Quick viewing(Text Mode)

SURVEY of METEORITE PHYSICAL PROPERTIES: DENSITY, POROSITY and MAGNETIC SUSCEPTIBILITY by ROBERT J. MACKE, S.J. S.B. Massachuset

SURVEY of METEORITE PHYSICAL PROPERTIES: DENSITY, POROSITY and MAGNETIC SUSCEPTIBILITY by ROBERT J. MACKE, S.J. S.B. Massachuset

SURVEY OF PHYSICAL PROPERTIES: DENSITY, POROSITY AND MAGNETIC SUSCEPTIBILITY

by

ROBERT J. MACKE, S.J. S.B. Massachusetts Institute of Technology, 1996 M.A. Washington University in St. Louis, 1999 M.A. Saint Louis University, 2006

A dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Department of Physics in the College of Sciences at the University of Central Florida Orlando, Florida

Fall Term 2010

Major Professor: Daniel T. Britt

ABSTRACT

The measurement of meteorite physical properties (i.e. density, porosity, magnetic susceptibility) supplements detailed chemical and isotopic analyses for small samples (thin sections or ~300 mg portions) by providing whole-rock data for samples massing in the tens of grams. With the advent of fast, non-destructive and non-contaminating measurement techniques including helium ideal-gas pycnometry for grain density, the Archimedean ―glass bead‖ method for bulk density and (with grain density) porosity, and the use of low-field magnetometry for magnetic susceptibility, all of which rely on compact and portable equipment, this has enabled a comprehensive survey of these physical properties for a wide variety of . This dissertation reports on the results of that survey, which spanned seven major museum and university meteorite collections as well as the Vatican collection. Bulk and grain densities, porosities and magnetic susceptibilities are reported for 1228 stones from 664 separate meteorites, including several rare meteorite types that are underrepresented in previous studies.

Summarized here are data for (carbonaceous, ordinary and enstatite) and stony .

Several new findings have resulted from this study. From the use of a ―weathering modulus‖ based on grain density and magnetic susceptibility to quantify weathering in finds, it is observed that the degree of weathering of ordinary chondrites is dependent on their initial porosity, which becomes reduced to less than ~8% for all finds, but for enstatite chondrites

ii weathering actually increases porosity. Grain density and magnetic susceptibility, which have been shown to distinguish H, L and LL ordinary chondrites, also may distinguish shergottites, and chassignites from each other, but the two groups of enstatite chondrites (EH and

EL) remain indistinguishable in these properties. H finds exhibit a slight negative trend in porosity with increasing petrographic type, and all chondrite falls together exhibit a pronounced negative trend in porosity spanning all petrographic types. The overall trend corresponds roughly to a positive trend in porosities with respect to both degree of oxidation and percentage of matrix. It also corresponds to the macroporosities of analogous . These traits constrain models of conditions in the solar nebula and the formation of chondrite parent- body precursors.

iii

Ad Majorem Dei Gloriam.

iv

ACKNOWLEDGMENTS

Work of this nature would not be possible without the support of a large number of people and institutions. To begin with, many curators and collections managers made their meteorite collections available to me and made me feel welcome at their institutions. These are

(in roughly chronological order) Br. Guy Consolmagno at the Vatican Observatory; Denton Ebel and Joe Boesenberg at the American Museum of Natural History; Glenn MacPherson, Linda

Welzenbach, Cari Corrigan and many others at the Smithsonian Institution’s National Museum of Natural History; Art Ehlmann and Teresa Moss at the Monnig collection at Texas Christian

University; Carl Agee and James Karner at the Institute of , University of New

Mexico; Meenakshi Wadhwa and Laurence Garvie at the Center for Meteorite Studies, Arizona

State University; Caroline Smith, Gretchen Benedix and Deborah Cassey at the Natural History

Museum, London; and Philipp Heck, James Holstein and Paul Sipiera at the Field Museum of

Natural History.

Then there is the tremendous support I have received from the remainder of the scientific community. In particular, I should mention Jon Friedrich, Tom Kohout, Phil McCausland, and

Melissa Strait, all of whom have strongly encouraged my work. The list here could go on and on, but for the sake of brevity I will only add a few others. George Flynn, Pierre Rochette, Alan

Rubin, and Melissa Strait provided valuable reviewer feedback for various papers which found their way into this dissertation.

v

Many others outside the field of meteoritics have supported me in this work as well. I wish to thank my Jesuit superiors, especially Fr. Tim McMahon (provincial) and Fr. David

Fleming (formation assistant) for approving my doctoral studies back in 2007, and the current provincial and formation assistant, Fr. Douglas Marcouiller and Fr. John Armstrong, who continue to support my studies. The Jesuits at Jesuit High School in Tampa, where I have resided while pursuing the Ph.D., have also been quite welcoming and supportive. Also, the

Jesuits at the Vatican Observatory have been quite encouraging in my work.

My thesis committee has been more than supportive; they have encouraged me to complete the dissertation in a timely manner and are all eager to see me become a productive member of the scientific community. They are Dan Britt, Br. Guy Consolmagno, Humberto

Campins and Joe Harrington. Dan and Guy are also responsible for my being here in the first place; Dan needed a graduate student to do this study for which he had grant funding, and Guy just happened to know a young Jesuit scientist who needed a Ph.D.

Speaking of grants, this work is funded by NASA Planetary Geology and Geophysics grants NNX09AD91G and NNG06GG62G. On top of that, the Smithsonian Institution funded my work there in the summer of 2008 with a 10-week Smithsonian Institution Graduate Research

Fellowship.

Last but not least, I want to thank my parents. Mom and Dad, where would I be without you?

vi

TABLE OF CONTENTS

LIST OF FIGURES ...... xii

LIST OF TABLES ...... xviii

LIST OF ABBREVIATIONS ...... xx

CHAPTER 1: INTRODUCTION ...... 1

1.1 Meteorite Basics...... 6

1.1.1 Shock...... 8

1.1.2 Terrestrial Weathering ...... 9

1.2 Some Science Questions ...... 12

1.2.1 Questions about Ranges and Classification ...... 13

1.2.2 Questions about Weathering ...... 16

1.2.3 Questions about Shock ...... 18

1.2.4 Big-Picture Questions ...... 18

1.2.5 Other Questions ...... 20

1.3 Organization of the Dissertation ...... 20

CHAPTER 2: MEASUREMENT METHODS ...... 22

vii

2.1 Background ...... 22

2.2 Grain density: Helium Ideal-gas Pycnometry ...... 24

2.2.1 Theory ...... 25

2.2.2 Measurement ...... 26

2.2.3 Other Considerations ...... 29

2.3 Bulk Density: Archimedean Glass Bead Method ...... 32

2.3.1 Theory ...... 32

2.3.2 Measurement ...... 33

2.3.3 Settling Methods and Systematic Error ...... 36

2.3.4. Further Considerations ...... 42

2.4 Magnetic Susceptibility ...... 44

2.4.1 Instrument ...... 46

2.4.2 Adjustments for Finite Sizes ...... 47

CHAPTER 3: ORDINARY CHONDRITES ...... 51

3.1 Introduction ...... 51

3.2 Falls ...... 54

3.3 H Chondrite Finds ...... 56

3.3.1 Model Porosities of H Finds ...... 56

3.3.2 Weathering Modulus for H Finds ...... 57

3.4 Falls ...... 60

viii

3.5 L Chondrite Finds ...... 60

3.5.1 Model Porosities of L Finds ...... 61

3.5.2 Weathering Modulus for L Finds ...... 61

3.6 LL Chondrite Falls ...... 62

3.7 LL Chondrite Finds ...... 63

3.7.1 Model Porosities of LL Finds ...... 63

3.7.2 Weathering Modulus for LL Finds ...... 64

3.8 Intermediate OC Types ...... 65

3.9 Porosities of Breccias and Non-Breccias ...... 65

3.10 Outliers among OC Falls ...... 67

3.11 Other Chondrite Types ...... 69

CHAPTER 4: CARBONACEOUS CHONDRITES ...... 71

4.1 Introduction ...... 71

4.2 Data ...... 73

4.2.1 The CR Clan: CR, CB and CH ...... 73

4.2.2 Aqueously Altered Carbonaceous Chondrites: CI and CM ...... 78

4.2.3 The Anhydrous Carbonaceous Chondrites: CO, CK and CV ...... 83

4.2.4 Other Carbonaceous Chondrites ...... 93

4.3 Discussion ...... 94

4.3.1 Petrographic Type ...... 95

ix

4.3.2 Shock...... 96

CHAPTER 5: ENSTATITE CHONDRITES ...... 98

5.1 Introduction ...... 98

5.2 Results ...... 101

5.2.1 Grain Density ...... 101

5.2.2 Bulk Density ...... 102

5.2.3 Porosity ...... 103

5.2.4 Magnetic Susceptibility ...... 104

5.3 Discussion ...... 105

5.3.1 Weathering Effects on Finds...... 105

5.3.2 Heterogeneity ...... 108

5.3.3 Grain Density, Magnetic Susceptibility and Metallic Iron Content ...... 111

5.4 Conclusions ...... 117

CHAPTER 6: ACHONDRITES ...... 118

6.1 Introduction ...... 118

6.2 Results ...... 121

6.2.1 Lunar Meteorites and Apollo Samples ...... 122

6.2.2 SNC ...... 123

6.2.3 HED ...... 126

6.2.4 ...... 129

x

6.2.5 ...... 132

6.2.6 ...... 133

6.2.7 and ...... 135

6.2.8 Other Primitive Achondrites: and ...... 136

6.3 Conclusions ...... 138

CHAPTER 7: SUMMARY AND CONCLUSIONS ...... 141

7.1 Weathering ...... 141

7.2 Homogeneity ...... 144

7.3 Grain Density and Magnetic Susceptibility as a Classification Tool ...... 147

7.4 and Iron Meteorites ...... 149

7.5 Overall Trends from Chondrite Falls ...... 150

7.5.1 Shock-Related Trends ...... 150

7.5.2 Trends by Petrographic Type ...... 151

7.5.3 Trends by Matrix Abundance and Degree of Oxidation ...... 152

7.5.4 Implications for the Solar Nebula ...... 154

7.6 Conclusion ...... 156

APPENDIX A: FIGURES ...... 158

APPENDIX B: TABLES ...... 252

APPENDIX C: COPYRIGHT PERMISSIONS ...... 295

LIST OF REFERENCES ...... 299

xi

LIST OF FIGURES

Figure 1: Basic Ideal-Gas Law Approach to Measurement of Volume...... 159

Figure 2: The Quantachrome Ultrapycnometer 1000...... 160

Figure 3: Diagram of the Quantachrome Ultrapycnometer 1000...... 161

Figure 4: Ramp-up of Grain Volume Measurements in the Ultrapycnometer 1000...... 162

Figure 5: Rare-Earth Element Concentrations for Glass Beads (Normalized to Chondritic)

With Comparison Data Taken from Common Brown and Blue Glass Bottles...... 163

Figure 6: The Bead Method Apparatus, Including Shake Platform and Nalgene Beaker...... 164

Figure 7: Data from Measurements on Quartz Utilizing a 77-cm3 Cup, Made Without

Employing a Bead Settling Method...... 165

Figure 8: Data from Measurements Made Using the ―Soft Tap‖ Settling Method...... 166

Figure 9: Data from Measurements Using the ―Free Shake‖ Settling Method...... 167

Figure 10: Data for Measurements Made Using the 5-second Secured Shake Method...... 168

Figure 11: Volume Discrepancy for the Secured-Shake Method Using 700-800 μm Diameter

Beads...... 169

Figure 12: Volume Discrepancy for the Secured-Shake Method Using 700-800 μm Diameter

Beads and the 155-cm3 Cup...... 170

Figure 13: Bulk Density of Small Glass Beads (ρbead ) vs. Relative Humidity...... 171

xii

Figure 14: The SM-30 Magnetic Susceptibility Meter. (a) Photograph of the Meter. (b)

Diagram...... 172

Figure 15: Operation of the SM-30 Magnetic Susceptibility Meter. (a) Photograph of the

Device as Utilized (Inverted, with the Meteorite Placed Atop and Centered Over

the Magnetic Coils). (b) Diagram of the Device in Operation...... 173

Figure 16: SM-30 Geometric Correction Factor α as a Function of Bulk Volume...... 174

Figure 17: Comparison of Magnetic Susceptibility Measurements Made Using the SM-30

(Vertical Axis) with Those Utilizing the KLY-2 (Horizontal Axis) for the Same

Stones in the Vatican Collection...... 175

Figure 18: Discrepancy in Log Units Between Measurements Made Using the SM-30 and

Those Utilizing the KLY-2 for the Same Stones...... 176

Figure 19: Grain Density vs. Magnetic Susceptibility for Falls in the

Vatican Collection...... 177

Figure 20: Grain Density vs. Magnetic Susceptibility for All Stones from OC Falls...... 178

Figure 21: Porosity as a Function of Shock for H Chondrite Falls...... 179

Figure 22: Porosity as a Function of Petrographic Type for H Chondrite Falls...... 180

Figure 23: Magnetic Susceptibility as a Function of Petrographic Type for H Chondrite

Falls...... 181

Figure 24: Grain Density/Bulk Density Relationship for H Falls, For Use in Determination of

Model Porosities for H Finds...... 182

Figure 25: Comparison of Model Porosities for H Finds with Actual Porosities for H Falls...... 183

Figure 26: Porosity as a Function of Petrographic Type for H Chondrites, Including

Measured Porosities of Falls and Model Porosities of Finds...... 184

xiii

Figure 27: Grain Density and Magnetic Susceptibility of H Finds...... 185

Figure 28: Porosity vs. Weathering Modulus for H Finds...... 186

Figure 29: Bulk Density vs. Weathering Modulus for H Finds...... 187

Figure 30: Model Porosity vs. Weathering Modulus for H Finds...... 188

Figure 31: Porosity vs. Petrographic Type for L Falls...... 189

Figure 32: Porosity as a Function of Shock Stage for L Falls...... 190

Figure 33: Comparison of Model Porosities for L Finds with Actual Porosities for L Falls...... 191

Figure 34: Porosity as a Function of Petrographic Type for L Chondrites, Including Measured

Porosities of Falls and Model Porosities of Finds...... 192

Figure 35: Grain Density and Magnetic Susceptibility of L Finds...... 193

Figure 36: Porosity vs. Weathering Modulus for L Finds...... 194

Figure 37: Bulk Density vs. Weathering Modulus for L Finds...... 195

Figure 38: Model Porosity vs. Weathering Modulus for L Finds...... 196

Figure 39: Porosity vs. Petrographic Grade for LL Falls...... 197

Figure 40: Magnetic Susceptibility as a Function of Petrographic Type for LL Falls...... 198

Figure 41: Comparison of Model Porosities for LL Finds with Actual Porosities for LL Falls. 199

Figure 42: Porosity as a Function of Petrographic Type for LL Chondrites, Including

Measured Porosities of Falls and Model Porosities of Finds...... 200

Figure 43: Grain Density and Magnetic Susceptibility of LL Finds...... 201

Figure 44: Porosity vs. Weathering Modulus for LL Finds...... 202

Figure 45: Bulk Density vs. Weathering Modulus for LL Finds...... 203

Figure 46: Model Porosity vs. Weathering Modulus for LL Finds...... 204

xiv

Figure 47: Histogram of Porosities for Brecciated and Non-Brecciated Ordinary Chondrite

Falls...... 205

Figure 48: Grain Density / Magnetic Susceptibility Plot for K and R Chondrites...... 206

Figure 49: Grain Densities of Carbonaceous Chondrites...... 207

Figure 50: Bulk Densities of Carbonaceous Chondrites...... 208

Figure 51: Porosities of Carbonaceous Chondrites...... 209

Figure 52: Magnetic Susceptibilities of Carbonaceous Chondrites...... 210

Figure 53: Grain Density and Magnetic Susceptibility for the CR Clan: CR, CB and CH

Carbonaceous Chondrites...... 211

Figure 54: Grain Density and Magnetic Susceptibility for CI and CM Carbonaceous

Chondrites...... 212

Figure 55: Grain Density and Magnetic Susceptibility for CO Carbonaceous Chondrites...... 213

Figure 56: Grain Density and Magnetic Susceptibility for CK Carbonaceous Chondrites...... 214

Figure 57: Grain Density and Magnetic Susceptibility for CV Carbonaceous Chondrites...... 215

Figure 58: Grain Density and Magnetic Susceptibility for Ungrouped Carbonaceous

Chondrites...... 216

Figure 59: Bulk Density vs. Petrographic Type for Falls...... 217

Figure 60: Grain Density vs. Petrographic Type for Carbonaceous Chondrite Falls...... 218

Figure 61: Porosity vs. Petrographic Type for Carbonaceous Chondrite Falls...... 219

Figure 62: Magnetic Susceptibility vs. Petrographic Type for Carbonaceous Chondrite Falls. ..220

Figure 63: Bulk Density vs. Shock for Carbonaceous Chondrites...... 221

Figure 64: Grain Density vs. Shock for Carbonaceous Chondrites...... 222

Figure 65: Porosity vs. Shock for Carbonaceous Chondrites...... 223

xv

Figure 66: Magnetic Susceptibility vs. Shock for Carbonaceous Chondrites...... 224

Figure 67: Grain Densities...... 225

Figure 68: Enstatite Chondrite Bulk Densities...... 226

Figure 69: Enstatite Chondrite Porosities...... 227

Figure 70: Enstatite Chondrite Magnetic Susceptibilities...... 228

Figure 71: Grain Density vs. Magnetic Susceptiblity for Enstatite Chondrites...... 229

Figure 72: Enstatite Chondrite Finds and Their Properties Grouped By Weathering: (a)

Selection of Groups Based on Grain Density and Magnetic Susceptibility; (b)

Porosity and Bulk Density of the Same Groups...... 230

Figure 73: Porosity as a Function of Weathering Modulus for EC Finds...... 231

Figure 74: Variability of EC Grain Densities by Mass...... 232

Figure 75: Mass-Weighted Average Grain Density vs. Magnetic Susceptibility for Falls

Exceeding Total Mass of 40 g...... 233

Figure 76: Grain Densities of Achondrites in this Study...... 234

Figure 77: Bulk Densities of Achondrites in this Study...... 235

Figure 78: Porosities of Achondrites in this Study...... 236

Figure 79: Magnetic Susceptibilities for Achondrites in this Study...... 237

Figure 80: Grain Density vs. Magnetic Susceptibility for SNC (Martian) Meteorites in this

Study...... 238

Figure 81: Grain Density vs. Magnetic Susceptibility for HED Meteorites in this Study...... 239

Figure 82: Grain Density vs. Magnetic Susceptibilities for Aubrites, Angrites and Ureilites in

this Study, as well as One Stone of an Enstatite (Zakłodzie)...... 240

xvi

Figure 83: Grain Density vs. Magnetic Susceptibility for Acapulcoites and Lodranites in this

Study...... 241

Figure 84: Grain Density vs. Magnetic Susceptibility for Primitive Achondrites in this Study. 242

Figure 85: Porosity vs. Shock State for All Chondrite Falls...... 243

Figure 86: Grain Density vs. Shock State for All Chondrite Falls...... 244

Figure 87: Magnetic Susceptibility vs. Shock State for All Chondrite Finds...... 245

Figure 88: Porosity vs. Petrographic Type for All Chondrite Falls...... 246

Figure 89: Grain Density vs. Petrographic Type for All Chondrite Falls...... 247

Figure 90: Bulk Density vs. Petrographic Type for All Chondrite Falls...... 248

Figure 91: Magnetic Susceptibility vs. Petrographic Type for All Chondrite Falls...... 249

Figure 92: Porosity vs. Oxidation State for All Chondrite Falls...... 250

Figure 93: Porosity vs. Percentage Matrix for All Chondrite Falls...... 251

xvii

LIST OF TABLES

Table 1: Electron-Probe Micro Analysis (EPMA) of Glass Beads...... 253

Table 2: Instrumental Neutron Activation Analysis (INAA) of Glass Beads...... 254

Table 3: Results Per Settling Method for Small Beads in the 77 cm3 Container...... 255

Table 4: Data for H Chondrite Falls...... 256

Table 5: Data for H Chondrite Finds...... 260

Table 6: Data for L Chondrite Falls...... 262

Table 7: Data for L Chondrite Finds...... 266

Table 8: Data for LL Chondrite Falls...... 269

Table 9: Data for LL Chondrite Finds...... 271

Table 10: Data for H/L and L/LL Chondrites...... 272

Table 11: Average Porosities of Brecciated and Non-Brecciated OC Falls...... 273

Table 12: Data for K and R Chondrites...... 274

Table 13: Data for Carbonaceous Chondrites...... 275

Table 14: Physical Properties of Carbonaceous Chondrite Falls by Petrographic Type...... 279

Table 15: Physical Properties of Carbonaceous Chondrites by Shock...... 280

Table 16: Data for Enstatite Chondrites...... 281

Table 17: Mass-weighted Averages by Meteorite for Enstatite Chondrites...... 282

Table 18: Data for Lunar Meteorites, Apollo Samples, and SNC...... 283

xviii

Table 19: Data for HEDs...... 284

Table 20: Data for Aubrites, Angrites and Ureilites...... 287

Table 21: Data for Primitive Achondrites...... 288

Table 22: Summary of Physical Property Averages for All Meteorite Groups...... 289

Table 23: Data for Mesosiderites, and Iron Meteorites...... 291

Table 24: Porosity Averages by Shock Stage for All Chondrite Falls...... 293

Table 25: Porosity Averages by Petrographic Type for Chondrite Falls...... 294

xix

LIST OF ABBREVIATIONS

Abbreviations for names of meteorite collections used in data tables:

AMNH American Museum of Natural History (New York, NY)

CMS Center for Meteorite Studies, University of Arizona (Tempe, AZ)

DuPont DuPont Meteorite Collection (at FMNH)

FMNH Field Museum of Natural History (, IL)

IOM Institute of Meteoritics, University of New Mexico (Albuquerque, NM)

LNHM The Natural History Museum (London, UK)

Monnig The Monnig Collection, Texas Christian University (Fort Worth, TX)

NASA NASA Lunar Sample Laboratory Facility, Johnson Space Center (Houston TX)

PSF Planetary Studies Foundation (at FMNH)

Vatican Vatican Meteorite Collection, Vatican Observatory (Vatican City State)

Abbreviations for common meteorite names:

ALH Allan Hills

HaH Hammadah al Hamra

JaH Jiddat al Harasis

NWA Northwest Africa

xx

Abbreviations for achondritic meteorite groups used in data tables:

Aca

Ang

Aub

Bra

Cha Chassignite

Dio

Enst-Ach Enstatite Achondrite (non-aubritic)

Euc

How

Lod

Lun or Apollo Sample

Nak

She Shergottite

Ure

Win

xxi

CHAPTER 1: INTRODUCTION

When it comes to understanding the origins and composition of the solar system, no current method of study exhibits more promise than the laboratory study of retrieved materials.

Nevertheless, sample-return missions to extraterrestrial bodies are expensive, resulting in a very limited number of celestial bodies to be sampled this way. To date, the only materials returned to Earth by this method are lunar samples from the Apollo and Luna programs, some dust from Wild 2 collected by the Stardust mission (cf. Brownlee et al., 2006), and possibly some dust from 25143 Itakawa collected by the Japanese Hayabusa mission (cf. Yoshikawa et al., 2010). This is not to say that there is a severe lack of extraterrestrial materials with which to work; given enough time and favorable orbits, much material finds its own way to Earth without the expenditure of a single dollar in mission costs. Some of this is in the form of interplanetary dust particles (IDPs), which are scooped from the upper atmosphere in aerogel (cf. Reitmeijer,

1998). Larger stones can be retrieved from the ground in the form of meteorites.

There are more than twenty-two thousand recovered meteorites (Grady, 2000) sampling from a large but indeterminate number of solar system objects. Most of the objects are asteroids, though some meteorites have been linked to the moon or Mars (Hutchison, 2004). This provides an abundant supply of solar system materials to study in the luxury of a laboratory setting.

Nevertheless, meteorites are not to be considered representative of solar system materials in

1 terms of their abundances; the Earth itself manufactures a selection bias in terms of the number and types of materials that reach this planet and survive atmospheric entry.

The factors that influence the availability of meteorites are favorable orbits for encountering Earth, favorable entry velocities, and internal strength of the rocks making up the asteroid. Near-Earth asteroids satisfy the first two criteria; because their orbital dynamics brings them in proximity to Earth, they are more likely to encounter the planet, and their orbital speeds are similar to that of the Earth, resulting in favorably low entry velocities. Asteroids and from farther out in the solar system, on the other hand, are less likely to meet the two criteria.

Even if an orbit gets perturbed so as to send an object on an Earth-crossing trajectory, the conversion of potential to kinetic energy means its relative velocity to the Earth will be much higher than for near-Earth objects (NEOs). In addition, outer system materials tend to be less coherent and have reduced internal strength, which means they have even further reduced odds of surviving atmospheric entry and impact with the ground. In one such example, the fall of the very friable carbonaceous chondrite , some material only survived intact because it landed in soft snow (Hildebrand et al., 2006).

Most (73%) of the meteorite falls recovered are of a class known as ordinary chondrites

(Grady, 2000). On the other hand, ordinary-chondrite analogous asteroids occupy only a small percentage of the population (Gaffey et al., 1993). They are more representative of the abundances (approximately 80%) of OC analogs found in NEOs (Bottke et al., 2002). Therefore, while the variety of meteorites available for study is considerable, the vast majority of available meteorites sample only a small portion of the solar system. Some meteorite types, such as those originating from Mars or from the outer asteroid belt, are much rarer than ordinary chondrites and due to their rarity may be more scientifically valuable, and undoubtedly other portions of the

2 solar system are entirely unrepresented in existing meteorites. Though meteorites present an otherwise unattainable supply of solar system material for the purpose of research, certain parts of the solar system are far better sampled than others.

These meteorites enable detailed analyses of mineralogy, composition, and isotopes that, among other things, provide useful information as to the age and geologic history of the parent bodies from which they originated. Most of the research performed on meteorites is of this variety, and generally involves the removal of some material from the meteorite for the creation of thin sections or for more destructive forms of analysis. In recent years, there has been growing recognization that, in addition to chemical and isotopic analyses, the study of bulk physical properties of whole stones provides a useful avenue of research without requiring destructive measures. One key such property is porosity, or the volume percentage of space within a rock that is not filled by solid matter. Contributors to porosity include microcracks, voids, or possibly even small gaps between adjacent crystals or inclusions. While this property may be studied in thin sections (e.g. Corrigan et al., 1997; Strait and Consolmagno 2002, 2005,

2010), it is best understood in terms of its whole-rock value. Porosity, being a structural rather than mineralogical property, is potentially influenced by a number of factors including (but not limited to) local gravitational compression, shock, and the range of physical sizes of minerals within the rock. This property will have been primarily influenced by the conditions under which the rock lithified, though subsequent history (including collisional events) would have further modified the structure of the rock and thus influenced porosity.

In addition to other useful information porosity reveals about these processes, meteorite porosity also yields clues to the structure of the meteorites’ asteroid analogs. Asteroids tend to be more porous than their analogous meteorites, which indicates a high degree of macroporosity,

3 or pore space that resides as larger-scale cracks and voids between the individual stones comprising the asteroid. Those asteroids beyond ~1020 kg tend to have negligible macroporisities, indicating that their self-gravity is strong enough to compress the pore space away. Many asteroids have macroporosities of around 20%, indicating that they exist structurally as ―rubble piles‖ composed of stones loosely bound together gravitationally, rather than as a coherent monolith. Other asteroids, in particular those objects in the outer asteroid belt that have been connected with carbonaceous chondritic meteorites, have macroporosities upward of 70%, indicating a very loose structure (Britt et al., 2002; Consolmagno et al., 2008).

The measurement of meteorite porosity (expressed as a volume percentage of pore space in the whole rock) requires two distinct density measurements, both of which are defined as total mass divided by volume, but for different volumes. Bulk density uses the complete volume enclosed by the bulk stone, while grain density uses the volume that is only occupied by the solid material making up the stone. The difference in the two volumes is due entirely to porosity.

Measurement of these two densities has historically involved destructive techniques such as immersion in a liquid or slicing the sample into a simple geometric form such as a parallelepiped, which because they were either contaminating or destructive limited their ability to be applied to large numbers of meteorites. The application of ideal-gas pycnometry to meteoritics (cf. Faeth and Willingham, 1955; Cadenhead et al., 1972) allowed for the non- destructive, non-contaminating measurement of grain density, but it was not until the advent of the Archimedean ―glass bead‖ method (Consolmagno and Britt, 1998) that bulk densities could also be measured quickly and efficiently without destruction or threat of serious contamination.

These methods (described in greater detail in Chapter 2) enabled the study for the first time of the density and porosity of very large numbers of meteorites of a wide variety of types,

4 especially the most underrepresented kinds of meteorites. Consolmagno and Britt (1998) tested and widely applied the techniques at the Vatican meteorite collection (also Consolmagno et al.,

2006), where they were developed with the intent of surveying meteorite porosities. The current study was instituted in 2007 with the purpose of expanding the survey to other major meteorite collections and also to include the measurement of magnetic susceptibility (another useful physical property that will be discussed later). The reassurance that the techniques were non- destructive and non-contaminating allowed curators and collections managers to permit their application to large portions of their collections, including rare meteorite types. Research took place at eight major meteorite collections. This included four museum collections (American

Museum of Natural History, Field Museum of Natural History, the Natural History Museum in

London, and the Smithsonian Institution’s National Museum of Natural History), three universities (Arizona State University, Texas Christian University, and the University of New

Mexico) and the Vatican collection. This dissertation is the primary report of the results of that survey. At least one physical property is reported for 1228 stones from 664 meteorites, of which

1019 stones had the full suite of bulk and grain density, porosity and magnetic susceptibility.

The following sections of this chapter provide (a) a basic meteorite primer that will describe classification, shock and weathering, and various terminology that will occur throughout the dissertation, and (b) a discussion of some of the many scientific questions that motivate this research and that this study may help answer.

5

1.1 Meteorite Basics

Meteorites, which originate from a wide range of sources within the solar system, are very diverse. They are most immediately subdivided, based on the amount and distribution of iron, into three categories: irons, stony-irons and stony meteorites. Irons, as the name suggests, are composed predominantly of iron metal (in the form of Fe-Ni alloys) and originated in the cores of differentiated parent bodies. Stony-irons are a roughly even mix of iron and stony material (mostly silicates), and are thought to originate somewhere near the core-mantle boundaries of differentiated parent bodies, though the nature of their formation is a subject of some debate. Stony-irons may be further subdivided into mesosiderites, which are resolidified silicate-and-metal mixtures, and pallasites, which are composed of large crystals of embedded in Fe-Ni metal. Iron and stony-iron meteorites were not extensively studied in this project, and so will not be discussed further here.

Stony meteorites, which are predominantly composed of silicates but many of which have iron abundances exceeding that of most terrestrial crustal rock, are subdivided into chondrites and achondrites. Chondrites, which are solidified aggregates composed of

(glassy melts created by localized heating of dust in the solar nebula over short time scales;

Connolly et al., 2006) embedded in a matrix of very fine particles, have chemical compositions that are very near solar average abundances except for (Hutchison, 2004). They are thought to be extremely primitive and have not been extensively processed since their initial formation. Chondrites are further subdivided into three classes: ordinary chondrites (OC), carbonaceous chondrites (CC), and enstatite chondrites (EC). In addition to the three primary classes, there are the rare Rumuruti-like (R) chondrites and Kakangari-like (K) chondrites.

6

Ordinary chondrites are subdivided into three further groups which vary by the amount of total Fe and metal and are named accordingly: H (high-Fe), L (low-Fe) and LL (both low Fe and low metal). Historically, they had more descriptive but also more confusing names such as olivine-bronzite (H), olivine-hypersthene (L) and olivine-pigeonite (LL; Mason, 1962). Enstatite chondrites are subdivided into two groups, which by way of analogy with ordinary chondrites were designated EH and EL (Sears et al., 1982). Carbonaceous chondrites, on the other hand, come in many varieties, with each group named after a type-specimen (Hutchison, 2004). These include CB (Bencubbin-type; Weisberg et al., 2001), CI (Ivuna-type), CK (Karoonda-type), CM

(Mighei-type), CO (-type), CR (Renazzo-type) and CV (Vigarano-type). There is an additional group of high-Fe carbonaceous chondrites called CH.

Chondrites have undergone varying degrees of thermal or aqueous processing, for which van Schmus and Wood (1967) devised a petrographic type scheme in order to quantify the kind and amount of processing a given meteorite has undergone. The scheme, which is based on the degree to which chondrules have equilibrated with surrounding matrix, ranges from 1 (most strongly aqueously altered) to 6 (most strongly thermally equilibrated), with a petrographic type of 3 representing a relatively unequilibrated meteorite. Subsequent developments of the type scheme have included the further gradation of type 3 into a range from 3.0 (unequilibrated) to 3.9

(just shy of type 4) and the addition of petrographic type 7, which applies in rare cases. It should be noted that thermal alteration does not rule out any aqueous alteration or vice versa – the type scheme expresses the primary mode of equilibration, not the only mode. All ordinary chondrites and enstatite chondrites are thermally equilibrated, and so occupy types 3-6 (or 7), while carbonaceous chondrites tend either to be aqueously altered (for example, CM and CI) or type 3

(CV, CO), though some are more thermally altered (for example, CK).

7

Achondrites are a catch-all category, meaning simply ―not chondrites.‖ They include some chemically primitive meteorites that have undergone varying degrees of processing, such as acapulcoites, lodranites and ureilites. Ureilites, though achondrites, are thought to be related to CV carbonaceous chondrites. Achondrites also include materials from varying depths in the crusts of differentiated parent bodies including the Moon, Mars, and the asteroid 4 Vesta. With the exception of , and (HEDs) which originate from 4 Vesta

(Consolmagno and Drake, 1977), achondrites are among the rarest meteorite types.

1.1.1 Shock

Impact events between celestial bodies can induce changes within the rocks involved, and the effects of shock metamorphism are observed in meteorites. Stöffler et al. (1988, 1991) devised a shock classification scheme for chondrites (adapted for enstatite chondrites by Rubin et al., 1997) based on fracturing and equilibration of olivine, Ca-poor , and plagioclase.

The effects on these minerals are dependent primarily on the maximum shock pressure. The scheme ranges from S1 (least shocked, with maximum pressure less than 5 GPa) to S6 (strongly shocked, with maximum pressure exceeding 75 GPa; Hutchison, 2004). There is an additional category for shock melts, which have been so severely shocked as to produce liquefaction and resolidification of whole rocks. The shock classification scheme is applicable to many other meteorites as well. A given meteorite’s shock classification only expresses the maximum impact pressure experienced by the object. Many meteorites show signs of multiple shock events.

8

Shock is important in terms of physical properties in a number of ways. In addition to causing fracturing and melting of minerals within the rock, it affects whole-rock structure.

Shock events tend to compress meteorites, potentially removing intergranular pore space. At the same time, the rapid compression and relaxation that occurs as a shock wave passes through rock induces the formation of cracks which themselves contribute to porosity. In a review of meteorite porosities, Consolmagno et al. (2008) observed that ordinary chondrite porosities drop considerably between shock states S1 and S3 or S4, and then tend to level off as additional shock creates as much pore space in the form of cracks as it destroys due to compression.

1.1.2 Terrestrial Weathering

From the moment a meteorite arrives on the Earth, interaction with the terrestrial environment begins to alter it. Most of the meteorites collected in recent years are finds. Many of those collected in hot desert regions have terrestrial ages of tens of thousands of years, as determined by the decay of cosmogenic nuclides (Bland et al., 1998b, 2006), and so have experienced considerable amounts of weathering. In order to properly interpret the data in this study, it is important to consider the effects of terrestrial weathering. In this dissertation, the term ―weathering‖ will refer exclusively to terrestrial weathering processes. It should be noted that the term ―weathering‖ also is used by planetary scientists to refer to various space- weathering processes, including processes that are relevant for meteorites and their parent- bodies.

9

The effects of weathering are many and varied, depending both on local environmental conditions and composition of the meteorite. For ordinary chondrites, which have been extensively studied (Bland et al., 1996, 1998b), one of the primary effects is oxidation (rusting) of Fe in all of its phases including Fe-bearing silicates, though the most pronounced effect is on

Fe-Ni metal and a somewhat lesser effect is the oxidation of (FeS). As such, H chondrites experience more severe weathering effects than L and LL. The products of Fe oxidation include goethite, magnetite and others (Bland et al., 1998b, 2006; Consolmagno et al.,

2008) which have a density roughly half that of Fe-Ni metal. The lower-density material occupies a greater volume than the original material, so it expands to fill existing pore space.

This has the effect of reducing total porosity. Oxidation of metal also reduces grain density, but since it does not significantly alter the total mass or bulk volume of the rock it does not have a significant effect on bulk density. Since metals have substantially higher magnetic susceptibility than iron oxides, another effect of weathering on Fe-rich meteorites is a reduction in magnetic susceptibility.

Other effects of weathering, such as the creation of carbonates, are minor in comparison with rusting for ordinary chondrites, but they may be more important for other meteorite types.

For example, some preliminary porosity data (Consolmagno et al., 2008) suggests that CO finds may have had their porosity reduced by carbonate expansion. Highly porous meteorites such as carbonaceous chondrites are also prone to the absorption of atmospheric water; experiments on dehydrated CCs exposed to air indicate rapid reabsorption of their original quantities of water.

Anecdotally, in the course of this study when masses were measured for some of the carbonaceous chondrites, the measured masses were slightly higher than those on the collection records, an effect that can be attributed to the absorption of a small mass of water.

10

At least for hot desert finds, ordinary chondrite weathering is a two-stage process. Stage one weathering occurs quite rapidly but at a rate that depends on local environmental conditions.

Fe metal oxidizes during this stage, expanding into existing pore space. The continued weathering of the interior of the meteorite depends on the ability of air and moisture to penetrate the interior, which it can do only through pore space. Once the porosity is reduced to the point that weathering agents can no longer reach the interior (or reactive grains have been ―protected‖ by an oxide buffer), weathering rates slow to almost nil for stage 2, in which the slow degradation of the meteorite exterior may continue but internal weathering has effectively stopped. Under the right conditions, a meteorite can survive stage 2 weathering for many thousands of years. For the purposes of this study, all finds can be thought to be in stage 2.

Weathering of Antarctic (cold desert) finds varies somewhat from those collected in hot deserts, in part because of small adjustments in weathering chemistry. Antarctic finds weather at a different rate and their long-term interaction with ice may affect them structurally.

Consolmagno et al. (1998) observe that Antarctic meteorites may have enhanced porosities over falls, indicating that their weathering mechanism is not merely different in magnitude but in kind from that of hot deserts. This study includes data from almost no Antarctic meteorites, and so for the purposes of this study only the hot desert weathering effects will be considered.

Weathering is strongest for finds, but it should be noted that even falls are weathered, though to a lesser degree. Bland et al. (1998a) conducted a study comparing stones of Holbrook

(L/LL6) collected shortly after they fell in 1912 with a stone collected in 1968. After a mere 56 years in the Arizona desert, the latter stone had experienced about 9.7% Fe conversion into Fe3+, which is about half the 20% maximum Fe conversion before porosity stabilization (Bland et al.,

11

2006). Even those meteorites which have been collected promptly after a fall experience some degree of terrestrial alteration while they reside on collection shelves.

Weathering classification for finds is varied and in many cases depends on the subjective judgment of the one making the classification. Antarctic meteorites are given a weathering grade of A, B or C depending on degree of rusty appearance of the exterior (A possessing little or no rustiness, and C being extremely rusty). A commonly-used scheme rates weathering grades of polished thin sections based on degree of oxidation of various phases. This system ranges from

W0 (unweathered) through degrees of oxidation of metal and troilite (W1-W4) and in more extreme cases (W5 and W6) alteration of mafic silicates (Jull et al., 1991; Wlotzka 1993; Al-

Kathiri et al., 2005). Another, more quantitative method of establishing weathering grades

(Bland et al., 1996, 1998b) uses 57Fe Mössbauer spectroscopy to measure the abundance of Fe3+, which is one of the products of oxidation processes that is not found in unweathered ordinary chondrites.

1.2 Some Science Questions

While meteorite physical properties have been studied to varying degrees (cf. Britt and

Consolmagno, 2003; Consolmagno and Britt, 1998; Consolmagno et al., 1998, 2006; Flynn and

Klock, 1998; Guskova 1985; Keil, 1962; Kohout et al., 2006; Kukkonen and Pesonen, 1983;

Matsui et al., 1980; Pesonen et al., 1993; Rochette et al., 2001, 2003, 2005, 2008, 2009, 2010;

Wilkison and Robinson, 2000; Yomogida and Matsui, 1981, 1983), the scale of this survey is unprecedented for density, porosity and magnetic susceptibility data collected under a consistent

12 methodology. It should enable the exploration of numerous questions for which sufficient statistics were previously lacking. That being said, this study was primarily limited by the availability of meteorites themselves. Though the rarest meteorite types were particularly targeted for study, their very rarity means that in many cases there still is not a statistically significant number of stones that have been measured. Nevertheless, for all meteorite types included in this survey, the total size of the available database marks an improvement over previous data, allowing at the very least the exploration of some questions merely hinted at in previous studies. In the remainder of this section, some of those questions will be discussed.

1.2.1 Questions about Ranges and Classification

1.2.1.1 Classification of Ordinary Chondrites

Consolmagno et al. (2004, 2006) observed that grain density and magnetic susceptibility together serve as a reliable tool for classifying ordinary chondrites into H, L and LL groups.

Differences in total Fe and in the amount of metal in the different OC groups results in differences in density (since Fe-Ni metal is more than twice as dense as most silicates) and magnetic susceptibility. It is not as robust as proper analysis, but the use of one or both physical properties can be applied to large numbers of stones and is good for first-look classification.

Among other things, this permits the identification of ―ringers‖ (misidentified or misclassified stones). Consolmagno et al (2004) measured multiple stones from individual meteorites, and

13 observed that for some large showers there were a few stones that did not conform to the grain density and magnetic susceptibility of the rest of the shower. They attributed this discrepancy to mislabeling, either by intention or accident. In particular, one stone of L’Aigle (L6) had grain density and magnetic susceptibility unambiguously in the LL region, despite having documented provenance. Unfortunately, it may be an insurmountable challenge to determine whether the

―ringers‖ are indeed mislabeled stones or perhaps genuine members of the shower as heterogeneous inclusions in the asteroid .

It would be good to know the extent to which this classification scheme is effective, whether a significant number of meteorites reside as ―outliers‖ or whether the regions exhibit significant overlap with the addition of data from hundreds of OC falls. As a side benefit, this project may identify more ―ringers.‖

1.2.1.2 Classification of Other Meteorites

While the use of grain density and magnetic susceptibility to classify OCs and identify potentially mislabeled stones is promising, it raises the question of whether the technique may be applied to other meteorite types. In particular, any two groups that differ in total Fe and iron metal should also be distinguishable at least in part by these physical properties. One intriguing class of meteorites to which to apply this technique is enstatite chondrites, which are grouped into ―high-Fe‖ (EH) and ―low-Fe‖ (EL). In a review paper, Consolmagno et al. (2008) reported grain density measurements for only 14 EC stones of which there is a small difference between the two groups (EH being about 0.09 g cm-3 higher in grain density on average than EL) with

14 considerable overlap in the data, though it should be noted that grain densities were measured by different investigators using different methods. On the other hand, in a study of magnetic susceptibilities for 72 separate EC meteorites, Rochette et al. (2008) reported no difference in average magnetic susceptibility between EH and EL. This survey provides further data for both grain density and magnetic susceptibility, conducted on the same samples, that may help shed light on whether and to what extent there is a difference in grain density and/or magnetic susceptibility, and what that says about the total iron and metal content of the groups.

It may also be asked whether other meteorite groups may be distinguished in grain density or magnetic susceptibility. In addition to the metal-rich meteorites, the metal-poor meteorites (HED, SNC, lunites, etc.) may be distinguishable. Because they are lower in total metal, small differences should stand out in magnetic susceptibility, and since some are higher in denser minerals (such as dunite) than others, differences in density may also appear, though since dunite is not strongly magnetic, density and magnetic susceptibility may not necessarily correlate as they do in higher-iron materials. For example, do shergottites differ from nakhlites, both of which are Martian in origin?

Also intriguing is the question of the ranges occupied by each group. Different classes of meteorites (for example, CK and OC) may experience considerable overlap in physical properties, rendering the technique useless for distinguishing different meteorite classes from each other. Nevertheless, each group occupies a region in grain density-magnetic susceptibility space. Better definition of those regions will provide better understanding about mineralogy and variability within the group.

15

1.2.2 Questions about Weathering

Since terrestrial weathering significantly alters meteorite chemistry and structure, it is important to be able to understand how this effect influences the physical properties so as to better interpret the data, separating out terrestrial from parent-body influences.

1.2.2.1 Weathering in Ordinary Chondrites

The effects of weathering on ordinary chondrite physical properties have been described

(Bland et al., 1996, 1998b, 2006; see Section 1.1.2) and studied to some degree (Consolmagno et al., 1998, 2006). With an increased database of ordinary chondrite finds, this opens the possibility of exploring how strongly porosity is affected by weathering. At what porosity or range of porosities is access of terrestrial weathering agents into the interior cut off or significantly restrained, marking the shift from rapid stage 1 weathering to slow stage 2 weathering? Is the cutoff porosity zero, or some small but finite amount?

Another useful avenue to explore in ordinary chondrite weathering is whether physical properties may help quantify the effect itself. Since grain density and magnetic susceptibility of ordinary chondrites are affected by weathering in a predictable manner, this suggests the possibility that these two properties might be used to quantify the degree of weathering, at least in terms of oxidation of Fe metal. Can such a quantifier be constructed, and if so, do any weathering-related trends appear in the other measured physical properties: bulk density and porosity?

16

1.2.2.2 Weathering in Other Stony Meteorites

If a weathering quantifier can be constructed for ordinary chondrites, can it be applied to other meteorite types? The effect of OC weathering on grain density and magnetic susceptibility depends strongly on the amount of metal available, so presumably a quantifier of this sort could only be applied to high-Fe meteorites. These should include enstatite chondrites, but for carbonaceous chondrites there may not be enough total metal available to have a substantial effect on these physical properties. Nevertheless, there are other questions that may be asked.

For example, do other meteorite types differ substantially between falls and finds in any of the four physical properties (grain and bulk density, porosity, and magnetic susceptibility)?

With the exception of ordinary chondrites, there has been up to now a lack of good statistics for any of the meteorite groups. For many groups, discerning weathering-related differences will prove impossible even with this study due simply to lack of available examples of either falls or finds, especially since Antarctic finds have been excluded. For example, K chondrites are represented by a small number of finds and only one fall (Kakangari). Nevertheless, in at least a few cases the question can begin to be explored.

For the carbonaceous chondrites, statistics in this study are generally much lower than for ordinary chondrites, but better than for most achondrites excepting HEDs. Among carbonaceous chondrites, it was observed that CO finds tend to have lower porosity than falls (Consolmagno et al., 2008), a trend which may be due to the formation of terrestrial carbonates within pore space

(Greenwood and Franchi, 2004). Nevertheless, this finding is based on five CO finds and only one fall. Is the result a true effect of weathering, or was the fall (Warrenton) unusually porous?

This study contains multiple CO falls which will help shed light on this question. Another

17 question is whether the other carbonaceous chondrite groups experience similar reductions in porosity or any other physical properties due to weathering.

1.2.3 Questions about Shock

As described above, impact shocks both reduce porosity by compression and may increase porosity by the introduction of cracks. Consolmagno et al. (1998) observed that porosity decreases with shock state, leveling off above S4 when presumably the introduction of new cracks balances the compressive effect of shock. With the addition of more data, does this result remain robust? Also, can it be extended beyond ordinary chondrites? Are similar affects observed in OC or EC? There are sufficient data in this dissertation to begin to address these questions.

1.2.4 Big-Picture Questions

1.2.4.1 Questions about Lithification

One of the big unanswered questions about meteorites is how primitive materials from the solar nebula became the solid rock of meteorites. Even the mechanisms by which planetesimals formed and grew follow a number of diverse theories (cf. Boss and Goswami,

2006; Chambers, 2006; Cuzzi and Weidenschilling, 2006), but almost all of these theories are

18 satisfied with the formation of loose aggregates. Almost nothing is known of the mechanisms by which the collections of chondrules, inclusions, dust and debris became bound together to form new rock. The best idea is that these aggregates lithified through shock-induced heating and compression (Wiedenschilling and Cuzzi, 2006).

Regolith breccias, which are broken material that recombined and re-lithified, are also probably lithified by shock processes. Shock lithification of breccias has been extensively studied (cf. Ashworth and Barber, 1976; Bischoff et al., 1983; Clark et al., 1992). Since such a process will affect porosity, it seems reasonable to expect that porosities of chondrites should exhibit similarities to those of regolith breccias. In our database are hundreds of ordinary chondrite falls, of which many tens are regolith breccias. A comparison of the porosities of brecciated and non-brecciated ordinary chondrites may yield some insights into whether the process by which ordinary chondrites lithified is comparable to the process by which regolith breccias re-lithified.

1.2.4.2 Questions about Solar System Stratigraphy and Asteroid/Meteorite Relationships

Based on comparison of meteorites with their asteroid analogs, we have a generally good sense of where (roughly) in the solar system the meteorites originated, and so can compare physical properties (in particular, porosity) to other meteoritic properties that vary based on the location of origin. This may shed some light upon the processes under which they formed and perhaps provide constraints for models of parent-body formation and lithification.

19

In addition, comparison of seemingly unrelated properties between asteroids and their meteorite analogs may prove useful for understanding how meteorites relate structurally to their asteroid parent-bodies. In particular, is there any kind of correlation between microporosity as measured in meteorites and asteroid macroporosity? Since the two properties are thought to relate to different aspects of parent-body formation (microporosity is related to lithification and subsequent shock events, and macroporosity is related to the way separate stones of a parent body are packed), any observed relationship would raise further questions about the relationship between parent-body formation and lithification.

1.2.5 Other Questions

The most important questions are often the unanticipated ones. In the end, it is hoped that this dissertation primarily serves as a tool for use by other investigators exploring questions that have not been anticipated or that require data from additional properties that are not currently part of the database.

1.3 Organization of the Dissertation

In the next chapter, the methods by which meteorite physical properties have been measured are described in detail. Chapters three through six together form a report of the data resulting from this study, including some interpretation of results and a discussion of the

20 significance of some of those results. Each of those chapters addresses a major class of meteorites. Chapter three focuses on ordinary chondrites, chapter four on carbonaceous chondrites, chapter five on enstatite chondrites, and chapter six on achondrites. The final chapter

(7) serves as a summary and conclusion in which some of the important results of the study are highlighted and in which bigger-picture questions that pertain to data from more than one chapter will be addressed.

21

CHAPTER 2: MEASUREMENT METHODS

2.1 Background

Measuring physical properties on the scale of this study requires techniques that are fast, non-destructive and non-contaminating. While magnetic susceptibility meets this criterion easily, the measurement of density and porosity has historically required methods that risk contamination or the deposition of residues, or involve altering the sample, and this in turn has limited the number of meteorites for which porosity has been measured. Porosity requires the determination of two different densities (bulk and grain) from their corresponding volumes. The problem here is that meteorites are irregularly-shaped solids, making determination of volume challenging. In addition, different methods will result in different volume determinations, depending on whether and to what extent the medium of measurement penetrates pore space within the sample. Bulk volume requires minimal pore penetration, and grain volume requires the opposite extreme, but many methods will result in something in-between.

Many of the methods employed for determination of density or volume utilize liquids and the principle of Archimedes. For example, Keil (1962) measured densities for 63 ordinary chondrites by submersing them in water, a technique also employed for measuring bulk density of 368 meteorites by Pesonen et al. (1993). Aside from contamination concerns, water may migrate into pore space after even a short exposure, resulting in an overestimate of bulk density.

22

Fujii et al. (1981) utilized a clever technique in which the samples were first thoroughly soaked in toluene to the point that effectively all pore space was filled. After submerging the sample in toluene to determine grain density, they then submerged the sample in water (immiscible to toluene) to measure bulk density.

Matsui et al. (1980) modified the Archimedean method by first wrapping their sample in clear plastic, then encasing it in clay before submerging. They varied the amount of clay, thus varying the total volume, and then extrapolated their data to zero clay to determine the volume of the meteorite without the clay wrapping. To avoid liquid contamination, other researchers have avoided Archimedean methods altogether. For example, Yomogida and Matsui (1981, 1982,

1983) cut their samples into precise parallelepipeds so their volumes could be determined by precise measurement of linear dimensions (length × width × height). A method that shows promise for the future is modeling volume through 3-D laser scans (e.g. Herd et al., 2003; Smith et al., 2006; McCausland et al., 2007), though at the moment there are some problems with the technique, both in terms of data acquisition (low-albedo surfaces such as fusion crusts are hard to image) and in terms of analysis (which can take hours per sample), that prevent its use on large numbers of meteorites. The use of x-ray microtomography (e.g. McCausland et al., 2010) for creating 3-dimensional models also shows considerable promise for future applications and avoids some of the problems of 3-D lasers, though the instrumentation is very expensive and requires qualified operators.

Helium ideal-gas pycnometry, a non-destructive and non-contaminating method in which the pressure change of a gas expanding over a specific volume yields volume displaced by the sample, has been widely available for measurement of grain volume/density since the introduction of an apparatus in 1955 (Faeth and Willingham, 1955). However, the development

23 of a reliable non-destructive and non-contaminating method for bulk density determination would have to wait until the late 1990s. Consolmagno and Britt (1998) developed an

Archimedean method in which spherical glass beads serve the role of the fluid. The glass beads are easily removed, do not interact chemically with the sample, and in the event of a few beads remaining after measurement are easily distinguished from native minerals in the sample. The technique has become widely employed since its introduction, and shows promise for application even beyond the field of meteoritics.

2.2 Grain density: Helium Ideal-gas Pycnometry

Measurement of grain density requires immersion of the sample into a medium that penetrates and fills pore space, so as to be displaced only by the solid material of the sample.

Gases are the preferred medium for such measurements, as they can more easily flow through small cracks and other openings, and are less likely to leave residues. The use of gas for density measurement is possible due to the ideal gas law, which expresses the relationship between the pressure of gas, its volume and temperature. The better the medium approximates an ideal gas, the better it is for this method. Gaseous helium is the best medium for this use. It is an inert monatomic gas which, due to the two electrons in its S1 valence shell, has the smallest atomic radius of any element, allowing it to easily penetrate very small voids, yet it does not penetrate the crystal lattice of silicates and metals.

Faeth and Willingham (1955) developed an apparatus for measuring grain densities using helium pycnometry, and a variant of their apparatus was employed in the early 1970s to measure

24 grain densities of some lunar samples from the Apollo program (e.g. Cadenhead et al., 1972;

Cadenhead and Stetter, 1975), during which many of the fundamental questions about the method were resolved. Yomogida and Matsui (1981, 1982, 1983) employed a commercially available pycnometer (Shimadzu Seisakusho model 1302) for measurements of grain density on numerous Antarctic meteorites from the Japanese expeditions. Other devices for the same technique have been extensively used in meteorite studies (e.g. Kuoppamäki et al., 1996;

Pesonen et al., 1997; Consolmagno and Britt, 1998; Flynn and Klock, 1998; Flynn et al., 1999;

McCausland and Flemming, 2006), and today the method is quite common for grain density measurements of meteorites.

2.2.1 Theory

The ideal gas law, expressed simply, is the basic relationship between the gas pressure P, the volume it occupies V, the temperature of the gas T, and the number of moles of gas n.

PV nRT (1)

where R is the gas constant (8.314472 J K-1 mol-1). In the specific case where the temperature and quantity of gas are held constant, the pressure and volume vary together:

PV constant (2)

25

PVi i P f V f (3)

By varying the volume that the gas occupies and measuring the variation in pressure, one can take advantage of this property to determine the volume Vs displaced by an object placed within the gas. Consider the case of a chamber that changes its volume from V1 to V2, both of which are known quantities (Figure 1). In this case, the volume occupied by the gas is (at the beginning)

V1-Vs, and (at the end) V2-Vs. Their corresponding pressures are P1 and P2. Equation 3 becomes:

PVVPVV1()() 1SS 2 2 (4)

Solving Equation 4 for VS yields:

PV1 1 PV 2 2 VS (5) PP12

2.2.2 Measurement

To perform pycnometry, we utilized a Quantachrome Ultrapycnometer 1000* (Figure 2), manufactured by Quantachrome Instruments (Boynton Beach, FL). This device employs two chambers milled out of solid aluminum blocks and connected through a valve (Figure 3).

*As of April 2010, we upgraded to a Quantachrome Ultrapyc 1200e. This device operates on the same principles using the same basic engineering (and same cell specifications) as the Ultrapycnometer 1000. The primary difference is in the electronic control systems which are more sophisticated and are compatible with computer interface. Measurements conducted at the Field Museum (Chicago) were performed with the 1200e. 26

Volume change is facilitated by the opening of the valve, transitioning the effective volume from that of the one chamber (known as the ―cell‖) to that of both the cell and the second chamber

(called VA, to maintain notation consistent with the operator’s manual). To complicate the matter a bit, we do not evacuate the chambers (a process that risks causing migration of material out of the meteorite), and the transducer measures pressures above local atmospheric pressure

PA.

Initially, the meteorite is placed in the cell, and atmosphere in both the cell and VA is replaced by helium at atmospheric pressure through a flow purgation process. Then the valves are closed and helium is pumped into the cell to an initial overpressure Pi. Thus, there are two equations of state to consider:

(PA P i )( Vcell V S ) n 1 RT (6)

PAA V n2 RT (7)

Each refers to a different quantity of helium (one for the cell and the other for VA). Adding

Equations 6 and 7 yields the equation of state for the whole system.

(PA P i )( Vcell V S ) P A V A ( n 1 n 2 ) RT (8)

After the valve is opened, the pressure equilibrates between the two chambers. The final pressure Pf is measured. This final state is:

(PA P f )( Vcell V A V S ) ( n 1 n 2 ) RT (9)

27

Comparing the initial to the final state results in the following:

(PPVVPVPPVVVA i )(cell S ) A A ( A f )( cell A S ) (10)

The atmospheric pressure PA cancels out of the equations, yielding:

PVVPVVVi()()cell S f cell A S (11)

This is effectively the same as the basic ideal-gas relationship in Equation 4, and the solution has the form of Equation 5, but with the appropriate terms substituted in:

PVicell P f() V cell V A P f VVVSAcell (12) PPPPi f() i f

To perform this calculation on real measurements, both Vcell and VA must be known.

Since these volumes include complicated features such as metallic tubing that may vary somewhat depending on temperature they must be measured directly rather than assumed based on geometry. These are measured to high precision using a set of stainless steel calibration spheres. There are in fact three cell sizes (large [4.8 cm dia. x 7.5 cm tall], medium [4.0 cm dia. x 3.9 cm tall] and small [2.4 cm dia. x 2.2 cm tall]) and two VA sizes (one used for both large and medium cells, and one for the small cell). Separate calibrations must be made for each of them.

28

2.2.3 Other Considerations

2.2.3.1 Ramp-up

The pycnometer has a documented ramp-up (Quantachrome, 2003), where initial volume measurements have lower value than subsequent ones, as can be seen in Figure 4. This is due in part to the sample not being in complete thermal equilibrium with the device, despite the practice of placing the sample in the cell storage compartment of the pycnometer which is designed to equilibrate with the rest of the device. It may also be due in part to an incomplete exchange of air inside the sample for helium. To accommodate this, each measurement included 15 consecutive runs, and the reported grain volume was taken from the average of the last five to six of them.

At this point, the ramp is not completely leveled off, but the slope and variability is greatly reduced after 10 runs. Calibration measurements are also based on a 15-run measurement, so sample volumes should be reliable. Measurements on zero-porosity quartz and topaz standards are self-consistent under separate calibrations, and are also consistent with volume measurements made using the bead method (Macke et al. 2010a).

2.2.3.2 Permeability and Penetration of He

While helium permeates even tiny openings, it is not 100% permeable through every substrate. (If it were, it would be impossible to store.) Pores that are fully enclosed within

29 minerals will not be permeated. However, most meteorites are shot through with microcracks, and helium has the ability to penetrate even very small cracks, and the high pressures employed

(>10 psi over atmospheric) assist the mobility of the gas, so this concern is minimized in most cases. We see in studies of large carbonaceous chondrite and ordinary chondrite showers

(Macke et al., 2008) very little variability in grain density between stones from the same meteorite, suggesting that helium is penetrating each stone to the same degree. If permeability were an issue, large stones would have demonstrably lower grain densities than small stones due to helium failing to penetrate the interior of the rock. This trend is not observed.

A related issue is that of air and other volatiles residing deep within the rocks. In theory, if these gases were not being replaced by helium in the measurement, it should affect results, again with a size-dependent bias. Cadenhead and Stetter (1975) recommended placing the sample under vacuum and baking it at above 45 C to force the outgassing of any volatiles. This approach has drawbacks, such as affecting future measurements of the same volatiles within the sample. Also, exposure to vacuum and removal of moist terrestrial air has been observed to force migration of certain water-soluble interior species to the exterior and to accelerate weathering (Gounelle and Zolensky, 2001). For these reasons, vacuum exposure is not acceptable for large-scale use. Given the somewhat high pressures used by the Ultrapycnometer and a sequence of multiple consecutive measurements per sample resulting in constant transport of helium in and out of the meteorite, atmospheric gases will be largely replaced by helium as of the final few measurements. Again, in our studies of large meteorite showers we see no evidence for size-dependent bias in grain density results that would be anticipated if this concern had significant effect.

30

2.2.3.3 Residual Helium Deposition

A concern when introducing any foreign material into a sample is whether the medium will leave a residue of any kind. This is an especially important consideration with helium because even a small amount of residual helium will affect possible trace noble gas measurements in the future. Of particular concern are carbonaceous chondrites, with their high carbon content that, in a manner similar to activated charcoal, may adsorb helium. This effect is expected to be negligible for this work (Steele and Halsey, 1954; Kini and Stacey, 1963).

Indeed, trace noble gas measurements were performed on a sample of ordinary chondrite before and after pycnometry, with no difference observed (Tim Swindle, private communication).

2.2.3.4 Measurement Uncertainty

Based on the repeatability of the pycnometry measurements on the same samples (quartz and topaz) under different machine calibrations, a uniform uncertainty was set for all measurements made with the same cell. For the large cell, uncertainty is 0.06 cm3. For medium, it is 0.04 cm3, and for the small cell it is 0.02 cm3. These uncertainties are generally much larger than the statistical variation among the last six runs in a set, and reflect repeatability of the instrument under varying conditions rather than precision of a specific measurement. In those rare instances where the standard deviation of the six runs used for grain density determination exceeds the preset uncertainty, the standard deviation is used instead. What is not factored in to the quoted measurement uncertainties is the effect of any impermeable pore space on the grain

31 volume. For grain densities, uncertainties are calculated by standard propagation of errors using uncertainties in sample mass and grain volume.

2.3 Bulk Density: Archimedean Glass Bead Method

Bulk volume measurements require a medium or technique that does not penetrate the interior of the meteorite, providing instead a measure of the volume enclosed by the outer boundary. To accomplish this, we have employed a method based on that developed by

Archimedes. In the Archimedean method, the object of unknown volume is submerged in an uncompressible fluid of known density. (Often water is used for the fluid.) The volume of fluid displaced is the same as the volume of the displacing object. The displaced volume can be found by measuring the mass of the system, as described in Section 2.3.1. We employ this method, but substitute non-contaminating glass beads to serve as the fluid. As well as being non- contaminating, the beads also do not penetrate interior pore space, thus yielding a reliable measure of bulk volume.

2.3.1 Theory

The basic principle operates as follows: Begin with a container of known mass mcup and volume Vcup. Fill it completely with the fluid, so as to overflow the container. Then insert the sample, after measuring its mass mS. Excess fluid will spill out of the container. (For the sake of

32 this argument, effects of surface tension such as a meniscus are ignored. They will be irrelevant anyway when the technique is applied to glass beads.) The mass mtotal of the filled cup is then measured. The volume displaced by the sample is the volume of the cup minus the volume of the fluid remaining in the cup.

VVVS cup fluid (13)

The volume of fluid remaining in the cup, however, is determined by the mass of fluid remaining and its density.

mfluid mtotal m cup mS Vfluid (14) ρρfluid fluid

Combining Equations 13 and 14 yields:

mtotal m cup mS VVS cup . (15) ρfluid

2.3.2 Measurement

As described in the introduction to this chapter, the use of an actual fluid of any type is highly problematic for bulk density measurements of meteorites, not only because of the possibility for contamination but also because liquids are likely to penetrate interior space, thus

33 throwing off the measurement. To get around this problem, Consolmagno and Britt (1998) developed a method for performing the measurement using small glass beads that collectively behave as a fluid. The beads, as used in this study, include two different size-sorts: 40-80 µm in diameter, and 700-800 µm. We use BALLOTINITM impact beads manufactured by Potters

Industries Inc. (Valley Forge, PA). These beads are produced in large quantities for industrial use, and so are widely available. (See Table 1 and Table 2, and Figure 5 for chemical analysis of the beads, courtesy R. Korotev, personal communication.) The smaller beads provide greater precision of measurement, but the larger beads are preferred for measurements at many collections because they can more easily be removed from the samples.

Replacing the term ―fluid‖ in Equation 15 with ―bead‖, the calculations do not vary:

mtotal m cup mS VVS cup (16) ρbead

One thing that does vary, however, is the density of the beads. It is not a perfectly incompressible fluid, and local environmental conditions may affect packing efficiency and hence bead density. The density of beads must then be measured directly through calibration measurements, in which the cup is filled entirely with beads but with no sample present. Bead density is then:

mmcalib cup ρbead (17) Vcup

34 where mcalib is the total mass of the bead-filled cup in the calibration measurement. Substituting

Equation 17 into Equation 16 yields a new form for the sample volume calculations:

mtotal m cup mS VVS cup (18) mmcalib cup

Vcup

This simplifies to:

mcalib mS m total VVS cup (19) mmcalib cup

The apparatus used in this study is seen in Figure 6. In this study, we made multiple measurements per sample in order to establish statistical uncertainty, since mtotal and mcalib may vary by as much as 1% between two measurements made under the same conditions. Typically, five sample measurements are sandwiched between two sets of five calibration measurements, with the averages used in calculations of sample bulk volume. Uncertainties for both mtotal and mcalib are determined statistically, and uncertainties in bulk volume and density are calculated by propagation of errors.

To avoid bias based on orientation of the meteorite, it was rotated between each measurement. Under some circumstances the number of measurements was varied. For example, for some friable meteorites the number of sample measurements was reduced to four or three to minimize the possibility that the beads would abrade the sample.

35

For each measurement, the beads are poured into the cup to the point of overflowing, and are encouraged to settle. Different settling methods were tested (see below), but we settled on a

―secured shake‖ method in which the sample is held down onto a vibrating platform that shakes for a duration of 5 seconds. At the cessation of the shaking, the surface is scraped level using a straight edge, and any beads sticking to the outside of the container due to static are carefully brushed off before the container is massed.

2.3.3 Settling Methods and Systematic Error

The glass bead method assumes that the beads behave as an incompressible fluid, but this is not the case. In particular, the beads do not flow perfectly as a fluid; nor do they maintain at all times the most efficient packing arrangement. Uneven flow around irregular samples may introduce inhomogeneities in bead density. Environmental conditions may affect packing efficiency. While the Archimedean glass bead method has become widely applied, little work has been done either to standardize the method (necessary for comparison of results), to establish the degree of bias present in a given method, or to establish which variations of the method minimize bias. Existing systematic error studies in the bead method have been limited to specific measurement methods employed by other investigators, which do not precisely match our method or methods applied elsewhere. For example, Wilkison and Robinson (2000) discuss an elaborate setup in which the bead-filled cup is suspended in a container filled with foam pellets. One of the first steps in this meteorite physical property survey was to study systematic error in our method and to establish a standard method to be applied throughout the survey. This

36 has been published in Macke et al. (2010a), and much of the following discussion is taken from that paper.2 In addition to limiting ourselves to methods compatible with our measurement apparatus which is notably simpler than Wilkison and Robinson’s, we sought a method that all investigators may apply, and so applied the criteria that the methods should be easily repeatable, simple to employ (not requiring an overly sophisticated apparatus) and minimally contaminating of either the beads or the sample.

For the most part, the key variable here is the method of settling the beads so they flow around the sample and achieve a good packing arrangement. To this end, I tested various methods of settling the beads, ranging from no settling to a short vigorous shake with the container secured to the shake platform. Because early work with the bead method employed the small bead sizes (40-80 µm), the most thorough testing has been performed on that size beads, and so most of the following discussion focuses on that size. By the time the 700-800 µm beads began to be employed, many of the other settling methods had already been ruled out as less accurate, and so only the dominant settling method was tested with the larger beads.

