You are on page 1of 26

–1–

Redetermination of the space weathering rate


using spectra of Iannini asteroid family members
arXiv:0802.2977v1 [astro-ph] 21 Feb 2008

Mark Willman1 , Robert Jedicke1 , David Nesvorný2 ,


Nicholas Moskovitz1 , Željko Ivezić3 , Ronald Fevig4

1
Institute for Astronomy, University of Hawai‘i at Manoa
2680 Woodlawn Drive, Honolulu, HI 96822
willman@ifa.hawaii.edu, 808-956-6989 tel, 808-956-9580 fax
jedicke@ifa.hawaii.edu, 808-956-9841 tel, 808-956-9580 fax
nmosko@ifa.hawaii.edu, 808-956-6700 tel, 808-956-9580 fax

2
Department of Space Studies, Southwest Research Institute
1050 Walnut Street, Suite 400, Boulder, CO 80302
davidn@boulder.swri.edu, 303-546-0023 tel, 303-546-9687 fax

3
Department of Astronomy, University of Washington
Box 351580, Seattle, WA 98195-1580
ivezic@astro.washington.edu, 206-543-9375 tel, 206-685-0403 fax

4
Lunar & Planetary Laboratory, University of Arizona
1629 University Blvd., Tucson, AZ 85721-0092
fevig@lpl.arizona.edu, 520-621-2692 tel, 520-621-4933 fax

27 pages

10 figures

3 tables
–2–

Running Head: Space weathering rate using Iannini spectra

Editorial correspondence to:


Mark Willman
Institute for Astronomy University of Hawaii’i at Manoa
2680 Woodlawn Drive, Honolulu, HI 96822
willman@ifa.hawaii.edu
808-956-6989 tel
808-956-9580 fax
–3–

ABSTRACT

We obtained moderate S/N (∼ 85) spectra at a realized resolution of R ∼ 100 for 11 members
of the Iannini family, until recently the youngest known family at under 5 million years of
age (Nesvorný et al. 2003). The spectra were acquired using the Echellette Spectrograph
and Imager in its low-resolution prism mode on the Keck II telescope. The family members
belong to the S-complex of asteroids with perhaps some K class members. The Iannini
family members’s average spectral slope, defined as the slope of the best-fit line constrained
to pivot about 1 at 550 nm, is (0.30 ± 0.04)/µm, matching the (0.26 ± 0.03)/µm reported
by Jedicke et al. (2004) using SDSS (Ivezić et al. 2002) color photometry. Using our spectra
for this family as well as new observations of Karin family members (Vernazza et al. 2006)
and new classifications of some older families we revised the space weathering rate of S-
complex asteroids originally determined by Jedicke et al. (2004). Following Jedicke et al.
(2004) we parameterize the space weathering rate of the principal component color of the
spectrum (P C1), which is correlated with the spectral slope, as P C1 (t) = P C1 (0)+∆P C1 [1−
exp−(t/τ ) ]. Our revised rate suggests that the characteristic time scale for space weathering
α

is τ = 570 ± 220 Myr and that new S-complex clusters will have an initial color of P C1 (0) =
0.31 ± 0.04. The revised time scale is in better agreement with lab measurements and
our measurements support the use of space weathering as a dating method. Assuming
all the spectra should be identical, since members derived from the same parent body are
presumably covered with similar regolith, we combined them into a high-S/N composite
family spectrum which is within the S-complex.

Keywords: asteroids, surfaces


–4–

1. Introduction

The likely source of the most common meteorites (Krot et al. 2003), the ordinary chon-
drites, is the population of S-complex asteroids near the ν6 resonance in the inner main
asteroid belt. However, remote spectra of the S-complex asteroids are more red, have
a lower albedo and a shallower 1µm absorption band than lab spectra of the chondrites
(Chapman and Salisbury 1973). The attempt to resolve this discrepancy led to the sugges-
tion of a space weathering effect, the alteration of surface spectra over time due to bom-
bardment by particles and radiation (Hapke 2001; Pieters et al. 2000; Adams and McCord
1971).
We consider the existence of space weathering on asteroids to be well established
(Clark et al. 2002). On the other hand, the mechanism of space weathering, whether it is
due to solar or cosmic irradiation or micro-particle bombardment, remains unknown (Hapke
2001; Sasaki et al. 2001; Pieters et al. 2000) as does its effect on different types of asteroid
surfaces and the rate at which it acts. One method of distinguishing between the possible
sources of space weathering is to correlate the observed rate of color change on asteroids with
lab measurements of irradiated meteoritic asteroid proxies. Regardless of the physical cause
of space weathering, if it is possible to determine the rate of space weathering on asteroid sur-
faces, perhaps as a function of their taxonomic class and/or heliocentric distance, it may be
possible to use an asteroid’s spectra to estimate its age (Jedicke et al. 2004; Nesvorný et al.
2005).
The youngest family known when this study began (Nesvorný et al. 2005) was the Ian-
nini cluster dated between one and five Myrs old (Nesvorný et al. 2003). Nesvorný et al.
(2005) and Jedicke et al. (2004) reported a characteristic time for the space weathering
rate of 25 ± 5 Gyr derived from Sloan Digital Sky Survey photometry of asteroid colors
(Ivezić et al. 2002) (3rd Data Release) that is in dramatic disagreement with the rates ob-
served in lab experiments (Brunetto et al. 2006; Vernazza et al. 2006; Strazulla et al. 2005;
Marchi et al. 2005; Nakamura et al. 2001; Sasaki et al. 2001). The Nesvorný et al. (2005)
and Jedicke et al. (2004) results are highly dependent on the colors of the two youngest
families known at that time - the Iannini and Karin families - but the SDSS provided only
0/3 and 3/6 asteroids (N1 /N2 , where N1 is the number of objects with color errors < 0.1
and N2 is the number with color errors < 0.3) from each of those families for their color
measurement.
Our goal was to obtain spectra of many Iannini family members thereby measuring
the color of relatively fresh unweathered S-complex regolith and refining the rate of space
weathering. The expectation was that the youngest S complex families would have the bluest
color which is equivalent to a shallower spectral slope. The 49 known Iannini family members
–5–

