Observation of the Crab Nebula with the HAWC Gamma-Ray Observatory

, , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , and

Published 2017 June 29 © 2017. The American Astronomical Society. All rights reserved.
, , Citation A. U. Abeysekara et al 2017 ApJ 843 39 DOI 10.3847/1538-4357/aa7555

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/843/1/39

Abstract

The Crab Nebula is the brightest TeV gamma-ray source in the sky and has been used for the past 25 years as a reference source in TeV astronomy, for calibration and verification of new TeV instruments. The High Altitude Water Cherenkov Observatory (HAWC), completed in early 2015, has been used to observe the Crab Nebula at high significance across nearly the full spectrum of energies to which HAWC is sensitive. HAWC is unique for its wide field of view, nearly 2 sr at any instant, and its high-energy reach, up to 100 TeV. HAWC's sensitivity improves with the gamma-ray energy. Above ∼1 TeV the sensitivity is driven by the best background rejection and angular resolution ever achieved for a wide-field ground array. We present a time-integrated analysis of the Crab using 507 live days of HAWC data from 2014 November to 2016 June. The spectrum of the Crab is fit to a function of the form $\phi {(E)={\phi }_{0}(E/{E}_{0})}^{-\alpha -\beta \cdot \mathrm{ln}(E/{E}_{0})}$. The data is well fitted with values of α = 2.63 ± 0.03, β = 0.15 ± 0.03, and ${\mathrm{log}}_{10}({\phi }_{0}\,{\mathrm{cm}}^{2}\,{\rm{s}}\,\mathrm{TeV})=-12.60\pm 0.02$ when E0 is fixed at 7 TeV and the fit applies between 1 and 37 TeV. Study of the systematic errors in this HAWC measurement is discussed and estimated to be ±50% in the photon flux between 1 and 37 TeV. Confirmation of the Crab flux serves to establish the HAWC instrument's sensitivity for surveys of the sky. The HAWC all-sky survey will be the deepest survey of the northern sky ever conducted in the multi-TeV band.

Export citation and abstract BibTeX RIS

1. Introduction

The Crab Pulsar Wind Nebula (the Crab Nebula or the Crab) occupies a place of special distinction in the history of high-energy astrophysics. It was the first high-confidence TeV detection in 1989 using the Whipple telescope (Weekes et al. 1989) and is the brightest steady source in the northern sky above 1 TeV. It has been observed with imaging atmospheric Cherenkov telescopes (IACTs) over the past two decades (Tanimori et al. 1998; Aharonian et al. 2004, 2006; Celik 2008; Aleksić et al. 2015).

In 2003, an alternative approach was demonstrated with the Milagro detection of the Crab Nebula (Atkins et al. 2003). This second approach consisted of a large-area particle detector that directly measured relativistic particles on the ground. The techniques are complementary: the IACT technique achieves generally higher instantaneous sensitivity, but with a narrow few-degree field of view and a ∼10% duty cycle. The direct-shower-sampling technique achieves a duty cycle in excess of 90% with a ∼2 sr instantaneous field of view. The Crab Nebula has been detected using the direct sampling technique with a number of technologies for the detector components: water-Cherenkov detectors (WCDs; Atkins et al. 2003), scintillation detectors (Amenomori et al. 2009), and resistive plate chambers (Bartoli et al. 2015).

The TeV emission in the Crab Nebula arises from inverse-Compton (IC) upscattering of low-energy ambient photons by energetic electrons accelerated in shocks surrounding the central pulsar (Atoyan & Aharonian 1996). Photons from synchrotron emission of the electrons themselves are likely the dominant IC target with subdominant contributions from the cosmic microwave background and the extragalactic background light (Martín et al. 2012). Despite rare flaring emission below 1 TeV (Abdo et al. 2011; Tavani et al. 2011) and a potential TeV flare (Aielli et al. 2010), the Crab is generally believed to be steady at higher energies (Abramowski et al. 2014; Aliu et al. 2014; Bartoli et al. 2015) when averaging over the the pulsations from the pulsar (Aliu et al. 2011; Ansoldi et al. 2016). Consequently, the Crab Nebula has been adopted as the reference source in TeV astronomy and is a reliable beam of high-energy photons to use for calibrating and understanding new TeV gamma-ray instruments.

The High Altitude Water Cherenkov (HAWC) observatory is a new instrument sensitive to multi-TeV hadron and gamma-ray air showers, operating at coordinates of 18° 59' 41''N, 97° 18' 27''W at an altitude of 4100 m in the Sierra Negra, Mexico. HAWC consists of a large 22,000 m2 area densely covered with 300 WCDs, of which 294 have been instrumented. Each WCD consists of a 7.3 m diameter, 5 m tall steel tank lined with a plastic bladder and filled with purified water. Figure 1 shows a schematic of the WCD and an overhead view of the full instrument. At the bottom of each WCD, three 8-inch Hamamatsu R5912 photomultiplier tubes (PMTs) are anchored in an equilateral triangle of side length 3.2 m, with one 10-inch high-quantum-efficiency Hamamatsu R7081 PMT anchored at the center.

Figure 1.

Figure 1. Left: schematic of a single HAWC WCD including the steel tank, the covering roof, the three 8-inch Hamamatsu R5912 PMTs, and one 10-inch Hamamatsu R7081-MOD PMT. The tanks are 7.5 m in diameter and 5 m high. Water is filled to a depth of 4.5 m with 4.0 m of water above each PMT. Right: layout of the completed HAWC instrument, covering 22,000 m2. The location of each WCD is indicated by a large circle, and PMTs are indicated with smaller circles. The gap in the center hosts a building with the data acquisition system.

Standard image High-resolution image

A high-energy photon impinging on the atmosphere above HAWC initiates an extensive electromagnetic air shower. The resulting mix of relativistic electrons, positrons, and gamma rays propagates to the ground in a thin tortilla of particles at nearly the speed of light. Energetic particles that reach the instrument can interact in the water and produce optical light via Cherenkov radiation. The high altitude of HAWC sets the scale for the photon energy that can be detected. At HAWC's altitude, the shower from a 1 TeV photon from directly overhead will have about 7% of the original photon energy left when the shower reaches the ground. The fraction of energy reaching the ground rises to ∼28% at 100 TeV. The detector is fully efficient to gamma rays with a primary energy above ∼1 TeV. Lower-energy photons can be detected when they fluctuate to interact deeper in the atmosphere than typical.

The voltages on the HAWC PMTs are chosen to match the PMT gains across the array. PMT pulses are amplified, shaped, and passed through two discriminators at approximately 1/4 and 4 PEs (Abeysekara et al. 2016) and digitized. The length of time that PMT pulses spend above these thresholds (time-over-threshold, or ToT) is used to estimate the total amount of charge collected in the PMT. Noise arises from a number of sources, including PMT afterpulsing, fragments of subthreshold air showers, PMT dark noise, and other sources. The 8-inch PMTs have a hit rate (a hit being each time the PMT signal crosses the 1/4 PE threshold) due to the combined effect of these sources of 20–30 kHz, and the 10-inch PMTs have a hit rate of 40–50 kHz.

The data from the front-end electronics are digitized with commercial time-to-digital converters (TDCs) and passed to a farm of computers for real-time triggering and processing. Events are preserved by the computer farm if they pass the trigger condition: a simple multiplicity trigger, requiring some number, Nthresh, of PMTs hit within 150 ns. Hits 500 ns prior to a trigger and up to 1000 ns after a trigger are also saved for reconstruction. During the operation of HAWC, Nthresh has varied between 20 and 50. The trigger rate at the time of writing, due primarily to hadronic cosmic-ray air showers, is ∼24 kHz with Nthresh = 28.