To perform the tests, we used zero-porosity standards composed of quartz, ranging in mass from 9.3 g to 89.9 g. Because quartz has a low density (2.62-2.65 g cm-3) compared to the

~3.3 g cm-3 of meteorites, we added three topaz samples (density 3.54 g cm-3) ranging in mass from 50 to 67 g. Both the quartz and topaz standards are fully crystalline, with no visible inclusions or voids. Because bulk and grain volumes for zero-porosity samples are the same, the actual volumes of the standards were determined via helium pycnometry. All of the bead method measurements were performed using two cup sizes: 77 cm3 and 155 cm3. For all methods, the 155-cm3 cup exhibited greater overall uncertainties, but volume measurements were within uncertainties of actual volumes. Smaller cups yield greater precision, so while the

2 Those portions of the dissertation that come from this source are used with permission. See Appendix C. 37

77cm3 cup does produce measurable systematic error, it is generally preferred over the larger container for small samples. Unless otherwise noted, the following refers to the 77-cm3 cup, which is the cup size employed for the majority of meteorite measurements in this survey. (The larger cup is employed only where the meteorite does not fit in the smaller cup.)

The four settling methods that were tested are as follows: Method #1 is a control, in which no physical settling method is employed. Beads are merely poured directly into the container, and the surface is leveled. Method #2 employs five taps on the side of the container using the bristles of a soft brush. It will be referred to hereafter as ―soft tap.‖ Method #3 employs a vigorous five-second shake using the shake platform, while the cup sits freely on the surface of the platform (―free shake‖). Method #4 is the same as method #3, except the cup is manually held securely onto the surface of the platform (―secured shake‖). Additional variations were also tested in an attempt to identify an optimal settling method, but none of those represented any improvement, and so are not reported here. For each method, the test involved multiple bulk volume determinations on each of the standards, and the results were compared with the volumes obtained through pycnometry.

2.3.3.1 Method #1. No Settling

Though it hasn’t yet been employed in our research, there are circumstances under which it may be desirable to skip the settling of beads altogether, for instance in the measurement of extremely friable meteorites. We also wanted to establish whether settling itself had an effect on measurement. For this method, we poured beads directly into the container and leveled it off

38 without any shaking or tapping. For some measurements, we placed a thin bed of beads into the bottom of the cup before inserting the sample. In other cases, we placed the sample into an empty cup and poured beads atop it. We found (Table 3) that this method tends to overestimate the bulk volume of the sample by as much as 3%. The average volume overestimate is 1.6% ±

2.4%, with the ―±‖ representing one standard deviation among the individual results unless otherwise stated. The average uncertainty in individual measurements was 2.0%. As will be apparent for all settling methods in this study, the two smallest quartz samples (masses 9.3 and

9.8 grams) exhibited highly unreliable results, with individual measurement uncertainties near 4 percent. This may be expected because precision breaks down at small sample volumes.

Eliminating them from consideration results in an average overestimate of 2.2% ± 1.5%.

This method does exhibit volume dependence (Figure 7). For samples sized 7 to 10 cm3, the bias is near 3 percent, but this falls with larger volume. Our largest sample, at 29.0 cm3, exhibited an overestimate of only 1.5% ± 0.5%. Eliminating the three smallest samples, the remaining data fit the following relationship with a correlation of 0.9999:

3 VVmeas1.009 actual 0.187cm (20)

2.3.3.2 Method #2. Soft Tap

The soft tap method using taps from a soft-bristled brush produced an overestimate of sample volume in both quartz and topaz (Table 3), though to a lesser degree in topaz. On average, quartz volumes were overestimated by 2.7% ± 2.6%. Topaz volumes were

39 overestimated by 1.4 % ± 0.7 %, though uncertainties in individual measurements range up to

1.4%. The data show no obvious mass or volume dependency (Figure 8), though the spread in results for the two smallest quartz crystals is significantly larger than that of the rest of the set.

Removing them, the average quartz volume overestimate is 2.9% ± 1.6%.

This method is strongly dependent on a number of factors that may be difficult to control.

For example, the strength of the tap and even the hardness or number density of the bristles may affect how well the beads settle. By experience we have found that it is difficult to maintain consistency between measurements in particular with regard to tap strength. This may explain the inconsistency in the results. In theory, this method should not produce different results for topaz than for quartz since the method should not allow for the sample itself to greatly influence the packing of the beads.

2.3.3.3 Method #3. Free Shake

Unsecured shaking of the cup tended to underestimate sample volumes in quartz (Table

3), but with a volume-dependent trend visible in the quartz data (Figure 9). At low masses (and hence small volumes), volumes are underestimated by approximately four percent, but this is reduced for masses above 40 grams to near zero. As with the soft tap, the two samples below 10 grams produced unreliable results. This curve can be approximated by a linear relationship between volume measurements:

3 VVmeas1.014 actual 0.48cm (21)

40

This relationship fits the existing data (not considering the three smallest quartz samples) quite well, with a correlation over 0.999. Nevertheless, it will probably not hold for samples significantly larger than those in the study, since the trend appears to be asymptotic to zero while the above relationship does not have that quality. For larger samples, the bias can be assumed to be zero.

2.3.3.4 Method #4. Secured Shake

Secured shaking is perhaps the most thoroughly studied method, since early indications showed it had the least systematic error over the range of quartz masses, though for masses below 10 grams it still suffers from large uncertainties. For quartz (not including the two masses below 10 g), volume discrepancy averages 0.5% ± 0.8%. This fits the average within the 0.4% to 1.1% range of uncertainties of most of the individual measurements. With topaz, the average volume bias is a slight underestimate of -0.5% ± 0.7%, again putting the average within the uncertainties of individual measurements. In short, this method does not appear to be a significant source of systematic error overall. No clear volume dependence is apparent either

(Figure 10).

41

2.3.4 Further Considerations

2.3.4.1 Larger Bead Size

Most of the meteorites in this survey were measured with glass beads of substantially larger size, with radii ranging from 700-800 μm in diameter, of the same manufacture as the 40-

80 μm beads. The larger bead size enhances their visibility, making them easy to remove completely from meteorites following measurement. The ―secured shake‖ settling method was employed for these measurements. Systematic error studies using the same quartz and topaz standards were performed using the larger beads in the ―secured shake‖ settling method. These larger beads exhibit a mass- or volume-independent systematic error (Figure 11). For the 77-cm3 cup, the volume was overestimated by an average 2.0% ± 1.0%. Large uncertainty is seen for not only the two smallest quartz samples, but also the 14 g sample, indicating that the method is best used for samples greater than 15-20 g (or a corresponding ~4 cm3). Eliminating data from these small samples, the volume overestimate is more consistently overestimated by 2.3% ±

0.4%. For the 155-cm3 cup (Figure 12), we eliminated the smallest samples as they were not measured using this method, and included measurements of combinations of the three topaz pieces to represent larger-volume samples. For the larger cup, the volume discrepancies all but disappeared, producing a slight average volume underestimate of 0.16% ± 1.34%, well within the individual measurement uncertainties of approximately 1.7%. As with the smaller beads, measurement precision using the 155-cm3 cup was significantly reduced compared to the 77-cm3 cup.

42

2.3.4.2 Environmental Effects

Three environmental factors (temperature, pressure, and relative humidity) were recorded as the bead method was applied, both during systematic error studies and during some meteorite measurements. This yielded 72 independent measurements of bead density ρbead for the small cup and small beads, covering a temperature range of 21.5 to 26 C, atmospheric pressure of 997 to 1026 mb, and humidity range from 31% to 68%. For the 155-cm3 cup and small beads, we have 25 measurements covering a similar range of conditions. No correlation was observed between bead density and temperature or pressure. However, a clear negative trend is observed with relative humidity. For each percentage increase in relative humidity, bead density (for small beads) decreases by approximately 1.×10-3 g cm-3 (see Figure 13). This correlation held for both cup sizes, with the ―secured shake‖ settling method. For the ―soft tap‖ method a negative trend is also observed, but there are too few data points for meaningful quantitative analysis. This negative density trend may be due to an increased cohesion between beads as humidity increases. This in turn would reduce the ability of beads to flow freely into the optimal packing arrangement, leaving them in a packing arrangement that has lower bulk density. The cohesion of beads in humid environments is supported by visual observation of a thin ―crust‖ developing at the surface of the bead pail after it has been left sitting for an extended period of time.

This raises the question of whether humidity influences bulk volume in sample measurements using the ~60 µm beads. It is not clear that there should be any such influence; after all, if calibration measurements are made under the same environmental conditions, bead density variations should be accounted for and no systematic error should be introduced. We

43 performed a Monte-Carlo simulation that confirms this intuition. However, there may be effects other than variations in bead density that would influence results. For example, if humidity impedes the smooth flow of beads, then they may not completely fill small gaps and cavities in the samples to be measured. This would result in an overestimate of volume that would increase with humidity. With current observations of zero-porosity standards, however, no such trend is observed. Nevertheless, the current data exist over a relatively small humidity range and further measurement may yield different results. Given the relatively small magnitude of the effect, we do not expect the influence to be noticeable beyond ordinary measurement uncertainties.

We also recorded environmental conditions for measurements performed with the ~750

µm beads. We have 208 measurements with the small cup and 47 with the 155-cm3 cup, covering a temperature range of 20 to 24.5 C, atmospheric pressure of 844 to 1022 mb (852 to

1021 mb for the 155-cm3 cup), and relative humidity range of 22% to 56%. As in the case with the small beads, there is no correlation between bead density and temperature or pressure. On the other hand, there is also no correlation between bead density and humidity for the larger size- sort of beads. Whatever phenomenon results in reduced bead density for small beads does not apply to the larger size. This emphasizes the fact that the effect is not a property of the glass itself, but results most likely from the high surface-area-to-volume ratio in the small beads.

2.4 Magnetic Susceptibility

As a medium is exposed to an external magnetic field, this will induce a magnetic response in the medium, based on the amount and type of magnetic materials within the medium. The magnetic response M relates to the imposed field B linearly by a factor χν.

44

M χBν (22)

This factor, the magnetic susceptibility, is a unitless intrinsic property of the material. Magnetic susceptibility of a rock is most strongly affected by those minerals contained within that have very high magnetic susceptibilities, such as ferromagnetic materials like iron metal. As such, this physical property serves as a reasonably good first-order indicator of the quantity of total metallic iron content within a specimen.

Because instruments for measuring magnetic susceptibility are widely available and most are quite portable, numerous studies have already been conducted on meteorite magnetic susceptibility. As early as the 1960s, Russian investigators surveyed magnetic susceptibilities of over 900 meteorites as well as numerous lunar samples from the Russian Luna program (see

Herndon et al., 1972, for a review of their results), though their work has remained largely unknown outside Russia. More recently, a collaborative effort including Pierre Rochette, Jérôme

Gattacceca and others has resulted in a substantial database of magnetic susceptibilities for representatives of most meteorites of all types (Rochette et al., 2003, 2008, 2009, 2010).

In this dissertation, I have included magnetic susceptibility primarily because it had already been shown to have correlations with grain density in ordinary chondrites (Consolmagno et al., 2006) and was worth further study. Many of the above studies, especially those by

Rochette et al., are more extensive in their scope and the total number of stones measured.

However, for comparison purposes it is important to have reliable measurements performed on the very same stones for which I measured densities and porosity. Magnetic susceptibility may vary stone-to-stone within a given meteorite, depending on degree of homogeneity, and so it would not be sufficient simply to use literature values for this property. To further explore

45 relationships between meteorite grain density and magnetic susceptibility or porosity and magnetic susceptibility, what was needed was a fresh measurement of magnetic susceptibility for each stone for which density measurements were performed.

2.4.1 Instrument

The instrument we use is a ZH-Instruments SM-30 handheld magnetic susceptibility meter (Figure 14). It was originally designed for measuring magnetic susceptibilities of large boulders in the field. This device exposes the sample to a low magnetic field created by an electric oscillator at a frequency of 9 kHz (Terraplus, 2003). The presence of magnetic materials within the sample induces a shift in frequency proportional to the total magnetic susceptibility.

Two measurements are performed: one with the sample, and a subsequent measurement on air.

The instrument determines the magnetic susceptibility of the sample by subtracting the air measurement from the sample measurement (Terraplus, 2003).

The device contains a magnetic pickup coil that lies up against the back surface of the machine, with the controls on the front (Figure 14b). For normal operation, the device rests against the object of interest, its back facing the object, with the controls facing the user. In performing measurements on small meteorites, the normal operation of the device is altered somewhat. It is inverted, placed face-down on the edge of a nonmetallic surface (such as a wooden table or plastic crate), sticking out from the edge enough to expose the pushbutton for activating a measurement. The meteorite is placed atop the device, on what is normally the backside and centered at the location of the coil (Figure 15). For the air measurement, the

46 meteorite is removed while the meter and everything else remains in place. The meter reports volume magnetic susceptibility χν, which is unitless in SI units. For consistency with magnetic susceptibility measurements reported in the literature, these are converted into mass magnetic susceptibility by dividing by bulk density:

χ χ ν (23) ρbulk

Mass magnetic susceptibilities are recorded in units of 10-9 m3 kg-1. Following measurement, further adjustments of the results must take place to accommodate the sample size and geometry.

These are discussed further in the following sections. Because meteorite magnetic susceptibilities vary over many orders of magnitude, the final results are reported in log units.

2.4.2 Adjustments for Finite Sizes

As stated above, the SM-30 is designed for use on large boulders, which approximate an infinite half-plane in the perspective of the meter. The magnetic field of the device remains strong (relative to maximum strength) throughout a cylindrical region approximately 8 cm high and 5 cm in diameter (Folco et al., 2006), extending well beyond the volume enclosed by most of the hand-sized meteorite stones in this study, which due to size limitations of the pycnometer chamber (see Section 2.2) do not extend more than 5 cm above the meter except in rare cases.

Because much of the magnetic field lies outside the sample, the magnetic susceptibility reading

47 will be much lower than the actual magnetic susceptibility of the sample. A correction factor α is required:

χ χ uncorrected (24) corrected α

The correction factor α was determined experimentally by Gattacceca et al (2004) by comparing

SM-30 measurements of 315 volcanic pebbles of varying size and mineralogy with measurements on the same pebbles made using a Kappabridge KLY2 Geofysika susceptometer

(Figure 16). The given correction factors are as follows:

0.78,V 1000 cm3 0.1972VV0.1984 , 200 cm 3 1000 cm 3 α 0.1744VV0.2228 , 40 cm 3 200 cm 3 (25) 0.0524VV0.5514 , 10 cm 3 40 cm 3 0.0254VV0.8776 , 10 cm 3

This correction alone does a reasonably good job of matching SM-30 results with expected magnetic susceptibilities, and has been field tested in Antarctica with good results

(Folco et al., 2006). In our database are measurements from a number of stones in the Vatican collection, many of which had also been measured by Pierre Rochette using the KLY-2

(Rochette, personal communication). The correlation is not bad, with most SM-30 measurements within about log χ = ±0.2 from KLY-2-derived values. However, as Gattacceca et al. (2004) note, measurements made by the SM-30 are dependent somewhat on shape as well as volume. Their correction assumes that the object is roughly ellipsoidal. The results are most

48 strongly affected when the shape of the object diverges significantly from an ellipsoid. The most notable affects occur when there are polished surfaces, which sit perfectly flat on the meter, presenting a much larger volume to the strongest part of the field than is assumed by the correction.

Further corrections can be made by first classifying stones based on simple shape. Stones in the Vatican collection can be characterized by three basic shapes: ―slabs‖, ―sliced chunks‖, and ―chunks‖. Slabs have been sliced on two opposing sides, presenting the entire mass of the stone in a relatively thin layer just above the strongest part of the field, with nothing but air above that. This group also includes ―end caps‖, for which one side is sliced and the opposite is rough, but when laid flat the entire stone is thin in the vertical dimension compared with the horizontal. Sliced chunks, many of which are stones cut in half, present a large flat surface upon which the stone rests during measurement, though the whole stone has a substantial extension in all dimensions. The remainder of stones fell into a catch-all category, but are characterized by an uneven contact surface and substantial extension in all dimensions.

For slabs, SM-30 measurements (after the first-order correction) were reduced by 0.105.

For sliced chunks, the reduction was 0.04. Chunks showed a magnetic-susceptibility-dependent offset that became more pronounced as magnetic susceptibility increased. This is consistent with a small non-linear response of the device for high magnetic susceptibilities (SM-30 manual), and is easily accommodated. If magnetic susceptibility was below 4, no correction was necessary.

Above 4, it was corrected according to the formula

logχχcorrected 1.037 log previous 0.097 . (26)

49

Figure 17 and Figure 18 compare the SM-30 data with the KLY-2 data before and after these corrections have been applied. As is apparent, the fit is much tighter after the corrections have been applied.

Reported measurement uncertainties for magnetic susceptibility measurements were based on one standard deviation of the discrepancy in log χ space for each geometrical type according to the Vatican data, with a minimum allowable uncertainty set at ±0.08 consistent with findings of Folco et al. (2006). For most stones, the uncertainty lay between 0.08 and 0.12. An exception was applied in the few cases where the uncertainty derived from the original correction factor exceeded that based on the shape, in which case the larger uncertainty was used.

50

CHAPTER 3: ORDINARY CHONDRITES

3.1 Introduction

Ordinary chondrites are by far the most abundant kind of , making up about

74% of all observed falls. However, in part because they are less likely to stand out against a background of terrestrial rocks than iron meteorites, and in part because when they weather they are more likely to disintegrate, they only constitute 63% of collected finds (Grady, 2000). Even that number is higher than historical reports; prior to the extensive collection of well-preserved meteorite finds in hot desert regions and Antarctica, Mason (1962) reported the percentage of chondrites among finds at less than 35%. As their name suggests, ordinary chondrites are chondrites, that is, they are characterized by the presence of spherules of predominantly olivine and other silicates that condensed from melt droplets. The origin of the chondrules is still a subject of study, but they may have been caused by localized and rapid heating and cooling of clumps of dust particles within the early solar nebula (Connolly et al., 2006). Space between chondrules in chondrites is filled by a matrix composed of fine-grained particles.

Chondrites are primitive meteorites; structurally, they have seen little change since the initial formation of their parent bodies. Nevertheless, they exhibit signs of post-formation processing. Most ordinary chondrites have at least some history of thermal alteration leading to partial equilibration of chondrules with surrounding matrix material, and giving them

51 petrographic types ranging from 3 to 6 (and in some cases 7). They are also affected by shock events, such as impacts, that compress existing pore space yet at the same time introduce new pores by forming cracks.

Ordinary chondrites are further subdivided into three main groups, plus two intermediate groups with smaller populations. These differ by total quantity of iron and ratio of metallic iron to FeO, and have been named to reflect that. H chondrites are high iron, typically 8 vol % Fe

(Righter et al., 2006) and Fe0/FeO of 0.58 (Weisberg et al., 2006). L chondrites are lower in iron

(about 3 vol% Fe and Fe0/FeO of 0.29), and LL chondrites are lower still (1.5 vol% Fe and

Fe0/FeO of 0.11). The intermediate groups, H/L and L/LL, exhibit intermediate quantities of iron and Fe0/FeO ratios.

Because the three groups differ so dramatically in their abundance of total iron and metallic iron, magnetic susceptibility has been posited as a good first-glance physical property for rough classification of ordinary chondrites (Rochette et al., 2003) and this has been tested in

Antarctica (Folco et al., 2006). Nevertheless, some overlap still exists between the groups.

Consolmagno et al. (2004, 2006), studying ordinary chondrite falls from the Vatican Observatory collection, found that when this property is combined with grain density in the construction of a grain density/ magnetic susceptibility plot (Figure 19), the three groups form three distinct regions. Thus, the combination of grain density and magnetic susceptibility provides an even more powerful tool for classifying ordinary chondrite falls. Because the techniques for measuring these properties are fast and non-destructive, they can be applied to large portions of a collection and may help to identify mislabeled or misclassified meteorites. Visible in Figure 19 are a couple of these so-called ―ringers‖ that plot in the wrong group for their type.

52

The situation is not as straightforward for finds, as Consolmagno et al. (2006) note.

Weathering has a significant effect on metal grains within ordinary chondrites. They oxidize, forming low-density minerals such as goethite. These weathering products fill available pore space. Once pore space is filled, however, the rate of weathering is significantly reduced because the paths to the interior become blocked (Bland et al., 1996, 1998b). The conversion of metallic iron into lower-density weathering product reduces grain density, though its expansion into pre-existing pore space means bulk density is not significantly altered (Consolmagno et al.,

1998). In addition, the weathering products are also less magnetic than metallic iron, thus reducing total magnetic susceptibility (Consolmagno et al., 2008). The varying degree to which a stone may be weathered means that grain density and magnetic susceptibility are not useful for classification of ordinary chondrite finds.

Ordinary chondrites have already been studied to some length using the same measurement techniques described here, especially during the development of these same techniques using the meteorite collection at the Vatican Observatory. As such, this chapter is not intended to be an in-depth analysis of the results of the study. Rather, it is for the most part a report of the data. I refer the reader to Consolmagno et al. (1998, 2006, 2008) for further detail and analysis. Nevertheless, some results from the extensive additional data this study has provided are unexpected. So as not to overlook important conclusions, some highlights as well as those unexpected findings will be presented along with the results.

That being said, this chapter does include one unique analytical approach in the area of weathering for ordinary chondrite finds. Currently, weathering classification schemes are non- standardized and rely to large degree on the personal judgment of the investigator. In addition, individual stones from the same meteorite may weather to differing degrees based on differences

53 in the stone size, the local environment, and even differences in curation practices at, for example, two different collections. Because of this, weathering grades reported in the literature are not necessarily useful. Since grain density and magnetic susceptibility both vary according to the degree of weathering, I present here a ―weathering modulus‖ based on the total extent to which these two properties differ from each population’s mean for unweathered falls. This modulus will then serve to demonstrate how the other two properties (bulk density and porosity) relate to degree of weathering.

3.2 H Chondrite Falls

In this study are included 207 stones from 116 H chondrite falls. The data are presented in Table 4. A plot of grain density and magnetic susceptibility showing all H, L and LL falls can be found in Figure 20. Grain densities for H falls average (by stone) 3.71 g cm-3, with a range from 3.18 g cm-3 to 4.14 g cm-3. Bulk densities averaged 3.35 g cm-3, ranging from 2.51 to 3.77 g cm-3. This yields an average porosity of 9.5%, with a range from 0 to 26.6%. Magnetic susceptibilities ranged from log χ = 4.57 to 5.64, with an average value in log space of log χ =

5.30.

Consolmagno et al. (1998) report a general dependence of porosity on shock, with the average porosity decreasing as shock increases due to compaction of pore space. The data reported here are consistent with that conclusion, with average porosity 9.3% for shock stage S1 and 12.0% for S2, dropping to 5.5% for S4, as exhibited in Figure 21. This analysis is based on whole-meteorite porosity values, not individual stones, since some meteorites have a large

54 number of stones in the database and thus might bias the results. Each meteorite’s porosity is determined by the mass-weighted average grain and bulk density for all stones from that meteorite. Unfortunately, this also reduces the number of data points available, yielding no more than four porosity measurements for all categories except S3, for which there are 20. The low statistics are partly responsible for the fact that porosity for S1 is less than S2; one meteorite with

2.8% porosity has thrown off the average. Omitting it yields an average porosity of 11.5%, similar to that for S2.

Consolmagno et al (1998) also report no observed trends in porosity with regard to petrographic type, a finding that is reiterated in Consolmagno et al. (2008). However, with the increased quantity of data now available for study, that finding may be questioned. Omitting type 3 meteorites, for which there are only two data points, I find no significant change in average porosity from petrographic type 4 to 5, which both reside near 10%, but for petrographic type 6 the average porosity drops to 7.8% (Figure 22). This drop corresponds to an increase in average bulk density from 3.33 to 3.42 g cm-3. The reduction in porosity for more thermally processed ordinary chondrites is understandable, since the increased equilibration between chondrules and matrix may result in the elimination of some of the interstitial spaces between them.

Curiously, there may also be a correlation between petrographic type and magnetic susceptibility, as the average magnetic susceptibility increases from log χ = 5.12 for petrographic type 3 to 5.34 for type 6 (Figure 23). Most of the difference for that property lies between type 3 and 4, but there is a steady rise from type 4 to 6. This may be indicative of a small increase in the metallic iron quantity within more thermally processed meteorites. Since relatively minor thermal processing should not result in migration of metals out of the meteorite, this may be

55 indicative of minor differences within parent bodies during the time of formation that may be consequent upon local conditions within the solar nebula where they formed.

3.3 H Chondrite Finds

For H finds, the database includes 79 stones from 63 meteorites. They are reported in

Table 5. The average bulk density was slightly higher than for falls at 3.43 g cm-3, but with a range from 2.86 to 4.21 g cm-3. It should be noted that the one stone measuring 4.21 g cm-3 is a piece of Acme with a mass of only 5.6 g, rendering the bulk density measurement unreliable.

Omitting that one, the next highest bulk density was 3.69 g cm-3. Grain densities were in general much lower than for falls, with an average value of 3.51 g cm-3 and a range from 3.19 to 3.79 g cm-3. Magnetic susceptibilities were also much lower than for falls. The average log χ was

5.05, with the population ranging from 4.41 to 5.61. Porosities were also significantly affected by weathering, as expected. The average porosity dropped to 2.8%, with many stones near zero and the highest value only 10.2%. The 5.6 g stone of Acme was omitted from this calculation, since its anomalously high bulk density measurement resulted in a negative porosity of -25%.

This serves as confirmation that its bulk density value is incorrect.

3.3.1 Model Porosities of H Finds

There is an additional column included in Table 5 labeled ―model porosity.‖ Because bulk density is not expected to vary considerably during weathering, it is still possible to estimate

56 the original porosity of the stone if the original grain density can be determined. This was determined by plotting grain density against bulk density for H falls and making a linear fit

(Figure 24). The fit, which is good to within about 3% for most stones in the population, is as follows:

ρρgrain0.417 bulk 3.215. (27)

Model porosities for the finds overlap actual porosities of the falls reasonably well

(Figure 25). Using the model porosities, it is possible to add further data to seek trends of porosity based on shock, petrographic type, etc. I have done so for petrographic type in Figure

26. The additional data continues to support the observed drop in average porosity for petrographic type 6, and using median values there is now visible a slight downward trend in porosities over all petrographic types.

3.3.2 Weathering Modulus for H Finds

In order to quantify degree of weathering of finds, we have developed a weathering modulus based on a distance in grain density / magnetic susceptibility space from the mean values for falls of a given population (Macke et al., 2010b). This is not, strictly speaking, a physical quantity, as it does not correlate directly to physical relationships, is unit-dependent, and is indeterminate in its own units (though not technically unitless). Nevertheless, it remains a

57 useful tool for establishing degree of weathering, from which weathering-related trends can be extracted from the data.

For the case of H finds, the weathering modulus (WH) follows the expression:

2 2 WH5 (logχ )00 log χ 1.2( ρ g ρ g ) (28)

where (log χ)0 and ρg0 represent the mean fall values for magnetic susceptibility (in log units of

10-9 m3 kg-1) and grain density (in g cm-3), respectively. The factor of 1.2 in the grain density term is there to better fit the ellipse occupied by the H fall population in grain density-magnetic susceptibility space. The factor of 5 is an arbitrary scaling factor, giving the weathering modulus a range from zero up to just over 5. Curves for unit values of the modulus have been plotted on

Figure 27.

Curiously, a plot of porosity as a function of weathering modulus (Figure 28) has little to no slope; that is, while all H finds have significantly reduced porosities as compared to H falls, their porosity does not vary with degree of weathering for modulus values over 1. Below 1, the degree of weathering is minimal but not dismissable. There are only a few stones with porosities greater than 8%, and two of these have modulus values below 1. However, there were no stones at all (below or above 1) with porosities greater than 10.3%.

There was no a priori expectation that bulk density should vary with degree of weathering, as the weathering products should fill existing volume without substantial alteration of mass. However (Figure 29), bulk density varies inversely with weathering modulus. This trend becomes more understandable if, instead of bulk density, model porosity is plotted as a

58 function of weathering modulus (Figure 30). While actual porosities do not vary noticeably with weathering modulus, the model porosities vary directly with it.

All of this is consistent with the chondrite weathering model of Bland et al. (1996,

1998b). That model includes two stages of weathering. The first stage operates on relatively short time scales3, during which oxidation reactions in metal cause low-density weathering products to fill existing pore spaces. Once porosity is reduced to near zero, this process comes to a halt primarily because weathering agents no longer have access to the interior of the specimen.

The second stage of weathering, during which the sample slowly fragments and crumbles, operates on much longer time scales that may be thousands of years. In general, any find will have completed stage one of weathering by the time it is collected. So, all stones will have had porosities reduced to near zero. (Notably, it appears that a stone can still have up to 8% porosity when stage one weathering ceases.) Those stones that originated with higher porosities and thus had higher model porosities would have more interior space into which weathering products might expand. In effect, they were less protected from weathering. Therefore, it stands to reason that they would be more strongly weathered, with a greater reduction of unweathered metal than stones with lower initial porosities.

3 As a side project, I have been studying this effect on , an L chondrite that fell in a suburb of Chicago in 2003. Some of the samples used in the study had been collected in 2010, after being exposed to the elements in northern for seven years. After such a short time, they already exhibit strongly reduced grain density and magnetic susceptibility. The study is at this time incomplete, and so is not included in this dissertation. 59

3.4 L Chondrite Falls

The database includes 216 stones from 122 L chondrite falls (Table 6). Average porosity for the group was slightly lower than for H chondrites at 8.0%, ranging from 0 to 16.3%. Grain density ranged from 3.39 g cm-3 to 3.90 g cm-3, with an average of 3.58 g cm-3, while bulk density averaged 3.30 g cm-3 and ranged from 2.98 to 3.86 g cm-3. Magnetic susceptibility averaged log χ = 4.87 and ranged from 3.81 to 5.47. However, the ranges include a few

―ringers‖ that may actually be H chondrite stones (for example, a 52 g stone of L’Aigle with magnetic susceptibility 5.23 and grain density 3.78 g cm-3) or LL chondrites (for instance, a stone labeled Ojuelos Altos with a magnetic susceptibility of 3.81).

Unlike with H chondrite falls, no discernible trend related to petrographic type is visible in any of the physical properties measured for L falls (see Figure 31 for porosity). However, if anything the inverse correlation between shock state and porosity is even more pronounced for L chondrites (Figure 32), dropping from a mean value of 11.9% for S2 to 3.4% for S6.

3.5 L Chondrite Finds

107 stones from 67 L finds were measured (Table 7). As with H finds, porosity is significantly reduced as compared to falls. L find porosities averaged 3.6%, ranging from zero to

12.1%. Bulk densities were 2.94 to 3.56 g cm-3, with an average of 3.34 g cm-3, while grain densities averaged 3.46 g cm-3 and ranged from 3.27 to 3.79 g cm-3. The average magnetic susceptibility was 4.62 in log units, with a range from 3.90 to 4.62.

60

3.5.1 Model Porosities of L Finds

Once again, a fit was made between L fall grain and bulk densities, yielding the relationship:

ρρgrain0.009 bulk 3.613. (29)

For most L’s this will result in a grain density within 0.01 g cm-3 of 3.58 g cm-3. Most actual grain densities for falls came to within 2% of their model values. Average model porosity for finds was 6.9% and ranged from zero to 17.9%, with good overlap with fall actual porosities.

(Figure 33).

Observing the relationship of model porosity and petrographic type (Figure 34), there is no obvious difference between types 4, 5, and 6. However, for type 3 stones the porosity is observably lower. The database contains ten L3 meteorites, so the result cannot be dismissed due to low statistics; its average is more than one standard deviation of the mean from the others.

3.5.2 Weathering Modulus for L Finds

As with that for H finds, a weathering modulus was determined for L finds (Figure 35).

In this case, the expression took the form:

2 2 WL5 (logχ )00 log χ 1.6( ρ g ρ g ) (30)

61

and the values for the centroids were also adjusted to match the averages for L falls. The same patterns described above for H finds are also seen with L finds: porosity does not exhibit any dependence on weathering (Figure 36), but bulk density is inversely correlated (Figure 37) and model porosity is positively correlated (Figure 38) with the weathering modulus.

3.6 LL Chondrite Falls

LL chondrites are much less abundant than H and L, and this is reflected in the fact that the database of measurements includes only 51 stones from 33 LL falls (Table 8). Bulk densities averaged 3.18 g cm-3, with a range from 2.80 to 3.51 g cm-3. Grain densities were tighter, ranging from 3.41 5o 3.63 g cm-3, with an average value of 3.52 g cm-3. This resulted in a porosity range from zero to 19.4%, averaging 9.5%. Magnetic susceptibilities were as low as

3.33, but in an extreme case (probably also a ringer) went as high as 5.15, and the average was

4.13. This particular group has greater variability in grain density and magnetic susceptibility than H and L, and the region it occupies in grain density/ magnetic susceptibility space (Figure

20) is less well defined or constrained than is that for the other two groups.

Looking at the data by petrographic type, no obvious trends are visible in porosity

(Figure 39). However, LL falls do exhibit a statistically significant negative trend in magnetic susceptibility with petrographic type (Figure 40), from an average log χ = 4.41 for LL3 to 3.93 for LL6 (and the one LL7 meteorite in the data has log χ = 3.33). Organized by shock state, no trends are apparent, though the relatively low number of LL meteorites with published shock

62 values (only one S1, and no S4 or S5s are in the database) has proven a hindrance for meaningful analysis of the data.

3.7 LL Chondrite Finds

The database contains 12 stones from 10 LL finds (Table 9). Average bulk density was

3.22 g cm-3, with a range from 2.93 to 3.41 g cm-3. Grain density ranged from 3.27 g cm-3 to

3.50 g cm-3, with a mean of 3.42 g cm-3. LL finds have an average porosity of 5.8%, with most stones below 12% porous, though Zerga has a porosity as high as 14.2%. The mean magnetic susceptibility was log χ = 4.05, with a high end of 4.65 and a low value of 3.66, excepting Zerga with an anomalously low magnetic susceptibility of log χ = 2.86.

3.7.1 Model Porosities of LL Finds

Based on data from LL falls, model grain densities follow the relation:

ρρgrain0.098 bulk 3.203. (31)

This yields grain densities that lie within 2% of the actual values for LL falls. From this, model porosities were calculated (Figure 41). They average 8.4%, with a range from 3.6% to 16.0%

(including Zerga), though excluding Zerga the model porosities remained below 12%.

63

As can be seen in Figure 42, the addition of model porosities does not help extract any petrographic-type based trends in the data. Due to the lack of LL finds with reported shock values, the inclusion of model porosities does not significantly alter that relationship either.

3.7.2 Weathering Modulus for LL Finds

Because LL finds occupy a much larger range, especially in magnetic susceptibility, than

H and L, and because the total number of falls is lower for LLs than for the other two populations, the region in grain density/magnetic susceptibility space occupied by LL falls is not densely populated enough to clearly define its shape. It also raises some doubts as to the usefulness of a weathering modulus when a relatively unweathered stone could in principle receive a large weathering value due to the fact that it lies far from the centroid of the region; or conversely a weathered stone could be reported as minimally weathered because it remains consistent with the main population of falls. Nevertheless, I attempted to determine a weathering modulus for LL finds. The modulus follows this form:

2 2 WLL1.5 (logχ )00 log χ 10( ρ g ρ g ) . (32)

Note that the ellipsoids of constant WLL are stretched in the magnetic susceptibility dimension and slightly compressed in the grain density dimension as compared to H and L weathering moduli (Figure 43).

64

As can be seen from Figure 44, Figure 45 and Figure 46, the lack of large numbers of LL finds in the data make extracting weathering trends a challenge. However, the data are not inconsistent with the same trends observed for H and L finds: No substantial correlation of porosity with weathering modulus, but an inverse correlation of bulk density and a positive correlation of model porosity with weathering modulus.

3.8 Intermediate OC Types

In addition to H, L and LL, some ordinary chondrites have been classified as intermediate

H/L or L/LL. For completeness, their data, covering 53 stones from 10 meteorites, have been included in Table 10. Of these, all but two stones from two meteorites are falls. For the two finds, model porosities have been included based on average grain densities for each category.

No attempt to establish a weathering modulus for these intermediate types was made.

3.9 Porosities of Breccias and Non-Breccias

The transformation of aggregates in parent bodies into the rocks comprising meteorites is still one of the unanswered questions in the formation of planetary bodies. It is thought that the lithification of ordinary chondrites may have proceeded by shock compression (Weidenschelling and Cuzzi, 2006) In order to better understand the relationship between shock and lithification, it is important to compore those meteorites that are known to have been lithified by shock

65 melting (i.e. breccias) with unbrecciated meteorites. If the process by which chondrites originally lithified is similar to the process by which breccias lithify, the two groups should exhibit similarities in porosity. The massive quantity of data available for ordinary chondrites enables a good analysis of brecciated and non-brecciated populations. For this analysis, breccias were identified based on the listing in the Catalogue of Meteorites (Grady, 2000). Grady (2000) also identifies a number of veined ordinary chondrites (most non-brecciated, but some also breccias), indicative of additional processing. To avoid further complication due to terrestrial weathering effects, only OC falls were considered in this analysis.

The meteorites were divided into brecciated and non-brecciated meteorites, though veined meteorites were excluded from the analysis. All of the analysis was conducted on whole- meteorite values (not individual stones). There were 54 breccias identified, 152 non-breccias and 55 veined meteorites. At first glance, porosities of the breccias and non-breccias were quite similar (Table 11). Both groups ranged from basically zero porosity to 26.6% porous. Breccias averaged 8.5% ± 0.6%, while unbrecciated OCs averaged 9.3% ± 0.4%, with error bars based on one standard deviation of the mean. Given that the error bars of the two populations overlap, this would seem to indicate that brecciated and unbrecciated OCs share similar porosities. The veined population had a slightly lower average porosity of 8.2% ± 0.5%, marking it lower on average than other non-breccias to a significance exceeding one sigma.

The similarity on average between breccias and non-breccias does not mean an overall similarity between the two populations. For one thing, among the 54 breccias only one (Sena) has a porosity exceeding 15%, while the non-breccias have eighteen out of 152 above 15% porous. Using Poisson statistics for the uncertainties, this means that the percentage of high- porosity breccias is 2% ± 2%, while for non-breccias it is significantly higher at 12% ± 3%.

66

With such large numbers of meteorites represented, it is possible to analyze the distribution of porosities over the whole population, as expressed in histogram form in Figure 47. While the non-breccias approximate a normal population distribution centered in the 7% - 9% porosity bin, the breccias exhibit increasing population numbers ranging from 0 to about 12%, then a sharp drop-off for porosities above that.

From the population distribution, it is clear that the breccias experienced different processes, likely stronger shock that compressed the higher-porosity materials. This indicates that the process by which the aggregate precursor material for ordinary chondrites lithified is different, either in kind or in magnitude, from the process by which breccias lithify.

3.10 Outliers among OC Falls

In the course of this study, a total of eighteen stones with labels attributing them to ordinary chondrite falls had grain densities and magnetic susceptibilities that were inconsistent with those of the main population for their group. These stones were identified as being more than three standard deviations from the centroid of their respective groups. They are plotted as orange symbols in Figure 20. Many of them reside squarely within the population of another group; for instance, stones labeled ―H‖ that are consistent with L falls. The outliers will remain unidentified until the curators have had a chance to examine the stones. (Curators have been notified of the outliers.) Some of them may be mislabeled or misidentified stones. It is also conceivable, though unlikely in most cases, that some may be stones of a different chondrite type

67 that were included in the original fall. Five of the stones are from the Vatican collection, and have been discussed elsewhere (Consolmagno et al., 2004, 2006)

Ten of the eighteen stones are labeled ―H‖, and of these five are consistent with L, with another three depleted in grain density or magnetic susceptibility. It is quite possible that some or all of these outliers may be the result of a measurable degree of weathering of the high-iron ordinary chondrites, despite their status as falls. Two other stones (Le Pressoir and Phû Hong, both from the Vatican collection) have very high grain densities and above average magnetic susceptibilities. Both of these stones are small (one 13.3 g, the other 4.0 g), so a likely explanation here is that they contain large metal grains that bias the results.

In a few cases, weathering can be clearly ruled out. For example, one stone of St.

Mesmin (LL) is enhanced in magnetic susceptibility. It rests within the L population but would not be inconsistent with an H/L either. One stone of L’Aigle (L) resides within the H region.

Discerning the source of such discrepancies may prove an impossible task. Potential causes include curatorial errors (a problem especially for historical collections), careless or unscrupulous collectors and dealers, the presence of pre-existing meteorites in the area of a fall, or even heterogeneous inclusions within the fall itself. The case of the L’Aigle stone is intriguing (Consolmagno et al., 2006), as it is accompanied by original documentation. This would rule out curatorial error, but does not rule out the possibility of an unscrupulous dealer who may have attached the name of this historically significant meteorite to a stone of lesser value.

In general, these outliers are limited to just one stone per meteorite. Two exceptions include L’Aigle and (H), with two outliers apiece. Curiously, Pultusk’s two outliers were located in two different meteorite collections. Both are approximately consistent with the L

68 falls, but further examination is necessary in order to determine whether these stones are L or are merely weathered.

3.11 Other Chondrite Types

The two other major classes of chondrite types, carbonaceous chondrites and enstatite chondrites, will be dealt with in subsequent chapters. However, there are a few sui generis chondrite types that do not fit into any of the three major classes. Of these, the database includes

9 stones of 6 Rumuruti-like (R) chondrites, and one stone from Kakangari, belonging to the

Kakangari-like (K) chondrite group. The data are presented in Table 12.

R chondrites, of which there 106 known meteorites (76 non-Antarctic) and just one known fall (Rumuruti), are chemically similar to H chondrites though with higher Olivine content (Hutchison, 2004), but have lower chondrite/matrix ratios (Kallemeyn et al., 1996), are more strongly oxidized and have unusual oxygen isotopic ratios, plotting below the terrestrial fractionation line.

All six of the R chondrites in this study are finds. They have an average grain density of

3.52 g cm-3 (range 3.45 g cm-3 to 3.59 g cm-3), average bulk density of 3.14 g cm-3 (range 2.79 g cm-3 to 3.32 g cm-3), porosity of 10.9% (range 5.6% to 19.2%), and magnetic susceptibility of log

χ = 3.11 (range 2.72 to 3.63). Some of the low density and magnetic susceptibility values (as compared to H chondrites, see Figure 48) may be accounted for by weathering, but not all. The entire magnetic susceptibility range is substantially less than that of even the lowest H find (as well as L finds). Even though total iron content of R chondrites is comparable to H chondrites,

69

0 the former are much more oxidized (prior to weathering), with a negligible Fe /Fetot ratio

(compared to a ratio of 0.6 for H chondrites; Hutchison, 2004). The oxidation and lack of metallic iron accounts for the very low magnetic susceptibilities.

None of the above discussion pertains to porosities. It is likely that R porosities are driven down artificially by weathering, and that fresh R falls would have even higher porosities than those reported here.

The K chondrite grouplet contains only three members (Kakangari, Lea County 002, and

Lewis Cliff 87232) of which Kakangari is the only fall. These meteorites do not fit into any of the three major classes (ordinary, carbonaceous, or enstatite), but have traits of each. They have high metal content (comparable to H chondrites), very high percentage matrix comparable to carbonaceous chondrites, but are enriched in enstatite. Their oxygen isotopic ratios lie below the terrestrial fractionation line. They have low oxidation states, though higher than R chondrites

(Weisberg et al., 1996).

In the database of this study, just one stone from Kakangari has been measured for physical properties. Its grain density (3.45 g cm-3 ± 0.01 g cm-3) and magnetic susceptibility

(5.09 ± 0.10) are most comparable to L chondrites, and are much higher than the R group due to lower degree of oxidation. The bulk density of 3.04 g cm-3 ± 0.05 g cm-3 gives it a porosity of

12.0% ± 1.5%.

70

CHAPTER 4: CARBONACEOUS CHONDRITES

4.1 Introduction

Carbonaceous chondrites are among the most primitive meteorites. They are also considerably less abundant than ordinary chondrites; there are 413 known non-Antarctic carbonaceous chondrites, of which there are 43 falls. Based on spectrographic studies, the best match for asteroidal analogs for carbonaceous chondrites are C-type asteroids (and perhaps some originate from X-type or K-type; Consolmagno et al., 2008; Burbine et al., 2002), which are low- albedo, colorless asteroids (Chapman et al., 1975) that generally reside in the outer portion of the asteroid belt, though it is unclear whether the asteroids originated in that part of the belt or migrated there over time. Regardless, carbonaceous chondrites probably originated farther out in the solar system than did ordinary chondrites.

Like ordinary chondrites, their elemental compositions (except for volatiles) closely resemble those of the Sun. The CI group of carbonaceous chondrites possesses the closest match of all meteorite types to Solar composition (Hutchison, 2004). Unlike ordinary chondrites, carbonaceous chondrites exhibit less thermal alteration overall, and many are aqueously altered.

In the petrographic-type scheme of van Schmus and Wood (1967), carbonaceous chondrites range from strongly aqueously altered (petrographic type 1) to moderately thermally altered

71

(petrographic type 4). There are a few minor exceptions; for example NWA 2388 is classified as a CK6 (Meteoritical Bulletin Database).

From preliminary studies of 52 carbonaceous chondrites (of which 17 had porosities),

Britt and Consolmagno (2003) observed a marked difference in grain and bulk densities between the aqueously-altered (hydrated) groups CM and CI, and the anhydrous groups CK, CO and CV.

Consolmagno et al. (2008) remark that the hydrated groups had low grain densities (below 3 g cm-3) as compared to typical grain densities above 3 g cm-3 for the anhydrous groups. Likewise, they observed hydrated CCs to have low bulk densities and (on average) higher porosities than anhydrous CCs. They did not carry this analysis further, however, by comparing porosity or other properties with petrographic type, presumably due to insufficient data.

Corrigan et al. (1997) also measured porosities of 26 carbonaceous chondrites using a combination of techniques. They measured matrix porosities using SEM point-counting techniques on thin sections, which applied to most of the meteorites in their study, while bulk porosity of larger (6.2 cm long) core samples were measured using a sonic profiling technique.

The latter, which are better for comparison purposes with whole-stone porosities described in this chapter, were performed on only nine meteorites.

Rochette et al (2008) measured magnetic properties for 594 stones of 222 carbonaceous chondrites, including numerous Antarctic meteorites. It should be noted that their database includes a number of meteorites for which the measured mass is significantly less than 10 g, and measurements may be biased at lower masses by minor variation in metal abundance (inclusion or exclusion of metal grains, etc.). Nevertheless, they observed that the different types of carbonaceous chondrites do exhibit different magnetic susceptibilities. These differences, however, are not sufficient for use as a classification tool. Rochette et al. (2008) also attempted

72 to couple their data with grain density from literature values, though their technique lacked density and magnetic susceptibility measurements conducted in a consistent manner and applied to the same stones.

This study includes data for 196 stones from 63 carbonaceous chondrites, a mild improvement over the 52 meteorites discussed in earlier works. Of those 196 stones, however, a full suite of measurements has been produced for 167 stones covering 60 of the 63 carbonaceous chondrites in the study, as opposed to only 17 meteorites before. This unprecedented database will permit testing of earlier observed trends. It also will permit the exploration of trends covering the entire population of carbonaceous chondrites (see Secton 4.3).

4.2 Data

The data for the 196 carbonaceous chondrites in the study are listed in Table 13 and are represented graphically in Figure 49 through Figure 52. The discussion of the data will be grouped under four subheadings, one for each of the following classes or major groupings of types: the CR clan, aqueously altered carbonaceous chondrites, anhydrous carbonaceous chondrites, and ungrouped carbonaceous chondrites.

4.2.1 The CR Clan: CR, CB and CH

CR, CB and CH carbonaceous chondrites are metal-rich and compositionally distinct from other carbonaceous chondrites, are believed to form in the same part of the solar nebula 73 under similar conditions as each other, and so were grouped together into a clan by Weisberg et al. (1995). Despite differences from other CCs, these meteorites are still classified carbonaceous due to refractory lithophile abundances. Carbonaceous chondrites have refractory lithophile/Mg abundance ratios (normalized to CI) greater than or (in the case of CI) equal to 1, while that for ordinary and enstatite chondrites is less than 1. Because CR-clan refractory lithophile/Mg ratios are near to or exceed 1, they are considered carbonaceous chondrites. In addition CR-clan chondrites have oxygen isotopic abundances that, like carbonaceous chondrites, lie below the terrestrial fractionation line (Krot et al., 2002).

4.2.1.1 CR

Renazzo-like carbonaceous chondrites (CR) tend to be moderately aqueously altered, and almost all are petrographic type 2. They were first recognized as a group by McSween (1977).

Of the three groups in the CR clan, CRs have the lowest total metal abundance at a maximum of about 8 vol% (Weisberg et al., 1993), with most of the metal residing in chondrules. Based on near-solar refractory lithophile abundances, they are thought to be very primitive, though they are depleted in volatiles (Weisberg et al., 2006). CRs are all breccias (Hutchison, 2004). There are 121 known CRs, of which 38 are non-Antarctic and only three (Al Rais, Kaidun, and

Renazzo) are observed falls.

In this study are included data for nine stones from seven CR meteorites, including one stone of Al Rais and three of Renazzo. Grain density averaged 3.42 g cm-3 (range 3.06 g cm-3 to

3.88 g cm-3), which coupled with an average bulk density of 3.11 g cm-3 (range 2.29 g cm-3 to

74

3.94 g cm-3) yields an average porosity of 9.5% (range zero to 25.0%), which is below typical porosities for most carbonaceous chondrites. Magnetic susceptibilities averaged log χ = 5.02, ranging from 4.64 to 5.34.

Of the CRs, Al Rais has an unusually low density (both grain and bulk) and an unusually high porosity. Omitting this meteorite, the lowest bulk density of the remaining CRs is 2.89 g cm-3, the lowest grain density is 3.30 g cm-3, and the highest porosity is 18.2%. This anomaly does not extend to magnetic susceptibility, however; Al Rais has a magnetic susceptibility of

4.89, which is below average but not outside the range of the remaining population.

In terms of grain density, CR falls are unusual in that they exhibit lower values than CR finds (Figure 53). It is not clear whether there is any porosity difference between falls and finds, though it is noteworthy that the highest-porosity stone (Al Rais) is a fall, while the lowest- porosity belongs to a find (Tafassasset). Since only two meteorite falls are represented in the data, it is unclear whether the observed difference is due to a genuine weathering effect or can be simply accounted for by natural variation among meteorites of this type. The range in grain densities of the CR meteorites and the variability in data from the three stones of Renazzo (with porosities ranging from 3.7% to 18.2%) suggest that their brecciated nature lends them a certain degree of heterogeneity.

4.2.1.2 CB

The CB group was first identified as a distinct group relatively recently by Weisberg et al. (2001). They are characterized by very high (more than 40 vol%) metal abundances occurring

75 in clasts, and CBs may also contain inclusions of material of other chondrite types. CBs are further subdivided into two groups: CBa is somewhat lower in metal, ranging 40-60%, and CBb has metal abundances exceeding 70 vol% (Krot et al., 2002). There are 14 known CB meteorites, of which 8 are non-Antarctic and only Gujba is an observed fall. In our database are only four stones from two meteorites (Bencubbin and Hammadah al Hamra 237), both finds, of which only one stone apiece has a full suite of physical property measurements. Bencubbin is a

CBa, and HaH 237 is a CBb. Petrographic types for CB meteorites are not listed in the

Meteoritical Bulletin Database.

Due to their very high metal content, CBs have the highest grain densities of any chondrites in the study. They averaged 5.65 g cm-3, grouping tightly at 5.63 g cm-3 and 5.66 g cm-3. They also are among the highest in magnetic susceptibility, with an average of log χ = 5.57 and a range from 5.31 to 5.79. The average is slightly lower than but statistically consistent with the mean value of 5.65 ± 0.04 reported by Rochette et al. (2008). On a grain density/magnetic susceptibility plot (Figure 53), CBs reside well above the regions occupied by other chondrites.

Bulk densities are also high (average 5.25 g cm-3, range 4.90 g cm-3 to 5.55 g cm-3), with a low average porosity of 3.9% (from two porosity measurements of 2.0% and 5.8%). No obvious difference between the CBa and the CBb stones is apparent in any of the physical properties measured.

76

4.2.1.3 CH

Like CB, CH carbonaceous chondrites are also very high in metal compared to other carbonaceous chondrites, though to a lesser extent than CB. While CB may be 60 vol% or more metal, CH is about 20 vol% metal (Weisberg et al., 1988). This metal abundance gives the group its designation (―H‖ for high metal, ―C‖ for carbonaceous chondrite; Krot et al., 2002). CH chondrites exhibit low alteration states, and all CHs listed in the Meteoritical Bulletin Database are petrographic type 3. There are 22 known CH meteorites (12 non-Antarctic and no falls).

Our database includes just two stones from one CH find (Acfer 214), though both stones have been measured for all four physical properties.

Acfer 214’s grain density (3.65 ± 0.02 g cm-3) is fully 2 g cm-3 less than the CB group, primarily the result of having less than half the total metal content, though it is within the overall range of the CR group. Magnetic susceptibility (log χ = 5.30 ± 0.12) is also less than CB and a little higher than the CR average. This is compatible with the value of 5.36 ± 0.10 reported by

Rochette et al. (2008). Grain density and magnetic susceptibility values are also consistent with

H ordinary chondrites and overlap those of CRs (Figure 53), demonstrating that while grain density and magnetic susceptibility can distinguish ordinary chondrite types from each other

(Consolmagno et al., 2006 and Chapter 3 of this dissertation), they are not equally useful for distinguishing carbonaceous chondrites from ordinary chondrites. The measure of bulk density for the stones may be a bit off. Acfer 214 averages 3.77 ± 0.08 g cm-3, yielding a negative porosity. However, it should be noted that porosity values of both stones, while negative, are within measurement uncertainty of zero. From this it should be inferred that Acfer 214 has effectively zero porosity. Whether this is indicative of CH as a group is doubtful; Acfer 214 is a

77 find, and a meteorite with as much metal as it possesses would likely experience weathering much like ordinary chondrites (Bland et al., 1998b) that would reduce porosity to near zero.

4.2.2 Aqueously Altered Carbonaceous Chondrites: CI and CM

In addition to the CR group, two other types of carbonaceous chondrite exhibit strong evidence of aqueous alteration, including secondary aqueous phases such as phyllosilicates

(Weisberg et al., 2006; Brearley, 2006). These are CI’s, all of which are petrographic type 1, and

CM’s, which are mostly petrographic type 2.

4.2.2.1 CI

Ivuna-type carbonaceous chondrites (CI) are highly porous and extremely friable. They are also quite rare (only eight meteorites of the type are known), and most collections only possess a modest mass of the material. Two of the key falls in this category, Alais and , both fell in France and hence their main masses are at the Muséum National d’Histoire Naturelle in Paris. A third key fall, Ivuna, fell in Tanzania and its main mass is held by the Tanzania

Geological Survey (Grady, 2000). Of the remaining meteorites, only Yamato 980115 is larger than 15 grams.

CI carbonaceous chondrites are chemically among the most primitive of meteorite types.

Elemental abundance ratios of CIs relative to Si are very near abundance ratios of the solar

78 photosphere, except in the case of the most volatile species (Anders and Grevesse, 1989). While they have retained the elemental signatures of the solar nebula quite well, this does not mean that they are unprocessed. CIs show signs of extreme aqueous alteration, with all known CIs being petrographic type 1. They also have no chondrules (with the possible exception of Yamato

82162; Hutchison, 2004) though whether that is original or the result of aqueous processing is unclear (Weisberg et al., 2006). Their matrix is composed largely of phyllosilicates and other secondary products resulting from aqueous alteration (Brearley, 2006).

Due to their extreme friability, even the relatively benign methods employed in this study present a possible risk of damage to the samples. Hence, curators have been wary of permitting their CI specimens to be subject to physical property measurements. The two CI samples in the database (both from Orgueil) are from the Vatican collection, and density and porosity measurements were conducted on just one of them by the curator of the collection, Guy

Consolmagno, using techniques compatible with those presented here. Magnetic susceptibility measurements on both samples were conducted by Jérôme Gattacceca and Pierre Rochette using a KLY2 magnetic susceptibility meter. The data were published originally in Britt and

Consolmagno (1996) and Rochette et al. (2008).

The grain density of Orgueil is 2.42 g cm-3, making it one of the least dense chondrites in this study. Together with a bulk density of 1.57 g cm-3, this gives it a porosity of 34.9%. The two samples in the study had an average magnetic susceptibility of log χ = 4.49, with the two measurements being 4.11 and 4.86. On a plot of grain density and magnetic susceptibility

(Figure 54), it resides well below the rest of the chondrites, but not to the lower left (low in both grain density and magnetic susceptibility) as basalts tend to do. There is sufficient metallic iron

79 in CI chondrites to give them magnetic susceptibilities comparable to L falls despite their low densities.

The porosity result is higher than some of the results based on liquid immersion techniques, which grossly underestimated the porosity (Consolmagno et al., 2008) by producing bulk density measurements much higher than their actual values. Also, Corrigan et al. (1997) measured porosity for Orgueil matrix of only 4% using SEM point-counting techniques on a thin section of matrix material, with similar results for matrix from Alais and Ivuna. It is likely that the production of thin sections either compressed the sample (destroying pore space) or, akin to observations of Strait and Consolmagno (2010), pore space appeared as voids that were large enough not to be counted.

4.2.2.2 CM

Mighei-type (CM) carbonaceous chondrites are also aqueously altered, though to a lesser degree than the CI’s. Most CMs are petrographic type 2, though a few are type 1. Unlike CIs, they posses chondrules (about 20 vol%), calcium-aluminum-rich inclusions (CAIs, about 5 vol%), and a small amount of metal. Nevertheless, their matrix abundance of about 70 vol% is second only to CIs (Brearley and Jones, 1998).

420 known CM meteorites are listed in the Meteoritical Bulletin Database, of which 36 are non-Antarctic and 15 are falls. Our database includes physical properties for 43 stones from

13 meteorites, including 14 stones from Murchison and 12 from Murray. All of the meteorites measured were falls except for one stone each of Cimarron and El-Quss Abu Said. The only

80 falls not represented were Boroskino, Erakot, Haripura, Maribo, and Sayama. The main mass of

Boroskino is at the Academy of Sciences in Moscow, with no more than 3.8 g available at any other institution. With the exception of a few grams, Erakot and Haripura are both at the

Geological Survey of India in Calcutta (Grady, 2000), and Sayama is privately held in Tokyo

(Meteoritical Bulletin Database). All of the meteorites in the study were also petrographic type 2.

Due to the large contributions from just two meteorites, all averages reported here are calculated by meteorite rather than by stone to avoid statistical bias.

Bulk densities are rather low, though not as low as Orgueil. They average 2.20 g cm-3, with a range from 1.88 g cm-3 to 2.47 g cm-3. Grain densities average 2.92 g cm-3, with a low of

2.74 g cm-3 and a maximum value of 3.26 g cm-3. CMs are quite porous, averaging 24.7% and ranging from 15.0% to 36.7%. Of the CMs, only Murchison was measured for bulk porosity by

Corrigan et al. (1997), and their result of 23% compares favorably to the average 22.1% porosity

(with a standard deviation among individual measurements of 2.2%) for the 14 stones measured in this study. Their point-counting-based matrix porosities for the remaining CMs averaged a low

6%.

Magnetic susceptibilities on average were log χ = 3.93, which is below that of Orgueil, but covered a wide range from 3.30 to 4.77. Curiously, the two finds (Cimarron and El-Quss

Abu Said) exhibit higher-than-average magnetic susceptibilities of 4.48 and 4.32, respectively, and exhibit no obvious weathering-related effects in any of the other physical properties that were measured. Individual measurements were almost all within uncertainties of the results of

Rochette et al. (2008), and their mean value for the entire group of log χ = 3.90 (σmean = 0.06) is also in very good agreement with the results posted here.

81

Figure 54 plots grain densities and magnetic susceptibilities for CMs as well as CIs. CM grain densities are higher than CIs, while their magnetic susceptibilities are lower. In terms of magnetic susceptibility, CMs are comparable to LL falls. Also visible on the plot is a cluster of

CM falls that lie between log χ = 3.3 and 4.0, and constrained in grain density between 2.7 and

3.0 g cm-3. These 26 stones belong to two falls: Murchison and Murray.

Because of the abundance of stones for Murchison and Murray, this provides an excellent opportunity to explore the homogeneity of stones from the same fall. Unlike extremely heterogeneous falls such as the ureiliite Almahata Sitta (Jenniskens et al., 2009; Bischoff et al.,

2010), Murchison and Murray are relatively uniform in texture, and so by comparing stones from each fall it will be possible to get a sense of the homogeneity of the parent body on scales approximating the size of the original , approximately decimeters to meters. In all parameters and for both meteorites, variation among stones (determined by one standard deviation) was less than 10% from the mean value, with the greatest degree of variability in the porosities. For Murchison, grain density ranged from 2.87 g cm-3 to 3.05 g cm-3 (mean 2.96 g cm-3), with a variability of 0.05 g cm-3, or 1.6% of the mean value. Bulk density averaged 2.31 g cm-3, ranging from 2.15 g cm-3 to 2.40 g cm-3. Variability in bulk density was 0.07 g cm-3, or

3.1% of the mean. Porosity ranged from 18.7% to 24.9%, with a variability of 2.2% (10.0% of the mean 22.1% porosity). Magnetic susceptibility averaged 3.73, with a range from 3.54 to

3.90. Variability was 0.13, or 3.6% of the average. It should be noted that the mean uncertainties for the individual measurements were 0.01 g cm-3 for grain density, 0.02 g cm-3 for bulk density, 0.9% for porosity, and 0.09 for magnetic susceptibility. This indicates that, while measurements did vary between stones, the differences were not many times larger than

82 measurement uncertainty. Overall, the stones from Murchison that were included in the study are homogeneous to within a few percent.

Murray produces similar results. For brevity, only standard deviations will be given.

Bulk density varied 0.05 g cm-3 (2.3% of the mean), grain density varied 0.02 g cm-3 (0.7% of the mean), porosity 1.8% (8.6% of the mean), and magnetic susceptibility 0.15 (4.0%). Mean measurement uncertainties for the stones were the same as for Murchison, but the overall variability is less. This indicates that Murray, at least for the stones in this survey, is more homogeneous overall than Murchison.

4.2.3 The Anhydrous Carbonaceous Chondrites: CO, CK and CV

Anhydrous carbonaceous chondrites may or may not exhibit signs of some aqueous alteration, but tend to have few if any of the secondary hydrous phases, such as phyllosilicates.

They also possess various secondary anhydrous phases that formed in the absence of water

(Brearley, 2006). In terms of the schema of van Schmus and Wood (1967), anhydrous CCs are almost all of petrographic type 3, with the exception of CKs that range into the thermally equilibrated petrographic types 4-6. They tend also to have lower total matrix abundances than the hydrous groups, ranging from 30-50% (Weisberg et al., 2006), although some sources put

CK matrix abundances as high as 75% (Hutchison, 2004; Brearley and Jones, 1998).

83

4.2.3.1 CO

Ornans-type (CO) carbonaceous chondrites are similar to CM in that they are mineralogically similar in anhydrous minerals, have similar chondruls size, similar refractory lithophile element abundances, and similar oxygen isotopic compositions, (Kallemeyn and

Wasson, 1982; Weisberg et al., 2006), though they show no signs of aqueous alteration

(Brearley, 2006) and are all petrographic type 3, though varying from petrographic type 3.0 to as high as 3.8. They have abundant matrix, but at ~34 vol% (Weisberg et al., 2006) the matrix is considerably less abundant than that of CMs. The similarities are enough that CO and CM have been placed together in a clan.

195 CO meteorites are recognized, with 116 non-Antarctic but just 6 falls. Among COs, we have measured 33 stones from 14 meteorites, though there are no more than 4 stones measured for any one meteorite. Of the six CO falls, five are represented, with only Moss

(which fell in 2006) missing from the database. Averaging by stone, bulk density is 3.03 g cm-3, grain density is 3.52 g cm-3, porosity is 13.6%, and magnetic susceptibility is log χ = 4.49.

Averaging by meteorite, the values are 3.06 g cm-3, 3.48 g cm-3, 11.6% and 4.48, respectively.

By stone, bulk density ranged from 2.18 g cm-3 to 3.48 g cm-3, grain density ranged from 2.99 g cm-3 to 3.78 g cm-3, porosity from 0 to 41.3%, and magnetic susceptibility from 3.91 to 5.00.

Corrigan et al. (1997) measured bulk porosities for two CO meteorites: Isna and Lancé.

Their 4% average porosity for Isna is substantially lower than the porosities of any of the three stones in this study, which in total average 14.5%. Their result for Lancé of 8.3% porosity is much closer to the 9.2% average of the four stones in our database, especially when the fact that two of the four stones in the study had porosities below 8.3% is taken into account. It is not clear

84 why there should be such a discrepancy between the results of Corrigan et al. (1997) and our results for Isna, though the authors report an unaccountable inability to produce useful results for

Ornans, another CO3, indicating that perhaps porous CO chondrites may present a unique hazard for sonic profiling techniques.

Of all carbonaceous chondrites, COs exhibit perhaps the greatest disparity between falls and finds. Six of the CO meteorites in this study were falls, while eight were finds, giving a good representation for each group. The difference is most apparent in porosity: falls averaged

19.4% porous (by meteorite), ranging 9.2% to 34.2% while finds averaged 5.7% and ranged from zero to 14.5%. This considerable drop in porosity for finds corresponds to a drop in average grain density (3.64 g cm-3 for falls, 3.35 g cm-3 for finds) and a modest drop in magnetic susceptibility (4.57 for falls, 4.41 for finds). On top of these more predictable effects, there is also a notable increase in average bulk density for finds (3.16 g cm-3) as compared with falls

(2.93 g cm-3). On a grain density-magnetic susceptibility plot (Figure 55), CO falls are somewhat clustered near L and LL falls, while CO are clearly separated from the main group of falls. Unlike OC finds, which are affected primarily by the oxidation of metal during weathering, the primary difference between CO falls and finds is in grain density rather than magnetic susceptibility. If the primary weathering product is carbonates rather than iron oxides, this explains the much smaller change in magnetic susceptibility.

Rochette et al. (2008) agrees well with the magnetic susceptibility data presented here.

They break down their averages for CO into finds and falls. In the case of CO finds, we add data for 023, 078 and 749, and for Rainbow, that are not listed in their data, though of course they add numerous Antarctic finds. They get an average magnetic susceptibility for CO finds of log χ = 4.49 ± 0.06 (σmean), which is a little higher than but within uncertainties of our

85 average find magnetic susceptibility of 4.41 ± 0.09. Likewise, their mean for CO falls is log χ =

4.54 ± 0.08, which is in good agreement with our 4.57 ± 0.08.

Despite other similarities between COs and CMs, physical properties of the two groups differ by a wide margin, both in the structure-related properties and in the mineralogically related properties. Considering falls alone, CMs on average have much higher porosities, but are lower in density (both grain and bulk) and have much lower magnetic susceptibilities. Comparing the two populations in Figure 54 and Figure 55, the difference between CO falls and CM falls is striking, though there is some overlap among finds.

4.2.3.2 CK

Karoonda-type carbonaceous chondrites (CK) are compositionally and texturally similar to CO and CV groups, though more closely related to CV. They are marked by very low carbon content (< 1 mg/g), high oxidization, high abundance of refractory lithophiles, and low abundance of refractory inclusions (Kallemeyn et al., 1991). Despite being strongly oxidized, they exhibit signs of anhydrous thermal metamorphism, making them quite unusual as carbonaceous chondrites go by having petrographic types greater than three, ranging from CK3 to CK6. They have abundant matrix and possess large chondrules (averaging ~600 µm diameter) occupying 10-15 vol% (Kallemeyn et al., 1991).

190 records for CKs exist in the Meteoritical Bulletin Database, of which 86 are non-

Antarctic. However, only two falls are known: Karoonda and Kobe. We have physical properties for 19 stones from 7 meteorites, including 12 stones of Karoonda itself. The average

86 bulk density by stone was 2.90 ± 0.05 g cm-3 (by meteorite, 3.00 ± 0.11 g cm-3). For grain density, the average by stone was 3.58 ± 0.02 g cm-3 (3.55 ± 0.04 g cm-3 by meteorite).

Porosities therefore averaged 17.8%, though this is heavily influenced by Karoonda with an average porosity of 21%. The average porosity by meteorite is a much lower 14.0%. In terms of magnetic susceptibility, the average value by stone (log χ = 4.67 ± 0.01) was roughly the same as the average by meteorite (4.66 ± 0.02).

