-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathchapter1.tex
921 lines (782 loc) · 45 KB
/
chapter1.tex
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
\chapter{Introduction}\label{ch:introduction}
\section{Outflows from BNS Mergers}\label{sec:bns}
Due to the joint electromagnetic and gravitational wave detection of the merger
event GW1701817 (see \cite{abbott_2018}), there has been renewed interest in SGRBs.
Specifically, this detection gives credence to the claim that the central engines of
SGRBs may be binary neutron star mergers (see \cite{narayan_1992}). There are
however other aspects of the process that are not clear, and several properties of
GW170817 have not been completely explained. The main concerns are as follows
(\cite{lazzati_2020}):
\begin{itemize}
\item The outflow of GRB170817A (the gamma-ray counterpart event to GW170817)
was lower in energy than a typical cosmological SGRB, by a factor of $10^4 -
10^5$, even though the event was one of the closest GW events recorded, at a
distance of $\sim$ 40 Mpc. The reasons for this could be twofold:
\begin{itemize}
\item The structured jet was viewed off-axis with respect to the
observer.
\item The internal engine powering this SGRB was intrinsically less
energetic, and differs from the one observed in other typical SGRBs.
\end{itemize}
\item A clear consensus has not been reached on \emph{how} the gamma-ray prompt
emission was produced. Models which have been considered to explain the jet
launching include :
\begin{itemize}
\item The structured outflow model, characterised by functions for the
Lorentz factor and the energy per unit solid angle, both of which
vary with the angle made with the polar jet axis, $\theta$. This
model produces detectable signals even at moderately large off-axis
angles.
\item The shock breakout model (\cite{gottlieb_2018}), wherein the
leading edge of the wind emits the prompt emission as it breaks out
of the cocoon of nuclear matter ejected before the jet was launched.
This model has been shown to explain the energetics and spectrum of
the prompt emission, although it does require a setup in which the
wind is fast enough so that it can reach a large enough distance at
breakout.
\end{itemize}
The uncertainty in the jet production mechanism partly comes from the fact
that the various delays which play major roles \emph{before} the observation
of the prompt emission have not been accurately constrained, although rough
estimates for each component exist.
\end{itemize}
More light can be shed on these questions by observing more such SGRBs, using both
the gravitational wave (GW) and electromagnetic (EM) windows. However, the
possibility of joint detections are slim, due to the fact that the EM observations
are highly dependent on the viewing angle of the system with respect to the observer
(due to relativistic beaming), whereas GW signal amplitudes depend on the distance
to the event (see \cite{seto_2015}).\\
Given that this is the case, it would be expedient to look for constraints on the
structure parameters of various models. Furthermore, it would be ideal to develop
models which are resilient to \emph{non}-detections, i.e. those that can produce
constraints on the parameters using even upper limits on the flux/fluence observed
by the various EM follow-up satellites, such as \textit{INTEGRAL},
\textit{FERMI-GBM} or \textit{Swift/BAT}.
For this purpose, the structured jet model is adopted in this report and used as a
starting point. Using this model as a springboard, the properties of prompt
emissions from NS mergers are investigated.\\
\subsection{Modelling outflows from BNS Mergers}\label{ssec:bns-outflows}
As mentioned before, the electromagnetic follow-up of the binary neutron star merger
event GW170817 helped (roughly) measure the various time delays between the time of
the GW signal trigger (which roughly is the merger time itself) and the time the
gamma-ray signals were picked up. This time delay is denoted $\Delta t_{GW-\gamma}$,
and was around 1.75 seconds for this event. The components which make up this delay
are as follows (\cite{lazzati_2020}):
\begin{itemize}
\item \textbf{Engine Delay} -- this is the delay due to a transition in the
central engine mechanism which powers the jet. For example, such a
transition could be a metastable, fast spinning neutron star collapsing,
post-merger, into a black hole when its rotation period increases (a process
that can take years). Or it could be the time elapsed in amplifying the
magnetic field to a value large enough for jet launching (a process which is
significantly faster, taking only seconds). This delay is denoted by $\Delta
t_{eng}$.
\item \textbf{Wind Delay} -- this is simply a delay in the launching of a
non-relativistic wind due to the neutron-rich matter from the progenitor(s)
being tidally shredded. For this reason, it can be \textit{negative} as
well, since the tidal shredding can occur before the merger itself. This is
denoted as $\Delta t_{wind}$.
\item \textbf{Breakout Delay} -- if the wind is ejected before the jet, then the
latter will have to propagate through the former. This happens at a
sub-relativistic speed, whereas the GW signal travels at a relativistic
speed. The delay due to this crossing is the breakout delay, and is denoted
$\Delta t_{BO}$. During this time, jet-wind interactions cause the
development of a structured outflow that maintains a bright core but also
has energetic wings at large polar angles.
\item \textbf{Photospheric Delay} -- once the jet has crossed the wind, it
still needs to propagate out to the photospheric radius, where the outflow
becomes transparent and the prompt gamma-ray emission is radiated. The delay
from the breakout radius to the photospheric radius is $\Delta t_{ph}$. For
GW170817, this is given by:
\begin{equation}
\label{eq:4}
\Delta t_{ph} \sim \dfrac{R_{ph}}{c \Gamma^2} =
1.4 \dfrac{R_{ph}}{2 \times 10^{12} \text{ cm}}
\left(
\dfrac{7}{\Gamma}
\right)^2 \text{ s}
\end{equation}
\item \textbf{Dissipation Delay} -- this is a requirement in some models, such
as the internal shock synchrotron model, wherein the outflow needs to travel
to the internal shock radius before the bulk energy of the flow is
dissipated and turned into radiation. The time required to get to this point
after crossing the photospheric radius is the dissipation delay, denoted
$\Delta t_{\gamma}$
\end{itemize}
Several attempts to constrain the various time delay components have been made.