The reconstruction of air showers involves determining the direction, the likelihood for the event to be a photon, and the event's size. A first-look reconstruction is applied at the HAWC site. In this analysis, all the data have been reconstructed again (the fourth revision, or Pass 4, of the reconstruction process) off-site in order to have a uniform data set and the best calibrations available. The chief background to gamma-ray observation is the abundant hadronic cosmic-ray population. Individual gamma-ray-induced air showers can be distinguished from cosmic-ray showers by their topology and the presence of deeply penetrating particles at the ground.

The strength of HAWC over the IACT technique is that photon showers may be detected across the entire ∼2 sr field of view of the instrument, day or night, regardless of weather conditions. As such, HAWC is uniquely suited to study the long-duration light curve of objects and to search for flaring sources in real time. Additionally, since sources are observed on every transit, HAWC obtains thousands of hours of exposure on each source, greatly improving the sensitivity to the highest-energy photons.

Section 2 outlines the algorithms by which the direction, size, and type (photon or hadron) of each shower are determined. Section 3 describes the identification of the gamma-ray signal from the Crab Nebula. The fit to the Crab energy spectrum, including a treatment of systematic errors, is described in Section 4. Finally, a discussion of the results is presented in Section 5, including a comparison to prior spectra measured by peer experiments and a computation of the sensitivity of the HAWC instrument, anchored in the agreement of the HAWC measurement with other experiments.

2. Air Shower Reconstruction

Events from the detector are reconstructed to determine the arrival direction of the primary particle and the size of the resulting air shower on the ground, a proxy for the primary particle's energy. In Section 2.1, the simulation is briefly described as it is key to evaluating the reconstruction process. Section 2.2 describes the calibration, by which the time and light level in individual PMTs are determined. Section 2.3 discusses the selection of PMT signals for reconstruction and the event size measurement. The direction reconstruction occurs in two steps: first the core reconstruction, described in Section 2.4, and then the direction determination, described in Section 2.5. The air shower core, the dense concentration of particles along the direction of the original primary, is needed to make the best reconstruction of the air shower's direction since the air shower arrival front is delayed from a pure plane, depending on the distance from the core. The identification of photon candidates is presented in Section 2.6. The directional fit is iterated to suppress noise, and this iteration is explained in Section 2.7.

2.1. Simulation

The HAWC instrument is modeled using a combination of community-standard simulation packages and custom software. The CORSIKA package (v7.4000) is used for simulation of air showers, propagating the primary particles through the atmosphere to the ground (Heck et al. 1998). At ground level, a GEANT 4 (Agostinelli et al. 2003) simulation (v4.10.00) of the shower particles is used to propagate the ground-level particles through the HAWC tanks and to track the Cherenkov photons to the faces of the PMTs.

The response of the PMTs and the calibration are approximated with a custom simulation that assumes that recorded light is faithfully detected with some efficiency and an uncertainty in the logarithm of the total charge recorded. Decorrelated single PE noise is added. The absolute PMT efficiency for detecting Cherenkov photons is established by scaling the simulated PMT response to vertical muons to match data. Most muons passing through HAWC are minimum ionizing with nearly constant energy loss. Vertical muons, therefore, are a nearly constant light source and are convenient for establishing to total PMT efficiency. Simulated events are subsequently reconstructed by the same procedure as experimental data to study the performance of the algorithms.

2.2. Calibration

Table 1 summarizes the steps in reconstruction of HAWC events. The first step in the reconstruction process is calibration, the processes by which true time and light level in each PMT are estimated from the TDC-measured threshold-crossing times of each PMT (Lauer 2013; Solares et al. 2016).

Table 1.  Steps in the HAWC Event Reconstruction

Step Description Hit Selection
1 Calibration
2 Hit selection
3 Center-of-mass core reconstruction Selected hits
4 SFCF core—first pass Selected hits
5 Direction—first pass Selected hits
6 SFCF core—second pass Selected hits within 50 ns of first-pass plane
7 Direction—second pass Selected hits within 50 ns of first-pass plane
8 Compactness, ${ \mathcal C }$ Selected hits within 20 ns of second-pass plane
9 PINCness, ${ \mathcal P }$ Selected hits within 20 ns of second-pass plane

Note. After calibration (Section 2.2) and the selection of hits (Section 2.3), the reconstruction is applied. The best shower core (Section 2.4) and direction (Section 2.5) reconstructions are applied with gradual narrowing of the hits used (Section 2.7). After the final core and direction determination, the photon/hadron discrimination algorithms, compactness, and PINCness are applied (Section 2.6).

Download table as:  ASCIITypeset image

The calibration associates the measured ToT in each PMT with the true number of PEs. To give a sense of scale, the ToT for a single PE crossing the low-threshold discriminator (about 1/4 PE) is ∼100 ns. Above a few PEs, the higher-threshold (about 4 PEs) ToT is used for charge assignment, and a high-threshold ToT of 400 ns roughly corresponds to a charge of 104 PEs. The timescale for these ToTs is determined by the shaping of the front-end electronics and is chosen to be longer than the characteristic arrival time distribution of PEs during an air shower so as to integrate the whole air shower arrival into one PMT hit.

In addition to the PE measurement, the calibration system accounts for electronic slewing of the PMT waveform: lower-PE waveforms cross the threshold later than contemporaneous high-PE waveforms.

Subsequent reconstruction algorithms treat all PMTs, the 8 inch and 10 inch, as identical, despite the larger size and greater efficiency of the 10-inch PMTs. To accommodate this, an effective charge Qeff is defined. For Qeff, PE values from the central 10-inch PMTs are scaled by a factor of 0.46 to place them on par with the 8-inch PMTs.

Finally, each PMT has a single calibrated timing offset that accounts for the different cable lengths and any other timing delays that may differ from PMT to PMT. These delays are measured to within a few nanoseconds by the calibration system and are refined to subnanosecond precision by iterated fits to hadronic air showers. Since the hadronic background is isotropic, the point of maximum cosmic-ray density is overhead and the PMT timing pedestals are chosen to ensure that this is true. A final small (∼0fdg2) rotation of all events is performed to ensure that the Crab Nebula appears in its known location.

Throughout the analysis, the Crab is assumed to be at a location of 83fdg63 right ascension and 22fdg01 declination, in the J2000.0 epoch, taken from Aharonian et al. (2004). While the pulsar position is known more precisely (e.g., Comella et al. 1969), this precision is sufficient for use in HAWC.

2.3. Hit Selection and Event Size Bins

As described in Section 1, the HAWC Data Acquisition System (DAQ) records 1.5 μs of data from all PMTs that have a hit during an air shower event. A subset of these hits are selected for the air shower fit. To be used for the air shower fit, hits must be found between −150 and +400 ns around the trigger time. Hits are removed if they occur shortly after a high-charge hit under the assumption that these hits are likely contaminated with afterpulses. Additionally, hits are removed if they have a pattern of TDC crossings that is not characteristic of real light; they cannot be calibrated accurately. Finally, each channel has an individual maximum calibrated charge, typically a few thousand PEs, but no more than 104 PEs, above which the PMTs are not used. Above ∼104 PEs, corresponding to a ToT of ∼400 ns, prompt afterpulsing in the PMTs can artificially lengthen the ToT measurement, giving a false measurement. Channels are considered available for reconstruction if they have a live PMT taking data that have not been removed by one of these cuts.

The angular error and the ability to distinguish photon events from hadron events are strongly dependent on the energy and size of events on the ground. We adopt analysis cuts and an angular resolution description that depends on this measured size. The data are divided into nine size bins, ${ \mathcal B }$, as outlined in Table 2. The size of the event is defined as the ratio of the number of PMT hits used by the event reconstruction to the total number of PMTs available for reconstruction, fhit. This definition allows for relative stability of the binning when PMTs are occasionally taken out of service.

