Geology and composition of the Orientale Basin impact melt sheet

Similar documents
Igneous activity in the southern highlands of the Moon

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

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

The Importance of Impact Melts

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

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

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

The Surprising Lunar Maria

Amazing Orientale Peaks and Valleys

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

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

Distribution and modes of occurrence of lunar anorthosite

Compositional studies of the Orientale, Humorum, Nectaris,

UV-V-NIR Reflectance Spectroscopy

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

ASTRONOMY. Chapter 9 CRATERED WORLDS PowerPoint Image Slideshow

Composition of the Moon's Crust

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

An indigenous origin for the South Pole Aitken basin thorium anomaly

The Moon. Part II: Solar System. The Moon. A. Orbital Motion. The Moon s Orbit. Earth-Moon is a Binary Planet

Signature: GEOLOGICAL SCIENCES 0050

Lecture 11 Earth s Moon January 6d, 2014

High resolution measurements of absolute thorium abundances on the lunar surface

Little Learners Activity Guide

LUNAR OLIVINE EXPOSURES: ORIGINS AND MECHANISMS OF TRANSPORT

Wed. Oct. 04, Makeup lecture time? Will Friday noon work for everyone? No class Oct. 16, 18, 20?

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

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

I n the Lunar Magma Ocean (LMO) hypotheses1, KREEP materials are proposed to lie between the mafic mantle

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

Global spatial deconvolution of Lunar Prospector Th abundances

The Apollo 17 Landing Site. posted February 12, 1997

Geologic history of the Mead impact basin, Venus

Lunar Glossary. Note to the User: Glossary

The distribution and purity of anorthosite across the Orientale basin: New perspectives from Moon Mineralogy Mapper data

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

Student Guide to Moon 101

Examining the Terrestrial Planets (Chapter 20)

What is the Moon? A natural satellite One of more than 96 moons in our Solar System The only moon of the planet Earth

Italian Lunar Science Studies and Possible Missions a.k.a. The Moon: an Italian Approach. Angioletta Coradini Istituto Nazionale di Astrofisica

Thorium abundances of basalt ponds in South Pole Aitken basin: Insights into the composition and evolution of the far side lunar mantle

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

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

Feasibility Assessment of All Science Concepts within South Pole-Aitken Basin

The Moon: Internal Structure & Magma Ocean

10. Our Barren Moon. Moon Data (Table 10-1) Moon Data: Numbers. Moon Data: Special Features 1. The Moon As Seen From Earth

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

Moon Formation. Capture Hypothesis Many Hypothesis Fission Hypothesis Double Impact Hypothesis Giant Impact Hypothesis

Written by Linda M. V. Martel Hawai'i Institute of Geophysics and Planetology

Signature: Name: Banner ID#:

New Perspectives on the Lunar Cataclysm from Pre-4 Ga Impact Melt Breccia and Cratering Density Populations

Detection of Adsorbed Water and Hydroxyl on the Moon

Moon 101. By: Seacrest School Moon Crew Blake Werab David Prue

MARINER VENUS / MERCURY 1973 STATUS BULLETIN

Mon. Oct. 09, Reading: For Friday. Andrews-Hanna et al (GRAIL Procellarium region)

PUBLICATIONS. Journal of Geophysical Research: Planets. Differentiation of the South Pole Aitken basin impact melt sheet: Implications

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

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

Block: Igneous Rocks. From this list, select the terms which answer the following questions.

Lunar Cratering and Surface Composition

SUPPLEMENTARY INFORMATION

Photogeologic Mapping of Mars

9/15/16. Guiding Questions. Our Barren Moon. The Moon s Orbit

Volcanic Features and Processes

Dana Felberg Steven Hester David Nielsen Zach Weddle Jack Williams

Science Concept 1: The Bombardment History of the Inner Solar System is Uniquely Revealed on the Moon

grams grams Crystalline matrix Breccia

Engineering Geology ECIV 2204

Vital Statistics. The Moon. The Tides The gravitational pull between the Earth, the Moon and the Sun causes three inter-related effects: Lunar Phases

Our Barren Moon. Chapter Ten. Guiding Questions

A) B) C) D) 4. Which diagram below best represents the pattern of magnetic orientation in the seafloor on the west (left) side of the ocean ridge?

Lunar Field Exploration Science

Photogeologic Mapping of the Moon Instructor Notes

page - Lab 13 - Introduction to the Geology of the Terrestrial Planets

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

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

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

Section 7. Reading the Geologic History of Your Community. What Do You See? Think About It. Investigate. Learning Outcomes

IMPACT-INDUCED MELTING OF NEAR-SURFACE WATER ICE ON MARS

A New Moon for the Twenty-First Century

1 Describe the structure of the moon 2. Describe its surface features 3. Summarize the hypothesis of moon formation

Lunar Impact Basins: New Data for the Western Limb and Far Side (Orientale and South Pole-Aitken Basins) from the First Galileo Flyby

EARTH SCIENCE. Geology, the Environment and the Universe. Chapter 5: Igneous Rocks

Dept., Univ. of ela as are, 'i?ewark, DE Approximately 130 low specific gravity ((2.601, high silica

We will apply two of these principles to features on the lunar surface in the following activities.

Lunar Geology of Apollo 11 Landing Site. Chenango Forks High School Sharon Hartzell Sarah Maximowicz Benjamin Daniels Sarah Andrus Jackson Haskell

D) outer core B) 1300 C A) rigid mantle A) 2000 C B) density, temperature, and pressure increase D) stiffer mantle C) outer core

Evidence for Young Volcanism on Mercury from the Third MESSENGER Flyby

Moon and Mercury 3/8/07

Lunar Crater Activity - Teacher Pages

Geology of the Moscoviense Basin

Analogue Mission Simulations

I. Introduction: II. Background:

Chapter 21. The Moon and Mercury: Comparing Airless Worlds

Cratering and the Lunar Surface

Essentials of Geology, 11e

Chapter 4 Rocks & Igneous Rocks

Name Class Date STUDY GUIDE FOR CONTENT MASTERY

MACRORYTHMIC GABBRO TO GRANITE CYCLES OF CLAM COVE VINALHAVEN INTRUSION, MAINE

crater density: number of craters per unit area on a surface

Transcription:

