Newer views of the Moon: Comparing spectra from Clementine and the Moon Mineralogy Mapper

Similar documents
Optical maturity variation in lunar spectra as measured by Moon Mineralogy Mapper data

DISCOVERY OF LUNAR OLIVINE WITH OH/H 2 O BAND AT MOSCOVIENSE BASIN USING HYPERSPECTRAL IMAGING SPECTROMETER DATA (M 3 )

Development, importance, and effect of a ground truth correction for the Moon Mineralogy Mapper reflectance data set

Thermal removal from near infrared imaging spectroscopy data of the Moon

Photometric modeling of the Moon using Lommel-Seeliger function and Chang E-1 IIM data

Iron and Titanium: Important Elements. posted October 20, References:

Mineralogical and Geomorphological Mapping of western central part of Mare Tranquillitatis using Hyperspectral Imager onboard Chandrayaan-1

Igneous activity in the southern highlands of the Moon

EFFECTS OF LASER SPACE WEATHERING ON DERIVED IRON OXIDE CONTENT IN SAN CARLOS OLIVINE, PYROXENE, AND ANORTHOSITE

Detection of Adsorbed Water and Hydroxyl on the Moon

Character and Spatial Distribution of OH/H 2 O on the Surface of the Moon Seen by M 3 on Chandrayaan-1

Using the modified Gaussian model to extract quantitative data from lunar soils

UV-V-NIR Reflectance Spectroscopy

Remote compositional analysis of lunar olivine rich lithologies with Moon Mineralogy Mapper (M 3 ) spectra

Statistical Analysis of the Links among Lunar Mare Soil Mineralogy, Chemistry, and Reflectance Spectra

Lunar soil characterization consortium analyses: Pyroxene and maturity estimates derived from Clementine image data

Lunar opposition effect as inferred from Chandrayaan-1 M 3 data

Thermal, Thermophysical, and Compositional Properties of the Moon Revealed by the Diviner Lunar Radiometer

STUDIES of the lunar regolith layer might provide the physical

Supporting Online Material for

Figure S1. CRISM maps of modeled mineralogy projected over CTX imagery (same

Composition of the Moon's Crust

Highland contamination in lunar mare soils: Improved mapping with multiple end-member spectral mixture analysis (MESMA)

PUBLICATIONS. Journal of Geophysical Research: Planets. Volcanic glass signatures in spectroscopic survey of newly proposed lunar pyroclastic deposits

SCIENCE CHINA Physics, Mechanics & Astronomy. The improvement of cross-calibration of IIM data and band selection for FeO inversion

Remote sensing and geologic studies of the Balmer-Kapteyn region of the Moon

providing 100-m per pixel resolution in nine ~1.0 µm wide infrared bands centered from

Constellation Program Office Tier 2 Regions of Interest for Lunar Reconnaissance Orbiter Camera (LROC) Imaging

The Surprising Lunar Maria

Science Return from Hayabusa

The absorption-peak map of Mare Serenitatis obtained by a hyper-spectral telescope

Global spatial deconvolution of Lunar Prospector Th abundances

Icarus 210 (2010) Contents lists available at ScienceDirect. Icarus. journal homepage:

Background Image: SPA Basin Interior; LRO WAC, NASA/GSFC/ASU

Compositional diversity and geologic insights of the Aristarchus crater from Moon Mineralogy Mapper data

The Importance of Impact Melts

Constellation Program Office Tier 1 Regions of Interest for Lunar Reconnaissance Orbiter Camera (LROC) Imaging

The Distribution of Mg- Spinel across the Moon and Constraints on Crustal Origin

Orbital Identification of Carbonate-Bearing Rocks on Mars

HYPERSPECTRAL IMAGING

EXERCISE 2 (16 POINTS): LUNAR EVOLUTION & APOLLO EXPLORATION

REVISED COORDINATES FOR APOLLO HARDWARE

Exploring the Lunar Surface

Cross-calibration of Geostationary Satellite Visible-channel Imagers Using the Moon as a Common Reference

Lunar mare volcanism in the eastern nearside region derived from Clementine UV/VIS data

Supporting Online Material for

Rilles Lunar Rilles are long, narrow, depressions formed by lava flows, resembling channels.

The Moon's Dark, Icy Poles

Photometric anomalies of the lunar surface: Results from Clementine data

Scattered light correction of Hayabusa/AMICA data and quantitative spectral comparisons of Itokawa

FLAT FIELDS FROM THE MOONLIT EARTH

AN UNLISTED DOME NEAR CRATER ARAGO located at longitude and latitude +7.66

Lunar Surface Material Composition Mapping

FANTASTIC!! MARINER VENUS / MERCURY 1973 STATUS BULLETIN BULLETIN NO. 27

Signature: Name: Banner ID#:

arxiv: v1 [astro-ph] 16 Aug 2008

Radiometric Calibration Stability and Inter-calibration of Solar-band Instruments in Orbit Using the Moon

UNIVERSITY COLLEGE LONDON

Supplementary Figure 1 Panoramic view of four sites (CE-0005, CE-0006, CE-0007, and CE-0008) measured by APXS and VNIS. Images (a), (b), and (d) were

The C1XS Lunar X-ray Spectrometer for Chandrayaan-1

C1XS - The Chandrayaan-1 X-ray Spectrometer

(Received 21 January 2015; revision accepted 23 October 2015)

Moon 101. Bellaire High School Team: Rachel Fisher, Clint Wu, Omkar Joshi

Mini-RF: An Imaging Radar for the Moon. Ben Bussey The Johns Hopkins University Applied Physics Laboratory

Lunar Poles. Status of Understanding a Potential Resource. Paul G. Lucey

Hyperspectral Atmospheric Correction

HICO Science Mission Overview

Spitzer Space Telescope

Chapter 4 Nadir looking UV measurement. Part-I: Theory and algorithm

The Moon. Tidal Coupling Surface Features Impact Cratering Moon Rocks History and Origin of the Moon

Olivine-Pyroxene Distribution of S-type Asteroids Throughout the Main Belt

Volatiles (H, C, N, F, S, Cl) in the lunar mantle, crust, and regolith: What questions remain and where to go next?