Table 2.  Cuts Used for the Analysis

${ \mathcal B }$ fhit ${\psi }_{68}(\deg )$ ${ \mathcal P }$ Maximum ${ \mathcal C }$ Minimum Crab Excess per Transit
1 6.7%–10.5% 1.03 <2.2 >7.0 68.4 ± 5.0
2 10.5%–16.2% 0.69 3.0 9.0 51.7 ± 1.9
3 16.2%–24.7% 0.50 2.3 11.0 27.9 ± 0.8
4 24.7%–35.6% 0.39 1.9 15.0 10.58 ± 0.26
5 35.6%–48.5% 0.30 1.9 18.0 4.62 ± 0.13
6 48.5%–61.8% 0.28 1.7 17.0 1.783 ± 0.072
7 61.8%–74.0% 0.22 1.8 15.0 1.024 ± 0.053
8 74.0%–84.0% 0.20 1.8 15.0 0.433 ± 0.033
9 84.0%–100.0% 0.17 1.6 3.0 0.407 ± 0.032

Note. The definition of the size bin ${ \mathcal B }$ is given by the fraction of available PMTs, fhit, that record light during the event. Larger events are reconstructed better, and ψ68, the angular bin that contains 68% of the events, reduces dramatically for larger events. The parameters ${ \mathcal P }$ and ${ \mathcal C }$ (Section 2.6) characterize the charge topology and are used to remove hadronic air shower events. Events with a ${ \mathcal P }$ less than indicated and a ${ \mathcal C }$ greater than indicated are considered photon candidates. The cuts are established by optimizing the statistical significance of the crab and trend toward harder cuts at larger size events. The number of excess events from the crab in each ${ \mathcal B }$ bin per transit is shown as well.

Download table as:  ASCIITypeset image

For this analysis, events are only used if they have more than 6.7% of the available PMTs' seeing light. Since typically 1000 PMTs are available, typically a minimum of 70 PMTs is needed for an event. This is substantially higher than the trigger threshold. The data between the trigger threshold and the threshold for ${ \mathcal B }=1$ in this analysis consist of real air showers, and techniques to recover these events and lower the energy threshold, beyond what is presented here, are under study.

Figure 2 shows the distribution of true energies as a function of the ${ \mathcal B }$ of the events. The distribution of energies naturally depends heavily on the source itself, both its spectrum and the angle at which it culminates during its transit. A pure power-law spectrum with a shape of ${E}^{-2.63}$ and a declination of 20° was assumed for this figure. As ${ \mathcal B }$ is a simple variable—containing no correction for zenith angle, impact position, or light level in the event—the energy distribution of ${ \mathcal B }$ bins is wide. Section 5.3 discusses planned improvements to this event parameter that will measure the energy of astrophysical gamma rays better.

Figure 2.

Figure 2. Fits to the true energy distribution of photons from a source with a spectrum of the form E−2.63 at a declination of +20°N for ${ \mathcal B }$ between 1 and 9, summed across a transit of the source. Better energy resolution and dynamic range can be achieved with a more sophisticated variable that takes into account the zenith angle of events and the total light level on the ground. The curves have been scaled to the same vertical height for display.

Standard image High-resolution image

Bin ${ \mathcal B }=9$ bears particular attention. It is an "overflow" bin containing events that have between 84% and 100% of the PMTs in the detector seeing light. Typically, a 10 TeV photon will hit nearly every sensor, and the ${ \mathcal B }$ variable has no dynamic range above this energy. This limit is not intrinsic to HAWC, and variables that utilize the light level seen in PMTs on the ground, similar to what was used in the original sensitivity study (Abeysekara et al. 2013), have a dynamic range above 100 TeV. These variables, not used in this analysis, will improve the identification of high-energy events. This is discussed further in Section 5.3.

2.4. Core Reconstruction

In an air shower, the concentration of secondary particles is highest along the trajectory of the original primary particle, termed the air shower core. Determining the position of the core on the ground is key to reconstructing the direction of the primary particle. In the sample event, Figure 3, the air shower core is evident in Figure 3(a). The image is an overhead view of the HAWC detector, with circles indicating the WCD location and the PMTs within the WCDs. The colors indicate the amount of light (measured in units of PEs) seen in each PMT. The air shower core is evident as the point of maximum PE density.

Figure 3.

Figure 3. These figures illustrate a high-confidence gamma ray from the Crab Nebula, taken from a data set with a better than 10:1 signal-to-background ratio. Panels (a) and (c) show an overhead view of the HAWC instrument, with large circles to indicate individual WCDs and small circles to indicate individual PMTs. Sporadic PMTs have been removed as described in Section 2.3. Panel (a) shows the PMT effective charge, defined in Section 2.2, for each PMT that recorded a hit during the event. The shower core is evident. Panel (c) shows the time each PMT recorded a hit. The color scale encodes the reconstructed arrival time; because the shower is inclined, a gradient in arrival time is observed. Panel (b) shows the lateral distribution function, the effective charge recorded as a function of the distance along the ground from the hit PMT to the reconstructed core. The fitted function from the core fit (Section 2.4) is overlaid. From this distribution, the photon/hadron separation parameters ${ \mathcal C }$ and ${ \mathcal P }$ are computed (Section 2.6), and the moving average used in the computation of ${ \mathcal P }$ is shown. Finally, panel (d) shows the time each PMT recorded a hit relative to a perfect shower plane (under the assumption that the photon came from the Crab, as explained in Section 2.5) as a function of the distance of the hit from the shower core. The need for a timing correction before the plane fit (due to the curvature and sampling effects) is evident in panel (d).

Standard image High-resolution image

The PE distribution on the ground is fit with a function that decreases monotonically with the distance from the shower core. The signal in the ith PMT, Si, is presumed to be

Equation (1)

where ${\boldsymbol{x}}$ is the core location, ${{\boldsymbol{x}}}_{i}$ is the location of the measurement, Rm is the Molière radius of the atmosphere, approximately 120 m at HAWC altitude, σ is the width of the Gaussian, and N is the normalization of the tail. Fixed values of σ = 10 m and N = 5 × 10−5 are used. This leaves three free parameters, the core location and overall amplitude, A.

The functional form used in this algorithm, termed the Super Fast Core Fit (SFCF), is a simplification of a modified Nishimura–Kamata–Greisen (NKG) function (Greisen 1960) and is chosen for rapid fitting of air shower cores. The NKG function has an additional free parameter, the shower age, and involves computationally intensive power-law and gamma function evaluation. The SFCF hypothesis in Equation (1) is similar, but numerical minimization can converge faster because the function is simpler, the derivatives are computed analytically, and there is a lack of a pole at the core location. This fit is seeded with a simple center-of-mass algorithm that computes the center of mass from the recorded charge in each PMT.

Figure 3(b) shows the recorded charge in each PMT as a function of the PMT's distance along the ground to the reconstructed shower core. While the full NKG function would describe the lateral distribution better, the SFCF form allows rapid identification of the center of showers, and this is sufficient for the present analysis. Cores can be localized to a median error of ∼2 m for large events (${ \mathcal B }=8$) and ∼4 m for small events (${ \mathcal B }=3$) for gamma-ray events with a core that lands on the HAWC detector. The error in reconstructing the shower core increases as the core moves farther from the array. For example, a shower with a core that is 50 m from the edge of the array will have an error in the location of the core of ∼35 m.

2.5. Direction Reconstruction

To first order, the air shower particles arrive on a plane defined by the speed of light and direction of the primary particle. In fact, the shower front is curved, since the shower particles originate from a common interaction point. Particles with paths along the shower axis have shorter travel distances compared with those that stray from it owing simply to their shorter path. In practice, this means that particles detected near the shower core are observed to arrive earlier than particles far from the core.