all have diameters > 1 km and have orbits clustered around semi-major axis a = 2.64 AU,
eccentricity e = 0.30, and inclination i = 11.8◦ . The parent body size is estimated to be 14
km (Nesvorný et al. 2003). We also used spectra for 24 Karin family members obtained by
Vernazza et al. (2006). They have orbits clustered around a = 2.87 AU, e = 0.08, and i = 1◦
with an estimated parent body size of 33 km (Nesvorný et al. 2006c).
Laboratory simulations of space weathering have estimated the time scale of space
weathering in the main belt. Sasaki et al. (2001) used olivine and pyroxene as material
representing ordinary chondrite meteorites and exposed it to pulsed laser radiation to simu-
late the energy deposited by the space weathering agent. Marchi et al. (2006) have suggested
that the source of the agent is the Sun and that the amount of space weathering depends
on an object’s heliocentric distance. Using 729 spectral slopes from SMASS (Bus and Binzel
2002; Binzel et al. 2004) and 44 from Marchi et al. (2006) (also Paolicchi et al. (2007)), ages
determined from a size-age relationship, and solar flux determined from orbital parameters,
they claim to have identified a significant correlation between total solar exposure and color.
If correct, their work suggests that the primary mechanism of space weathering is either ion
bombardment from the solar wind or solar radiation as opposed to micrometeorite impacts
or cosmic rays.
Calibrating the rate of space weathering requires a collection of asteroids with known
ages. Asteroid families provide an opportunity to determine the surface age of all the mem-
bers since the collision took place essentially instantaneously and then ‘reset’ the surface
of each of the family members as they presumably accumulated fresh regolith in the after-
math of the catastrophic disruption that formed the family. Rather than displaying a range
of asteroid spectra as might be expected in the aftermath of the disuption of a differenti-
ated asteroid, family members presumably all tend to have homogeneous spectra. It is now
believed (Asphaug et al. 2002) that even small asteroids can retain a regolith layer so it is
presumed that virtually all asteroids within a family will exhibit essentially identical spectra.
There are already several dynamical methods of dating asteroid families (Nesvorný et al.
2005). The combination of these dynamical methods provides age estimates for about 20
families derived from S or C complex parent bodies. While there is a reasonable sample of
families dated at tens or hundreds of Myrs only a few families are known to be younger than
ten Myrs old. Prior to 2006 there were only two such S-complex families, the Karin and
Iannini clusters, but four small genetic groups of asteroids with ages under a million years
were identified in 2006 (Nesvorný et al. 2006a; Nesvorný and Vokrouhlický 2006b). Since
these rare young families provide all the leverage on the young end in the determination of
the space weathering rate they have influence beyond their numbers in that calculation.
In this work we present new spectra for 11 members of the Iannini family and use these
–6–

results along with other published data to reassess the space weathering rate as determined
by Nesvorný et al. (2005) and Jedicke et al. (2004). We also combine the spectra of these
extremely young asteroids to search for unusual spectral features that may be weathered
away on older asteroid surfaces.

Fig. 1.— The distribution of absolute magnitude for Iannini (solid) family members and
SMASS (Bus and Binzel 2002)(dotted) asteroids. The apparent magnitude at opposition
for asteroids in the Iannini family are about 3.2 magnitudes fainter than their absolute
magnitude.

2. Data acquisition and reduction

2.1. Spectra acquisition

Collisions resulting in catastophic disruptions occur more frequently on small bodies


than on large bodies (Bottke et al. 2004). Therefore, young families caused by recent col-
lisions are more likely to be composed of small members. One of the major problems in
–7–

achieving our goal of revising the space weathering rate was that the young Iannini asteroid
family is composed of small asteroids for which it is difficult to obtain spectra with telescope
apertures typically employed in planetary science. The Iannini family has 49 known mem-
bers with mean absolute magnitude H = 15.4, compared to 1447 SMASS (Bus and Binzel
2002) asteroids with mean H = 11.4. Figure 1 compares the overlapping absolute mag-
nitude distributions for these mutually exclusive populations. Since the SMASS data was
acquired on two-meter class telescopes, to achieve similar S/N in a reasonable time on objects
about 40 times fainter requires a ten-meter class telescope. Fortunately, at the University of
Hawaii’s Institute for Astronomy we were able to obtain time on the Keck II telescope with
its Echellette Spectrograph and Imager (ESI) in its low-resolution prism mode.
Observations of 11 members of the Iannini asteroid family with the 10-m Keck II tele-
scope were acquired between 10 November 2004 UT and 1 October 2006 UT as shown in
Table 1. The ESI, a 2048 × 4096 visual and near infrared pixel array (Sheinis et al. 2002), was
used in low-dispersion prism mode which provided a single order spectral range of 390−1100
nm in a single exposure. The 15 µm pixels provide a plate scale of 0.154 arcsec yielding a
raw spectral resolution (R) ranging from 6000 − 1000 across the spectrum.
These are apparently the first asteroid spectra obtained with this facility and our tech-
niques and procedures evolved rapidly over the seven nights of observations as we learned how
to use the telescope and instrument. In general, we followed standard asteroid spectra acqui-
sition procedures as employed and described by Fevig and Fink (2007) and Bus and Binzel
(2002). Similarly, our data reduction pipeline followed common spectra reduction proce-
dures and employed standard IRAF spectra analysis routines along with some customized
IDL programs to handle ESI-specific issues.
Bias frames, dome flats and arc calibration spectra were acquired at the beginning and
end of each night. We used a slit width of 1′′ or 1.25′′ for arc lamp spectra to obtain the
best wavelength calibration. Mercury-neon and xenon arc lamps were used for wavelength
calibration in the visible and near IR, respectively, with spectral lines ranging from 405−905
nm.
The realized PSF on the ESI detector was excellent with typical seeing in the range of
0.8 ≤FWHM≤ 1.0′′ . Typically three exposures of 900 seconds each were taken with adjust-
′′

ments for magnitude and weather conditions in order to obtain S/N > ∼ 40 for raw asteroid
spectra.
Solar analog stars were used to divide out the incident solar spectrum leaving only the
spectrum due to asteroid reflectance. The solar analogs were chosen from two sources, the
first being a list of 11 analogs recommended by Bobby Bus for which extensive compar-
–8–