However, no conclusions have been arrived upon except for relative comparisons. For
example, one can only say that the photospheric delay is the major component out of
all the delays, and that wind delay (if non-zero) has to be lesser than the jet
delay, so that the jet catches up to the wind and the jet-wind interaction generates
the structured outflow. Fig. \ref{fig:jet_delay} summarises these delays in the
broader context.
\begin{figure}[H]
\centering
\includegraphics[width=12cm]{jet_delay}
\caption[Relative Positions of Jet Delays]
{
Two possible scenarios for the relative positioning of the delays in
time, which contribute to $\Delta t_{GW-\gamma}$. Owing to the
requirement of a structured outflow, GW170817 possibly follows the top
timeline. The relative contributions of the various delays are debated,
but it is agreed that $\Delta t_{wind} < \Delta t_{jet} \ll 1$ s,
$\Delta t_{BO} \ll 1$ s, $\Delta t_{\gamma} \sim 0$ and $\Delta t_{ph}
\sim \Delta t_{GW-\gamma}$.
}
\label{fig:jet_delay}
\end{figure}
Due to the uncertainties in the delay terms, several models for the jet can explain
the energetics and observed structure. Numerical simulations are also unequivocal
about their favouring of one model over the other (see \cite{shibata_2019}).
Some models try to explain the \textit{apparent} structure of the jet, which are the
observables seen by a particular observer at a particular viewing angle. Other
models are used to explain the \textit{intrinsic} structure, such as the polar angle
variation of the bulk Lorentz factor and the energy across the solid angle, in the
jet co-moving frame. See \cite{salafia_2015} for a detailed discussion of
the differences between the two structures. Some of the models considered are
described below (see also Figs. \ref{fig:tophat} and \ref{fig:jet_models}):
\begin{itemize}
\item Top-hat -- This model, as used in \cite{saleem_2020B}, assumes
that the bulk Lorentz and energy functions drop to zero beyond some cut-off
angle, $\theta_j$. Below this threshold, the functions are at their
respective on-axis values.
\item Gaussian -- This model is widely used, in some contexts to explain the
apparent jet structure (by \cite{hayes_2020}), and in others the
intrinsic jet structure (by \cite{saleem_2020B}). The former is
simply given by $y_{GJ}(\theta) = e^{- \frac{1}{2} \left(
\frac{\theta}{\theta_{\sigma}} \right)^2}$, since the authors consider only
the apparent jet structure, as explained above and $\theta_\sigma$ is a
structure parameter which is inferred by the authors' Bayesian inference
code.\\
In the latter, as the authors consider the intrinsic jet structure, they
assume that $\Gamma \beta (\theta) = \Gamma_0 \beta_0 \exp\left(- \theta^2 /
2\theta_c^2\right)$ and that $\epsilon (\theta) \propto \exp(- \theta^2 /
\theta_c^2)$\footnote{
This is the normalised energy profile function. The normalisation
constant is estimated by the condition $2\pi \int d(\cos \theta)
\epsilon(\theta) = E_{tot., \gamma}$, where $E_{tot., \gamma}$ is the
total energy in gamma-rays.
}, and derive the observed properties (see below).
\item Power Law -- This model is used by \cite{hayes_2020} to explain
the apparent structure of the jet, assuming that any variation in the energy
is simply because of relativistic beaming and the jet being viewed off-axis.
It is given using the shape function $y(\theta)$ (which is multiplied with
the on-axis isotropic equivalent energy $E_{iso, 0}$ to give
$E_{iso}(\theta)$\footnote{Using the equation $E_{iso}(\theta_v) = E_{iso,
0} \cdot y(\theta_v)$}):
\begin{equation}
\label{eq:5}
y(\theta) = \begin{cases}
1,
& 0 \leq \theta \leq \theta_c, \\
(\theta/\theta_c)^{-2},
& \theta_c < \theta \leq \theta, \\
0,
& \theta_j < \theta
\end{cases}
\end{equation}
Here $\theta_c$ and $\theta_j$ are simply structure parameters, inferred
using Bayesian methods.
\end{itemize}
\begin{figure}[H]
\centering
\includegraphics[width=10cm]{tophat}
\caption[Tophat jet structure model]{
Functional form of the tophat jet structure model, as considered in
\cite{saleem_2020B}. The dashed line denotes the jet angle
$\theta_j = 15^{\circ}$.
}
\label{fig:tophat}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=\textwidth]{jet_models}
\caption[Jet structures as in \cite{hayes_2020}]{
Functional forms of the jet structure models, as considered by
\cite{hayes_2020}. (Left) The gaussian jet structure with
a width $\theta_{\sigma} = 20^{\circ}$, also marked by the dashed
line. (Right) The power-law structure with a core angle $\theta_c =
8^\circ$ and a jet angle $\theta_j = 32^\circ$.
