Sie sind auf Seite 1von 423

KIMBERLITES,

ORANGEITES, AND
RELATED ROCKS
KIMBERLITES,
ORANGEITES, AND
RELATED ROCKS

Roger Howard Mitchell


Lakehead University
Thunder Bay, Ontario
Canada

SPRINGER SCIENCE+BUSINESS MEDIA, LLC


Library of Congress Cataloging-in-Publication Data

On file

ISBN 978-1-4613-5822-0 ISBN 978-1-4615-1993-5 (eBook)


DOI 10.1007/978-1-4615-1993-5

© 1995 by Springer Science+Business Media New York


Originally published by Plenum Press New York in 1995
Softcover reprint of the hardcover 1st edition 1995

Ali rights reserved

No part of this book may be reproduced, stored in a retrieval system, or transmitted in any form or by any
means, electronic, mechanical, photocopying, microfilming, recording, or otherwise, without written
permission from the Publisher
This work is dedicated to my wife
VALERIE ANNE DENNISON
in recognition of her support and
encouragement of my studies of
alkaline rocks, and for providing
an environment favorable to the
preparation of this book
What stuff ~is made, whereof it is born, 1am to learn.
The Merchant of Veil ice, Act I, Sc.1
W. Shakespeare

PREFACE

This is the final volume of a trilogy of monographs devoted to the petrology of primary
diamond-bearing rocks. It should be noted that, in common with the previous volumes,
Kimberlites and Petrology of Lamproites, the book is not about upper mantle xenoliths
or exploration for diamond.
The principal objective of this work is to present a revised terminology for primary
diamond-bearing rocks. To this end orangeites are recognized as a group of rocks distinct
in mineralogy, geochemistry, and petrology from archetypal kimberlites and lamproites.
The name orangeite is used in recognition of the initial discovery of these rocks in
the Orange Free State of South Africa. The name is not a new one: In 1928, Percy Wagner
suggested that the term orangite [sic] be used to describe the rocks that are currently
known as micaceous or group II kimberlites.
Much of this monograph is devoted to summarizing all that is known of the
mineralogy and geochemistry of orangeites. The work incorporates several thousand new
analyses of minerals from orangeites and kimberlites, together with new trace element
geochemical data. These data are used to compare and contrast orangeites with kimber-
lites and lamproites. The work also presents a revised textural genetic classification of
kimberlites and reviews some of the advances in kimberlite petrology since 1986.
The work is a critical synthesis of existing data, not merely a summary of received
concepts, although current hypotheses for the genesis of kimberlite and orangeite are
critically reviewed.
These rocks present particular challenges for petrological investigation because of
their mineralogical complexity and the absence of modern equivalents. Curiously, despite
over 100 years of study, we remain far from achieving a complete understanding of the
mineralogy of kimberlites and orangeites. Hence, it is not surprising that hypotheses
regarding the nature and evolution of their parental magmas remain, to this day, highly
speculative.
Although primary diamond-bearing rocks are relatively rare rock types, they have
an economic and petrological importance which far outweighs their relative obscurity
and therefore justifies continued study oftheir petrogenesis. Thus, it is particularly hoped
that the novel ideas and concepts advanced in the work will stimulate much further study
of the character and evolution of primary diamond-bearing rocks.
vii
viii PREFACE

Many people have contributed to the production of this book, and I wish to
acknowledge the following colleagues who have, over the past 15 years, contributed much
discussion, preprints, thin sections, rock samples, and unpublished and/or difficult-to-
obtain information: Steve Bergman, Roger Clement, Howard Coopersmith, Barry
Dawson, Alan Edgar, Tony Erlank, Steve Haggerty, Barry Hawthorne, Bram Janse,
Viktoria Komilova, Sergei Kostrovitskii, Henry Meyer, Peter Nixon, Nick Rock, Mike
Skinner, Patricia Sheahan, Simon Shee, Barbara Scott Smith, Andy Spriggs, Ken Tainton,
Larry Taylor, Nikolai Vladykin, Allan Woolley, and Peter Wyllie.
Special thanks go to Henry Meyer, for providing many hours of microprobe time at
Purdue University, and to Mike Skinner, for samples and the opportunity to examine the
Anglo-American Research Laboratory collection of orangeites. Particular thanks are
expressed to Ken Tainton for permission to quote data from his Ph.D. thesis.
Particular gratitude is expressed to Sam Spivak for drafting and photographic work
and to Anne Hammond for preparing many polished thin sections of these difficult rocks.
Their dedication, skills, and attention to detail are greatly appreciated by the author.
Others from Lakehead University who helped materially during the production of this
work include Reino Viitala (thin sections), Alan MacKenzie (electron microscopy), and
Shelley Moogk-Pickard (trace element analysis). Carl Hager is thanked for assistance in
using the Purdue microprobe.
Critical reviews of all or portions of Chapters 1, 2, and 3 were provided by Henry
Meyer, Barbara Scott Smith, lain Downie, Craig Smith, Alan Edgar, and Howard
Coopersmith. Valerie Dennison is thanked for proofreading innumerable drafts of this
manuscript, typing tables, collating the citations, and improving my written English!
The Natural Sciences and Engineering Research Council of Canada, Lakehead
University, and De Beers Consolidated Mines are acknowledged for financial and
logistical assistance during the course of preparation of this work.

Roger H. Mitchell
Thunder Bay
CONTENTS

Chapter 1. Kimberlites and Orangeites . . . . . . . . . . • • . . . . . • . • • 1


1.1. Etymology of Group I and II Kimberlites . . . . . 1
1.2. Definitions of Cryptogenic and Primary Phases . . 5
1.3. The Hybrid Nature of Kimberlites and Orangeites 5
1.4. Philosophy and Principles of Classification . 7
1.4.1. Modal versus Genetic Classifications 7
1.4.2. Petrological Clans. . . . . . . . . . . 8
1.4.3. The Lamprophyre Clan . . . . . . . . 9
1.4.4. Mineralogical-Genetic Nomenclature within
Petrological Clans. . . . . . . . . . . . . . . . 10
1.5. Mineralogical Comparisons between Kimberlites and Orangeites II
1.6. Definitions of Orangeites and Kimberlites . 14
1.6.1. Orangeites . . . . . . . . . . 14
1.6.2. Kimberlites . . . . . . . . . 14
1.7. Age and Distribution of Orangeites 16
1.8. Occurrences of Orangeites . 18
1. 8.1. Finsch . . . . . . . . 18
1.8.2. Barkly West Region. 21
1.8.2.1. Bellsbank. 21
1.8.2.2. Sover . . . 25
1.8.2.3. Newlands. 27
1.8.2.4. Pniel . . . 27
1.8.3. Boshof District . . . 27
1.8.3.1. Roberts Victor 28
1.8.3.2. New Elands 28
1.8.4. Winburg District ... 28
1.8.5. Kroonstad District .. 29
1.8.6. Swartruggens District. 30
1.8.7. Dokolwayo .. 31
1.8.8. Prieska District 32
1.8.9. Summary ... 35

ix
x CONTENTS

1.9. Textural-Genetic Classifications of Petrological Clans . 35


1.9.1. Kimberlites . . . . . . . . 37
1.9.1.1. Crater Facies . . . 37
1.9.1.2. DiatremeFacies . 41
1.9.1.3. Hypabyssal Facies 48
1.9.1.4. Spatial Relationships between Diatreme and
Hypabyssal Facies Kimberlites. . 51
l.9.2. Orangeites . . . . . . . . . . . . . 58
l.9.3. Melilitite Clan . . . . . . . . . . . . . . . 58
1.10. Petrographic Characteristics of Orangeite . . . . 60
1.11. Petrographic Differences with Respect to Kimberlites . . 74
1.12. Petrographic Differences with Respect to Lamproites . . 79

Chapter 2. Mineralogy of Orangeites • . • • • • . . . • • • • • . • . . • • •• 91


2.1. Mica . . . . . . . . . . . . . . . . . . 91
2.1.1. Paragenesis . . . . . . . . . . 91
2.1.2. Composition of Primary Mica. 94
2.1.2.]. Ah03-Ti02 Variation 97
2.1.2.2. Ah03-FeOT Variation 104
2.1.2.3. Macrocrysts versus Microphenocrysts . 109
2.] .2.4. Minor Elements. . . . . . 111
2.1.2.5. Trace Elements. . . . . . 114
2.1.3. Aluminous Mica-Microxenoliths 115
2.1.4. Aluminous Biotite Macrocrysts . . 117
2.1.5. Micas from the Swartruggens Male Lamprophyre 118
2.1.6. Summary of Mica Compositional Variation . 119
2.1.7. Solid Solutions in Orangeite Mica 122
2.1.8. Mica in Kimberlites . . 126
2.1.8.1. Macrocrysts . . . . . . . . 127
2.] .8.2. Primary Micas . . . . . . 128
2.1.8.3. Summary of Kimberlite Mica Compositional
Variation . 155
2.1.9. Mica in Lamproites . . . . . . . . 157
2.1.10. Mica in Minettes. . . . . . . . . . 160
2.1.11. Mica in Ultramafic Lamprophyres 161
2.2. Clinopyroxene . . . . 166
2.2.1. Paragenesis .,. 166
2.2.2. Composition... 166
2.2.2.1. Diopside 166
2.2.2.2. Titanian Aegirine . 171
2.2.2.3. Minor Elements . . ] 72
2.2.3. Pyroxenes in the Swartruggens Male Lamprophyre .. 176
2.2.4. Megacrystal Pyroxenes . . . . . . . . . . . . . . . .. 177
CONTENTS xi

2.2.5. Comparison with Pyroxenes in Kimberlites 178


2.2.6. Comparisons with Pyroxenes in Lamproites 179
2.2.7. Comparisons with Pyroxenes in Ultramafic
Lamprophyres . . . . . . . . . . . . . . . . 180
2.2.8. Comparisons with Pyroxenes from Minettes . 181
2.3. Olivine . . . . . . . 181
2.3.1. Paragenesis . . . . . . . . . . . . . . . . . 181
2.3.2. Composition . . . . . . . . . . . . . . . . 183
2.3.3. Comparisons with Olivines in Kimberlites . 185
2.3.4. Comparisons with Olivines in Lamproites . 187
2.4. Spinel . . . . . . . . 188
2.4.1. Paragenesis . . . . . . . . . . . . . . . 188
2.4.2. Composition . . . . . . . . . . . . . . 189
2.4.3. Comparisons with Kimberlite Spinels . 195
2.4.4. Spinel Compositional Variation in Lamproites and
Lamprophyres . . . . 198
2.5. Potassium Barium Titanates 200
2.5.1. Hollandite . . . . . . 200
2.5 .1.1. Paragenesis . 200
2.5.1.2. Composition 201
2.5.1.3. Comparison with Hollandites from Lamproites,
Kimberlites, and Other Potassic Rocks 207
2.5.2. Potassium Triskaidecatitanate . 213
2.5.3. Barium Pentatitanate 216
2.6. Perovskite . . . . . 216
2.6.1. Paragenesis . . . . . 216
2.6.2. Composition .... 218
2.6.3. Comparison with Perovskites from Kimberlite 221
2.6.4. Comparison with Lamproite Perovskite 223
2.7. Phosphates . . . . . . . . . 225
2.7.1. Apatite . . . . . . . 225
2.7.1.1. Paragenesis. 225
2.7.1.2. Composition 225
2.7.1.3. Comparison with Kimberlite and Lamproite
Apatite . 225
2.7.2. Daqingshanite . . . . 227
2.7.3. Monazite . . . . . . 228
2.7.4. Sr-REE Phosphate . 229
2.8. Amphiboles-Potassium Richterite 229
2.8.1. Paragenesis . . . . . . . . . 229
2.8.2. Composition . . . . . . . . 230
2.8.3. Comparison with Potassium Richterite in Lamproite and
Other Potassic Rocks 233
2.9. Potassium Feldspar 235
2.10. Ilmenite . . . . . . . . . . 237
xii CONTENTS

2.10.1. Comparison with Groundmass I1menites from


Kimberlites . . . . . . . . . . . . . . . . . . . . 240
2.10.2. Comparison with Ilmenites in Lamproites . . . . 241
2.11. Rutile . . . . . . . . . . . . . 241
2.12. Zirconium Silicates 243
2.12.1. Zircon .. 243
2.12.2. Wadeite .. 243
2.12.3. Zirconium-Bearing Garnet . . . . .. 244
2.12.4. Calcium Zirconium Silicate . . . . . 245
2.13. Carbonates . . . . . .......... . 245
2.13.1. Calcite . . . . . . . . . . . . . . . . 245
2.13.2. Dolomite . . . . . . . 245
2.13.3. Other Carbonates. 246
2.14. Other Minerals 247
2.15. Summary . . . . . . . . . 247

Chapter 3. Geochemistry of Orangeites • • • • • • • • • • • • • • • • • • • • • 249


3.1. Contamination and Alteration . . . . . . . . . 250
3.2. Primary Magma Compositions. . . . . . . . . . . . . . . . . . 252
3.3. Major Element Geochemistry . . . . . . . . . . . . . . . . . . 252
3.3.1. Unevolved Orangeites . . . . . . . . . . . . . . . 253
3.3.2. Mineralogical Controls on the Major Element
Geochemistry . . . . . . . . . . . . . . 255
3.3.3. Evolved Orangeites . . . . . . . . . . . . . . . 257
3.3.4. Comparison with Kimberlites . . . . . . . . . 258
3.3.5. Comparison with Lamproites . . . . . . . . . . . . 261
3.4. First-Period Transition Elements. . . . . . . . . . 262
3.5. Incompatible Elements . . . . . . . . . . . . . . . . . . . 264
3.5.1. Alkaline Earths . . . . . . . . . . . . . . . . . . . 265
3.5.2. Second- and Third-Period Transition Elements .. 265
3.5.2.1. Zirconium and Hafnium . . . . . . . . . . 265
3.5.2.2. Niobium and Tantalum . . . . . . . . . . 268
3.5.3. Thorium and Uranium. . . . . . . . . 271
3.5.4. Rare Earth Elements . . . . . . . . . . . . 272
3.5.5. Alkali Elements . . . . . . . . . . . 277
3.5.6. Lead . . . . . . . . . . . . . . . . . . . . . 279
3.6. Inter-Element Relationships . . . . . . . . . . . . . . . . . . . 280
3.6.1. Extended Incompatible Element Distribution Diagrams 280
3.6.2. CeIY and La/Yb versus Zr/Nb . . . 286
3.7. Peridotite Mixing and Assimilation . . . . 288
3.8. Radiogenic Isotopes .. . . . . . . . . . . 292
3.8.1. Strontium and Neodymium . . . . . 292
3.8.2. Lead . . . . . . . . . . . . . . . . . . . . . . . . . . . 296
CONTENTS xiii

3.9. Stable Isotopes . . . . . . 298


3.10. Summary . . . . . . . . . 301

Chapter 4. Petrogenesis of Orangeites and Kimberlites • • • • • • • • • • •• 303


4.1. Geochemical Models of Orangeite Petrogenesis Involving
Limited Partial Melting of Lherzolitic Sources . . . . . . . .. 304
4.1.1. Earlier Hypotheses . . . . . . . . . . . . . . . . . . .. 304
4.1.2. Melting of Enriched Mantle and Peridotite Entrainment 305
4.1.3. Three-Stage Processes-Depletion, Enrichment,
and Melting . . . . . . . . . . . . . . . . . . . . . . .. 306
4.2. Experimental Evidence Pertaining to Orangeite Petrogenesis. 310
4.2.1. Liquidus Experiments on Orangeite Compositions .. 310
4.2.2. Liquidus Experiments on Lamproite Compositions .. 311
4.2.3. Melting of Mica Pyroxenites . . . . . . . . . . . . .. 312
4.2.4. Phase Relations in the System: Phlogopite-Potassium
Richterite-Apatite . . . . . . . . . . . . . . . . . . . 314
4.3. Petrogenesis of Archetypal Kimberlites-Recent Hypotheses.. 315
4.3.1. Carbonated Lherzolite Sources . . . . . . . . . . . 315
4.3.1.1. Volatile Fluxing-Diapiric Model. . . . . 317
4.3.1.2. Partial Melting of Magnesite Peridotite . . 320
4.3.1.3. Partial Melting of Carbonated Phlogopite
Lherzolite . . . . . . . . . . . . . . . . . . . 322
4.3.1.4. Carbonates in the Mantle? . . . . . . . . . .. 323
4.3.2. liquidus Experimental Studies at High Pressures . .. 325
4.3.2.1. Liquidus Studies of Natural Kimberlite. . .. 325
4.3.2.2. Liquidus Studies of Synthetic Kimberlite . .. 327
4.3.2.3. Summary-A Cautionary Note . . . . . . .. 329
4.4. Geodynamic Models of Kimberlite and Orangeite Genesis .. 329
4.4.1. Transition Zone Melting . . . . . . . . . . . . . . .. 331
4.4.2. Metasome Melting and Mantle Plumes. . . . . . . .. 333
4.4.3. Hot-Spot Melting . . . . . . . . . . . . . . . . 335
4.4.4. Partial Melting of Heterogeneous Lithosphere . . . .. 338
4.4.5. Redox Melting . . . . . . . . . . . . . . . .. 341
4.5. Petrogenesis of the Orangeite Clan . . . . . . . . . . . . . . 343
4.5.1. Development of the Source . . . . . . . . . . . . . .. 344
4.5.1.1. Continental Roots . . . . . . . . . 345
4.5.1.2. Depth of Origin of Orangeite Magmas .. 346
4.5.1.3. Compositional Heterogeneities-Veined
Harzburgites . . . . . . . . . . . . . . . . 348
4.5.2. Melting of the Source. . . . . . . . . . . . . . . . .. 351
4.5.2.1. Causes of Melting . . . . . . . . . . . . . . .. 352
4.5.2.2. Melting of Veined Lithosphere . . . . . . . . . 358
4.5.3. Melt Segregation, Contamination, and Ascent . . . . .. 362
xiv CONTENTS

4.5.4. Low-Pressure and Post-Emplacement Crystallization. 364


4.5.5. Summary . . . . . . . . . . . . . . . . . . 366
4.6. Petrogenesis of the Kimberlite Clan . . . . . . . . 366
4.6.1. Nature of the Source and Depth of Melting 367
4.6.2. The Megacryst Problem . . . . . . . . . . 370
4.6.3. Contamination of Kimberlites in the Mantle. 371
4.6.4. Post-Emplacement Crystallization . . . . . . 373
4.6.5. Summary . . . . . . . . . . . . . . . . . . . . 373
4.7. Relationships of Orangeites to Kimberlites, Lamproites, and
Other UItrapotassic Magmas . . 375
4.7.1. Kimberlites . . . . . . . . . 375
4.7.2. Lamproites . . . . . . . . . 376
4.7.3. Other U1trapotassic Magmas 377
4.7.4. Summary . . . . . . 380
4.8. Primary Diamond Deposits . . . . 380

Postscript . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 385

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 407
Working towards precision in nomenclatural problems is
not simply play for bureaucratic minds but. on the
contrary. it is a suitable exercise which should lead to less
ambiguous communication.
Gian Gaspare Zujfa (1991)

KIMBERLITES AND ORANGEITES

The principal objective of this chapter is to compare, contrast, and illustrate the mineral-
ogy and petrology of archetypal kimberlites, also known as group I kimberlites, with
those of the group of diamond-bearing rocks which, in this work, are termed "orangeites."
The latter rocks have previously been termed "micaceous kimberlites" or "group II
kimberlites." The discussion, in conjunction with detailed mineralogical studies de-
scribed in Chapter 2, will demonstrate conclusively that kimberlites and orangeites cannot
be derived from the same parental magma and thus are not genetically related.
The second objective is to present suggestions for a revised textural-genetic classi-
fication of kimberlites and to show that this scheme is applicable to orangeites and
melilitoids, although it should be clearly realized that none of these is cogenetic.

1.1. ETYMOLOGY OF GROUP I AND II KIMBERLITES


Diamonds derived from kimberlites were first found in South Africa in October 1869
on the farms Bultfontein and Dorstfontein (Dutoitspan) and then in July 1870 at Koffie-
fontein and Jagersfontein (Roberts 1976). They were found in muddy material, excavated
from small quarries located adjacent to shallow water-filled depressions known in South
Africa as "pans." It was not recognized at that time that the pans were the surface
expression of cylindrical intrusions of igneous rock. Subsequently, the discovery of
significant quantities of diamonds on the farm Vooruitzigt in May 1871 led to the
discovery of three other diamond-rich deposits in the same area. Exploitation of these
deposits resulted in the establishment of four major diamond mines. It was around these
mines that the town of Kimberley was established. The story of these early discoveries
and the development ofthe region is described in detail by Roberts (1976), Lenzen (1980),
and Wilson (1982).
The diamonds occurring in the pans were initially considered to be of alluvial origin,
but, as the deposits were excavated, it became clear that they were derived from a highly
altered decomposed rock. This material, locally termed "yellow ground," was found to
give way, with increasing depth, to fresher competent rock termed "blue ground." By
1872 it was recognized that the deposits were not alluvial in origin and occurred in
cylindrical pipe-like structures. The blue ground was eventually recognized as an altered
igneous rock and the primary source of the diamonds.
1
2 CHAPTER!

The diamond-bearing rocks were not given a petrographic name until 1887, when
Henry Carvill Lewis, at a meeting of the British Association for the Advancement of
Science in Manchester, stressed the unique character of the rock. Lewis (1887, 1888)
described the rock as a porphyritic mica-bearing peridotite and recognized it as a type of
volcanic breccia. Following the type locality nomenclature rules of the day, this rock was
named, from its occurrence at Kimberley, kimberlite.
Many other occurrences of kimberlite were quickly discovered as knowledge of the
geological character of the original deposits at Kimberley was disseminated. By the end
of the nineteenth century kimberlites had been located throughout the Cape Province, the
Orange Free State, and the Transvaal. Typically, prospectors referred to any igneous rock
containing diamond as kimberlite. Curiously, geologists followed this practice, and
identification of a rock as kimberlite came to be based more upon the presence of this
trace accessory mineral than on the major mineral assemblage present. This practice
survives to this day in many parts of the world. Thus, diamond-bearing olivine melilitites
occurril)g on the eastern flanks of the Anabar Shield are termed "kimberlites" by Russian
petrologists (Kornilova et al. 1983). Similarly, in the Arkhangelsk diamond province,
rocks containing diamond are termed "kimberlite," whereas similar diamond-free rocks
are known as "picrites" (V. Tretyachenko, pers. comm.).
The Diamond Fields of South Africa by Percy Wagner, published in 1914, was the
first comprehensive summary of the occurrences of kimberlite in South Africa. This
extremely influential work contained some important petrological observations, but also
unfortunately set the stage for much of the confusion which was to follow concerning the
nature of kimberlite.
Wagner (1914, p. 78) proposed that basaltic and micaceous or /amprophyric kim-
berlites could be recognized. The latter variety was further divided into subtypes based
upon the presence or absence of augite (Wagner 1914, p. 107). Although Wagner noted
substantial petrographic differences between the two major groups, it should be realized
that his terminology is based primarily upon the macroscopic appearance of the rocks.
The only common factor linking these petrographically disparate rocks is the presence of
diamond and olivine macrocrysts.
Subsequently, Wagner (1928, p. 140) referred to the micaceous kimberlite constitut-
ing the Lion Hill Dyke (Orange Free State) as orangite (sic). As a footnote to this initial
use of the term he stated: "This name will be proposed by the writer in a forthcoming
publication to designate what has hitherto been known as mica-rich or lamprophyric
kimberlite" (Wagner 1928, p. 148). Clearly, Wagner (1928) recognized the fundamental
differences between his two varieties of "kimberlite" and believed there were sufficient
grounds for the reclassification of one variety as a new rock type.
Unfortunately, Wagner died not long after the publication of the 1928 paper, and the
promised article was either not written or never published. Consequently, the new name
never entered the petrographic lexicon. Instead, Wagner's (1914) classification continued
to be widely accepted, and until recently remained unchallenged, although minor refine-
ments were made by some petrologists (Williams 1932, Bobrievich et al. 1959a,b,
Milashev 1963, Dawson 1967, Frantsesson 1968, 1970, Vladimirov etal. 1981,1990,
Kornilova et al. 1983).
KlMBERLITES AND ORANGElTES 3

In the 1970s, modem mineralogical studies (see below) led to revisions in kimberlite
nomenclature and will ultimately, it is expected, lead to the demise of the Wagner (1914)
classification. Unfortunately, this classification is still used indiscriminantly by petrolo-
gists and geochemists who are unfamiliar with recent revisions to kimberlite terminology.
Consequently, any mica-rich kimberlite is commonly referred to as a micaceous kimber-
lite, with the implication that it is similar to Wagner's (1914) group of South African
micaeous kimberlites.
The first major revision to Wagner's (1914) classification scheme was made by
Mitchell (1970), who recommended that the term "basaltic kimberlite" be abandoned
because kimberlites do not contain feldspar and are neither mineralogically nor geneti-
cally related to basalts. Mitchell (1970) proposed that three mineralogical varieties of
kimberlite were recognizable on the basis of the dominance of olivine, phlogopite, and
calcite in their modes. These were kimberlite (equivalent to Wagner's basaltic kimberlite),
micaceous kimberlite (equivalent to Wagner's lamprophyric kimberlite), and calcite
kimberlite. The latter variety was introduced as a new name in recognition of the presence
of primary magmatic calcite in kimberlite. Prior to this all calcite in kimberlites was
considered to be secondary in origin.
Skinner and Clement (1979) and Clement et al. (1984) subsequently devised a modal
classification ofkimberlites which completely superseded Wagner's (1914) terminology.
Their approach was to classify kimberlites on the basis of the primary groundmass modal
mineralogy. The method is based upon the premise that the ubiquitous presence and
relative abundance of olivine is oflimited use for classification purposes. This conclusion
stems from the observation of Skinner and Clement (1979, p. 131) that it is difficult to
determine the relative amounts of phenocrystal and xenocrystal olivine in kimberlites.
Skinner and Clement (1979) noted that diopside, monticellite, phlogopite, calcite,
and serpentine are the five primary major groundmass constituents of the majority of
kimberlites. Hence, they proposed five basic subdivisions of kimberlite, named after the
groundmass mineral that is modally dominant. Olivine was considered to be ubiquitous,
and, although in modal abundance it varies widely within and between kimberlites, its
presence plays no role in the classification scheme for the reasons noted. In the Skinner
and Clement (1979) classification, most of Wagner's (1914) and Mitchell's (1970)
micaceous kimberlites are reclassified as phlogopite kimberlites.
The Skinner and Clement (1979) Classification has proven to be of great use because
it permits the petrographic comparison of kimberlites from diverse localities. A modified
version of the scheme forms the basis of the kimberlite classification used in this work
(1.11 ). The classification may be utilized only when it has been determined that the sample
being described is actually a kimberlite.
In 1983, Craig Smith, using samples selected on a petrographic basis by E. Michael
W. Skinner (De Beers), demonstrated that monticellite calcite serpentine kimberlite and
phlogopite kimberlite from the Kaapvaal craton (South Africa) possess distinctive Sr and
Nd isotopic compositions. On the basis of these data, Smith (1983) suggested that
monticellite serpentine calcite kimberlites, termed group I kimberlites, and phlogopite
kimberlites, termed group II kimberlites, were derived from asthenospheric and li-
thospheric mantle sources respectively. It should be particularly noted that Smith (1983)
4 CHAFfERl

introduced the terms group I and group II as a means of classifying kimberlites primarily
on an isotopic basis and not upon their petrographic character.
As a consequence of these isotopic studies, Smith et al. (1985b) and Skinner (1986,
1989) proposed that kimberlites can be divided into two distinct groups, I and II, on the
basis of differences in their distribution patterns, age, petrography, content of mantle-de-
rived xenocrysts, xenoliths and megacrysts, isotopic character, and whole rock geochem-
istry. It was noted by Skinner (1989) that petrographically and isotopically defined group
II kimberlites are found only in South Africa, whereas group I kimberlites are found
throughout the world. In southern Africa, group II kimberlites are typically older (Smith
et al. 1985a,b, 1994) than most of the geographically associated group I kimberlites.
Interestingly, Skinner's (1989) group II kimberlites correspond to occurrences of
micaceous kimberlites (sensu Wagner 1914). In retrospect, it is unfortunate that Wagner
(1914) was so influenced by the presence of diamond as a means of identifying kimberlite
that he did not initially follow existing petrographic practice and propose a type locality
name for this petrographically distinctive suite of "kimberlites."
Subsequent to the recognition of group I and II kimberlites there has been an interest
in characterizing the mineralogy and geochemistry of group II rocks (Fraser et al. 1985,
Fraser 1987, Dawson 1987, Mitchell and Meyer 1989a, Skinner 1989, Tainton and
Browning 1991, Tainton 1992, Fraser and Hawkesworth 1992, Skinner et al. 1994,
Tainton and McKenzie 1994, Mitchell 1994a). From these studies it has become evident
that group I and II "kimberlites" are mineralogically and geochemically quite distinct and
that group II rocks have closer affinities to lamproites than to group I kimberlites. On the
basis of this evidence, Mitchell (1991a,b, 1994a) has suggested thatthe rocks are derived
from genetically distinct parental magmas, and group II rocks should not be regarded as
a variety of kimberlite but as rocks belonging to an entirely different petrological lineage.
If this conclusion is correct, group II rocks should not be designated as kimberlites. Thus,
Mitchell (1989, 1991a, 1994a), Mitchell and Meyer (1989a), and Mitchell and Bergman
(1991) proposed the revival of Wagner's (1928) term "orangite" as a potential name for
these rocks. Following Wagner (1928), the name is given in recognition of their initial
discovery in the Orange Free State of South Africa. If this proposal is accepted, then there
exist three distinct major primary occurrences of diamond: kimberlite, orangeite, and
olivine lamproite.
This chapter examines the petrographic grounds for recognizing the term "orangeite"
as a useful rock name. Currently, few petrologists who actively study kimberlites and
related rocks question that group I and II rocks are distinctive and have different origins.
Resistance to eliminating the term group II kimberlite appears to be related more to
preserving the status quo than to be based upon petrological argument and evidence.
The differences between group I and II "kimberlites" have been recognized by the
International Union of Geological Sciences Subcommission on the Systematics of
Igneous Rocks (Woolley et al. 1995). Although the Subcommission does not sanction the
term "orangeite," it finds no compelling grounds to accept or reject the term until the
rocks in question have been sufficiently characterized. Woolley et al. (1995) provide a
definition of kimberlite (group I) based on Mitchell's (1986) definition and data obtained
during the preparation of this work. A preliminary definition of the rocks currently known
as "group II kimberlite," based upon data presented in this monograph, is also presented.
KIMBERLITES AND ORANGEITES 5

1.2. DEFINITIONS OF CRYPTOGENIC AND PRIMARY PHASES


In general, kimberlites and orangeites exhibit a distinctive inequigranular texture due
to the presence of large rounded-to-anhedral crystals set in a finer-grained matrix. The
origin of many of these crystals has not been satisfactorily determined. Some are without
doubt xenocrysts, but others may be either phenocrysts or xenocrysts. In recognition of
this ambiguity, Clement et al. (1984) and Mitchell (1986) have recommended that such
cryptogenic pseudophenocrystal phases be referred to as megacrysts and macrocrysts,
terms devoid of genetic inferences. In this work they are defined as follows:
Megacrysts are rounded-to-anhedral crystals greater than 1.0 cm in their maximum
dimension. Megacrystal kimberlites 01' mega crystal orangeites are arbitrarily defined as
containing greater than 5 vol % of such crystals.
Macrocrysts are rounded-to-anhedral crystals 0.5-10 mm in maximum dimension.
Many macrocrysts are merely fragments of megacrysts. Macrocrystal kimberlites or
macrocrystalorangeites are arbitrarily defined as containing greater than 5 vol % of these
crystals.
Small (<0.5 mm, commonly 1-500 Ilm), anhedral crystals which are compositionally
similar to megacrysts and macrocrysts may be found scattered throughout the ground-
mass. Such crystals are interpreted as fragments of disaggregated megacrysts and macro-
crysts. To indicate their provenance and set them apart from bonafide groundmass phases,
they are termed macrocrystal clasts or microcrysts.
Aphanitic kimberlites or aphanitic orangeites are varieties in which megacrysts and
macrocrysts are absent, or present, only in small quantities (<5 vol %; Apter et al. 1984).
Crystals of subhedral-to-euhedral habit, considered to be early-forming primary
liquidus phases, are termed phenocrysts (>0.5 mm) and microphenocrysts (0.1-0.5 mm).
Phenocrystal kimberlites are rare. Aphanitic kimberlites containing abundant olivine
microphenocrysts may be described as microporphyritic kimberlites. Orangeites are
typically characterized by the presence of microphenocrystal phlogopite and may also be
described as microporphyritic.
The term "porphyritic" should never be used to describe the characteristic inequi-
granular texture of macrocrystal kimberlite or orangeite. Much of the Russian literature
which refers to porphyritic kimberlite (Milashev et al. 1963, Bobrievich et al. 1964,
Komilova et al. 1983, Vladimirov et al. 1990) is in reality describing macrocrystal
kimberlite.
Small (typically <0.1 mm) euhedral-to-anhedral primary minerals which constitute
the bulk of the fine-grained groundmass are termed groundmass phases. These crystals
may be set in a very fine-grained-to-optically unresolvable primary matrix or mesostasis.
Note that the mineralogy of the mesostasis and the groundmass is different. The distinc-
tion is made to emphasize the conclusion, based on petrographic evidence, that crystal-
lization of groundmass minerals typically ceases before formation of the mesostasis.