JOURNAL OF GEOPHYSICAL RESEARCH: PLANETS, VOL. 119, 19 29, doi:10.1002/2013je004521, 2014 Geology and composition of the Orientale Basin impact melt sheet Paul D. Spudis, 1 Dayl J. P. Martin, 1,2 and Georgiana Kramer 1 Received 5 September 2013; revised 1 November 2013; accepted 21 November 2013; published 8 January 2014. [1] The Orientale Basin is one of the largest (930 km diameter) and youngest (~3.8 Ga) impact craters on the Moon. As the basin is only partly flooded by mare lava, its floor materials expose a major portion of the basin impact melt sheet, which some previous work has suggested might have undergone igneous differentiation. To test this idea, we remapped the geology of the Orientale Basin using images and topography from the Lunar Reconnaissance Orbiter, mineralogical information from the Chandrayaan-1 Moon Mineralogy Mapper, and elemental concentration maps from Clementine multispectral imaging and Lunar Prospector gamma ray data. The Maunder Formation (impact melt sheet of the basin) is uniform in chemical composition (equivalent to anorthositic norite ) in at least the upper 2 km of the deposit. The deepest sampling of the basin melt sheet (maximum depths of ~3 5 km by the crater Maunder, 55 km in diameter) shows a variety of lithologies, but these rock types (anorthosite, anorthositic norite melt rocks, mare basalt, and gabbro) are not those predicted by the differentiation model. We conclude that no differentiation of the Orientale Basin melt sheet has occurred and that such a process is not evident from new remote sensing data for the Moon or in the Apollo lunar samples. Citation: Spudis, P. D., D. J. P. Martin, and G. Kramer (2014), Geology and composition of the Orientale Basin impact melt sheet, J. Geophys. Res. Planets, 119, 19 29, doi:10.1002/2013je004521. 1. Introduction [2] As Orientale is the youngest and best preserved multiring impact basin on the Moon, it has long served as a prototype basin and has been studied for clues to the processes and histories of older, more degraded features [e.g., Head, 1974; Moore et al., 1974; McCauley, 1977; Wilhelms, 1987]. The most recent geological map of the Orientale Basin was made using data from the Lunar Orbiter and Zond 8 missions [Scott et al., 1978]. Recent missions such as the Lunar Reconnaissance Orbiter have provided high-resolution images of the surface as well as compositional and physical information on basin deposits, allowing detailed study and new understanding of lunar processes and history to be developed. [3] Large impact events produce copious volumes of shock melted, liquid rock, most of which lines the transient crater and ends up as the floor deposit of craters [Howard and Wilshire, 1975; Grieve et al., 1977; Cintala and Grieve, 1998]. It was recognized that the distinctive cracked and smooth deposits of the inner Orientale Basin (i.e., the Maunder Formation (abbreviated as Fm.) of McCauley [1977]) was probably an exposure of the impact melt sheet of the basin. New images and topographic data permit the full extent of the Maunder Fm. to be mapped and estimate of its thickness, volume, and composition to the made. Thus, 1 Lunar and Planetary Institute, Houston, Texas, USA. 2 School of Earth, Atmospheric and Environmental Sciences, University of Manchester, Manchester, UK. Corresponding author: P. D. Spudis, Lunar and Planetary Institute, 3600 Bay Area Blvd., Houston, TX 77058, USA. (spudis@lpi.usra.edu) 2013. American Geophysical Union. All Rights Reserved. 2169-9097/14/10.1002/2013JE004521 we can fully characterize the properties of a body of impact melt associated with the Orientale impact, a major event in lunar history [Wilhelms, 1987]. [4] Some cratering studies have emphasized the very large volumes of impact melt generated during basin-forming events [e.g., Cintala and Grieve, 1998]. In addition, the idea that large bodies of impact melt may cool slowly and undergo igneous differentiation is currently popular; this notion was encouraged by the hypothesis that the Sudbury Igneous Complex, a large igneous body associated with the terrestrial Sudbury Basin, is a differentiated impact melt sheet [Grieve et al., 1991]. This idea has been thought to be applicable to the largest basins of the Moon, including Orientale [Vaughan et al., 2013]. An early study of the compositional variation of the Orientale melt sheet by Budney et al. [1998] found no evidence for differentiation, but this work was done before the new spacecraft data were obtained, and the question has assumed renewed importance and urgency in light of ongoing efforts to model the differentiation of the Orientale melt sheet [Vaughan et al., 2013] and those of other large basins, such as South Pole-Aitken [e.g., Vaughan and Head, 2013]. [5] Because the Orientale Basin is sparsely filled with mare basalt, the original basin floor is widely exposed and offers the opportunity to study a well-preserved melt sheet of a major impact feature. Using morphology to geologically classify the basin deposits [e.g., Head, 1974; Moore et al., 1974; McCauley, 1977], we can recognize and make a new geological map of the impact melt sheet from images and estimate its extent and thickness. Moreover, using this map, we can directly determine its elemental composition using Clementine spectral reflectance and Lunar Prospector gamma ray maps [e.g., Lucey et al., 2000; Gillis et al., 2004]. 19

Figure 1. Orientale Basin interior with Maunder Fm. morphologies illustrated in the breakout images. M is the mare basalt fill of Mare Orientale, S is the smooth member, and R is the rough member of the Maunder Fm. The crater Maunder (55 km diameter; 14.6, 93.8 ) is illustrated by the arrow. Area of Montes Rook Fm. shown in Figure 4 is indicated by box. 1994; Bussey and Spudis, 1997] enabled the distinction of major compositional units of the basin. [8] Geological units were defined using characteristics such as position within the basin, surface texture, structure, and stratigraphic position. We adapted the unit names of the Orientale Group [McCauley, 1977], which defined a series of stratigraphic formations on the basis of their morphology and position. Stratigraphic formations were further subdivided into members [e.g., Wilhelms, 1987] on the basis of detailed surface textures and position. Although we are in the process of mapping the geology of the entire basin, for this effort, we focus only on basin interior deposits, particularly the inner basin [6] The aim of this work is to map the geology of the Orientale Basin using images and data from the Lunar Reconnaissance Orbiter Camera (LROC), the Moon Mineralogy Mapper (M3) of Chandrayaan-1, Clementine and Lunar Prospector (LP) FeO and TiO2 maps, and the Lunar Orbiter Laser Altimeter (LOLA) topographic data and to determine the extent of differentiation of the basin impact melt sheet (if any). The updated geological map shows a more accurate representation of the distribution of units both inside and outside of the basin; it also provides a clearer insight into the distribution of the melt sheet, the geological context of impact melt inside the basin, and the nature of the basin-forming impact. Determining if the Orientale Basin melt sheet has differentiated can help us understand the behavior of large, shallow magma bodies on the Moon and how this behavior might differ from impact melt on Earth. 2. Methods [7] The global image mosaics at 100 m/pixel of the LROC Wide Angle Camera (WAC) [Robinson et al., 2010] were used to create a new geological map of the Orientale Basin. An orthographic projection of this mosaic centered on Orientale ( 19, 93 ) was used as a base map, and units were rendered in ArcGIS 10.0. The GLD100 global topographic map (derived from LOLA and WAC stereo images) [Scholten et al., 2012] was used to aid in the identification and mapping of various units and to estimate the thickness of interior deposits. In addition, Clementine FeO and TiO2 maps [Lucey et al., 2000] and LP iron maps [Gillis et al., 2004] were used to help distinguish areas of mare material (high FeO content) from texturally similar highlands plains material of the basin melt sheet. Clementine RGB false color composites [Pieters et al., Figure 2. Geologic map of the inner Orientale Basin. The Maunder Fm. is largely restricted within the Outer Rook ring. Table 1 summarizes compositional data for these units. 20

