Articles

A MULTIVARIATE FIT LUMINOSITY FUNCTION AND WORLD MODEL FOR LONG GAMMA-RAY BURSTS

Published 2013 March 15 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Amir Shahmoradi 2013 ApJ 766 111 DOI 10.1088/0004-637X/766/2/111

0004-637X/766/2/111

ABSTRACT

It is proposed that the luminosity function, the rest-frame spectral correlations, and distributions of cosmological long-duration (Type-II) gamma-ray bursts (LGRBs) may be very well described as a multivariate log-normal distribution. This result is based on careful selection, analysis, and modeling of LGRBs' temporal and spectral variables in the largest catalog of GRBs available to date: 2130 BATSE GRBs, while taking into account the detection threshold and possible selection effects. Constraints on the joint rest-frame distribution of the isotropic peak luminosity (Liso), total isotropic emission (Eiso), the time-integrated spectral peak energy (Ep, z), and duration (T90, z) of LGRBs are derived. The presented analysis provides evidence for a relatively large fraction of LGRBs that have been missed by the BATSE detector with Eiso extending down to ∼1049 erg and observed spectral peak energies (Ep) as low as ∼5 keV. LGRBs with rest-frame duration T90, z ≲ 1 s or observer-frame duration T90 ≲ 2 s appear to be rare events (≲ 0.1% chance of occurrence). The model predicts a fairly strong but highly significant correlation (ρ = 0.58 ± 0.04) between Eiso and Ep, z of LGRBs. Also predicted are strong correlations of Liso and Eiso with T90, z and moderate correlation between Liso and Ep, z. The strength and significance of the correlations found encourage the search for underlying mechanisms, though undermine their capabilities as probes of dark energy's equation of state at high redshifts. The presented analysis favors—but does not necessitate—a cosmic rate for BATSE LGRBs tracing metallicity evolution consistent with a cutoff Z/Z ∼ 0.2–0.5, assuming no luminosity–redshift evolution.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Ever since the discovery of the first gamma-ray burst (GRB) by the Vela satellites in 1967 (Klebesadel et al. 1973), there has been tremendous effort and attempts to constrain the energetics, luminosity function (LF), and the underlying mechanism responsible for these events. Early observations by the Konus (Mazets & Golenetskii 1981) and Ginga (Fenimore et al. 1988; Nishimura 1988) gamma/X-ray instruments suggested a possible link between GRBs and neutron stars with output energy ranges of the order of ∼1040 erg. With the launch of the Compton Gamma-Ray Observatory (CGRO), the Burst And Transient Source Experiment (BATSE) on board CGRO dramatically changed the understanding of GRBs. While previous catalogs (e.g., Atteia et al. 1987) indicated an isotropic distribution of GRB sources, the BATSE observations extended this isotropy down to the weakest bursts. The non-homogenous (e.g., Fenimore et al. 1993) and isotropic spacial event distribution (e.g., Meegan et al. 1992; Briggs 1993; Fishman et al. 1994) provided, for the first time, strong support for a cosmological versus galactic origin of GRBs, undermining neutron stars in the local universe as the potential candidates for some—if not all—classes of gamma-ray events. Furthermore, the joint duration–hardness distribution of GRBs provided a direct evidence for at least two separate classes of GRBs: long-soft versus short-hard (e.g., Kouveliotou et al. 1993, see also Figure 1 here).

The possibility of a cosmological origin for GRBs indicated an enormous output energy on the order of ∼1051 erg (e.g., Dermer 1992). Nevertheless, an accurate description of the GRB LF also required knowledge of the GRB cosmic rate, information that could not be extracted from BATSE observations alone. This became possible only with the launch of the Italian-Dutch X-ray satellite BeppoSax (Boella et al. 1997) and the identification of the first GRB with firmly measured cosmological redshift (Metzger et al. 1997) that marked the beginning of the afterglow era in the field of GRBs. The launch of the Swift satellite (Gehrels et al. 2004) was another milestone that revolutionized the study of GRBs by facilitating the X-ray afterglow observations (Burrows et al. 2005) and further ground-based follow-ups for redshift measurement.

Alongside the observational triumphs over a few decades, several theoretical models have stood up against the rivals based on the available evidence and GRB data. Most prominently, the Collapsar model (e.g., Woosley 1993) has been relatively successful in linking the long-duration class of gamma-ray bursts (LGRBs) to the final stages in the lives of massive stars, while the short-duration class of bursts (SGRBs) is generally attributed to the coalescence of compact binary systems (e.g., Paczynski 1986; Nakar 2007 and references therein). The two classes of SGRBs and LGRBs in this work correspond to Type-I and Type-II GRBs, respectively, according to the physical classification scheme of Zhang et al. (2007) and Bloom et al. (2008). Further refinement of the potential candidates, as the progenitors and the emission mechanism for both classes, requires more rigorous analysis of observational data in all possible energy frequencies. In particular, the prompt gamma-ray emission of LGRBs has been subject of intense observational and theoretical studies.

Beginning with BATSE observations, numerous authors have examined the prompt emission of LGRBs searching for potential underlying correlations among the spectral parameters (e.g., Nemiroff et al. 1994; Fenimore et al. 1995; Mallozzi et al. 1995; Petrosian & Lee 1996; Brainerd 1997; Dezalay et al. 1997; Petrosian et al. 1999; Lloyd et al. 2000; Norris et al. 2005). The lack of known redshifts for BATSE events and poor knowledge of LGRB cosmic rates, however, strongly limited the prediction power of such analyses. Instead, the first direct evidence for potential correlations and constraints on the distributions of the prompt-emission spectral parameters came with a few LGRBs detected by BATSE, BeppoSax, IPN, or HETE-II satellites with measured redshifts (e.g., Reichart & Lamb 2001; Amati et al. 2002; Ghirlanda et al. 2004; Yonetoku et al. 2004) and was further developed by the inclusion of Swift LGRBs (e.g., Schaefer 2007; Gehrels et al. 2009). Such findings, however, have been criticized for relying primarily on a handful of events with high signal-to-noise ratio (S/N) required for spectral analysis with afterglows sufficiently bright for redshift measurement, arguing that the proposed joint distributions of the spectral properties do not represent the entire underlying population of LGRBs (e.g., Band & Preece 2005; Nakar & Piran 2005; Li 2007; Butler et al. 2007, 2009; Shahmoradi & Nemiroff 2009, 2010, 2011). Responding to criticisms, attempts were made to model the effects of different gamma-ray instruments' detection thresholds and the limiting effects of spectral analysis (e.g., Ghirlanda et al. 2008; Nava et al. 2008; see Shahmoradi & Nemiroff 2011 for a review of relevant literature).

Despite significant progress, difficulties in modeling the complex effects of detector threshold on the multivariate distribution of the prompt-emission spectral properties and the lack of a sufficiently large sample of uniformly detected LGRBs has led the community to focus on individual spectral variables, most importantly the LF (e.g., Petrosian 1993; Schmidt 1999, 2001, 2009; Kommers et al. 2000; Kumar & Piran 2000; Band 2001; Porciani & Madau 2001; Sethi & Bhargavi 2001; Stern et al. 2002; Guetta et al. 2005; Salvaterra & Chincarini 2007; Salvaterra et al. 2009; Campisi et al. 2010; Wanderman & Piran 2010; Salvaterra et al. 2012). Recently, Butler et al. (2010, hereafter B10) presented an elaborate multivariate analysis of Swift LGRBs, including the potential correlations among three temporal and spectral variables: total isotropic energy emission (Eiso (erg)), a definition of duration (Tr45 (s)), and the spectral peak energy (Ep, z (keV)) of the bursts. Focusing their analysis on the EisoEp, zTr45 interrelation, B10 find a strong and significant correlation—but with a broad scatter—between the isotropic emission and the peak energy of the Swift LGRBs. Also realized by B10 is the possibility of a positive—and perhaps strong—correlation between the duration and the total energy output of the bursts. Moreover, to accommodate the potential existence of a large population of sub-luminous events, a broken power-law LF is used by B10.

It is known that the energy budget of GRBs must be limited and a turnover in the LF at the low-luminosity tail of the population is expected. However, the choice of the broken power law as the candidate LF is justified by the fact that current observational data cannot constrain this turnover point, which is expected to be far below the detection threshold of current gamma-ray instruments. The existence of a turnover point in the LF can have important clues for the underlying physics of LGRBs. The low-luminosity tail of the LF in such a case would be the result of the convolution of the stellar mass distribution with the LGRB rate as a function of the properties of massive dying stars, imposed by the mechanism.

Motivated by the search for the potential shape of the LGRB LF at low luminosity and the flurry of recent reports on the possible existence of correlations among LGRBs' temporal and spectral variables, the author of this paper has considered a wide variety of different statistical models that could incorporate and explain all observed correlations and spectral distributions, while preserving the prediction power of the model for the spectral properties of a potentially large fraction of LGRBs that could go undetected by current gamma-ray detectors. Here it will be shown that it is indeed possible to construct an LGRB world model capable of describing most (if not all) prompt-emission spectral and temporal properties observed in the current data catalogs. Toward this, the presented analysis is mainly focused on the largest catalog of GRBs available to date: the BATSE catalog of 2130 GRBs (Paciesas et al. 1999). The author shows that there is still an enormous amount of information buried in BATSE observations that need to be dug out by GRB researchers.

In the following sections, an example of such data mining on BATSE catalog will be presented: Section 2.1 presents an elaborate method for classifying GRBs into the two known subclasses of SGRBs and LGRBs. In Section 2.2 an LGRB world model capable of describing BATSE observations is presented, followed by a discussion of the procedure for fitting the model to data in Section 2.3. Univariate and multivariate goodness-of-fit (GoF) tests are performed to ensure that the model predicts BATSE data accurately (Section 2.4). The results from model fitting and implications on the multivariate distribution of the prompt-emission spectral properties are discussed in Section 3. A summary of the analysis and important outcomes of the proposed LGRB world model will be presented in Section 4.

2. GRB WORLD MODEL

2.1. Sample Selection

Depending on their detection criteria, some gamma-ray detectors might facilitate detection of one class of bursts over the others. For example, the specific detector sensitivity of the Burst Alert Telescope (BAT) on board the Swift satellite (Gehrels et al. 2004) results in better detections of LGRBs over SGRBs (e.g., Band 2003, 2006). Therefore, a simple GRB classification method, such as a cutoff in the observed duration distributions of GRBs (T90 ∼ 3 s) generally results in LGRB samples that are minimally contaminated by SGRBs (e.g., B10). Compared to BAT, however, BATSE Large Area Detectors (LADs) had an increased relative sensitivity to SGRBs (e.g., Band 2003). Knowing that many temporal and spectral properties of LGRBs and SGRBs, most importantly T90, overlap, identification of LGRBs in the BATSE catalog solely based on their T90 will likely result in a significant number of misclassifications (e.g., Figure 1, center panel).

Figure 1.