1.3. THE HYBRID NATURE OFKIMBERLITES AND ORANGEITES


Kimberlites and orangeites are petrographically complex rocks. Both are hybrids
consisting of crystals originating from three distinct sources: mantle-derived xenoliths,
6 CHAPTER!

the Cr-poor megacrystlmacrocryst suite, and primary phases crystallizing from the
magma. The relative contribution to the overall mineralogy of any given rock from each
of these sources varies widely and accounts for the wide petrographic variation observed
in kimberlites and orangeites. Note that the Cr-poor megacrystlmacrocryst suite is absent
from most orangeites.
The principal mantle-derived xenoliths found in kimberlites and orangeites are garnet
lherzolite, garnet harzburgite, chromite harzburgite, spinel lherzolite, websterite, eclogite
and grosspydite, metasomatized peridotites (containing potassic richterite, phlogopite,
and the titanates, yimengite, and hawthorneite), and the MARID (mica-amphibole-
rutile-ilmenite-diopside) suite of rocks. Detailed descriptions of the mineralogy of
these rocks may be found in Sobolev (1977), Dawson (1980), and Nixon (1987).
Disaggregation of these xenoliths during incorporation into, and transportation by,
kimberlite or orangeite magmas results in the addition of a wide variety of xenocrysts to
the magma. The majority ofthese xenocrysts can be easily identified on the basis of their
compositional equivalence with the minerals in the xenoliths, e.g., chrome diopside and
Cr-pyrope derived from garnet lherzolite; subcalcic knorringitic garnets from garnet
harzburgite; jadeitic pyroxenes, kyanite, and grossular-rich garnets from eclogites.
Other xenocrysts, in particular olivine and phlogopite, have compositions which are
identical to those of minerals considered to have crystallized from the magma as primary
phases. As there are no simple textural or compositional means of identifying these
minerals as xenocrysts, they are commonly included in the macrocryst suite.
The Cr-poor megacrystlmacrocryst or discrete nodule suite consists principally of
single crystals of magnesian ilmenite, Cr-poor titanian pyrope, Cr-poor subcalcic diop-
side, enstatite, phlogopite, and zircon. Coarse-grained lamellar intergrowths of ilmenite
with clinopyroxene or orthopyroxene are common, as are small inclusions of one mineral
within larger crystals of another. Megacrysts are common constituents of kimberlites but
are rarely present in orangeites (Skinner 1989, Smith et al. 1985b).
The compositional variation and textural relations within the suite suggest that the
megacrysts represent a series of crystals precipitating from a differentiating magma. The
megacryst assemblage found within any given kimberlite is considered to be a hybrid
formed by the mixing of crystals derived from several episodes of crystallization of the
magma which was parental to the megacrysts. Megacrysts are not in eqUilibrium with
their transporting magmas at low pressures (Shee 1984, Pasteris 1980).
Megacrysts are considered to be either xenocrysts, unrelated to kimberlite (Nixon
and Boyd 1973, Pasteris et al. 1979, Hops et al. 1992), or cognate crystals formed in the
upper mantle from kimberlite magma (Harte and Gurney 1981, Hunter and Taylor 1984,
Mitchell 1986, 1987, Canil and Scarfe 1990). Reviews of the origin and nature of the
Cr-poor megacryst assemblage are given by Mitchell (1986) and Schulze (1987).
Primary phases include phenocrysts, microphenocrysts, and crystals of subhedral-
to-euhedral habit which represent minerals crystallizing in situ to form the groundmass
and mesostasis. The assemblage of these minerals together with their compositional
variation are considered to reflect the character of their parental magma. By detailed
studies of these assemblages it is possible to classify correctly rocks which, petrographi-
cally, are superficially similar but actually of diverse provenance.
KIMBERLITES AND ORANGElTES 7

1.4. pmLOSOPHY AND PRINCIPLES OF CLASSIFICATION


Igneous rocks are commonly classified using a combination of modal and textural
criteria. While such methods are appropriate for common igneous rocks, they are
considered by Mitchell and Bergman (1991), Scott Smith (1992), Mitchell (1994a), and
Woolley et al. (1995) to be inappropriate for kimberlites, lamproites, orangeites, and
lamprophyres. Many of these rocks cannot be unambiguously identified using standard
modal-textural classifications, either in the field or the laboratory, as they are very similar
in their macroscopic and petrographic appearance. Altered ultramafic alkaline rocks
present particular challenges, and it is often extremely difficult, if not impossible, to
classify correctly such rocks using simple petrographic criteria.
The similarity of modal and textural criteria has commonly resulted in rocks
belonging to different magmatic series being incorrectly classified. Clearly, this situation
is inappropriate for petrogenetic purposes, and modally based classifications of rocks
have led to suggestions of consanguineity where none actually exist, e.g., the original
description of olivine lamproites as kimberlites (McCulloch et al. 1983) and the classifi-
cation of sanidine-rich lamproites as minettes (Middlemost et al. 1988).
Mitchell (1994a) has noted that a particular assemblage of minerals arises from the
operation of petrogenetic processes and is not a fortuitous random association. In addition,
the type and composition of minerals crystallizing from a given magma type must be
controlled by, and reflect, the composition of that magma. Hence, it is suggested that
ambiguities of nomenclature may be resolved by incorporating into classification
schemes other significant descriptive criteria, i.e., mineral composition and accessory
mineral assemblages, believed to be of petrogenetic significance. This suggestion stems
from the belief of Mitchell (1994a) that classifications should have a genetic significance
if progress is to be made in understanding the origins of, and relationships between,
diverse ultramafic alkaline igneous rock suites.

1.4.1. Modal versus Genetic Classifications


Modal classifications are based upon the texture and volumetric percentages of the
dominant, or major, minerals present in a rock. No consideration is given to the compo-
sition of the minerals, with the exception of the feldspars, and accessory phases are
typically ignored. This approach to nomenclature has no genetic significance, and a given
rock is named without consideration of the character of coexisting comagmatic rocks.
Genetic or mineralogical classifications use textural and modal information in
conjunction with compositional data for some or all ofthe minerals present. Classification
is based upon assemblages of typomorphic or characteristic minerals, some of which
would be relegated to accessory status in purely modal classifications. This approach has
genetic implications, in that rocks are assigned to a petrological clan or suite of consan-
guineous rocks of widely varying texture and modal mineralogy. The mineral composi-
tional data permit discrimination between rocks which, on a modal petrographic basis,
would be given the same name, but which are actually genetically different, having
crystallized from different parental magmas.
Note that this type of genetic classification does not imply that we know how a
particular magma type was formed or how individual rocks in a comagmatic series
8 CHAFfER 1

GENETIC MODAL GENETIC

Ti - PHLOGOPITE TO Ti - TETRAFERRI - + MICA (45) • AI - PHLOGOPITE TO Ti-BIOTITE


PHLOGOPITE
( Fe ,K ) AI Si 30 e + ALKALI FELDSPAR (35)
• (Na,KIAISi 30e
TITANIAN POTASSIUM RICHTERITE + AMPHIBOLE (4)
• ARFVEDSONITE
PRIDERITE + OPAQUES (I)
• ILMENITE

+ + +
SAMDINE PHLOGOPITE LAMPROITE II "MINETTE"
II MINETTE

Figure 1.1. Modal versus genetic classifications of a feldspar and mica-bearing hypabyssal rock.

originated. Classification is based upon directly observable characteristics of the mineral


phases present and not upon inferred genetic criteria. Thus, the nomenclature is inde-
pendent of revisions to, or modifications of, hypotheses advanced to explain the petro-
genesis of a given comagmatic series.
The differences between modal and genetic classifications may be appreciated by
considering the names given to a hypabyssal or effusive rock consisting of mica (45
vol %), alkali feldspar (35 vol %), clinopyroxene (10 vol %), and accessory minerals (5
vol %). In a modal classification this rock would be termed a "minette" on the basis of
the dominance of mica and feldspar. However, Figure 1.1. demonstrates how the rock
may be termed either a "sanidine phlogopite lamproite" or a "minette" when the compo-
sition of the minerals and the accessory phases are taken into consideration. Clearly,
incorrect classification would have significant repercussions if the object of the exercise
is for petrogenetic or exploration purposes.
For the identification of a given rock, genetic classifications of the type proposed
here commonly require more information than is available from standard transmitted/re-
flected light studies. It is necessary to employ electron microbeam methods for the
analysis and identification of minerals, as experience has shown that many typomorphic
phases cannot be readily identified by thin-section petrography. The use ofbackscattered
electron imagery coupled with energy dispersive X-ray spectrometric microanalysis
(Mitchell 1995) is a prerequisite of these investigations. Use of electron microbeam
techniques may be considered by some exploration geologists, or petrologists who do not
specialize in alkaline rocks, to be a significant impediment to rock identification.
However, these analytical techniques are now routine, and their application to the problem
is mandatory if advances are to be made in the classification of alkaline rocks leading to
a better understanding of their petrogenesis or economic potential.

1.4.2. Petrological Clans


The concept of a petrological clan was introduced by Scottish petrologists at the end
of the nineteenth century and popularized by Reginald Daly in his influential text Igneous
Rocks and Their Origin (Daly 1914). It was first realized at this time that differentiation
and crystallization of particular magma types led to the formation of characteristic suites
of rock types. Rocks which are so related form a consanguineous or comagmatic series.
In this work a petrological clan is regarded as a group of rocks derived from a
particular type of parental magma which has been produced repeatedly in time and space.
KIMBERLITES AND ORANGEITES 9

Individual petrological provinces are composed of comagmatic rocks derived from


specific batches of this magma type (Mitchell 1994a). Note that rocks of diverse modal
character and chemical composition may be included in such a clan as they are genetically
related.
The term "clan" as a genetic concept, as used above, follows the original (eleventh
to twelfth century) form of the Scottish Highland clan system, in which membership of
a clan was defined in terms of actual or purported descent from a common ancestor
(Donaldson 1974, pp. 161-164, Oxford English Dictionary 1989, The Concise Scots
Dictionary 1985). The word "clan" is derived from the Gaelic "clann," meaning children.
This definition of clan is different from that given in standard geological reference
works, namely "a group of igneous rocks that are closely related in composition"
(American Geological Institute Glossary of Geology, The Encyclopedia of Igneous and
Metamorphic Petrology). This definition is particularly unsatisfactory because it has no
petrogenetic significance and "closely related" is not defined. Rocks which are "closely
related" in composition may be merely heteromorphs. Of greater significance is the fact
that rocks belonging to a comagmatic series are certainly not closely related in composi-
tion, e.g., olivine lamproite and sanidine richterite lamproite are both members of the
lamproite clan, yet they differ significantly in composition.

1.4.3. The Lamprophyre Clan


Rock (1986,1990) has linked kimberlites (and orangeites), lamproites, and lampro-
phyres (sensu lato) into a supergroup of rocks termed the "lamprophyre clan." Rock used
the term to refer to a group of rocks he believed looked superficially similar, which are
commonly associated in the field and have a number of petrological characteristics in
common, e.g., richness in volatiles, porphyritic texture, occurrence as minor intrusions.
This usage follows a looser and deri vative version of the original Scottish clan concept,
namely that a "clan" is group united by a common trait and not related by blood
(Donaldson 1974, Oxford English Dictionary 1989). Thus, Rock (1990) recognized that
members of his "lamprophyre clan" have distinct origins.
Mitchell (1994a,c) has discussed Rock's lamprophyre clan concept at length and
noted that it is a misnomer, since members of the clan are genetically unrelated. The clan
merely unites rocks which have crystallized under volatile-rich conditions. Rocks belong-
ing to different comagmatic series which are united by such a common characteristic are
better considered as an example of the facies concept (Cas and Wright 1987). Conse-
quently, Mitchell (1994c) proposes the recognition of a lamprophyre facies as a means
of conveying the concept that some members of a petrological clan crystallized under
different, i.e., volatile-rich, conditions than other members of that clan. Thus, Mitchell
(l994c) recommends that the term "lamprophyre clan" be abandoned because there is no
"lamprophyre magma type." Woolley et al. (1995) have also recommended that kimber-
lites and lamproites should not be regarded as members of a "lamprophyre clan."
Mitchell (l994c) recommends that the adjective "lamprophyric" be used to describe
a facies of rocks derived from a particular parent magma. This usage retains the original
meaning of "lamprophyric" as a description of observable characteristics of a particular
group of rocks within a petrological clan which set them apart from other associated rocks.
10 CHAPTER 1

Recognition of a lamprophyre facies does not imply that we know how rocks belonging
to the facies originated. However, recognition of the facies ultimately has genetic
significance, as it serves to draw our attention to rocks which must have formed by
processes specific to and/or different from other members of the comagmatic suite.
Although it is possible to recognize a lamprophyric facies of the kimberlite, orangeite,
melilitite, or lamproite clans, there are no sound petrogenetic reasons for gathering these
rocks under a single petrological banner, as they may form by diverse processes in very
different environments.

1.4.4. Mineralogical-Genetic Nomenclature within Petrological Clans


Mineralogical-genetic classifications of the type used in this work were first pro-
posed for kimberlites by Skinner and Clement (1979). The methodology was sub-
sequently applied to lamproites by Scott Smith and Skinner (1984), Mitchell (1985), and
Mitchell and Bergman (1991). Practical experience in using this style of nomenclature
by both academic and exploration petrologists during the last decade has demonstrated
its effectiveness in the description of kimberlites and lamproites (Jaques et al. 1986,
Mitchell 1986, Mitchell and Bergman 1991, Scott Smith 1992).
In mineralogical-genetic classifications, rocks are named on the basis of the nature
of the parental magma of the clan. This magma type is stated in a root name (equivalent
to a genus), e.g., kimberlite or lamproite. Individual rocks, or subdivisions of the clan,
are described by mineral names (equivalent to a species) given as prefixes to the root
name, e.g., leucite diopside lamproite. These compound names reflect the modal abun-
dance of the major phases present. Prefixes are given in order of increasing modal
abundance. Skinner and Clement (1979) recommended including in the name those
minerals which are present in amounts exceeding two thirds of the volumetric abundance
of the dominant mineral. Thus, a monticellite serpentine kimberlite would consist
predominantly of serpentine in combination with monticellite amounting to more than
two thirds of the serpentine abundance. Multiple prefixes may be added according to the
limits defined by Skinner and Clement (1979). Strict application of this style of classifi-
cation requires accurate determination of the mode. However, in practice, prefix names
are more commonly given simply in order of increasing modal abundance. This approach
reflects, in part, the fact that it is commonly very tedious, difficult, and time consuming
to determine the modes of fine-grained altered ultramafic alkaline rocks.
It should be stressed that mineralogical nomenclature may only be" utilized when a
rock has been identified as belonging to a particular petrological clan on the basis of the
overall mineralogy and/or other criteria. Mineralogical-genetic classifications of some
alkaline rocks have only become possible because of recent detailed petrological studies
of type suites of rocks derived from a particular magma type, e.g., the lamproite clan
(Scott Smith and Skinner 1984, Jaques et al. 1986, Mitchell and Bergman 1991). These
studies have enabled petrologists to recognize the extent of modal and compositional
variation in a given suite of consanguineous rocks. Prior to these studies it was difficult,
if not impossible, to characterize many isolated samples of alkaline ultrabasic igneous
rocks obtained during grassroots exploration or regional mapping programs.
KIMBERLITES AND ORANGEITES 11

In practice, mineralogical classification of a rock is usually based upon samples of


hypabyssal facies material, as crystallization is usually sufficiently slow to allow devel-
opment of typomorphic mineral assemblages. The classification of other textural varieties
is far more difficult as rapidly quenched lapilli usually have not crystallized diagnostic
groundmass minerals. In addition, crater and diatreme facies rocks, especially in tropical
environments, are particularly prone to alteration and/or weathering.
The existing mineralogical-genetic classifications of kimberlites and lamproites
have recently been endorsed by an lUGS Subcommission on the Systematics oflgneous
Rocks (Woolley et al. 1995). Further discussion of the mineralogical nomenclature of
orangeites and kimberlites may be found in Sections 1.10 and 1.11, respectively.

1.5. MINERALOGICAL COMPARISONS BETWEEN KIMBERLITES AND


ORANGEITES
Mitchell (1986, 1991a, 1994a) and Mitchell and Bergman (1991) have suggested
that the mineralogical and compositional differences between the rocks currently known
as group I and II kimberlites, are so profound that they must represent rocks derived from
different magma types. If this contention is correct, it follows that the rocks should possess
readily identifiable petrographic and mineralogical characteristics permitting them to be
classified according to the genetic principles described above.
Important mineralogical features of the rocks are compared in Table 1.1, from which
it is apparent that they differ greatly with respect to their megacrystlmacrocryst and
primary mineral assemblages. Table 1.1 is based upon previous mineralogical studies of
archetypal kimberlites, summarized by Mitchell (1986) and Skinner (1989), and the new
data for orangeites and kimberlites presented in Chapter 2 of this work.
Orangeites differ from kimberlites in that they
• Do not characteristically contain Cr-poor Ti pyrope, magnesian ilmenite, subcalcic
diopside, and enstatite megacrystslmacrocrysts. Although some orangeites in the
Prieska area 0.8.9) contain Mg-ilmenite and Cr-poor pyroxene macrocrysts (Skinner
et al. 1994), and others contain Cr-poor Ti pyrope megacrysts (Moore and Gurney
1991, Bell and Rossman 1992), they are not abundant.
• Contain primary microphenocrystal and groundmass diopside (see 2.2).
• Do not contain magnesian ulvospinel and spinels belonging to kimberlite spinel
evolutionary trend 1 (see 2.4.3).
• Do not contain monticellite.
• Are characterized by micas which evolve from phlogopite to tetraferriphlogopite.
Ba-rich aluminous micas belonging to the phlogopite-kinoshitalite solid solution
series are absent.
• Contain K-Ba titanates (2.5) and zirconium silicates (2.12).
• Rarely contain groundmass sanidine and potassium richterite.
• Contain perovskites and apatites which are characteristically enriched in rare earth
elements and strontium relative to the compositions of these minerals in kimber-
lites (see 2.6, 2.7).
12 CHAPTER 1

Table 1.1. Comparison of the Mineralogy of Kimberlites and Orangeites


Kimberlite Orangeite
Olivine
Macrocrysts Abundant-principally xenocrysts Common to rare-principally
xenocrysts
Phenocrys ts Common (FoS7_90) subhedral/euhedral Minor (F091 - 93 )
subhedral/euhedraVdog's tooth
Mica
Macrocrysts Minor, phlogopite-cryptogenic Common, phlogopite cognate
Microphenocrysts Rare, phlogopite Common, phlogopite
Groundmass Common, phlogopite-kinoshitalite Common, phlogopite-
reticulate laths tetraferriphlogopite
Poikilitic plates
Spinels Abundant, large (O.D1-O.1 mm) Minor to rare. small (<0.001-0.02 mm)
Typically Mg-chromite zoned to Mg- Euhedral Mg-chromite common,
ulvospinel (Trend I). Atoll spinels rarely zoned to Ti-magnetite (Trend
very common. Trend 2 spinels rare, 2). Atoll spinels rare. Mg-ulvospinels
only in varieties with macrocrystal absent
mica
Monticellite Common, may be pseudomorphed by Absent
carbonate and/or serpentine
Diopside Primary diopside absent, may occur in Microphenocryst. Common to rare.
contaminated groundmasses Commonly resorbed. Zoned to Ti-
aegirine
Perovskite Common, rounded to euhedral SrO- «I Rare, subhedral to poikilitic srO- «1-6
wt %) and (REE)zOrpoor «7 wt %) wt %) and (REE}z03 rich (3-16 wt
%)
Apatite Common to rare, euhedral prisms or Common euhedral prisms and poikilitic
acicular radiating aggregates in plates srO- (3-22 wt %) and
serpentine---calcite segregations. srO- (REEhOrrich (<1-10 wt %)
«1 wt %) and (REE}z0rpoor(<1 wt
%).
Carbonates Simple assemblages, common calcite, Common calcite, common Sr-Mn-Fe
minor dolomite. Rare Sr-REE dolomites, minor witherite, ancylite,
carbonates very evolved types strontianite, norsethite
Serpentine Abundant secondary and common Common secondary
primary in segregations
Sanidine Absent Rare groundmass
K-richterite Absent Rare groundmass
K-Ba hollandite Very rare, only evolved types Common
K 2Ti 130 27 Absent Common
Mn ilmenite Rare Common
Zr-silicates Very rare, only evolved types Common
Leucite Absent Rare pseudomorphs in poikilitic mica
REE-phosphates Absent Minor monazite, daqingshanite, Sr-
REE phosphate
Barite Rare Common
Quartz Absent Minor, groundmass
Macrocryst suite Characteristic Absent to extremely rare
KIMBERLITES AND ORANGEITES 13

• Contain a more varied assemblage of carbonates, including norsethite, strontian-


ite, and witherite, than is found in kimberlites (2.13).
Significant petrographic differences are also apparent, e.g., the modes of orangeites
are dominated by microphenocrystal and ground mass phlogopite, whereas phlogopite-
rich kimberlites are rare. Groundmass spinels and perovskites are typically less abundant
and finer grained than spinels and perovskites in kimberlites (Skinner 1989). Illustrations
of the petrographic characteristics of orangeites and kimberlites are given in Sections 1.1 0
and 1.11, respectively.
Geochemical differences (Chapter 3) include the distinctive isotopic compositions
of the rocks (Smith 1983) and the elevated abundances of incompatible elements (Sr, Ba,
REE, Zr) in orangeites relative to those of kimberlites.
Mitchell (1994a) has stated that, given the very different mineralogical character of
micaceous (group II) kimberlites relative to archetypal monticellite-bearing primary
diopside-free group I kimberlites, it is unlikely that, if this group of rocks were to be
discovered today, they would be termed "kimberlites."
It is concluded, on the basis of the evidence presented in Table 1.1, that orangeites
are not merely petrographic variants of archetypal kimberlites and thus must be derived
from a distinct magma type. If the two rock groups are not genetically related there is no
rational petrological reason for continuing to refer to group II rocks as kimberlites.
Perpetuation of the name will only suggest false petrogenetic relationships between these
rocks and bona fide kimberlites. Consequently, different names should be given to the
two groups of "kimberlites" to reflect their fundamentally different origins.
Hence, Mitchell (1994a), following Wagner (1928), has proposed that the term
"orangeite" be used to describe the rocks currently known as group II kimberlites, and
the terms group I and II kimberlites be abandoned. Group I kimberlites are best referred
to simply as "kimberlites." The group I prefix is not necessary as the clan name alone
conveys all of the required nomenclatural (and genetic) information.
The introduction of a new name is justified in that it recognizes a newly identified,
distinct magma type. Rocks crystallizing from this magma belong to the orangeite clan
and may be described by compound names of the type described in Section 1.4.4, e.g.,
apatite orangeite, richterite orangeite. Orangeite is chosen as a name for the clan because
the type localities of these rocks occur primarily within the Orange Free State of South
Africa.
All of the mineralogical criteria required to identify orangeites and distinguish them
from kimberlites, lamproites, and lamprophyres are readily observable by using a
combination of petrographic and electron microbeam methods. Recognition is not based
upon inferred or interpreted genetic criteria. However, the latter serve to confirm that
orangeites and kimberlites are generated from different magma types whose sources are
located at different depths within the mantle (see 3.8.1, 4.5, 4.6).
The value of the recognition of an orangeite clan may be appreciated by consideration
of rocks recently described from southern Africa by Tainton (1992), Tainton and Brown-
ing (1991), Clarke et al. (1991), and Skinner et al. (1994). These studies showed that
differentiation of certain group II kimberlites (Pniel, Sover North, Sweetput-Soutput,
Besterskraal) leads to the formation of rocks containing groundmass potassium feldspar
14 CHAPl'ER 1

and potassium titanium richterite. These rocks might be termed "sanidine richterite
lamproites" if they were classified in isolation from their consanguineous antecedents
with which they are modally gradational. However, it is clear from field relationships and
their mineralogy that they are genetically related to less-evolved olivine- and phlogopite-
rich rocks. Using existing terminology they could justifiably be termed "sanidine
richterite group IT kimberlites." This designation, although meaningful to kimberlite
petrologists who are aware of the mineralogical distinctions between the two groups, is
guaranteed to breed confusion among nonspecialists, who might not realize that there are
"kimberlites" and "kimberlites." Consequently, reference to these late differentiates as
sanidine richterite orangeites makes clear the genetic distinctions between these rocks,
kimberlites, and lamproites.

1.6. DEFINITIONS OF ORANGEITES AND KIMBERLITES


1.6.1. Orangeites
There is no previous detailed mineralogical definition of orangeite, as mineralogical
characteristics of these rocks have been included in definitions ofkimberlites (sensu lato).
A preliminary definition is given in Woolley et al. (1995). On the basis of the data
presented here and in Chapters 2 and 3, the following definition is suggested:
Orangeites are a clan of ultrapotassic peralkaline volatile-rich (dominantly H20-rich)
rocks, characterized by the presence of phlogopite macrocrysts and microphenocrysts,
together with groundmass micas, which vary in composition from phlogopite to tetrafer-
riphlogopite. Rounded olivine macrocrysts and euhedral primary olivines are common,
but are not always major constituents. Characteristic primary groundmass phases include
diopside, commonly zoned to, and mantled by, titanian aegirine, spinels ranging in
composition from Mg-chromite to Ti-magnetite, Sr- and REE-rich perovskite, Sr-rich apatite,
REE-rich phosphates (monazite, daqingshanite), potassian harlan titanates belonging to the
hollandite group, potassium triskaidecatitanates, Nb-rutile, and Mn-ilmenite. These are
set in a mesostasis which may contain calcite, dolomite, ancylite, and other rare earth
carbonates, witherite, norsethite, and serpentine. Evolved members of the group contain
groundmass sanidine and potassium richterite. Zirconium silicates (wadeite, zircon,
kimzeyitic garnets, Ca-Zr silicate) are common as late-stage groundmass minerals.
Quartz may occur rarely as a mesostasis mineral. Barite is a common secondary mineral.
Orangeites may be distinguished from kimberlites by the absence of monticellite,
magnesian ulvospinel, and Ba-rich micas belonging to the barian phlogopite-kinoshi-
talite series. In addition, orangeites, in common with kimberlites and lamproites, do not
contain melilite, alkali feldspar, plagioclase, kalsilite, or nepheline.

