Fast Radio Bursts - I: Initial Cogitation

Similar documents
arxiv: v1 [astro-ph.he] 24 Nov 2015

arxiv: v1 [astro-ph.he] 20 Aug 2015

Fast Radio Bursts. Laura Spitler Max-Planck-Institut für Radioastronomie 11. April 2015

SETI and Fast Radio Bursts

FAST RADIO BURSTS. Marta Burgay

arxiv: v1 [astro-ph.he] 24 Dec 2015

FRB : A Repeating Fast Radio Burst. Laura Spitler 20. June 2016 Bonn Workshop IX

arxiv: v1 [astro-ph.he] 21 Dec 2018

Fast Radio Bursts. The chase is on. Sarah Burke Spolaor National Radio Astronomy Observatory

arxiv:astro-ph/ v1 10 Nov 1999

PoS(ISKAF2010)083. FRATs: a real-time search for Fast Radio Transients with LOFAR

Fast Radio Burts - a literature review

A SEARCH FOR FAST RADIO BURSTS WITH THE GBNCC SURVEY. PRAGYA CHAWLA McGill University (On Behalf of the GBNCC Collaboration)

arxiv: v1 [astro-ph] 14 Jul 2008

Fast Radio Transients and Next- Generation Instruments In Search of the Rare and Elusive. Jean-Pierre Macquart

Limits on Fast Radio Bursts at 145 MHz with ARTEMIS, a real-time software backend

THE HIGH TIME RESOLUTION UNIVERSE. A survey for pulsars & fast transients

Discovery of fast radio transients at very low frequencies

Fast Transients with the Square Kilometre Array and its Pathfinders: An Indian Perspective

FAST RADIO BURSTS. Marta Burgay

Cosmological Fast Radio Bursts from Binary White Dwarf Mergers

The FRB Population: Observations and Theory

arxiv: v2 [astro-ph.he] 31 Aug 2017

arxiv: v1 [astro-ph.he] 23 Nov 2018

A Population of Fast Radio Bursts at Cosmological Distances

Discovery of two pulsars towards the Galactic Centre

Radio Aspects of the Transient Universe

Search for Neutrino Emission from Fast Radio Bursts with IceCube

The High Time Resolution Universe Pulsar Survey III. Single-pulse searches and preliminary analysis

The millisecond radio sky: transients from a blind single-pulse search

Estimate of an environmental magnetic field of fast radio bursts

Fast Radio Bursts: Recent discoveries and future prospects

THE PARKES MULTIBEAM PULSAR SURVEY

arxiv: v3 [astro-ph.he] 22 Mar 2014

arxiv: v1 [astro-ph.co] 27 Sep 2010

arxiv: v1 [astro-ph.he] 2 Mar 2016

Mike Garrett. Sir Bernard Lovell Chair, Prof. of Astrophysics. Director Jodrell Bank Centre for Astrophysics

Survey for Pulsars and Extra-galactic Radio Bursts

A New Model for the Beams of Radio Pulsars

The Mystery of Fast Radio Bursts and its possible resolution. Pawan Kumar

arxiv: v1 [astro-ph.he] 10 Oct 2018

A SEARCH FOR FAST RADIO BURSTS AT LOW FREQUENCIES WITH MURCHISON WIDEFIELD ARRAY HIGH TIME RESOLUTION IMAGING

1. GAMMA-RAY BURSTS & 2. FAST RADIO BURSTS

Transient radio bursts from rotating neutron stars

HI Galaxy Science with SKA1. Erwin de Blok (ASTRON, NL) on behalf of The HI Science Working Group

arxiv: v1 [astro-ph.he] 1 Nov 2018

Fast Radio Bursts from neutron stars plunging into black holes arxiv: v1 [astro-ph.he] 24 Nov 2017

Pulsar Surveys Present and Future: The Arecibo-PALFA Survey and Projected SKA Survey

DISCOVERY OF 14 RADIO PULSARS IN A SURVEY OF THE MAGELLANIC CLOUDS

Recent Radio Observations of Pulsars

Pulsar Overview. Kevin Stovall NRAO

ASKAP Array Configurations: Options and Recommendations. Ilana Feain, Simon Johnston & Neeraj Gupta Australia Telescope National Facility, CSIRO

