-Z{\'u}{\~n}iga}, Carlos and Rose, Benjamin and Ross, Ashley J. and Rossi, Graziano and Rowlands, Kate and Rubin, Kate H. R. and Salvato, Mara and S{\'a}nchez, Ariel G. and {S{\'a}nchez-Menguiano}, Laura and {S{\'a}nchez-Gallego}, Jos{\'e} R. and Sayres, Conor and Schaefer, Adam and Schiavon, Ricardo P. and Schimoia, Jaderson S. and Schlafly, Edward and Schlegel, David and Schneider, Donald P. and Schultheis, Mathias and Schwope, Axel and Seo, Hee-Jong and Serenelli, Aldo and Shafieloo, Arman and Shamsi, Shoaib Jamal and Shao, Zhengyi and Shen, Shiyin and Shetrone, Matthew and Shirley, Raphael and Aguirre, V{\'i}ctor Silva and Simon, Joshua D. and Skrutskie, M. F. and Slosar, An{\v z}e and Smethurst, Rebecca and Sobeck, Jennifer and Sodi, Bernardo Cervantes and Souto, Diogo and Stark, David V. and Stassun, Keivan G. and Steinmetz, Matthias and Stello, Dennis and Stermer, Julianna and {Storchi-Bergmann}, Thaisa and Streblyanska, Alina and Stringfellow, Guy S. and Stutz, Amelia and Su{\'a}rez, Genaro and Sun, Jing and {Taghizadeh-Popp}, Manuchehr and Talbot, Michael S. and Tayar, Jamie and Thakar, Aniruddha R. and Theriault, Riley and Thomas, Daniel and Thomas, Zak C. and Tinker, Jeremy and Tojeiro, Rita and Toledo, Hector Hernandez and Tremonti, Christy A. and Troup, Nicholas W. and Tuttle, Sarah and {Unda-Sanzana}, Eduardo and Valentini, Marica and {Vargas-Gonz{\'a}lez}, Jaime and {Vargas-Maga{\~n}a}, Mariana and {V{\'a}zquez-Mata}, Jose Antonio and Vivek, M. and Wake, David and Wang, Yuting and Weaver, Benjamin Alan and Weijmans, Anne-Marie and Wild, Vivienne and Wilson, John C. and Wilson, Robert F. and Wolthuis, Nathan and {Wood-Vasey}, W. M. and Yan, Renbin and Yang, Meng and Y{\`e}che, Christophe and Zamora, Olga and Zarrouk, Pauline and Zasowski, Gail and Zhang, Kai and Zhao, Cheng and Zhao, Gongbo and Zheng, Zheng and Zheng, Zheng and Zhu, Guangtun and Zou, Hu},
year = {2020},
month = jul,
volume = {249},
pages = {3},
doi = {10.3847/1538-4365/ab929e},
abstract = {This paper documents the 16th data release (DR16) from the Sloan Digital Sky Surveys (SDSS), the fourth and penultimate from the fourth phase (SDSS-IV). This is the first release of data from the Southern Hemisphere survey of the Apache Point Observatory Galactic Evolution Experiment 2 (APOGEE-2); new data from APOGEE-2 North are also included. DR16 is also notable as the final data release for the main cosmological program of the Extended Baryon Oscillation Spectroscopic Survey (eBOSS), and all raw and reduced spectra from that project are released here. DR16 also includes all the data from the Time Domain Spectroscopic Survey and new data from the SPectroscopic IDentification of ERosita Survey programs, both of which were co-observed on eBOSS plates. DR16 has no new data from the Mapping Nearby Galaxies at Apache Point Observatory (MaNGA) survey (or the MaNGA Stellar Library ``MaStar''). We also preview future SDSS-V operations (due to start in 2020), and summarize plans for the final SDSS-IV data release (DR17).},
adsnote = {Provided by the SAO/NASA Astrophysics Data System},
adsurl = {https://ui.adsabs.harvard.edu/abs/2020ApJS..249....3A},
archiveprefix = {arXiv},
eid = {3},
eprint = {1912.02905},
eprinttype = {arxiv},
keywords = {1174,1378,1624,1630,2002,786,83,Astronomy databases,Astrophysics - Astrophysics of Galaxies,Astrophysics - Cosmology and Nongalactic Astrophysics,Astrophysics - Instrumentation and Methods for Astrophysics,Galactic abundances,Infrared astronomy,Optical telescopes,Redshift surveys,Stellar properties,Stellar spectral lines},
number = {1},
primaryclass = {astro-ph.GA}
}
@article{Albareti.AllendePrieto.ea2017,
title = {The 13th {{Data Release}} of the {{Sloan Digital Sky Survey}}: {{First Spectroscopic Data}} from the {{SDSS}}-{{IV Survey Mapping Nearby Galaxies}} at {{Apache Point Observatory}}},
shorttitle = {The 13th {{Data Release}} of the {{Sloan Digital Sky Survey}}},
author = {Albareti, Franco D. and Allende Prieto, Carlos and Almeida, Andres and Anders, Friedrich and Anderson, Scott and Andrews, Brett H. and {Arag{\'o}n-Salamanca}, Alfonso and {Argudo-Fern{\'a}ndez}, Maria and Armengaud, Eric and Aubourg, Eric and {Avila-Reese}, Vladimir and Badenes, Carles and Bailey, Stephen and Barbuy, Beatriz and Barger, Kat and {Barrera-Ballesteros}, Jorge and Bartosz, Curtis and Basu, Sarbani and Bates, Dominic and Battaglia, Giuseppina and Baumgarten, Falk and Baur, Julien and Bautista, Julian and Beers, Timothy C. and Belfiore, Francesco and Bershady, Matthew and {Bertran de Lis}, Sara and Bird, Jonathan C. and Bizyaev, Dmitry and Blanc, Guillermo A. and Blanton, Michael and Blomqvist, Michael and Bolton, Adam S. and Borissova, J. and Bovy, Jo and Brandt, William Nielsen and Brinkmann, Jonathan and Brownstein, Joel R. and Bundy, Kevin and Burtin, Etienne and Busca, Nicol{\'a}s G. and Orlando Camacho Chavez, Hugo and Cano D{\'i}az, M. and Cappellari, Michele and Carrera, Ricardo and Chen, Yanping and Cherinka, Brian and Cheung, Edmond and Chiappini, Cristina and Chojnowski, Drew and Chuang, Chia-Hsun and Chung, Haeun and Cirolini, Rafael Fernando and Clerc, Nicolas and Cohen, Roger E. and Comerford, Julia M. and Comparat, Johan and {Correa do Nascimento}, Janaina and Cousinou, Marie-Claude and Covey, Kevin and Crane, Jeffrey D. and Croft, Rupert and Cunha, Katia and Darling, Jeremy and Davidson, Jr., James W. and Dawson, Kyle and Da Costa, Luiz and Da Silva Ilha, Gabriele and Deconto Machado, Alice and Delubac, Timoth{\'e}e and De Lee, Nathan and {De la Macorra}, Axel and {De la Torre}, Sylvain and {Diamond-Stanic}, Aleksandar M. and Donor, John and Downes, Juan Jose and Drory, Niv and Du, Cheng and {Du Mas des Bourboux}, H{\'e}lion and Dwelly, Tom and Ebelke, Garrett and Eigenbrot, Arthur and Eisenstein, Daniel J. and Elsworth, Yvonne P. and Emsellem, Eric and Eracleous, Michael and Escoffier, Stephanie and Evans, Michael L. and {Falc{\'o}n-Barroso}, Jes{\'u}s and Fan, Xiaohui and Favole, Ginevra and {Fernandez-Alvar}, Emma and {Fernandez-Trincado}, J. G. and Feuillet, Diane and Fleming, Scott W. and {Font-Ribera}, Andreu and Freischlad, Gordon and Frinchaboy, Peter and Fu, Hai and Gao, Yang and Garcia, Rafael A. and {Garcia-Dias}, R. and {Garcia-Hern{\'a}ndez}, D. A. and Garcia P{\'e}rez, Ana E. and Gaulme, Patrick and Ge, Junqiang and Geisler, Douglas and Gillespie, Bruce and Gil Marin, Hector and Girardi, L{\'e}o and Goddard, Daniel and Gomez Maqueo Chew, Yilen and {Gonzalez-Perez}, Violeta and Grabowski, Kathleen and Green, Paul and Grier, Catherine J. and Grier, Thomas and Guo, Hong and Guy, Julien and Hagen, Alex and Hall, Matt and Harding, Paul and Harley, R. E. and Hasselquist, Sten and Hawley, Suzanne and Hayes, Christian R. and Hearty, Fred and Hekker, Saskia and Hernandez Toledo, Hector and Ho, Shirley and Hogg, David W. and {Holley-Bockelmann}, Kelly and Holtzman, Jon A. and Holzer, Parker H. and Hu, Jian and Huber, Daniel and Hutchinson, Timothy Alan and Hwang, Ho Seong and {Ibarra-Medel}, H{\'e}ctor J. and Ivans, Inese I. and Ivory, KeShawn and Jaehnig, Kurt and Jensen, Trey W. and Johnson, Jennifer A. and Jones, Amy and Jullo, Eric and Kallinger, T. and Kinemuchi, Karen and Kirkby, David and Klaene, Mark and Kneib, Jean-Paul and Kollmeier, Juna A. and Lacerna, Ivan and Lane, Richard R. and Lang, Dustin and Laurent, Pierre and Law, David R. and Leauthaud, Alexie and Le Goff, Jean-Marc and Li, Chen and Li, Cheng and Li, Niu and Li, Ran and Liang, Fu-Heng and Liang, Yu and Lima, Marcos and Lin, Lihwai and Lin, Lin and Lin, Yen-Ting and Liu, Chao and Long, Dan and Lucatello, Sara and MacDonald, Nicholas and MacLeod, Chelsea L. and Mackereth, J. Ted and Mahadevan, Suvrath and Maia, Marcio Antonio Geimba and Maiolino, Roberto and Majewski, Steven R. and Malanushenko, Olena and Malanushenko, Viktor and Mallmann, N{\'i}colas Dullius and Manchado, Arturo and Maraston, Claudia and {Marques-Chaves}, Rui and Martinez Valpuesta, Inma and Masters, Karen L. and Mathur, Savita and McGreer, Ian D. and Merloni, Andrea and Merrifield, Michael R. and M{\'e}sz{\'a}ros, Szabolcs and Meza, Andres and Miglio, Andrea and Minchev, Ivan and Molaverdikhani, Karan and {Montero-Dorta}, Antonio D. and Mosser, Benoit and Muna, Demitri and Myers, Adam and Nair, Preethi and Nandra, Kirpal and Ness, Melissa and Newman, Jeffrey A. and Nichol, Robert C. and Nidever, David L. and Nitschelm, Christian and O'Connell, Julia and Oravetz, Audrey and Oravetz, Daniel J. and Pace, Zachary and Padilla, Nelson and {Palanque-Delabrouille}, Nathalie and Pan, Kaike and Parejko, John and Paris, Isabelle and Park, Changbom and Peacock, John A. and Peirani, Sebastien and {Pellejero-Ibanez}, Marcos and Penny, Samantha and Percival, Will J. and Percival, Jeffrey W. and {Perez-Fournon}, Ismael and Petitjean, Patrick and Pieri, Matthew and Pinsonneault, Marc H. and Pisani, Alice and Prada, Francisco and Prakash, Abhishek and {Price-Jones}, Natalie and Raddick, M. Jordan and Rahman, Mubdi and Raichoor, Anand and Barboza Rembold, Sandro and Reyna, A. M. and Rich, James and Richstein, Hannah and Ridl, Jethro and Riffel, Rogemar A. and Riffel, Rog{\'e}rio and Rix, Hans-Walter and Robin, Annie C. and Rockosi, Constance M. and {Rodr{\'i}guez-Torres}, Sergio and Rodrigues, Tha{\'i}se S. and Roe, Natalie and Roman Lopes, A. and {Rom{\'a}n-Z{\'u}{\~n}iga}, Carlos and Ross, Ashley J. and Rossi, Graziano and Ruan, John and Ruggeri, Rossana and Runnoe, Jessie C. and {Salazar-Albornoz}, Salvador and Salvato, Mara and Sanchez, Sebastian F. and Sanchez, Ariel G. and {Sanchez-Gallego}, Jos{\'e} R. and Santiago, Bas{\'i}lio Xavier and Schiavon, Ricardo and Schimoia, Jaderson S. and Schlafly, Eddie and Schlegel, David J. and Schneider, Donald P. and Sch{\"o}nrich, Ralph and Schultheis, Mathias and Schwope, Axel and Seo, Hee-Jong and Serenelli, Aldo and Sesar, Branimir and Shao, Zhengyi and Shetrone, Matthew and Shull, Michael and Silva Aguirre, Victor and Skrutskie, M. F. and Slosar, An{\v z}e and Smith, Michael and Smith, Verne V. and Sobeck, Jennifer and Somers, Garrett and Souto, Diogo and Stark, David V. and Stassun, Keivan G. and Steinmetz, Matthias and Stello, Dennis and Storchi Bergmann, Thaisa and Strauss, Michael A. and Streblyanska, Alina and Stringfellow, Guy S. and Suarez, Genaro and Sun, Jing and {Taghizadeh-Popp}, Manuchehr and Tang, Baitian and Tao, Charling and Tayar, Jamie and Tembe, Mita and Thomas, Daniel and Tinker, Jeremy and Tojeiro, Rita and Tremonti, Christy and Troup, Nicholas and Trump, Jonathan R. and {Unda-Sanzana}, Eduardo and Valenzuela, O. and {Van den Bosch}, Remco and {Vargas-Maga{\~n}a}, Mariana and Vazquez, Jose Alberto and Villanova, Sandro and Vivek, M. and Vogt, Nicole and Wake, David and Walterbos, Rene and Wang, Yuting and Wang, Enci and Weaver, Benjamin Alan and Weijmans, Anne-Marie and Weinberg, David H. and Westfall, Kyle B. and Whelan, David G. and Wilcots, Eric and Wild, Vivienne and Williams, Rob A. and Wilson, John and {Wood-Vasey}, W. M. and Wylezalek, Dominika and Xiao, Ting and Yan, Renbin and Yang, Meng and Ybarra, Jason E. and Yeche, Christophe and Yuan, Fang-Ting and Zakamska, Nadia and Zamora, Olga and Zasowski, Gail and Zhang, Kai and Zhao, Cheng and Zhao, Gong-Bo and Zheng, Zheng and Zheng, Zheng and Zhou, Zhi-Min and Zhu, Guangtun and Zinn, Joel C. and Zou, Hu},
year = {2017},
month = dec,
volume = {233},
pages = {25},
doi = {10.3847/1538-4365/aa8992},
abstract = {The fourth generation of the Sloan Digital Sky Survey (SDSS-IV) began observations in 2014 July. It pursues three core programs: the Apache Point Observatory Galactic Evolution Experiment 2 (APOGEE-2), Mapping Nearby Galaxies at APO (MaNGA), and the Extended Baryon Oscillation Spectroscopic Survey (eBOSS). As well as its core program, eBOSS contains two major subprograms: the Time Domain Spectroscopic Survey (TDSS) and the SPectroscopic IDentification of ERosita Sources (SPIDERS). This paper describes the first data release from SDSS-IV, Data Release 13 (DR13). DR13 makes publicly available the first 1390 spatially resolved integral field unit observations of nearby galaxies from MaNGA. It includes new observations from eBOSS, completing the Sloan Extended QUasar, Emission-line galaxy, Luminous red galaxy Survey (SEQUELS), which also targeted variability-selected objects and X-ray-selected objects. DR13 includes new reductions of the SDSS-III BOSS data, improving the spectrophotometric calibration and redshift classification, and new reductions of the SDSS-III APOGEE-1 data, improving stellar parameters for dwarf stars and cooler stars. DR13 provides more robust and precise photometric calibrations. Value-added target catalogs relevant for eBOSS, TDSS, and SPIDERS and an updated red-clump catalog for APOGEE are also available. This paper describes the location and format of the data and provides references to important technical papers. The SDSS web site, http://www.sdss.org, provides links to the data, tutorials, examples of data access, and extensive documentation of the reduction and analysis procedures. DR13 is the first of a scheduled set that will contain new data and analyses from the planned {$\sim$}6 yr operations of SDSS-IV.},
journal = {ApJS},
keywords = {atlases,catalogs,surveys}
}
@article{Anderson.Hogg.ea2018,
ids = {Anderson.Hogg.ea2018a},
title = {Improving {{Gaia Parallax Precision}} with a {{Data}}-Driven {{Model}} of {{Stars}}},
author = {Anderson, Lauren and Hogg, David W. and Leistedt, Boris and {Price-Whelan}, Adrian M. and Bovy, Jo},
year = {2018},
month = oct,
volume = {156},
pages = {145},
issn = {0004-6256},
doi = {10.3847/1538-3881/aad7bf},
abstract = {Converting a noisy parallax measurement into a posterior belief over distance requires inference with a prior. Usually, this prior represents beliefs about the stellar density distribution of the Milky Way. However, multiband photometry exists for a large fraction of the Gaia-TGAS Catalog and is incredibly informative about stellar distances. Here, we use 2MASS colors for 1.4 million TGAS stars to build a noise-deconvolved empirical prior distribution for stars in color-magnitude space. This model contains no knowledge of stellar astrophysics or the Milky Way but is precise because it accurately generates a large number of noisy parallax measurements under an assumption of stationarity; that is, it is capable of combining the information from many stars. We use the Extreme Deconvolution (XD) algorithm\textemdash which is an empirical-Bayes approximation to a full-hierarchical model of the true parallax and photometry of every star\textemdash to construct this prior. The prior is combined with a TGAS likelihood to infer a precise photometric-parallax estimate and uncertainty (and full posterior) for every star. Our parallax estimates are more precise than the TGAS catalog entries by a median factor of 1.2 (14\% are more precise by a factor {$>$}2) and they are more precise than the previous Bayesian distance estimates that use spatial priors. We validate our parallax inferences using members of the Milky Way star cluster M67, which is not visible as a cluster in the TGAS parallax estimates but appears as a cluster in our posterior parallax estimates. Our results, including a parallax posterior probability distribution function for each of 1.4 million TGAS stars, are available in companion electronic tables.},
journal = {AJ},
keywords = {catalogs,Hertzsprung–Russell and C–M diagrams,methods: statistical,parallaxes}
}
@article{Ando.Osaki1975,
title = {Nonadiabatic Nonradial Oscillations - an Application to the Five-Minute Oscillation of the Sun},
author = {Ando, H. and Osaki, Y.},
year = {1975},
volume = {27},
pages = {581--603},
issn = {0004-6264},
abstract = {A solution is obtained for the equations of linear nonadiabatic nonradial oscillations for acoustic modes trapped in the solar convection zone on the basis of a solar envelope model. The growth or damping rates of oscillations are determined from the imaginary part of eigenfrequencies. The radiative transfer of perturbations is treated by the Eddington approximation which is applicable both for the optically thick and thin cases. Results of stability analysis are shown in the diagnostic diagram and compared with the observed power spectra of the five-minute oscillation. It is found that many of the acoustic modes are overstable mainly due to the so-called Kappa-mechanism of the hydrogen ionization zone, while atmospheric radiation loss plays an important part in the stability of higher overtones. Unstable modes occupy a long mountain-range-like region in the diagnostic diagram centered with a period of 300 sec and with a wide range of wavelength.},
journal = {PASJ},
keywords = {Acoustic Instability,Astrophysics,Eddington Approximation,Hydrogen Ions,Magnetoacoustic Waves,Nonadiabatic Theory,Nonstabilized Oscillation,Radiative Transfer,Solar Atmosphere,Stellar Envelopes}
}
@article{Appourchaux.Antia.ea2015,
title = {A Seismic and Gravitationally Bound Double Star Observed by {{Kepler}}. {{Implication}} for the Presence of a Convective Core},
author = {Appourchaux, T. and Antia, H. M. and Ball, W. and Creevey, O. and Lebreton, Y. and Verma, K. and Vorontsov, S. and Campante, T. L. and Davies, G. R. and Gaulme, P. and R{\'e}gulo, C. and Horch, E. and Howell, S. and Everett, M. and Ciardi, D. and Fossati, L. and Miglio, A. and Montalb{\'a}n, J. and Chaplin, W. J. and Garc{\'i}a, R. A. and Gizon, L.},
year = {2015},
month = oct,
volume = {582},
pages = {A25},
issn = {0004-6361},
doi = {10.1051/0004-6361/201526610},
abstract = {Context. Solar-like oscillations have been observed by Kepler and CoRoT in many solar-type stars, thereby providing a way to probe stars using asteroseismology. Aims: The derivation of stellar parameters has usually been done with single stars. The aim of the paper is to derive the stellar parameters of a double-star system (HIP 93511), for which an interferometric orbit has been observed along with asteroseismic measurements. Methods: We used a time series of nearly two years of data for the double star to detect the two oscillation-mode envelopes that appear in the power spectrum. Using a new scaling relation based on luminosity, we derived the radius and mass of each star. We derived the age of each star using two proxies: one based upon the large frequency separation and a new one based upon the small frequency separation. Using stellar modelling, the mode frequencies allowed us to derive the radius, the mass, and the age of each component. In addition, speckle interferometry performed since 2006 has enabled us to recover the orbit of the system and the total mass of the system. Results: From the determination of the orbit, the total mass of the system is 2.34-0.33+0.45 M{$\odot$}. The total seismic mass using scaling relations is 2.47 {$\pm$} 0.07 M{$\odot$}. The seismic age derived using the new proxy based upon the small frequency separation is 3.5 {$\pm$} 0.3 Gyr. Based on stellar modelling, the mean common age of the system is 2.7-3.9 Gyr. The mean total seismic mass of the system is 2.34-2.53 M{$\odot$} consistent with what we determined independently with the orbit. The stellar models provide the mean radius, mass, and age of the stars as RA = 1.82-1.87R{$\odot$}, MA = 1.25-1.39 M{$\odot$}, AgeA = 2.6-3.5 Gyr; RB = 1.22-1.25 R{$\odot$}, MB = 1.08-1.14 M{$\odot$}, AgeB = 3.35-4.21 Gyr. The models provide two sets of values for Star A: [1.25-1.27] M{$\odot$} and [1.34-1.39] M{$\odot$}. We detect a convective core in Star A, while Star B does not have any. For the metallicity of the binary system of Z {$\approx$} 0.02, we set the limit between stars having a convective core in the range [1.14-1.25] M{$\odot$}. Appendices are available in electronic form at http://www.aanda.org},
