Brought to you by:

OSSOS. VIII. The Transition between Two Size Distribution Slopes in the Scattering Disk

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

Published 2018 April 19 © 2018. The American Astronomical Society. All rights reserved.
, , Citation S. M. Lawler et al 2018 AJ 155 197 DOI 10.3847/1538-3881/aab8ff

Download Article PDF
DownloadArticle ePub

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

1538-3881/155/5/197

Abstract

The scattering trans-Neptunian Objects (TNOs) can be measured to smaller sizes than any other distant small-body population. We use the largest sample yet obtained, 68 discoveries, primarily by the Outer Solar System Origins Survey (OSSOS), to constrain the slope of its luminosity distribution, with sensitivity to much fainter absolute H-magnitudes than previous work. Using the analysis technique in Shankman et al., we confirm that a single slope for the H-distribution is not an accurate representation of the scattering TNOs and Centaurs, and that a break in the distribution is required, in support of previous conclusions. A bright-end slope of αb = 0.9 transitioning to a faint-end slope αf of 0.4–0.5 with a differential number contrast c from 1 (a knee) to 10 (a divot) provides an acceptable match to our data. We find that break magnitudes Hb of 7.7 and 8.3, values both previously suggested for dynamically hot Kuiper Belt populations, are equally non-rejectable for a range of αf and c in our statistical analysis. Our preferred divot H-distribution transitions to αf = 0.5 with a divot of contrast c = 3 at Hb = 8.3, while our preferred knee H-distribution transitions to αf = 0.4 at Hb = 7.7. The intrinsic population of scattering TNOs required to match the OSSOS detections is 3 × 106 for Hr < 12, and 9 × 104 for Hr < 8.66 (D ≳ 100 km), with Centaurs having an intrinsic population two orders of magnitude smaller.

Export citation and abstract BibTeX RIS

1. Introduction

The populations of small bodies in our solar system are incrementally grinding themselves into dust through mutual collisions. On short timescales, collisions are infrequent, though on occasion, the aftermath can be directly observed (e.g., Jewitt et al. 2010). Over the age of the solar system, collisions may be the main force that shaped the observed size distribution of all but the largest trans-Neptunian objects (TNOs; Schlichting et al. 2013), or the size distribution may be a result of formation conditions (Fraser et al. 2014). As dynamical evolution is not size-dependant for these small TNOs, we do not expect the size distribution to be affected by removal of TNOs from the scattering population due to interactions with the giant planets. The size distribution of populations that are shaped by collisions can be described by a power law of the form $\tfrac{{dN}}{{dD}}\propto {D}^{-q}$, where an idealized infinite collisional cascade will produce an exponent of q = 3.5 (Dohnanyi 1969).

In the outer solar system, the luminosity distribution must be used as a proxy for the size distribution, because TNOs are unresolved. Luminosity is measured as an apparent magnitude, which can be directly converted to an absolute magnitude H when combined with a measured distance. The H-magnitude can then be directly mapped to diameter, as long as an albedo is measured (or assumed). A handful of small (H ∼ 9–14) TNOs and Centaurs have had their albedos measured observationally, and they range from 4% to 16% (Duffard et al. 2014). The size distribution can be written in terms of absolute magnitude H as $\tfrac{{dN}}{{dH}}\propto {10}^{\alpha H}$, where the size distribution exponent q is related to the H-magnitude exponent α by q = 5α + 1 (assuming albedo is size-independent; Irwin et al. 1995; Fraser et al. 2008; Petit et al. 2008).

Measuring the size distribution of a small-body population tells us about their composition, collisional processes that shape them, and may also provide information on their formation. Collisional simulations of the asteroid belt (e.g., Bottke et al. 2005; Pan & Schlichting 2012) have found that the sizes of the largest asteroids are set by the initial formation sizes, which in combination with mass depletion of the asteroid belt (caused by Jupiter's migration), sets any structure in the size distribution. The size distribution of the asteroids can be measured to much smaller sizes (larger H-magnitudes) than the TNOs due to the fact that it is much closer, and thus smaller objects will be above survey detection limits. The asteroid size distribution at smaller sizes shows intriguing structure, which collisional simulations have shown to likely be caused by a combination of formation size and the initial number density of the asteroid belt; the transition between primordial and collisionally evolved populations happens at ∼10–100 km (Bottke et al. 2005; Morbidelli et al. 2009a). By measuring the size distribution of the Kuiper Belt across several orders of magnitude in size, as has been done in the asteroid belt, we may gain an additional constraint on the timing and manner of Neptune's migration, which severely depleted the mass of the Kuiper Belt (Malhotra 1995; Gomes et al. 2004; Nesvorný 2015).

The magnitude distribution of the Kuiper Belt has long been modeled as a single slope at large sizes (e.g., Jewitt et al. 1996). Gladman et al. (2001) found that the smallest TNOs had a size distribution inconsistent with a single power law. Later, Bernstein et al. (2004) measured a rollover, proving that a single power law was not adequate to describe the observed Kuiper Belt. Surveys are now reaching deep enough and detecting enough TNOs that additional structure in the size distribution is required to match observations (Fuentes & Holman 2008; Shankman et al. 2013, 2016; Fraser et al. 2014; Alexandersen et al. 2016).