}
\label{fig:jet_models}
\end{figure}
This report deals exclusively with the intrinsic jet structure, and given the
functional forms relevant for describing the intrinsic jet structure, the observed
jet structure is computed. In order to do this, following \cite{granot_2002},
consider the emission profile of a point source moving at some angle with the
observer, essentially rendering this scenario off-axis. This geometry will affect
the prompt jet emission, as well as the initial afterglow, and thus warrants careful
analysis. Now, let the initial jet opening angle be $\theta_0$ and let the observer
be at an angle $\theta_{obs}$. In general, for a point source moving at any angle
$\theta$ with respect to the observer, the observed flux is given by :
\begin{equation}
\label{eq:1}
F_{\nu} =
\dfrac{L^{\prime}_{\nu^{\prime}}}{4 \pi d_L^2}
\left( \dfrac{\nu}{\nu^\prime}\right)^3
=
\dfrac{1 + z}{4 \pi d_L^2}
\dfrac{L^{\prime}_{\nu^{\prime}}}{\gamma^3 (1 - \beta \cos \theta)^3}
\end{equation}
Here, $L^{\prime}_{\nu^{\prime}}$ and $\nu^{\prime}$ are the jet comoving frame
spectral luminosity and frequency, $d_L$ is the luminosity distance, $\gamma = (1 -
\beta^2)^{-1/2}$ is the jet Lorentz factor. If $t$ and $\nu$ are the observed time
and frequency for an observer at $\theta$, and $t_0$ and $\nu_0$ are those for an
observer on the axis, then:
\begin{equation}
\label{eq:2}
\dfrac{t_0}{t} =
\dfrac{\nu}{\nu_0}
=
\dfrac{(1 - \beta)}{(1 - \beta \cos \theta)}
\equiv a
\approx \dfrac{1}{(1 + \gamma^2 \theta^2)}
\end{equation}
And finally putting Eq. \ref{eq:2} into Eq. \ref{eq:1} and expanding using a Taylor
series approximation up to the leading order:
\begin{equation}
\label{eq:3}
F_{\nu}(\theta_{obs}, t) = a^3 F_{\nu/a}(0, at)
\end{equation}
This equation can hence be used to relate observed off-axis quantities to the
on-axis ones. Furthermore, this enables a transfer from an intrinsic structure to an
observed one, which is what was required.\\
Since in this report, major focus is given to the EM outflows from NSBH mergers, the
background presented here about the possible EM outflows from BNS mergers suffices.
The interested reader is referred to \cite{ioka_2019}, \cite{mohan_2019} \&
\cite{lee_2007} for more detailed discussions on the properties of EM outflows from
BNS mergers.
\section{Outflows from NSBH Mergers}\label{sec:nsbh}
The major difference in the NSBH merger pathway to SGRBs, compared to the case of
BNS mergers, is that though there is theoretical and simulational support for the
launching of SGRB jets from the merger of a neutron star and a black hole of
appropriate mass (see for example \cite{ruiz_2020}, \cite{shibata_2019},
\cite{foucart_2020}), there has not been strong observational evidence for the same.
In the first half of the third observing run of the LVC (also known as O3a), there
have been several triggers which are reportedly confident NSBH triggers. However,
there were no counterpart EM signals picked up, which decreases the credibility of
NSBH mergers as the progenitors of SGRBs.\\
The electromagnetic component from NSBH mergers, is largely decided based on the
amount of mass left post-merger, outside the horizon of the remnant black hole.
This decides how much matter participates in the subsequent processes, which may be
the rapid neutron-capture process which gives rise to the kilonova signal
(\cite{metzger_2019}) or the magnetic field amplification process via the
Magneto-Rotational Instability (MRI) which leads to a SGRB jet
(\cite{postnov_2019}).\\
Qualitatively, for a binary where the neutron star is treated as a test mass and the
black hole's spin is aligned with the orbital angular momentum of the binary, the
innermost-stable circular orbit radius $r_{ISCO}$ scales as $r_{ISCO} \sim
f(\chi_{BH}) G M_{BH}/c^2$ (where f is a function ranging from 1 to 9, decreasing
for increasing (prograde) spins; see \cite{bardeen_1972}) and the radius at which
the tidal disruption of the neutron star occurs, $r_{dis}$ scales as $r_{dis} \sim k
(M_{BH}/M_{BNS})^{1/3} R_{NS}$ (where k is a constant with a dependence on the black
hole spin and the equation of state). Only requiring that $r_{dis} \gtrsim
r_{ISCO}$, as a rough requirement that disruption should occur before the neutron
star plunges into the black hole, leads to the conclusion that for disruption, one
requires:
\begin{enumerate}[label=\alph*.)]
\item Low-mass black holes, since $r_{dis}/r_{ISCO} \propto M_{BH}^{-2/3}$.
\item Larger NS radii, since $r_{dis} \propto R_{NS}$. This also implies
that the `softer' equations of state can disfavour tidal disruptions. This
is because a softer equation of state will lead to smaller NS radii, and
thus not much disruption would occur before the NS plunges into the black
hole.
\item Higher prograde BH spins, since $r_{ISCO}$ decreases for higher prograde
BH spins.