This magnetic susceptibility average is just a little higher than the mean of 4.62 ± 0.03 recorded by Rochette et al. (2008), but error bars of each group (based on σmean) overlap. The agreement is remarkable given the difference in which stones are represented. Aside from

Karoonda, Maralinga and Dar al Gani 275 and 431, which agree among the two datasets within uncertainties, the results sample entirely different sets of meteorites.

The overall variability of stones in this class is low in the mineralogically dependent physical properties but is much higher in structure-dependent properties. Magnetic susceptibility only ranged from 4.59 to 4.77 by stone (4.60 to 4.72 by meteorite). Grain densities ranged from

3.37 g cm-3 to 3.66 g cm-3 (both by stone and by meteorite). In grain density-magnetic susceptibility space (Figure 56), CKs reside in the same region as L/LL ordinary chondrites such as Holbrook, and with the exception of a couple CK finds are tightly clustered. Bulk densities ranged from 2.54 g cm-3 to 3.39 g cm-3 (by meteorite, 2.66 g cm-3 to 3.39 g cm-3) because porosities covered the range from zero to 23.4% (23.3% by meteorite). The considerable variability in porosities may be due in part to the influence of weathering. Finds were on average low in porosity, ranging from zero to 17.6%, while falls were all above 16%, most above 20%.

It should be noted that all of the fall stones belonged to just one meteorite: Karoonda, making it

87 uncertain whether the observed effect is due to weathering or just a basic difference between

Karoonda and other CKs.

As was the case with Murchison and Murray for the CMs, the abundance of measured stones of Karoonda permits the exploration of homogeneity for that fall. Again, the 12 stones exhibited homogeneity in all properties measured to within 10%. Standard deviations were: for grain density, 0.04 g cm-3 (1.0% of the mean 3.60 g cm-3); for bulk density, 0.07 g cm-3 (2.4% of the mean 2.85 g cm-3); for porosity, 2.1% (10.0% of the mean 21.1% porosity); and for magnetic susceptibility 0.04 (0.9% of the mean 4.67). Average measurement uncertainties were 0.10 g cm-3 for bulk density, 0.04 g cm-3 for grain density, 2.7% for porosity, and 0.08 for magnetic susceptibility. Therefore, variability among the stones measured was within the bounds of measurement uncertainty. One particular outlier skews some of the results: a 28.1 g fragment from the American Museum of Natural History (reference 3970A) has a grain density noticeably lower than all of the others, yielding a low porosity of 16.4%. If this is omitted (though it should not be omitted in a proper analysis), the standard deviation for the grain density of the remaining stones reduces to 0.02 g cm-3, and for porosity it reduces to 1.4%.

4.2.3.3 CV

Vigarano-type (CV) carbonaceous chondrites have excesses in refractory lithophiles, due to an abundance of large calcium-aluminum-rich inclusions (CAIs) (Hutchison, 2004). They exhibit various signs of some aqueous or thermal metamorphism (Hutchison, 2004; Brearley,

2006) but with one exception (the CV2 meteorite Mundrabila 012) all are petrographic type 3

88

(Meteoritical Bulletin Database). Among carbonaceous chondrites, they are relatively abundant with 176 recorded meteorites, of which 93 are non-Antarctic and 7 are observed falls. Perhaps the best known CV is Allende, of which more than 2 metric tons fell in the state of Chihuahua,

Mexico in 1969 (King et al., 1969; Hutchison, 2004). Allende contains CAIs that have been dated as among the oldest material in the solar system at 4.568 ± 0.003 Ga (Allègre et al., 1995).

CAIs from other CVs and CMs have been dated with comparable dates (Russell et al., 2006).

CVs can be further subdivided. McSween (1977) divided the CV chondrites into oxidized (CVo) and reduced (CVr) subgroups based on metal and Ni contents. The oxidized subgroup formed in a predominantly oxidizing environment, and meteorites of this type are characterized among other things by fayalitic olivine rims around chondrules and CAIs, higher total Fe, Ni enrichment in metal, and higher magnetite to metal ratios, all compared to the reduced subroup (Krot et al., 1995). CVs as a whole have only undergone minor metamorphism after parent body formation (almost all are CV3 in the petrographic schema of van Schmus and

Wood, 1967), and the oxidation of CVo’s occurred after formation of chondrules and CAI’s.

Weisberg et al. (1997) proposed a further subdivision of the oxidized subgroup into Allende-like

(CVoA) and Bali-like (CVoB) meteorites, but for the purposes of this analysis that distinction will not be made. MacPherson and Krot (2002) suggest that, because CVr’s may have been more compacted, they should have less porosity than CVo, and Consolmagno et al. (2008) note that

CVo porosities may exceed CVr porosities by a considerable margin, but caution that their results were based on only two meteorites in each category.

In this study are data from 75 stones of 12 CVs, including 52 stones of Allende alone.

This includes 4 falls and 8 finds. It also includes five meteorites designated CVo and three CVr, according to Krot et al. (1995). (Four meteorites had not been designated in either group.) The

89 following averages and ranges are all by meteorite rather than by stone. Grain densities for the entire population of CVs averaged 3.54 g cm-3, with a range from 3.25 g cm-3 to 3.68 g cm-3.

Bulk densities averaged 3.03 g cm-3 (range 2.59 g cm-3 to 3.46 g cm-3), and porosities averaged

14.6%, with a range from 0.6% to 27.7%. Magnetic susceptibilities averaged log χ = 4.08, with a range from 3.23 to 4.86. This is consistent with an average 4.12 ± 0.07 reported by Rochette et al. (2008).

Of the meteorites included in the porosity study of Corrigan et al. (1997), there is a mix of agreement and discrepancy. Our average porosity for Allende was 21.9% with a standard deviation of 2.9%, in good agreement with both their bulk porosity (20%) and matrix porosity

(25%) for the meteorite. Likewise, there is good agreement with Mokoia (27.7% vs. their 24% bulk and 30% matrix porosities) and Leoville (2.1% vs. their bulk 2% porosity). Efremovka and

Vigarano results, on the other hand, disagree. They measured bulk porosities of 7% for

Efremovka and 1.9 to 7% for various lithologies in Vigarano, while our data are 0.5% and 8.3%, respectively, with variances on the order of 1%. Both of these are low-porosity meteorites, and small variations may be magnified by sample selection effects.

Subdividing the CVs into their oxidized and reduced subgroups provides some tantalizing insights into the differences between the two groups. While their grain densities are similar

(both have an average grain density of 3.50 g cm-3, slightly lower than ungrouped CV meteorites,

-3 which average 3.62 g cm ), their bulk densities are very different. The average CVo bulk

-3 -3 density is 2.87 g cm , while for CVr it is 3.37 g cm . Because the grain densities are the same on average, the difference is entirely structural, due to a much higher porosity for the CVo group than for the CVr group. Indeed, CVo meteorites average 19.7% porous, while CVr’s average only 3.6%, an even greater disparity between the two groups than Consolmagno et al. (2008)’s

90 initial report. There is also a considerable difference in magnetic susceptibilities between the two groups. The average CVo magnetic susceptibility is log χ = 3.84, while for CVr the average is 4.70. This difference may be accountable for the abundance of unoxidized metal in CVr as compared to CVo. Rochette et al. (2008) do not distinguish averages for oxidized and reduced subgroups, but do note that CVr’s tend to higher magnetic susceptibilities while CVo meteorites lie at both extremes of the data.

The differences listed above incorporate both falls and finds. However, there may be some weathering effect present. Most notably, the CVo find Nova 002 has a grain density of

3.25 g cm-3, much lower than average, and a correspondingly low porosity of 6.0%. In Figure

57, which plots grain density against magnetic susceptibility for CVs, Nova 002 lies far outside the region occupied by other CVo’s. Omitting it and the other CVo find, Axtell (which exhibits no obvious weathering effects), yields an average CVo porosity of 24.8%. The average magnetic susceptibility for CVo falls is slightly higher when finds are omitted, at 3.95, but it should be recognized that Nova 002’s magnetic susceptibility is higher than average at 4.12. If the same is done for the CVr subgroup, the only fall for which data have been collected is Vigarano, which has a porosity averaged among all stones that were measured of 8.3% but a lower-than-average

(for CVr) magnetic susceptibility of 4.41.

The four CV meteorites in this study that are not explicitly grouped into CVo or CVr (Dar al Gani 1040, Sahara 98044, and NWAs 2140 and 3118) are not necessarily on their own. They probably belong to either the oxidized or reduced subgroups, but have not yet been designated to one or the other. In the absence of more definitive analysis, grain density and magnetic susceptibility may be used for tentative group assignments. DaG 1040, NWA 2140 and NWA

3118 each lie clearly within the CVo region (Figure 57). By way of confirmation, each also has

91 porosities exceeding 20%, more closely resembling the expected porosities of CVo rather than

CVr. Sahara 98044, on the other hand, resides ambiguously between the two regions. Given its low porosity of 5.2% ± 2.0%, it likely should be assigned to the CVr subgroup.

Of all of the meteorites in this study, no single meteorite is better represented than

Allende, with 52 total stones included, 40 of which possess the full suite of measurements. The stones group fairly well in all properties, which is illustrated for grain density and magnetic susceptibility in Figure 57. Standard deviations among the population are: for grain density, 0.06 g cm-3 (1.6% of 3.65 g cm-3); for bulk density, 0.08 g cm-3 (2.7% of 2.86 g cm-3); for porosity,

2.7% (12.7% of 21.5% mean porosity); and for magnetic susceptibility, 0.06 (1.6% of 3.65).

Average measurement uncertainties are 0.03 g cm-3, 0.05 g cm-3, 1.5%, and 0.10, respectively, indicating that variability is small but exceeds measurement uncertainties for densities and porosity, though not for magnetic susceptibility. 39 of the stones were measured from the collection at the American Museum of Natural History, while 12 stones were at the Smithsonian

Institution-National Museum of Natural History and one was in the Vatican collection. The two different collections (AMNH and NMNH) yielded slight discrepancies for all of the properties measured. Grain densities were higher for AMNH stones (3.66 ± 0.01 g cm-3) than for NMNH

(3.62 ± 0.02 g cm-3), as were magnetic susceptibilities (3.66 ± 0.01 and 3.61 ± 0.02, respectively). Bulk densities were lower for AMNH (2.84 ± 0.01) than for NMNH (2.93±0.01), resulting in higher average porosities (23.4% ± 0.4% and 19.0% ± 0.5%, respectively), with all uncertainties given as σmean. While differences in any one property may be accountable to methodological errors, especially since work at AMNH was performed before measurement techniques were perfected, the fact that differences are apparent in every physical property included in the study suggests that the populations as a whole differ. This may be attributable to

92 differences in their curation history at two different institutions under different climate conditions for the past forty years.

One very likely culprit is humidity: carbonaceous chondrites can act like a sponge, absorbing water from the air. By way of illustration, one stone of the CM2 Murray at the Field

Museum was measured at 0.16 g more than its catalog listing. The extra water adds mass, raising bulk density. Because the additional material has low density (1.0 g cm-3), it will reduce the overall grain density of the stone, and by filling space it will also reduce porosity. Also, water is weakly diamagnetic, which could negatively influence magnetic susceptibility measurements.

4.2.4 Other Carbonaceous Chondrites

A few carbonaceous chondrites do not fit neatly into subgroups. Of these, we have data for Adelaide, Coolidge, Ningqiang, Dar al Gani 430, Loongana 001, and Northwest Africa 1152.

Aside from Ningqiang, all are finds. Because they do not form a subgroup of their own, it is not particularly meaningful to express average values for the physical properties measured. Their data are included in Table 13, with grain densities and magnetic susceptibilities also presented graphically in Figure 58. They range in grain density from 3.27 g cm-3 (Adelaide) to 3.66 g cm-3

(Ningqiang), and in bulk density from 2.80 g cm-3 (Ningqiang) to 3.55 g cm-3 (Coolidge).

Ningqiang has the highest grain density of the group and the lowest bulk density, giving it the highest total porosity (23.6%). Magnetic susceptibilities for this group ranged from 4.32

(Loongana 001) to 5.14 (Coolidge).

93

Earlier attempts at classifying Ningqiang had placed it as an anomalous CV (Rubin et al.,

1988), though Kallemeyn et al. (1991) later dubbed it an anomalous CK due to petrographic and chemical similarities, calling it closer to CK than CV. Unlike most CKs, it is relatively unequilibrated and is assigned petrographic type 3. Its grain density and magnetic susceptibility are quite consistent with that of the CK group (Figure 56 and Figure 58) These properties also reside within the region defined by the oxidized subgroup of CVs (Figure 57 and Figure 58), though its magnetic susceptibility (log χ = 4.69) is considerably higher than the average of 3.84 for CVo. Bulk density and porosity are also similar to CVo falls such as Allende, but not inconsistent with the one CK fall, Karoonda.

NWA 1152 had also been originally classified as a CV (Russell et al., 2002) and is included among CV3s in the data of Rochette et al (2008), though the Meteoritical Bulletin

Database currently lists it as ungrouped. Its density (bulk density 2.84 g cm-3, grain density 3.49

-3 g cm ) and porosity (18.9%) are consistent with CVo or ungrouped CV, though like Ningqiang it has an unusually high magnetic susceptibility of 4.81. Of the remaining meteorites in this group,

Adelaide had at one time been considered CM (Grady, 2000), though its grain density and magnetic susceptibility exceed the range of CM, while its porosity (3.1%) is much lower than typical for CMs.

4.3 Discussion

The carbonaceous chondrites are a diverse group, as their physical properties indicate.

Nevertheless, there are a few general conclusions that can be pulled from the data. First,

94 carbonaceous chondrites are more porous and less dense on average than ordinary chondrites.

This is no surprise, as it has been observed before (cf. Consolmagno et al., 2008). Even more insights may be gleaned by looking at trends in OC physical properties as a function of other properties such as petrographic type and shock.

4.3.1 Petrographic Type

For the purposes of this aspect of the study, only CC falls were included so as to avoid the influences of any possible weathering effects in the finds. When the data are organized by petrographic type, this naturally groups them by different chondrite types. For example, the one meteorite with a petrographic type of 1 is the one CI (Orgueil) in this study. Type 2 consists of

CM and CR falls. Type 3 is occupied by CV, some CO, and Ningqiang, and type 4 consists of other CO falls and the CK fall Karoonda. (CH and CB are not represented, since there are no falls in the study.) When grouped in this way, however, their physical properties begin to reveal trends. The average values of the data are presented in Table 14.

Bulk and grain densities (Figure 59 and Figure 60) are both very low for the aqueously altered petrographic types (1 and 2), with type 1 much lower than type 2 in both physical properties, and both show signs of leveling out for petrographic types 3 and above (though unfortunately no representatives of thermally altered carbonaceous chondrite falls were sampled above type 4). Porosity follows an inverse trend (Figure 61), with porosity decreasing on average with increasing petrographic type. In Chapter 7, this result will be extended to include all chondrites: carbonaceous, enstatite, and ordinary, and it will be discussed further there.

95

Unlike the other three physical properties, magnetic susceptibility exhibits no overall trend. With the exception of the type 2 group (CMs and CRs), which are low compared to the others, all groups have magnetic susceptibility averages at approximately log χ = 4.5.

4.3.2 Shock

For this analysis, there were too few falls with shock states recorded in the literature

(Grady, 2000; Meteoritical Bulletin Database) to get any meaningful data; almost all falls were

S1, with only one S2, two S3, and nothing higher than that. In order to increase the number of higher-shock meteorites in the analysis, it became necessary to expand it to include finds as well.

This may result in some interference in the data due to weathering-related effects. The addition of finds increased the number of S2 to 12, with three S3 and two meteorites of S4. Results are summarized in Table 15. The bias toward low-shock meteorites itself raises an interesting question: is there a selection effect at play here, or do carbonaceous chondrites suffer relatively little shock? That question, unfortunately, lies outside the scope of this dissertation.

Bulk density follows an increasing trend (Figure 63) from S1 to S4, with grain density staying relatively level at least over the interval S2 to S4 (Figure 64). This implies that shock compression has taken place, a fact that is not so obvious in the porosity data (Figure 65) primarily because of the inclusion of the S3 meteorite Felix (CO3) with a porosity of 21%. With only two S3 meteorites with porosity values in the sampling, this one high-porosity exception (of which three separate stones were measured, so the porosity is not erroneous) hides what should otherwise be a reduction of porosity with increasing shock. The fact that this one meteorite that

96 has suffered substantial shock has not been compressed as virtually all other shocked meteorites have is itself an intriguing question that will require careful structural analysis to resolve (also outside the scope of this dissertation).

As for magnetic susceptibility, the lack of hard data for S3 and S4 makes it impossible to determine if any trend is present. One noticeable feature in the magnetic susceptibility data

(Figure 66) is the difference in magnetic susceptibility between S1 and S2, each of which are represented by a statistically meaningful number of meteorites. Since this difference does not continue into shock stages 3 and 4, it is likely that there is some sampling bias present. The S1 population samples a large number of different meteorite types, including most of the CM and

CVo that have low magnetic susceptibilities, while the S2 population is dominated by CO finds.

97

CHAPTER 5: ENSTATITE CHONDRITES

Note: The contents of this chapter are taken from a paper that is in press (Macke et al.,

2010c)4 and has been inserted here with minor modifications, including the inclusion of updated data, the removal of the abstract, the removal of the acknowledgments and measurement sections

(which are redundant with information already provided in this dissertation), a new discussion of weathering modulus applied to ECs, and typographical adjustments. Please see Chapter 2 of this dissertation for more detail regarding the measurement techniques.

5.1 Introduction

Enstatite chondrites (ECs), characterized in part by an abundance of nearly FeO-free enstatite are relatively rare among known meteorites, with only 17 recorded falls (a little over 1% of all meteorite falls) and about 400 finds, mostly Antarctic with only 110 non-Antarctic finds known (Grossman, 2009). Most of the iron in ECs is contained within metal grains (Dodd,

1981). By analogy of the H-L-LL nomenclature adopted for groups of ordinary chondrites which differ in total iron content, the two groups of ECs were named EH and EL for high-Fe and low-Fe by Sears et al. (1982), though differences between the two groups in chemistry,

4 Meteoritics & Planetary Science © 2010 by the . Used with permission (see Appendix C). 98 mineralogy and texture were understood at least as early as the 1960s (Anders, 1964; Keil,

1968), and the two groups exhibit sufficient differences to establish their origins on separate parent bodies (Keil, 1989). The two groups also differ in degree of thermal metamorphism, with most EHs being type 4 or 5 on the petrographic type scheme of van Schmus and Wood (1967) and most ELs type 6, though representatives for both groups covering the range from type 3 to type 6 are known (e.g. Sears et al., 1984; Prinz et al., 1984; Zhang et al., 1993).

Though many studies on chemistry and mineralogy of enstatite chondrites have been conducted, there have been relatively few studies of their physical properties, such as density and porosity. To date the only systematic studies have been Guskova (1985) and Rochette et al.

(2008). Guskova (1985) measured density and magnetic properties for 21 stones from ten meteorites in the collections of the Soviet Union, many of which were small and one less than 1 g. Rochette et al. (2008) measured magnetic susceptibilities of ~150 stones from 72 meteorites, including many Antarctic finds. No study of porosity conducted under a consistent methodology for a statistically meaningful number of enstatite chondrites has been conducted before now.

Physical properties, generally measured on bulk samples of greater than 10 grams, yield information about the whole rock at scales generally not studied in more detailed chemical analyses. Grain density, the density of the solid component of the meteorite, is effectively a mass-weighted average of the mineral species composing the sample. Porosity is the percentage of the sample occupied by pore space (cracks, voids, and intergranular spaces), and is determined by comparison of grain density with bulk density (the density determined by the total volume enclosed by the sample). Physical properties such as porosity are informative of physical conditions under which the meteorite parent bodies formed and under which they were physically altered, as well as providing information about large-scale structure of analogous

99 asteroids. Magnetic susceptibility, the degree to which a sample exposed to a magnetic field will respond to that field, is also an important physical property. It depends primarily on the quantity of paramagnetic and ferromagnetic materials in the sample, so serves as a good first-order indicator of metallic iron quantity. It is not a perfect indicator, as nonmetallic materials, including phosphides such as , may contribute to magnetic susceptibility.

Nevertheless, these phases constitute less than 1 wt% of the typical enstatite chondrite (Keil,

1968) and so are negligible compared to the contribution of metallic iron. In the case of ordinary chondrites, magnetic susceptibility in conjunction with grain density (also largely dependent on the quantity of dense iron metal relative to less-dense silicates and other phases) has been shown to serve as a viable tool for rapid classification of stones into H, L, and LL subgroups

(Consolmango et al., 2006). The same should occur for EH and EL if the difference in their iron content is significant.

The dearth of studies of enstatite chondrite physical properties is due in part to their relative rarity and to the lack until recently of methods for measuring bulk density that would avoid contaminating or destroying specimens, as would be the case with typical fluid immersion.

Consolmagno and Britt (1998) developed a fast, non-destructive, and non-contaminating method for performing these studies on hand-sized stones using small glass beads that collectively behave as an Archimedean fluid. The development of this technique, in conjunction with other fast, non-destructive and non-contaminating techniques for measuring grain density and magnetic susceptibility, has enabled large surveys of many hand-sized stones in major meteorite collections.

With this goal in mind, we have visited eight major meteorite collections: the Natural

History Museum in London, the Vatican meteorite collection in Italy, the Smithsonian Institution

100 in Washington DC, the American Museum of Natural History in New York, the Field Museum in Chicago, and three university collections at Arizona State University, the University of New

Mexico, and Texas Christian University. We have performed physical property measurements on 1228 stones from over 664 individual meteorites, of which 46 stones were from enstatite chondrites. All but one of the stones from enstatite chondrites exceeded 10 g (The exception, a

7.3 g piece of Pillistfer from the Vatican collection, has only been measured for magnetic susceptibility). Twenty-three distinct EC meteorites were sampled, including 11 of the 17 known falls and a large percentage of the non-Antarctic finds with > 10 g stones available for study. Our measurements included 18 stones from nine EH meteorites (five falls, four finds) as well as 28 stones from 14 EL meteorites (six falls, eight finds). This data set provides sufficient statistics of enstatite chondrite physical properties to establish trends caused by weathering and to study whether the EH/EL distinction is manifest in grain density and magnetic susceptibility.

5.2 Results

5.2.1 Grain Density

A summary of all data can be seen in Table 16 and Figure 67 through Figure 70. Grain density ranged from 3.17 to 4.46 g cm-3, with the majority of stones falling between 3.5 and 3.8 g cm-3. This is slightly lower than found in H chondrite falls (3.6 to 3.9 g cm-3) and roughly comparable to L falls (3.5 to 3.9 g cm-3). Enstatite chondrite finds exhibit significantly reduced

101 grain density compared with falls (Figure 67). The average grain density for finds is 3.51 g cm-3, which is significantly lower than the average for falls at 3.66 g cm-3. That being said, one stone of the EL find Blithfield had an anomalously high grain density of 4.46 g cm-3. This drop in grain density in finds has not previously been observed in enstatite chondrites due to low statistics (Consolmagno et al., 2008). Nevertheless, it is generally well-understood as the result of weathering of iron which expands as it oxidizes and has been well-studied in ordinary chondrites (Bland et al., 1998b, 2006; Consolmagno et al., 1999).

Since our data indicate that finds are no longer representative of the original densities of the meteorites, we restrict further discussion of grain density to falls. EH and EL grain densities, while not identical, are quite similar and exhibit substantial overlap (Figure 67). The average grain density for EH falls was 3.66 g cm-3, with a standard deviation of 0.06 g cm-3, while for EL falls it was also 3.66 g cm-3 with a larger standard deviation of 0.16 g cm-3. Even eliminating the somewhat anomalous sample of Khairpur (4.17 g cm-3), the average EL fall is 3.63 g cm-3 with a standard deviation of 0.10 g cm-3. The range for EH falls is from 3.52 g cm-3 to 3.76 g cm-3, and for EL falls (not counting the aforementioned sample of Khairpur) is 3.45 g cm-3 to

3.80 g cm-3.

5.2.2 Bulk Density

Bulk density results follow the same trends established under grain density. Bulk density for the entire population ranged from 2.89 g cm-3 to 4.51 g cm-3, with the majority of stones falling between 3.35 g cm-3 and 3.65 g cm-3. As with grain density, enstatite chondrite finds

102 exhibit a significant reduction in bulk density as compared with falls (Figure 68), with the average bulk density for all falls being 3.49 g cm-3, and the average for finds is 3.32 g cm-3

(including the anomalously high bulk density of the aforementioned Blithfield stone). This is unlike comparable results for ordinary chondrites, where no substantial change in bulk density is anticipated or observed as a result of weathering (Consolmagno et al, 1998).

Again, EH and EL falls do not exhibit strong differences in this property. The average bulk density for EH falls is 3.56 ± 0.12 g cm-3 (with ―±‖ representing one standard deviation of sample values), and for all EL falls it is 3.58 ± 0.21 g cm-3. Eliminating the high value for the same Khairpur stone as before, it is 3.55 ± 0.16 g cm-3. The range for EH falls is 3.25 g cm-3 to

3.73 g cm-3, and for EL falls (minus Khairpur) it is 3.15 g cm-3 to 3.78 g cm-3.

5.2.3 Porosity

Porosities for enstatite chondrites are on the low end for the chondrite group. Measured porosities in this survey ranged from effectively zero to 13%, with most stones falling below 7%.

By comparison, ordinary chondrites average about 9% porosity with a range that exceeds 20%, and carbonaceous chondrites tend to much higher porosities. Consolmagno et al. (2008) noted the possibility of two populations of enstatite chondrites; one with relatively high porosity

(above 10%) and one with low (below ~6%), but cited the need for more statistics. We do not see such a trend in these data. Only four stones exceeded 10%, two of which belong to meteorites with multiple stones included in the survey, but are the only members of their group to have high porosity. We do, however, see an unexpected trend with regard to finds. Rather

103 than a reduction in porosity, finds exhibit on average an enhanced porosity (Figure 69). The average enstatite chondrite fall has a porosity of 2.4%, with a range from zero to 11.7%. The average find has a porosity of 5.5%, with a range from zero to 13.2%, and with most stones exhibiting porosities greater than 2%. Finds include the two stones (Sahara 97096 and North

West Forrest) with the highest porosity measured in this survey. Porosity enhancement in finds correlates well with the reduction in bulk density mentioned above.

EH and EL fall populations have very similar porosities. Both range between zero and

11.7%, with average values of 2.8% and 2.1%, respectively. All but one of the EH falls in this study exhibit porosities below 5%, while all but one of EL falls lies below 7%, with three stones between 5% and 7%. Given measurement uncertainties typically exceeding 1%, we do not recommend reading too much into this small discrepancy.

5.2.4 Magnetic Susceptibility

Magnetic susceptibilities are reported in log units of 10-9 kg3 m-1. Values ranged from

4.16 to 5.72, with most stones between 5.3 and 5.6. The effect of weathering on finds is most observable in magnetic susceptibility as a severe reduction (Figure 70), due to the loss of magnetic components in the material. In a plot of grain density against magnetic susceptibility

(Figure 71), both of which vary with metallic iron content, it is apparent that the reduction in magnetic susceptibility correlates with that of grain density, though the effect is much more pronounced in magnetic susceptibility. Falls averaged 5.48, with a range from 5.30 to 5.68, while finds averaged 5.03, with a range from 4.16 to 5.72. EH and EL falls are again quite

104 similar. EH fall magnetic susceptibilities averaged 5.45, ranging from 5.30 to 5.58, and EL fall values averaged 5.50, ranging from 5.30 to 5.68. The averages for the two groups of falls are within measurement uncertainties of those reported by Rochette et al. (2008) for falls in their survey.

5.3 Discussion

5.3.1 Weathering Effects on Finds.

As noted, weathered finds exhibit on average a reduced grain density, bulk density and magnetic susceptibility, and an increased porosity. It is difficult to correlate these effects to established weathering states, for a number of reasons. First, only a few meteorites from this set have had determinations of weathering states made. Even in those cases, weathering determinations were made for stones other than those measured in this survey, and weathering effects can differ for stones of differing volumes. In addition, criteria for determining weathering states are not well-established and rely to some degree on the personal judgment of the investigator.

Though we cannot easily link our data to established weathering states, we can determine a rough degree of weathering for all stones, and establish how the degree of weathering affects trends in physical properties. Because the weathering effect on magnetic susceptibility due to oxidation of metallic iron in iron-rich meteorites is well understood (Consolmagno et al., 2006)

105 and is the most pronounced of the four effects, we arbitrarily subdivided the enstatite chondrite finds into four groups based on that property, as in Figure 72a. Group 0 exhibits minimal alteration, and group 3 is the most severely altered. By coincidence of the data, there are recognizable gaps between each of the groups. The three meteorites with recorded weathering states are in the following groups: Ilafegh 009 (W0/1) and Sahara 97096 (W1) and are in group

0, and NWA 3132 (W4) is in group 2.

The weathering effects of all four properties in this survey correlate with each other.

Grain density correlates somewhat with magnetic susceptibility (Figure 72a), though the effect on grain density is clearly less pronounced that that of magnetic susceptibility. This is quite reasonable if the primary effect of weathering is oxidation of metals. As iron oxidizes (reducing magnetic susceptibility), it expands, thus increasing grain volume without significantly affecting total mass, and so grain density is reduced.

Trends in bulk density and porosity are also related to degree of weathering (Figure 72b).

Group 0, as expected, aligns with falls. Groups 1 through 3 are clearly distinguished on the bulk density/ porosity plot, and progress along a trend toward decreased bulk density and increased porosity as weathering state increases. This is the first time such a trend has been observed, because it is the first time enough statistics have been gathered to make the trend visible.

This result can be better illustrated by use of the weathering modulus, as described in

Chapter 3 of this dissertation. For ECs, the weathering modulus takes the following form:

2 2 WEC4 (logχ )00 log χ 1.2( ρ g ρ g ) (33)

106

In a plot of porosity vs. weathering modulus (Figure 73), there is an obvious positive correlation between porosity and degree of weathering. It should be noted that this is measured porosity and not model porosity. Since EC weathering behavior is obviously not the same as that of OCs, it cannot be assumed that bulk density remains relatively unaltered in the weathering process; in fact, the average EC find bulk density is 3.32 g cm-3, or 0.25 g cm-3 less than EC falls. Thus, any model porosities calculated from EC find bulk densities would be at best misleading.

The results for bulk density and porosity, unlike those for grain density and magnetic susceptibility, are somewhat unexpected. To first order, oxidized metals should expand into available pore space, reducing porosity while not substantially affecting bulk density, as is observed in ordinary chondrites (Bland et al., 1998b, Consolmagno et al., 1998). Two possible explanations exist for the observed trends in enstatite chondrites. The first is an extension of the standard model for weathering. Enstatite chondrites begin with very low porosity, so that even minor weathering would fill what pore space exists. With further weathering, the expanding iron oxides would force apart the meteorite, cracking it and introducing new pore space while at the same time increasing the bulk volume and hence decreasing bulk density. The question here is how much of this expansion can take place before the structural integrity of the stone is compromised and the stone disintegrates. None of the samples in this survey were excessively friable, and so this explanation begs the question of why that would be the case. The second possibility is that materials are leached out during weathering. One obvious candidate mineral is (CaS), which is so water soluble that samples of the EH4 fall Parsa exhibited signs of oldhamite loss due simply to moisture in the environment where it was stored (Bhandari et al.,

1980). ―Pits‖ left over from oldhamite leaching were observed in Yilmia (Buseck and

Holdsworth, 1972). Oldhamite alone is insufficient to account for the observed effects, however.

107

First, it is not so abundant to account for the excess porosity; literature values place its abundance in falls between 0-2 wt% (Keil, 1968; Bhandari et al., 1980; Rubin and Keil, 1983;

Rubin et al., 1997; Rubin, 1983a,c). Assuming the maximum 2% original abundance and total loss, this can account for less than 3% porosity. This also assumes that oxidation products from metals do not fill in the new cavities, which is not a reasonable assumption as oxidation products have been observed even in mildly weathered falls (cf. Bhandari et al., 1980). Second, oldhamite is a low-density phase compared to the average EC grain density, so its loss would actually increase grain density. Finally, it would not be able to account for the considerable drop in magnetic susceptibility over 1.5 orders of magnitude. To account for all of the observed effects through leaching, it is necessary that metals rather than merely accessory phases are leached out during weathering. These two possibilities (cracking vs. leaching) can be resolved by further studies aimed at characterizing the amount of oxides present in weathered finds.

5.3.2 Heterogeneity

Where multiple stones from an individual meteorite have been measured, results have varied widely, much more than is typical in other chondrites. For those meteorites with multiple stones measured, the bulk density range was on average 0.3 g cm-3, grain density range was 0.2 g cm-3, porosity 6%, and magnetic susceptibility 0.23, all of which exceed individual measurement uncertainties. The range is exaggerated among finds compared to falls, not a surprising result given the variable effect of weathering on different sized stones, though weathering is not the only source of heterogeneity.

108

For particular meteorites, variability is even more extreme than described above. The most extreme example, Blithfield, ranges about 1 g cm-3 in both bulk density and grain density, and ranges in magnetic susceptibility by about 0.41 (a factor of about 2.6). This inhomogeneity is not atypical for Blithfield, which is an impact breccia with large troilite-rich clasts (Rubin,

1984); one small sample studied by Keil (1968) had an anomalously low iron content of 12.9 wt%, much lower than the ~25 wt% average for the other Type II’s in his study. This has been attributed to the fact that the sample studied by Keil was one of these troilite-rich clasts (Rubin,

1984). For a truly representative sample of Blithfield, as much as 200 g may be required (Rubin, personal communication). While Blithfield may be the most inhomogeneous meteorite in this study, it is not the only one to exhibit large variability. Many other enstatite chondrites are breccias, including (Dawson et al., 1960; Rubin and Keil, 1983), Adhi Kot (Rubin, 1983a),

Atlanta (Rubin, 1983b), Hvittis (Rubin, 1983c), and Happy Canyon (Rubin et al., 1997), which may account for much of the inhomogeneity, though non-brecciated ECs also exhibit considerable variability between stones. Khairpur is not listed as a breccia, but of the three stones measured, one 24.9 g stone is anomalously high in both grain density and magnetic susceptibility, as described in preceding sections of this paper.

Hutson (1996) compiled literature data from seven separate studies of enstatite chondrite elemental abundances and observed a large degree of intra-meteorite variability in bulk chemical abundances, in particular in iron and sulfur. These were attributed to the fact that all of the analyses had been performed on samples below 8 g. Jarosewich (1990) observed in the context of ordinary chondrites that for meteorites with coarse iron grains the inclusion or exclusion of individual iron grains may skew results, and recommended a minimum mass of 10 g for homogeneity. Many enstatite chondrites have very coarse metal grains, with EL grains much

109 coarser than EH (Easton, 1983), and so this effect must be taken into consideration. (Easton’s work expresses a difference due to petrograpic type, since all EL’s in the work were type 6 and all EH’s were type 4-5, but this does not invalidate the statement, since as a population ELs are dominated by type 6 and EHs are dominated by lower petrographic types, and the ―definitive‖ studies of the chemical and mineralogical differences between EH and EL were performed on

EL6 and EH4-5 stones. In addition, there are indications that the size difference may not be solely due to petrologic type and even EL3 metal grains are larger than EH3 grains [Schneider et al, 1998].) Nevertheless, all but one of the stones in our survey exceeded 10 g, some by an order of magnitude, and so we conclude that the variability we observe is due to larger-scale causes.

We have compared the variation in grain density with sample mass for EC falls (Figure 74), and find not only that the standard deviation of grain density results increases with decreasing mass, but that it exceeds similar results for H and L falls. The data exhibit scale-dependent variability that remains above 0.08 g cm-3 for all masses below ~ 40 g and does not appear to level out until at least ~ 60 g. This indicates that the samples exhibit scale-dependent inhomogeneities at low masses, and we consider the minimum mass for a representative homogeneous sample to be at least 40 g. Also plotted in Figure 74 are EH and EL fall data, from which it is apparent that EL falls are considerably more inhomogeneous than EH falls. We observe a similar effect in bulk density, for masses up to ~ 60 g, though that physical property may be strongly influenced by porosity variations as well as mineralogical differences and so is not included here. Curiously, we do not see any scale-dependent effects in magnetic susceptibility, and Rochette et al (2008) only observed heterogeneity below 1 g. The reason for the discrepancy eludes us, though grain density data alone are sufficient to establish that EC falls are heterogeneous below 40 g.

110

5.3.3 Grain Density, Magnetic Susceptibility and Metallic Iron Content

Since both grain density and magnetic susceptibility vary with the amount of metallic iron present in a sample, together they give a respectable first-order indicator of relative amounts of iron metal present in various meteorites. In the case of ordinary chondrites, the difference in metal between H, L, and LL falls results in three distinct populations on a grain density/ magnetic susceptibility plot (Figure 20). If the EH-EL distinction is truly analogous to the H-L-

LL distinction among ordinary chondrites, then they too should group in two distinct populations. We do not observe this, as can be seen in Figure 75. For that figure, we used mass- weighted meteorite averages (Table 17) of only those meteorites for which we measured more than 40 g. This has the disadvantage of considerably reduced statistics over the use of individual stones (there are only four EH and five EL data points), but it resolves doubts that may arise from the inherent variability among stones. The basic result does not change if individual stones are included. The two populations overlap so closely as to be unable to distinguish whether the small observed difference in population averages is real or a result of statistics of small numbers.

They do not spread into distinct groups.

Taking our results to be representative, we see a slight difference between EH and EL populations, with EH slightly higher in both grain density and EL slightly higher in magnetic susceptibility, but with a substantial overlap between the two. Grain density averaged (by meteorite, not by stone) 3.69 g cm-3 for EH, with one standard deviation of 0.06 g cm-3, and averaged 3.68 g cm-3 with standard deviation of 0.13 g cm-3 for EL. Magnetic susceptibility averaged 5.45 ± 0.05 for EH and 5.50 ± 0.05 for EL. The centroids and range for the data are particularly sensitive to small variations in the data; in an earlier analysis of the data that

111 included fewer stones, the magnetic susceptibility for EL was lower than for EH and the EL grain density was as low as 3.63 g cm-3. The similarities between EH and EL in both magnetic susceptibilities and grain densities implies no substantial difference in metallic iron quantities.

We point out also that some ELs have grain density and magnetic susceptibility values that exceed that of some EHs, indicating that in those instances the ELs actually have more metallic iron than the EHs.

Rochette et al (2008) were the first to note the almost identical magnetic susceptibility values for EH and EL populations and to infer from the data that the groups do not differ in quantity of iron metal. They did not question the literature with regard to the total iron quantity of the two populations, and they posited that the reported differences in iron were likely due to nonmetallic sulfide-bearing phases such as troilite (FeS). They also cited the similarities in grain density between the two groups (based on literature data) as confirmation of this. However, given the fact the density of troilite at 4.9 g cm-3 is higher than the average EC density of ~3.64 g cm-3, any sizeable difference in the quantity of the mineral should appear as a recognizable difference in meteorite grain density.

In order to better understand how grain density and iron content vary with mineralogy in enstatite chondrites, we constructed a simple model based on a mixture of nearly pure enstatite at

3.1 g cm-3, pure (7.9 g cm-3) to represent the metal, troilite (4.9 g cm-3), and 8 wt% plagioclase (2.7 g cm-3), based on abundances given in Mason (1966). We varied the kamacite and troilite percentages, leaving enstatite as the dependent quantity. We chose to use the same metallic iron quantity (21.5%, close to values determined by Keil, 1968) to represent both EH and EL, and varied the amount of troilite. In this case, a 3.5 wt% difference in the abundance of troilite (from 7.5% to 11%) accounted for the (largest reported) difference between the average

112 grain densities of the two groups of 3.69 and 3.63 g cm-3. This difference corresponds to a difference of only 2.2 wt% in total Fe between the two groups (26.2% and 24.0%, respectively).

(If we consider the actual average grain densities for meteorites with > 40 g measured, the differences between the populations fall to less than 1% for troilite and less than 0.7% for total

Fe.) This model is relatively insensitive to the abundance values chosen; for a specified metallic abundance, the observed difference in grain densities can be achieved by varying the troilite abundance by approximately 3.5-4%.

This model is not perfect, as it omits a number of less-abundant mineral phases. To exemplify this, the sulfur abundances according to the models are 4% (EH) and 2.7% (EL), while values from the literature (compiled in Hutson 1996) average 5.3% and 3.1%, respectively.

Some of this difference can be accounted for with the incorporation of varying amounts of other sulfide-bearing phases such as oldhamite (CaS). Nevertheless, a few important conclusions can still be drawn from the model. First, if the observed difference in average grain density is to be achieved by iron metal alone, the difference in metal quantity must be very small (about 2%).

Also, in that case the troilite variation between the two groups must also be very small (much less than 3.5%) which is not consistent with the literature (cf. Mason 1966). Second, even keeping the metallic iron quantities the same, the variation in troilite is not very dramatic, and the resulting difference total iron abundance of just over 2% is nowhere near as pronounced as some researchers have implied.

We would like to point out also that the meteorite-to-meteorite variation implies a considerable range of iron abundances within both EC subgroups. Based on grain densities alone, the range in total Fe abundance between EH falls for which we measured more than 40 g is anywhere from 4% to 5%, depending whether it is metal or troilite that varies, and for EL falls

113 the range is anywhere from 6% to 7% (omitting Khairpur). With smaller stones taken into account, EH falls vary by as much as 8-9% Fe and EL falls vary by as much as 12-13%. Much of this variation is likely due to differences in metal between individual meteorites, as exemplified by the corresponding range of magnetic susceptibilities, though some of the variation is also likely due to troilite.

Our results here are consistent with some other findings as well. Most notably, Keil

(1968) performed a study of chemical abundances in enstatite chondrites using electron microprobe x-ray analysis of meteorite thin sections. He analyzed 3 ―Type I‖ (EH) meteorites, and 7 ―Type II‖ (EL). He found an average 25.3 wt% Fe for type I, and 23.3 wt% Fe for type II, with substantial variation among the two groups – with standard deviations of 3.6 wt% for type I and 6.0 wt% for type II. His data include finds as well as falls, and one sample of Blithfield is clearly anomalous, with only 12.9 wt% Fe and a total metal abundance an order of magnitude lower than all other samples in the study. Eliminating this, the average iron in Keil’s type II group increases to 25.1 wt% (standard deviation 4.4%). We also note that the sample in Keil’s study with the highest measured iron content belonged to an EL, not an EH. Using Keil (1968)’s mineral abundances with fresh averages calculated for the EH and EL falls, we modified our simple model to include other minor phases. This produced average grain densities of 3.63 g cm-

3 for EL and 3.68 g cm-3 for EH and total Fe abundances at 24.1 wt% and 26.0 wt% for EL and

EH, respectively, with a standard deviation among individual meteorites of 5 wt% and 3 wt%, respectively. This indicates both that Keil’s data are consistent with our own and that our model produces Fe abundances that are reasonably consistent with the modal analysis. In addition,

Hutson (1996) measured elemental abundances in large-area thin-sections of an EH (Qingzhen)

114 and an EL (MacAlpine Hills 88136), each of petrographic type 3 and with minimal weathering, and found iron abundances within measurement uncertainties of each other.

This of course raises the question of why some earlier studies have exhibited more pronounced differences in iron quantity. In one of the definitive studies of chemical abundances in ECs, Kallemeyn and Wasson (1986) observe a notable hiatus in Fe/Mg ratios between EH and

EL, which differ from each other by a factor of almost 2. Fe/Si ratios also differ at 0.9 and 0.6, respectively. We note that the magnesium-normalized abundances are in part influenced not only by absolute Fe, but by absolute Mg as well which exaggerates the difference. Wasson and

Kallemeyn (1988) list individual element abundances, and Mg is more abundant in EL than EH by a factor of 1.3 on average. Si abundances follow a similar trend. They still observed a difference in absolute Fe between the groups (290 mg/g for EH and 220 mg/g for EL). This is a difference of 7 wt% in absolute abundance, and a relative difference between the two of about

30%. This discrepancy is still large enough to beg for an explanation.

We do not question the instrumental accuracy of the neutron activation analysis that they and other researchers performed. One possible explanation is that the excess Fe is hidden in nonmetallic phases that do not significantly affect grain density or magnetic susceptibility, though the studies by Keil (1968) and Hutson (1996) should then testify to the difference if it is present, which they do not. Another possibility that must be seriously considered is that of sample bias when preparing small quantities of material for analysis. We remind the reader that

Jarosewich (1990) observed in studies of ordinary chondrites the need for representative samples of 10 g or more to properly accommodate metal grain sizes. We present one possible scenario in which biases could be introduced in sample preparation. The INAA techniques employed by

Kallemeyn and Wasson, which are described in detail in Kallemeyn et al. (1989), make use of

115

250-300 mg sample sizes. They note that, due to inclusion or omission of metal grains, Fe and siderophile abundances in ordinary chondrites may vary by 10% (Kallemeyn et al., 1989). Given large metal grains in ECs, this may influence results of ECs to an even greater extent. Hutson

(1996) posits that in studies performed on small EC samples, coarse metal grains are biased against in the sampling. In describing their method for sample preparation, Kallemeyn and

Wasson (1986) note that on the unsawn surfaces of their 250-300 mg chunks, they removed any visible ―rusty patches‖. These rusty patches may have originally been metal grains, and their removal (coupled with natural removal of surface metal grains due to weathering) may negatively influence total Fe. Since EL6 metal grains are larger than EH4-5, it is at least conceivable that more Fe was removed from EL than EH, exaggerating the difference.

The question of the true degree of Fe difference and where the difference resides can be better resolved through modal abundance studies utilizing large thin sections of representative sizes. None of these results should be taken to imply that there is no difference between the two groups of enstatite meteorites classified as EH and EL. That there are two distinct populations of enstatite chondrites was recognized by Anders (1964) long before the EH/EL nomenclature was adopted. Given their significant systematic trace element differences, it is almost certain that they even come from different parent bodies (Keil, 1989). However, the density and magnetic susceptibility trends now indicate that these differences are not related to a systematic difference in Fe or metal content.

116

5.4 Conclusions

Enstatite chondrites have not yet divulged all they have to tell us. We see in this study intriguing trends that call for further study of this class of meteorite. The counterintuitive weathering trend in bulk density and porosity of finds itself raises important questions, such as the underlying cause of the trend of increasing porosity with weathering and how, if the weathering is due to expansion of oxidized iron to form new cracks, the stone maintains its structural integrity and how far the process can continue before disintegration.

The similarity we observe in these properties, especially grain density and magnetic susceptibility, between EH and EL subgroups at the scale of bulk stones also calls for further study of iron as well. If there is a genuine difference in iron, why do we not see it? Is it fully accounted for by the presence of nonmetallic iron-bearing minerals such as troilite? Or is the discrepancy between our work and the literature due to sample bias for small fragments used in many of the past analyses?

Our homogeneity results present a word of caution for investigators. Many forms of analysis, such as neutron-activation analysis, provide high-precision results for very small samples, and so have become favored for such analyses, but when the meteorite in question exhibits variability at decagram scales, the possibility for bias in small samples cannot be ignored. We consider the need for representative sizes for future studies. One viable method is

SEM analysis of thin sections of surface area comparable to the cross-section of a typical decagram stone or larger.

117

CHAPTER 6: ACHONDRITES

Note: The contents of this chapter are taken from a paper that is in press (Macke et al.,

2010d)5 with minor modifications, including the inclusion of updated data, a minor expansion of

Section 6.2.1 to incorporate data from Apollo samples and their preliminary analysis, the removal of the abstract, the removal of the acknowledgments and measurement sections (which are redundant with information already provided in this dissertation), and typographical adjustments. Please see Chapter 2 of this dissertation for more detail regarding the measurement techniques.

6.1 Introduction

Stony achondritic meteorites constitute approximately 3% of all meteorites. Their parent bodies have undergone considerably more processing than those of the chondrites, but to different degrees. Some, including lunites, shergottites, nakhlites and chassignites (SNCs) and howardites, eucrites and diogenites (HEDs) come from the upper mantle and crust of differentiated parent bodies. Others including ureilites, brachinites, winonaites, acapulcoites and lodranites retain much of the basic chemistry and mineralogy of chondrites, indicating little if

5 Meteoritics & Planetary Science © 2010 by the Meteoritical Society. Used with permission (see Appendix C). 118 any differentiation in the parent body. Meteorites among most of the achondrite types, with the notable exception of HEDs, are rare on Earth, depriving us of good statistics in the analysis of any one group.

Because these meteorites originate on parent bodies that have undergone considerable processing or differentiation, they form a valuable piece of the puzzle that is our solar system.

They can provide clues as to how parent bodies form and differentiate, and can give some insight as to the similarities and differences between Earth and other planetary bodies in the solar system.

Much has been done to study the chemistry and mineralogy of achondrites, but relatively few have sought to systematically study physical properties such as density, magnetic susceptibility, and porosity. A group from the Geological Survey of Finland (Kukkonnen and

Pesonen, 1983; Terho et al., 1991, 1993; Pesonen et al., 1993) studied bulk density and magnetic properties for 368 meteorites of all classes from Finnish collections, including 19 achondrites.

Rochette et al. (2009) conducted an extensive study of magnetic susceptibilities for 291 different achondrites, including numerous Antarctic stones, though their study did not include density or porosity for the same specimens.

In order to determine porosity, two distinct density measurements are required: grain density and bulk density, both of which are determined by sample mass divided by a volume.

Grain density, or the density due solely to contributions from solid matter within the sample, uses only that portion of the sample volume that actually contains solid matter, omitting any pore space. Bulk density, on the other hand, uses the total volume enclosed by the exterior shape of the meteorite, including all pore space. Determining these two measures, especially bulk density, has historically required techniques such as liquid immersion (cf. Keil, 1962; Pesonen et

119 al., 1993) or slicing the sample into simple geometric shapes (cf. Yomogida and Matsui, 1981) that either damage or risk contaminating otherwise pristine samples. This had limited density measurements to a select few samples designated for the purpose, leaving the vast majority of meteoritic specimens unavailable for density surveys.

Consolmagno and Britt (1998) established a rapid, non-destructive and non- contaminating method for measuring bulk density by immersion in a ―fluid‖ composed of glass beads, and this in combination with helium ideal-gas methods for determining grain densities allowed large-scale surveys of density and porosity for hand-sized stones (Consolmagno et al.,

1998, 2006, 2008). Initial work focused primarily on ordinary and carbonaceous chondrites due to their ready availability in statistically-significant numbers, but the methods are just as applicable to other meteorite types.

Consolmagno et al. (2006) found that grain density in conjunction with magnetic susceptibility, both of which vary with metallic iron content, serve as a useful first-order classification scheme for ordinary chondrites into H, L and LL subgroups, while at the same time they identified a few mislabeled meteorites. This demonstrates the value of measuring physical properties, especially those such as grain density and magnetic susceptibility that depend on bulk mineralogy, when more detailed chemical analysis is not possible for a given sample either due to lack of time, availability of equipment or expertise, or due to the undesirability of sacrificing a portion of a given sample for the analysis. Intra-group and inter-group variability in these properties may be indicative of mineralogical differences. Variability in these properties among stones from the same meteorite may be indicative of inhomogeneities that may have been missed by chemical analysis of just one stone.

120

With the desire to explore further any trends that may be present in porosity or other physical properties for other meteorite types, we undertook a survey of density and porosity, as well as magnetic susceptibility, for stony meteorites of all types at major collections throughout the , as well as the Natural History Museum in London and the Vatican meteorite collection. The U.S. collections included the American Museum of Natural History, the

Smithsonian Institution, Arizona State University, University of New Mexico, and the Monnig collection at Texas Christian University. To date, our database includes 1228 stones from 664 meteorites, of which there are 201 stones from 116 achondrites, including five Apollo lunar samples. This provides us with sufficient numbers of stones for most achondritic subgroups to begin observing their different ranges, similarities and differences in each property. This paper is intended to serve primarily as a report of our data, with only first-order attempts at interpretation given.

6.2 Results

At least one physical property is reported for 201 stones from 116 meteorites, of which there is a complete set of measurements for 181 stones. The data are summarized in Table 18 through Table 21 and in Figure 76 through Figure 79.

121

6.2.1 Lunar Meteorites and Apollo Samples

Data for four lunar meteorites (all finds) and five Apollo samples are included in this study. The four meteorites are Dhofar 081, Northwest Africa (NWA) 482, NWA 773, and NWA

5000, all of which are lunar breccias. The Apollo samples include low-Ti basalts (12051,19 and

15555,62; Meyer, 2005, 2009c), one brecciated gabbroic anorthosite (15418,179; Meyer, 2008), and two crystalline-rich impact breccias from the rim of the Fra Mauro crater (14303,14 and

14321,220; Meyer, 2009a, 2009b). Because they represent a wide range of lithologies, lunar materials exhibit a considerable variation in physical properties (Table 18). Nevertheless, it is apparent (Figure 76 and Figure 77) that lunar materials tend to the low end in grain and bulk densities, reflecting the remarkably low density of the lunar crust and low total iron content.

Magnetic susceptibilities (Figure 79) are also low, but are similar to those of other materials from differentiated parent bodies such as SNCs and HEDs.

Lunar meteorites and Apollo samples in this study differ considerably in porosity.

Among lunar meteorites, porosities range from almost nil to about 11% (for NWA 773). On the other hand, Apollo samples range from about 2% up to 22%. The average meteorite porosity was 5%, while the average for Apollo was 12.7%. The low porosity of lunar meteorites compared with Apollo samples may indicate that the stress of the and terrestrial atmospheric entry selects for stronger materials; most higher-porosity materials may not survive the journey. If so, this serves as a word of caution for interpreting other meteorite data; there may be a sampling bias due to the stresses of removing rocks from parent bodies and delivering them to the Earth.

122

The preliminary data are based on a relatively small number of sources representing a range of lithologies. A much larger sampling is necessary to extract trends and differences among different types of lunar materials. Further data will also help confirm a difference in porosities between lunar return materials and lunar meteorites. The acquisition of further lunar data is the subject of ongoing research and extends beyond the limits of this dissertation.

6.2.2 SNC

Shergottites, nakhlites and chassignites (SNCs) are igneous meteorites, some volcanic and some of plutonic origin. SNCs were linked to a common parent body by O isotopic compositions that lie along a common fractionation line (Clayton and Mayeda, 1983). Based on their late crystallization ages (McSween et al., 1979; Wasson and Wetherill, 1979), isotopic ratios of trapped gases (Bogard and Johnson, 1983, Bogard et al., 1984) and other evidence it has become accepted that the common parent body of SNCs is the planet Mars. Nakhlites are clinopyroxenites predominantly composed of magnesian augite with some Fe-rich olivine, and chassignites are dunites dominated by Fe-rich olivine (Hutchison, 2004). Shergottites are a diverse group, including basaltic shergottites that in some cases represent surface or near-surface flows, lherzolites, and olivine-phyric basalts (Hutchison, 2004; Goodrich, 2002).

Several investigators have explored physical properties of Martian meteorites in recent years. Britt and Consolmagno (2003) summarized density and porosity data from the literature for 8 stones from 4 SNCs, and by the subsequent review (Consolmagno et al., 2008) the database had grown to 12 stones from 7 SNCs. Coulson et al. (2007) used SEM analysis to determine

123 porosity for numerous thin sections, supplementing the data with He-pycnometry-measured grain densities for 4 stones and bulk density using the glass bead method for 2 stones. Rochette et al.

(2001, 2005, 2009) measured magnetic properties (magnetic susceptibility and magnetic remanence) for numerous SNC samples.

In this study, we measured 16 stones from 10 SNC meteorites. This includes 9 stones from 5 shergottites, 6 stones from 4 nakhlites, and one chassignite stone (). These are about 15% of the 68 known non-Antarctic SNC’s (56 shergottites, all but 3 of which are numbered African finds; 6 nakhlites; 2 chassignites; and 4 ungrouped). Data are summarized in

Table 18. Grain density averaged 3.37 g cm-3, ranging from 3.08 g cm-3 to 3.73 g cm-3. Bulk density averaged 3.08 g cm-3, ranging from 2.83 g cm-3 to 3.48 g cm-3. Porosity ranged from 3% to 17%, averaging about 9%. Average magnetic susceptibility was log χ = 3.11, with a range from 2.79 to 3.68.

The individual groups (shergottite, nakhlite and chassignite) exhibit clear differences in density and magnetic susceptibility, though no clear differences are apparent in porosity. The differences are most apparent in magnetic susceptibility. Shergottites group tightly around an average magnetic susceptibility of log χ = 2.85, with a standard deviation of 0.05. One meteorite is excepted here: Los Angeles lies well outside the rest of the population with a magnetic susceptibility of 3.52 ± 0.12. Though Los Angeles is a basaltic shergottite, its mineralogy is unusual as being much more ferroan than other basaltic shergottites (Rubin et al., 2000). It also has an abundance of titanomagnetite and is the most magnetic of the Martian meteorites

(Rochette et al., 2001), so the high magnetic susceptibility is not unexpected. Nakhlites group somewhat more loosely than shergottites, with an average magnetic susceptibility of 3.40

124

(standard deviation 0.19), and the one chassignite in the study has a magnetic susceptibility of

2.98 ± 0.08.

Grain densities also differ, though shergottites and nakhlites are more similar in this property than in magnetic susceptibility. They have respective averages of 3.31 g cm-3 (standard deviation 0.07 g cm-3) and 3.42 g cm-3 (standard deviation 0.07 g cm-3). Again, Los Angeles was omitted from the shergottite average, since it is exceptionally low at 3.08 ± 0.02 g cm-3.

Chassigny exhibits a grain density (3.73 ± 0.04 g cm-3) that is much higher than that of the other two groups. It is also considerably higher than the typical density of terrestrial dunite (3.2-3.4 g cm-3) which tends to be dominated by magnesium-rich olivine (forsterite). The difference is probably due to higher percentage Fe-rich ferrisillite in the Chassigny olivine.

Bulk densities reflect somewhat the trends in grain densities, with shergottites and nakhlites similar (3.01, standard deviation 0.14; and 3.10, standard deviation 0.12, respectively), and Chassigny exhibiting a much higher bulk density of 3.48 ± 0.08 g cm-3. In this case, Los

Angeles was not exceptionally different from the other shergottites, and so was included in the population average. Because bulk density roughly trends with grain density for the groups, their porosities are similar. Shergottite porosities averaged 8.7%, with a standard deviation of 3.3%.

The average nakhlite porosity was 9.3% with a slightly larger standard deviation at 4.9%.

Chassigny’s porosity was 6.8% ± 2.3%, also consistent with the other two groups.

The three populations are clearly distinct on a plot of grain density and magnetic susceptibility (Figure 80), with shergottites and nakhlites near each other but clearly separate, and Chassigny notably distant from the other stones. Los Angeles’ anomalous grain density and magnetic susceptibility also are exhibited, with the meteorite plotting to the right of and below the rest of the shergottites. Curiously, the population of shergottites includes the basaltic

125 shergottites Shergotty, Zagami, and Los Angeles as well as the olivine-phyric shergottites Dar al

Gani 476 and Sayh al Uhaymir 005 (Goodrich, 2002). Lherzolitic shergottites are not represented, as all known examples are Antarctic finds.

Also apparent on the plot are the differences between falls (solid shapes) and finds (open shapes). At this point we caution the reader that the total number of stones represented in this analysis is small, so the low statistics should be taken into consideration. The differences between falls and finds are most apparent in grain density for shergottites (Los Angeles excepted) and nakhlites, with finds exhibiting higher grain densities than falls. Most of the iron present in SNCs is already heavily oxidized, so weathering will not further affect it in ways detectable in the physical properties included in this study. Secondary effects, such as the leaching of a low-density mineral component, may account for the unusual increase in density in the finds. We also note that the shergottite falls in this study are all basalts, while the finds are all olivine-phyric, so the difference in grain density between falls and finds for shergottites may be accounted for by the small mineralogic differences between the two groups rather than by weathering effects.

6.2.3 HED

A common parent body for howardites, eucrites and diogenites was speculated as far back as 1918 (Hutchison, 2004), and that parent body is currently believed to be the asteroid 4

Vesta (Consolmagno and Drake, 1977; Ruzicka et al., 1997). Diogenites are orthopyroxene cumulates, while eucrites are basalts, some of which are also cumulate. Eucrites and diogenites

126 may be unbrecciated, monomict or polymict, but howardites, being mixtures of eucritic and diogenitic material, are all polymict breccias (Bischoff et al., 2006; Hutchison, 2004).

We measured 108 stones from 58 HEDs, including 23 stones of 16 howardites, 65 stones of 31 eucrites, and 20 stones of 11 diogenites. Our data are summarized in Table 19. Average grain density for the entire set was 3.25 g cm-3, ranging from 2.96 g cm-3 to 3.51 g cm-3. Bulk density averaged 2.89 g cm-3 and ranged between 2.61 g cm-3 and 3.37 g cm-3. Porosity averaged

10.9%, ranging from zero to 20%. Magnetic susceptibility covered a large range, from 2.56 to

4.44, with an average of 3.12. These values are based on individual stones, but using meteorite averages the results do not differ significantly. Most of the range in magnetic susceptibility is due to eucrites, with howardites and diogenites having values from 2.62 to 3.67. Most eucrites are breccias (some polymict) and would thus be expected to exhibit more inhomogeneity than other types. The average magnetic susceptibility for eucrite stones (3.07) nearly matches that of diogenites (3.04), though both are lower than the average for howardites (3.34). Nevertheless, all populations overlap in magnetic susceptibility, and the range for howardites is entirely contained within the range for both diogenites and eucrites. Rochette et al (2009) also observed this trend in magnetic susceptibilities, attributing the higher average values for howardites to a metal enrichment during the process of regolith formation.

Though magnetic susceptibilities for eucrites vary widely, they occupy a tight range of grain densities, from 2.99 g cm-3 to 3.34 g cm-3. Eucrite grain densities are lower than diogenites

-3 -3 (Figure 76), which range from 3.36 g cm to 3.51 g cm . Diogenites are dominated by En67-85 (Hutchison, 2004), which at a density of 3.4-3.5 g cm-3 is consistent with the meteorite values. Howardites, consisting of a mixture of eucritic and diogenitic minerals,

127 naturally have grain densities that lie between and overlap the two other groups, though they more strongly overlap eucrites than diogenites. They lie between 2.96 g cm-3 and 3.36 g cm-3.

Bulk densities follow the same trend as grain densities, with diogenites higher in bulk density than eucrites and howardites lying between. One significant difference is that in bulk density the range of eucrites overlaps that of diogenites by quite a bit, a difference again accountable by variations in the structure of eucritic breccias. All three groups occupy similar ranges in porosities, averaging near 10% and ranging up to 20%, though the range for diogenites only reaches as high as 15% porosity.

Each of the three populations was well-represented with both falls and finds, with the lowest representation being diogenites with only 36% (4) finds. There is a small difference in average porosity between falls and finds in both howardites and diogenites, each being higher for falls by 5% and 3% respectively, comparable to their respective standard deviations. Grain densities of the falls and finds for each group are nearly identical (Figure 81), so the porosity difference is related to degree of compaction rather than any chemical change due to weathering.

Since the same result is not seen in eucrites, for which a much larger population is represented, this small porosity difference may be an artifact of sampling. We conclude from this that weathering processes do not significantly alter the physical properties of HEDs.

Eighteen of the 31 eucrites in our data are monomict breccias, but we also have represented seven polymict breccias, five cumulates, and two other eucrites (Agoult and NWA

2690). We compared the physical properties of each group to see if there was any visible effect due to brecciation. No difference in grain density is discernible, either between breccias and non-breccias or monomict and polymict breccias. Breccias exhibit lower average bulk density

(2.83 g cm-3) than non-breccias (2.92 g cm-3) due to a somewhat higher average porosity among

128 breccias (11.4%) than non-breccias (7.8%). We point out that the populations overlap considerably in these properties, with standard deviations in porosity of about 5% each.

Polymict breccias possess slightly higher average porosity (12.9%, standard deviation 4.0%) than monomict breccias (11.0%, standard deviation 4.5%). The only significant difference between brecciated and non-brecciated eucrites is found in magnetic susceptibility. In this case, monomict breccias, cumulates and ordinary eucrites each exhibit similar averages (overall, 2.98 with a standard deviation of 0.43). Polymict breccias, on the other hand, average 3.45 with a standard deviation of 0.47. We also note that magnetic susceptibilities of both groups of breccias extend to higher values than for non-breccias, which only range up to log χ = 3.3. Both polymict and monomict populations extend above log χ = 4, though the monomict range is more densely populated at low values than that of the polymict breccias. A modest enhancement in magnetic susceptibility for polymict breccias is also reported by Rochette et al (2009) and is given the same explanation as that for the enhancement found among howardites; namely the enhancement in metal during regolith-forming processes. While our average magnetic susceptibility for unbrecciated eucrites and cumulates is essentially the same as that of 2.93 ± 0.3 reported by

Rochette et al. (2009), our average for polymict eucrites exceeds the 3.03 ± 0.3 that they report; nevertheless, they may have omitted as ―outliers‖ some of the highest-valued meteorites such as

Camel Donga from their average.

6.2.4 Aubrites

Aubrites are 75-98 vol% FeO-free enstatite (Watters and Prinz, 1979; Hutchison, 2004) and may be related to enstatite chondrites (ECs), though they originated on separate parent

129 bodies (Keil, 1989). Like ECs, the most abundant metal is kamacite, though the total abundance of metal in aubrites is substantially less than that of their chondritic relatives, ranging from trace quantities to 0.7 vol% (Watters and Prinz, 1979; Mittlefehldt et al., 1998), as opposed to near 10 vol% (20-25 wt%) for ECs (Keil, 1968).

We measured 19 stones from 9 of the 22 known non-Antarctic aubrites, including 8 of the

9 falls. They are tightly grouped in grain densities, ranging from 3.11 g cm-3 to 3.29 g cm-3 , averaging 3.21 g cm-3 (with one outlier—a 4 g piece of Bishopville—at 3.44 g cm-3 ) consistent with a dominance of enstatite (density 3.1 g cm-3) mixed with smaller amounts of higher-density phases. Their bulk densities are not so tightly constrained, spanning 0.62 g cm-3 (2.53 g cm-3 to

3.15 g cm-3, average 2.90 g cm-3), thus yielding porosities ranging from 2% to 21.5% (average

9%). Magnetic susceptibilities also span a wide range of values, from log χ = 2.94 to 4.72

(average 3.58).

Unlike the other properties, we see considerable intrameteorite variability in magnetic susceptibility. For example, we measured five stones from Cumberland Falls. Between the highest and lowest values for grain density is only about a 2% difference, and only about 4% for bulk density. On the other hand, for magnetic susceptibility in log units the five stones range from 3.24 to 4.26, or a difference of 30% in log space (an absolute difference of a factor of 10) and well beyond measurement uncertainties. Rochette et al. (2009) also reported considerable variation in aubrite magnetic susceptibilities, with standard deviations among measurements for individual meteorites as large as 0.70 in log units, with most aubrite standard deviations in the

0.3-0.4 range. This is substantially higher than standard deviations of meteorites in other achondrite categories, most of which were lower than 0.2. Variations in the quantity of kamacite may easily account for this phenomenon. Kamacite is strongly magnetic, so small variations in

130 its abundance on the order of a few tenths vol% (effectively doubling or tripling the abundance of the mineral) may have disproportionate affects on magnetic susceptibility. At the same time, the low absolute abundance of the dense metal means that such small variations would not have equally visible effects on grain density (Figure 82).

Physical properties of aubrites are not at all comparable to the same properties of ECs

(Macke et al., 2010c and Chapter 5 of this dissertation). This is largely the result of different metal abundances. At an average of 3.66 g cm-3 for falls, enstatite chondrites have higher grain density. Likewise, magnetic susceptibility (averaging 5.47) is also much higher. These values are similar to those of H chondrites. More intriguing than density and magnetic susceptibility, however, is porosity. ECs have very low porosity, with less than 7% porosity for falls, while aubrites have porosities ranging upwards of 20%. This suggests a considerable difference in the conditions under which aubrites formed as opposed to ECs which enabled aubrites to remain largely uncompacted.

In our survey, we have one non-aubrite enstatite achondrite (Zakłodzie, a find). It is known to have much higher metal content than aubrites (Stepniewski et al., 2000), and consequently has much higher grain density and magnetic susceptibility (Figure 82). In fact, its grain density and magnetic susceptibility are similar to those of EC finds (Macke et al., 2010c).

This is consistent with speculation that Zakłodzie formed from melting of an enstatite chondritic precursor (Przylibski et al., 2005).