Object Date Time Total Exposure H V S/N


(UT) (UT) (seconds)
2000 OZ7 2004 Nov 10 05:44-06:16 2320 14.0 17.9 57
1999 RE 2004 Nov 10 14:26-15:15 2700 15.5 20.6 31
2000 XZ42 2004 Nov 10 06:49-07:43 2700 15.9 18.0 144
2000 OH45 2006 Jan 04 10:35-11:38 3600 14.1 17.1 115
2000 OZ7 2006 Jan 04 09:06-10:10 3600 14.0 18.1 57
2000 RO76 2006 Jan 04 12:14-13:18 3600 15.2 18.2 73
2000 RZ81 2006 Jan 04 14:24-15:13 3600 15.1 19.1 59
2000 RZ81 2006 Jan 04 15:49-16:16 1200 15.1 19.1 59
2000 RA80 2006 Jan 05 11:40-13:20 5400 15.8 19.3 70
2000 RP4 2006 Jan 05 13:57-16:14 6300 14.8 19.5 87
2003 AF89 2006 Aug 23 14:47-15:19 1800 15.5 18.2 158
2003 AF89 2006 Oct 01 10:18-10:50 1800 15.5 17.3 158
2000 HU23 2007 Mar 09 06:21-07:11 2700 13.8 18.8 79
2001 QO282 2007 Mar 09 07:37-07:52 900 15.1 19.7 30

Table 1: Summary of observations of eleven Iannini family member. H refers to the object’s
absolute magnitude, V is apparent magnitude (as reported by JPL horizons), and the signal
to noise (S/N) is for the final binned spectra in 10 nm bins averaged over the range 440 − 920
nm. Note that some objects were acquired on multiple nights providing an opportunity to
cross check our data reduction procedure.

isons have been made against two widely accepted solar analogs (Hardorp 1978), 16 Cyg B
(HD186427, MV = 6.2, G5V ) and Hyades 64 (HD28099, MV = 8.1, G8V). These include
Hyades 64 itself, SA 93-101, SA 98-978, SA 102-1081, SA 105-56, SA 107-684, SA 107-998,
SA 110-361, SA 112-1333, SA 113-276, and SA 115-271. The second source was a list of
analogs provided by Fevig and Fink (personal communication).
Using analogs brighter than 10th magnitude with Keck caused the ESI camera to saturate
even for a one second exposure, so the individual component mirrors were fanned out in
primary mirror focus mode (PMFM); instead of all 36 mirrors bringing light to a common
focus the mirrors were misaligned to create a hexagonal pattern of identical images. We
then moved the telescope so that multiple images appeared on the slit resulting in multiple
identical spectra. While this procedure added some complication to the reduction procedure
it also had the benefit of providing multiple identical spectra suitable for co-adding to produce
very high S/N analog spectra.
–9–

Our choice of slit size evolved over the course of the program. Eventually we settled on
using a 1′′ slit for the arc lamp calibration spectra to get the sharpest emission lines and 6′′
for both asteroid and analog spectra. The 6′′ slit allowed easier centering and alignment for
the PMFM analog stellar spectra and consistency led to also using it for the asteroids. Any
systematic difference in the slit centers for arc lamp wavelengths were reduced to less than
0.1 nm by correcting the O2 A-band position.
Asteroid spectra were almost always obtained within a couple hours of crossing the
meridian and were never performed at an airmass > 1.3. Solar analogs at matching air mass
were taken before and after each set of asteroid spectra. For both the solar analogs and the
asteroids the slit was aligned with the parallactic angle (PA) to avoid spectral distortion due
to refraction of light at an angle to the slit. Near zenith this is less critical than at larger
air mass where refraction is stronger. During a sequence of observations we set the slit at a
fixed PA corresponding to the value for the object near the midpoint of the exposures.
Three previously-studied bright asteroids were observed as controls on our data reduc-
tion pipeline: 15 Eunomia and 33 Polyhymnia, both S complex, and 814 Tauris which is C
complex. Our reduced spectra for these obsects were in good agreement with previous work
(Bus and Binzel 2002).

2.2. Data reduction

The process of reducing the raw images to final spectra is summarized here:

• Raw images were bias subtracted, rotated to make the 2-dimensional spectra horizontal,
and vertically trimmed to a convenient size that included the spectrum and sufficient
area for background apertures.

• Emission lines in the 2-D arc lamp spectra were used to create a transform map to
straighten the 2-D asteroid and analog spectra. This is necessary because the raw
images from the ESI in low-dispersion mode produce strongly curved emission lines.

• Dome flats were median-combined into a column normalized master flat. i.e. the
average of all pixels in any column (at the same wavelength) in the master dome flat
was fixed at unity to correct for pixel-to-pixel variations in the quantum efficiency at
all points on the CCD at the same wavelength. This master flat was then divided into
the data frames.

• Since the ESI prism-mode spectra extends to about 1 µm we found it necessary to


correct for fringing at long wavelengths. We created a master fringe frame piecemeal
– 10 –

from spectrum-free sections of many spectra containing asteroids. The data frames
were divided by the master fringe frame to minimize fringing effects past 8500 A.

• All images were cleaned of cosmic rays before further processing.

• In some exposures we found that the object drifted off the slit, or perhaps cirrus drifted
through the field, such that the S/N for the target asteroid was dramatically lower in
those frames. Thus, the signal-to-noise on the source was estimated for each data frame
to determine if the data should be included in the final combined spectra.

• One dimensional solar analog spectra were then extracted from the two-dimensional
analog spectra using apertures 200 pixels high for both the analog and sky background.
1-D asteroid spectra were extracted from the 2-D spectra using apertures 19 pixels high
for both the asteroid and sky background spectra. The aperture used for the analog
spectrum extraction must be larger than that for the asteroid due to the multiple analog
spectra created by the PMFM operation of the Keck telescope. The ideal aperture
size of 19 pixels for asteroid spectra was determined experimentally, numerically, and
analytically; all three methods agreed to within 10%.

• Residual variation in the wavelength calibrations was minimized by measuring the


position of the observed O2 A-band and adjusting the reference position to place the
A-band at 762.5 nm.

• The 1-D spectra were normalized to 1 at 550 nm.

• Multiple 1-D spectra of the same asteroid were median combined into master asteroid
spectra to improve the S/N.

• Following Bus and Binzel (2002) we combined 1-D solar analogs into a master analog
frame. The individual master asteroid spectra were then divided by the master analog
spectrum.