Figure 1. Classification of 1966 BATSE LGRBs according to the most suitable clustering algorithm and set of GRB variables (Section 2.1): fuzzy C-means classification on Ep and T90. Red and blue colors represent LGRB and SGRB classes, respectively, in all three plots. Top: the joint T90Ep distribution. The uncertainties in LGRBs are derived from the empirical Bayes model discussed in Appendix C. Center: T90 distribution. Bottom: Ep distribution. Ep estimates are taken from Shahmoradi & Nemiroff (2010). Compare this plot to the plot of Figure 13 of Shahmoradi & Nemiroff (2010), where the entire univariate Ep distributions of BATSE GRBs were fit by a two-component Gaussian mixture.

Standard image High-resolution image

Thus, to ensure correct analysis of a long-duration class of GRBs, it is first necessary to collect the least biased sample of events, all belonging to LGRB category. The word "bias" here refers to the bias that might be introduced when using the traditional definition of GRB classes, based on a sharp cutoff on the duration variable T90 (Kouveliotou et al. 1993), as it is generally assumed by many GRB researchers (e.g., Guetta et al. 2005; B10; Campisi et al. 2010; Wanderman & Piran 2010) or not assumed or explicitly discussed (e.g., Salvaterra & Chincarini 2007; Salvaterra et al. 2009, 2012). This goal is achieved by testing extensive varieties of classification and clustering methods, most importantly, the fuzzy clustering algorithms of Rousseeuw et al. (1996) and the method of fuzzy C-means (Dunn 1973; Bezdek 1981). Each BATSE-catalog GRB is assigned a probability (i.e., class coefficient) of belonging to the LGRB (versus SGRB) population according to the choice of clustering algorithm and the set of GRB variables used. This can be any combination of the peak flux (P50–300 photons s−1 cm−2) in the BATSE detection energy range, in three different timescales: 64, 256, and 1024 ms; bolometric fluence (Sbol(erg cm−2)); the spectral peak energy (Ep (keV)) estimates by Shahmoradi & Nemiroff (2010); duration (T90 (s)); and the fluence-to-peak-flux ratio (FPR (s)). Then GRBs with LGRB class coefficient >0.5 are flagged as long-duration class bursts. Overall, the fuzzy C-means classification method with the two GRB variables Ep and T90 is preferred over other clustering methods and sets of GRB variables (cf. Appendix A). This leads to the selection of 1376 events as LGRBs out of the 1966 BATSE GRBs having measured temporal and spectral parameters mentioned above.1

As a further safety check to ensure minimal contamination of the sample by SGRBs, the light curves of 291 bursts among 1966 BATSE GRBs with LGRB class coefficients in the range of 0.3–0.7 are visually inspected in the four main energy channels of BATSE LADs. This leads to reclassification of 17 events (originally flagged as LGRB by the clustering algorithm) to potentially SGRB or soft gamma repeater (SGR) events, and reclassification of seven events (originally flagged as SGRB by the clustering algorithm) to LGRBs. The result is a reduction in the size of the original LGRB sample from 1376 to 1366 (Table 1). It is notable that the inclusion of the uncertainties on the two GRB variables T90 and Ep turns out to not have significant effects on the derived samples of the two GRB classes discussed above. Also, a classification based on T50 instead of T90 results in about the same samples for the two GRB classes with only a negligible difference of ∼0.7%.

Table 1. 1366 BATSE Catalog Triggers Classified as LGRBs

Trigger Trigger Trigger Trigger Trigger Trigger Trigger Trigger Trigger Trigger Trigger Trigger Trigger Trigger
105 107 109 110 111 114 121 130 133 143 148 160 171 179
204 211 214 219 222 223 226 228 235 237 249 257 288 332
351 394 398 401 404 408 414 451 465 467 469 472 473 493
501 516 526 540 543 548 549 559 563 577 591 594 606 630
647 658 659 660 673 676 678 680 685 686 690 692 704 717
741 752 753 755 761 764 773 795 803 815 816 820 824 825
829 840 841 869 907 914 927 938 946 973 999 1009 1036 1039
1042 1046 1085 1086 1087 1114 1120 1122 1123 1125 1126 1141 1145 1148
1150 1152 1153 1156 1157 1159 1167 1190 1192 1196 1197 1200 1204 1213
1218 1221 1235 1244 1279 1288 1291 1298 1303 1306 1318 1382 1384 1385
1390 1396 1406 1416 1419 1425 1432 1439 1440 1446 1447 1449 1452 1456
1458 1467 1468 1472 1492 1515 1533 1540 1541 1551 1552 1558 1559 1561
1567 1574 1578 1579 1580 1586 1590 1601 1604 1606 1609 1611 1614 1623
1625 1626 1628 1642 1646 1651 1652 1653 1655 1656 1657 1660 1661 1663
1664 1667 1676 1687 1693 1700 1701 1704 1711 1712 1714 1717 1730 1731
1733 1734 1740 1742 1806 1807 1815 1819 1830 1883 1885 1886 1922 1924
1956 1967 1974 1982 1989 1993 1997 2018 2019 2035 2047 2053 2061 2067
2069 2070 2074 2077 2079 2080 2081 2083 2087 2090 2093 2101 2102 2105
2106 2110 2111 2112 2114 2119 2122 2123 2129 2133 2138 2140 2143 2148
2149 2151 2152 2156 2181 2187 2188 2189 2190 2191 2193 2197 2202 2203
2204 2207 2211 2213 2219 2228 2230 2232 2233 2240 2244 2252 2253 2254
2267 2276 2277 2287 2298 2304 2306 2309 2310 2311 2315 2316 2321 2324
2325 2328 2329 2340 2344 2345 2346 2347 2349 2362 2367 2371 2373 2375
2380 2381 2383 2385 2387 2391 2392 2393 2394 2405 2419 2423 2428 2429
2430 2432 2435 2436 2437 2438 2440 2441 2442 2443 2446 2447 2450 2451
2452 2453 2458 2460 2472 2476 2477 2482 2484 2495 2496 2500 2505 2508
2510 2511 2515 2519 2522 2528 2530 2533 2537 2541 2551 2560 2569 2570
2581 2586 2589 2593 2600 2603 2606 2608 2610 2611 2619 2620 2628 2634
2636 2640 2641 2660 2662 2663 2664 2665 2671 2677 2681 2688 2691 2695
2696 2697 2700 2703 2706 2709 2711 2719 2725 2727 2736 2749 2750 2751
2753 2767 2770 2774 2775 2780 2790 2793 2797 2798 2812 2815 2825 2830
2831 2843 2848 2850 2852 2853 2855 2856 2857 2862 2863 2864 2877 2880
2889 2890 2891 2897 2898 2900 2901 2913 2916 2917 2919 2922 2924 2925
2927 2929 2931 2932 2944 2945 2947 2948 2950 2951 2953 2958 2961 2980
2984 2985 2986 2990 2992 2993 2994 2996 2998 3001 3003 3005 3011 3012
3015 3017 3026 3028 3029 3032 3035 3040 3042 3055 3056 3057 3067 3068
3070 3071 3072 3074 3075 3076 3080 3084 3085 3088 3091 3093 3096 3100
3101 3102 3103 3105 3109 3110 3115 3119 3120 3127 3128 3129 3130 3131
3132 3134 3135 3136 3138 3139 3141 3142 3143 3153 3156 3159 3166 3167
3168 3171 3174 3177 3178 3193 3212 3217 3220 3227 3229 3237 3238 3241
3242 3245 3246 3247 3255 3256 3257 3259 3267 3269 3276 3279 3283 3284
3287 3290 3292 3301 3306 3307 3319 3320 3321 3322 3324 3330 3336 3339
3345 3347 3350 3351 3352 3356 3358 3364 3369 3370 3378 3403 3405 3407
3408 3415 3416 3436 3439 3448 3458 3465 3471 3472 3480 3481 3485 3486
3488 3489 3491 3493 3503 3505 3509 3511 3512 3514 3515 3516 3523 3527
3528 3552 3567 3569 3588 3593 3598 3608 3618 3634 3637 3648 3649 3654
3655 3658 3662 3663 3664 3671 3717 3733 3740 3745 3765 3766 3768 3771
3773 3776 3779 3788 3792 3800 3801 3805 3807 3811 3814 3815 3819 3840
3843 3853 3860 3869 3870 3871 3875 3879 3886 3890 3891 3892 3893 3899
3900 3901 3903 3905 3906 3908 3909 3912 3913 3914 3916 3917 3918 3924
3926 3929 3930 3935 3941 3954 4039 4048 4095 4146 4157 4216 4251 4312
4350 4368 4388 4556 4569 4653 4701 4710 4745 4814 4939 4959 5080 5255
5304 5305 5379 5387 5389 5407 5409 5411 5412 5415 5416 5417 5419 5420
5421 5423 5428 5429 5433 5434 5447 5450 5451 5454 5463 5464 5465 5466
5470 5472 5473 5474 5475 5476 5477 5478 5479 5480 5482 5483 5484 5486
5487 5489 5490 5492 5493 5494 5495 5497 5503 5504 5507 5508 5510 5512
5513 5515 5516 5517 5518 5523 5524 5526 5530 5531 5538 5539 5540 5541
5542 5545 5548 5551 5554 5555 5559 5563 5565 5566 5567 5569 5571 5572
5573 5574 5575 5581 5585 5589 5590 5591 5593 5594 5597 5601 5603 5604
5605 5606 5608 5610 5612 5614 5615 5617 5618 5621 5622 5624 5626 5627
5628 5632 5635 5637 5640 5644 5645 5646 5648 5654 5655 5667 5697 5704
5706 5713 5715 5716 5718 5719 5721 5723 5725 5726 5729 5731 5736 5773
5867 5890 5955 5983 5989 5995 6004 6082 6083 6090 6098 6100 6101 6102
6103 6104 6111 6113 6115 6118 6119 6124 6127 6128 6131 6137 6139 6141
6147 6151 6152 6154 6158 6159 6165 6167 6168 6176 6186 6188 6189 6190
6194 6198 6206 6222 6223 6225 6226 6227 6228 6233 6234 6241 6242 6243
6244 6249 6266 6267 6269 6270 6271 6272 6273 6274 6279 6280 6283 6285
6288 6295 6298 6300 6303 6304 6305 6306 6308 6309 6315 6317 6319 6320
6321 6322 6323 6328 6329 6330 6334 6335 6337 6339 6344 6345 6346 6349
6351 6353 6355 6369 6370 6375 6380 6388 6390 6395 6396 6397 6399 6400
6404 6405 6408 6409 6413 6414 6419 6422 6425 6435 6437 6440 6444 6446
6448 6450 6451 6453 6454 6472 6487 6489 6490 6498 6504 6519 6520 6521
6522 6523 6525 6528 6529 6531 6533 6534 6536 6538 6539 6544 6546 6550
6551 6552 6554 6557 6560 6564 6566 6576 6577 6578 6582 6583 6585 6587
6589 6590 6592 6593 6598 6600 6601 6602 6605 6610 6611 6613 6615 6616
6619 6620 6621 6622 6625 6629 6630 6631 6632 6642 6648 6649 6655 6657
6658 6665 6666 6670 6672 6673 6674 6676 6678 6683 6686 6694 6695 6698
6702 6707 6708 6720 6745 6762 6763 6764 6767 6774 6782 6796 6802 6814
6816 6830 6831 6853 6877 6880 6882 6884 6891 6892 6903 6911 6914 6917
6930 6935 6938 6963 6987 6989 7000 7012 7028 7030 7064 7087 7108 7110
7113 7116 7130 7147 7164 7167 7170 7172 7178 7183 7185 7191 7206 7207
7209 7213 7219 7228 7230 7247 7250 7255 7263 7285 7293 7295 7298 7301
7310 7318 7319 7322 7323 7328 7335 7343 7357 7358 7360 7369 7371 7374
7376 7377 7379 7381 7386 7387 7390 7403 7404 7429 7432 7433 7446 7451
7452 7457 7460 7464 7469 7475 7477 7481 7485 7486 7487 7488 7491 7493
7494 7497 7500 7502 7503 7504 7509 7515 7517 7518 7520 7523 7527 7528
7529 7532 7533 7535 7548 7549 7550 7551 7552 7560 7563 7564 7566 7567
7568 7573 7575 7576 7579 7580 7587 7588 7597 7598 7603 7604 7605 7606
7607 7608 7609 7614 7615 7617 7619 7625 7630 7635 7638 7642 7645 7648
7654 7656 7657 7660 7662 7677 7678 7683 7684 7688 7695 7701 7703 7705
7707 7711 7727 7729 7741 7744 7749 7750 7752 7762 7766 7769 7770 7780
7781 7785 7786 7788 7790 7794 7795 7798 7802 7803 7810 7818 7822 7825
7831 7835 7838 7840 7841 7843 7845 7858 7862 7868 7872 7884 7885 7886
7888 7900 7902 7903 7906 7918 7923 7924 7929 7932 7934 7936 7938 7942
7948 7954 7963 7968 7969 7973 7976 7984 7987 7989 7992 7994 7997 7998
8001 8004 8008 8009 8012 8019 8022 8026 8030 8036 8039 8045 8049 8050
8054 8059 8061 8062 8063 8064 8066 8073 8075 8084 8086 8087 8098 8099
8101 8102 8105 8110 8111 8112 8116 8121