Of the aubrites in our survey, we have only one find (Shallowater). It exhibits no obvious effects of weathering on physical properties; even its magnetic susceptibility is higher than average for the population. Among the finds reported in Rochette et al. (2009), only one (LaPaz

Icefield 03719) has anomalously low magnetic susceptibility (1.96). The others are within the

131 normal range of magnetic susceptibilities for aubrites, including some of the higher values.

Inter-meteorite and intra-meteorite magnetic susceptibility variation swamps any effect that weathering may have had. From this we conclude that weathering of aubrites has negligible effect on these physical properties.

6.2.5 Angrites

Angrites are igneous basalts containing Ca-Al-Ti-rich pyroxene, Ca-rich olivine, and anorthitic plagioclase. (Weisberg et al., 2006) They are very rare, with only 14 non-Antarctic stones known, of which Angra dos Reis is the only fall. In part because of their rarity, we only have two angrite stones in our survey, one for NWA 4590 and the other for NWA 4801. They are similar in all properties measured. Grain densities were 3.48 g cm-3 and 3.37 g cm-3, respectively. Bulk densities were 3.24 g cm-3 and 3.18 g cm-3. Porosities were 7% and 6%, with uncertainties of order 1% each. Magnetic susceptibilities were 3.15 and 2.77, marking the largest difference between the two stones. These properties are comparable to those of diogenites (Figure 76, Figure 79 and Figure 82).

Rochette et al. (2009) report magnetic susceptibilities for 10 angrites, with most values ranging from 2.61 to 3.14, though they report an anomalously high value of 4.54 for the paired

NWA 3164-5167. For NWA 4590 and 4801, they report respective values of 2.82 (lower than our 3.15) and 3.14 (higher than our 2.77). Since their measurements were performed on different stones from different collections, this may reflect some degree of heterogeneity in magnetic susceptibilities. We rule out the possibility of accidentally switching samples; if the raw

132 measurement values are switched, the adjusted final values still do not compare well with

Rochette et al. (2009) and the difference between our two stones is exaggerated. In addition, we see higher grain density for the sample with higher magnetic susceptibility, consistent with expectations. We note that our sample of NWA 4801 is visibly weathered with a reddish-brown coating on the exterior, which we do not see on NWA 4590. Therefore, we do not discount the possibility that the discrepancy may be accounted for by variable amounts of weathering between the different samples.

6.2.6 Ureilites

159 non-Antarctic ureilites are known, of which only six are falls. Ureilites are composed of olivine and pyroxene embedded in a carbon-rich matrix (Hutchison, 2004; Weisberg et al.,

2006) Their oxygen isotopes fall along a mixing line common to CM and CV carbonaceous chondrites (Clayton et al., 1976; Clayton and Mayeda 1988) rather than a mass-dependent fractionation line, indicative of a low degree of processing (Goodrich, 1992). They may in fact be related to the CVs, possibly originating from CV-like precursors (Rubin, 1988).

Our study includes 20 stones from 11 ureilites, including one fall (Novo-Urei). These range in grain density from 3.25 to 3.53 g cm-3 (average 3.36 g cm-3), and in bulk density from

3.04 to 3.36 g cm-3 (average 3.22 g cm-3). This group possesses relatively low porosities that range from zero to about 12.5%, with a low average value (4%) and with most stones below 8%.

Magnetic susceptibility is higher for ureilites as a group than for the groups discussed so far, and ranges over almost one and a half orders of magnitude, with the minimum value of 3.93 and a

133 maximum 5.28 (average 4.60). This range overlaps that of all of the ordinary chondrite groups

(see Figure 79).

Our average magnetic susceptibility is a little higher than the average of log χ = 4.39 reported by Rochette et al (2009), though the range of their data was larger, with values as low as

3.61 and as high as 5.32. Considering only those meteorites for which our data sets overlap, their average (4.54) and range (4.06 to 5.22) are more consistent with values reported here.

Meteorite-by-meteorite comparison yields only one notable inconsistency: all three of the stones of Nova 001 we measured are more than 0.1 higher than the 4.14 average reported by Rochette et al. (2009).

Curiously, the ureilites as a group are comparable in grain density and magnetic susceptibility to a particular subgroup of CV carbonaceous chondrites that originated in reducing environments, though not that of Allende, which originated in oxidizing environments (Figure

82), though grain densities of reduced CVs on average are slightly higher than that of ureilites.

Structurally, the low porosity of ureilites is also reflected in reduced CVs, though not in oxidized

CVs which have porosities on the order of 20%. This suggests that, if there is indeed an original relationship between ureilites and CVs, they are likely more closely related to those that originated in a reducing environment.

Due to the fact that ureilites have higher iron content and correspondingly higher magnetic susceptibility, it stands to reason that they should exhibit weathering effects comparable to that of ordinary chondrites. We have only one fall represented in our study

(Novo-Urei), though it is represented by four stones. Its magnetic susceptibility (4.96) is above average for the population, as is its grain density (3.43). Its porosity, averaging 7.4%, is slightly above the average for the population. This indicates that the other ureilites may be weathered

134 somewhat, though the low statistics of falls in the population prevent us from speaking conclusively. We note that Goalpara, a find, exceeds Novo-Urei in grain density, magnetic susceptibility and porosity. Rochette et al. (2009) add the falls Haverö and Dyalpur. All of their falls have magnetic susceptibility values (averaged per meteorite) of at least 4.90, increasing the likelihood that values substantially below 4.9 in finds may result from weathering. In the event that weathering of finds does have an effect on ureilite physical properties, it should still not greatly affect the comparison with reduced CV carbonaceous chondrites, since most of the CVs in the comparison were also finds and should have been affected in the same manner.

This dissertation does not contain measurements of Almahata Sitta, the ureilite that fell recently in the Sudan after being tracked in space as asteroid 2008 TC3. This meteorite is an anomalous polymict ureilite that is quite heterogeneous (Jenniskens et al., 2009) and has been reported to contain stones that classify as enstatite chondrites and ordinary chondrites

(Horstmann and Bischoff, 2010; Bischoff et al., 2010). Some preliminary physical property measurements have been performed (Kohout et al., forthcoming), but because of its unusual and heterogeneous nature, we consider it a priority to study numerous stones from the fall.

6.2.7 Acapulcoites and Lodranites

Acapulcoites and lodranites are granular-textured primitive achondrites with mineralogy similar to that of H chondrites, including an abundance of metal (Hutchison, 2004). We measured ten stones from eight meteorites in this group (six stones from four acapulcoites and four stones from three lodranites), including representatives from the only two known falls,

Acapulco and Lodran. The acapulcoites are more tightly grouped than the lodranites, with grain

135 density averaging 3.69 g cm-3 (range from 3.55 to 3.88 g cm-3), bulk density averaging 3.46 g cm-3 (range from 3.33 to 3.59 g cm-3), and magnetic susceptibility averaging 5.20 (from 4.99 to

5.51). These values are comparable to those of H and L ordinary chondrites (Figure 83) and reflect the high iron content of these primitive achondrites. Acapulcoite porosities ranged from

2% to 12%. On the other hand, lodranite grain densities ranged from 3.38 to 4.16 g cm-3, bulk densities from 3.24 to 3.82 g cm-3, magnetic susceptibilities from 4.74 to 5.68, and porosities from 0 to 9%. Their averages (grain density 3.74 g cm-3, bulk density 3.53 g cm-3, magnetic susceptibility 5.24 g cm-3, and porosity 5.4%) are all within the range occupied by acapulcoites.

The large spread in lodranite physical properties is likely related to the considerable variation in mineralogy, including metal (0.5-20%; Hutchison, 2004) between meteorites in this group. It is possible that some of that variation may be the result of weathering processes.

While acapulcoites show no significant difference between falls and finds in any of the properties measured (cf. Figure 83), the two lodranite falls (both stones from Lodran) are much higher in all four properties than the two finds. Grain density, magnetic susceptibility and porosity trends are all consistent with weathering patterns expected for meteorites with iron quantities comparable to ordinary chondrites (Bland et al., 2006).

6.2.8 Other Primitive Achondrites: Brachinites and Winonaites

Brachinites are olivine-rich meteorites with very little to no Ca-poor pyroxene, and are somewhat depleted in metal-sulfide compared with ordinary chondrites. Their origins are uncertain and may vary by meteorite (Mittlefehldt et al., 1998), though one possible origin is the

136 partial melting of a CI-like precursor (Nehru et al., 1996). Winonaites are similar in composition and texture to silicate inclusions in IAB iron meteorites, indicating that winonaites originated as

IAB inclusions (Hutchison, 2004). In fact, the winonaite Mt. Morris () may be paired with the IAB iron Pine River (Bevan and Grady, 1988). Winonaites likely originate from a parent body that partially differentiated above ~1000 C, may have gone through a stage of break- up and reassembly to facilitate mixing of silicates and metal, and cooled slowly (Benedix et al.,

2000). Both brachinites and winonaites have approximately ordinary chondrite composition.

From this set we measured 15 stones from 8 meteorites. This includes 8 stones from 4 brachinites, 6 stones from 3 winonaites, and one stone of an ungrouped .

Only 14 non-Antarctic brachinites are known, and of winonaites only 13 are known, of which

Pontlyfni is the only observed fall. Brachinite averages were: grain density, 3.55 g cm-3

(standard deviation 0.06 g cm-3); bulk density, 3.48 g cm-3 (standard deviation 0.26 g cm-3); and magnetic susceptibility 4.06 (standard deviation 0.19). All stones had porosities below 7% except Eagle’s Nest, which was 15% porous. All of the brachinites in this study were finds, making an analysis of weathering impossible. Brachinite densities and magnetic susceptibilities were comparable to those of LL ordinary chondrites (Figure 84).

Winonaites exhibited on average higher grain density (3.60 g cm-3, standard deviation

0.25); and magnetic susceptibility (4.90, standard deviation 0.40) than brachinites, though the average bulk density (3.24 g cm-3, standard deviation 0.19) was lower than that of brachinites.

This resulted in porosities that were considerably higher. Porosity ranged from 4 to 13%, averaging roughly 8%. Winonaite physical properties were roughly consistent with that of H and

L ordinary chondrites (Figure 84), as well as acapulcoites and lodranites. The one winonaite fall in the study (Pontlyfni) had higher grain density, magnetic susceptibility and porosity than the

137 finds, indicating that the finds were subject to a possible weathering effect similar to the possible effect seen in lodranites, though as with lodranites the lack of a statistically significant number of falls means we cannot rule out the possibility that Pontlyfni simply originated with more metallic iron than the others. The normal bulk density and high porosity of Pontlyfni compared to winonaite finds does not appear to support this hypothesis, but rather provides support for the conclusion that winonaites weather in a manner similar to ordinary chondrites.

6.3 Conclusions

The achondrites in this study can be subdivided into two categories: primitive achondrites that retain most of the chemistry of their chondritic precursors but nevertheless have been substantially processed, and other achondrites, most of which originate from the upper mantle and crust of differentiated parent bodies.

Grain densities and magnetic susceptibilities of the primitive achondrites such as acapulcoites, lodranites, brachinites and winonaites compare favorably with those of ordinary chondrites (and ureilites compare well with CVr). Even so, different types within this category exhibit differences in averages and ranges for these values that are indicative of different mineralogies. These differences are not enough to classify stones based on their physical properties, as is the case with OC falls (Consolmagno et al., 2006), but the technique may be used to eliminate incompatible meteorite types without removal of a portion of the stone, as is required in standard classification. For example, an achondrite with grain density and magnetic susceptibility in the H-chondrite range may be a winonaite, acapulcoite, or possibly lodranite, but

138 is unlikely to be a ureilite or brachinite. This distinction may be especially useful for ―triage‖ when confronted with a large shower from a particularly inhomogeneous fall such as Almahata

Sitta. In the case of Almahata Sitta, grain density and magnetic susceptibility should enable rapid identification of those stones which are not ureilitic, since H chondrites as well as enstatite chondrites will have higher grain density and magnetic susceptibility than ureilite stones. L chondrites will also have higher density, though magnetic susceptibility will be consistent with ureilites.

Grain densities and magnetic susceptibilities for the achondrites originating from differentiated parent bodies, such as lunites, HEDs, SNCs, and angrites, are on average lower than those of ordinary chondrites or primitive achondrites, primarily due to a much lower quantity of metallic iron. This difference is most pronounced in magnetic susceptibility. Due to the considerable overlap in physical properties between different meteorite types, they do not form a useful classification tool for this category, though as with the primitive achondrites they may be used to eliminate certain meteorite types. Also, if a stone is known to be an HED and if the grain density is at the high end or low end of the HED spectrum, this may establish whether it is a eucrite or diogenite, though intermediate densities will remain ambiguous. Howardites cannot be identified solely on the basis of grain density and magnetic susceptibility. If a stone is known to be Martian, on the other hand, density and magnetic susceptibility together may be used to distinguish shergottites, nakhlites and chassignites.

With regard to bulk density and porosity, trends are less apparent or informative. Bulk density is in part influenced by grain density, and so roughly follows the trends reported for that property above. However, it is also dependent on structure, and porosity is fully dependent on structure. Porosity shows no clear trend or any notable difference between primitive and other

139 achondrites as a whole. Indeed, variations in porosity within a given meteorite type are far greater than differences on average between any two groups. Achondrites result from considerable physical and chemical processing, which will destroy any primitive structure. With some imagination, one may see minute trends within stones from differentiated parent bodies such as SNCs and HEDs. Nakhlites appear less porous on average than shergottites, and chassignites even less so, and diogenites appear less porous than eucrites. This may be indicative of greater compression of those materials formed deeper within the parent body. However, we caution that the differences are small and may not be genuine.

Due to the rarity of most achondrites and the present dearth of samples from many groups, this study will not remain the final word on their physical properties. With luck, more stones will fall or be discovered and will be made available for study, thus helping to provide better population statistics.

140

CHAPTER 7: SUMMARY AND CONCLUSIONS

The wealth of information produced by a study of this sort (summarized in Table 22) presents an opportunity to explore trends and relationships not merely within meteorite groups, but between them as well. We can begin to draw some conclusions, but at the same time the new data points to other areas that require further study. This study is a start, but not the end point.

In this chapter, I will examine the various scientific questions presented in the introduction, and will discuss what we have learned and where there is need for more research. In the last section

(7.5), I will examine the overall trends among all chondrite falls and what these may indicate about the formation of their parent bodies and possibly about conditions within the solar nebula.

7.1 Weathering

Since the majority of recovered meteorites are finds, many of which exhibit substantial amounts of terrestrial weathering, it is important to understand how weathering affects physical properties and, insofar as it may be possible, to account or correct for it. This study indicates that weathering varies by meteorite type. This is understandable, since the chemical reactions that take place during the weathering process are largely dependent on the mineral species present. Achondrites from differentiated parent bodies, including SNCs and HEDs, exhibit

141 minimal effects due to weathering on density, porosity or magnetic susceptibility. This is explained by the low abundance of metal. That being said, shergottites and nakhlites may have a slight increase in grain density and/or magnetic susceptibility as a result of weathering, possibly due to the loss of a low-density nonmetallic phase, though more research is necessary to establish the veracity of this effect. Among primitive achondrites, the lack of statistically significant numbers of both falls and finds for most populations prevents conclusive claims regarding weathering patterns. For the more metal-rich meteorite types such as ureilites, finds appear to have lower grain density, magnetic susceptibility and porosity than falls, a trend that is consistent with the effect on ordinary chondrites as will be discussed later.

For carbonaceous chondrites, the issue is complicated by the low numbers of finds and the fact that each group of carbonaceous chondrites differs widely from others in physical properties, so each population has to be treated separately. There is no observable difference between falls and finds for most groups, with one notable exception. CO finds have strongly reduced grain densities and much lower porosities than CO falls. This result is not new; it was reported for much lower numbers of samples in Consolmagno et al. (2008) and is thought to be caused by the expansion of weathering-related carbonates into pore space. It should be noted that CO is the best represented carbonaceous chondrite group in the data, with 14 different meteorites of which six are falls and eight are finds. This, along with the tremendous degree to which weathering affects this particular type, is why we can draw conclusions about COs but not about other CC groups. One thing is clear; if weathering does affect other CC groups, the effect is much smaller than for CO, and so this calls for further research and more data. We do have better statistics for the other groups than were present in other studies, but this is still insufficient for a weathering analysis.

142

The effect of weathering on ordinary chondrites, however, has been well understood for some time (Bland et al., 1996, 1998b). It thus came as no surprise that the OC results in this study agreed with those of other studies; OC finds exhibit reduced grain density, porosity and magnetic susceptibility as a result of oxidation of metal and its subsequent expansion into existing pore space. While it had been hypothesized that this effect may be used to quantify the degree of weathering, the closest approach to this prior to this study is Rochette et al.’s (2003) attempt to quantify weathering according to magnetic susceptibility. Weathering can be even better quantified by considering both grain density and magnetic susceptibility, which I have done in the construction of a weathering modulus.

Using the weathering modulus, I have been able to explore correlations of degree of weathering on bulk density and porosity. An unexpected result from this analysis is that bulk density exhibits a negative correlation with degree of weathering. Conversely, model porosities constructed from bulk densities correlate positively with weathering, indicating that the most weathered stones are those that started with the highest porosities. In retrospect, this result makes sense, since the higher-porosity stones have more space into which the weathering products can expand, but it was nevertheless unexpected.

Enstatite chondrites produced the most unexpected weathering results by far. As with ordinary chondrites, bulk densities for ECs correlated negatively with degree of weathering (as determined by the weathering modulus), but unlike OCs, EC porosities (not just model porosities) exhibited a positive correlation with weathering. Somehow, the more weathered the stone, the more pore space was created as a result of weathering. This raises the question of its cause; is it the introduction of new cracks in a stone that has very little original pore space into which weathering product may expand, or is the weathering process causing excessive leaching

143 of some mineral phase? If the latter, what phase is it? Further detailed analyses of sections of weathered ECs may reveal the answer to this question.

7.2 Homogeneity

In order to better understand how indicative the data in this study are for the materials making up the parent bodies from which meteorites come, it is important to have some grasp of the homogeneity of the populations. Two questions are implicit here: first, how homogeneous are meteorite groups? That is, how similar is one meteorite from a population to another from the same population? Second, how homogeneous are stones from individual meteorites? The former question focuses on entire populations, while the latter focuses on the homogeneity of the original source asteroid from which an individual meteorite fell. On top of that is the question of scale-dependent homogeneity; some meteorites contain clasts of varying density or metal content that may strongly affect results if the size of the sample is too small. The question here is, what scale is necessary for a representative sample? To avoid further complication due to weathering- related effects, finds are omitted from consideration in this discussion.

For the question of population homogeneity, the answer varies by meteorite type and by which physical property is studied. Ordinary chondrite falls (in particular H and L) are tightly constrained in grain density and magnetic susceptibility, though they exhibit considerable variation in porosity. This suggests mineralogical heterogeneity is low, but structural variation does occur. The key factor here is likely shock; more strongly shocked meteorites will have substantially reduced porosity, though their grain density and magnetic susceptibility are not

144 strongly affected. With enstatite chondrites, the population as a whole is tightly constrained in grain density/magnetic susceptibility space, but with somewhat more variability than ordinary chondrites. EC falls, however, are much coarser mineralogically and exhibit scale-dependent variations in grain density and magnetic susceptibility below masses of ~40-60 g, a feature that is not observed in ordinary chondrites. In contrast, their porosities vary (by population) much less than OCs, which may be attributed to the fact that ECs as a whole have very low porosities to begin with.

For carbonaceous chondrites, this question depends on the individual subgroups, with the most tightly constrained being CK, for which both falls and finds reside in grain density/magnetic susceptibility space between H and L OC populations. Porosity varies more for the population, but if finds are omitted only one meteorite (Karoonda) remains. The CV group occupies the other end of the homogeneity spectrum. The CVr subgroup, with low porosity, moderately high magnetic susceptibility and low grain density, is more tightly constrained than the CVo subgroup, whose magnetic susceptibility range covers almost two orders of magnitude.

Among achondrites, the paucity of falls limits the possibility for drawing strong conclusions for most groups, with the exception of HEDs for which there are ample numbers of samples for this purpose. Each of the subgroups (howardites, eucrites, and diogenites) are tightly constrained in grain density, but because of their low average metal content the low magnetic susceptibility values are particularly sensitive to small variations in metal and other magnetic materials and so varies more than grain density. Taking advantage of this fact, we can see a greater degree of variation among eucrite breccias than among non-brecciated eucrites, howardites or diogenites. Among SNCs, shergottites and nakhlites may be tightly constrained in grain density and magnetic susceptibility, but the different lithologies vary between each other.

145

Also, the unusual shergottite Los Angeles provides a note of caution when drawing conclusions; while a population as a whole may be homogeneous, one must always allow for the presence of outliers. Lunar materials, on the other hand, are quite heterogeneous in all properties, including porosity. This is in part because they sample very different lithologies from different regions on the lunar surface. If anything, this only underscores the need to study very large numbers of lunar materials to obtain representative data for each of the major lunar lithologies.

Regarding the question of intra-meteorite homogeneity, the studies of large numbers of stones from certain ordinary and carbonaceous chondrites, including Holbrook, Gold Basin,

Pultusk, Allende, Murchison, Murray, and Karoonda, allow for better understanding of the decimeter-scale homogeneity of the parent asteroids from which they came. By and large, stones from these meteorites varied less than four percent in all of the physical properties measured except porosity (for which small variations in density became exaggerated), indicating very little heterogeneity. Most telling of all is the fact that stones of Allende from two collections exhibited measurable differences in all four properties, indicating that Allende is homogeneous to the degree that small environmental influences became significant in the analysis.

While ordinary chondrite and carbonaceous chondrite parent bodies are largely homogeneous on the decimeter scale, the same cannot be necessarily said about other kinds of meteorites, although more study is needed in order to establish their degrees of homogeneity. In large part, the lack of study here so far is due to the lack of large numbers of stones from individual meteorites from these groups. One telling point is the observed scale-dependent intra- meteorite heterogeneity of enstatite chondrites, which begs for further study of large stones.

Also, among achondrites, the vast range of magnetic susceptibilities covered by just a few stones of the aubrite Cumberland Falls is undeniable. Another tantalizing object for future study is

146

Almahata Sitta, the meteorite remnants of asteroid 2008-TC3 that fell in the Sudan in 2008. As mentioned in Chapter 6, this object is known to be quite heterogeneous, and is currently understood to be an aggregation of materials from ureilite, ordinary chondrite and enstatite chondrite parent bodies (Jenniskens, 2009; Horstmann and Bischoff, 2010; Bischoff et al., 2010).

It may yield clues to processes by which parent bodies break up and recombine.

7.3 Grain Density and Magnetic Susceptibility as a Classification Tool

Since Consolmagno et al. (2004, 2006) demonstrated that the use of grain density and magnetic susceptibility together serves as a good first-order classification tool for ordinary chondrite falls, this naturally led to the question of whether the technique could be applied to other meteorite types. The results here were quite mixed. Most notably, the two classes of enstatite chondrites (EH and EL) were completely indistinguishable in either grain density or magnetic susceptibility. While it rendered null the possibility of using the technique for classification, this result did provide solid grounds for examining the question of whether the two types of enstatite chondrite have any difference in total iron and/or metal content.

Other groups for which the technique may be useful are SNCs and HEDs, though with caveats for each. For HEDs, the three constituent groups (howardites, eucrites and diogenites) overlap in magnetic susceptibility, but each occupies a tightly-constrained range in grain densities. In this sense, eucrites and diogenites are distinguishable from each other, but howardites complicate the question by overlapping both, especially eucrites. Eucrite or diogenite stones whose grain densities or magnetic susceptibilities fall outside the howardite region are

147 distinguishable, but otherwise these physical properties are not useful. With SNCs, shergottites, nakhlites and the one chassignite in the study each (excepting Los Angeles) occupied distinct regions in grain density/magnetic susceptibility space. The number of SNCs surveyed was relatively low, so it is necessary to increase the statistics before drawing solid conclusions. Of course, given the rarity of Martian meteorites, it may prove impossible to substantially expand the database.

Regardless of the usefulness of this technique for classification, one thing that is necessary a priori is some knowledge of the general type of meteorite. Many different meteorite types occupy overlapping regions in grain density/magnetic susceptibility space. For example, low-metal meteorites such as SNCs, HEDs, angrites, and lunar meteorites each overlap to varying degrees. Intermediate and high-metal meteorites such as carbonaceous chondrites, ordinary chondrites and primitive achondrites also overlap. One cannot, for instance, distinguish a CH from an H chondrite on the basis of grain density and magnetic susceptibility alone.

Nevertheless, the technique remains useful as a negative indicator. Each meteorite type occupies a distinct region in grain density/magnetic susceptibility space, although some of the regions are quite large. A meteorite whose physical properties lie substantially outside the region occupied by a given meteorite type can be ruled out as a candidate for that type. In one example, these measurements were performed on an unclassified meteorite at one of the museum collections. (Since it has not been classified, it is not included among the meteorites reported in this study.) This meteorite had an unusual texture that caused the collections manager to suspect it might belong to the SNCs, but its magnetic susceptibility of 5.3 ruled it out as a . With a grain density of 3.6 g cm-3, the meteorite may have been an H chondrite, an

148 acapulcoite, a CH chondrite, or one of a number of other meteorite types consistent with those measurements, but it could not have been an SNC.

7.4 Mesosiderites and Iron Meteorites

Not included in the analysis of this dissertation were the mesosiderites and iron meteorites. A few meteorites of these types were analyzed in the process of the research, and for completeness have been included in Table 23. Nevertheless, two primary issues related to the very high metal content of these groups prevented more detailed study. First, the samples were very high in density (5 to 8 g cm-3), which had an unpredictable effect on the settling behavior of the glass beads when they were immersed. This in turn caused bulk density measurements to have large uncertainties and to be unreliable, which also meant unreliable porosities. Second, due to the excess metal the response of the magnetic susceptibility meter was in a highly nonlinear regime, effectively maxed out. No meaningful comparison of magnetic susceptibilities could be made.

There is room for further analysis of mesosiderites and iron meteorites in the future. In particular, grain density may be a useful analytical tool. It may prove useful for determining the metal-to-silicate ratio in falls, or could be used on iron finds to establish degree of weathering.

149

7.5 Overall Trends from Chondrite Falls

Within Chapters 3, 4 and 5 of this dissertation, the three major types of chondrites

(ordinary, carbonaceous and enstatite) have been treated separately. In large part, this is to be expected, as their different compositions and formation histories influence their densities and magnetic susceptibilities in quite different ways. Nevertheless, there is something to be said for considering all chondrites together as a single group. In particular, all chondrites are to some degree remnants of the early planet-forming processes in the solar nebula. By comparing them with each other, perhaps some clues to conditions within various parts of the solar nebula may stand out.

7.5.1 Shock-Related Trends

In relation to ordinary chondrites (Chapter 3), it was apparent that shock is one of the major influences on porosity, as shock compresses material and causes transport of metals into existing pore space. At the same time, new pore space in the form of cracks is created, albeit not to the same extent as the pre-existing porosity is destroyed. This pattern holds true for all chondrites (Figure 85 and Table 24) as well. It also presents a word of caution for examining trends related to other characteristics such as petrographic type; porosities will naturally scatter and tend lower due to varying shock histories. A large spread in porosities for these other characteristics should then be expected.

150

Grain density, on the other hand, does not appear to exhibit any shock-related trends

(Figure 86). There is a decrease on average for S1 compared to other groups, but this is accounted for by the disproportionate number of S1 carbonaceous chondrites like CM with very low grain densities; the groups are not evenly distributed among all shock states. The lack of a general trend conforms to expectations, since shock should not significantly alter the overall mineral composition of the meteorite except in extreme cases where it induces thermal metamorphism. The same goes for magnetic susceptibility (Figure 87), though low-shock (S1) meteorites are somewhat lower on average than higher-shock (S2-6) meteorites. Again, this may be the result of the carbonaceous-chondrite bias. Lesser number of carbonaceous chondrites serve as low-χ ―outliers‖ for the S2 and S3 populations, as can be seen in Figure 87.

7.5.2 Trends by Petrographic Type

In Chapter 3, it was noted that L chondrite falls exhibit a previously-unobserved porosity trend from petrographic type 3 to 6. However, when all chondrite falls are included (Table 25 and Figure 88), the trend reverses, going from highest porosity at low petrographic types (types 1 and 2) to lowest porosity at high petrographic types (5 and 6). This is a trend that ranges from extremely aqueously altered to extremely thermally processed, and gives some indication that porosity may be affected by alteration history or more likely by the localized conditions under which the parent body formed and which subsequently led to aqueous or thermal alteration. This will be discussed further in Section 7.5.4.

151

In terms of other physical properties, both bulk density and grain density (Figure 89 and

Figure 90) are substantially lower for petrographic types 1 and 2 (all of which are carbonaceous chondrites) than for the other groups, but average densities tend to level out for the thermally- processed petrographic types 3-6. With magnetic susceptibility (Figure 91), each petrographic type differs from the other, but there is no specific trend.

7.5.3 Trends by Matrix Abundance and Degree of Oxidation

The true value of the data in this work may well be found when it is considered not by itself, but in the context of the extensive work performed by other investigators. In particular, studies involving processed samples such as thin sections provide supplementary data with which physical properties, especially porosity, may be compared. To illustrate this further, I have selected two properties: matrix abundance and oxidation state (expressed as a ratio of oxidized Fe [Feox] to Feox + MgO, where Feox includes contributions from both FeO and Fe2O3).

These data are not readily available for all meteorites in this study, and so for the purposes of this analysis it is limited to averages by meteorite type. Average matrix abundances were taken from

Brearley and Jones (1998) and average Feox/(Feox+MgO) was taken from Wasson and Kallemeyn

(1988). Neither source made the distinction between oxidized and reduced CV. In addition,

Wasson and Kallemeyn (1988) do not include data for CK, CR or K.

The reason for exploring porosity as a function of oxidation state originated as a basic observation: oxidized CV meteorites in this study are substantially more porous than reduced

CV, and enstatite chondrites, which are the most reduced chondrite group, also have extremely

152 low porosities. When porosities are compared with their oxidation states (Figure 92), this basic observation points to a trend by which porosity correlates to degree of oxidation. The most reduced (EH and EL) are also the least porous, and the most heavily oxidized (CI) are the most porous.

Comparing porosity to matrix abundance is also logical; after all, much of the pore space may exist as intergranular spaces within the matrix. Porosity does indeed correlate positively with percentage matrix (Figure 93) in an approximately linear manner. This indicates that, shock effects notwithstanding, porosity increases as percentage matrix increases to a maximum value of approximately 35% at 100% matrix. If one assumes that matrix grains can be approximated by spheres of uniform size, optimal packing of the grains would result in only 26% porosity. On the other hand, random packing of uniform spheres in shaken containers does produce approximately 36% porosity (Torquato et al., 2000). Therefore one may infer from the data that matrix is assembled more or less randomly, and serves as the key contributor to meteorite porosity. This result, of course, is only preliminary. Matrix grains are neither uniform in size nor in shape. A substantial variation in grain sizes allows small grains to fill pore spaces left by larger grains; therefore, grain size variation will reduce total pore space. Irregular shapes, on the other hand, may affect porosity either positively or negatively, depending on the manner in which they are packed. If irregular edges jam into each other, then they will tend to leave larger pore spaces open, but the irregular shapes may also manage to fit into gaps that would otherwise have been left open, resulting in reduced porosity.

153

7.5.4 Implications for the Solar Nebula

The three observed porosity trends reported in this section (petrographic type, matrix abundance and oxidation) are further remarkable when one considers the relationships between these properties and their further relationships to the solar nebula. The most aqueously altered meteorites (CMs and CIs) are also among the highest in matrix abundance and also are the most heavily oxidized. CIs are at the extreme end here, and the one sample of a CI (Orgueil) included in this study has one of the highest porosities of any meteorite that was measured. At the other extreme are enstatite chondrites, which are among the most reduced chondrites, have low matrix abundance, and ELs tend to be strongly thermally equilibrated. They have the lowest average porosities of any of the chondrite groups in this study. Ordinary chondrites lie between the extremes in terms of oxidation and matrix abundance, though they admittedly are well represented in all petrographic types for thermally equilibrated meteorites from 3 to 6 (and some

7). Unsurprisingly, they also have intermediate porosities.

The overall porosity trend also follows some models for the locations of parent bodies in the early solar system, with EH and EL parent bodies forming closer to the Sun, ordinary chondrite parent bodies at intermediate distances (nearer Earth), and carbonaceous chondrites somewhat further out. Hutchison (2004) further connects this model with the availability of water in different regions of the solar nebula, from almost no available water near the Sun to much greater amounts available further out. In the range of CI’s, water availability reaches higher than 10 wt% (about 30 vol%, close to the measured CI porosity). Using this, it may be possible to extract from this study a few implications for conditions of formation within the solar nebula.

154

Before proceeding, it is important to introduce a caveat. Researchers have been tackling questions about the early solar system and the formation of early planetesimals and meteorite parent bodies, and have almost as many theories as there are people exploring the question. It would be extremely naïve to assume that the addition of a relatively small datum to the existing body of knowledge would tell the tale conclusively, but it may point to some questions and ideas worth pursuing.

In this case, one thing worth pursuing is the role of water in the lithification of early solar system materials. It is possible (T. McCoy, personal communication) that matrix grains were encased in water ice, especially farther from the Sun. It is within the realm of possibility that the water may have played a role in the binding of grains together. The departure of water through sublimation or vaporization at some time after the formation of parent-body precursor material would leave pore space. Prior to vaporization, aqueous alteration could take place in those meteorite precursors where water was abundant enough. Water may have also provided the oxygen necessary for increased oxidation of Fe.

This hypothesis appears to adequately handle aqueously-altered carbonaceous chondrites, but leaves a few questions remaining, especially for the most reduced materials such as enstatite chondrites. If water was key to aggregation of materials prior to lithification, how did such reduced meteorites lithify? This indicates that, while water may have been important, it is not the whole story. Other factors, such as van der Waals forces, must not be discounted in modeling the formation of meteorite parent body precursors.

Whatever model is produced, this study provides a key constraint related to porosity.

Objects forming near the Sun had low porosity, and the further out they were, the higher their

155 porosity. Any reasonable model of the formation of bodies from the solar nebula must also match that characteristic.

It is interesting to note that the microporosity trend in meteorites is matched by a similar macroporosity trend in asteroids from about 2-4 AU; those closer to the Sun have low macroporosities while those farther out have higher macroporosities (Consolmagno et al., 2008;

Consolmagno and Britt, 2008). While the former can be described by the lithification of individual stones, the latter is a result of larger-scale processes that bring existing stones together to form asteroids. The fact that both follow a similar trend is not to be expected a priori, but suggests that there may be a link between the two processes.

7.6 Conclusion

This study is the most comprehensive survey of meteorite density, porosity and magnetic susceptibility performed to date, and offers the further advantage of consistent measurement techniques and equipment across meteorites from a wide range of collections. With these data, it has been possible to corroborate trends already observed in the literature, to discover some new trends, and in a few cases to contradict pre-existing observations that were based on less extensive observations.

While the analyses that have been presented in this dissertation are in themselves scientifically valuable, the true value of the research is the production of a database that will be made available to other investigators. The process of mining the database and comparing these

156 data with different measurements performed by other researchers promises to be fruitful for years to come.

157

APPENDIX A: FIGURES

158

Vs P1

V1

Vs P2

V2

As the volume of the chamber expands from V1 to V2, the pressure drops from P1 to P2. From the pressure change, the volume VS displaced by the sample can be determined.

Figure 1: Basic Ideal-Gas Law Approach to Measurement of Volume.

159

Figure 2: The Quantachrome Ultrapycnometer 1000.

160

transducer

PA+ Pi Pi PA valve Vs

Vcell VA

PA+ Pf Pf PA+ Pf

Vs

Vcell VA

The instrument contains two chambers connected through a valve. When the valve is opened, the gas expands from the cell into VA, causing a drop in pressure from Pi to Pf. Pressures (above ambient atmospheric pressure PA) are measured by the transducer attached to the cell. Since the chamber volumes are known, the volume Vs displaced by the sample can be calculated.

Figure 3: Diagram of the Quantachrome Ultrapycnometer 1000.

161

The vertical axis is, for each subsequent run number, the average pycnometer measurement for all the meteorite samples in the survey divided by the final reported volume for each. Error bars are 1 standard deviation among the individual measurements. Reported volumes are taken from the average of the last six runs for each sample.

Figure 4: Ramp-up of Grain Volume Measurements in the Ultrapycnometer 1000.

162

REE Analysis of Glass Beads 100

10

1 ConcentrationChondrites/ 60 um beads 750 um beads Bass Ale RedHook Schlafly Pale Ale Guinness Fat Tire Boulevard Bridgeview chardonnay Schmitt-Sohne Riesling Giovello pinot grigio 0.1 La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu REE Atomic Number

(Data for this figure courtesy R. Korotev)

Figure 5: Rare-Earth Element Concentrations for Glass Beads (Normalized to Chondritic) With Comparison Data Taken from Common Brown and Blue Glass Bottles.

163

Figure 6: The Bead Method Apparatus, Including Shake Platform and Nalgene Beaker.

164

The small dots represent a fit (linear on a volume-volume curve) to the data after eliminating the three smallest samples. Low-mass samples (below about 15 g) have very large error bars and are unreliable. Above 15 g, samples exhibit a roughly 2% volume overestimate that depends on sample size.

Figure 7: Data from Measurements on Quartz Utilizing a 77-cm3 Cup, Made Without Employing a Bead Settling Method.

165

Figure 8: Data from Measurements Made Using the ―Soft Tap‖ Settling Method.

166

The small dots represent a fit (linear in volume-volume) to all but the three smallest samples. This settling method produces an overall underestimate in volume, with the effect reducing as volume increases, approaching zero for samples over ~30 cm3.

Figure 9: Data from Measurements Using the ―Free Shake‖ Settling Method.

167

Excepting the two quartz samples under 10 g, data indicate very small discrepancies from the actual volume. Error bars represent the variation among individual measurements, but individual measurement uncertainties in most cases are comparable to or larger than error bars shown.

Figure 10: Data for Measurements Made Using the 5-second Secured Shake Method.

168

Figure 11: Volume Discrepancy for the Secured-Shake Method Using 700-800 μm Diameter Beads.

169

Figure 12: Volume Discrepancy for the Secured-Shake Method Using 700-800 μm Diameter Beads and the 155-cm3 Cup.

170

Figure 13: Bulk Density of Small Glass Beads (ρbead ) vs. Relative Humidity.

171

control buttons

pickup coil

Top View

Side View (a)

(b)

Figure 14: The SM-30 Magnetic Susceptibility Meter. (a) Photograph of the Meter. (b) Diagram.

172

meteorite

Table

(a) (b)

Figure 15: Operation of the SM-30 Magnetic Susceptibility Meter. (a) Photograph of the Device as Utilized (Inverted, with the Meteorite Placed Atop and Centered Over the Magnetic Coils). (b) Diagram of the Device in Operation.

173

Based on Gattacceca et al. (2004). Note the log-log axes. The correction factor overlays data comparing measurements on pebbles made with the SM-30 (M) with KLY2-derived reference susceptibilities for the same pebbles (Kr). Pebble data provided by J. Gattacceca (private communication). Figure 16: SM-30 Geometric Correction Factor α as a Function of Bulk Volume.

174

(a) (b) SM-30 measurements have been adjusted for bulk volume in both plots. In (a) no shape correction has been applied, and (b) has been corrected for shape.

Figure 17: Comparison of Magnetic Susceptibility Measurements Made Using the SM-30 (Vertical Axis) with Those Utilizing the KLY-2 (Horizontal Axis) for the Same Stones in the Vatican Collection.

175

(a) (b) The horizontal axis contains KLY-2 values. SM-30 measurements have been adjusted for bulk volume in both plots. In (a) no shape correction has been applied, and (b) has been corrected for shape.

Figure 18: Discrepancy in Log Units Between Measurements Made Using the SM-30 and Those Utilizing the KLY-2 for the Same Stones.

176

Ordinary Chondrites - Vatican Stones 4.20 H L 4.10 LL

4.00

3.90

)

3 -

3.80

3.70

3.60

Grain Grain Density cm (g 3.50

3.40

3.30

3.20 3.00 3.50 4.00 4.50 5.00 5.50 6.00 Magnetic Susceptibility (log χ)

This plot follows Consolmagno et al. (2006), though original data are used. Note that H, L, and LL fall primarily into separate regions (encircled). There are a few stones that appear in the wrong regions. These are possibly misclassified or mislabeled stones.

Figure 19: Grain Density vs. Magnetic Susceptibility for Ordinary Chondrite Falls in the Vatican Collection.

177

Outliers are highlighted in yellow.

Figure 20: Grain Density vs. Magnetic Susceptibility for All Stones from OC Falls.

178

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 21: Porosity as a Function of Shock for H Chondrite Falls.

179

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 22: Porosity as a Function of Petrographic Type for H Chondrite Falls.

180

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 23: Magnetic Susceptibility as a Function of Petrographic Type for H Chondrite Falls.

181

H Chondrite Falls 4.30

4.10

3.90

)

3 -

3.70 y = 0.1465x + 3.2148

3.50 GrainDensity (gcm 3.30

3.10

2.90 2.40 2.60 2.80 3.00 3.20 3.40 3.60 3.80 4.00 Bulk Density (g cm-3)

Figure 24: Grain Density/Bulk Density Relationship for H Falls, For Use in Determination of Model Porosities for H Finds.

182

H Find Model Porosities 30.00%

H Fall Porosity H Find Model Porosity 25.00%

20.00%

15.00% Porosity

10.00%

5.00%

0.00% 2.50 2.70 2.90 3.10 3.30 3.50 3.70 3.90 Bulk Density (g cm-3)

Figure 25: Comparison of Model Porosities for H Finds with Actual Porosities for H Falls

183

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 26: Porosity as a Function of Petrographic Type for H Chondrites, Including Measured Porosities of Falls and Model Porosities of Finds.

184

Contours labeled (1), (2), (3) and (4) represent weathering moduli at those values.

Figure 27: Grain Density and Magnetic Susceptibility of H Finds.

185

Figure 28: Porosity vs. Weathering Modulus for H Finds.

186

Figure 29: Bulk Density vs. Weathering Modulus for H Finds.

187

Figure 30: Model Porosity vs. Weathering Modulus for H Finds.

188

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 31: Porosity vs. Petrographic Type for L Falls.

189

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 32: Porosity as a Function of Shock Stage for L Falls.

190

L Find Model Porosities 20.00%

L Fall Porosities 18.00% L Find Model Porosities 16.00%

14.00%

12.00%

10.00% Porosity 8.00%

6.00%

4.00%

2.00%

0.00% 2.70 2.90 3.10 3.30 3.50 3.70 3.90 Bulk Density (g cm-3)

Figure 33: Comparison of Model Porosities for L Finds with Actual Porosities for L Falls.

191

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 34: Porosity as a Function of Petrographic Type for L Chondrites, Including Measured Porosities of Falls and Model Porosities of Finds.

192

Contours labeled (1), (2), (3) and (4) represent weathering moduli at those values.

Figure 35: Grain Density and Magnetic Susceptibility of L Finds.

193

Figure 36: Porosity vs. Weathering Modulus for L Finds.

194

Figure 37: Bulk Density vs. Weathering Modulus for L Finds.

195

Figure 38: Model Porosity vs. Weathering Modulus for L Finds.

196

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below each category are the number of data points for that category.

Figure 39: Porosity vs. Petrographic Grade for LL Falls.

197

Numbers in parentheses below each category are the number of data points for that category.

Figure 40: Magnetic Susceptibility as a Function of Petrographic Type for LL Falls.

198

LL Find Model Porosities 25.00% LL Fall Porosities LL Find Model Porosities

20.00%

15.00% Porosity 10.00%

5.00%

0.00% 2.50 2.70 2.90 3.10 3.30 3.50 3.70 Bulk Density (g cm-3)

Figure 41: Comparison of Model Porosities for LL Finds with Actual Porosities for LL Falls.

199

Numbers in parentheses below each category are the number of data points for that category.

Figure 42: Porosity as a Function of Petrographic Type for LL Chondrites, Including Measured Porosities of Falls and Model Porosities of Finds.

200

Contours labeled (1), (2), (3) and (4) represent weathering moduli at those values.

Figure 43: Grain Density and Magnetic Susceptibility of LL Finds.

201

Figure 44: Porosity vs. Weathering Modulus for LL Finds.

202

Figure 45: Bulk Density vs. Weathering Modulus for LL Finds.

203

Figure 46: Model Porosity vs. Weathering Modulus for LL Finds.

204

OC Falls: Regolith Breccia Porosities

30% Normal OCs Breccias 25%

20%

15%

10% Percentage Percentage ofPopulation

5%

0% 0 - 3% 4% - 6% 7% - 9% 8% - 12% 12% - 15% 16% - 18% 19%+ Porosity Bin

Bins are inclusive (e.g. ―4% - 6%‖ spans 4.00% to 6.99%).

Figure 47: Histogram of Porosities for Brecciated and Non-Brecciated Ordinary Chondrite Falls.

205

For reference, the regions occupied by H, L, and LL ordinary chondrite falls are displayed in yellow.

Figure 48: Grain Density / Magnetic Susceptibility Plot for K and R Chondrites.

206

Average value and σmean 50% of population Median Individual meteorites

Ordinary Chondrite Range

Numbers in parentheses below group names are number of meteorites represented in each group.

Figure 49: Grain Densities of Carbonaceous Chondrites.

207

Average value and σmean 50% of population Median Individual meteorites

Ordinary Chondrite Range

Numbers in parentheses below group names are number of meteorites represented in each group.

Figure 50: Bulk Densities of Carbonaceous Chondrites.

208

Average value and σmean 50% of population Median Individual meteorites

Ordinary Chondrite Range

Numbers in parentheses below group names are number of meteorites represented in each group.

Figure 51: Porosities of Carbonaceous Chondrites.

209

H Range

L Range

LL Range

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group names are number of meteorites represented in each group.

Figure 52: Magnetic Susceptibilities of Carbonaceous Chondrites.

210

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 53: Grain Density and Magnetic Susceptibility for the CR Clan: CR, CB and CH Carbonaceous Chondrites.

211

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 54: Grain Density and Magnetic Susceptibility for CI and CM Carbonaceous Chondrites.

212

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 55: Grain Density and Magnetic Susceptibility for CO Carbonaceous Chondrites.

213

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 56: Grain Density and Magnetic Susceptibility for CK Carbonaceous Chondrites.

214

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 57: Grain Density and Magnetic Susceptibility for CV Carbonaceous Chondrites.

215

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 58: Grain Density and Magnetic Susceptibility for Ungrouped Carbonaceous Chondrites.

216

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 59: Bulk Density vs. Petrographic Type for Carbonaceous Chondrite Falls.

217

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 60: Grain Density vs. Petrographic Type for Carbonaceous Chondrite Falls.

218

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 61: Porosity vs. Petrographic Type for Carbonaceous Chondrite Falls.

219

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 62: Magnetic Susceptibility vs. Petrographic Type for Carbonaceous Chondrite Falls.

220

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 63: Bulk Density vs. Shock for Carbonaceous Chondrites.

221

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 64: Grain Density vs. Shock for Carbonaceous Chondrites.

222

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 65: Porosity vs. Shock for Carbonaceous Chondrites.

223

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 66: Magnetic Susceptibility vs. Shock for Carbonaceous Chondrites.

224

Average value and σmean 50% of population Median Individual stones

Ordinary Chondrite Range

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 67: Enstatite Chondrite Grain Densities.

225

Average value and σmean 50% of population Median Individual stones

Ordinary Chondrite Range

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 68: Enstatite Chondrite Bulk Densities.

226

Average value and σmean 50% of population Median Individual stones

Ordinary Chondrite Range

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 69: Enstatite Chondrite Porosities.

227

H Range

L Range

Average value and σmean 50% of population Median LL Range Individual stones

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 70: Enstatite Chondrite Magnetic Susceptibilities.

228

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow

Figure 71: Grain Density vs. Magnetic Susceptiblity for Enstatite Chondrites.

229

(a) (b)

Figure 72: Enstatite Chondrite Finds and Their Properties Grouped By Weathering: (a) Selection of Groups Based on Grain Density and Magnetic Susceptibility; (b) Porosity and Bulk Density of the Same Groups.

230

Figure 73: Porosity as a Function of Weathering Modulus for EC Finds.

231

For comparison purposes, H and L ordinary chondrite data are overlaid. Data were ordered by mass, and running bins created (ten stones per bin for EC, five for EH and EL, and twenty each for H and L, with bin sizes based on total number of stones per group), from which the standard deviation of grain densities was calculated. The sample mass is the average mass per bin.

Figure 74: Variability of EC Grain Densities by Mass.

232

The ovals represent 1-σ and 2-σ from the mean for (solid) EH and (dashed) EL.

Figure 75: Mass-Weighted Average Grain Density vs. Magnetic Susceptibility for Falls Exceeding Total Mass of 40 g.

233

Ordinary Chondrite Range

Average value and σmean 50% of population Median Individual stones

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 76: Grain Densities of Achondrites in this Study.

234

Average value and σmean 50% of population Median Individual stones

Ordinary Chondrite Range

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 77: Bulk Densities of Achondrites in this Study.

235

Average value and σmean

50% of population Median Individual stones

OC Range

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 78: Porosities of Achondrites in this Study.

236

Average value and σmean 50% of population H Range Median Individual stones L Range

LL Range

Numbers in parentheses below group labels are number of stones represented in each group.

Figure 79: Magnetic Susceptibilities for Achondrites in this Study.

237

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 80: Grain Density vs. Magnetic Susceptibility for SNC (Martian) Meteorites in this Study.

238

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow

Figure 81: Grain Density vs. Magnetic Susceptibility for HED Meteorites in this Study.

239

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow and the region occupied by reduced CVs is in blue.

Figure 82: Grain Density vs. Magnetic Susceptibilities for Aubrites, Angrites and Ureilites in this Study, as well as One Stone of an Enstatite Achondrite (Zakłodzie).

240

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 83: Grain Density vs. Magnetic Susceptibility for Acapulcoites and Lodranites in this Study.

241

For reference, the regions occupied by H, L and LL ordinary chondrite falls are displayed in yellow.

Figure 84: Grain Density vs. Magnetic Susceptibility for Primitive Achondrites in this Study.

242

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 85: Porosity vs. Shock State for All Chondrite Falls.

243

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 86: Grain Density vs. Shock State for All Chondrite Falls.

244

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 87: Magnetic Susceptibility vs. Shock State for All Chondrite Finds.

245

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 88: Porosity vs. Petrographic Type for All Chondrite Falls.

246

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 89: Grain Density vs. Petrographic Type for All Chondrite Falls.

247

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 90: Bulk Density vs. Petrographic Type for All Chondrite Falls.

248

Average value and σmean 50% of population Median Individual meteorites

Numbers in parentheses below group labels are number of meteorites represented in each group.

Figure 91: Magnetic Susceptibility vs. Petrographic Type for All Chondrite Falls.

249

Chondrite Falls: Porosity vs. Oxidation State 40.0% CI CK 35.0% CM CO CR CV 30.0% H L LL 25.0% EH EL

20.0% Porosity

15.0%

10.0%

5.0%

0.0% 0 5 10 15 20 25 30 35 40 45 50

Average Feox / (Feox + MgO)

Data points are group averages, and ―error bars‖ are one standard deviation of the population.

Figure 92: Porosity vs. Oxidation State for All Chondrite Falls.

250

Chondrite Falls: Porosity vs. Matrix Abundance 40.0% CI CK 35.0% CM CO 30.0% CR CV H 25.0% L LL 20.0% EH

Porosity EL 15.0% K

10.0%

5.0%

0.0% 0 20 40 60 80 100 Percentage Matrix (vol %)

Data points are group averages, vertical ―error bars‖ are one standard deviation of the population, and horizontal bars represent ranges given in Brearley and Jones (1988).

Figure 93: Porosity vs. Percentage Matrix for All Chondrite Falls.

251

APPENDIX B: TABLES

252

Table 1: Electron-Probe Micro Analysis (EPMA) of Glass Beads.

Sample SiO2 TiO2 Al2O3 Cr2O3 FeO MnO MgO CaO Na2O K2O TOTAL 5 Density beads, 750 µm 73.02 0.02 0.12 0.02 0.00 0.03 4.37 8.75 13.69 0.08 100.09 5 Density beads, 750 µm 73.24 0.00 0.15 0.00 0.04 0.00 4.51 8.89 13.68 0.05 100.57 5 Density beads, 750 µm 73.72 0.06 0.58 0.03 0.15 0.06 4.27 8.62 13.43 0.10 101.02 5 Density beads, 750 µm 73.49 0.03 0.94 0.02 0.09 0.01 3.69 9.12 13.54 0.21 101.15 5 Density beads, 750 µm 74.35 0.03 0.18 0.02 0.10 0.00 3.91 8.54 13.74 0.10 100.96 5 Density beads, 750 µm 73.24 0.02 0.20 0.02 0.01 0.02 4.35 8.77 13.62 0.07 100.30 6 Density beads, 60 µm 73.70 0.04 0.20 0.00 0.12 0.00 4.32 8.79 13.44 0.05 100.67 6 Density beads, 60 µm 74.63 0.06 0.09 0.00 0.56 0.00 4.02 8.73 13.32 0.00 101.42 6 Density beads, 60 µm 73.67 0.00 0.19 0.01 0.13 0.09 4.30 8.71 13.43 0.03 100.56 6 Density beads, 60 µm 74.08 0.24 0.11 0.01 0.67 0.04 4.17 8.62 13.46 0.02 101.42 6 Density beads, 60 µm 73.62 0.01 0.09 0.00 0.06 0.00 3.06 9.89 13.90 0.04 100.66 7 Density beads, 750 µm (original bead to check calibration) 72.08 0.01 0.14 0.00 0.08 0.00 4.33 8.67 13.62 0.07 99.02 17 Density beads as a check (last check) 73.01 0.00 0.08 0.03 0.09 0.00 3.92 8.87 13.30 0.01 99.31 Average (all) 73.53 0.04 0.24 0.01 0.16 0.02 4.09 8.84 13.55 0.06 100.55 Standard Deviation 0.651 0.062 0.248 0.011 0.207 0.030 0.387 0.347 0.174 0.053

Average (large) 73.31 0.03 0.33 0.01 0.07 0.02 4.20 8.76 13.62 0.10 100.44 Average (Small) 73.79 0.06 0.13 0.01 0.27 0.02 3.96 8.93 13.47 0.03 100.67 Data courtesy R. Korotev.

253

Table 2: Instrumental Neutron Activation Analysis (INAA) of Glass Beads.

60 µm beads 750 µm beads

Na2O % 14.07 ± 0.14 13.95 ± 0.14

K2O % n.d. n.d. CaO % 8.95 ± 0.18 8.51 ± 0.15 Sc ppm 0.201 ± 0.0021 0.334 ± 0.0033 Cr ppm 12.74 ± 0.2 2.99 ± 0.19 FeO % 0.218 ± 0.0025 0.096 ± 0.0026 Co ppm 7.32 ± 0.073 1.65 ± 0.037 Ni ppm n.d. n.d. Zn ppm 28.4 ± 1.4 11.5 ± 1.5 As ppm <0.9 ± 0.22 0.43 ± 0.15 Se ppm 0.56 ± 0.12 0.9 ± 0.26 Br ppm 1.11 ± 0.18 0.87 ± 0.14 Rb ppm n.d. n.d. Sr ppm 84 ± 5.7 109 ± 8 Zr ppm 59 ± 7.9 799 ± 18 Ag ppm 6.42 ± 0.091 1.58 ± 0.091 Sb ppm 0.825 ± 0.01 3.09 ± 0.031 Cs ppm 0.07 ± 0.021 3.65 ± 0.041 Ba ppm 14.6 ± 2 20 ± 3.1 La ppm 1.848 ± 0.018 2.1 ± 0.023 Ce ppm 53.1 ± 0.53 19.64 ± 0.2 Nd ppm 1.5 ± 0.37 1.9 ± 0.55 Sm ppm 0.274 ± 0.0034 0.341 ± 0.0044 Eu ppm 0.044 ± 0.0083 0.063 ± 0.012 Tb ppm 0.032 ± 0.0057 0.054 ± 0.011 Yb ppm 0.175 ± 0.0056 0.58 ± 0.0099 Lu ppm 0.0295 ± 0.0013 0.109 ± 0.0021 Hf ppm 1.51 ± 0.026 18.22 ± 0.18 Ta ppm 0.058 ± 0.016 0.65 ± 0.063 W ppm <1.0 ± 0.45 <0.8 ± 0.39 Ir ppb <0.6 ± 0.29 <1.4 ± 0.51 Au ppb 2.8 ± 0.34 <1.3 ± 0.62 Th ppm 0.448 ± 0.013 0.73 ± 0.029 U ppm 0.33 ± 0.022 0.81 ± 0.036 Data courtesy R. Korotev

254

Table 3: Results Per Settling Method for Small Beads in the 77 cm3 Container.

Sample Mass Actual volume Volume Volume Volume Volume (g) (cm3) overestimate: overestimate: overestimate: overestimate: No Settling (%) Soft Tap (%) Free Shake (%) Secured Shake (%) Quartz 1 9.30 3.549 ± 0.004 0.66 ± 3.99 1.71 ± 6.36 -4.94 ± 4.60 -1.10 ± 2.56 Quartz 2 9.76 3.723 ± 0.005 -0.15 ± 5.18 2.35 ± 5.33 -1.80 ± 2.14 -2.69 ± 2.62 Quartz 3 14.04 5.325 ± 0.032 1.34 ± 2.92 2.98 ± 0.93 -2.28 ± 2.37 0.68 ± 0.18 Quartz 4 18.97 7.200 ± 0.006 3.13 ± 2.13 4.03 ± 1.53 -4.67 ± 1.63 0.36 ± 0.43 Quartz 5 26.47 10.04 ± 0.02 2.71 ± 0.91 2.96 ± 2.54 -2.35 ± 2.26 -0.10 ± 1.10 Quartz 6 50.14 19.01 ± 0.03 1.69 ± 0.41 2.04 ± 1.64 -0.66 ± 0.90 1.11 ± 0.45 Quartz 7 76.61 28.97 ± 0.04 1.47 ± 0.48 2.69 ± 1.31 -0.09 ± 0.40 0.63 ± 0.75 Quartz 8 46.96 17.89 ± 0.01 2.59 ± 0.43 5.34 ± 0.72 -0.82 ± 0.73 -0.58 ± 0.60 Topaz 1 67.04 18.88 ± 0.01 n.d. 1.47 ± 1.33 -1.75 ± 0.50 -0.33 ± 0.19 Topaz 2 49.99 14.12 ± 0.01 n.d. 1.69 ± 0.73 -3.47 ± 0.88 -0.01 ± 0.98 Topaz 3 56.31 15.92 ± 0.05 n.d. 1.03 ± 0.04 -1.75 ± 0.31 -1.13 ± 0.15 The symbol ―n.d.‖ indicates no data available.

255

Table 4: Data for H Chondrite Falls.