\end{enumerate}
These conclusions can be reached from Fig. \ref{fig:nsbh_disruption_condition} as
well. However, for quantitative results, simulations need to be performed such the
effect of the various components in the problem are correctly taken into account
(for a summary, see Fig. \ref{fig:rest_mass_fraction}). As seen from the
literature, where the requisite general-relativistic magnetohydrodynamic simulations
are carried out, the matter left over post-merger heavily depends on:
\begin{itemize}
\item \textbf{The mass ratio of the system}. This is defined as $\mathcal{Q} =
M_{BH} / M_{NS}$ so that $\mathcal{Q} > 1$ always. Fully general
relativistic, magnetohydrodynamic simulations (such as \cite{ruiz_2020})
show that in cases where the mass ratio is 3:1, regardless of the neutron
spin, a collimated outflow is observed, whereas the same is not realised in
all cases where the mass ratio is 5:1 or higher, and only occurs for high
black hole spin cases.
\item \textbf{The spin of the components of the system}. In geometrized units
(where $G = c = 1$), these are prescribed in terms of $a_{BH} / M_{BH}$ or
$a_{NS} / M_{NS}$. Whether these two spins align (prograde) or are
anti-aligned (retrograde) decides whether the neutron star would be tidally
disrupted (and hence participate in the processes mentioned previously) or
not, respectively. Via simulations, it is seen that the more the prograde
spin of the neutron star, the farther out it is tidally disrupted, albeit
this is only observed for the case of $\mathcal{Q}$ = 3:1 (comparing say,
Figs. \ref{fig:nsbh_jet} and \ref{fig:nsbh_5to1}). Also, this leads to long
tidal tails which produces a baryon-loaded environment and thus, the
magnetic field of the tidally disrupted matter must overcome the baryon ram
pressure to launch the jet. This process hence delays the launching of the
jet.
\end{itemize}
\begin{figure}[H]
\centering
\includegraphics[width=9cm]{nsbh_disruption_condition}
\caption[Disruption condition in a NSBH binary]
{
Maximum value of the mass-ratio ($M_{BH}/M_{NS}$) for which a NSBH
system disrupts, as a function of the neutron star radius $R_{NS}$,
and the aligned component of the dimensionless black hole spin
$\chi_{BH}$, assuming $M_{NS} = 1.35 M_{\odot}$. Results for other
neutron star masses can be obtained by rescaling considering the
disruption condition at constant compaction $C_{NS} = GM_{NS}/R_{NS}
c^2$. From \cite{foucart_2020}.
}
\label{fig:nsbh_disruption_condition}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=11cm]{rest_mass_fraction}
\caption[Rest mass outside black hole horizon, as a function of time]
{
Fraction of rest-mass of the NS outside the apparent horizon of the black
hole as a function of coordinate time, for the various configurations
considered in \cite{ruiz_2020}. The inset figures report the same for
non-magnetized cases, and the coordinate time is shifted such that the
merger time coincides with 0.
}
\label{fig:rest_mass_fraction}
\end{figure}
Aside from the SGRB jet, which requires magnetic field amplification (via MRI) as
well as thermal pair production (from the disk remnant) followed by the
Blandford-Znajek process, there is a possibility that NSBH mergers can produce
kilonovae signatures (\cite{barbieri_2019b}). For this, the dynamically ejected mass
has to be between $10^{-4.5} - 10^{-2} (M_{NS}/1.4 M_{\odot}) M_{\odot}$ (see
\cite{ruiz_2020} for more details), and this will lead to kilonovae potentially
detectable by the Large Synoptic Survey Telescope (LSST).
\begin{figure}[H]
\centering
\includegraphics[width=\textwidth]{nsbh_jet}
\caption[Tidal disruption of a NS in a 3:1 NSBH binary]
{
Volume rendering of the rest mass density ($\rho_0$) (in log scale),
normalized to the NS maximum value $\rho_0 = 8.92 \times 10^{14} (1.4
M_{\odot}/M_{NS})^2 \text{ g/cm}^3$, for particular times for a magnetized
neutron star, with $\mathcal{Q}$ = 3:1 and a prograde NS spin of 0.23. Top
three panels highlight the inspiral and tidal disruption, whereas the
bottom three panels highlight the appearance of the magnetically-driven jet.
White lines denote the magnetic field, arrows denote the fluid velocity and
the BH's apparent horizon is the black sphere. Here M = $2.5 \times 10^{-2}
(M_{NS}/M_{1.4M_{\odot}}) \text{ ms} = 7.58 (M_{NS}/M_{1.4M_{\odot}}) \text{
km}$ (in geometrized units). From \cite{ruiz_2020}.
}
\label{fig:nsbh_jet}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=\textwidth]{nsbh_5to1}
\caption[Tidal disruption of a NS in a 5:1 NSBH binary]
{
Similar to Fig. \ref{fig:nsbh_jet}, however with the NS spin being
0.33, the BH spin being 0 and q = 5:1. In this case, no strong
collimation of the magnetic field is observed from the merger
remnant, and so a magnetically-driven jet is also not observed.
From \cite{ruiz_2020}.
}
\label{fig:nsbh_5to1}
\end{figure}
\subsection{Modelling outflows from NSBH Mergers}\label{ssec:outflows-nsbh}
As mentioned before, the outflows from NSBH mergers depend on the amount of mass
left outside the event horizon post-merger. In the work done by \cite{foucart_2018},
the authors consider a suite of 75 numerical relativity simulations of NSBH mergers
over the parameter space $\mathcal{Q} \in [1, 7]$, $ \chi_{BH} \in [-0.5, 0.97]$,
$\mathcal{C}_{NS} \in [0.13, 0.182]$\footnote{
Here, $\mathcal{Q} = M_{BH}/M_{NS}$ is the mass-ratio, $\chi_{BH} =
c|\mathbf{S}|/GM_{BH}^2$ is the effective spin of the black hole and $C_{NS} =
GM_{NS}/R_{NS}c^2$ is the compactness of the neutron star.