1.6.2. Kimberlites
Previous definitions of kimberlite (Clement et al. 1984, Mitchell 1979, 1986) in-
cluded the mineralogical characteristics of orangeites. The definition presented below is
based upon the definition of Mitchell (1986), which has been modified by the elimination
of minerals characteristic of the orangeite clan and incorporation of recent mineralogical
studies of groundmass micas in kimberlites (see 2.1.9). This definition of kimberlite has
KIMBERLITES AND ORANGEITES 15

been endorsed by the lUGS Subcommission on the Systematics of Igneous Rocks


(Woolley et al. 1995).
Kimberlites are a group of volatile-rich (dominantly C02) potassic ultrabasic rocks
commonly exhibiting a distinctive inequigranular texture resulting from the presence of
macrocrysts (and in some instances megacrysts), set in a fine-grained matrix. The
megalmacrocryst assemblage consists of anhedral crystals of olivine, magnesian ilmenite,
Cr-poor titanian pyrope, diopside (commonly subcalcic), phlogopite, enstatite, and
Ti-poor chromite. Olivine macrocrysts are a characteristic constituent in all but fraction-
ated kimberlites. The matrix contains a second generation of primary euhedral-to-subhe-
dral olivine which occurs together with one or more of the following primary minerals:
monticellite, phlogopite, perovskite, spinel (magnesian ulvospinel-Mg-chromite-
ulvospinel-magnetite solid solutions), apatite, and serpentine. Many kimberlites contain
late-stage poikilitic micas belonging to the barian phlogopite-kinoshitalite series. Nicke-
liferous sulfides and rutile are common accessory minerals. The replacement of earlier-
formed olivine, phlogopite, monticellite, and apatite by deuteric serpentine and calcite is
common. Evolved members of the group may be poor in, or devoid of, macrocrysts and/or
composed essentially of second-generation olivine, calcite, serpentine, and magnetite,
together with minor phlogopite, apatite, and perovskite.
Kimberlites are best identified using the typomorphic assemblage of primary min-
erals referred to in the above definition. It is particularly important to make a distinction
between cryptogenic, macrocrystal subcalcic diopside and primary groundmass diopside
and note that kimberlites do not contain the latter. When present, groundmass diopside is
a secondary phase, the crystallization of which is induced by the assimilation of siliceous
xenoliths (Clement 1982, Scott Smith et al. 1984). Macrocrystal diopside is included in
the definition in recognition of its common, but not characteristic, occurrence in kimber-
lites. Reference to this and other members of the macrocryst suite should be deleted from
the definition if they are subsequently proven to be entirely of xenocrystal origin. For this
reason bonafide xenocrysts such as diamond are not included in the definition.
Note that perovskites and apatites in kimberlites are poor in Sr and REE relative to
the compositions of these minerals in orangeites. Potassium feldspar and potassium
richterite are not found in the groundmass of kimberlites. The majority of kimberlites
examined to date also lack the suite of complex K-Ba titanates, K-Zr silicates, and Sr-,
Ba-, Mg-, and REE-rich carbonates which occur in the groundmass of orangeites. A
similar accessory mineral assemblage has been found only in highly evolved calcite
kimberlites, e.g., Benfontein, Wesselton sills (Mitchell 1994b). However, these kimber-
lites contain kinoshitalite and magnesian ulvospinel, i.e., typomorphic minerals which
serve to distinguish them from orangeites.
Scott Smith (1989) has reported from kimberlite pipes 2 and 5 of the Andra Pradesh
(India) province a Ca-aluminosilicate pseudomorphing a lath-shaped mineral which is
considered to have been originally melilite. Murthy et at. (1994) have found a similar
mineral in pipe 10 of this field and claim this to be fresh melilite; however, the
composition is not in accord with that of gehlenite. These particular Andra Pradesh
kimberlites also contain pectolite and have unusual olivine morphologies (Scott Smith
1989). Pseudoprimary pectolite in kimberlites elsewhere has been shown to result from
contamination (Scott Smith et ai. 1983), a conclusion which suggests that if melilite were
16 CHAPTER 1

originally present its formation may also have been due to contamination. Although these
occurrences suggest that melilite may occur in some kimberlites, it is concluded here that
as yet there are no proven occurrences of melilite in kimberlites or orangeites.

1.7. AGE AND DISTRIBUTION OF ORANGEITES


Figure 1.2 (after Skinner 1989, Skinner et al. 1994) illustrates the age and distribution
of orangeites and kimberlites in southern Africa. Skinner et al. (1992) have estimated that
there are 229 occurrences of orangeite (termed group II kimberlites) as compared with
580 occurrences of archetypal (or group I) kimberlite. Figure 1.2 shows that most
orangeites are distributed within a 400 x 1250 km belt trending southwest from Doko1-
wayo (Swaziland) to Eendekuil, near Sutherland (Cape Province), most of the occur-
rences being within the Kaapvaal craton as clusters of intrusions in the southwestern part.
Only in the Preiska area, located at the southwest margin of the craton, do a few atypical
orangeites occur in an off-craton setting (Clarke et al. 1991, Skinner et al. 1994). In
contrast, kimberlites are more widely distributed and characteristically occur in on- and
off-craton settings.

,....0500
o! 500 km
!
.......·0 ........ -
S:> 90 ~~~~ r'
O~S /,/
~ .
J
~50 .r'
h.. .r~ 0 I
i\ ~/-.' . . . . ..."..) .~- ~ 1:5 e . i
i I 90
O~~D
0,.
~oo
Kaapvaal
~
1200 (fe I
~.

nbI oo~.20
A.e " . 1 4 5 : J 'r-'
Craton .
115 (X)

(\'~. . ._.__ .J1S i09"·~::';~


CO 100 120, ~o
0 F~tl90
It~ 90 .'"
C3'c) .J

e liO
~137
SOUTH AFRICA o KIMBERLITES
e ORANGE ITES
~ Craton Margin
165= Age in M.Y.

Figure 1.2. Distribution of orangeites and kimberlites in southern Africa relative to the Kaapvaal craton. Data
points may represent clusters of several intrusions (after Skinner 1989).
KIMBERLITES AND ORANGEITES 17

Orangeites are unknown elsewhere in the world (Skinner 1989), whereas kimberlites,
identical in geochemical and mineralogical character to archetypal southern African
kimberlites, are commonly found on all continents. Rocks which have been ''termed
micaceous kimberlites," e.g., the Zagodochnaya kimberlite (Yakutia) and the Tunraq
kimberlite (Canada), are now considered in this work merely to represent kimberlites
which have been modally enriched in macrocrystal micas. Such rocks have been termed
"micaceous kimberlites" primarily on their macroscopic appearance. Unfortunately,
without further detailed mineralogical and geochemical studies, this appellation is com-
monly taken to imply that the rocks are equivalent to the southern African orangeites.
Orangeites in southern Africa range in age from about 200 Ma to 110 Ma (Skinner
1989). Within the belt of orangeite intrusions, ages appear to decrease progressively from
the northeast to the southwest (Figure 1.2). Skinner (1989) and Ie Roex (1986) consider
this trend is suggestive of passage of the craton over a hot spot, although Mitchell (1986,
this work) and Bailey (1993) disagree with this interpretation (see 4.4.3).
It is significant that the age trend is not well defined in the main clusters of Lower
Cretaceous orangeites occurring in the southwestern part of the craton (Figure 1.2).
Intrusions appear to have been emplaced contemporaneously over a period of 25 Ma in
different fields rather than sequentially. These youngest occurrences include the miner-
alogically atypical orangeites found at the margins of the craton in the Preiska area
(Skinner 1989, Skinner et al. 1994, Smith et al. 1994).
As an alternative explanation of the data it is suggested that there were two episodes
of orangeite magmatism rather than one continuous period of activity. Thus, Lower
Cretaceous (125-110 Ma) orangeites form a broadly contemporaneous group of intru-
sions, among which there are no obvious temporal and spatial correlations. The relatively
few Upper Jurassic (165-145 Ma) intrusions in the central part of the craton may be
regarded as separate events rather than representatives of the initial portions of a
chronological trend. The 200-Ma Dokolwayo intrusion is temporally anomalous relative
to all other orangeites in that it predates Karroo volcanism. The age and status of this
intrusion as an orangeite require reinvestigation (see 1.8.7).
In contrast to orangeites, southern African kimberlites exhibit a very wide range in
age and tectonic setting (Figure 1.2). While the majority of those in on-craton tectonic
settings are 100-80 Ma, three other significantly older periods of kimberlite magmatism
have been found: Kuruman 1600 Ma, Premier 1200 Ma, and Zimbabwe 500 Ma (Shee
et al. 1989). Kimberlites emplaced in mobile belts surrounding the Kaapvaal craton give
consistently younger ages, on the order of 54-72 Ma (Dawson 1989).
The Mayeng and Frank Smith kimberlites of the Bellsbank area have ages of 117
and 114 Ma (Apter et al. 1984, Smith et al. 1985a), respectively. These ages are similar
to those of the Bellsbank and Newlands orangeite dikes, 119 Ma and 114 Ma, respectively
(Smith et al. 1985a). Skinner (1989) has suggested that in cases where the differences in
age between geographically closely related orangeites and kimberlites, e.g., Bellsbank,
Preiska, are not great, the orangeites are less micaceous than orangeites from elsewhere,
and the slightly younger kimberlites exhibit isotopic signatures transitional between those
of orangeites and kimberlites.
18 CHAPTER 1

1.8. OCCURRENCES OF ORANGEITES


It is not the object of this work to provide a gazetteer or detailed descriptions of all
occurrences of orangeite. Unfortunately, few modem textural-genetic studies of orangeite
intrusions have been undertaken. Brief descriptions of the nine major districts (Figure
1.3) referred to in this work are given to illustrate the overall style of magmatism.

1.8.1. Finsch
The Finsch intrusion (Clement 1982), located near Postmasburg, Cape province, 170
km west of Kimberley (Figure 1.3), was discovered in 1960 and is the second largest pipe
in South Africa after the Premier kimberlite. The surface expression is roughly circular
with an area of 17.9 ha (Figure 1.4). The original area at the time of emplacement is
estimated to be about 100 ha. From the current erosion surface to a depth of about 140 m
the intrusion is emplaced in banded ironstones of the Asbesheuwels subgroup of the
Precambrian Griqualand West Supergroup. These are underlain by 30 m of interbedded
siliceous, dolomitic, and ferruginous sediments known as the Passage Beds. The lower

Kaapvaa/
./
/'
J Craton
r· ../
GIBEON • . \ r· )/ ( \ 13 3
I" . '\. ... . '...J • PRETORP 5
.r-.

NAMIBIA i L._ ..-" SWARTRUGG;NS 9 J~HAN~~~~~~:Y~'~


I 55 *-.;; ~'<; 0 \.
1'\ J r---q;~; /4 8 13· KROONSTAD
, ') POSTMASBURG.. Y? 0 0
I., ,,",-,~ . ../' I!INSCH .....,.." .BOSHOP.WINBURG
2.-' »-___.
~ KO~FIEFONTE~
KIMBE!!:P37
PRIESK~O ..,.,/

BRITSTOWN .J
4~
VICTORIA WEST

SOUTH AFRICA

o 500 km

Figure 1.3. Distribution of orangeite fields and locations referred to in the text. Numbers adjacent to field
boundaries indicate number of intrusions in the field (after Skinner et al. 1992),
KIMBERLITES AND ORANGEITES 19

A
N

100 m
L..........J

Datum 1590,136m. AMSL


Surface
. ': :.',:
....
....
: .. ,": .
'
"
' ... '.
: • : • II : : :1
t~-f~ .J.J
I I
ORANGEITE
.c Satellite~
L Orangeite -.I
I "~
1 I

if;;um .
'%

Figure 1.4. Size and shape of the Finsch diatreme (after Clement 1982).

part of the pipe (170-600 m) is emplaced in the Chuniespoort (Ghaa Plateau) dolomite.
The Rb-Sr age of the intrusion is 118 rna (Smith 1983).
Figure 1.4 shows that the sides of the diatreme dip steeply at about 80 0 and the
cross-sectional area decreases regularly with depth. Mining has not yet reached any
complex root zone.
The pipe consists of eight discrete intrusions between the surface and 348-m level
(Figure 1.5) of the open pit. These are numbered according to the sequence in which they
were discovered. F1, a pelletal-textured volcaniclastic heterolithic orangeite breccia, is
volumetrically the most important intrusion. Other significant intrusions are F2, F3, and
F7, although their relative importance depends upon the level at which they are observed
in the pipe. Thus, F7 is of minor importance between 196 and 220 m relative to F2,
whereas the reverse is observed at 348 m. Clement (1982) has determined that the F6
dolerite-poor hypabyssal intrusion cuts the F5 dolerite-rich hypabyssal intrusion. Both
intrusions are restricted in the open pit to a small projection outward from the pipe margin
20 CHAPTER I

F6

Figure 1.5. Composite plan illustrating the distribution of intrusions (FI-F8) and megaxenoliths (floating reef)
in the Finsch diatreme between the 196- and 220-m levels of the open pit (after Clement 1982). Megaxenoliths
consist of Ventersdorp lava (stippled) and Karoo sediments (black).

and have been truncated by the Fi volcaniclastic unit. F3 is also a marginal earlier
hypabyssal intrusion whose age, relative to F5 and F6, is not known. F2, F4, F7, F7 A,
F7B, and F8 all cut Fl and are internal hypabyssal intrusions. F2 is characterized by the
presence of abundant globular segregations. F4, F7 A, and F7B are all thin curvilinear
internal dikes. The hypabyssal units appear to have been emplaced in the sequence F7,
F2, F4. F2 is younger than F8, but age relations between F7 and F8 are not known. From
these data it appears that an early period of hypabyssal activity, i.e., embryonic pipe
activity, was succeeded by diatreme-producing volcaniclastic activity, and that hypabys-
sal units subsequently intrude the diatreme facies rocks during the waning stages of
activity. Given the increasing abundance of hypabyssal rocks with increasing depth, it
would appear that the deep levels of the pipe will consist largely of a hypabyssal root
zone-dike complex.
The Fi volcaniclastic unit contains many un metamorphosed xenoliths of Karroo age
basalts, dolerites, shales, and sandstones, although no Karroo rocks are preserved in situ
in the vicinity of the pipe. Extremely large xenoliths of basaltic lava, known locally as
floating reefs, are concentrated near the margins of the pipes (Figure 1.5). These are also
encountered in exploratory drill holes at depths of 600 m. Clement (1982) estimated that
lava inclusions at Finsch must have sunk as much as 1000 m within the diatreme.
The Finsch intrusion is rich in diamonds and has an average recovery grade of about
1 metric carat/tonne (Shee etal. 1982). Clement (1982) noted that there is no variation
in the diamond grade of the Fl unit from the surface to 348-m depth. The diamonds
contain inclusions belonging primarily (86%) to the peridotitic suite (Gurney et al.
1979a). Mantle-derived ultramafic inclusions are not common at Finsch. Shee et al.
KIMBERLITES AND ORANGEITES 21

(1982) noted the presence of two diamondiferous peridotites in a suite of 80 xenoliths


found at a single location in a minor body of kimberlite intersected at the 172-184-m
level. The Fl unit is devoid of ultrabasic xenoliths, although high-Cr, Ca-poor knorringitic
garnets occur in the heavy mineral concentrates produced during diamond extraction
(Gurney and Switzer 1973). Megacrystal Ti-pyrope, subcalcic diopside, and Mg-ilmenite
are not present at Finsch (Clement 1982).
Despite the size and economic importance of the Finsch intrusion there have been
few published studies of the geology or petrology of the rocks. The weathered material
occurring in the upper levels of the pipe is described by Ruotsala (1975). Descriptions of
the morphology of the pipe and the various intrusions are given by Clement (1982).
Geochemical and isotopic studies have been undertaken by Smith (1983), Smith et al.
(1985b), Fraser (1987), Fraser and Hawkesworth (1992), and Kirkley et al. (1989). Prior
to this work the only mineral compositional data were those obtained by Fraser (1987)
in a limited study of olivine, spinel, and phlogopite.

1.8.2. Barkly West Region


A large number of dikes and small pipes occur about 100 km northwest of Kimberley
in the district of Barkly West, Cape Province. The intrusions form a NNW-to-SSE-trend-
ing zone from Bellsbank through Sover and Newlands to Pniel (Figure 1.6).

1.8.2.1. Bellsbank
The Bellsbank orangeite field (Figure 1.7) consists of several subparallel dikes,
known locally as fissures, striking N300E and approximately parallel to the escarpment
of the Campbell Rand Plateau. Since their discovery in 1952, mining of portions of the
Main and Bobbejaan Fissures has been undertaken intermittently. The Intermediate and
Water Fissures have lower diamond grades and are infrequently mined. The grade of the
West Fissures is unknown.
The dikes intrude the Campbell Rand dolomite and the underlying Black Reef Group
quartzite and shale of the Griqualand West Supergroup. The orientation of the dikes is
strongly controlled by the regional fracture pattern (Bosch 1971, Tainton 1992). The
Rb-Sr phlogopite age of the dikes is 121 Ma (Smith et al. 1985a).
The Bellsbank Main Fissure is approximately 4.2 km in length and consists of a
discontinuous series of en-echelon thin dikes «1 m), dipping 85° east (Bosch 1971,
Tainton 1992). At the extremities of each segment of the dikes they decrease in width
from a single dike to four or five veinlets before pinching out. Three small "blows" are
developed on the Main Fissure system, the North, Middle, and South Blows, and a fourth
(East Blow) is found 200 m to the east of the Middle Blow on the adjacent Kosmos
Fissure. The North and Middle Blows contain macrocrystal orangeites which do not differ
in petrographic character to the orangeites of the Main Fissure. Xenoliths of country rock
are absent. The North Blow consists of four pipe-like swellings which narrow into a dike
0.5 m in width at a depth of 100 m. The South Blow orangeite contains abundant xenoliths
of granite and schist derived from the underlying basement, in addition to rounded
xenoliths of dolomite, but is in other respects similar to that of the Main dike. Karroo age
xenoliths are not present, and there is no evidence in the xenolith suite to support the
22 CHAPTER!

20·
Kalahari Supergroup

m
r· ... ·,.·'\
20·----~-----'~·~_r--f_t_
Karoo Sequence
')~......,\
1.1 t~ Postmasburg Group

~ Asbesheuwels Subgroup

~ Campbell Rand Subgroup


·•'••..: Ventersdorp CAPE
TOWN
·. . Supergroup oI loookm
I

Figure 1.6. Geology of the Barkly West District and locations of orangeites discussed in the text (after Tainton
1992).

contentions ofTainton (1992) and Bosch (1971) that the South Blow represents a diatreme
that extended to the original surface. All the blows are best regarded as hypabyssal facies
root zones.
The Bobbejaan Fissure dips 85° east, is 2.3 km in length, and has an average width
of 0.5-0.8 m but may be up to 1.2 m in places. Figure 1.8 illustrates the structure ofthe
dike with respect to depth and shows that it consists of a series of en-echelon lenses.
Compared with the Main Fissure, this dike is relatively continuous, and only a few lateral
offsets of up to 15 m are found. Blows are not developed on this dike.
KIMBERLITES AND ORANGEITES 23

C/)

D
w
Ghaap Group a::
::J
C/)

1m
Kalahari ~
IL.
Supergroup
Calcrete and l-
alluvium (/)
w
-----Road 3:

- - Farm Boundary

! 1000 m!

Figure 1.7. Distribution of dikes of orangeite in the Bellsbank field (after Bosch 1971, Tainton 1992).

The Intennediate, Water, and West Fissures are blowfree dikes similar to the Main
and Bobbejaan Fissures. Their surface expression is marked by a discontinuous zone of
calcrete, and little is known of their petrology.
The Main and Bobbejaan dikes are strongly weathered at their surface outcrops and
where the dike is less than 0.6 m in width. Wide variations in color, from brown through
green to gray along strike, reflect different intensity of alteration (Bosch 1971). Fresh
material obtained from deeper levels of the dikes consists of hypabyssal facies orangeite
characterized by numerous large olivine macrocrysts set in a fine-grained black matrix
consisting predominantly of microphenocrystal mica with minor amounts of primary
olivine, spinel, apatite, serpentine, and dolomite. Flow differentiation has commonly
24 CHAPTER 1

No.7 LEVEL

No.6 LEVEL
L- ~ No.5 LEVEL __ B_
No.1 LEVEL
& No.3 LEVEL

U/
&
SHAFT SECTION
LOOKING
A SECTION LOOKING WEST SURFACE B C SOUTH 0
~--~~~~~~~~~~~~
I I
I I
I I
I I
I
I
I /
I /

e
\... ,/E;'::",=.,"=/;;=:= No.1

. . . . .U---:;' NO 2
.
",- ---- ... ......... ,
,,
./' ' I No.3
\
\\ \
\ ,/
.I
"
/ \
\ ,,"'"
...= No.4
.""","

J,-----""~t~~=='=
.I~' I'
/' , I
," / I I
/ / ==="¢!.:=:==== No.5
,,,,,.I "
'.... - ----,
-,,(
I
I

o1..._ _......................501 m

Figure 1.8. Plans and sections illustrating the distribution of en-echelon lenses of orangeite along a portion of
the Bobbejaan dike. Bellsbank field (after Clement et al. 1973).

concentrated the larger macrocrysts at the centers of the dikes. The macrocrysts display
undulose extinction and may have recrystallized to strain-free neoblasts.
Tainton (1992) noted that with increasing alteration the abundance of dolomite in the
matrix of the rock increases relative to that of serpentine. In the Bobbejaan Fissure this
replacement has gone to completion and only dolomite is present in the matrix.
The Intermediate Fissures consist of macrocrystal orangeites of similar character to
those of the Main Fissure, but are poorer in spinel, apatite, and perovskite (Tainton 1992).
The West Fissures are relatively poor in macrocrystal olivine, although complex parallel
aggregates of olivine prisms are present in addition to subhedral phenocrysts. Olivines of
KIMBERLITES AND ORANGEITES 2S

this morphology are common in lamproites (Mitchell and Bergman 1991). The ground-
mass consists of diopside and phlogopite set in a matrix of serpentine. Tainton (1992)
considered that the serpentine matrix was originally sanidine and that the West Fissures
are more evolved than others at Bellsbank and are of "lamproitic affinity."
Mineralogical and geochemical studies of the Bellsbank orangeites may be found in
Bosch (1971), Fesq etal. (1975), Kable etal. (1975), Boctor and Boyd (1982), Smith
et al. (1985b), Kirkley et al. (1989), and Tainton (1992).

1.8.2.2. Sover
Two kimberlite dikes occur on the farms Sover, Doornkloof, and Mitchemanskraal
(Figures 1.6 and 1.9). They are emplaced in the lower Karroo shales, which are underlain
at depths greater than 100m by Ventersdorp lavas. The dikes, labeled A and B by Bosch

- Orangeite Dyke

E;;;';:'I Dolerite} Karoo Sequence


O Shale
-------- Road
- - - Farm Boundary

A
N

Figure 1.9. Distribution of dikes and intrusions in the Sover field (after Bosch 1971. Tainton 1992).
26 CHAPTER 1

(1971), strike NI5°E and N30oE, with lengths of 4 and 6.5 kIn, respectively. They appear
to intersect between the Du Plessis No. 1 and No.2 mine shafts (Figure 1.9). However,
Bosch (1971) claims that two preexisting intersecting fractures were filled with magma
of different diamond content and the eastern dike B on the Sover farm is actually the
northern extension of dike A.
The dikes, which are discontinuous lenses that regularly pinch out along strike, have
an average width of about 0.8 m, reaching 1.5 m in places. Individual lenses may be
slightly offset relative to each other (Figure 1.9), but the dikes do not form a well-defined
en-echelon suite. Chilled margins are absent, and macrocrysts are concentrated by flow
differentiation in the center of the dikes. Tainton (1992) suggests that in places the dikes
are multiple, with the internal contacts being marked by veins of carbonate and serpentine.
Orangeites from this dike system exhibit a wide range in color, reflecting different
degrees of alteration and/or carbonate content of the groundmass. Gray fresh rocks occur
in the centers of the dikes and are isolated from the wall rocks by red-brown altered
material. Country rock shale xenoliths are rare, and the xenolith suite consists primarily
of rare lower crustal rocks and eclogites. Many of the latter are diamond bearing.
The intrusions are composed of macrocrystal hypabyssal orangeite. Microphe-
nocrysts of phlogopite are set in a fine-grained groundmass of mica, spinel, perovskite,
and apatite. The mesostasis consists of serpentine and dolomite.
The Sover North intrusion is an elliptical body (600 m2) located approximately 1.5
km NW of the Sover Mine (Tainton 1992). It intrudes Karroo shales which have been
upturned at the margins of the intrusion. The morphology of the pipe is unknown. The
main intrusion, termed SNI by Tainton (1992), consists principally of hypabyssal
orangeite containing xenoliths of Ventersdorp andesite. A small intrusion consisting of
xenolith-poor orangeite, termed SN2 by Tainton (1992), does not outcrop, lies at a depth
of 30 m within the SNI breccia, and appears to have intruded the latter. All samples of
Sover North material studied in this work are derived from the SN2 intrusion.
SN1 is hypabyssal facies lithic orangeite breccia. Xenoliths have reacted with
magma, the andesites being replaced by phlogopite, amphibole, and clinopyroxene, the
shales by sanidine and phlogopite. The rock consists of an assemblage of minerals
(olivine, sanidine, diopside, perovskite, K-Ba titanates, rare richterite) and exhibits
textures ("dog's tooth" olivines, poikilitic mica) similar to those occurring in madupitic
olivine lamproites. Thus, Tainton (1992) classifies the rock as a "madupitic olivine
lamproite."
The SN2 intrusion has a similar mineralogy but is relatively poor in olivine and rich
in potassian richterite. Tainton (1992) describes SN2 as a "poorly macrocrystal K-
richterite lamproite." Mineralogical differences between SNI and SN2 are considered by
Tainton (1992) to be due to heteromorphism rather than different bulk composition.
Tainton (1992) considers the main Sover dike system and the Sover North intrusion to be
consanguineous. Thus, the Sover North intrusion is considered here to represent a
differentiate of an orangeite magma. The mineralogy of Sover North is illustrated and
discussed further in Section 1.10.
The Sover dike system has been investigated only by Bosch (1971) and Tainton
(1992).
KlMBERLITES AND ORANGEITES 27

1.8.2.3. Newlands
The Newlands orangeites (Figure 1.6) occur as five small blows aligned over 600 m
along a northeast-south west-trending series of fissures. The Rb-Sr age of the intrusion
is 114 Ma (Smith et al. 1985a). The orangeite intrudes calcrete of the Kalahari Super-
group, which forms a veneer over Karroo shales and mudstones which are partially capped
by the remnants of a diabase sill. The blows are numbered 1 to 5 from east to west. Blow
1, also termed the North Blow, consists of up to 90% shale xenoliths. Blow 2 is small and
inadequately characterized due to poor exposure. Blow 3 contains many gneissic base-
ment xenoliths. Blow 4 has a high content of lava xenoliths believed to be derived from
the Karroo Stormberg series. These lavas are no longer preserved in the Karroo sequence
in this region. Blow 5 (South Blow) is rich in xenoliths of shale. The fissure linking the
blows is deeply weathered and replaced by calcrete. The orangeite filling the dike and
forming the matrix of the blows is a macrocrystal hypabyssal orangeite. The petrographic
character of the rock is illustrated in Section 1.10. The Newlands complex may be
regarded as the lower levels of a diatreme root zone that is transitional to a feeder dike
system.
Newlands orangeites contain a wide range of ultramafic and eclogitic xenoliths,
including diamondiferous varieties. Xenocrysts of green knorringitic garnet are common
at Newlands (Clarke and Carswell 1977, Schulze 1989).
Previously published studies of the geology and mineralogy of Newlands are those
of Bonney (1899), Wagner (1914), and Tainton (1992). Isotopic studies have been
undertaken by Smith et al. (1985a), Kirkley et al. (1989), and Tainton (1992).

1.8.2.4. Pniel
The Pniel, also known as Aaron's Prospect, intrusion was emplaced in Karroo shales
(Figure 1.6). The geology of the occurrence is not well characterized as it is not of
economic significance. Tainton (1992) states that it consists of a central shale-rich blow
cut by numerous later dikes and sills. Wagner (1914) describes the occurrence as a
complicated network of weathered dikes and sills which have indurated the shale.
The intrusion consists of macrocrystal diopside richterite orangeite. The presence of
groundmass K-Ti richterite was originally noted by Erlank (1973), and the intrusion was
considered to be a possible lamproite by Mitchell and Bergman (1991). Tainton (1992)
describes the rock as an "amphibole-bearing phlogopite kimberlite," rather than a
lamproite, because sanidine is not present in the rock. The petrography and mineralogy
of the Pniel occurrence are discussed further in Sections 1.10 and 2.8.

1.8.3. BoshofDistrict
Numerous northeasterly trending dikes occur approximately 40 km to the east of the
town of Boshof, Orange Free State (Figure 1.3). Most of the dikes appear to be
uneconomic, although three significant diamond mines (Roberts Victor, New Elands, and
Blaaubosch) have been developed at points where some of the dikes blossom into
root-zone complexes. The dikes in the field are undoubtedly consanguineous as they yield
similar Rb-Sr ages (Smith eta/. 1985a) of 127 ± 5Ma (New Elands), 128 ± 15Ma
(Roberts Victor), and 133 ± 27Ma (Blaaubosch).
28 CHAPFER 1

Unfortunately, very little infonnation is available regarding the geology and petrol-
ogy of the intrusive rocks found in this field, although there have been many detailed
investigations of the eclogite xenoliths, which are particularly abundant at Roberts Victor.

1.8.3.1. Roberts Victor


The Roberts Victor intrusion (Clement et al. 1973) consists of two parallel fissures
(Nos. 1 and 2) whose orientation has been controlled by the regional joint system. Number
1 fissure strikes 028° and dips 76-86° east. Three pipes and three blows have developed
along No. 1 fissure. The fissures are covered by 12 m of Kalahari sand and are intrusive
into Ecca shales and underlying Ventersdorp lavas. Wagner (1914) claims that the No.1
and No.3 pipes, which are separated by a distance of only 50 m at the surface, are different
pipes developed upon the same fissure. Number 1 pipe contains xenoliths of Karroo shale
and lavas and is devoid of xenoliths derived from units higher in the Karroo succession.
In contrast, pipe No.3, in addition to Ecca shales, contains large xenoliths of Beaufort
sandstone which must have been downrafted into the pipe. In contrast, Clement et al.
(1973) contend that only one pipe is present as the intrusions apparently coalesce at depth.
Insufficient petrological data are presented to assess the validity of this claim.
The matrix of both the fissure and the blows is a macrocrystal hypabyssal orangeite.
Unfortunately many of the orangeites found within the intrusion are highly altered.
Consequently, detailed mineralogical or geological studies of these rocks have not been
undertaken. Some data may be found in C.B. Smith et al. (1985b),J.V. Smithet al. (1978),
and Kirkley et al. (1989).

1.8.3.2. New Elands


The New Elands Mine is located on a blow developed where two dikes of orangeite
intersect. Later intrusions of orangeite, relatively richer in mica with respect to the earlier
dikes, cut this blow. The country rocks consist of Ecca shale and the main pipe contains
xenoliths of these and downrafted Beaufort sandstone (Wagner 1914). The dikes consist
of macrocryst-poor hypabyssal orangeite (see 1.10).
Detailed mineralogical studies of the New Elands orangeite have been presented by
Mitchell and Meyer (1989a). Other mineralogical and geochemical data may be found in
J.v. Smith et al. (1978) and C.B. Smith et al. (1985a,b).