Here, we focus our analysis on the scattering TNOs and Centaurs. Because they come closer to the Sun than most TNOs, we can observe smaller TNOs within this population than any other in the Kuiper Belt. Scattering TNOs and Centaurs are part of the dynamically "hot" population. TNOs in the dynamically hot population have had their orbits excited to higher inclinations and eccentricities by scattering off Neptune or past/current entanglement with mean-motion resonances (Gladman 2005). Previous work has demonstrated that the hot population, due to its different collisional and formation history, has a different size distribution than the dynamically cold population of the main classical Kuiper Belt (Petit et al. 2011; Fraser et al. 2014). We specifically exclude those TNOs that are currently resonant from the analysis presented in this manuscript, as they are likely to have experienced a different pathway to dynamical excitation than the scattering TNOs and Centaurs (i.e., Gladman et al. 2012).

Shankman et al. (2016) used scattering TNOs detected in four well-characterized surveys to measure the scattering TNO H-distribution to great precision. In this work, we provide an update for the measurement of the scattering TNO H-distribution with the inclusion of the full discovery data set of the Outer Solar System Origins Survey (OSSOS; Bannister et al. 2016; the full data set is in Bannister et al. 2018). OSSOS has completed its observing, more than tripling the sample of scattering TNOs and Centaurs since the analysis of Shankman et al. (2016).

The analysis here builds on the work of Shankman et al. (2016), using the same methodologies. We first discuss the OSSOS survey, summarizing the mechanics of the Survey Simulator, which allows us to forward-bias our model to allow statistical comparison with our observational sample of TNOs. In Section 3, we summarize the statistical analysis that we use to find the range of acceptable H-magnitude distributions allowed by our observed sample. Section 4 includes our population measurements, and in Section 5, we discuss how our measurements of the scattering disk fit into the larger context of the solar system.

2. Scattering Sample Selection

Because scattering TNOs and Centaurs have high eccentricities, and their pericenter distances can range from nearly Jupiter-crossing to >40 au, the observing biases are extreme and must be accounted for carefully; e.g., small Centaurs and TNOs with closer pericenters are far more likely to be detected in magnitude-limited surveys (as is visible in Figure 1). By using only TNOs detected by well-characterized surveys in this analysis, where the magnitude limits, pointings, and tracking efficiencies are known and published,10 we are able to forward-bias models of the scattering disk and statistically compare the resulting biased simulated detections with the real TNO discoveries.

Figure 1.

Figure 1. Orbital properties (pericenter distance q and inclination i) of the 68 TNOs detected by the OSSOS ensemble of surveys (see Table 3 in the Appendix, and Table 3 in Bannister et al. 2018) that are classified as scattering (circles) or Centaurs (squares; see Section 2 for details on classification). Semimajor axis a is shown via point color, most of these TNOs have a < 200 au. Point sizes are proportional to diameter (assuming the same albedo); note that the closest objects are preferentially small due to discovery biases. Outside q > 37 au (noted with dotted line in plot), scattering TNOs have preferentially larger a and are more weakly bound. See the text for further discussion.

Standard image High-resolution image

OSSOS is a large program on the Canada–France–Hawaii Telescope over five years to discover TNOs while carefully characterizing tracking fractions, detection efficiencies, and pointing directions, allowing the survey biases to be fully quantified (Bannister et al. 2016). This methodology has been followed for three other large Kuiper Belt surveys: The Canada–France Ecliptic Plane Survey (CFEPS; Petit et al. 2011), the CFEPS high latitude component (HiLat; Petit et al. 2017), and the survey of (Alexandersen et al. 2016, hereafter referred to as MA). Combining these three surveys with OSSOS gives a well-characterized set of surveys (which we refer to throughout this paper as the "OSSOS ensemble"), whose combined detected TNOs provide powerful constraints on the intrinsic TNO orbital distributions and populations when used in combination with the Survey Simulator. This statistical reproduction of the survey biases is discussed extensively in other works, (e.g., Kavelaars et al. 2009; Petit et al. 2011; Bannister et al. 2016; Shankman et al. 2016; Lawler et al. 2018).

Shankman et al. (2016) analyzed a scattering TNO sample of 22 objects from CFEPS, HiLat, MA, and the first two (of eight) observing blocks of OSSOS. OSSOS has since detected dozens of new scattering TNOs and Centaurs, bringing the full sample available for analysis to 68 TNOs (17 Centaurs, 51 scattering). The orbital elements of the full sample analyzed here are shown in Figure 1 and Table 3 in the Appendix, and further detail is available in Table 3 (ensemble catalog) in Bannister et al. (2018).

Here, we use the dynamical classification scheme of Gladman et al. (2008) to determine membership in the scattering and Centaur classes. These two classes are both unstable on timescales much shorter than the age of the solar system. The distinction between them is semimajor axis a relative to Neptune's orbit; Centaurs have smaller a and scattering TNOs have larger (shown by different symbols in Figure 1). Their changes in a over time are usually due to close encounters with one of the giant planets, but can also be due to dynamical diffusion for the more weakly bound TNOs (Bannister et al. 2017): those that have largest pericenter distances q > 37 au tend to also have the largest semimajor axes of the sample (see Figure 1). The Centaurs show similar evolution in semimajor axis and represent the low-a tail of the scattering population (e.g., Gomes et al. 2008), thus it is expected that they will share the same H-distribution.