journal = {A\&A},
keywords = {asteroseismology,astrometry,binaries: general,stars: evolution,stars: solar-type}
}
@article{Appourchaux.Chaplin.ea2012,
title = {Oscillation Mode Frequencies of 61 Main-Sequence and Subgiant Stars Observed by {{Kepler}}},
author = {Appourchaux, T. and Chaplin, W. J. and Garc{\'i}a, R. A. and Gruberbauer, M. and Verner, G. A. and Antia, H. M. and Benomar, O. and Campante, T. L. and Davies, G. R. and Deheuvels, S. and Handberg, R. and Hekker, S. and Howe, R. and R{\'e}gulo, C. and Salabert, D. and Bedding, T. R. and White, T. R. and Ballot, J. and Mathur, S. and Silva Aguirre, V. and Elsworth, Y. P. and Basu, S. and Gilliland, R. L. and {Christensen-Dalsgaard}, J. and Kjeldsen, H. and Uddin, K. and Stumpe, M. C. and Barclay, T.},
year = {2012},
month = jul,
volume = {543},
pages = {A54},
issn = {0004-6361},
doi = {10.1051/0004-6361/201218948},
abstract = {Context. Solar-like oscillations have been observed by Kepler and CoRoT in several solar-type stars, thereby providing a way to probe the stars using asteroseismology Aims: We provide the mode frequencies of the oscillations of various stars required to perform a comparison with those obtained from stellar modelling. Methods: We used a time series of nine months of data for each star. The 61 stars observed were categorised in three groups: simple, F-like, and mixed-mode. The simple group includes stars for which the identification of the mode degree is obvious. The F-like group includes stars for which the identification of the degree is ambiguous. The mixed-mode group includes evolved stars for which the modes do not follow the asymptotic relation of low-degree frequencies. Following this categorisation, the power spectra of the 61 main-sequence and subgiant stars were analysed using both maximum likelihood estimators and Bayesian estimators, providing individual mode characteristics such as frequencies, linewidths, and mode heights. We developed and describe a methodology for extracting a single set of mode frequencies from multiple sets derived by different methods and individual scientists. We report on how one can assess the quality of the fitted parameters using the likelihood ratio test and the posterior probabilities. Results: We provide the mode frequencies of 61 stars (with their 1-{$\sigma$} error bars), as well as their associated \'echelle diagrams. Appendices are available in electronic form at http://www.aanda.org},
journal = {A\&A},
keywords = {asteroseismology,stars: oscillations,stars: solar-type}
}
@inproceedings{Asplund.Grevesse.ea2005,
title = {The Solar Chemical Composition},
booktitle = {Cosmic Abundances as Records of Stellar Evolution and Nucleosynthesis},
author = {Asplund, M. and Grevesse, N. and Sauval, A. J.},
editor = {Barnes, Thomas G., III and Bash, Frank N.},
year = {2005},
month = sep,
volume = {336},
pages = {25},
series = {Astronomical Society of the Pacific Conference Series}
}
@article{Asplund.Grevesse.ea2009,
title = {The {{Chemical Composition}} of the {{Sun}}},
author = {Asplund, Martin and Grevesse, Nicolas and Sauval, A. Jacques and Scott, Pat},
year = {2009},
month = sep,
volume = {47},
pages = {481--522},
issn = {0066-4146},
doi = {10.1146/annurev.astro.46.060407.145222},
abstract = {The solar chemical composition is an important ingredient in our understanding of the formation, structure, and evolution of both the Sun and our Solar System. Furthermore, it is an essential reference standard against which the elemental contents of other astronomical objects are compared. In this review, we evaluate the current understanding of the solar photospheric composition. In particular, we present a redetermination of the abundances of nearly all available elements, using a realistic new three-dimensional (3D), time-dependent hydrodynamical model of the solar atmosphere. We have carefully considered the atomic input data and selection of spectral lines, and accounted for departures from local thermodynamic equilibrium (LTE) whenever possible. The end result is a comprehensive and homogeneous compilation of the solar elemental abundances. Particularly noteworthy findings are significantly lower abundances of C, N, O, and Ne compared to the widely used values of a decade ago. The new solar chemical composition is supported by a high degree of internal consistency between available abundance indicators, and by agreement with values obtained in the Solar Neighborhood and from the most pristine meteorites. There is, however, a stark conflict with standard models of the solar interior according to helioseismology, a discrepancy that has yet to find a satisfactory resolution.},
journal = {ARA\&A}
}
@article{AstropyCollaboration.Price-Whelan.ea2018,
title = {The {{Astropy Project}}: {{Building}} an {{Open}}-Science {{Project}} and {{Status}} of the v2.0 {{Core Package}}},
shorttitle = {The {{Astropy Project}}},
author = {{Astropy Collaboration} and {Price-Whelan}, A. M. and Sip{\H o}cz, B. M. and G{\"u}nther, H. M. and Lim, P. L. and Crawford, S. M. and Conseil, S. and Shupe, D. L. and Craig, M. W. and Dencheva, N. and Ginsburg, A. and VanderPlas, J. T. and Bradley, L. D. and {P{\'e}rez-Su{\'a}rez}, D. and {de Val-Borro}, M. and Aldcroft, T. L. and Cruz, K. L. and Robitaille, T. P. and Tollerud, E. J. and Ardelean, C. and Babej, T. and Bach, Y. P. and Bachetti, M. and Bakanov, A. V. and Bamford, S. P. and Barentsen, G. and Barmby, P. and Baumbach, A. and Berry, K. L. and Biscani, F. and Boquien, M. and Bostroem, K. A. and Bouma, L. G. and Brammer, G. B. and Bray, E. M. and Breytenbach, H. and Buddelmeijer, H. and Burke, D. J. and Calderone, G. and Cano Rodr{\'i}guez, J. L. and Cara, M. and Cardoso, J. V. M. and Cheedella, S. and Copin, Y. and Corrales, L. and Crichton, D. and D'Avella, D. and Deil, C. and Depagne, {\'E}. and Dietrich, J. P. and Donath, A. and Droettboom, M. and Earl, N. and Erben, T. and Fabbro, S. and Ferreira, L. A. and Finethy, T. and Fox, R. T. and Garrison, L. H. and Gibbons, S. L. J. and Goldstein, D. A. and Gommers, R. and Greco, J. P. and Greenfield, P. and Groener, A. M. and Grollier, F. and Hagen, A. and Hirst, P. and Homeier, D. and Horton, A. J. and Hosseinzadeh, G. and Hu, L. and Hunkeler, J. S. and Ivezi{\'c}, {\v Z}. and Jain, A. and Jenness, T. and Kanarek, G. and Kendrew, S. and Kern, N. S. and Kerzendorf, W. E. and Khvalko, A. and King, J. and Kirkby, D. and Kulkarni, A. M. and Kumar, A. and Lee, A. and Lenz, D. and Littlefair, S. P. and Ma, Z. and Macleod, D. M. and Mastropietro, M. and McCully, C. and Montagnac, S. and Morris, B. M. and Mueller, M. and Mumford, S. J. and Muna, D. and Murphy, N. A. and Nelson, S. and Nguyen, G. H. and Ninan, J. P. and N{\"o}the, M. and Ogaz, S. and Oh, S. and Parejko, J. K. and Parley, N. and Pascual, S. and Patil, R. and Patil, A. A. and Plunkett, A. L. and Prochaska, J. X. and Rastogi, T. and Reddy Janga, V. and Sabater, J. and Sakurikar, P. and Seifert, M. and Sherbert, L. E. and {Sherwood-Taylor}, H. and Shih, A. Y. and Sick, J. and Silbiger, M. T. and Singanamalla, S. and Singer, L. P. and Sladen, P. H. and Sooley, K. A. and Sornarajah, S. and Streicher, O. and Teuben, P. and Thomas, S. W. and Tremblay, G. R. and Turner, J. E. H. and Terr{\'o}n, V. and {van Kerkwijk}, M. H. and {de la Vega}, A. and Watkins, L. L. and Weaver, B. A. and Whitmore, J. B. and Woillez, J. and Zabalza, V. and {Astropy Contributors}},
year = {2018},
month = sep,
volume = {156},
pages = {123},
issn = {0004-6256},
doi = {10.3847/1538-3881/aabc4f},
abstract = {The Astropy Project supports and fosters the development of open-source and openly developed Python packages that provide commonly needed functionality to the astronomical community. A key element of the Astropy Project is the core package astropy, which serves as the foundation for more specialized projects and packages. In this article, we provide an overview of the organization of the Astropy project and summarize key features in the core package, as of the recent major release, version 2.0. We then describe the project infrastructure designed to facilitate and support development for a broader ecosystem of interoperable packages. We conclude with a future outlook of planned new features and directions for the broader Astropy Project. .},
journal = {AJ},
keywords = {astronomy,methods: data analysis,methods: miscellaneous,methods: statistical,python,reference systems}
}
@article{AstropyCollaboration.Robitaille.ea2013,
title = {Astropy: {{A}} Community {{Python}} Package for Astronomy},
shorttitle = {Astropy},
author = {{Astropy Collaboration} and Robitaille, Thomas P. and Tollerud, Erik J. and Greenfield, Perry and Droettboom, Michael and Bray, Erik and Aldcroft, Tom and Davis, Matt and Ginsburg, Adam and {Price-Whelan}, Adrian M. and Kerzendorf, Wolfgang E. and Conley, Alexander and Crighton, Neil and Barbary, Kyle and Muna, Demitri and Ferguson, Henry and Grollier, Fr{\'e}d{\'e}ric and Parikh, Madhura M. and Nair, Prasanth H. and Unther, Hans M. and Deil, Christoph and Woillez, Julien and Conseil, Simon and Kramer, Roban and Turner, James E. H. and Singer, Leo and Fox, Ryan and Weaver, Benjamin A. and Zabalza, Victor and Edwards, Zachary I. and Azalee Bostroem, K. and Burke, D. J. and Casey, Andrew R. and Crawford, Steven M. and Dencheva, Nadia and Ely, Justin and Jenness, Tim and Labrie, Kathleen and Lim, Pey Lian and Pierfederici, Francesco and Pontzen, Andrew and Ptak, Andy and Refsdal, Brian and Servillat, Mathieu and Streicher, Ole},
year = {2013},
month = oct,
volume = {558},
pages = {A33},
issn = {0004-6361},
doi = {10.1051/0004-6361/201322068},
abstract = {We present the first public version (v0.2) of the open-source and community-developed Python package, Astropy. This package provides core astronomy-related functionality to the community, including support for domain-specific file formats such as flexible image transport system (FITS) files, Virtual Observatory (VO) tables, and common ASCII table formats, unit and physical quantity conversions, physical constants specific to astronomy, celestial coordinate and time transformations, world coordinate system (WCS) support, generalized containers for representing gridded as well as tabular data, and a framework for cosmological transformations and conversions. Significant functionality is under activedevelopment, such as a model fitting framework, VO client and server tools, and aperture and point spread function (PSF) photometry tools. The core development team is actively making additions and enhancements to the current code base, and we encourage anyone interested to participate in the development of future Astropy versions.},
journal = {A\&A},
keywords = {astronomy,methods: data analysis,methods: miscellaneous,python,virtual observatory tools}
}
@article{Aver.Olive.ea2015,
title = {The Effects of {{He I}} {$\Lambda$}10830 on Helium Abundance Determinations},
author = {Aver, Erik and Olive, Keith A. and Skillman, Evan D.},
year = {2015},
month = jul,
volume = {07},
pages = {011},
issn = {1475-7516},
doi = {10.1088/1475-7516/2015/07/011},
abstract = {Observations of helium and hydrogen emission lines from metal-poor extragalactic H II regions, combined with estimates of metallicity, provide an independent method for determining the primordial helium abundance, Yp. Traditionally, the emission lines employed are in the visible wavelength range, and the number of suitable lines is limited. Furthermore, when using these lines, large systematic uncertainties in helium abundance determinations arise due to the degeneracy of physical parameters, such as temperature and density. Recently, Izotov, Thuan, \& Guseva (2014) have pioneered adding the He I {$\lambda$}10830 infrared emission line in helium abundance determinations. The strong electron density dependence of He I {$\lambda$}10830 makes it ideal for better constraining density, potentially breaking the degeneracy with temperature. We revisit our analysis of the dataset published by Izotov, Thuan, \& Stasi\&aposnska (2007) and incorporate the newly available observations of He I {$\lambda$}10830 by scaling them using the observed-to-theoretical Paschen-gamma ratio. The solutions are better constrained, in particular for electron density, temperature, and the neutral hydrogen fraction, improving the model fit to data, with the result that more spectra now pass screening for quality and reliability, in addition to a standard 95\% confidence level cut. Furthermore, the addition of He I {$\lambda$}10830 decreases the uncertainty on the helium abundance for all galaxies, with reductions in the uncertainty ranging from 10-80\%. Overall, we find a reduction in the uncertainty on Yp by over 50\%. From a regression to zero metallicity, we determine Yp = 0.2449 {$\pm$} 0.0040, consistent with the BBN result, Yp = 0.2470 {$\pm$} 0.0002, based on the Planck determination of the baryon density. The dramatic improvement in the uncertainty from incorporating He I {$\lambda$}10830 strongly supports the case for simultaneous (thus not requiring scaling) observations of visible and infrared helium emission line spectra.},
journal = {J. Cosmol. Astropart. Phys.}
}
@inproceedings{Baglin.Auvergne.ea2006,
title = {{{CoRoT}}: A High Precision Photometer for Stellar Ecolution and Exoplanet Finding},
booktitle = {36th {{COSPAR}} Scientific Assembly},
author = {Baglin, A. and Auvergne, M. and Boisnard, L. and {Lam-Trong}, T. and Barge, P. and Catala, C. and Deleuil, M. and Michel, E. and Weiss, W.},
year = {2006},
month = jan,
volume = {36},
pages = {3749},
address = {{Beijing}, {China}}
}
@article{Bahcall.Pinsonneault.ea1995,
title = {Solar Models with Helium and Heavy-Element Diffusion},
author = {Bahcall, John N. and Pinsonneault, M. H. and Wasserburg, G. J.},
year = {1995},
month = oct,
volume = {67},
pages = {781--808},
issn = {0034-6861},
doi = {10.1103/RevModPhys.67.781},
abstract = {Helium and heavy-element diffusion are both included in precise calculations of solar models. In addition, improvements in the input data for solar interior models are described for nuclear reaction rates, the solar luminosity, the solar age, heavy-element abundances, radiative opacities, helium and metal diffusion rates, and neutrino interaction cross sections. The effects on the neutrino fluxes of each change in the input physics are evaluated separately by constructing a series of solar models with one additional improvement added at each stage. The effective 1 {$\sigma$} uncertainties in the individual input quantities are estimated and used to evaluate the uncertainties in the calculated neutrino fluxes and the calculated event rates for solar neutrino experiments. The calculated neutrino event rates, including all of the improvements, are 9.3+1.2-1.4 SNU for the 37Cl experiment and 137+8-7 SNU for the 71Ga experiments. The calculated flux of 7Be neutrinos is 5.1 (1.00+0.06-0.07)\texttimes 109 cm-2 s-1 and the flux of 8B neutrinos is 6.6(1.00+0.14-0.17)\texttimes 106 cm-2 s-1. The primordial helium abundance found for this model is Y=0.278. The present-day surface abundance of the model is Ys=0.247, in agreement with the helioseismological measurement of Ys=0.242+/-0.003 determined by Hernandez and Christensen-Dalsgaard (1994). The computed depth of the convective zone is R=0.712Rsolar, in agreement with the observed value determined from p-mode oscillation data of R=0.713+/-0.003Rsolar found by Christensen-Dalsgaard et al. (1991). Although the present results increase the predicted event rate in the four operating solar neutrino experiments by almost 1 {$\sigma$} (theoretical uncertainty), they only slightly increase the difficulty of explaining the existing experiments with standard physics (i.e., by assuming that nothing happens to the neutrinos after they are created in the center of the sun). For an extreme model in which all diffusion (helium and heavy-element diffusion) is neglected, the event rates are 7.0+0.9-1.0 SNU for the 37Cl experiment and 126+6-6 SNU for the 71Ga experiments, while the 7Be and 8B neutrino fluxes are, respectively, 4.5(1.00+0.06-0.07)\texttimes 109 cm-2 s-1 and 4.9(1.00+0.14-0.17)\texttimes 106 cm-2 s-1. For the no-diffusion model, the computed value of the depth of the convective zone is R=0.726Rsolar, which disagrees with the observed helioseismological value. The calculated surface abundance of helium, Ys=0.268, is also in disagreement with the p-mode measurement. The authors conclude that helioseismology provides strong evidence for element diffusion and therefore for the somewhat larger solar neutrino event rates calculated in this paper.},
journal = {Rev. Mod. Phys.}
}
@article{Bahcall.Ulrich1988,
title = {Solar Models, Neutrino Experiments, and Helioseismology},
author = {Bahcall, John N. and Ulrich, Roger K.},
year = {1988},
month = apr,
volume = {60},
pages = {297--372},
issn = {0034-6861},
doi = {10.1103/RevModPhys.60.297},
abstract = {The event rates and their recognized uncertainties are calculated for eleven solar neutrino experiments using accurate solar models. The same solar models are used to evaluate the frequency spectrum of the p and g oscillation modes of the Sun and to compare with existing observations. A numerical table of the characteristics of the standard solar model is presented. Improved values have been calculated for all of the neutrino absorption cross sections evaluating the uncertainties for each neutrino source and detector as well as the best estimates. The neutrino capture rate calculated from the standard solar model for the 37Cl experiment is (7.91{$\pm$}0.33) SNU, which spans the total theoretical range; the rate observed by Davis and his associates is (2.0{$\pm$}0.3) SNU. The various proposed solutions to the solar neutrino problem are reviewed. A solar model is constructed with a drastically altered nuclear energy generation that eliminates entirely the important high-energy 8B and 7Be neutrinos, but which affects by less than 0.01\% the calculated p-mode oscillation frequencies.},