The Square Kilometre Array Epoch of Reionisation and Cosmic Dawn Experiment

Radio Transients, Stellar End Products, and SETI Working Group Report

Pulsars and Radio Transients. Scott Ransom National Radio Astronomy Observatory / University of Virginia

FRBs, Perytons, and the hunts for their origins

Transient Cosmic Phenomena and their Influence on the Design of the SKA Radio Telescope

The high angular resolution component of the SKA

The Dynamic Radio Sky: On the path to the SKA. A/Prof Tara Murphy ARC Future Fellow

arxiv: v1 [astro-ph.sr] 9 Nov 2009

arxiv:astro-ph/ v1 16 Nov 1999

High Energy Astrophysics

arxiv: v1 [astro-ph.sr] 7 Dec 2015

Fast Transient Science in the VLASS

154 MHz detection of faint, polarised flares from UV Ceti

Simultaneous single-pulse observations of radio pulsars. III. The behaviour of circular polarization

arxiv: v2 [astro-ph.he] 30 Mar 2016

Pulsars. in this talk. Pulsar timing. Pulsar timing. Pulsar timing. Pulsar timing. How to listen to what exotic. are telling us! Paulo César C.

arxiv: v1 [astro-ph.he] 4 Feb 2014

arxiv: v1 [astro-ph] 3 Nov 2008

Gravity with the SKA

A real-time fast radio burst: polarization detection and multiwavelength follow-up

The pulsar population and PSR/FRB searches with MeerKAT Slow boring pulsars

HI Surveys and the xntd Antenna Configuration

FLUX DENSITIES AND RADIO POLARIZATION CHARACTERISTICS OF TWO VELA-LIKE PULSARS

LOFAR for Pulsar. Ben Stappers ASTRON/UvA

arxiv: v1 [astro-ph.sr] 27 Jun 2012

Radio timing observations of the pulsar by Kaspi et al. (1994) have

Search Strategies. Basic Problem: where to look? Possible Scenarios Powerful, omnidirectional beacons

Much Ado About Nothing: Several Large-Area Surveys For Radio Pulsars From Arecibo

Unusual orbits in the Andromeda galaxy Post-16

SKADS Virtual Telescope: Pulsar Survey IV: Globular Cluster Pulsars

The Search for Extraterrestrial Intelligence (SETI) What can SETI

Astronomy 421. Lecture 23: End states of stars - Neutron stars

arxiv: v1 [astro-ph.ga] 22 Dec 2009

arxiv: v1 [astro-ph.im] 30 Oct 2012

arxiv: v1 [astro-ph.he] 4 Mar 2016 Accepted XXX. Received YYY; in original form ZZZ

The Square Kilometer Array. Presented by Simon Worster

Rotating RAdio Transients (RRATs) ApJ, 2006, 646, L139 Nature, 2006, 439, 817 Astro-ph/

The (obscene) Challenges of Next-Generation Pulsar Surveys

Determining the Specification of an Aperture Array for Cosmological Surveys

The Parkes Multibeam Pulsar Survey - III. Young Pulsars and the Discovery and Timing of 200 Pulsars

Observatory. On behalf of RANGD project & Radio Astronomy Group and S. Koichiro (NAOJ)

The Eight-meter-wavelength Transient Array

Design Reference Mission for SKA1 P. Dewdney System Delta CoDR

Correcting for the solar wind in pulsar timing observations: the role of simultaneous and low-frequency observations

The dispersion-brightness relation for fast radio bursts from a wide-field survey

arxiv: v1 [astro-ph] 3 Jun 2008

ETA Observations of Crab Pulsar Giant Pulses

The Parkes Multibeam Pulsar Survey II. Discovery and timing of 120 pulsars

Transcription:

Mon. Not. R. Astron. Soc. 000, 000 000 (0000) Printed 23 September 2014 (MN LATEX style file v2.2) Fast Radio Bursts - I: Initial Cogitation E.F. Keane 1,2 1 Centre for Astrophysics and Supercomputing, Swinburne University of Technology, Mail H30, PO Box 218, VIC 3122, Australia. 2 ARC Centre of Excellence for All-sky Astrophysics (CAASTRO). arxiv:1409.6125v1 [astro-ph.he] 22 Sep 2014 23 September 2014 1 INTRODUCTION ABSTRACT Fast radio bursts (FRBs) are millisecond-duration janskyflux density signals discovered by single dish radio telescopes operating at frequencies of 1.4 GHz (Lorimer et al. 2007; Keane et al. 2011; Thornton et al. 2013; Spitler et al. 2014; Burke-Spolaor & Bannister 2014; Petroff et al. 2014b). All but one were detected with the 64-m Parkes Telescope; the other with the 300-m dish at Arecibo. The bursts exhibit a frequency-dependent time delay which obeys a quadratic form so strictly that the signals could only have traversed low density regions, such as the interstellar and intergalactic media, en route to Earth (Dennison 2014). The magnitude of this delay parametrised by the dispersion measure (DM), which is the integrated electron density along the line of sight is so large that cosmological distances are inferred for the sources of the FRBs (Ioka 2003; Inoue 2004). There are two main questions of interest in relation to FRBs: (i) what are they? (ii) what can they be used for? The source of the FRB signals is hotly debated in the literature with suggested progenitors ranging from terrestrial interference and flare stars (Burke-Spolaor et al. 2011; Loeb et al. 2014) to neutron star-neutron star mergers, gammaray bursts and planetary companions to extragalactic pulsars (Totani 2013; Zhang 2014; Mottez & Zarka 2014). However, the two leading candidate theories are that FRB bursts are associated with giant magnetar flares (Kulkarni et al. 2014; Lyubarsky 2014), or that they are blitzars occurring when a neutron star just above the Tolmann-Oppenheimer- Volkoff limit collapses to a black hole (Falcke & Rezzolla 2014). Amongst other observables which would distinguish between these two possibilities (see e.g. Ravi & Lasky 2014) Fast radio bursts (FRBs) are millisecond-duration radio signals thought to originate from cosmological distances. Many authors are endeavouring to explain their progenitors, with others outlining their potential uses as cosmological probes. Here we describe some sub-optimal performance in existing FRB search software, which can reduce the volume probed by over 20%, and result in missed discoveries and incorrect flux densities and sky rates. Recalculating some FRB flux densities, we find that FRB 010125 was approximately 50% brighter than previously reported. Furthermore we consider incompleteness factors important to the population statistics. Finally we make data for the archival FRBs easily available, along with software to analyse these. Key words: surveys intergalactic medium methods: data analysis we simply note that magnetar flares repeat whereas blitzars are one-off events. For an expansive discussion of FRB origin possibilities we refer the reader to Kulkarni et al. (2014). The uses of FRBs are many: they have the potential to be used as astophysical/cosmological tools to make several important measurements. FRBs can be used to weigh the missing baryons, as it is the ionised component of these same baryons which cause the dispersion of the FRB signal (McQuinn 2014). The measurement of the rotation measure of a linearly polarised FRB would tell us the magnetic field strength of the intergalactic medium, analagously to what has been done in the Milky Way using pulsars, see e.g. Noutsos et al. (2008). If a sufficiently large sample of FRBs were identified, with independent redshift measurements from observations at other wavelengths the distribution of the DM values as a function of redshift can be used as an independent measure of the dark energy equation of state (Gao et al. 2014; Zhou et al. 2014). Determining progenitors and fully exploiting the exciting cosmological possibilities demands that we detect more FRBs. To do this we have undertaken a large-scale FRB search of existing archival data, which we will describe in a series of papers. In this first paper, we set the scene, by discussing, in 2, commonly used search techniques including some flaws therein that can unnecessarily reduce the sensitivity of the search, and the impacts of this. In 3 we discuss three known archival FRBs. In 4 we discuss additional incompleteness factors one ought to consider when deriving population statistics. Finally, in 5 we present our conclusions and highlight our online data and code repositories which we hope will encourge further FRB analysis.