• The resulting intrinsic asteroid spectrum (with solar spectrum removed) still showed
incomplete cancellation of the water absorption bands redward of 800 nm. This was
minimized by creating a water band model. Two solar analog spectra from the same
star showing the greatest difference in water band amplitudes were selected. Dividing
these images allowed us to enhance the water bands for that solar analog and combining
such cases from different stars produced the master water band spectrum. This was
then divided into the asteroid spectrum created in the last step. Since the strength of
the water bands in any individual spectrum may be different from the master water
band model we allowed the strength of the master to vary in order to minimize the
distortion due to water bands in the final spectrum.
– 11 –

• The O2 A-band at 762.5 nm and any residual cosmic rays were excised. The spectra
were binned into 10 nm bins in the manner of Bus and Binzel (2002) changing the
realized resolution dominated by binning to R∼ 44 − 92 over the wavelength range of
interest from 440 − 920 nm.

Figure 2 shows the final spectra of all 11 Iannini members obtained in this study. Note
the characteristic S-complex or K class spectra and note that many of the spectra show an
anomalous feature near 550nm.

Fig. 2.— Eleven Iannini spectra binned in 10 nm bins and shown over the wavelength
range of interest for slope calculations from 440 − 920 nm. Spectra are offset vertically for
comparison.
– 12 –

3. Results & Discussion

The 11 Iannini family member spectra share a close family resemblance as shown in Fig.
2. The family spectral similarity suggests that none of the observed asteroids are dynamical
interlopers but we need to establish the spectral class for this family before proceeding.
In the past, the determination of an asteroid’s taxonomic classification has been a sub-
jective procedure achieved by visual comparison of the object’s spectra to a set of standards.
To standardize and quantify this technique we have developed a software package that uses
the SMASS classes (Bus and Binzel 2002) as standards and calculates the difference between
each SMASS class and a comparison asteroid spectrum. Since the current era of asteroid
spectroscopy is witnessing a shift from filter photometry surveys towards full spectra surveys
the SMASS classes were chosen as the standard due to the large number of available asteroid
spectra (over 1400), the uniform techniques used for observation and reduction, and consis-
tency with the Tholen classification system. To classify an asteroid spectra we determine
the sX
spectral dif f erencec = N × (A(λk ) − Sc (λk ))2 (1)
k

where N is an arbitrary normalization factor, the sum is over reflectance values (A) for
the asteroid spectrum at discrete wavelengths (λk ), and Sc (λk ) is the interpolated SMASS
reflectance for the class (c) at the same wavelength. The spectral differences are ranked by
increasing difference and plotted against the ordinal ranking of the SMASS classes giving a
convenient visual and quantitative measure of class relationship. The code (smassclass.pro)
is available from the author (MW) by request.
The spectral differences for the combined spectrum of the 11 observed Iannini members
are shown in Figure 3. The class with the smallest difference is Sk. The second-ranked class
(Xc) has about 30% greater difference and is roughly tied with Sq, the third-ranked class.
As the spectral difference increases the range of possible classes related to the mean Iannini
spectrum also increases but we believe there is sufficient evidence at this time to assign this
family to the S-complex with Sk being the most likely type.
Since space weathering boosts the relative reflectance of silicate surfaces in the long
wavelength visual and near infrared range (Binzel et al. 2002) around the 1 µm absorption
band and tends to increase the slope of the region between that band and the UV band,
the combined Iannini spectrum of Fig. 4 is consistent with minimally weathered S-complex
surfaces as expected for such a young family.
– 13 –

Fig. 3.— The 26 SMASS classes are shown ranked in order of increasing spectral difference
with the combined spectrum of the 11 observed Ianninis. The calculation of the difference
(eq. 1) is described in the text. The ordinate gives the quantitative value of the difference
with class Sk being the most likely match.

3.1. Space weathering rate

To determine the rate of color change of S-complex asteroids due to space weathering
we require 1) a set of S-complex asteroids with known surface ages and 2) a consistent set of
color data for the same asteroids. Nesvorný et al. (2005) have determined the ages of many
asteroid families and used the early data release (Stoughten 2002) from the SDSS to provide
asteroid colors for each family. While one of our goals in obtaining spectra of the Iannini
family members was the re-determination of the space weathering rate we also wanted to
explore the possibility of other unexpected features in the spectra of the youngest asteroid
surfaces. Thus, we chose to obtain spectra rather than the SDSS ugriz bands for the Iannini
family members. This had the unfortunate consequence of making it more difficult for us
to directly compare our results with previous work on the space weathering rate due to
– 14 –

Fig. 4.— The 11 combined Iannini spectra over the range of interest for slope calculations
(440 − 920 nm). The best SMASS class match to the spectrum is of type Sk.

incomplete overlap of our spectral wavelength coverage with the SDSS z filter and almost no
overlap with the u filter.
Recognizing that different asteroid surface materials may weather at different rates it is
important to ensure that we compare asteroids of the same mineralogy when measuring the
space weathering rate. For instance, Nesvorný et al. (2005) have proposed that C complex
asteroid colors becoming bluer over time in the opposite sense to the S complex, . Thus,
we must selectively screen candidate families for a given class and it would be preferrable
to select only those asteroids/families that are within specific sub-types of a class (e.g. SIV
within the S-complex). However, given the current level of mineralogical understanding of
asteroid surfaces through their relatively coarse taxonomic classification, the best we have
been able to do is ensure that all 11 families used in this work are members of the S-complex.
Some of our choices for inclusion in this study may be disputed. For instance, it is unde-
– 15 –