journal = {Rev. Mod. Phys.},
keywords = {Absorption Cross Sections,Computational Astrophysics,Helioseismology,Monte Carlo Method,P Waves,Solar Neutrinos,Solar Spectra,Stellar Models,Sun}
}
@article{Ball.Gizon2014,
title = {A New Correction of Stellar Oscillation Frequencies for Near-Surface Effects},
author = {Ball, W. H. and Gizon, L.},
year = {2014},
month = aug,
volume = {568},
pages = {A123},
issn = {0004-6361},
doi = {10.1051/0004-6361/201424325},
abstract = {Context. Space-based observations of solar-like oscillations present an opportunity to constrain stellar models using individual mode frequencies. However, current stellar models are inaccurate near the surface, which introduces a systematic difference that must be corrected. Aims: We introduce and evaluate two parametrizations of the surface corrections based on formulae given by Gough (1990, LNP, 367, 283). The first we call a cubic term proportional to {$\nu$}3/ {$\mathscr{I}$} and the second has an additional inverse term proportional to {$\nu$}-1/ {$\mathscr{I}$}, where {$\nu$} and {$\mathscr{I}$} are the frequency and inertia of an oscillation mode. Methods: We first show that these formulae accurately correct model frequencies of two different solar models (Model S and a calibrated MESA model) when compared to observed BiSON frequencies. In particular, even the cubic form alone fits significantly better than a power law. We then incorporate the parametrizations into a modelling pipeline that simultaneously fits the surface effects and the underlying stellar model parameters. We apply this pipeline to synthetic observations of a Sun-like stellar model, solar observations degraded to typical asteroseismic uncertainties, and observations of the well-studied CoRoT target HD 52265. For comparison, we also run the pipeline with the scaled power-law correction proposed by Kjeldsen et al. (2008, ApJ, 683, L175). Results: The fits to synthetic and degraded solar data show that the method is unbiased and produces best-fit parameters that are consistent with the input models and known parameters of the Sun. Our results for HD 52265 are consistent with previous modelling efforts and the magnitude of the surface correction is similar to that of the Sun. The fit using a scaled power-law correction is significantly worse but yields consistent parameters, suggesting that HD 52265 is sufficiently Sun-like for the same power-law to be applicable. Conclusions: We find that the cubic term alone is suitable for asteroseismic applications and it is easy to implement in an existing pipeline. It reproduces the frequency dependence of the surface correction better than a power-law fit, both when comparing calibrated solar models to BiSON observations and when fitting stellar models using the individual frequencies. This parametrization is thus a useful new way to correct model frequencies so that observations of individual mode frequencies can be exploited.},
journal = {A\&A},
keywords = {asteroseismology,stars: individual: HD 52265,stars: oscillations}
}
@article{Ball.Gizon2017,
title = {Surface-Effect Corrections for Oscillation Frequencies of Evolved Stars},
author = {Ball, W. H. and Gizon, L.},
year = {2017},
month = apr,
volume = {600},
pages = {A128},
issn = {0004-6361},
doi = {10.1051/0004-6361/201630260},
abstract = {Context. Accurate modelling of solar-like oscillators requires that modelled mode frequencies are corrected for the systematic shift caused by improper modelling of the near-surface layers, known as the surface effect. Several parametrizations of the surface effect are now available but they have not yet been systematically compared with observations of stars showing modes with mixed g- and p-mode character. Aims: We investigate how much additional uncertainty is introduced to stellar model parameters by our uncertainty about the functional form of the surface effect. At the same time, we test whether any of the parametrizations is significantly better or worse at modelling observed subgiants and low-luminosity red giants. Methods: We model six stars observed by Kepler that show clear mixed modes. We fix the input physics of the stellar models and vary the choice of surface correction between five parametrizations. Results: Models using a solar-calibrated power law correction consistently fit the observations more poorly than the other four corrections. Models with the remaining four corrections generally fit the observations about equally well, with the combined surface correction by Ball \& Gizon perhaps being marginally superior. The fits broadly agree on the model parameters within about the 2{$\sigma$} uncertainties, with discrepancies between the modified Lorentzian and free power law corrections occasionally exceeding the 3{$\sigma$} level. Relative to the best-fitting values, the total uncertainties on the masses, radii and ages of the stars are all less than 2, 1 and 6 per cent, respectively. Conclusions: A solar-calibrated power law, as formulated by Kjeldsen et al., appears unsuitable for use with more evolved solar-like oscillators. Among the remaining surface corrections, the uncertainty in the model parameters introduced by the surface effects is about twice as large as the uncertainty in the individual fits for these six stars. Though the fits are thus somewhat less certain because of our uncertainty of how to manage the surface effect, these results also demonstrate that it is feasible to model the individual mode frequencies of subgiants and low-luminosity red giants, and hence also use these individual stars to help to constrain stellar models.},
journal = {A\&A},
keywords = {asteroseismology,stars: oscillations}
}
@article{Balser2006,
title = {The {{Chemical Evolution}} of {{Helium}}},
author = {Balser, Dana S.},
year = {2006},
month = dec,
volume = {132},
pages = {2326--2332},
issn = {0004-6256},
doi = {10.1086/508515},
abstract = {We report on measurements of the 4He abundance toward the outer Galaxy H II region S206 with the NRAO Green Bank Telescope. Observations of hydrogen and helium radio recombination lines between 8 and 10 GHz were made toward the peak radio continuum position in S206. We derive 4He/H=0.08459+/-0.00088 (random)+/-0.0010 (known systematic), 20\% lower than optical recombination line results. It is difficult to reconcile the large discrepancy between the optical and radio values even when accounting for temperature, density, and ionization structure or for optical extinction by dust. Using only M17 and S206 we determine {$\Delta$}Y/{$\Delta$}Z=1.41+/-0.62 in the Galaxy, consistent with standard chemical evolution models. High helium abundances in the old stellar population of elliptical galaxies can help explain the increase in UV emission with shorter wavelength between 2000 and 1200 \AA, called the ``UV upturn'' or UVX. Our lower values of {$\Delta$}Y/{$\Delta$}Z are consistent with a normal helium abundance at higher metallicity and suggest that other factors, such as a variable red giant branch mass loss with metallicity, may be important. When combined with 4He abundances in metal-poor galaxy H II regions, Magellanic Cloud H II regions, and M17 that have been determined from optical recombination lines, including the effects of temperature fluctuations, our radio 4He/H abundance ratio for S206 is consistent with a helium evolution of {$\Delta$}Y/{$\Delta$}Z=1.6. A linear extrapolation to zero metallicity predicts a 4He/H primordial abundance ratio about 5\% lower than that given by the Wilkinson Microwave Anisotropy Probe and standard big bang nucleosynthesis. The measured 4He abundances may be systematically underestimated by a few percent if clumping exists in these H II regions.},
journal = {AJ},
keywords = {ISM: Abundances,ISM: H II Regions,Radio Lines: ISM}
}
@article{Baraffe.Pratt.ea2017,
title = {Lithium {{Depletion}} in {{Solar}}-like {{Stars}}: {{Effect}} of {{Overshooting Based}} on {{Realistic Multi}}-Dimensional {{Simulations}}},
shorttitle = {Lithium {{Depletion}} in {{Solar}}-like {{Stars}}},
author = {Baraffe, I. and Pratt, J. and Goffrey, T. and Constantino, T. and Folini, D. and Popov, M. V. and Walder, R. and Viallet, M.},
year = {2017},
month = aug,
volume = {845},
pages = {L6},
doi = {10.3847/2041-8213/aa82ff},
abstract = {We study lithium depletion in low-mass and solar-like stars as a function of time, using a new diffusion coefficient describing extra-mixing taking place at the bottom of a convective envelope. This new form is motivated by multi-dimensional fully compressible, time-implicit hydrodynamic simulations performed with the MUSIC code. Intermittent convective mixing at the convective boundary in a star can be modeled using extreme value theory, a statistical analysis frequently used for finance, meteorology, and environmental science. In this Letter, we implement this statistical diffusion coefficient in a one-dimensional stellar evolution code, using parameters calibrated from multi-dimensional hydrodynamic simulations of a young low-mass star. We propose a new scenario that can explain observations of the surface abundance of lithium in the Sun and in clusters covering a wide range of ages, from {$\sim$}50 Myr to {$\sim$}4 Gyr. Because it relies on our physical model of convective penetration, this scenario has a limited number of assumptions. It can explain the observed trend between rotation and depletion, based on a single additional assumption, namely, that rotation affects the mixing efficiency at the convective boundary. We suggest the existence of a threshold in stellar rotation rate above which rotation strongly prevents the vertical penetration of plumes and below which rotation has small effects. In addition to providing a possible explanation for the long-standing problem of lithium depletion in pre-main-sequence and main-sequence stars, the strength of our scenario is that its basic assumptions can be tested by future hydrodynamic simulations.},
journal = {Astrophys. J. Lett.},
keywords = {convection,hydrodynamics,stars: evolution,stars: pre-main sequence,stars: rotation,stars: solar-type}
}
@article{Barber.Dobkin.ea1996,
title = {The Quickhull Algorithm for Convex Hulls},
author = {Barber, C. Bradford and Dobkin, David P. and Huhdanpaa, Hannu},
year = {1996},
month = dec,
volume = {22},
pages = {469--483},
issn = {0098-3500},
doi = {10.1145/235815.235821},
abstract = {The convex hull of a set of points is the smallest convex set that contains the points. This article presents a practical convex hull algorithm that combines the two-dimensional Quickhull algorithm with the general-dimension Beneath-Beyond Algorithm. It is similar to the randomized, incremental algorithms for convex hull and delaunay triangulation. We provide empirical evidence that the algorithm runs faster when the input contains nonextreme points and that it used less memory. computational geometry algorithms have traditionally assumed that input sets are well behaved. When an algorithm is implemented with floating-point arithmetic, this assumption can lead to serous errors. We briefly describe a solution to this problem when computing the convex hull in two, three, or four dimensions. The output is a set of ``thick'' facets that contain all possible exact convex hulls of the input. A variation is effective in five or more dimensions.},
journal = {ACM Trans. Math. Softw.},
keywords = {convex hull,Delaunay triangulation,halfspace intersection,Voronoi diagram},
number = {4}
}
@article{Basu.Antia1995,
title = {Helium Abundance in the Solar Envelope},
author = {Basu, Sarbani and Antia, H. M.},
year = {1995},
month = oct,
volume = {276},
pages = {1402--1408},
doi = {10.1093/mnras/276.4.1402},
abstract = {The abundance of helium in the solar envelope can be determined using the variation of the adiabatic index of the stellar material in the second helium ionization zone. All techniques for inferring helium abundance from the observed frequencies of solar p modes are known to be sensitive to the equation of state used in the reference models. The sensitivity of inferred helium abundance to the equation of state is studied by using different reference models with MHD and OPAL equations of state. Recent observations of high-degree solar p-mode frequencies yield a helium abundance Y=0.246 when determined using reference models with the MHD equation of state and Y=0.249 using the OPAL equation of state. Further, the models constructed with the OPAL equation of state are found to be in better agreement with the inferred sound speed below the HeII ionization zone.},
journal = {MNRAS},
keywords = {EQUATION OF STATE,SUN: ABUNDANCES,SUN: OSCILLATIONS}
}
@article{Basu.Antia2004,
title = {Constraining {{Solar Abundances Using Helioseismology}}},
author = {Basu, Sarbani and Antia, H. M.},
year = {2004},
month = may,
volume = {606},
pages = {L85-L88},
doi = {10.1086/421110},
abstract = {Recent analyses of solar photospheric abundances suggest that the oxygen abundance in the solar atmosphere needs to be revised downward. In this study, we investigate the consequence of this revision on helioseismic analyses of the depth of the solar convection zone and the helium abundance in the solar envelope and find no significant effect. We also find that the revised abundances along with the current OPAL opacity tables are not consistent with seismic data. A significant upward revision of the opacity tables is required to make solar models with lower oxygen abundance consistent with seismic observations.},
journal = {Astrophys. J. Lett.},
keywords = {Sun: Abundances,Sun: Interior,Sun: Oscillations}
}
@article{Basu.Mazumdar.ea2004,
title = {Asteroseismic Determination of Helium Abundance in Stellar Envelopes},
author = {Basu, Sarbani and Mazumdar, Anwesh and Antia, H. M. and Demarque, Pierre},
year = {2004},
month = may,
volume = {350},
pages = {277--286},
doi = {10.1111/j.1365-2966.2004.07644.x},
abstract = {Intermediate degree modes of the solar oscillations have previously been used to determine the solar helium abundance to a high degree of precision. However, we cannot expect to observe such modes in other stars. In this work we investigate whether low degree modes that should be available from space-based asteroseismology missions can be used to determine the helium abundance, Y, in stellar envelopes with sufficient precision. We find that the oscillatory signal in the frequencies caused by the depression in {$\Gamma$}1 in the second helium ionization zone can be used to determine the envelope helium abundance of low-mass main-sequence stars. For frequency errors of one part in 104, we expect errors {$\sigma$}Y in the estimated helium abundance to range from 0.03 for 0.8-Msolar stars to 0.01 for 1.2-Msolar stars. The task is more complicated in evolved stars, such as subgiants, but is still feasible if the relative errors in the frequencies are less than 10-4.},
journal = {MNRAS},
keywords = {stars: abundances,stars: oscillations}
}
@article{Bellinger.Angelou.ea2016,
title = {Fundamental {{Parameters}} of {{Main}}-{{Sequence Stars}} in an {{Instant}} with {{Machine Learning}}},
author = {Bellinger, Earl P. and Angelou, George C. and Hekker, Saskia and Basu, Sarbani and Ball, Warrick H. and Guggenberger, Elisabeth},
year = {2016},
month = oct,
volume = {830},
pages = {31},
doi = {10.3847/0004-637X/830/1/31},
abstract = {Owing to the remarkable photometric precision of space observatories like Kepler, stellar and planetary systems beyond our own are now being characterized en masse for the first time. These characterizations are pivotal for endeavors such as searching for Earth-like planets and solar twins, understanding the mechanisms that govern stellar evolution, and tracing the dynamics of our Galaxy. The volume of data that is becoming available, however, brings with it the need to process this information accurately and rapidly. While existing methods can constrain fundamental stellar parameters such as ages, masses, and radii from these observations, they require substantial computational effort to do so. We develop a method based on machine learning for rapidly estimating fundamental parameters of main-sequence solar-like stars from classical and asteroseismic observations. We first demonstrate this method on a hare-and-hound exercise and then apply it to the Sun, 16 Cyg A and B, and 34 planet-hosting candidates that have been observed by the Kepler spacecraft. We find that our estimates and their associated uncertainties are comparable to the results of other methods, but with the additional benefit of being able to explore many more stellar parameters while using much less computation time. We furthermore use this method to present evidence for an empirical diffusion-mass relation. Our method is open source and freely available for the community to use.6},
journal = {ApJ},
keywords = {methods: statistical,stars: abundances,stars: fundamental parameters,stars: low-mass,stars: oscillations,stars: solar-type}
}
@article{Bellinger.Hekker.ea2019,
ids = {Bellinger.Hekker.ea2019a},
title = {Stellar Ages, Masses, and Radii from Asteroseismic Modeling Are Robust to Systematic Errors in Spectroscopy},
author = {Bellinger, E. P. and Hekker, S. and Angelou, G. C. and Stokholm, A. and Basu, S.},
year = {2019},
month = feb,
volume = {622},
pages = {A130},
issn = {0004-6361, 1432-0746},
doi = {10.1051/0004-6361/201834461},
abstract = {Methods. We used the Stellar Parameters in an Instant (SPI) pipeline to estimate the parameters of nearly 100 stars observed by Kepler and Gaia, many of which are confirmed planet hosts. We adjusted the reported spectroscopic measurements of these stars by introducing faux systematic errors and, separately, artificially increasing the reported uncertainties of the measurements, and quantified the differences in the resulting parameters. Results. We find that a systematic error of 0.1 dex in [Fe/H] translates to differences of only 4\%, 2\%, and 1\% on average in the resulting stellar ages, masses, and radii, which are well within their uncertainties ({$\sim$}11\%, 3.5\%, 1.4\%) as derived by SPI. We also find that increasing the uncertainty of [Fe/H] measurements by 0.1 dex increases the uncertainties of the ages, masses, and radii by only 0.01 Gyr, 0.02 M , and 0.01 R , which are again well below their reported uncertainties ({$\sim$}0.5 Gyr, 0.04 M , 0.02 R ). The results for Teff at 100 K are similar. Conclusions. Stellar parameters from SPI are unchanged within uncertainties by errors of up to 0.14 dex or 175 K. They are even more robust to errors in Teff than the seismic scaling relations. Consequently, the parameters for their exoplanets are also robust.},