Table 1. FeO and TiO 2 Concentrations for Orientale Basin Geologic Units (Figure 2) Geologic unit FeO wt % TiO 2 wt % Maria 11.0 ± 3.3 2.3 ± 1.4 Maunder Fm. (smooth member) 4.5 ± 1.9 0.6 ± 0.3 Maunder Fm. (rough member) 4.4 ± 2.0 0.6 ± 0.3 Montes Rook Fm. (knobby member) 4.6 ± 1.1 0.5 ± 0.1 Montes Rook Fm. (smooth member) 5.1 ± 1.1 0.5 ± 0.1 Massifs (average) 4.1 ± 1.6 0.5 ± 0.2 Massifs (anorthosite, average of eight) 0.6 ± 0.5 0.3 ± 0.8 Maunder Formation [McCauley, 1977], which has been convincingly interpreted by most workers as a remnant of the basin impact melt sheet [Head, 1974; Moore et al., 1974; McCauley, 1977;Wilhelms, 1987]. [9] The new geological map of the basin interior was used to create a stencil to extract compositional information of a given unit. The stencil and compositional image are registered in ArcGIS and statistics collected. Thus, we are able to estimate the chemical and mineralogical makeup of the surface of the impact melt sheet of the Orientale Basin. In addition to this information, we mapped the position and extent of the ejecta from over 300 superposed impact craters on the Maunder Fm. Using this stencil, we are able to collect information on the composition of the ejecta from these craters, which are derived from beneath the surface by varying amounts (depths up to 5 km). By this method, we use impact craters as natural drill holes to study possible compositional variations in vertical section. [10] In addition to compositional data from Clementine and LP, the M 3 obtained spectral data for the deposits of the Orientale Basin [Green et al., 2011], from which we can infer the dominant mineralogy [Pieters et al., 2008]. We have examined M 3 data for several fresh craters on the Maunder Fm. and in particular, the deposits of the crater Maunder. At 55 km diameter, Maunder is the largest postbasin impact crater in the interior of Orientale (Figure 1) and hence a source of the deepest stratum (~3 5 km depth) of melt sheet lithologies. When information on the dominant mineralogy is complemented with estimates of chemical composition, we can infer the composition of mapped geological units in terms of known lunar rock types. As the computational models of differentiation make specific predictions about rock types as a function of depth in the melt sheet [Vaughan et al., 2013], we use this information to test those models. 3. Geology of the Basin Interior [11] Orientale is only partly filled by mare basalt, so its original floor configuration can be clearly seen over most of the basin interior (Figure 1). We have remapped the geology of the basin interior utilizing the stratigraphic nomenclature of McCauley [1977], which has been largely unchanged except that some formations have been subdivided into members on the basis of surface texture. Several geological formations collectively make up the Orientale Group; in the interior of the basin, these formations are the Maunder, Montes Rook, and (of extremely limited extent here) Hevelius Formations. In addition to these formally named units, we also map two additional (informal) rock stratigraphic units: mare and massif material (Figure 2). The Orientale Basin interior displays a number of small melt ponds that were previously unmapped (and partly revealed by comparison with the Clementine FeO map) [Lucey et al., 2000]. [12] The mare basalts of Orientale have been mapped and described in detail by previous workers [Head et al., 1993; Gillis, 1998; Whitten et al., 2011]. The basalts thinly cover the innermost center of the basin and occur as small patches at the base of and concentric to major rings. Submare Maunder Fm. crops out in central Mare Orientale as small kipukas (Figure 2) and as rim material of the crater Hohmann (16 km diameter; 17.9, 94.1 ), indicating that the basalts here must thin to a featheredge. The nearly pure basaltic composition of ejecta from the crater Il in (13 km diameter; 17.8, 97.5 ) indicates that the mare basalts here must be on the order of at least 1 km thick. Thus, the inner basin floor is highly irregular in relief, displaying a wide range of basalt thicknesses. The basalts of Mare Orientale appear to be moderate in titanium content (~2.3 wt %, Table 1, but see also Whitten et al. [2011]), which is relatively low Ti compared to the Apollo samples but higher than the Ti content of other typical farside maria [Gillis, 1998]. [13] The other informal rock unit is made up of the massifs (mountains) of the inner basin rings (Figure 2). Massifs typically make up the two intermediate rings (i.e., the Inner and Outer Montes Rook). The other two basin rings (inner shelf ring and Montes Cordillera; see below) are more scarp like. Massifs can be as large as 20 30 km across, are typically equant in plan, cluster in some locales and occur as isolated peaks in others. Parts (but not all) of the Inner Rook ring are composed of massifs made up of pure anorthosite [Spudis et al., 1984; Bussey and Spudis, 1997; Hawke et al., 2003; Cheek et al., 2013]. In some cases, these anorthosites are shocked to levels of at least 20 GPa [Spudis et al., 1984], but less than 30 GPa, as evidenced by the presence of the 1250 nm plagioclase absorption feature [Cheek et al., 2013]. [14] The Orientale Basin interior displays at least four distinct concentric rings (Figure 3) [Hartmann and Wood, 1971; Moore et al., 1974; Wilhelms, 1987]. The innermost ring (320 km diameter) is expressed as a simple scarp covered by the Maunder Fm., with up to 3 km of relief between the upper shelf and the flat, mare-covered floor of the basin. Figure 3. Ring configuration of the inner basin. Three prominent rings occur within the basin topographic rim and the fourth makes up that rim. GLD100 topographic image. 21