cided whether family 606 Brangane should be considered K (Tholen 1984) or T, S, or D class
(Bus and Binzel 2002) while Bus and Binzel (2002) suggest that family 808 Merxia should
be in S-complex sub-type Sq. Resolution of these classification and mineralogical issues will
require better spectroscopic or in situ data enabling more detailed mineral identifications.
Infrared observations at 2µm could be particularly useful in distinguishing olivine from py-
roxene. Measurements of the space weathering rate along these lines would also benefit by
a much larger sample of asteroid families and also by more accurate determinations of their
ages. In the mean time, we use the available data and make the most consistent selections
possible.
In order to combine our results with those of Nesvorný et al. (2005) we note that their
principal component analysis of the SDSS asteroid colors found that the first two compo-
nents, P C1 and P C2, accounted for almost all the variation among the points in ugriz color
space. P C1 and P C2 in turn were strongly correlated with the spectral slope and curvature
respectively for asteroids in the SDSS database. Thus, by measuring the spectral slope for
the Iannini members we expected to determine the corresponding P C1 for comparison with
the Nesvorný et al. (2005) results.
We determined the spectral slope (s) for the Iannini members in the manner of SMASS
(Bus and Binzel 2002). The reduced asteroid spectra were binned into 10 nm segments and
a 3rd-order polynomial spline smoothing function was fit to each binned spectrum. We then
determined the spectral slope as that of a straight line fit to the smoothed spectrum while
constrained to equal unity at 550 nm.
One of the problems with this procedure is that it makes it difficult to determine the
error on the derived spectral slope. To this end we created synthetic spectra with a shape
specified by the 3rd-order polynomial spline smoothing function for each asteroid. Random
guassian noise with known S/N was added to the smoothed function and the spectral slope
determination procedure was repeated in order to determine the slope of the synthetic spec-
tra. In the real ESI spectra the S/N varies as a function of wavelength but this is a second
order effect in terms of determining the error on the slope. We defined the error on the slope
to be the RMS of the slope values in a narrow range centered on the actual S/N value for
the asteroid under consideration. In general, we found the error on the slope due to this
fitting method to be on the order of ±0.01.
There are at least two other significant sources of error in the slope determination for an
individual asteroid: natural slope differences due to hemispherical variations on the surface
and systematic analysis errors due to factors such as changing atmospheric conditions and
solar analog differences. We can estimate the level of systematic error due to these factors
by considering repeated observations on the same object. One such case yielded two slope
– 16 –

measurements differing by 0.05 suggesting that systematic error tends to overwhelm the
smaller measurement errors described above. The systematic error SMASS used was 0.07
which was derived from a much larger sample. We therefore conservatively adopted 0.07 for
the individual slope errors (statistical+systematic) for our analysis (shown in Fig. 5).
The resulting average slope for the Iannini family members is 0.30 ± 0.14(RMS)/µm
while the standard error on the mean of the average slope is ±0.04/µm.

Fig. 5.— The relationship between SMASS-style spectral slopes and P C1 (as defined by
Nesvorný et al. (2005)) for 133 asteroids common to SDSS and SMASS with ∆P C1 ≤ 0.09.
The slope errors on each data point are fixed at our adopted value of 0.07 (see text for
details). The values and errors on PC1 are calculated directly from the SDSS data according
to the formula of Nesvorný et al. (2005).

The correlation between spectral slope and P C1 identified by Nesvorný et al. (2005)
was determined by comparing their P C1 to the slope of a line fit to the reflectance in each
of the five ugriz bands. Thus, in order to combine our results with Nesvorný et al. (2005)
we required a conversion between our calculated SMASS-style (Bus and Binzel 2002) slopes
– 17 –

derived from continuous spectra (which did not include the u band) to the Nesvorný et al.
(2005) style P C1.
The SDSS 3rd Data Release and SMASS have 141 asteroids in common and Fig. 5
shows P C1 (Nesvorný et al. (2005)) vs. our spectral slope (s) for a sub-sample of 133 of
these objects. Since our desire was to obtain the best possible determination of the slope
(m) of the P C1 -s line we selected only those SDSS objects with δP C1 < δP C1,max where
δP C1,max was chosen to minimize the derived error on the slope, δm. For large δP C1,max
decreasing δP C1,max decreases δm while simultaneously decreasing the sample size. However,
as the sample size decreases eventually the smaller statistics cause δm to increase. We found
that δP C1,max = 0.09 minimized δm. The transformation between our Iannini spectral slopes
(s[/µm]) and the P C1 as determined by Nesvorný et al. (2005) is given by
P C1 = (0.87 ± 0.02) × s + (0.082 ± 0.012) (2)
Finally, the mean Iannini family P C1 = 0.35 ± 0.04 (using the standard error on the mean
slope).
Our value for the P C1 of the Iannini family is in surprising agreement with the value of
P C1 = 0.33 ± 0.05 from Nesvorný et al. (2005) based on SDSS colors. The surprise is that
the SDSS result derives from only three objects with a large scatter in their colors while our
mean color derives from spectra of almost four times as many objects.
Now that we have established a new mean P C1 value for the Iannini family we re-
examine the colors and inclusion of other S-type families in the space weathering rate determi-
nation. In particular, the Karin family is the only other cluster used to anchor the young end
of the P C1 − age relationship. Vernazza et al. (2006) provided us recent spectral data from
their Karin family survey over the full SMASS wavelength range. We have determined the
spectral slope of each of the asteroids in the same manner as described above and find a mean
spectral slope for the Karin family of 0.28 ± 0.03/µm corresponding to P C1 = 0.32 ± 0.02,
statistically indistinguishable from the photometric value of P C1 = 0.39 ± 0.08 determined
by Nesvorný et al. (2005), although now having better precision.
Our new value for the mean P C1 for the Iannini family, the new value derived for the
Karin family spectra of Vernazza et al. (2006), and the previously known values for older
families with the exception of Eos are shown in Fig. 6. Eos has been excluded due to its
reclassification as K class by Bus and Binzel (2002). The data continue to support the idea
that an age-related color dependence is evident for the S-complex asteroids.
We envision the space weathering process as converting unweathered surface (U) to
weathered surface through the action of some as-yet undefined agent. For the purpose of
this work it is unimportant how the agent operates (e.g. solar or cosmic irradiation, heating
– 18 –

Fig. 6.— First principal color component (P C1) vs age for the S-complex asteroid families
as first shown by Nesvorný et al. (2005) with the following modifications: our new Iannini
P C1 value, a new Karin P C1 derived from Vernazza et al. (2006) and the removal of the Eos
family now considered to be K class. P C1 errors are the standard error on the mean for the
family. The new fit (dashed) to the space weathering function (see text for details) for our
data is shown along with the fit (dotted) originally reported by Nesvorný et al. (2005).

by micro-meteoroid bombardment, etc.). At the same time, an asteroid’s surface will be


‘gardened’ by the impact of larger meteoroids that bring deeper, unweathered regolith to
the surface. Over a large area and long period of time we expect that the rate of both space
weathering and gardening is nearly constant. Space weathering should be proportional to
the amount of unweathered surface: −dU/dt ∝ U, i.e. as a decaying exponential. Gardening
should be proportional to the amount of weathered surface: −dW/dt ∝ W , where W = 1−U.
This model is independent of the particular physical causes of weathering and gardening,
depending only on the assumption of uniform rates.
The above assumptions are mathematically equivalent (except for the introduction of
– 19 –

a generalizing exponent, α) to parameterizing the color of the surface as a function of time


(Jedicke et al. 2004) as:

P C1 (t) = P C1 (0) + ∆P C1 (1 − exp[−(t/τ )α ]) (3)

where P C1 (0) is the unweathered color of fresh asteroid surface material later identified as
S-complex, ∆P C1 is the magnitude of the weathering color change after a long period of
time, and τ is the characteristic time for space weathering. In this model τ is a combination
of the space weathering (τw ) and gardening (τg ) timescales with the shorter time having the
dominant influence according to:
τw τg
τ= . (4)
τw + τg
Disentangling the two characteristic times (and other possible effects) remains a matter for
further research. The exponent α generalizes the function but we would expect a priori that
α = 1.