Multiple scattering of shower particles also leads to variation of arrival times at the observation level. Generally, the shower front is several nanoseconds thick near the core and wider far from the core. Since the HAWC electronics only measures the arrival time of the earliest arriving particle, there is an unavoidable pulse-amplitude-dependent timing correction. If only a single particle is detected, its most probable time is at the center of the shower front, but when many particles are detected, the measured time is the time of the first particle, which is likely near the leading edge of the shower front. Since particle density is higher near the core than in the shower periphery, there is a resulting position-dependent shower timing offset. We refer to this phenomenon as sampling. Generally, it is hard to disentangle curvature and sampling, so in the reconstruction they are regarded as a single effect.

The curved shape of the air shower front can be easily seen in data with the sample event in Figure 3. Figure 3(c) shows, for each PMT in the sample event, the calibrated time each PMT saw light. The color trend is due to the inclined direction of the air shower. Taking out this inclination, we can see the curved shower front. The event in Figure 3 was chosen for display because—taken from a high signal-to-background sample of data—it is very likely a photon from the Crab. If we assert that the origin of this particle is the Crab Nebula, we know the air shower plane precisely and can make Figure 3(d): we adjust the times of each PMT hit assuming a pure planar air shower from the direction of the Crab Nebula. Figure 3(d) shows the plane-corrected time as a function of the PMT distance from the core of the air shower along the ground. A pure planar shower would correspond to a horizontal line on this figure. The delay of particles far from the core is evident. Note that the deviation as a function of distance is about 5–10 ns/100 m, more than a degree of deviation from a plane. Understanding and correcting for shower sampling and curvature is the main challenge in achieving the best possible angular resolution of 0.15 deg.

A combined curvature/sampling correction—a function of the distance of hits from the shower core and the total charge recorded in the PMT—is utilized to account for the deviation of the front shape compared to a plane. The curvature/sampling correction is based on a combination of simulation and Crab observations. The functional form of the correction was derived from studies of simulated showers, but the parameters of the function were optimized using a subset of our Crab data set and then tested using a separate subset. We use data instead of simulations to optimize the shower front shape, since the subtitles of the timing calibration methodology and PMT and electronics response are not well reproduced by our simulation.

After correcting for the sampling and curvature, the angular fit is a simple χ2 planar fit and has been described before (Atkins et al. 2003).

2.6. Photon/Hadron Separation

Hadronic cosmic rays are the most abundant particles producing air showers in HAWC and constitute the chief background to high-energy photon observation. The air showers produced by high-energy cosmic rays and gamma rays differ: gamma-ray showers are pure electromagnetic showers with few muons or pions. Conversely, hadronic cosmic rays produce hadronic showers rich with pions, muons, other hadronic secondaries, and structure due to the showering of daughter particles created with high transverse momentum. In HAWC, these two types of showers appear quite different, particularly for showers above several TeV.

Figure 4 shows the lateral distributions for two showers, an obvious cosmic ray (left) and a strong photon candidate (right) from the Crab Nebula. The effective light level Qeff falls off for hits farther from the shower core in both showers, but in the hadronic shower there are sporadic high-charge hits far from the air shower's center. This clumpiness is characteristic of hadronic showers and arises from a combination of penetrating particles (primarily muons) and hadronic subshowers that are largely absent in photon-induced showers.

Figure 4.

Figure 4. Lateral distribution functions of an obvious cosmic ray (left) and a photon candidate from the Crab Nebula (right). The cosmic ray has isolated high-charge hits far from the shower core, due to penetrating particles in the hadronic air shower. These features are absent in the gamma-ray shower.

Standard image High-resolution image

Two parameters are used to identify cosmic-ray events. The first parameter, compactness, was used in the sensitivity study (Abeysekara et al. 2013). The variable CxPE40 is the effective charge measured in the PMT with the largest effective charge outside a radius of 40 m from the shower core. We then define the compactness, ${ \mathcal C }$, as

Equation (2)

where Nhit is the number of PMT hits during the air shower. CxPE40 is typically large for a hadronic event, so ${ \mathcal C }$ is small.

In addition to the largest hit outside the core, the "clumpiness" of the air shower is quantified with a parameter ${ \mathcal P }$, termed the PINCness of an event (short for Parameter for Identifying Nuclear Cosmic rays). ${ \mathcal P }$ is defined using the lateral distribution function of the air shower, seen in Figure 4. Each of the PMT hits, i, has a measured effective charge ${Q}_{\mathrm{eff},i}$. ${ \mathcal P }$ is computed using the logarithm of this charge ${\zeta }_{i}={\mathrm{log}}_{10}({Q}_{\mathrm{eff},i})$. For each hit, an expectation is assigned $\langle {\zeta }_{i}\rangle $ by averaging the ζi in all PMTs contained in an annulus containing the hit, with a width of 5 m, centered at the core of the air shower.

${ \mathcal P }$ is then calculated using the χ2 formula:

Equation (3)

The errors ${\sigma }_{{\zeta }_{i}}$ are assigned from a study of a sample of strong gamma-ray candidates in the vicinity of the Crab.

The ${ \mathcal P }$ variable essentially requires axial smoothness. Figure 4 shows the moving average $\langle {\zeta }_{i}\rangle $ for two sample events. The hadronic event in Figure 4 is "clumpy" and has several hits that differ sharply from the moving average, yielding a large ${ \mathcal P }.$

The ${ \mathcal C }$ and ${ \mathcal P }$ variables are well modeled in simulation. Figure 5 shows the measured distribution of ${ \mathcal C }$ in the vicinity of the Crab Nebula (the Crab region) and in an annular reference region around the Crab (the background region). The background region is scaled to have the same solid angle as the Crab region. The distributions in the vicinity of the Crab are made of a combination of hadronic cosmic rays and true photons from the Crab. Figures 5(c) and (d) show these distributions with the background distribution subtracted. The subtraction yields the data-measured distribution of ${ \mathcal C }$ for gamma rays from the Crab. Figure 6 is a comparable figure for ${ \mathcal P }$. Figures 5 and 6 are compared to a simulation prediction from the final fitted flux from Section 4; the simulation agreement is evident.

Figure 5.

Figure 5. Compactness (${ \mathcal C }$) distribution for events in the vicinity of the Crab for ${ \mathcal B }=3$ (left panels) and ${ \mathcal B }=8$ (right panels). The top two figures show the raw recorded events in the vicinity of the Crab itself. Events within 1.5ψ68 from Table 2 are used to define this "Crab region." A background expectation is generated by scaling the compactness distribution from a large annulus around the Crab. The bottom figures show the measured distribution from the Crab with this background level subtracted. Predictions from photon simulation are overlaid, showing that these variables are well modeled. Larger photon showers are typically easier to identify from hadrons, as evidenced in the figures. In panel (d), the error bars in the background-dominated region are quite large and zoomed to show the photon distribution; an inset shows the entire distribution.

Standard image High-resolution image
Figure 6.

Figure 6. Similar to Figure 5, but for the PINCness (${ \mathcal P }$) distribution for events in the vicinity of the Crab for ${ \mathcal B }=3$ (left panels) and ${ \mathcal B }=8$ (right panels). Since ${ \mathcal P }$ is essentially a χ2 computation, the ${ \mathcal P }$ variable is ∼1–2 for true photons.

Standard image High-resolution image

2.7. Noise and Fit Refinement

HAWC's outer 8-inch PMTs individually trigger at some 20–30 kHz, and the 10-inch central PMTs trigger at 40–50 kHz. Of this random noise, roughly half (with large uncertainties) is believed to be due to real shower fragments and roughly half due to nonshower sources like radioactive contaminants in the PMT glass or PMT afterpulsing. Approximately 1 kHz of the noise is "high-PE" noise from individual accidental air shower muons (10–200 PEs) and can be correlated between the PMTs within a WCD that the muon hits.