Notes. Temporal and spectral data for these triggers are available in the BATSE 4B and Current Catalogs. The spectral peak energy (Ep) estimates of the above triggers and the rest of the 2130 BATSE Catalog GRBs are provided by Shahmoradi & Nemiroff (2010). The full conditional Ep probability density functions are available for download at https://sites.google.com/site/amshportal/research/aca/in-the-news/lgrb-world-model.

Download table as:  ASCIITypeset images: 1 2

2.2. Model Construction

The goal of the presented analysis is to derive a multivariate model that is capable of reproducing the observational data of the 1366 BATSE LGRBs. Examples of multivariate treatment of LGRB data are rare in GRB literature, with the most recent (and perhaps the only) example of such work presented by B10. Conversely, many authors have focused primarily on the univariate distribution of the spectral parameters, most importantly on the LF. A variety of univariate models have been proposed as the LGRB LF and fit to data by approximating the complex detector threshold as a step function (Schmidt 1999) or an efficiency grid (e.g., the four-interval efficiency modeling of Guetta et al. 2005) or by other approximation methods. A more accurate modeling of the LF, however, requires at least two LGRB observables incorporated in the model: the bolometric peak flux (Pbol) and the observed peak energy (Ep). The parameter Ep is required, since most gamma-ray detectors are photon counters, a quantity that depends on not only Pbol but also Ep of the burst. This leads to the requirement of using a bivariate distribution as the minimum acceptable model to begin with, for the purpose of constraining the LF. The choice of model can be almost anything (e.g., Kommers et al. 2000; Porciani & Madau 2001; Sethi & Bhargavi 2001; Schmidt 2009; Campisi et al. 2010; Wang & Dai 2011), since current theories of LGRB prompt emission do not set strong limits on the shape and range of the LF or any other LGRB spectral or temporal variables.

Here, the multivariate log-normal distribution is proposed as the simplest natural candidate model capable of describing data. The motivation behind this choice of model comes from the available observational data that closely resemble a joint multivariate log-normal distribution for the four most widely studied temporal and spectral parameters of LGRBs in the observer frame: Pbol, Sbol (bolometric fluence), Ep, T90: since most LGRBs originate from moderate redshifts z ∼ 1–3, a fact known thanks to Swift satellite (e.g., B10; Racusin et al. 2011), the convolution of these observer-frame parameters with the redshift distribution results in negligible variation in the shape of the rest-frame joint distribution of the same LGRB parameters. Therefore, the redshift-convoluted four-dimensional (4D) rest-frame distribution can be well approximated as a linear translation from the observer-frame parameter space to the rest-frame parameter space, keeping the shape of the distribution almost intact. This implies that the joint distribution of the intrinsic LGRB variables: the isotropic peak luminosity (Liso), the total isotropic emission (Eiso), the rest-frame time-integrated spectral peak energy (Ep, z), and the rest-frame duration (T90, z) might be indeed well described by a multivariate log-normal distribution.

In general, models with higher nonzero moments than the log-normal model can also be considered for fitting, such as a multivariate skew-lognormal (e.g., Azzalini 1985) or variants of multivariate stable distributions (e.g., Press 1972). This, however, requires fitting for a higher number of free parameters, which is practically impossible given BATSE data with no available redshift information.

Following the discussion above, the process of LGRB observation can be therefore considered as a non-homogeneous Poisson process whose mean rate parameter—the cosmic LGRB differential rate, $\mathcal {R}_{{\rm cosmic}}$—is the product of the differential comoving LGRB rate density $\dot{\zeta }(z)$ with a p = 4D log-normal probability density function (pdf), $\mathcal {LN}$, of four LGRB variables: Liso, Eiso, Ep, z, and T90, z, with location vector μ and the scale (i.e., covariance) matrix Σ,

Equation (1)

where the factor (1 + z) in the denominator accounts for cosmological time dilation and the comoving volume element per unit redshift, dV/dz, is

Equation (2)

with DL being the luminosity distance,

Equation (3)

assuming a flat ΛCDM cosmology, with parameters set to h = 0.70, ΩM = 0.27, and $\Omega _\Lambda =0.73$ (Jarosik et al. 2011). Here, C and H0 = 100 h (Km s−1 MPc−1) stand for the speed of light and the Hubble constant, respectively.

The 4D log-normal distribution of Equation (1), $\mathcal {LN}$, has an intimate connection to multivariate Gaussian distribution in the logarithmic space of LGRB observable parameters (cf. Appendix D).

One could generalize the LGRB rate of Equation (1) to incorporate a redshift evolution of the LGRB variables in the form of μi(z) = μ0, i + αilog (1 + z), (i = 1, ..., 4), where α has to be constrained by observational data. This is, however, impractical for BATSE data due to unknown redshifts, as the fitting results in degenerate values for α. Nevertheless, the multivariate analysis of Swift LGRBs presented by B10 strongly rejects the possibility of redshift evolution of the LF, a fact that further legitimizes the absence of redshift–luminosity evolution in Equation (1).

As for the comoving rate density $\dot{\zeta }(z)$, it is assumed that LGRBs trace the star formation rate (SFR) in the form of a piecewise power-law function of Hopkins & Beacom (2006, hereafter HB06):

Equation (4)

with parameters (z0, z1, γ0, γ1, γ2) set to best-fit values (0.97, 4.5, 3.4, −0.3, −7.8) of HB06, and also to the best values (0.993, 3.8, 3.3, 0.055, −4.46) of an updated SFR fit by Li (2008). Alternatively, the bias-corrected redshift distribution of LGRBs derived from Swift data (B10) with best-fit parameter values (0.97, 4.00, 3.14, 1.36, −2.92) can be employed as $\dot{\zeta }(z)$. This parameter set is consistent with an LGRB rate scenario tracing metallicity-corrected SFR with a cutoff Z/Z ∼ 0.2–0.5 (Figure 10 and Equation (8) in B10; Li 2008). The hypothesis of LGRB rate evolving with cosmic metallicity is both predicted by the Collapsar model of LGRBs (e.g., Woosley & Heger 2006) and supported by observations of LGRB host galaxies (e.g., Stanek et al. 2006; Levesque et al. 2010a), although the metallicity–rate connection and the presence of a sharp metallicity cutoff have been challenged by few recent host galaxy observations (e.g., Levesque et al. 2010c, 2010b) and possible unknown observational biases (e.g., Levesque 2012).

The cosmic LGRB rate, $\mathcal {R}_{{\rm cosmic}}$, in Equation (1), although quantified correctly, does not represent the observed rate ($\mathcal {R}_{{\rm obs}}$) of LGRBs detected by BATSE LADs, unless convolved with an accurate model of BATSE trigger efficiency, η, as a function of the burst redshift and rest-frame parameters, discussed in Appendix B,

Equation (5)

2.3. Model Fitting

Now, with a statistical model for the observed rate of LGRBs at hand (i.e., Equation (5)), the best-fit parameters can be obtained by the method of maximum likelihood. This is done by maximizing the likelihood function of the model, given the observational data, using a variant of the Metropolis–Hastings Markov chain Monte Carlo (MCMC) algorithm discussed in detail in Appendix C. As mentioned before, the fitting is performed for three redshift-distribution scenarios of HB06, B10, and Li (2008).

It is also known that the T90 of LGRBs are potentially subject to estimation biases. To ensure that the reported T90 of BATSE LGRBs do not bias the fitting results for the rest of the parameters, model fitting was also performed by considering only three spectral variables of BATSE LGRBs: the bolometric 1 s peak flux (Pbol), bolometric fluence (Sbol), and observed peak energy (Ep), excluding duration (T90, z) variable from the model, thus reducing the dimension of the model by one. Only after the fitting was performed did it become clear that the inclusion of the T90 of BATSE LGRBs in the fitting does not significantly affect the resulting best-fit parameters of the model. Therefore, only results from the full model fitting are presented here, as in Table 2.