Figure 4. Flow lobe within the Montes Rook Fm. (arrow) suggesting that impact melt makes up at least part of this unit. Montes Rook Fm. laps up onto Cordillera scarp here in southwestern corner of Orientale Basin. Field of view is about 200 km across; LROC WAC mosaic. The next larger ring (480 km diameter) is the Inner Montes Rook, an irregular but broadly circular arrangement of equant massifs and massif clusters. Some of these massifs are composed of nearly pure anorthosite [Spudis et al., 1984; Hawke et al., 2003; Cheek et al., 2013]. The next larger ring is the Outer Montes Rook (620 km diameter), again made up of massive, blocky mountains, roughly arranged in a circular pattern. The outer ring (and main topographic rim of the basin) is the Montes Cordillera (930 km diameter), which has the morphology of a gigantic scarp. In most cases, the rings demark the limit of exposure of units of the inner Orientale Group (e.g., the Montes Rook Fm. is largely confined between the Cordillera and Outer Rook rings) but exceptions occur locally (Figure 2). [15] The Maunder Formation (Figure 1) has been split into two members a smooth (plains-forming) member and a rough member (Figure 2). The Maunder smooth member is relatively flat and appears to occur within topographic lows. The Maunder rough member has a relatively large range of surface relief and appears to be draped over the preexisting topography; it is extensively fractured in some areas (termed corrugated facies by Head [1974]). These cracks may have been created as the melt sheet settled and cooled over the rugged parts of the basin floor. On the basis of their position with respect to the basin rim and their resemblance to the floor deposits of fresh complex craters such as Tycho [e.g., Howard and Wilshire, 1975], both members of the Maunder Fm. are interpreted as the remnant of the impact melt sheet of the basin [McCauley, 1977; Wilhelms, 1987]. The smooth plains member likely represents small local areas where the melt has ponded and filled in hollows and small valleys in the floor. [16] Moving outward from the basin center, the next unit is the Montes Rook Formation (Figure 2). This unit has also been split into two members a rough, knobby member ( domical facies of Head [1974]; knobby terrain of Moore et al. [1974]) and a smoother, plains-like member. The knobby member appears to contain large, equant, bulbous hummocks of material giving these areas a blocky and uneven appearance. The Montes Rook Fm. is similar in appearance to the Alpes Fm. of the nearside Imbrium Basin [Wilhelms, 1987], but unlike the Alpes, the Montes Rook Fm. is largely restricted to the Orientale Basin interior, with minor extensions beyond the Cordillera Basin rim crest. Flow lobes within the Montes Rook Fm. have been observed against some of the massifs or against the Cordillera Ring in some areas (Figure 4), suggesting some type of fluid movement of material during late stages of emplacement. The smooth member is mainly concentrated in the southwestern quadrant of the basin interior. [17] The origin of the knobby surface texture displayed by the Montes Rook Fm. is unknown. Head [1974] suggested that seismic shaking during megaterrace formation of the Cordillera ring induced a domical morphology to the unit. Moore et al. [1974] equated the Montes Rook Fm. with the Maunder Fm. and suggested that both units were largely composed of impact melt created during the basin-forming event. McCauley [1977] proposed that the knobs are the surface manifestation of coherent, ejecta megablocks quarried from deep stratigraphic horizons. None of these suggestions Figure 5. (left) The Inner Montes Rook ring (480 km diameter) and the topographic relief along its circumference (from A clockwise to B). (right) Relief is up to 6.5 km from highest to lowest segments; as Maunder Fm. partly covers this ring, this value may be taken as an approximate limit to the thickness of the basin melt sheet. 22

Figure 6. Geologic stencil, WAC mosaic, and Clementine FeO image used to derive the elemental composition of the Maunder Formation. 500 600 km in diameter and has an average thickness of ~6 km or less, possibly thickening locally to 8 9 km in the center of the basin. A melt sheet of this extent and thickness would have a volume of ~8.5 105 km3, a value roughly comparable to the independent estimates of one million cubic kilometers by Potter et al. [2013] and Vaughan et al. [2013]. This estimate may represent a minimum value for the total amount of melt generated in the Orientale Basin-forming impact, as significant amounts of impact melt would be ejected as crater size increases [Cintala and Grieve, 1998]. explains the widespread occurrence of this surface texture across much of the lunar nearside [Spudis et al., 2011]. The idea that the Montes Rook Fm. is composed of impact melt at least in part is supported by the presence of the flow lobes in the southwestern corner of the basin interior (Figure 4). [18] The basin exterior deposits are collectively named the Hevelius Formation [McCauley, 1977; Wilhelms, 1987]. This unit displays a variety of morphologies, including radially lineated, swirl-like, transverse (to the basin rim), and hummocky. At ranges beyond about one half to one basin radii, the continuous Hevelius Fm. transitions into a set of discontinuous units, including smooth, Cayley-like [Wilhelms, 1987] plains and abundant secondary impact craters, including singles, open clusters, and sequential chains oriented radial to the basin center. As our focus in this study is on the basin impact melt sheet, we will defer detailed discussion of the morphology and distribution of the Hevelius Fm. for a later contribution. The vast bulk of the Hevelius Fm. occurs beyond the Cordillera rim, but we have identified a few small occurrences of this material inside the basin rim (Figure 2). [19] Vaughan et al. [2013] used LOLA topography and thermal model considerations to estimate the thickness of the Orientale Basin melt sheet (i.e., the Maunder Fm.). They estimated an initial thickness of about 15 km near the basin center. As an alternative, we examined the topographic relief concentrically within the Inner Montes Rook ring (Figure 5), a feature that is partly embayed by the melt sheet. If the exposed western section of the Inner Montes Rook is representative of the currently buried eastern segments (which seems likely), then the maximum amount of relief within this exposed portion of the ring represents the likely maximum thickness of the melt sheet at this radial position. The observed maximum relief within the exposed Inner Montes Rook ring is about 6.3 km. [20] As the Maunder Fm. sometimes completely covers the Inner Montes Rook ring (near the southeastern sections; Figure 1) and sometimes laps up against it, we suggest that the thickness of the melt sheet is no greater than about 6 km. The thickness of the melt sheet in the basin center cannot be directly determined but is not likely to be significantly greater than this estimate plus the elevation difference between the Maunder shelf and basin center, about 2 km [Vaughan et al., 2013]. Thus, we suggest that the Maunder Fm. is about 4. Composition of the Maunder Formation [21] The new geological mapping (Figure 1) described above can be used to determine the chemical composition of the Orientale Basin melt sheet. Using the map of the extent of the Maunder Fm., a stencil was created to mask out data of non-maunder Fm. units. The remaining pixels represent the Figure 7. Iron concentrations of the Maunder Fm. surface derived from geologic map stencil over the Clementine FeO image. Curve is well-behaved, narrow peak with mean FeO of 4.4 ± 2.0 wt %, indicating a very uniform composition. See Table 1 for details. 23