journal = {A\&A},
keywords = {asteroseismology,exoplanets,planets and satellites: fundamental parameters,spectroscopy,stars: abundances,stars: evolution,stars: low-mass,stars: oscillations},
language = {en}
}
@article{Bennett.Larson.ea2013,
title = {Nine-Year {{Wilkinson Microwave Anisotropy Probe}} ({{WMAP}}) {{Observations}}: {{Final Maps}} and {{Results}}},
shorttitle = {Nine-Year {{Wilkinson Microwave Anisotropy Probe}} ({{WMAP}}) {{Observations}}},
author = {Bennett, C. L. and Larson, D. and Weiland, J. L. and Jarosik, N. and Hinshaw, G. and Odegard, N. and Smith, K. M. and Hill, R. S. and Gold, B. and Halpern, M. and Komatsu, E. and Nolta, M. R. and Page, L. and Spergel, D. N. and Wollack, E. and Dunkley, J. and Kogut, A. and Limon, M. and Meyer, S. S. and Tucker, G. S. and Wright, E. L.},
year = {2013},
month = oct,
volume = {208},
pages = {20},
doi = {10.1088/0067-0049/208/2/20},
abstract = {We present the final nine-year maps and basic results from the Wilkinson Microwave Anisotropy Probe (WMAP) mission. The full nine-year analysis of the time-ordered data provides updated characterizations and calibrations of the experiment. We also provide new nine-year full sky temperature maps that were processed to reduce the asymmetry of the effective beams. Temperature and polarization sky maps are examined to separate cosmic microwave background (CMB) anisotropy from foreground emission, and both types of signals are analyzed in detail. We provide new point source catalogs as well as new diffuse and point source foreground masks. An updated template-removal process is used for cosmological analysis; new foreground fits are performed, and new foreground-reduced CMB maps are presented. We now implement an optimal C -1 weighting to compute the temperature angular power spectrum. The WMAP mission has resulted in a highly constrained {$\Lambda$}CDM cosmological model with precise and accurate parameters in agreement with a host of other cosmological measurements. When WMAP data are combined with finer scale CMB, baryon acoustic oscillation, and Hubble constant measurements, we find that big bang nucleosynthesis is well supported and there is no compelling evidence for a non-standard number of neutrino species (N eff = 3.84 {$\pm$} 0.40). The model fit also implies that the age of the universe is t 0 = 13.772 {$\pm$} 0.059 Gyr, and the fit Hubble constant is H 0 = 69.32 {$\pm$} 0.80 km s-1 Mpc-1. Inflation is also supported: the fluctuations are adiabatic, with Gaussian random phases; the detection of a deviation of the scalar spectral index from unity, reported earlier by the WMAP team, now has high statistical significance (ns = 0.9608 {$\pm$} 0.0080); and the universe is close to flat/Euclidean (\textbackslash Omega \_k = -0.0027\^\{+ 0.0039\}\_\{-0.0038\}). Overall, the WMAP mission has resulted in a reduction of the cosmological parameter volume by a factor of 68,000 for the standard six-parameter {$\Lambda$}CDM model, based on CMB data alone. For a model including tensors, the allowed seven-parameter volume has been reduced by a factor 117,000. Other cosmological observations are in accord with the CMB predictions, and the combined data reduces the cosmological parameter volume even further. With no significant anomalies and an adequate goodness of fit, the inflationary flat {$\Lambda$}CDM model and its precise and accurate parameters rooted in WMAP data stands as the standard model of cosmology.},
journal = {ApJS},
keywords = {cosmic background radiation,cosmology: observations,dark matter,early universe,instrumentation: detectors,space vehicles,space vehicles: instruments,telescopes}
}
@article{Berger.Huber.ea2018,
title = {Revised {{Radii}} of {{{\emph{Kepler}}}} {{Stars}} and {{Planets Using}} {{{\emph{Gaia}}}} {{Data Release}} 2},
author = {Berger, Travis A. and Huber, Daniel and Gaidos, Eric and {van Saders}, Jennifer L.},
year = {2018},
month = oct,
volume = {866},
pages = {99},
issn = {1538-4357},
doi = {10.3847/1538-4357/aada83},
abstract = {One bottleneck for the exploitation of data from the Kepler mission for stellar astrophysics and exoplanet research has been the lack of precise radii and evolutionary states for most of the observed stars. We report revised radii of 177,911 Kepler stars derived by combining parallaxes from the Gaia Data Release 2 with the DR25 Kepler Stellar Properties Catalog. The median radius precision is {$\approx$}8\%, a typical improvement by a factor of 4\textendash 5 over previous estimates for typical Kepler stars. We find that {$\approx$}67\% ({$\approx$}120,000) of all Kepler targets are main-sequence stars, {$\approx$}21\% ({$\approx$}37,000) are subgiants, and {$\approx$}12\% ({$\approx$}21,000) are red giants, demonstrating that subgiant contamination is less severe than some previous estimates and that Kepler targets are mostly main-sequence stars. Using the revised stellar radii, we recalculate the radii for 2123 confirmed and 1922 candidate exoplanets. We confirm the presence of a gap in the radius distribution of small, close-in planets, but find that the gap is mostly limited to incident fluxes {$>$}200 F\AA, and its location may be at a slightly larger radius (closer to {$\approx$}2 R{$\oplus$}) when compared to previous results. Furthermore, we find several confirmed exoplanets occupying a previously described ``hot super-Earth desert'' at high irradiance, show the relation between a gas-giant planet's radius and its incident flux, and establish a bona fide sample of eight confirmed planets and 30 planet candidates with Rp {$<$} 2 R{$\oplus$} in circumstellar ``habitable zones'' (incident fluxes between 0.25 and 1.50 F\AA ). The results presented here demonstrate the potential for transformative characterization of stellar and exoplanet populations using Gaia data.},
journal = {ApJ},
keywords = {distances,fundamental parameters,Gaia,Kepler,parallax},
language = {en},
number = {2}
}
@article{Berger.Huber.ea2020,
title = {The {{Gaia}}-{{Kepler Stellar Properties Catalog}}. {{I}}. {{Homogeneous Fundamental Properties}} for 186,301 {{Kepler Stars}}},
author = {Berger, Travis A. and Huber, Daniel and {van Saders}, Jennifer L. and Gaidos, Eric and Tayar, Jamie and Kraus, Adam L.},
year = {2020},
month = jun,
volume = {159},
pages = {280},
doi = {10.3847/1538-3881/159/6/280},
abstract = {An accurate and precise Kepler Stellar Properties Catalog is essential for the interpretation of the Kepler exoplanet survey results. Previous Kepler Stellar Properties Catalogs have focused on reporting the best-available parameters for each star, but this has required combining data from a variety of heterogeneous sources. We present the Gaia-Kepler Stellar Properties Catalog, a set of stellar properties of 186,301 Kepler stars, homogeneously derived from isochrones and broadband photometry, Gaia Data Release 2 parallaxes, and spectroscopic metallicities, where available. Our photometric effective temperatures, derived from \$g\textbackslash,\textbackslash mathrm\{to\}\textbackslash,\{K\}\_\{s\}\$ colors, are calibrated on stars with interferometric angular diameters. Median catalog uncertainties are 112 K for \$\{T\}\_\{\textbackslash mathrm\{eff\}\}\$ , 0.05 dex for \$\textbackslash mathrm\{log\}\textbackslash,g\$ , 4\% for \$\{R\}\_\{\textbackslash star \}\$ , 7\% for \$\{M\}\_\{\textbackslash star \}\$ , 13\% for \$\{\textbackslash rho \}\_\{\textbackslash star \}\$ , 10\% for \$\{L\}\_\{\textbackslash star \}\$ , and 56\% for stellar age. These precise constraints on stellar properties for this sample of stars will allow unprecedented investigations into trends in stellar and exoplanet properties as a function of stellar mass and age. In addition, our homogeneous parameter determinations will permit more accurate calculations of planet occurrence and trends with stellar properties.},
journal = {AJ},
keywords = {205,484,555,Catalogs,Exoplanet systems,Fundamental parameters of stars}
}
@article{Blanton.Bershady.ea2017,
title = {Sloan {{Digital Sky Survey IV}}: {{Mapping}} the {{Milky Way}}, {{Nearby Galaxies}}, and the {{Distant Universe}}},
shorttitle = {Sloan {{Digital Sky Survey IV}}},
author = {Blanton, Michael R. and Bershady, Matthew A. and Abolfathi, Bela and Albareti, Franco D. and Allende Prieto, Carlos and Almeida, Andres and {Alonso-Garc{\'i}a}, Javier and Anders, Friedrich and Anderson, Scott F. and Andrews, Brett and {Aquino-Ort{\'i}z}, Erik and {Arag{\'o}n-Salamanca}, Alfonso and {Argudo-Fern{\'a}ndez}, Maria and Armengaud, Eric and Aubourg, Eric and {Avila-Reese}, Vladimir and Badenes, Carles and Bailey, Stephen and Barger, Kathleen A. and {Barrera-Ballesteros}, Jorge and Bartosz, Curtis and Bates, Dominic and Baumgarten, Falk and Bautista, Julian and Beaton, Rachael and Beers, Timothy C. and Belfiore, Francesco and Bender, Chad F. and Berlind, Andreas A. and Bernardi, Mariangela and Beutler, Florian and Bird, Jonathan C. and Bizyaev, Dmitry and Blanc, Guillermo A. and Blomqvist, Michael and Bolton, Adam S. and Boquien, M{\'e}d{\'e}ric and Borissova, Jura and {van den Bosch}, Remco and Bovy, Jo and Brandt, William N. and Brinkmann, Jonathan and Brownstein, Joel R. and Bundy, Kevin and Burgasser, Adam J. and Burtin, Etienne and Busca, Nicol{\'a}s G. and Cappellari, Michele and Delgado Carigi, Maria Leticia and Carlberg, Joleen K. and Carnero Rosell, Aurelio and Carrera, Ricardo and Chanover, Nancy J. and Cherinka, Brian and Cheung, Edmond and G{\'o}mez Maqueo Chew, Yilen and Chiappini, Cristina and Choi, Peter Doohyun and Chojnowski, Drew and Chuang, Chia-Hsun and Chung, Haeun and Cirolini, Rafael Fernando and Clerc, Nicolas and Cohen, Roger E. and Comparat, Johan and {da Costa}, Luiz and Cousinou, Marie-Claude and Covey, Kevin and Crane, Jeffrey D. and Croft, Rupert A. C. and {Cruz-Gonzalez}, Irene and Garrido Cuadra, Daniel and Cunha, Katia and Damke, Guillermo J. and Darling, Jeremy and Davies, Roger and Dawson, Kyle and {de la Macorra}, Axel and Dell'Agli, Flavia and De Lee, Nathan and Delubac, Timoth{\'e}e and Di Mille, Francesco and {Diamond-Stanic}, Aleks and {Cano-D{\'i}az}, Mariana and Donor, John and Downes, Juan Jos{\'e} and Drory, Niv and {du Mas des Bourboux}, H{\'e}lion and Duckworth, Christopher J. and Dwelly, Tom and Dyer, Jamie and Ebelke, Garrett and Eigenbrot, Arthur D. and Eisenstein, Daniel J. and Emsellem, Eric and Eracleous, Mike and Escoffier, Stephanie and Evans, Michael L. and Fan, Xiaohui and {Fern{\'a}ndez-Alvar}, Emma and {Fernandez-Trincado}, J. G. and Feuillet, Diane K. and Finoguenov, Alexis and Fleming, Scott W. and {Font-Ribera}, Andreu and Fredrickson, Alexander and Freischlad, Gordon and Frinchaboy, Peter M. and Fuentes, Carla E. and Galbany, Llu{\'i}s and {Garcia-Dias}, R. and {Garc{\'i}a-Hern{\'a}ndez}, D. A. and Gaulme, Patrick and Geisler, Doug and Gelfand, Joseph D. and {Gil-Mar{\'i}n}, H{\'e}ctor and Gillespie, Bruce A. and Goddard, Daniel and {Gonzalez-Perez}, Violeta and Grabowski, Kathleen and Green, Paul J. and Grier, Catherine J. and Gunn, James E. and Guo, Hong and Guy, Julien and Hagen, Alex and Hahn, ChangHoon and Hall, Matthew and Harding, Paul and Hasselquist, Sten and Hawley, Suzanne L. and Hearty, Fred and Gonzalez Hern{\'a}ndez, Jonay I. and Ho, Shirley and Hogg, David W. and {Holley-Bockelmann}, Kelly and Holtzman, Jon A. and Holzer, Parker H. and Huehnerhoff, Joseph and Hutchinson, Timothy A. and Hwang, Ho Seong and {Ibarra-Medel}, H{\'e}ctor J. and {da Silva Ilha}, Gabriele and Ivans, Inese I. and Ivory, KeShawn and Jackson, Kelly and Jensen, Trey W. and Johnson, Jennifer A. and Jones, Amy and J{\"o}nsson, Henrik and Jullo, Eric and Kamble, Vikrant and Kinemuchi, Karen and Kirkby, David and Kitaura, Francisco-Shu and Klaene, Mark and Knapp, Gillian R. and Kneib, Jean-Paul and Kollmeier, Juna A. and Lacerna, Ivan and Lane, Richard R. and Lang, Dustin and Law, David R. and Lazarz, Daniel and Lee, Youngbae and Le Goff, Jean-Marc and Liang, Fu-Heng and Li, Cheng and Li, Hongyu and Lian, Jianhui and Lima, Marcos and Lin, Lihwai and Lin, Yen-Ting and {Bertran de Lis}, Sara and Liu, Chao and {de Icaza Lizaola}, Miguel Angel C. and Long, Dan and Lucatello, Sara and Lundgren, Britt and MacDonald, Nicholas K. and Deconto Machado, Alice and MacLeod, Chelsea L. and Mahadevan, Suvrath and Geimba Maia, Marcio Antonio and Maiolino, Roberto and Majewski, Steven R. and Malanushenko, Elena and Malanushenko, Viktor and Manchado, Arturo and Mao, Shude and Maraston, Claudia and {Marques-Chaves}, Rui and Masseron, Thomas and Masters, Karen L. and McBride, Cameron K. and McDermid, Richard M. and McGrath, Brianne and McGreer, Ian D. and Medina Pe{\~n}a, Nicol{\'a}s and Melendez, Matthew and Merloni, Andrea and Merrifield, Michael R. and Meszaros, Szabolcs and Meza, Andres and Minchev, Ivan and Minniti, Dante and Miyaji, Takamitsu and More, Surhud and Mulchaey, John and {M{\"u}ller-S{\'a}nchez}, Francisco and Muna, Demitri and Munoz, Ricardo R. and Myers, Adam D. and Nair, Preethi and Nandra, Kirpal and {Correa do Nascimento}, Janaina and Negrete, Alenka and Ness, Melissa and Newman, Jeffrey A. and Nichol, Robert C. and Nidever, David L. and Nitschelm, Christian and Ntelis, Pierros and O'Connell, Julia E. and Oelkers, Ryan J. and Oravetz, Audrey and Oravetz, Daniel and Pace, Zach and Padilla, Nelson and {Palanque-Delabrouille}, Nathalie and Alonso Palicio, Pedro and Pan, Kaike and Parejko, John K. and Parikh, Taniya and P{\^a}ris, Isabelle and Park, Changbom and Patten, Alim Y. and Peirani, Sebastien and {Pellejero-Ibanez}, Marcos and Penny, Samantha and Percival, Will J. and {Perez-Fournon}, Ismael and Petitjean, Patrick and Pieri, Matthew M. and Pinsonneault, Marc and Pisani, Alice and Poleski, Rados{\l}aw and Prada, Francisco and Prakash, Abhishek and Queiroz, Anna B{\'a}rbara de Andrade and Raddick, M. Jordan and Raichoor, Anand and Barboza Rembold, Sandro and Richstein, Hannah and Riffel, Rogemar A. and Riffel, Rog{\'e}rio and Rix, Hans-Walter and Robin, Annie C. and Rockosi, Constance M. and {Rodr{\'i}guez-Torres}, Sergio and {Roman-Lopes}, A. and {Rom{\'a}n-Z{\'u}{\~n}iga}, Carlos and Rosado, Margarita and Ross, Ashley J. and Rossi, Graziano and Ruan, John and Ruggeri, Rossana and Rykoff, Eli S. and {Salazar-Albornoz}, Salvador and Salvato, Mara and S{\'a}nchez, Ariel G. and Aguado, D. S. and {S{\'a}nchez-Gallego}, Jos{\'e} R. and Santana, Felipe A. and Santiago, Bas{\'i}lio Xavier and Sayres, Conor and Schiavon, Ricardo P. and {da Silva Schimoia}, Jaderson and Schlafly, Edward F. and Schlegel, David J. and Schneider, Donald P. and Schultheis, Mathias and Schuster, William J. and Schwope, Axel and Seo, Hee-Jong and Shao, Zhengyi and Shen, Shiyin and Shetrone, Matthew and Shull, Michael and Simon, Joshua D. and Skinner, Danielle and Skrutskie, M. F. and Slosar, An{\v z}e and Smith, Verne V. and Sobeck, Jennifer S. and Sobreira, Flavia and Somers, Garrett and Souto, Diogo and Stark, David V. and Stassun, Keivan and Stauffer, Fritz and Steinmetz, Matthias and {Storchi-Bergmann}, Thaisa and Streblyanska, Alina and Stringfellow, Guy S. and Su{\'a}rez, Genaro and Sun, Jing and Suzuki, Nao and Szigeti, Laszlo and {Taghizadeh-Popp}, Manuchehr and Tang, Baitian and Tao, Charling and Tayar, Jamie and Tembe, Mita and Teske, Johanna and Thakar, Aniruddha R. and Thomas, Daniel and Thompson, Benjamin A. and Tinker, Jeremy L. and Tissera, Patricia and Tojeiro, Rita and Hernandez Toledo, Hector and {de la Torre}, Sylvain and Tremonti, Christy and Troup, Nicholas W. and Valenzuela, Octavio and Martinez Valpuesta, Inma and {Vargas-Gonz{\'a}lez}, Jaime and {Vargas-Maga{\~n}a}, Mariana and Vazquez, Jose Alberto and Villanova, Sandro and Vivek, M. and Vogt, Nicole and Wake, David and Walterbos, Rene and Wang, Yuting and Weaver, Benjamin Alan and Weijmans, Anne-Marie and Weinberg, David H. and Westfall, Kyle B. and Whelan, David G. and Wild, Vivienne and Wilson, John and {Wood-Vasey}, W. M. and Wylezalek, Dominika and Xiao, Ting and Yan, Renbin and Yang, Meng and Ybarra, Jason E. and Y{\`e}che, Christophe and Zakamska, Nadia and Zamora, Olga and Zarrouk, Pauline and Zasowski, Gail and Zhang, Kai and Zhao, Gong-Bo and Zheng, Zheng and Zheng, Zheng and Zhou, Xu and Zhou, Zhi-Min and Zhu, Guangtun B. and Zoccali, Manuela and Zou, Hu},
year = {2017},
month = jul,
volume = {154},
pages = {28},
doi = {10.3847/1538-3881/aa7567},
abstract = {We describe the Sloan Digital Sky Survey IV (SDSS-IV), a project encompassing three major spectroscopic programs. The Apache Point Observatory Galactic Evolution Experiment 2 (APOGEE-2) is observing hundreds of thousands of Milky Way stars at high resolution and high signal-to-noise ratios in the near-infrared. The Mapping Nearby Galaxies at Apache Point Observatory (MaNGA) survey is obtaining spatially resolved spectroscopy for thousands of nearby galaxies (median z{$\sim$} 0.03). The extended Baryon Oscillation Spectroscopic Survey (eBOSS) is mapping the galaxy, quasar, and neutral gas distributions between z{$\sim$} 0.6 and 3.5 to constrain cosmology using baryon acoustic oscillations, redshift space distortions, and the shape of the power spectrum. Within eBOSS, we are conducting two major subprograms: the SPectroscopic IDentification of eROSITA Sources (SPIDERS), investigating X-ray AGNs and galaxies in X-ray clusters, and the Time Domain Spectroscopic Survey (TDSS), obtaining spectra of variable sources. All programs use the 2.5 m Sloan Foundation Telescope at the Apache Point Observatory; observations there began in Summer 2014. APOGEE-2 also operates a second near-infrared spectrograph at the 2.5 m du Pont Telescope at Las Campanas Observatory, with observations beginning in early 2017. Observations at both facilities are scheduled to continue through 2020. In keeping with previous SDSS policy, SDSS-IV provides regularly scheduled public data releases; the first one, Data Release 13, was made available in 2016 July.},
journal = {AJ},
keywords = {cosmology: observations,galaxies: general,Galaxy: general,instrumentation: spectrographs,stars: general,surveys}
}
@article{Bohm-Vitense1958,
title = {\"Uber Die Wasserstoffkonvektionszone in Sternen Verschiedener Effektivtemperaturen Und Leuchtkr\"afte. {{Mit}} 5 Textabbildungen},
author = {{B{\"o}hm-Vitense}, E.},
year = {1958},
month = jan,
volume = {46},
pages = {108},
journal = {Z. F\"ur Astrophys.}
}
@article{Bonaca.Tanner.ea2012,
title = {Calibrating {{Convective Properties}} of {{Solar}}-like {{Stars}} in the {{Kepler Field}} of {{View}}},