}
and fit the results for the \textit{remnant mass} $M_{\mathrm{rem}}$ (sometimes also
denoted as $M_{\mathrm{out}}$), as a function of the binary parameters (masses,
spins of the components, tidal deformability of the neutron star etc). This fit is
given as follows:
\begin{equation}
M_{\mathrm{out}} =
M_{NS}^b \cdot
\max \left( \alpha \dfrac{1-2\rho}{\eta^{1/3}} -
\beta \hat{R}_{\mathrm{ISCO}} \dfrac{\rho}{\eta} + \gamma, 0 \right)^\delta
\label{eq:m_out}
\end{equation}
where:
\begin{itemize}
\item The baryonic mass of the neutron star is given by the equation $$M_{NS}^b
= M_{NS} \left(1 + \dfrac{0.6 C_{NS}}{1 + 0.5 C_{NS}} \right)$$
\item The tidal deformability of the neutron star is given by $\Lambda_{NS}$ and
$\rho = (15 \Lambda_{NS})^{-1/5}$. It is also related to the compactness of
the neutron star via the C--Love relation (see \cite{yagi_2017}):
\begin{equation}
\label{eq:c_love}
C_{NS} = \sum_{k=0}^{2} a_k (\ln \Lambda_{NS})^k
\end{equation}
Where $a_0 = 0.360, a_1 = -0.0335, a_2 = 0.000705$.
\item $\eta$ is the symmetric mass ratio, given by $ \eta =
\dfrac{\mathcal{Q}}{(1 + \mathcal{Q})^2} $.
\item $\hat{R}_{ISCO} = c^2 R_{ISCO} / GM_{BH}$ is the normalized ISCO radius
for a spinning black hole, given in \cite{bardeen_1972} as :
\begin{align}
\label{eq:r_isco}
\hat{R}_{ISCO} &=
3 +
Z_2 -
\mathrm{sgn}(\chi_{BH}) \sqrt{(3 - Z_1)(3+Z_1 + 2Z_2)} \\
\hookrightarrow Z_1 &=
1 +
(1 - \chi_{BH}^2)^{1/3}
[
(1 + \chi_{BH})^{1/3} + (1 - \chi_{BH})^{1/3}
] \nonumber \\
\hookrightarrow Z_2 &=
\sqrt{3\chi_{BH}^2 + Z_1^2} \nonumber
\end{align}
\item $(\alpha, \beta, \gamma, \delta) \equiv (0.308, 0.124, 0.283, 1.536)$ are
the fit coefficients.
\end{itemize}
Similarly, \cite{kawaguchi_2016} fit to the results of 45 numerical relativity
simulations over the parameter space $\mathcal{Q} \in [3,7]$, $\chi_{\mathrm{BH}}
\in [0, 0.90]$, $C_{\mathrm{NS}} \in [0.138, 0.180]$. This fit produces a formula
for the \textit{dynamic} mass $M_{\mathrm{dyn}}$, which is the unbound mass ejected
at the time of disruption, in terms of the binary parameters. This fit is given as
follows:
\begin{multline}
\dfrac{M_{dyn}}{M_{NS}^b} =
\max \biggl\{
a_1 Q^{n_1}(1 - 2C_{NS})C^{-1}_{NS} -
a_2 Q^{n_2} \hat{R}_{ISCO}(\chi_{BH}) + \\
a_3\left(1 - \dfrac{M_{NS}}{M^b_{NS}}\right) +
a_4, 0
\biggr\}
\end{multline}
Where the symbols have their usual meanings, and additionally:
\begin{align*}
a_1 &= 4.464 \times 10^{-2} & a_2 &= 2.269 \times 10^{-3} \\
a_3 &= 2.431 & a_4 &= -0.4159 \\
n_1 &= 0.2497 & n_2 &= 1.352
\end{align*}
From these two quantities, one can derive the disc mass $M_{\mathrm{disc}}$ as :
\begin{equation}
M_{\mathrm{disc}} = \max\{M_{\mathrm{out}} - M_{\mathrm{dyn}}, 0\}
\label{eq:disc_mass}
\end{equation}
However, due to the fact that these two fits are derived from simulations over
different regions of the input parameter space, care must be taken while applying
them together. This is to ensure that $M_{\mathrm{dyn}} \leq M_{\mathrm{out}}$
always, so that the disc mass is non-negative. This validation is performed by
considering the ratio $M_{\mathrm{out}} / M_{\mathrm{dyn}}$. Another constraint is
imposed, which is motivated by the fact that NSBH simulations carried out by
\cite{foucart_2019} in the near-equal mass ratio regime found an unbound component
no more massive than roughly 30\% of the total remnant mass (note that one expects
maximal tidal disruption in this regime, given a fast spinning black hole). Thus one
sets:
\begin{equation}
\label{eq:constraint}
M_{\mathrm{dyn, max}} = f \cdot M_{\mathrm{rem}} = 0.3 \cdot M_{\mathrm{rem}}
\end{equation}
Additionally, the masses of the other wind ejecta, namely the neutrino-driven and
viscosity-driven wind ejecta, are derived from that of the disc mass:
\begin{align}
\begin{split}
M_{\mathrm{vis}} &=