3. Measuring the True H-distribution

Previous work has shown that there is a sharp transition in the H-distribution of the TNOs, though the form of the transition is unclear (Shankman et al. 2013, 2016; Fraser et al. 2014; Alexandersen et al. 2016). We parameterize this transition using a bright-end slope αb, a faint-end slope αf, a break magnitude Hb, and a differential contrast c. We use the terminology that c = 1 is a knee, c > 1 is a divot. We refer the reader to Figure 9 in Shankman et al. (2016) for a graphical demonstration of the effect of these two different transitions on the cumulative and differential number distributions in H.

The slope at the bright end of the TNO H-distribution αb, a range of H ≃ 4–7, is well-probed by previous work (e.g., Fraser & Kavelaars 2009; Petit et al. 2011). Our OSSOS ensemble detections range from Hr values of 6 to 14.5 because of the very close pericenter distances of some of these TNOs, and thus this analysis is sensitive to a much fainter Hr range than previous work. We note that several of the scattering TNOs included in this sample were not observed in r-band because some blocks of CFEPS observed only in g. These have had their g-band and Hg magnitudes transposed to r by assuming that g − r = 0.7, which is at the neutral end of the observed color range of dynamically excited TNOs (Tegler et al. 2016). Shankman et al. (2016) used g − r = 0.7, and also demonstrated that using g − r values ranging from 0.5 to 0.9 makes no difference to the statistical analysis performed below (see Figure 8 in Shankman et al. 2016).

In this analysis, as in previous work (Shankman et al. 2013, 2016), we seek to measure the slope of the faint end of the H-distribution αf, the contrast of the transition c, and the H-magnitude where the break occurs Hb. We test H-distributions from a grid covering αf from 0.1 to 0.9, and c from 1 to 100, with two different break magnitudes, Hb = 8.3 (preferred break magnitude from Shankman et al. 2016) and Hb = 7.7 (preferred break magnitude from Fraser et al. 2014).

3.1. The Survey Simulator and Statistical Analysis

Our method of forward-biasing a model distribution with different H-distributions is discussed in detail in Shankman et al. (2013) and Shankman et al. (2016). Briefly, we start with a version of the scattering distribution modeled by the emplacement simulation of Kaib et al. (2011), with the dynamically hotter inclination distribution used in Shankman et al. (2016). We then draw orbits from this simulation. Orbits are randomly oriented (random ω and Ω), and objects are placed with a random mean anomaly on these orbits (which sets the distance), and are given an H-magnitude from within a chosen H-distribution, and then an r-magnitude is calculated. The Survey Simulator then determines if that r-magnitude, rate of motion, and on-sky position was detectable in the OSSOS survey ensemble. This process is continued until a large number (hundreds) of simulated detections are created. The cumulative distributions of simulated detections are then statistically compared with the cumulative distributions of the 68 real Centaurs and scattering TNOs in semimajor axis a, inclination i, r-magnitude mr, pericenter distance q, distance at detection d, and H-magnitude in r-band Hr. These six cumulative distributions are shown in Figure 2 for the real TNOs as well as simulated detections using three different H-distributions.

Figure 2.

Figure 2. Cumulative distributions across six parameters for the 68 observed scattering TNOs and Centaurs (red step-function), and three candidate H-distributions. Panels (A–F) correspond to the semimajor axis a, inclination i (see Section 3.1.1), magnitude at detection in r-band mr, pericenter q, distance at detection d, and H-magnitude in r-band, respectively. The rightmost panel provides schematics for three different H-distributions: (1) our preferred (c, αf) pair (solid blue line) (2) our preferred knee distribution (dotted–dashed purple line) and (3) the best-fit knee distribution from Fraser et al. (2014) (dashed green line).

Standard image High-resolution image

The statistical analysis is described in detail in Shankman et al. (2016), and we summarize below. We first calculate the Anderson–Darling (AD) statistic (Anderson & Darling 1954), comparing the observed TNOs and the simulated detections for a given H-distribution. An AD statistic is computed for each parameter. From previous work, we found that the most powerful lever arms for this analysis (because they vary most for different modeled H-distributions) come from using the parameters q, d, and Hr, so we sum the AD statistics calculated for each of these three distributions (following the analysis method of Parker 2015). This summed AD statistic is bootstrapped by selecting at random 68 objects from the distribution of simulated detections, calculating the AD statistic between this random sample and the simulated detections in each parameter, and summing them. This random selection and AD statistic calculation is repeated hundreds of times. The distribution of summed AD statistics for random samples of the simulated distribution is then compared to the summed AD statistic for the real TNOs. If that AD statistic or larger occurs for <5% of the random distributions, we can reject that distribution with >2σ (>95%) confidence. To explain in another way, if <95% of random subsets of the model are farther from the parent model than the observations are, then the model cannot be rejected.

3.1.1. The Scattering Inclination Distribution

Figure 2 shows a good match between the observations and the preferred model for five of the six parameters measured; the inclination distribution has a rather poor fit at high inclinations (this is true for all H-distributions tested). The paucity of high-inclination objects in the model as compared with observations was noted and discussed in Shankman et al. (2016). The difficulty of generating high-inclination objects in emplacement models is a well-noted problem (e.g., Kaib et al. 2011), and suggests that a small fraction of scattering TNOs may require a different emplacement pathway in order to match the real Kuiper Belt. Suggested mechanisms in the literature include diffusion from the Oort Cloud (Kaib et al. 2009; Brasser et al. 2012), interaction with a distant massive planet (Gomes et al. 2015), and interaction with a rouge planet that was later ejected from the solar system (Gladman & Chan 2006). Creating dynamical emplacement models of the Kuiper Belt that obtain a realistic inclination distribution is currently an area of active research.