Table 2. Mean Best-fit Parameters of LGRB World Model, for the Three Redshift Distribution Scenarios

Parameter HB06 Li (2008) B10
Redshift parameters (Equation (4))
z0 0.97 0.993 0.97
z1 4.5 3.8 4.00
γ0 3.4 3.3 3.14
γ1 −0.3 0.0549 1.36
γ2 −7.8 −4.46 −2.92
Location parameters
log (Liso) 51.35 ± 0.20 51.50 ± 0.19 51.73 ± 0.19
log (Eiso) 51.82 ± 0.20 51.94 ± 0.20 52.03 ± 0.21
log (Ep, z) 2.43 ± 0.05 2.47 ± 0.05 2.54 ± 0.05
log (T90, z) 0.99 ± 0.03 0.96 ± 0.03 0.80 ± 0.03
Scale parameters
$\log (\sigma _{L_{{\rm iso}}})$ −0.22 ± 0.06 −0.23 ± 0.06 −0.20 ± 0.05
$\log (\sigma _{E_{{\rm iso}}})$ −0.07 ± 0.03 −0.07 ± 0.03 −0.04 ± 0.03
$\log (\sigma _{E_{p,z}})$ −0.44 ± 0.02 −0.44 ± 0.02 −0.44 ± 0.02
$\log (\sigma _{T_{90,z}})$ −0.38 ± 0.01 −0.39 ± 0.01 −0.40 ± 0.01
Correlation coefficients
$\rho _{L_{{\rm iso}}-E_{{\rm iso}}}$ 0.93 ± 0.01 0.94 ± 0.01 0.96 ± 0.01
$\rho _{L_{{\rm iso}}-E_{p,z}}$ 0.47 ± 0.07 0.45 ± 0.07 0.44 ± 0.08
$\rho _{L_{{\rm iso}}-T_{90,z}}$ 0.52 ± 0.08 0.59 ± 0.09 0.75 ± 0.07
$\rho _{E_{{\rm iso}}-E_{p,z}}$ 0.58 ± 0.04 0.58 ± 0.04 0.59 ± 0.04
$\rho _{E_{{\rm iso}}-T_{90,z}}$ 0.63 ± 0.05 0.66 ± 0.05 0.74 ± 0.04
$\rho _{E_{p,z}-T_{90,z}}$ 0.34 ± 0.04 0.37 ± 0.04 0.50 ± 0.04
BATSE LGRB detection efficiency (Equation (A5))
μthresh −0.44 ± 0.02 −0.45 ± 0.02 −0.44 ± 0.02
log (σthresh) −0.88 ± 0.05 −0.90 ± 0.05 −0.88 ± 0.05

Notes. The full Markov chain sampling of the above parameters from the 16-dimensional parameter space of the likelihood function (Appendix C) is available for download at https://sites.google.com/site/amshportal/research/aca/in-the-news/lgrb-world-model for each of the three redshift distributions.

Download table as:  ASCIITypeset image

Due to lack of redshift information for BATSE GRBs, the resulting parameters of the model exhibit strong covariations with each other. This is illustrated in the example correlation matrix of the LGRB world model in Table 3. All location parameters appear to correlate strongly positively with each other, and so do the scale parameters. The location parameters however negatively correlate with the scale parameters, meaning that an increase in the average values of the rest-frame parameters reduces the half-width of the corresponding distributions of the variables. In general, it is also observed that the correlations among the four variables weaken with increasing the location parameters. An exception to this is the correlation of T90, z with Liso and Eiso which tends to increase with location parameters. Since an excess in the cosmic rates of LGRBs at high redshifts generally results in an increase in the values of location parameters, it can be said that "given BATSE LGRBs data, a higher rate for LGRBs at distant universe generally implies weaker EisoEp, z and LisoEp, z correlations and stronger covariation of T90, z with the three other parameters."

Table 3. Correlation Matrix of the Parameters of the LGRB World Model, for the Median Case of an LGRB Cosmic Rate Tracing Star Formation Rate in Li (2008)

Parameter log (Liso) log (Eiso) log (Ep, z) log (T90, z) $\log (\sigma _{L_{{\rm iso}}})$ $\log (\sigma _{E_{{\rm iso}}})$ $\log (\sigma _{E_{p,z}})$ $\log (\sigma _{T_{90,z}})$ $\rho _{L_{{\rm iso}}-E_{{\rm iso}}}$ $\rho _{L_{{\rm iso}}-E_{p,z}}$ $\rho _{L_{{\rm iso}}-T_{90,z}}$ $\rho _{E_{{\rm iso}}-E_{p,z}}$ $\rho _{E_{{\rm iso}}-T_{90,z}}$ $\rho _{E_{p,z}-T_{90,z}}$ μthresh log (σthresh)
log (Liso) 1.00 0.99 0.90 0.34 −0.91 −0.86 −0.59 −0.14 −0.10 −0.45 0.51 −0.52 0.45 0.05 −0.68 −0.44
log (Eiso)   1.00 0.92 0.42 −0.91 −0.90 −0.62 −0.15 −0.20 −0.54 0.45 −0.56 0.40 0.00 −0.67 −0.43
log (Ep, z)     1.00 0.38 −0.82 −0.84 −0.79 −0.15 −0.26 −0.77 0.40 −0.77 0.38 0.07 −0.61 −0.40
log (T90, z)       1.00 −0.37 −0.52 −0.32 −0.16 −0.52 −0.32 −0.48 −0.33 −0.50 −0.56 −0.17 −0.09
$\log (\sigma _{L_{{\rm iso}}})$         1.00 0.94 0.59 0.12 0.14 0.53 −0.53 0.57 −0.46 0.03 0.50 0.30
$\log (\sigma _{E_{{\rm iso}}})$           1.00 0.67 0.17 0.41 0.63 −0.33 0.66 −0.32 0.14 0.46 0.28
$\log (\sigma _{E_{p,z}})$             1.00 0.12 0.36 0.84 −0.22 0.84 −0.25 −0.06 0.37 0.24
$\log (\sigma _{T_{90,z}})$               1.00 −0.03 0.10 −0.09 0.11 0.00 0.03 0.11 0.07
$\rho _{L_{{\rm iso}}-E_{{\rm iso}}}$                 1.00 0.47 0.43 0.41 0.22 0.26 −0.01 −0.01
$\rho _{L_{{\rm iso}}-E_{p,z}}$                   1.00 −0.15 0.97 −0.21 −0.05 0.29 0.19
$\rho _{L_{{\rm iso}}-T_{90,z}}$                     1.00 −0.18 0.95 0.57 −0.32 −0.20
$\rho _{E_{{\rm iso}}-E_{p,z}}$                       1.00 −0.22 0.06 0.29 0.18
$\rho _{E_{{\rm iso}}-T_{90,z}}$                         1.00 0.64 −0.26 −0.17
$\rho _{E_{p,z}-T_{90,z}}$                           1.00 −0.09 −0.07
μthresh                             1.00 0.79
log (σthresh)                               1.00

Download table as:  ASCIITypeset image

2.4. Goodness-of-fit Tests

In any statistical fitting problem, perhaps more important than the model construction is to provide tests showing how good the model fit is to input data. For many univariate studies of the GRB LF, this is done by employing well-established statistical tests such as the Kolmogorov–Smirnov (K-S; e.g., Kolmogoroff 1941; Smirnov 1948) or Pearson's χ2 (e.g., Fisher 1924) tests. In the case of multivariate studies (e.g., B10), a combination of visual inspection of the fitting results, K-S test on the marginal, and bivariate distributions and variants of χ2 (e.g., likelihood ratio) tests have been used.

In general, univariate tests on the marginal distributions of multivariate fits provide only necessary—but not sufficient—evidence for a good multivariate fit. Alternatively one could assess the similarity by using nonparametric multivariate GoF tests. Such tests, although existing, have been rarely discussed and treated in statistics due to difficulties in the interpretation of the test statistic (e.g., Peacock 1983; Press et al. 1992; Justel et al. 1997). Ideally, one can always use the Pearson's χ2 GoF test for any multivariate distribution. However, for the special case of BATSE 1366 LGRBs, one would need an observed sample consisting of N ≫ 1366 observations to avoid serious instabilities that occur in χ2 tests due to small sample sizes (e.g., Cochran 1954).

To ensure a good fit to the observational data in all—and not only univariate—levels of the multivariate structure of data, an assessment of similarity can be obtained by scanning and comparing the model and data along their principal axes, in addition to univariate tests on the marginal distributions. Although statistically not a sufficient condition for the multivariate similarity of the model prediction to data, this can provide strong evidence in favor of a good fit, at a much higher confidence than tests performed only on the marginal distributions.

Following from above, Figure 2 presents the model predictions for marginal distributions of the four LGRB variables in the observer frame. The K-S test probabilities for the similarity of the model predictions to the marginal distributions of BATSE LGRB variables are also reported on the top right of each plot. All three redshift-distribution parameters of HB06, Li (2008), and B10 (Equation (4)) result in relatively similar fits to BATSE data in the observer frame. Thus, for brevity, only plots for one representative (median) redshift distribution (i.e., Li 2008) are presented.

Figure 2.

Figure 2. Marginal distribution predictions (solid blue lines) of the LGRB world model given BATSE detection efficiency, superposed on BATSE 1366 LGRB data (red histograms). The solid gray lines represent the model predictions for the entire LGRB population (detected and undetected), with no correction for BATSE sky exposure and the beaming factor (fb). The 90% confidence intervals (dashed green lines) represent random Poisson fluctuations expected in the BATSE LGRB-detection process. Bottom: the differential (left panel) and cumulative (right panel) rate of LGRBs as a function of peak photon flux in the BATSE nominal detection energy range (50–300 keV). For a comparison with the Swift sample of LGRBs and the proposed multivariate model of B10, the two dashed gray lines—taken from Figure 1 in B10—represent the predictions of the broken power-law luminosity function in Equation (2) of B10, in the observer frame. The Kolmogorov–Smirnov (K-S) test probabilities for the goodness-of-fit of the model predictions to BATSE data are reported at the top-right corner of each plot.

Standard image High-resolution image

As the second level of GoF tests, the joint bivariate model predictions are compared to BATSE LGRB data, presented in Figures 35. This method of scanning model and data along the principal axes of the joint bivariate distributions can be generalized to trivariate and quadruvariate joint distributions. For brevity, however, only the bivariate tests are presented.

Figure 3.

Figure 3. Top: BATSE 1366 LGRB data (red dots) superposed on the joint bivariate distribution predictions (black dots) of the LGRB world model for BATSE detection efficiency. The gray background dots represent the model predictions for the entire LGRB population (detected and undetected). The uncertainties in the BATSE LGRBs' variables are derived from the empirical Bayes model discussed in Appendix C. Center and bottom: gauging the goodness-of-fit of the bivariate model to data by scanning and comparing the joint distributions of the model and BATSE data along their principal axes. Colors bear the same meaning as the top panels, in addition to the 90% confidence intervals (green dashed lines) that represent random Poisson fluctuations expected in BATSE LGRB-detection process.

Standard image High-resolution image
Figure 4.

Figure 4. Top: BATSE 1366 LGRB data (red dots) superposed on the joint bivariate distribution predictions (black dots) of the LGRB world model for BATSE detection efficiency. The background gray dots represent the model predictions for the entire LGRB population (detected and undetected). The uncertainties in the BATSE LGRBs' variables are derived from the empirical Bayes model discussed in Appendix C. Center and bottom: gauging the goodness-of-fit of the bivariate model to data by scanning and comparing the joint distributions of the model and BATSE data along their principal axes. Colors have the same meaning as the top panels, in addition to the 90% confidence intervals (green dashed lines) that represent random Poisson fluctuations expected in the BATSE LGRB-detection process.