2 E.F. Keane 2 SINGLE PULSE SEARCHES FRBs are detected as follows: (i) Acquisition: Radio telescopes record incoherent filterbank data: these are timefrequency-flux density data cubes. Thusfar the data wherein FRBs have been discovered have been centred at 1.4 GHz with bandwidths ranging from 288 400 MHz, frequency resolutions ranging from 0.336 3 MHz and time resolutions ranging from 0.064 1 ms; (ii) Cleaning: The data are cleaned of radio frequency interference signals in various ways; (iii) Dedispersion: The data are dedispersed at a number of trial dispersion measure (DM) values to remove frequency-dependent delays imparted by the interstellar and intergalactic media; (iv) Search: Each dedispersed time series is subjected to a single pulse (SP) search, which is a matched filter search to a number of trial boxcar widths. Usually events down to a level which is well within the noise floor are recorded; (v) Refinement: Upon detection optimised DM and width values of the pulse are derived. In steps (ii) (iv) there is the potential for a loss in sensitivity. All of these are avoidable, but accuracy is sometimes sacrificed for processing speed. The DM parameter is always covered in a scalloped fashion, where the next DM trial is chosen so as to limit the sensitivity loss of a narrow pulse falling between DM trials. Typically the choice is to lose no more than 10% of the sensitivity for bursts narrower than the sampling time. However FRBs are typically much wider than the sampling time, and the observed width is dominated by dispersion smearing so that the loss in sensitivity to FRBs is typically much less than this. As DM corresponds to the volume probed in a line-of-sight dependent way, the actual volume probed can be quite uncertain, especially for lines-of-sight closer to the Galactic plane. The searching step can be subject to the root 2 problem, which manifests itself when performing a decimation search. This is a procedure in which a time series is searched for events of 1 sample in width. It is then down-sampled by a factor of 2, averaging adjacent samples. This process is repeated a number of times in order to search for a range of pulse widths (Cordes & McLaughlin 2003). However, this search method is not optimal. For example, let s consider a time series with samples i, i=1,2,3,4,..., and a top-hat pulse which is 4 bins wide, occupying bins 3,4,5 and 6. This pulse is out of phase with respect to the down-sampling procedure and it is clear that the derived S/N will be too low by a factor of 2. The optimal way to search a time series is to run a sliding boxcar along the time series. Figure 1 illustrates this problem by showing the results of searching for a synthesised FRB as it is moved along a time series in single time sample steps up to one pulse width in total. The results of several commonly used SP search codes are shown. In particular these are heimdall 1, dedisperse all 2, seek 2 destroy 3. Here a typical FRB with an intrinsic pulse width of 2 ms (16 bins for the simulated time sampling value of 125 µs), a DM of 1000 cm 3 pc, and an injected S/N of 16, is used. The data are centred at a frequency of 1374 MHz with spectral resolution of 3 MHz meaning the dispersion smearing will 1 http://sourceforge.net/projects/heimdall-astro/ 2 see e.g. https://github.com/sixbynine/psrsoft 3 https://github.com/evanocathain/destroy gutted Recovered S/N 16 14 12 10 8 6 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 Boxcar Phase (arbitrary zero point) Heimdall Dedisperse All Destroy Seek Figure 1. Here we show the detected S/N of a simulated FRB (intrinsic pulse width 2 ms, DM of 1000 cm 3 pc observed at 1.4 GHz) for four commonly used SP search codes, as a function of boxcar phase. The injected S/N in this case is 16. Due to dispersion smearing, and searching for power-of-2 boxcars the maximum theoretical detection S/N is 15.4. Here we plot the mean recovered S/N values for 100 random realisations. We can see that heimdall (circles) and destroy (crosses) recover the correct S/N regardless of the position of the phase. dedisperse all (squares) and seek (triangles) give S/N values which are strongly dependent on phase, with a maximum loss factor of 2. The rms deviation of the recovered S/N values (not shown) are in all cases 1, as expected. Also shown are various thresholds: a 10-σ threshold which is often used for realtime FRB searches (Petroff et al. 2014b), a 8-σ threshold more commonly used for offline processing, two false-alarm thresholds for typical pulsar search parameters: the lower (higher) dashed line for observations with 2 21 (2 23 ) samples, 1000 DM trials and 10 width trials. make the observed pulse width 7.4 ms (59 bins). Thus, for a power-of-2 boxcar search the optimal S/N we expect to find is 16 max( 32/59, 59/64) = 16 59/64 = 15.4. We can see that dedisperse all reaches the maximum theoretical S/N when the pulse is in phase and reaches a minimum S/N when the pulse is out of phase. seek has the same problem, although to a lesser extent, and is relatively rotated in phase and with a response curve which repeats at twice the frequency. These differences are because seek performs an extra 2-bin smoothing step to the data prior to each down-sampling step. heimdall and destroy give the correct result for all phases. We have verified that the curves in Figure 1 scale directly with the injected S/N. One can quickly make a simple estimate of the effects of these results on a survey, e.g. for an N S 3/2 min law seek probes 86% and dedisperse all only 78% of the volume probed by heimdall and destroy. The true volume probed will be even lower than this however as here the exact DM of the pulse has been used, so that the scallop loss factor (which would effect all four codes equally) has been removed. Crucially this can mean that FRBs detectable in our data are never detected. Some of these errors can also result in incorrect flux density and volumetric rate estimates for those FRBs which are detected.