1.8.4. Winburg District


Numerous east-west-striking dikes of orangeite occur in the Winburg district (Figure
1.3), but geological descriptions of the majority of these occurrences have not been
published. A major mine, the Star Mine, is located upon a series of parallel dikes near
Theron's Siding, about 15 km north ofTheunissen. Here several dikes (Star, Wynandsfon-
tein, Burns) occur in a zone about 100 m in width and have been traced over 15 km along
an east-west strike. A small pipe, the Phoenix (or Lion Hill) pipe, occurs at the western
end of the Star Mine property on the Wynandsfontein dike. The dikes are typically <1 m
in width and have intruded shales of the Lower Karroo Beaufort Series.
The dikes consist of macrocrystal hybabyssal orangeite (see 1.10) with each one
differing subtly in petrographic character and mode. The only published detailed minera-
KIMBERLITES AND ORANGEITES 29

logical studies are by Mitchell and Meyer (1989a). Other mineralogical and geochemical
information may be found in Wagner (1914) and Smith et al. (1978).

1.8.5. Kroonstad District


Two large pipes, Voorspoed and Lace (Crown), several smaller ones, and many small
dikes and sills (Voorspoed, Rhenosterkop, Besterskraal, Normandiend) of orangeite have
been identified in the Kroonstad district (Figure 1.3). Of the latter, the Besterskraal
occurrence is particularly interesting in that it is a sanidine richterite orangeite (see Section
1.10). The age of the intrusions is on the order of 145 Ma (Skinner 1989). Modern
geological and mineralogical investigations of this region have not been published.

Voorspoed
The Voorspoed Pipe occurs at the intersection of two fracture systems and crosscuts
a northwest-striking antecedent orangeite dike. Clasts of this dike are found within the
pipe, which is itself cut by a northeast-striking subsequent orangeite dike. The pipe and
dikes are emplaced in Ecca Series shales and sandstones (Wagner 1914).
At the current level of erosion over 50% of the surface area of the pipe is occupied
by a very large xenolith (6 ha) of Drakensberg lava. Figure 1.10 illustrates the size of this
"floating reef' relative to the pipe. Because of the size and location of the xenolith, Wagner

30m Level

A B G H
o -

11~~~t\
30 m-

137 m-

','."

340 m-

Figure 1.10. Plan of, and vertical sections through, the Voorspoed diatreme, Kroonstad District, showing the
downrafted megaxenolith (stippled) of Drakensberg lava (after Clement 1982).
30 CHAPl'ER 1

(1914) believed that the Voorspoed orangeite must have intruded a preexisting volcanic
neck. However, recent drilling has demonstrated that the basaltic rocks form a real
xenolith that persists to a depth of 300 m in the pipe (Figure 1.10; Clement 1982).
The orangeite in the pipe is a highly altered, relatively mica-poor volcaniclastic
orangeite breccia. The antecedent dike is a hyabyssal diopside orangeite.

1.8.6. Swartruggens District


Orangeites, in the region of Swartruggens, Transvaal, occur as a series of subparallel,
en-echelon dikes (Figures 1.3 and 1.11) emplaced in lavas of the Pretoria Series. The
principal dike swarm has an overall strike trend of 110° over a distance of about 5 km.
Several small outcrops of orangeite, found along the Elands River to the north of the main
dike system, suggest the existence of other dike swamis in the area.
Within the main dike swarm individual dikes have a sinuous outcrop and may strike
between 100° and 120°. The thickness ranges within short distances, from 2 cm to 2 m.
Many are composite, with individual members being as thin as 0.5 cm. It is impossible
to trace anyone dike as a continuous unit along strike throughout the whole system. Dikes
encountered in the mines commonly do not outcrop. Indi vidual dike segments differ with
respect to their degree of alteration and petrographic character. Thus, the Changehouse
dike is highly altered, whereas the Main dike is relatively fresh. Individual dikes
encountered at different levels in the mines are petrographically different as a conse-
quence of the complexity resulting from multiple intrusion. Hence, each dike is not a
single petrographic unit formed by the crystallization of a single batch of magma. The
dike system appears to have a character similar to that of Bellsbank and is thus considered
to be a discontinuous series of en-echelon lensoid bodies.
Individual dike segments have been named primarily after their discoverers with
some apparently renamed by various mining companies. Consequently, dike nomencIa-

NOOITGEDACHT 405 A
N

o
I
I
I
km
- 5

Figure 1.11. Distribution of orangeite dikes in the vicinity of Swartruggens. Transvaal (after Fourie 1958).
KIMBERLITES AND ORANGEITES 31

ture is confusing. All Swartruggens samples studied in this work are derived from
different mining levels intersecting the Main dike occurring in the Helam Mine.
The Swartruggens dikes vary with respect to their diamond content and petrographic
character. All diamond-bearing dikes are classified here as macrocryst-poor hypabyssal
orangeites. Other diamond-free dikes are olivine-bearing minettes. The principal occur-
rence ofthe latter is known as the Male lamprophyre (N.B. "Male" [mah-lay] in Afrikaans
means barren). This rock consists of serpentine pseudomorphs after oli vine set in a matrix
of pyroxene, mica, and potassium feldspar. The compositions of the pyroxenes (2.2.3)
are not significantly different to those of pyroxenes in the Swartruggens orangeites.
However, the micas have different compositions and evolutionary trends from those of
orangeite mica, suggesting that these lamprophyres have no simple genetic relationship
to orangeites.
It was recognized in earlier studies (Fourie 1958) that Swartruggens orangeites differ
from those in other districts in that they contain considerable amounts of quartz. Interest-
ingly, quartz veins occur parallel to the strike of the dike swarm. Fourie (1958) states that
these veins are post-dike in age, but this conclusion does not rule out their being formed
from hydrothermal residual fluids. Alternatively, the quartz may be simply of secondary
origin.
Age determinations have not established the relationship between the orangeite and
lamprophyre dikes. Allsopp and Barrett (1975) give a Rb-Sr mica isochron age of 147 ±
4 Ma for micas from the Main dike. However, this age is unreliable as it is essentially a
two-point isochron. These data, together with others for micas from the Main dike, were
used by Smith et al. (1985a) to construct a Rb-Sr errorchron that gave an age of 156 ±
13 Ma. This errorchron included one sample from the Male dike. As the datum lay within
the limits of the errorchron, it may be concluded that the Male lamprophyre dike may be
of similar age to the orangeites. However, the extremely different mineralogy and
petrographic character of the two dikes suggest that they are not consanguineous. It is
suggested here that they are merely contemporaneous and have exploited the same
fracture system during emplacement.
Previous studies of the geology of the Swartruggens dikes are by Haughton (1935)
and Fourie (1958). Mineralogical and geochemical data are given by Mitchell and
Brunfelt (1975), Allsopp and Barrett (1975), J.Y. Smith et al. (1978), C.B. Smith et at.
(1985a,b), Skinner and Scott (1979), and Kirkley et at. (1989).

1.8.7. Dokolwayo
The Dokolwayo diatreme, Swaziland (Figure 1.3), considered to be a "group II
kimberlite" by Skinner (1989), occurs as a single intrusion emplaced in late Archean
granite-gneiss country rocks. It lies outside the present limits of the Karroo basin but
contains downrafted xenoliths of Karroo age. No accurate radiometric age determinations
for the intrusion are available. Skinner (1989) gives an age of 200 Ma (method unspeci-
fied), and Hawthorne et al. (1979) suggest an age of 190 Ma, based upon field relations
and xenolith provenance. No other diamondiferous rocks, apart from a few minor dikes,
occur in the area (Hawthorne et al. 1979).
32 CHAPTERl

Hawthorne et al. (1979) have published a general description of the Dokolwayo


diatreme and the nearby Hlane alluvial diamond deposit. The diatreme is an elongate
structure (600 m x 50--100 m) with a northeasterly strike. The area increases from 2.8 ha
at the surface to 3.8 ha at a depth of 120 m, narrowing again at greater depths. The
upper-level bulge is capped on the northern flank by breccias lacking an igneous matrix,
which are interpreted (this work) to be contact breccias (Clement 1982). The central parts,
and the bulk, of the intrusion consist of several varieties of diatreme facies volcaniclastic
heterolithic breccias (this work). Some units with extremely high contents of xenocrystal
quartz are described as crystallinoclastic breccias (Clement 1982). Xenoliths range in size
from> 1 m to microscopic clasts and consist primarily of sandstones and coals of middle
Ecca (Lower Permian) age together with locally derived Archean granite-gneisses.
Xenoliths of crater facies resedimented volcaniclastic (epiclastic) material are present in
one of the diatreme-facies volcaniclastic units (this work). Pelletal-textured rocks are
absent, and the intrusion is considered here to have been eroded to depths at which the
diatreme proper enters the root zone level.
Altered hypabyssal rocks are found at the margins of the intrusion. Their textures
have been described as "porphyritic" by Hawthorne et al. (1979). They consist of altered
subhedral olivines set in a groundmass of spinel, phlogopite, serpentine, and calcite.
Groundmass micas occur as fine-grained «0.1-0.3 mm), irregular partly-altered plates
and laths. Euhedral grains are present locally, and much of the mica has been altered to
serpentine. Because of the high modal amounts (30 vol %) of groundmass mica, Haw-
thorne et al. (1979) term the rock a "phlogopite kimberlite" (sic), using the mineralogical
classification of Skinner and Clement (1979).
Examination of the altered hypabyssal facies rock during the preparation of this work
showed that the groundmass micas are colorless phlogopites which are commonly
intergrown in a reticulate pattern of interlocking laths. The texture is reminiscent of that
observed in many archetypal kimberlites (Figure 1.74). Detailed mineralogical studies of
the rocks have not been undertaken; however, on a petrographic basis, there are no
grounds for considering them to be orangeites, as the micas do not have the textural or
optical characteristics of orangeite micas. Thus, the petrological status of the Dokolwayo
intrusion requires reexamination.

1.S.S. Prieska District


The geologically complex southwestern margin of the Kaapvaal craton in the
neighborhood of Prieska (Figure 1.3) has 130 known intrusions of kimberlite and
orangeite (Skinner et al. 1994). The region represents an area where a geographical
association of temporally distinct kimberlite and orangeite magmatism can be found.
Skinner et al. (1994) have divided the region into five domains (Figure 1.12) on the
basis of the disposition of the intrusions, relative to the craton margin, major faults, and
petrographic and isotopic character of the rocks. Each domain is characterized by
intrusions of predominantly one petrological and isotopic type. Thus, intrusions in
domains I and III are predominantly "group 2 kimberlites" (Le., orangeites), whereas
those in domains II and IV are "group 1 kimberlites" (Le., archetypal kimberlites).
Intrusions in domain V have mineralogical and isotopic characteristics considered by
KlMBERLITES AND ORANGEITES 33

KIO(l14)
~26
KII

DO
.K32 DOMAIN
A
N
O~ .. K5(126)
i1t(9, ••
~ ~t9G' , •
\ ~~ K27 • • Prieska
, ~ DOMAIN III
" ij\ 1": +6'>.......... K25
o 6/" ~ I/~ . , -300

• • //.9 ~ .........
K56K20• ~ ~! •
-...K24 ..... ? .I'K46
\.. • . . . ~K52 K25 .. aoK45
KI5.' , . t•
K41 0" 0 K42 \
(127),. KI9 (liS) . ' K40/ ..
••• K9, • .1 (114) ..
KI30 :.. K64 • • I 0 ...... DOMAIN II
• .. 1 oo~..;,;.;.:...:..;.;,.;.....:.:..
K690 K66~K6 ,~" t ....
K2(IOI) ;-'O--'.. K65 K35(74)
o
KI (71) .........
K22 (S2)
Q, 0 .........
El it! 0 " 0 DOMAIN IV , - 31 0
K25 0 0 , *KI6(I03) .0 ..........
KI5
KI(74)~ 0 o Kimberlite
o 0K5 0 • Orangeite
' / - - - - - - - - - - K23
. *KII K9**. " " ? .. Unclassified
K19* DOMAIN V .
* Isotopically
Transitional
Major fault
o
!
50km
! ~
..
",
Inferred craton
boundary
, ~,,-,' Domain boundaries

Figure 1.12. Tectonic domains and the distribution of orangeites and kimberlites in the Prieska District, Cape
Province (after Skinner et al. 1994).

Skinner et al. (1994) to be transitional between "group 1 and 2 kimberlites" (see below
and 3.8.1). Figure 1.12 demonstrates that the orangeite intrusions occur primarily in an
on-craton tectonic setting, whereas the kimberlites straddle the craton margin.
The intrusions of domains I and III are predominantly hypabyssal dikes which are
commonly extensively altered. Skinner et al. (1994) have described them by optical
petrographic studies, but detailed mineralogical investigation by electron microbeam
34 CHAPTERl

methods has not yet been undertaken. Only a few of the intrusions are petrographically
similar to bona fide on-craton orangeites such as occur at Barkly West, the olivine
macrocryst-rich intrusions of Sandrift (25/K5) and Omdraaisvlei (26/K25) being the best
examples.
Many of the rocks contain groundmass spinels and perovskites which are relatively
coarse-grained compared to those occurring in other on-craton orangeites. Many also
contain significant amounts ofxenocrystal ilmenite (Albertshop 25K125 , Markt 26/K15,
Welgevonden 26/K67, Silvery Home 27/Kl1), chrome-poor garnet macrocrysts (Middel-
water-Ol 26/K19, Wielpan 26/K52, Nietgedacht 261K65), Ti02-rich garnet macrocrysts
(Middelwater-Ol 26/K19, Kafferskolk 26/K20, Melton Wold 27/K9), and Cr-poor cli-
nopyroxene macrocrysts (Wielpan South 26/K60, Silvery Home 271K 11). Although the
presence of these macrocrysts is atypical of orangeites, Skinner et al. (1994) nevertheless
consider their host intrusions to be petrographically and isotopically "group 2 kimber-
lites," i.e., orangeites.
Skinner etat. (1994) noted that many of the rocks exhibit petrographic features
similar to those observed in olivine lamproites. These include the common late crystal-
lization of amphibole and sanidine, and the presence of madupitic or oikocrystal
coarse-grained, orange-brown, strongly pleochroic ground mass phlogopite and
glomeroporphyritic aggregates of olivine. Rocks exhibiting such features are termed
"evolved group 2 kimberlite" by Skinner et at. (1994), who also note the similarity of
these rocks to amphibole- and sanidine-bearing orangeites in the Barkly West and
Kroonstad Districts. Note that the Prieska District rocks could be termed "lamproites"
using the terminology of Tainton (1992) and Tainton and Browning (1991). Describing
the rocks as "orangeites" eliminates the previous ambiguities in nomenclature (see also
1.5).
Rocks occurring in domain V contain unusually coarse-grained and abundant cli-
nopyroxene and perovskite in addition to having unusual isotopic signatures (3.8.1). They
are petrographically unlike typical on-craton orangeites, although they contain abundant
groundmass mica. The presence of primary amphibole and possible sanidine is considered
by Skinner et al. (1994) as sufficient grounds to justify their designation as varieties of
"group 2 kimberlite" (orangeite).
Archetypal kimberlites occurring in domains II and IV are no different in their
mineralogy, macrocryst suites, and isotopic character from kimberlites occurring else-
where in southern Africa in on- and off-craton tectonic settings.
Skinner et at. (1994) and Smith et at. (1994) have shown that the orangeites of
domains I and II range in age from 114 to 127 Ma and are thus apparently contempora-
neous with the main body of orangeite magmatism in the central parts of the craton.
Intrusions in domain V are significantly older (136-167 Ma), although geochronological
data are less precise. The kimberlites are all younger than 114 Ma (Skinner et at. 1994).
Although some (Witberg 251KI0, domain I) are slightly older than the main period of
contemporaneous kimberlitic on-craton magmatism (90 Ma) in southern Africa, the
majority appear to have relatively young «80 Ma) ages, similar to those of other
off-craton kimberlites emplaced in mobile belts.
In this work, the orangeites emplaced in domains I, II, and III, together with the
isolated intrusion (Mount Pierre 27/K15) in domain IV, are considered to compose a single
KIMBERLITES AND ORANGElTES 35

consanguineous field. The intrusions in this field were derived from magmas which were
apparently more evolved than those in the Barkly West and Boshof Districts. Further it
is suggested that intrusions occurring in domain V originated during a distinctly different
and older magmatic event and were not derived from orangeite parental magmas. They
may represent archetypal kimberlite magmas which have been contaminated by melts
derived from ancient enriched lithospheric mantle.

1.8.9. Summary
It is concluded from the above descriptions of several fields of orangeite magmatism
that

1. Orangeites, in common with kimberlites and melilitoids, form bonafide diatre-


mes. They do not form the wide, shallow vents characteristic of lamproitic
magmatism. Diatreme formation is preceded by the crystallization of hypabyssal
rocks which are subsequently incorporated into the volaniclastic rocks of the
diatreme. Magma intruded into the lower parts of the diatreme subsequently
crystallizes as a second generation of late-stage hypabyssal rocks.
2. Orangeite magmas ascend to high crustal levels as dikes utilizing preexisting
fracture systems. Diatremes and blows of hypabyssal orangeite are developed
above dike systems.
3. Many orangeites occur as en-echelon dike swarms, a style of intrusion not
typical of kimberlites but characteristic of many suites of lamprophyres (Currie
and Ferguson 1970). Composite intrusions are common.
4. Orangeite magmas differentiate to residual magmas which crystallize amphi-
bole, sanidine, and possibly leucite (see 1.10).
5. Individual fields of orangeite differ with respect to their mineralogy and degree
of evolution. Thus, the Barkly West field is characterized by olivine macrocrystal
orangeites, whereas other fields are relatively poor in olivine macrocrysts. The
Kroonstad and Prieska fields are relatively evolved compared to the Barkly
West, Boshof, and Winburg fields. The Swartruggens orangeites are atypical in
containing quartz and many zirconium-based minerals (see 2.12).
6. Lavas and plutonic rocks belonging to the orangeite clan have not been recog-
nized. Clasts consisting of pyroxene and phlogopite or phlogopite alone have
been recognized in several orangeites (Skinner 1989, this work). Dawson and
Smith (1977) noted that xenoliths consisting of mica, amphibole, and diopside
occur at Newlands but do not appear to be MARID xenoliths (Skinner 1989).
They may represent disrupted cumulates formed in the magma prior to intrusion.

1.9. TEXTURAL-GENETIC CLASSIFICATIONS OF PETROLOGICAL


CLANS
The extensive mineralogical and textural variations exhibited by kimberlites have
given rise to a plethora of textural classifications (see Wagner 1914, Kovalskii 1963,
Rabkin et al. 1961, Artsybasheva et al. 1964, Frantsesson 1968, 1970, Kornilova et al.
1983, Clement and Skinner 1985, Vladimirov et al. 1981, 1990). These classifications
36 CHAPTER 1

VOLCANICLASTICS

Figure 1.13. Model of an idealized kimberlite magmatic system. illustrating the relationships between crater
diatreme and hypabyssal facies rocks (not to scale). Hypabyssal rocks include sills. dikes. root zone. and "blow"
(after Mitchell 1986).

are commonly mutually exclusive, making comparison of kimberlites derived from


different provinces difficult or impossible. Russian petrologists in particular have taken
the erection of independent classifications to an extreme, with some groups devising
different classification schemes for the same kimberlites.
Recognizing these problems, Clement and Skinner (1979, 1985) and Clement (1982)
attempted to bring order to the chaos by devising a textural-genetic classification scheme
based upon the concept that different kimberlite facies may be recognized. Kimberlites
are considered to be a clan of rocks which formed from a volatile-rich magma. This
magma, in common with others, undergoes differentiation and crystallization in diverse
temperature and pressure regimes, giving rise to a spectrum oftexturally and petrographi-
cally different rocks. Classification is based on the premise that macroscopic and
microscopic characteristics observed in textural variants of kimberlites are mainly related
to different near-surface emplacement and crystallization processes. Thus, Clement and
KIMBERLITES AND ORANGEITES 37

Skinner (1985) recognized three textural genetic groups ofkimberlite associated with a
particular style of magmatic activity, namely crater, diatreme, and hypabyssal facies
rocks.
In principle, the facies approach to classification of Cas and Wright (1987) and
Clement and Skinner (1985) is applicable to any other magma type. Thus, Mitchell and
Bergman (1991) have devised a similar textural-genetic classification for the lamproite
clan which recognizes lava, crater and pyroclastic, hypabyssal, and plutonic facies.
Analogous schemes could be devised for the members of the melilitoid clan, ultrapotassic
rocks of the kamafugite series and some minettes.
The revised textural-genetic classification of kimberlites presented below is a
development of the Clement and Skinner (1985) model that is also applicable to
orangeites and other diatreme-forming alkaline ultrabasic igneous rocks.

1.9.1. Kimberlites
Figure 1.13 illustrates an idealized model of a kimberlitic magmatic system (Mitchell
1986) based on observations of southern African kimberlites (Mannard 1962, Dawson
1971, Hawthorne 1975, Clement 1982), showing the spatial relationships between the
different facies.

1.9.1.1. Crater Facies


Crater facies rocks are divided into lavas, pyroclastic rocks, and resedimented
volcaniclastic rocks (Figure 1.14; Mitchell 1986). The latter have previously been termed
"epiclastic kimberlites" (see below).

1.9.1.1.a. Lavas. Kimberlite lavas have not yet been conclusively identified. The
only potential candidates are a series of small lava flows comprising the Igwisi Hills,
Tanzania (Dawson 1994). Lavas may be described by standard terminology.

KIMBERLITE LAVA STANDARD TEXTURAL DESCRIPTIONS

STANDARD PARTICLE SIZE AND


PYROCLASTIC KIMBERLITE
TEXTURAL TERMINOLOGY

VOLCANICLASTIC
ROCKS
RESEOIMENTED STANDARD PARTICLE
VOLCANICLASTIC SIZE AND TEXTURAL
KIMBERLITE TERMINOLOGY

PSEUDO-CRATER FACIES ROCKS CONSISTING OF TERRUGINOUS


METACHRONOUS COMPONENTS AND REWORKED VOLCANICLASTIC MATERIAL
VOLCANOGENIC FILLING EROSIONAL DEPRESSIDNS AND DEPOSITED UPON ANCIENT
SEDIMENTARY DEPOSITS VOLCANIC DIATREMES OR THE HYPABYSSAL ROOT ZONES OF
OIATREMES ARE NOT CRATER FACIES DEPOSITS.

Figure 1.14. Textural genetic classification of crater facies kimberlites.


38 CHAPTERl

1.9.1.I.b. Pyroclastic Rocks. In situ examples of pyroclastic kimberlites are rare


because the craters of most kimberlitic volcanoes have been destroyed by erosion.
Unfortunately, there are no published modern descriptions of the pyroclastic kimberlites
known to occur in Tanzania and Botswana (Mannard 1962, Hawthorne 1975). It is to be
expected that subaerial and subaqueous tephra deposits, together with resedimented
volcaniclastic deposits (see below), will eventually be described from Botswana crater
facies kimberlites.
Standard pyroclastic terminology (Fisher and Schminke 1984, Cas and Wright 1987)
may be applied to pyroclastic kimberlites. Following Fisher and Schmincke (1984),
pyroclastic fragments are considered to be produced by processes associated with
volcanic eruptions. They represent particles expelled through volcanic vents without
reference to the causes of eruption or origin of the particles. Pyroclastic deposits may be
interpreted in terms of standard volcanic facies models (Cas and Wright 1987).
Crater facies may also contain hydroclastic rocks, such as base surge deposits,
resulting from phreatomagmatic activity. These are most likely to be found outside the
crater, as base surges within craters are primarily erosional rather than depositional.
Hydroclastic is used, following Fisher and Schminke (1984), to describe the products of
explosions due to steam derived from water of any kind. Hydroclastic crater facies rocks
are different in character and origin to hydroclastic diatreme facies rocks (see below).

1.9.1.I.c. Resedimented Volcaniclastic (Epiclastic) Rocks. Current usage of the


term "epiclastic" in kimberlite petrology (and elsewhere) is unsatisfactory. The term has
traditionally been used by sedimentologists to describe igneous material weathered and
eroded from older source rocks. Thus, epiclastic rocks are commonly considered to be
sedimentary units with clasts of lithified pyroclastic or volcanic rock. Recently, vol-
canologists have extended the term to include material derived from contemporaneous
and unconsolidated sources with or without lithic (terrigenous) components. Kimberlite
petrologists have typically used the term "epiclastic" in this context rather that in the older
sedimentological sense. Unfortunately, there is no agreement among volcanologists as to
a definition of "epiclastic" (see Fisher and Schminke 1984, p. 89, Cas and Wright 1987,
pp. 8-9, Cas 1991, Allen 1991, McPhie et al. 1993). A significant further problem with
the existing terminology is that it gives no indication as to when a particular "epiclastic"
deposit was formed relative to the age of the juvenile clasts.
Recently, McPhie et al. (1993) have proposed the elimination of the term as a
consensus has not yet been agreed upon regarding a redefinition. Thus, "epiclastic rocks"
derived from contemporaneous pyroclastic rocks are renamed "resedimented volcaniclas-
tic deposits." The resedimentation of juvenile pyroclastics (and effusives) may take place
by mass flow, traction, or suspension.
Discussion of the merits of the nomenclature of McPhie et al. (1993) versus the
traditional epiclastic nomenclature are beyond the scope of this work, but it is suggested
here that epiclastic kimberlites may be better described using the terminology of McPhie
et al. (1993), as this gives a clearer indication of their origin. Hence, in this work crater
facies "epiclastic" rocks are referred to as resedimented volcaniclastic kimberlites (sensu
McPhie et al. 1993). However, many readers of this monograph may prefer to continue
using the older terminology until volcanologists agree upon the terminology ofthese types
KIMBERLITES AND ORANGEITES 39

of rocks. Consequently, the term "epiclastic" is also given in parentheses wherever the
new terminology is used, to provide continuity with existing literature.
Resedimented volcaniclastic deposits are named by McPhie et at. (1993) as resedi-
mented ash-rich mudstone, resedimented ash-rich sandstone, etc., following the standard
particle size terminology employed in volcanology (Fisher and Schminke 1984). In the
case of kimberlites they may be termed "resedimented kimberlitic ash-rich sandstone,"
etc. However, the particle size terminology was adopted from that devised for clastic
sediments (Wentworth 1922). The use of a name which has both grain size and compo-
sitional implications causes many volcanologists to be resistant to terming a resedimented
pyroclastic deposit which does not contain significant quantities of detrital lithic material
a "sandstone." Conversely, sedimentologists are loathe to term such deposits exhibiting
sedimentary structures a "tuff." Unfortunately, there is as yet no agreed upon solution to
this problem (see McPhie et al. 1993, Fritz 1993, Zuffa 1991).
Resedimented volcaniclastic rocks currently described as "epiclastic kimberlites" are
commonly preserved in the slightly eroded kimberlites of Tanzania and Botswana
(Mannard 1962, Hawthorne 1975). Xenoliths of these rocks are also found as downrafted
blocks in the diatremes (Clement 1982). Despite their relatively common occurrence,
their character and genesis are poorly understood as they have not been investigated by
modem petrological techniques and are extensively altered. The majority of the crater
facies resedimented volcaniclastic (epiclastic) kimberlitic rocks which have been studied
to date contain juvenile and extraneous lithic clasts in addition to authigenic components.
They have been usually interpreted to have formed in crater lakes formed above diatremes
(Mannard 1962, Hawthorne 1975, Mitchell 1986).
Two principal varieties of resedimented volcaniclastic (epiclastic) deposit may be
recognized: syn-eruptive and post-eruptive. The former are produced during periods of
quiescence during the lifetime of the vent and may be interbedded with pyroclastic rocks
(see below). Post-eruptive resedimented volcaniclastics (epiclastics) are formed imme-
diately after eruptions have ceased and represent the products of destruction of the
volcanic edifice by contemporaneous erosion. They may occur within the crater as crater
lakes become filled with resedimented pyroclastic material. Infilling may occur soon after
termination of activity at a time when the tuff ring and crater walls still exist. Such
resedimented volcaniclastics (epiclastics) may contain a high proportion of juvenile
clasts. Other post-eruptive deposits may form contemporaneously outside the crater (see
below).
When volcanism is episodic, resedimented volcaniclastic (epiclastic) rocks may be
interbedded with subaerial pyroclastic rocks. Fragmentation of these rocks by later
volcanic events will result in the recycling of lithic and juvenile debris into newly formed
pyroclastic rocks (Houghton and Smith 1993). Reworking of the latter will generate new
resedimented volcaniclastic (epiclastic) rocks. The proportions of resedimented volcani-
clastic (epiclastic) to pyroclastic rocks will vary as the vent (or vents) evolves. Identifi-
cation of a given unit as pyroclastic or resedimented volcaniclastic (epiclastic) becomes
extremely difficult when components have been recycled several times. Note that
determination of the origin of crater facies units from drill-core samples is particularly
difficult as the lithological relationships between units cannot be observed.
40 CHAPTER!

Most discussions of crater facies rocks, in the kimberlitic environment have been
restricted to those occurring within the crater. This is a direct consequence of the absence
of modern kimberlitic volcanism and the generally extensive erosion of kimberlite
intrusions. Base surge deposits, pyroclastic rocks, and resedimented volcaniclastic (epi-
clastic) rocks, such as debris flows, lahars, and lacustrine deposits, formed outside craters
have not been previously considered as they are not typically preserved. However, such
extracraier rocks may comprise portions of kimberlite intrusions at Fort a la Corne,
Saskatchewan (Lehnert-Thiel et al. 1992), and Bakwanga, Zaire (Bardet 1974). Unfor-
tunately detailed modern mineralogical and petrological studies of these rocks are not
available.
Lacustrine and other extra-crater resedimented volcaniclastic (epiclastic) rocks can
be classified and described according to standard volcano-sedimentological terms and
facies analyses as described by Cas and Wright (1987).