PSRD: Recent Gas Escape from the Moon

Mercury Data (Table 11-1) 11a. Sun-Scorched Mercury. Mercury Data: Numbers

Galileo solid state imaging data, and the mixing of mare

The Mairan domes: Silicic volcanic constructs on the Moon

MESSENGER: Exploring New Views of Mercury s Exosphere and Surface

Thickness of proximal ejecta from the Orientale Basin from Lunar Orbiter Laser Altimeter (LOLA) data: Implications for multi ring basin formation

GASEOUS JETS IN COMET HALE BOPP (1995 O1) 1. Introduction

The Tagish Lake Meteorite: A Possible Sample from a D-Type Asteroid

arxiv: v2 [astro-ph.ep] 30 Jan 2014

LRO-LOLA: Measurements of Lunar Altimetry and Surface Conditions

The Apollo 17 Landing Site. posted February 12, 1997

Volcanic Features and Processes

Estimation of ocean contribution at the MODIS near-infrared wavelengths along the east coast of the U.S.: Two case studies

Hyperspectral Data as a Tool for Mineral Exploration

Q. Some rays cross maria. What does this imply about the relative age of the rays and the maria?

High resolution measurements of absolute thorium abundances on the lunar surface

OH/H 2 0 Enhancements associated with Some Silicic Volcanism as detected by M 3

Accessing the Lunar Poles for Human Exploration Missions

MARINER VENUS / MERCURY 1973 STATUS BULLETIN

Student Guide to Moon 101

Malapert Mountain: A Recommended Site for a South Polar Outpost

OSIRIS-REX OVERVIEW PRESENTATION TO THE PLANETARY SCIENCE SUBCOMMITTEE

LUNAR OLIVINE EXPOSURES: ORIGINS AND MECHANISMS OF TRANSPORT

Hansteen Alpha: A volcanic construct in the lunar highlands

Presented by: Sydney Brewer, Erin Manuel, Julie Ponton, Shannon Smith, and Caroline Stasiowski Seton Keough High School, Baltimore, Maryland

ASTRONOMY. Chapter 9 CRATERED WORLDS PowerPoint Image Slideshow

The origin of lunar crater rays

624 SNPP VIIRS Solar Diffuser BRDF Degradation Trend Changes in Early Evan Haas* and Frank De Luccia The Aerospace Corporation, El Segundo CA

Transcription:

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 116,, doi:10.1029/2010je003728, 2011 Newer views of the Moon: Comparing spectra from Clementine and the Moon Mineralogy Mapper Georgiana Y. Kramer, 1 Sebastien Besse, 2 Jeffrey Nettles, 3 Jean Philippe Combe, 1 Roger N. Clark, 4 Carlé M. Pieters, 3 Matthew Staid, 5 Erik Malaret, 6 Joseph Boardman, 7 Robert O. Green, 8 James W. Head III, 3 and Thomas B. McCord 1 Received 31 August 2010; revised 17 December 2010; accepted 18 January 2011; published 9 April 2011. [1] The Moon Mineralogy Mapper (M 3 ) provided the first global hyperspectral data of the lunar surface in 85 bands from 460 to 2980 nm. The Clementine mission provided the first global multispectral maps the lunar surface in 11 spectral bands across the ultraviolet visible (UV VIS) and near infrared (NIR). In an effort to understand how M 3 improves our ability to analyze and interpret lunar data, we compare M 3 spectra with those from Clementine s UV VIS and NIR cameras. The Clementine mission provided the first global multispectral maps the lunar surface in 11 spectral bands across the UV VIS and NIR. We have found that M 3 reflectance values are lower across all wavelengths compared with albedos from both of Clementine s UV VIS and NIR cameras. M 3 spectra show the Moon to be redder, that is, have a steeper continuum slope, than indicated by Clementine. The 1 mm absorption band depths may be comparable between the instruments, but Clementine data consistently exhibit shallower 2 mm band depths than M 3. Absorption band minimums are difficult to compare due to the significantly different spectral resolutions. Citation: Kramer, G. Y., et al. (2011), Newer views of the Moon: Comparing spectra from Clementine and the Moon Mineralogy Mapper, J. Geophys. Res., 116,, doi:10.1029/2010je003728. 1. Introduction [2] The Clementine mission mapped the lunar surface between 19 February and 3 May 1994 and provided the first high spatial resolution, global multispectral imaging of the entire lunar surface captured in 11 spectral bands via the ultraviolet visible (UV VIS) and near infrared (NIR) cameras [Nozette et al., 1994; McEwen and Robinson, 1997]. Clementine s UV VIS and NIR cameras had limited capabilities inherent in their low spectral resolution and the fact that a continuous UV VIS NIR spectrum had to be divided between two cameras. In addition, calibration of both data sets were fraught with difficulties, some of which delayed the release of the global NIR Digital Image Model (DIM) as fully calibrated and resampled to fit with the UV VIS DIM until 2007 (L. Gaddis et al., The Clementine NIR global lunar mosaic, 2007; P. Lucey, USGS Clementine NIR global 1 Bear Fight Institute, Winthrop, Washington, USA. 2 Department of Astronomy, University of Maryland, College Park, Maryland, USA. 3 Department of Geological Sciences, Brown University, Providence, Rhode Island, USA. 4 U.S. Geological Survey, Denver, Colorado, USA. 5 Planetary Science Institute, Napa Valley, California, USA. 6 Applied Coherent Technology, Herndon, Virginia, USA. 7 Analytical Imaging and Geophysics, Boulder, Colorado, USA. 8 Jet Propulsion Laboratory, Pasadena, California, USA. Copyright 2011 by the American Geophysical Union. 0148 0227/11/2010JE003728 mosaic, http://astrogeology.usgs.gov/projects/clementinenir/, 2007]. Nevertheless, for more than 15 years these two instruments have been the foundation upon which multispectral data analysis has advanced our knowledge of the Moon. Considering the wealth of scientific results and significant advancements in lunar science, as well as the important contributions to Earth and Planetary Science in general, this is quite an achievement for one small mission. [3] The release of new, global imaging spectrometer data from the Moon Mineralogy Mapper (M 3 ), marks the passing of the torch to the next generation of lunar spectroscopy and our new understanding of the Moon. In an effort to understand how M 3 improves our ability to analyze and interpret lunar data, we compare M 3 spectra with those from Clementine s UV VIS and NIR cameras of representative regions and lithologies across the lunar surface. Specifically, we compare the recently calibrated and photometrically corrected M 3 images [Hicks et al., 2011; R. O. Green et al., The Moon Mineralogy Mapper (M3) imaging spectrometer for lunar science: Instrument, calibration, and on orbit measurement performance, submitted to Journal of Geophysical Research, 2011] to Clementine DIMs [McEwen et al., 1998; McEwen and Robinson, 1997]. 2. Mission and Instrument Comparison [4] The M 3 spectrometer was a guest instrument on Chandrayaan 1, the Indian Space Research Organization s (ISRO) first mission to the Moon [Goswami and Annadurai, 1of11