FRBs I 3 3 ARCHIVAL FRBS Here we look at the three archival FRBs reported so far in the literature. We do not examine the bursts reported by Thornton et al. (2013), Spitler et al. (2014) or Petroff et al. (2014b). The FRBs, which all occurred in 2001, are discussed in the chronological order of their discovery which, as it happens, is the reverse chronological order to when they actually occurred. FRB 010724 FRB 010724, the so-called Lorimer Burst, the first FRB to be discovered, was by far the strongest of any FRB detection to date (see Figure 2). It saturated the 1-bit digitiser in the main beam in which it was detected, and was additionally detected in 3 other beams (Lorimer et al. 2007; Burke-Spolaor et al. 2011). Using destroy we determine the S/N value and corresponding width values to be > 100 and 20 ms (beam 6), 16(1) and 9 +12 2 ms (beam 7), 6(1) and 33 +12 28 ms (beam 12), and 27(1) and 15+4 3 ms (beam 13), and < 5 over a wide range of widths in all other beams. The S/N estimate for beam 6 is a lower limit due to the digitiser saturation (Lorimer et al. 2007). The corresponding detected peak flux density values are > 1500 mjy (beam 6), 357 +73 138 554 +88 80 mjy (beam 7), 83+165 24 mjy (beam 12), and mjy (beam 13). The limits for the other beams are, assuming the same width as for the weakest detection in beam 6, < 54 mjy (beam 1), < 58 mjy (inner ring, 6, 7) and < 69 (outer ring, 12, 13). We note that we have not included the additional source of uncertainty (which is at the 20 30% level) which arises from the scaling to flux density units. One can use the multiple detections to try to refine the localisation of the burst. Noting the beam configuration and invoking a simple Gaussian beam model (Staveley- Smith et al. 1996) it is straightforward to show that for two beams with relative detected flux densities r, full-width halfmaxima f and beam separation s, the angular offset to the true position of the source, along the line joining the two beams, is given by: s/2 + (1/s)(f/2.355) 2 ln r. The offset in the perpendicular direction is unconstrained for a single pair. Using two pairs of beams one can define a point (or region, when the uncertainties are included). Using a third, or more, pairs gives a consistency check. Applying this method, using the values quoted above, and ignoring the uncertainties in the flux densities, defines a region on the sky whose centre is at (α, δ) = (19.154, 75.155), where α is in degrees of right ascension and δ is in degrees of arc. The offset with respect to the centre of beam 6 is ( α, δ ) = ( 0.371, 0.050), i.e. the position is WSW of beam 6. The errors on the position are large, i.e. of the same order as this offset. They are at least 5.1 arcmin: (σ α, σ δ ) = (0.332, 0.085), and are about double this if the uncertainty in the flux densities (mentioned above) are included. So it seems that FRB 010724 occured somewhere within the full-width half-power range of beam 6. In comparison to the estimate of Burke-Spolaor et al. (2011) we are in agreement, with our uncertainties slightly larger 4. Assuming our above position would imply a true 4 The uncertainty in right ascension in Burke-Spolaor et al. flux density of at least > 1.5 Jy/G(6.4 arcmin) > 2.7 Jy, where G denotes the angular response of the beam. Our estimate and that of Burke-Spolaor et al. (2011) differ from the kite-shaped region determined by Kulkarni et al. (2014), perhaps because that analysis did not include the detection in beam 12. In theory, with knowledge of the frequencydependent beam response one could also estimate the true spectral index (which is highly degenerate with position) if sufficient S/N existed to measure the spectral index in multiple beams. We do not attempt this (and our analysis implicitly assumes a flat spectral index) as (i) we do not have such a beam response model; and (ii) our position is less localised than that of Burke-Spolaor et al. (2011), who determined a spectral index in the range 2.5 to 0.6, the uncertainty we would obtain would therefore be worse than this. Clearly the uncertainty in FRB spectra is huge, and dramatically so for all FRBs detected in just one beam, as illustrated well in Spitler et al. (2014) where an uncertainty of an order of magnitude in the spectral index is noted. We also note the existence of a previously unreported narrowband signal trailing the FRB (see Figure 2). This is seen in a single frequency channel only (between frequencies 1395 and 1398 MHz), is 20(2) ms in duration, and trails the FRB by 46(2) ms. The S/N of this pulse is difficult to determine given the saturation, but is nominally 10. This channel is not known to have been subject to radio frequency interference, e.g. in the Parkes Multi-beam Pulsar Survey (Manchester et al. 2001), which used this receiver over the course of several years around the time of the FRB. This pulse is not seen in any of the other 12 beams. On the face of it, this is statistically significant, although we can offer no explanation for its origin. To allow the reader make an informed assessment we have not removed the known bad channels in Figure 2. FRB 010621 FRB 010621 was the second FRB to be identified (Keane et al. 2011) and was a much weaker detection than FRB 010724. The S/N value and corresponding width values are 18(1) and 8 +4 2 ms (beam 10) and < 5 over a wide range of widths in all other beams. The corresponding flux density values were 505 +111 115 mjy (beam 10), < 111 mjy (beam 1), < 118 mjy (inner ring), < 140 mjy (outer ring, 10), where again we do not include the flux density scaling uncertainty in this estimate. As per all flux density values based upon detections within a single beam, this is a lower-limit and the true value is larger by the inverse of the gain curve at the (unknown) offset position. As a single-beam detection the positional uncertainty can only be refined using its nondetection in the closest surrounding beams (3 and 4): this is a poor constraint which yields only that the pulse must be within a 12.2 arcmin radius of the centre of the beam in which it was discovered. The burst s DM is nominally extragalactic but Keane et al. (2012) could only find Galactic solutions compatible with this burst, and Bannister & Madsen (2014) determined that there is a high probability that the electron density along this line of sight is underestimated (2011) was given in degrees of arc, rather than degrees of right ascension (Burke-Spolaor, priv. comm.).