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) b H5 Fall LNHM BM19971 31.24 3.29 ± 0.04 3.59 ± 0.02 8.5% ± 1.2% 4.83 ± 0.10 Agen H5 Fall Vatican 8 9.18 n.d. n.d. n.d. 5.16 ± 0.08† Agen H5 Fall Vatican 6 36.8 3.37 ± 0.02 3.77 ± 0.04 10.5% ± 1.1% 5.20 ± 0.09 Agen H5 Fall Average 3.33 3.69 9.6% 5.07 Alessandriab H5 Fall AMNH 2224 8.83 3.27 ± 0.10 3.77 ± 0.03 13.3% ± 2.7% 5.42 ± 0.08 H5 S1 W0 Fall AMNH 371 92.81 3.06 ± 0.05 3.73 ± 0.01 18.0% ± 1.3% 5.32 ± 0.09 Allegan H5 S1 W0 Fall NMNH 2632 26.089 3.11 ± 0.05 3.64 ± 0.02 14.6% ± 1.5% 5.23 ± 0.08 Allegan H5 S1 W0 Fall NMNH 2633 17.974 2.97 ± 0.06 3.68 ± 0.03 19.3% ± 1.7% 5.33 ± 0.08 Allegan H5 S1 W0 Fall Vatican 25 145.37 2.94 ± 0.03 n.d. n.d. 5.29 ± 0.09 Allegan H5 S1 W0 Fall Vatican 24 294.96 3.39 ± 0.03‡ n.d. n.d. 5.12 ± 0.08† Allegan H5 S1 W0 Fall Average 3.19 3.71 14.0% 5.26 Ambapur Nagla H5 Fall AMNH 401 82.51 3.07 ± 0.05 3.66 ± 0.01 16.2% ± 1.3% 5.40 ± 0.09 H5 Fall CMS 1046-1 32.02 3.13 ± 0.04 3.63 ± 0.02 13.8% ± 1.3% 5.50 ± 0.09 H6 Fall AMNH 4000 46.26 3.47 ± 0.02 3.60 ± 0.01 3.7% ± 0.7% 5.22 ± 0.10 H4 S2 Fall NMNH 6882 14.031 2.88 ± 0.39 3.18 ± 0.03 9.4% ± 12.4% 5.18 ± 0.10 Avanhandava H4 S2 Fall IOM C 135.1a 35.34 3.04 ± 0.03 3.67 ± 0.02 17.0% ± 0.9% 5.22 ± 0.12 Avanhandava H4 S2 Fall Average 2.99 3.51 14.8% 5.20 Barbotanb H5 S3 Fall AMNH 3904 84.22 3.44 ± 0.01 3.71 ± 0.01 7.2% ± 0.4% 5.15 ± 0.09 Barbotan H5 S3 Fall Vatican 82 83.19 3.53 ± 0.02 3.78 ± 0.04 6.6% ± 1.0% 5.27 ± 0.09 Barbotan H5 S3 Fall Average 3.49 3.75 6.9% 5.21 Batha H4 Fall Vatican 91 37.715 3.48 ± 0.02 3.89 ± 0.04 10.5% ± 1.1% 5.14 ± 0.10 Bath H4 Fall Vatican 90 156.8 3.43 ± 0.07‡ n.d. n.d. n.d. Bath H4 Fall Vatican 89 477.1 3.41 ± 0.07‡ n.d. n.d. n.d. Bath H4 Fall Average 3.42 3.89 12.0% 5.14 Beardsley H5 S3 Fall NMNH 2658 36.509 3.28 ± 0.20 3.64 ± 0.01 9.9% ± 5.5% 5.27 ± 0.09 Beardsley H5 S3 Fall Vatican 97 9.87 n.d. n.d. n.d. 5.03 ± 0.08† Beardsley H5 S3 Fall Vatican 96 17.42 3.41 ± 0.03 n.d. n.d. 5.29 ± 0.08 Beardsley H5 S3 Fall Average 3.32 3.64 8.8% 5.20 Beaver Creek H5 S3 Fall NMNH 2659 17.42 3.24 ± 0.09 3.62 ± 0.03 10.4% ± 2.6% 5.30 ± 0.08 Beaver Creek H5 S3 Fall Vatican 98 204.26 3.14 ± 0.03 n.d. n.d. 5.27 ± 0.10 Beaver Creek H5 S3 Fall Average 3.15 3.62 13.1% 5.28 Benldb H6 Fall CMS 1173-1 88.6 3.46 ± 0.05 3.70 ± 0.01 6.6% ± 1.5% 5.26 ± 0.12 Bielokrynitschieb H4 Fall CMS 150-1 18.13 3.39 ± 0.06 3.59 ± 0.03 5.5% ± 1.7% 5.16 ± 0.08 Bielokrynitschie H4 Fall LNHM BM66213 47.12 3.49 ± 0.05 3.67 ± 0.02 4.8% ± 1.4% 5.34 ± 0.10 Bielokrynitschie H4 Fall Average 3.46 3.65 5.0% 5.25 Binningup H5 Fall LNHM BM1989,M13 34.35 3.19 ± 0.04 3.72 ± 0.02 14.3% ± 1.1% 5.29 ± 0.09 Bjelaja Zerkov H6 Fall Vatican 118 13.26 n.d. n.d. n.d. 5.41 ± 0.08† Bur-Gheluai H5 S3 Fall IOM C 202.1 31.1 3.54 ± 0.05 3.66 ± 0.02 3.2% ± 1.4% 5.29 ± 0.08 Bur-Gheluai H5 S3 Fall Vatican 152 1287.49 3.54 ± 0.07‡ 3.64 ± 0.04‡ 2.7% ± 2.2% 5.38 ± 0.08 Bur-Gheluai H5 S3 Fall Average 3.54 3.64 2.8% 5.33 Burnwell H4 S3 Fall LNHM BM2000,M4 65.24 3.51 ± 0.04 3.73 ± 0.01 5.9% ± 1.1% 5.34 ± 0.12 Cañellasa H4 Fall CMS 790-1 28.06 3.32 ± 0.04 3.68 ± 0.02 9.8% ± 1.3% 5.30 ± 0.10 Cangas de Onis H5 S3 Fall IOM C 199.2 19.06 3.64 ± 0.09 3.65 ± 0.03 0.3% ± 2.5% 5.12 ± 0.10 Cangas de Onis H5 S3 Fall Vatican 167 46.76 3.50 ± 0.02 3.82 ± 0.04 8.4% ± 1.1% 5.27 ± 0.09 Cangas de Onis H5 S3 Fall Vatican 168 12.46 n.d. n.d. n.d. 5.25 ± 0.08† Cangas de Onis H5 S3 Fall Average 3.54 3.77 6.2% 5.21 Cape Girardeau H6 Fall Vatican 181 97.91 3.42 ± 0.04 n.d. n.d. 5.40 ± 0.09 Capilla del Monte H6 Fall LNHM BM1964,68 22.79 3.34 ± 0.08 3.70 ± 0.02 9.7% ± 2.2% 5.27 ± 0.10 Carancas H4-5 S3 W0 Fall CMS 1615-1 19.46 3.17 ± 0.07 3.68 ± 0.03 13.9% ± 1.9% 5.23 ± 0.08 Carancas H4-5 S3 W0 Fall Vatican 1486 32.38 3.06 ± 0.03 n.d. n.d. 5.12 ± 0.09 Carancas H4-5 S3 W0 Fall Average 3.10 3.68 15.8% 5.17 Castaliaa H5 Fall CMS 347-1 24.06 3.37 ± 0.05 3.62 ± 0.02 6.9% ± 1.6% 5.05 ± 0.10 Castalia H5 Fall LNHM BM50804 26.68 3.40 ± 0.05 3.70 ± 0.02 8.1% ± 1.4% 5.14 ± 0.08 Castalia H5 Fall Average 3.38 3.66 7.5% 5.10 Ceresetoa H5 Fall LNHM BM33297 21.88 3.23 ± 0.06 3.67 ± 0.02 12.0% ± 1.7% 5.21 ± 0.10 Cereseto H5 Fall Vatican 199 11.99 n.d. n.d. n.d. 5.29 ± 0.08† Cereseto H5 Fall Average 3.23 3.67 12.0% 5.25 Charsonvilleb H6 S4 Fall CMS 538-2 58.35 3.54 ± 0.09 3.71 ± 0.01 4.5% ± 2.5% 5.39 ± 0.10 Charsonville H6 S4 Fall Vatican 206 11.89 n.d. n.d. n.d. 5.64 ± 0.08† Charsonville H6 S4 Fall Average 3.54 3.71 4.5% 5.51 Charwallis H6 Fall CMS 562-2-1 44.18 3.39 ± 0.04 3.60 ± 0.01 5.6% ± 1.0% 5.33 ± 0.09 Collescipoli H5 Fall Vatican 233 9.24 n.d. n.d. n.d. 5.37 ± 0.08†

256

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Dashoguz H5 S3 W0/1 Fall Monnig M1148.1 97.26 3.14 ± 0.05 3.74 ± 0.01 16.0% ± 1.3% 5.32 ± 0.09 Dhajala H4 S1 Fall NMNH 5832 49.564 3.21 ± 0.03 3.72 ± 0.01 13.6% ± 1.0% 5.18 ± 0.09 Dhajala H4 S1 Fall IOM C 158.1 41.65 3.23 ± 0.04 3.73 ± 0.01 13.4% ± 1.0% 5.16 ± 0.09 Dhajala H4 S1 Fall Average 3.22 3.72 13.5% 5.17 Djati-Pengilon H6 Fall LNHM BM1964,564 77.29 3.64 ± 0.03 3.73 ± 0.01 2.4% ± 0.9% 5.19 ± 0.09 Djati-Pengilon H6 Fall Vatican 271 10.54 n.d. n.d. n.d. 5.14 ± 0.08† Djati-Pengilon H6 Fall Average 3.64 3.73 2.4% 5.16 Donga Kohrod H6 Fall LNHM BM85671 35.16 3.42 ± 0.05 3.70 ± 0.02 7.6% ± 1.4% 5.37 ± 0.09 Doroninska H5-7 Fall CMS 1103-1 68.12 3.52 ± 0.03 3.72 ± 0.01 5.4% ± 0.8% 5.47 ± 0.09 Dwalenia,b H4-6 Fall CMS 841-1 27 3.32 ± 0.05 3.71 ± 0.02 10.4% ± 1.5% 5.29 ± 0.08 Ehole H5 Fall Monnig M1153.1 18.87 3.37 ± 0.07 3.74 ± 0.03 9.8% ± 2.0% 5.49 ± 0.08 Eichstädt H5 Fall LNHM BM84188 33.31 2.51 ± 0.03 3.42 ± 0.01 26.6% ± 0.9% 5.25 ± 0.09 Épinal H5 Fall Vatican 1105 3.43 3.48 ± 0.12‡ 3.71 ± 0.08‡ 6.3% ± 4.0% 5.36 ± 0.08† Farmville H4 S3 Fall NMNH 937 4 38.949 3.50 ± 0.09 3.70 ± 0.02 5.5% ± 2.6% 5.38 ± 0.10 Favars H5 Fall Vatican 330 14.83 n.d. n.d. n.d. 5.50 ± 0.08† Fermoa H3-5 Fall Vatican 1186 75.79 3.39 ± 0.02 3.88 ± 0.04 12.7% ± 1.0% 5.22 ± 0.09 Florencea H3 Fall CMS 218-1-1 20.69 3.14 ± 0.05 3.66 ± 0.03 14.0% ± 1.5% 5.38 ± 0.12 Forest Citya H5 S2 Fall Vatican 341 36.09 3.43 ± 0.03 3.80 ± 0.04 9.8% ± 1.1% 5.33 ± 0.09 Forest City H5 S2 Fall Vatican 340 88.35 3.45 ± 0.02 n.d. n.d. 5.33 ± 0.09 Forest City H5 S2 Fall Average 3.44 3.80 9.5% 5.33 Forest Vale H4 Fall Monnig M1247.1 29.88 2.96 ± 0.08 3.69 ± 0.03 19.7% ± 2.3% 5.33 ± 0.12 Gao-Guenie H5 Fall Vatican 1368 21.06 3.55 ± 0.05 3.60 ± 0.04 1.4% ± 1.8% 5.26 ± 0.12 Gladstone (stone) b H4 S3 Fall IOM C 13.25 25.44 3.48 ± 0.08 3.54 ± 0.02 1.7% ± 2.3% 5.07 ± 0.10 Gross-Divina H5 Fall Vatican 374 9.51 n.d. n.d. n.d. 5.37 ± 0.08† Grünebergb H4 Fall LNHM BM35179 21.19 3.55 ± 0.09 3.73 ± 0.03 4.7% ± 2.6% 5.34 ± 0.10 Guareña H6 S1 W3 Fall CMS 787-1 51.98 3.61 ± 0.04 3.71 ± 0.02 2.8% ± 1.1% 5.24 ± 0.10 Hessle H5 Fall LNHM BM1927,1288 37.65 3.21 ± 0.04 3.71 ± 0.01 13.3% ± 1.3% 5.29 ± 0.09 Hessle H5 Fall Vatican 392 10.28 n.d. n.d. n.d. 5.19 ± 0.08† Hessle H5 Fall Vatican 391 64.81 3.29 ± 0.02 3.78 ± 0.04 12.8% ± 1.0% 5.29 ± 0.09 Hessle H5 Fall Average 3.26 3.75 13.0% 5.26 Ipiranga H6 S3 Fall IOM C 163.3 66.42 3.42 ± 0.05 3.74 ± 0.01 8.5% ± 1.4% 5.34 ± 0.10 Jilin H5 Fall IOM C 205.3 33.56 3.49 ± 0.06 3.70 ± 0.02 5.7% ± 1.8% 5.39 ± 0.09 Juancheng H5 S2 Fall LNHM BM1999,M23 43.54 3.61 ± 0.04 3.63 ± 0.01 0.7% ± 1.2% 5.23 ± 0.09 Juancheng H5 S2 Fall Vatican 1434 201.92 3.53 ± 0.03 n.d. n.d. 5.25 ± 0.08 Juancheng H5 S2 Fall Average 3.55 3.63 2.4% 5.24 Kabo H4 Fall CMS 977-1 21.49 3.43 ± 0.04 3.63 ± 0.02 5.4% ± 1.4% 5.27 ± 0.12 Kerilis H5 Fall Vatican 494 47.59 3.40 ± 0.04 3.88 ± 0.04 12.4% ± 1.4% 5.33 ± 0.12 Kernouvéb H6 S1 Fall Vatican 499 29.69 3.40 ± 0.03 3.77 ± 0.04 9.7% ± 1.1% 5.49 ± 0.08 Kernouvé H6 S1 Fall Vatican 1379 17.96 3.65 ± 0.07 3.74 ± 0.04 2.5% ± 2.0% 5.51 ± 0.08† Kernouvé H6 S1 Fall Average 3.49 3.76 7.1% 5.50 Kesen H4 S3 Fall NMNH 3329 85.651 3.47 ± 0.03 3.68 ± 0.01 5.6% ± 0.9% 5.32 ± 0.10 Kesen H4 S3 Fall IOM C 48.5 60.52 3.43 ± 0.07 3.64 ± 0.01 5.7% ± 1.9% 5.46 ± 0.09 Kesen H4 S3 Fall Vatican 503 184.44 3.48 ± 0.03 n.d. n.d. 5.37 ± 0.09 Kesen H4 S3 Fall Average 3.47 3.66 5.2% 5.38 Kilbourn H5 Fall CMS 250-1 25.82 3.30 ± 0.05 3.65 ± 0.02 9.7% ± 1.4% 5.21 ± 0.12 Kilbourn H5 Fall Vatican 504 18.72 3.41 ± 0.05 3.81 ± 0.04 10.5% ± 1.6% 5.21 ± 0.12 Kilbourn H5 Fall Average 3.34 3.72 10.0% 5.21 Lançonb H6 Fall Monnig M 588.3 62 3.49 ± 0.04 3.77 ± 0.01 7.4% ± 1.2% 5.38 ± 0.09 Lançon H6 Fall Vatican 546 22.43 3.39 ± 0.03 3.76 ± 0.04 9.9% ± 1.1% 5.42 ± 0.08 Lançon H6 Fall Average 3.46 3.77 8.1% 5.40 Le Pressoir H6 Fall Vatican 550 13.25 n.d. 4.02 ± 0.04 n.d. 5.53 ± 0.08† Leighton H5 Fall CMS 1192-1 42.09 3.54 ± 0.07 3.59 ± 0.02 1.5% ± 2.0% 5.17 ± 0.12 Limerickb H5 S3 Fall NMNH 4836 173.15 3.52 ± 0.05 3.70 ± 0.00 5.0% ± 1.5% 5.29 ± 0.12 Limerick H5 S3 Fall Vatican 558 33.93 3.46 ± 0.02 3.77 ± 0.04 8.3% ± 1.1% 5.33 ± 0.09 Limerick H5 S3 Fall Vatican 559 13.78 n.d. n.d. n.d. 5.40 ± 0.08† Limerick H5 S3 Fall Average 3.51 3.71 5.5% 5.34 Lixnab H4 Fall CMS 943-1 23.37 3.42 ± 0.05 3.68 ± 0.02 7.1% ± 1.5% 5.30 ± 0.08 Lixna H4 Fall LNHM BM1985,M28 67.51 3.46 ± 0.04 3.67 ± 0.01 5.7% ± 1.0% 5.45 ± 0.09 Lixna H4 Fall Average 3.45 3.67 6.1% 5.38 Macáub H5 Fall CMS 1105-1 42.08 3.43 ± 0.04 3.55 ± 0.02 3.4% ± 1.1% 5.35 ± 0.09 Marilia H4 S3 Fall IOM C 133.1 45.81 3.38 ± 0.03 3.72 ± 0.01 9.1% ± 0.9% 5.30 ± 0.09

257

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Marilia H4 S3 Fall LNHM BM1982,M5 33.41 3.36 ± 0.07 3.77 ± 0.02 10.9% ± 1.8% 5.46 ± 0.09 Marilia H4 S3 Fall Average 3.37 3.74 9.9% 5.38 Menow H4 S2 W1 Fall Vatican 609 27.51 2.99 ± 0.02 3.55 ± 0.04 15.8% ± 1.0% 5.31 ± 0.10 Misshof H5 Fall Vatican 627 7.32 n.d. n.d. n.d. 5.41 ± 0.08† Molinaa H5 Fall CMS 220a 98.21 3.51 ± 0.06 3.64 ± 0.01 3.6% ± 1.6% 5.20 ± 0.10 Monroea H4 S2 Fall Vatican 645 37.966 3.61 ± 0.03 3.76 ± 0.04 4.1% ± 1.3% 5.47 ± 0.09 Monroe H4 S2 Fall Vatican 646 26.783 3.54 ± 0.02 3.77 ± 0.04 6.0% ± 1.1% 5.27 ± 0.08 Monroe H4 S2 Fall Average 3.58 3.77 4.9% 5.37 Mooresfort H5 Fall Monnig M1173.1 34.21 3.33 ± 0.04 3.74 ± 0.02 11.0% ± 1.2% 5.38 ± 0.09 Mooresfort H5 Fall LNHM BM61309 46.77 3.42 ± 0.03 3.68 ± 0.01 7.1% ± 1.0% 5.27 ± 0.09 Mooresfort H5 Fall Average 3.38 3.70 8.8% 5.33 Mornansb H5 Fall LNHM BM63551 35.68 3.36 ± 0.04 3.72 ± 0.02 9.8% ± 1.1% 5.47 ± 0.09 Mount Browne H6 S3 W2-3 Fall CMS 1184-1 43.68 3.30 ± 0.04 3.70 ± 0.02 10.8% ± 1.2% 5.21 ± 0.10 Mount Browne H6 S3 W2-3 Fall LNHM BM1920,325 52.98 3.44 ± 0.03 3.68 ± 0.02 6.4% ± 1.0% 5.34 ± 0.10 Mount Browne H6 S3 W2-3 Fall Average 3.38 3.69 8.4% 5.28 Nanjemoy H6 Fall Monnig M1161.1 33.11 3.40 ± 0.04 3.77 ± 0.02 9.7% ± 1.1% 5.37 ± 0.09 Naoki H6 Fall CMS 614-1-1 42.46 3.37 ± 0.03 3.69 ± 0.02 8.5% ± 1.1% 5.34 ± 0.09 Noblesvillea H4-6 Fall CMS 1582-1 34.32 3.37 ± 0.10 3.67 ± 0.02 8.2% ± 2.7% 5.18 ± 0.12 Nuevo Mercurio H5 Fall NMNH 6088 75.681 3.04 ± 0.03 3.72 ± 0.01 18.2% ± 0.8% 5.48 ± 0.09 Nuevo Mercurio H5 Fall NMNH 6089 58.325 3.01 ± 0.03 3.70 ± 0.01 18.8% ± 0.8% 5.47 ± 0.09 Nuevo Mercurio H5 Fall NMNH 6092 37.455 3.01 ± 0.03 3.72 ± 0.01 19.2% ± 0.8% 5.38 ± 0.09 Nuevo Mercurio H5 Fall NMNH 6093 42.661 3.05 ± 0.03 3.72 ± 0.01 18.0% ± 0.8% 5.45 ± 0.09 Nuevo Mercurio H5 Fall NMNH 6094 39.561 3.08 ± 0.03 3.72 ± 0.01 17.2% ± 0.9% 5.47 ± 0.09 Nuevo Mercurio H5 Fall NMNH 6097 43.532 3.09 ± 0.06 3.72 ± 0.01 17.1% ± 1.5% 5.39 ± 0.09 Nuevo Mercurio H5 Fall NMNH 6098 59.477 3.10 ± 0.03 3.72 ± 0.01 16.7% ± 0.8% 5.48 ± 0.09 Nuevo Mercurio H5 Fall NMNH 6100 26.831 3.10 ± 0.03 3.73 ± 0.02 16.9% ± 1.0% 5.43 ± 0.08 Nuevo Mercurio H5 Fall NMNH 6110 22.414 3.05 ± 0.04 3.72 ± 0.02 18.0% ± 1.2% 5.43 ± 0.08 Nuevo Mercurio H5 Fall NMNH 6120 82.929 3.03 ± 0.03 3.75 ± 0.01 19.2% ± 0.8% 5.41 ± 0.09 Nuevo Mercurio H5 Fall Average 3.05 3.72 18.1% 5.44 Nullesa H6 S3 Fall NMNH 1470 100.764 3.24 ± 0.22 3.70 ± 0.01 12.4% ± 5.9% 5.23 ± 0.10 Ochanska H4 S3 Fall NMNH 1788 27.637 3.46 ± 0.05 3.62 ± 0.02 4.5% ± 1.4% 4.97 ± 0.08 Ochansk H4 S3 Fall Monnig M 199.2 47.06 3.32 ± 0.05 3.75 ± 0.01 11.6% ± 1.5% 5.37 ± 0.09 Ochansk H4 S3 Fall Vatican 706 183.77 3.29 ± 0.03 n.d. n.d. 5.33 ± 0.09 Ochansk H4 S3 Fall Vatican 705 209.83 3.28 ± 0.03 n.d. n.d. 5.38 ± 0.08 Ochansk H4 S3 Fall Vatican 704 1026.57 3.23 ± 0.07‡ n.d. n.d. n.d. Ochansk H4 S3 Fall Average 3.25 3.70 12.2% 5.26 Ogi H6 S2 W2 Fall NMNH 616 37.654 3.49 ± 0.04 3.69 ± 0.01 5.4% ± 1.1% 5.24 ± 0.12 Ohaba H5 Fall CMS 1079-1 28.42 3.27 ± 0.04 3.68 ± 0.03 11.3% ± 1.3% 5.23 ± 0.08 Orvinioa H6 S3 Fall NMNH 308 49.684 3.47 ± 0.03 3.62 ± 0.01 4.2% ± 0.9% 5.34 ± 0.12 Orvinio H6 S3 Fall Vatican 726 382 3.67 ± 0.09‡ n.d. n.d. n.d. Orvinio H6 S3 Fall Vatican 1494 155.15 3.37 ± 0.03 n.d. n.d. 5.28 ± 0.09 Orvinio H6 S3 Fall Average 3.57 3.62 1.3% 5.31 Oum Dreygaa H3-5 S4 W0 Fall Monnig M1147.1 75.43 3.47 ± 0.06 3.71 ± 0.01 6.4% ± 1.6% 5.24 ± 0.09 H Fall CMS 1599-1 35.54 3.40 ± 0.09 3.68 ± 0.02 7.7% ± 2.4% 5.39 ± 0.09 Paitan H6 Fall LNHM BM1987,M2 50.65 3.53 ± 0.04 3.60 ± 0.02 1.9% ± 1.2% 5.30 ± 0.09 Peekskilla H6 Fall Vatican 1165 32.05 3.30 ± 0.05 3.87 ± 0.04 14.8% ± 1.4% 5.25 ± 0.12 Phû Hongb H4 Fall Vatican 1044 3.97 3.49 ± 0.17 4.14 ± 0.22 15.5% ± 6.0% 5.35 ± 0.08† Pultuska,b H5 S3 Fall NMNH 463 A 43.503 3.48 ± 0.05 3.70 ± 0.01 5.9% ± 1.3% 5.27 ± 0.09 Pultusk H5 S3 Fall NMNH 463 B 41.511 3.47 ± 0.04 3.67 ± 0.01 5.5% ± 1.0% 5.28 ± 0.09 Pultusk H5 S3 Fall NMNH 463 C 32.987 3.54 ± 0.04 3.73 ± 0.02 5.0% ± 1.2% 5.36 ± 0.09 Pultusk H5 S3 Fall NMNH 812 38.633 3.55 ± 0.04 3.69 ± 0.01 3.8% ± 1.1% 5.28 ± 0.09 Pultusk H5 S3 Fall NMNH 1117 A 61.655 3.51 ± 0.03 3.69 ± 0.01 5.0% ± 0.9% 5.28 ± 0.09 Pultusk H5 S3 Fall NMNH 1117 B 22.586 3.53 ± 0.05 3.69 ± 0.02 4.3% ± 1.4% 5.25 ± 0.08 Pultusk H5 S3 Fall NMNH 1615 93.678 3.40 ± 0.03 3.70 ± 0.01 8.1% ± 0.8% 5.28 ± 0.09 Pultusk H5 S3 Fall NMNH 2560 30.509 3.57 ± 0.03 3.70 ± 0.02 3.6% ± 1.0% 5.26 ± 0.08 Pultusk H5 S3 Fall NMNH 3003 80.504 3.52 ± 0.03 3.69 ± 0.01 4.6% ± 0.8% 5.32 ± 0.09 Pultusk H5 S3 Fall NMNH 3004 53.506 3.28 ± 0.03 3.54 ± 0.01 7.4% ± 0.9% 4.89 ± 0.09 Pultusk H5 S3 Fall Vatican 769 8.14 3.75 ± 0.10 3.79 ± 0.04 0.9% ± 2.8% 4.57 ± 0.08† Pultusk H5 S3 Fall Vatican 770 8.201 3.57 ± 0.08 3.80 ± 0.04 6.1% ± 2.3% 5.25 ± 0.08† Pultusk H5 S3 Fall Vatican 762 19.98 3.46 ± 0.03 3.83 ± 0.04 9.6% ± 1.2% 5.35 ± 0.08 Pultusk H5 S3 Fall Vatican 767 8.35 3.66 ± 0.10 3.81 ± 0.04 3.9% ± 2.7% 5.28 ± 0.08† Pultusk H5 S3 Fall Vatican 771 5.81 3.36 ± 0.10 3.82 ± 0.04 12.1% ± 2.7% 5.32 ± 0.08† Pultusk H5 S3 Fall Vatican 514 9.653 3.63 ± 0.10 3.77 ± 0.04 3.8% ± 2.7% 5.35 ± 0.08† Pultusk H5 S3 Fall Vatican 761 20.07 3.52 ± 0.03 3.72 ± 0.04 5.5% ± 1.2% 5.17 ± 0.08 Pultusk H5 S3 Fall Vatican 772 4.78 3.57 ± 0.13 3.89 ± 0.19 8.2% ± 5.7% 5.39 ± 0.08† Pultusk H5 S3 Fall Vatican 768 8.18 3.22 ± 0.06 3.54 ± 0.04 9.1% ± 2.0% 5.40 ± 0.08† Pultusk H5 S3 Fall Vatican 763 13.04 3.77 ± 0.08 3.78 ± 0.04 0.3% ± 2.3% 5.42 ± 0.08† Pultusk H5 S3 Fall Vatican 766 8.45 3.37 ± 0.06 3.81 ± 0.09 11.6% ± 2.5% 5.45 ± 0.08†

258

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Pultusk H5 S3 Fall Vatican 760 63.48 3.43 ± 0.07 3.85 ± 0.04 10.9% ± 2.1% n.d. Pultusk H5 S3 Fall Vatican 759 101.83 3.41 ± 0.03 3.80 ± 0.04 10.1% ± 1.3% 5.39 ± 0.09 Pultusk H5 S3 Fall Vatican 758 186.12 3.50 ± 0.03 n.d. n.d. 5.39 ± 0.09 Pultusk H5 S3 Fall Vatican 757 510.18 3.39 ± 0.07‡ n.d. n.d. n.d. Pultusk H5 S3 Fall Average 3.44 3.72 7.5% 5.27 Quenggouk H4 Fall Vatican 783 51.72 3.24 ± 0.02 3.81 ± 0.04 15.1% ± 1.0% 5.23 ± 0.09 Ranchapur H4 Fall LNHM BM1925,441 64.95 3.45 ± 0.03 3.61 ± 0.01 4.3% ± 1.0% 5.22 ± 0.10 Rancho de la Presaa H5 Fall CMS 98.1x 16.9 3.48 ± 0.19 3.64 ± 0.03 4.3% ± 5.3% 5.27 ± 0.12 Richardtonb H5 S2 W2 Fall NMNH 3362 34.504 3.02 ± 0.03 3.70 ± 0.02 18.3% ± 1.0% 5.24 ± 0.10 Richardton H5 S2 W2 Fall LNHM BM1937,1390 65.41 3.13 ± 0.05 3.71 ± 0.01 15.7% ± 1.4% 5.38 ± 0.10 Richardton H5 S2 W2 Fall Average 3.09 3.71 16.6% 5.31 Searsmont H5 Fall CMS 105s 25.8 3.24 ± 0.05 3.72 ± 0.02 12.9% ± 1.4% 5.25 ± 0.08 Senaa H4 Fall CMS 785 26.97 2.61 ± 0.03 3.55 ± 0.03 26.6% ± 1.1% 5.44 ± 0.09 Sharps H3.4 S3 Fall NMNH 640 B 32.932 3.47 ± 0.05 3.60 ± 0.02 3.6% ± 1.3% 4.88 ± 0.12 Shupiyana H6 Fall LNHM BM1915,143 65.31 3.35 ± 0.06 3.71 ± 0.01 9.8% ± 1.6% 5.29 ± 0.10 Simmern H6 Fall CMS 591.1 65.66 3.44 ± 0.03 3.72 ± 0.01 7.5% ± 0.9% 5.31 ± 0.09 St. Germain-du-Pinel H6 S3 Fall Vatican 900 27.295 3.33 ± 0.06 3.76 ± 0.04 11.5% ± 1.8% 5.33 ± 0.12 Ställdalena H5 S3 Fall CMS 185a 51.77 3.64 ± 0.04 3.70 ± 0.02 1.7% ± 1.2% 5.44 ± 0.09 Ställdalen H5 S3 Fall Vatican 906 10.08 n.d. n.d. n.d. 5.35 ± 0.08† Ställdalen H5 S3 Fall Average 3.64 3.70 1.7% 5.40 Supuheea H6 Fall Vatican 916 278.43 3.55 ± 0.03 n.d. n.d. 5.35 ± 0.10 Thuathe H4 S2-3 Fall Vatican 1448 24.792 3.55 ± 0.04 3.84 ± 0.04 7.5% ± 1.5% 5.37 ± 0.08 Timochin H5 S3 Fall LNHM BM373 42.3 2.84 ± 0.03 3.65 ± 0.01 22.1% ± 0.9% 5.22 ± 0.10 Timochin H5 S3 Fall LNHM BM35183 36.76 3.25 ± 0.04 3.71 ± 0.02 12.5% ± 1.1% 5.37 ± 0.10 Timochin H5 S3 Fall Vatican 929 27.08 3.30 ± 0.03 3.80 ± 0.04 13.2% ± 1.2% 5.35 ± 0.08 Timochin H5 S3 Fall Average 3.09 3.71 16.8% 5.31 Tirupati H6 Fall CMS 616.1 30.24 3.24 ± 0.04 3.76 ± 0.02 13.8% ± 1.1% 5.51 ± 0.08 Tjabe H6 Fall CMS 1144 22.85 3.30 ± 0.04 3.68 ± 0.02 10.2% ± 1.3% 5.47 ± 0.08 Torino H6 Fall Vatican 1396 66.24 3.49 ± 0.03 3.86 ± 0.04 9.7% ± 1.3% n.d. Torino H6 Fall Vatican 962 94 3.32 ± 0.07 n.d. n.d. n.d. Torino H6 Fall Vatican 961 107.81 3.30 ± 0.07 3.83 ± 0.04 13.8% ± 1.9% n.d. Torino H6 Fall Average 3.35 3.84 12.8% n.d. Torrington H6 Fall CMS 510.1x 47.3 3.50 ± 0.04 3.72 ± 0.01 5.8% ± 1.1% 5.35 ± 0.10 Toulouseb H6 Fall Vatican 966 34.737 3.65 ± 0.03 3.71 ± 0.04 1.5% ± 1.3% 5.18 ± 0.08 Toulouse H6 Fall Vatican 967 13.37 n.d. n.d. n.d. 5.34 ± 0.08† Toulouse H6 Fall Average 3.65 3.71 1.5% 5.26 Trenzanob H3/4 Fall LNHM BM1985,M49 23.83 3.08 ± 0.06 3.71 ± 0.02 16.9% ± 1.6% 5.37 ± 0.08 Trenzano H3/4 Fall Vatican 973 11.71 n.d. n.d. n.d. 5.59 ± 0.08† Trenzano H3/4 Fall Average 3.08 3.71 16.9% 5.48 Tysnes Islanda H4 Fall Vatican 976 11.62 n.d. n.d. n.d. 5.24 ± 0.08† Uberabab H5 S3 Fall Vatican 978 24.91 3.33 ± 0.03 3.83 ± 0.04 13.0% ± 1.1% 5.29 ± 0.10 Uberaba H5 S3 Fall Vatican 977 89.74 3.32 ± 0.04 n.d. n.d. 5.32 ± 0.09 Uberaba H5 S3 Fall Average 3.32 3.83 13.3% 5.31 Udipib H5 Fall LNHM BM1985,M50 79.63 3.36 ± 0.03 3.68 ± 0.01 8.7% ± 0.9% 5.44 ± 0.09 Weston H4 Fall LNHM BM1985,M52 60.79 3.24 ± 0.03 3.73 ± 0.01 13.1% ± 0.9% 5.24 ± 0.10 Weston H4 Fall Vatican 1023 14.85 n.d. n.d. n.d. 5.23 ± 0.08† Weston H4 Fall Vatican 1022 74.821 3.20 ± 0.03 3.79 ± 0.04 15.4% ± 1.2% 5.18 ± 0.09 Weston H4 Fall Average 3.22 3.76 14.4% 5.22 Yatoor H5 Fall CMS 259s 51.31 3.26 ± 0.05 3.69 ± 0.02 11.6% ± 1.5% 5.42 ± 0.09 Yonozu H4/5 Fall IOM C 49.5 24.81 3.14 ± 0.03 3.48 ± 0.02 9.7% ± 1.0% 5.07 ± 0.10 Zaga H3-6 S3 W0/1 Fall IOM C 350.1 78.67 3.54 ± 0.03 3.73 ± 0.01 5.0% ± 0.9% 5.29 ± 0.09 Zag H3-6 S3 W0/1 Fall IOM C 350.2 33.66 3.58 ± 0.04 3.71 ± 0.02 3.6% ± 1.2% 5.30 ± 0.08 Zag H3-6 S3 W0/1 Fall Average 3.56 3.73 4.6% 5.29 Zaoyang H5 Fall LNHM BM1999,M16 28.07 3.09 ± 0.03 3.72 ± 0.02 17.0% ± 1.0% 5.40 ± 0.08 Zebrak H5 Fall CMS 860 23.04 3.23 ± 0.05 3.67 ± 0.02 12.0% ± 1.5% 5.32 ± 0.12 Zebrak H5 Fall LNHM BM76153 75.53 3.27 ± 0.03 3.69 ± 0.01 11.4% ± 0.7% 5.35 ± 0.09 Zebrak H5 Fall Average 3.26 3.69 11.5% 5.33 Zhovtnevyi H5 S3 Fall CMS 743a 61.04 3.35 ± 0.03 3.64 ± 0.01 7.9% ± 1.0% 5.24 ± 0.09 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study. aBreccia (Grady, 2000) bVeined (Grady, 2000)

259

Table 5: Data for H Chondrite Finds.

Bulk Grain Model Magnetic Weathering Meteorite Type Fall Collection Catalog Mass Density Density Porosity Porosity Susceptibility Modulus (g) (g cm-3) (g cm-3) (log χ) Abbott a H6 S3 Find AMNH 4748 86.42 3.56 ± 0.01 3.61 ± 0.01 1.5% ± 0.4% 4.8% 5.13 ± 0.10 1.03 Abbott H6 S3 Find IOM C 75.38 50.86 3.42 ± 0.04 3.50 ± 0.01 2.1% ± 1.1% 7.9% 4.91 ± 0.09 2.32 Abbott H6 S3 Find IOM C 75.39 42.68 3.30 ± 0.03 3.48 ± 0.02 5.2% ± 1.1% 10.8% 4.85 ± 0.09 2.63 Abbott H6 S3 Find Average 3.46 3.55 2.6% 7.1% 4.97 - Acfer 028 H3.8 Find AMNH 4806 14.8 3.30 ± 0.06 3.54 ± 0.03 6.8% ± 2.0% 10.8% 5.13 ± 0.08 1.35 Acfer 166 H3-5 Find AMNH 4809 21.6 3.65 ± 0.05 3.62 ± 0.02 -0.7% ± 1.6% 2.7% 5.21 ± 0.10 0.67 Acme H5 Find AMNH 3990 A 5.64 4.21 ± 0.35 3.35 ± 0.04 -25.6% ± 10.5% -10.0% 4.74 ± 0.08 3.54 Acme H5 Find AMNH 3990 B 4.97 3.67 ± 0.31 3.41 ± 0.05 -7.6% ± 9.3% 2.2% 4.83 ± 0.08 2.97 Acme H5 Find LNHM BM1959,1009 29.63 3.32 ± 0.04 3.39 ± 0.02 2.0% ± 1.3% 10.3% 4.64 ± 0.10 3.82 Acme H5 Find Average 3.46 3.39 -2.3% 7.0% 4.74 - Alamogordo H5 S3 Find NMNH 5853 20.048 3.51 ± 0.07 3.52 ± 0.02 0.1% ± 2.2% 5.8% 5.11 ± 0.12 1.50 Alamogordo H5 S3 Find IOM C 11.2 51.39 3.48 ± 0.08 3.59 ± 0.02 3.1% ± 2.3% 6.5% 5.27 ± 0.09 0.71 Alamogordo H5 S3 Find Average 3.49 3.57 2.2% 6.3% 5.19 - Ashmore H5 S3 Find NMNH 5597 14.211 3.68 ± 0.08 3.70 ± 0.04 0.6% ± 2.4% 2.0% 5.18 ± 0.12 0.59 Ashmore H5 S3 Find IOM C 169.2d 21.89 3.16 ± 0.04 3.31 ± 0.02 4.3% ± 1.4% 14.0% 4.75 ± 0.08 3.68 Ashmore H5 S3 Find Average 3.35 3.45 3.0% 9.6% 4.96 - Attica H4 S1 Find NMNH 7006 26.738 3.20 ± 0.26 3.45 ± 0.02 7.1% ± 7.4% 13.0% 4.94 ± 0.10 2.38 Belmontb H6 S3 Find IOM C 40.3c1 28.6 3.52 ± 0.05 3.53 ± 0.02 0.2% ± 1.6% 5.6% 5.11 ± 0.08 1.43 Burdett H5 Find Vatican 153 18.101 3.44 ± 0.03 3.53 ± 0.04 2.7% ± 1.3% 7.6% 5.14 ± 0.12 1.31 Chico Hills H4 S4 Find IOM C 77.40 41.99 3.42 ± 0.04 3.53 ± 0.02 2.9% ± 1.1% 7.9% 5.01 ± 0.10 1.82 Chico Hills H4 S4 Find IOM C 77.43 46.59 3.69 ± 0.04 3.58 ± 0.01 -3.1% ± 1.3% 1.8% 5.04 ± 0.10 1.54 Chico Hills H4 S4 Find IOM C 77.42 37.15 3.45 ± 0.04 3.52 ± 0.02 1.9% ± 1.3% 7.3% 4.84 ± 0.10 2.60 Chico Hills H4 S4 Find Average 3.52 3.54 0.5% 5.6% 4.96 - Clovis (no. 1) H3.6 Find Vatican 221 33.26 3.44 ± 0.02 3.50 ± 0.03 1.7% ± 1.2% 7.6% 4.87 ± 0.12 2.51 Cobija H6 Find NMNH 419 69.065 3.40 ± 0.04 3.50 ± 0.01 2.8% ± 1.1% 8.4% 4.97 ± 0.10 2.06 Correo H4 Find LNHM BM1986,M11 24.04 3.19 ± 0.05 3.47 ± 0.02 8.0% ± 1.5% 13.4% 4.81 ± 0.10 2.86 Daraj 114 H4 Find LNHM BM1988,M53 66.4 3.34 ± 0.03 3.43 ± 0.01 2.7% ± 0.9% 9.8% 4.59 ± 0.10 3.93 Davy (b) H4 S2 W2 Find NMNH 6863 B 20.953 3.47 ± 0.11 3.45 ± 0.03 -0.6% ± 3.3% 6.8% 4.91 ± 0.12 2.51 Dimmitta H4 S3 W2 Find IOM C 10.1 34.96 3.50 ± 0.04 3.42 ± 0.02 -2.2% ± 1.3% 6.2% 4.80 ± 0.12 3.05 H4 S3 W2 Find Vatican 270 113.68 3.42 ± 0.03 3.56 ± 0.04 3.8% ± 1.4% 7.9% 4.96 ± 0.10 1.92 Dimmitt H4 S3 W2 Find Average 3.44 3.52 2.4% 7.5% 4.88 - Doyleville H5 Find LNHM BM1959,857 36.29 3.44 ± 0.04 3.68 ± 0.02 6.3% ± 1.2% 7.4% 5.26 ± 0.09 0.29 El Hammami H5 S2 Find IOM C 346.2 32.62 3.40 ± 0.05 3.79 ± 0.02 10.2% ± 1.3% 8.4% 5.37 ± 0.08 0.60 Elm Creek H4 Find Vatican 290 52.86 n.d. n.d. n.d. n.d. 5.03 ± 0.08† n.d. Estacado H6 S1 W1 Find Vatican 311 13.84 n.d. n.d. n.d. n.d. 5.54 ± 0.08† n.d. Estacado H6 S1 W1 Find Vatican 309 142.88 3.59 ± 0.04 n.d. n.d. 4.1% 5.61 ± 0.09 n.d. Estacado H6 S1 W1 Find Vatican 310 114.26 3.57 ± 0.04 n.d. n.d. 4.5% 5.57 ± 0.09 n.d. Estacado H6 S1 W1 Find Average 3.58 n.d. n.d. 4.3% 5.57 - Faucett H5 S4 Find NMNH 5689 14.728 3.47 ± 0.07 3.43 ± 0.05 -1.2% ± 2.4% 6.7% 5.14 ± 0.12 1.86 Faucett H5 S4 Find IOM C 90.1 19.64 3.61 ± 0.09 3.57 ± 0.03 -0.9% ± 2.6% 3.7% 5.22 ± 0.12 0.91 Faucett H5 S4 Find Average 3.55 3.51 -1.1% 5.0% 5.18 - Franconia H5 S3 W2 Find CMS 1511.1 31.74 3.43 ± 0.05 3.58 ± 0.02 4.3% ± 1.7% 7.7% 5.24 ± 0.12 0.81 Gilgoin H5 S4 Find NMNH 509 A 28.703 3.56 ± 0.05 3.57 ± 0.02 0.4% ± 1.5% 4.8% 5.19 ± 0.10 0.99 Goose Creek H5 S1 W5 Find CMS 1422 29.65 2.86 ± 0.05 3.19 ± 0.02 10.2% ± 1.7% 21.2% 4.62 ± 0.12 4.62 Gruver H4 Find Vatican 379 11.44 n.d. n.d. n.d. n.d. 4.93 ± 0.08† n.d. HaH 019 H6 S2 W3 Find CMS 1533D 28.07 3.28 ± 0.03 3.36 ± 0.02 2.4% ± 0.9% 11.3% 4.41 ± 0.10 4.91 Hartsel H4 S3 W2 Find CMS 1550 20.77 3.33 ± 0.04 3.57 ± 0.02 6.5% ± 1.4% 10.0% 5.24 ± 0.12 0.91 Howe H5 S3 Find IOM C 19.1&2 26.25 3.54 ± 0.07 3.48 ± 0.02 -1.8% ± 2.1% 5.1% 5.13 ± 0.08 1.63 Kimbolton H4 Find LNHM BM1988,M48 24.45 3.41 ± 0.08 3.43 ± 0.02 0.7% ± 2.4% 8.2% 4.88 ± 0.08 2.69 La Villa H4 S1 Find IOM C 255.1 14.62 3.29 ± 0.05 3.42 ± 0.03 3.7% ± 1.8% 11.0% 5.08 ± 0.08 2.09 Landreth Draw H5 S2 W2 Find NMNH 6978 2 12.854 3.67 ± 0.09 3.63 ± 0.04 -1.1% ± 2.9% 2.2% 5.17 ± 0.12 0.80 Little Spring Creek H5 S1 W1 Find CMS 1551 20.2 3.57 ± 0.07 3.68 ± 0.03 3.1% ± 1.9% 4.6% 5.21 ± 0.12 0.50 Mayfield H5 S3 Find Vatican 1062 19.31 3.40 ± 0.04 3.61 ± 0.04 5.8% ± 1.4% 8.4% 4.98 ± 0.12 1.73 Miami H5 S2 Find CMS 399.1x 26.95 3.32 ± 0.05 3.37 ± 0.03 1.4% ± 1.6% 10.4% 4.62 ± 0.08 3.97 NWA 130 H3.7 S2 W3 Find CMS 1431 28.56 3.25 ± 0.07 3.38 ± 0.02 3.8% ± 2.1% 12.0% 4.89 ± 0.12 2.87 NWA 3130 H5 S1 W2-3 Find CMS 1504 17.35 3.57 ± 0.05 3.46 ± 0.03 -3.1% ± 1.6% 4.5% 4.94 ± 0.10 2.35 Oakley (stone) H6 Find Vatican 703 117.06 3.67 ± 0.04 n.d. n.d. 2.2% 5.21 ± 0.12 n.d.

260

Bulk Grain Model Magnetic Weathering Meteorite Type Fall Collection Catalog Mass Density Density Porosity Porosity Susceptibility Modulus (g) (g cm-3) (g cm-3) (log χ) Octave Mine H5 S1 W3 Find CMS 1566 22.29 3.40 ± 0.04 3.39 ± 0.02 -0.4% ± 1.4% 8.3% 4.82 ± 0.10 3.09 Oczeretnab H4 Find LNHM BM48368 85.06 3.41 ± 0.03 3.70 ± 0.01 7.8% ± 0.9% 8.1% 5.21 ± 0.10 0.43 O'Donnell H5 S2 W2 Find CMS 1470-2 19.33 3.43 ± 0.09 3.39 ± 0.04 -1.2% ± 2.7% 7.6% 4.94 ± 0.12 2.63 Orlovka H5 Find LNHM BM1933,272 52.92 3.52 ± 0.04 3.60 ± 0.01 2.3% ± 1.1% 5.6% 5.15 ± 0.09 1.00 Ovid (B) H5 S1 W2 Find CMS 1562 25.17 3.68 ± 0.05 3.54 ± 0.03 -3.7% ± 1.6% 2.1% 5.07 ± 0.12 1.53 Ozona H6 Find LNHM BM1953,154 54.44 3.28 ± 0.03 3.39 ± 0.01 3.4% ± 1.1% 11.3% 4.84 ± 0.09 2.98 Pipe Creek H6 Find Vatican 746 25.257 3.37 ± 0.02 3.63 ± 0.04 7.2% ± 1.1% 9.2% 5.45 ± 0.12 0.90 Pipe Creek H6 Find Vatican 744 100.35 3.45 ± 0.03 n.d. n.d. 7.4% 5.55 ± 0.10 n.d. Pipe Creek H6 Find Vatican 745 79.26 3.43 ± 0.03 n.d. n.d. 7.8% 5.54 ± 0.12 n.d. Pipe Creek H6 Find Average 3.43 3.63 5.6% 7.8% 5.52 - Plainview (1917) a H5 S3 Find IOM C 33.2 78.54 3.51 ± 0.06 3.64 ± 0.01 3.7% ± 1.7% 5.9% 5.27 ± 0.10 0.42 Plainview (1917) H5 S3 Find IOM C 33.3a 62.88 3.49 ± 0.07 3.64 ± 0.01 4.1% ± 1.9% 6.3% 5.24 ± 0.12 0.54 Plainview (1917) H5 S3 Find Average 3.50 3.64 3.9% 6.1% 5.25 - Prairie Dog Creek H3.8 Find Vatican 754 154.15 3.32 ± 0.03 n.d. n.d. 10.4% 4.87 ± 0.09 n.d. Sacramento Wash 001 H4 S2 W3 Find CMS 1546 28.13 3.15 ± 0.04 3.32 ± 0.02 5.2% ± 1.4% 14.3% 4.63 ± 0.10 4.06 Saline H5 Find Vatican 816 30.82 3.56 ± 0.02 3.69 ± 0.04 3.4% ± 1.2% 4.7% 5.16 ± 0.12 0.73 Saline H5 Find Vatican 815 219.57 3.56 ± 0.03 n.d. n.d. 4.8% 5.41 ± 0.08 n.d. Saline H5 Find Average 3.56 3.69 3.5% 4.8% 5.28 - Tatum H4 S2 W4 Find IOM C 264.1&2 30.34 3.05 ± 0.07 3.21 ± 0.01 5.1% ± 2.3% 16.8% 4.45 ± 0.09 5.20 Texline H5 S4 WA Find IOM C 39.1 29.8 3.44 ± 0.10 3.65 ± 0.03 5.8% ± 2.8% 7.6% 5.37 ± 0.12 0.50 Tolar H4 S2 W5 Find CMS 1510 54.68 3.26 ± 0.04 3.34 ± 0.01 2.2% ± 1.3% 11.6% 4.55 ± 0.10 4.37 Tomhannock Creeka H5 Find Vatican 957 15.67 n.d. n.d. n.d. n.d. 5.18 ± 0.08† n.d. Travis County (a) H5 S4 W3 Find CMS 225ax 44.61 3.41 ± 0.06 3.49 ± 0.02 2.2% ± 1.8% 8.1% 5.15 ± 0.12 1.51 Travis County (a) H5 S4 W3 Find Vatican 969 71.11 3.45 ± 0.04 n.d. n.d. 7.2% 5.22 ± 0.12 n.d. Travis County (a) H5 S4 W3 Find Average 3.44 3.49 1.5% 7.6% 5.19 - Travis County (b) H4 S2 W3 Find NMNH 6871 65.127 3.41 ± 0.04 3.57 ± 0.01 4.5% ± 1.2% 8.3% 5.12 ± 0.10 1.24 Two Buttes H5 Find FMNH ME 2633 #1 31.08 3.29 ± 0.06 3.46 ± 0.02 4.7% ± 1.9% 10.9% 4.78 ± 0.12 3.01 Wairarapa Valley H5 Find FMNH ME 1882 #1 62.03 3.55 ± 0.04 3.57 ± 0.01 0.8% ± 1.1% 5.0% 5.10 ± 0.10 1.30 Warm Springs Wilderness H4-6 S2 W1 Find CMS 1545 35.23 3.52 ± 0.04 3.62 ± 0.02 2.9% ± 1.2% 5.6% 5.08 ± 0.10 1.22 Wikieup H5 Find FMNH ME 2870 #1 25.89 3.25 ± 0.07 3.59 ± 0.03 9.4% ± 2.0% 11.9% 5.22 ± 0.12 0.83 Witchelina H4 Find FMNH ME 2702 #1 18.15 3.36 ± 0.05 3.49 ± 0.03 3.7% ± 1.6% 9.4% 4.70 ± 0.12 3.28 Wray H4 Find FMNH ME 2387 #1 24.43 3.50 ± 0.05 3.63 ± 0.02 3.5% ± 1.4% 6.1% 5.23 ± 0.12 0.59 Wray (a) H4 Find LNHM BM1959,905 60.85 3.48 ± 0.04 3.62 ± 0.01 3.8% ± 1.1% 6.5% 5.27 ± 0.12 0.56 Wynella H4 Find FMNH ME 2724 #2 95.69 3.38 ± 0.06 3.46 ± 0.01 2.2% ± 1.6% 8.8% 4.87 ± 0.09 2.61 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. aBreccia (Grady, 2000) bVeined (Grady, 2000)

261

Table 6: Data for L Chondrite Falls.

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Achiras L6 Fall AMNH 3933 A 26.84 3.48 ± 0.04 3.58 ± 0.02 3.1% ± 1.1% 4.92 ± 0.12 Achiras L6 Fall AMNH 3933 B 17.72 3.50 ± 0.05 3.53 ± 0.04 0.8% ± 1.8% 4.88 ± 0.12 Achiras L6 Fall Average 3.48 3.56 2.2% 4.90 Aguada L6 Fall AMNH 3934 A 52.85 3.25 ± 0.06 3.54 ± 0.01 8.1% ± 1.6% 4.79 ± 0.10 Aguada L6 Fall AMNH 3934 B 21.67 3.44 ± 0.04 3.56 ± 0.02 3.4% ± 1.3% 4.72 ± 0.08 Aguada L6 Fall Average 3.30 3.54 6.8% 4.76 Aïr L6 Fall CMS 1326 36.69 3.38 ± 0.05 3.50 ± 0.01 3.4% ± 1.4% 4.96 ± 0.12 Alfianello L6 S5 W2 Fall NMNH 2210 23.593 3.26 ± 0.04 3.54 ± 0.02 7.8% ± 1.4% 4.88 ± 0.08 Alfianello L6 S5 W2 Fall IOM C 61.1 33.69 3.21 ± 0.04 3.54 ± 0.02 9.3% ± 1.1% 5.03 ± 0.09 Alfianello L6 S5 W2 Fall Vatican 20 66.582 3.27 ± 0.03 3.59 ± 0.04 8.9% ± 1.3% 4.78 ± 0.10 Alfianello L6 S5 W2 Fall Vatican 19 152.91 3.25 ± 0.03 n.d. n.d. 4.93 ± 0.09 Alfianello L6 S5 W2 Fall Average 3.25 3.56 8.7% 4.91 Andover L6 Fall LNHM BM86762 19.78 3.12 ± 0.07 3.59 ± 0.03 13.1% ± 1.9% 5.03 ± 0.08 L6 Fall AMNH 1049 18.86 2.98 ± 0.04 3.51 ± 0.03 15.0% ± 1.2% 4.25 ± 0.08 Aptb L6 S4 Fall Vatican 37 50.208 3.31 ± 0.02 3.61 ± 0.04 8.4% ± 1.0% 4.99 ± 0.09 Atoka L6 Fall Monnig M 40.1 36.76 3.35 ± 0.03 3.58 ± 0.02 6.5% ± 1.1% 5.02 ± 0.09 Atoka L6 Fall Vatican 1085 17.43 3.21 ± 0.03 n.d. n.d. 4.89 ± 0.08 Atoka L6 Fall Average 3.30 3.58 7.8% 4.95 Aumaleb L6 S4 Fall AMNH 407 56.11 3.35 ± 0.02 3.55 ± 0.01 5.9% ± 0.7% 5.14 ± 0.09 Aumale L6 S4 Fall Vatican 60 121.51 3.36 ± 0.03 n.d. n.d. 4.85 ± 0.09 Aumale L6 S4 Fall Average 3.35 3.55 5.6% 4.99 Aumièresb L6 Fall AMNH 652 45.37 3.23 ± 0.02 3.53 ± 0.02 8.6% ± 0.7% 4.66 ± 0.09 Aumières L6 Fall Vatican 62 72.34 3.19 ± 0.02 3.61 ± 0.04 11.6% ± 1.0% 4.68 ± 0.09 Aumières L6 Fall Vatican 64 22.3 3.21 ± 0.03 3.59 ± 0.04 10.5% ± 1.3% 4.69 ± 0.08 Aumières L6 Fall Vatican 65 15.72 3.15 ± 0.03 n.d. n.d. 4.70 ± 0.08 Aumières L6 Fall Average 3.20 3.58 10.6% 4.68 Ausson L5 S2 Fall AMNH 1051 13.88 3.25 ± 0.07 3.54 ± 0.05 8.3% ± 2.4% 5.47 ± 0.08 Ausson L5 S2 Fall NMNH 2645 19.805 3.21 ± 0.03 3.55 ± 0.03 9.6% ± 1.2% 5.01 ± 0.08 Ausson L5 S2 Fall Vatican 69 48.397 3.21 ± 0.02 3.69 ± 0.04 12.8% ± 1.0% 4.76 ± 0.20 Ausson L5 S2 Fall Vatican 67 104.026 3.23 ± 0.03 3.66 ± 0.04 11.9% ± 1.2% 4.96 ± 0.09 Ausson L5 S2 Fall Vatican 70 21.996 3.20 ± 0.03 3.63 ± 0.04 11.9% ± 1.2% 4.94 ± 0.08 Ausson L5 S2 Fall Average 3.22 3.65 11.6% 5.03 Aztec L6 Fall Monnig M 855.1 28.71 3.42 ± 0.04 3.59 ± 0.02 4.9% ± 1.2% 4.95 ± 0.08 Bachmut L6 Fall AMNH 409 22.31 3.30 ± 0.04 3.52 ± 0.02 6.3% ± 1.3% 5.03 ± 0.08 Bald Mountainb L4 Fall CMS 302.1x 84 3.33 ± 0.03 3.56 ± 0.01 6.5% ± 0.9% 5.17 ± 0.10 Baldwynb L6 Fall AMNH 4830 28.55 3.39 ± 0.05 3.56 ± 0.02 4.6% ± 1.5% 4.78 ± 0.08 Barwell L5 S3 Fall NMNH 2612 36.958 3.36 ± 0.08 3.60 ± 0.02 6.6% ± 2.2% 4.87 ± 0.10 Bath Furnace L6 Fall AMNH 616 49.84 3.41 ± 0.05 3.49 ± 0.01 2.4% ± 1.6% 5.14 ± 0.09 Bath Furnace L6 Fall Vatican 93 44.194 3.28 ± 0.04 3.62 ± 0.04 9.4% ± 1.4% 4.96 ± 0.12 Bath Furnace L6 Fall Vatican 92 45 3.86 ± 0.04 3.90 ± 0.04 0.9% ± 1.4% n.d. Bath Furnace L6 Fall Average 3.50 3.66 4.3% 5.05 Berlanguillasb L6 Fall Vatican 109 10.91 n.d. n.d. n.d. 4.94 ± 0.08† Beustea L5 Fall LNHM BM64340 33.41 3.26 ± 0.04 3.60 ± 0.02 9.3% ± 1.3% 5.10 ± 0.10 Blanket L6 Fall CMS 203bx 81.9 3.29 ± 0.03 3.52 ± 0.01 6.6% ± 0.9% 4.99 ± 0.09 Bocas L6 Fall LNHM BM92564 32.98 3.19 ± 0.04 3.57 ± 0.02 10.5% ± 1.2% 4.84 ± 0.09 Borib L6 Fall Vatican 132 23.564 3.27 ± 0.04 3.61 ± 0.04 9.5% ± 1.3% 4.99 ± 0.08 L3 Fall Monnig M1172.1 32.37 3.47 ± 0.04 3.59 ± 0.02 3.6% ± 1.4% 5.06 ± 0.12 Bruderheim L6 S4 Fall NMNH 2081 24.573 3.33 ± 0.04 3.55 ± 0.02 6.1% ± 1.2% 5.12 ± 0.08 Bruderheim L6 S4 Fall IOM C 50.11 61.62 3.38 ± 0.03 3.58 ± 0.01 5.6% ± 1.0% 4.97 ± 0.10 Bruderheim L6 S4 Fall Vatican 151 11.51 n.d. n.d. n.d. 4.98 ± 0.08† Bruderheim L6 S4 Fall Average 3.36 3.57 5.7% 5.02 Buschhofb L6 Fall CMS 958s 55.3 3.17 ± 0.03 3.54 ± 0.01 10.5% ± 1.0% 4.80 ± 0.09 Cabezo de Mayoa L6 Fall Vatican 158 61.72 3.26 ± 0.02 3.60 ± 0.04 9.6% ± 1.0% 4.81 ± 0.09 Çanakkale L6 Fall CMS 731 34.05 3.29 ± 0.04 3.51 ± 0.02 6.3% ± 1.3% 4.71 ± 0.09 Chadong L6 S3 W0/1 Fall Monnig M1103.1 50.2 3.26 ± 0.03 3.62 ± 0.02 10.1% ± 0.9% 5.07 ± 0.09 Chandakapura L5 Fall CMS 729s 118.67 3.30 ± 0.05 3.51 ± 0.01 6.1% ± 1.5% 4.77 ± 0.10 Chantonnaya L6 S6 Fall Vatican 201 25.052 3.46 ± 0.04 3.55 ± 0.04 2.4% ± 1.5% 4.96 ± 0.08 Château-Renardb L6 Fall LNHM BM1920,295 59.78 3.30 ± 0.03 3.58 ± 0.01 7.8% ± 0.9% 4.99 ± 0.09 Château-Renard L6 Fall Vatican 211 46.78 3.37 ± 0.03 3.57 ± 0.04 5.8% ± 1.3% 5.01 ± 0.09 Château-Renard L6 Fall Average 3.33 3.58 6.9% 5.00

262

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Chervettaz L5 Fall LNHM BM86761 25.97 3.02 ± 0.04 3.57 ± 0.02 15.5% ± 1.2% 4.80 ± 0.10 Chitado L6 Fall CMS 1239-1 17.66 3.08 ± 0.06 3.51 ± 0.03 12.2% ± 1.9% 4.85 ± 0.10 Cilimus L5 Fall LNHM BM1980,M16 45.59 3.34 ± 0.04 3.63 ± 0.02 8.0% ± 1.1% 4.95 ± 0.09 Clohars L4 Fall LNHM BM1960,330 35.77 3.27 ± 0.03 3.50 ± 0.01 6.7% ± 1.0% 4.70 ± 0.10 Colby (Wisconsin) b L6 S3 Fall NMNH 618 21.819 3.38 ± 0.04 3.56 ± 0.02 5.0% ± 1.3% 4.98 ± 0.08 Dandapurb L6 Fall CMS 440 21.55 3.15 ± 0.04 3.49 ± 0.02 9.9% ± 1.3% 4.73 ± 0.08 Dandapur L6 Fall LNHM BM53321 69.52 3.38 ± 0.03 3.49 ± 0.01 3.3% ± 1.0% 4.96 ± 0.09 Dandapur L6 Fall Average 3.32 3.49 4.9% 4.85 Demina L6 Fall LNHM BM1956,320 39.27 3.26 ± 0.05 3.52 ± 0.01 7.4% ± 1.4% 4.91 ± 0.09 Drake Creeka,b L6 Fall Monnig M 624.3 44.76 3.18 ± 0.03 3.60 ± 0.01 11.6% ± 0.9% 4.59 ± 0.10 Duralab L6 Fall Vatican 1130 31.801 3.29 ± 0.03 3.62 ± 0.04 9.2% ± 1.3% 4.90 ± 0.09 Elenovka L5 S2 W0 Fall NMNH 5952 15.783 3.06 ± 0.04 3.62 ± 0.03 15.6% ± 1.3% 4.87 ± 0.08 Elenovka L5 S2 W0 Fall IOM C 26.5 22.42 3.02 ± 0.05 3.56 ± 0.02 15.3% ± 1.6% 4.89 ± 0.12 Elenovka L5 S2 W0 Fall Average 3.03 3.59 15.4% 4.88 Ergheo L5 S3 Fall Vatican 303 107.61 3.35 ± 0.03 n.d. n.d. 4.17 ± 0.10 Ergheo L5 S3 Fall Vatican 305 17.301 3.50 ± 0.08 3.50 ± 0.03 0.1% ± 2.4% 4.81 ± 0.10 Ergheo L5 S3 Fall Vatican 304 32.681 3.50 ± 0.03 3.49 ± 0.03 -0.2% ± 1.4% 4.68 ± 0.12 Ergheo L5 S3 Fall Average 3.39 3.49 2.9% 4.56 Farmingtona L5 S4 Fall Vatican 326 109.5 3.39 ± 0.03 3.48 ± 0.03 2.5% ± 1.4% 4.90 ± 0.12 Farmington L5 S4 Fall Vatican 327 47.901 3.40 ± 0.04 3.55 ± 0.04 4.0% ± 1.6% 5.14 ± 0.09 Farmington L5 S4 Fall Vatican 325 154.22 3.64 ± 0.08‡ n.d. n.d. n.d. Farmington L5 S4 Fall Vatican 324 213.75 3.32 ± 0.07‡ n.d. n.d. n.d. Farmington L5 S4 Fall Vatican 323 518.35 3.25 ± 0.08‡ n.d. n.d. n.d. Farmington L5 S4 Fall Average 3.34 3.50 4.4% 5.02 Fisherb L6 S5 Fall NMNH 212 2 38.176 3.41 ± 0.07 3.51 ± 0.02 2.8% ± 2.1% 4.85 ± 0.12 Fisher L6 S5 Fall Vatican 338 127.08 3.31 ± 0.03 n.d. n.d. 4.88 ± 0.09 Fisher L6 S5 Fall Average 3.33 3.51 5.1% 4.86 Forsythb L6 Fall LNHM BM1985,M82 67.12 3.23 ± 0.03 3.55 ± 0.01 8.8% ± 0.9% 4.87 ± 0.09 Fukutomib L5 S3 Fall IOM C 172.1 32.65 3.46 ± 0.05 3.54 ± 0.02 2.2% ± 1.6% 4.73 ± 0.12 Futtehpurb L6 Fall LNHM BM1985,M83 35.82 3.16 ± 0.04 3.59 ± 0.02 12.0% ± 1.3% 4.91 ± 0.09 Gambatb L6 Fall Vatican 349 98.76 3.30 ± 0.03 n.d. n.d. 5.04 ± 0.10 Gifu L6 S4 Fall IOM C 47.5 21.31 3.37 ± 0.09 3.52 ± 0.02 4.3% ± 2.6% 5.03 ± 0.10 Girgentib L6 S4 Fall Vatican 360 52.491 3.33 ± 0.02 3.61 ± 0.04 7.6% ± 1.1% 4.93 ± 0.09 Grossliebenthalb L6 Fall Vatican 377 19.901 3.29 ± 0.05 3.62 ± 0.04 9.1% ± 1.6% 4.72 ± 0.08 Grossliebenthal L6 Fall Vatican 376 21.46 3.23 ± 0.04 3.61 ± 0.04 10.5% ± 1.4% 4.71 ± 0.08 Grossliebenthal L6 Fall Average 3.26 3.61 9.8% 4.71 Guangrao L6 Fall CMS 1518D 46.8 3.30 ± 0.04 3.59 ± 0.02 8.0% ± 1.1% 5.00 ± 0.09 Hedjaza L3.7 S4 W1 Fall Vatican 386 24.442 3.55 ± 0.04 3.58 ± 0.04 1.0% ± 1.6% 4.90 ± 0.08 Hedjaz L3.7 S4 W1 Fall Vatican 387 17.41 3.39 ± 0.06 n.d. n.d. 5.00 ± 0.08† Hedjaz L3.7 S4 W1 Fall Average 3.48 3.58 2.8% 4.95 Homesteada L5 S4 Fall Vatican 437 14.01 n.d. n.d. n.d. 5.05 ± 0.08† L5 S4 Fall Vatican 436 29.657 3.37 ± 0.04 3.61 ± 0.04 6.7% ± 1.4% 5.12 ± 0.08 Homestead L5 S4 Fall Vatican 435 262.16 3.43 ± 0.08‡ n.d. n.d. n.d. Homestead L5 S4 Fall Vatican 434 583.71 3.36 ± 0.07‡ n.d. n.d. n.d. Homestead L5 S4 Fall Vatican 433 1090.59 3.32 ± 0.07‡ n.d. n.d. n.d. Homestead L5 S4 Fall Average 3.35 3.61 7.3% 5.09 Honolulub L5 S3 Fall LNHM BM25460 24.17 3.25 ± 0.05 3.55 ± 0.02 8.4% ± 1.4% 4.67 ± 0.12 Honolulu L5 S3 Fall Vatican 441 101.25 n.d. 3.62 ± 0.04 n.d. 4.78 ± 0.08† Honolulu L5 S3 Fall Average 3.25 3.60 9.9% 4.72 Jackalsfontein L6 Fall CMS 836 19.71 3.37 ± 0.06 3.51 ± 0.03 4.1% ± 1.8% 4.90 ± 0.12 Jhung L5 S3 Fall NMNH 2157 20.495 3.19 ± 0.05 3.53 ± 0.04 9.7% ± 1.6% 4.91 ± 0.12 Kediri L4 Fall Monnig M1212.1 120.64 3.27 ± 0.05 3.51 ± 0.01 6.7% ± 1.5% 4.74 ± 0.12 Kendletona L4 S3 Fall Monnig M 32.12a 51.01 3.39 ± 0.03 3.59 ± 0.01 5.6% ± 0.8% 5.08 ± 0.09 Kendleton L4 S3 Fall IOM C 288.1 81.57 3.38 ± 0.04 3.56 ± 0.01 5.1% ± 1.0% 4.94 ± 0.09 Kendleton L4 S3 Fall Vatican 1083 121.15 3.35 ± 0.03 3.57 ± 0.04 6.2% ± 1.3% 4.90 ± 0.09 Kendleton L4 S3 Fall Average 3.36 3.57 5.7% 4.97 Khohar L3.6 S4 Fall NMNH 1252 83.209 3.38 ± 0.06 3.53 ± 0.01 4.1% ± 1.7% 4.88 ± 0.09 Kuleschovkab L6 Fall LNHM BM44774 32.16 3.19 ± 0.04 3.56 ± 0.02 10.2% ± 1.3% 4.80 ± 0.09 Kuttippuram L6 Fall Monnig M1401.1 43.2 3.26 ± 0.03 3.57 ± 0.01 8.7% ± 0.9% 4.71 ± 0.09 Kyushub L6 S5 Fall NMNH 568 44.914 3.45 ± 0.04 3.58 ± 0.02 3.7% ± 1.2% 5.09 ± 0.09 Kyushu L6 S5 Fall IOM C 51.5 44.42 3.36 ± 0.04 3.51 ± 0.01 4.2% ± 1.3% 5.04 ± 0.09 Kyushu L6 S5 Fall LNHM BM1905,68 34.39 3.29 ± 0.06 3.51 ± 0.01 6.1% ± 1.7% 4.92 ± 0.10 Kyushu L6 S5 Fall Vatican 521 42.62 3.32 ± 0.02 3.60 ± 0.04 7.9% ± 1.1% 4.89 ± 0.09

263

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Kyushu L6 S5 Fall Average 3.36 3.55 5.4% 4.99 L’Aiglea L6 S4 Fall Vatican 525 48.78 3.31 ± 0.02 3.44 ± 0.03 4.0% ± 1.1% 4.59 ± 0.09 L’Aigle L6 S4 Fall Vatican 523 215.18 3.39 ± 0.03 n.d. n.d. 4.78 ± 0.08 L’Aigle L6 S4 Fall Vatican 526 24.13 3.25 ± 0.03 3.66 ± 0.04 11.3% ± 1.2% 4.81 ± 0.08 L’Aigle L6 S4 Fall Vatican 528 9.45 3.35 ± 0.07 3.71 ± 0.04 9.6% ± 2.1% 4.96 ± 0.08† L’Aigle L6 S4 Fall Vatican 524 52.02 3.56 ± 0.02 3.78 ± 0.04 5.8% ± 1.1% 5.23 ± 0.09 L’Aigle L6 S4 Fall Vatican 527 19.5 n.d. 3.70 ± 0.04 n.d. n.d. L’Aigle L6 S4 Fall Average 3.39 3.63 6.8% 4.87 La Bécasse L6 Fall Vatican 530 10.56 n.d. n.d. n.d. 4.84 ± 0.08† La Criolla L6 S4 W2 Fall NMNH 6294 30.814 3.17 ± 0.03 3.57 ± 0.02 11.4% ± 1.0% 4.89 ± 0.09 Lalitpura,b L6 Fall LNHM BM63058 79.11 3.19 ± 0.03 3.56 ± 0.01 10.5% ± 0.8% 4.81 ± 0.09 Leedey L6 S3 Fall CMS 489.15 34.59 3.26 ± 0.09 3.55 ± 0.01 8.4% ± 2.6% 4.99 ± 0.09 Leedey L6 S3 Fall Vatican 1081 22.59 3.28 ± 0.03 3.66 ± 0.04 10.3% ± 1.2% 4.95 ± 0.08 Leedey L6 S3 Fall Average 3.27 3.60 9.1% 4.97 Leighlinbridge L6 S3 Fall Monnig M 908.1 65.69 3.34 ± 0.03 3.59 ± 0.01 6.9% ± 0.9% 5.02 ± 0.09 Leighlinbridge L6 S3 Fall LNHM BM2001,M1 73.07 3.31 ± 0.03 3.56 ± 0.01 7.0% ± 0.8% 5.15 ± 0.09 Leighlinbridge L6 S3 Fall Average 3.32 3.57 7.0% 5.08 Lesves L6 Fall LNHM BM81535 56.43 3.15 ± 0.03 3.57 ± 0.01 11.7% ± 0.8% 4.91 ± 0.09 Lissab L6 Fall Vatican 562 21.09 3.59 ± 0.07 3.55 ± 0.04 -1.2% ± 2.2% 4.94 ± 0.08† Lissa L6 Fall Vatican 561 28.99 3.27 ± 0.02 3.67 ± 0.04 10.9% ± 1.1% 4.92 ± 0.08 Lissa L6 Fall Average 3.40 3.62 6.1% 4.93 Little Piney L5 Fall LNHM BM24005 97.96 3.21 ± 0.03 3.83 ± 0.01 16.3% ± 0.7% 4.70 ± 0.10 Lua L5 Fall LNHM BM1928,480 35.63 3.30 ± 0.04 3.56 ± 0.01 7.2% ± 1.1% 5.03 ± 0.09 Lundsgård L6 Fall CMS 973s 89.76 3.29 ± 0.05 3.54 ± 0.01 7.1% ± 1.5% 5.03 ± 0.09 Lundsgård L6 Fall Vatican 574 17.57 3.29 ± 0.05 n.d. n.d. 4.99 ± 0.08 Lundsgård L6 Fall Average 3.29 3.54 7.1% 5.01 Madrida,b L6 Fall IOM C 363.1 57.28 3.30 ± 0.03 3.55 ± 0.01 6.9% ± 1.0% 4.88 ± 0.12 Marion (Iowa) b L6 Fall Vatican 589 30.65 3.21 ± 0.02 3.64 ± 0.04 11.9% ± 1.1% 4.71 ± 0.10 Marmande L5 Fall Vatican 592 25.1 3.15 ± 0.02 3.66 ± 0.04 13.8% ± 1.1% 4.97 ± 0.08 Mascombes L6 Fall Vatican 594 3.93 n.d. n.d. n.d. 4.81 ± 0.08† Mauerkirchen L6 Fall Vatican 596 2.29 n.d. n.d. n.d. 4.95 ± 0.08† Mernb L6 Fall Vatican 612 14.46 n.d. n.d. n.d. 4.86 ± 0.08† Mezö-Madarasa L3.7 S2 Fall Vatican 614 29.9 3.40 ± 0.02 3.51 ± 0.04 3.1% ± 1.2% 4.82 ± 0.08 Mezö-Madaras L3.7 S2 Fall Vatican 613 38.93 3.40 ± 0.02 3.51 ± 0.04 3.2% ± 1.1% 4.78 ± 0.10 Mezö-Madaras L3.7 S2 Fall Average 3.40 3.51 3.2% 4.80 Milena L6 Fall CMS 335s 53.19 3.27 ± 0.03 3.57 ± 0.01 8.5% ± 0.8% 4.91 ± 0.09 Mócsb L5-6 S3-5 Fall NMNH 1781 30.938 3.30 ± 0.03 3.54 ± 0.02 6.8% ± 1.0% 4.66 ± 0.10 Mócs L5-6 S3-5 Fall NMNH 467 1 32.415 3.37 ± 0.04 3.57 ± 0.02 5.8% ± 1.1% 4.82 ± 0.08 Mócs L5-6 S3-5 Fall NMNH 467 2 33.093 3.22 ± 0.04 3.54 ± 0.02 9.0% ± 1.1% 4.80 ± 0.09 Mócs L5-6 S3-5 Fall NMNH 467 3 23.933 3.24 ± 0.04 3.53 ± 0.02 8.3% ± 1.1% 4.88 ± 0.08 Mócs L5-6 S3-5 Fall NMNH 467 4 31.311 3.27 ± 0.03 3.54 ± 0.02 7.7% ± 0.9% 4.81 ± 0.08 Mócs L5-6 S3-5 Fall NMNH 467 5 28.742 3.24 ± 0.03 3.53 ± 0.02 8.2% ± 0.9% 4.84 ± 0.08 Mócs L5-6 S3-5 Fall NMNH 1717 1 25.705 3.22 ± 0.04 3.54 ± 0.02 8.9% ± 1.1% 4.80 ± 0.08 Mócs L5-6 S3-5 Fall NMNH 1717 2 13.998 3.35 ± 0.09 3.51 ± 0.04 4.4% ± 2.8% 4.94 ± 0.08 Mócs L5-6 S3-5 Fall NMNH 1717 3 22.939 3.26 ± 0.04 3.53 ± 0.02 7.5% ± 1.2% 4.81 ± 0.08 Mócs L5-6 S3-5 Fall NMNH 1717 4 17.628 3.25 ± 0.05 3.48 ± 0.03 6.8% ± 1.6% 4.72 ± 0.08 Mócs L5-6 S3-5 Fall Vatican 635 66.32 3.27 ± 0.02 3.68 ± 0.04 11.2% ± 1.0% 4.90 ± 0.09 Mócs L5-6 S3-5 Fall Vatican 636 55.38 3.24 ± 0.02 3.70 ± 0.04 12.5% ± 1.0% 4.79 ± 0.09 Mócs L5-6 S3-5 Fall Vatican 632 100.85 3.27 ± 0.01 3.65 ± 0.04 10.4% ± 1.0% 4.95 ± 0.09 Mócs L5-6 S3-5 Fall Vatican 634 74.49 3.28 ± 0.01 3.66 ± 0.04 10.3% ± 1.0% 4.98 ± 0.09 Mócs L5-6 S3-5 Fall Vatican 639 25.62 3.24 ± 0.02 3.68 ± 0.04 11.9% ± 1.0% 4.85 ± 0.08 Mócs L5-6 S3-5 Fall Vatican 640 22.34 3.23 ± 0.03 3.69 ± 0.04 12.5% ± 1.2% 4.78 ± 0.10 Mócs L5-6 S3-5 Fall Vatican 631 101.94 3.23 ± 0.03 3.69 ± 0.04 12.5% ± 1.2% 4.95 ± 0.09 Mócs L5-6 S3-5 Fall Vatican 637 36.52 3.22 ± 0.02 3.64 ± 0.04 11.5% ± 1.0% 4.84 ± 0.09 Mócs L5-6 S3-5 Fall Vatican 633 92.56 3.23 ± 0.04 3.67 ± 0.04 11.9% ± 1.3% 4.97 ± 0.09 Mócs L5-6 S3-5 Fall Average 3.26 3.63 10.3% 4.85 Monte Milone L5 S3 Fall Vatican 648 8.49 n.d. n.d. n.d. 4.69 ± 0.08† Monze L6 Fall IOM C 29.2 21.72 3.31 ± 0.04 3.50 ± 0.02 5.3% ± 1.3% 4.93 ± 0.12 Monze L6 Fall Vatican 650 165.37 3.28 ± 0.03 n.d. n.d. 5.01 ± 0.09 Monze L6 Fall Average 3.28 3.50 6.1% 4.97 Nerftb L6 Fall Vatican 684 39.61 3.39 ± 0.03 3.64 ± 0.04 6.9% ± 1.2% 5.11 ± 0.09 New Concordb L6 S4 Fall Vatican 695 63.66 3.28 ± 0.01 3.64 ± 0.04 9.8% ± 1.0% 4.95 ± 0.09 Nikolskoe L4 S2 W1 Fall NMNH 1732 23.777 3.04 ± 0.04 3.60 ± 0.02 15.6% ± 1.3% 4.98 ± 0.08 Oesel L6 Fall Vatican 713 22.301 3.13 ± 0.02 3.64 ± 0.04 14.2% ± 1.0% 4.84 ± 0.08 Ojuelos Altosa L6 Fall CMS 782 71.5 3.08 ± 0.05 3.50 ± 0.01 11.8% ± 1.5% 3.81 ± 0.10