\xi_{\mathrm{vis}}M_{\mathrm{disc}} =
0.2M_{\mathrm{disc}} \\
M_{\nu} &=
\xi_{\nu}M_{\mathrm{disc}} =
0.01 M_{\mathrm{disc}}
\end{split}
\end{align}
To model the SGRB jet, the procedure of \cite{zhu_2020} is followed. The
kinetic energy of the jet is decided by the disc mass and the black hole spin as
follows:
\begin{equation}
E_{\mathrm{K, jet}} =
\epsilon(1 - \xi_{\mathrm{vis}} - \xi_{\nu})
M_{\mathrm{disc}} c^2 \Omega_H^2 f(\Omega_H)
\label{eq:e_kin_jet}
\end{equation}
Where:
\begin{itemize}
\item The dimensionless angular velocity at the horizon of the black hole is
given by:
\begin{equation}
\Omega_H = \dfrac{\chi_{BH}}{2(1 + \sqrt{1 + \chi_{BH}^2}}
\label{eq:omega_func}
\end{equation}
\item $f(\Omega_H)$ is a high-spin correction factor given by:
\begin{equation}
f(\Omega_H) = 1 + 1.38\Omega_H^2 - 9.2 \Omega_H^4
\label{eq:Omega_h}
\end{equation}
\item $\epsilon$ is a fudge factor which depends on the large-scale geometry of
the magnetic field, disc aspect ratio and the ratio of the magnetic field
energy density to disc pressure at saturation.\\ In order to set it to a
definite value, it is noted that the maximum disc mass cannot exceed the
total NS baryonic mass i.e. $M_{\mathrm{disc}} \lesssim 2M_\odot$. Also, the
spin-dependent factor $\Omega_H^2f(\Omega_H)$ cannot exceed 0.2 (since
$\chi_{BH} \in [-1, 1]$). Furthermore, the most energetic of SGRBs has had a
$E_{\gamma, \mathrm{iso}} \sim 7.4 \times 10^{52}$ erg, and if one assumes a
10\% conversion efficiency of kinetic to gamma-ray energy along with a
typical jet opening angle of 5$^{\circ}$, this corresponds to a kinetic
energy of $E_{\mathrm{K, jet}} \sim 3 \times 10^{52}$ erg.\\
Based on this, one can calculate $\boxed{\epsilon \approx 0.015}$.
\end{itemize}
However, note that in Eq. \ref{eq:omega_func}, the value of $\chi_{BH}$ is the
\emph{spin of the remnant BH}, which can be computed solving the following implicit
equation (\cite{pannarale_2013}):
\begin{equation}
\chi_{BH, f} = \dfrac
{
\chi_{BH, i}M_{BH}^2 + l_z(\hat{R}_{ISCO, f}, \chi_{BH, f}) M_{BH}
[ (1 - f(\eta))M_{NS} + f(\eta) M_{b, NS} - M_{out}) ]
}
{
[M (1 - ( 1 - e(\hat{R}_{ISCO, i}, \chi_{BH, i}) ) \eta ) -
e(\hat{R}_{ISCO, i}, \chi_{BH, i}) M_{out}]
}
\end{equation}
For $\chi_{BH, f}$, where:
\begin{itemize}
\item $\chi_{BH, i}$ and $\chi_{BH, f}$ are the initial and final spins of the
BH and BH remnant respectively.
\item $M_{BH}$ and $M_{NS}$ are the masses of the BH and NS respectively. $M_{b,
NS}$ and $M_{out}$ are the baryonic mass of the NS and the remnant mass for
the particular NSBH merger calculated using Eq. \ref{eq:m_out} respectively.
Finally, $M$ is the total (initial) mass of the binary, and $\eta =
\frac{\mathcal{Q}}{(1 + \mathcal{Q})^2}$ is the initial symmetric mass
ratio. All mass quantities are expressed in geometrized units, i.e. where $G
= c = 1$.
\item $l_z(\hat{R}_{ISCO, f}, \chi_{BH, f})$ is the orbital angular momentum per
unit mass of a test particle orbiting the remnant BH at the ISCO. Note that
here $\hat{R}_{ISCO}$ plays the role of the (dimensionless) radial
Boyer-Lindquist coordinate, useful for describing the orbits in the Kerr
spacetime around this rotating BH remnant. Additionally, $\hat{R}_{ISCO, f}$
refers to the value computed by using the final BH spin in Eq.
\ref{eq:r_isco}. Also for a given radial coordinate, $\hat{R}$ and a BH
spin, $\chi$ this orbital angular momentum is given as:
\begin{equation}
l_z(\hat{R}, \chi) = \mathrm{sgn}(\chi)
\left[
\dfrac
{
\hat{R}^2 - \mathrm{sgn}(\chi) \cdot 2 \chi \sqrt{\hat{R}} +
\chi^2
}
{
\sqrt{\hat{R}}
(
\hat{R}^2 - 3\hat{R} + \mathrm{sgn}(\chi) \cdot 2 \chi
\sqrt{\hat{R}}
)^{1/2}
}
\right]
\label{eq:AM_boyerlindquist}
\end{equation}
Where $\mathrm{sgn}(\chi)$ is positive/negative for prograde/retrograde
orbits.