Data P C1 (0)(/µm) ∆P C1 (/µm) τ α


Jedicke et al. (2004) 0.32 ± 0.01 1.0 ± 0.1 25 ± 5 Gyr 0.50 ± 0.05
This work 0.31 ± 0.04 0.31 ± 0.07 570 ± 220 Myr 0.53 ± 0.19

Table 2: Space weathering rate parameters. P C1 (0) is the P C1 value for new S-complex
regolith, ∆P C1 is the increment in P C1 after infinite time, τ is the timescale for the weath-
ering+gardening process, and α is a generalizing exponent.

The Jedicke et al. (2004) parameterization and our result are both shown in Fig. 6 and
the fit parameters are provided in Table 2. We empahsize that these parameters are model
dependent and although we believe the model is reasonable it is not the only possible one.
The characteristic time τ for the process is 570 ± 220 Myr - a dramatic decrease from the
value of 25, 000 ± 5, 000 Myr reported by Jedicke et al. (2004). The difference results from a
combination of changes in the data set and different solution constraints. Recall that in this
analysis we removed the Eos family from consideration and that we adopt updated colors for
the two young Iannini and Karin families. Furthermore, we believe that the algorithm we
used to identify the best fit solution is superior to that of Jedicke et al. (2004) who fixed two
of the parameters in Eq. 3, P C1(0) and ∆P C1 , at what seemed like reasonable values and
allowed the remaining parameters to vary. We explored the entire phase space of possible
solutions to identify a region near the global minimum in χ2 . The phase space is complicated
by the presence of multiple local minima and the fact that τ and α are tightly coupled. We
– 20 –

then performed a full four parameter fit within a restricted region near the global minimum
that resulted in the parameters provided in Table 2.
One possible problem with our analysis, and also for Jedicke et al. (2004) and Nesvorný et al.
(2005), is that we have not corrected for the fact that families receive different amount of
solar flux in the same period of time due to their different orbits. While this work has not
established the origin of the agent responsible for space weathering, Marchi et al. (2006)
suggest an asteroid color dependence on heliocentric distance implying a solar origin for this
effect. If the Sun is actually responsible for the differences in S-complex asteroid colors pre-
sented here then we can correct for the solar exposure received by each family. We calculate
a ‘normalized solar exposure age’ corresponding to the time the family would have to be
exposed to solar radiation on an Earth-like orbit (a = 1, e = 0) to equal the amount of
flux they received at their actual (a,e). These adjusted ages give a solution with larger χ2
and parameter errors than without adjustment for solar flux. Since there are only eleven S-
complex families in our analysis with semi-major axes distributed over the relatively narrow
2.41 − 2.87 AU range our data may not be an appropriate sample for testing this hypothesis.
The weathering rate presented here is in good agreement with lab measurements and
other space weathering rate estimates. Sasaki et al. (2001) used a pulsed laser on olivine
and pyroxene pellets to simulate the effects of space weathering and measured a charac-
teristic time scale of about 100 Myr (at 1 AU, increasing by a factor of seven at Iannini’s
a = 2.64 AU) for the color change due to the production of nano-scale iron production in
micro-heating events. They suggest this may be a lower limit because “... the timescale
to reach space weathering maturity may depend more on the timescale for developing an
equilibrium regolith than for just changing colors of inert materials” (Chapman 2004). A
similar timescale of 100-800 Myr was proposed by Pieters et al. (1994) who used the geology
of dated lunar craters to estimate the optical maturity of the moon’s regolith though it may
not be appropriate to compare the weathering rate on the material and distance of the moon
to that of S-complex asteroids in the main belt. An anomalously short timescale for space
weathering color changes of only 50 Kyr was determined by Hapke (2001) based on irradi-
ating silicate powders in the lab. They suggested regolith gardening would make the actual
time much longer but we cannot speculate on whether it could account for the four orders
of magnitude difference between their measurement and our ∼570 Myr time scale. If the
Hapke (2001) space weathering rate is correct then it is difficult to explain how there could
be Q class asteroids (thought to be unweathered S-complex objects) in the NEA population
since their dynamical lifetime is on the order of 10 Myr (Gladman et al. 2000). There may
also be different mechanisms at work with different timescales (Jedicke et al. 2004).
One caveat to our measured space weathering rate is suggested by the recent work by
– 21 –