Standard image High-resolution image
Figure 5.

Figure 5. Top: BATSE 1366 LGRB data (red dots) superposed on the joint bivariate distribution predictions (black dots) of the LGRB world model for BATSE detection efficiency. The gray background dots represent the model predictions for the entire LGRB population (detected and undetected). The uncertainties in BATSE LGRBs' variables are derived from the empirical Bayes model discussed in Appendix C. Center and bottom: gauging the goodness-of-fit of the bivariate model to data by scanning and comparing the joint distributions of the model and BATSE data along their principal axes. Colors have the same meaning as the top panels, in addition to the 90% confidence intervals (green dashed lines) that represent random Poisson fluctuations expected in the BATSE LGRB-detection process.

Standard image High-resolution image

3. RESULTS AND DISCUSSION

It is observed in the plots of Figures 25 that the model provides excellent fit to data, within the uncertainties caused by random Poisson fluctuations in the BATSE LGRB observed rate. These random fluctuations in BATSE detections are encompassed in each graph by the green dashed lines that represent the 90% confidence intervals (CI) on BATSE LGRB detections (blue solid lines), derived by repeated sampling from the model.

Unfortunately, the same methods for a comparison of data and model cannot be applied in the LGRB rest frame, due to lack of redshift information for the BATSE sample of LGRBs. Nevertheless, a comparison of the model with observational data of other instruments—with measured redshifts—can provide clues on the underlying joint distribution of LGRB temporal and spectral variables in the rest frame compared to LGRB detections of different gamma-ray instruments, as will be done in the following sections.

3.1. LGRB Luminosity Function and log(N)–log(P) Diagram

The log (N)–log (P) diagram of GRBs has been subject of numerous studies in the BATSE era, primarily for the purpose of finding signatures of cosmological (versus galactic) origins in the LGRB rate. The cosmological origin of LGRBs is now well established. Nevertheless, the log (N)–log (P) diagram can still provide useful information for future gamma-ray experiments.

Figure 2 (bottom) depicts the prediction of the LGRB world model for the traditional log (N)–log (P) diagram for 1 s peak photon flux in the BATSE nominal detection energy range 50–300 keV, both for the differential (left panel) and the cumulative (right panel) LGRB rate. For all three LGRB cosmic rates considered in this work—as in Table 2—the differential log (N)–log (P) diagram shows a peak in the rate at P50–300 ∼ 0.1 photons s−1 cm−2. Such a peak in the LGRB rate results in a relative flattening at the dim end of the cumulative log (N)–log (P) diagram, as compared to its bright end. This observation has already been reported by B10 for the Swift sample of LGRBs, although an entirely different LF—a broken power-law LF—was used by B10 in their multivariate LGRB world model (cf. Equation (2) in B10).

On the other hand, the peak in the observer-frame LGRB rate translates to three relatively different (at ∼1σ level) peaks in the LF of LGRBs. In general, it is observed that the peak of the LF (i.e., the average 1 s peak luminosity of LGRBs) increases with increasing cosmic rate of LGRBs at high redshift. This effect is well depicted in the top left panel of Figure 6 for the three LGRB cosmic rates considered: HB06, Li (2008), and B10.

Figure 6.

Figure 6. Top and center: the marginal distribution predictions (blue dotted/dashed/solid lines) of the LGRB world model for BATSE LGRBs in the burst's rest frame. The gray dotted/dashed/solid lines—corresponding to the three LGRB redshift distributions HB06/Li (2008)/B10—represent model predictions for the entire LGRB population (detected and undetected), with no correction for BATSE sky exposure and the beaming factor (fb). For comparison with the Swift sample of LGRBs, refer to Figures 2, 6, and 7 of B10. Bottom: the three differential (left panel) and cumulative (right panel) rates of LGRBs at a given redshift (z).

Standard image High-resolution image

Compared to the predictions of B10's LGRB world model (cf. bottom plot of Figure 6 in B10), the log-normal model suggests a lower peak for BATSE LGRB LF (∼52.3 here versus ∼52.7 in B10) for the same redshift distribution of LGRBs. Averaging over the three redshift distributions considered, the model predicts a dynamic 3σ range of observer-frame brightness log (Pbol(erg s−1 cm−2)) ∈ [ − 7.11 ± 2.66] corresponding to Pbol(erg s−1 cm−2) ∈ [1.70 × 10−10, 3.58 × 10−5] for LGRBs. This translates to an average dynamic 3σ range—in the rest frame—of log (Liso(erg s−1)) ∈ [51.53 ± 1.99] corresponding to Liso(erg s−1) ∈ [3.46 × 1049, 3.38 × 1053].

3.2. Isotropic Emission and Peak Energy Distributions

A comparison of the top-right panel of Figure 6 with Swift observations of LGRBs (e.g., B10, Figures 7 and 8) indicates the similarity of the peak rates in BATSE and Swift samples of total isotropic emission (Eiso) distributions. The distributions for both instruments show a similar peak at log (Eiso) ∼ 52.7. The Swift Eiso distribution, however, spans relatively lower Eiso as compared to BATSE. This observation is not surprising if one takes into account the strong trivariate correlations of the three LGRB variables: Liso, Eiso, and Ep, z, all of which play a role in LGRB detection by Swift and BATSE. In fact, a comparison of the spectral peak energy (Ep, z) distributions of Swift (e.g., Figure 2 in B10) with the model's prediction for BATSE LGRBs (center left panel of Figure 6) reveals the relatively high sensitivity of the Swift BAT detector to dim soft LGRBs as compared to BATSE LADs. Such difference between the two detectors has already been discussed frequently by different authors (e.g., Band 2003, 2006).

In sum, averaging over the three redshift distributions considered, the LGRB world model predicts a dynamic 3σ range of observer-frame bolometric LGRB fluence log (Sbol(erg cm−2)) ∈ [ − 6.16 ± 3.01] corresponding to Sbol(erg cm−2) ∈ [6.82 × 10−10, 7.01 × 10−4]. This translates to an average 3σ range—in the rest frame—of log (Eiso (erg)) ∈ [51.93 ± 2.71] corresponding to Eiso (erg) ∈ [1.66 × 1049, 4.46 × 1054].

As for the spectral peak energy (Ep and Ep, z) distributions, the model predicts a 3σ range of observer-frame LGRB spectral peak energy log (Ep (keV)) ∈ [1.93 ± 1.22] corresponding to Ep (keV) ∈ [5, 1427]. This translates to an average 3σ range—in the rest frame—of log (Ep, z (keV)) ∈ [2.48 ± 1.12] corresponding to Ep, z (keV) ∈ [23, 4006].

3.3. Duration Distribution

GRBs are traditionally flagged as a long-duration class of bursts if their observed durations (T90) exceed 2 s (e.g., Kouveliotou et al. 1993). Such a classification, however, has long been known to be ambiguous close to the cutoff set at T90 = 2 s. It will be therefore useful to explore how accurate such classification is for the entire LGRB population (including non-triggered LGRBs). Figure 2 (center right plot) depicts the underlying population versus the 1366 BATSE LGRB observed durations (T90). As implied by the model, the shape of the T90 distribution of LGRBs is not significantly affected by the triggering process of BATSE, since both BATSE and entire LGRB population distributions show a similar peak at T90 ∼ 30 s. There is however a slight difference (<0.1 dex) in the predicted observer-frame peak of LGRBs T90 distribution, depending on the underlying LGRB redshift distribution assumed. The difference is magnified to ∼0.2 dex in the rest-frame (T90, z) duration distribution of LGRBs, for the two extreme cases of HB06 and B10 redshift distributions. In general, a higher LGRB rate at high redshifts (as in the case of B10 redshift distribution) results in a shift to shorter durations in the duration distribution of LGRBs, in both the observer and the rest frames (Figure 6, center right plot).

Overall, the model predicts an average dynamic 3σ range of observer-frame log (T90 (s)) ∈ [1.47 ± 1.32] corresponding to T90 (s) ∈ [1.4, 620] for LGRBs. This translates to an average dynamic 3σ range of rest-frame log (T90, z (s)) ∈ [0.92 ± 1.24] corresponding to T90, z (s) ∈ [0.47, 145].

3.4. Temporal and Spectral Correlations

Ever since the launch of the Swift gamma-ray detector satellite, there has been a flurry of reports on the discovery of strong and significant correlations among the spectral parameters of LGRBs, most prominently among the rest-frame spectral peak energy and the total isotropic emission or peak luminosity of LGRBs (e.g., Amati et al. 2008; Ghirlanda et al. 2008). Despite the lack of measured redshifts for BATSE GRBs, signatures of such correlations had been found by earlier works in the BATSE era through careful analysis of observer-frame spectral properties of LGRBs (e.g., Lloyd et al. 2000). Nevertheless, the strength and significance of these correlations were undermined by analyzing larger samples of the BATSE catalog of GRBs (e.g., Nakar & Piran 2005; Band & Preece 2005; Shahmoradi & Nemiroff 2009, 2011) or the Swift sample of GRBs (e.g., Butler et al. 2007, 2009; B10), all arguing that the sample of bursts used to construct the claimed spectral relations is representative of only bright-soft LGRBs. These arguments have been responded to by others (e.g., Ghirlanda et al. 2005, 2008; Nava et al. 2008; see Shahmoradi & Nemiroff 2011 for a complete history of the debate).

A multivariate analysis of the BATSE LGRB data that carefully eliminates potential biases at the detection process can therefore greatly help us understand the strength and significance of the reported correlations. Figure 7 shows the predictions of the LGRB world model for the two widely discussed spectral correlations: the LisoEp, z (the Yonetoku) and EisoEp, z (the Amati) relations. As indicated by the model, a large fraction of BATSE LGRBs (and a much larger fraction of the entire LGRB population) on the dim-hard regions of the plots appear to be underrepresented by the sample of LGRBs used for the construction of these relations. Nevertheless, given the three redshift distributions considered for the model, a relatively strong (Pearson's correlation coefficient $\rho _{E_{{\rm iso}}-E_{p,z}}=0.58\pm 0.04$) and highly significant (>14σ) correlation is predicted between the Eiso and Ep, z of LGRBs. The slope of the two relations suggested by the model also differs significantly from the original reports of the relations (Schaefer 2007; Ghirlanda et al. 2008). Generally, in regression modeling, it is assumed that there are two independent and dependent variables. In the case of the proposed relations, none of the variables is known to depend theoretically on the other. Therefore, due to the large statistically unexplained variances of the two LGRB variables, different regression methods, such as ordinary least squares, OLS(Y|X) and OLS(X|Y), result in entirely different slopes for the relations. The best-fit power-law relations and the conditional variances of the regress and given the regressor can be easily obtained from the parameters of the multivariate log-normal model in Table 2 (e.g., Section 2.11 in Kutner et al. 2004).

Figure 7.