264

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Pacula a L6 Fall CMS 96a 65.3 3.37 ± 0.04 3.61 ± 0.01 6.7% ± 1.0% 5.06 ± 0.09 Park Forest L6 Fall FMNH ME 3353 #1 159.4 3.40 ± 0.06 3.55 ± 0.00 4.3% ± 1.6% 4.92 ± 0.09 Pavlograd L6 Fall Vatican 738 40.87 3.28 ± 0.02 3.63 ± 0.04 9.6% ± 1.0% 4.86 ± 0.09 Perpeti L6 Fall Monnig M1403.1 46.18 3.24 ± 0.03 3.61 ± 0.02 10.3% ± 1.0% 4.86 ± 0.12 Pricetown L6 Fall Vatican 756 30.19 3.18 ± 0.02 3.64 ± 0.04 12.5% ± 1.0% 4.74 ± 0.08 Putinga L6 Fall Vatican 780 13.08 n.d. n.d. n.d. 5.03 ± 0.08† Rakovka L6 Fall Vatican 787 15.34 n.d. n.d. n.d. 4.93 ± 0.08† Ramsdorfa L6 Fall CMS 703.2 48.81 3.14 ± 0.03 3.53 ± 0.02 11.0% ± 0.9% 4.89 ± 0.10 Reliegos L5 Fall CMS 786 33.67 3.26 ± 0.09 3.52 ± 0.02 7.4% ± 2.5% 5.02 ± 0.09 Renqiu L6 Fall CMS 1289 78.49 3.39 ± 0.06 3.39 ± 0.01 0.2% ± 1.7% 4.76 ± 0.09 Sallesb L5 Fall LNHM BM1985,M38 71.33 3.23 ± 0.03 3.58 ± 0.01 9.6% ± 0.9% 4.78 ± 0.09 Saratov L4 S2 W1 Fall NMNH 5956 28.076 3.07 ± 0.04 3.58 ± 0.02 14.1% ± 1.1% 5.02 ± 0.08 Saratov L4 S2 W1 Fall IOM C 329.1 35.22 3.18 ± 0.04 3.58 ± 0.02 11.3% ± 1.3% 4.92 ± 0.09 Saratov L4 S2 W1 Fall Vatican 1082 90.68 3.07 ± 0.03 n.d. n.d. 4.93 ± 0.09 Saratov L4 S2 W1 Fall Average 3.10 3.58 13.5% 4.96 Schönenbergb L6 Fall Vatican 857 14.37 n.d. n.d. n.d. 4.80 ± 0.08† Schönenberg L6 Fall LNHM BM67208 39.47 3.25 ± 0.03 3.59 ± 0.01 9.6% ± 1.0% 5.03 ± 0.09 Schönenberg L6 Fall Average 3.25 3.59 9.6% 4.92 Segowlie L6 Fall Vatican 864 13.02 n.d. n.d. n.d. 4.47 ± 0.08† Sevrukovo L5 Fall Vatican 866 19.91 3.65 ± 0.08 3.54 ± 0.04 -3.1% ± 2.4% 4.95 ± 0.10 Shelburnea,b L5 S4 Fall Monnig M 260.4 37.68 3.46 ± 0.05 3.61 ± 0.02 4.3% ± 1.6% 4.86 ± 0.09 Shelburne L5 S4 Fall Vatican 873 13.44 n.d. n.d. n.d. 4.75 ± 0.08† Shelburne L5 S4 Fall Vatican 871 67.83 3.28 ± 0.03 3.66 ± 0.04 10.4% ± 1.3% 5.00 ± 0.10 Shelburne L5 S4 Fall Average 3.34 3.64 8.3% 4.87 St. Christophe-la-Chartreuse L6 Fall Vatican 892 38.81 3.44 ± 0.04 3.63 ± 0.04 5.3% ± 1.5% 4.97 ± 0.10 St. Michel L6 Fall Vatican 904 38.77 3.37 ± 0.03 3.62 ± 0.04 6.9% ± 1.2% 4.81 ± 0.09 St. Michel L6 Fall Vatican 905 32.42 3.47 ± 0.06 3.66 ± 0.04 5.0% ± 1.8% 4.98 ± 0.08 St. Michel L6 Fall Average 3.42 3.64 6.1% 4.90 Suizhou L6 Fall IOM C 334.1 40.02 3.22 ± 0.03 3.56 ± 0.02 9.5% ± 1.0% 4.75 ± 0.10 Tadjera L5 Fall LNHM BM71574 36.19 3.39 ± 0.04 3.52 ± 0.01 3.6% ± 1.2% 4.95 ± 0.12 Tadjera L5 S4 Fall Vatican 918 31.42 3.58 ± 0.08 3.51 ± 0.04 -2.0% ± 2.5% 4.92 ± 0.12 Tadjera L5 S4 Fall Average 3.48 3.52 1.1% 4.94 Tenhamb L6 S4 Fall NMNH 2472 19 48.873 3.35 ± 0.03 3.46 ± 0.01 3.3% ± 1.0% 4.77 ± 0.09 L6 S4 Fall IOM C 119.1&3 37.07 3.30 ± 0.04 3.49 ± 0.01 5.5% ± 1.1% 4.84 ± 0.09 Tenham L6 S4 Fall Vatican 1132 20.75 3.38 ± 0.05 3.55 ± 0.04 4.7% ± 1.8% 4.80 ± 0.12 Tenham L6 S4 Fall Average 3.34 3.49 4.4% 4.80 Tennasilmb L4 S3 W2-3 Fall NMNH 483 60.46 3.20 ± 0.03 3.57 ± 0.01 10.6% ± 0.9% 5.01 ± 0.09 Tennasilm L4 S3 W2-3 Fall Vatican 925 57.301 3.29 ± 0.03 3.70 ± 0.04 10.9% ± 1.1% 4.82 ± 0.09 Tennasilm L4 S3 W2-3 Fall Average 3.24 3.63 10.7% 4.91 Tourinnes-la-Grosse L6 S3 Fall Vatican 968 45.43 3.27 ± 0.04 3.68 ± 0.04 11.0% ± 1.3% 5.01 ± 0.09 Umm Ruaba L5 Fall CMS 1051 57.64 3.32 ± 0.05 3.52 ± 0.01 5.7% ± 1.6% 5.08 ± 0.12 Valdinizza L6 Fall Monnig M 832.1 27.24 3.25 ± 0.04 3.56 ± 0.02 8.6% ± 1.3% 5.11 ± 0.08 Vouilléb L6 S5 Fall Vatican 1007 124.01 3.32 ± 0.03 n.d. n.d. 5.01 ± 0.09 Woolgorong L6 Fall CMS 712 78.7 3.19 ± 0.05 3.56 ± 0.01 10.3% ± 1.5% 4.83 ± 0.09 Zavida L6 Fall IOM C 200.1 25.79 3.32 ± 0.04 3.57 ± 0.02 6.8% ± 1.2% 5.07 ± 0.08 Zavid L6 Fall Vatican 1040 69.23 3.23 ± 0.04 3.58 ± 0.04 9.8% ± 1.3% 4.93 ± 0.09 Zavid L6 Fall Average 3.26 3.58 9.0% 5.00 Zhaodong L4 Fall Monnig M1266.1 66.32 3.46 ± 0.03 3.60 ± 0.01 3.9% ± 0.9% 4.97 ± 0.10 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study. aBreccia (Grady, 2000) bVeined (Grady, 2000)

265

Table 7: Data for L Chondrite Finds.

Bulk Grain Model Magnetic Weathering Meteorite Type Fall Collection Catalog Mass Density Density Porosity Porosity Susceptibility Modulus (g) (g cm-3) (g cm-3) (log χ) Akron (1961) L6 Find AMNH 4160 41.42 3.35 ± 0.06 3.45 ± 0.02 3.0% ± 1.9% 6.4% 4.51 ± 0.12 2.06 ALH 76001 L6 WA Find LNHM BM1999,M51 19.39 3.24 ± 0.10 3.57 ± 0.03 9.3% ± 2.9% 9.6% 4.77 ± 0.12 0.51 ALH 76009 L6 WB Find CMS 1149 70.82 3.20 ± 0.03 3.53 ± 0.01 9.4% ± 0.8% 10.6% 4.84 ± 0.10 0.39 Amherst L6 Find AMNH 4593 40.9 3.54 ± 0.06 3.48 ± 0.02 -1.8% ± 1.8% 1.1% 4.79 ± 0.12 0.91 Arapahoe L5 Find AMNH 4271 52.34 3.43 ± 0.02 3.52 ± 0.01 2.3% ± 0.6% 4.1% 5.11 ± 0.09 1.29 Armel (Colorado) L5 Find AMNH 4461 61.48 3.48 ± 0.06 3.43 ± 0.01 -1.4% ± 1.7% 2.9% 4.84 ± 0.12 1.21 Barratta L4 S4 W3 Find NMNH 720 A 36.348 3.38 ± 0.03 3.42 ± 0.02 1.0% ± 1.1% 5.6% 4.92 ± 0.10 1.34 Barratta L4 S4 W3 Find Vatican 87 88.894 3.40 ± 0.04 3.49 ± 0.03 2.8% ± 1.4% 5.2% 5.03 ± 0.10 1.06 Barratta L4 S4 W3 Find Average 3.39 3.47 2.3% 5.3% 4.98 - Bayard L5 Find LNHM BM1984,M10 24.71 3.28 ± 0.06 3.36 ± 0.02 2.4% ± 1.8% 8.6% 3.92 ± 0.09 5.08 Beaver-Harrison L6 S5 Find IOM C 178.1a 27.59 3.32 ± 0.05 3.44 ± 0.02 3.7% ± 1.6% 7.4% 4.58 ± 0.08 1.79 Bluff (a) a L5 S4 W2 Find NMNH 135 78.26 3.47 ± 0.06 3.47 ± 0.01 0.0% ± 1.8% 3.2% 4.86 ± 0.10 0.90 Bluff (a) L5 S4 W2 Find CMS 13b 49.57 3.46 ± 0.06 3.55 ± 0.02 2.4% ± 1.7% 3.4% 5.18 ± 0.12 1.59 Bluff (a) L5 S4 W2 Find Vatican 126 57.649 3.36 ± 0.03 3.47 ± 0.03 3.2% ± 1.4% 6.1% 4.81 ± 0.12 0.90 Bluff (a) L5 S4 W2 Find Vatican 125 62.245 3.46 ± 0.04 3.50 ± 0.04 1.2% ± 1.5% 3.4% 4.91 ± 0.09 0.65 Bluff (a) L5 S4 W2 Find Average 3.44 3.49 1.6% 4.0% 4.94 - Bouse L4-6 S1-4 W1 Find CMS 1564 19.96 3.22 ± 0.05 3.49 ± 0.02 7.8% ± 1.5% 10.2% 4.87 ± 0.08 0.72 Bruceville L6 S5 W4 Find CMS 1420 63.57 2.94 ± 0.16 3.27 ± 0.01 10.0% ± 5.0% 17.9% 3.96 ± 0.10 5.17 Calliham L6 Find Vatican 1102 12.501 n.d. n.d. n.d. n.d. 4.40 ± 0.08† n.d. Carraweena L3.9 S4 Find NMNH 2367 43.956 3.54 ± 0.04 3.40 ± 0.02 -3.9% ± 1.2% 1.2% 4.65 ± 0.12 1.80 Chico L6 S6 Find IOM C 87.8 22.8 3.34 ± 0.09 3.34 ± 0.02 0.0% ± 2.9% 6.7% 5.20 ± 0.33 2.50 Claytonville L5 W2 Find Vatican 219 19.15 3.29 ± 0.05 3.50 ± 0.03 6.0% ± 1.7% 8.3% 4.56 ± 0.12 1.68 Coon Buttea L6 Find Vatican 235 8.09 n.d. n.d. n.d. n.d. 4.78 ± 0.08† n.d. Densmore (1879) L6 Find Vatican 263 21.768 3.32 ± 0.04 3.39 ± 0.03 2.0% ± 1.6% 7.3% 4.23 ± 0.12 3.54 Dougherty L6 S3 W1 Find CMS 1508 31.91 3.31 ± 0.04 3.51 ± 0.02 5.7% ± 1.4% 7.7% 4.54 ± 0.12 1.77 Eli Elwah L6 Find LNHM BM1927,1275 46.6 3.26 ± 0.03 3.39 ± 0.01 3.9% ± 0.9% 9.0% 4.24 ± 0.09 3.48 Etter L5 S5 Find IOM C 73.2a 51.47 3.44 ± 0.07 3.44 ± 0.01 -0.1% ± 2.2% 3.9% 4.81 ± 0.12 1.16 Etter L5 S5 Find Vatican 320 32.99 3.37 ± 0.04 3.44 ± 0.03 2.0% ± 1.5% 5.9% 4.75 ± 0.12 1.29 Etter L5 S5 Find Vatican 321 24.107 3.47 ± 0.04 3.47 ± 0.03 0.1% ± 1.4% 3.1% 4.82 ± 0.12 0.88 Etter L5 S5 Find Average 3.43 3.45 0.6% 4.3% 4.79 - Felt (b) a L3.5-5 S4-5 W1 Find IOM C 353.1 27.75 3.47 ± 0.16 3.44 ± 0.03 -0.9% ± 4.7% 3.2% 4.75 ± 0.12 1.30 Franklinville L6 Find LNHM BM1988,M3 20.28 3.24 ± 0.05 3.38 ± 0.02 4.1% ± 1.5% 9.6% 4.43 ± 0.12 2.71 Garraf L6 S2 Find Vatican 352 8.48 n.d. n.d. n.d. n.d. 4.48 ± 0.08† n.d. Garraf L6 S2 Find Vatican 351 17.74 n.d. 3.53 ± 0.04 n.d. n.d. 4.69 ± 0.08† 0.99 Garraf L6 S2 Find Average n.d. 3.53 n.d. n.d. 4.59 - Gila Bend L5 S3 W1 Find CMS 1483 27.37 3.31 ± 0.04 3.47 ± 0.02 4.7% ± 1.3% 7.7% 4.80 ± 0.08 0.93 Gold Basin L4 W2-3 Find NMNH 7081 16 73.909 3.36 ± 0.03 3.50 ± 0.01 3.9% ± 0.9% 6.2% 4.86 ± 0.09 0.67 Gold Basin L4 W2-3 Find NMNH 7081 17 46.741 3.39 ± 0.05 3.47 ± 0.01 2.3% ± 1.4% 5.5% 4.76 ± 0.09 1.05 Gold Basin L4 W2-3 Find NMNH 7081 21 31.992 3.43 ± 0.06 3.45 ± 0.01 0.6% ± 1.8% 4.2% 4.83 ± 0.08 1.02 Gold Basin L4 W2-3 Find NMNH 7081 22 53.746 3.44 ± 0.03 3.51 ± 0.01 1.8% ± 1.0% 3.9% 4.95 ± 0.09 0.70 Gold Basin L4 W2-3 Find NMNH 7081 25 91.788 3.40 ± 0.03 3.51 ± 0.01 3.1% ± 0.9% 5.2% 5.00 ± 0.09 0.88 Gold Basin L4 W2-3 Find NMNH 7081 26 52.895 3.39 ± 0.03 3.48 ± 0.01 2.6% ± 0.9% 5.3% 4.80 ± 0.09 0.88 Gold Basin L4 W2-3 Find NMNH 7081 27 64.84 3.43 ± 0.03 3.51 ± 0.01 2.2% ± 1.0% 4.2% 4.88 ± 0.09 0.57 Gold Basin L4 W2-3 Find NMNH 7081 28 16.171 3.39 ± 0.05 3.45 ± 0.03 1.7% ± 1.6% 5.3% 4.92 ± 0.08 1.07 Gold Basin L4 W2-3 Find NMNH 7081 29 37.698 3.34 ± 0.04 3.45 ± 0.02 3.1% ± 1.3% 6.7% 4.81 ± 0.09 1.09 Gold Basin L4 W2-3 Find NMNH 7081 31 16.009 3.37 ± 0.05 3.44 ± 0.03 2.2% ± 1.6% 6.1% 4.68 ± 0.08 1.45 Gold Basin L4 W2-3 Find NMNH 7081 32 57.551 3.40 ± 0.03 3.53 ± 0.01 3.8% ± 1.0% 5.2% 4.92 ± 0.09 0.49 Gold Basin L4 W2-3 Find Vatican 1089 22.72 3.38 ± 0.04 3.61 ± 0.04 6.3% ± 1.5% 5.7% 5.00 ± 0.08 0.70 Gold Basin L4 W2-3 Find Vatican 1088 22.71 3.42 ± 0.04 3.66 ± 0.04 6.5% ± 1.4% 4.4% 4.97 ± 0.08 0.83 Gold Basin L4 W2-3 Find Average 3.40 3.50 3.0% 5.2% 4.88 - Gretna L5 Find Vatican 371 146.22 3.29 ± 0.03 n.d. n.d. 8.1% 4.61 ± 0.09 n.d. Gunlock L3.2 Find Monnig M1369.1 46.65 3.44 ± 0.03 3.53 ± 0.02 2.7% ± 1.0% 4.0% 4.94 ± 0.12 0.53 Hajmah (c) L5/6 Find LNHM BM1980,M22 57.92 3.33 ± 0.03 3.41 ± 0.01 2.1% ± 1.0% 6.9% 4.78 ± 0.10 1.46 HaH 173 L6 S5 W1 Find LNHM BM1996,M9 32.62 3.21 ± 0.04 3.49 ± 0.02 8.1% ± 1.3% 10.3% 4.78 ± 0.09 0.82 Harding County L4 Find LNHM BM1959,868 30.83 3.21 ± 0.05 3.40 ± 0.02 5.4% ± 1.5% 10.3% 4.75 ± 0.12 1.58 Hermitage Plains L6 Find Vatican 390 499.70 3.30 ± 0.07‡ 3.61 ± 0.06‡ 8.7% ± 2.5% 7.9% n.d. n.d. Hub L5 S4 W2 Find CMS 1424 47.89 3.37 ± 0.06 3.43 ± 0.01 1.8% ± 1.9% 6.0% 4.54 ± 0.10 2.02 JaH 055 L4-5 S2 W3 Find CMS 1500 39.96 3.18 ± 0.05 3.38 ± 0.02 5.7% ± 1.6% 11.1% 4.72 ± 0.12 1.78 JaH 055 L4-5 S2 W3 Find Vatican 1459 27.42 3.50 ± 0.04 n.d. n.d. 2.4% 4.83 ± 0.08 n.d.

266

Bulk Grain Model Magnetic Weathering Meteorite Type Fall Collection Catalog Mass Density Density Porosity Porosity Susceptibility Modulus (g) (g cm-3) (g cm-3) (log χ) JaH 055 L4 -5 S2 W3 Find Average 3.30 3.38 2.2% 7.8% 4.78 -

Julesburg L3.6 S3 W1.5 Find Vatican 1080 56.346 3.44 ± 0.04 n.d. n.d. 3.8% 4.93 ± 0.12 n.d. Kermichel L6 Find Vatican 1077 37.66 3.28 ± 0.03 3.44 ± 0.03 4.7% ± 1.4% 8.4% 3.90 ± 0.12 4.96 Kermichel L6 Find Vatican 1076 61.68 3.28 ± 0.04 3.48 ± 0.03 5.6% ± 1.4% 8.4% 4.08 ± 0.12 4.02 Kermichel L6 Find Vatican 497 220.7 3.27 ± 0.07‡ n.d. n.d. 8.8% n.d. n.d. Kermichel L6 Find Vatican 496 426.38 3.25 ± 0.07‡ n.d. n.d. 9.2% n.d. n.d. Kermichel L6 Find Average 3.26 3.47 5.9% 9.0% 3.99 - Kingfisher L5 S6 Find NMNH 1521 78.717 3.32 ± 0.06 3.36 ± 0.01 1.2% ± 1.7% 7.2% 4.54 ± 0.12 2.37 Kingfisher L5 S6 Find IOM C 124.3 19.93 3.31 ± 0.09 3.35 ± 0.02 1.2% ± 2.8% 7.6% 4.51 ± 0.12 2.59 Kingfisher L5 S6 Find Average 3.32 3.36 1.2% 7.3% 4.53 - Lakewood L6 Find Vatican 539 26.97 3.30 ± 0.04 3.47 ± 0.03 4.9% ± 1.6% 8.0% 4.52 ± 0.12 1.98 Long Islandb L6 S4 Find IOM C 281.1 110.13 3.34 ± 0.05 3.43 ± 0.01 2.7% ± 1.6% 6.8% 4.43 ± 0.09 2.48 Long Island L6 S4 Find Vatican 566 77.25 3.24 ± 0.03 3.44 ± 0.03 5.8% ± 1.4% 9.5% 3.98 ± 0.10 4.60 Long Island L6 S4 Find Vatican 568 41.42 3.36 ± 0.02 3.51 ± 0.04 4.3% ± 1.1% 6.3% 4.38 ± 0.09 2.52 Long Island L6 S4 Find Vatican 567 50.51 3.36 ± 0.02 3.50 ± 0.04 4.0% ± 1.1% 6.2% 4.24 ± 0.10 3.21 Long Island L6 S4 Find Average 3.32 3.46 4.0% 7.4% 4.26 - Loongana 003 L5 S3 W4 Find IOM C 310.12 15.9 3.22 ± 0.05 3.46 ± 0.03 7.1% ± 1.7% 10.2% 4.66 ± 0.08 1.41 Lutschaunig's L6 Find LNHM BM36107 31.55 3.35 ± 0.04 3.50 ± 0.02 4.4% ± 1.2% 6.6% 4.79 ± 0.10 0.76 Stone Macy L6 Find LNHM BM1986,M30 53.24 3.11 ± 0.03 3.41 ± 0.01 8.8% ± 0.8% 13.2% 4.45 ± 0.09 2.47 Mainzb L6 Find Vatican 1463 15.14 n.d. n.d. n.d. n.d. 4.68 ± 0.08† n.d. McCook L6 S2 W2 Find CMS 1423 23.22 3.24 ± 0.08 3.55 ± 0.03 8.8% ± 2.4% 9.5% 4.62 ± 0.12 1.28 McKinney L4 S6 Find Vatican 603 35.59 3.49 ± 0.02 3.53 ± 0.04 1.0% ± 1.2% 2.6% 4.74 ± 0.10 0.78 McKinney L4 S6 Find Vatican 602 96.21 3.48 ± 0.03 3.52 ± 0.04 1.2% ± 1.4% 2.9% 4.66 ± 0.10 1.16 McKinney L4 S6 Find Vatican 604 29.4 3.09 ± 0.10 n.d. n.d. 13.9% n.d. n.d. McKinney L4 S6 Find Vatican 601 141.26 3.46 ± 0.03 n.d. n.d. 3.4% 4.82 ± 0.12 n.d. McKinney L4 S6 Find Vatican 600 368.89 3.56 ± 0.07‡ n.d. n.d. 0.5% n.d. n.d. McKinney L4 S6 Find Vatican 599 384.57 3.45 ± 0.08‡ n.d. n.d. 3.7% n.d. n.d. McKinney L4 S6 Find Average 3.48 3.52 1.1% 2.8% 4.74 - Melrose (a) L5 Find LNHM BM1937,385 41.21 3.29 ± 0.03 3.41 ± 0.02 3.7% ± 1.0% 8.3% 4.45 ± 0.10 2.47 Moorabie L3.8 S4-5 Find NMNH 6302 127.98 3.49 ± 0.05 3.52 ± 0.01 1.0% ± 1.6% 2.6% 5.20 ± 0.12 1.73 Ness County (1894) L6 S6 Find Vatican 687 119.04 n.d. n.d. n.d. n.d. 4.08 ± 0.08† n.d. Ness County (1894) L6 S6 Find Vatican 688 31.72 3.29 ± 0.03 3.53 ± 0.04 6.9% ± 1.2% 8.2% 4.16 ± 0.12 3.55 Ness County (1894) L6 S6 Find Vatican 689 28.79 3.40 ± 0.03 3.51 ± 0.04 3.3% ± 1.3% 5.2% 4.71 ± 0.10 0.99 Ness County (1894) L6 S6 Find Vatican 686 697.47 3.35 ± 0.07‡ 3.35 ± 0.03‡ -0.2% ± 2.3% 6.4% 4.42 ± 0.08 2.90 Ness County (1894) L6 S6 Find Average 3.35 3.36 0.2% 6.4% 4.34 - NWA 1696 L3-6 S1-3 W1 Find CMS 1502 20.47 3.28 ± 0.06 3.49 ± 0.02 6.0% ± 1.8% 8.4% 4.84 ± 0.12 0.72 NWA 5243 L3 S2 W2 Find CMS 1595-1 48.02 3.01 ± 0.03 3.30 ± 0.01 8.8% ± 0.8% 16.0% 3.98 ± 0.10 4.99 Pampa (b) L4-5 S4 Find IOM C 284.1 104.75 3.19 ± 0.05 3.35 ± 0.01 4.7% ± 1.5% 10.9% 4.25 ± 0.09 3.59 Pierceville (Stone) L6 S3 Find IOM C 32.2 33.69 3.29 ± 0.08 3.33 ± 0.02 1.1% ± 2.6% 8.2% 4.06 ± 0.12 4.55 Pinto Mountains L6 S5 Find IOM C 93.22 60.7 3.37 ± 0.03 3.44 ± 0.01 2.2% ± 1.0% 6.0% 4.72 ± 0.10 1.33 Quinçay L6 Find Vatican 1059 29.44 3.34 ± 0.02 3.56 ± 0.04 6.1% ± 1.1% 6.7% 4.69 ± 0.08 0.90 Roosevelt Co. 106 L6 S6 W2 Find CMS 1436 74.72 3.24 ± 0.03 3.43 ± 0.01 5.6% ± 0.9% 9.5% 4.43 ± 0.10 2.49 Roy (1933) L5 S3 Find NMNH 2290 A 35.72 3.42 ± 0.06 3.40 ± 0.02 -0.7% ± 2.0% 4.5% 4.40 ± 0.10 2.75 Roy (1933) L5 S3 Find IOM C 37.2 23.42 3.17 ± 0.08 3.38 ± 0.03 6.2% ± 2.6% 11.4% 4.23 ± 0.12 3.54 Roy (1933) L5 S3 Find Vatican 806 11.34 n.d. n.d. n.d. n.d. 4.20 ± 0.08† n.d. Roy (1933) L5 S3 Find Average 3.32 3.39 2.2% 7.4% 4.28 - Taiban (b) L6 W3 Find IOM C 153.2 22.02 3.37 ± 0.09 3.37 ± 0.03 -0.1% ± 2.7% 5.9% 4.42 ± 0.12 2.84 Temple L6 Find FMNH ME 2560 #1 33.39 3.40 ± 0.05 3.45 ± 0.01 1.5% ± 1.5% 5.1% 4.28 ± 0.10 3.10 Tryon L6 Find FMNH ME 2890 #1 50.61 3.32 ± 0.03 3.41 ± 0.01 2.6% ± 1.0% 7.5% 3.90 ± 0.10 5.03 Tulia (B) L6 Find FMNH ME 2322 #4 32.6 3.25 ± 0.04 3.37 ± 0.02 3.8% ± 1.3% 9.4% 4.33 ± 0.09 3.18 Valkeala L6 Find FMNH ME 2974 #1 33.13 3.37 ± 0.03 3.43 ± 0.01 1.8% ± 1.1% 6.0% 4.29 ± 0.09 3.14 Valle de Allende L Find FMNH ME 3086 #1 38.14 3.35 ± 0.03 3.47 ± 0.02 3.3% ± 1.1% 6.4% 4.62 ± 0.12 1.54 Wacondaa L6 S4 Find NMNH 6561 65.627 3.18 ± 0.06 3.50 ± 0.01 9.3% ± 1.9% 11.3% 4.77 ± 0.09 0.78 Waconda L6 S4 Find Vatican 1011 126.01 3.17 ± 0.03 3.61 ± 0.04 12.1% ± 1.2% 11.5% 4.81 ± 0.09 0.39 Waconda L6 S4 Find Vatican 1012 27.27 n.d. 3.79 ± 0.04 n.d. n.d. 4.89 ± 0.08† 1.71 Waconda L6 S4 Find Vatican 1010 156.4 3.33 ± 0.03 n.d. n.d. 7.0% 4.54 ± 0.09 n.d. Waconda L6 S4 Find Average 3.24 3.60 9.9% 9.5% 4.75 - Waldo L6 Find FMNH ME 2831 #1 84.21 3.38 ± 0.03 3.44 ± 0.01 1.6% ± 1.0% 5.5% 4.37 ± 0.10 2.73 Waltman L4 Find FMNH ME 3143 #1 38.17 3.27 ± 0.06 3.49 ± 0.02 6.4% ± 1.8% 8.8% 4.66 ± 0.12 1.28

267

Bulk Grain Model Magnetic Weathering Meteorite Type Fall Collection Catalog Mass Density Density Porosity Porosity Susceptibility Modulus (g) (g cm-3) (g cm-3) (log χ) Wardswell Draw L6 Find FMNH ME 2840 #1 27.24 3.36 ± 0.05 3.51 ± 0.02 4.3% ± 1.4% 6.2% 4.51 ± 0.10 1.85 Zulu Queen L3 S3 Find IOM C 291.1 66.34 3.47 ± 0.03 3.41 ± 0.01 -1.8% ± 0.9% 3.2% 4.63 ± 0.10 1.86 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study. aBreccia (Grady, 2000) bVeined (Grady, 2000)

268

Table 8: Data for LL Chondrite Falls.

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ)

Alta'ameem LL5 Fall NMNH 5964 A 71.399 3.16 ± 0.06 3.51 ± 0.01 10.1% ± 1.7% 4.05 ± 0.09 Appley Bridgeb LL6 S2 Fall NMNH 614 18.147 3.21 ± 0.05 3.44 ± 0.03 6.8% ± 1.7% 3.58 ± 0.12 Bandong LL6 Fall AMNH 410 27.59 3.30 ± 0.05 3.51 ± 0.02 6.0% ± 1.5% 3.78 ± 0.09 Bandong LL6 Fall NMNH 966 48.666 3.23 ± 0.03 3.48 ± 0.01 7.2% ± 0.8% 3.76 ± 0.09 Bandong LL6 Fall Vatican 81 25.09 3.21 ± 0.03 3.62 ± 0.04 11.3% ± 1.3% 3.92 ± 0.09 Bandong LL6 Fall Average 3.24 3.52 7.9% 3.82 Benares (a) LL4 Fall NMNH 1538 72.388 2.97 ± 0.02 3.45 ± 0.01 14.0% ± 0.7% 3.61 ± 0.10 Bensour LL6 Fall LNHM BM2002,M9 32.54 3.17 ± 0.04 3.52 ± 0.02 10.0% ± 1.1% 3.71 ± 0.10 Bensour LL6 Fall PSF PSF02-006 66.86 3.23 ± 0.04 3.56 ± 0.01 9.1% ± 1.2% 3.70 ± 0.10 Bensour LL6 Fall Average 3.21 3.55 9.4% 3.70 Benton LL6 Fall NMNH 1493 B 14.344 3.13 ± 0.05 3.50 ± 0.03 10.6% ± 1.7% 3.72 ± 0.12 Bholaa LL3-6 Fall NMNH 1806 87.613 3.17 ± 0.05 3.50 ± 0.01 9.5% ± 1.5% 4.26 ± 0.09 Chainpur LL3.4 S1 Fall NMNH 1251 96.106 2.80 ± 0.05 3.41 ± 0.01 17.8% ± 1.4% 4.51 ± 0.09 Cherokee Springs LL6 Fall NMNH 1319 A 12.298 3.06 ± 0.07 3.52 ± 0.04 13.0% ± 2.1% 4.40 ± 0.10 Cherokee Springs LL6 Fall Monnig M 638.1 53.72 3.03 ± 0.03 3.53 ± 0.01 14.1% ± 1.0% 4.44 ± 0.09 Cherokee Springs LL6 Fall Average 3.04 3.53 13.9% 4.42 Dhurmsala LL6 S3 W3 Fall NMNH 1015 56.152 3.33 ± 0.04 3.49 ± 0.01 4.4% ± 1.2% 4.28 ± 0.09 Dhurmsala LL6 S3 W3 Fall Vatican 267 85.99 3.32 ± 0.03 3.57 ± 0.04 7.1% ± 1.3% 4.29 ± 0.09 Dhurmsala LL6 S3 W3 Fall Vatican 268 82.134 3.35 ± 0.04 n.d. n.d. 4.26 ± 0.12 Dhurmsala LL6 S3 W3 Fall Vatican 269 29.28 3.30 ± 0.04 3.54 ± 0.04 6.5% ± 1.4% 4.28 ± 0.08 Dhurmsala LL6 S3 W3 Fall Average 3.33 3.54 5.8% 4.28 Ensisheima LL6 S2 Fall NMNH 5953 13.669 3.51 ± 0.07 3.50 ± 0.04 -0.1% ± 2.2% 4.15 ± 0.08 Ensisheim LL6 S2 Fall Vatican 297 33.56 3.46 ± 0.03 3.53 ± 0.04 2.1% ± 1.4% 4.14 ± 0.09 Ensisheim LL6 S2 Fall Average 3.47 3.52 1.4% 4.15 Hamlet LL4 S3 Fall NMNH 3455 65.641 3.22 ± 0.06 3.50 ± 0.01 8.0% ± 1.8% 4.43 ± 0.10 Hamlet LL4 S3 Fall Monnig M 708.2 44.52 3.28 ± 0.04 3.53 ± 0.02 7.1% ± 1.2% 4.46 ± 0.12 Hamlet LL4 S3 Fall Average 3.24 3.51 7.6% 4.45 Jelicaa LL6 S3 W1 Fall Vatican 457 130.07 3.12 ± 0.03 3.56 ± 0.04 12.3% ± 1.2% 3.54 ± 0.09 Jelica LL6 S3 W1 Fall Vatican 458 12.39 n.d. n.d. n.d. 3.72 ± 0.08† Jelica LL6 S3 W1 Fall Average 3.12 3.56 12.3% 3.63 Karatu LL6 Fall Monnig M 965.1 29.22 3.22 ± 0.06 3.49 ± 0.03 7.6% ± 1.9% 3.84 ± 0.12 Kilaboa LL6 Fall PSF PSF02-036 d 58.06 3.14 ± 0.03 3.51 ± 0.01 10.4% ± 0.8% 3.66 ± 0.09 Krymka LL3.2 S3 Fall NMNH 6457 25.534 3.23 ± 0.05 3.42 ± 0.02 5.6% ± 1.6% 4.38 ± 0.08 Manbhoom LL6 Fall LNHM BM1985,M137 26.1 3.09 ± 0.04 3.48 ± 0.02 11.1% ± 1.2% 3.46 ± 0.09 Mangwendia LL6 Fall Vatican 586 92.31 3.23 ± 0.03 3.63 ± 0.04 10.8% ± 1.3% 4.04 ± 0.10 Ngawia LL3.6 Fall NMNH 2483 82.92 3.11 ± 0.05 3.44 ± 0.01 9.4% ± 1.5% 4.19 ± 0.10 Olivenza LL5 S3 Fall NMNH 2978 51.075 3.11 ± 0.04 3.50 ± 0.01 11.2% ± 1.1% 3.76 ± 0.09 Olivenza LL5 S3 Fall Monnig M 170.1 102.43 3.05 ± 0.05 3.54 ± 0.01 13.8% ± 1.5% 3.90 ± 0.09 Olivenza LL5 S3 Fall Average 3.07 3.53 12.9% 3.83 Ottawaa LL6 Fall CMS 870 41.39 3.00 ± 0.03 3.49 ± 0.02 14.1% ± 0.9% 4.19 ± 0.09 Ottawa LL6 Fall LNHM BM1920,331 64.75 3.05 ± 0.03 3.54 ± 0.01 13.9% ± 0.8% 4.13 ± 0.09 Ottawa LL6 Fall Average 3.03 3.52 14.0% 4.16 Oued el Hedjar LL6 S2 W0/1 Fall Monnig M 890.1 22.86 3.23 ± 0.08 3.53 ± 0.02 8.5% ± 2.2% 4.12 ± 0.10 Paragould LL5 Fall Monnig M 293.2 36.98 3.41 ± 0.04 3.51 ± 0.01 2.8% ± 1.2% 4.63 ± 0.09 Parambu LL5 S3 Fall IOM C 143.2 77.82 3.15 ± 0.03 3.51 ± 0.01 10.1% ± 0.8% 3.95 ± 0.09 Parnallee LL3.6 Fall NMNH 1110 55.162 3.25 ± 0.07 3.46 ± 0.01 6.0% ± 2.0% 4.73 ± 0.09 Parnallee LL3.6 S2 Fall Vatican 733 89.12 3.20 ± 0.03 3.52 ± 0.04 9.1% ± 1.3% 4.45 ± 0.10 Parnallee LL3.6 S2 Fall Vatican 734 5.68 n.d. 3.56 ± 0.04 n.d. 4.49 ± 0.08† Parnallee LL3.6 S2 Fall Vatican 735 3.93 n.d. 3.58 ± 0.04 n.d. 4.61 ± 0.08† Parnallee LL3.6 S2 Fall Average 3.22 3.50 8.0% 4.57 Savtschenskoje LL4 Fall Vatican 854 14.94 3.36 ± 0.08 3.62 ± 0.04 7.2% ± 2.3% 4.32 ± 0.08† Sienaa LL5 Fall CMS 554s 39.53 3.30 ± 0.03 3.46 ± 0.01 4.7% ± 1.0% 4.65 ± 0.09 Soko-Banjaa LL4 Fall NMNH 3078 48.395 3.02 ± 0.03 3.47 ± 0.01 12.8% ± 0.9% 4.38 ± 0.09 Soko-Banja LL4 Fall Vatican 888 9.77 n.d. n.d. n.d. 4.22 ± 0.08† Soko-Banja LL4 Fall Vatican 887 27.16 3.16 ± 0.03 3.58 ± 0.04 11.6% ± 1.3% 4.34 ± 0.08 Soko-Banja LL4 Fall Average 3.07 3.51 12.4% 4.31 St. Mesmina LL6 Fall Monnig M 259.1 24.52 3.29 ± 0.05 3.60 ± 0.02 8.6% ± 1.5% 5.15 ± 0.08 St. Mesmin LL6 Fall Vatican 901 52.34 3.10 ± 0.05 3.56 ± 0.04 12.9% ± 1.6% 4.26 ± 0.12 St. Mesmin LL6 Fall Average 3.16 3.57 11.6% 4.71 Tuxtuac LL5 S2 Fall NMNH 6197 14.076 3.11 ± 0.08 3.49 ± 0.03 11.0% ± 2.5% 4.16 ± 0.12 Uden LL7 Fall NMNH 5931 38.255 3.07 ± 0.03 3.41 ± 0.02 10.0% ± 1.1% 3.33 ± 0.09

269

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ)

Vavilovka LL6 Fall Vatican 997 26.701 3.25 ± 0.05 3.54 ± 0.04 8.4% ± 1.6% 3.52 ± 0.09 Witsand Farm LL4 Fall LNHM BM1948,295 71.83 2.88 ± 0.02 3.57 ± 0.01 19.4% ± 0.7% 4.66 ± 0.09 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. aBreccia (Grady, 2000) bVeined (Grady, 2000)

270

Table 9: Data for LL Chondrite Finds.

Bulk Grain Model Magnetic Weathering Meteorite Type Fall Collection Catalog Mass Density Density Porosity Porosity Susceptibility Modulus (g) (g cm-3) (g cm-3) (log χ) Arcadia LL6 Find FMNH ME 2301 #2 48.26 3.10 ± 0.06 3.27 ± 0.01 5.2% ± 1.8% 11.5% 3.66 ± 0.12 3.77 Kelly LL4 S3 Find NMNH 1240 95.254 3.24 ± 0.05 3.41 ± 0.01 4.9% ± 1.6% 8.1% 3.72 ± 0.10 1.83 Kelly LL4 S3 Find IOM C 298.1&2 27.12 3.31 ± 0.05 3.35 ± 0.02 1.2% ± 1.6% 6.1% 3.89 ± 0.09 2.51 Kelly LL4 S3 Find Average 3.25 3.39 4.1% 7.6% 3.81 - Lake Labyrinth LL6 S4 Find NMNH 3333 28.102 3.30 ± 0.04 3.47 ± 0.02 5.0% ± 1.3% 6.4% 4.18 ± 0.10 0.69 Lake Labyrinth LL6 S4 Find FMNH ME 2236 #1 104.3 3.13 ± 0.05 3.48 ± 0.01 10.1% ± 1.5% 10.9% 4.18 ± 0.09 0.64 Lake Labyrinth LL6 S4 Find Average 3.16 3.48 9.1% 10.0% 4.18 - Lazbuddie LL5 Find FMNH ME 2863 #2.4 20.91 3.41 ± 0.06 3.45 ± 0.02 1.1% ± 2.0% 3.7% 4.53 ± 0.10 1.27 Oberlin LL5 Find CMS 283ax 45.85 3.28 ± 0.06 3.46 ± 0.02 5.1% ± 1.7% 6.9% 4.50 ± 0.12 1.08 Oubaria LL6 Find FMNH ME 2988 #1 20.11 3.09 ± 0.05 3.48 ± 0.02 11.3% ± 1.7% 11.8% 4.07 ± 0.12 0.54 Pevensey LL5 Find CMS 1277 41.44 3.22 ± 0.03 3.39 ± 0.01 5.0% ± 1.0% 8.5% 3.85 ± 0.12 1.99 Richfield LL3.7 Find CMS 1573.1 44.84 3.37 ± 0.06 3.39 ± 0.02 0.6% ± 1.7% 4.6% 4.48 ± 0.12 1.99 Saint Lawrence LL6 Find FMNH ME 2972 #1 12.22 3.28 ± 0.07 3.50 ± 0.04 6.3% ± 2.3% 6.9% 4.65 ± 0.12 0.83 Zerga LL6 Find NMNH 5740 71.934 2.93 ± 0.05 3.42 ± 0.01 14.2% ± 1.4% 15.9% 2.86 ± 0.09 2.45 The symbol ―n.d.‖ indicates no data available for that particular stone. aBreccia (Grady, 2000)

271

Table 10: Data for H/L and L/LL Chondrites.

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Bremervörde a H/L3.9 S2 Fall NMNH 630 26.194 3.33 ± 0.03 3.57 ± 0.02 6.9% ± 1.1% 4.92 ± 0.08 Bremervörde H/L3.9 S2 Fall Vatican 139 16.6 3.12 ± 0.03 n.d. n.d. 5.11 ± 0.08 Bremervörde H/L3.9 S2 Fall Average 3.24 3.57 9.2% 5.02 Haxtun H/L4 W4 Find CMS 1409 43.94 3.08 ± 0.06 3.28 ± 0.01 6.1% ± 1.7% 4.35 ± 0.09 Tieschitz H/L3.6 W0 Fall NMNH 1157 12.292 2.99 ± 0.06 3.60 ± 0.04 16.8% ± 2.0% 4.91 ± 0.08 L/LL4 Fall LNHM BM55387 47.99 3.04 ± 0.03 3.51 ± 0.02 13.2% ± 0.9% 4.40 ± 0.10 Bjurböle L/LL4 S1 W2 Fall IOM C 7.6 47.36 2.80 ± 0.03 3.49 ± 0.02 19.8% ± 1.0% 4.49 ± 0.09 Bjurböle L/LL4 S1 W2 Fall Vatican 119 183.68 n.d. n.d. n.d. 4.55 ± 0.08† Bjurböle L/LL4 S1 W2 Fall Average 2.80 3.49 19.8% 4.52 Cynthiana L/LL4 S3 Fall NMNH 2651 86.331 3.15 ± 0.06 3.49 ± 0.01 9.7% ± 1.6% 4.68 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 442 A 38.224 3.23 ± 0.04 3.55 ± 0.01 9.1% ± 1.2% 4.71 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 442 B 22.827 3.23 ± 0.04 3.55 ± 0.02 9.0% ± 1.3% 4.74 ± 0.08 Holbrook L/LL6 S2 Fall NMNH 2817 78.069 3.21 ± 0.03 3.56 ± 0.01 9.8% ± 0.9% 4.75 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 437 28.165 3.41 ± 0.04 3.53 ± 0.03 3.4% ± 1.5% 4.69 ± 0.12 Holbrook L/LL6 S2 Fall NMNH 569 A 64.564 3.13 ± 0.03 3.55 ± 0.01 11.8% ± 0.8% 4.75 ± 0.10 Holbrook L/LL6 S2 Fall NMNH 569 B 50.296 3.14 ± 0.03 3.56 ± 0.01 11.6% ± 0.9% 4.71 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 569 C 71.107 3.11 ± 0.02 3.54 ± 0.01 12.2% ± 0.7% 4.71 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 569 D 46.938 3.12 ± 0.03 3.53 ± 0.01 11.6% ± 0.8% 4.62 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 569 E 40.58 3.13 ± 0.04 3.53 ± 0.01 11.3% ± 1.1% 4.65 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 569 F 43.562 3.17 ± 0.03 3.49 ± 0.01 9.3% ± 0.9% 4.60 ± 0.09 Holbrook L/LL6 S2 Fall NMNH 569 G 18.475 3.19 ± 0.04 3.55 ± 0.03 10.1% ± 1.4% 4.74 ± 0.08 Holbrook L/LL6 S2 Fall NMNH 569 H 84.403 3.18 ± 0.03 3.55 ± 0.01 10.4% ± 0.8% 4.88 ± 0.09 Holbrook L/LL6 S2 Fall CMS H70 A 45.21 3.14 ± 0.06 3.39 ± 0.02 7.1% ± 1.9% 4.16 ± 0.12 Holbrook L/LL6 S2 Fall CMS H70 B 39.71 3.17 ± 0.06 3.37 ± 0.02 5.9% ± 1.8% 4.25 ± 0.10 Holbrook L/LL6 S2 Fall CMS H59 42.4 3.17 ± 0.03 3.42 ± 0.01 7.5% ± 1.0% 4.22 ± 0.09 Holbrook L/LL6 S2 Fall CMS H72 30.66 3.10 ± 0.05 3.51 ± 0.02 11.7% ± 1.4% 4.63 ± 0.09 Holbrook L/LL6 S2 Fall CMS H52 A 36.45 3.23 ± 0.05 3.42 ± 0.02 5.4% ± 1.7% 4.27 ± 0.12 Holbrook L/LL6 S2 Fall CMS H52 B 28.35 3.15 ± 0.03 3.40 ± 0.02 7.6% ± 1.1% 4.10 ± 0.08 Holbrook L/LL6 S2 Fall CMS H52 C 20.74 3.17 ± 0.05 3.40 ± 0.02 6.8% ± 1.7% 4.25 ± 0.08 Holbrook L/LL6 S2 Fall CMS CMS57A 35.53 3.00 ± 0.03 3.47 ± 0.01 13.5% ± 1.0% 4.46 ± 0.09 Holbrook L/LL6 S2 Fall Vatican 410 18.21 3.54 ± 0.04 3.59 ± 0.04 1.5% ± 1.4% 4.23 ± 0.08† Holbrook L/LL6 S2 Fall Vatican 402 34.65 3.22 ± 0.02 3.52 ± 0.04 8.6% ± 1.1% 4.39 ± 0.09 Holbrook L/LL6 S2 Fall Vatican 413 9.15 3.28 ± 0.09 3.69 ± 0.04 11.1% ± 2.5% 4.63 ± 0.08† Holbrook L/LL6 S2 Fall Vatican 403 33.81 3.25 ± 0.04 3.64 ± 0.04 10.8% ± 1.3% 4.64 ± 0.09 Holbrook L/LL6 S2 Fall Vatican 399 53.83 3.20 ± 0.01 3.78 ± 0.04 15.3% ± 0.9% 4.70 ± 0.09 Holbrook L/LL6 S2 Fall Vatican 406 24.801 3.23 ± 0.03 3.64 ± 0.04 11.2% ± 1.2% 4.62 ± 0.08 Holbrook L/LL6 S2 Fall Vatican 407 23.92 3.13 ± 0.03 3.66 ± 0.04 14.6% ± 1.2% 4.75 ± 0.08 Holbrook L/LL6 S2 Fall Vatican 415 9.143 3.32 ± 0.08 3.67 ± 0.04 9.5% ± 2.3% 4.67 ± 0.08† Holbrook L/LL6 S2 Fall Vatican 419 7.267 3.44 ± 0.08 3.65 ± 0.04 5.7% ± 2.4% 4.68 ± 0.08† Holbrook L/LL6 S2 Fall Vatican 401 40.45 3.20 ± 0.02 3.61 ± 0.04 11.5% ± 1.0% 4.67 ± 0.09 Holbrook L/LL6 S2 Fall Vatican 398 77.78 3.19 ± 0.01 3.63 ± 0.04 12.0% ± 1.0% 4.67 ± 0.09 Holbrook L/LL6 S2 Fall Vatican 400 49.65 3.19 ± 0.02 3.64 ± 0.04 12.4% ± 1.0% 4.62 ± 0.09 Holbrook L/LL6 S2 Fall Vatican 405 27.27 3.19 ± 0.03 3.72 ± 0.04 14.1% ± 1.2% 4.59 ± 0.08 Holbrook L/LL6 S2 Fall Vatican 408 21.32 3.16 ± 0.04 3.72 ± 0.04 15.0% ± 1.3% 4.74 ± 0.08 Holbrook L/LL6 S2 Fall Vatican 416 8.301 3.28 ± 0.12 3.69 ± 0.04 11.0% ± 3.5% 4.78 ± 0.08† Holbrook L/LL6 S2 Fall Vatican 397 148.02 3.17 ± 0.03 n.d. n.d. 4.70 ± 0.09 Holbrook L/LL6 S2 Fall Average 3.18 3.55 10.4% 4.58 Knyahinyaa L/LL5 S3 W2-3 Fall NMNH 2855 87.408 3.35 ± 0.06 3.45 ± 0.01 2.8% ± 1.9% 4.60 ± 0.09 Knyahinya L/LL5 S3 W2-3 Fall Vatican 510 45.78 3.35 ± 0.02 3.54 ± 0.04 5.3% ± 1.1% 4.80 ± 0.10 Knyahinya L/LL5 S3 W2-3 Fall Vatican 509 60.96 3.41 ± 0.03 3.53 ± 0.04 3.3% ± 1.4% 4.82 ± 0.10 Knyahinya L/LL5 S3 W2-3 Fall Vatican 511 27.19 3.35 ± 0.02 3.54 ± 0.04 5.4% ± 1.1% 4.69 ± 0.08 Knyahinya L/LL5 S3 W2-3 Fall Vatican 764 9.36 3.45 ± 0.07 3.51 ± 0.04 1.5% ± 2.1% 4.74 ± 0.08† Knyahinya L/LL5 S3 W2-3 Fall Vatican 1126 9.42 3.62 ± 0.15 3.54 ± 0.04 -2.3% ± 4.5% 4.76 ± 0.08† Knyahinya L/LL5 S3 W2-3 Fall Vatican 508 491.73 3.35 ± 0.07‡ n.d. n.d. n.d. Knyahinya L/LL5 S3 W2-3 Fall Average 3.36 3.50 4.0% 4.73 Sultanpur L/LL6 Fall NMNH 1255 58.898 3.46 ± 0.04 3.57 ± 0.01 3.2% ± 1.1% 4.96 ± 0.09 Vera L/LL4 Find NMNH 2220 26.863 3.48 ± 0.13 3.46 ± 0.03 -0.4% ± 3.9% 4.91 ± 0.12 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study. aBreccia (Grady, 2000)

272

Table 11: Average Porosities of Brecciated and Non-Brecciated OC Falls.

No. of Meteorites Average Porosity σ σmean Min Max Breccias 54 8.5% 4.5% 0.6% 1.3% 26.6% Veined 55 8.2% 3.5% 0.5% 1.5% 16.9% Non-Breccias 152 9.3% 4.6% 0.4% 0.0% 26.6%

273

Table 12: Data for K and R Chondrites.

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Kakangari K3 Fall LNHM BM69062 89.03 3.04 ± 0.05 3.45 ± 0.01 12.0% ± 1.5% 5.09 ± 0.10 NWA 5469 R5 S4 W0 Find CMS 1619-2 21.21 3.32 ± 0.08 3.57 ± 0.04 7.0% ± 2.6% 3.63 ± 0.12 NWA 753 R3.9 S2 W2 Find Monnig M1038.4 24.49 3.30 ± 0.05 3.41 ± 0.02 3.1% ± 1.6% 3.13 ± 0.10 NWA 753 R3.9 S2 W2 Find Vatican 1450 15.35 3.28 ± 0.04 3.63 ± 0.09 9.5% ± 2.4% 3.23 ± 0.09 NWA 753 R3.9 S2 W2 Find Average 3.30 3.49 5.6% 3.18 NWA 800 R Find Monnig M1042.1 45.96 3.04 ± 0.03 3.54 ± 0.02 14.1% ± 0.9% 2.87 ± 0.09 NWA 851 R4 W4 Find Monnig M1182.1 14.69 3.18 ± 0.06 3.51 ± 0.05 9.2% ± 2.2% 3.15 ± 0.12 NWA 978 R3.8 S3 W2 Find Monnig M1040.2 14.4 3.40 ± 0.10 3.61 ± 0.04 6.0% ± 2.9% 3.03 ± 0.10 NWA 978 R3.8 S3 W2 Find Monnig M1040.3 29.16 3.15 ± 0.03 3.58 ± 0.02 11.9% ± 1.1% 3.18 ± 0.09 NWA 978 R3.8 S3 W2 Find Average 3.23 3.59 10.0% 3.11 Ouzina R4 S2 W4 Find Monnig M 914.2 31.51 2.80 ± 0.03 3.44 ± 0.02 18.7% ± 0.9% 2.72 ± 0.09 Ouzina R4 S2 W4 Find LNHM BM2000,M28 27.59 2.78 ± 0.05 3.46 ± 0.02 19.7% ± 1.5% 2.72 ± 0.09 Ouzina R4 S2 W4 Find Average 2.79 3.45 19.2% 2.72

274

Table 13: Data for Carbonaceous Chondrites.

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Ningqiang C3 Fall AMNH 4893 A 71.18 2.80 ± 0.11 n.d. n.d. 4.67 ± 0.10 Ningqiang C3 Fall NMNH 6546 37.04 2.79 ± 0.03 3.66 ± 0.02 23.9% ± 0.9% 4.72 ± 0.09 Ningqiang C3 Fall Average 2.80 3.66 23.64% 4.69 Adelaide C2 Find LNHM BM1995,M4 23.15 3.26 ± 0.07 3.31 ± 0.02 1.8% ± 2.1% 4.88 ± 0.10 Adelaide C2 Find FMNH ME 2923 #1 14.75 3.04 ± 0.09 3.20 ± 0.04 5.0% ± 3.0% 4.79 ± 0.12 Adelaide C2 Find Average 3.17 3.27 3.07% 4.83 Coolidge C4 S2 Find Monnig M 633.1 22.19 3.55 ± 0.10 3.51 ± 0.02 -1.0% ± 3.1% 5.14 ± 0.12 Dar al Gani 430 C3 Find LNHM BM2000,M36 35.25 2.92 ± 0.03 3.61 ± 0.01 19.3% ± 0.9% 4.79 ± 0.12 Loongana 001 C4 Find Monnig M1259.1 50.48 3.10 ± 0.06 3.28 ± 0.01 5.2% ± 2.0% 4.32 ± 0.09 NWA 1152 C3 S2 W2/3 Find Monnig M 924.1 21.06 2.83 ± 0.05 3.51 ± 0.02 19.5% ± 1.5% 4.81 ± 0.08 NWA 1152 C3 S2 W2/3 Find LNHM BM2001,M26 9.53 2.86 ± 0.09 3.46 ± 0.05 17.5% ± 2.7% 4.82 ± 0.08 NWA 1152 C3 S2 W2/3 Find Average 2.84 3.49 18.88% 4.81 Bencubbin CBa Find AMNH 4206 182.66 4.90 ± 0.08 n.d. n.d. 5.79 ± 0.09 Bencubbin CBa Find Monnig M 888.1 58.03 5.55 ± 0.09 5.66 ± 0.03 2.0% ± 1.7% 5.56 ± 0.12 Bencubbin CBa Find Average 5.05 5.66 10.84% 5.68 HaH 237 CBb S3 W2 Find LNHM BM2000,M9 31.04 5.31 ± 0.10 5.63 ± 0.04 5.8% ± 1.9% 5.61 ± 0.12 HaH 237 CBb S3 W2 Find Vatican 1166 8.63 n.d. n.d. n.d. 5.31 ± 0.08† HaH 237 CBb S3 W2 Find Average 5.31 5.63 5.82% 5.46 Acfer 214 CH3 S2 W2 Find Monnig M1315.1 31.51 3.74 ± 0.09 3.66 ± 0.02 -2.3% ± 2.4% 5.22 ± 0.12 Acfer 214 CH3 S2 W2 Find LNHM BM1998,M28 11.26 3.84 ± 0.19 3.65 ± 0.05 -5.1% ± 5.3% 5.39 ± 0.08 Acfer 214 CH3 S2 W2 Find Average 3.77 3.65 -3.07% 5.30 Orgueil CI1 Fall Vatican 719 14.00 n.d. n.d. n.d. 4.11 ± 0.08† Orgueil CI1 Fall Vatican 718 47.20 1.57 ± 0.03‡ 2.42 ± 0.06‡ 34.9% ± 2.1% 4.86 ± 0.08† Orgueil CI1 Fall Average 1.57 2.42 34.93% 4.49 Karoonda CK4 S1 Fall AMNH 3970 A 28.10 2.94 ± 0.11 3.51 ± 0.04 16.4% ± 3.3% 4.69 ± 0.09 Karoonda CK4 S1 Fall AMNH 3970 B 24.80 2.80 ± 0.11 3.61 ± 0.04 22.3% ± 3.1% 4.62 ± 0.08 Karoonda CK4 S1 Fall AMNH 3970 C 23.70 2.76 ± 0.11 3.59 ± 0.04 23.2% ± 3.1% 4.67 ± 0.08 Karoonda CK4 S1 Fall AMNH 3970 D 23.60 2.88 ± 0.11 3.60 ± 0.04 20.2% ± 3.2% 4.66 ± 0.08 Karoonda CK4 S1 Fall AMNH 3970 E 22.70 2.77 ± 0.11 3.60 ± 0.04 23.0% ± 3.1% 4.72 ± 0.08 Karoonda CK4 S1 Fall AMNH 3970 F 19.90 2.76 ± 0.11 3.60 ± 0.04 23.4% ± 3.1% 4.67 ± 0.08 Karoonda CK4 S1 Fall AMNH 3970 G 24.60 2.87 ± 0.11 3.65 ± 0.04 21.5% ± 3.1% 4.71 ± 0.08 Karoonda CK4 S1 Fall AMNH 3970 H 18.60 2.83 ± 0.11 3.57 ± 0.04 20.7% ± 3.2% 4.72 ± 0.08 Karoonda CK4 S1 Fall AMNH 4774 37.50 2.83 ± 0.11 n.d. n.d. 4.66 ± 0.09 Karoonda CK4 S1 Fall AMNH 20.20 2.97 ± 0.11 n.d. n.d. n.d. Karoonda CK4 S1 Fall NMNH 5281 35.57 2.87 ± 0.04 3.60 ± 0.01 20.4% ± 1.1% 4.68 ± 0.09 Karoonda CK4 S1 Fall FMNH ME 2666 #1 20.15 2.90 ± 0.04 3.62 ± 0.03 19.9% ± 1.1% 4.59 ± 0.08 Karoonda CK4 S1 Fall Average 2.85 3.60 20.79% 4.67 Dar al Gani 275 CK4/5 S2 W4 Find Monnig M 602.1 19.52 3.28 ± 0.13 3.60 ± 0.03 8.8% ± 3.7% 4.65 ± 0.12 Dar al Gani 431 CK3 Find Monnig M 952.1 21.76 3.39 ± 0.06 3.37 ± 0.03 -0.6% ± 2.1% 4.60 ± 0.12 Maralinga CK4 S2 Find AMNH 4742 24.99 2.54 ± 0.05 n.d. n.d. 4.77 ± 0.12 Maralinga CK4 S2 Find LNHM BM1991,M23 12.40 2.94 ± 0.07 3.46 ± 0.04 15.0% ± 2.4% 4.68 ± 0.12 Maralinga CK4 S2 Find Average 2.66 3.46 23.25% 4.72 NWA 2388 CK6 Find Monnig M1274.1 14.32 3.13 ± 0.10 3.66 ± 0.06 14.3% ± 3.1% 4.62 ± 0.12 NWA 2867 CK4 S1 W0/1 Find Monnig M 376.1 48.83 2.98 ± 0.03 3.62 ± 0.01 17.6% ± 0.9% 4.70 ± 0.09 NWA 5515 CK4 Find Vatican 1490 431.58 2.70 ± 0.08 n.d. n.d. n.d. Banten CM2 Fall NMNH 6017 24.99 2.40 ± 0.03 2.95 ± 0.01 18.8% ± 1.0% 4.10 ± 0.10 Banten CM2 Fall LNHM BM1984,M5 16.59 2.40 ± 0.05 2.95 ± 0.02 18.5% ± 1.9% 4.11 ± 0.08 Banten CM2 Fall Average 2.40 2.95 18.66% 4.11 Cold Bokkeveld CM2 S1 Fall NMNH 6332 18.12 2.42 ± 0.03 2.80 ± 0.02 13.7% ± 1.1% 3.74 ± 0.09 Cold Bokkeveld CM2 S1 Fall LNHM BM1985,M147 42.38 2.39 ± 0.02 2.77 ± 0.01 13.7% ± 0.9% 3.76 ± 0.09 Cold Bokkeveld CM2 S1 Fall FMNH ME 1736 #1 16.02 2.25 ± 0.03 2.80 ± 0.02 19.6% ± 1.1% 3.53 ± 0.09 Cold Bokkeveld CM2 S1 Fall Average 2.36 2.78 15.01% 3.68 Crescent CM2 S1 Fall NMNH 6927 14.99 2.03 ± 0.03 2.99 ± 0.02 32.0% ± 1.1% 4.40 ± 0.12 Crescent CM2 S1 Fall Monnig M 16.1 24.32 2.09 ± 0.03 3.01 ± 0.02 30.7% ± 0.9% 4.50 ± 0.10 Crescent CM2 S1 Fall Average 2.07 3.00 31.19% 4.45 Essebi CM2 S1 Fall NMNH 3200 46.07 n.d. 2.80 ± 0.01 n.d. 4.77 ± 0.09 Mighei CM2 S1 Fall NMNH 3483 27.07 1.93 ± 0.02 2.85 ± 0.01 32.2% ± 0.8% 3.41 ± 0.10 Mighei CM2 S1 Fall Vatican 616 1.88 n.d. 4.70 ± 1.18 n.d. 4.04 ± 0.08† Mighei CM2 S1 Fall Average 1.93 2.93 33.96% 3.72 Murchison CM2 S1-2 Fall AMNH 4376 56.90 2.23 ± 0.09 2.96 ± 0.03 24.5% ± 3.0% n.d. Murchison CM2 S1-2 Fall AMNH 4377 A 6.96 2.34 ± 0.06 3.05 ± 0.05 23.3% ± 2.3% 3.44 ± 0.09 Murchison CM2 S1-2 Fall AMNH 4720 40.70 2.30 ± 0.02 3.04 ± 0.03 24.3% ± 1.1% n.d. Murchison CM2 S1-2 Fall Britt 47.90 2.32 ± 0.01 3.03 ± 0.03 23.4% ± 0.9% n.d. Murchison CM2 S1-2 Fall NMNH 5349 38.48 2.15 ± 0.02 2.87 ± 0.01 24.9% ± 0.6% 3.54 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5356 49.38 2.35 ± 0.02 2.97 ± 0.01 20.9% ± 0.7% 3.90 ± 0.09