Vokrouhlický et al. (2006) showing a systematic overestimate of age due to the initial velocity
field produced by the collision creating the family. This may account for ∼ 30 − 50% of the
semimajor axis dispersion for families between 0.1 and 1 Gyr old. For instance the Massalia
and Merxia families are estimated to have corrected ages making them approximately 1/3
younger. Correcting for this effect would lead to a reduced characteristic time.
The initial color that we find for unweathered S-complex surface regolith, P C1 (0) =
0.31 ± 0.04, is identical within error bars to the value provided by (Nesvorný et al. 2005).
However, ∆P C1 = 0.31 ± 0.07, the amount of color change expected over a very long period
of time due to space weathering, is much smaller than the Nesvorný et al. (2005) value of
1.0 ± 0.1. Since their results were obtained by an ad hoc guess at ∆P C1 while our value is
determined by a fit to the function and because laboratory measurements suggest that color
saturation with increasing irradiation is approached on a time-scale of a few hundred million
years (Sasaki et al. (2001)) we propose that our solution is more robust and consistent with
lab measurements. On the other hand, like Nesvorný et al. (2005), we are unable to explain
why our value of the exponent α = 0.53 ± 0.19 is much different from unity as expected from
simple physical arguments.
Our results suggest that all young S-complex asteroid families should have P C1 essen-
tially identical to our revised Iannini and Karin values: P C1(0) = 0.31 ± 0.04. Binzel et al.
(2004) suggest that the close match between spectra of Q class asteroids and ordinary chon-
drite meteorites makes the Q class the starting point of a continuum of spectral shapes
that proceeds from Q to S over time. The Q class mean slope of 0.19 ± 0.03 translates to
P C1 = 0.25 ± 0.04 which is compatible with our model (P C1 (0)).
Of course, it is possible that different modes of space weathering act at different rates.
There is as yet no reason to believe that all the known space weathering effects (reddening
of the continuum, decreasing albedo, and reduced depth of the 1µm feature) occur at the
same rate. Thus, it would be interesting to examine the spectra or colors of other young
S-complex families. Recently, clusters even younger than the Iannini family (1 − 5 Myr)
have been discovered (Nesvorný et al. 2006a; Nesvorný and Vokrouhlický 2006b) including
the Datura (∼450 kyr), Lucascavin (300 − 800 kyr), Emilkowalski (∼220 kyr) and 1992
YC2 (50 − 250 kyr) clusters. We plan on obtaining spectra or colors of all these families to
determine their taxonomy and then target those S-complex clusters to test our prediction
for the color of young S-complex surfaces.
– 22 –

4. Conclusions

We have obtained spectra for 11 members of the Iannini family, till very recently the
youngest known family at under 5 million years of age. The combined Iannini spectrum
is calculated by a new difference algorithm to be in the S-complex with Sk the single best
matching SMASS class. The Iannini family members’s average spectral slope is 0.30±0.04µm
corresponding to a first principal component color of P C1 = 0.35 ± 0.04 matching that
reported by Jedicke et al. (2004) using SDSS (Ivezić et al. 2002) color photometry. Using our
spectra for this family as well as new observations of Karin family members (Vernazza et al.
2006) we have revised the space weathering rate of S-complex asteroids originally determined
by Jedicke et al. (2004). Our calculated time scale for space weathering is τ = 570±220 Myr
and we predict that new S-complex clusters will have an initial color of P C1 (0) = 0.31±0.04.
The revised time scale is in much better agreement with lab measurements and supports the
use of space weathering as an asteroid surface dating method.

5. Acknowledgments

MW and RJ acknowledge useful collaboration and discussion with Bobby Bus on re-
duction techniques, Greg Wirth on instrument issues, Ron Fevig on data reduction, Dale
Kocevski, Bin Yang and Hai Fu on IRAF procedures. Pierre Vernazza kindly provided the
full-spectrum Karin family data for slope calculations and constructive suggestions during
the review process. The data presented herein were obtained at the W.M. Keck Observatory,
which is operated as a scientific partnership among the California Institute of Technology,
the University of California and the National Aeronautics and Space Administration. The
Observatory was made possible by the generous financial support of the W.M. Keck Foun-
dation. We recognize the cultural role and reverence that the summit of Mauna Kea has
within the indigenous Hawaiian community and appreciate the opportunity to observe from
this mountain. This work is supported under NSF grant AST04-07134.

REFERENCES

Adams, J.B., McCord, T.B., 1971. Alteration of lunar optical properties: age and composi-
tion effects. Science, 171, 567-571.

Asphaug, E., Ryan, E.V., Zuber, M.T., 2002. Asteroid interiors. In: Bottke, W.F. Cellino,
A., Paolicchi, P. Binzel, R.P., Editors, 2002. Asteroids III, Univ. of Arizona Press,
Tucson, AZ, pp. 463-484.
– 23 –

Binzel, R.P., Dmitrij, F.L.,Martino, M.D., Whiteley, R.J., Hahn, G.J., 2002. Physical prop-
erties of near-Earth objects. In: Bottke, W.F., Cellino, A., Paolicchi, P., Binzel, R.P.
(Eds.), Asteroids III, Univ. of Arizona Press, Tucson, pp. 255-271.

Binzel, R.P., Rivkin, A.S., Stuart, J.S., Harris, A.W., Bus, S.J., Burbine, T.H., 2004. Ob-
served spectral properties of near-Earth objects: results for population distribution,
source regions, and space weathering processes. Icarus 170, 259-294.

Bottke, W.F., Durda, D.D., Nesvorný, D., Jedicke, R., Morbidelli, A., Vokrouhlický, D.,
Levison, H., 2004. The fossilized size distribution of the main asteroid belt. Icarus
175, 1, 111-140.

Brunetto, R., Vernazza, P., Marchi, S., Birlan, M., Fulchignoni, M., Orofino, V., Strazzulla,
G., 2006. Modelling asteorid surfaces from observations and irradiation experiments:
the case of 832 Karin. Icarus 184, 327-337.

Bus, S.J., Binzel, R.P., 2002. Phase II of the Small Main-Belt Asteroid Spectroscopic Survey,
The Observations. Icarus 158, 106-145.

Bus, S.J., Binzel, R.P., 2002. Phase II of the Small Main-Belt Asteroid Spectroscopic Survey,
A Feature-Based Taxonomy. Icarus 158, 146-177.

Cellino, A., Zappala, V., Doressoundiram, A., Di Martino, M., Bendjoy, Ph., Dotto, E.
Migliorini, F., 2001. The Puzzling Case of the Nysa-Polana Family. Icarus 152, 225-
237.

Chapman, C.R., 2004. Space Weathering of Asteroid Surfaces, Annu. Rev. Earth Planet.
Sci., 32, 550-551.

Chapman C.R., Salisbury, J.W., 1973. Comparison of meteorite and asteroid spectral reflec-
tivities. Icarus 19, 507-522.

Chapman, C.R., Enke, B., Merline, W.J., Nesvorný, D., Tamblyn,.P., Young, E.F., Olkin,
C., 2007. Young Asteroid 832 Karin Shows no rotational spectral variations. Icarus
in press.

Clark B. F., Hapke B., Pieters C., Britt D., 2002. Asteroid Space Wearthing and Regolith
Evolution. In: Bottke, W.F., Cellino, A., Paolicchi, P., Binzel, R.P. (Eds.), Asteroids
III, Univ. of Arizona Press, Tucson, pp. 585-599.