author = {Bonaca, Ana and Tanner, Joel D. and Basu, Sarbani and Chaplin, William J. and Metcalfe, Travis S. and Monteiro, M{\'a}rio J. P. F. G. and Ballot, J{\'e}r{\^o}me and Bedding, Timothy R. and Bonanno, Alfio and Broomhall, Anne-Marie and Bruntt, Hans and Campante, Tiago L. and {Christensen-Dalsgaard}, J{\o}rgen and Corsaro, Enrico and Elsworth, Yvonne and Garc{\'i}a, Rafael A. and Hekker, Saskia and Karoff, Christoffer and Kjeldsen, Hans and Mathur, Savita and R{\'e}gulo, Clara and Roxburgh, Ian and Stello, Dennis and Trampedach, Regner and Barclay, Thomas and Burke, Christopher J. and Caldwell, Douglas A.},
year = {2012},
month = aug,
volume = {755},
pages = {L12},
issn = {0004-637X},
doi = {10.1088/2041-8205/755/1/L12},
abstract = {Stellar models generally use simple parameterizations to treat convection. The most widely used parameterization is the so-called mixing-length theory where the convective eddy sizes are described using a single number, {$\alpha$}, the mixing-length parameter. This is a free parameter, and the general practice is to calibrate {$\alpha$} using the known properties of the Sun and apply that to all stars. Using data from NASA's Kepler mission we show that using the solar-calibrated {$\alpha$} is not always appropriate, and that in many cases it would lead to estimates of initial helium abundances that are lower than the primordial helium abundance. Kepler data allow us to calibrate {$\alpha$} for many other stars and we show that for the sample of stars we have studied, the mixing-length parameter is generally lower than the solar value. We studied the correlation between {$\alpha$} and stellar properties, and we find that {$\alpha$} increases with metallicity. We therefore conclude that results obtained by fitting stellar models or by using population-synthesis models constructed with solar values of {$\alpha$} are likely to have large systematic errors. Our results also confirm theoretical expectations that the mixing-length parameter should vary with stellar properties.},
journal = {Astrophys. J. Lett.},
keywords = {stars: fundamental parameters,stars: interiors,stars: oscillations}
}
@article{Bonanno.Schlattl.ea2002,
title = {The Age of the {{Sun}} and the Relativistic Corrections in the {{EOS}}},
author = {Bonanno, A. and Schlattl, H. and Patern{\`o}, L.},
year = {2002},
month = aug,
volume = {390},
pages = {1115--1118},
issn = {0004-6361},
doi = {10.1051/0004-6361:20020749},
abstract = {We show that the inclusion of special relativistic corrections in the revised OPAL and MHD equations of state has a significant impact on the helioseismic determination of the solar age. Models with relativistic corrections included lead to a reduction of about 0.05-0.08 ; Gyr with respect to those obtained with the old OPAL or MHD EOS. Our best-fit value is tseis = (4.57 +/- 0.11) ; Gyr which is in remarkably good agreement with the meteoritic value for the solar age. We argue that the inclusion of relativistic corrections is important for probing the evolutionary state of a star by means of the small frequency separations delta nu l,n=nu l,n-nu l+2,n-1, for spherical harmonic degrees l = 0,1 and radial order n {$>>$} l.},
journal = {A\&A},
keywords = {equation of state,Sun: interior,Sun: oscillations}
}
@article{Borucki.Koch.ea2010,
title = {Kepler {{Planet}}-{{Detection Mission}}: {{Introduction}} and {{First Results}}},
shorttitle = {Kepler {{Planet}}-{{Detection Mission}}},
author = {Borucki, William J. and Koch, David and Basri, Gibor and Batalha, Natalie and Brown, Timothy and Caldwell, Douglas and Caldwell, John and {Christensen-Dalsgaard}, J{\o}rgen and Cochran, William D. and DeVore, Edna and Dunham, Edward W. and Dupree, Andrea K. and Gautier, Thomas N. and Geary, John C. and Gilliland, Ronald and Gould, Alan and Howell, Steve B. and Jenkins, Jon M. and Kondo, Yoji and Latham, David W. and Marcy, Geoffrey W. and Meibom, S{\o}ren and Kjeldsen, Hans and Lissauer, Jack J. and Monet, David G. and Morrison, David and Sasselov, Dimitar and Tarter, Jill and Boss, Alan and Brownlee, Don and Owen, Toby and Buzasi, Derek and Charbonneau, David and Doyle, Laurance and Fortney, Jonathan and Ford, Eric B. and Holman, Matthew J. and Seager, Sara and Steffen, Jason H. and Welsh, William F. and Rowe, Jason and Anderson, Howard and Buchhave, Lars and Ciardi, David and Walkowicz, Lucianne and Sherry, William and Horch, Elliott and Isaacson, Howard and Everett, Mark E. and Fischer, Debra and Torres, Guillermo and Johnson, John Asher and Endl, Michael and MacQueen, Phillip and Bryson, Stephen T. and Dotson, Jessie and Haas, Michael and Kolodziejczak, Jeffrey and Van Cleve, Jeffrey and Chandrasekaran, Hema and Twicken, Joseph D. and Quintana, Elisa V. and Clarke, Bruce D. and Allen, Christopher and Li, Jie and Wu, Haley and Tenenbaum, Peter and Verner, Ekaterina and Bruhweiler, Frederick and Barnes, Jason and Prsa, Andrej},
year = {2010},
month = feb,
volume = {327},
pages = {977},
doi = {10.1126/science.1185402},
abstract = {The Kepler mission was designed to determine the frequency of Earth-sized planets in and near the habitable zone of Sun-like stars. The habitable zone is the region where planetary temperatures are suitable for water to exist on a planet's surface. During the first 6 weeks of observations, Kepler monitored 156,000 stars, and five new exoplanets with sizes between 0.37 and 1.6 Jupiter radii and orbital periods from 3.2 to 4.9 days were discovered. The density of the Neptune-sized Kepler-4b is similar to that of Neptune and GJ 436b, even though the irradiation level is 800,000 times higher. Kepler-7b is one of the lowest-density planets (\textasciitilde{} 0.17 gram per cubic centimeter) yet detected. Kepler-5b, -6b, and -8b confirm the existence of planets with densities lower than those predicted for gas giant planets.},
journal = {Science}
}
@article{Bossini.Vallenari.ea2019,
title = {Age Determination for 269 {{{\emph{Gaia}}}} {{DR2}} Open Clusters},
author = {Bossini, D. and Vallenari, A. and Bragaglia, A. and {Cantat-Gaudin}, T. and Sordo, R. and {Balaguer-N{\'u}{\~n}ez}, L. and Jordi, C. and Moitinho, A. and Soubiran, C. and Casamiquela, L. and Carrera, R. and Heiter, U.},
year = {2019},
month = mar,
volume = {623},
pages = {A108},
issn = {0004-6361, 1432-0746},
doi = {10.1051/0004-6361/201834693},
abstract = {Context. The Gaia Second Data Release provides precise astrometry and photometry for more than 1.3 billion sources. This catalog opens a new era concerning the characterization of open clusters and test stellar models, paving the way for better understanding of the disk properties.},
journal = {A\&A},
keywords = {bayesian inference,cluster ages,clusters,statistics},
language = {en}
}
@article{Bovy.Rix.ea2016,
title = {On {{Galactic Density Modeling}} in the {{Presence}} of {{Dust Extinction}}},
author = {Bovy, Jo and Rix, Hans-Walter and Green, Gregory M. and Schlafly, Edward F. and Finkbeiner, Douglas P.},
year = {2016},
month = feb,
volume = {818},
pages = {130},
issn = {0004-637X},
doi = {10.3847/0004-637X/818/2/130},
abstract = {Inferences about the spatial density or phase-space structure of stellar populations in the Milky Way require a precise determination of the effective survey volume. The volume observed by surveys such as Gaia or near-infrared spectroscopic surveys, which have good coverage of the Galactic midplane region, is highly complex because of the abundant small-scale structure in the three-dimensional interstellar dust extinction. We introduce a novel framework for analyzing the importance of small-scale structure in the extinction. This formalism demonstrates that the spatially complex effect of extinction on the selection function of a pencil-beam or contiguous sky survey is equivalent to a low-pass filtering of the extinction-affected selection function with the smooth density field. We find that the angular resolution of current 3D extinction maps is sufficient for analyzing Gaia sub-samples of millions of stars. However, the current distance resolution is inadequate and needs to be improved by an order of magnitude, especially in the inner Galaxy. We also present a practical and efficient method for properly taking the effect of extinction into account in analyses of Galactic structure through an effective selection function. We illustrate its use with the selection function of red-clump stars in APOGEE using and comparing a variety of current 3D extinction maps.},
journal = {ApJ},
keywords = {dust,extinction,Galaxy: kinematics and dynamics,Galaxy: structure,methods: data analysis,stars: statistics,surveys}
}
@article{Brogaard.Bruntt.ea2011,
title = {Age and Helium Content of the Open Cluster {{NGC}} 6791 from Multiple Eclipsing Binary Members. {{I}}. {{Measurements}}, Methods, and First Results},
author = {Brogaard, K. and Bruntt, H. and Grundahl, F. and Clausen, J. V. and Frandsen, S. and Vandenberg, D. A. and Bedin, L. R.},
year = {2011},
month = jan,
volume = {525},
pages = {A2},
issn = {0004-6361},
doi = {10.1051/0004-6361/201015503},
abstract = {Context. Models of stellar structure and evolution can be constrained by measuring accurate parameters of detached eclipsing binaries in open clusters. Multiple binary stars provide the means to determine helium abundances in these old stellar systems, and in turn, to improve age estimates. Aims: Earlier measurements of the masses and radii of the detached eclipsing binary V20 in the open cluster NGC 6791 were accurate enough to demonstrate that there are significant differences between current stellar models. Here we improve on those results and add measurements of two additional detached eclipsing binaries, the cluster members V18 and V80. The enlarged sample sets much tighter constraints on the properties of stellar models than has hitherto been possible, thereby improving both the accuracy and precision of the cluster age. Methods: We employed (i) high-resolution UVES spectroscopy of V18, V20 and V80 to determine their spectroscopic effective temperatures, [Fe/H] values, and spectroscopic orbital elements; and (ii) time-series photometry from the Nordic Optical Telescope to obtain the photometric elements. Results: The masses and radii of the V18 and V20 components are found to high accuracy, with errors on the masses in the range 0.27-0.36\% and errors on the radii in the range 0.61-0.92\%. V80 is found to be magnetically active, and more observations are needed to determine its parameters accurately. The metallicity of NGC 6791 is measured from disentangled spectra of the binaries and a few single stars to be [Fe/H] = +0.29 {$\pm$} 0.03 (random) {$\pm$} 0.07 (systematic). The cluster reddening and apparent distance modulus are found to be E(B-V) = 0.160{$\pm$}0.025 and (m-M)V = 13.51 {$\pm$} 0.06. A first model comparison shows that we can constrain the helium content of the NGC 6791 stars, and thus reach a more accurate age than previously possible. It may be possible to constrain additional parameters, in particular the C, N, and O abundances. This will be investigated in Paper II. Conclusions: Using multiple, detached eclipsing binaries for determining stellar cluster ages, it is now possible to constrain parameters of stellar models, notably the helium content, which were previously out of reach. By observing a suitable number of detached eclipsing binaries in several open clusters, it will be possible to calibrate the age-scale and the helium enrichment parameter {$\Delta$}Y/{$\Delta$}Z, and provide firm constraints that stellar models must reproduce. Based on observations carried out at the Nordic Optical Telescope at La Palma and ESO's VLT/UVES ESO, Paranal, Chile (75.D-0206A, 77.D-0827A, 081.D-0091).Tables 11-22 are only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsarc.u-strasbg.fr/viz-bin/qcat?J/A+A/525/A2},
journal = {A\&A},
keywords = {binaries: eclipsing,binaries: spectroscopic,open clusters and associations: individual: NGC 6791,stars: evolution,techniques: photometric,techniques: spectroscopic}
}
@article{Brogaard.VandenBerg.ea2012,
title = {Age and Helium Content of the Open Cluster {{NGC}} 6791 from Multiple Eclipsing Binary Members. {{II}}. {{Age}} Dependencies and New Insights},
author = {Brogaard, K. and VandenBerg, D. A. and Bruntt, H. and Grundahl, F. and Frandsen, S. and Bedin, L. R. and Milone, A. P. and Dotter, A. and Feiden, G. A. and Stetson, P. B. and Sandquist, E. and Miglio, A. and Stello, D. and {Jessen-Hansen}, J.},
year = {2012},
month = jul,
volume = {543},
pages = {A106},
issn = {0004-6361},
doi = {10.1051/0004-6361/201219196},
abstract = {Context. Models of stellar structure and evolution can be constrained by measuring accurate parameters of detached eclipsing binaries in open clusters. Multiple binary stars provide the means to determine helium abundances in these old stellar systems, and in turn, to improve estimates of their age. Aims: In the first paper of this series, we demonstrated how measurements of multiple eclipsing binaries in the old open cluster NGC 6791 sets tighter constraints on the properties of stellar models than has previously been possible, thereby potentially improving both the accuracy and precision of the cluster age. Here we add additional constraints and perform an extensive model comparison to determine the best estimates of the cluster age and helium content, employing as many observational constraints as possible. Methods: We improve our photometry and correct empirically for differential reddening effects. We then perform an extensive comparison of the new colour-magnitude diagrams (CMDs) and eclipsing binary measurements to Victoria and DSEP isochrones in order to estimate cluster parameters. We also reanalyse a spectrum of the star 2-17 to improve [Fe/H] constraints. Results: We find a best estimate of the age of \textasciitilde 8.3 Gyr for NGC 6791 while demonstrating that remaining age uncertainty is dominated by uncertainties in the CNO abundances. The helium mass fraction is well constrained at Y = 0.30{$\pm$}0.01 resulting in {$\Delta$}Y/{$\Delta$}Z \textasciitilde{} 1.4 assuming that such a relation exists. During the analysis we firmly identify blue straggler stars, including the star 2-17, and find indications for the presence of their evolved counterparts. Our analysis supports the RGB mass-loss found from asteroseismology and we determine precisely the absolute mass of stars on the lower RGB, MRGB = 1.15{$\pm$}0.02 M{$\odot$}. This will be an important consistency check for the detailed asteroseismology of cluster stars. Conclusions: Using multiple, detached eclipsing binaries for determining stellar cluster ages, it is now possible to constrain parameters of stellar models, notably the helium content, which were previously out of reach. By observing a suitable number of detached eclipsing binaries in several open clusters, it will be possible to calibrate the age-scale and the helium enrichment parameter {$\Delta$} Y/{$\Delta$} Z, and provide firm constraints that stellar models must reproduce. Based on observations carried out at the Nordic Optical Telescope at La Palma and ESO's VLT/UVES ESO, Paranal, Chile (75.D-0206A, 77.D-0827A, 81.D-0091).Photometric data are only available at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsarc.u-strasbg.fr/viz-bin/qcat?J/A+A/543/A106},
journal = {A\&A},
keywords = {binaries: eclipsing,eclipsing binaries,helium enrichment,open clusters and associations: individual: NGC 6791,stars: abundances,stars: evolution,stellar ages,techniques: photometric,techniques: spectroscopic}
}
@article{Broomhall.Chaplin.ea2011,
title = {Solar-Cycle Variations of Large Frequency Separations of Acoustic Modes: Implications for Asteroseismology},
shorttitle = {Solar-Cycle Variations of Large Frequency Separations of Acoustic Modes},
author = {Broomhall, A.-M. and Chaplin, W. J. and Elsworth, Y. and New, R.},
year = {2011},
month = jun,
volume = {413},
pages = {2978--2986},
issn = {0035-8711},
doi = {10.1111/j.1365-2966.2011.18375.x},
abstract = {We have studied solar-cycle changes in the large frequency separations that can be observed in Birmingham Solar Oscillations Network (BiSON) data. The large frequency separation is often one of the first outputs from asteroseismic studies because it can help constrain stellar properties like mass and radius. We have used three methods for estimating the large separations: use of individual p-mode frequencies, computation of the autocorrelation of frequency-power spectra, and computation of the power spectrum of the power spectrum. The values of the large separations obtained by the different methods are offset from each other and have differing sensitivities to the realization noise. A simple model was used to predict solar-cycle variations in the large separations, indicating that the variations are due to the well-known solar-cycle changes to mode frequency. However, this model is only valid over a restricted frequency range. We discuss the implications of these results for asteroseismology.},
journal = {MNRAS},
keywords = {methods: data analysis,Sun: helioseismology,Sun: oscillations}
}
@article{Broomhall.Miglio.ea2014,
title = {Prospects for Asteroseismic Inference on the Envelope Helium Abundance in Red Giant Stars},
author = {Broomhall, A.-M. and Miglio, A. and Montalb{\'a}n, J. and Eggenberger, P. and Chaplin, W. J. and Elsworth, Y. and Scuflaire, R. and Ventura, P. and Verner, G. A.},
year = {2014},
month = may,
volume = {440},
pages = {1828--1843},
doi = {10.1093/mnras/stu393},
abstract = {Regions of rapid variation in the internal structure of a star are often referred to as acoustic glitches since they create a characteristic periodic signature in the frequencies of p modes. Here we examine the localized disturbance arising from the helium second ionization zone in red giant branch and clump stars. More specifically, we determine how accurately and precisely the parameters of the ionization zone can be obtained from the oscillation frequencies of stellar models. We use models produced by three different generation codes that not only cover a wide range of stages of evolution along the red giant phase but also incorporate different initial helium abundances. To study the acoustic glitch caused by the second ionization zone of helium we have determined the second differences in frequencies of modes with the same angular degree, l, and then we fit the periodic function described by Houdek \& Gough to the second differences. We discuss the conditions under which such fits robustly and accurately determine the acoustic radius of the second ionization zone of helium. When the frequency of maximum amplitude of the p-mode oscillations was greater than 40 {$\mu$}Hz a robust value for the radius of the ionization zone was recovered for the majority of models. The determined radii of the ionization zones as inferred from the mode frequencies were found to be coincident with the local maximum in the first adiabatic exponent described by the models, which is associated with the outer edge of the second ionization zone of helium. Finally, we consider whether this method can be used to distinguish stars with different helium abundances. Although a definite trend in the amplitude of the signal is observed any distinction would be difficult unless the stars come from populations with vastly different helium abundances or the uncertainties associated with the fitted parameters can be reduced. However, application of our methodology could be useful for distinguishing between different populations of red giant stars in globular clusters, where distinct populations with very different helium abundances have been observed.},