4 E.F. Keane Amp. (Arb. Units) Frequency (GHz) 0.80 0.60 0.40 0.20 0.00 1.50 1.45 1.40 1.35 1.30 Detected Peak S ν (Jy) 100 10 1 0.1 1 Jy ms 0.1 Jy ms 10 Jy ms 100 Jy ms Parkes FRBs Spitler et al. FRB Constant S/N Constant Fluences 1.25 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Time (seconds) 0.01 0.1 1 10 100 Width obs (ms) Figure 2. The pulse profile (top) dedispersed to 1374 MHz, and time-frequency sweep (bottom) of the Lorimer Burst, the exemplar FRB. In this, the main beam (beam number 6) in which the burst was most significantly detected, previously unshown, the signal was so strong as to saturate the 1-bit analogue digitisers. The unexplained extra trailing signal in the 1395 1398 MHz channel, which lasts for 20(2) ms, is evident 46(2) ms after the main pulse. so that a Galactic source seems most favoured. The actual progenitor is unlikely to be definitively identified unless it is seen to repeat. FRB 010125 FRB 010125 has recently been reported by Burke-Spolaor & Bannister (2014). The S/N value and corresponding width value is 25(1) and 10 +3 2 ms (beam 5) and < 5 over a wide range of widths in all other beams. The corresponding flux density values were 530 ± 85 mjy (beam 5), < 100 mjy (beam 1), < 106 mjy (inner ring, 5), < 125 mjy (outer ring), where again we do not include the flux density scaling uncertainty in this estimate. The non-detection in surrounding beams tells us only that the true position is no closer to the central beam than 11.9 arcmin, and no closer to the closest inner (outer) ring beams than 12.0 ( 12.1) arcmin. The reported S/N by Burke-Spolaor & Bannister (2014) was 17(1) but our value is 2 times higher. Our minimum flux densities are more so in disagreement, due to differing width estimates, although the appropriate width to use is vague: Burke-Spolaor & Bannister (2014) use the fullwidth at half the maximum value whereas we are essentially using the equivalent width. The true discrepency is about a factor of 2, which we postulate to be due to the root 2 problem discussed in 2. 4 SELECTION EFFECTS IN FRB SEARCHES The estimate for FRBs detectable by the current setup at Parkes is 10 4 FRBs/(4π sr)/day (Thornton et al. 2013). This number is simply the observed rate of 4 FRBs in 23 days, extrapolated, from the 0.55 deg 2 half-power field-ofview of the multi-beam receiver at Parkes (Staveley-Smith et al. 1996), to the entire sky. In addition to the obvious Figure 3. The detected flux density-observed width parameter space. Lines of constant signal-to-noise (dashed) and constant fluence (solid) are shown. Events above the thick black flux density limit line are detectable at Parkes. In this case, and if all FRBs are less than some maximum putative width (here denoted by the thick vertical line), we only have completeness in terms of fluence above the thick black line. In the shaded triangle we have incomplete coverage of fluence. Note that here the Arecibo FRB (green triangle) is included for illustration only. The sensitivity curves from Arecibo differ to those from Parkes; different incompleteness regions apply to different observing configurations. caveats of such an extrapolation, the meaning of this rate must be interpreted carefully, for a number of reasons. Fluence & Width Incompleteness?: One might suggest that the all-sky rate is that above the flux-density threshold at the half-power beam-width of Parkes. Using the system configuration relevant to the data wherein the 3 bursts discussed here were found this corresponds to flux densities of 1.01/ W/ms Jy (1.07/ W/ms Jy or 1.27/ W/ms Jy) for the central (inner or outer ring) receivers, for an 8-σ signal. Even though the physically relevant parameter is the fluence, i.e. the area under the pulse curve in the dedispersed time series, our sensitivity depends also on the pulse width. Thus, pulses with the same fluence but different widths (due to traversing different paths in the IGM/ISM) are not equally detectable. Figure 3 shows the sensitivity to bursts in the flux density-width plane. We are always incomplete to wide bursts. However if FRBs do not exist above some maximum width we are still left with an incomplete sampling of fluence. Such selection effects should be considered when determining population estimates, especially when extrapolating to lower frequencies where observed pulses are likely to be wider (Trott et al. 2013; Coenen et al. 2014). Latitude dependence?: Some additional selection effects have been identified empirically, e.g. FRBs seem to be much more difficult to detect at lower Galactic latitudes, despite intensive efforts (Petroff et al. 2014a; Burke-Spolaor & Bannister 2014). This points towards a Galactic obscuration effect but at present none of the known possibilities are sufficient to explain the paucity of low-latitude detections. The possibility remains that the rate (extrapolated from 4 events, see above) is in fact too high. What is an FRB?: There is also the question as to how