This noise can bias the air shower reconstruction, and the muon noise has the potential, if not removed, to confuse the identification of gamma-ray showers because a single muon is enough to indicate that a shower is of hadronic origin.

In order to achieve the best direction reconstruction and to avoid falsely rejecting true photons, the SFCF core reconstruction and the plane fit are each performed twice, as outlined in Table 1. During the first pass, all selected hits are used to locate the core and initial direction. After this first "rough" fit, hits that are more than ±50 ns from the curvature/sampling-corrected air shower plane are removed, and the shower is fit a second time.

The computation of photon/hadron separation variables is done with hits that are within ±20 ns of the curvature/sampling-corrected air shower plane. With these cuts, only ∼4% of gamma-ray events will have an accidental muon contributing to the photon/hadron variable computation and risk being falsely rejected.

3. Crab Nebula Signal

Once individual events are reconstructed, the identification and characterization of sources proceed. Section 3.1 discusses the first 553 days of data taking. Section 3.2 describes how the residual cosmic-ray background, after gamma/hadron separation cuts, is estimated in the vicinity of the Crab. Section 3.3 describes the validation of the angular resolution. Section 3.4 describes the optimization of photon/hadron discrimination cuts.

3.1. Data Set

We consider here data taken by HAWC between 2014 November 26 and 2016 June 2, a total elapsed time of 553 days. The detector was not taking data for a cumulative time of 40 days for various operational reasons. Additionally, a further 7 days of data were rejected owing to trigger rate instability. This yields a total livetime of 506.7 days, an average fractional livetime of 92%. Figure 7 shows the fractional livetime achieved in blocks of 10 days. With the exception of one period of extended downtime (due to a failure of the power transformer at the site), during no single 10-day block was the detector live less than 75% of the time.

Figure 7.

Figure 7. Left: livetime fraction of HAWC during the first 553 days of data taking. Data are shown averaged over 10-day increments along with the average from the entire data set. Right: exposure deviation of the instrument, measured with the reconstructed event direction, as a function of right ascension. The overhead sky is nearly uniformly exposed with a maximum deviation from uniformity of less than ±2%.

Standard image High-resolution image

The occasional downtime does not heavily bias the exposure. Figure 7 shows the relative exposure of the instrument, measured as the fractional deviation of the number of shower events observed as a function of reconstructed right ascension. The anisotropy in cosmic-ray arrival direction is subdominant at ∼10−3 (Abdo et al. 2009), and we can conclude that the exposure is flat to within ±2%. During the 507-day livetime, the Crab is visible above a zenith angle of 20° for ∼1400 hr and above a zenith angle of 45° for ∼3200 hr.

HAWC is operated 90% of the time. The sensitivity depends very weakly on the season or on local weather. Our exposure in R.A. is uniform to the few-percent level, and the background rate for the analysis bins, a measure of the sensitivity, has a day-to-day variability of only 2%, which is much lower than other systematic errors discussed later. For this reason, we treat the sky exposure for this data set as uniform in R.A. and do not attempt to correct for weather or seasonal variability. With these assumptions, the exposure is completely described by a source's declination and the total number of operational days in the data set.

3.2. Background Estimation

Even with strict photon/hadron discrimination cuts, the data are still dominated by hadronic cosmic-ray events. Fortunately, the directions of hadronic cosmic rays are randomized by magnetic deflection in transit from their sources, and the population of cosmic rays is isotropic to a few parts in 103 (Abdo et al. 2009). Gamma-ray sources, by comparison, appear as localized "bumps" on this smooth background. In order to identify gamma-ray sources, this background contamination must be estimated.

Figure 8 shows the background computed in the declination of the Crab across the entire sky in the example bins ${ \mathcal B }=3$ and ${ \mathcal B }=8$. The background is estimated using an algorithm developed for analysis of Milagro data (Abdo et al. 2012) called direct integration. The algorithm assumes that, for 2 hr blocks of time, the all-sky rate is independent of the spatial arrival distribution of events in the detector. This assumption is accurate to a few parts in 103, the level of the cosmic-ray anisotropy. The background at any R.A., decl. point in the sky, during the 2 hr interval, is determined by convolving the all-sky rate with the spatial distribution of events in detector coordinates. The convolution integral output is averaged over 0fdg5 patches of sky to account for sparse events in higher-${ \mathcal B }$ bins. Events from known strong sources, the Galactic plane, Mrk 421 and Mrk 501, and the Crab Nebula, are excluded from the background computation.

Figure 8.

Figure 8. For ${ \mathcal B }$ bins 3 and 8, the density of events recorded as a function of right ascension for the band of declination within 0fdg5 of the true Crab Nebula location. The density of recorded events is shown along with the direct-integration background. The Crab Nebula is evident as a single excess at an R.A. of 83fdg6. The relative exposure is clear as a slight drop in the number of events at ∼R.A. = 100°. An inset in panel (a) shows a zero-suppressed zoom-in of the region around the Crab Nebula.

Standard image High-resolution image

This convolution naturally handles sporadic downtime in the detector, natural rate drifts as the atmosphere changes, and has been shown to accurately determine the background to a few parts in 103. Figure 8(a) shows the slight underexposure of events from R.A. ≈ 100. The background estimate accurately captures this feature.

Finally, being completely determined by the data themselves, the background estimate is wholly independent of any simulation of the background. Any remaining systematic error is subdominant and is neglected in the eventual systematic error tabulation.

3.3. Point-spread Function

The point-spread function (PSF), ρ(ψ), describes how accurately the directions of gamma-ray events are reconstructed. Here ψ is the space-angle difference between the true photon arrival direction and the reconstructed direction. To a good approximation, the PSF of HAWC is the sum of two 2D Gaussians with different widths

Equation (4)

where Gi is a Gaussian distribution with width σi,

Equation (5)

which is normalized to unity across the unit sphere.

Figure 9 exhibits the measured angular resolution in HAWC data in two size bins ${ \mathcal B }=3$ and ${ \mathcal B }=8$ obtained by assuming that the Crab Nebula is a point source and that all the angular spread observed in HAWC is due to instrumental precision. The solid-angle density of recorded events dN/dΩ in the vicinity of the Crab is shown as a function of ψ2. Bins of ψ2 have constant solid angle (in the small-angle approximation), so any remaining cosmic-ray background shows up as a flat component and the gamma rays are evident as a peak near ψ2 = 0. The improvement in angular resolution for larger events is clear.

Figure 9.

Figure 9. Maps of the sky around the Crab Nebula for events ${ \mathcal B }=3$ (left) and ${ \mathcal B }=8$ (right) after photon/hadron discrimination in equatorial coordinates. The pixelation is based on the HEALpix library (Gorski et al. 2005), and each pixel corresponds to 9.99 × 10−7 sr. The top panels show the recorded number of events with pixels on the sky much smaller than the HAWC angular resolution. The Crab is readily evident. The bottom panels show the number of recorded events per steradian (dN/dΩ) as a function of the distance from the Crab. At higher ${ \mathcal B }$, the angular resolution and background rejection improve dramatically.

Standard image High-resolution image

Fits to this functional form of Equation (4) can have highly coupled parameters. It is more useful and traditional to quantify the resulting fits with the 68% containment radius, ψ68, the angular radius around the true photon direction in which 68% of events are reconstructed. Figure 10 shows ψ68, for each ${ \mathcal B }$ of the analysis, measured on the Crab and predicted from simulation. At best, events are localized to within 0fdg17, the best angular resolution achieved for a wide-field ground array detector directly sampling the particle cascade on the ground.

Figure 10.

Figure 10. Measured angular resolution, the angular bin required to contain 68% of the photons from the Crab, as a function of the event size, ${ \mathcal B }$. The measurements are compared to simulation. The measured and predicted angular resolutions are close enough that using the simulated angular resolution for measuring spectra is a subdominant systematic error.