journal = {MNRAS},
keywords = {asteroseismology,stars: abundances,stars: interiors,stars: oscillations}
}
@article{Buldgen.Salmon.ea2016,
title = {In-Depth Study of {{16CygB}} Using Inversion Techniques},
author = {Buldgen, G. and Salmon, S. J. A. J. and Reese, D. R. and Dupret, M. A.},
year = {2016},
month = dec,
volume = {596},
pages = {A73},
issn = {0004-6361},
doi = {10.1051/0004-6361/201628773},
abstract = {Context. The 16Cyg binary system hosts the solar-like Kepler targets with the most stringent observational constraints. Indeed, we benefit from very high quality oscillation spectra, as well as spectroscopic and interferometric observations. Moreover, this system is particularly interesting since both stars are very similar in mass but the A component is orbited by a red dwarf, whereas the B component is orbited by a Jovian planet and thus could have formed a more complex planetary system. In our previous study, we showed that seismic inversions of integrated quantities could be used to constrain microscopic diffusion in the A component. In this study, we analyse the B component in the light of a more regularised inversion. Aims: We wish to analyse independently the B component of the 16Cyg binary system using the inversion of an indicator dedicated to analyse core conditions, denoted tu. Using this independent determination, we wish to analyse any differences between both stars due to the potential influence of planetary formation on stellar structure and/or their respective evolution. Methods: First, we recall the observational constraints for 16CygB and the method we used to generate reference stellar models of this star. We then describe how we improved the inversion and how this approach could be used for future targets with a sufficient number of observed frequencies. The inversion results were then used to analyse the differences between the A and B components. Results: The inversion of the tu indicator for 16CygB shows a disagreement with models including microscopic diffusion and sharing the chemical composition previously derived for 16CygA. We show that small changes in chemical composition are insufficient to solve the problem but that extra mixing can account for the differences seen between both stars. We use a parametric approach to analyse the impact of extra mixing in the form of turbulent diffusion on the behaviour of the tu values. We conclude on the necessity of further investigations using models with a physically motivated implementation of extra mixing processes including additional constraints to further improve the accuracy with which the fundamental parameters of this system are determined.},
journal = {A\&A},
keywords = {asteroseismology,stars: fundamental parameters,stars: interiors,stars: oscillations}
}
@article{Campante.Lund.ea2016,
title = {Spin-{{Orbit Alignment}} of {{Exoplanet Systems}}: {{Ensemble Analysis Using Asteroseismology}}},
shorttitle = {Spin-{{Orbit Alignment}} of {{Exoplanet Systems}}},
author = {Campante, T. L. and Lund, M. N. and Kuszlewicz, J. S. and Davies, G. R. and Chaplin, W. J. and Albrecht, S. and Winn, J. N. and Bedding, T. R. and Benomar, O. and Bossini, D. and Handberg, R. and Santos, A. R. G. and Van Eylen, V. and Basu, S. and {Christensen-Dalsgaard}, J. and Elsworth, Y. P. and Hekker, S. and Hirano, T. and Huber, D. and Karoff, C. and Kjeldsen, H. and Lundkvist, M. S. and North, T. S. H. and Silva Aguirre, V. and Stello, D. and White, T. R.},
year = {2016},
month = mar,
volume = {819},
pages = {85},
issn = {0004-637X},
doi = {10.3847/0004-637X/819/1/85},
abstract = {The angle {$\psi$} between a planet's orbital axis and the spin axis of its parent star is an important diagnostic of planet formation, migration, and tidal evolution. We seek empirical constraints on {$\psi$} by measuring the stellar inclination Is via asteroseismology for an ensemble of 25 solar-type hosts observed with NASA's Kepler satellite. Our results for Is are consistent with alignment at the 2{$\sigma$} level for all stars in the sample, meaning that the system surrounding the red-giant star Kepler-56 remains as the only unambiguous misaligned multiple-planet system detected to date. The availability of a measurement of the projected spin-orbit angle {$\lambda$} for two of the systems allows us to estimate {$\psi$}. We find that the orbit of the hot Jupiter HAT-P-7b is likely to be retrograde (\textbackslash psi =116\textbackslash buildrel\{\textbackslash circ\}\textbackslash over\{.\} \{4\}-14.7+30.2), whereas that of Kepler-25c seems to be well aligned with the stellar spin axis (\textbackslash psi =12\textbackslash buildrel\{\textbackslash circ\}\textbackslash over\{.\} \{6\}-11.0+6.7). While the latter result is in apparent contradiction with a statement made previously in the literature that the multi-transiting system Kepler-25 is misaligned, we show that the results are consistent, given the large associated uncertainties. Finally, we perform a hierarchical Bayesian analysis based on the asteroseismic sample in order to recover the underlying distribution of {$\psi$}. The ensemble analysis suggests that the directions of the stellar spin and planetary orbital axes are correlated, as conveyed by a tendency of the host stars to display large values of inclination.},
journal = {ApJ},
keywords = {asteroseismology,methods: statistical,planetary systems,planets and satellites: general,stars: solar-type,techniques: photometric}
}
@article{Campante.Lund.ea2016a,
title = {Spin-{{Orbit Alignment}} of {{Exoplanet Systems}}: {{Ensemble Analysis Using Asteroseismology}}},
shorttitle = {Spin-{{Orbit Alignment}} of {{Exoplanet Systems}}},
author = {Campante, T. L. and Lund, M. N. and Kuszlewicz, J. S. and Davies, G. R. and Chaplin, W. J. and Albrecht, S. and Winn, J. N. and Bedding, T. R. and Benomar, O. and Bossini, D. and Handberg, R. and Santos, A. R. G. and Van Eylen, V. and Basu, S. and {Christensen-Dalsgaard}, J. and Elsworth, Y. P. and Hekker, S. and Hirano, T. and Huber, D. and Karoff, C. and Kjeldsen, H. and Lundkvist, M. S. and North, T. S. H. and Silva Aguirre, V. and Stello, D. and White, T. R.},
year = {2016},
month = mar,
volume = {819},
pages = {85},
doi = {10.3847/0004-637X/819/1/85},
abstract = {The angle {$\psi$} between a planet's orbital axis and the spin axis of its parent star is an important diagnostic of planet formation, migration, and tidal evolution. We seek empirical constraints on {$\psi$} by measuring the stellar inclination Is via asteroseismology for an ensemble of 25 solar-type hosts observed with NASA's Kepler satellite. Our results for Is are consistent with alignment at the 2{$\sigma$} level for all stars in the sample, meaning that the system surrounding the red-giant star Kepler-56 remains as the only unambiguous misaligned multiple-planet system detected to date. The availability of a measurement of the projected spin-orbit angle {$\lambda$} for two of the systems allows us to estimate {$\psi$}. We find that the orbit of the hot Jupiter HAT-P-7b is likely to be retrograde (\textbackslash psi =116\textbackslash buildrel\{\textbackslash circ\}\textbackslash over\{.\} \{4\}-14.7+30.2), whereas that of Kepler-25c seems to be well aligned with the stellar spin axis (\textbackslash psi =12\textbackslash buildrel\{\textbackslash circ\}\textbackslash over\{.\} \{6\}-11.0+6.7). While the latter result is in apparent contradiction with a statement made previously in the literature that the multi-transiting system Kepler-25 is misaligned, we show that the results are consistent, given the large associated uncertainties. Finally, we perform a hierarchical Bayesian analysis based on the asteroseismic sample in order to recover the underlying distribution of {$\psi$}. The ensemble analysis suggests that the directions of the stellar spin and planetary orbital axes are correlated, as conveyed by a tendency of the host stars to display large values of inclination.},
journal = {ApJ},
keywords = {asteroseismology,methods: statistical,planetary systems,planets and satellites: general,stars: solar-type,techniques: photometric}
}
@article{Casagrande.Flynn.ea2007,
title = {The Helium Abundance and {{$\Delta$Y}}/{{$\Delta$Z}} in Lower Main-Sequence Stars},
author = {Casagrande, Luca and Flynn, Chris and Portinari, Laura and Girardi, Leo and Jimenez, Raul},
year = {2007},
month = dec,
volume = {382},
pages = {1516--1540},
issn = {0035-8711},
doi = {10.1111/j.1365-2966.2007.12512.x},
abstract = {We use nearby K dwarf stars to measure the helium-to-metal enrichment ratio {$\Delta$}Y/{$\Delta$}Z, a diagnostic of the chemical history of the solar neighbourhood. Our sample of K dwarfs has homogeneously determined effective temperatures, bolometric luminosities and metallicities, allowing us to fit each star to the appropriate stellar isochrone and determine its helium content indirectly. We use a newly computed set of Padova isochrones which cover a wide range of helium and metal content. Our theoretical isochrones have been checked against a congruous set of main-sequence binaries with accurately measured masses, to discuss and validate their range of applicability. We find that the stellar masses deduced from the isochrones are usually in excellent agreement with empirical measurements. Good agreement is also found with empirical mass-luminosity relations. Despite fitting the masses of the stars very well, we find that anomalously low helium content (lower than primordial helium) is required to fit the luminosities and temperatures of the metal-poor K dwarfs, while more conventional values of the helium content are derived for the stars around solar metallicity. We have investigated the effect of diffusion in stellar models and the assumption of local thermodynamic equilibrium (LTE) in deriving metallicities. Neither of these is able to resolve the low-helium problem alone and only marginally if the cumulated effects are included, unless we assume a mixing-length which is strongly decreasing with metallicity. Further work in stellar models is urgently needed. The helium-to-metal enrichment ratio is found to be {$\Delta$}Y/{$\Delta$}Z = 2.1 +/- 0.9 around and above solar metallicity, consistent with previous studies, whereas open problems still remain at the lowest metallicities. Finally, we determine the helium content for a set of planetary host stars.},
journal = {MNRAS},
keywords = {binaries: general,etc.),Hertzsprung-Russell (HR) diagram,luminosities,masses,radii,stars: abundances,stars: fundamental parameters (colours,stars: interiors,stars: late-type,temperatures}
}
@article{Casagrande.Ramirez.ea2010,
title = {An Absolutely Calibrated {{Teff}} Scale from the Infrared Flux Method. {{Dwarfs}} and Subgiants},
author = {Casagrande, L. and Ram{\'i}rez, I. and Mel{\'e}ndez, J. and Bessell, M. and Asplund, M.},
year = {2010},
month = mar,
volume = {512},
pages = {A54},
issn = {0004-6361},
doi = {10.1051/0004-6361/200913204},
abstract = {Various effective temperature scales have been proposed over the years. Despite much work and the high internal precision usually achieved, systematic differences of order 100 K (or more) among various scales are still present. We present an investigation based on the infrared flux method aimed at assessing the source of such discrepancies and pin down their origin. We break the impasse among different scales by using a large set of solar twins, stars which are spectroscopically and photometrically identical to the Sun, to set the absolute zero point of the effective temperature scale to within few degrees. Our newly calibrated, accurate and precise temperature scale applies to dwarfs and subgiants, from super-solar metallicities to the most metal-poor stars currently known. At solar metallicities our results validate spectroscopic effective temperature scales, whereas for [Fe/H]{$\lnapprox$} -2.5 our temperatures are roughly 100 K hotter than those determined from model fits to the Balmer lines and 200 K hotter than those obtained from the excitation equilibrium of Fe lines. Empirical bolometric corrections and useful relations linking photometric indices to effective temperatures and angular diameters have been derived. Our results take full advantage of the high accuracy reached in absolute calibration in recent years and are further validated by interferometric angular diameters and space based spectrophotometry over a wide range of effective temperatures and metallicities. Table 8 is only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/512/A54},
journal = {A\&A},
keywords = {infrared: stars,stars: abundances,stars: atmospheres,stars: fundamental parameters,techniques: photometric}
}
@article{Chan.Bovy2020,
title = {The {{Gaia DR2}} Parallax Zero-Point: Hierarchical Modelling of Red Clump Stars},
shorttitle = {The {{Gaia DR2}} Parallax Zero-Point},
author = {Chan, Victor C. and Bovy, Jo},
year = {2020},
month = apr,
volume = {493},
pages = {4367--4381},
issn = {0035-8711},
doi = {10.1093/mnras/staa571},
abstract = {The systematic offset of Gaia parallaxes has been widely reported with Gaia's second data release, and it is expected to persist in future Gaia data. In order to use Gaia parallaxes to infer distances to high precision, we develop a hierarchical probabilistic model to determine the Gaia parallax zero-point offset along with the calibration of an empirical model for luminosity of red clump stars by combining astrometric and photometric measurements. Using a cross-matched sample of red clump stars from the Apache Point Observatory Galactic Evolution Experiment and Gaia Data Release 2 (DR2), we report the parallax zero-point offset in DR2 to be {$\pi\_$}0 = -48 {$\pm$} 1 {$\mu$} as. We infer the red clump absolute magnitude to be -1.622 {$\pm$} 0.004 in Ks, 0.435 {$\pm$} 0.004 in G, -1.019 {$\pm$} 0.004 in J, and -1.516 {$\pm$} 0.004 in H. The intrinsic scatter of the red clump is {$\sim$}0.09 mag in J, H, and Ks, or {$\sim$} 0.12 mag in G. We tailor our models to accommodate more complex analyses such as investigating the variations of the parallax zero-point with each source's observed magnitude, observed colour, and sky position. In particular, we find fluctuations of the zero-point across the sky to be of order or less than a few 10s of {$\mu$} as.},
journal = {MNRAS},
keywords = {catalogues,methods: statistical,parallaxes,stars: distances,stars: horizontal branch,surveys}
}
@book{Chandrasekhar1939,
title = {An Introduction to the Study of Stellar Structure},
author = {Chandrasekhar, Subrahmanyan},
year = {1939}
}
@article{Chaplin.Elsworth.ea2007,
title = {Solar P-{{Mode Frequencies}} over {{Three Solar Cycles}}},
author = {Chaplin, W. J. and Elsworth, Y. and Miller, B. A. and Verner, G. A. and New, R.},
year = {2007},
month = apr,
volume = {659},
pages = {1749--1760},
issn = {0004-637X},
doi = {10.1086/512543},
abstract = {We analyze thirty years of solar oscillations data collected by the Birmingham Solar Oscillations Network (BiSON). Estimates of the mean frequency shifts of low-degree p-modes have been extracted over a period spanning solar cycles 21-23. Two methods of analysis are used to extract the frequency shifts: one method uses results on fitted frequencies of individual modes, which are then averaged to give mean frequency shifts; the other method uses cross-correlations of power frequency spectra made from subsets of the data shifted in time. The frequency shifts are correlated against six proxies of solar activity, which are sensitive to magnetic and irradiance variability at a range of locations from the photosphere to the corona. We find proxies that have good sensitivity to the effects of weak-component magnetic flux-which is more widely distributed in latitude than the strong flux in the active regions-are those that follow the frequency shifts most consistently over the three cycles. This list includes the Mg II H and K core-to-wing data, the 10.7 cm radio flux, and the He I equivalent width data. While the two methods of analysis give consistent results, use of the cross-correlation function to measure mean frequency shifts returns less precise values in cases in which the duty cycle is greater than 30\%. Estimation of uncertainties from the cross-correlation method also requires that proper allowance be made for strong correlations in the data.},
journal = {ApJ},
keywords = {Sun: Activity,Sun: Helioseismology,Sun: Magnetic Fields}
}
@article{Chaplin.Miglio2013,
title = {Asteroseismology of {{Solar}}-{{Type}} and {{Red}}-{{Giant Stars}}},
author = {Chaplin, William J. and Miglio, Andrea},
year = {2013},
month = aug,
volume = {51},
pages = {353--392},
doi = {10.1146/annurev-astro-082812-140938},
abstract = {We are entering a golden era for stellar physics driven by satellite and telescope observations of unprecedented quality and scope. New insights on stellar evolution and stellar interiors physics are being made possible by asteroseismology, the study of stars by the observation of natural, resonant oscillations. Asteroseismology is proving to be particularly significant for the study of solar-type and red-giant stars. These stars show rich spectra of solar-like oscillations, which are excited and intrinsically damped by turbulence in the outermost layers of the convective envelopes. In this review we discuss the current state of the field, with a particular emphasis on recent advances provided by the Kepler and COROT (Convection, Rotation \& Planetary Transits) space missions and the wider significance to astronomy of the results from asteroseismology, such as stellar populations studies and exoplanet studies.},
journal = {ARA\&A}
}
@article{Chaplin.Serenelli.ea2020,
title = {Age Dating of an Early {{Milky Way}} Merger via Asteroseismology of the Naked-Eye Star {$\nu$} {{Indi}}},