to algorithmically define what an FRB is, in contrast to a RRAT : pulsars which show detectable single pulses of radio emission only very occasionally (Keane & McLaughlin 2011). If we detect a single pulse from a RRAT how do we distinguish it from an FRB? A RRAT may eventually repeat but the main difference is that RRATs are Galactic whereas FRBs are believed to be extragalactic. We decide upon this based on the ratio of the DM of a detected event to the maximum Galactic DM contribution along the line of sight, DM max. If x = DM/DM max > 1 then the source is extragalactic, but the model we use to determine DM max is quite uncertain, especially at high Galactic latitudes (Cordes & Lazio 2002). This has resulted in ad hoc selection rules such as selecting events with x > 0.9 (Petroff et al. 2014a), but, as noted in Spitler et al. (2014) there is a rather wide grey area in this parameter space. As the uncertainty on x depends both on the DM and the line of sight in very asystematic ways this is difficult to quantify and it is quite conceivable that many FRBs have already been detected and falsely labelled as RRATs, and the converse may already be the case for FRB 010621. In effect there is a low-redshift blindness to FRBs, analogous to the low-dm blindness in Galactic SP searches (Keane et al. 2010). 5 CONCLUSIONS, DATA & SOFTWARE RELEASE With so many selection effects evident, large uncertainties in flux density and fluence estimates (see 3), essentially no knowledge of FRB spectra, and when dealing with such small number statistics, serious population analyses are precluded. These obstacles will be overcome only when a much larger number of FRBs are discovered. For example, to remove the fluence incompleteness one could simply discard all FRBs below the incompleteness value, but this is only practical (see Figure 3) when a much larger sample is obtained. To remove the low-redshift blindness we would benefit from independent distance estimates. New means of estimating the distance, such as the method of Bannister & Madsen (2014), could shed some light on this if they are applied to a large sample of RRATs and FRBs. The clear scientific potential of FRBs further motivates a search for more, and as such we are currently performing a large-scale search of the existing archival data, some of which has either never been searched previously, or has only been searched to a limited extent. We will present the results of this search in subsequent papers. Beyond that the only way to discover hundreds to thousands of FRBs is to use high sensitivity wide field-of-view telescopes with a large amount of on-sky time, e.g. through piggy-back transient observations with MeerKAT and SKA1-Mid (Fender 2015). We have made the data easily available for the three