Table 1. Instrument Specifications for Clementine s UV VIS and NIR Cameras and the Moon Mineralogy Mapper UV VIS NIR M 3a Focal plane array Thompson CCD Amber InSb HgCdTe Pixel array 384 228 1 256 256 1 600 1 260 (T) (samples lines bands) 300 1 85 (G) Pixel size (mm) 23 23 38 38 27 27 Field of view (deg) 5.6 4.2 5.6 5.6 24 Spatial resolution b 100 200 150 500 70 (T) (m/pixel) 140 (G) Spectral range 415 ± 20 1100 ± 30 460 2980 ± 10 (T) and resolution c (nm) 750 ± 5 1250 ± 30 460 700 ± 40 (G) 900 ± 15 1500 ± 30 730 1550 ± 20 (G) 950 ± 15 2000 ± 30 1580 2980 ± 40 (G) 1000 ± 15 2600 ± 30 400 1000 2792 ± 146 Power (W) 4.5 11 15 a T, target mode, and G, global mode. b Clementine at 425 km altitude and M 3 at 100 km altitude. c Clementine, band pass; filters ± FWHM. 2009]. Although Chandrayaan 1 s did not complete its scheduled mission, M 3 imaged almost the entire lunar surface [cf. Boardman et al., 2011] in global mode, which consisted of 85 spectral bands between 460 to 2980 (cf. Green et al., submitted manuscript, 2010). Orbital observations from M 3 are divided into optical periods (OP) that differ mainly with spacecraft altitude and beta angle (the angle between the plane of the spacecraft s orbit and the vector from the sun). M 3 s spatial resolution is 140 m/pixel for images obtained while in a 100 km orbit, between OP1A and OP2B, and 280 km/pixel in a 200 km orbit, OP2C OP2D [cf. Boardman et al., 2011]. For this comparative analysis we used level 1B calibrated M 3 data (radiometrically calibrated to radiance) from OP1A OP1B. The data were selenographically registered by ray tracing each M 3 spatial element on the lunar surface to a LOLA derived lunar reference frame [cf. Boardman et al., 2011]. The reader is referred to the several companion articles in this issue for thorough discussions of the M 3 instrument and calibration (Green et al., submitted manuscript, 2010), geometric control [Boardman et al., 2011], and photometric correction [Hicks et al., 2011], which are all fundamental for the comparison presented herein. [5] M 3 calibration is primarily based on preflight laboratory acquired calibration information, and is improved through adjustments to the flat field and stray light calibrations using in flight data (Green et al., submitted manuscript, 2010). Radiance calibration is still undergoing refinement, and we expect subsequent calibrations (such as will be released to the PDS) will have better addressed known issues, such as a scattered light component below 1 mm, that may cause artifacts in the work presented here. Our analysis used K calibrated, photometrically corrected reflectance data at standard illumination geometry (incidence angle, i = phase angle, a = 30 ) [cf. Boardman et al., 2011; Hicks et al., 2011] to bring the data to a level of processing similar to that of Clementine s DIMs (Figure 2). This photometric correction does not include topography, instead it assumes a smooth sphere of radius = 1734 km, which is the same as the photometric calibration for Clementine. M 3 sk calibration contained residual spectral structure still present even after these described conversions (cf. Green et al., submitted manuscript, 2010). We used a correction factor (specific to each wavelength) to smooth M 3 spectra, which is described by Clark et al. [2011]. Spectra presented here are truncated at 2.6 mm to avoid stretching of the plot due to the high reflectances at longer wavelengths due to thermal emission. Thermal corrections are also still being refined, and so have not been applied to the data used in this analysis. This is also in keeping with our endeavor to compare data sets at similar levels of calibration as Clementine data were never thermally corrected. As a result, beyond 2.2 mm the data may diverge from a true reflectance spectrum due to an influence from thermal emission, and it should be noted that these do not necessarily result from calibration differences. [6] Clementine s 11 band spectra were achieved by passing light entering the optics through one of six spectral filters on a rotating wheel [Nozette et al., 1994], one wheel for each camera. The six band pass filters for the UV VIS camera and six for the NIR camera are listed in Table 1, and their spectral response curves are shown in Figure 1. Figure 1. Band passes of Clementine filters for the UV VIS and NIR cameras. In addition to the 10 filters shown here, the UV VIS camera had a broadband filter, and the NIR camera had a filter centered near 2800 nm. From Nozette et al. [1994]. 2of11

Figure 2. Zoom of small crater located at 3.5 S and 35.5 E (pink ring in Figure 3a). Each frame is 6.8 km across. Selected to compare spatial resolutions of (a) M 3 at 750 nm reflectance and (b) Clementine s UV VIS camera at 750 nm reflectance, (c) M 3 at 2020 nm, and (d) Clementine s NIR camera at 2000 nm; (e) M 3 band ratio color composite (based on Clementine s UV VIS channels). Note the east to west gradient, which is due to the photometry correction used here not yet being fully adapted. (f) Clementine band ratio color composite. Images created from M 3 frame M3G20090106T152345 and Clementine DIM frame ui03s033. Green crescent in center of Figure 2f (western crater wall) is due to saturated pixels (which were thus recorded as null values) in Clementine s 415 nm channel. [7] Full resolution Clementine DIMs are available through the U.S. Geological Survey, currently as two data sets: one from Clementine s UV VIS camera [Eliason et al., 1999] and the other from the NIR camera [Eliason et al., 2003]. Both data sets have been radiometrically corrected [McEwen et al., 1998], geometrically controlled, spatially resampled to 100 m/pixel [Eliason et al., 2003], and photometrically normalized to form uniformly illuminated mosaics of the lunar surface. Coverage of the Moon by Clementine s UV VIS and NIR cameras are close to 100% and essentially identical. The spatial resolution of the UV VIS data set ranged from 100 to 200 m/pixel (Figure 2b) and the NIR data set is 500 m/pixel (Figure 2d), but both were spatially resampled to 100 m/pixel so that combined, every pixel contains a continuous and cartographically distinct spectrum. [8] Early tests of the UV VIS data being returned from Clementine using the preflight derived calibration quickly demonstrated the need for in flight calibration observations. Starting at orbit 94, Clementine made systematic observations of Vega and empty space for calibration purposes on almost every orbit. It is from these observations that the Clementine team derived the point spread function for each spectral filter, adjusted instrument gains and offsets, subtracted the dark current, applied the flat field normalization, and converted digital number (DN) to radiance [Malaret et al., 1998; McEwen et al., 1998]. 3of11