275

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Murchison CM2 S1-2 Fall NMNH 5366 27.42 2.34 ± 0.02 2.94 ± 0.01 20.4% ± 0.9% 3.72 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5369 26.92 2.40 ± 0.02 2.95 ± 0.01 18.7% ± 0.9% 3.73 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5370 27.46 2.40 ± 0.02 2.97 ± 0.01 19.2% ± 0.9% 3.85 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5373 36.55 2.35 ± 0.02 2.96 ± 0.01 20.7% ± 0.7% 3.84 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5375 24.34 2.27 ± 0.02 2.93 ± 0.01 22.7% ± 0.8% 3.72 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5381 46.81 2.39 ± 0.02 2.94 ± 0.01 18.8% ± 0.7% 3.74 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5384 33.63 2.24 ± 0.03 2.93 ± 0.01 23.7% ± 0.9% 3.74 ± 0.09 Murchison CM2 S1-2 Fall NMNH 5398 20.66 2.34 ± 0.04 2.96 ± 0.02 20.9% ± 1.4% 3.79 ± 0.09 Murchison CM2 S1-2 Fall Average 2.31 2.96 22.10% 3.73 Murray CM2 S1 Fall NMNH 1769 18 62.05 2.30 ± 0.02 2.90 ± 0.01 20.7% ± 0.7% 3.36 ± 0.09 Murray CM2 S1 Fall NMNH 1769 21 50.79 2.35 ± 0.02 2.88 ± 0.01 18.4% ± 0.8% 3.61 ± 0.09 Murray CM2 S1 Fall NMNH 1769 26 43.05 2.30 ± 0.02 2.91 ± 0.01 21.2% ± 0.7% 3.83 ± 0.09 Murray CM2 S1 Fall NMNH 1769 27 42.50 2.32 ± 0.02 2.88 ± 0.01 19.4% ± 0.8% 3.67 ± 0.09 Murray CM2 S1 Fall NMNH 1769 28 36.18 2.23 ± 0.02 2.89 ± 0.01 22.8% ± 0.7% 3.67 ± 0.09 Murray CM2 S1 Fall NMNH 1769 29 36.14 2.28 ± 0.02 2.93 ± 0.01 22.2% ± 0.8% 3.83 ± 0.09 Murray CM2 S1 Fall NMNH 1769 33 32.45 2.43 ± 0.03 2.94 ± 0.01 17.2% ± 1.1% 3.85 ± 0.09 Murray CM2 S1 Fall NMNH 1769 37 28.38 2.26 ± 0.02 2.89 ± 0.01 21.9% ± 0.8% 3.66 ± 0.09 Murray CM2 S1 Fall NMNH 1769 42 23.30 2.31 ± 0.03 2.89 ± 0.01 20.0% ± 1.0% 3.71 ± 0.09 Murray CM2 S1 Fall NMNH 1769 45 23.04 2.34 ± 0.03 2.91 ± 0.01 19.4% ± 1.1% 3.48 ± 0.10 Murray CM2 S1 Fall IOM Cr 1.10 56.77 2.26 ± 0.02 2.93 ± 0.01 23.0% ± 0.7% 3.53 ± 0.09 Murray CM2 S1 Fall FMNH ME 2612 #1 93.35 2.30 ± 0.04 2.92 ± 0.01 21.3% ± 1.2% 3.71 ± 0.09 Murray CM2 S1 Fall Average 2.30 2.91 20.82% 3.66 Nawapali CM2 Fall LNHM BM82968 18.47 2.17 ± 0.03 2.74 ± 0.02 20.9% ± 1.2% 3.30 ± 0.09 Nogoyá CM2 S1 Fall LNHM BM1985,M154 20.90 2.17 ± 0.04 2.78 ± 0.02 21.9% ± 1.5% 3.67 ± 0.09 Nogoyá CM2 S1 Fall FMNH ME 2670 #1 34.58 2.14 ± 0.02 2.79 ± 0.01 23.3% ± 0.8% 3.66 ± 0.09 Nogoyá CM2 S1 Fall Average 2.15 2.79 22.75% 3.66 Pollen CM2 S1 Fall LNHM BM1964,496 19.53 2.22 ± 0.03 2.82 ± 0.02 21.1% ± 1.2% 3.65 ± 0.10 Santa Cruz CM2 S1 Fall LNHM BM1959,782 10.80 1.88 ± 0.03 2.96 ± 0.03 36.7% ± 1.4% 3.61 ± 0.10 Cimarron CM2 Find AMNH 4861 A 6.94 2.14 ± 0.05 3.03 ± 0.03 29.3% ± 1.8% 4.48 ± 0.08 El-Quss Abu Said CM2 S1 W0 Find LNHM BM2002,M24 24.98 2.47 ± 0.04 3.26 ± 0.02 24.2% ± 1.2% 4.32 ± 0.09 Felix CO3.3 S3 Fall NMNH 235 10.72 3.00 ± 0.06 3.67 ± 0.05 18.1% ± 1.9% 4.56 ± 0.12 Felix CO3.3 S3 Fall Monnig M 734.1 25.66 2.83 ± 0.03 3.65 ± 0.03 22.5% ± 1.1% 4.60 ± 0.12 Felix CO3.3 S3 Fall FMNH ME 1330 #15 20.47 2.89 ± 0.04 3.67 ± 0.03 21.2% ± 1.3% 4.67 ± 0.08 Felix CO3.3 S3 Fall Average 2.88 3.66 21.25% 4.61 Kainsaz CO3.2 S1 Fall NMNH 2486 22.26 3.20 ± 0.04 3.70 ± 0.02 13.5% ± 1.2% 50 ± 0.08 Kainsaz CO3.2 S1 Fall Monnig M 528.2 126.39 3.22 ± 0.05 3.30 ± 0.01 2.6% ± 1.6% 4.58 ± 0.09 Kainsaz CO3.2 S1 Fall Monnig M 528.1 93.81 3.19 ± 0.06 3.56 ± 0.01 10.2% ± 1.6% 4.75 ± 0.12 Kainsaz CO3.2 S1 Fall FMNH ME 2755 #3 90.45 3.09 ± 0.05 3.70 ± 0.01 16.4% ± 1.4% 4.91 ± 0.10 Kainsaz CO3.2 S1 Fall Average 3.17 3.50 9.32% 4.81 Lancé CO4 S1 Fall NMNH 2874 35.01 3.30 ± 0.04 3.54 ± 0.01 6.9% ± 1.1% 4.45 ± 0.09 Lancé CO4 S1 Fall IOM Cr 2.1 27.22 3.18 ± 0.04 3.56 ± 0.02 10.5% ± 1.3% 4.52 ± 0.08 Lancé CO4 S1 Fall Vatican 540 7.63 3.48 ± 0.19 3.78 ± 0.04 7.8% ± 5.1% 4.35 ± 0.08† Lancé CO4 S1 Fall FMNH ME 589 #1 48.72 3.18 ± 0.03 3.54 ± 0.01 10.2% ± 0.8% 4.41 ± 0.10 Lancé CO4 S1 Fall Average 3.23 3.56 9.17% 4.43 Moss CO3.6 S2 Fall Monnig M 175.2 19.30 3.04 ± 0.06 3.74 ± 0.03 18.6% ± 1.7% 4.79 ± 0.08 Ornans CO3.4 S1 Fall Monnig M 253.1 17.88 2.63 ± 0.04 3.66 ± 0.03 28.1% ± 1.2% 4.23 ± 0.12 Ornans CO3.4 S1 Fall IOM Cr 5.1 16.33 2.61 ± 0.04 3.62 ± 0.03 27.8% ± 1.4% 4.40 ± 0.08 Ornans CO3.4 S1 Fall Vatican 723 25.26 2.18 ± 0.05 3.71 ± 0.04 41.3% ± 1.5% 4.29 ± 0.08† Ornans CO3.4 S1 Fall Average 2.42 3.67 34.23% 4.31 Warrenton CO3.7 S1 Fall AMNH 376 56.50 2.75 ± 0.10 3.75 ± 0.04 26.7% ± 2.9% 4.44 ± 0.09 Warrenton CO3.7 S1 Fall NMNH 1177 25.55 2.78 ± 0.03 3.67 ± 0.02 24.1% ± 1.0% 4.44 ± 0.08 Warrenton CO3.7 S1 Fall Vatican 1018 10.91 3.23 ± 0.10 n.d. n.d. 4.46 ± 0.08† Warrenton CO3.7 S1 Fall FMNH ME 1720 #1 78.25 2.82 ± 0.04 3.67 ± 0.01 23.3% ± 1.2% 4.50 ± 0.09 Warrenton CO3.7 S1 Fall Average 2.81 3.70 23.96% 4.46 Colony CO3.0 S1 Find AMNH 4595 22.30 2.90 ± 0.11 3.33 ± 0.04 13.0% ± 3.5% 4.58 ± 0.08 Colony CO3.0 S1 Find NMNH 6264 28.24 3.03 ± 0.06 3.22 ± 0.02 5.9% ± 2.1% 4.52 ± 0.12 Colony CO3.0 S1 Find FMNH ME 3089 #1 25.91 2.71 ± 0.04 2.99 ± 0.02 9.4% ± 1.4% 4.47 ± 0.12 Colony CO3.0 S1 Find Average 2.87 3.17 9.22% 4.52 Dar al Gani 005 CO3 S2 W2 Find AMNH 4916 36.50 3.21 ± 0.12 3.35 ± 0.03 4.0% ± 3.8% 4.62 ± 0.09 Dar al Gani 023 CO3 S2 W2 Find AMNH 4931 24.40 3.24 ± 0.12 3.31 ± 0.04 2.1% ± 3.9% 4.63 ± 0.08 Dar al Gani 078 CO3 S2 W2 Find AMNH 4917 21.50 3.14 ± 0.12 3.34 ± 0.04 6.0% ± 3.8% 4.80 ± 0.08 Dar al Gani 078 CO3 S2 W2 Find IOM Cr 23.1 14.16 3.16 ± 0.08 3.32 ± 0.03 4.8% ± 2.6% 4.59 ± 0.12 Dar al Gani 078 CO3 S2 W2 Find Average 3.15 3.33 5.53% 4.70 Dar al Gani 749 CO3 W2 Find Monnig M 972.1 51.35 3.30 ± 0.03 3.38 ± 0.01 2.3% ± 0.9% 4.63 ± 0.09 Dar al Gani 749 CO3 W2 Find PSF PSF00-002 35.77 3.29 ± 0.04 3.38 ± 0.01 2.8% ± 1.4% 4.59 ± 0.09 Dar al Gani 749 CO3 W2 Find Average 3.30 3.38 2.51% 4.61 Isna CO4 Find NMNH 5890 A 91.32 3.07 ± 0.02 3.54 ± 0.01 13.3% ± 0.7% 3.91 ± 0.12 Isna CO4 Find Monnig M 733.1 22.47 3.08 ± 0.04 3.57 ± 0.02 13.7% ± 1.3% 3.96 ± 0.12 Isna CO4 Find FMNH ME 2799 #1 17.99 2.88 ± 0.05 3.66 ± 0.03 21.2% ± 1.4% 4.21 ± 0.12

276

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Isna CO4 Find Average 3.04 3.56 14.53% 4.03 NWA 062 CO3.4 S1 Find Monnig M1058.1 21.11 3.18 ± 0.06 3.55 ± 0.04 10.2% ± 1.9% 4.22 ± 0.12 Rainbow CO3.2 Find NMNH 6981 16.35 3.26 ± 0.07 3.19 ± 0.04 -2.1% ± 2.6% 3.96 ± 0.12 Al Rais CR2 S1 Fall LNHM BM1971,289 10.82 2.29 ± 0.04 3.06 ± 0.04 25.0% ± 1.6% 4.89 ± 0.12 Renazzo CR2 S1-3 Fall NMNH 2196 15.86 3.19 ± 0.06 3.31 ± 0.03 3.7% ± 2.0% 5.06 ± 0.10 Renazzo CR2 S1-3 Fall LNHM BM41105 10.94 3.06 ± 0.08 3.30 ± 0.04 7.3% ± 2.8% 5.17 ± 0.08 Renazzo CR2 S1-3 Fall Vatican 792 12.18 2.89 ± 0.08 3.54 ± 0.04 18.2% ± 2.3% 5.10 ± 0.08† Renazzo CR2 S1-3 Fall Average 3.05 3.37 9.50% 5.11 Acfer 097 CR2 S2 W2 Find AMNH 4808 15.44 3.06 ± 0.12 3.54 ± 0.04 13.7% ± 3.5% 5.06 ± 0.08 Acfer 270 CR2 S2 W2 Find Monnig M 966.1 27.24 3.26 ± 0.07 3.46 ± 0.02 6.0% ± 2.2% 5.07 ± 0.10 Dhofar 1432 CR2 S2 W mod Find Monnig M1358.1 36.93 3.19 ± 0.06 3.37 ± 0.02 5.5% ± 1.9% 4.64 ± 0.12 El Djouf 001 CR2 S2 W2 Find Monnig M1316.1 23.77 3.09 ± 0.05 3.35 ± 0.02 7.7% ± 1.6% 4.89 ± 0.08 Tafassasset CR Find NMNH 7100 27.29 3.94 ± 0.05 3.88 ± 0.02 -1.5% ± 1.5% 5.34 ± 0.08 Allende CV3o S1 Fall AMNH 4286 70.20 2.89 ± 0.05 3.68 ± 0.04 21.4% ± 1.6% 3.61 ± 0.10 Allende CV3o S1 Fall AMNH 4287 A 82.43 2.78 ± 0.05 3.69 ± 0.04 24.7% ± 1.6% 3.72 ± 0.10 Allende CV3o S1 Fall AMNH 4287 B 152.89 2.83 ± 0.05 n.d. n.d. 3.58 ± 0.10 Allende CV3o S1 Fall AMNH 4288 95.32 2.82 ± 0.05 3.65 ± 0.04 22.6% ± 1.6% 3.69 ± 0.10 Allende CV3o S1 Fall AMNH 4290 A 118.99 2.81 ± 0.05 n.d. n.d. 3.63 ± 0.10 Allende CV3o S1 Fall AMNH 4290 B 123.56 2.89 ± 0.05 3.65 ± 0.04 20.8% ± 1.6% 3.62 ± 0.10 Allende CV3o S1 Fall AMNH 4293 A 55.84 n.d. 3.65 ± 0.04 n.d. 3.59 ± 0.10 Allende CV3o S1 Fall AMNH 4293 B 46.04 2.88 ± 0.06 3.64 ± 0.04 21.1% ± 1.7% 3.63 ± 0.10 Allende CV3o S1 Fall AMNH 4297 A 25.41 2.98 ± 0.05 3.66 ± 0.04 18.6% ± 1.5% n.d. Allende CV3o S1 Fall AMNH 4297 B 40.21 2.91 ± 0.03 3.69 ± 0.04 21.1% ± 1.2% n.d. Allende CV3o S1 Fall AMNH 4297 C 35.40 2.87 ± 0.03 3.68 ± 0.04 22.1% ± 1.2% n.d. Allende CV3o S1 Fall AMNH 4297 D 27.54 2.93 ± 0.05 3.73 ± 0.04 21.5% ± 1.5% n.d. Allende CV3o S1 Fall AMNH 4298 111.42 2.88 ± 0.05 3.65 ± 0.04 21.0% ± 1.6% n.d. Allende CV3o S1 Fall AMNH 4299 A 88.75 2.79 ± 0.06 3.66 ± 0.04 23.7% ± 1.7% 3.68 ± 0.10 Allende CV3o S1 Fall AMNH 4299 B 83.03 2.81 ± 0.06 3.64 ± 0.04 22.8% ± 1.8% 3.68 ± 0.10 Allende CV3o S1 Fall AMNH 4300 A 32.48 2.79 ± 0.05 3.67 ± 0.04 24.0% ± 1.7% 3.70 ± 0.10 Allende CV3o S1 Fall AMNH 4300 B 25.14 2.74 ± 0.07 3.80 ± 0.04 27.9% ± 2.0% 3.70 ± 0.10 Allende CV3o S1 Fall AMNH 4305 A 87.22 2.84 ± 0.05 3.69 ± 0.04 23.2% ± 1.6% 3.63 ± 0.10 Allende CV3o S1 Fall AMNH 4305 B 67.79 2.82 ± 0.01 3.71 ± 0.04 24.2% ± 0.9% 3.63 ± 0.10 Allende CV3o S1 Fall AMNH 4306 A 60.32 2.88 ± 0.03 3.66 ± 0.04 21.2% ± 1.2% 3.65 ± 0.10 Allende CV3o S1 Fall AMNH 4306 B 49.43 2.95 ± 0.04 3.62 ± 0.04 18.7% ± 1.3% 3.72 ± 0.10 Allende CV3o S1 Fall AMNH 4307 A 75.95 2.84 ± 0.01 3.66 ± 0.04 22.5% ± 0.9% 3.65 ± 0.10 Allende CV3o S1 Fall AMNH 4307 B 38.67 2.66 ± 0.12 3.66 ± 0.04 27.5% ± 3.4% 3.65 ± 0.10 Allende CV3o S1 Fall AMNH 4308 A 101.25 2.78 ± 0.05 3.65 ± 0.04 23.7% ± 1.5% 3.61 ± 0.10 Allende CV3o S1 Fall AMNH 4308 B 53.79 2.68 ± 0.06 3.63 ± 0.04 26.1% ± 1.8% 3.72 ± 0.10 Allende CV3o S1 Fall AMNH 4309 A 32.45 2.91 ± 0.05 3.65 ± 0.04 20.2% ± 1.7% n.d. Allende CV3o S1 Fall AMNH 4309 B 43.90 3.00 ± 0.05 3.65 ± 0.04 17.6% ± 1.5% n.d. Allende CV3o S1 Fall AMNH 4310 A 50.25 2.80 ± 0.11 3.86 ± 0.04 27.3% ± 2.9% 3.64 ± 0.10 Allende CV3o S1 Fall AMNH 4310 B 43.02 2.76 ± 0.11 3.66 ± 0.04 24.7% ± 3.0% 3.65 ± 0.10 Allende CV3o S1 Fall AMNH 4314 A 42.18 2.93 ± 0.04 3.66 ± 0.04 20.0% ± 1.4% 3.64 ± 0.10 Allende CV3o S1 Fall AMNH 4314 B 21.87 2.97 ± 0.08 3.64 ± 0.04 18.3% ± 2.4% 3.68 ± 0.10 Allende CV3o S1 Fall AMNH 4315 A 66.23 2.81 ± 0.02 3.64 ± 0.04 22.6% ± 0.9% 3.70 ± 0.10 Allende CV3o S1 Fall AMNH 4315 B 77.42 2.78 ± 0.02 3.66 ± 0.04 24.0% ± 1.0% 3.69 ± 0.10 Allende CV3o S1 Fall AMNH 4317 A 49.23 2.87 ± 0.04 3.65 ± 0.04 21.3% ± 1.4% 3.65 ± 0.10 Allende CV3o S1 Fall AMNH 4317 B 20.01 2.77 ± 0.08 3.61 ± 0.04 23.3% ± 2.4% 3.73 ± 0.10 Allende CV3o S1 Fall AMNH 4318 A 47.30 2.83 ± 0.04 3.61 ± 0.04 21.7% ± 1.4% n.d. Allende CV3o S1 Fall AMNH 4318 B 44.90 2.85 ± 0.04 3.60 ± 0.04 20.8% ± 1.5% n.d. Allende CV3o S1 Fall AMNH 4324 A 60.18 2.79 ± 0.06 3.66 ± 0.04 23.7% ± 1.9% 3.73 ± 0.10 Allende CV3o S1 Fall AMNH 4324 B 49.77 2.89 ± 0.08 3.63 ± 0.04 20.5% ± 2.3% 3.74 ± 0.10 Allende CV3o S1 Fall NMNH 5556 63.42 2.92 ± 0.02 3.60 ± 0.01 18.9% ± 0.7% 3.61 ± 0.09 Allende CV3o S1 Fall NMNH 5557 61.04 2.97 ± 0.03 3.60 ± 0.01 17.7% ± 0.8% 3.56 ± 0.09 Allende CV3o S1 Fall NMNH 5559 67.07 2.94 ± 0.02 3.59 ± 0.01 18.1% ± 0.7% 3.55 ± 0.09 Allende CV3o S1 Fall NMNH 5567 2 37.00 2.93 ± 0.04 3.62 ± 0.01 18.9% ± 1.2% 3.67 ± 0.09 Allende CV3o S1 Fall NMNH 5567 3 27.32 2.99 ± 0.03 3.60 ± 0.02 16.9% ± 1.0% 3.54 ± 0.09 Allende CV3o S1 Fall NMNH 5567 4 21.42 2.94 ± 0.04 3.59 ± 0.02 17.9% ± 1.2% 3.69 ± 0.09 Allende CV3o S1 Fall NMNH 5568 34.89 2.92 ± 0.03 3.60 ± 0.01 19.0% ± 0.8% 3.72 ± 0.09 Allende CV3o S1 Fall NMNH 5570 1 32.97 2.90 ± 0.03 3.59 ± 0.02 19.1% ± 0.9% 3.53 ± 0.09 Allende CV3o S1 Fall NMNH 5570 3 59.79 2.90 ± 0.03 3.59 ± 0.01 19.2% ± 0.7% 3.62 ± 0.09 Allende CV3o S1 Fall NMNH 5570 4 55.46 2.91 ± 0.03 3.60 ± 0.01 19.2% ± 0.9% 3.65 ± 0.09 Allende CV3o S1 Fall NMNH 5570 5 65.65 2.91 ± 0.03 3.59 ± 0.01 19.2% ± 0.8% 3.60 ± 0.09 Allende CV3o S1 Fall NMNH 5570 7 46.77 2.91 ± 0.03 3.61 ± 0.01 19.3% ± 0.9% 3.53 ± 0.09 Allende CV3o S1 Fall Vatican 28 168.74 2.93 ± 0.03 3.85 ± 0.08 23.8% ± 1.7% 3.69 ± 0.09 Allende CV3o S1 Fall Average 2.86 3.66 21.93% 3.65 Grosnaja CV3o S3 Fall Vatican 373 11.12 3.18 ± 0.11 n.d. n.d. 3.96 ± 0.08† Mokoia CV3o S1 Fall LNHM BM1910,729 14.71 2.59 ± 0.04 3.58 ± 0.04 27.7% ± 1.3% 4.24 ± 0.08 Vigarano CV3r S1-2 Fall AMNH 2226 B 10.62 3.41 ± 0.06 3.38 ± 0.04 -1.1% ± 2.3% 4.30 ± 0.12 Vigarano CV3r S1-2 Fall NMNH 3137 9.02 3.15 ± 0.07 3.50 ± 0.05 10.0% ± 2.4% 4.49 ± 0.12 Vigarano CV3r S1-2 Fall NMNH 477 150.94 3.16 ± 0.04 n.d. n.d. 4.49 ± 0.10 Vigarano CV3r S1-2 Fall Vatican 1002 82.22 3.31 ± 0.10 3.53 ± 0.04 6.2% ± 3.0% 4.46 ± 0.12

277

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Vigarano CV3 r S1-2 Fall Vatican 1496 10.38 n.d. n.d. n.d. 4.39 ± 0.08† Vigarano CV3r S1-2 Fall FMNH ME 782 #1.1 32.73 3.22 ± 0.03 3.49 ± 0.02 7.8% ± 1.0% 4.32 ± 0.10 Vigarano CV3r S1-2 Fall FMNH ME 782 #1.2 23.77 3.22 ± 0.03 3.51 ± 0.02 8.2% ± 1.1% 4.43 ± 0.08 Vigarano CV3r S1-2 Fall Average 3.22 3.51 8.27% 4.41 Axtell CV3o Find IOM Cr 22.1 13.81 2.69 ± 0.10 3.51 ± 0.04 23.4% ± 3.0% 3.23 ± 0.12 Dar al Gani 1040 CV3 S low W mod Find Monnig M1296.1 21.64 2.93 ± 0.06 3.67 ± 0.03 20.2% ± 1.6% 4.84 ± 0.12 Efremovka CV3r S4 Find AMNH 4221 A 106.82 3.51 ± 0.05 3.50 ± 0.01 -0.3% ± 1.6% 4.89 ± 0.10 Efremovka CV3r S4 Find AMNH 4221 B 13.94 3.70 ± 0.08 3.49 ± 0.04 -6.2% ± 2.5% 4.89 ± 0.10 Efremovka CV3r S4 Find AMNH 4221 C 28.65 3.60 ± 0.02 3.54 ± 0.02 -1.8% ± 0.7% 4.93 ± 0.10 Efremovka CV3r S4 Find NMNH 6456 124.12 3.36 ± 0.04 3.44 ± 0.01 2.5% ± 1.1% 4.69 ± 0.09 Efremovka CV3r S4 Find NMNH 7029 15.67 3.35 ± 0.05 3.49 ± 0.03 4.0% ± 1.6% 4.83 ± 0.08 Efremovka CV3r S4 Find Monnig M1051.1 39.24 3.54 ± 0.04 3.50 ± 0.01 -1.1% ± 1.2% 4.91 ± 0.09 Efremovka CV3r S4 Find Average 3.46 3.48 0.55% 4.86 Leoville CV3r S3 Find AMNH 4337 9.30 2.94 ± 0.12 3.39 ± 0.05 13.1% ± 3.8% 4.73 ± 0.08 Leoville CV3r S3 Find NMNH 3535 36.69 3.57 ± 0.06 3.53 ± 0.02 -1.3% ± 1.8% 4.95 ± 0.12 Leoville CV3r S3 Find Average 3.42 3.50 2.09% 4.84 Nova 002 CV3o Find NMNH 6828 12.83 3.06 ± 0.04 3.25 ± 0.03 6.0% ± 1.5% 4.12 ± 0.12 NWA 2140 CV3 Find Monnig M 295.2 15.26 2.85 ± 0.06 3.68 ± 0.05 22.5% ± 2.0% 3.28 ± 0.12 NWA 3118 CV3 W low Find Monnig M1167.1 54.44 2.75 ± 0.05 3.57 ± 0.01 23.1% ± 1.5% 3.29 ± 0.12 Sahara 98044 CV3 S4 W4 Find Monnig M 893.1 25.68 3.37 ± 0.07 3.56 ± 0.03 5.2% ± 2.0% 4.29 ± 0.08 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study.

278

Table 14: Physical Properties of Carbonaceous Chondrite Falls by Petrographic Type.

Petrographic Type No. Meteorites Bulk Density (g cm-3) Grain Density (g cm-3) Porosity Magnetic Susceptibility (log χ) max max max max Average Average Average Average min min min min 1 1 1.57 2.42 34.9% 4.49

3.05 3.37 36.7% 5.11 2 13 2.26 ± 0.09 2.93 ± 0.05 23.1% ± 2.2% 4.03 ± 0.16 1.88 2.74 9.5% 3.30

3.22 3.74 34.2% 4.81 3 10 2.90 ± 0.08 3.63 ± 0.03 21.0% ± 2.7% 4.39 ± 0.12 2.42 3.50 8.3% 3.65

3.23 3.60 20.8% 4.67 4 4 3.04 ± 0.19 3.58 ± 0.02 15.0% ± 5.8% 4.55 ± 0.12 2.85 3.56 9.2% 4.43

279

Table 15: Physical Properties of Carbonaceous Chondrites by Shock.

Shock Stage No. Meteorites Bulk Density (g cm-3) Grain Density (g cm-3) Porosity Magnetic Susceptibility (log χ) max max max max Average Average Average Average min min min min 3.23 3.70 36.7% 4.89 S1 20 2.57 ± 0.10 3.25 ± 0.08 21.6% ± 2.0% 4.17 ± 0.11 1.88 2.78 9.2% 3.28

3.55 3.74 23.3% 5.14 S2 12 3.13 ± 0.07 3.46 ± 0.04 9.4% ± 2.2% 4.81 ± 0.05 2.66 3.31 -1.0% 4.62

3.42 3.66 21.2% 4.84 S3 3 3.16 ± 0.16 3.58 ± 0.08 11.7% ± 9.6% 4.47 ± 0.26 2.22 2.91 15.0% 3.66

3.46 3.56 5.2% 4.86 S4 2 3.42 ± 0.04 3.52 ± 0.04 2.9% ± 2.3% 4.57 ± 0.28 3.37 3.48 0.6% 4.29

280

Table 16: Data for Enstatite Chondrites.

Bulk Grain Magnetic Weathering Meteorite Type Fall Collection Catalog Mass Density Density Porosity Susceptibility Modulus (g) (g cm-3) (g cm-3) (log χ) Abee a EH4 S2-5 Fall NMNH 2096 A 138.65 3.48 ± 0.05 3.63 ± 0.01 4.0% ± 1.5% 5.58 ± 5.58 Abee EH4 S2-5 Fall NMNH 2096 B 30.497 3.46 ± 0.06 3.52 ± 0.02 1.8% ± 1.7% 5.40 ± 5.40 Abee EH4 S2-5 Fall Vatican 1129 66.92 3.60 ± 0.04 3.68 ± 0.04 2.2% ± 1.5% 5.55 ± 5.55 Abee EH4 S2-5 Fall FMNH ME 2474 #6 78.26 3.54 ± 0.04 3.62 ± 0.01 2.2% ± 1.0% 5.37 ± 5.37 Abee EH4 S2-5 Fall Average 3.52 3.62 2.96% 5.47 Adhi Kota EH4 Fall AMNH 3993 56.94 3.58 ± 0.14 3.76 ± 0.04 4.7% ± 3.8% 5.58 ± 5.58 Adhi Kot EH4 Fall NMNH 2156 16.017 3.66 ± 0.09 3.71 ± 0.03 1.5% ± 2.5% 5.41 ± 5.41 Adhi Kot EH4 Fall Average 3.60 3.75 3.98% 5.49 Indarch EH4 S4 Fall AMNH 2237 13.44 3.25 ± 0.13 3.67 ± 0.05 11.7% ± 3.7% 5.30 ± 5.30 Indarch EH4 S4 Fall NMNH 3482 44.481 3.63 ± 0.05 3.62 ± 0.02 -0.3% ± 1.5% 5.53 ± 5.53 Indarch EH4 S4 Fall Monnig M 731.1 51.5 3.59 ± 0.03 3.65 ± 0.02 1.7% ± 1.0% 5.39 ± 5.39 Indarch EH4 S4 Fall LNHM BM1921,23 77.89 3.56 ± 0.03 3.64 ± 0.01 2.2% ± 1.0% 5.37 ± 5.37 Indarch EH4 S4 Fall FMNH ME 1404 #5 23.5 3.73 ± 0.10 3.70 ± 0.02 -1.0% ± 2.9% 5.34 ± 5.34 Indarch EH4 S4 Fall Average 3.58 3.65 1.90% 5.38 Saint-Sauveur EH5 Fall NMNH 7213 31.172 3.63 ± 0.04 3.66 ± 0.03 0.7% ± 1.4% 5.57 ± 5.57 St. Mark's EH5 Fall LNHM BM1916,59 86.46 3.57 ± 0.03 3.74 ± 0.01 4.5% ± 0.9% 5.45 ± 5.45 Bethune EH4/5 Find AMNH 4201 12.23 3.14 ± 0.07 3.33 ± 0.04 5.7% ± 2.5% 4.49 ± 4.49 4.2 Bethune EH4/5 Find LNHM BM1959,847 26.35 2.90 ± 0.05 3.17 ± 0.02 8.5% ± 1.5% 4.49 ± 4.49 4.6 Bethune EH4/5 Find Average 2.97 3.22 7.67% 4.49 Kota-Kota EH3 S4 Find NMNH 7048 16.454 3.53 ± 0.06 3.55 ± 0.03 0.4% ± 2.0% 5.25 ± 5.25 1.0 Sahara 97096 EH3 S2 W1 Find AMNH 4940 44.8 3.24 ± 0.05 3.73 ± 0.04 13.2% ± 1.7% 5.72 ± 5.72 1.0 Sahara 97158 EH3 Find LNHM BM1997,M8 123.83 3.55 ± 0.06 3.66 ± 0.01 2.9% ± 1.7% 5.64 ± 5.64 0.6 Daniel's Kuil EL6 Fall AMNH 4241 A 23.87 3.78 ± 0.04 3.73 ± 0.02 -1.2% ± 1.3% 5.59 ± 5.59 Daniel's Kuil EL6 Fall LNHM BM1985,M143 62.67 3.47 ± 0.03 3.70 ± 0.01 6.2% ± 0.9% 5.43 ± 5.43 Daniel's Kuil EL6 Fall Average 3.55 3.71 4.29% 5.51 Eagle EL6 S2 Fall AMNH 4739 B 14.31 3.50 ± 0.06 3.45 ± 0.03 -1.6% ± 2.1% 5.45 ± 5.45 Hvittisb EL6 S2 Fall AMNH 4067 36.5 3.35 ± 0.06 3.59 ± 0.04 6.6% ± 1.8% 5.66 ± 5.66 Hvittis EL6 S2 Fall NMNH 400 68.182 3.61 ± 0.03 3.58 ± 0.01 -0.9% ± 0.9% 5.44 ± 5.44 Hvittis EL6 S2 Fall Monnig M 753.1a 48.92 3.50 ± 0.06 3.54 ± 0.01 1.1% ± 1.9% 5.30 ± 5.30 Hvittis EL6 S2 Fall Vatican 443 51.17 3.60 ± 0.03 3.58 ± 0.04 -0.8% ± 1.3% 5.53 ± 5.53 Hvittis EL6 S2 Fall Average 3.54 3.57 1.03% 5.48 Jajh deh Kot Lalu EL6 S2 Fall AMNH 3954 13.8 3.15 ± 0.12 3.57 ± 0.04 11.7% ± 3.6% 5.54 ± 5.54 Jajh deh Kot Lalu EL6 S2 Fall NMNH 1260 14.913 3.62 ± 0.05 3.53 ± 0.03 -2.4% ± 1.8% 5.43 ± 5.43 Jajh deh Kot Lalu EL6 S2 Fall LNHM BM1928,479 31.91 3.37 ± 0.04 3.53 ± 0.02 4.6% ± 1.2% 5.32 ± 5.32 Jajh deh Kot Lalu EL6 S2 Fall Average 3.37 3.54 4.69% 5.43 Khairpur EL6 S2 Fall NMNH 1055 24.934 4.11 ± 0.06 4.17 ± 0.03 1.4% ± 1.7% 5.60 ± 5.60 Khairpur EL6 S2 Fall LNHM BM51366 29.62 3.63 ± 0.05 3.80 ± 0.02 4.3% ± 1.4% 5.68 ± 5.68 Khairpur EL6 S2 Fall FMNH ME 1538 #2 24.13 3.70 ± 0.06 3.65 ± 0.02 -1.3% ± 1.9% 5.32 ± 5.32 Khairpur EL6 S2 Fall Average 3.79 3.86 1.70% 5.53 Pillistfer EL6 S2 Fall AMNH 526 42.6 3.54 ± 0.13 3.74 ± 0.04 5.2% ± 3.7% 5.66 ± 5.66 Pillistfer EL6 S2 Fall NMNH 2993 25.571 3.76 ± 0.05 3.66 ± 0.02 -2.5% ± 1.6% 5.55 ± 5.55 Pillistfer EL6 S2 Fall Monnig M 544.1 44.14 3.59 ± 0.06 3.65 ± 0.02 1.5% ± 1.6% 5.46 ± 5.46 Pillistfer EL6 S2 Fall Vatican 743 7.28 n.d. n.d. n.d. 5.43 ± 5.43† Pillistfer EL6 S2 Fall FMNH ME 1646 #1 37.81 3.61 ± 0.06 3.75 ± 0.02 3.7% ± 1.6% 5.67 ± 5.67 Pillistfer EL6 S2 Fall Average 3.61 3.70 2.44% 5.55 Atlantab EL6 Find Monnig M 863.1 27.27 3.44 ± 0.06 3.51 ± 0.02 2.0% ± 1.7% 5.22 ± 5.22 1.2 Atlanta EL6 Find LNHM BM1959,1001 30.77 3.41 ± 0.09 3.63 ± 0.03 6.2% ± 2.6% 5.42 ± 5.42 0.2 Atlanta EL6 Find Average 3.42 3.57 4.23% 5.32 Blithfieldb EL6 S2 Find AMNH 646 111.8 3.45 ± 0.04 3.50 ± 0.04 1.3% ± 1.5% 5.18 ± 5.18 1.4 Blithfield EL6 S2 Find NMNH 534 37.259 4.51 ± 0.09 4.46 ± 0.03 -1.1% ± 2.2% 5.59 ± 5.59 3.9 Blithfield EL6 S2 Find Average 3.67 3.70 0.80% 5.39 Happy Canyona EL6/7 S2 Find NMNH 5820 10.593 3.12 ± 0.06 3.32 ± 0.04 6.1% ± 2.2% 4.53 ± 4.53 4.1 Ilafegh 009 EL7 S4 W0/1 Find AMNH 4757 35.3 3.54 ± 0.06 3.68 ± 0.04 3.8% ± 1.9% 5.69 ± 5.69 0.9 North West Forrest EL6 Find LNHM BM1989,M27 78.47 2.89 ± 0.05 3.28 ± 0.01 12.0% ± 1.5% 4.16 ± 4.16 5.6 NWA 2965 EL6/7 Find Monnig M1326.1 40.98 3.08 ± 0.04 3.33 ± 0.02 7.7% ± 1.2% 4.43 ± 4.43 4.5 NWA 3132 EL3 S2 W4 Find CMS 1506 21.26 2.94 ± 0.05 3.23 ± 0.02 9.0% ± 1.5% 4.74 ± 4.74 3.6 Yilmia EL6 Find IOM C 89.2 72.21 3.14 ± 0.06 3.31 ± 0.01 5.0% ± 1.7% 4.94 ± 4.94 2.7 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. a Impact melt breccia (Rubin et al., 1997). b Breccia.

281

Table 17: Mass-weighted Averages by Meteorite for Enstatite Chondrites.

Meteorite Type Fall Number of Stones Total Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Abee EH4 S2-5 Fall 4 314.33 3.52 3.62 2.96% 5.47 Adhi Kot EH4 Fall 2 72.96 3.60 3.75 3.98% 5.49 Indarch EH4 S4 Fall 5 210.81 3.58 3.65 1.90% 5.38 Saint-Sauveur EH5 Fall 1 31.17 3.63 3.66 0.67% 5.57 St. Mark's EH5 Fall 1 86.46 3.57 3.74 4.49% 5.45 Bethune EH4/5 Find 2 38.58 2.97 3.22 7.67% 4.49 Kota-Kota EH3 S4 Find 1 16.45 3.53 3.55 0.42% 5.25 Sahara 97096 EH3 S2 W1 Find 1 44.80 3.24 3.73 13.18% 5.72 Sahara 97158 EH3 Find 1 123.83 3.55 3.66 2.92% 5.64 Daniel's Kuil EL6 Fall 2 86.54 3.55 3.71 4.29% 5.51 Eagle EL6 S2 Fall 1 14.31 3.50 3.45 -1.60% 5.45 Hvittis EL6 S2 Fall 4 204.77 3.54 3.57 1.03% 5.48 Jajh deh Kot Lalu EL6 S2 Fall 3 60.62 3.37 3.54 4.69% 5.43 Khairpur EL6 S2 Fall 3 78.68 3.79 3.86 1.70% 5.53 Pillistfer EL6 S2 Fall 5 157.40 3.61 3.70 2.44% 5.55 Atlanta EL6 Find 2 58.04 3.42 3.57 4.23% 5.32 Blithfield EL6 S2 Find 2 149.06 3.67 3.70 0.80% 5.39 Happy Canyon EL6/7 S2 Find 1 10.59 3.12 3.32 6.07% 4.53 Ilafegh 009 EL7 S4 W0/1 Find 1 35.30 3.54 3.68 3.77% 5.69 North West Forrest EL6 Find 1 78.47 2.89 3.28 11.99% 4.16 NWA 2965 EL6/7 Find 1 40.98 3.08 3.33 7.70% 4.43 NWA 3132 EL3 S2 W4 Find 1 21.26 2.94 3.23 9.03% 4.74 Yilmia EL6 Find 1 72.21 3.14 3.31 4.97% 4.94

282

Table 18: Data for Lunar Meteorites, Apollo Samples, and SNC.

Magnetic Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Susceptibility (g) (g cm-3) (g cm-3) (log χ) Dhofar 081 Lun Find Vatican 1441 6.96 n.d. 2.64 ± 0.03‡ n.d. n.d. NWA 482 Lun Find IOM A 49.1 9.87 2.82 ± 0.10 2.85 ± 0.03 1.1% ± 3.7% 2.62 ± 0.12 NWA 5000 Lun Find Monnig M1431.1 7.00 2.73 ± 0.07 2.80 ± 0.05 2.7% ± 3.1% 3.31 ± 0.12 NWA 773 Lun S2 W1 Find LNHM BM2001,M23 13.22 2.87 ± 0.06 3.24 ± 0.03 11.4% ± 2.1% 3.36 ± 0.12 Apollo 12051 Lun Field NASA 12051 19 12.19 3.27 ± 0.05 3.32 ± 0.02 1.8% ± 1.7% 2.79 ± 0.10 Apollo 14303 Lun Field NASA 14303 14 22.26 2.52 ± 0.03 3.05 ± 0.01 17.5% ± 1.0% 3.33 ± 0.10 Apollo 14321 Lun Field NASA 14321 220 10.01 2.36 ± 0.04 3.03 ± 0.03 22.1% ± 1.5% 3.14 ± 0.10 Apollo 15418 Lun Field NASA 15418 179 28.68 2.65 ± 0.02 3.12 ± 0.01 14.9% ± 0.7% 2.90 ± 0.10 Apollo 15555 Lun Field NASA 15555 62 32.98 3.11 ± 0.03 3.35 ± 0.01 7.1% ± 0.9% 2.87 ± 0.10 Dar al Gani 476 She Find Vatican 1168 18.1 3.18 ± 0.06‡ n.d. n.d. 2.90 ± 0.08† Los Angeles She Find CMS 1418 24.54 2.83 ± 0.03 3.08 ± 0.02 8.1% ± 1.0% 3.52 ± 0.12 Sayh al Uhaymir 005 She Find Monnig M 944.1 23.94 3.12 ± 0.04 3.35 ± 0.04 6.8% ± 1.6% 2.92 ± 0.12 Sayh al Uhaymir 005 She Find LNHM BM2000,M40 19.04 3.07 ± 0.04 3.27 ± 0.02 6.2% ± 1.4% 2.88 ± 0.12 Sayh al Uhaymir 005 She Find Vatican 1169 19.28 3.13 ± 0.03 3.44 ± 0.03‡ 9.0% ± 1.2% 2.83 ± 0.12 Sayh al Uhaymir 005 She Find Average 3.11 3.35 7.3% 2.87 Shergotty She Fall LNHM BM41021 57.34 2.83 ± 0.02 3.27 ± 0.01 13.5% ± 0.8% 2.89 ± 0.09 Zagami She Fall IOM A 23.6a 9.45 3.13 ± 0.12 3.25 ± 0.05 3.5% ± 4.0% 2.79 ± 0.12 Zagami She Fall CMS 1306.2 23.78 2.92 ± 0.04 3.30 ± 0.03 11.4% ± 1.3% 2.83 ± 0.12 Zagami She Fall LNHM BM1966,54 32.32 2.91 ± 0.04 3.29 ± 0.01 11.4% ± 1.2% 2.79 ± 0.09 Zagami She Fall Average 2.95 3.28 10.3% 2.80 Governador Valadares Nak Find LNHM BM1999,M64 21.38 3.11 ± 0.04 3.43 ± 0.02 9.4% ± 1.4% 3.27 ± 0.12 Governador Valadares Nak Find FMNH ME 3216 #1 9.84 3.07 ± 0.05 3.51 ± 0.05 12.3% ± 2.0% 3.31 ± 0.12 Governador Valadares Nak Find Average 3.10 3.45 10.3% 3.29 Lafayette (stone) Nak Find CMS 167ax 16.01 3.21 ± 0.04 3.41 ± 0.03 5.8% ± 1.5% 3.49 ± 0.12 Lafayette (stone) Nak Find LNHM BM1959,755 26.65 3.16 ± 0.04 3.45 ± 0.03 8.3% ± 1.3% 3.50 ± 0.12 Lafayette (stone) Nak Find Average 3.18 3.43 7.4% 3.50 Nakhla Nak Fall Vatican 675 153.7 3.19 ± 0.03 3.29 ± 0.06‡ 3.0% ± 2.1% 3.17 ± 0.09 NWA 998 Nak Find Monnig M1063.1 17.66 2.88 ± 0.05 3.47 ± 0.03 16.9% ± 1.5% 3.68 ± 0.09 Chassigny Cha Fall Vatican 208 15.66 3.48 ± 0.08‡ 3.73 ± 0.04‡ 6.8% ± 2.3% 2.98 ± 0.08† The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study.

283

Table 19: Data for HEDs.

Bulk Grain Magnetic Meteorite Type Fall Collection Catalog Mass Density Density Porosity Susceptibility (g) (g cm-3) (g cm-3) (log χ) Dar al Gani 779 How S2 W1 Find LNHM BM2001,M3 21.38 2.77 ± 0.05 3.36 ± 0.02 17.6% ± 1.5% 3.15 ± 0.09

Dhofar 018 How Find Vatican 1444 2.84 2.61 ± 0.07‡ 2.96 ± 0.04‡ 12.0% ± 2.7% n.d.

Dhofar 485 How S4 W2 Find Monnig M1447.1 39.56 2.95 ± 0.06 3.21 ± 0.02 8.2% ± 1.9% 3.09 ± 0.12

Frankfort (stone) How Fall AMNH 351 10.51 2.90 ± 0.08 3.32 ± 0.02 12.7% ± 2.5% 3.36 ± 0.09 Frankfort (stone) How Fall Vatican 346 9.21 n.d. n.d. n.d. 3.49 ± 0.08† Frankfort (stone) How Fall Average 2 stones 2.90 3.32 12.7% 3.43

Hughes 004 How Find AMNH 4845 14.62 30 ± 0.04 3.33 ± 0.03 9.7% ± 1.3% 3.09 ± 0.12

Jodzie How Fall FMNH ME 1371 #1 16.53 2.77 ± 0.03 3.24 ± 0.03 14.5% ± 1.3% 3.53 ± 0.10

Kapoeta How Fall AMNH 3924 A 18.92 2.94 ± 0.04 3.26 ± 0.02 9.9% ± 1.3% 3.67 ± 0.10 Kapoeta How Fall CMS 827 54.31 2.87 ± 0.03 3.30 ± 0.01 12.9% ± 1.0% 3.50 ± 0.09 Kapoeta How Fall FMNH ME 3129 #3 41.00 2.87 ± 0.03 3.35 ± 0.02 14.5% ± 1.0% 3.95 ± 0.10 Kapoeta How Fall Average 3 stones 2.88 3.31 13.0% 3.71

Le Teilleul How Fall NMNH 7209 21.996 2.77 ± 0.04 3.28 ± 0.02 15.5% ± 1.2% 3.46 ± 0.09

Luotolax How Fall AMNH 4932 14.41 2.77 ± 0.07 3.22 ± 0.03 13.9% ± 2.2% 3.21 ± 0.10

Muckera 001 How Find NMNH 6301 18.974 3.15 ± 0.05 3.21 ± 0.02 2.0% ± 1.7% 3.60 ± 0.10

Muckera 002 How Find CMS 1394.2 36.33 3.02 ± 0.04 3.17 ± 0.01 4.6% ± 1.2% 3.63 ± 0.10

NWA 1942 How Find Monnig M1113.1 36.61 2.96 ± 0.03 3.30 ± 0.02 10.3% ± 1.1% 3.07 ± 0.09

NWA 2696 How Find Monnig M1270.1 69.91 2.82 ± 0.02 3.20 ± 0.01 11.9% ± 0.8% 3.06 ± 0.09

Pavlovka How Fall LNHM BM55255 66.8 2.77 ± 0.02 3.33 ± 0.01 16.7% ± 0.8% 3.29 ± 0.10 Pavlovka How Fall Vatican 1470 8.14 n.d. 3.29 ± 0.03‡ n.d. 3.51 ± 0.08† Pavlovka How Fall FMNH ME 1377 #1 41.31 2.84 ± 0.03 3.32 ± 0.01 14.6% ± 1.1% 3.55 ± 0.09 Pavlovka How Fall FMNH ME 1378 #1 93.36 2.79 ± 0.05 3.30 ± 0.01 15.3% ± 1.4% 3.29 ± 0.09 Pavlovka How Fall Average 4 stones 2.80 3.31 15.6% 3.41

Washougal How Fall LNHM BM1959,754 16.91 2.62 ± 0.04 3.26 ± 0.03 19.6% ± 1.4% 3.19 ± 0.10

Yurtuk How Fall NMNH 1422 29.669 2.83 ± 0.02 3.25 ± 0.01 13.0% ± 0.8% 2.90 ± 0.09 Yurtuk How Fall LNHM BM1956,321 21.57 2.84 ± 0.04 3.30 ± 0.02 14.0% ± 1.3% 2.81 ± 0.09 Yurtuk How Fall Average 2 stones 2.83 3.27 13.4% 2.85

Agoult Euc Find Monnig M1060.1 51.65 3.06 ± 0.03 3.19 ± 0.01 4.1% ± 1.0% 2.61 ± 0.12

Béréba Euc mmict Fall NMNH 6703 15.35 2.98 ± 0.05 3.17 ± 0.03 6.2% ± 1.8% 3.29 ± 0.10

Bialystok Euc pmict Fall NMNH 332 11.389 2.77 ± 0.04 3.20 ± 0.04 13.5% ± 1.6% 2.73 ± 0.09

Binda Euc Find AMNH 4008 7.95 2.94 ± 0.07 3.34 ± 0.06 11.8% ± 2.6% 3.03 ± 0.12 Binda Euc Find Monnig M 68.1 35.66 2.97 ± 0.03 3.32 ± 0.01 10.6% ± 0.9% 3.18 ± 0.10 Binda Euc Find Average 2 stones 2.96 3.32 10.8% 3.11

Bouvante Euc mmict Find AMNH 4550 29.95 2.80 ± 0.01 3.23 ± 0.01 13.3% ± 0.6% 3.11 ± 0.10 Bouvante Euc mmict Find CMS 1290.2 68.69 2.73 ± 0.05 3.21 ± 0.01 14.8% ± 1.4% 3.43 ± 0.09 Bouvante Euc mmict Find Average 2 stones 2.75 3.21 14.4% 3.27

Cachari Euc mmict Find NMNH 6708 24.249 2.92 ± 0.03 3.19 ± 0.02 8.5% ± 1.1% 2.57 ± 0.10 Cachari Euc mmict Find IOM A 14.1 19.01 2.78 ± 0.04 3.21 ± 0.02 13.6% ± 1.3% 2.80 ± 0.09 Cachari Euc mmict Find Average 2 stones 2.85 3.20 10.8% 2.69

Caldera Euc Find NMNH 6394 28.915 n.d. 3.14 ± 0.01 n.d. 2.64 ± 0.09

Camel Donga Euc mmict Find AMNH 4597 51.55 2.83 ± 0.01 3.20 ± 0.01 11.4% ± 0.4% 4.17 ± 0.10 Camel Donga Euc mmict Find NMNH 6430 54.253 2.82 ± 0.02 3.16 ± 0.01 10.8% ± 0.8% 4.28 ± 0.09 Camel Donga Euc mmict Find FMNH ME 3091 #1 44.65 2.83 ± 0.03 3.20 ± 0.01 11.4% ± 0.9% 4.44 ± 0.10 Camel Donga Euc mmict Find Average 3 stones 2.83 3.18 11.2% 4.30

Haraiya Euc mmict Fall AMNH 4062 49.5 2.63 ± 0.05 3.25 ± 0.03 19.0% ± 1.8% 2.86 ± 0.09

Ibitira Euc mmict Fall NMNH 6860 13.333 3.12 ± 0.07 3.18 ± 0.03 2.0% ± 2.4% 2.78 ± 0.12 Ibitira Euc mmict Fall IOM A 7.2 15.47 2.88 ± 0.06 3.17 ± 0.03 9.0% ± 2.1% 2.83 ± 0.09

284

Bulk Grain Magnetic Meteorite Type Fall Collection Catalog Mass Density Density Porosity Susceptibility (g) (g cm-3) (g cm-3) (log χ) Ibitira Euc mmict Fall Average 2 stones 2.98 3.17 5.9% 2.81

Jonzac Euc mmict Fall NMNH 7216 39.568 2.74 ± 0.03 3.20 ± 0.02 14.2% ± 0.9% 2.62 ± 0.09 Jonzac Euc mmict Fall Vatican 466 31.86 2.74 ± 0.02 3.26 ± 0.03 15.8% ± 1.0% 2.70 ± 0.09 Jonzac Euc mmict Fall Average 2 stones 2.74 3.22 14.9% 2.66

Juvinas Euc mmict Fall AMNH 466 18.14 2.82 ± 0.03 3.21 ± 0.02 12.2% ± 1.1% 3.08 ± 0.09 Juvinas Euc mmict Fall NMNH 1051 19.089 2.86 ± 0.04 3.18 ± 0.02 9.9% ± 1.3% 2.88 ± 0.09 Juvinas Euc mmict Fall Monnig M 202.2 14.39 2.86 ± 0.04 3.21 ± 0.03 11.1% ± 1.4% 2.71 ± 0.09 Juvinas Euc mmict Fall Monnig M 200.1 75.35 2.96 ± 0.02 3.20 ± 0.01 7.5% ± 0.8% 3.34 ± 0.09 Juvinas Euc mmict Fall Vatican 470 70.94 3.07 ± 0.01 3.22 ± 0.03 4.8% ± 1.0% 2.77 ± 0.09 Juvinas Euc mmict Fall Vatican 469 212.66 3.07 ± 0.03 2.99 ± 0.03‡ -2.5% ± 1.4% 3.05 ± 0.09 Juvinas Euc mmict Fall Vatican 473 16.49 3.00 ± 0.07 n.d. n.d. n.d. Juvinas Euc mmict Fall Vatican 471 31.36 2.73 ± 0.02 n.d. n.d. 2.58 ± 0.09 Juvinas Euc mmict Fall FMNH ME 2649 #5 46.16 2.83 ± 0.03 3.22 ± 0.01 12.0% ± 1.0% 2.96 ± 0.09 Juvinas Euc mmict Fall Average 9 stones 2.98 3.10 4.0% 2.92

Kirbyville Euc mmict Fall NMNH 6924 22.691 2.93 ± 0.03 3.19 ± 0.02 8.0% ± 1.1% 2.66 ± 0.12

Lakangaon Euc mmict Fall LNHM BM1915,142 81.91 2.70 ± 0.04 3.24 ± 0.01 16.7% ± 1.4% 2.88 ± 0.09

Macibini Euc pmict Fall AMNH 4560 14.3 2.82 ± 0.06 3.16 ± 0.03 10.7% ± 1.9% 3.20 ± 0.10 Macibini Euc pmict Fall NMNH 1344 14.184 2.85 ± 0.06 3.13 ± 0.03 9.1% ± 2.2% 3.26 ± 0.09 Macibini Euc pmict Fall CMS 1578 29.44 2.73 ± 0.03 3.19 ± 0.02 14.5% ± 1.0% 3.41 ± 0.10 Macibini Euc pmict Fall Average 3 stones 2.78 3.17 12.3% 3.29

Millbillillie Euc mmict Fall AMNH 4698 64.57 2.74 ± 0.02 n.d. n.d. 2.66 ± 0.09 Millbillillie Euc mmict Fall AMNH 4699 142.99 2.84 ± 0.04 3.20 ± 0.00 11.3% ± 1.4% 2.59 ± 0.10 Millbillillie Euc mmict Fall NMNH 6452 46.306 2.88 ± 0.03 3.18 ± 0.01 9.5% ± 1.1% 2.69 ± 0.10 Millbillillie Euc mmict Fall IOM A 3.3 51.85 2.77 ± 0.03 3.20 ± 0.01 13.3% ± 1.0% 2.76 ± 0.10 Millbillillie Euc mmict Fall Average 4 stones 2.81 3.19 12.0% 2.67

Moama Euc cum Find LNHM BM1973,M11 1 66.94 2.78 ± 0.03 3.21 ± 0.01 13.5% ± 0.8% 2.59 ± 0.09 Moama Euc cum Find LNHM BM1973,M11 2 25.15 2.68 ± 0.04 3.03 ± 0.01 11.3% ± 1.3% 3.17 ± 0.09 Moama Euc cum Find Average 2 stones 2.75 3.16 12.9% 2.88

Moore County Euc cum Fall AMNH 4471 A 23.01 2.92 ± 0.03 3.10 ± 0.02 5.7% ± 1.1% 2.87 ± 0.09 Moore County Euc cum Fall AMNH 4471 B 13.91 3.03 ± 0.04 3.08 ± 0.03 1.5% ± 1.7% 3.06 ± 0.09 Moore County Euc cum Fall AMNH 4471 C 10.04 2.98 ± 0.05 3.07 ± 0.04 2.9% ± 2.0% 2.99 ± 0.09 Moore County Euc cum Fall Average 3 stones 2.97 3.09 3.9% 2.97

Nuevo Laredo Euc mmict Find NMNH 1783 84.329 2.76 ± 0.05 3.21 ± 0.01 14.1% ± 1.5% 2.83 ± 0.09

NWA 1109 Euc pmict Find Monnig M1064.1 42.35 2.76 ± 0.03 3.25 ± 0.01 15.1% ± 1.1% 3.22 ± 0.09

NWA 2690 Euc W2 Find Monnig M 375.2 36.88 2.71 ± 0.06 3.23 ± 0.02 16.0% ± 1.8% 3.30 ± 0.09

Padvarninkai Euc mmict Fall CMS 812.2 17.42 2.96 ± 0.06 3.15 ± 0.02 6.1% ± 2.1% 2.83 ± 0.09 Padvarninkai Euc mmict Fall LNHM BM1931,108 47.93 2.83 ± 0.03 3.16 ± 0.01 10.6% ± 1.0% 3.50 ± 0.09 Padvarninkai Euc mmict Fall Average 2 stones 2.86 3.16 9.4% 3.17

Palo Blanco Creek Euc mmict Find IOM A 6.2 46.63 2.95 ± 0.06 3.14 ± 0.01 6.1% ± 1.9% 3.01 ± 0.12 Palo Blanco Creek Euc mmict Find IOM A 6.5b 14.34 2.89 ± 0.04 3.14 ± 0.03 8.0% ± 1.6% 2.58 ± 0.12 Palo Blanco Creek Euc mmict Find Average 2 stones 2.93 3.14 6.6% 2.79

Pasamonte Euc pmict Fall NMNH 6318 63.288 2.66 ± 0.02 3.19 ± 0.01 16.6% ± 0.7% 3.40 ± 0.09 Pasamonte Euc pmict Fall Monnig M 10.1 52.7 2.61 ± 0.02 3.22 ± 0.01 18.9% ± 0.8% 3.08 ± 0.09 Pasamonte Euc pmict Fall FMNH ME 2623 #1 88.9 2.61 ± 0.04 3.21 ± 0.01 18.8% ± 1.3% 3.37 ± 0.10 Pasamonte Euc pmict Fall Average 3 stones 2.63 3.21 18.2% 3.28

Petersburg Euc pmict Fall AMNH 388 11.64 3.00 ± 0.06 3.19 ± 0.04 6.1% ± 2.0% 4.43 ± 0.08 Petersburg Euc pmict Fall NMNH 438 10.483 2.88 ± 0.06 3.13 ± 0.04 8.0% ± 2.2% 3.84 ± 0.09 Petersburg Euc pmict Fall Monnig M 498.1 24.58 2.97 ± 0.07 3.26 ± 0.02 9.0% ± 2.3% 4.19 ± 0.08 Petersburg Euc pmict Fall Average 3 stones 2.96 3.21 8.1% 4.16

Rancho Blanco Euc mmict Find CMS 1316 20.68 2.92 ± 0.07 3.17 ± 0.03 7.8% ± 2.4% 3.24 ± 0.12

Serra de Magé Euc Fall NMNH 839 21.938 3.08 ± 0.04 3.11 ± 0.02 0.9% ± 1.5% 2.56 ± 0.09

Sioux County Euc mmict Fall NMNH 836 10.331 2.73 ± 0.05 3.19 ± 0.04 14.4% ± 1.9% 3.69 ± 0.10 Sioux County Euc mmict Fall Monnig M 210.2 32.18 2.67 ± 0.02 3.20 ± 0.01 16.7% ± 0.8% 2.90 ± 0.09 Sioux County Euc mmict Fall DuPont JMD 658 20.96 2.62 ± 0.04 3.23 ± 0.02 18.8% ± 1.2% 3.31 ± 0.10 Sioux County Euc mmict Fall Average 3 stones 2.66 3.21 17.0% 3.30

285

Bulk Grain Magnetic Meteorite Type Fall Collection Catalog Mass Density Density Porosity Susceptibility (g) (g cm-3) (g cm-3) (log χ) Smara Euc pmict Find LNHM BM2001,M4 19.72 2.75 ± 0.15 3.19 ± 0.03 13.8% ± 4.9% 3.64 ± 0.12

Stannern Euc mmict Fall AMNH 4362 13.94 2.79 ± 0.06 3.21 ± 0.03 13.2% ± 2.1% 2.65 ± 0.09 Euc mmict Fall NMNH 141 30.47 2.85 ± 0.03 3.18 ± 0.01 10.5% ± 0.9% 2.57 ± 0.09 Stannern Euc mmict Fall Vatican 907 18.54 2.80 ± 0.02 3.33 ± 0.03 16.0% ± 1.1% 2.62 ± 0.10 Stannern Euc mmict Fall Vatican 908 14.98 n.d. n.d. n.d. 2.66 ± 0.08† Stannern Euc mmict Fall Average 4 stones 2.82 3.23 12.8% 2.63

Dar al Gani 411 Euc pmict S2 W0 Find Vatican 1167 19.45 2.89 ± 0.03 3.32 ± 0.03 13.0% ± 1.3% 3.04 ± 0.12

Aïoun el Atrouss Dio pmict Fall AMNH 4466 76.68 2.93 ± 0.06 3.45 ± 0.03 15.2% ± 1.9% 2.68 ± 0.12 Aïoun el Atrouss Dio pmict Fall IOM A 12.1 30.47 3.04 ± 0.04 3.41 ± 0.02 10.8% ± 1.1% 2.66 ± 0.12 Aïoun el Atrouss Dio pmict Fall LNHM BM1977,M9 34.39 3.04 ± 0.04 3.41 ± 0.01 10.9% ± 1.1% 2.63 ± 0.12 Aïoun el Atrouss Dio pmict Fall Average 3 stones 2.98 3.43 13.2% 2.66

Bilanga Dio Fall Monnig M 915.2 110.1 3.10 ± 0.05 3.39 ± 0.01 8.5% ± 1.5% 2.79 ± 0.09 Bilanga Dio Fall Vatican 1398 14.58 2.95 ± 0.06 n.d. n.d. 2.87 ± 0.10 Bilanga Dio Fall Average 2 stones 3.09 3.39 9.1% 2.83

Dhofar 700 Dio Find Monnig M1169.1 94.54 3.37 ± 0.03 3.46 ± 0.01 2.5% ± 0.9% 2.83 ± 0.09

Garland Dio pmict Fall NMNH 2140 66.705 2.92 ± 0.02 3.36 ± 0.01 13.0% ± 0.8% 3.67 ± 0.10

Johnstown Dio Fall AMNH 2497 A 9.75 3.21 ± 0.10 3.41 ± 0.05 6.0% ± 3.2% 3.15 ± 0.12 Johnstown Dio Fall AMNH 2497 B 7.53 2.99 ± 0.05 3.43 ± 0.06 12.8% ± 2.1% 3.20 ± 0.12 Johnstown Dio Fall NMNH 6633 12.832 3.16 ± 0.08 3.38 ± 0.04 6.6% ± 2.6% 3.30 ± 0.09 Johnstown Dio Fall Monnig M 460.2 18.63 3.03 ± 0.03 3.42 ± 0.04 11.3% ± 1.4% 3.36 ± 0.12 Johnstown Dio Fall FMNH ME 2059 #1 64.24 3.10 ± 0.03 3.40 ± 0.01 8.9% ± 0.9% 3.22 ± 0.09 Johnstown Dio Fall Average 5 stones 3.10 3.41 9.1% 3.25

NWA 1877 Dio Find Monnig M1269.1 60.56 3.05 ± 0.03 3.48 ± 0.01 12.5% ± 0.9% 3.57 ± 0.09

NWA 2038 Dio Find Monnig M1164.1 32.64 3.08 ± 0.04 3.45 ± 0.01 10.9% ± 1.2% 3.05 ± 0.09

NWA 4664 Dio pmict Find Monnig M1363.2 37.11 3.32 ± 0.08 3.37 ± 0.02 1.4% ± 2.4% 3.35 ± 0.12

Roda Dio Fall Vatican 797 7.42 n.d. 3.50 ± 0.03‡ n.d. 3.00 ± 0.08†

Shalka Dio Fall NMNH 244 46.875 3.11 ± 0.03 3.42 ± 0.01 9.1% ± 0.9% 3.06 ± 0.09 Shalka Dio Fall LNHM BM33761 54.9 3.03 ± 0.03 3.45 ± 0.01 12.2% ± 0.9% 2.75 ± 0.09 Shalka Dio Fall Vatican 868 18.42 3.11 ± 0.06 3.51 ± 0.04 11.4% ± 1.9% 3.12 ± 0.09 Shalka Dio Fall Average 3 stones 3.07 3.45 10.9% 2.98

Tatahouine Dio Fall NMNH 6338 20.966 3.31 ± 0.06 3.36 ± 0.02 1.7% ± 1.8% 2.62 ± 0.09 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study. cum: cumulate mmict: monomict breccia pmict: polymict breccia

286

Table 20: Data for Aubrites, Angrites and Ureilites.

Bulk Grain Magnetic Meteorite Type Fall Collection Catalog Mass Density Density Porosity Susceptibility (g) (g cm-3) (g cm-3) (log χ) Aubres Aub Fall NMNH 7047 24.386 3.09 ± 0.03 3.15 ± 0.02 1.7% ± 1.3% 3.66 ± 0.10 Bishopville Aub Fall NMNH 971 21.445 2.69 ± 0.03 3.18 ± 0.02 15.5% ± 1.0% 3.35 ± 0.09 Bishopville Aub Fall Vatican 115 4.27 n.d. 3.44 ± 0.03‡ n.d. 3.04 ± 0.08† Bishopville Aub Fall Average 2 stones 2.69 3.22 16.6% 3.20 Cumberland Falls Aub Fall AMNH 2222 62.86 3.06 ± 0.01 3.22 ± 0.01 4.9% ± 0.4% 4.12 ± 0.10 Cumberland Falls Aub Fall AMNH 664 A 45.13 3.15 ± 0.05 3.23 ± 0.01 2.7% ± 1.7% 3.24 ± 0.12 Cumberland Falls Aub Fall AMNH 664 B 27.65 3.10 ± 0.02 3.25 ± 0.02 4.6% ± 0.9% 3.37 ± 0.12 Cumberland Falls Aub Fall NMNH 2266 37.806 3.01 ± 0.06 3.22 ± 0.02 6.5% ± 2.0% 4.26 ± 0.10 Cumberland Falls Aub Fall Vatican 255 23.34 3.03 ± 0.03 3.29 ± 0.03 7.9% ± 1.3% 3.89 ± 0.09 Cumberland Falls Aub Fall Average 5 stones 3.07 3.24 5.1% 3.78 Khor Temiki Aub Fall AMNH 3973 69.66 n.d. 3.18 ± 0.01 n.d. 3.19 ± 0.09 Khor Temiki Aub Fall NMNH 1551 80.128 n.d. 3.14 ± 0.01 n.d. 3.30 ± 0.09 Khor Temiki Aub Fall Monnig M 762.2 20.08 2.53 ± 0.03 3.22 ± 0.03 21.4% ± 1.3% 3.29 ± 0.12 Khor Temiki Aub Fall Average 3 stones 2.53 3.16 20.0% 3.26 Mayo Belwa Aub Fall AMNH 4465 26.1 2.93 ± 0.02 3.17 ± 0.02 7.6% ± 0.8% 3.54 ± 0.09 Mayo Belwa Aub Fall FMNH ME 2807 #1 15.32 2.88 ± 0.07 3.11 ± 0.03 7.4% ± 2.5% 3.93 ± 0.09 Mayo Belwa Aub Fall Average 2 stones 2.91 3.15 7.5% 3.74 Norton County Aub Fall AMNH 4702 51.14 2.53 ± 0.02 n.d. n.d. 2.94 ± 0.09 Norton County Aub Fall NMNH 2110 16.242 2.80 ± 0.06 3.18 ± 0.03 12.2% ± 1.9% 3.41 ± 0.09 Norton County Aub Fall Average 2 stones 2.59 3.18 18.7% 3.17 Peña Blanca Spring Aub Fall AMNH 4852 B 22.21 2.86 ± 0.04 3.18 ± 0.02 10.0% ± 1.4% 3.39 ± 0.12 Peña Blanca Spring Aub Fall Vatican 1128 30.84 2.83 ± 0.02 3.17 ± 0.03 10.6% ± 1.0% 3.61 ± 0.34 Peña Blanca Spring Aub Fall Average 2 stones 2.84 3.17 10.4% 3.50 Pesyanoe Aub Fall NMNH 1425 58.379 2.74 ± 0.03 3.14 ± 0.01 12.8% ± 1.1% 3.75 ± 0.09 Shallowater Aub Find NMNH 1206 12.5 3.15 ± 0.06 3.27 ± 0.03 3.8% ± 2.0% 4.72 ± 0.10 Zaklodzie Enst-Achond Find Monnig M 932.2 24.15 3.38 ± 0.05 3.51 ± 0.02 3.8% ± 1.5% 5.37 ± 0.12 NWA 4590 Ang Find Monnig M1423.1 23.33 3.24 ± 0.03 3.48 ± 0.02 6.8% ± 1.1% 3.15 ± 0.09 NWA 4801 Ang Find Monnig M1395.1 20.43 3.18 ± 0.04 3.37 ± 0.02 5.7% ± 1.4% 2.77 ± 0.09 Dar al Gani 1010 Ure W2 Find PSF PSF03-023 23.47 3.23 ± 0.05 3.39 ± 0.02 4.6% ± 1.5% 4.17 ± 0.08 Goalpara Ure Find NMNH 1545 78.63 n.d. 3.43 ± 0.01 n.d. 5.20 ± 0.10 Goalpara Ure Find Monnig M 850.1 28.86 3.09 ± 0.06 3.53 ± 0.02 12.3% ± 1.7% 5.28 ± 0.12 Goalpara Ure Find Average 2 stones 3.09 3.46 10.5% 5.24 Hammadah al Hamra 126 Ure Find AMNH 4889 19.79 3.34 ± 0.03 3.32 ± 0.02 -0.6% ± 1.2% 4.55 ± 0.12 Kenna Ure Find NMNH 5825 49.12 3.31 ± 0.03 3.36 ± 0.01 1.3% ± 0.9% 4.61 ± 0.09 Kenna Ure Find CMS 1049 26.5 3.31 ± 0.03 3.34 ± 0.03 0.8% ± 1.2% 4.68 ± 0.12 Kenna Ure Find FMNH ME 2841 #1 33.09 3.30 ± 0.10 3.43 ± 0.02 3.8% ± 2.8% 4.25 ± 0.12 Kenna Ure Find Average 3 stones 3.31 3.38 1.9% 4.51 Nilpena Ure pmict Find LNHM BM1982,M14 13.14 3.15 ± 0.09 3.29 ± 0.03 4.3% ± 2.9% 3.93 ± 0.10 Nova 001 Ure Find AMNH 4823 20.5 3.31 ± 0.28 3.25 ± 0.04 -1.7% ± 8.8% 4.48 ± 0.08 Nova 001 Ure Find Monnig M 917.1 27.63 3.09 ± 0.03 3.25 ± 0.02 4.8% ± 1.2% 4.25 ± 0.12 Nova 001 Ure Find LNHM BM1993,M23 24.82 3.27 ± 0.10 3.26 ± 0.03 -0.1% ± 3.2% 4.26 ± 0.12 Nova 001 Ure Find Average 3 stones 3.21 3.25 1.4% 4.33 Novo-Urei Ure Fall AMNH 514 10.12 3.36 ± 0.12 3.40 ± 0.05 1.4% ± 3.9% 5.01 ± 0.08 Novo-Urei Ure Fall NMNH 307 73.27 3.17 ± 0.03 3.43 ± 0.01 7.6% ± 0.9% 4.97 ± 0.09 Novo-Urei Ure Fall LNHM BM63625 21.47 3.14 ± 0.06 3.47 ± 0.02 9.4% ± 1.9% 5.00 ± 0.08 Novo-Urei Ure Fall FMNH ME 1678 #1 35.97 3.16 ± 0.03 3.41 ± 0.01 7.3% ± 1.0% 4.87 ± 0.10 Novo-Urei Ure Fall Average 4 stones 3.17 3.43 7.4% 4.96 NWA 2634 Ure Find Monnig M1409.1 21.63 3.04 ± 0.04 3.36 ± 0.02 9.5% ± 1.2% 4.47 ± 0.10 NWA 3140 Ure Find Monnig M1255.1 42.09 3.18 ± 0.03 3.33 ± 0.01 4.5% ± 0.9% 4.61 ± 0.09 NWA 3140 Ure Find PSF PSF04-019 27.84 3.11 ± 0.03 3.27 ± 0.02 4.8% ± 1.0% 4.45 ± 0.08 NWA 3140 Ure Find Average 2 stones 3.15 3.31 4.6% 4.53 NWA 766 Ure S2 W2 Find LNHM BM2002,M26 23.69 3.28 ± 0.07 3.34 ± 0.02 1.9% ± 2.1% 4.35 ± 0.08 Reid 016 Ure pmict Find AMNH 4882 18.43 3.32 ± 0.04 3.31 ± 0.02 -0.3% ± 1.5% 4.66 ± 0.12 The symbol ―n.d.‖ indicates no data available for that particular stone. †Magnetic susceptibilities for some Vatican stones are those recorded by P. Rochette and J. Gattacceca. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study. pmict: polymict breccia

287

Table 21: Data for Primitive Achondrites.