author = {Chaplin, William J. and Serenelli, Aldo M. and Miglio, Andrea and Morel, Thierry and Mackereth, J. Ted and Vincenzo, Fiorenzo and Kjeldsen, Hans and Basu, Sarbani and Ball, Warrick H. and Stokholm, Amalie and Verma, Kuldeep and Mosumgaard, Jakob R{\o}rsted and Silva Aguirre, Victor and Mazumdar, Anwesh and Ranadive, Pritesh and Antia, H. M. and Lebreton, Yveline and Ong, Joel and Appourchaux, Thierry and Bedding, Timothy R. and {Christensen-Dalsgaard}, J{\o}rgen and Creevey, Orlagh and Garc{\'i}a, Rafael A. and Handberg, Rasmus and Huber, Daniel and Kawaler, Steven D. and Lund, Mikkel N. and Metcalfe, Travis S. and Stassun, Keivan G. and Bazot, Mich{\"a}el and Beck, Paul G. and Bell, Keaton J. and Bergemann, Maria and Buzasi, Derek L. and Benomar, Othman and Bossini, Diego and Bugnet, Lisa and Campante, Tiago L. and Orhan, Zeynep {\c C}elik and Corsaro, Enrico and {Gonz{\'a}lez-Cuesta}, Luc{\'i}a and Davies, Guy R. and Di Mauro, Maria Pia and Egeland, Ricky and Elsworth, Yvonne P. and Gaulme, Patrick and Ghasemi, Hamed and Guo, Zhao and Hall, Oliver J. and Hasanzadeh, Amir and Hekker, Saskia and Howe, Rachel and Jenkins, Jon M. and Jim{\'e}nez, Antonio and Kiefer, Ren{\'e} and Kuszlewicz, James S. and Kallinger, Thomas and Latham, David W. and Lundkvist, Mia S. and Mathur, Savita and Montalb{\'a}n, Josefina and Mosser, Benoit and Bed{\'o}n, Andres Moya and Nielsen, Martin Bo and {\"O}rtel, Sibel and Rendle, Ben M. and Ricker, George R. and Rodrigues, Tha{\'i}se S. and Roxburgh, Ian W. and Safari, Hossein and Schofield, Mathew and Seager, Sara and Smalley, Barry and Stello, Dennis and Szab{\'o}, R{\'o}bert and Tayar, Jamie and Theme{\ss}l, Nathalie and Thomas, Alexandra E. L. and Vanderspek, Roland K. and {van Rossem}, Walter E. and Vrard, Mathieu and Weiss, Achim and White, Timothy R. and Winn, Joshua N. and Y{\i}ld{\i}z, Mutlu},
year = {2020},
month = jan,
volume = {4},
pages = {382--389},
issn = {2397-3366},
doi = {10.1038/s41550-019-0975-9},
abstract = {Over the course of its history, the Milky Way has ingested multiple smaller satellite galaxies1. Although these accreted stellar populations can be forensically identified as kinematically distinct structures within the Galaxy, it is difficult in general to date precisely the age at which any one merger occurred. Recent results have revealed a population of stars that were accreted via the collision of a dwarf galaxy, called Gaia-Enceladus1, leading to substantial pollution of the chemical and dynamical properties of the Milky Way. Here we identify the very bright, naked-eye star {$\nu$} Indi as an indicator of the age of the early in situ population of the Galaxy. We combine asteroseismic, spectroscopic, astrometric and kinematic observations to show that this metal-poor, alpha-element-rich star was an indigenous member of the halo, and we measure its age to be 11.0 {$\pm$}0.7 ? (stat) {$\pm$}0.8 ? (sys) billion years. The star bears hallmarks consistent with having been kinematically heated by the Gaia-Enceladus collision. Its age implies that the earliest the merger could have begun was 11.6 and 13.2 billion years ago, at 68\% and 95\% confidence, respectively. Computations based on hierarchical cosmological models slightly reduce the above limits.},
journal = {Nat. Astron.}
}
@article{Chiappini.Anders.ea2015,
title = {Young [ {\emph{{$\alpha$}}} /{{Fe}}]-Enhanced Stars Discovered by {{CoRoT}} and {{APOGEE}}: {{What}} Is Their Origin?},
shorttitle = {Young [ {\emph{{$\alpha$}}} /{{Fe}}]-Enhanced Stars Discovered by {{CoRoT}} and {{APOGEE}}},
author = {Chiappini, C. and Anders, F. and Rodrigues, T. S. and Miglio, A. and Montalb{\'a}n, J. and Mosser, B. and Girardi, L. and Valentini, M. and Noels, A. and Morel, T. and Minchev, I. and Steinmetz, M. and Santiago, B. X. and Schultheis, M. and Martig, M. and {da Costa}, L. N. and Maia, M. A. G. and Allende Prieto, C. and {de Assis Peralta}, R. and Hekker, S. and Theme{\ss}l, N. and Kallinger, T. and Garc{\'i}a, R. A. and Mathur, S. and Baudin, F. and Beers, T. C. and Cunha, K. and Harding, P. and Holtzman, J. and Majewski, S. and M{\'e}sz{\'a}ros, Sz. and Nidever, D. and Pan, K. and Schiavon, R. P. and Shetrone, M. D. and Schneider, D. P. and Stassun, K.},
year = {2015},
month = apr,
volume = {576},
pages = {L12},
issn = {0004-6361, 1432-0746},
doi = {10.1051/0004-6361/201525865},
abstract = {We report the discovery of a group of apparently young CoRoT red-giant stars exhibiting enhanced [{$\alpha$}/Fe] abundance ratios (as determined from APOGEE spectra) with respect to solar values. Their existence is not explained by standard chemical evolution models of the Milky Way, and shows that the chemical-enrichment history of the Galactic disc is more complex. We find similar stars in previously published samples for which isochrone-ages could be reliably obtained, although in smaller relative numbers. This might explain why these stars have not previously received attention. The young [{$\alpha$}/Fe]-rich stars are much more numerous in the CoRoT-APOGEE (CoRoGEE) inner-field sample than in any other high-resolution sample available at present because only CoRoGEE can explore the inner-disc regions and provide ages for its field stars. The kinematic properties of the young [{$\alpha$}/Fe]-rich stars are not clearly thick-disc like, despite their rather large distances from the Galactic mid-plane. Our tentative interpretation of these and previous intriguing observations in the Milky Way is that these stars were formed close to the end of the Galactic bar, near corotation \textendash{} a region where gas can be kept inert for longer times than in other regions that are more frequently shocked by the passage of spiral arms. Moreover, this is where the mass return from older inner-disc stellar generations is expected to be highest (according to an inside-out disc-formation scenario), which additionally dilutes the in-situ gas. Other possibilities to explain these observations (e.g., a recent gas-accretion event) are also discussed.},
journal = {A\&A},
keywords = {alpha-rich,red giants},
language = {en}
}
@article{Chiosi.Matteucci1982,
title = {The Helium to Heavy Element Enrichment Ratio, {{Delta Y}}/{{Delta Z}}},
author = {Chiosi, C. and Matteucci, F. M.},
year = {1982},
month = jan,
volume = {105},
pages = {140--148},
issn = {0004-6361},
abstract = {The effects of mass loss by stellar wind in stars of more than one solar mass, and of different choices for the initial mass function on the helium-to-heavy element enrichment ratio, are studied on the way to the proposal of a model for the solar vicinity's chemical evolution which obeys all major observational constraints while taking account of star mass outflow and the dependence of the mass function on the gas metallicity. It is found that a Saltpeter-like mass function cannot be safely used to reproduce the observed Delta Y/Delta Z ratios unless massive star mass loss at substantial rates is invoked, since this function equally weighs both helium and metal producers. In such a case the ratio is easily matched but the observed young star metallicity and gas cannot be reproduced. It is concluded that present stellar and galactic evolution data can explain the observed Delta Y/Delta Z ratio.},
journal = {A\&A},
keywords = {Abundance,Astronomical Models,Chemical Evolution,Galactic Evolution,Heavy Elements,Helium,Mass Ratios,Stellar Evolution,Stellar Mass Ejection,Sun}
}
@article{Choi.Dotter.ea2016,
title = {Mesa {{Isochrones}} and {{Stellar Tracks}} ({{MIST}}). {{I}}. {{Solar}}-Scaled {{Models}}},
author = {Choi, Jieun and Dotter, Aaron and Conroy, Charlie and Cantiello, Matteo and Paxton, Bill and Johnson, Benjamin D.},
year = {2016},
month = jun,
volume = {823},
pages = {102},
doi = {10.3847/0004-637X/823/2/102},
abstract = {This is the first of a series of papers presenting the Modules for Experiments in Stellar Astrophysics (MESA) Isochrones and Stellar Tracks (MIST) project, a new comprehensive set of stellar evolutionary tracks and isochrones computed using MESA, a state-of-the-art open-source 1D stellar evolution package. In this work, we present models with solar-scaled abundance ratios covering a wide range of ages (5{$\leq$}slant \{log\}(\{Age\}) [\{year\}]{$\leq$}slant 10.3), masses (0.1{$\leq$}slant M/\{M\}{$\odot$} {$\leq$}slant 300), and metallicities (-2.0{$\leq$}slant [\{\{Z\}\}/\{\{H\}\}]{$\leq$}slant 0.5). The models are self-consistently and continuously evolved from the pre-main sequence (PMS) to the end of hydrogen burning, the white dwarf cooling sequence, or the end of carbon burning, depending on the initial mass. We also provide a grid of models evolved from the PMS to the end of core helium burning for -4.0{$\leq$}slant [\{\{Z\}\}/\{\{H\}\}]\textbackslash lt -2.0. We showcase extensive comparisons with observational constraints as well as with some of the most widely used existing models in the literature. The evolutionary tracks and isochrones can be downloaded from the project website at http://waps.cfa.harvard.edu/MIST/.},
journal = {ApJ},
keywords = {stars: evolution,stars: general,stars: interiors}
}
@article{Christensen-Dalsgaard.Gough1980,
title = {Is the Sun Helium-Deficient},
author = {{Christensen-Dalsgaard}, J. and Gough, D. O.},
year = {1980},
month = dec,
volume = {288},
pages = {544--547},
doi = {10.1038/288544a0},
abstract = {The recent observations of solar 5-min oscillations of low degree agree approximately with the predictions of a standard solar model with normal abundances of helium and heavy elements. Much of the apparent discrepancy noticed when the observations were first announced was a result of having neglected the influence of the sun's atmosphere in the normal mode analysis of the theoretical models. Our standard solar models are not in perfect agreement with observation, but it seems that major modifications will not be necessary to remove the remaining small discrepancies.},
journal = {Nature},
keywords = {Abundance,Heavy Elements,Helium,Solar Atmosphere,Solar Oscillations,Stellar Models}
}
@article{Christensen-Dalsgaard.Proffitt.ea1993,
title = {Effects of Diffusion on Solar Models and Their Oscillation Frequencies},
author = {{Christensen-Dalsgaard}, J. and Proffitt, C. R. and Thompson, M. J.},
year = {1993},
month = feb,
volume = {403},
pages = {L75-L78},
doi = {10.1086/186725},
abstract = {Settling and diffusion of helium have significant effects on the structure of solar models and their oscillation frequencies. We examine these effects in considerably more detail than has been done before, and we compare the computed frequencies with an extensive set of observed frequencies. We find that inclusion of diffusion results in a significant improvement in the agreement between theory and observations.},
journal = {Astrophys. J. Lett.},
keywords = {Asymptotic Properties,Frequencies,Helium,Plasma Diffusion,Solar Interior,Solar Oscillations,Stellar Models}
}
@article{Christensen-Dalsgaard1982,
title = {Seismological Studies of the Sun and Other Stars},
author = {{Christensen-Dalsgaard}, J.},
year = {1982},
volume = {2},
pages = {11--19},
doi = {10.1016/0273-1177(82)90250-2},
abstract = {Aspects of solar oscillations are addressed, emphasizing their utility as probes of solar interior structure. The results of observations of solar oscillations are reviewed, and a simplified version of asymptotic analyses of these oscillations is presented. The interpretation of five-minute oscillations of high and low degree and of rotationally split modes is discussed, and the prospects for constructing a stellar seismology based on observations of five-minute oscillations and theoretical estimates of their stellar analogues using stochastic excitation are examined. Expected advances in the study of solar oscillations are discussed, mentioning the particular projects that may lead to these advances.},
journal = {Adv. Space Res.},
keywords = {Bibliographies,Frequencies,Periodic Variations,Seismology,Solar Interior,Solar Oscillations,Stellar Models,Stellar Oscillations,Stellar Structure}
}
@article{Christensen-Dalsgaard1984,
title = {What {{Will Asteroseismology Teach}} Us},
author = {{Christensen-Dalsgaard}, J.},
year = {1984},
pages = {11},
abstract = {In the present paper the author considers the study of stars by means of their oscillation frequencies. Assuming high-order modes, the oscillation frequencies are conveniently discussed in terms of asymptotic theory. This is briefly reviewed. Then the author makes some remarks on observations of stellar oscillations and presents results of frequency calculations for a set of main sequence stars: such calculations are useful in giving an indication of the information that may be obtained from stellar oscillation frequencies.}
}
@article{Christensen-Dalsgaard2008,
title = {{{ASTEC}}\textemdash the {{Aarhus STellar Evolution Code}}},
author = {{Christensen-Dalsgaard}, J{\o}rgen},
year = {2008},
month = aug,
volume = {316},
pages = {13--24},
doi = {10.1007/s10509-007-9675-5},
abstract = {The Aarhus code is the result of a long development, starting in 1974, and still ongoing. A novel feature is the integration of the computation of adiabatic oscillations for specified models as part of the code. It offers substantial flexibility in terms of microphysics and has been carefully tested for the computation of solar models. However, considerable development is still required in the treatment of nuclear reactions, diffusion and convective mixing.},
journal = {Ap\&SS}
}
@article{Clevert.Unterthiner.ea2015,
title = {Fast and Accurate Deep Network Learning by Exponential Linear Units ({{ELUs}})},
author = {Clevert, Djork-Arn{\'e} and Unterthiner, Thomas and Hochreiter, Sepp},
year = {2015},
month = nov,
abstract = {We introduce the "exponential linear unit" (ELU) which speeds up learning in deep neural networks and leads to higher classification accuracies. Like rectified linear units (ReLUs), leaky ReLUs (LReLUs) and parametrized ReLUs (PReLUs), ELUs alleviate the vanishing gradient problem via the identity for positive values. However, ELUs have improved learning characteristics compared to the units with other activation functions. In contrast to ReLUs, ELUs have negative values which allows them to push mean unit activations closer to zero like batch normalization but with lower computational complexity. Mean shifts toward zero speed up learning by bringing the normal gradient closer to the unit natural gradient because of a reduced bias shift effect. While LReLUs and PReLUs have negative values, too, they do not ensure a noise-robust deactivation state. ELUs saturate to a negative value with smaller inputs and thereby decrease the forward propagated variation and information. Therefore, ELUs code the degree of presence of particular phenomena in the input, while they do not quantitatively model the degree of their absence. In experiments, ELUs lead not only to faster learning, but also to significantly better generalization performance than ReLUs and LReLUs on networks with more than 5 layers. On CIFAR-100 ELUs networks significantly outperform ReLU networks with batch normalization while batch normalization does not improve ELU networks. ELU networks are among the top 10 reported CIFAR-10 results and yield the best published result on CIFAR-100, without resorting to multi-view evaluation or model averaging. On ImageNet, ELU networks considerably speed up learning compared to a ReLU network with the same architecture, obtaining less than 10\% classification error for a single crop, single model network.},
adsnote = {Provided by the SAO/NASA Astrophysics Data System},
archiveprefix = {arXiv},
eid = {arXiv:1511.07289},
eprint = {1511.07289},
eprinttype = {arxiv},
journal = {ArXiv e-prints},
keywords = {Computer Science - Machine Learning},
primaryclass = {cs.LG}
}
@article{Coc.Uzan.ea2013,
title = {Standard Big-Bang Nucleosynthesis after Planck},
author = {Coc, Alain and Uzan, Jean-Philippe and Vangioni, Elisabeth},
year = {2013},
month = jul,
abstract = {Primordial or Big Bang nucleosynthesis (BBN) is one of the three historical strong evidences for the Big-Bang model together with the expansion of the Universe and the Cosmic Microwave Background radiation (CMB). The recent results by the Planck mission have slightly changed the estimate of the baryonic density Omega\_b, compared to the previous WMAP value. This article updates the BBN predictions for the light elements using the new value of Omega\_b determined by Planck, as well as an improvement of the nuclear network and new spectroscopic observations. While there is no major modification, the error bars of the primordial D/H abundance (2.67+/-0.09) x 10\^\{-5\} are narrower and there is a slight lowering of the primordial Li/H abundance (4.89\^+0.41\_-0.39) x 10\^\{-10\}. However, this last value is still -0.5ex\textasciitilde 3 times larger than its observed spectroscopic abundance in halo stars of the Galaxy. Primordial Helium abundance is now determined to be Y\_p = 0.2463+/-0.0003.},
adsnote = {Provided by the SAO/NASA Astrophysics Data System},
archiveprefix = {arXiv},
eid = {arXiv:1307.6955},
eprint = {1307.6955},
eprinttype = {arxiv},
journal = {ArXiv e-prints},
keywords = {Astrophysics - Cosmology and Extragalactic Astrophysics},
primaryclass = {astro-ph.CO}
}
@article{Connelly.Bizzarro.ea2012,
title = {The {{Absolute Chronology}} and {{Thermal Processing}} of {{Solids}} in the {{Solar Protoplanetary Disk}}},
author = {Connelly, James N. and Bizzarro, Martin and Krot, Alexander N. and Nordlund, {\AA}ke and Wielandt, Daniel and Ivanova, Marina A.},
year = {2012},
month = nov,
volume = {338},
pages = {651},
doi = {10.1126/science.1226919},
abstract = {Transient heating events that formed calcium-aluminum-rich inclusions (CAIs) and chondrules are fundamental processes in the evolution of the solar protoplanetary disk, but their chronology is not understood. Using U-corrected Pb-Pb dating, we determined absolute ages of individual CAIs and chondrules from primitive meteorites. CAIs define a brief formation interval corresponding to an age of 4567.30 {$\pm$} 0.16 million years (My), whereas chondrule ages range from 4567.32 {$\pm$} 0.42 to 4564.71 {$\pm$} 0.30 My. These data refute the long-held view of an age gap between CAIs and chondrules and, instead, indicate that chondrule formation started contemporaneously with CAIs and lasted \textasciitilde 3 My. This time scale is similar to disk lifetimes inferred from astronomical observations, suggesting that the formation of CAIs and chondrules reflects a process intrinsically linked to the secular evolution of accretionary disks.},
journal = {Science}
}
@article{Cooke.Fumagalli2018,
title = {Measurement of the Primordial Helium Abundance from the Intergalactic Medium},