We perform a simple experiment to make sure that the inclination distribution does not severely affect the three parameters we test (H, d, and q) by doubling and halving all of the inclinations in the model and re-running our statistical test. We find that the bootstrapped AD values only vary by 1%–2% for these two very different inclination distributions, and so we conclude that while the inclination distribution shown in Figure 2 does not provide a great match to observations, the other properties of the model still provide an excellent fit to the real scattering TNOs.

3.2. Preferred H-distribution

Using the 68 detected scattering TNOs and Centaurs from the OSSOS ensemble, we find that the least-rejectable H-distribution is for αf = 0.5 and c = 3.2, using αb = 0.9. This H-distribution is shown as a blue solid line in Figure 2, and by a blue star in Figure 3.

Figure 3.

Figure 3. Contours of the rejectability for the tested faint-end slope αf and contrast c pairs with a break located at Hb = 8.3 (left) and Hb = 7.7 (right); all models tested use αb = 0.9. The contours represent the 1σ, 2σ and 3σ rejectability levels with white being rejectable at >3σ, red being rejectable at >2σ, and orange and yellow not statistically rejectable. Stars highlight notable (αf, c) pairs: the dark blue, green, and purple stars show models that are also plotted in the same color in Figure 2. The dark blue star denotes our preferred (αf, c) pair (see Section 3.2), the green star (with 1σ error bars) denotes the best-fit knee model for dynamically hot TNOs from Fraser et al. (2014), and the purple star is our preferred knee model. For comparison with previous work, the white star denotes the preferred (αf, c) pair from Shankman et al. (2016), and the black star denotes a single slope of α = 0.9 (identical in both plots), and is strongly rejectable.

Standard image High-resolution image

We are unable to statistically reject a knee distribution. A transition to a faint slope αf = 0.4 at Hb = 7.7 is non-rejectable at 3σ significance in our analysis; this preferred knee H-distribution is shown by a purple dashed–dotted line in Figure 2 and by a purple star in Figure 3. For comparison, the best-fit knee distribution from Fraser et al. (2014) is shown with a green star, including 1σ error bars.

The preferred divot H-distribution from Shankman et al. (2016) remains a viable explanation for the scattering TNO H-distribution (white star in Figure 3), but the analysis here increases the number of rejectable models, more tightly constraining the acceptable parameter space of αf and c. As in Shankman et al. (2013, 2016), a single power law (c = 1, αf = αb = 0.9) is rejectable at >3σ significance (shown with a black star in both plots in Figure 3).

Interestingly, we are not able to rule out either break magnitude Hb we tested. We tested two different values of Hb: 8.3 and 7.7, based on predictions from previous work (Fraser et al. 2014; Shankman et al. 2016). The yellow contours in Figure 3 highlight the H-distributions, which are rejectable by our analysis at the lowest significance (i.e., least-rejectable distributions). Contours of <1σ rejectability occur for both Hb values that we tested, and (αf = 0.5, c = 3.2) are the least-rejectable H-distributions for both values of Hb.

4. The Intrinsic Population Size

We use the Survey Simulator to determine the number of scattering TNOs and Centaurs brighter than a given H-magnitude that must be drawn from the Kaib et al. (2011) scattering TNO model to allow 68 detections (Table 1: scattering TNOs), or 17 detections from the a < 30 au subset of the Kaib et al. (2011) scattering TNO model (Table 2: Centaurs). Error bars on these intrinsic populations are calculated by running this experiment many times and finding the populations that bracket 95% of the estimates; the error bars given are thus 95% confidence intervals on the intrinsic population.

Table 1.  Population Estimates for Scattering TNOs

      Hr < 8.66 Hr < 10 Hr < 12
Hb αf c Population Population Population Comment
8.3 0.5 3.2 (0.9 ± 0.2) × 105 (2.9 ± 0.7) × 105 (2.7 ± 0.7) × 106 preferred divot, this work
7.7 0.4 1 (0.8 ± 0.2) × 105 $({3.5}_{-0.6}^{+0.9})\times {10}^{5}$ $({2.4}_{-0.4}^{+0.6})\times {10}^{6}$ preferred knee, this work
8.3 0.5 5.6 $({1.0}_{-0.2}^{+0.3})\times {10}^{5}$ $({2.6}_{-0.5}^{+0.7})\times {10}^{5}$ $({2.1}_{-0.4}^{+0.6})\times {10}^{6}$ preferred, Shankman et al. (2016)
8.3 0.4 1 (0.8 ± 0.2) × 105 (4.0 ± 0.9) × 105 $({2.8}_{-0.7}^{+0.6})\times {10}^{6}$ least-rejectable knee, Hb = 8.3
7.7 0.5 3.2 (0.7 ± 0.2) × 105 $({2.8}_{-0.6}^{+0.7})\times {10}^{5}$ $({2.7}_{-0.6}^{+0.7})\times {10}^{6}$ least-rejectable divot, Hb = 7.7
Previously Published Population Estimates
8.3 0.5 5.6     ∼1 × 106 estimate from Shankman et al. (2013)
8.3 0.5 5.6     (2.4–8.3) × 105 estimate from Shankman et al. (2016)
0.8 $({5}_{-3}^{+5})\times {10}^{3}$   $({4}_{-3}^{+4})\times {10}^{6}$ CFEPS estimate (Petit et al. 2011)

Note. Error bars on population estimates are 95% confidence intervals.