Figure 7. Joint bivariate distribution predictions (black dots) of the LGRB world model for BATSE detection efficiency, assuming the LGRB rate tracing the cosmic star formation rate (SFR), updated by Li (2008). The gray background dots represent model predictions for the entire LGRB population (detected and undetected). Superposed on the model predictions are the well-known proposed Yonetoku (LisoEp, z) and Amati (EisoEp, z) relations. The red lines in each plot represent the best-fit power-law relations derived by the respective authors. Neither the Yonetoku nor the Amati relations appear to be consistent with the predicted joint distributions for the entire LGRB population (gray background dots) or 1366 BATSE LGRBs (black dots). This implies strong bias and selection effects in the detection process, redshift determination, and spectral analysis of the LGRB samples that were used to construct the two relations.

Standard image High-resolution image

A partial correlation analysis of the two LisoEp, z and EisoEp, z relations (Figure 8) reveals that the moderate correlation of the isotropic peak luminosity with the time-integrated peak energy is entirely due to the strong association of Liso with the total isotropic emission from the burst. As seen in the top left plot of Figure 8 there is indeed a negative correlation between the Liso and Ep, z of GRBs for a fixed isotropic emission Eiso and burst duration T90, z. Conversely, the model indicates a highly significant correlation of Eiso with the time-integrated Ep, z, even after elimination of the effects of Liso and T90, z from the EisoEp, z relation.

Figure 8.

Figure 8. Partial correlation analysis of LGRB variables. Each plot depicts the posterior distributions of the zeroth-, first-, and second-order (partial) correlation coefficients among pairs of LGRB variables, for the median case of LGRB redshift distribution of Li (2008). Top left: keeping T90, z and Ep, z fixed, a strong correlation between Liso and Eiso is still observed, indicating an intrinsic strong association between the total isotropic emission (Eiso) and the 1 s isotropic peak luminosity (Liso) of LGRBs. Top right: eliminating the strong dependence of Liso on Eiso reveals that there is indeed a moderate negative correlation (ρ ∼ −0.32 ± 0.1) between the Liso and Ep, z of LGRBs. A similar argument also holds for the correlation of Liso with T90, z (as seen in center left plot). Conversely, the center right plot shows a highly significant positive correlation (ρ ∼ 0.49 ± 0.06) of Eiso with Ep, z even after dissociation of Eiso and Ep, z from the two other parameters of the model, Liso and T90, z, suggesting the existence of a true underlying link between the total isotropic prompt emission (Eiso) and hardness (Ep, z) of LGRBs. Posterior distributions of the correlation coefficients in the case of the two other LGRB redshift distributions of HB06 and B10 also exhibit similar behavior.

Standard image High-resolution image

As observed in Table 2, the model predicts positive correlation among all four LGRB variables. The isotropic peak luminosity (Liso) and the total isotropic emission (Eiso) appear to be strong indicators of each other reciprocally. Surprisingly, it is also observed that the rest-frame duration T90, z of LGRBs strongly correlates with both Eiso and Liso. The existence of a possible positive correlation between the isotropic emission and the duration of LGRBs has been implied by the analysis of Swift LGRBs (B10), though only weakly present there. Such correlations can be enlightening for the early studies of time-dilation signatures in BATSE GRBs. A positive duration–brightness correlation is also opposite to—but not necessarily in contradiction with—the negative duration–brightness correlation in pulses of individual GRBs (e.g., Fenimore et al. 1995; Nemiroff 2000; Ramirez-Ruiz & Fenimore 2000). Combination of the two correlations implies that the number of pulses in individual LGRBs should be positively correlated with the peak luminosity (or equivalently, the total isotropic emission) of the bursts. This is indeed in qualitative agreement with the observed inequality relation between the isotropic peak luminosity and the number of pulses in Swift LGRBs (Figure 6 in Schaefer 2007). The strength of the correlations found encourages the search for the underlying physical mechanism that could give rise to these relations. This is, however, beyond the scope of this paper (cf. Rees & Mészáros 2005; Ryde et al. 2006; Thompson et al. 2007; Giannios 2012; Dado & Dar 2012 for example discussions).

It is also worth mentioning that the T90 duration of BATSE LGRBs strongly correlates with the bolometric fluence to bolometric 1 s peak flux ratio (FPR), with Pearson's correlation coefficient ρFPR–T90 ≈ 0.67. A comparison of BATSE data with the predictions of the model for the bivariate distribution of FPR and T90 is given in Figure 9 (left panel). Interestingly, the model predicts the same correlation strength of ρFPR–T90 ≈ 0.67 for the entire LGRB population, implying that the detection process does not bias the FPR–T90 relation. Such strong correlation indicates an underlying intrinsic interrelation between the three variables: Pbol, Sbol, and T90, also among their corresponding rest-frame counterparts. In fact, B10 use a variant of this trivariate correlation to define an effective peak flux in terms of fluence and duration, discarding the traditional definition of peak flux as the peak photon counts in 1 s time interval.

Figure 9.

Figure 9. Left panel: the joint bivariate distribution of the observed duration (T90) vs. the ratio of bolometric fluence to 1 s bolometric peak flux (FPR) of 1366 BATSE LGRBs (red dots) superposed on the predictions of the LGRB world model (black dots) for BATSE LGRB-detection efficiency. The gray background dots represent model predictions for the entire LGRB population (detected and undetected). Both BATSE LGRB data and the LGRB world model exhibit the same (Pearson's) correlation strength of $\rho _{{\rm FPR}\hbox{--}T_{90}}\approx 0.67$, indicating that the observed joint distribution is not a byproduct of selection effects in the detection process. Right panel: differential duration distribution (dN/dT90) of 1966 BATSE GRBs. The gray background histogram and curve represent the entire sample of 1966 BATSE GRBs and the bimodal log-normal fit to this sample, while the blue and red histograms and curves represent the two classes of short-hard and long-soft GRBs. The black curve represents the "asymptotic" prediction of the LGRB world model for the underlying distribution of LGRBs' duration (T90) distribution in the observer frame. The green dashed lines represent the 90% confidence interval on the "asymptotic" prediction of the LGRB world model for 1366 BATSE LGRBs (red curve). The observed flatness in the duration distribution of LGRBs (and also SGRBs) can be attributed to sample incompleteness and the skewed nature of the log-normal distribution when plotted as dN/dT90 on logarithmic axes. This interpretation is inconsistent with the argument provided by Bromberg et al. (2012) who point to the observed flatness in T90 distribution of BATSE GRBs as a direct evidence of the Collapsar model of LGRBs.

Standard image High-resolution image

4. SUMMARY AND CONCLUDING REMARKS

The primary goal of the presented analysis was to model and constrain the LF, temporal and spectral correlations, and energetics of long-duration class of GRBs by exploiting the wealth of information that has been buried and untouched in the BATSE GRB catalog to this date. Below is a summary of steps taken to construct the LGRB world model, detailed in Section 2.

  • 1.  
    A sample of 1366 LGRBs is carefully selected by using the fuzzy C-means clustering algorithm based on temporal and spectral parameters and inspection of the individual light curves of BATSE-catalog GRBs (Figure 1; Section 2.1).
  • 2.  
    It is proposed that the BATSE LGRB data might be very well consistent with being drawn from a multivariate log-normal population of LGRBs in four rest-frame LGRB variables: the bolometric isotropic 1 s peak luminosity (Liso), the bolometric isotropic emission (Eiso), the spectral peak energy (Ep, z), and the duration (T90, z). Therefore, the observed joint distribution of the four LGRB variables: the bolometric 1 s peak flux (Pbol), the bolometric fluence (Sbol), the observed spectral peak energy (Ep) and the observed duration (T90) results from the convolution of the rest-frame multivariate log-normal population with the cosmic rate (i.e., the redshift distribution) of LGRBs, truncated by the complex LGRB trigger threshold of BATSE LADs, as illustrated in Section 2.2, Equations (1)–(5). A prescription for modeling BATSE LAD detection efficiency is given in Appendix B.
  • 3.  
    The LGRB model (Equation (1)) is fit to BATSE data by maximizing the likelihood function of the model (Section 2.3 and Appendix C; Equation (C2)). In order to derive the best-fit parameters of the model and their corresponding uncertainties, an adaptive Metropolis–Hastings MCMC (AMH-MCMC) algorithm is set up to efficiently sample from the 16D likelihood function. The best-fit parameters are obtained for three LGRB cosmic rates: SFR of HB06, SFR of Li (2008), and the predicted LGRB redshift distribution of B10 which is consistent with the LGRB rate tracing cosmic metallicity with a cutoff Z/Z ∼ 0.2–0.5.
  • 4.  
    To ensure the model provides adequate fit to observational data, multivariate GoF tests are presented (Section 2.4 and Figures 25).

Summarized below are the principal conclusions drawn from the analysis based on the proposed LGRB world model.