Figure 3. (a) Clementine 750 nm reflectance image depicting locations from which sampled spectra from both Clementine and M 3 were derived. Colors of squares correspond to colors of spectra in Figures 3b 3e; colors of rings correspond to colors of spectra in Figures 3f 3i. The open blue and lavender squares depict locations of the craters and spectra detailed in Figures 4 and 5, respectively. (b) M 3 spectra scaled to 750 nm albedo of surface regolith (squares in Figure 3a). (c) Clementine spectra scaled to 750 nm albedo of surface regolith (same squares in Figure 3a). (d) Continuum removed M 3 spectra; continuum calculated at 750 nm and 1510 nm. Continuum removed Clementine spectra; continuum calculated at 750 nm and 1500 nm. (f i) Same as for Figures 3b 3e except spectra derive from immature craters (rings in Figure 3a). [9] Clementine s conversion to reflectance did not use the preferred method of normalizing the radiance calibration to the incident solar irradiance. Instead they used an absolute calibration based on observations of the Apollo 16 landing site and laboratory reflectance measurements of Apollo 16 soil sample, 62231 [McEwen et al., 1998; Pieters, 1999]. The Apollo 16 site was selected as a calibration target because the region is relatively homogeneous, and since it is largely battered anorthositic highland material, the spectrum is bright and relatively free of absorption features that could introduce calibration artifacts. The soil sample measurements were relative to a halon standard and acquired at laboratory standard geometry (i = a = 30,e = 0) at the RELAB facilities at Brown University [Pieters, 1983; Pieters and Hiroi, 2004]. To convert to reflectance, gain, offset, and flat field corrected DNs were photometrically corrected to the same viewing geometry as the laboratory measurements of the Apollo 16 soil sample. The Apollo 16 soil sample was normalized to the calibrated DN to obtain a spectral calibration correction factor that can be applied to any photometrically corrected lunar region. [10] The Apollo 16 soil measurements were made out to 2500 nm, and so could be convolved for all UV VIS bands and the first four bands of the NIR data (1100 2000 nm) [Eliason et al., 2003]. The derived NIR reflectance values were then compared with the same Apollo 16 landing site to obtain a normalization factor. The 2600 and 2780 nm bands were not included in this calibration because reflectance information was not available, and because the reflectance signal is complicated by thermal emission at these wavelengths. [11] Efforts to develop a lunar photometric function based on Galileo [McEwen et al., 1996] and Earth based telescopic observations [Hillier et al., 1999] demonstrated early on the difficulties associated with trying to completely remove the effects of viewing geometry. Although photometric behavior does vary with terrain, the photometric correction used to create Clementine s DIMs does not vary 4of11

Figure 4. (a) M 3 750 nm albedo image of Franck Crater (location north of Mare Tranquillitatis shown as blue open square in Figure 3a). (b) Locations in crater from which pixel data were averaged to create spectra shown in Figures 4c 4f. Colorsof features in Figure 4b correspond to colors of spectra in Figures 4c 4f: dark blue, crater floor; medium blue, crater wall; light blue, light streaks on wall. (c) M 3 spectra scaled to 750 nm albedo. (d) Clementine spectra scaled to 750 nm albedo. (e) Continuum removed M 3 spectra; continuum calculated at 750 nm and 1510 nm. (f) Continuum removed Clementine spectra; continuum calculated at 750 nm and 1500 nm. as a function of albedo. Clementine DIMs use a reversible phase function and include a full resolution map of phase angle for each pixel. This was done with the intention that one could use the phase map to remove the existing phase correction and apply a new phase correction, or apply a second correction, specific to a desired terrain or phase angle. [12] Clementine absolute reflectance values are significantly higher (by a factor of 1.5 2) than what is expected based on calibrated telescopic measurements of the Moon from Earth [Shkuratov et al., 2001]. This difference in albedo is likely a Clementine calibration issue linked to the necessity to use an Apollo 16 soil sample measured in the laboratory to calibrate Clementine DN to reflectance. The comparative albedos between Clementine and telescopic data revealed that lab measurements of a mature soil are brighter than natural remote measurements of a comparable mature soil. The cause of the discrepancy is unclear, but may be related to the calibration correction for the halon reference standard [Blewett et al., 1997], scattered light, and/or to differences in roughness and compaction between the studied soil samples and the actual lunar surface [Hillier et al., 1999; Shkuratov et al., 2001]. Therefore, we expect Clementine spectra to be brighter than M 3 ; even when M 3 data are photometrically corrected to the same geometry as Clementine. To remove these albedo effects and properly compare spectral differences between the two data sets we scaled Clementine and M 3 spectra to their respective 750 nm reflectance values. [13] Even after normalizing both data sets to their respective 750 nm albedo, we expect there will still be a contrast difference because many areas of the UV VIS data have a significant additive signal due to scattered light. P. Lucey (USGS Clementine NIR global mosaic, 2007) developed a set of corrections to better match the UV VIS and NIR data to telescopic measurements near Aristarchus Crater. These corrections may help to improve the scattered light and other photometric effects, but do not fix Clementine s high albedo issue. We have opted not to use these correction factors in this analysis to keep with our effort to compare the instruments at similar levels of calibration. The K calibrated M 3 data used for the analysis reported here used no spectral correction based on spectral data from another instrument. 3. Analytical Techniques [14] To compare spectra between Clementine and M 3 we selected regions on the lunar nearside that are easily recognizable, selenographically identifiable, cover both mare and highland lithologies, and where image data from both instruments were acquired under reasonably similar illumination geometries. Knowledge of the spacecraft s viewing geometry, in particular the phase angle, is essential for proper conversion to reflectance, cartographic, and photometric applications. The solar phase angle is the angle between the vector to the Sun and the vector to the spacecraft from a given point on the Moon s surface. The phase angle that describes a mapping orbit (and as discussed herein) refers to the angle between the vector to the Sun and the intersection of the spacecraft s orbit with the Moon s equator. The optimal solar phase angle for mapping lunar mineralogy is between 20 and 40 degrees. At lower angles the data begin to suffer from the opposition effect: a rapid 5of11