archival bursts discussed in this paper. The data can be accessed via a dropbox repository 5. Additionally analysis software is available and can be accessed via a github repository 6. These have been used in the preparation of this paper. It is our hope that others can use these to directly access 5 https://www.dropbox.com/sh/l6kr87oxsxpg367/aaaqdwv-zdqqcmpxozmfk0g3a 6 https://github.com/evanocathain/useful FRB stuff FRBs I 5 and analyse the raw data collected at the telescope. In this way uncertainties and misinterpretations can be minimised, and new predictions and analysis tools can be quickly tested. ACKNOWLEDGEMENTS EFK thanks E. Petroff & E. Barr for extensive helpful comments. EFK acknowledges the support of the Australian Research Council Centre of Excellence for All-sky Astrophysics (CAASTRO), through project number CE110001020. REFERENCES Bannister K. W., Madsen G. J., 2014, MNRAS, 440, 353 Burke-Spolaor S., Bailes M., Ekers R., Macquart J., Crawford F., 2011, ApJ, 727, 18 Burke-Spolaor S., Bannister K. W., 2014, ApJ, 792, 19 Coenen, T. et al., 2014, A&A in press (astro-ph/1408.0411) Cordes J. M., Lazio T. J. W., 2002 (astro-ph/0207156) Cordes J. M., McLaughlin M. A., 2003, ApJ, 596, 1142 Dennison B., 2014, MNRAS, 443, L11 Falcke H., Rezzolla L., 2014, A&A, 562, A137 Fender R. P., 2015, Ch. 51, to appear in Advancing Astrophysics with the Square Kilometre Array, PoS Gao H., Li Z., Zhang B., 2014, ApJ, 788, 189 Inoue S., 2004, MNRAS, 348, 999 Ioka K., 2003, ApJ, 598, L79 Keane E. F., Kramer M., Lyne A. G., Stappers B. W., McLaughlin M. A., 2011, MNRAS, 415, 3065 Keane E. F. et al., 2010, MNRAS, 401, 1057 Keane E. F., McLaughlin M. A., 2011, Bulletin of the Astronomical Society of India, 39, 333 Keane E. F., Stappers B. W., Kramer M., Lyne A. G., 2012, MNRAS, 425, L71 Kulkarni S. R., Ofek E. O., Neill J. D., Zheng Z., Juric M., 2014, ApJ submitted (astro-ph/1402.4766) Loeb A., Shvartzvald Y., Maoz D., 2014, MNRAS, 439, L46 Lorimer D. R., Bailes M., McLaughlin M. A., Narkevic D. J., Crawford F., 2007, Science, 318, 777 Lyubarsky Y., 2014, MNRAS, 442, L9 Manchester R. N. et al., 2001, MNRAS, 328, 17 McQuinn M., 2014, ApJ, 780, L33 Mottez F., Zarka P., 2014, A&A, in press (astroph/1408.1333) Noutsos A., Johnston S., Kramer M., Karastergiou A., 2008, MNRAS, 386, 1881 Petroff E. et al., 2014a, ApJ, 789, L26 Petroff E. et al., 2014b, MNRAS, submitted Ravi V., Lasky P. D., 2014, MNRAS, 441, 2433 Spitler L. G. et al., 2014, ApJ, 790, 101 Staveley-Smith L. et al., 1996, Proc. Astr. Soc. Aust., 13, 243 Thornton D. et al., 2013, Science, 341, 53 Totani T., 2013, PASJ, 65, L12 Trott C. M., Tingay S. J., Wayth R. B., 2013, ApJ, 776, L16 Zhang B., 2014, ApJ, 780, L21 Zhou B., Li X., Wang T., Fan Y.-Z., Wei D.-M., 2014, Phys- RevD, 89, 107303