samples of the Maunder Fm., yielding its surface composition and variations (Figures 6 and 7). Both FeO and TiO2 concentration maps were prepared from the Clementine spectral reflectance images using the algorithms of Lucey et al. [2000] and Gillis et al. [2004]. Results of this analysis are given in Table 1, where mean iron and titanium values are given for all inner basin geological units. The iron content of the surface of the Maunder Fm. is very uniform (Figure 7), at around 4.5 wt % FeO and 0.6 wt % TiO2. Low-Ca pyroxene dominates the spectra observed in M3 data in fresh exposures of the Maunder Fm., but as pyroxene is a spectrally dominant mineral, its influence on a reflectance spectrum is not proportional to its petrologic modal abundance. These characteristics lead to the conclusion that the surface of the Maunder Fm. is made up of anorthositic norite (Al2O3 ~28 wt %, estimated from average FeO values) [see Spudis et al., 2002; Zellner et al., 2002], a rock type common to the upper crustal composition of the lunar highlands. A chemical and mineralogical analog to this composition in the Apollo collections are the samples 68415 and 68416, impact melts collected from Station 8 at the Apollo 16 landing site [Meyer, 2010]. [22] While the surface composition of the Maunder Fm. is very uniform (Figure 7), is this composition representative of the unit as a whole? Because of its great age (older than 3.8 Ga), the Orientale Basin has a large number of impact craters superposed on its units. We mapped the ejecta deposits of over 300 small (0.5 18 km diameter) impact craters on the Maunder Fm. This map was used to create another stencil so that we could assess the variation in iron and titanium composition of the melt sheet with depth. Results are shown in Figure 8. These plots show that the composition of the Maunder Fm. is remarkably uniform and homogeneous down to depths of almost 2 km. Virtually all compositions fall within one standard deviation of the mean value of the surface composition except for a population at the smallest Figure 8. Composition of the ejecta deposits of craters on Maunder Fm. Solid horizontal lines are the mean values of the surface; shaded areas are one standard deviation. With increasing crater size, deeper stratigraphic horizons are excavated. Yet the largest craters (excavating the deepest) show no significant differences in composition with the surface of the melt sheet, suggesting that there are no changes in composition to depths of about 2 km. Figure 9. Maunder (55 km diameter; 14.6, 93.8 ), an Eratosthenian complex crater whose impact target was the Maunder Fm. overlying the inner basin shelf ring. The LROC WAC image shows morphology, the M3 composite image (red = band depth at 950 nm, green = band depth at 1050 nm, and blue = band depth at 1250 nm) displays mineralogy, and the Clementine iron map shows FeO weight concentrations (colors as in Figure 6). In the M3 composite image, the yellow/orange colors indicate pyroxene-rich lithologies, blue is crystalline plagioclase, and white (small fresh craters, mostly on the SW side) is olivine-rich. Note the bilateral symmetry in ejecta, with mafic ejecta in the SW and NE quadrants and feldspathic ejecta in the NW and SE sectors. 24

Figure 10. Geologic map of Maunder Crater. Units are recognized by morphology and position; ejecta deposits have been subdivided according to FeO content. Mineralogy is estimated from extracted M 3 spectra of fresh small (~100 m) craters superposed on a given unit. See Table 2 for unit compositions and interpretations. diameters, which seem to be contaminated with a fraction of more mafic members. As these points show both enhanced iron and titanium, they are probably caused by the inclusion of near-surface mare basalt flows and feeder dikes related to the emplacement of the lavas of Mare Orientale. The fact that these mafic anomalies are restricted to the smallest crater diameters suggests that we were not totally successful in restricting our geological mapping solely to the Maunder Fm. at the smallest scales, probably caused by the moderate resolution (~100 m/pixel) of the images used for mapping. [23] The largest craters sampled were approximately 18 km across and excavated material from ~1.5 2 km depth within the melt sheet [Croft, 1980, 1985]. As there is no apparent correlation between crater size (and hence depth of excavation) and elemental composition, we conclude that the upper ~2 km of the Orientale Basin impact melt sheet is chemically homogeneous and identical to the surface composition. A recent study that modeled the crystallization of the Orientale melt body produced a hypothetical melt sheet stratigraphy of ~1 km of anorthosite debris overlying a layer of norite that extends to depths of about 4 km [Vaughan et al., 2013]. Such composition and structure are not observed in our data. We find evidence for spatially isolated outcrops of pure anorthosite (i.e., FeO content <1 wt %) only within some isolated massifs of the Inner Rook basin ring and in a few small impact craters near this ring. But the Maunder Fm. itself is not anorthosite but anorthositic norite (FeO ~4.5 wt %), and this composition manifests itself to depths of at least 2 km. 5. Composition of the Deposits of Maunder Crater [24] The crater Maunder (55 km diameter; 14.6, 93.8 ) is the largest postbasin crater whose target consisted of inner basin deposits, in this case, the Maunder Fm. and part of the inner basin (scarp) ring (Figures 9 and 10). As such, the composition of its ejecta, rim deposits, melt sheet, and central peak can provide critical information about the composition and nature of the Orientale melt sheet at depth. Thus, we undertook a special study of this crater and its geology to better understand the nature of the interior of the basin melt sheet. [25] Maunder is a complex crater, showing a flat floor and central peak, and is of Late Eratosthenian age [Scott et al., 1978; Wilhelms, 1987]. On the basis of its diameter and morphology, the excavation cavity of the crater was on the order of 30 35 km diameter and maximum depth of excavation would be a bit more than one tenth that number (3 4 km) [Croft, 1980, 1985]. Central peaks of complex craters represent slightly deeper stratigraphic horizons, with the Maunder central peak coming from roughly 5 5.5 km depth [Cintala and Grieve, 1998]. Thus, Maunder would have both excavated and brought up through structural uplift the deepest stratigraphic horizons of the melt sheet. If the hypothesized mafic cumulates exist, we might expect them to make up part of the Maunder deposits. [26] The spectral parameters used to generate the M 3 color composite in Figure 9 (middle image) are the depths of the absorption bands at 950 nm (red), 1050 nm (green), and 1250 nm (blue). The position of the mafic absorption band is indicative of mineralogy and mineral composition, and these parameters were selected to exploit that fact. Low-Ca pyroxene has a maximum absorption near 950 nm. The wavelength of maximum absorption increases with increasing Ca to near 1050 nm. Therefore, orange and yellow regions in Figure 9 (middle) show locations of low- to high-ca pyroxene-rich lithologies. Crystalline anorthosite (>90% plagioclase) containing minor amounts of FeO (at least ~0.5 wt %) has a diagnostic absorption centered near 1250 nm. Bright blue areas in Figure 9 (middle) show the locations of anorthosite exposures. Three closely spaced absorptions in olivine overlap to appear as one broad absorption band centered near 1100 nm [Burns, 1993; Sunshine and Pieters, 1998] but still absorb a significant amount light at 950 and 1250 nm. Therefore, white and pale yellow in Figure 9 (middle) represent olivine-rich material. Figure 11. Representative spectra of the principal geological units of Maunder Crater; numbers and colors correspond to units shown in mineral key at top right of Figure 10. 25