\item $e(\hat{R}_{ISCO}, \chi_{BH, f})$ is the energy per unit mass of a test
particle orbiting the remnant BH at the ISCO. For equatorial Kerr orbits at
a radial coordinate, $\hat{R}$ around a BH with a spin, $\chi$ it
is given as:
\begin{equation}
e_z(\hat{R}, \chi) =
\dfrac
{
\hat{R}^2 - 2 \hat{R} + \mathrm{sgn}(\chi) \cdot \chi
\sqrt{\hat{R}}
}
{
\hat{R}
(
\hat{R}^2 - 3\hat{R} + \mathrm{sgn}(\chi) \cdot 2 \chi
\sqrt{\hat{R}}
)^{1/2}
}
\label{eq:E_boyerlindquist}
\end{equation}
\item $f(\eta)$ is a transition function to smooth the behaviour of this
equation in the transition from large ($\mathcal{Q} \geq 4$) mass ratio
mergers to smaller ($\mathcal{Q} \in (2, 4]$) mass ratio mergers. This
function is not uniquely defined, but \cite{pannarale_2013} defines it using
the following properties which are validated by comparing with numerical
relativity simulations:
\begin{itemize}
\item $f(\eta \geq 2/9) = 1$ and $f(\eta \leq 0.16) = 0$. This means
that for mergers with relatively large mass ratios ($\mathcal{Q}
\geq 4 \Leftrightarrow \eta \leq 0.16$) the function `switches off'
and for lower mass ratios ($\mathcal{Q} \leq 2 \Leftrightarrow
\eta \geq 2/9$) the function `switches on'.
\item $\dfrac{df}{d\eta} \geq 0 \quad \forall \eta \in (0, 0.25]
\Leftrightarrow \mathcal{Q} \in (1, \infty) $. This means that the
transition function is monotonically increasing. Additionally
for simplicity, it is assumed to be $C^\infty$.
\end{itemize}
With these constraints in mind, the transition function is defined as:
\begin{equation}
f(\eta) =
\begin{cases}
0, & \eta \leq 0.16 \\
\dfrac{1}{2}\left[
1 - \cos\left(
\dfrac{\pi(\eta - 0.16)}{2/9 - 0.16}
\right)
\right], & 0.16 < \eta < 2/9 \\
1, & 2/9 \leq \eta \leq 0.25
\end{cases}
\label{eq:transition_func}
\end{equation}
\end{itemize}
Now, given the jet energetics using the equations above, one can define the
structure of the SGRB jet, given by the following equations:
\begin{equation}
\dfrac{dE(\theta)}{d\Omega} =
\dfrac{E_{\mathrm{k, jet}}}{\pi \theta_{\mathrm{c, E}}^2}
e^{-(\theta/\theta_{\mathrm{c, E}})^2}
\label{eq:dE_dOmega}
\end{equation}
\begin{equation}
\Gamma(\theta) = (\Gamma_c - 1)e^{-(\theta/\theta_{\mathrm{c, E}})^2} + 1
\end{equation}
\begin{equation}
E_{\mathrm{iso}}(\theta_v) =
\eta \int \dfrac{\delta^3}{\Gamma} \dfrac{dE}{d\Omega} d\Omega
\label{eq:eiso}
\end{equation}
where:
\begin{itemize}
\item $\Gamma_c = 100, \theta_{\mathrm{c, E}} = 0.1, \theta_{\mathrm{c}, \Gamma}
= 0.2$. See \cite{salafia_2015} and \cite{barbieri_2019a}.
\item $\eta$ is the conversion efficiency of gamma-ray energy to kinetic energy,
which is traditionally taken to be 10\%.
\item $\delta$ is the Doppler factor, given by $$\delta = \dfrac{1}{\Gamma[1 -
\beta \cos \alpha_v]}$$ where $\alpha_v$ is the angle between the jet
element at $(\theta, \phi)$ and the observer's direction.