Standard image High-resolution image

Knowing the angular resolution is critical to subsequent steps of the analysis. Figure 9 indicates that the simulated angular resolution is in good agreement with measurements of the Crab Nebula. This is important because the angular resolution of HAWC for objects at declinations above and below the Crab will differ. While the measured PSF at the position of the Crab cannot be easily extrapolated to other declinations, the simulation can be used to predict the shape of the PSF at any declination. Therefore, the data–simulation agreement shown in Figure 9 is an important verification step. The angular resolution is weakly dependent on the assumed spectral index, and the impact on the analysis is described later in Section 4.3.

3.4. Cut Selection and Gamma-Ray Efficiency

The two parameters described in Section 2.6, the compactness ${ \mathcal C }$ and PINCness ${ \mathcal P }$, are used to remove hadrons and keep gamma rays. Events are removed using simple cuts on these variables, and the cuts depend on the size bin, ${ \mathcal B }$, of the event. The cuts are chosen to maximize the statistical significance with which the Crab is detected in the first 337 days of the 507-day data set. Concerns of using the data themselves for optimizing the cuts are minimal with a source as significant as the Crab.

Table 2 shows the cuts chosen for each ${ \mathcal B }$ bin. The rates of events across the entire sky going into the 9 bins, after hadron rejection cuts, vary dramatically, from ∼500 Hz for ${ \mathcal B }=1$ to ∼0.05 Hz for ${ \mathcal B }=9.$ Figure 11 shows the predicted efficiency for gamma rays (from simulation) along with the measured efficiency for hadronic background under these cuts. The efficiency of photons is universally greater than 30% while keeping, at best, only 2 in 103 hadrons. The efficacy of the cuts is a strong function of the event size, primarily because larger cosmic-ray events produce many more muons than gamma-ray events of a similar size.

Figure 11.

Figure 11. Fraction of gamma rays and background hadron events passing photon/hadron discrimination cuts as a function of the event size, ${ \mathcal B }$. Good efficiency for photons is maintained across all event sizes, with hadron efficiency approaching 1 × 10−3 for high-energy events.

Standard image High-resolution image

The limiting rejection at high energies is better than predicted in the sensitivity design study (Abeysekara et al. 2013). The original study was conservative in estimating the rejection power that HAWC would ultimately achieve. With more than a year of data, we now know the hadron rejection of the cuts and can accurately compute the background efficiency.

4. Spectral Fit

Knowing the angular resolution and the background in each ${ \mathcal B }$, the energy spectrum of the Crab Nebula may be inferred from the measured data. Section 4.1 describes the likelihood fit to the data, Section 4.2 describes the resulting measurement, and Section 4.3 describes the systematic errors to which this measurement is subject.

4.1. Likelihood Analysis

The HAWC data are fit using the maximum likelihood approach to find the physical flux of photons from the Crab (Wilks 1938; Younk et al. 2016). In this approach, the likelihood of observations is found under two "nested" hypotheses where some number of free parameters are fixed in one model. We form the likelihood of our observations under the null hypothesis, ${{ \mathcal L }}_{\mathrm{null}}$, and an alternative hypothesis, ${{ \mathcal L }}_{\mathrm{alt}}$, with additional free parameters that are not in the null-hypothesis model. If the null model is true, the Test Statistic, $\mathrm{TS}=-2\cdot \mathrm{ln}({L}_{\mathrm{null}}/{L}_{\mathrm{alt}})$, will be distributed as a χ2 distribution with a number of degrees of freedom equal to the number of additional free parameters in the alternative model, allowing a quantitative description of how much improvement the additional parameters provide.

The likelihood function is formed over the small (on the scale of the angular resolution) spatial pixels within 2° of the Crab. Each pixel p has an expected number of background events of Bp and, for a specific flux model, an expected number of true photons ${S}_{p}({\boldsymbol{a}})$, where ${\boldsymbol{a}}$ denotes the parameters of our spectral model of the Crab. The predicted photon counts fall off from the source according to the assumed PSF. The likelihood ${ \mathcal L }({\boldsymbol{a}})$ is then the simple Poisson probability of obtaining the measured events in each pixel, Mp, under the assumption of the flux given by ${\boldsymbol{a}}$. The ${ \mathcal B }$ dependence of each term in Equation (6) is suppressed,

Equation (6)

Specifically, we fit a differential photon flux ϕ(E) of the log parabola (LP) form:

Equation (7)

Here ϕ0 is the flux at E0, α is the primary spectral index, and β is a second spectral index that governs the changing spectral power change of spectral shape across the energy range of the fit. In this formulation, E0 is not fitted but is chosen to minimize correlations between the free parameters in the fit. When fit to an LP function, E0 = 7 TeV produces good results.

4.2. Fit Results

We find the parameters for ${\boldsymbol{a}}$ that maximize the likelihood function under signal and background hypotheses and quantify the error region of ${\boldsymbol{a}}$ using Wilks's theorem (Wilks 1938). Figure 12 shows the corresponding regions of α, β, and ϕ0 for the LP fit that are consistent with HAWC data at 1σ and 2σ. The maximum likelihood occurs at α = 2.63 ± 0.03, β = 0.15 ± 0.03, and ${\mathrm{log}}_{10}({\phi }_{0}\,{\mathrm{cm}}^{2}\,{\rm{s}}\,\mathrm{TeV})=-12.60\pm 0.02$. At this best flux the TS, compared to the background-only hypothesis, is 11,225, a more than 100σ detection.

Figure 12.

Figure 12. Likelihood space around Crab best fit. Shown are the best-fit flux and the region of fluxes allowed at 1σ and 2σ. The space shown is the ϕ0, α, and β space from Equation (7) with a pivot energy of E0 = 7 TeV.

Standard image High-resolution image

Figure 13 confirms that the fit is a faithful representation of the HAWC data. We show the excess number of events above background in each ${ \mathcal B }$ bin. These excesses come from a fit to the total number of events in the Crab in each ${ \mathcal B }$ bin. The measured number of events in each bin agrees well with the prediction from simulation at the fitted spectrum for the Crab.

Figure 13.

Figure 13. Measured, background-subtracted number of photons from the Crab in each ${ \mathcal B }$ bin. To get the total number of photons, the signal from the Crab is fit for each ${ \mathcal B }$ separately. The measurements are compared to prediction from simulation assuming that the Crab spectrum is at the HAWC measurement. The fitted spectrum is a good description of the data, with no evidence of bias in the residuals.

Standard image High-resolution image

The TS between an unbroken power-law hypothesis (with β = 0) and the full LP fit is 142, so the spectrum is inconsistent with an unbroken power law at 12σ. The best-fit hypothesis with an assumption that β = 0 is α = 2.57 ± 0.02 and ${\mathrm{log}}_{10}({\phi }_{0}\,{\mathrm{cm}}^{2}\,{\rm{s}}\,\mathrm{TeV})=-12.72\pm 0.01.$

We quantify the energy range of this fit in two ways. First, we take the spectral fit solution and compute the lower 10% quantile of true energy for ${ \mathcal B }=1$ and the upper 90% quantile of true energy for ${ \mathcal B }=9.$ These are 375 GeV and 85 TeV, respectively. This is the energy range over which, under the fitted hypothesis, most of the measured photons are expected to lie.

This energy range is representative of the true energies over which we believe that HAWC can detect the Crab. However, the energy analysis is rudimentary. Though we certainly detect the Crab Nebula for a ${ \mathcal B }$ of 1 through 9, the energies these bins correspond to can be uncertain by nearly an order of magnitude, depending on the spectral shape of the source. Hence, while 375 GeV–85 TeV represents the energy range over which the Crab is detected under the assumed functional form, we cannot definitively say that we detect either 375 GeV or 85 TeV photons with the current analysis.