Cloutis, E.A., 1994. Brown University Keck/NASA Relab Spectra Catalog,


http://lf314-rlds.geo.brown.edu/.
– 24 –

Fevig, R., Fink, U., 2007. Spectral observations of 19 weathered and 23 fresh NEAs and their
correlations with orbital parameters. Icarus 188, 175-188.

Gladman, B., Michel, P., Froeschle, C., 2000. The Near-Earth Object Population. Icarus
146, 176-189.

Hapke, B., 2000. Space weathering in the asteroid belt. Lunar and Planetary Science, 31,
1087.

Hardorp, J., 1978. The Sun among the Stars. A&A 63, 383-390.

Hinrichs, J.,Lucey, P., 2002. Temperature-Dependent Near-Infrared Spectral Properties of


Minerals, Meteorites, and Lunar Soil. Icarus 155, 169-180.

Hiroi, T., Vilas, F., Sunshine, J., 1996. Discovery and Analysis of Minor Absorption Bands
in S-Asteroid Visible Reflectance Spectra. Icarus 119, 202-208.

Ivezić, Ž., Jurić, M., Lupton, R. H., Tabachnik, S., Quinn, T., 2002. Asteroids Observed by
The Sloan Digital Survey. In: Survey and Other Telescope Technologies and Discov-
eries, Tyson, J.A., Wolff, S. (Eds.), Proc. SPIE, 4836, pp. 98-103.

JPL Small-Body Database, 2006. http://ssd.jpl.nasa.gov/sbdb.cgi.

Jedicke, R., Nesvorný, D., Whiteley, R.J., Ivezić, Ž., Jurić, M., 2004. An age-colour relation-
ship for main-belt S-complex asteroids. Nature, 429, 275-277.

Krot, A.N., Keil, K., Goodrich, C.A., Scott, E.R.D., Weisberg, M.K., 2003. Classification
of Meteorites. In: Davis, A.M., Holland, H.D., Turekian, K.K., (Eds.), Meteorites,
Comets and Planets, Vol. 1, Treatise on Geochemistry, Elsevier, Oxford, pp. 86-116.

Marchi, S., Brunetto, R., Magrin, S., Lazzarin, M., Gandolfi, D., 2006. Space weathering
of near-Earth and main belt silicate-rich asteroids: observations and ion irradiation
experiments. A&A, 443, 769-775.

Marchi, S., Paolicchi, P., Lazzarin, M., Magrin, S., 2006. A general spectral slope-exposure
relation for S-type main belt and near-earth asteroids. AJ, 131, 1138-1141.

Nakamura, K., Sasaki, S., Hamabe, Y., Kurahashi, E., Hiroi, T., 2001. Laboratory simulation
of space weathering: A transmission electron microscopic study - microstructures of
the laser irradiated samples. Lunar and Planetary Science Conference XXXII abstract
#1547.
– 25 –

Nesvorný, D., Bottke, W.F., Dones, L., Levison, H.F., 2002. The recent breakup of an
asteroid in the main-belt region. Nature 417, 6890, 720-771.

Nesvorný, D., Bottke, W.F., Levison, H.F., Dones, L. 2003. Recent origin of the solar system
dust bands. ApJ 591, 486-497. 720-771.

Nesvorný, D., Jedicke, R., Whiteley, R.J., Ivezić, Ž., 2005. Evidence for asteroid space
weathering from the Sloan Digital Sky Survey. Icarus 173, 132-152.

Nesvorný, D., Vokrouhlický, D., Bottke, W.F., 2006. The Breakup of a Main-Belt Asteroid
450 Thousand Years Ago. Science, 312, 1490.

Nesvorný, D., Vokrouhlický, D., 2006. New Candidates for Recent Asteroid Breakups. AJ,
132, 1950-1958.

Nesvorný, D., Enke, B., Bottke, W.F., Durda, D., Asphaug, E., Richardson, D., 2006. Karin
cluster formation by asteroid impact. Icarus, 183, 296-311.

Paolicchi, P., Marchi, S., Nesvorny, D., Magrin, S., Lazzarin, M., 2007. Towards a general
model of space weathering of S-complex asteroids and ordinary chondrites. A&A 464,
1139-1146.

Pieters, C.M., Staid, M.I., Fischer, E.M., Tompkins, S., He, G., 1994. A Sharper View of
Impact Craters from Clementine Data. Science 266, 5192, 1844-1848.

Pieters, C.M., Taylor, L.A., Noble, S.K., Lindsay, P.K., Hapke, B., Morris, R.V., Allen, C.C.,
McKay, D.S., Wentworth, S., 2000. Space weathering on airless bodies: Resolving a
mystery with lunar samples. Meteorit. Planet. Sci. 35, 1101-1107.

Sasaki, S., Nakamura, K., Hamabe, Y., Kurahashi, E., Hiroi, T., 2001. Production of Iron
Nanoparticles by Laser Irradiation in a Simulation of Lunar-like Space Weathering.
Nature, 410, 555-557.

Sheinis, A.I., Bolte, M., Epps, H.W., Kibrick, R.I., Miller, J.S., Radovan, M.V., Bigelow,
B.C., Sutin, B.M., 2002. ESI, A New Keck Observatory Echellette Spectrograph and
Imager. PASP, 114, 851-865.

Stoughten, C., 191 colleagues, 2002. Sloan Digital Sky Survey: early data release. AJ 123,
485-548.

Strazzulla, G., Dotto, E., Binzel, R., Brunetto, R., Barucci, M.A., Blanco, A., Orofino, V.,
2005. Spectral alteration of the meteorite Epinal (H5) induced by heavy ion irradi-
ation: A simulation of space weathering effects on near-Earth asteroids, Icarus 174,
31-35.
– 26 –

Tholen, D.J., 1984. Ph.D. Dissertation. Univ. of Arizona, p. 95.

Vernazza, P., Birlan, M., Rossi, A., Dotto, E., Nesvorný, D, Brunetto, R., Fornasier, S.,
Fulchignoni, M., Renner, S., 2006. Physical Characterization of the Karin Family.
A&A 460, 945-951.

Vokrouhlický, D., Broz, M., Bottke, W.F., Nesvorný, D., Morbidelli, A., 2006.
Yarkovsky/YORP chronology of asteroid families. Icarus 182, 118-142.

This preprint was prepared with the AAS LATEX macros v5.2.

You might also like