\end{itemize}
\section{NS mergers in GW regime}
\label{sec:ns_in_gw}
Consider any astrophysical source emitting gravitational waves, which come in two
polarizations, namely the \emph{plus} $h_{+}(t; \mathbf{\Theta}_{GW})$ and the
\emph{cross} $h_{\times}(t; \mathbf{\Theta}_{GW})$ polarizations. Here
$\mathbf{\Theta}_{GW}$ is the parameter vector, typically taken to be $\{m_1, m_2,
\mathbf{\chi}_1, \mathbf{\chi}_2, D_L, \iota, t_c, \phi_c\}$ which are the component
masses, component spins, the binary's luminosity distance, the inclination angle
of the orbital plane with respect to the line of sight, and two constants of
integration: the time and phase of coalescence, respectively.\\
A detector's response is recorded as the GW strain such waves produce, but this is
done in the frequency domain and so the input signals are Fourier transformed before
processing. Also the detector's antenna patterns (sensitivity as a function of the
source location on the sky) and location phase factor (effect of the earth's
rotation) play a role in the response. Thus, the detector response to these
gravitational waves is of the form:
\begin{multline}
H(f; \mathbf{\Theta}) =
F_{lp}(f; \alpha, \delta) \cdot
[
H_{+}(f; \mathbf{\Theta}_{GW}) F_{+}(f; \alpha, \delta, \psi) + \\
H_{\times}(f; \mathbf{\Theta}_{GW}) F_{\times}(f; \alpha, \delta, \psi)
]
\end{multline}
Where $F_{lp}$ is the location phase factor as a function of the frequency and the
source right-ascension and declination, $H_{+/\times}$ are the frequency domain
waveforms, and $F_{+/\times}$ are the detector antenna patterns for each
polarization. Also $\mathbf{\Theta} = \{\mathbf{\Theta}_{GW}, \alpha, \delta,
\psi\}$.\\ The sensitivity of a detector is given by the detector's noise $n(t)$ and
its autocorrelation $\kappa = \overline{n(t_1)n(t_2)}$. Usually, the noise is
assumed to be stationary, zero-mean and Gaussian. Thus, one can define the one-sided
power spectral density $S_n(f)$ as the Fourier transform of the autocorrelation.\\
From this, one can define the `overlap' between two GW signals (for e.g.: the
detector responses for two different waveforms) using the noise-weighted scalar
product:
\begin{equation}
\langle H, G \rangle =
2 \int_{0}^{\infty} \dfrac{H(f)G^{\ast}(f) + H^{\ast}(f)G(f)}{S_n(f)} df
\end{equation}
And using this definition of the scalar product, the signal-to-noise ratio is
defined as:
\begin{equation}
\rho^2 =
\langle H, H \rangle =
4 \int_0^\infty \dfrac{|H(f)|^2}{S_n(f)} df
\end{equation}
Now, since the noise $n(t) = s(t) - h(t)$ is assumed to a zero-mean Gaussian, its
Fourier transform also behaves the same way, and thus the probability of noise can
be written down as :
\begin{equation}
\label{eq:probability}
p(\mathbf{\Theta}) =
p^0(\mathbf{\Theta})
e^{
-\frac{1}{2}
\langle S - H(\mathbf{\Theta}), S - H(\mathbf{\Theta}) \rangle
}
\end{equation}
Where $p^0$ is the prior on the parameter vector of the detector response. Assuming
that an event signal S has a high SNR, the value of $\mathbf{\Theta}$ at peak
probability is a good estimate of the true value $\mathbf{\Theta}^{\ast}$.
Additionally peak probability occurs when the exponential $E = \langle S - H, S - H
\rangle$ is the largest. Expanding it around the maximum value:
\begin{equation}
E(\mathbf{\Theta}) =
E(\mathbf{\Theta}^{\ast}) +
\dfrac{1}{2}
\dfrac{\partial^2 E(\mathbf{\Theta})}{\partial \Theta_i \partial \Theta_j}
\bigg\rvert_{\mathbf{\Theta} =
\mathbf{\Theta}^{\ast}}
\Delta\Theta_i \Delta\Theta_j +
\cdots
\end{equation}
Where $\Delta \Theta_i = \Theta_i - \Theta_i^\ast$. The Hessian given by:
\begin{equation}
\dfrac{\partial^2 E(\mathbf{\Theta})}{\partial \Theta_i \partial \Theta_j} =
2 \langle
\partial_{\Theta_i} H(\mathbf{\Theta}),
\partial_{\Theta_j} H(\mathbf{\Theta})
\rangle +
\langle
\partial_{\Theta_i} \partial_{\Theta_j} H(\mathbf{\Theta}),
N
\rangle
\end{equation}
Can be simplified for large SNR, where second-order differentials become negligible.
This leads to the definition of the Fisher Information Matrix $\Gamma$:
\begin{equation}
\Gamma_{ij} =
\langle
\partial_{\Theta_i} H(\mathbf{\Theta}),
\partial_{\Theta_j} H(\mathbf{\Theta})
\rangle
\end{equation}
And hence, Eqn. \ref{eq:probability} becomes:
\begin{equation}
p(\mathbf{\Theta}) \sim
\exp\left(
-\dfrac{1}{2}
\Gamma_{ij}
\Delta\Theta_i
\Delta\Theta_j
\right)
\end{equation}
Which implies that the assumption of Gaussian noise helps associate the FIM to the
inverse of the covariance matrix $\Sigma \equiv \Gamma^{-1}$. This also means that
the diagonal and off-diagonal elements of $\Gamma^{-1}$ denote the variances and
covariances of the parameters, respectively, with 1$\sigma$ estimates of the error
are given as $\sigma_{\Theta_i} = \sqrt{\Sigma_{ii}}$.\\
This Fisher Information Matrix (FIM) formalism is a method of rapid GW data
analysis, which approaches the accuracy of traditional Bayesian parameter estimation
for events with large SNR. Sometimes this method is also referred to as the Fisher
Information Formalism (FIF), in which case the Fisher information is not a matrix
but is instead the variance of the partial derivative with respect to the parameter
vector, of the natural logarithm of the likelihood function for the random variable
whose parameters are to be estimated.\\
Software tools such as \textbf{GWBENCH} (see \cite{borhanian_2020}), leverage this
formalism to compute the Fisher matrix for a particular NS merger, given the network
configuration and binary parameters. In this way, it enables rapid calculations to
benchmark detector upgrades as well as forecast the confidence with which parameters
may be estimated for the NS mergers in question.\\
\section{Summary}
NS mergers can present an ideal testing environment for physical theories under
extreme gravity, and by observing them in both the EM and GW windows, current
theories can be better understood and refined. Several questions also remain about
the exact mechanisms which power the outflows from these NS mergers.\\
In this report, the focus is mainly on SGRB jets from NSBH mergers and in pursuit of
the same, population synthesis studies are carried out to infer the conditions for
and implications of observing a SGRB jet from NSBH mergers.