A more conservative approach can be made focusing on the lowest and highest energies where HAWC data can definitively claim detection of photons. To do this, we separately fit functions of the form

and

to find the highest Elow and lowest Ehigh that are, at 1σ, inconsistent with the HAWC observation. This is a contrived, simplified test with the sole objective of making conservative estimates of the maximum and minimum energy of photons contributing to the Crab result. With this approach, we believe that we have positive detection of photons from the Crab between 1 and 37 TeV. That is, the ${ \mathcal B }=1$ detection means photons of at least 1 TeV, and the ${ \mathcal B }=9$ detection means photons of no lower than 37 TeV. This is not to say that higher- or lower-energy photons cannot be a part of the HAWC observation, but using the event size ${ \mathcal B }$ to measure the energy of photons limits the dynamic range of the observation. Other sources at other declinations may yield different high and low energies.

4.3. Systematic Errors

Table 3 summarizes the major systematic errors contributing to the measurement of the Crab spectrum with HAWC. These systematic errors have been investigated by computing the spectrum from the Crab under varying assumptions to study the stability of the results under perturbation of the assumptions.

Table 3.  Summary of Primary Contributions to HAWC Systematic Error in Measuring Photon Fluxes

Systematic Overall Flux Spectral Index log10(E)
Charge resolution/relative quantum efficiency ±20% ±0.05 <±0.1
PMT absolute quantum efficiency ±15% ±0.05 <±0.1
Time dependence, PMT layout, and Crab optimization ±10% ±0.1 <±0.1
Angular resolution ±20% ±0.1
Late light simulation ±40% ±0.15 <±0.15
Total flux ±50% ±0.2 <0.2

Note. The different effects are described in the text. Systematics in the overall flux, the spectral index of sources, and the energy scale are shown. The systematics claims are conservative and are likely to improve with more understanding and better modeling.

Download table as:  ASCIITypeset image

For spectral measurements, a systematic error in three quantities is shown: the overall flux, the spectral index measured, and the energy scale. The errors are summed in quadrature to arrive at a total systematic error. In addition to these systematic errors, a systematic error in the absolute pointing of the instrument has been studied.

4.3.1. Charge Resolution and Relative Quantum Efficiency

The charge resolution is a quantity that captures by how much individual PMT charge measurements can vary, for fixed input light, and is estimated to be 10%–15% from studies using the HAWC calibration system. Additionally, PMTs vary in their photon detection efficiency by 15%–20%. The relative detection efficiency has been estimated from laboratory measurements prior to deployment, variations in the measured charge from vertical muons (selected by their time), and variations in the rate at which individual PMTs participate in events at different charge levels. These factors are not simple numbers, but can vary for different light levels in the detector and can change with the arrival time distribution during air showers. Varying these assumptions, the ${ \mathcal P }$ and ${ \mathcal C }$ change and impact the event passing rates, impacting the spectrum.

4.3.2. PMT Absolute Quantum Efficiency

PMTs have an efficiency for converting photons impinging on their surface into PEs detected by the PMT, typically between 20% and 30%. Of course, a single "efficiency" number vastly simplifies the situation: the efficiency is divided between the efficiency for producing a PE and the efficiency for collecting a PE, varies across the face of the PMT, and is wavelength dependent. Additionally, the absorption of the water itself is wavelength dependent. Much of this is modeled, but the simulation carries uncertainties in the treatment and is difficult to validate. The calibration system, in particular, cannot yield the absolute PMT efficiency because it requires establishing the efficiency of the calibration system's optical path to the PMTs much more precisely than is known. Furthermore, the laser for the calibration system is green light and must be extrapolated for application to blue Cherenkov light.

Instead, the absolute efficiency is established by selecting vertical muons in HAWC tanks by their timing properties. Vertical muons are typically minimum ionizing with a relatively constant energy loss. The simulated response to vertical muons is scaled to match data. Nevertheless, we estimate a ±10% uncertainty in the absolute PMT efficiency that propagates to the errors in the spectra of sources.

4.3.3. Time Dependence, PMT Layout, and Crab Optimization

The HAWC instrument has changed over time. The main change is that PMTs or channels are occasionally removed during maintenance. Repaired channels have been recalibrated, and the calibration constants have been changed occasionally for other reasons. The event size bins, ${ \mathcal B }$, are based on fractions of available PMTs to mitigate the impact of the varying numbers of PMTs. Furthermore, a single simulation with a single representative PMT layout is used to model the detector, and this simplification results in a corresponding systematic error.

Different detector layouts were simulated to bound the impact of sporadic tubes being added or removed. As confirmation, the passing rate of background cosmic rays through the photon/hadron discrimination cuts was studied and shows comparable drifts to the simulation studies.

Additionally, the cuts used in the analysis were established by maximizing the statistical significance of the observations of the Crab Nebula during the first 337 days of data. While strictly not an a priori measurement, the Crab is strong enough in HAWC data that there is very little bias to the final measurement from this optimization process (and no bias for other weaker sources).

To investigate any potential bias, the data were divided into two sections, a 337-day and a 170-day data set. The 337-day data set corresponds to the period over which the cut optimization was done and is the only data set that could have an overoptimization bias. The Crab spectrum was then measured separately in each data set. The fitted spectra differ by ±10% in the flux and ±0.1 in the spectral index, similar to what is expected from the varying number of PMTs.

It is unclear whether the different Crab spectra in the 337-day and 170-day data sets are due to overtuning the Crab or the changing detector later in the data taking. They are similar size effects. Whatever the origin of the effect, it is a subdominant, but non-negligible, systematic error.

4.3.4. Angular Resolution

The chief uncertainties in the angular resolution arise from a mismatch between the data and the simulation and spectral dependence of the angular resolution. The impact of angular resolution has been studied by reconstructing the Crab spectrum under different angular resolution hypotheses.

4.3.5. Late Light Simulation

The single largest source of systematic error is how late light in the air shower is treated. Simulation suggests that the arrival time distribution of PEs at the PMTs should be well within ∼10 ns. Nevertheless, the distributions of ${ \mathcal C }$ and ${ \mathcal P }$ in background cosmic rays, as well as the raw PE distributions themselves, suggest some mismodeled effect above about 50 PEs.

Dedicated studies of late light, using the the calibration system, have been seen to extend ToT measurements, thereby distorting the measured charge in PMTs, but an arrival time distribution much wider than expected from simulation is needed to explain the data.

Efforts to understand this systematic are aimed at measuring the entire PMT waveform for a sample of PMTs to better understand the arrival time distribution of PEs without requiring simulation. It is likely that this systematic error will be better understood in the future, but currently it dominates.

4.3.6. Absolute Pointing

The absolute pointing error is estimated to be no more than 0fdg1 for sources that transit above 45° in HAWC. It is estimated to be no more than 0fdg3 for higher-inclination sources.

The absolute pointing of the HAWC instrument is impacted by the timing calibration, as discussed in Section 2.2. Each PMT has a calibrated offset to account for different cable lengths and other timing delays. These offsets are established coarsely by repeated reconstructions to force the peak of maximum cosmic-ray density to be overhead. With this correction, the location of the Crab is within 0fdg2 of its true location. A final detailed alignment is performed to put the Crab Nebula in its known location. The Crab itself must be in the correct location, by construction.

The absolute pointing error on other sources has been studied two ways. First, the Crab location has been fit using only events in bands of reconstructed zenith angle. The Crab location drifts by no more than 0fdg1 up to a zenith angle of 45°. Above that inclination, the Crab is weakly detected, and we cannot independently demonstrate better than 0fdg3 absolute pointing error. Furthermore, the Crab location has been reconstructed separately using data from each of the nine ${ \mathcal B }$ bins, and they agree to within 0fdg1.

Finally, other bright known sources, the blazars Mrk 421 and Mrk 501, agree with their known locations to within 0fdg1.