Table 2. Average FeO and TiO 2 Contents, Mineralogy (lpx = Low-Ca Pyroxene, cpx = Clinopyroxene, ol = Olivine, and plg = Plagioclase), and Interpreted Lithology of Maunder Crater Deposits a Unit FeO (wt %) TiO 2 (wt %) Dominant Mineral Interpretation Central peak 9.7 ± 1.6 0.6 ± 0.2 lpx, cpx norite, gabbro Crater floor 9.3 ± 1.4 0.8 ± 0.1 mafic glass basaltic impact melt Wall materials 4.5 ± 1.6 0.6 ± 0.1 lpx, plg norite, anorthosite, basin melt sheet Feldspathic ejecta 4.7 ± 1.5 0.6 ± 0.2 plg basin melt sheet rocks, feldspathic massif materials Mafic ejecta 8.3 ± 2.2 1.2 ± 0.4 ol, mafic glass olivine vitrophyre mare basalt plus basin melt sheet debris Distal ejecta 6.0 ± 2.3 0.9 ± 0.4 cpx, ol basin melt sheet debris on basalts of Mare Orientale a See geologic map of Figure 10 for location of units. [27] A geological map of Maunder is shown in Figure 10. We used M 3 data to examine the spectral properties of Maunder units [Whitten et al., 2011] and extracted point spectra for small, fresh craters superposed on Maunder units (Figures 10 and 11, color symbols and spectra). From Clementine images, the iron and titanium concentrations were determined for each mapped unit and used in conjunction with mineral information to interpret the lithological makeup of Maunder materials (Table 2). It is important to note that M 3 data alone do not uniquely determine lithology, although they can often be used to infer the composition of the spectrally dominant mineral [e.g., Burns, 1993; King and Ridley, 1987; Sunshine and Pieters, 1998; Isaacson et al., 2011; Klima et al., 2011]. Likewise, geochemical information alone cannot uniquely determine what rock types are present. However, both sources of information together can distinguish between contrasting (and drastically different) alternative lithologic interpretations. For example, an olivine-bearing unit containing FeO ~10 wt % could be dunite. However, if that unit also contains 1.2 wt % TiO 2, it is more likely to be made up of olivine-rich mare basalts mixed with highland materials. While the latter is not a unique determination, it is more geologically reasonable than a dunite with an anomalously high abundance of titanium. [28] From the mapping and compositional data presented here (Figures 10 and 11 and Table 2), we suggest the following interpretation of the geology of Maunder Crater. The impact occurred on Maunder Fm. (basin melt sheet), which here overlies the scarp of the innermost basin ring. Two principal lithologies make up the deep subsurface (>2 km depth) of the melt target here: a large norite pluton or massif striking NW-SE and intrusive bodies of mare basalt (or gabbro), likely of the same composition as the surface flows of Mare Orientale and possibly making up feeder dikes for those lavas. These bodies are overlain by basin impact melt, which has a composition of anorthositic norite, as described above. These rock types were excavated during the formation of Maunder Crater, creating the heterogeneous and mixed ejecta blanket that surrounds the feature. The central peak is made up mostly of norite and lesser amounts of gabbro, suggesting a complex petrologic structure at depth. Floor material (impact melt sheet of Maunder Crater) is a highland basalt (FeO ~10 wt %) displaying spectra indicative of a mafic glass. It is probably Figure 12. Comparison of the reconstructed stratigraphy of the Maunder Fm. of Orientale Basin in the case of simple equilibrium crystallization in the large melt sea of Vaughan et al. [2013] and inferred subsurface structure derived from the mapping and analysis of this study. 26

similar to the basaltic impact melts ubiquitous in the Apollo collections; although its KREEP content is unknown, it is probably low. There is no evidence in any of the crater deposits for ultramafic rock types such as pyroxenite and dunite; if such products exist, they must occur at depths greater than 5 6 km. The crater exterior ejecta displays two principal compositions: a mafic, olivine-bearing unit that probably consists of a mixture of fragments of the basin melt sheet (feldspathic) and mare basalt and another more feldspathic unit, largely composed of basin melt sheet fragments with minor blocks of anorthosite (Figures 9 and 10). A reconstruction of the inferred subsurface stratigraphy of the Maunder Fm. and comparison to the configuration proposed in Vaughan et al. [2013] are shown in Figure 12. 6. Discussion [29] Our studies of the Maunder Fm. of the Orientale Basin have shown that although partly covered by mare basalt, the impact melt sheet is exposed and widespread in the center of the basin, with an estimated maximum thickness of ~6 km. It has a remarkably uniform surface composition (Figure 7) and this homogeneity continues to depths of at least 2 km (Figure 8). Its bulk composition is significantly more feldspathic than norite, corresponding to an anorthositic norite (Al 2 O 3 ~ 28 wt %). Our one data point for deeper stratigraphic levels is provided by the deposits of Maunder Crater, which show a complex association of norite, gabbro, mare basalt, and feldspathic basin melt rocks. [30] The idea that large bodies of impact melt might differentiate in place has gained some currency in lunar science [e.g., Cintala and Grieve, 1998; Morrison, 1998; Vaughan et al., 2013]. This supposition is based on analogy to the terrestrial Sudbury Igneous Complex, which has been hypothesized to represent a differentiated impact melt sheet [Grieve et al., 1991]. Large melt sheets supposedly contain lesser amounts of cold clasts relative to the impact melt sheets of smaller craters and hence would retain more heat for long periods of time [Cintala and Grieve, 1998]. Previous studies of the melt sheets at other terrestrial craters [Grieve et al., 1977; Floran et al., 1978] and lunar samples [Ryder and Wood, 1977] led to a different understanding that impact melts mix energetically during crater formation and thoroughly homogenize diverse target lithologies into a very uniform, blended melt composition. (A corollary of this idea is that chemically defined groups of impact melts must have been derived from distinct impact events; see Spudis [1993, p. 172 180].) Rapid cooling is precipitated by the physical incorporation of abundant unmelted clasts as the crater grows, which are intimately mixed into the melt during cratering flow [Simonds et al., 1976]. While the impact melt remains mobile enough to settle, drape topography, pond, and partly flow for short distances, it cools relatively quickly, resulting in the production of fine-grained, clast-laden, chemically uniform melt rocks [Simonds et al., 1976; Spudis and Ryder, 1981]. Which (if either) of these two scenarios are correct and under what conditions? [31] A recent model of the differentiation of an Orientale Basin-sized melt body produced an igneous body having a stratigraphy of norite on top (8 km thick) overlying pyroxenite (4 km) and basal dunite (2 km) [Vaughan et al., 2013]. Although the upper surface of the Orientale melt sheet is anorthositic norite (significantly more feldspathic than norite), there is little evidence for the presence of the other predicted mafic compositions at depth (Table 2). Given the estimated initial thickness of the melt sheet which ranges from 6 km (this study) up to 15 km [Vaughan et al., 2013] at least some level of differentiation would have occurred by analogy with the largest impact melt sheets on Earth (e.g., the Sudbury Igneous Complex; Grieve et al. [1991]). The results of this study indicate no differentiation of the Orientale melt sheet along the lines predicted by the Vaughan et al. [2013] modeling. Although some lithologic diversity appears to be present in the subsurface beneath the crater Maunder, it is not the diversity predicted by the differentiation model but is consistent with a complex and heterogeneous basin floor overlain by a homogeneous melt sheet. [32] In a more philosophical vein, this work calls into question the whole idea of the igneous differentiation of lunar impact melt sheets. At nearly 1000 km diameter, Orientale is one of the largest impact basins on the Moon [Wilhelms, 1987]. As such, it would be expected, at least to some degree, to manifest the features and phenomena characteristic of the biggest impacts. The lack of evidence for differentiation of its impact melt sheet calls into question assumptions that the impact melt produced by the largest lunar impact crater, the South Pole-Aitken Basin, underwent differentiation [e.g., Morrison, 1998; Vaughan et al., 2013]. A lack of such differentiation is supported by some interpretations of remote sensing data [e.g., Pieters et al., 2001; Nakamura et al., 2009]. [33] To look at it from another perspective, if the melt sheets in lunar basins undergo widespread differentiation, what is the origin of the ubiquitous fine-grained basaltic impact melts in the Apollo collection? These rocks occur at every landing site and in many lunar meteorites, contain clasts of at least middle (if not deep) crustal origin (i.e., nonsurficial material), all date from the time of heavy bombardment (circa 3.9 Ga), have distinctive meteoritic siderophile signatures, and are found proximate to the latest and largest impact basins of the lunar near side [e.g., Spudis, 1993]. In bulk composition, they are more mafic than at least the upper few kilometers of the lunar crust, so they cannot be derived from the melt sheets of typical highland craters. These facts seem to point to an origin in a large (basin-forming) impact for most of these rocks, even if the basin to which each rock belongs is questioned [cf. Spudis and Ryder, 1981;Haskin, 1998]. As the plutonic rocks in the lunar samples appear to be of endogenous origin (most have crystallization ages (~4.5 to 4.1 Ga) prior to the ages youngest basins and extremely low meteoritic siderophile element concentrations), we have no obvious candidates for melt sheet differentiates in the lunar samples. If the basaltic melt rocks are ejecta, they would say nothing about the possibility of melt sheet differentiation, except that the starting compositions of such complexes might be quite different from the ones currently assumed by existing modeling [e.g., Vaughan et al., 2013]. [34] Although the idea of differentiation of impact melt sheets is intriguing, at present there is no direct evidence that such a process occurs on the Moon. It would require multiple sample returns from a complex area, such as the floor of South Pole-Aitken Basin, to establish this process as an important contributor to lunar surface evolution. Melt sheet differentiation is inconsistent with the geology and compositions of the Orientale Basin and with the characteristics of the lunar sample collection. 27