Figure 5. (a) M 3 750 nm albedo image of Rosse Crater (location in Mare Nectaris shown as lavender open square in Figure 3a). (b) Locations in crater from which pixel data were averaged to create spectra shown in Figures 5c 5f. Colors of features in Figure 5b correspond to colors of spectra in Figures 5c 5f: green, crater floor; dark purple, light crater wall; light purple, dark material at top of crater wall (east and west); dark lavender, dark material at top of crater wall (north); salmon, short dark material spilling down proximal ejecta blanket. (c) M 3 spectra scaled to 750 nm albedo. (d) Clementine spectra scaled to 750 nm albedo. (e) Continuum removed M 3 spectra; continuum calculated at 750 nm and 1510 nm. (f) Continuum removed Clementine spectra; continuum calculated at 750 nm and 1500 nm. increase in albedo as the source of illumination moves behind the observer (the phase angle approaches zero). At angles greater than 40, shadows obscure too much of the image. 30 degrees is the standard phase angle used for laboratory measurements of reflectance spectra, so for ease of comparison, Clementine s orbit was selected to maximize the time period in which the solar phase angle was within 30 degrees [Eliason et al., 1999]. M 3 viewed the Moon from a wider variety of phase angles, and there are few locations where Clementine and M 3 viewed the Moon with the same illumination geometry. The Tranquillitatis Nectaris region (Figure 3) was selected for this study because it was mapped by both Clementine and M 3 at approximately 30 degrees phase angle. [15] The Clementine mosaics used in this study (Figure 3a) are resampled to 140 m/pixel to match the spatial resolution of the comparable data from M 3. Our technique samples from several pixels and used the average of these pixels spectra, which should mitigate problems associated with imperfect registration between the 2 data sets, as well as known issues merging spectra between Clementine s UV VIS and NIR cameras. [16] We sampled surface regolith (Figures 3b 3e) and features within two larger ( 10 km diameter) craters, Franck (Figures 4b 4e) and Rosse (Figure 5), which were delineated based on albedo and/or morphology. Each spectrum is an average from several pixels (accurately depicted in Figures 3a and 4a) of no less than 4 pixels in one spatial dimension. The 4 pixel restriction ensures we are not sampling from less than the true spatial resolution of Clementine NIR data (500 m/pixel). Averaging several pixels from a single feature reduces noise, and provides a more robust spectrum. Combined with the Small Crater Rim and Ejecta Probing (SCREP) procedure [Kramer, 2010], we obtained 50 spectra for each instrument, where each spectrum is an average of several pixels from (what can be spatially resolved as) a single lithology. [17] Spectra shown in Figures 3f 3i were obtained using SCREP to extract compositional information from pixels on the rims and proximal ejecta of small, immature craters (1 25 km in diameter). This was done to explore the idea that these craters can be used see through the mature regolith, and observe the spectral character of the underlying lithology that is exposed in their ejecta [McCord and Adams, 1973; Staid and Pieters, 2000]. Analysis is focused on the rims and proximal ejecta of the craters because it is thought that this location best exposes the pristine bedrock while simultaneously avoiding uncorrected photometric effects due to steep slopes [Kramer et al., 2008; Kramer, 2010]. This area represents the thickest part of the crater ejecta, and thus consists of the most concentrated, or highest proportion of native material compared to foreign material that collectively make up the regolith [e.g., Arvidson et al., 1975; Melosh, 1989; Li and Mustard, 2000]. Furthermore, the rim and proximal ejecta suffer the least amount of postimpact regolith buildup [Kramer, 2010]. For each selected crater, SCREP defines and extracts spectral information from pixels that describe a ring, the inner circumference of which delineates the crater rim and the outer circumference such that the ring thickness is 1/10 of the crater diameter. These data are then averaged to obtain a single spectrum for each crater that is meant to closely approximate the composition of the pristine, 6of11

Figure 6. Comparison of M 3 albedos (y axes) versus Clementine albedos (x axes) for the 11 Clementine band canters. Note that several of the plots do not share the same axes values. All plot axes do have a 1:1 relationship except 2780 nm, which is 1:3. M 3 was resampled to Clementine spectral resolution to create these plots. Since M 3 s first spectral channel is 461 nm and Clementine s is 415 nm, we calculated Clementine s albedo at 461 assuming a straight line fit between 415 and 750 nm. subregolith lithology exposed in that location. In this way, the method reduces potential errors in interpretation of spectral features compared with using spectra derived from only a single pixel. 4. Discussion 4.1. Spatial Comparison [18] Figure 2 shows M 3 and Clementine images of a small impact crater in the highland/light plains peninsula between Mare Tranquillitatis and Mare Fecunditatis. Figure 2 demonstrates the differences in spatial resolution between M 3 and Clementine s UV VIS (Figures 2a and 2b) and NIR cameras (Figures 2c and 2d). The M 3 data used for this work are from OP1b, so the spatial resolution across its spectral range is 140 m/pixel [Boardman et al., 2011]. Clementine s spatial resolutions also varied by orbital altitude, but also by camera (see Table 1). Although Clementine UV VIS DIMs are sampled at 100 m/pixel, and can be that good for the 750 nm channel, morphological features in Figure 2a are much clearer than in Figure 2b. Since the images from both instruments are taken at similar illumination conditions, the difference cannot be explained by shadows. The blurry appearance of Clementine images (compared with M 3 images) could be a consequence of the on board compression algorithms, which were necessary for Clementine to map the entire Moon [McEwen and Robinson, 1997]. In the NIR, M 3 reflectance data (Figure 2c shows M 3 channel 2020 nm) at 140 m/pixel shows features remarkably clearer than the comparable channel for Clementine s NIR camera (2000 nm) at 500 m/pixel (Figure 2d). 7of11