5. Discussion

5.1. Comparison to Other Experiments

Figure 14 shows the Crab spectrum measured with HAWC between 1 and 37 TeV compared to the spectrum reported by other experiments. It is consistent with prior measurements within the systematic errors of the HAWC measurement.

Figure 14.

Figure 14. Crab photon energy spectrum measured with HAWC and compared to other measurements using other instruments (Aleksić et al. 2015; Amenomori et al. 2009, 2015; Bartoli et al. 2015; Holler et al. 2015; Meagher 2016). The red band shown for HAWC is the ensemble of fluxes allowed at 1σ, and the best fit is indicated with a dark red line. The light-red band indicates the systematic extremes of the HAWC flux.

Standard image High-resolution image

A number of improvements to the HAWC measurement—most notably the inclusion of a proper energy reconstruction—will reduce the systematic errors and increase the dynamic range of our measurements. The observation of the Crab validates this analysis for subsequent application to the sources across the rest of the sky.

5.2. Performance Figures

Figure 15 shows the effective area of HAWC, in this analysis, to photons arriving within 13° from overhead. The effective area is defined as the geometrical area over which events are detected, convolved with the efficiency for detecting events, and is determined from simulations. The exact conditions for a photon to be considered detected are complicated for this analysis because, since we are performing a likelihood fit of all pixels in the vicinity of the Crab, photons that are poorly reconstructed play some role in the analysis. In order to have a well-formed effective area, we consider only photons defined within the 68% containment radius, from Table 1. For comparison, Figure 15 includes the progression of cuts, from the effective area without any photon/hadron discrimination without a strong angular accuracy cut to the full analysis cuts. The effective area can exceed the geometrical area of the instrument (about 2 × 104 m2) because events with a core location off the detector occasionally pass the imposed cuts.

Figure 15.

Figure 15. Effective area for HAWC for events within 13° from overhead. To show the progression of analysis cuts, we show curves without any photon/hadron discrimination, insisting that events only reconstruct within 4° of their true direction. Requiring events to be reconstructed within their 68% containment radius lowers the effective area, and photon/hadron discrimination cuts lower it further. With a requirement that events be reconstructed on the detector, the effective area flattens at roughly half the physical area of the instrument.

Standard image High-resolution image

The effective area computed in this work is smaller than in Abeysekara et al. (2012). For this analysis, developed for steady, multi-TeV sources, we employ tight angular cuts and have a higher energy threshold than used in the initial design study. Furthermore, many small events are excluded by demanding that they are larger than ${ \mathcal B }=1$. Excluding these events limits the effective area below 1 TeV.

Figure 16 shows the computed differential sensitivity of HAWC to sources at the declination of the Crab utilizing the procedure of Abeysekara et al. (2013) with the analysis presented here. A point source of differential photon spectrum ${E}^{-2.63}$ is simulated and fitted using the full likelihood fit. The flux required to be detected at 5σ 50% of the time is shown for each bin, ${ \mathcal B }$. The lines for each ${ \mathcal B }$ are shown with a width corresponding to the width required to contain 68% of the events under the ${E}^{-2.63}$ hypothesis. A correction is made to adjust the ${ \mathcal B }$ separation to a quarter decade in true energy, and the result is fitted. The sensitivity prediction from Abeysekara et al. (2013) suffered from uncertain background at the highest energies. Now, with more than a year of data, we know the background precisely and can set the cuts appropriately. This has resulted in a more accurate (and more sensitive) analysis above 10 TeV. Below about 1 TeV, for a number of reasons, the sensitivity is somewhat worse than predicted in the original study. The background is larger than the original simulation-only prediction. Furthermore, in the current analysis we employ a relatively high cut (defined by ${ \mathcal B }=1$) so that improperly modeled noise can be neglected.

Figure 16.

Figure 16. Quasi-differential sensitivity of HAWC to a point source at a declination of +22°N as a function of photon energy, compared to existing IACTS (Aleksić et al. 2016; Holler et al. 2015; Park 2016) and Large Area Telescope on the Fermi Gamma-Ray Space Telescope (Pass 8 Sensitivity: https://www.slac.stanford.edu/exp/glast/groups/canda/lat_Performance.htm). Also shown is an inferred quarter-decade differential upper limit from the Crab at 141 TeV from the CASA-MIA experiment (Borione et al. 1997). We show the flux, assuming a source with a differential energy spectrum E−2.63, required to produce a 5σ detection 50% of the time. This flux is shown in light red for each of the nine ${ \mathcal B }$ bins, with a width in energy corresponding to the central 68% containment energies in each bin. These values are adjusted to find the equivalent quarter-decade-separated flux sensitivities, and a fit to these values is shown in dark red. The 507-day observation of HAWC corresponds to ∼3000 hr of a source at a declination of 22° within HAWC's field of view. HAWC's 1 yr sensitivity surpasses a 50 hr observation by current-generation IACTs at ∼10 TeV.

Standard image High-resolution image

5.3. Anticipated Improvements

The main limitation of this analysis is the reliance on the number of PMTs, used for the definition of ${ \mathcal B }$, to simultaneously constrain the energy of photons, angular resolution, and photon/hadron efficiency. Figure 2 shows that this is a poor energy estimation, with each bin ${ \mathcal B }$ spanning roughly an order of magnitude of energy. More critically, an overhead ∼10 TeV photon can trigger nearly every PMT in HAWC if the core lands near the center of the detector. Consequently, ${ \mathcal B }=9$ is an overflow bin of everything above 10 TeV. These limitations can be removed with an event parameter that accounts for the light level in the event and the specific geometry and inclination angle of events. Approaches like this are under development. The planned deployment of a sparse "outrigger" array should further increase the sensitivity to photons above 10 TeV (Sandoval 2016).

Additionally, the principal systematic error (the modeling of late light) is conservatively estimated here and is being studied using the calibration system. It is likely that the effects of late light will be better modeled in the future.

Finally, the threshold for this analysis is established by including only events where more than 6.7% of the PMTs detect light. The typical number of live, calibrated PMTs is ∼1000, corresponding to a threshold of ∼70 PMTs. Events with 20–30 PMTs could be reconstructed if the noise could be confidently identified. A relatively high event size threshold is used in this analysis to reduce its dependence on the modeling of noise hits. Planned improvements in the modeling should lower the energy threshold of the spectrum analysis in future studies.

The HAWC instrument is performing well, with survey sensitivity exceeding current-generation instruments above 10 TeV, sensitivity that HAWC maintains across much of its field of view. The all-sky survey Abeysekara et al. 2017 conducted by HAWC probes unique flux space and reveals the highest-energy photon sources in the northern sky. Understanding the Crab gives confidence in the survey results.

We acknowledge support from the U.S. National Science Foundation (NSF); the U.S. Department of Energy Office of High-Energy Physics; the Laboratory Directed Research and Development (LDRD) program of Los Alamos National Laboratory; Consejo Nacional de Ciencia y Tecnología (CONACyT), México (grants 271051, 232656, 260378, 179588, 239762, 254964, 271737, 258865, 243290, 132197), Laboratorio Nacional HAWC de rayos gamma; L'OREAL Fellowship for Women in Science 2014; Red HAWC, México; DGAPA-UNAM (grants IG100317, IN111315, IN111716-3, IA102715, 109916, IA102917); VIEP-BUAP; PIFI 2012, 2013, PROFOCIE 2014, 2015; the University of Wisconsin Alumni Research Foundation; the Institute of Geophysics, Planetary Physics, and Signatures at Los Alamos National Laboratory; Polish Science Centre grant DEC-2014/13/B/ST9/945; and Coordinación de la Investigación Científica de la Universidad Michoacana. Thanks to Luciano Díaz and Eduardo Murrieta for technical support.

Please wait… references are loading.
10.3847/1538-4357/aa7555