7. Conclusions [35] An updated geological map of the Orientale Basin shows the relations of basin units and has been used to study the variation in composition of the basin s impact melt sheet. We find that the Maunder Fm. (basin melt sheet) varies from 1 6 km thick and has a near-uniform composition of anorthositic norite (FeO ~4.5 wt %). This composition does not change over the upper 2 km of the unit. Deeper horizons of the Maunder Fm. in at least one location show a greater diversity of composition but include lithologies neither expected from nor consistent with impact melt sheet differentiation. Results of this study confirm predictions of the previous model of rapidly quenched impact melt of uniform chemical composition. Study of the new data returned by the fleet of exploration missions over the last few years have greatly increased our ability to address these and other complex questions in lunar science. [36] Acknowledgments. The work of DJPM is partly supported by the LPI Summer Intern Program. We thank A. Fagan, N. Petro, and B. Sharpton for review and comment on an earlier version of this paper. This is Lunar and Planetary Institute contribution 1764. This work is partly supported by NASA Lunar Science Institute contract NNA09DB33A (PI David A. Kring). References Budney, C. J., G. J. Taylor, and P. G. Lucey (1998), Did the Orientale melt sheet differentiate?, Lunar Planet. Sci., 29, 1537. Burns, R. G. (1993), Mineralogical Applications of Crystal Field Theory, Cambridge Univ. Press, Cambridge UK, doi:10.1017/cbo9780511524899. Bussey, D. B. J., and P. D. Spudis (1997), Compositional analysis of the Orientale Basin using full resolution Clementine data: Some preliminary results, Geophys. Res. Lett., 24, 445 448. Cheek, L. C., K. L. Donaldson Hanna, C. M. Pieters, J. W. Head, and J. L. Whitten (2013), Distribution and purity of anorthosite across the Orientale Basin: New perspectives from Moon Mineralogy Mapper data, J. Geophys. Res. Planets, 118, 1805 1820, doi:10.1002/jgre.20126. Cintala, M. J., and R. A. F. Grieve (1998), Scaling impact melting and crater dimensions: Implications for the lunar cratering record, Meteorol. Planet. Sci., 33, 889 912. Croft, S. K. (1980), Cratering flow fields: Implications for the excavation and transient expansion stages of crater formation, Proc. Lunar Planet. Sci., 11, 2347 2378. Croft, S. K. (1985), The scaling of complex craters. Proceedings, lunar and planetary science conference 15, J. Geophys. Res., 90, C828 C842. Floran, R. J., R. A. F. Grieve, W. C. Phinney, J. L. Warner, C. H. Simonds, D. P. Blanchard, and M. R. Dence (1978), Manicouagan impact melt, Quebec, 1, Stratigraphy, petrology and chemistry, J. Geophys. Res., 83(B6), 2737 2759. Gillis, J. J. (1998), The composition and geologic setting of mare deposits on the far side of the Moon, PhD dissertation, Rice Univ., Houston TX, 257 pp. Gillis, J. J., B. L. Jolliff, and R. L. Korotev (2004), Lunar surface geochemistry: Global concentrations of Th, K, and FeO as derived from Lunar Prospector and Clementine data, Geochim. Cosmochim. Acta, 68(18), 3791 3805. Green, R. O., et al. (2011), The Moon Mineralogy Mapper (M 3 ) imaging spectrometer for lunar science: Instrument description, calibration, onorbit measurements, science data calibration and on-orbit validation, J. Geophys. Res., 116, E00G19, doi:10.1029/2011je003797. Grieve, R. A. F., M. R. Dence, and P. B. Robertson (1977), Cratering processes: As interpreted from the occurrence of impact melts, in Impact and Explosion Cratering, edited by D. J. Roddy, R. O. Pepin, and R. B. Merrill, pp. 791 814, Pergamon Press, NY. Grieve, R. A. F., D. Stöffler, and A. Deutsch (1991), The Sudbury structure: Controversial or misunderstood?, J. Geophys. Res., 96(E5), 22,753 22,764. Hartmann, W. K., and C. A. Wood (1971), Moon: Origin and evolution of multi-ring basins, Moon, 3, 3 78. Haskin, L. A. (1998), The Imbrium impact event and the thorium distribution at the lunar highlands surface, J. Geophys. Res., 103(E1), 1679 1689. Hawke, B. R., C. A. Peterson, D. T. Blewett, D. B. J. Bussey, P. G. Lucey, G. J. Taylor, and P. D. Spudis (2003), Distribution and modes of occurrence of lunar anorthosite, J. Geophys. Res., 108(E6), 5050, doi:10.1029/2002je001890. Head, J. W. (1974), Orientale multi-ring basin interior and implications for the petrogenesis of lunar highland samples, Moon, 11, 327 356. Head, J. W., S. Murchie, J. F. Mustard, C. M. Pieters, G. Neukum, A. McEwen, R. Greeley, E. Nagel, and M. J. S. Belton (1993), Lunar impact basins: New data from the western limb and far side (Orientale and South Pole-Aitken Basins) from the first Galileo flyby, J. Geophys. Res., 98, 17,149 17,181. Howard, K. A., and H. G. Wilshire (1975), Flows of impact melt in lunar craters, J. Res. U.S. Geol. Surv., 3, 237 251. Isaacson, P. J., et al. (2011), Remote compositional analysis of lunar olivinerich lithologies with Moon Mineralogy Mapper (M3) spectra, J. Geophys. Res., 116, E00G11, doi:10.1029/2010je003731. King, T. V. V., and W. I. Ridley (1987), Relation of the spectroscopic reflectance of olivine to mineral chemistry and some remote sensing implications, J. Geophys. Res., 92, 11,457 11,469, doi:10.1029/ JB092iB11p11457. Klima, R. L., et al. (2011), New insights into lunar petrology: Distribution and composition of prominent low-ca pyroxene exposures as observed by the Moon Mineralogy Mapper (M3), J. Geophys. Res., 116, E00G06, doi:10.1029/2010je003719. Lucey, P. G., D. T. Blewett, and B. L. Jolliff (2000), Lunar iron and titanium abundance algorithms based on final processing of Clementine ultraviolet visible images, J. Geophys. Res., 105(E8), 20,297 20,305. McCauley, J. F. (1977), Orientale and Caloris, Phys. Earth Planet. Inter., 15, 220 250. Meyer, C. (2010), Sample 68415 and 68416: Basaltic impact melt. Lunar Sample Compendium, NASA JSC, http://curator.jsc.nasa.gov/lunar/ lsc/68415.pdf. Moore, H. J., C. A. Hodges, and D. H. Scott (1974), Multiringed basins illustrated by Orientale and associated features, Proc. Lunar Sci. Conf., 5, 71 100. Morrison, D. A. (1998), Did a thick South Pole-Aitken Basin melt sheet differentiate to form cumulates?, Lunar Planet. Sci., 29, 1657. Nakamura, R., et al. (2009), Ultramafic impact melt sheet under the South Pole-Aitken Basin on the Moon, Geophys. Res. Lett., 36, L22202, doi:10.1029/2009gl040765. Pieters, C. M., M. I. Staid, E. Fischer, S. Tompkins, and G. He (1994), A sharper view of impact craters from Clementine data, Science, 266(5192), 1844 1848. Pieters, C. M., J. W. Head, L. Gaddis, B. Jolliff, and M. Duke (2001), Rock types of the South Pole-Aitken Basin and extent of basaltic volcanism, J. Geophys. Res., 106(E11), 28,001 28,022. Pieters, C. M., et al. (2008), The Moon Mineralogy Mapper (M3) on Chandrayaan-1, Curr. Sci. (India), 96(4), 500 505. Potter, R. W. K., D. A. Kring, G. S. Collins, W. S. Kiefer, and P. J. McGovern (2013), Numerical modeling of the formation and structure of the Orientale impact basin, J. Geophys. Res. Planets, 118, 963 979, doi:10.1002/ jgre.20080. Robinson, M. S., et al. (2010), Lunar Reconnaissance Orbiter Camera (LROC) instrument overview, Space Sci. Rev., 150, 81 124. Ryder, G., and J. A. Wood (1977), Serenitatis and Imbrium impact melts: Implications for large-scale layering in the lunar crust, Proc. Lunar Planet. Sci. Conf., 8, 655 668. Scholten, F., J. Oberst, K. D. Matz, T. Roatsch, M. Wählisch, E. J. Speyerer, and M. S. Robinson (2012), GLD100: The near-global lunar 100 m raster DTM from LROC WAC stereo image data, J. Geophys. Res., 117, E00H17, doi:10.1029/2011je003926. Scott, D. H., J. F. McCauley, and M. N. West (1978), Geologic map of the west side of the Moon. U.S. Geol. Survey Map I-1034. Simonds, C. H., J. L. Warner, W. C. Phinney, and P. E. McGee (1976), Thermal model for impact breccia lithification: Manicouagan and the Moon, Proc. Lunar Sci. Conf., 7, 2509 2528. Spudis, P. D. (1993), The Geology of Multi-Ring Basins: The Moon and Other Planets, pp. 263, Cambridge Univ. Press, New York and Cambridge. Spudis, P. D., and G. Ryder (1981), Apollo 17 impact melts and their relation to the Serenitatis Basin. In multi-ring basins, Proc. Lunar Planet. Sci. Conf., 12A, 133 148. Spudis, P. D., B. R. Hawke, and P. Lucey (1984), Composition of Orientale Basin deposits and implications for the lunar basin-forming process. Proceedings, lunar and planetary science conference 15, J. Geophys. Res., 89, C197 C210. Spudis, P. D., N. Zellner, J. Delano, D. C. B. Whittet, and B. Fessler (2002), Petrologic mapping of the Moon: A new approach, Lunar Planet. Sci., XXXIII, 1104. Spudis P. D., D. E. Wilhelms, and M. S. Robinson (2011), The Sculptured Hills of the Taurus Highlands: Implications for the relative age of Serenitatis, basin chronologies and the cratering history of the Moon, J. Geophys. Res., 116, E00H03, doi:10.1029/2011je003903. 28