Bulk Grain Magnetic Meteorite Type Fall Collection Catalog Mass Density Density Porosity Susceptibility (g) (g cm-3) (g cm-3) (log χ) Acapulco Aca Fall AMNH 4492 10.2 3.59 ± 0.10 3.67 ± 0.05 2.0% ± 3.0% 5.18 ± 0.08 Acapulco Aca Fall LNHM BM1978,M25 37.94 3.33 ± 0.05 3.63 ± 0.02 8.1% ± 1.5% 5.04 ± 0.12 Acapulco Aca Fall FMNH ME 2790 #1 17.96 3.53 ± 0.10 3.69 ± 0.03 4.3% ± 2.8% 5.06 ± 0.12 Acapulco Aca Fall Average 3 stones 3.42 3.65 6.2% 5.09 NWA 2871 Aca W3 Find Monnig M1273.1 23.01 3.36 ± 0.04 3.55 ± 0.02 5.3% ± 1.4% 4.99 ± 0.10 NWA 725 Aca Find Monnig M1061.5 19.9 3.40 ± 0.08 3.88 ± 0.03 12.4% ± 2.2% 5.51 ± 0.12 TIL 99002 Aca WA Find PSF PSF00-027 42.43 3.55 ± 0.04 3.73 ± 0.01 4.8% ± 1.1% 5.40 ± 0.09 Gibson Lod Find CMS 1569 16.87 3.24 ± 0.05 3.38 ± 0.03 4.2% ± 1.6% 4.90 ± 0.12 Lodran Lod Fall LNHM BM1985,M184 20.88 3.68 ± 0.08 4.05 ± 0.03 9.1% ± 2.1% 5.63 ± 0.08 Lodran Lod Fall LNHM BM44003 53.7 3.82 ± 0.04 4.16 ± 0.01 8.1% ± 1.0% 5.68 ± 0.09 Lodran Lod Fall Average 2 stones 3.78 4.12 8.4% 5.66 NWA 4478 Lod Find Monnig M1424.1 34.59 3.38 ± 0.04 3.39 ± 0.01 0.2% ± 1.3% 4.74 ± 0.10 Eagle's Nest Bra Find AMNH 4850 15.26 3.10 ± 0.12 3.65 ± 0.04 15.1% ± 3.5% 3.79 ± 0.12 Hughes 026 Bra Find AMNH 4883 A 12.61 3.78 ± 0.08 3.54 ± 0.04 -6.9% ± 2.7% 3.99 ± 0.12 Hughes 026 Bra Find AMNH 4883 B 8.1 3.90 ± 0.17 3.52 ± 0.06 -10.7% ± 5.1% 3.97 ± 0.12 Hughes 026 Bra Find AMNH 4909 17.16 3.37 ± 0.13 n.d. n.d. 3.95 ± 0.12 Hughes 026 Bra Find Average 3 stones 3.60 3.53 -2.1% 3.97 Nova 003 Bra Find AMNH 4815 13.82 3.33 ± 0.13 3.54 ± 0.04 5.9% ± 3.9% 4.26 ± 0.12 Nova 003 Bra Find NMNH 6828 35.12 3.50 ± 0.08 3.47 ± 0.02 -0.9% ± 2.4% 4.29 ± 0.12 Nova 003 Bra Find LNHM BM1993,M11 92.17 3.50 ± 0.07 3.54 ± 0.01 1.3% ± 2.1% 4.29 ± 0.10 Nova 003 Bra Find Average 3 stones 3.48 3.52 1.2% 4.28 Reid 013 Bra Find Monnig M1305.1 24.22 3.35 ± 0.06 3.59 ± 0.02 6.7% ± 1.9% 3.92 ± 0.10 Pontlyfni Win Fall LNHM BM1975,M6 18.05 3.29 ± 0.08 3.79 ± 0.03 13.0% ± 2.1% 5.55 ± 0.08 Tierra Blanca Win Find AMNH 4561 A 32.9 2.96 ± 0.06 n.d. n.d. 4.44 ± 0.10 Tierra Blanca Win Find LNHM BM1999,M37 22.4 3.10 ± 0.06 3.24 ± 0.02 4.3% ± 1.8% 4.56 ± 0.10 Tierra Blanca Win Find Average 2 stones 3.02 3.24 6.9% 4.50 Winona Win Find AMNH 4158 A 96.2 3.22 ± 0.06 n.d. n.d. 4.79 ± 0.09 Winona Win Find AMNH 4158 C 16.8 3.39 ± 0.13 3.76 ± 0.04 10.0% ± 3.6% 5.05 ± 0.08 Winona Win Find NMNH 854 13.38 3.47 ± 0.08 3.62 ± 0.04 4.2% ± 2.4% 5.00 ± 0.08 Winona Win Find Average 3 stones 3.26 3.70 11.7% 4.95 NWA 3133 Primitive Achondrite Find Monnig M1309.1 55.62 3.34 ± 0.08 3.45 ± 0.01 3.2% ± 2.3% 4.50 ± 0.10 Divnoe Achondrite - Ungrouped Find LNHM BM1999,M65 44.07 3.42 ± 0.04 3.67 ± 0.02 6.7% ± 1.2% 5.14 ± 0.09 The symbol ―n.d.‖ indicates no data available for that particular stone.

288

Table 22: Summary of Physical Property Averages for All Meteorite Groups.

No. Magnetic Susceptibility Type No. Stones Bulk Density (g cm-3) Grain Density (g cm-3) Porosity Meteorites (log χ) max max max max Average Average Average Average min min min min 3.77 4.14 26.6% 5.64 H fall 207* 116 3.35 ± 0.01 3.71 ± 0.01 9.5% ± 0.4% 5.30 ± 0.01 2.51 3.18 0.3% 4.57 3.65 4.14 26.6% 5.53 207 116* 3.35 ± 0.02 3.70 ± 0.01 9.4% ± 0.5% 5.31 ± 0.01 2.51 3.42 1.3% 4.88 3.69 3.79 10.2% 5.61 H find 79* 63 3.42 ± 0.02 3.51 ± 0.01 2.9% ± 0.4% 5.05 ± 0.03 2.86 3.19 -3.7% 4.41 3.68 3.79 10.2% 5.57 79 63* 3.41 ± 0.02 3.51 ± 0.02 3.0% ± 0.4% 5.02 ± 0.03 2.86 3.19 -3.7% 4.41 3.86 3.90 16.3% 5.47 L fall 216* 122 3.30 ± 0.01 3.58 ± 0.01 8.0% ± 0.3% 4.87 ± 0.01 2.98 3.39 -3.1% 3.81 3.65 3.83 16.3% 5.17 216 122* 3.29 ± 0.01 3.57 ± 0.01 8.0% ± 0.3% 4.88 ± 0.02 2.98 3.39 -3.1% 3.81 3.56 3.79 12.1% 5.20 L find 107* 67 3.34 ± 0.01 3.46 ± 0.01 3.6% ± 0.3% 4.62 ± 0.03 2.94 3.27 -3.9% 3.90 3.54 3.61 10.0% 5.20 107 67* 3.32 ± 0.01 3.45 ± 0.01 3.7% ± 0.4% 4.58 ± 0.04 2.94 3.27 -3.9% 3.90 3.51 3.63 19.4% 5.15 LL fall 51* 33 3.18 ± 0.02 3.52 ± 0.01 9.5% ± 0.6% 4.13 ± 0.05 2.80 3.41 -0.1% 3.33 3.47 3.63 19.4% 4.71 51 33* 3.17 ± 0.02 3.51 ± 0.01 9.8% ± 0.7% 4.08 ± 0.07 2.80 3.41 1.4% 3.33 3.41 3.50 14.2% 4.65 LL find 12* 10 3.22 ± 0.04 3.42 ± 0.02 5.8% ± 1.2% 4.05 ± 0.14 2.93 3.27 0.6% 2.86 3.41 3.50 14.2% 4.65 12 10* 3.21 ± 0.05 3.42 ± 0.02 6.2% ± 1.4% 4.06 ± 0.17 2.93 3.27 0.6% 2.86 5.55 5.66 5.8% 5.79 CB 4* 2 5.25 ± 0.19 5.65 ± 0.01 3.9% ± 1.9% 5.57 ± 0.10 4.90 5.63 2.0% 5.31 5.31 5.66 10.8% 5.68 4 2* 5.18 ± 0.13 5.65 ± 0.01 8.3% ± 2.5% 5.57 ± 0.11 5.05 5.63 5.8% 5.46 3.84 3.66 -2.3% 5.39 CH 2* 1 3.79 ± 0.05 3.65 ± 0.00 -3.7% ± 1.4% 5.30 ± 0.08 3.74 3.65 -5.1% 5.22 4.86 CI 2* 1 1.57 2.42 34.9% 4.49 ± 0.38 4.11 3.39 3.66 23.4% 4.77 CK 19* 7 2.90 ± 0.05 3.58 ± 0.02 17.7% ± 1.7% 4.67 ± 0.01 2.54 3.37 -0.6% 4.59 3.39 3.66 23.3% 4.72 19 7* 3.00 ± 0.11 3.55 ± 0.04 14% ± 3.6% 4.66 ± 0.02 2.66 3.37 -0.6% 4.60 2.47 4.70 36.7% 4.77 CM 43* 13 2.27 ± 0.02 2.96 ± 0.04 22.2% ± 0.7% 3.80 ± 0.05 1.88 2.74 13.7% 3.30 2.47 3.26 36.7% 4.77 43 13* 2.20 ± 0.05 2.92 ± 0.04 24.7% ± 1.9% 3.93 ± 0.12 1.88 2.74 15.0% 3.30 3.48 3.78 41.3% 5.00 CO 33* 14 3.03 ± 0.05 3.52 ± 0.03 13.6% ± 1.7% 4.49 ± 0.04 2.18 2.99 -2.1% 3.91 3.30 3.74 34.2% 4.81 33 14* 3.06 ± 0.06 3.48 ± 0.05 11.6% ± 2.7% 4.48 ± 0.07 2.42 3.17 -2.1% 3.96 3.30 3.66 21.2% 4.80 CO fall 19* 6 3.10 ± 0.07 3.36 ± 0.02 7.6% ± 2.3% 4.41 ± 0.05 2.71 2.99 -2.1% 3.91 3.30 3.56 14.5% 4.70 19 6* 3.16 ± 0.12 3.35 ± 0.04 5.7% ± 4.2% 4.41 ± 0.09 2.87 3.17 -2.1% 3.96 3.48 3.78 41.3% 5.00 CO find 14* 8 2.98 ± 0.05 3.64 ± 0.05 18.3% ± 1.7% 4.55 ± 0.08 2.18 3.30 2.6% 4.23 3.23 3.74 34.2% 4.81 14 8* 2.93 ± 0.05 3.64 ± 0.05 19.4% ± 1.9% 4.57 ± 0.10 2.42 3.50 9.2% 4.31 3.94 3.88 25.0% 5.34 CR 9* 7 3.11 ± 0.14 3.42 ± 0.08 9.5% ± 2.7% 5.02 ± 0.07 2.29 3.06 -1.5% 4.64 3.94 3.88 25.0% 5.34 9 7* 3.13 ± 0.18 3.44 ± 0.09 9.4% ± 3.1% 5.00 ± 0.08 2.29 3.06 -1.5% 4.64 3.70 3.86 27.9% 4.95 CV (all) 75* 12 2.97 ± 0.03 3.61 ± 0.01 17.7% ± 1.0% 3.91 ± 0.06 2.59 3.25 -6.2% 3.23 3.46 3.68 27.7% 4.86 75 12* 3.03 ± 0.09 3.54 ± 0.04 14.6% ± 3.1% 4.08 ± 0.18 2.59 3.25 0.6% 3.23 3.18 3.86 27.9% 4.24 CVo 56* 5 2.87 ± 0.01 3.64 ± 0.01 21.4% ± 0.5% 3.67 ± 0.02 2.59 3.25 6.0% 3.23 3.18 3.66 27.7% 4.24 56 5* 2.87 ± 0.11 3.50 ± 0.09 19.7% ± 4.7% 3.84 ± 0.18 2.59 3.25 6.0% 3.23 3.70 3.54 13.1% 4.95 CVr 15* 3 3.36 ± 0.06 3.48 ± 0.01 3.1% ± 1.6% 4.65 ± 0.06 2.94 3.38 -6.2% 4.30 3.46 3.51 8.3% 4.86 15 3* 3.37 ± 0.08 3.50 ± 0.01 3.6% ± 2.4% 4.70 ± 0.15 3.22 3.48 0.6% 4.41 3.73 3.76 13.2% 5.72 EH fall 13* 5 3.48 ± 0.05 3.61 ± 0.03 3.7% ± 0.9% 5.36 ± 0.08 2.90 3.17 -1.0% 4.49 3.63 3.75 13.2% 5.72 13 5* 3.46 ± 0.07 3.62 ± 0.05 4.2% ± 1.3% 5.39 ± 0.12 2.97 3.22 0.4% 4.49

289

No. Magnetic Susceptibility Type No. Stones Bulk Density (g cm-3) Grain Density (g cm-3) Porosity Meteorites (log χ) max max max max Average Average Average Average min min min min 3.55 3.73 13.2% 5.72 EH find 5* 4 3.27 ± 0.12 3.49 ± 0.10 6.1% ± 2.2% 5.12 ± 0.27 2.90 3.17 0.4% 4.49 3.55 3.73 13.2% 5.72 5 4* 3.32 ± 0.14 3.54 ± 0.11 6.0% ± 2.8% 5.27 ± 0.28 2.97 3.22 0.4% 4.49 4.11 4.17 11.7% 5.68 EL fall 18* 6 3.58 ± 0.05 3.66 ± 0.04 2.1% ± 1.0% 5.50 ± 0.03 3.15 3.45 -2.5% 5.30 3.79 3.86 4.7% 5.55 18 6* 3.56 ± 0.06 3.64 ± 0.06 2.1% ± 0.9% 5.49 ± 0.02 3.37 3.45 -1.6% 5.43 4.51 4.46 12.0% 5.69 EL find 10* 8 3.35 ± 0.15 3.52 ± 0.11 5.2% ± 1.2% 4.99 ± 0.16 2.89 3.23 -1.1% 4.16 3.67 3.70 12.0% 5.69 10 8* 3.22 ± 0.10 3.43 ± 0.07 6.1% ± 1.2% 4.90 ± 0.19 2.89 3.23 0.8% 4.16 3.27 3.35 22.1% 3.36 Lun 9* 9 2.79 ± 0.10 3.04 ± 0.08 9.8% ± 2.8% 3.04 ± 0.10 2.36 2.64 1.1% 2.62 3.18 3.44 13.5% 3.52 She 9* 5 3.01 ± 0.05 3.28 ± 0.04 8.7% ± 1.2% 2.93 ± 0.08 2.83 3.08 3.5% 2.79 3.21 3.51 16.9% 3.68 Nak 6* 4 3.1 ± 0.05 3.42 ± 0.03 9.3% ± 2.0% 3.40 ± 0.08 2.88 3.29 3.0% 3.17 Cha 1* 1 3.48 3.73 6.8% 2.98 3.15 3.36 19.6% 3.95 How 23* 16 2.85 ± 0.03 3.26 ± 0.02 12.5% ± 0.9% 3.34 ± 0.06 2.61 2.96 2.0% 2.81 3.12 3.34 19.0% 4.44 Euc 65* 31 2.84 ± 0.02 3.19 ± 0.01 10.9% ± 0.6% 3.07 ± 0.06 2.61 2.99 -2.5% 2.56 3.37 3.51 15.2% 3.67 Dio 20* 11 3.10 ± 0.03 3.43 ± 0.01 9.2% ± 1.0% 3.04 ± 0.07 2.92 3.36 1.4% 2.62 3.15 3.44 21.4% 4.72 Aub 19* 9 2.90 ± 0.05 3.21 ± 0.02 8.7% ± 1.4% 3.58 ± 0.10 2.53 3.11 1.7% 2.94 3.24 3.48 6.8% 3.15 Ang 2* 2 3.21 ± 0.03 3.42 ± 0.05 6.2% ± 0.6% 2.96 ± 0.19 3.18 3.37 5.7% 2.77 3.36 3.53 12.3% 5.28 Ure 20* 11 3.22 ± 0.02 3.36 ± 0.02 4.0% ± 0.9% 4.60 ± 0.08 3.04 3.25 -1.7% 3.93 3.59 3.88 12.4% 5.51 Aca 6* 4 3.46 ± 0.04 3.69 ± 0.05 6.2% ± 1.5% 5.20 ± 0.09 3.33 3.55 2.0% 4.99 3.82 4.16 9.1% 5.68 Lod 4* 3 3.53 ± 0.13 3.74 ± 0.21 5.4% ± 2.0% 5.24 ± 0.24 3.24 3.38 0.2% 4.74 3.90 3.65 15.1% 4.29 Bra 8* 4 3.48 ± 0.09 3.55 ± 0.02 1.5% ± 3.3% 4.06 ± 0.07 3.10 3.47 -10.7% 3.79 3.47 3.79 13.0% 5.55 Win 6* 3 3.24 ± 0.08 3.60 ± 0.13 7.9% ± 2.2% 4.90 ± 0.16 2.96 3.24 4.2% 4.44 * Denotes whether the row is averaged over stones or whole meteorite values. ―±‖ values are uncertainties in the averages, based on the standard deviation of the mean.

290

Table 23: Data for Mesosiderites, Pallasites and Iron Meteorites.

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Arispe Iron ICD Find Vatican 40 12.84 6.94 ± 0.26‡ n.d. n.d. n.d. Arispe Iron ICD Find Vatican 39 87.16 7.65 ± 0.08‡ n.d. n.d. n.d. Arispe Iron ICD Find Vatican 38 162.85 7.32 ± 0.02‡ n.d. n.d. n.d. Arispe Iron ICD Find Average 7.40 - Augustinovka Iron IIIAB Find Vatican 44 1199.13 6.99 ± 0.17‡ 7.13 ± 0.07‡ 1.9% ± 2.5% n.d. Canyon Diablo Iron IAB Find Vatican 174 212.26 7.84 ± 0.18‡ n.d. n.d. n.d. Canyon Diablo Iron IAB Find Vatican 172 737.81 7.21 ± 0.16‡ n.d. n.d. n.d. Canyon Diablo Iron IAB Find Average 7.34 - Chisenga Iron IIIAB Fall LNHM BM1993,M4 91.11 7.64 ± 0.09 7.85 ± 0.04 2.7% ± 1.2% 5.38 ± 0.09 Indian Valley Iron IIAB Find Vatican 453 576.98 7.17 ± 0.17‡ 7.71 ± 0.21‡ 6.9% ± 3.3% n.d. Kayakent Iron IIIAB Fall LNHM BM1969,9 99.6 7.33 ± 0.24 7.77 ± 0.04 5.7% ± 3.1% 5.38 ± 0.09 Mount Morris (Wisconsin) Iron IAB Find CMS 389.1x 29.4 3.06 ± 0.04 3.28 ± 0.01 6.8% ± 1.3% 4.42 ± 0.08 Muzaffarpur Iron IAB Fall LNHM BM1983,M12 30.45 7.50 ± 0.20 7.81 ± 0.08 4.0% ± 2.8% 5.37 ± 0.08 N'goureyma Iron Fall IOM I 27.10 36.45 7.88 ± 0.29 7.67 ± 0.06 -2.8% ± 3.9% 5.35 ± 0.12 N'Kandhla Iron IID Fall Monnig M 826.1 86.2 7.82 ± 0.13 7.86 ± 0.03 0.5% ± 1.6% 5.04 ± 0.12 Nyaung Iron IIIAB Fall LNHM BM1983,M13 11.29 6.24 ± 0.42 7.69 ± 0.21 18.9% ± 5.9% 5.20 ± 0.12 Odessa (iron) Iron IAB Find Vatican 712 91.101 7.90 ± 0.20‡ n.d. n.d. n.d. Odessa (iron) Iron IAB Find Vatican 711 480.501 7.01 ± 0.14‡ n.d. n.d. n.d. Odessa (iron) Iron IAB Find Average 7.14 - Santa Catharina Iron IAB Find Vatican 836 50.15 5.00 ± 0.12‡ n.d. n.d. n.d. Santa Catharina Iron IAB Find Vatican 833 102.1 4.03 ± 0.08‡ n.d. n.d. n.d. Santa Catharina Iron IAB Find Vatican 834 83.23 3.10 ± 0.08‡ n.d. n.d. n.d. Santa Catharina Iron IAB Find Vatican 832 118.08 4.35 ± 0.09‡ n.d. n.d. n.d. Santa Catharina Iron IAB Find Vatican 830 183.43 6.36 ± 0.13‡ n.d. n.d. n.d. Santa Catharina Iron IAB Find Vatican 831 132.45 2.73 ± 0.06‡ n.d. n.d. n.d. Santa Catharina Iron IAB Find Average 4.02 - Scottsville Iron IIAB Find Vatican 859 489.72 7.52 ± 0.16‡ 7.77 ± 0.36‡ 3.3% ± 4.9% n.d. Sikhote-Alin Iron IIAB Fall Monnig M 350.11 66.4 7.61 ± 0.10 7.84 ± 0.04 3.0% ± 1.4% 5.32 ± 0.08 Sikhote-Alin Iron IIAB Fall Monnig M 350.8 45.31 7.44 ± 0.12 7.87 ± 0.05 5.5% ± 1.7% 5.42 ± 0.08 Sikhote-Alin Iron IIAB Fall Monnig M 350.9 84.47 7.97 ± 0.20 7.73 ± 0.04 -3.2% ± 2.6% 5.18 ± 0.09 Sikhote-Alin Iron IIAB Fall Monnig M 350.10 76.68 7.66 ± 0.14 7.76 ± 0.03 1.2% ± 1.8% 5.33 ± 0.09 Sikhote-Alin Iron IIAB Fall Average 7.70 7.79 1.1% 5.31 Steinbach Iron IVA Find AMNH 315 73.38 4.55 ± 0.02 5.78 ± 0.03 21.3% ± 0.5% 5.51 ± 0.10 Steinbach Iron IVA Find Vatican 913 40.16 4.19 ± 0.10‡ n.d. n.d. n.d. Steinbach Iron IVA Find Average 4.42 5.78 23.6% 5.51 Tamentit Iron IIIAB Find Vatican 922 577.01 7.55 ± 0.15‡ 7.32 ± 0.13‡ -3.2% ± 2.8% n.d. Acfer 063 Mes Find AMNH 4759 58.43 4.53 ± 0.08 4.38 ± 0.02 -3.5% ± 1.9% 5.87 ± 0.12 Bondoc Mes B4 Find AMNH 4616 85.81 7.17 ± 0.13 6.81 ± 0.04 -5.3% ± 2.0% 5.55 ± 0.12 Clover Springs Mes A2 Find AMNH 4391 11.9 4.20 ± 0.10 4.12 ± 0.06 -2.0% ± 2.9% 5.61 ± 0.12 Crab Orchard Mes Find Vatican 248 178.18 4.37 ± 0.11‡ n.d. n.d. n.d. Crab Orchard Mes Find Vatican 247 246.57 4.40 ± 0.09‡ n.d. n.d. n.d. Crab Orchard Mes Find Vatican 246 544.26 4.09 ± 0.09‡ n.d. n.d. n.d. Crab Orchard Mes Find Average 4.21 - Emery Mes A3 Find AMNH 4367 A 30.25 4.58 ± 0.04 4.42 ± 0.04 -3.7% ± 1.3% 5.78 ± 0.12 Emery Mes A3 Find AMNH 4367 B 24.19 4.65 ± 0.06 4.48 ± 0.05 -3.7% ± 1.8% 5.71 ± 0.12 Emery Mes A3 Find Average 4.61 4.45 -3.7% 5.74 Estherville Mes A3/4 Fall Monnig M 223.8 46.45 5.40 ± 0.05 5.45 ± 0.03 0.9% ± 1.1% 5.49 ± 0.08 Ilafegh 002 Mes Find AMNH 4760 A 62.28 4.05 ± 0.01 4.08 ± 0.02 0.9% ± 0.5% 5.40 ± 0.10 Ilafegh 002 Mes Find AMNH 4760 B 26.57 4.37 ± 0.25 4.24 ± 0.03 -3.1% ± 6.0% 5.54 ± 0.10 Ilafegh 002 Mes Find Average 4.14 4.13 -0.3% 5.47 Łowicz Mes A3 Fall Monnig M 885.1 58.98 5.15 ± 0.06 5.17 ± 0.02 0.3% ± 1.3% 5.53 ± 0.10 Mincy Mes S1 Find Vatican 623 74.61 4.31 ± 0.09‡ 4.62 ± 0.05‡ 6.6% ± 2.2% n.d. Mincy Mes S1 Find Vatican 622 410.32 4.19 ± 0.10‡ n.d. n.d. n.d. Mincy Mes S1 Find Average 4.21 4.62 9.0% - Morristown Mes Find AMNH 305 C 47.59 4.36 ± 0.02 4.27 ± 0.02 -2.0% ± 0.7% 5.63 ± 0.10 Morristown Mes S1-2 Find Vatican 655 190.18 4.36 ± 0.11‡ n.d. n.d. n.d. Morristown Mes S1-2 Find Vatican 654 235.8 4.51 ± 0.11‡ n.d. n.d. n.d. Morristown Mes S1-2 Find Vatican 653 893.87 4.04 ± 0.09‡ n.d. n.d. n.d. Morristown Mes S1-2 Find Average 4.17 4.27 2.5% 5.63 Vaca Muerta Mes A1 Find IOM A 34.1 63.3 3.14 ± 0.03 3.17 ± 0.01 1.0% ± 1.0% 3.59 ± 0.10 Vaca Muerta Mes A1 Find IOM A 34.2 63.3 3.08 ± 0.03 3.12 ± 0.01 1.5% ± 1.1% 3.52 ± 0.09 Vaca Muerta Mes A1 Find Average 3.11 3.14 1.2% 3.55

291

Meteorite Type Fall Collection Catalog Mass Bulk Density Grain Density Porosity Magnetic Susceptibility (g) (g cm-3) (g cm-3) (log χ) Ahumada Pal Find Vatican 9 62.96 4.88 ± 0.26‡ n.d. n.d. n.d. Australia Pal Find AMNH 439 21.04 2.53 ± 0.03 3.17 ± 0.02 20.2% ± 1.2% 4.86 ± 0.08 Pal Find Vatican 145 101.66 4.88 ± 0.13‡ n.d. n.d. n.d. Brenham Pal Find Vatican 144 132.72 4.72 ± 0.24‡ n.d. n.d. n.d. Brenham Pal Find Vatican 143 246.87 4.66 ± 0.16‡ n.d. n.d. n.d. Brenham Pal Find Average 4.73 - Eagle Station Pal Find Vatican 279 69.41 4.41 ± 0.11‡ n.d. n.d. n.d. Finmarken Pal Find Vatican 331 321.65 4.78 ± 0.14‡ 5.07 ± 0.18‡ 5.9% ± 4.3% n.d. Glorieta Mountain Pal Find Vatican 367 25.51 6.56 ± 0.30‡ n.d. n.d. n.d. Glorieta Mountain Pal Find Vatican 366 51.99 7.22 ± 0.27‡ n.d. n.d. n.d. Glorieta Mountain Pal Find Vatican 365 53.77 7.03 ± 0.23‡ n.d. n.d. n.d. Glorieta Mountain Pal Find Vatican 364 111.5 7.46 ± 0.21‡ n.d. n.d. n.d. Glorieta Mountain Pal Find Vatican 363 203.62 7.75 ± 0.17‡ n.d. n.d. n.d. Glorieta Mountain Pal Find Vatican 362 558.14 7.71 ± 0.16‡ n.d. n.d. n.d. Glorieta Mountain Pal Find Average 7.59 - Pal Find Vatican 282 29.66 5.19 ± 0.18‡ n.d. n.d. n.d. Imilac Pal Find Vatican 280 67.7 4.97 ± 0.12‡ n.d. n.d. n.d. Imilac Pal Find Average 5.03 - Marjalahti Pal Fall IOM P 4.1 17.87 6.61 ± 0.33 7.26 ± 0.12 9.0% ± 4.7% 5.66 ± 0.08 Mount Vernon Pal Find Monnig M 447.1 28.35 4.06 ± 0.14 4.24 ± 0.04 4.1% ± 3.5% 5.59 ± 0.12 The symbol ―n.d.‖ indicates no data available for that particular stone. ‡Some bulk or grain density data for Vatican stones were provided by G. Consolmagno and predate R. Macke’s involvement in the study.

292

Table 24: Porosity Averages by Shock Stage for All Chondrite Falls.

Chondrite Shock Stage Type S1 S2 S3 S4 S5 S6 CC* 21.6% ± 2.0% 9.4% ± 2.2% 11.7% ± 9.6% 2.9% ± 2.3% - - H 9.3% ± 2.7% 12.0% ± 3.3% 8.0% ± 1.0% 5.5% ± 1.0% - - L - 11.9% ± 2.3% 7.9% ± 0.9% 6.4% ± 0.8% 6.4% ± 1.2% 2.4% LL 17.8% 7.1% ± 1.6% 9.0% ± 1.3% - - - EC - 1.7% ± 1.0% - 1.9% - - ALL 19.7% ± 2.1% 8.3% ± 1.3% 8.4% ± 0.6% 5.8% ± 0.6% 6.4% ± 1.2% 2.4% *Both falls and finds were included in carbonaceous chondrite data. ―±‖ values are uncertainties in the averages, based on the standard deviation of the mean.

293

Table 25: Porosity Averages by Petrographic Type for Chondrite Falls.

Chondrite Petrographic Type Type 1 2 3 4 5 6 CC 34.9% 23.1% ± 2.2% 21.0% ± 2.7% 15.0% ± 5.8% - - H - - 8.8% ± 5.2% 10.2% ± 1.3% 10.1% ± 0.8% 7.8% ± 0.8% ( + model ) - - 9.7% ± 1.5% 10.0% ± 0.8% 9.0% ± 0.6% 7.7% ± 0.7% L - - 3.4% ± 0.3% 8.7% ± 1.5% 8.0% ± 1.0% 8.1% ± 0.3% ( + model ) - - 3.7% ± 0.6% 6.7% ± 1.3% 6.6% ± 0.8% 6.8% ± 0.3% LL - - 10.2% ± 2.7% 12.1% ± 2.2% 8.6% ± 1.6% 9.4% ± 0.8% ( + model ) - - 9.1% ± 2.3% 11.4% ± 2.0% 7.9% ± 1.2% 9.8% ± 0.7% EC - - - 2.9% ± 0.6% 2.6% ± 1.9% 1.7% ± 0.9% ALL 34.9% 23.1% ± 2.2% 13.6% ± 1.9% 10.2% ± 0.9% 6.2% ± 0.5% 4.4% ± 0.3% ―+ model‖ includes both measured porosities for falls and model porosities for finds. ―±‖ values are uncertainties in the averages, based on the standard deviation of the mean.

294

APPENDIX C: COPYRIGHT PERMISSIONS

295

Permission for use of: R. J. Macke, D. T. Britt and G. J. Consolmagno (2010) Analysis of systematic error in ―bead method‖ measurements of meteorite bulk volume and density.

Planetary and Space Science 58, 421-426.

The journal Planetary and Space Science is published by Elsevier Press, who permit authors to use their own content in a dissertation. The permission is expressed on the following web address (accessed Aug 22, 2010): http://www.elsevier.com/wps/find/authorsview.authors/copyright#whatrights.

This page states, ―… These rights are retained and permitted without the need to obtain specific permission from Elsevier. These include… the right to use the journal article, in full or in part, in a thesis or dissertation.‖

296

297

298

LIST OF REFERENCES

Al-Kathiri, A., Hofmann, B. A., Jull, A. J., & Gnos, E. (2005). Weathering of meteorites from Oman: Correlation of chemical/mineralogical weathering proxies with 14C terrestrial ages and the influence of soil chemistry. Meteoritics & Planetary Science, 40, 1215-1239.

Allègre, C. J., Manhès, G., & Gopel, C. (1995). The age of the Earth. Geochimica et Cosmochimica Acta, 59, 1445-1456.

Anders, E. (1964). Origin, age and composition of meteorites. Space Science Review, 3, 583-714.

Anders, E., & Grevesse, N. (1989). Abundances of the elements - Meteoritic and solar. Geochimica et Cosmochimica Acta, 53, 197-214.

Ashworth, J. R., & Barber, D. J. (1976). Lithification of gas-rich meteorites. Earth and Planetary Science Letters, 30, 222-233.

Benedix, G. K., McCoy, T. J., Keil, K., & Love, S. G. (2000). A petrologic study of the IAB iron meteorites: Constraints on the formation of the IAB-Winonaite parent body. Meteoritics & Planetary Science, 35, 1127-1141.

Bevan, A. W., & Grady, M. M. (1988). Mount Morris (Wisconsin): A fragment of the IAB iron Pine River? Meteoritics, 23, 349-352.

Bhandari, N., Shah, V. B., & Wasson, J. T. (1980). The Parsa enstatite chondrite. Meteoritics, 15, 225-233.

Bischoff, A., Rubin, A. E., Keil, K., & Stöffler, D. (1983). Lithification of gas-rich chondrite regolith breccias by grain boundary and localized shock melting. Earth and Planetary Science Letters, 66, 1-10.

Bischoff, A., Scott, E. R., Metzler, K., & Goodrich, C. A. (2006). Nature and Origins of Meteoritic Breccias. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 679-712). Tucson: University of Arizona Press.

Bischoff, A., Horstmann, M., Laubenstein, M., & Haberer, S. (2010). Asteroid 2008 TC3 - Almahata Sitta: Not only a ureilitic meteorite, but a breccia containing many different achondritic and chondritic lithologies (abstract #1763). 41st Lunar and Planetary Science Conference.

299

Bland, P. A., Berry, F. J., Smith, T. B., Skinner, S. J., & Pillinger, C. T. (1996). The flux of meteorites to the Earth and weathering in hot desert ordinary chondrite finds. Geochimica et Cosmochimica Acta, 60, 2053-2059.

Bland, P. A., Berry, F. J., & Pillinger, C. T. (1998a). Rapid weatering in Holbrook: An iron-57 Mössbauer spectroscopy study. Meteoritics & Planetary Science, 33, 127-129.

Bland, P. A., Sexton, A. S., Jull, A. J., Bevan, A. W., Berry, F. J., Thornley, D. M., et al. (1998b). Climate and rock weathering: A study of terrestrial age dated ordinary chondritic meteorites from hot desert regions. Geochimica et Cosmochimica Acta, 62, 3168-3184.

Bland, P. A., Zolensky, M. E., Benedix, G. K., & Sephton, M. A. (2006). Weathering of Chondritic Meteorites. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 853-867). Tucson: University of Arizona Press.

Bogard, D. D., & Johnson, P. (1983). Martian gases in an Antarctic meteorite. Science, 221, 651- 654.

Bogard, D. D., Nyquist, L. E., & Johnson, P. (1984). Noble gas contents of shergottites and implications for the martian origin of SNC meteorites. Geochimica et Cosmochimica Acta, 48, 1723-1739.

Boss, A. P., & Goswami, J. N. (2006). Presolar Cloud Collapse and the Formation and Early Evolution of the Solar Nebula. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 171-186). Tucson: University of Arizona Press.

Bottke, W. F., Cellino, A., Paolicchi, P., & Binzel, R. P. (2002). An Overview of the Asteroids: The Asteroids III Perspective. In W. Bottke, A. Cellino, P. Paolicchi, & R. Binzel (Eds.), Asteroids III (pp. 3-15). Tucson: University of Arizona Press.

Brearley, A. J. (2006). The Action of Water. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 587-624). Tucson: University of Arizona Press.

Brearley, A. J., & Jones, R. H. (1998). Chondritic Meteorites. In J. J. Papike (Ed.), Planetary Materials (pp. 3-1 - 3-398). Washington, D.C.: Mineralogical Society of America.

Britt, D. T., & Consolmagno, G. J. (2003). Stony meteorite porosities and densities: A review of the data through 2001. Meteoritics & Planetary Science, 38, 1161-1180.

Britt, D. T., Yeomans, D., Housen, K., & Consolmagno, G. (2002). Asteroid density, porosity and structure. In W. Bottke, A. Cellino, P. Paolicchi, & R. Binzel (Eds.), Asteroids III (pp. 485-500). Tucson: University of Arizona Press.

Brownlee, D., Tsou, P., Aléon, J., Alexander, C. M., Araki, T., Bajt, S., et al. (2006). Comet 81P/Wild2 under a microscope. Science, 314, 1711-1716.

300

Burbine, T. H., McCoy, T. J., Meibom, A., Gladman, B., & Keil, K. (2002). Meteoritic Parent Bodies: Their Number and Identification. In W. Bottke, A. Cellino, P. Paolicchi, & R. Binzel (Eds.), Asteroids III (pp. 653-667). Tucson: University of Arizona Press.

Buseck, P. R., & Holdsworth, E. F. (1972). Mineralogy and petrology of the Yilmia enstatite chondrite. Meteoritics, 7, 429-447.

Cadenhead, D. A., & Stetter, J. R. (1975). Specific gravities of lunar materials using helium pycnometry. Proceedings of the 6th Lunar Science Conference (pp. 3199-3206). New York: Pergamon Press, Inc.

Cadenhead, D. A., Wagner, N. J., Jones, B. R., & Stetter, J. R. (1972). Some surface characteristics and gas interactions of Apollo 14 fines and rock fragments. Proceedings of the Third Lunar Science Conference (pp. 2243-2257). Cambridge, MA: M.I.T. Press.

Chambers, J. (2006). Meteoritic Diversity and Planetesimal Formation. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 487-497). Tucson: University of Arizona Press.

Chapman, C. R., Morrison, D., & Zellner, B. (1975). Surface properties of asteroids: A synthesis of polarimetry, radiometry and spectrometry. Icarus, 25, 104-130.

Clark, B. E., Fanale, F. P., & Salisbury, J. W. (1992). Meteorite-asteroid spectral comparison: The effects of comminution, melting and recrystallization. Icarus, 97, 288-297.

Clayton, R. N., & Mayeda, T. K. (1983). Oxygen isotopes in eucrites, shergottites, nakhlites, and chassignites. Earth and Planetary Science Letters, 62, 1-6.

Clayton, R. N., & Mayeda, T. K. (1988). Formation of ureilites by nebular processes. Geochimica et Cosmochimica Acta, 52, 1313-1318.

Clayton, R. N., Mayeda, T. K., Onuma, N., & Shearer, J. (1976). Oxygen-isotopic composition of minerals in the Kenna ureilite. Geochimica et Cosmochimica Acta, 40, 1475-1476.

Connolly, H. C., Desch, S. J., Ash, R. D., & Jones, R. H. (2006). Transient Heating Events in the Protoplanetary Nebula. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 383-398). Tucson: University of Arizona Press.

Consolmagno, G. J., & Britt, D. T. (1998). The density and porosity of meteorites from the Vatican collection. Meteoritics & Planetary Science, 33, 1231-1241.

Consolmagno, G. J., & Britt, D. T. (2008). Asteroids and the stratigraphic sequence of the solar nebula (abstract #8180). Asteroids, Comets and Meteors 2008, LPI Contribution No. 1405.

Consolmagno, G. J., & Drake, M. J. (1977). Composition and evolution of the eucrite parent body: Evidence from rare earth elements. Geochimica et Cosmochimica Acta, 41, 1271- 1282.

301

Consolmagno, G. J., Britt, D. T., & Stoll, C. P. (1998). The porosities of ordinary chondrites: Models and interpretation. Meteoritics & Planetary Science, 33, 1221-1229.

Consolmango, G. J., Macke, R. J., & Britt, D. T. (2004). How homogeneous are stones from ordinary chondrite showers (abstract #5048). Meteoritics & Planetary Science, 67 Supplement.

Consolmagno, G. J., Macke, R. J., Rochette, P., Britt, D. T., & Gattacceca, J. (2006). Density, magnetic susceptibility, and the characterization of ordinary chondrite falls and showers. Meteoritics & Planetary Science, 41, 331-342.

Consolmagno, G. J., Britt, D. T., & Macke, R. J. (2008). The significance of meteorite density and porosity. Chemie der Erde - Geochemistry, 68, 1-29.

Corrigan, C. M., Zolensky, M. E., Dahl, J., Long, M., Weir, J., Sapp, C., et al. (1997). The porosity and permeability of chondritic meteorites and interplanetary dust particles. Meteoritics & Planetary Science, 32, 509-515.

Coulson, I. M., Beech, M., & Nie, W. (2007). Physical properties of Martian meteorites: Porosity and density measurements. Meteoritics & Planetary Science, 42, 2043-2054.

Cuzzi, J. N., & Weidenschilling, S. J. (2006). Particle-Gas Dynamics and Primary Accretion. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 353-382). Tucson: University of Arizona Press.

Dawson, K. R., Maxwell, J. A., & Parsons, D. E. (1960). A description of the meteorite which fell near Abee, Alberta, Canada. Geochimica et Cosmochimica Acta, 21, 127-144.

Dodd, R. T. (1981). Meteorites: A Petrologic-Chemical Synthesis. Cambridge: Cambridge University Press.

Easton, A. J. (1983). Grain-size distribution and morphology of metal in E-chondrites. Meteoritics, 18, 19-27.

Faeth, P. A., & Willingham, C. B. (1955). The assembly, calibration and operation of a gas adsorption apparatus for the measurement of surface area, pore volume distribution, and density of finely divided solids. In Technical Bulletin of Physical Chemistry. Pittsburgh: Carnegie Mellon Institute of Research.

Flynn, G. J., & Klock, K. (1998). Densities and porosities of meteorites: Implications for the porosities of asteroids (abstract #1112). Lunar and Planetary Science Conference XXIX.

Flynn, G. J., Moore, L. B., & Klock, W. (1999). Density and porosity of stone meteorites: Implications for the density, porosity, cratering, and collisional disruption of asteroids. Icarus, 142, 92-105.

302

Folco, L., Rochette, P., Gattacceca, J., & Perchiazzi, N. (2006). In situ identification, pairing, and classification of meteorites from Antarctica through magnetic susceptibility measurements. Meteoritics & Planetary Science, 41, 343-353.

Fujii, N., Miyamoto, M., Kobayashi, Y., & Ito, K. (1981). Differences of relative strength among chondrites measured by the vibrational fracturing rate. Memoirs of the National Institiute of Polar Research, 20, 362-371.

Gaffey, M. J., Burbine, T. H., & Binzel, R. P. (1993). Asteroid spectroscopy: Progress and perspectives. Meteoritics, 28, 161-187.

Gattacceca, J., Eisenlohr, P., & Rochette, P. (2004). Calibration of in situ magnetic susceptibility measurements. Geophysical Journal International, 158, 42-49.

Goodrich, C. A. (1992). Ureilites - A critical review. Meteoritics, 27, 327-352.

Goodrich, C. A. (2002). Olivine-phyric martian basalts: A new type of shergottite. Meteoritics & Planetary Science, 37 Supplement, B31-B34.

Gounelle, M., & Zolensky, M. E. (2001). A terrestrial origin for sulfate veins in CI1 chondrites. Meteoritics & Planetary Science, 36, 1321-1329.

Grady, M. M. (2000). Catalogue of Meteorites (5th ed.). Cambridge: Cambridge University Press.

Greenwood, R. C., & Franchi, I. A. (2004). Alteration and metamorphism of CO3 chondrites: Evidence from oxygen and carbon isotopes. Meteoritics & Planetary Science, 39, 1823- 1838.

Grossman, J. (Ed.). (2009). Meteoritical Bulletin Database. Retrieved September 21, 2009, from http://tin.er.usgs.gov/meteor/metbull.php

Guskova, E. G. (1985). Magnetic properties of enstatite chondrites. Meteoritika, 44, 111-118.

Herd, R., Spray, J., Samson, C., Miller, S., & Christie, I. (2003). 3D imaging and modelling with a space-qualified laser camera system: Development of terrestrial applications and potential for planetary exploration (abstract #1718). 34th Lunar and Planetary Science Conference.

Herndon, J. M., Rowe, M. W., Larson, E. E., & Watson, D. E. (1972). Magnetism of meteorites: A review of Russian studies. Meteoritics, 7, 263-284.

Hildebrand, A. R., McCausland, P. J., Brown, P. G., Longstaffe, F. J., Russell, S. D., Tagliaferri, E., et al. (2006). The fall and recovery of the Tagish Lake meteorite. Meteoritics & Planetary Science, 41, 407-431.

Horstmann, M., & Bischoff, A. (2010). Characterization of spectacular lithologies from the Almahata Sitta breccia (abstract #1784). 41st Lunar and Planetary Science Conference.

303

Hutchison, R. (2004). Meteorites: A Petrologic, Chemical and Isotopic Synthesis. Cambridge: Cambridge University Press.

Hutson, M. L. (1996). Chemical studies of enstatite chondrites. Ph.D. thesis, University of Arizona, Tucson, Arizona, USA.

Jarosewich, E. (1990). Chemical analysis of meteorites: A compilation of stony and analyses. Meteoritics, 25, 323-337.

Jenniskens, P., Shaddad, M. H., Numan, D., Elsir, S., Kudoda, A. M., Zolensky, M. E., et al. (2009). The impact and recovery of asteroid 2008 TC3. Nature, 458, 485-488.

Jull, A. J., Wlotzka, F., & Donohue, D. J. (1991). Terrestrial ages and petrologic description of Roosevelt County meteorites (abstract). Lunar and Planetary Science XXIV, 667-668.

Kallemeyn, G. W., & Wasson, J. T. (1982). The compositional classification of chondrites: III. Ungrouped carbonaceous chondrites. Geochimica et Cosmochimica Acta, 46, 2217-2228.

Kallemeyn, G. W., & Wasson, J. T. (1986). Compositions of enstatite (EH3, EH4,5 and EL6) chondrites: Implications regarding their formation. Geochimica et Cosmochimica Acta, 50, 2153-2164.

Kallemeyn, G. W., Rubin, A. E., Wang, D., & Wasson, J. T. (1989). Ordinary chondrites: Bulk compositions, classification, lithophile element fractionations, and composition- petrographic type relationships. Geochimica et Cosmochimica Acta, 53, 2747-2767.

Kallemeyn, G. W., Rubin, A. E., & Wasson, J. T. (1991). The compositional classification of chondrites: V. The Karoonda (CK) group of carbonaceous chondrites. Geochimica et Cosmochimica Acta, 55, 881-892.

Kallemeyn, G. W., Rubin, A. E., & Wasson, J. T. (1996). The compositional classification of chondrites VII: The R chondrite group. Geochimica et Cosmochimica Acta, 60, 2243- 2256.

Keil, K. (1962). Quantitativ-erzmikroskopische Integrationsanalyse der Chondrite. Chemie der Erde, 22, 281-348.

Keil, K. (1968). Mineralogical and chemical relationships among enstatite chondrites. Journal of Geophysical Research, 73, 6945-6976.

Keil, K. (1989). Enstatite meteorites and their parent bodies. Meteoritics, 24, 195-208.

King, E. A., Schonfeld, E., Richardson, K. A., & Eldridge, J. S. (1969). Meteorite fall at Pueblito de Allende, Chihuahua Mexico: Preliminary information. Science, 163, 928-929.

Kini, K. A., & Stacey, W. O. (1963). The adsorption of Helium by carbonaceous solids. Carbon, 1, 17-24.

304

Kohout, T., Elbra, T., Pesonen, L. J., Schnabl, P., & Slechta, S. (2006). Study of the physical properties of meteorites using mobile laboratory facility (abstract #1607). Lunar and Planetary Science Conference XXXVII.

Kohout, T., Kiuru, R., Montonen, M., Sheirich, P., Britt, D., Macke, R., et al. (Forthcoming). 2008 TC3 asteroid internal structure and physical properties inferred from study of the Almahata Sitta meteorite. Icarus.

Krot, A. N., Scott, E. R., & Zolensky, M. E. (1995). Mineralogical and chemical modification of components in CV3 chondrites: Nebular or asteroidal processing? Meteoritics, 30, 748- 775.

Krot, A. N., Meibom, A., Weisberg, M. K., & Keil, K. (2002). The CR chondrite clan: Implications for early solar system processes. Meteoritics & Planetary Science, 37, 1451- 1490.

Kukkonen, I., & Pesonen, L. J. (1983). Classification of meteorites by petrophysical methods. Bulletin of the Geological Society of Finland, 55, 157-177.

Kuoppamäki, K., Kivekäs, L., Timonen, J., Hartikainen, J., Hartikainen, K., & Pesonen, L. J. (1996). Determination of porosity of meteorites--Comparison of two methods. Meteoritics & Planetary Science, 31 suppl., A75.

Macke, R., Britt, D., & Consolmagno, G. (2008). Density and porosity of shower meteorites as indicators of meter-scale asteroid homogeneity (abstract #33.02). Bulletin of the American Astronomical Society, 40, 455.

Macke, R. J., Britt, D. T., & Consolmagno, G. J. (2010a). Analysis of systematic error in "bead method" measurements of meteorite bulk volume and density. Planetary and Space Science, 58, 421-426.

Macke, R. J., Britt, D. T., & Consolmagno, G. J. (2010b). Using grain density and magnetic susceptibility to quantify weathering in chondrite finds (abstract #5155). Meteoritics & Planetary Science, 45 Supplement, A124.

Macke, R. J., Consolmagno, G. J., Britt, D. T., & Hutson, M. L. (2010c). Enstatite chondrite density, magnetic susceptibility and porosity. Meteoritics & Planetary Science, 45, In press.

Macke, R. J., Britt, D. T., & Consolmagno, G. J. (2010d). Density, porosity and magnetic susceptibility of achondritic meteorites. Meteoritics & Planetary Science, 45, In press.

MacPherson, G. J., & Krot, A. N. (2002). Distribution of Ca-Fe-silicates in CV3 chondrites: Possible controls by parent-body compaction (abstract). Meteoritics & Planetary Science, 37 Supplement, A91.

Mason, B. (1962). Meteorites. New York: John Wiley & Sons, Inc.

305

Mason, B. (1966). The enstatite chondrites. Geochimica et Cosmochimica Acta, 30, 23-26.

Matsui, T., Hamano, Y., & Honda, M. (1980). Porosity and compressional-wave velocity measurement of Antarctic meteorites. Memoirs of the National Institute of Polar Resarch (special issue), 17, 268-275.

McCausland, P. J., & Flemming, R. L. (2006). Preliminary bulk and grain density measurements of Martian, HED and other achondrites (abstract #1574). 37th Lunar and Planetary Science Conference.

McCausland, P. J., Samson, C., & DesLauriers, A. (2007). Bulk densities of meteorites via visible light 3D laser imaging. Meteoritics & Planetary Science Supplement, 42, 5066.

McCausland, P. J. A., Brown, P. G. & Holdsworth, D. W. (2010) Rapid, reliable acquisition of meteorite volumes and internal features by laboratory X-ray micro-CT scanning (abstract #2584). 41st Lunar and Planetary Science Conference.

McSween, H. Y. (1977). Petrographic variations among carbonaceous chondrites of the Vigarano type. Geochimica et Cosmochimica Acta, 41, 1777-1790.

McSween, H. Y., Stolper, E. M., Taylor, L. A., Muntean, R. A., O'Kelley, G. D., Eldridge, J. S., et al. (1979). Petrogenetic relationship between and other achondrites. Earth and Planetary Science Letters, 45, 275-284.

Meyer, C. (2005). 12051: Ilmenite basalt. Retrieved March 2010, from Lunar Sample Compendium: http://curator.jsc.nasa.gov/lunar/compendium.cfm

Meyer, C. (2008). 15418: Heavily-shocked and brecciated granulite. Retrieved March 2010, from Lunar Sample Compendium: http://curator.jsc.nasa.gov/lunar/compendium.cfm

Meyer, C. (2009a). 14303: Crystalline matrix breccia. Retrieved March 2010, from Lunar Sample Compendium: http://curator.jsc.nasa.gov/lunar/compendium.cfm

Meyer, C. (2009b). 14321: Clast-rich, Crystalline matrix breccia. Retrieved March 2010, from Lunar Sample Compendium: http://curator.jsc.nasa.gov/lunar/compendium.cfm

Meyer, C. (2009c). 15555: Olivine-normative basalt. Retrieved March 2010, from Lunar Sample Compendium: http://curator.jsc.nasa.gov/lunar/compendium.cfm

Mittlefehldt, D. W., McCoy, T. J., Goodrich, C. A., & Kracher, A. (1998). Non-chondritic Meteorites from Asteroidal Bodies. In J. J. Papike (Ed.), Planetary Materials (pp. 4-1 - 4- 195). Washington, D.C.: Mineralogical Society of America.

Nehru, C. E., Pring, M., Weisberg, M. K., Ebihara, M. E., Clayton, R. N., & Mayeda, T. K. (1996). A new brachinite and petrogenesis of the group. Lunar and Planetary Science, 27, 943-944.

306

Pesonen, L., Terho, M., & Kukkonnen, I. T. (1993). Physical properties of 368 meteorites: Implications for meteorite magnetism and planetary geophysics. Proceedings of the NIPR Symposium of Antarctic Meteorites. 6, pp. 401-416. Tokyo: National Institute of Polar Resarch.

Pesonen, L. J., Kuoppamäki, K., Timonen, J., Hartikainen, J., Terho, M., & Hartikainen, K. (1997). On the porosity of L and H chondrites (abstract #1684). 28th Lunar and Planetary Science Conference.

Prinz, M., Nehru, C. E., Weisberg, M. K., & Delaney, J. S. (1984). Type 3 enstatite chondrites: A newly recognized group of unequilibrated enstatite chondrites (uec's) (abstract #1331). Lunar and Planetary Science Conference XV, 653-654.

Przylibski, T. A., Zagożdżon, P. P., Kryza, R., & Pilski, A. S. (2005). The Zaklodzie enstatite meteorite: Mineralogy, petrology, origin, and classification. Meteoritics & Planetary Science, 40 suppl., 185-200.

Quantachrome. (2003). Ultrapycnometer 1000 Micro-ultrapycnometer 1000 Operating Manual. Boynton Beach, FL.

Reitmeijer, F. J. (1998). Interplanetary Dust Particles. In J. J. Papike (Ed.), Reviews in Mineralogy and Geochemistry (Vol. 36: Planetary Materials, pp. 2-01 - 2-95). Washington, DC: Mineralogical Society of America.

Righter, K., Drake, M. J., & Scott, E. (2006). Compositional Relationships Between Meteorites and Terrestrial Planets. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 803-828). Tucson: University of Arizona Press.

Rochette, P., Lorand, J.-P., Fillion, G., & Sautter, V. (2001). Pyrrhotite and the remanent magnetization of SNC meteorites: A changing perspective on Martian magnetism. Earth and Planetary Science Letters, 190, 1-12.

Rochette, P., Sagnotti, L., Bourot-Denise, M., Consolmagno, G., Folco, L., Gattacceca, J., et al. (2003). Magnetic classification of stony meteorites: 1. Ordinary chondrites. Meteoritics & Planetary Science, 38, 251-268.

Rochette, P., Gattacceca, J., Chevrier, V., Hoffmann, V., Lorand, J.-P., Funaki, M., et al. (2005). Matching Martian crustal magnetization and magnetic properties of Martian meteorites. Meteoritics & Planetary Science, 40, 529-540.

Rochette, P., Gattacceca, J., Bonal, L., Bourot-Denise, M., Chevrier, M., Clerc, J.-P., et al. (2008). Magnetic classification of stony meteorites: 2. Non-ordinary chondrites. Meteoritics & Planetary Science, 43, 959-980.

Rochette, P., Gattacceca, J., Bourot-Denise, M., Consolmagno, G., Folco, L., Kohout, T., et al. (2009). Magnetic classification of stony meteorites: 3. Achondrites. Meteoritics & Planetary Science, 44, 405-427.

307

Rochette, P., Gattacceca, J., Ivanov, A. V., Nazarov, M. A., & Bezaeva, N. S. (2010). Magnetic properties of lunar materials: Meteorites, Luna and Apollo returned samples. Earth and Planetary Science Letters, 292, 383-391.

Rubin, A. E. (1983a). The Adhi Kot breccia and implications for the origin of chondrules and silica-rich clasts in enstatite chondrites. Earth and Planetary Science Letters, 64, 201- 212.

Rubin, A. E. (1983b). The Atlanta enstatite chondrite breccia. Meteoritics, 18, 113-121.

Rubin, A. E. (1983c). Impact melt-rock clasts in the Hvittis enstatite chondrite breccia: Implications for a genetic relationship between EL chondrites and Aubrites. (Proceedings of the 14th Lunar and Planetary Science Conference). Journal of Geophysical Research, 88 Supplement, B293-B300.

Rubin, A. E. (1984). The Blithfield meteorite and the origin of sulfide-rich, metal-poor clasts and inclusions in brecciated enstatite chondrites. Earth and Planetary Science Letters, 67, 273-283.

Rubin, A. E. (1988). Formation of ureilites by impact-melting of carbonaceous chondritic material. Meteoritics, 23, 333-337.

Rubin, A. E., & Keil, K. (1983). Mineralogy and petrology of the Abee enstatite chondrite breccia and its dark inclusions. Earth and Planetary Science Letters, 62, 118-131.

Rubin, A. E., Wang, D., Kallemeyn, G. W., & Wasson, J. T. (1988). The Ningqiang meteorite: Classification and petrology of an anomalous CV chondrite. Meteoritics, 23, 13-23.

Rubin, A. E., Scott, E. R., & Keil, K. (1997). Shock metamorphism of enstatite chondrites. Geochimica et Cosmochimica Acta, 61, 847-858.

Rubin, A. E., Warren, P. H., Greenwood, J. P., Verish, R. S., Leshin, L. A., Hervig, R. L., et al. (2000). Los Angeles: The most differentiated basaltic Martian meteorite. Geology, 28, 1011.

Russell, S. S., Zipfel, J., Grossman, J. N., & Grady, M. M. (2002). The Meteoritical Bulletin, No. 86. Meteoritics & Planetary Science, 37 Supplement, A157-A184.

Russell, S. S., Hartmann, L., Cuzzi, J., Krot, A. N., Gounelle, M., & Weidenschilling, S. (2006). Timescales of the Solar Protoplanetary Disk. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 233-251). Tucson: University of Arizona Press.

Ruzicka, A., Snyder, G. A., & Taylor, L. A. (1997). Vesta as the howardite, eucrite and diogenite parent body: Implications for the size of a core and for large-scale differentiation. Meteoritics & Planetary Science, 32, 825-840.

308

Schneider, D. M., Akridge, D. G., & Sears, D. W. (1988). Size distribution of metal grains and chondrules in enstatite chondrites (abstract). Meteoritics & Planetary Science, 36 Supplement, A136-A137.

Sears, D. W., Kallemeyn, G. W., & Wasson, J. T. (1982). The compositional classification of chondrites: II The enstatite chondrite groups. Geochimica et Cosmochimica Acta, 46, 597-608.

Sears, D. W., Weeks, K. S., & Rubin, A. E. (1984). First known EL5 chondrite: Evidence for dual genetic sequence for enstatite chondrites. Nature, 308, 257-259.

Smith, D. L., Samson, C., Herd, R., DesLauriers, A., Sink, J.-E., Christie, I., et al. (2006). Measuring the bulk density of meteorites nondestructively using three-dimensional laser imaging. Journal of Geophysical Research, 111, E10002.

Steele, W. A., & Halsey, G. D. (1954). The interaction of rare gas atoms with surfaces. Journal of Chemical Physics, 22, 979-984.

Stepniewski, M., Borucki, J., Durakiewicz, T., Giro, L., & Sharp, Z. D. (2000). Preliminary study of a new enstatite meteorite from Zaklodzie (southeast Poland). Meteoritics & Planetary Science, 35 suppl., A152.

Stöffler, D., Bischoff, A., Buchwald, V. F., & Rubin, A. E. (1988). Shock effects in meteorites. In J. F. Kerridge, & M. S. Matthews (Eds.), Meteorites and the Early Solar System (pp. 165-202). Tucson: University of Arizona Press.

Stöffler, D., Keil, K., & Scott, E. R. (1991). Shock metamorphism of ordinary chondrites. Geochimica et Cosmochimica Acta, 55, 3845-3867.

Strait, M. M., & Consolmagno, G. J. (2002). The nature and origin of meteorite porosity: Evidence from thin section analysis (abstract). Meteoritics & Planetary Science, 37 Supplement, A137.

Strait, M. M., & Consolmagno, G. J. (2005). Validation of methods used to determine microcrack porosity in meteorites (abstract #2073). Lunar and Planetary Science Conference XXXVI.

Strait, M. M., & Consolmagno, G. J. (2010). Visualizing porosity in meteorites (abstract #1533). 41st Lunar and Planetary Science Conference.

Terho, M., Pesonen, L. J., & Kukkonen, I. T. (1991). The petrophysical classification of meteorites: New results. Geological Survey of Finland Report Q29, 1/91/1.

Terho, M., Pesonen, L. J., Kukkonen, I. T., & Bukovanská, M. (1993). The petrophysical classification of meteorites. Studia Geophysica et Geodaetica, 37, 65-82.

Terraplus USA, Inc. (2003). Magnetic Susceptibility Meter SM-30 User’s Manual. Littleton, CO.

309

Torquato, S., Truskett, T. M., & Benedetti, P. G. (2000). Is random close packing of spheres well defined? Physical Review Letters, 84, 2065-2067. van Schmus, W. R., & Wood, J. A. (1967). A chemical-petrologic classification for the chondritic meteorites. Geochimica et Cosmochimica Acta, 31, 747-765.

Wasson, J. T., & Kallemeyn, G. W. (1988). Compositions of chondrites. Philosophical Transactions of the Royal Society of London A, 325, 535-544.

Wasson, J. T., & Wetherill, G. W. (1979). Dynamical, chemical and isotopic evidence regarding the formation locations of asteroids and meteorites. In T. Gehrels (Ed.), Asteroids (pp. 926-974). Tucson, AZ: University of Arizona Press.

Watters, T. R., & Prinz, M. (1979). Aubrites: Their origin and relationship to enstatite chondrites. Proceedings of the 10th Lunar and Planetary Science Conference (pp. 1073- 1093). New York: Pergamon Press, Inc.

Weidenschilling, S. J., & Cuzzi, J. N. (2006). Accretion Dynamics and Timescales: Relation to Chondrites. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 473-486). Tucson: University of Arizona Press.

Weisberg, M. K., Prinz, M., & Nehru, C. E. (1988). Petrology of ALH 85085: A chondrite with unique characteristics. Earth and Planetary Science Letters, 91, 19-32.

Weisberg, M. K., Prinz, M., Clayton, R. N., & Mayeda, T. K. (1993). The CR (Renazzo-type) carbonaceous chondrite group and its implications. Geochimica et Cosmochimica Acta, 57, 1567-1586.

Weisberg, M. K., Prinz, M., Clayton, R. N., Mayeda, T. K., Grady, M. M., & Pillinger, C. T. (1995). The CR chondrite clan. Proceedings of the NIPR Symposium on Antarctic Meteorites, 8, 11-32.

Weisberg, M. K., Prinz, M., Clayton, R. N., Mayeda, T. K., Grady, M. M., Franchi, I., et al. (1996). The K (Kakangari) chondrite grouplet. Geochimica et Cosmochimica Acta, 60, 4253-4263.

Weisberg, M. K., Prinz, M., Clayton, R. N., & Mayeda, T. K. (1997). The CV3 chondrites: Three subgroups, not two (abstract). Meteoritics & Planetary Science, 32 Supplement, A138- A139.

Weisberg, M. K., Prinz, M., Clayton, R. N., Mayeda, T. K., Sugiura, N., Zashu, S., et al. (2001). A new metal-rich chondrite grouplet. Meteoritics & Planetary Science, 36, 401-418.

Weisberg, M. K., McCoy, T. J., & Krot, A. N. (2006). Systematics and Evaluation of . In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the Early Solar System II (pp. 19-52). Tucson: University of Arizona Press.

310

Wilkison, S. L., & Robinson, M. S. (2000). Bulk density of ordinary chondrite meteorites and implications for asteroidal internal structure. Meteoritics & Planetary Science, 35, 1203- 1213.

Wlotzka, F. (1993). A weathering scale for the ordinary chondrites (abstract). Meteoritics, 28, 460.

Yomogida, K., & Matsui, T. (1981). Physical properties of some Antarctic meteorites. Memoirs of the National Institute of Polar Research, 20, 284-294.

Yomogida, K., & Matsui, T. (1982). Physical properties of some unequilibrated Antarctic ordinary chondrites. Memoirs of the National Institute of Polar Research (special issue), 25, 308-318.

Yomogida, K., & Matsui, T. (1983). Physical properties of ordinary chondrites. Journal of Geophysical Research, 88, 9513-9533.

Yoshikawa, M., Kawaguchi, J., & Yano, H. (2010). Asteroid sample return mission Hayabusa. its engineering challenges and scientific results (abstract #2746). 41st Lunar and Planetary Science Conference.

Zhang, Y., Benoit, P. H., & Sears, D. W. (1993). LEW 88180, LEW 87119, and ALH 85119: New EH6, EL7, and EL4 enstatite chondrites. Meteoritics, 28, 468.

311