Table 2. Gains and Offsets of M 3 Data Relative to Clementine Data for Each Wavelength a l(nm) Gain Offset s gain s offset 461 0.6566 0.0046 0.0066 0.0153 750 0.6469 0.0106 0.0101 0.0241 900 0.7177 0.0153 0.0100 0.0193 950 0.6772 0.0092 0.0093 0.0202 1000 0.6923 0.0120 0.0104 0.0217 1100 0.6676 0.0086 0.0111 0.0250 1250 0.6549 0.0032 0.0119 0.0278 1500 0.5946 0.0135 0.0125 0.0355 2000 0.5842 0.0230 0.0113 0.0331 2600 0.5506 0.0533 0.0123 0.0405 2780 0.2163 0.0685 0.0050 0.1067 a The M 3 data were resampled to Clementine spectral resolution. Data points are from the 50 SCREP derived and averaged regolith spectra plotted in Figure 6. The gains and offsets were calculated using a least absolute deviation method to minimize the influence of outliers in the data. Errors were calculated from the contributions of the gains and offsets to the absolute deviation of the straight line fit. [19] The color composite images (Figures 2e and 2f) are both based on the band ratios used for Clementine, specifically the 415, 750 and 950 channels, which is described in detail by Greeley et al. [1993] and Pieters et al. [1994]. Three ratios were selected to highlight petrologic characteristics of the surface. The red channel is the 750/415 nm ratio, which is high when the spectrum contains a steep continuum slope, increasing reflectance with increasing wavelength. A steep continuum slope is characteristic of a glass spectrum [e.g., Tompkins and Pieters, 2010] and mature soils, the latter due to increasing amounts of nanophase Fe 0 and agglutinates produced by space weathering [McCord and Adams, 1973; McKay et al., 1991; Taylor et al., 2001; Hawke et al., 2001]. [20] The green channel is the 750/950 nm ratio, which reflects the abundance of Fe 2+ in minerals and glasses. Absorption of the 950 nm wavelength is caused by the presence of ferrous iron [Burns, 1982] and is stronger for immature mafic surfaces. Thus, the green channel is enhanced where there is a strong mafic absorption in the soil spectrum, such as where fresh, FeO rich mare basalts dominate. [21] The blue channel is the 415/750 nm ratio. Although this ratio is the inverse of the ratio assigned to the red channel, the 415/750 nm ratio reflects compositions beyond simply the diminished influence of the surface properties to which the red channel is sensitive. 750/415 nm values are high when there is a strong continuum slope, and its inverse has high values when the continuum is flatter. A flatter, or shallow, spectral slope reflects the influence from opaque mineral phases, such as ilmenite (dark and flat spectrum) [Charette et al., 1974] or fresh highlands (bright and flat spectrum). Thus, in the red channel high 750/415 nm pixels are bright (red) and lower 750/415 nm values are darker, or rather, less red, allowing the two other color channels to come through. By inverting the 750/415 nm ratio and assigning the blue channel to that ratio, what would be dark in the red channel are now bright blue. In this way, the variety of hues that result from mixing between the three color channels provide information beyond the 3 end member ratios. [22] The band ratio color composite images provide clear distinctions between pyroclastic deposits and/or impact melt, ejecta rays, and units of different mineralogy. To recreate a Clementine style color composite image using M 3 data we averaged the reflectance values from three adjacent bands that correspond as closely to the Clementine bands as possible: R 415M 3 ¼ meanfr 460:99 ; R 500:92 ; R 540:84 g R 750M 3 ¼ meanfr 730:48 ; R 750:44 ; R 770:40 g R 950M 3 ¼ meanfr 930:10 ; R 950:06 ; R 970:02 g [23] Some of the differences between the M 3 and Clementine color ratio images are clearly associated with the different spatial resolutions, as discussed above. Also, the channels used to create the RGB image are affected by scattered light in both instruments in different ways. The vertical striping in Figure 2e is a result of dividing two bands that both have residual noise (even after the flat field correction) in the same spatial pixels. There is also a gradient from east to west which is related to the photometry correction used for this study, which is still being improved [Hicks et al., 2011]. Such features are expected to be improved in later calibrations, such as the data recently released to the PDS, as well as future calibrations. Despite the accentuated striping, the M 3 color composite shows more detail and petrologic complexity in the crater s ejecta over Clementine s color composite (Figure 2f). 4.2. Spectral Comparison [24] There are no consistent differences in shape of the spectrum due to band depth, band center, or overall albedo, which is evident in the comparison of data captured by both instruments from the same location. However, the continuum slope is consistently redder in M 3 data compared with Clementine (Figures 3, 4, and 5). The discrepancy in the continuum slopes from each instrument reflects Clementine s reflectance calibration based on Apollo 16 soil sample measurements, which are now known to be brighter than Earthbased measurements of a comparable mature soil [Blewett et al., 1997; Hillier et al., 1999; Shkuratov et al., 2001]. [25] The Moon Mineralogy Mapper and Clementine s UV VIS and NIR cameras are very different instruments, so it is unsurprising that data returned by them would have noteworthy differences. M 3 is a spectrometer, which separates incoming light via a diffraction grating, and sends the entire spectrum of each spatial pixel of a one dimensional array to a two dimensional detector. Clementine captured a two dimensional spatial scene one spectral channel at a time via a color filter wheel. Clementine s radiance calibration used orbital observations of Vega for radiometric calibration, and laboratory spectra of Apollo 16 soil samples to obtain reflectance data. In contrast, M 3 s calibration was based largely on preflight laboratory tests (cf. Green et al., submitted manuscript, 2010), so does not have issues inherent in an assumption about the lunar continuum. [26] Figures 3, 4, and 5 compare M 3 and Clementine spectra from immature craters and mature regolith, and demonstrate the value in higher spectral resolution for mineralogical interpretation. M 3 s increased spectral sampling means we can resolve a wavelength within ±10 nm (±20 nm or ±40 nm, depending on the wavelength; see Table 1), which improves our ability to not only identify mineralogy, but also compositions with a mineral solution series [Adams, 8of11

spectrum did not always properly line up. Such a nonlinearity between the two cameras can sometimes be observed as a deviant slope between 1000 and 1100 nm (e.g., yellow spectra in Figures 3c and 3e), although it is not always evident. Regardless, no correction has been found to smooth this continuum mismatch and therefore precludes reliable mineralogical assessment with Clementine data. [28] As stated above, Clementine reflectance spectra are expected to be too bright. Therefore, all the spectra shown that compare M 3 and Clementine have been normalized to their respective 750 nm reflectance for the purpose of an appropriate comparison. To demonstrate the differences between M 3 and Clementine spectra we resampled M 3 to Clementine wavelengths and plotted the albedos for each wavelength from all 50 of the averaged SCREP and regolith spectra extracted from each data set (not all averaged spectra are shown in Figures 3 5). These averaged spectra were used as comparison instead of individual pixels because even with spatial resampling, the particular feature covered in a single pixel will not be the same between the compared instruments, and the overall albedo of a pixel is a function of the albedos within that pixel. Therefore, we compared the averaged spectra for improved accuracy. This explains at least some of the deviations (error bars in Figure 6) in the gains and offsets. Since M 3 s first spectral channel is 461 nm and Clementine s is 415 nm, we calculated Clementine s albedo at 461 nm assuming a straight line fit between 415 nm and 750 nm. Using the 50 SCREP derived and averaged regolith spectra (plotted in Figure 6) we calculated gains and offsets of M 3 relative to Clementine for each wavelength (M 3 data resampled to Clementine spectral resolution). The gains and offsets were calculated using a least absolute deviation method to minimize the influence of outliers in the data (Tables 1 and 2 and Figure 7). Figure 7. Trends in the plots of Figure 6 shown as (a) gains and (b) offsets as a function of wavelength. Shows gains and offsets of M 3 data from Clementine data. The gains and offsets were calculated using a least absolute deviation method to minimize the influence of outliers in the data. Data for plots are given in Table 2. 1974; Klima et al., 2011]. Clementine reflectance data made high resolution, global mapping of mafic mineralogy possible [e.g., Staid and Pieters, 2000; Lucey, 2004]. However, the ability to discern mineralogy or differences in mineral compositions (e.g., low Ca pyroxene versus high Ca pyroxene) was mostly limited to three channels in the UV VIS. Expanding the spectral range by appending NIR camera data provided a single channel (2000 nm) with which to confirm the presence of pyroxene. [27] Olivine has a broad 1 mm absorption that can still be observed at wavelengths as long as 1300 nm (see the turquoise spectrum in Figures 3h and 3i). The extended spectral range of the combined UV VIS and NIR cameras could have provided this potentially diagnostic feature had the merged data sets been consistently smooth. However, calibration parameters were complex and inconsistent between the two cameras and over time, so that a merged 5. Conclusions [29] The reflectance values of both of Clementine s UV VIS and NIR cameras are higher across all wavelengths compared with M 3 reflectance. When both data sets are normalized to their respective 750 nm reflectance values (Figures 3b, 3c, 3f, 3g, 4c, 4d, 5c, and 5d) differences in the continuum slope and spectral shape are more readily apparent. Continuum removed spectra (Figures 3d, 3e, 3h, 3i, 4e, 4f, 5e, and 5f) accentuate absorption band features, such as band minimum, depth, and width. M 3 spectra show the Moon to be redder, or have steeper continuum slope, than indicated by Clementine. 1 mm absorption band depths may be comparable between the instruments, but Clementine consistently has shallower 2 mm band depths than M 3. Absorption band minimums are difficult to compare due to the significantly different spectral resolutions. [30] Despite the inherent differences between the two instruments, and limitations in calibrations and photometric corrections, data from M 3 and Clementine s two multispectral cameras demonstrate some remarkable consistencies. The data sets display similar spectral shapes and relative albedos between the different pixel averaged SCREP craters (Figures 3, 4, and 5). Deviations from these two spectral similarities are associated with a sampled crater s location near the edge of one (or both) of a data set s orbital strip. These are locations where both Clementine s 9of11