Sunshine, J. M., and C. M. Pieters (1998), Determining the composition of olivine from reflectance spectroscopy, J. Geophys. Res., 103, 13,675 13,688, doi:10.1029/98je01217. Vaughan, W., and J. W. Head (2013), Modeling the South Pole-Aitken Basin subsurface, Lunar Planet. Sci., 44, 2012. Vaughan, W., J. W. Head, L. Wilson, and P. C. Hess (2013), Geology and petrology of enormous volumes of impact melt on the Moon: A case study of the Orientale Basin impact melt sea, Icarus, 223, 749 765. Whitten, J., J. W. Head, M. Staid, C. M. Pieters, J. Mustard, R. Clark, J. Nettles, R. L. Klima, and L. Taylor (2011), Lunar mare deposits associated with the Orientale impact basin: New insights into mineralogy, history, mode of emplacement and relation to Orientale Basin evolution from Moon Mineralogy Mapper (M 3 ) data from Chandrayaan-1, J. Geophys. Res., 116, E00G09, doi:10.1029/2010je003736. Wilhelms, D. E. (1987), The geologic history of the Moon, USGS Prof. Paper 1348, 302 pp. Zellner, N. E. B., P. D. Spudis, J. W. Delano, and D. C. B. Whittet (2002), Impact glasses from the Apollo 14 landing site and implications for regional geology, J. Geophys. Res., 107(E11), 5102, doi:10.1029/ 2001JE001800. 29