Download table as:  ASCIITypeset image

Table 2.  Population Estimates for Centaurs

      Hr < 8.66 Hr < 10 Hr < 12
Hb αf c Population Population Population Comment
8.3 0.5 3.2 ${110}_{-40}^{+60}$ ${390}_{-150}^{+200}$ ${3500}_{-1400}^{+1800}$ preferred divot, this work
7.7 0.4 1 ${130}_{-70}^{+80}$ ${550}_{-290}^{+340}$ ${3700}_{-2000}^{+2300}$ preferred knee, this work
Previously Published Population Estimates
7.7 0.2 1 ≤75,000 $({2.8}_{-2.5}^{+10.0})\times {10}^{4}$   calculated from Uranian co-orbitalsa
8.5 0.5 6 ≤75,000 $({2.8}_{-2.5}^{+13.0})\times {10}^{4}$   calculated from Uranian co-orbitalsa
7.7 0.2 1 ${2500}_{-2100}^{+11,000}$ ${7100}_{-6800}^{+32,000}$   calculated from Neptunian co-orbitalsa
8.5 0.5 6 ${2900}_{-2500}^{+11,000}$ ${7500}_{-7100}^{+32,000}$   calculated from Neptunian co-orbitalsa

Note. Error bars on population estimates are 95% confidence intervals.

aCalculated from observations and models of Alexandersen et al. (2013, 2016). Note that populations here are actually for the a < 34 au scattering population, a large fraction of which will be Centaurs; see the text.

Download table as:  ASCIITypeset image

4.1. The Size of the Scattering TNO Population

Using our preferred H-distribution (Hb = 8.3, αf = 0.5, c = 3.2), the population must be $({2.7}_{-0.5}^{+0.6})\times {10}^{6}$ for Hr < 12 (which corresponds to D ≳ 20 km for an albedo of 0.04), and $(8\pm 2)\times {10}^{4}$ for Hr < 8.66 (which corresponds to D ≳ 100 km for an albedo of 0.04). Interestingly, using other statistically acceptable H-distributions does not cause the population to vary by more than a very small factor; the population estimates from all statistically acceptable H-distributions are consistent within the 95% error bars.

Table 1 lists population estimates using several different H-distributions that are statistically acceptable in our analysis, as well as comparisons with previously published scattering population estimates. Our population estimates here are slightly higher than those reported in Shankman et al. (2013) and Shankman et al. (2016). CFEPS (Petit et al. 2011) estimates a population of ${5000}_{-3000}^{+5000}$ scattering TNOs for Hg < 9.16 (Hr ≲ 8.66), much smaller than our population estimate. However, after scaling by the assumed single slope of α = 0.8 down to Hr < 12 gives $({4}_{-3}^{+4})\times {10}^{6}$, consistent with our population estimates, albeit with very large error bars.

Assuming that this size distribution holds for another order of magnitude smaller in TNO size, we can scale our population estimates up to include TNOs at very small sizes (H < 18), and compare with the number of scattering TNOs that are required to supply the observed population of Jupiter Family Comets (JFCs). However, this close-in population has been measured to have slightly shallower slopes (Snodgrass et al. 2011; Fernández et al. 2013; Bauer et al. 2017) than the faint slope αf found in this analysis, so this may not be a valid assumption. With our preferred H-distribution, we find that the scattering population down to H < 18 should include 3 × 109 objects, which is a large enough supply to be the origin of the Jupiter Family Comets (Volk & Malhotra 2008).

4.2. The Size of the Centaur Population

The intrinsic Centaur population is about two orders of magnitude smaller than the intrinsic scattering TNO population, consistent with their shorter dynamical lifetime. In Table 2, we compare our Centaur population estimates with the population estimates of temporary Uranian and Neptunian co-orbitals in Alexandersen et al. (2016) and the abundance of these relative to a < 34 au scattering objects estimated in Alexandersen et al. (2013). Alexandersen et al. (2013) gives the fraction of the a < 34 au scattering population that must be trapped as temporary co-orbitals with Neptune and Uranus at any given time. The orbital distributions from Parker (2015) and Alexandersen et al. (2013) are combined with a knee H-distribution similar to the best fit of Fraser et al. (2014) and a divot distribution similar to the preferred H-distribution from Shankman et al. (2016) to calculate the population estimates in Table 2. The Centaur population estimates from our analysis are much smaller, but are not inconsistent when taking into account the (very large) error bars and upper limits from Alexandersen et al. (2013, 2016).

Another way we can make use of the Survey Simulator is to estimate how many relatively large Centaurs should exist based on our preferred H-distribution. Using this methodology, we find that the expected number of Hr < 6 Centaurs is ≤1 with 95% confidence. Reassuringly, the largest known Centaur, (10199) Chariklo, has an Hr magnitude of 6.82 ± 0.02 (assuming a linear spectrum and no phase correction; Peixinho et al. 2015).