1.9.1.1.d. Volcanogenic Sedimentary Deposits. In some cases, sedimentary rocks of


similar appearance to syn- and post-eruptive resedimented volcaniclastic (epiclastic)
rocks may represent material washed into depressions formed above vents or diatremes,
long after volcanism has ceased. In these examples the original volcanic edifice has been
removed by erosion prior to the formation of the sedimentary unit, and thus a crater lake
is not present. The depressions may be formed either by subsidence of the material in the
vent or differential erosion. Examples oflakes occupying such depressions are the "pans"
and post-glacial lakes formed above kimberlites in southern Africa and Canada, respec-
tively.
Material washed into the depressions may be devoid of juvenile components, being
entirely derived from the terrain outside the depression. Such deposits should not be
termed either "crater facies rocks" or "resedimented volcaniclastic (epiclastic) rocks," as
they are merely normal lacustrine sediments.
In other cases, rocks containing a juvenile component may occupy the depressions.
These may be formed by the reworking of preexisting igneous rocks coupled with input
of extraneous lithic material. The time of deposition of these materials may be millions
of years after the formation of the vents. Erosion of the vent to hypabyssal or lower root-
zone facies may have occurred and the sediments deposited directly upon these rocks.
The presence in the rocks of igneous clasts qualifies them as an "epiclastic" deposit in
the original sedimentological sense of the term, in that the juvenile material is derived
from lithified rocks. However, they cannot be regarded as crater facies rocks as they did
not accumulate in, or adjacent to, an active volcanic crater and are not directly related to
the volcanic activity. In this work such pseudo crater facies rocks are termed metachro-
nous volcanogenic sediments (epiclastics). Metachronous (Greek meta = after +
khronos =time) is used to indicate that they formed a long time after the magmatic event.
The rocks are termed "volcanogenic" rather than "resedimented volcaniclastic" to indi-
cate that the juvenile component was derived from ancient lithified igneous rock (McPhie
et al. 1993).
It is very important to distinguish these rocks from bona fide synchronous (syn- and
post-eruptive) crater facies rocks of similar appearance. Failure to do this can lead to
incorrect interpretations of the evolution of kimberlite intrusions (see 1.9.1.4).
KIMBERLITES AND ORANGElTES 41

Pseudo crater facies volcanogenic sedimentary (epiclastic) rocks are known from the
Kelsey Lake kimberlite, Colorado (Coopersmith 1993), several Yakutian kimberlites (see
1.9.1.4), and diatremes in the Arkhangelsk diamond province. In the Kelsey Lake
occurrence, Recent metachronous volcanogenic sedimentary (epiclastic) rocks form at
the current erosional surface above Devonian kimberlites which contain downrafted bona
fide resedimented volcaniclastic (epiclastic) xenoliths. This is an excellent example of
the deficiencies of the older terminology, as the same name is applied to genetically and
temporally different rocks.
Topographic depressions in the upper levels of diatremes of the Arkhangelsk Prov-
ince (Kudrjavtseva et al. 1991, Kharkiv 1992, Sinitsyn et al. 1992) are filled with
lacustrine breccias, the components of which are commonly derived entirely from the
Vendian country rocks. Petrographic studies reveal a few examples containing a very
minor juvenile igneous component consisting of altered phlogopite and olivine macro-
crysts or very thin layers of tuff. Russian petrologists term these metachronous volcano-
genic sedimentary (epiclastic) rocks "compensational sediments" and interpret them as
crater facies rocks, even though they recognize that they were formed subsequent to the
cessation of volcanic activity.
Brief general descriptions of crater facies kimberlites have been provided by Man-
nard (1962), Edwards and Howkins (1966), Rolf (1973), Hawthorne (1975), Clement
(1982), and Clement and Skinner 1985). A review of the geology of crater facies
kimberlites has been provided by Mitchell (1986).

1.9.1.2. Diatreme Facies


To reflect the association of diatremes with a variety of rocks, including kimberlitic,
minette, and melilititic volcanism, Mitchell (1986, p. 74) defined diatremes as
cone shaped. downward tapering. inclined or vertical structural units. composed wholly or partly
of angular. or rounded. clasts of cognate or xenolithic origin. with or without a matrix. Xenolithic
clasts may be derived from the walls or roof of the body. They are commonly well-mixed, and
some xenoliths have apparently sunk within the diatreme. Diatremes are volcanic features
associated with volatile-rich magmatism commonly of an ultrabasic composition.
Note that the term "diatreme" is not synonymous with "volcanic vent."
Kimberlite diatremes are long carrot-shaped bodies underlying crater facies rocks
(Figure 1.13). They have near-vertical axes and steeply dipping margins (80-85°). Their
elliptical or roughly circular cross-sectional area decreases regularly with depth until they
terminate in a region known as the root zone. This zone (Figure 1.13) is marked by the
expansion, contraction, or splitting up of the diatreme into an irregularly shaped multi-
phase intrusion of hypabyssal kimberlite. Detailed descriptions of kimberlite diatremes
are provided by Hawthorne (1975), Clement (1982), and Mitchell (1986). Note that
kimberlitic diatremes in the Yakutian kimberlite province might not conform in structure
to the South African model (see 1.9.1.4) as diatreme morphology may be strongly
influenced by the fluid content, type, and rigidity of the country rock.
Diatreme facies kimberlites are currently classified as "tuffisitic kimberlite" and
"tuffisitic kimberlite breccia" (Clement and Skinner 1985, Clement 1982). The latter
differ from the former only in that they contain more than 15 vol % clasts greater than 4
mm in maximum dimension. The rocks consist of lithic fragments and discrete mineral
42 CHAPTER 1

grains of diverse origin cemented by extremely fine-grained minerals which are inter-
preted by Clement (1982) as quench products derived from the condensates of vapor
phases in short-lived, vapor-liquid-solid fluidized systems. Clement (1982) thus regards
the rocks as intrusive tuffs or tuffisites. Use of the term "tuffisite" (Cloos 1941) has
definite genetic connotations and implies formation of a diatreme by gas-solid fluidiza-
tion processes, i.e., tuffisitization. Clement and Skinner's (1985) terminology is correctly
applied if all diatreme facies rocks form by such a process. However, there is no agreement
on how diatreme facies rocks are formed. Currently, opinions cover the entire spectrum
between hydrovolcanic and fluidization origins (see reviews by Clement 1982, Mitchell
1986, Clement and Reid 1989). Thus, while the use of genetic-specific terms is appropri-
ate for a textural-genetic classification, they should not be applied indiscriminantly. A
specific genetic term should only be used when the origin of a particular diatreme facies
rock is understood.
As the nature of the volcanic processes giving rise to most diatreme facies rocks is
not known, they are better termed volcaniclastic kimberlites. The general nongenetic term
volcaniclastic was introduced by Fisher (1961) to include alI clastic volcanic materials
formed by any process of fragmentation, dispersed in any transporting agent, and
deposited in any kind of environment.
Thus, it is suggested that Clement and Skinner's (1985) genetic term "tuffisitic" be
abandoned as a general term for diatreme facies kimberlites and they be described as
volcaniclastic kimber/ites and volcaniclastic kimberlite breccias. Only if the origins of
the clasts can be determined should a second level of classification be employed, e.g.,
tuffisitic kimberlite, hydroclastic kimberlite. Figure 1.15 presents the recommended
nomenclature of diatreme facies kimberlites. Note that this nomenclature is applicable to
orangeites and melilitoids but not to lamproites. Other textural terms which have been
inadequately defined in previous works include pelIetal lapilli, autoliths, and nucleated
autoliths.

DIATREME FACIES

VOLCANICLASTIC KIMBERLITE BRECCIA


(15 %; )4 mm Clasts ) 15 %; )4 mm Clasts

) 1 0 % PeHetal Lapilli

No

Ve.

PELLETAL-TEXTURED

Figure 1.15. Textural-genetic classification of diatreme facies kimberlites_


KIMBERLITES AND ORANGEITES 43

Pelletal lapilli (Clement 1973, 1982, Clement and Skinner 1985) are particularly
characteristic of diatreme rocks and may be regarded as the "hallmark" for the recognition
of diatreme environments. Especially important is that pelletallapilli are not confined to
kimberlites and may be found in melilitite (Cloos 1941, Rust 1937, Reed and Sinclair
1991) and orangeite diatremes, e.g., Finsch (Clement 1982). Pelletal lapilli are thus
characteristic products of a particular process operating during the formation of diatremes
but are not indicators of a particular magma type. However, all rocks containing pelletal
lapilli appear to be derived from magmas that are C02"fich relative to common magma
types.
Pelletal lapilli are discrete spherical-to-elliptical lapilli-sized (2-64 mm) clasts
consisting of fine-grained primary igneous material. Lapilli commonly contain, at their
centers, a single, relatively large euhedral crystal or crystal fragment (Figure 1.16). These
cores, or kernels (Clement 1973), consist typically of olivine (fresh or pseudomorphed)
and, less commonly, phlogopite or other macrocrysts. Country rock clasts very rarely
form the cores oflapilli, and juxtaposed country rock clasts are typically devoid of igneous
mantles.
The mantles consist of very fine-grained microphenocrystal material. Minerals found
within the mantle are primary groundmass and mesostasis phases characteristic of the
parental magma. Thus, in kimberlites the mantle consists of microphenocrysts of euhedral
olivine and/or laths of phlogopite set in a groundmass of perovskite, spinel, calcite, and
serpentine. Prismatic minerals are commonly oriented around the kernel of the lapilli and
a poorly-to-well-developed concentric structure may be discernible (Figure 1.16).
Pelletallapilli within a given sample exhibit a wide range of textural charaCteristics
and degree of crystallinity. They have the textural characteristics of having been formed
by the rapid crystallization of volatile-poor magma containing phenocrysts and may be
described as microphenocrystal hypabyssal kimberlite. They are not, on the basis of their
mineralogy and texture, examples of accretionary lapilli.
The origins of pelletallapilli have been reviewed by Clement (1982), Lorenz (1979),
and Mitchell (1986). Currently, they are believed to represent magma droplets formed by
the explosive fragmentation or frothing of magma due to either rapid expulsion of
dissolved volatiles (Clement 1973, 1982, Dawson 1980) or interaction with groundwater
(Lorenz 1979, Mitchell 1986). In the latter mechanism, known as a fluid-coolant
(magma-water) interaction (Sheridan and Wohletz 1983), groundwater is believed to
flash instantaneously into steam, leading to the explosive ~isruption of the magma into
an "aerosol-like" cloud of rotating droplets of silicate liquid supported by superheated
steam.
Regardless of the mechanism of formation, the presence of olivine nuclei indicates
that, prior to the formation of the lapilli, either some crystallization of the magma had
occurred or that it contained abundant olivine macrocrysts. Subsequent to their genera-
tion, surface tension effects must have resulted in magma adhering to these earlier-formed
crystals. The shapes and textures of the lapilli undoubtedly result from surface tension
effects combined with rotation of the drops during transport. The magma which formed
the mantles was initially fluid, and their textures demonstrate that crystallization contin-
ued after formation of the lapilli. Rapid quenching and the formation of glass are not a
44 CHAPTER!

Figure 1.16. (a) Pelletal lapillus. Postmasburg orangeite. Oriented laths of mica surrounding a kernel of altered
olivine. (Field of view (FOY)) = 1 nun; (b) pelletallapilli. Finsch orangeite (FOY = 4 nun); (c) pelletallapilli.
Chomur kimberlite. Russia (FOY= 4 nun); (d) pelletal lapillus. Urach melilitite (FOY = 2.25 nun).
KIMBERLITES AND ORANGEITES 45

Figure 1.16. (continued)


46 CHAPTER 1

part of the lapilli-forming process in kimberlites, although glassy matrices might have
been originally present in melilititic pelletallapilli.
Autoliths are defined here as angular-to-subrounded, lapilli- to ash-sized clasts
formed by the fragmentation of preexisting solidified material. The term is used in the
original sense of referring to fragments of solidified kimberlite found within· a younger
kimberlite (Rabkin et al. 1961). Autoliths can be derived from any facies and therefore
have any texture. However, autoliths of hypabyssal material appear to be the most
common variety.
The textures of hypabyssal autoliths are identical to those of hypabyssal intrusive
rocks and they may be classified and described by terminology appropriate to the latter
(1.9.1.3). No preferred orientation of minerals is evident within the clasts. Commonly,
fractured microphenocrysts and/or macrocrysts located at the margins of the clasts
provide indisputable evidence that they originated from the fragmentation of solid
material. Autoliths of differing mineralogy and textural type may coexist, and they may
or may not occur with pelletal lapilli (Figure 1.17). Subrounded autoliths may be
impossible to distinguish from pelletallapilli with eccentric kernels that exhibit only weak
concentric structures.
Considerable confusion exists in kimberlite literature concerning the meaning of the
term "autolithic kimberlite." In many Russian publications autolithic kimberlites are
synonymous with pelletal-textured kimberlites, and no distinction is made between
pelletal lapilli and autoliths, even though both may occur in the same sample. In other
cases, diatreme facies rocks are said to contain "nucleated autoliths," an imprecisely
defined term that may refer to either bona fide pelletal lapilli or large (up to 8 cm)
spheroidal masses of hypabyssal rock with cores of country or basement rocks. The latter
have been described by Ferguson et al. (1973) and Danchin et al. (1975) as "nucleated
autoliths." They are not regarded by Clement (1982) or Mitchell (1986) as varieties of
pelletallapilli (see below).
In this work it is suggested that
• Pelletal-textured volcaniclastic diatreme facies rocks are those which contain
significant amounts (> 15 vol %) of pelletallapilli as defined above, e.g., pelle-
tal-textured volcaniclastic kimberlite. Pelletallapilli are primary magmatic con-
stituents and not xenolithic clasts.
• Autolithic volcaniclastic diatreme facies rocks are those which contain significant
amounts (> 15 vol %) of autoliths as defined above, e.g., autolithic volcaniclastic
kimberlite breccia.
• The term "nucleated autolith" (sensu Danchin et al. 1975) be abandoned (see
below).
A given diatreme facies rock may contain significant amounts of both pelletallapilli
and autoliths and should be termed a pelletal-textured autolithic volcaniclastic kimberlite
breccia. Note that a rock containing only pelletallapilli is not in itself a breccia, although
pelletal-textured rocks which contain significant amounts (> 15 vol %) of country rock
clasts may be termed pelletal-textured lithic volcaniclastic kimberlite breccias. Rocks
containing both autoliths and country rock clasts may be described as heterolithic
volcaniclastic kimberlite breccias. Representative idealized examples of these diverse
KIMBERLITES AND ORANGEITES 47

Pelletal ~ Macrocrystal
Lapilli "-2...:::) Olivi ne

~..... Autoliths
'QJ.1:
.: '

Figure 1.17. Idealized representations of the textures of diatreme facies volcaniclastic kimberlite and volcani-
clastic kimberlite breccias (VKB): (A) pelletal-textured volcaniclastic kimberlite. Note this rock is not a breccia;
(B) autolithic VKB; (C) pelletal-textured autolithic VKB; (D) pelletal-textured heterolithic VKB; (E)
pelletal-textured lithic VKB; (F) heterolithic VKB.

textural varieties of diatreme facies rocks are depicted in Figure 1.17. Note that although
olivine (and other) macrocrysts are common in diatreme facies volcaniclastic rocks, their
presence is usually not considered in the terminology of diatreme facies rocks.
Pelletal lapilli, autoliths and country rock clasts are set in a matrix that may be
described as having a uniform or segregationary texture (Clement and Skinner 1985). The
terms refer to whether or not the matrix minerals constitute a uniform aggregate or have
crystallized into discrete patches of differing mineralogy. The interclast matrices are very
fine-grained and consist predominantly, when fresh, of micro- to cryptocrystalline
diopside and serpophitic serpentine. Such matrices are particularly prone to subsolidus
hydrothermal alteration and/or weathering, and are commonly replaced by an optically
unresolvable mixture of serpentine, chlorite, and clay mineral. Interclast matrices do not
contain primary groundmass typomorphic mineral assemblages. Thus, in kimberlitic
48 CHAPTER 1

diatreme facies rocks the interclast matrices are devoid of second-generation olivine,
spinel, perovskite, monticellite, and calcite.
Textural features of the matrix are commensurate with rapid nonequilibrium depo-
sition from a hydrothermal fluid (Mitchell 1986) or the condensate derived from a
vapor-solid fluidized system (Clement 1982). Variations in the temperature, composition,
and rate of crystallization of such fluids can account for gradations between the uniform
and segregationary groundmass textures. In some instances, crystallization of the fluid
commences with the nucleation of diopside upon substrates provided by the clasts and
macrocrysts, the residue crystallizing subsequently as serpentine. Associated pelletal
lapilli and autoliths do not contain diopside.
Mitchell (1986) has explained the abundance of diopside and the absence of mon-
ticellite in the interclast matrices as a consequence of contamination of the matrix -forming
fluids with silica derived from country rock clasts. The process is considered to be
analogous to the contamination of hypabyssal kimberlites by country rock xenoliths. The
latter are commonly surrounded by fringes of diopside which forms as a result of reaction
between the silica-rich xenolith and the magma. Assimilation of the xenolith is believed
to raise the silica activity of the magma to levels which support the crystallization of
diopside in preference to monticellite.

1.9.1.3. Hypabyssal Facies


Hypabyssal kimberlites comprise the root zones of diatremes and occur as dikes and
sills (Figure 1.13, Mitchell 1986, Clement and Skinner 1985, Clement 1982, Dawson
1980). The recommended textural-genetic terminology is given in Figure 1.18. Depend-
ing upon the presence or absence of clasts, rocks may be described as hypabyssal
kimberlite or hypabyssal kimberlite breccia. The latter are defined as containing more

HYPABYSSAL FACIES

APHANITIC

MACROCRYSTAL

PORPHYRITIC

SEGREGATIONARY

Figure 1.18. Textural-genetic classification of hypabyssal facies kimberlites.


KIMBERLITES AND ORANGEITES 49

than 15 vol % of clasts greater than 4 mm in maximum dimension (Clement and Skinner
1985). The clasts may comprise country rocks and/or kimberlitic autoliths. The latter are
regarded as fragments of previously consolidated earlier generations of hypabyssal
kimberlites. Thus, lithic, autolithic, and heterolithic hypabyssal kimberlite breccias may
be recognized. Note that autoliths in the diatreme and hypabyssal environment may be
mineralogically similar but texturally different, with the latter being typically coarser
grained and less altered. Pelletallapilli do not occur in hypabyssal facies rocks, although
spherical masses of kimberlite, known as "globular segregations," are relatively common
(see below).
Kimberlites and kimberlite breccias may be further described as aphanitic or macro-
crystal in character, with porphyritic kimberlites being extremely rare (see 1.2). Previous
textural-genetic classifications have not incorporated any information as to the nature of
the macrocrysts. Although olivine is by far the most abundant macrocryst, there do exist
kimberlites (De Beers, Tunraq, Zagodachnaya, Koidu) which contain significant amounts
of macrocrystal phlogopite, ilmenite, or garnet. It is recommended that the term "macro-
crystal" when used without prefix refers to a macrocrystal assemblage dominated by
olivine. Prefixes may be added to recognize the presence of significant modal amounts
(>15 vol %) of other macrocrystal phases, e.g., a phlogopite macrocrystal kimberlite
would contain macrocrysts of both phlogopite and olivine. The recognition ofphlogopite
macrocrystal kimberlites is very important as these rocks are not petrologically synony-
mous with, and cannot be termed, "phlogopite kimberlites," as this term is reserved for
describing kimberlites rich in primary groundmass phlogopite.
The groundmass may be described as uniform or segregationary, depending upon
whether the primary groundmass phases and the mesostasis have crystallized together or
separately. Rocks with a uniform groundmass may be described by standard terminology
and named using the mineralogical-genetic principles outlined in Section 1.4.4., e.g.,
hypabyssal macrocrystal apatite monticellite calcite kimberlite heterolithic breccia.
Segregationary-textured kimberlites have a nonuniform distribution of groundmass
minerals and mesostasis. The groundmass is commonly identical in texture and character
to that of uniformly textured hypabyssal kimberlite. The segregations are amoeboid-to-
spherical discrete regions consisting of relatively coarse-grained phlogopite-kinoshitalite
solid solutions, apatite, calcite, and primary serpentine. Segregations consisting of calcite
and serpentine are particularly common. Although the mineralogy of the mesostasis of
the uniformly textured portions of the rock and the segregations may be similar, the latter
typically lack perovskite and spinel. Segregations may be described as being bounded or
gradational, depending upon whether their contact with the uniformly textured ground-
mass is sharp « 50 Jlm) or gradational (20-500 Jlm). Commonly, euhedral crystals of
apatite, calcite, and phlogopite project into the serpentine-filled centers of segregations.
The segregations result from the separation of late-crystallizing components from
the silicate oxide groundmass into discrete masses. In some instances, e.g., the Benfontein
and Skinner's Sills, segregations appear to have migrated through their partially crystal-
lized parental liquids prior to complete consolidation of the magma. The origin and
mobilization of segregations has been discussed at length by Dawson and Hawthorne
(1973), Donaldson and Reid (1982), Mitchell (1984a, 1986), and Clement (1982). A
consensus regarding their genesis has not been reached and segregations are variously
50 CHAPTER 1

regarded as gas condensates in vesicles, filled breached vesicles, immiscible liquids, or


low-temperature residual fluids. In the latter case the segregations form as a result of
surface tension effects between the water-rich segregation and more viscous crystal-rich
silicate oxide groundmass (Mitchell 1986).
Globular segregations are spherical masses of hypabyssal material which may range
up to 100 mm in diameter. When present in large numbers, they confer a pseudocon-
glomeritic appearance to the rock. In kimberlites they consist of relatively fine-grained
hypabyssal kimberlite and may be found locally in coarser grained, but otherwise similar,
uniformly textured hypabyssal kimberlite. Clement (1982) and Mitchell (1986) have
suggested that the segregations are generated by surface tension effects in boiling magmas
in near-surface hypabyssal environments.
There are also known spherical masses of kimberlite which contain distinct cores of
xenolithic material. The xenoliths appear to be predominately basement gneisses, and
relatively few cores consist of high-level country rocks. These objects are typically found
irregularly distributed within diatreme facies volcaniclastic rocks. Commonly, they are
strongly concentrically zoned, the zonation resulting from the presence of many thin
concentric shells of fine-grained hypabyssal kimberlite of differing thickness and mode.
The size of the kimberlitic mantle bears no systematic relation to the size of the xenolithic
core. Examples are known from several Lesothan (Ferguson et al. 1973) and Yakutian
(Legkaya, Zarnitsa) kimberlites.
The objects have been termed "nucleated autoliths" by Danchin et al. (1975) and
Ferguson et al. (1973). This term is inappropriate as they are not fragments of preexisting
rocks. They have not crystallized in their current hosts and must represent a transported
assemblage. In contrast, the globular segregations described from Finsch and Dutoitspan
(Clement 1982) or Mukorob (Mitchell 1986) appear to have crystallized in situ in the
hypabyssal environment.
Whether the two varieties of spherical "segregations" have the same origin, and thus
should have the same name, is debatable, as they have not yet been sufficiently studied.
Until further information regarding their character and genesis becomes available, it is
recommended that the term "nucleated autolith" (sensu Danchin et al. 1975) be aban-
doned and these objects be referred to by the nongenetic term nucleated globules. In
contrast, spherical objects which appear to have been generated in situ by local devolati-
zation, or boiling of magma in closed systems in near-surface hypabyssal environments
(Mitchell 1986, Clement 1982) should retain the appellation globular segregation.
Note that globular segregations are not confined to kimberlites and are known from
lamproite dikes (Pilot Butte, Wyoming, this work) and orangeites (Finsch, Clement 1982),
suggesting that a common mechanism is responsible for their formation in these diverse
volatile-rich magmas. Thus, the terminology, while genetic in character, is not magma
specific.
Globular segregations have petrographic similarities to some pelletallapilli (Clement
1982); however, they are typically coarser grained and commonly lack a macrocrystal
nucleous. Clement (1982) has suggested that some coarse-grained pelletallapilli grew as
small globular segregations and were subsequently mixed with fine-grained "spray" or
"aerosol" pelletal lapilli that were generated during the diatreme-forming process. This
hypothesis assumes that pelletallapilli were also formed primarily by volatile degassing.
KIMBERLITES AND ORANGElTES 51

However, lapilli may have hydroclastic origins (Mitchell 1986), thus ruling out the
existence of a genetic continuum between pelletallapilli and globular segregations. Of
course, petrographically similar, but genetically different, lapilli may coexist as a hybrid
assemblage. Clearly, much further work on these problems is required.

1.9.1.4. Spatial Relationships between Diatreme and Hypabyssal Facies Kimberlites


Current ideas regarding the spatial relationships between diatreme and hypabyssal
facies rocks are based upon data obtained from the deep mining of several southern
African kimberlites. According to this model (Figure 1.13), hypabyssal rocks occur only
in the root rones of diatremes and as sills and dikes. The high levels of diatremes lack
intrusive hypabyssal rocks, and in this regime they are present only as clasts of disaggre-
gated preexisting rocks. There are no obvious physicochemical reasons precluding the
formation of hypabyssal rocks at high levels in diatremes. Such rocks are common in
melilitite diatremes and typical of lamproitic vents.
Recently, information has become available regarding the structure of some Yakutian
diatremes which seems to be at variance with the southern African model. Thus, diatremes
in the Daldyn-Alakit field differ from those in southern Africa in that they are commonly
multiple intrusions in which massive macrocrystal hypabyssal kimberlite occurs at higher
levels in the diatreme.
Figures 1.19 to 1.21 illustrate the structures of the Sitikanskaya, Krasnopresnen-
skaya, and Udachnaya intrusions of the Daldyn-Alakit field. These are all double
intrusions which coalesce into a single complex body at high structural levels. On the
basis of these data, Russian geologists believe that many geographically closely related

--.=:-.=-- - - - - --~-=-:::
- --=---~-=-:::-::: - - --=-:::-
- -:::-=-:::-:::-:::-
- - - - --:::-::=-':--~-:-==--p:'
-- -- - - -- -- T:::-:::-=-:::--:::---_-_-_-_
-- - - -:... __
-~-$;::;;:;:;;:;:;:;:;:~

- 230

Figure 1.19. Cross section of the Sitikanskaya diatreme, Alakit field, Yakutia. This double pipe consists of an
earlier intrusion of volcaniclastic kimberlite breccia and a later intrusion ofmacrocrystal hypabyssal kimberlite.
Note the downrafted block of Silurian sediment (S 1), the Carboniferous-Permian metachronous volcanogenic
(epiclastic) sedimentary rocks (C-P) and the capping of Permo-Triassic (P-T) basaltic lavas (after data provided
by the Amakinsk Geological Expedition, Aikhal, Yakutia).
52 CHAPTER I

Figure 1.20. Isometric diagram illustrating the fonn of the Krasnopresnenskaya kimberlite, Alakit field,
Yakutia. This Paleozoic age double pipe (stippled) has been intruded by a basaltic sill (dashed) of Penno-Tri-
assic age. Bowl-shaped depressions in the upper parts of the pipes are filled with metachronous volcanogenic
(epiclastic) sedimentary rocks and covered with younger basaltic volcanic rocks (see Figure 1.23 for details)
(after Kriuchkov et al. 1994).

intrusions represent the eroded roots of such double-, or even multiple-, pipe systems.
This approach is illustrated in Figure 1.22 which shows that four intrusions, exposed at
the current level of erosion in the Ukukitskoye field, are believed to have formed a large
coalescing vent at the time of intrusion.
Cross sections of the Krasnopresnenskaya (Figure 1.23) and Yubileinaya (Figure
1.24) intrusions illustrate further significant differences with respect to southern African
pipes. The upper part of the Krasnopresnenskaya intrusion (Kriuchkov et at. 1994) flares
rapidly into a wide bowl-shaped body above the cylindrical, smooth-sided main feeder
vent. The sides of the bowl-shaped unit dip at 50-60° adjacent to the feeder and at 15-25°
at the margins. The bowl is filled with Lower Carboniferous metachronous volcanogenic
(epidastic) lacustrine sediments which may be up to 50 m thick and which cover most of
WEST EAST

Figure 1.21. Cross section of the Udachnya kimberlite. Each pipe is filled with different varieties of
volcaniclastic and macrocrystal kimberlites as indicated by the different ornamentation (after Milashev 1984).

BALATINSKAYA

FESTIVALNAYA VASILYEOSTROVSKAYA

PETROGRADSKAYA
Figure 1.22. Postulated structure of a group of kimberlite diatremes in the Ukukitskoye field. Yakutia, prior
to erosion. At the current level of exposure there occur four apparently unconnected intrusions (after V.
Kornilova. pers. comm.).
S3
54 CHAPTER 1

Figure 1.23. Cross section of the Krasnopresnenskaya diatreme. Alakit field. Yakutia. Kl and K2 are
volcaniclastic kimberlite breccias. Note that K2 contains abundant xenoliths of wall rock Silurian limestone
together with xenoliths of Ordovician marl which have been transported upward in the diatreme. The
bowl-shaped depression above K2 contains Carboniferous metachronous volcanogenic (epiclastic) sedimentary
rocks (KCF). The intrusion is capped by a variety of Permo-Triassic basaltic rocks and cut by a basaltic sill of
the same age (after Kriuchkov et al. 1994).

the underlying diatreme. Late Carboniferous sedimentary rocks unconformably overlie


the volcanogenic (epiclastic) sedimentary rocks and cover the entire diatreme. The
kimberlites and the overlying sedimentary rocks have been intruded, and covered, by
Permo-Triassic basaltic dikes, sills, and "tuffaceous units."
KIMBERLITES AND ORANGElTES ss

w E
o

400
+
+
+
600 +
+
+ + +
+. +
800
VKB + VKB + +
• +
+ + + + + +
1000
m
Figure 1.24. Cross section of the Yubileinaya kimberlite, Alakit field. Yakutia. The intrusion consists of an
early volcaniclastic kimberlite breccia (VKB) which has been intruded by a macrocrystal hypabyssal kimberlite
breccia (MKB). Note the "flaring" of the latter unit at the top of the pipe. The bowl-shaped depression above
the MKB is filled with metachronous volcanogenic (epiclastic) sedimentary rocks (ME), and the whole intrusion
is capped by Perrno-Carboniferous (P-C) sediments (after Kharkiv 1990).