  • 1.  
    Energetics. It is expected that the peak brightness distribution of LGRBs has the effective range of log (Pbol(erg s−1 cm−2)) ∈ [ − 7.11 ± 2.66] corresponding to Pbol(erg s−1 cm−2) ∈ [1.70 × 10−10, 3.58 × 10−5]. This translates to a dynamic 3σ range—in the rest frame—of log (Liso(erg s−1)) ∈ [51.53 ± 1.99], corresponding to Liso(erg s−1) ∈ [3.46 × 1049, 3.38 × 1053]. In addition, a turnover is predicted in the differential log (N)–log (P) diagram of LGRBs at P50–300 ∼ 0.1 photons s−1 cm−2 in the BATSE nominal detection energy range (50–300 keV). This is consistent with and further extends the apparent flattening in the cumulative log (N)–log (P) diagram of Swift LGRBs reported recently by B10.As for the bolometric fluence and the total isotropic emission distributions, a range of log (Sbol(erg cm−2)) ∈ [ − 6.16 ± 3.01] corresponding to Sbol(erg cm−2) ∈ [6.82 × 10−10, 7.01 × 10−4] is indicated. This translates to an average dynamic 3σ range—in the rest frame—of log (Eiso (erg)) ∈ [51.93 ± 2.71] corresponding to Eiso (erg) ∈ [1.66 × 1049, 4.46 × 1054] (Sections 3.2 and 3.1; Table 2).
  • 2.  
    Durations and spectral peak energies. The rest-frame spectral peak energies (Ep, z) of LGRBs are likely well described by a log-normal distribution with an average 3σ range of log (Ep, z (keV)) ∈ [2.48 ± 1.12] corresponding to Ep, z (keV) ∈ [23, 4006] with peak LGRB rate at Ep, z ∼ 300 (keV). This translates to an effective observer-frame peak energy range of log (Ep (keV)) ∈ [1.93 ± 1.22] corresponding to Ep (keV) ∈ [5, 1427] with peak LGRB rate at Ep ∼ 85 keV. It is also observed that the observer-frame T90 duration of LGRBs peaks at T90 ∼ 30 s with a 3σ range of T90 (s) ∈ [1.4, 620]. This translates to an average 3σ range of rest-frame log (T90, z (s)) ∈ [0.92 ± 1.24] corresponding to T90, z (s) ∈ [0.47, 145] with a peak rate at T90, z ∼ 10 s (Section 3.3; Table 2).Recently, Bromberg et al. (2012) proposed the apparent flatness in the duration distribution of BATSE LGRBs—when plotted in the form of dN/dT90 instead of dN/dlog (T90)—as the first direct evidence of the Collapsar model of LGRBs. The results of presented analysis are inconsistent with a flat T90 distribution of LGRBs at short durations (Figure 2, center right panel and Figure 9, right panel). The observed flat T90 distribution of LGRBs at short durations can be explained away in terms of the skewed nature of the log-normal distribution subject to sample incompleteness. It is therefore expected that a significantly larger sample of LGRBs that will be detected by future gamma-ray satellites will smear out the apparent flatness at the short tail of the duration distribution of LGRBs. A similar flat distribution is also observed for SGRBs at very short durations (Figure 9, right panel) which might be hard to reconcile with the Collapsar interpretation of the observed flatness in the LGRB T90 distribution, proposed by Bromberg et al. (2012).
  • 3.  
    Temporal and spectral correlations. All four LGRB variables: Liso, Eiso, Ep, z, and T90, z appear to be either moderately or strongly positively correlated with each other. In particular, a relatively strong and "broad" but highly significant correlation strength (Pearson's correlation coefficient $\rho _{E_{{\rm iso}}-E_{p,z}}=0.58\pm 0.04$) is predicted between Eiso and Ep, z of long-duration class of GRBs. Surprisingly, T90, z appears to evolve with Liso and Eiso such that brighter bursts generally tend to have longer durations (Section 3.4; Table 2). This prediction of the model together with the previously reported negative correlation of the brightness and the duration of individual pulses in LGRBs (e.g., Fenimore et al. 1995; Nemiroff 2000; Ramirez-Ruiz & Fenimore 2000) might possibly indicate that intrinsically brighter LGRBs contain, on average, a higher number of pulses.There is a slight chance that a small fraction (<50) of BATSE LGRBs were misclassified as SGRBs by the automated pattern recognition methods exploited in this analysis (see Figure 3, center right panel). If true, it will most likely affect (if significant at all) the constraints derived on the LF of LGRBs and the correlation of Ep, z with Liso.
  • 4.  
    Redshift distribution. The lack redshift information for the BATSE GRBs strongly limits the prediction power of the presented analysis for the cosmic rate of LGRBs. Nevertheless, based on the Markov chain sampling of the likelihood function for the three LGRB redshift distributions considered here (Section 2.2 and Figure 10), it is observed that BATSE data potentially, but not necessarily, favor an LGRB rate consistent with cosmic metallicity evolution with a cutoff Z/Z ∼ 0.2–0.5 (cf. B10), with no luminosity–redshift evolution.Assuming that LGRBs track SFR, only a tiny fraction (i.e., ∼2–3) of 1366 BATSE LGRBs are expected to have originated from high redshifts (z ≳ 5). In the case of an LGRB rate tracing cosmic metallicity evolution (e.g., B10), the fraction increases by one order of magnitude to ∼2%, corresponding to ∼27 bursts out of 1366 BATSE LGRBs. For comparison, the expected fraction of Swift and EXIST LGRBs with z ≳ 5 are ∼6% and ∼7% (B10). The discrepancy is well explained by the fact that both Swift and EXIST are more sensitive to long-soft bursts—characteristic of high-redshift LGRBs—due to their lower gamma-ray trigger energy window, compared to BATSE DADs (cf. Gehrels et al. 2004; Band et al. 2008; Grindlay & the EXIST Team 2009).
Figure 10.

Figure 10. Left: a comparison of the model-predicted BATSE trigger efficiency of 1366 BATSE LGRBs with the nominal trigger efficiency estimates of BATSE 4B catalog. The discrepancy between the two curves is primarily due to different methodologies and GRB samples used to derive BATSE-catalog efficiency curve2 and the trigger efficiency of the LGRB world model. Right: the normalized sampling distribution of the likelihood function (Equation (C2)) from Markov chain for three LGRB redshift distributions (Section 2.2; Equation (4)). The fact that the redshift distribution of B10 generally results in larger likelihoods as compared to the two other can provide evidence—but does not necessitate in any way—that BATSE data favor an LGRB rate tracing cosmic metallicity evolution over SFR. A firm decision can only be made with complete knowledge of BATSE LGRB redshifts.

Standard image High-resolution image

Although fitting is performed for the rest-frame variables, it is notable that the overall shape of the resulting observer-frame distribution of the variables also resembles a multivariate log-normal (see Figures 27). In other words, the redshift convolution of the rest-frame population distribution approximately acts as a linear transformation from the rest frame of LGRBs to the observer frame. This is primarily due to the narrow redshift distribution of LGRBs—as compared to the width of the LGRB rest-frame temporal and spectral distributions—with almost 90% of the population originating from intermediate redshifts, z ∈ [1, 4.3]. Balazs et al. (2003) provide an elegant discussion on the potential effects of redshift convolution on the observed distribution of LGRB durations and spectral parameters.

As implied by the model, there is no evidence for a significant population of bright-hard LGRBs that could have been missed in the BATSE catalog of GRBs. Conversely, a large population of low luminosity with moderate-to-low spectral peak energies seem to have gone undetected by BATSE LAD detectors. It should be emphasized that the apparent lack of very bright-soft LGRBs has a true physical origin according to the analysis presented here and is not an artifact of the detection process or spectral fitting models (e.g., the Band model, CPL or SBPL models) used by GRB researchers. Whether the X-Ray Flashes (XRF), X-ray-rich, and the sub-luminous GRBs (e.g., Strohmayer et al. 1998; Kippen et al. 2003; Tikhomirova et al. 2006) can be incorporated into a unified class of events described by a single model remains an open question in this work. At present, these events can be either considered as a separate class of cosmological events or as the soft-dim tail of the LGRB world model that have been mostly out of BATSE detection range and missed (see Figure 3 of Kippen et al. 2003 for a comparison with the predictions of the LGRB world model here in Figures 25). A definite answer to this question requires knowledge of the true rate of sub-luminous bursts and XRFs based on the observed rates of these events convolved with complex detection thresholds of different instruments used for observations.

B10 have presented an elaborate multivariate analysis of Swift LGRBs. While providing reasonable fit to Swift data, the model of B10 is primarily aimed at the discovery of the potential sub-luminous events that mostly go undetected by gamma-ray detectors. Such a model, capable of accounting for a large population of undetected sub-luminous bursts, is proposed at the cost of throwing away parts of information stored in the spectral parameters of LGRBs in the analysis of B10 (cf. Ghirlanda et al. 2012). Nevertheless, the presented analysis indicates that the apparent correlation of the isotropic peak luminosity (Liso) with the time-integrated spectral peak energy (Ep, z) of LGRBs is peripheral to the more fundamental relation between the total isotropic emission (Eiso) and Ep, z, and that the relation can be created by defining an effective luminosity based on the two GRB variables Eiso and duration (e.g., T90, z) (Figures 8 and 9, left panel). In light of the analysis presented by Ghirlanda et al. (2012), it can be therefore suggested that a new definition of luminosity based on the Eiso and T90, z of GRBs drawn from the LGRB world model of B10 will alleviate the apparent discrepancy between the observed LisoEp, z relation of LGRBs and the predicted relation from the LGRB model of B10. It is also expected that a better definition of peak luminosity that is not limited to a specific timescale in the observer frame of the bursts would result in an improvement in the correlations of the time-integrated spectral peak energy (Ep, z) and the isotropic emission (Eiso) with the luminosity variable (Liso). Given the above arguments, the presence of Liso as an independent variable in the LGRB world model—which was unfortunate due to the dependence of the BATSE trigger algorithm on the GRB's peak luminosity—might be viewed as overfitting and unnecessary.

The proposed multivariate log-normal model while requiring minimal free parameters compared to any other statistical model considered in GRB literature to date provides an accurate comprehensive description of the largest catalog of long-duration GRBs, serving as a powerful probe to explore the population properties of a large fraction of LGRBs that are missed in spectral analyses due to low-quality data or the lack of measured redshift, or simply go undetected due to the instrument's gamma-ray detector threshold. Data from future gamma-ray experiments will enable us to further confirm, improve, or invalidate predictions of the presented model.

The author owes special thanks to Robert J. Nemiroff (NASA GSFC Astrophysicist and professor of Physics at Michigan Tech University) for his generous support and help through numerous discussions over more than a year, from the early stages of developing the basic ideas behind this work to the core of the analysis and final steps of manuscript preparation.

I thank William H. Press (professor of Computer Science and Integrative Biology at The University of Texas at Austin), Lars Koesterke (at Texas Advanced Computing Center), and Mehdi Mortazavi (at MTU) for helpful comments on some statistical and computational aspects of this work. I am very grateful to Swadesh Mahajan (professor at Institute for Fusion Studies) and Richard Hazeltine (professor and chair at the Department of Physics, The University of Texas at Austin) for their generous support in the final stages of preparing this manuscript. I also thank Giancarlo Ghirlanda (at INAF-Osservatorio Astronomico di Brera) for his timely feedback and comment on the GRB terminology used in this manuscript. Greatly appreciated were helpful comments and detailed criticisms from the anonymous referee of this manuscript.

This work would have not been accomplished without the vast time and effort spent by many scientists and engineers who designed, built, and launched the Compton Gamma-Ray Observatory and were involved in the analysis of GRB data from BATSE Large Area Detectors.

APPENDIX A: GRB CLASSIFICATION

It is well known that the traditional classification of GRBs based on a sharp cutoff in the observed duration (T90) distribution of GRBs—usually set at T90 = 2 s[50–300 keV]—is insufficient and can be misleading close to the sharply defined border. The apparent long-soft-bright to short-hard-dim trend observed in the prompt-emission properties of BATSE GRBs (e.g., Figure 1, top panel; see also Figure 8 of Shahmoradi & Nemiroff 2010) necessitates the use of a rigorous classification scheme based on all available spectral properties of GRBs, in addition to duration.

Despite a rich literature on the classification methodologies for GRBs (e.g., Hakkila et al. 2000a, 2003, and references therein), the choice of a classification method to separate the BATSE catalog of GRBs into two subgroups of long and short durations with minimal misclassifications remains a difficult task. This is primarily due to the significant overlap (or similarity) in all (or some) spectral and temporal properties of the two classes of GRBs, in addition to the heterogeneity of objective functions that might differ considerably from one classification algorithm to another.

Here, fuzzy (soft) clustering algorithms are preferred over hard clustering methods, since they provide a probability of the event belonging to each specific subgroup, in contrast to hard classifications that return only binary probabilities of either 0 or 1. The choice of fuzzy algorithm greatly facilitates identification of bursts that might have been potentially misclassified (Section 2.1).