5. Discussion

Although we are unable to formally reject either a knee or divot distribution, the power of forward-biasing combined with statistical analysis of the full OSSOS data set has vastly reduced the allowed parameter space compared to previous analyses (Shankman et al. 2013, 2016). But even with the earlier much smaller number of detections, this analysis technique is powerful. While the range of parameter space that was non-rejectable in Shankman et al. (2013) was many times larger than in our analysis here, the preferred divot from the analysis in Shankman et al. (2013) still provides a good agreement to the fit obtained here, even though that analysis only included 11 TNOs, while the analysis here contains over six times as many TNOs.

5.1. Knee or Divot?

This analysis has shown that a divot fits the data slightly better than a knee distribution, but knees cannot be rejected for several values of αf. For the break at larger TNO sizes, Hb = 7.7, αf of 0.4–0.5 and c from 1 (knee) to 5.6 are non-rejectable. For the Hb = 8.3 break, a slightly larger parameter space is non-rejectable, encompassing αf from 0.3 to 0.6, and c from 1 to 10.

Our preferred knee distribution has a slightly steeper slope (αf = 0.4) than the best-fit knee H-distribution found by the analysis of Fraser et al. (2014) (αf = 0.2). However, the 1σ uncertainties the published uncertainties on the Fraser et al. (2014) faint-end slope fit allow up to αf = 0.3, which is just inside the contour of non-rejectability (Figure 3, right panel), and is thus non-rejectable by our analysis.

5.2. Comparison with Other TNO Populations

This analysis is in broad agreement with the luminosity functions found for other dynamically hot populations in the Kuiper Belt.

Fraser et al. (2014) reports a αf slope of 0.36, with a break magnitude Hb = 8.4 for the Trojan asteroids, which is acceptable in our analysis and would thus allow a common H-distribution for the two populations. If the Kuiper Belt was emplaced by scattering off the giant planets during a period of instability (Morbidelli et al. 2005; Nesvorný et al. 2013), the Trojans would also be drawn from this population and should have the same size distribution (Morbidelli et al. 2009b). Determining whether or not the Trojans and dynamically hot Kuiper Belt populations share a size distribution is an important test of this model and is an area of active research (e.g., Wong & Brown 2015; Yoshida & Terai 2017).

While the number of detected Neptune Trojans is small, previous surveys have noted that there appears to be a lack of small members (Sheppard & Trujillo 2010; Parker 2015), which would be consistent with a divot in the size distribution.

The plutinos (TNOs in the 3:2 mean-motion resonance with Neptune) constitute the closest well-populated resonance, so studies are able to probe the size distribution down to smaller sizes than any other resonance. The well-characterized survey of Alexandersen et al. (2016) performs a similar analysis to this work and found that a break is required in the plutino size distribution, with a range of contrasts (including 1), break magnitudes, and faint-end slopes that match their H-distribution of plutino detections. Their preferred divot H-distribution is similar to the preferred divot of this work with a steeper faint-end slope: c = 6 and αf = 0.8 at Hb = 8.4 (though their non-rejectable parameter space covers a large range of αf and c values, see Figure 10 in Alexandersen et al. 2016). Knee distributions also provide a statistically acceptable match to their plutino detections, with their preferred fit exactly matching ours (αf = 0.4 at Hb = 7.7) and consistent with the best fit in Fraser et al. (2014).

Volk et al. (2016), which used detections only from the first two (of eight) OSSOS observing blocks, find no evidence in favor of a break in the size distribution, but show that this could be an effect of small number statistics. The analysis of the plutinos in the full OSSOS survey has several times more detected plutinos, and a transition is required in the H-distribution to match these observations; however, both a knee or divot transition provide reasonable matches to the data (Volk et al. 2017).

We note that previous analysis of the dynamically hot classical TNOs prefers a bright-end slope αb = 0.8 (95% confidence range 0.6–1.1, see Figure 5 in Petit et al. 2011), and ongoing analysis on the OSSOS discoveries indicates perhaps an even shallower slope provides a better fit to the larger TNOs (J.-M. Petit et al. 2018, in preparation). Our bright-end slope of αb = 0.9 is consistent with our data and with previous analysis of dynamically hot populations (e.g., Gladman et al. 2012), but as more relatively bright TNOs are discovered by current and future all-sky surveys (e.g., Holman et al. 2018), the best fit for the bright-end slope should be revisited.

5.3. Comparison with the Cratering Record

The distribution of craters on a planetary surface can be used to infer the distribution of impactor sizes if one understands the orbital distribution (and thus planetary impact speed distribution) of the projectiles. Due to its orbital inclination and Kozai oscillation while inside the 3:2 mean-motion, Pluto spends a large fraction of its time at latitudes above the dynamically cold classical belt, and its orbital eccentricity results in it spending little time passing through the cold classical Kuiper Belt (Greenstreet et al. 2015); the majority of its impacting projectiles thus come from dynamically hot populations, and it is therefore the dynamically hot population's size distribution that will be encoded in the crater counts.

Using imaging from the New Horizons spacecraft's Pluto encounter (Stern et al. 2015; Moore et al. 2016) and crater-rate production calculations (Greenstreet et al. 2015), the distribution of impactor sizes has been estimated to arise from an H-distribution with α ≃ 0.4 for projectiles with H = 13–19 (projectile diameters of 1–20 km; Singer et al. 2016). This H range just barely overlaps with our present analysis, which covers H ≃ 6–13, but this joint data set implies a roughly constant index α could extend from the break near D ∼ 100 km down to H ≃ 19 (D ≃ 1 km). If the faint-end slope we measure does indeed continue to D ≃ 1 km, this is additional support for the scattering disk being the sole source of the JFCs, as this assumption was made above (Section 4) to calculate the population size that was in agreement with this requirement. For even smaller objects, recent results of the Charon crater-field analysis indicate that the Kuiper Belt's α becomes even shallower (Singer et al. 2018), but sub-km TNOs are beyond the reach of ground-based and even space-based near-Earth telescopes.