The vent is filled with two types of kimberlites. The upper phase is termed a
"kimberlite tuff breccia," the lower phase a "autolithic tuff breccia." Samples of these
rocks are not yet available for study, and consequently it is not possible to describe them
in terms of current textural-genetic classifications.
Figure 1.24 shows that the upper portions of the Yubileinaya intrusion (Kharkiv
1990) also flare out rapidly into a bowl-shaped body. The intrusion consists of at least
two kimberlite types with the youngest of these forming the flared body. Russian
petrologists consider that each intrusion consists of "autolithic kimberlite breccia" and
"kimberlite with a massive cement" (sensu Komilova et at. 1983). The former is equiva-
lent to autolithic kimberlite breccia, and the latter to macrocrystal hypabyssal kimberlite
as used in this work. Note that hypabyssal kimberlites are apparently present at high
structural levels in this intrusion as the erosion level in this field is not great enough to
expose bonafide root zones and feeder dike systems.
The upper levels of the intrusion are covered by metachronous volcanogenic (epi-
clastic) sedimentary rocks that are considered by Kharkiv (1990) to be crater facies rocks
(see below).
The Sitikanskaya (Figure 1.19) pipe shows similarities to the Yubileinaya intrusion.
Here a complex multiphase unit of pelletal-textured and volcaniclastic diatreme facies
kimberlite breccias is intruded by a late-stage unit of macrocrystal hypabyssal kimberlite
56 CHAPTER 1

• • • • • • • • • •
• • • • • • • • •

• • • • • DOLERITE
• • • ••
• •
• •
P2- T 3 •

• • •
• • •

LIMESTONE

Figure 1.25. Cross section of the Odintsov pipe. Alakit field. Yakutia (after Kriuchkov er al. 1994). K-I:
porphyritic kimberlite; K-2: autolithic kimberlite breccia; K-3: supra-pipe breccia-limestone + kimberlite.

and autolithic kimberlite breccia. This latter unit again flares out into a shallow-dipping
body in the upper parts of the vent. Metachronous volcangenic (epiclastic) sedimentary
rocks are preserved in a bowl-shaped depression above the northeastern intrusion.
The Odintsov pipe (Kriuchkov et al. 1994) appears to be a proto- or blind diatreme
(Clement 1982, Mitchell 1986), although this intrusion has no exact counterpart in the
southern African kimberlite province. Russian geologists consider the body to be a
cryptovolcanic structure. The "vent" consists of a complex breccia unit termed the
"carbonate cap," which is underlain by a pipe of kimberlite (Figure 1.25). The carbonate
breccia consists of local country rock limestones which have been fragmented by
explosive volcanism. Clasts in this breccia range in size from 5 to 10 cm together with
large (several meters) xenoliths. The clasts are cemented together by kimberlite-derived
material and the cap rock is extensively veined in the lower regions by massive kimberlite.
The cap rock clearly represents the fragmentation of the country rock above an advancing
pulse of kimberlite magma. The body is described as a "semi blind kimberlite" by Russian
KIMBERLITES AND ORANGEITES 57

geologists as the kimberlite-derived fluids are believed to have penetrated to the surface,
while the kimberlite did not actually erupt.
Intrusive kimberlites are present below the cap rock. These consist of an early
generation of "porphyritic kimberlite" which has been intruded by a later "autolithic
kimberlite" (sensu Kriuchkov et al. 1994). The former probably corresponds to macro-
crystal hypabyssal kimberlite, although the exact nature of the latter is uncertain. If it is
a volcaniclastic diatreme facies unit, the sequence of intrusion would be quite unlike that
seen in southern Africa.
Multiple pipes are not, with the exception of the Mir-Sputnik pair, characteristic of
the Malo-Botuobinsk field, and diatremes (International, 23CPC) in these fields are
similar to southern African diatremes. The kimberlite fields in the northern parts of the
Yakutian province, e.g., the Kuoiskoye, Molodinskoye, Toluopskoye fields, are more
deeply eroded, and only root zones and the feeder dike systems, consisting of hypabyssal
kimberlites, are preserved. Many of the small hypabyssal intrusions, e.g., Obnazhennaya,
Russlovaya, Festival, Anomaliya 23, in the Kuoiskoye field, are similar to "blows"
developed along dike systems in the southern African fields, e.g., pipe 200 (Kresten and
Dempster 1973). Russian geologists explain these differences from the Daldyn-Alakit
and Chomur fields on the basis of the relative degree of erosion of the fields. Thus,
multiple and flared pipes are not present in the above fields due to the greater extent of
erosion.
In summary, the character of high-level intrusions in the Daldyn-Alakit fields
suggests that
• Macrocrystal hypabyssal facies kimberlite may exist in the upper levels of
diatremes.
• Volcaniclastic kimberlites may intrude hypabysssal kimberlites.
• Multiple coalescing intrusions are common.
• The upper parts of intrusions commonly flare out into wide bowl-shaped units
which subsequently form erosional depressions which may be filled with pseudo
crater facies, metachronous volcanogenic (epiclastic) sedimentary rocks.
With respect to the last conclusion, which contradicts Kharkiv (1990), it is apparent
that volcanogenic (epiclastic) sedimentary rocks in the Daldyn-Alakit field are not true
crater facies rocks. This may be demonstrated at Sitikanskaya where these rocks are found
to rest directly upon a downrafted block of Silurian limestone enclosed within volcani-
clastic diatreme facies rocks (Figure 1.19). Thus, the volcanogenic sedimentary units
cannot be either in situ or downrafted crater facies material. They merely occupy an
erosional depression formed in the pipe subsequent to erosion of the actual crater and
uppermost parts of the diatreme (see 1.9.1.1). They and other pseudo crater facies rocks
found at Aikhal, Krasnopresnenskaya, and Yubileinaya formed on a Lower Carboniferous
paleoerosion surface, and have been preserved only because of their subsequent burial
beneath a protective cover of Permo-Triassic basalts. Consequently, although high
structural levels of diatremes are found in the Daldyn-Alakit region, true crater facies
rocks are not preserved in the Yakutian province.
The differences in the intrusive style of the Yakutian diatremes relative to the southern
African diatremes are significant, but as yet unexplained. Paired or triple diatremes which
58 CHAPTERl

coalesce at higher structural levels are known from southern Africa, e.g., Venetia,
Jwaneng. Clement (1982) has suggested that the Koffiefontein-Ebenhaezer and Dutoit-
span-Bultfontein diatremes may also have coalesced at higher levels. However, such
occurrences are not typical of the province as a whole.
As the petrographic character of the kimberlites in the two provinces is similar, it is
apparent that the differences in intrusive style must reflect differences in the nature and
water contents of the intruded rocks rather than differences in the parental magmas.

1.9.2. Orangeites
Orangeites occur as diatremes, dikes, and sills (l.8). These intrusions, on the basis
of our current knowledge, apparently do not differ in style from those of archetypal
kimberlites. The textural-genetic classification developed for kimberlites in this work
(Figures 1.14, 1.15, and 1.18) is therefore considered to be entirely applicable to
orangeites. Thus, the diatreme facies Fl unit at Finsch (Clement 1982) may be described
as a pelletal-textured heterolithic volcaniclastic orangeite breccia. Individual hypabyssal
facies orangeites may be described according to the mineralogical-genetic classifications
outlined in Section 1.4.4 and described in Section 1.1 O.
The majority of orangeite intrusions appear to belong to the hypabyssal and root-zone
facies. Large diatremes are represented only by the Finsch pipe, and crater facies rocks
have not been described.

1.9.3. Melilitite Clan


Melilitite magmas form diatremes, dikes, and sills in addition to occurring as lavas
and plutonic rocks. The textural-genetic classification developed for kimberlites is
applicable to rocks belonging to this clan with the addition of a plutonic facies. Most
melilitoids can be described by standard igneous petrological terms.
Melilititic diatremes are common, e.g., Swabia (Cloos 1941, Lorenz 1979), Missouri
(Singewald and Milton 1930), Montana (Hearn 1968), Sutherland, South Africa (McIver
and Ferguson 1979), and James Bay Lowlands, Ontario (Janse et al. 1989). These display
the full range of textures exhibited by diatreme facies kimberlites (and orangeites). Thus
pelletal-textured autolithic volcaniclastic melilitite breccias (Figure 1.16) may be found
in such diatremes. Particular care must be taken in exploration programs in distinguishing
these rocks from superficially similar volcaniclastic kimberlites. This cannot be achieved
using macroscopic observations alone, and correct assessment of petrological clan affinity
requires detailed mineralogical studies using a combination of optical and electron
microbeam methods (Mitchell 1995).
Although detailed discussion of the nature of melilititic diatremes is beyond the scope
of this work, note that they differ from most kimberlite diatremes in exhibiting
well-developed internal structures consisting of bedded volcaniclastic units and
central conduits of hypabyssal material (Hearn 1968, Lorenz 1975, 1984). The occur-
rence of high-level magmatic rocks is similar to the presence of high-level hypabyssal
facies material in some of the Yakutian diatremes, although the latter lack well-defined
internal structures.
KIMBERLITES AND ORANGEITES 59

Figure 1.26. Macrocrystal orangeite. Star dike. M =phlogopite; 0 =olivine (TL. FOY =4 mm).

Figure 1.27. Macrocrystal orangeite. Sover. M = phlogopite; 0 = olivine (TL. FOV = 4 mm).
60 CHAPTER 1

1.10. PETROGRAPmC CHARACTERISTICS OF ORANGEITE


Orangeites may be described using mineralogical-genetic classifications, as de-
scribed in 1.4.4. The majority of orangeites do not differ greatly from one another in their
petrographic character within and between intrusions. The pri!1cipal differences are with
respect to the amount of macrocrystal olivine and the ratio of macrocrystal to microphe-
nocrystal phlogopite. Only evolved orangeites contain significant modal amounts of
diopside and/or sanidine.
Although diatreme and hypabyssal facies orangeites are known, it is only from the
latter that samples suitable for undertaking detailed petrographic and mineralogical
studies may be obtained. In common with kimberlites, diatreme facies rocks are typically
altered and usually do not contain typomorphic minerals because of their rapid crystal-
lization (Scott Smith 1992). Diatreme facies rocks are thus best identified as orangeites
by their consanguineous association with hypabyssal facies rocks.
The principal varieties of hypabyssal orangeite are the following:

Figure 1.28. Macrocrystal orangeite, Bellsbank. Note the region consisting entirely of mica and apatite (white)
at lower right. Backscattered electron image (BSE-image). 0 = olivine.
KIMBERLITES AND ORANGEITES 61

Figure 1.29. Macrocrystal orangeite. Sover. Note the areas rich in apatite (white) in the fine-grained phlo-
gopite-rich matrix. BSE image. 0 =olivine.

Orangeite, a rock consisting principally of microphenocrystal phlogopite set in a


fine-grained groundmass consisting essentially of phlogopite-tetraferriphlogopite and
minor apatite, chromite, Mn-ilmenite, and perovskite, with a mesostasis of calcite and/or
dolomite together with serpentine (Figures 1.26-1.45). The ratio of microphenocrysts to
groundmass mica and/or mesostasis varies widely. Some rocks consist predominantly
(85-90 vol %) of closely packed tablets of microphenocrystal phlogopite (Figure 1.33).
Many of the microphenocrysts and groundmass micas are strongly-zoned from pale
yellow phlogopite cores to bright-red tetraferriphlogopite margins (Figure 1.39).
Phlogopite macrocrysts are common (1-10 vol %) and typically deformed and
broken (Figures 1.26, 1.27, 1.31). Typically, their composition is broadly similar to that
of microphenocrystal phlogopite. Minor reverse zoning and mantling may be present.
Cryptogenic macrocrysts of green or brown pleochroic biotite are rarely «<1 vol %)
present. These are typically mantled by colorless phlogopite (Figure 1.32).
Relative to kimberlites the groundmass of orangeites is fine-grained, poor in spinels
and perovskites (see 1.11), and rich in apatite. The latter may occur as euhedral prisms
62 CHAPTER 1

Figure 1.30. Macrocrystal orangeite. Boshof. Note that the matrix consists primarily of microphenocrystal
phlogopite (gray) and euhedral apatite (white). BSE image. 0 =olivine.

(Figures 1.37, 1.41) or poikilitic plates. Rare relatively coarse-grained patches (?segre-
gations) in the ground mass consist predominantl y of phlogopite and apatite set in a calcite
mesostasis (Figure 1.41).
Small euhedral-to-subhedral primary olivines may be present but are not typically
abundant and are commonly completely serpentinized.
Macrocrystal orangeite is a rock identical to the above but characterized by a
pronounced inequigranular texture due to the presence of olivine macrocrysts (Figures
1.26-1.30). The latter may be fresh, or partially or completely serpentinized. As a
consequence of the extremely inhomogeneous distribution of macrocrysts within a given
intrusion, there is a complete gradation from macrocrystal orangeite to rocks entirely
lacking olivine macrocrysts. The majority of the macrocrysts, on the basis of their habit,
undulose extinction, and occurrence as multiple-grain aggregates, are considered to be
xenocrysts.
Diopside orangeite and macrocrystal diopside orangeite (Figures 1.46, 1.49) are
similar to orangeite and macrocrystal orangeite as described above, but differ in contain-
Figure 1.31. Orangeite, New Elands. Macrocrysts of phlogopite (M) set in a very fine-grained groundmass
consisting of euhedral spinels (black) and fine-grained phlogopite (TL, FOY =4 mm).

Figure 1.32. Orangeite, Star. Macrocrysts of phlogopite (PHL) and dark-colored aluminous biotite (B) together
with euhedral, mantled microphenocrystal phlogopite (M), and opaque spinel (TL, FOY '" 2.25 mm).
63
Figure 1.33. Orangeite, Swartruggens. Closely packed aggregate of microphenocrystal phlogopite (TL, FOY =
1 mm).

Figure 1.34. Orangeite, New Elands. Microphenocrysts of phlogopite and subhedral opaque spinels set in
carbonate-rich mesostasis (TL, FOY = 1 mm).
64
Figure 1.35. Orangeite, Finsch. Plates of groundmass phlogopite enclosing opaque subhedral spinels (TL,
FOY =2.25 mm).

Figure 1.36. Orangeite, Makganyene. Flow-aligned microphenocrysts of phlogopite (TL, FOY =2.25 mm).
65
~

(":)
Figure 1.37. Macrocrystal orangeite, Bellsbank. Mantled olivine macrocryst Figure 1.38. Orangeite, Swartruggens. Macrocrystal (PHL) and microphe- ==
(0) set in a matrix of microph enocrystaI and groundmass phlogopite (gray). The nocrystal (M) phlogopite set in a matrix of calcite, phlogopite, and apatite. Dark
groundmass contains abundant euhedral prisms of apatite (white). Mesostasis is laths are serpentine/chlorite pseudomorphs after an unknown primary mineral. ~
:;c
calcite (dark gray). BSE image. BSE image. ....
~
gg
::c
5
~
~
t::I
o
~
t"l

Figure 1.39. Orangeite, Swartruggens. Microphenocrystal phlogopite (M) Figure 1.40. Orangeite, Swartruggens. Distorted, closely packed microphe-
mantled by tetraferriphlogopite (light gray). Other groundmass minerals nocrysts of phlogopite (M), and subhedral apatite (white).
present include apatite and spinel (white). Mesostasis (dark gray) is calcite.
BSEimage. ~
~

120 urn

Figure 1.41. Orangeite, Swartruggens. Euhedral microphenocrysts of zoned and Figure 1.42. Orangeite, New Elands. Chloritized, distorted microphenocrysts of
=
mantled phlogopite (M) and euhedral crystals of apatite (A) set in a serpentine-<:al- phlogopite (M) set in a matrix of apatite (white) and calcite (C). BSE image. ~
i:I::I
cite mesostasis (C). BSE image. ...
150 urn 120 urn ~
=
til
~
>
~
o

i
~

Figure 1.43. Orangeite, Lace. Microphenocrysts of phlogopite (M) set in a Figure 1.44. Orangeite, Bellsbank. Chloritized (dark bands), microphenocrysts
matrix of subhedral apatite (white) and calcite (C). BSE image. of phlogopite (M) mantled by tetraferriphlogopite (light gray) and subhedral
groundmass apatite (A) set in a mesostasis of carbonate (mottled) and serpentine
(dark gray). BSE image. ~
....<:>

("')

Figure 1.45. Orangeite, Swartruggens. Euhedral groundmass apatite (white). Figure 1.46. Diopside orangeite, Postmasburg. Mantled microphenocrysts of phlo-
Inclusions (dark) in the apatite are serpentine pseudomorphs after an unknown gopite (M) set in a matrix of diopside (d), apatite (white), calcite (dark gray), and
mineral. C = calcite. M = microphenocrystal phlogopite. BSE image. potassium feldspar plus serpentine (black). BSE image.
~
....
~
t:I:I

"-l
~
>
~
o
~
~
"-l

Figure 1.47. Diopside orangeite, Postmasburg. Zoned and mantled microphe- Figure 1.48. Diopside sanidine orangeite. Postmasburg. EuhedraI crystals of
nocrysts of phlogopite-tetraferriphlogopite (m) and prismatic diopside (d) set in groundmass diopside (D) with mantles of titan ian aegirine (light gray rims). Also
a matrix of apatite (white), calcite (gray), and serpentine (black). BSE image. present is subhedral apatite (white). Dark gray mesostasis is sanidine (S). BSE
-..I
image.
-
72 CHAPTER 1

Figure 1.49. Diopside-bearing orangeite, Swartruggens. Typical appearance of a diopside microphenocryst (d)
in an unevolved orangeite. M =phlogopite microphenocryst (TL, FOY =1 mm).

ing widely ranging amounts of microphenocrystal diopside (1-20 vol %). The diopside
phenocrysts are typically not zoned and are commonly resorbed. Groundmass potassium
feldspar is absent. Diopside is common in many, but not all, dikes in the Winburg, Boshof,
and Kroonstad areas. Diopside-bearing orangeites seem to be typically absent from the
Sover-Doomkloof, Main Bellsbank, Bobbejaan, and Newlands dikes (Wagner 1914,
Dawson et al. 1977, Tainton 1992).
Sanidine diopside orangeite is a rock similar to orangeite as described above but
containing less phlogopite and characterized by the presence of abundant euhedral-to-
subhedral prisms of diopside (20-50 vol %). The latter may be zoned from colorless
diopside toward greenish titanian aegirine and are commonly set in a matrix of potassium
feldspar (Figure 1.48). Examples are known from the Voorspoed, Postmasburg, Mak-
ganyene, and Prieska region orangeites.
Richterite orangeite is known from Pniel (Tainton 1992) and consists of macrocrystal
olivine with reaction coronas of phlogopite which, together with diopside and chromite,
are set in a groundmass of phlogopi te-tetraferriphlogopite. In this rock, richterite occurs
as anhedral tablets which may subpoikilitically enclose phlogopite and diopside (Figure
1.52). Sanidine is absent from the Pniel samples examined by Tainton (1992) and Erlank
(1973). Its absence, and the similarity of the rock to amphibole-free "group 2 kimberIites"
in the Barkly West region, led Tainton (1992) to describe the rock as an "amphibole-bear-
ing phlogopite kimberlite."
KIMBERLITES AND ORANGEITES 73

Figure 1.50. Macrocrystal sanidine orangeite, Besterskraal. Olivine macrocrysts (0) and microphenocrystal
phlogopite (M) set in a matrix of apatite (white) and sanidine (black).

Very similar rocks to the above occur at Besterskraal, Sover North, Lace, and
Makganyene. In these, richterite also crystallizes after tetraferriphlogopite as a late-stage
groundmass mineral. The rocks differ in that the richterite is intergrown with colorless
potassium feldspar (Figure 1.53). The latter is commonly replaced by clay minerals and
serpentine-like material. Tainton (1992) classified these rocks as "macrocrystal K-
richterite phlogopite lamproites," regardless of their association with typical "group 2
kimberlites." In this work, and Mitchell (1994), these richterite sanidine orangeites (see
1.5, 1.12) are regarded as the products of differentiation of orangeite magma and termed
evolved orangeites.
Richterite sanidine orangeites from Sover North and Besterskraal commonly contain
poikilitic plates of groundmass phlogopite (Figure 1.54). These oikocrysts contain
chadacrysts of chromite, diopside, apatite, and subspherical inclusions now consisting of
potassium feldspar and/or serpentine-like material. Such poikilitic micas have been found
only in the most evolved orangeites which are rich in K-Ba titanates, sanidine, and
richterite. The spherical inclusions, on the basis of their morphology, are considered by
74 CHAPTER!

Figure 1.51. Macrocrystal diopside sanidine orangeite, Sover North. Macrocrysts of olivine (0) with phlo-
gopite reaction rims, subhedral diopside (D), and apatite (white) set in a matrix of groundmass phlogopite and
sanidine (black).

Tainton (1992) to have been originally leucite. The texture of the rocks is reminiscent of
that ofmadupitic olivine lamproites (Mitchell and Bergman 1991, p. 34), and it is in these
rocks that the mineralogical convergence toward lamproitic mineral assemblages is best
developed (see 1.12).
Photomicrographs illustrating the petrographic characteristics of orangeites are
presented in Figures 1.26-1.54. These should be compared and contrasted with the
illustrations of kimberlites given in Figures 1.55-1.76.

1.11. PETROGRAPHIC DIFFERENCES WITH RESPECT TO KIMBERLITES


Hypabyssal kimberlites are typically characterized by the presence of macrocrysts
and subhedral microphenocrysts of olivine set in a groundmass consisting of spinel,
perovskite, monticellite, phlogopite-kinoshitalite, apatite, serpentine, and calcite (Fig-
ures 1.55-1.76). Modes vary widely as a consequence of differentiation and alteration.
Kimberlites are described according to the abundances of the primary groundmass
~
~

l'"J
~
til

~
1:;1
o

til
!

Figure 1.52. Richterite orangeite, Pniel. Groundmass phlogopite (M) with diopside and apatite inclusions set in a matrix of potassium richterite (R)
-..J
(TL, FOV = I mm). VI
76 CHAPTER 1

Figure 1.53. Richterite sanidine orangeite, Besterskraal. Subhedral richterite (R) and diopside (D) are set in a
matrix of phlogopite (M) and altered sanidine (S) (TL, FOV =2.25 mm).

minerals, as described in Section 1.4.4. using the principles established by Skinner and
Clement (1979).
Microphenocrystal mica is not a characteristic mineral ofkimberlites, although many
contain minor amounts «1 vol %) of macro crystal phlogopite (Figure 1.56). When mica
is present, it occurs primarily as late-stage colorless poikilitic plates and laths of phlo-
gopite-kinoshitalite (Figures 1.66, 1.67, 1.74, 1.75) and less commonly as small ground-
mass tablets of phlogopite. Although such rocks may be described as phlogopite
kimberlites, it is very important to realize that this term is not used in the original sense
of Skinner and Clement (l979), as they did not discriminate between orangeites and
archetypal kimberlites rich in macrocrystal phlogopite. Hence, in their terminology both
varieties were termed "phlogopite kimberlite."
Kimberlites differ substantially from orangeites with respect to their overall texture
and the type and abundance of primary minerals, as described in Section 1.5 and Table
1.1.
Many kimberlites are characterized by segregation-textured groundmasses (1.9.1.3;
Figures 1.69, 1.71-1.73). Such calcite-serpentine segregations appear to be absent from
orangeites. Spinels and perovskites are abundant in kimberlite and coarse-grained relative
to those in orangeites (1.5; Skinner 1989). Sheafs of quench-textured apatite are common
in kimberlites but not typical of orangeites. In the latter, apatites occur as euhedral prisms
KIMBERLITES AND ORANGEITES 77

Figure 1.54. Evolved orangeite, Sover North. Groundmass poikilitic phlogopite with spherical pseudomorphs
of serpentine which are considered by Tainton (1992) to be formed after leucite. BSE image.

or poikilitic groundmass plates. Atoll-textured spinels (Figure 1.76; Mitchell 1986, pp.
217-219), although common in kimberlites, appear to be absent from orangeites.
The only kimberlites which are petrographically similar to unevolved macrocrystal
orangeites are those which are modally enriched in macrocrystal mica, e.g., Tunraq,
Koidu, Zagodochnaya, Ngopoetsu, Chicken Park, as a consequence of concentration of
this mineral by differentiation processes. These kimberlites typically exhibit other textural
and mineralogical features characteristic of kimberlite which serve to set them apart from
orangeites, i.e., they are oxide-rich, contain atoll spinels, and have segregation textures
while tetraferriphlogopite and diopside are absent.
Kimberlites do not contain richterite, sanidine, or primary diopside and consequently
are easily distinguished from evolved orangeites containing these minerals.
Photomicrographs of hypabyssal kimberlites illustrating their characteristic texture
and mineralogy are given in Figures 1.55-1.76 and should be compared and contrasted
with those of orangeites in Figures 1.26-1.54.
Figure 1.55. Macrocrystal kimberlite, Wesselton Mine. Note the presence of euhedral-to-subhedral primary
groundmass olivine. Unresolved dark matrix consists of spinel, perovskite, carbonate, and serpentine. (TL,
FOV=4nun).

Figure 1.56. Macrocrystal kimberlite with a phlogopite macrocryst (M), De Beers Mine. Note the presence of
abundant small primary olivines and opaque spinels in the groundmass (TL. FOV = 4 nun).

78
KIMBERLITES AND ORANGEITES 79

Figure 1.57. Macrocrystal kimberlite, Wesselton Mine. Typical appearance of the oxide-rich groundmass of
an archetypal kimberlite consisting of subhedral primary olivine (0) and opaque minerals (spinel and
perovskite) set in a uniform mesostasis of serpentine and calcite (TL. FOV =2.25 rum).

1.12. PETROGRAPmC DIFFERENCES WITH RESPECT TO LAMPROITES


Orangeites and macrocrystal orangeites are petrographically so different from most
lamproites that they are easily distinguished from the latter. However, evolved orangeites
contain minerals (richterite, sanidine, diopside, K-Ba titanates) similar in type and
composition to some of those found in lamproites. Thus, considered in isolation and
without reference to their consanguineous antecedents, some orangeites might indeed be
petrographically classified as lamproites, e.g., previous descriptions of rocks from Sover
North (1.5, 1.11; Tainton 1992) or Pniel (Mitchell and Bergman 1991) as "lamproite."
Richterite sanidine orangeites differ texturally from all bona fide extrusive phlo-
gopite lamproites (see Mitchell and Bergman 1991) in that the latter contain abundant
microphenocrystal sanidine, diopside, leucite, and only minor olivine. Mineralogical and
petrographic similarities are closest with respect to some hypabyssal sanidine richterite
lamproites such as are found at Endlich Hill (Wyoming), Cancarix (Spain), and Mount
North (Australia). These rocks are characterized by intergrown groundmass sanidine and
richterite. However, they typically lack olivine, perovskite, spinel, and calcite and are
relatively rich in leucite. Mineral compositions provide further discriminants (see 2.5,
2.8). Further, all evolved orangeites are poor in shcherbakovite and wadeite relative to
such lamproites.
The presence of oikocrysts of Ti-phlogopite containing inclusions of (?)pseudo-
leucite in the Sover North and Besterskraal orangeites further emphasize the affinities of
~

("J
Figure 1.58. Macrocrystal kimberlite, Wesselton Mine. Mantled macrocrysts of Figure 1.S? Macrocrystal kimberlite, Wesselton Mine. Higher magnification
olivine (0) together with small subhedral primary groundmass olivines lacking BSE image of the groundmass of Figure 1.58 showing subhedral primary olivines
mantles set in an oxide-rich (white) serpentine calcite matrix. BSE image. (P) set in a matrix of spinel and perovskite (white), laths of phlogopite-kinoshi- ~
~
talite (gray), calcite (gray), and serpentine (black). ....
Figure 1.60. Groundmass of a serpentine calcite kimberlite, Lipa Pipe, Alakit field, Russia, showing euhedral
primary olivine and subhedral opaque spinels set in a mesostasis of serpentine (S) and calcite (C) (TL, FOV =2.25
mm).

Figure 1.61. Groundmass of an oxide-rich serpentine calcite kimberlite, De Beers Mine, showing subhedral
primary olivine (0) within a groundmass of subhedral opaque spinel and perovskite set in a uniform
calcite-serpentine mesostasis.

81
Figure 1.62. Groundmass of an oxide-rich serpentine calcite kimberlite, Michigan (U.S.A.), showing primary
olivine (0) with a mantle (M) containing rutile and spinel inclusions, together with abundant euhedral opaque
spinel and perovskite set in a uniform mesostasis of calcite and serpentine (TL, FOV = 2.25 mm).

Figure 1.63. Groundmass of a perovskite-rich serpentine calcite kimberlite, Wesselton Mine. Perovskites are
rounded crystals (P). Unresolved uniform matrix consists of monticellite, calcite, and serpentine. OM = olivine
macrocryst; 0 =primary olivine (TL, FOV 2.25 mm).

82
Figure 1.64. Groundmass of a monticellite calcite kimberlite. Pipe 200 (Lesotho), showing small rounded-to-
subhedral crystals of monticellite (high relief phase with fluid inclusions), rounded crystals of perovskite (P),
and subhedral opaque spinels set in a mesostasis of calcite (TL, FOY =2.25 rom).

Figure 1.65. Segregation-textured groundmass, Ham kimberlite, Somerset Island (Canada). Segregations
consist of resorbed calcite rhombs (high relief) set in a serpophitic serpentine matrix (S). Note the very high
concentration of opaque oxides in the silicate oxide portion of the groundmass and their absence in the
segregations (TL, FOY =2.25 rom).
83
Figure 1.66. Oxide-rich serpentine calcite kimberlite. Xi-Yu (China). with poikilitic laths of phlogopite-ki-
noshitalite (TL. FOY =2.25 mm).

Figure 1.67. Oxide-rich serpentine calcite kimberlite. Iron Mountain (Wyoming. U.S.A.) with poikilitic laths
of phlogopite-kinoshitalite (TL. FOY = I mm).

84
~
~
5
trJ
~

>
Z
t:l
o

~
trJ

Figure 1.69. Groundmass of a segregationary textured kimberlite, from the


Figure 1.68. Groundmass of unifonn-textured kimberlite, Wesselton Mine. Wesselton Mine, shown at the same scale as Figure 1.68. The groundmass
Anhedral macrocrystal and subhedral smaller primary olivines (dark gray) consists of irregularly shaped calcite segregations (C) within an oxide-rich
are set in an unresolved unifonn mixture of spinel and perovskite (white), (white) unifonn serpentine calcite matrix. Anhedral macrocrystal and subhedral
calcite, and serpentine. BSE image. primary olivines (dark gray) are also present. BSE image. ~
~

Figure 1.70. Monticellite kimberlite, Wesselton, Mine. Subhedral primary oli- Figure 1.71. Segregation-textured phlogopite kimberlite, Iron Mountain (Wyo- n
vines (dark gray) are set in a matrix consisting ofmonticellite (gray) and spinel ming, U.s.A.). Segregations consist of laths of phlogopite-kinoshitalite (P) set
(white). BSE image.
==
in a calcite matrix (dark gray). Similar laths crowded with spinel inclusions
(white) occur in the oxide-silicate-calcite uniform portions of the groundmass.
0= primary groundmass olivine. BSE image.
~
~
~
5==
to:!
Vl
>
~
o

~
to:!