author = {Cooke, Ryan J. and Fumagalli, Michele},
year = {2018},
month = oct,
volume = {2},
pages = {957--961},
issn = {2397-3366},
doi = {10.1038/s41550-018-0584-z},
abstract = {Almost every helium atom in the Universe was created just a few minutes after the Big Bang through a process commonly referred to as Big Bang nucleosynthesis1,2. The amount of helium that was made during Big Bang nucleosynthesis is determined by combining particle physics and cosmology3. The current leading measures of the primordial helium abundance (YP) are based on the relative strengths of H uc(i) and He uc(i) emission lines emanating from star-forming regions in local metal-poor galaxies4-7. As the statistical errors on these measurements improve, it is essential to test for systematics by developing independent techniques. Here we report a determination of the primordial helium abundance based on a near-pristine intergalactic gas cloud that is seen in absorption against the light of a background quasar. This gas cloud, observed when the Universe was just one-third of its present age (zabs = 1.724), has a metal content around 100 times less than that of the Sun, and at least 30\% less metal content than the most metal-poor H uc(ii) region currently known where a determination of the primordial helium abundance is possible. We conclude that the helium abundance of this intergalactic gas cloud is Y =0.25 0-0.025+0.033 , which agrees with the standard model primordial value8-10, YP = 0.24672 {$\pm$} 0.00017. Our determination of the primordial helium abundance is not yet as precise as that derived using metal-poor galaxies, but our method has the potential to offer a competitive test of physics beyond the standard model during Big Bang nucleosynthesis.},
journal = {Nat. Astron.}
}
@article{Corsaro.DeRidder.ea2015,
title = {High-Precision Acoustic Helium Signatures in 18 Low-Mass Low-Luminosity Red Giants: {{Analysis}} from More than Four Years of {{{\emph{Kepler}}}} Observations{$\star$}},
shorttitle = {High-Precision Acoustic Helium Signatures in 18 Low-Mass Low-Luminosity Red Giants},
author = {Corsaro, E. and De Ridder, J. and Garc{\'i}a, R. A.},
year = {2015},
month = jun,
volume = {578},
pages = {A76},
issn = {0004-6361, 1432-0746},
doi = {10.1051/0004-6361/201525922},
abstract = {Context. High-precision frequencies of acoustic modes in red giant stars are now available thanks to the long observing length and high quality of the light curves provided by the NASA Kepler mission, thus allowing the interior of evolved cool low-mass stars to be probed with an unprecedented level of detail.},
journal = {A\&A},
keywords = {asteroseismology,helium,statistics},
language = {en}
}
@book{Cox.Giuli1968,
title = {Principles of Stellar Structure},
author = {Cox, J. P. and Giuli, R. T.},
year = {1968}
}
@article{Cyburt.Fields.ea2016,
title = {Big Bang Nucleosynthesis: {{Present}} Status},
shorttitle = {Big Bang Nucleosynthesis},
author = {Cyburt, Richard H. and Fields, Brian D. and Olive, Keith A. and Yeh, Tsung-Han},
year = {2016},
month = jan,
volume = {88},
pages = {015004},
issn = {0034-6861},
doi = {10.1103/RevModPhys.88.015004},
abstract = {Big bang nucleosynthesis (BBN) describes the production of the lightest nuclides via a dynamic interplay among the four fundamental forces during the first seconds of cosmic time. A brief overview of the essentials of this physics is given, and new calculations presented of light-element abundances through 6Li and 7Li, with updated nuclear reactions and uncertainties including those in the neutron lifetime. Fits are provided for these results as a function of baryon density and of the number of neutrino flavors N{$\nu$}. Recent developments are reviewed in BBN, particularly new, precision Planck cosmic microwave background (CMB) measurements that now probe the baryon density, helium content, and the effective number of degrees of freedom Neff. These measurements allow for a tight test of BBN and cosmology using CMB data alone. Our likelihood analysis convolves the 2015 Planck data chains with our BBN output and observational data. Adding astronomical measurements of light elements strengthens the power of BBN. A new determination of the primordial helium abundance is included in our likelihood analysis. New D/H observations are now more precise than the corresponding theoretical predictions and are consistent with the standard model and the Planck baryon density. Moreover, D/H now provides a tight measurement of N{$\nu$} when combined with the CMB baryon density and provides a 2 {$\sigma$} upper limit N{$\nu<$}3.2 . The new precision of the CMB and D/H observations together leaves D/H predictions as the largest source of uncertainties. Future improvement in BBN calculations will therefore rely on improved nuclear cross-section data. In contrast with D/H and 4He, 7Li predictions continue to disagree with observations, perhaps pointing to new physics. This paper concludes with a look at future directions including key nuclear reactions, astronomical observations, and theoretical issues.},
journal = {Rev. Mod. Phys.}
}
@article{Das.Sanders2019,
title = {{{MADE}}: A Spectroscopic Mass, Age, and Distance Estimator for Red Giant Stars with {{Bayesian}} Machine Learning},
shorttitle = {{{MADE}}},
author = {Das, Payel and Sanders, Jason L.},
year = {2019},
month = mar,
volume = {484},
pages = {294},
doi = {10.1093/mnras/sty2776},
abstract = {We present a new approach (MADE) that generates mass, age, and distance estimates of red giant stars from a combination of astrometric, photometric, and spectroscopic data. The core of the approach is a Bayesian artificial neural network (ANN) that learns from and completely replaces stellar isochrones. The ANN is trained using a sample of red giant stars with mass estimates from asteroseismology. A Bayesian isochrone pipeline uses the astrometric, photometric, spectroscopic, and asteroseismology data to determine posterior distributions for the training outputs: mass, age, and distance. Given new inputs, posterior predictive distributions for the outputs are computed, taking into account both input uncertainties, and uncertainties in the ANN parameters. We apply MADE to 10 000 red giants in the overlap between the 14th data release from the APO Galactic Evolution Experiment (APOGEE) and the Tycho-Gaia astrometric solution (TGAS). The ANN is able to reduce the uncertainty on mass, age, and distance estimates for training-set stars with high output uncertainties allocated through the Bayesian isochrone pipeline. The fractional uncertainties on mass are \< 10 per cent and on age are between 10 to 25 per cent. Moreover, the time taken for our ANN to predict masses, ages, and distances for the entire catalogue of APOGEE-TGAS stars is of a similar order of the time taken by the Bayesian isochrone pipeline to run on a handful of stars. Our resulting catalogue clearly demonstrates the expected thick- and thin-disc components in the [M/H]-[{$\alpha$}/M] plane, when examined by age.},
journal = {MNRAS},
keywords = {asteroseismology,machine learning,neural network,red giants,spectroscopy},
language = {en}
}
@article{Davies.SilvaAguirre.ea2016,
title = {Oscillation Frequencies for 35 {{{\emph{Kepler}}}} Solar-Type Planet-Hosting Stars Using {{Bayesian}} Techniques and Machine Learning},
author = {Davies, G. R. and Silva Aguirre, V. and Bedding, T. R. and Handberg, R. and Lund, M. N. and Chaplin, W. J. and Huber, D. and White, T. R. and Benomar, O. and Hekker, S. and Basu, S. and Campante, T. L. and {Christensen-Dalsgaard}, J. and Elsworth, Y. and Karoff, C. and Kjeldsen, H. and Lundkvist, M. S. and Metcalfe, T. S. and Stello, D.},
year = {2016},
month = feb,
journal = {\mnras},
volume = {456},
number = {2},
pages = {2183--2195},
issn = {0035-8711, 1365-2966},
doi = {10.1093/mnras/stv2593},
urldate = {2019-10-15},
abstract = {Kepler has revolutionized our understanding of both exoplanets and their host stars. Asteroseismology is a valuable tool in the characterization of stars and Kepler is an excellent observing facility to perform asteroseismology. Here we select a sample of 35 Kepler solar-type stars which host transiting exoplanets (or planet candidates) with detected solar-like oscillations. Using available Kepler short cadence data up to Quarter 16 we create power spectra optimized for asteroseismology of solar-type stars. We identify modes of oscillation and estimate mode frequencies by `peak bagging' using a Bayesian Markov Chain Monte Carlo framework. In addition, we expand the methodology of quality assurance using a Bayesian unsupervised machine learning approach. We report the measured frequencies of the modes of oscillation for all 35 stars and frequency ratios commonly used in detailed asteroseismic modelling. Due to the high correlations associated with frequency ratios we report the covariance matrix of all frequencies measured and frequency ratios calculated. These frequencies, frequency ratios, and covariance matrices can be used to obtain tight constraint on the fundamental parameters of these planet-hosting stars.},
langid = {english},
keywords = {asteroseismology,planetary systems,planets and satellites: fundamental parameters,stars: evolution,stars: fundamental parameters,stars: oscillations}
}
@article{Davies.Chaplin.ea2015,
title = {Asteroseismic Inference on Rotation, Gyrochronology and Planetary System Dynamics of 16 {{Cygni}}},
author = {Davies, G. R. and Chaplin, W. J. and Farr, W. M. and Garc{\'i}a, R. A. and Lund, M. N. and Mathis, S. and Metcalfe, T. S. and Appourchaux, T. and Basu, S. and Benomar, O. and Campante, T. L. and Ceillier, T. and Elsworth, Y. and Handberg, R. and Salabert, D. and Stello, D.},
year = {2015},
month = jan,
volume = {446},
pages = {2959--2966},
issn = {0035-8711},
doi = {10.1093/mnras/stu2331},
abstract = {The solar analogues 16 Cyg A and B are excellent asteroseismic targets in the Kepler field of view and together with a red dwarf and a Jovian planet form an interesting system. For these more evolved Sun-like stars we cannot detect surface rotation with the current Kepler data but instead use the technique of asteroseimology to determine rotational properties of both 16 Cyg A and B. We find the rotation periods to be 23.8\^\{+1.5\}\_\{-1.8\} and 23.2\^\{+11.5\}\_\{-3.2\} d, and the angles of inclination to be 56\^\{+6\}\_\{-5\}\textdegree{} and 36\^\{+17\}\_\{-7\}\textdegree, for A and B, respectively. Together with these results we use the published mass and age to suggest that, under the assumption of a solar-like rotation profile, 16 Cyg A could be used when calibrating gyrochronology relations. In addition, we discuss the known 16 Cyg B star-planet eccentricity and measured low obliquity which is consistent with Kozai cycling and tidal theory.},
journal = {MNRAS},
keywords = {asteroseismology,planet-star interactions,stars: oscillations,stars: rotation,stellar rotation}
}
@article{Davies.Miglio2016,
title = {Asteroseismology of Red Giants: {{From}} Analysing Light Curves to Estimating Ages},
shorttitle = {Asteroseismology of Red Giants},
author = {Davies, G. R. and Miglio, A.},
year = {2016},
month = sep,
volume = {337},
pages = {774},
issn = {0004-6337},
doi = {10.1002/asna.201612371},
abstract = {Asteroseismology has started to provide constraints on stellar properties that will be essential to accurately reconstruct the history of the Milky Way. Here we look at the information content in data sets representing current and future space missions (CoRoT, Kepler, K2, TESS, and PLATO) for red giant stars. We describe techniques for extracting the information in the frequency power spectrum and apply these techniques to Kepler data sets of different observing length to represent the different space missions. We demonstrate that for KIC 12008916, a low-luminosity red giant branch star, we can extract useful information from all data sets, and for all but the shortest data set we obtain good constraint on the g-mode period spacing and core rotation rates. We discuss how the high precision in these parameters will constrain the stellar properties of stellar radius, distance, mass and age. We show that high precision can be achieved in mass and hence age when values of the g-mode period spacing are available. We caution that tests to establish the accuracy of asteroseismic masses and ages are still ``work in progress''.},
journal = {Astron. Nachrichten},
keywords = {ages,analysis,asteroseismology,red giants,stars: fundamental parameters,stars: individual (KIC 12008916),stars: interiors,stars: oscillations}
}
@article{Deubner.Gough1984,
title = {Helioseismology: {{Oscillations}} as a {{Diagnostic}} of the {{Solar Interior}}},
shorttitle = {Helioseismology},
author = {Deubner, Franz-Ludwig and Gough, Douglas},
year = {1984},
volume = {22},
pages = {593--619},
doi = {10.1146/annurev.aa.22.090184.003113},
abstract = {Contents: (1) Introduction. (2) Resonant cavities in the Sun. (3) Observational methods. (4) High-degree modes: Solar structure. Subphotospheric velocities. (5) Low-degree modes: High-order p modes. Gravity modes. Solar structure. (6) Five-minute modes of intermediate degree. (7) Limb observations. (8) Problems for the immediate future. Appendix: Classification of stellar oscillations.},
journal = {ARA\&A}
}
@article{Deubner.Ulrich.ea1979,
title = {Solar P-Mode Oscillations as a Tracer of Radial Differential Rotation},
author = {Deubner, F.-L. and Ulrich, R. K. and Rhodes, Jr., E. J.},
year = {1979},
month = feb,
volume = {72},
pages = {177--185},
issn = {0004-6361},
abstract = {Photoelectric observations of solar p-modes obtained with improved wavenumber and frequency resolution are presented. The observations are compared with model calculations of the p-modes, and the degree of spatial and temporal coherence of the observed wave pattern is investigated. It is found that the p-mode oscillations pervade the visible surface of the sun with a high degree of coherence in space and time, so that the whole complex pattern of standing waves with its nodes and antinodes can be regarded as a fixed pattern corotating with the solar surface layers. The p-modes are introduced as a tracer of solar rotational flow velocities. The equatorial differential rotation is estimated as a function of effective depth on the basis of the theoretical contribution functions for the p-modes recently derived by Ulrich et al. (1978). The results strongly indicate that the angular speed of rotation is not uniform even in the relatively shallow layer extending about 20,000 km below the photosphere.},
journal = {A\&A},
keywords = {Angular Velocity,Atmospheric Models,Chromosphere,Flow Velocity,Solar Oscillations,Solar Rotation,Standing Waves,Vibration Mode}
}
@article{Dillon.Langmore.ea2017,
ids = {Dillon.Langmore.ea2017a},
title = {{{TensorFlow Distributions}}},
author = {Dillon, Joshua V. and Langmore, Ian and Tran, Dustin and Brevdo, Eugene and Vasudevan, Srinivas and Moore, Dave and Patton, Brian and Alemi, Alex and Hoffman, Matt and Saurous, Rif A.},
year = {2017},
month = nov,
abstract = {The TensorFlow Distributions library implements a vision of probability theory adapted to the modern deep-learning paradigm of end-to-end differentiable computation. Building on two basic abstractions, it offers flexible building blocks for probabilistic computation. Distributions provide fast, numerically stable methods for generating samples and computing statistics, e.g., log density. Bijectors provide composable volume-tracking transformations with automatic caching. Together these enable modular construction of high dimensional distributions and transformations not possible with previous libraries (e.g., pixelCNNs, autoregressive flows, and reversible residual networks). They are the workhorse behind deep probabilistic programming systems like Edward and empower fast black-box inference in probabilistic models built on deep-network components. TensorFlow Distributions has proven an important part of the TensorFlow toolkit within Google and in the broader deep learning community.},
adsnote = {Provided by the SAO/NASA Astrophysics Data System},
archiveprefix = {arXiv},
eid = {arXiv:1711.10604},
eprint = {1711.10604},
eprinttype = {arxiv},
journal = {ArXiv e-prints},
keywords = {Computer Science - Artificial Intelligence,Computer Science - Machine Learning,Computer Science - Programming Languages,Statistics - Machine Learning},
primaryclass = {cs.LG}
}
@article{Dotter.Conroy.ea2017,
title = {The {{Influence}} of {{Atomic Diffusion}} on {{Stellar Ages}} and {{Chemical Tagging}}},
author = {Dotter, Aaron and Conroy, Charlie and Cargile, Phillip and Asplund, Martin},
year = {2017},
month = may,
volume = {840},
pages = {99},
doi = {10.3847/1538-4357/aa6d10},
abstract = {In the era of large stellar spectroscopic surveys, there is an emphasis on deriving not only stellar abundances but also the ages for millions of stars. In the context of Galactic archeology, stellar ages provide a direct probe of the formation history of the Galaxy. We use the stellar evolution code MESA to compute models with atomic diffusion\textemdash with and without radiative acceleration\textemdash and extra mixing in the surface layers. The extra mixing consists of both density-dependent turbulent mixing and envelope overshoot mixing. Based on these models we argue that it is important to distinguish between initial, bulk abundances (parameters) and current, surface abundances (variables) in the analysis of individual stellar ages. In stars that maintain radiative regions on evolutionary timescales, atomic diffusion modifies the surface abundances. We show that when initial, bulk metallicity is equated with current, surface metallicity in isochrone age analysis, the resulting stellar ages can be systematically overestimated by up to 20\%. The change of surface abundances with evolutionary phase also complicates chemical tagging, which is the concept that dispersed star clusters can be identified through unique, high-dimensional chemical signatures. Stars from the same cluster, but in different evolutionary phases, will show different surface abundances. We speculate that calibration of stellar models may allow us to estimate not only stellar ages but also initial abundances for individual stars. In the meantime, analyzing the chemical properties of stars in similar evolutionary phases is essential to minimize the effects of atomic diffusion in the context of chemical tagging.},
journal = {ApJ},
keywords = {stars: abundances,stars: evolution}
}
@article{Dotter2016,
title = {{{MESA Isochrones}} and {{Stellar Tracks}} ({{MIST}}) 0: {{Methods}} for the {{Construction}} of {{Stellar Isochrones}}},