and M 3 s (current) photometric corrections suffer the greatest error. [31] Some of the inaccuracies in Clementine reflectance data stems from our limited understanding of the effect of lunar environmental conditions (e.g., illumination geometry) on remote reflectance data, such as its effect on the spectral continuum slope. Clementine reflectance data set would have been more accurate had it not required calibration based on laboratory measurements of Apollo 16 soil samples. New information from M 3, Kaguya, and Lunar Reconnaissance Orbiter could be used to revisit Clementine calibration. [32] M 3 and Clementine data will always be very different from each other. Notwithstanding the inherent differences between the instruments and their calibrations, both missions are over, and much of their data for any given location were acquired under different illumination geometries. To bring both data sets to comparable levels would have require rigorous preflight testing and characterization for each instrument and measurements of the changes that were inflicted upon the instrument beginning from launch to mission completion. However, even with that, images would need to be acquired from (lunar) morning and afternoon illuminations to obtain absolute spectral information for all regions of the Moon. Last, the thermal and photometric corrections would need to be capable of accurately correcting all conditions that affect them because inherent difference between the instruments (such as field of view, spectral resolution) would preclude their ability to capture precisely the same regions under precisely the same illumination conditions. [33] Photometric models, including corrections for the thermal contributions, are still being improved as is our understanding of the causes and effects illumination conditions and the space environment have on reflectance data. This type of data comparison is an important step in that process. This comparison is not meant to highlight the flaws or inaccuracies in one instrument over another, but instead to relish in our accomplishments of the past, delight in our progress, and explore our ability for further improvement. It is from this type of comparative analysis and future, more detailed ones that we will learn and understand how to bring remote measurements closer to reality and improve the return of scientific knowledge from spacecraft data. The Moon, as our nearest celestial neighbor is the best recorder of our solar system s history and perfect example of naked exposure to space. Between 2006 and 2010, we increased the number of spacecraft that acquired global data of our Moon by half, compared with all years previous; the available data volume has several data sets with which to refine our understanding. [34] Acknowledgments. The authors wish to thank the Indian Space Agency, ISRO, for providing the Moon Mineralogy Mapper the opportunity to ride as a guest instrument on Chandrayaan 1. The authors also thank the reviewer for insightful comments and suggestions. References Adams, J. B. (1974), Visible and near infrared diffuse reflectance spectra of pyroxenes as applied to remote sensing of solid objects in the solar system, J. Geophys. Res., 79, 4829 4836, doi:10.1029/jb079i032p04829. Arvidson, R., R. J. Drozd, C. M. Hohenberg, C. J. Morgan, and G. Poupeau (1975), Horizontal transport of the regolith, modification of features, and erosion rates on the lunar surface, Moon, 13, 67 79, doi:10.1007/ BF00567508. Blewett, D. T., P. G. Lucey, B. R. Hawke, G. G. Ling, and M. S. Robinson (1997), A comparison of Mercurian reflectance and spectral quantities with those of the Moon, Icarus, 129, 217 231, doi:10.1006/ icar.1997.5785. Boardman, J. W., C. M. Pieters, R. O. Green, S. Lundeen, P. Varansi, J. Nettles, N. E. Petro, P. J. Isaacson, S. Besse, and L. A. Taylor (2011), Measuring moonlight: An overview of the spatial properties, lunar coverage, selenolocation and related level 1B products of the Moon Mineralogy Mapper, J. Geophys. Res., doi:10.1029/2010je003730, in press. Burns, R. G. (1982), Mineralogical Applications of Crystal Field Theory, 2nd ed., Cambridge Univ. Press, New York, NY. Charette, M. P., T. B. McCord, and C. Pieters (1974), Application of remote spectral reflectance measurements to lunar geology classification and determination of titanium content of lunar soils, J. Geophys. Res., 79, 1605 1613, doi:10.1029/jb079i011p01605. Clark, R. N., C. M. Pieters, R. O. Green, J. W. Boardman, and N. E. Petro (2011), Thermal removal from near infrared imaging spectroscopy data of the Moon, J. Geophys. Res., doi:10.1029/2010je003751, in press. Eliason, E. M., et al. (1999), Digital processing for a global multispectral map of the Moon from the Clementine UVVIS imaging instrument, Lunar Planet. Sci., XXX, Abstract 1933. Eliason, E. M., et al. (2003) A near infrared (NIR) global multispectral map of the Moon from Clementine, Lunar Planet. Sci., XXXIV, Abstract 2093. Goswami, J. N., and M. Annadurai (2009), Chandrayaan 1: India s first planetary science mission to the Moon, Lunar Planet. Sci., XL, Abstract 2571. Greeley, R., M. J. S. Belton, J. W. Head, A. S. McEwen, C. M. Pieters, G. Neukum, T. L. Becker, E. M. Fischer, S. D. Kadel, and M. S. Robinson (1993), Galileo imaging results from the second Earth Moon flyby: Lunar maria and related units, Proc. Lunar Planet. Sci. Conf., 24, 565 566. Hawke, B. R., P. G. Lucey, G. A. Smith, G. J. Taylor, D. J. Lawrence, and P. D. Spudis (2001), Remote sensing studies of selected spectral anomalies on the Moon, Lunar Planet. Sci., XXXII, Abstract 1241. Hicks, M. D., B. J. Buratti, J. Nettles, M. Staid, J. Sunshine, C. M. Pieters, S. Besse, and J. Boardman (2011), A photometric function for analysis of lunar images in the visual and infrared based on Moon Mineralogy Mapper observations, J. Geophys. Res., doi:10.1029/2010je003733, in press. Hillier, J. K., B. J. Buratti, and K. Hill (1999), Multispectral photometry of the Moon and absolute calibration of the Clementine UV/Vis camera, Icarus, 141, 205 225, doi:10.1006/icar.1999.6184. Klima, R. L., M. D. Dyar, and C. M. Pieters (2011), Near infrared spectra of clinopyroxenes: Effects of calcium content and crystal structure, Meteorit. Planet. Sci., doi:10.1111/j.1945-5100.2010.01158.x, in press. Kramer, G. Y. (2010), Characterizing bedrock lithologies using small crater rim and ejecta probing (SCREP), Adv. Space Res., 45(10), 1257 1267, doi:10.1016/j.asr.2009.12.006. Kramer, G. Y., B. L. Jolliff, and C. R. Neal (2008), Distinguishing highalumina mare basalts using Clementine UVVIS and Lunar Prospector GRS data: Mare Moscoviense and Mare Nectaris, J. Geophys. Res., 113, E01002, doi:10.1029/2006je002860. Li, L., and J. F. Mustard (2000), Compositional gradients across marehighland contacts: Importance and geological implication of lateral transport, J. Geophys. Res., 105(E8), 20,431 20,450, doi:10.1029/ 1999JE001168. Lucey, P. G. (2004), Mineral maps of the Moon, Geophys. Res. Lett., 31, L08701, doi:10.1029/2003gl019406. Malaret, E., L. Perez, and H. Taylor (1998), Clementine s in flight calibration results (UVVIS camera), ACT s final report to NASA NASW 5014, Appl. Coherent Technol. Corp., Herndon, Va. McCord, T. B., and J. B. Adams (1973), Progress in remote optical analysis of lunar surface composition, Moon, 7, 453 474, doi:10.1007/ BF00564646. McEwen, A., E. Eliason, P. Lucey, E. Malaret, C. Pieters, M. Robinson, and T. Sucharski (1998), Summary of radiometric calibration and photometric normalization steps for the Clementine UVVIS images, Lunar Planet. Sci., XXIX, Abstract 1466. McEwen, A. S., and M. S. Robinson (1997), Mapping of the Moon by Clementine, Adv. Space Res., 19(10), 1523 1533, doi:10.1016/s0273-1177(97)00365-7. McEwen, A. S., M. S. Robinson, and R. A. Craddock (1996), A precise lunar photometric function, Proc. Lunar Planet. Sci. Conf., 27, 841 842(abstract). McKay, D. S., G. H. Heiken, A. Basu, G. Blanford, S. Simon, R. Reedy, B. M. French, and J. Papike (1991), The lunar regolith, in Lunar Sourcebook, pp. 285 356, Cambridge Univ. Press, New York. Melosh, H. J. (1989), Impact Cratering: A Geologic Process, Oxford Univ. Press, New York. Nozette, S., et al. (1994), The Clementine mission to the Moon: Scientific overview, Science, 266, 1835 1839, doi:10.1126/science.266.5192.1835. 10 of 11