~
Vl

Figure 1.72. Segregation-textured kimberlite. WesseIton Mine. Segregations Figure 1.73. Segregation-textured kimberlite. Wesselton Mine. The large segre-
consist of euhedral-to-rounded (resorbed) calcite crystals (C) set in a serpophitic gation consists of euhedral calcite (C) and laths of strongly wned phlogopite-ki-
serpentine matrix (S). The uniform portions of the groundmass consist of opaque noshitaIite set in a matrix of serpophitic serpentine (S). The remainder of the
oxides (white). phlogopite-kinoshitalite. calcite. and serpentine. BSE image. groundmass contains amoeboid calcite-rich segregations. atoll-textured spinels. and
laths of phlogopite-kinoshitalite set in a calcite serpentine mesostasis. BSE image. ~
gg

100 urn

("J

Figure 1.74. Poikilitic lath of phlogopite-kinoshitalite (P) enclosing atoll-tex- Figure 1.75. Poikilitic laths of phlogopite-kinoshitalite (P) enclosing spinels
tured spinels (white). Iron Mountain kimberlite (Wyoming. U.S.A). BSE image. and perovksites (white). Mesostasis is calcite (C). BSE image.
~...
KIMBERLITES AND ORANGEITES 89

Figure 1. 76. Atoll-textured spinels, Wesselton Mine kimberlite. This characteristic texture of many kimberlitic
spinels consists of cores of magnesiochromite-magnesian ulvospinel solid solution separated from thin rims of
Ti-magnetite by zones of serpentine and calcite. The texture is believed to be formed by the resorption of portions
of an original complexly mantled spinel during the later stages of crystallization of the groundmass. BSE image.

some orangeites with lamproites rather than kimberlites. However, these orangeites differ
from madupitic lamproites in that the latter are olivine lamproites which do not typically
contain sanidine or richterite.
Some evolved orangeites contain olivines which exhibit complex parallel growth (or
resorption) morphologies, e.g., Sover North, Bellsbank West Fissure (Tainton 1992). The
habit is similar to that of olivines found in some olivine lamproites. However, this
morphological feature cannot be used to assert that the rocks are lamproites because it
has no genetic significance, i.e., similar olivines are found in rocks derived from other
magmas types, e.g., melilitites (McIver 1981, Moore and Erlank 1979).
Evolved orangeites are taxonomically challenging rocks in that they cannot be
classified without recourse to detailed mineralogical (and geochemical) investigations.
They can only be unambiguously recognized as belonging to the orangeite clan by their
association with less-evolved consanguineous rocks. They demonstrate the problems of
90 CHAPTERl

identifying rocks by simple petrographic methods using only isolated specimens.


Richterite sanidine orangeites illustrate the mineralogical convergence of the orangeite
clan with rocks of the lamproite clan that results from differentiation of similar but
genetically different potassic peralkaline magmas. The relationships of orangeites to
lamproites are discussed further in Section 4.7.2.
In order to know an object. I must not know its external but
all its internal qualities
Ludwig Wittgenstein (1922)
Tractatus Logico·Phiiosophicus

MINERALOGY OF ORANGEITES

The mineralogy of orangeites has not been extensively studied as data for these rocks
have typically been incorporated with studies of kimberlites (sensu lato). Thus, previous
studies did not attempt to delineate mineral assemblages and compositional trends which
could be used to discriminate between each group of rocks. This chapter is based upon
many new analyses of minerals in archetypal kimberlites and orangeites. The study of the
latter is far from definitive as detailed studies have been made only of the major silicate
and oxide minerals and not all occurrences of orangeites have been examined. The minor
and accessory minerals require much further investigation, and the data presented here
should be regarded merely as a reconnaissance study.
The information presented in this chapter summarizes all that is currently known
about the mineralogy of orangeites and compares these data with the parageneses and
compositions of similar minerals in kimberlites and lamproites. The minerals, with the
exception of olivine, are discussed in their approximate sequence of crystallization.
Particular emphasis is given to mica, as this mineral dominates the mode of most
orangeites and exhibits extensive compositional variation.
Xenocrysts, including diamond, are not discussed, as this work is concerned only
with primary minerals. Megacrysts belonging to the Cr-poor suite, which are charac-
teristic of archetypal kimberlites, are discussed primarily in Sections 1.5 and 4.6.2, as
they are only rarely found in orangeite (see 1.8.8,2.2.3).
The locations of the orangeites referred to in the text are given in Figure 1.3.
Individual occurrences are described in Section 1.8.

2.1. MICA

2.1.1. Paragenesis
Mica is ubiquitous in orangeites. Typically, primary mica forms closely packed
mosaics of tabular-to-square cross section, pale-brown, weakly pleochroic microphe-
nocrysts (0.01-0.3 mm). Modal abundances (30-90 vol %) vary widely within and
between intrusions and primarily reflect variations in the macrocrystal olivine content.
Flow alignment is present in many occurrences. Microphenocrysts may be zoned from
colorless or pale brown cores, to margins exhibiting stronger brown or red-brown
pleochroism. Zoning may be continuous or discrete. Many of the microphenocrysts are
91
92 CHAPTER 2

mantled by groundmass, red, reversely pleochroic tetraferriphlogopite. Some microphe-


nocrysts consist of discrete cores containing abundant vermiform fluid inclusions and
colorless, inclusion-free mantles. Glomeroporphyritic aggregates of microphenocrysts
are common.
Groundmass micas occur as small (0.02-0.1 mm) strongly pleochroic brown-to-red-
brown subhedral crystals. Many comprise the outermost mantles or rims of microphe-
nocrysts, although discrete subhedral-to-euhedral crystals are commonly found
throughout the mesostasis. The outer reversely pleochroic tetraferriphlogopite-rich mar-
gins of the crystals, which are immersed in the mesostasis, may be euhedral and fresh or
ragged and resorbed. Unlike the microphenocrysts, groundmass micas typically contain
inclusions of spinel, apatite, and diopside.
Large (0.2-0.5 mm) poikilitic plates of groundmass mica are found only in evolved
orangeites, e.g., Sover North, Besterskraal. These are strongly pleochroic from yellowish
red to deep red-brown and may be simply twinned. Inclusions of earlier-formed crystals

AI

EAST
I
*----- ---- -~SID
I

-/
~I
~-:::----x~ / o~ ~ ~
1 00 0
0
0
0

/
70 .~
•• I
~ .~
,. 0
()) PHL ••
*,---... :. ~~~
.. ~--.-- -if - .... ANN
00

lao \ · ..
~ 75
it I ~ AI
, • I. 0 -

\ . :1
\ ••• • C!>/I-
• • • 0 0
\ •••• • ilin
0.1 EAST
\ ~" ••• I PHL
1~--------"7
• I .: I Mg L -.......... _ _- - - "
TFP TFA eT
F
• MICROPHENOCRYSTS \
AND GROUNDMASS
\
T
•"
.,.
.:J
~.

Mg t..=.===r:===:::;:====:;:::==~--*FP • •• ., • •
--:-• ....;.....,/----./..----...,/~---I.~ FeT
5 10 15 20 25 30 35 40 45

Figure 2.1. Compositions (atomsl22 oxygens) of micas from orangeites plotted in the ternary system AI-Mg-
Fer. Total Fe expressed as Fe2+. EAST = "eastonite." SID = siderophyllite. PHL = phlogopite. ANN = annite.
TFP = tetraferriphlogopite. TFA = tetraferriannite.
MINERAWGY OF ORANGEITES 93

Table 2.1. Representative Compositions of the Cores of Microphenocrystal Phlogopites of the


Extreme AI-Depletion Trenda
Wt% 2 3 4 5 6 7 8 9 lO
Si02 39.96 41.57 40.16 40.02 41.88 39.95 41.21 39.89 39.80 40.28
1102 1.85 1.73 2.08 2.04 0.82 3.25 1.55 2.84 1.53 1.80
AI 20 3 12.13 9.29 12.36 9.10 11.96 12.81 12.89 12.15 12.64 11.29
Cr203 1.34 0.29 1.24 0.11 0.92 0.30 0.48 0.16 1.47 0.31
FeOr 4.68 7.43 4.55 8.98 2.50 6.32 4.59 5.47 4.51 6.30
MoO 0.11 0.09 0.09 0.11 0.03 0.07 0.02 0.04 0.03 0.03
MgO 24.78 24.76 24.04 24.82 26.17 22.11 24.26 23.47 24.40 24.83
Na20 0.09 0.18 0.07 0.05 0.15 0.34 0.18 0.12 0.05 0.04
K20 lO.29 10.06 10.19 9.90 lO.93 lO.41 lO.31 10.25 lO.27 10.26
BaO 0.39 0.22 0.12 0.34 n.a o.a 0.09 0.52 0.09 0.16
NiO 0.02 0.07 0.17 0.07 0.13 0.29 0.19 0.14 0.09 0.15
--
95.64 95.69 95.07 95.54 95.49 95.85 95.77 95.05 94.88 95.45

Structural formulae based on 22 oxygens


Si 5.727 5.999 5.766 5.851 5.922 5.733 5.847 5.763 5.726 5.804
AI 2.049 1.580 2.091 1.568 1.993 2.167 2.155 2.069 2.143 1.917
Ti 0.199 0.188 0.225 0.224 0.087 0.351 0.165 0.309 0.166 0.195
Cr 0.152 0.033 0.141 0.013 0.103 0.034 0.054 0.018 0.167 0.035
Fe 0.561 0.897 0.546 1.098 0.296 0.759 0.545 0.661 0.543 0.759
Mo 0.013 0.011 0.011 0.014 0.004 0.009 0.002 0.005 0.004 0.006
Mg 5.294 5.326 5.144 5.409 5.516 4.729 5.130 5.054 5.232 5.333
Na 0.025 0.050 0.020 0.014 0.041 0.095 0.050 0.034 0.014 0.011
K 1.881 1.852 1.866 1.846 1.972 1.906 1.866 1.889 1.885 1.886
Ba 0.022 0.012 0.007 0.020 0.005 0.029 0.014 0.009
Ni 0.002 0.008 0.020 0.008 O.ot5 0.033 0.022 0.016 O.OlO 0.017
mg 0.904 0.856 0.904 0.831 0.949 0.862 0.904 0.884 0.906 0.875
aFeOr = total Fe expressed as FeO; n.a. = not analyzed; 1-2, Sover Mine; 3-4, Lace; 5-6, New Elands; 7-8, Swartruggens
level 6; 9-10, Finsch F4 and F7 intrusions, respectively. All data this work.

are characteristic. In these rocks mica also occurs as reaction mantles around olivine
macrocrysts.
The common macrocrysts (>0.5 mm) are irregular, rounded, distorted, and weakly-
pleochroic pale-brown phlogopites which are optically identical to the microphenocrysts.
These macrocrysts are interpreted (2.1.7) as having crystallized from orangeitic magmas
prior to the emplacement of their current host and are thus considered to be cogenetic
with the microphenocrysts.
Other, less common, cryptogenic macrocrysts include strongly-pleochroic very dark-
brown or dark-green biotites. These are commonly mantled by pale-brown phlogopite
which is optically and compositionally identical to phlogopite microphenocrysts and
macrocrysts.
Detailed descriptions of mica paragenesis in orangeites may be found in Smith et ai,
(1978), Skinner and Scott (1979), Mitchell and Meyer (1989a), and Tainton (1992). The
94 CBAPTER2

Table 2.2. Representative Compositions of Groundmass Micas of the Extreme AI-Depletion


Trenda
Wt% 2 3 4 5 6 7 8 9 10
Si02 40.34 41.65 40.15 40.31 39.72 38.66 37.90 40.07 40.98 39.69
Ti02 0.86 1.02 1.82 1.97 1.38 1.45 1.11 0.89 0.31 0.49
AI20 3 7.60 0.03 4.25 1.18 4.40 2.12 0.66 0.48 0.34 0.26
Cr203 0.09 0.07 0.08 0.07 0.02 0.07 0.07 n.d 0.10 0.13
FeOT 7.80 14.59 14.31 15.42 15.05 18.70 24.71 16.29 16.14 17.61
MnO 0.08 0.22 0.14 0.11 0.13 0.20 0.47 0.08 0.21 0.15
MgO 26.15 25.40 23.48 24.19 22.12 22.58 18.41 25.05 25.54 24.99
Na20 0.10 0.22 0.08 0.09 0.06 0.07 0.17 0.21 0.01 0.25
K20 9.33 9.96 9.95 9.61 9.63 9.98 9.26 10.12 9.14 9.67
BaO 0.11 0.29 0.41 0.31 0.33 0.25 0.19 0.22 n.a 0.59
NiO 0.10 0.19 0.15 0.09 0.12 0.14 0.22 0.11 0.01 0.04
92.56 93.64 94.82 93.35 92.96 94.22 93.17 93.52 92.78 93.87

Structural formulae based on 22 oxygens


Si 6.027 6.451 6.116 6.280 6.179 6.095 6.243 6.288 6.406 6.259
Al 1.338 0.006 0.763 0.217 0.807 0.395 0.128 0.089 0.063 0.058
Ti 0.097 0.119 0.209 0.231 0.161 0.172 0.138 0.105 0.036 0.048
Cr 0.011 0.009 0.009 0.009 0.003 0.009 0.010 0.012 0.016
Fe 0.975 1.889 1.823 2.009 1.958 2.469 3.404 2.138 2.110 2.322
Mn 0.010 0.029 0.018 0.Dl5 0.017 0.027 0.066 0.011 0.028 0.020
Mg 5.823 5.863 5.331 5.617 5.129 5.315 4.520 5.859 5.951 5.874
Na 0.029 0.066 0.024 0.027 0.018 0.021 0.054 0.064 0.003 0.076
K 1.778 1.968 1.933 1.910 1.911 2.010 1.946 2.026 1.823 1.945
Ba 0.064 0.018 0.025 0.019 0.020 0.016 0.012 0.014 0.037
Ni 0.012 0.024 0.018 0.011 0.015 0.D18 0.029 0.014 0.001 0.005
mg 0.857 0.756 0.745 0.737 0.724 0.683 0.570 0.733 0.738 0.717
"FeOr =total Fe expressed as FeO; n.d. =not detected; n.a. =not analyzed. 1-2, Finsch internal dike F4;~, Saltpeterpan;
5-7, Swanruggens level 6; 8, Lace; 9, Star; 10, New Elands. All data this work.

paragenesis of mica is discussed further in Section 1.10 and illustrated in Figures


1.26-1.54.

2.1.2. Composition of Primary Mica


Although of a reconnaissance nature, the study of nine random samples of orangeite
by Smith et al. (1978) served to establish the overall character of the micas present. The
principal findings were the recognition of relatively rare dark-colored Fe-rich mica
(termed type I) of unknown origin, and common microphenocrystal micas zoned from
phlogopite cores (termed type II) to tetraferriphlogopite margins. Subsequent reconnais-
sance studies of the Bellsbank dikes by Boctor and Boyd (1982), and detailed studies by
Skinner and Scott (1979) and Mitchell and Meyer (1989a) of the Swartruggens and New
MINERALOGY OF ORANGElTES 9S

Table 2.3. Representative Compositions of Micas from Besterskraal and Sover NorthQ
Wt% 2 3 4 5 6 7 8 9 10
Si02 40.89 40.68 41.34 40.59 40.99 40.80 39.58 40.06 39.43 40.01
TI02 4.92 5.34 5.66 8.09 6.79 6.23 1.76 1.49 2.58 1.29
AI20 3 6.08 6.25 5.57 7.33 5.78 3.86 13.32 7.38 12.28 7.97
Cr203 0.03 0.03 0.04 0.07 0.32 0.04 1.29 0.06 0.02 0.02
FeOT 9.39 9.78 10.18 8.04 12.92 12.25 4.72 13.52 7.37 13.59
MoO 0.06 0.07 0.13 0.09 0.15 0.14 0.10 0.24 0.14 0.28
MgO 20.94 20.70 20.57 20.06 17.49 20.69 23.95 21.43 22.97 21.47
Na20 0.74 0.69 0.72 0.52 0.72 0.51 0.05 0.09 0.07 0.07
K20 9.86 9.93 9.86 9.63 9.90 9.38 10.29 9.93 10.15 9.94
BaO 1.08 1.08 1.04 1.28 0.40 1.16 0.36 0.26 0.59 0.32
NiO 0.05 om 0.10 0.03 0.10 0.01 0.04 0.06 0.08 0.05
94.04 94.62 95.21 95.73 95.56 95.07 95.46 94.52 95.68 95.01

Structural fonnulae based 00 22 oxygeos


Si 6.138 6.085 6.152 5.943 6.133 6.149 5.676 6.057 5.713 6.085
AI 1.076 1.l04 0.977 1.265 1.019 0.686 2.251 1.315 2.097 1.267
Ti 0.555 0.601 0.633 0.891 0.764 0.706 0.189 0.169 0.281 0.148
Cr 0.004 0.004 0.005 0.008 0.038 0.005 0.146 0.007 0.002 0.002
Fe 1.179 1.224 1.267 0.984 1.617 1.544 0.566 1.710 0.893 1.728
Mo 0.008 0.009 0.016 0.011 0.019 0.018 0.012 0.031 0.017 0.036
Mg 4.686 4.615 4.562 4.378 3.900 4.648 5.119 4.830 4.965 4.867
Na 0.215 0.200 0.208 0.148 0.209 0.149 0.014 0.026 0.019 0.021
K 1.888 1.845 1.872 1.799 1.890 1.803 1.883 1.915 1.876 1.928
Ba 0.064 0.063 0.061 0.073 0.024 0.069 0.020 0.D15 0.034 0.019
Ni 0.006 0.008 0.012 0.004 0.010 0.012 0.005 0.007 0.009 0.006
mg 0.799 0.790 0.783 0.816 0.707 0.751 0.900 0.739 0.848 0.738
"FeOT =total Fe expressed as FeO; 1-3, groundmass poikilitic mica, Besterskraal; 4-6, groundmass poikilitic mica, Sover
North; 7-8 and 9-10, cores and rims of microphenocrystaJ mica, Sover North.

Elands occurrences, respectively, confirmed and amplified the observations of Smith


et at. (1978). Recently, Tainton (1992) presented compositional data for micas in
orangeites from the Barkly West region.
Discussion below of the compositional variation of mica in orangeites is based, in
part, upon these earlier studies, but principally upon approximately 1000 new analyses
of mica obtained during the preparation of this monograph. In this work, Smith et at. 's
(1978) type I micas are termed AI-biotite macrocrysts, and their type II micas are referred
to as primary phlogopite macrocrysts and microphenocrysts.
Figure 2.1 illustrates the major element compositional variation exhibited by mica
in orangeites. The data plotted are intended to be representative of the principal varieties
of mica present and not to reflect their relative abundances. The figure demonstrates that
primary microphenocrystal and groundmass micas are members of the phlogopite-
tetraferriphlogopite series. Compositions plot close to, and parallel to, this Fe2+-free
compositional join, as there is only minor solid solution toward biotite. Solid solution
96 CHAPTER 2

Table 2.4. Representative Compositions of Micas Exhibiting the Moderate AI-Depletion


Trenda
Wt% 2 3 4 5 6 7 8 9 10

SiOz 40.31 39.95 41.50 39.82 40.48 38.64 39.55 38.83 40.75 38.16
TiO z 1.91 3.18 1.45 6.06 1.63 2.23 1.74 2.69 3.09 3.95
Al z0 3 12.62 10.76 12.87 8.94 12.07 9.80 13.09 10.39 8.69 9.68
CrZ03 1.82 0.12 0.47 0.08 0.32 0.09 0.56 0.06 0.20 0.18
FeOT 4.38 9.14 4.70 11.53 4.80 14.60 4.68 14.07 16.82 19.09
MnO 0.03 0.06 0.03 0.09 0.15 0.25 0.04 0.21 0.26 0.26
MgO 24.29 21.72 23.77 18.78 24.22 18.22 24.73 18.40 15.93 13.41
NazO 0.18 0.17 0.20 0.25 0.08 0.12 0.10 0.14 0.07 0.02
KzO 10.23 9.88 10.18 9.47 10.26 9.65 10.05 10.01 9.57 9.37
BaO 0.22 0.36 0.14 0.44 0.33 0.37 0.40 0.09 n.d. n.d.
NiO 0.17 0.14 0.25 0.12 0.01 0.Q4 0.12 0.03 0.05 0.03

96.16 95.48 95.56 95.58 94.35 94.01 95.06 94.92 95.43 94.15

Structural formulae based on 22 oxygens


Si 5.725 5.835 5.897 5.883 5.857 5.913 5.686 5.847 6.140 5.920
Al 2.113 1.852 2.155 1.557 2.058 1.768 2.218 1.844 1.544 1.770
Ti 0.213 0.349 0.155 0.673 0.177 0.257 0.188 0.305 0.350 0.461
Cr 0.204 0.014 0.053 0.009 0.037 0.011 0.058 0.076 0.024 0.022
Fe 0.530 1.116 0.559 1.425 0.581 1.805 0.563 1.772 2.120 2.477
Mn 0.004 0.007 0.004 0.011 0.006 0.032 0.005 0.027 0.033 0.034
Mg 5.142 4.728 5.035 4.135 5.224 4.156 5.299 4.130 3.579 3.102
Na 0.050 0.048 0.055 0.072 0.022 0.036 0.028 0.041 0.021 0.006
K 1.853 1.841 1.845 1.785 1.894 1.884 1.843 1.923 1.840 1.855
Ba 0.012 0.021 0.008 0.026 0.019 0.022 0.023 0.005
Ni 0.019 0.016 0.029 0.016 0.001 0.017 0.014 0.004 0.006 0.004
mg 0.908 0.809 0.900 0.744 0.900 0.697 0.904 0.700 0.628 0.556
aFeOr = total Fe expressed as FeO; n.d. = not detected; 1-2,3-4, cores and rims of microphenocrystal micas, Voorspoed; 5--6,
7-8, cores and rims of microphenocrystal micas, Postmasburg PK37; 9-10, groundmass micas, Postmasburg PK36. All data
this work.

toward "eastonite," siderophyllite, tetraferriannite, and annite is not significant. Further


discussion of the solid solutions present is given in Section 2.1.8.
Aluminous micas are rarely present in orangeites as microxenoliths (see 2.1.3) or
dark-colored macrocrysts (see 2.1.4). The former are aluminous phlogopites, and the latter
aluminous biotites which are intermediate members of a solid solution between phlo-
gopite, "eastonite," and siderophyllite (Figure 2.1).
Orangeite micas exhibit extensive compositional variation with respect to their AI,
Fe, Ti, and Mg contents (Tables 2.1-2.4). Other elements show only minor variation and
are not present in significant quantities (see 2.1.2.3). Consequently, mica compositional
variation is best illustrated by plots of Ah03 versus Ti02 or FeaT.
MINERAWGY OF ORANGEITES 97

SWARTRUGGENS
LEVEL :3-4 LEVEL 6 LEVEL 7
14
• PHENOCRYSTS
13
o GROUNDMASS .f

12
1"\ .
:...,~.
II

t 10

-
~ 8
o
~ 7
..,
o 6 o
..!:'
<t o
5

3
0
2 00
0
0

0 0
0 0

2 3 4 2 3 4 2 3 4

Ti0 2 wt. % ~

Figure 2.2. AI203 versus Ti02 compositional variation of micas from the Swartruggens orangeites. Samples
were collected from different levels in the mine and represent distinct intrusions in this multiple dike system.

2.1.2.1. Ah03-TI02 Variation


Tables 2.1, 2.3, and 2.4, together with Figures 2.2 to 2.7, show that microphenocryst
cores typically contain from 13.5 to 9.0 wt % Ah03 and 1 to 3 wt % Ti02. Intra- and
inter-intrusion compositional differences are not significant. Thus, the geographically and
temporally isolated Swartruggens dikes contain microphenocrysts (Figure 2.2) of similar
composition to those in the younger Sover, Lace, and New Elands dikes (Figure 2.3).
Micas in samples from different levels of the Swartruggens Mine are identical in
composition (Figure 2.2). It is unlikely that these samples were derived from the same
dike, given the complexity of the multiple-dike system at Swartruggens 0.8.6), suggest-
ing that individual dikes cannot be distinguished on the basis of mica composition. Similar
98 CHAPTER 2

16 NEW ELANDS LACE SOVER

15

14 []
[]

13
,
12

t
II

10

~ 9
..:
~
8
If)
0
~ 7
«
6 o
o

5
o AI-BIOTITE MICROPHENOCRYSTS X MICROXENOLITHS
4 XENOCRYSTS + (03) o AI-BIOTITE
• MICROPHENOCRYSTS • (07) XENOCRYSTS
3 o GROUNDMASS GROUNDMASS • MICROPHENOCRYSTS

o 0 (02) [] GROUNDMASS
2
o
00
o

2 3 4 5 6 2 :3 4 5 2 3 4 5

Ti0 2 wt. % ~

Figure 2.3. Al203 versus 1102 compositional variation of micas from the New Elands. Lace. and Sover Mine
orangeites.

conclusions were reached by Skinner and Scott (1979) on the basis of fewer data.
Similarly, the different intrusions within the Finsch pipe cannot be distinguished on the
basis of their microphenocrystal mica compositions (Figure 2.4). Microphenocrysts in
the Sydney-on-Vaal dike (Figure 2.4) are relatively rich in Ti02 (2.4-4.9 wt %) and Ah03
(11.2-14.2 wt %).
Mitchell and Meyer (1989a) have demonstrated that pale-brown microphenocrysts
rich in fluid inclusions and found in the New Elands dikes, are slightly richer in Ti02
=
(1.0-3.5 wt %) and FeOT (4.0-7.0 wt %, mg 0.85-0.91) than coexisting colorless
= =
microphenocrysts (Ti02 0.15-2.5 wt %, FeOT 2.4-6.5 wt %, mg 0.87-0.96). =
Tables 2.2 and 2.4 together with Figures 2.2 to 2.4 demonstrate that, in general,
groundmass micas are very poor in Al relative to microphenocrystal micas. Zonation and
MINERALOGY OF ORANGElTES 99

FINSCH
14
?
8

12

t
~
10

..: 8 +
~

It)
0

j
~
<t 6

4 0


+
0 F2
+0 • F4
2
• • + F4E
F4ID
0 F7

Figure 2.4. Al203 versus Ti02 compositional •


variation of phenocrystal and groundmass micas 0
from the Finsch orangeite and phenocrystal micas l'
from the Sydney-on-Vaal orangeite. F-numbers 2 3 4 5
represent petrographically discrete intrusions at
Finsch. Ti0 2 Wt. % ---.

mantling trends of microphenocrystal mica typically involve depletion in Al at constant


or decreasing Ti contents. The outermost rims of microphenocrysts have compositions
similar to those of the most AI-depleted groundmass micas. Individual intrusions differ
only with respect to the degree of Al depletion and Ti content.
Similar zonation trends (not illustrated) are found in the Bellsbank (North Blow,
Southern Extension), Newlands, Roberts Victor, Star, and Saltpeterpan occurrences.
Ti-rich groundmass micas have been recognized only in highly evolved rocks from
Besterskraal (this work) and Sover North (this work, Tainton 1992). These are poikilitic
groundmass micas containing 4.5-9.0 wt % Ti02 (Table 2.3, Figure 2.5). Individual
100 CHAPTER 2

15
SOVER NORTH
• PHENOCRYST CORES AND RIMS
14

13

12

"

\
7 \

5
~++
4

LAMPROITE GROUNDMASS
3

o
MICAS
Leucite Hills
2 L_
I
I
1- -"I
I.- _...J
W. Kimberley ---_I

• ZONATION TREND

2 3 4 5 6 7 8 9 10

Figure 2.5. AI203 versus Ti02 compositional variation of micas from the Sover North and Besterskraal
evolved orangei tes. Data for phenocryst cores and rims, and for poikilitic groundmass micas, are from different
samples of the Sover North intrusion. Fields delineating compositional variation of lamproite micas from
Mitchell and Bergman (199\).
MINERALOGY OF ORANGElTES 101

14
POSTMASBURG
13

12 CORE-RIM I GROUNDMASS
• 0

t
II

10

If)
9
0
«'" 8

6
8ESTERSKRAAL~ SOVER.
NORTH
5

2 3 4 5 6 7 8 9

Figure 2.6. Al203 versus Ti02 compositional variation of micas from the Postmasburg sanidine-bearing
orangeite. Also shown are compositional fields for poikilitic groundmass micas from Sover North and
Besterskraal.

crystals are not zoned with respect to their Ti content, although distinct intergrain
compositional variation is evident.
The Sover North intrusion also contains rocks which are petrographically similar to
typical un evolved orangeites, such as occur at New Elands or Sover Mine. Microphe-
nocrystal micas in these rocks have compositions (Table 2.3) identical to those in other
orangeites and are zoned toward rims which are moderately depleted in Al at constant Ti
contents (Figure 2.5, Table 2.3). Rocks containing micas intermediate in composition
102 CHAPTER 2

14

13
VOORSPOED K 1/110 •
12 KI/III +
ZONATION TREND

i
II

10

It)
9
0
t\I
« 8

5
BESTERSKRAAL~ SOVER
NORTH

2 3 4 5 6 7 8 9

Figure 2.7. AIz03 versus Ti02 compositional variation of micas from two distinct intrusions in the sanidine-
bearing Voorspoed orangeite.

between the Ti-rich groundmass variety and the Ti-poor microphenocrystal type have not
yet been found at Sover North. However, a compositional trend of decreasing Al and
increasing Ti is present in micas from the Postmasburg and Voorspoed sanidine-bearing
orangeites (Table 2.4 analyses 4 and 5; Figures 2.6 and 2.7). The evolutionary trend of
these micas is toward the compositions of the Ti-rich poikilitic groundmass micas from
Sover North and Besterskraal, suggesting the latter micas might have originated by the
extensive fractional crystallization ofVoorspoed-type precursor magmas.
14
~
13

12 7 PHENOCRYSTS
SWARTRUGGENS (LEVEL 6) ~
II • PHENOCRYSTS g
o PHENOCRYST RIMS I GROUNDMASS
10
- - - ZONATION TREND ~
f 9 ~
ae LEVEL 3- 4
PHENOCRYSTS ~
~ 8
~
10
0 7
~ 7 RIMS I GROUND MASS
« 6

2 LEVEL 3 - 4

o o

3 4 5 6 7 8 9 10 II 12 13 14 15 16 17 18 19 20 21 22 23 24 25

- - FeO T wt. %
Figure 2.8. AI 20 3 versus FeOT compositional variation of micas from the Swartruggens orangeites. -e
104 CHAPTER 2

2.1.2.2. Ah03-FeOr Variation


Orangeite micas are strongly zoned with respect to their FeOr (total Fe expressed as
FeO) content. Increasing Fe is accompanied by Al depletion which may be either very
pronounced (13 to <1 wt % Ah03; Table 2.2) or relatively weak (13->8 wt % Ah03;
Table 2.4).
Figures 2.8-2.11 illustrate the FeOT enrichment and extreme Al depletion trend as
observed in the Swartruggens, Lace, New Elands, and Sover Mine occurrences, respec-
tively. The cores of microphenocrystal phlogopites in these examples are of relatively
uniform composition with respect to their Alz03 (10.0-13.5 wt %) and FeOT (2-6 wt %)
contents. Individual occurrences differ with regard to the range in FeOr in the cores and
the extent and nature of the weak zoning present.