Investigation of different fuzzy algorithms available in the literature leads us to two prominent candidates: the SAND method of Rousseeuw et al. (1996) and the fuzzy C-means discussed by Dunn (1973) and Bezdek (1981). While fuzzy C-means is especially useful for cases where subgroups are known to be approximately symmetric, the SAND algorithm is superior to C-means for its lack of sensitivity to different subgroup sizes, orientations, and asymmetries. Nevertheless, the presence of a handful of SGRs in the BATSE catalog—with spectral properties comparable to that of LGRBs—results in relatively poor classification by the SAND method as compared to C-means. Besides the choice of algorithm, the GRB variables—by which the classification is done—are selected such that the resulting relative sizes of the two SGRB and LGRB populations correspond to those found by Shahmoradi & Nemiroff (2010) through a different approach that the author believes to be less prone to biases (cf. Figure 13 and Table 4 in Shahmoradi & Nemiroff 2010).

APPENDIX B: BATSE TRIGGER EFFICIENCY

Before the LGRB world model of Equation (1) is fit to BATSE observational data, it is necessary to convolve the model with the BATSE trigger threshold (Equation (5)). The study of the BATSE detection efficiency is well documented in a series of articles by the BATSE team (e.g., Pendleton et al. 1995, 1998; Paciesas et al. 1999; Hakkila et al. 2003; cf. Shahmoradi & Nemiroff 2011 for further discussion and references). Here, based on the observation that almost all 1366 BATSE LGRBs have durations of T90 > 1 s, the primary trigger timescale for BATSE LGRBs is assumed to be 1024 ms. This eliminates the relatively complex dependence of the detection probability (η in Equation (5)) on the duration of the events. The probability of detection for an LGRB is then modeled by the cumulative density function (CDF) of log-normal distribution,

Equation (B1)

where P(Liso, Ep, z, z) is the 1 s peak photon flux in the BATSE nominal detection energy range: 50–300 keV, and μthresh and σthresh are the detection threshold parameters to be determined by the model. The link between the 1 s peak photon flux (P) and the LGRB rest-frame variables (Liso, Ep, z) and redshift (z) is provided by fitting a smoothly broken power law known as the Band model (Band et al. 1993) to the differential photon spectra of LGRBs,

Equation (B2)

such that

Equation (B3)

It has been shown by B10 that fixing the high-energy and low-energy photon indices of the Band model (Equation (B2)) to the corresponding population average α = −1.1, β = −2.3 produces only a negligible error of <0.05 dex in the resulting flux estimates. Given the uncertainties in the BATSE LGRB variables, in particular Ep estimates, such an assumption provides reasonable approximation for the calculation of the peak fluxes. A more accurate treatment, however, would be to include possible weak correlations that are observed between the Band model photon indices and the spectral peak energies of the bursts (cf. Shahmoradi & Nemiroff (2010, 2011) for a discussion of the correlations and potential origins).

The goodness of the log-normal CDF assumption for BATSE detection efficiency can be checked by a comparison of the resulting model predictions with the 1366 BATSE LGRBs' distribution of peak fluxes (Figure 2). The best-fit model prediction of BATSE trigger efficiency for the long-duration class of GRBs is compared to the nominal trigger efficiency of the BATSE 4B catalog for the class of soft-long bursts in Figure 10 (left panel). Although the difference between the two curves is significant, it does not necessarily imply a contradiction, given the fact that different methodologies and GRB samples were used to derive the two efficiency curves.3

APPENDIX C: LIKELIHOOD FUNCTION

To obtain the joint posterior for the unknown parameters of the LGRB world model of Equation (1) given BATSE data, the likelihood function of the model must be, in principle, constructed by correctly accounting for uncertainties in observational data (e.g., Eddington 1913; Jeffreys 1938). In addition, it is known that astronomical surveys at low S/Ns close to survey threshold can be potentially biased (e.g., Hogg & Turner 1998). A Bayesian multilevel methodology (e.g., Hobson et al. 2010) can incorporate the above corrections required to construct the likelihood function: under the assumption of normality for the uncertainties of LGRB variables in the BATSE catalog, each LGRB event—denoted by Oi, standing for the ith LGRB observation—has the likelihood $\mathcal {L}_i$ of having the true parameters $\boldsymbol O_i\equiv [P_{{\rm bol},i},S_{{\rm bol},i},E_{p,i},T_{90,i}]$ that is described by a 4D Gaussian pdf:

Equation (C1)

where the location (μi, 0) and the scale (Σi, 0) parameters of the pdf are to be determined by the model and individual event data from the BATSE GRB catalog. Given $\mathcal {L}_i$ and the LGRB world model (Equation (1)) convolved with BATSE trigger efficiency (Equation (5)), the full Poisson likelihood function can be written as

Equation (C2)

where N = 1366, and $\mathcal {A}$ is a factor that properly normalizes the cosmic and the observed rates ($\mathcal {R}_{{\rm cosmic}}$ and $\mathcal {R}_{{\rm obs}}$). The term $\mathcal {R}_{{\rm cosmic}}$ in Equation (C2) acts as a prior for $\mathcal {L}_i$. In the absence of knowledge of the prior (i.e., $\mathcal {R}_{{\rm cosmic}}$, as in the case here), the empirical Bayes approach can provide an alternative solution, in which an ad hoc estimate of the model parameters based on the observed data (excluding uncertainties) serves as the prior for same data at the second level of analysis. Calculation of the normalization factor $\mathcal {A}$ involves an integration of the LGRB world model over the 5D space of LGRB variables and redshift, with a complex integration limit set by the BATSE trigger efficiency modeled in Equation (B1), as a function of Liso, Ep, z, and z. In addition, since almost no redshift information is available for BATSE catalog of GRBs, the probability for each LGRB has to be marginalized over redshift, for which a range of z ∈ [0.1, ] is considered. These integrations make the maximization of the likelihood for all unknown parameters an extremely difficult task, perhaps challenging current computational technologies. Moreover, it has been known that GRB fluences and durations are likely underestimated close to detection threshold due to the so-called fluence–duration bias (e.g., Hakkila et al. 2000b, 2003). Such bias makes the overestimation correction of the fluence and duration as prescribed by Hogg & Turner (1998) unjustified before the fluence–duration bias effects are well quantified. The algorithms for calculating peak fluxes, however, appear to result in less biased measurements—even down to detector threshold—with negligible uncertainties (e.g., Stern et al. 2001). Given the above lines of reasoning and the computational limitations, the uncertainties on three BATSE LGRB variables, Pbol, Sbol, and T90, are excluded from the likelihood function (Equation (C2)). The uncertainties on the Ep estimates of BATSE LGRBs are, however, significant compared to the three former variables and must be incorporated in the calculations of the likelihood. Nevertheless, it was realized after likelihood maximization that the exclusion of the Ep uncertainties—by fixing Ep values to the "bisector" estimates of Shahmoradi & Nemiroff (2010)—results in only negligible (≪1σ) changes in the model's best-fit parameters. In general, the use of the bisector line of OLS regression lines (e.g., Isobe et al. 1990) for estimation purposes is unfavored due to lack of a maximum likelihood interpretation. The special case of BATSE LGRBs here, however, turns out to be an exception.

In addition to cosmological time-dilation correction, it is common practice to make an energy correction to the temporal variables of Swift GRBs (e.g., Gehrels et al. 2006; B10), such as T90, when translating the variable from the observer frame to the rest frame. For BATSE LGRBs, this energy correction is likely negligible, given the fact that the T90 durations are calculated based on the total photon counts in the BATSE LAD energy range (e.g., Kouveliotou et al. 1993; Fishman et al. 1994), 20–2000 keV, which can be practically considered as bolometric. Nevertheless, it is expected that such an energy correction, if needed, would slightly relax the strong correlation of the rest-frame duration (T90, z) with the total isotropic emission (Eiso) and the peak isotropic luminosity (Liso).

The joint posterior distribution of the model parameters is obtained by maximizing the likelihood function of Equation (C2) convolved with a non-informative uniform prior on the location parameters and the standard choice of Jeffreys prior on the scale parameters (Jeffreys 1946). In order to efficiently sample from the 16D posterior density function, a variant of the MCMC methods known as AMH-MCMC is employed (e.g., Haario et al. 2001). The choice of an adaptive (versus classical) MH algorithm is very important, since the model parameters exhibit strong covariance (Table 3). To reduce the simulation runtime, all algorithms including AMH-MCMC are implemented in Fortran, which is by far the fastest and most efficient programming language for many intensive scientific calculations and number crunching (Loh 2010). In addition, the numerical integration in the definition of the luminosity distance (Equation (3))—encountered on the order of ≳ 109 times during the full MCMC sampling—is greatly simplified by the analytical approximation method of Wickramasinghe & Ukwatta (2010). Due to the intrinsic sequential character of MCMC sampling methods, the parallelization of simulation algorithms (on either shared or distributed memory architecture machines) is impractical or at best inefficient for a single Markov chain. Nevertheless, to increase the MCMC sample size and more importantly, to ensure convergence to the global (versus local) extremum, the chain is initiated simultaneously at 10 random starting points in the parameter space on a 12-core desktop CPU. In general, convergence and good mixing occur within the first few thousands of iterations (burn-in period) given a suitable initial guess for the covariance matrix of the proposal distribution—here chosen to be multivariate Gaussian. The resulting mean and 1σ standard deviations of the model parameters are tabulated in Table 2.

APPENDIX D: LGRB MONTE CARLO UNIVERSE

The prediction power and consistency of the presented LGRB world model—based on BATSE data—can be easily checked against observational data from current and future gamma-ray experiments, in particular the Fermi satellite. All it takes is to construct a Monte Carlo universe of LGRBs, based on the best-fit parameters of the LGRB world model in Table 2, and compare the outcome with observational data. Although straightforward, the steps for such simulation and comparison are summarized below.

  • 1.  
    A random redshift for each simulated LGRB is drawn from the redshift distribution of Equation (4) with parameters taken from Table 2. It is recommended to repeatedly randomly draw the set of model parameters from the full Markov chain samples4 instead of fixing the parameters to the mean values reported in Table 2.
  • 2.  
    The four LGRB variables: Liso, Eiso, Ep, z, T90, z are randomly drawn from a 4D log-normal distribution with location (μ) and scale (i.e., the covariance matrix: Σ) parameters constructed from fitting results in Table 2. This can be easily and quickly done by noting that a multivariate log-normal distribution is equivalent to a multivariate Gaussian distribution in the logarithmic space of the above variables,
    Equation (D1)
    such that the 4D log-normal density function, $\mathcal {LN}$, of Equation (1) can be exactly replaced by a 4D Gaussian distribution,
    Equation (D2)
    for which the cosmic LGRB differential rate ($\mathcal {R}_{{\rm cosmic}}$) of Equation (1) will be
    Equation (D3)
  • 3.  
    The above Monte Carlo universe of LGRBs can be then measured according to the instrument's detection efficiency (Section 2.2 and Equation (5)). For the case of BATSE LAD detectors, the trigger efficiency can be modeled according to the prescription in Appendix B.

Footnotes

Please wait… references are loading.
10.1088/0004-637X/766/2/111