6. Summary and Conclusions

This work is an exploration of the scattering TNO H-distribution with the full OSSOS sample, expanding on the analysis of Shankman et al. (2016) with a threefold larger set of detections (68 rather than 22 TNOs) and including fainter Hr magnitudes than previous work. We have demonstrated that existing models (H-distributions with either a divot or knee transition from bright- to faint-end slopes) provide acceptable matches for the H-distribution observed for scattering TNOs, but we have greatly constrained the allowed parameter space of possible faint slopes αf and contrasts of the transition. Our preferred H-distribution has a bright-end slope αb = 0.9, a faint slope αf = 0.5, and a divot of contrast c = 3.2, though a knee distribution with αf = 0.4 is also acceptable. The H-magnitude at the break is not important to our fit, and we find equally statistically acceptable H-distributions for Hb = 7.7 or 8.3, both of which were proposed by previous analyses. Large surveys such as Pan-STARRS and LSST will detect many new TNOs, especially at relatively bright H-magnitudes, and that will likely provide more statistical constraint on exactly where the break magnitude is, providing more information on the initial planetesimal formation size and collisional history of the Kuiper Belt.

We find that the shallower slope at faint magnitudes makes populations that are consistent with both the cratering record on Pluto and the population required to be the source of the Jupiter Family Comets.

A full exploration of possible size distributions would be best done in the context of a formation and evolutionary model of the solar system. The current degeneracy across potential break locations and divot or knee distributions may be addressed through additional constraints from formation theories. In order to explore this, one must understand the conditions under which accretion takes place, e.g., born big (Morbidelli et al. 2009a) or pebble accretion (Shannon et al. 2016), and must also understand the dynamical excitation process, e.g., whether Neptune's migration was smooth (Hahn & Malhotra 2005), grainy (Nesvorný & Vokrouhlický 2016), or chaotic (Tsiganis et al. 2005). By using these dynamical constraints, we can understand the process that emplaced the hot TNOs and shut off collisional grinding, leaving the Kuiper Belt with the size distribution we observe today.

The authors acknowledge the sacred nature of Maunakea and appreciate the opportunity to observe from the mountain. CFHT is operated by the National Research Council (NRC) of Canada, the Institute National des Sciences de l'Universe of the Centre National de la Recherche Scientifique (CNRS) of France, and the University of Hawaii, with OSSOS receiving additional access due to contributions from the Institute of Astronomy and Astrophysics, Academia Sinica, Taiwan. Data are produced and hosted at the Canadian Astronomy Data Centre; processing and analysis were performed using computing and storage capacity provided by the Canadian Advanced Network For Astronomy Research (CANFAR).

S.M.L. gratefully acknowledges support from the NRC-Canada Plaskett Fellowship. M.T.B. appreciates support from UK STFC grant ST/L000709/1. W.F. acknowledges support from Science and Technology Facilities Council grant ST/P0003094/1. This project was funded by the National Science and Engineering Research Council and the National Research Council of Canada.

Facility: CFHT (MegaPrime). -

Software: matplotlib (Hunter 2007), scipy (Jones et al. 2001).

Appendix:

The Appendix comprises Table 3.