Pieters, C. M. (1983), Strength of mineral absorption features in the transmitted component of near infrared reflected light first results from RELAB, J. Geophys. Res., 88, 9534 9544, doi:10.1029/jb088ib11p09534. Pieters, C. M. (1999) The Moon as a spectral calibration standard enabled by lunar samples: The Clementine example, paper presented at Workshop on New Views of the Moon II: Understanding the Moon Through the Integration of Diverse Datasets, Lunar and Planet. Inst., Flagstaff, Ariz. Pieters, C. M., and T. Hiroi (2004), RELAB (reflectance experiment laboratory): A NASA multiuser spectroscopy facility, Lunar Planet. Sci., XXXV, Abstract 1720. Pieters, C. M., M. I. Staid, E. M. Fischer, S. Tompkins, and G. He (1994), A sharper view of impact craters from Clementine data, Science, 266, 1844 1848, doi:10.1126/science.266.5192.1844. Shkuratov, Y. G., V. G. Kaydash, M. A. Kreslavsky, and N. V. Opanasenko (2001), Absolute calibration of the Clementine UVVIS data: Comparison with ground based observation of the Moon, Space Sci. Rev., 35(1), 29 34. Staid, M. I., and C. M. Pieters (2000), Integrated spectral analysis of mare soils and craters: Application to eastern nearside basalts, Icarus, 145, 122 139, doi:10.1006/icar.1999.6319. Taylor, L. A., C. M. Pieters, L. P. Keller, R. V. Morris, and D. S. McKay (2001), Lunar mare soils: Space weathering and the major effects of surface correlated nanophase Fe, J. Geophys. Res., 106(E11), 27,985 27,999, doi:10.1029/2000je001402. Tompkins,S.,andC.M.Pieters(2010),Spectralcharacteristicsof lunar impact melts and inferred mineralogy, Meteorit. Planet. Sci., 45, 1152 1169, doi:10.1111/j.1945-5100.2010.01074.x. S. Besse, Department of Astronomy, University of Maryland, College Park, MD 20742 2421, USA. J. Boardman, Analytical Imaging and Geophysics LLC, 4450 Arapahoe Ave., Ste. 100, Boulder, CO 80305, USA. R. N. Clark, U.S. Geological Survey, MS964, Box 25046, Federal Center, Denver, CO 80225, USA. J. P. Combe, G. Y. Kramer, and T. B. McCord, Bear Fight Institute, 22 Fiddler s Rd., Box 667, Winthrop, WA 98862, USA. (kramer@lpi. usra.edu) R. O. Green, Jet Propulsion Laboratory, 4800 Oak Grove Dr., MS 306 438, Pasadena, CA 91109 8099, USA. J. W. Head III, J. Nettles, and C. M. Pieters, Department of Geological Sciences, Brown University, Providence, Rhode Island, USA. E. Malaret, Applied Coherent Technology, 112 Elden St., Suite K, Herndon, VA 20170, USA. M. Staid, Planetary Science Institute, 1700 E. Fort Lowell, Ste. 106, Tucson, AZ 85719, USA. 11 of 11