Table 3.  Centaurs and Scattering TNOs in the OSSOS Survey Ensemble

Dynamical Survey r-band Hrb Distance at a e i MPC
Classa Name Magnitudeb   Discovery (au) (au)   (degrees) desig.c
sca o3e01 21.50 7.73 23.291 34.416173 0.589571 7.711 K02GG6G
sca o3e11 23.60 7.86 36.851 86.729341 0.609269 18.362 K13GD6Z
sca o3l01 23.06 10.89 16.046 55.817595 0.719066 22.246 K13U15R
sca o3l65 24.14 7.51 45.138 44.608588 0.277799 11.207 K13U16Z
sca o3o14 23.54 8.00 35.456 143.317456 0.754854 8.580 K13J64O
sca o3o16 23.92 8.34 35.680 57.383825 0.435939 13.701 K13J64P
sca o3o17 24.31 8.71 35.811 77.572262 0.540647 10.459 K13J64R
sca o3o36 23.73 6.09 57.342 49.020848 0.544507 34.879 K13J64Q
sca o4h03 22.69 9.55 20.758 49.901041 0.779420 5.679 K14UM9Q
sca o4h04 24.59 11.23 21.916 35.028185 0.376088 31.276 K14UM9A
sca o4h67PD 23.07 9.49 22.886 38.083254 0.654259 4.960 K06QI0P
sca o5c002 23.74 11.18 17.958 33.555021 0.524814 15.414
sca o5c022 23.68 8.30 34.284 71.897316 0.528607 5.612
sca o5c101 23.79 6.58 52.291 98.388020 0.646235 4.287
sca o5d002 24.95 10.36 28.844 41.040976 0.301831 34.818
sca o5d020 24.54 9.14 34.655 44.202910 0.278006 7.719
sca o5d025 24.19 8.60 36.217 68.621838 0.487903 2.105
sca o5d034 23.91 8.08 38.181 115.493325 0.777844 22.481
sca o5m03 23.94 12.85 12.879 89.174138 0.873805 38.666
sca o5m04 24.38 10.19 26.018 32.488890 0.225414 7.026
sca o5m52 24.27 8.12 41.057 680.202784 0.940468 13.994 K15KG3G
sca o5p009 24.07 9.20 30.845 184.132849 0.919622 53.315
sca o5p019 22.94 7.55 34.605 31.378013 0.302694 28.288
sca o5p021 24.71 9.27 35.180 45.967151 0.249269 11.745
sca o5p024 22.80 7.30 35.900 94.674918 0.629258 24.631
sca o5p025 22.66 7.08 36.250 100.871870 0.642654 4.771
sca o5p060 24.46 8.34 40.983 311.768577 0.876807 8.795 K15G50T
sca o5p146 24.09 6.47 57.872 85.613291 0.604676 14.247
sca o5s06 22.90 8.53 26.576 56.481339 0.531203 13.304 K15RO5W
sca o5s10 24.22 8.89 33.472 101.338298 0.687477 18.054
sca o5s11 24.54 9.14 33.969 50.814125 0.394573 15.159
sca o5s13 24.55 9.09 34.254 226.592608 0.861874 6.031 K15RO5Y
sca o5s20 24.04 8.24 37.139 42.894088 0.241075 6.932
sca o5t04 22.99 9.32 22.722 30.988803 0.289815 13.747 K15RO5U
sca o5t05 24.16 8.80 33.518 126.448249 0.735055 19.83
sca o5t06 24.20 8.79 33.933 72.064128 0.534457 12.327
sca o5t50 24.32 7.12 51.422 59.872018 0.688095 30.267
sca o5t52 24.13 6.10 62.394 425.861136 0.893065 12.138 K15RO5X
sca L3h08 23.59 7.66 38.445 159.681973 0.761413 15.500 K03H57B
sca L3q01 23.30 7.46 38.171 51.054204 0.484715 6.922 K03QB3W
sca L4k09 22.94 8.63 26.634 30.191945 0.185168 13.586 K04K18V
sca L4m01 23.05 8.05 31.360 33.467236 0.332719 8.205 K04M08W
sca L4p07 21.71 6.96 29.586 39.953648 0.280856 23.545 K04PB7Y
sca L4v04 23.44 8.39 31.848 64.100391 0.506381 13.642 K04VD1G
sca L4v11 23.49 9.24 26.757 60.035908 0.629283 11.972 K04VD1H
sca L4v15 21.77 8.21 22.950 68.385618 0.698262 14.032 K04VD1M
sca L7a03 23.14 6.41 46.991 59.613266 0.439491 4.575 K06BS4S
sca HL7j2 23.37 7.50 37.377 133.932936 0.725235 34.197 K07L38H
sca HL8a1 22.93 6.29 44.517 32.392864 0.374396 42.826 K08AD8U
sca HL8n1 23.73 8.52 31.849 41.531221 0.491379 103.447 K08K42V
sca HL9m1 21.13 9.57 12.872 348.905416 0.968470 68.016 K09M09S
cen o3l02 23.91 11.47 17.045 19.327805 0.127022 32.476 K13U17C
cen o3l03 24.39 10.25 25.336 25.872108 0.249698 8.515 K13U17U
cen o3o01 23.39 11.95 13.774 22.144387 0.378570 32.021 K13J64C
cen o4h01 22.74 10.29 17.756 23.195009 0.377843 21.319 K14UM5J
cen o4h02 24.33 11.47 19.526 27.954961 0.440821 12.242 K14UM9G
cen o5c001 23.72 11.75 15.857 28.529138 0.457119 36.539
cen o5d001 23.93 12.74 13.286 28.271438 0.542533 5.729
cen o5p001 24.05 13.40 12.029 12.048082 0.082638 24.112
cen o5p003 21.39 10.15 13.563 18.145145 0.269879 3.070
cen o5p004 23.92 12.68 13.563 20.995607 0.420656 1.628
cen o5p005 24.34 10.67 23.501 22.225868 0.257298 11.401
cen o5s04 24.51 13.11 13.441 20.915615 0.508346 10.109
cen o5s05 23.21 10.10 19.884 21.981271 0.479320 15.389 K15RO5V
cen o5t02 24.91 14.51 10.616 21.692667 0.519340 0.927
cen o5t03 23.27 10.48 18.515 25.967473 0.288012 18.849
cen mah01 24.45 10.86 22.432 30.072429 0.259122 53.886 K12UH7W
cen mal01 22.58 9.57 20.296 19.091885 0.176854 10.811 K11Q99F

Notes. All decimal places listed are significant. The full data set is available in Bannister et al. (2018).

aScattering TNOs are designated by "sca," Centaurs by "cen." These and all dynamical classifications within OSSOS use the classification scheme from Gladman et al. (2008). bAs noted in Section 3, all measurements have been transposed to r-band assuming g − r = 0.7. c https://www.minorplanetcenter.net/iau/info/PackedDes.html

Download table as:  ASCIITypeset images: 1 2

Footnotes

  • 10 

    Survey Simulator code and OSSOS ensemble survey pointings are publicly available at https://github.com/OSSOS/SurveySimulator, and properties of TNOs detected by the OSSOS ensemble are published in Bannister et al. (2018).

Please wait… references are loading.
10.3847/1538-3881/aab8ff