13 +

12

r:
II

8
~


4
LACE
PHENOCRYSTS
+ (03) 0
2 • (07)
GROUNDMASS
o (02) o 0
0

2 3 4 5 6 7 8 9 10 II 12 13 14

FeOr Wt. % ~

Figure 2.9. Al203 versus FeOT compositional variation of micas from the Lace orangeite.
MINERAWGY OF ORANGElTES 105

15

14 o

13

12

t II

10

0~ 9
.:
~ 8

0'"
N 7
<i
6
o
5

4 NEW ELANDS
o AI - BIOTITE MACROCRYSTS
3
o Fe - POOR MACROCRYSTS
2 • PHENOCRYSTS
o GROUNDMASS o
- - - - - ZONING TREND

2 3 4 5 6 7 8 9 10 II 12 13 14 15 16 17 18 19

FeOT wt. % •

Figure 2.10. A1203 versus FeOr compositional variation of micas from the New Elands orangeite.

Figure 2.12 illustrates the complexity of zonation found in microphenocryst cores in


the New Elands orangeite. Within this intrusion, microphenocrysts exhibit normal and
reverse zoning with respect to Fe and increasing or decreasing Al content. Micas of
different zonation type are commonly juxtaposed within a single specimen. Figure 2.13
shows that similar complex zoning patterns exist with regard to Ti and Fe content.
Different intrusions in the Swartruggens dike system cannot be distinguished on the
basis of the composition of microphenocryst cores (Figure 2.8). Similarly, the Finsch F4
and F7 intrusions contain microphenocrystal micas of similar composition (Figure 2.14).
Groundmass poikilitic plates in Finsch F2 are rich in Fe and poor in Al relative to the
phenocrysts.
Increasing intensity of red pleochroism in microphenocryst rims and groundmass
plates reflects increasing Fe and decreasing Al contents. The core-rimlgroundmass
compositional trends, illustrated in Figures 2.8-2.11 and 2.14, culminate in the formation
of groundmass poikilitic micas containing very little Ah03 «2.0 wt %). Importantly, the
extreme Fe enrichment is not accompanied by a concomitant depletion in MgO (Table
2.1), and thus the mg numbers of even the most Fe-rich micas remain greater than 0.75.
106 CHAPTER 2

15

14

13

12

1 II

10

0~
9


.:
8
If)
0
.J:J 7
~
6

5
SOVER MINE
4 c AI - BIOTITE MACROCRYSTS
X MICROXENOLITHS
3
o Fe - POOR MACROCRYSTS
2 • PHENOCRYSTS
o GROUNDMASS o C§ 0
o
- - ZONING TREND

2 3 4 5 6 7 8 9 10 II 12 13 14 15 16 17 18 19

FeOT wt. % ~

Figure 2.11. AI203 versus FeOr compositional variation of micas from the Sover Mine orangeite.

This observation is interpreted to indicate that the AI-poor micas are tetraferriphlogopites
exhibiting only minor solid solution toward biotite (see 2.1.2).
Individual intrusions differ regarding the extent of Fe enrichment attained in the
groundmass mica, e.g., Swartruggens (level 6, 17-25 wt % FeOT; level 7, 18-20 wt %
FeOT; levels 3 and 4, 14.5-16.0wt % FeOT; see Figure 2.8), SoverMine (15.0-18.5 wt %
FeOT; see Figure 2.11), and Saltpeterpan (l5.~16.0 wt % FeOT). As all of these intrusions
contain microphenocrysts of similar composition, this observation is interpreted to
indicate that the post-intrusion crystallization history of each occurrence was slightly
different.
Figures 2.15-2.17 illustrate the extent of FeOT enrichment for the moderate-to-weak
AI-depletion trend observed in the Postmasburg, Voorspoed, and Sover North occur-
rences, respectively. Microphenocryst cores have compositions (Table 2.4) similar to
those of microphenocrysts in orangeites whose micas evolve along the extreme
AI-depletion trend. Compositional trends in each intrusion differ with respect to the
degree of Al depletion and Fe enrichment. A wide range of compositions is found even
13'0

1". M

If)

q.
«
11'5

I"' 0 I----N-EW--E-LA-N-O-S-""
M - Fe - POOR MACROCRYSTS
• • 0
10'5 CORE RIM
ZONATION TRENO

2'0 2'5 3'0 3'5 4'0 4'5 5'0 5'5 6'0

FeOT wt. % ~

Figure 2.12. Al203 versus FeOr compositional variation of macrocrystal and microphenocrystal micas from
the New Elands orangeites.

3 NEW ELANDS •
CORE TO MANTLE

t
ZONATION TREND
• •

~ •
...:
~ 2
0
N

l- •

3 4 5 6 7 8 9

FeOr Wt. % •
Figure 2.13. Ti02 versus FeOr compositional variation of microphenocrystai micas from the New Elands
orangeites.

107
188 CHAPTER 2

12
ca' FINSCH

r 10

-
~

If)
8

6
+

0
~
« F2
(;)

4 • F4 +
+ F4E

0
F4ID
F7
2 ~
CORE TO RIM
ZONATION TREND

2 4 6 8 10 12 14 16

FeOT wt. %
Figure 2.14. A1203 versus FeOr compositional variation of micas from the Hnsch orangeites. F-numbers refer
to petrographically distinct intrusions.

14 POSTMASBURG

r
13
CORE TO RIM
ZONATION K37
12 MICROPHENOCRYSTS
0~
.; II
~
If)
10
0
~
« 9

GROUNDMASS
8
K36

3 4 5 6 7 8 9 10 II 12 13 14 15 16 17 18 19

FeOT Wt. % ~

Figure 2.1S. Al203 versus FeOr compositional variation of micas from the Postmasburg sanidine-bearing
orangeites. K37 and K36 are different intrusions in the Postmasburg area.
MINERAWGY OF ORANGElTES 109

t
13 VOORSPOED
MICROPHENOCRYSTS
12 CORE TO RIM ZONATION

ae
...: II
~
fC)
010
~
«
9

8
3 4 5 6 7 8 9 10 II 12 13

FeOT wt. % •
Figure 2.16. Ah03 versus FeOr compositional variation of micas from the Voorspoed sanidine-bearing
orangeites.

within a single specimen, e.g., Sover North (Figure 2.17). The extent of Fe enrichment is
similar to that found in the AI-poor tetraferriphlogopites: however, this iron enrichment
is accompanied by magnesium depletion (Table 2.4), representing solid solution toward
biotite and tetraferriphlogopite (see 2.1.8)
Figure 2.17 shows that groundmass poikilitic plates in the Sover North evolved rocks
are moderately AI depleted, but not strongly enriched, in Fe (Table 2.3). Individual
crystals are typically not zoned, although inter-grain compositional variation is evident.
Similar compositions have been reported by Tainton (1992). Groundmass micas of these
compositions (Table 2.3, Figure 2.17) also occur in the Besterskraal orangeite.
Figure 2.17 also shows that the rims of some microphenocrysts in the Sover North
microporphyritic rocks are of identical composition, with respect to Fe, but not Ti (Figure
2.5), to the poikilitic groundmass micas in the evolved rocks.

2.1.2.3. Macrocrysts Versus Microphenocrysts


Weakly pleochroic phlogopite micas, classified as macrocrysts on a textural basis,
cannot in most instances be distinguished from typical microphenocrysts on the basis of
their AI, Fe, and n contents (Figure 2.18). Macrocrysts may be zonation- and/or mantle-
free, or mantled by mica having greater (Figure 2.12) or lesser FeOr than the cores.
Macrocrysts may represent either relatively large primary/cognate micas or
xenocrysts derived by the fragmentation of upper mantle-derived xenoliths. Unfortu-
nately, the compositions of many mantle-derived (Delaney et al. 1980, Dawson and Smith
1977) and microphenocrystal micas are similar, and it is not possible to distinguish
unambiguously the origin of the macrocrysts on the basis of their major or minor element
compositions.
14
[22l PHENOCRYSTS LEUCITE HILLS
13

12

t
II

10

*
..:
~
9

8
If)
0
N 7
<i
6

.
3
SOVER NORTH
PHENOCRYST CORES AND RIMS
2
+ GROUNDMASS
BESTERSKRAAL

2 3 4 5 6 7 8 9 10 II 12 13 14 15 16

Figure 2.17. AI203 versus FeOT compositional variation of micas from the Sover North and Besterskraal
orangeites. Data for phenocrystal and poikilitic groundmass micas from Sover North are from different samples.
Fields delineating lamproite mica compositional variation from Mitchell and Bergman (199\).

t
,
• MICROPHENOCRYSTS

• 7:

~.. • MACROCRYSTS

4".. :.f ,.· • •


13 -

:
.
~
.,;
~ 12 - *W..*'. •• •.....,..,:. • •
~

• • •••
• ...•
tet·, ... • •• • • •• • ••
If')
0
•••
"-
C\I
«
• •

IO-L--r-I-r-I~I--~I~ I I I I I I I I I
I 2 4 5 6 7

Ti0 2 Wt. %. FeOT wt. %


Figure 2.18. Ah03 versus Ti02 and FeOT compositional variation of macrocrysts and microphenocrystal
micas from the Swartruggens orangeites.

110
MINERALOGY OF ORANGEITES 111

Relatively Fe-poor «3.0 wt % FeOT), Cr-rich (0.5-1.2 wt % Cr203) macrocrysts,


which may represent xenocrysts of lherzolite-derived primary mica, appear to be absent
or very rare in orangeites. When present they comprise much less than 1% of the mica
population. Such micas are present in the New Elands (Figure 2.10) and Sover Mine
(Figure 2.11) mica populations.
Unfortunately, it is not possible to distinguish between relatively Fe-rich mantle-
derived xenocrysts and primary microphenocrysts on the basis oftheircomposition. This
observation suggests that some of the macrocryst population may actually be xenocrysts.
The problem of identification of xenocrystal and cognate mica is the same as that
encountered in the characterization of the olivine population (see 2.3.2).

2.1.2.4. Minor Elements


Of the minor elements occurring in orangeite micas only Cr, Ba, F, and rarely Na,
are present in significant amounts. Ni contents are uniformly low, being typically less
than 0.2 wt % NiO, and commonly less than 0.1 wt % NiO.
The Cr content of orangeite mica varies widely. Representative Cr contents are given
in Table 2.5. Cr203 contents greater than 2.0 wt % have not been found. Typically, Cr
contents decrease with increasing Fe content and groundmass tetraferriphlogopites
contain very low levels of Cr.
Compositional variations within macrocryst and microphenocryst populations are
complex and no simple trends are evident. Macrocrysts may be Cr rich (>1.0 Cr203) or
Cr poor «0.5 wt % Cr203). Commonly, the cores ofCr-poor macrocrysts are mantled by
Cr-rich (1.0-1.5 wt % Cr203) mica. Typically, adjacent microphenocrysts differ widely
with respect to their Cr contents. Figures 2.19-2.21 illustrate Cr-Ti variation in coexisting
microphenocrysts from the Lace orangeite (this work) and the New Elands and Star-
Burns dikes (Mitchell and Meyer 1989a). These figures show that Cr may decrease or
increase as Ti decreases or increases.
Microphenocrystal micas characteristically contain less than 1.0 wt % BaO and
commonly have less than 0.5 wt % BaO (Table 2.5). Groundmass micas exhibit a wide
range in Ba content and are typically enriched in Ba relative to microphenocrysts. Ba
contents do not exceed 2.0 wt % BaO, and only groundmass micas from Besterskraal
have consistently high inter-grain BaO contents. High Ba contents are not associated with
increased Al contents, demonstrating the absence of solid solution toward kinoshitalite
(see 2.1.9).
The majority of microphenocrystal and groundmass micas contain less than 0.5 wt %
Na20 (Table 2.5). Relatively high Na contents (Table 2.5) are found only in the poikilitic
groundmass micas from Sover North and Besterskraal. Na enrichment is positively
correlated with Ba enrichment in these occurrences only. In contrast, groundmass micas
from Lace have high Ba contents but are poor in Na (Table 2.5).
Little is known about the F content of orangeite micas. Data obtained in this study
show that orangeite micas typically contain less than 1.0 wt % F. In the Swartruggens
dikes, microphenocrysts contain from 0.61 to 0.14 wt % F and are zoned toward
decrea<;ing F contents from core to margin. Groundmass tetraferriphlogopites contain
0.38-0.12 wt % F. Diverse samples of Lace orangeite show no consistent zoning trend
and individual coexisting microphenocrysts may be zoned toward increasing or decreas-
112 CHAPTER 2

Table 2.5. Cr203, BaO and Na20 Contents (wt %) of Microphenocrystal and Groundrnass
Micas a
Occurrence Cr203 BaO Na20
Bellsbank Southern Extension
microphenocrysts 0.28 -1.72 0.10 - 0.39 n.d. -0.24
groundmass 0.05 - 0.81 n.d. -0.20 n.d. -0.20
Besterskraal
groundmass n.d. 0.95 - 1.37 0.54- 0.74
Finsch
microphenocrysts (F4) 0.15 - 1.58 0.08 - 0.58 n.d. -0.10
microphenocrysts (F7) 0.09 - 1.69 0.09 - 0.31 n.d. - 0.10
groundmass (F2A) 0.02 -0.36 0.03 - 0.83 n.d. - 0.13
groundmass (F4-ID) n.d. - 0.08 0.05 - 0.76 n.d. - 0.31
Lace
microphenocrysts 0.15 - 1.84 n.d. - 0.21 0.10 - 0.22
groundmass n.d. - 0.31 0.06 - 1.85 0.05 - 0.39
Makganyane
microphenocrysts 0.05 -1.78 n.d. - 0.61 0.05 - 0.13
New Elands
microphenocrysts 0.10 -1.02 n.d. - 1.04 n.d-O.20
groundmass n.d. -0.05 0.05-0.97 0.08 - 0.32
Postmasburg
microphenocrysts (K37) n.d. - 1.85 0.13 - 0.73 0.05 - 0.12
groundmass (K36) 0.20 -0.25 n.a. 0.03-0.07
Saltpeterpan
microphenocrysts 0.28 - 1.85 n.d.-0.68 0.07 - 0.22
groundmass 0.02 -0.30 0.14- 0.36 n.d. -0.29
SoverMine
microphenocrysts 0.12 - 2.39 n.d -0.43 n.d. -0.16
groundmass n.d. - 0.18 0.07 - 0.55 n.d. -0.12
SoverNorth
groundmass 0.01 -0.32 0.40 - 0.45 0.10-0.72
Swartruggens
microphenocrysts 0.09 -1.82 n.d. - 0.43 n.d. -0.25
groundmass n.d. - 0.10 0.15 - 1.09 0.07 -0.41
Sydney-on-Vaal
microphenocrysts 0.10-1.81 n.d. -0.65 n.d. - 0.41
Voorspoed
microphenocrysts n.d. - 1.83 0.05 - 0.44 n.d. -0.29
an.d. = not detected; n.a. = not analyzed.
0·5 1·0 1·5 2·0 2·5

Ti0 2 wt. % ..
Figure 2.19. Cf203 versus Ti02 compositional variation of microphenocrystal micas from the Lace orangeite.

1·5
NEW ELANDS
• 0 MACROCRYSTS

1 1·0
0 •


• •

MICROPHENOCRYSTS

FLUID INCLUSION - RICH


~ MICROPHENOCRYSTS

+=~ 0 GROUNDMASS

If)
0
0
N
~

U 0·5

0·5 1·0 1·5 2·0 2·5 3·0 3·5 4·0 4·5

Ti02 wt. % II

Figure 2.20. Cf203 versus Ti02 compositional variation of micas from the New Elands orangeite (after
Mitchell and Meyer 1989a).

113
114 CHAPTER 2

STAR DIKES
1·5
• BURNS
•• WYNANDSFONTEIN

t
0 NEW STAR
..
~
..: 1·0
~
rt)
0
(\J
~

0·5

0·5 1·0 1·5 2·0 2·5


Ti0 2 wt. % •
Figure 2.21. CT203 versus Ti02 compositional variation of micas from the Star dike system.

ing F from core to margin. F contents range from 0.5 to 1.09 wt %. Some groundmass
micas contain no detectable F, while others exhibit an inter-grain range of 1.5 to 0.13
wt % F. Microphenocrystal micas in the Sover Mine orangeite contain from 0.53 to 0.14
wt % F and are zoned toward decreasing F from core to margin. Groundmass tetrafer-
riphlogopites contain from 0.45 to 0.11 wt % F. At Bellsbank Southern Extension,
macrocrysts are slightly richer in F (0.6-0.3 wt %) than microphenocrysts (0.5-0.20 wt %
F). However, groundmass tetraferriphlogopites encompass this range and contain 0.6-0.2
wt % F. Data for Besterskraal and Sover North micas are not available.

2.1.2.5. Trace Elements


The trace element content of orangeite micas has not been studied in detail. Smith
et al. (1979) reported that microphenocrystal micas contain 90-820 ppm Rb20 and
470-5800 ppm BaO. KlRb ratios range from 112 to 1039, with the majority being
MINERALOGY OF ORANGEITES 115

100-300 (average 251). K/Ba ratios range from 16 to 199 and average 74. Allsopp and
Barrett (1975) report wide variations in the Rb (202-952 ppm) and Sr (55-281 ppm)
content of microphenocrystal micas from Roberts Victor and Swartruggens. K/Rb ratios
range from 100 to 160.

2.1.3. Aluminous Mica-Microxenoliths


Micas which are enriched in AI, relative to the majority of most macrocrystal and
microphenocrystal micas, occur as microxenoliths at Swartruggens. These micas contain
17.5-20.0 wt % Ah03 and are aluminous phlogopites exhibiting solid solution toward
"eastonite" (Figure 2.1). Each microxenolith contains micas which are distinct with
respect to their Ti and Fe contents (Table 2.6, Figure 2.22). The micas exhibit wide ranges
in Cr203 (0.21-3.3 wt %) and Ti02 (0.1-1.2 wt %) contents within and between xenoliths.
Single macrocrysts of these compositions, which might be derived by the disaggregation
of microxenoliths, have not yet been found at Swartruggens.

Table 2.6. Representative Compositions of Aluminous Phlogopitea


Wt% 2 3 4 5 6 7 8
Si02 38.66 38.56 38.31 36.32 33.93 37.67 35.6 37.38
n02 0.32 0.08 0.13 1.24 0.57 0.00 3.15 0.48
AI2D3 17.41 18.48 19.29 18.42 19.84 19.95 20.0 17.31
Cn03 3.25 1.53 0.80 0.34 0.21 3.77 0.40 2.48
FeOr 3.72 4.04 4.89 8.24 11.22 4.58 4.19 3.40
MnO 0.05 0.09 0.13 0.14 0.16 0.12 n.a. 0.09
MgO 22.92 23.13 22.18 20.62 17.19 21.09 21.5 22.51
Na20 0.35 0.18 0.24 0.30 0.40 0.59 0.43 n.a.
K20 9.89 10.12 10.12 9.84 9.79 9.88 9.45 10.91
BaO 0.08 0.01 0.07 0.51 0.19 0.09 n.a. n.a
NiO 0.04 0.05 0.05 0.05 0.03 0.0 n.a. n.a
96.69 96.64 96.21 96.02 93.53 97.74 94.72 94.69

Structural formulae based on 22 oxygens


Si 5.436 5.427 5.407 5.256 5.116 5.267 5.094 5.390
AI 2.885 3.065 3.209 3.142 3.526 3.287 3.373 2.942
n 0.034 0.008 0.014 0.135 0.065 0.339 0.052
Cr 0.361 0.170 0.089 0.039 0.025 0.417 0.045 0.284
Fe 0.437 0.476 0.577 0.997 1.416 0.536 0.501 0.410
Mn 0.006 0.011 0.016 0.017 0.020 0.014 0.011
Mg 4.804 4.852 4.667 4.448 3.863 4.395 4.585 4.838
Na 0.095 0.049 0.066 0.084 0.177 0.160 0.119
K 1.774 1.817 1.822 1.817 1.863 1.762 1.725 2.007
Ba 0.004 0.006 0.004 0.029 0.011 0.005
Ni 0.005 0.005 0.006 0.006 0.004 O.ot5
mg 0.917 0.911 0.890 0.817 0.732 0.891 0.901 0.921
QFeOr =total Fe expressed as FeO; n.a. =not analyzed. 1-5. aluminous phlogopites from microxenoliths. Swartruggens (this
work); 6, secondary phlogopite from harzburgite (Delaney et al. 1980); 7-8, aluminous phlogopites from microxenoliths,
Leucile Hills (Barton and van Bergen 1981, Mitchell and Bergman 1991, respectively).
116 CHAPTER 2

20
. . . MICROXENOLlTHS . . .

t
(j)
18
@ AI-BIOTITE
MACROCRYSTS

~
J
...: 16
~

SWARTRUGGENS
~
0
~
<[
14

12 MICROPHENOCRYSTS

2 4 5 6 7 8 9 10 II
- - FeOT wt. % •
Figure 2.2l. Compositional variation of aluminous phlogopites from three microxenoliths in the Swartruggens
orangeites.

The relationship of the microxenoliths to their host is unclear, but they are considered
to be more likely genetically related to orangeite than being xenoliths of mantle material.
This hypothesis is based upon two observations:
Primary phlogopites in mantle-derived Iherzolites are relatively poor in Ah03
(12.4-14.5 wt %; Delaney et al. 1980). Most secondary micas in lherzolites and harzbur-
gites also have relatively low Al contents, with the exception of a single mica from a
harzburgite which is similar in composition to some of the Swartruggens Fe-poor
microxenolithic micas (Table 2.6, anal. 6). Micas from MARID xenoliths contain only
8-13 wt % Ah03 (Dawson and Smith 1977). Thus, the low Al content of most mantle-
derived micas implies that derivation of the microxenoliths from such a source is unlikely.
Phlogopites of similar paragenesis and composition (Table 2.6, Figure 2.22) have
been described by Barton and van Bergen (1981) and Mitchell and Bergman (1991) from
the Hallock Butte, South Table Mountain, and Hatcher Mesa lamproites of the Leucite
Hills. These micas form a compositional continuum with phenocrystal micas oflower Al
content, and are thus interpreted to be high-pressure phenocrysts.
MINERALOGY OF ORANGEITES 117

The wide range in Fe contents (Table 2.6) of the microxenolith micas suggests that
orangeite magmas might undergo considerable high-pressure differentiation if the mi-
croxenoliths found at Swartruggens have a similar origin to that proposed for the Leucite
Hills mica-rich xenoliths.

2.1.4. Aluminous Biotite Macrocrysts


Representative compositions of AI- and Fe-rich mica macrocrysts are given in Table
2.7. Figure 2.1 shows that the micas are aluminous biotites whose compositions may be
regarded as solid solutions between phlogopite and siderophyllite. Although individual
macrocrysts are of uniform composition there is considerable intergrain compositional
variation (Figures 2.1, 2.3, 2.10, 2.11). Macrocrysts are markedly enriched in Fe and Al
(Figures 2.1, 2.10, 2.11) and slightly in Ti (Figure 2.3) relative to microphenocrysts. In
many instances the macrocrysts are mantled by pale-colored mica identical in composi-

Table 2.7. Representative Compositions of Aluminous Biotite Xenocrystsa


Wt% 2 3 4 5 6 7· 8 9 10
Si02 36.84 36.38 36.94 35.76 35.94 37.15 35.6 35.70 36.75 36.70
Ti02 2.47 3.03 3.75 3.09 2.05 4.28 3.2 3.2 5.65 1.81
AI 20 3 17.08 15.67 15.62 15.27 16.77 14.70 15.3 15.5 14.23 5.97
Cr203 0.07 0.04 0.04 0.09 n.d. 0.04 n.d. n.d. 0.04 n.d.
FeOT 15.68 17.60 13.38 18.84 17.71 17.16 18.9 21.7 16.11 12.91
MnO 0.15 0.13 0.03 0.25 0.13 0.25 n.a. n.a. 0.15 n.a.
MgO 12.88 12.27 14.94 10.63 12.39 12.30 10.5 9.9 12.70 16.21
Na20 0.62 0.65 0.17 0.31 0.26 0.25 0.3 0.1 0.12 0.27
K20 9.37 9.67 9.85 9.63 9.66 9.62 9.2 9.5 9.82 10.10
BaO 0.39 0.44 0.17 0.16 0.06 0.40 n.a. n.a. n.a. n.a.
NiO 0.06 0.08 0.04 0.05 0.02 0.03 n.a. n.a. n.a. n.a.
--
95.61 95.96 94.93 94.08 94.99 96.18 93.0 95.6 95.57 93.97

Structural formulae based on 22 oxygens


Si 5.518 5.511 5.521 5.555 5.474 5.589 5.567 5.564 5.533 5.530
AI 3.015 2.798 2.751 2.796 3.010 2.607 2.820 2.847 2.525 2.836
Ti 0.278 0.345 0.421 0.361 0.235 0.484 0.376 0.258 0.640 0.205
Cr 0.008 0.005 0.005 0.Q11 0.005
Fe 1.964 2.230 1.672 2.448 2.256 2.159 2.472 2.828 2.029 1.627
Mn 0.019 0.017 0.004 0.033 0.017 0.032 0.019
Mg 2.875 2.771 3.328 2.461 2.813 2.753 2.448 2.299 2.850 3.640
Na 0.180 0.191 0.049 0.093 0.077 0.073 0.091 0.030 0.035 0.079
K 1.790 1.869 1.878 1.908 1.877 1.846 1.835 1.889 1.886 1.941
Ba 0.023 0.026 0.010 0.008 0.004 0.024
Ni 0.007 0.009 0.005 0.006 0.002 0.004
mg 0.594 0.554 0.666 0.501 0.555 0.561 0.498 0.449 0.584 0.691
aFeOr =total Fe calculated as FeO; n.a. =not analyzed; n.d. =not determined. 1-2, Swartruggens; 3-4, Sover Mine; 5,
Bellsbank; 6, New Elands (l..{') this work); 7, Saltpeterpan; 8, Monteleo (Smith et al. 1978); 9, Postmasburg K35; 10, Newlands
(Tainton 1992).
118 CHAPTER 2

tion to microphenocrystal phlogopite (Figures 2.3, 2.1 0, 2.11; this work, Smith et al.
1978) and rarely by tetraferriphlogopite (Figure 2.10).
Macrocrysts exhibit wide intergrain compositional variation with respect to Ti and
Ba contents, e.g., New Elands (1.71-4.28 wt % Ti02, 0.4-0.6 wt % BaO), Swartruggens
(0.44-3.1 wt % Ti02, 0.05-0.44 wt % BaO), Sover Mine (3.1-3.90 wt % Ti02, 0.07-0.38
wt % BaO). Cr203 and Na20 contents are typically below 0.1 and 0.3 wt %, respectively.
The few data available (this work) suggest that the macrocrysts are F poor, e.g., Sover
Mine, <0.1-0.8 wt % F (11 analyses); Swartruggens, 0.1-0.3 wt % F (4 analyses).
Little is known of the trace element content of these micas. Smith et al. (1979) have
determined that they contain 240-{)30 ppm Rb20 (average of 5 analyses =470 ppm) and
1030-4320 ppm BaO (average of 5 analyses = 2670 ppm). KlRb ratios range from 137
to 348 (average 209) and KlBa from 20 to 88 (average 42).
The origins of the AI-biotite macrocrysts are unknown. Smith et al. (1978), while
recognizing that the micas may be simply xenocrysts, suggest derivation from the
"intrusion of another magma just prior to intrusion of the kimberlite." This magma was
suggested to be carbonatitic. However, Smith et al. (1978) did not present any evidence
to support this hypothesis, other than the observation that some micas in carbonatites are
of similar composition to the aluminous biotites.
Mitchell and Meyer (1989a), emphasizing the constant association of the macrocrysts
with orangeites, and the absence of polymineralic mantle-derived microxenoliths con-
taining mica of similar composition, suggested that the micas are unlikely to be simply
xenocrysts. The Al-biotites found in the Swartruggens Male lamprophyre (2.1.5) are Al
poor relative to the macrocrysts. Hence, derivation from such a lamprophyric source is
considered unlikely.
The absence of similar macrocrysts in archetypal kimberlites emplaced in the
Kaapvaal craton (and elsewhere) suggests the Al-biotites have a genetic affinity with
orangeite magmas or relationship to their source regions.
Determining the origin of the macrocrysts is analogous to the problem of the origin
of green Fe-rich pyroxenes in other alkaline magmas (see reviews by Duda and Schminke
1985, Bedard et al. 1988, Dobosi and Fodor 1992). Despite intensive investigation the
origins of these pyroxenes remain ambiguous. They are commonly interpreted, largely
on subjective criteria, to be high-pressure phenocrysts derived from differentiated batches
of magma.

2.1.5. Micas from the Swartmggens Male Lamprophyre


The cores of microphenocrystal micas in the Swartruggens Male lamprophyric dike
have similar Ah03 contents to microphenocryst cores in the Swartruggens and other
orangeites. The micas differ in being relatively richer in Ti02 and zoned primarily with
respect to FeOT content (Table 2.8). The continuous zonation trend is one of increasing
FeOT, decreasing MgO, and slightly decreasing Ah03 at essentially constant Ti02 content
(Figures 2.23, 2.24). The compositional trend is similar to that found for the Postmasburg
and Sover North micas (Figure 2.24) and represents an evolutionary trend from phlo-
gopite toward AI-biotite. The absence of tetraferriphlogopite in these rocks suggests that
MINERAWGY OF ORANGEITES 119

Table 2.8. Representative Compositions of Micas in the Swartruggens Male LamprophyreQ


2 3

Wt% C R C R C R
SiOz 38.68 37.56 38.34 36.38 38.16 36.00
TiOz 3.69 4.01 3.72 4.22 4.07 4.11
AIlO 11.96 11.98 12.23 10.75 12.49 9.74
CrZ0 3 0.02 n.d. n.d. n.d. 0.03 n.d.
FeOT 8.47 11.21 7.89 16.65 7.29 19.89
MnO 0.04 0.09 0.08 0.25 0.05 0.28
MgO 21.80 19.50 21.96 15.91 22.09 14.61
NazO 0.11 0.14 0.10 0.17 0.08 0.17
KzO 9.99 9.90 10.02 9.39 10.06 9.14
BaO 0.23 0.47 0.38 0.66 0.52 0.69
NiO 0.01 n.d. 0.04 n.d. n.d. n.d.
95.00 94.86 94.76 94.38 94.82 94.63
Structural formulae based on 22 oxygens.
Si 5.654 5.593 5.616 5.612 5.575 5.634
Al 2.061 2.104 2.113 1.955 2.152 1.798
Ti 0.406 0.449 0.410 0.489 0.447 0.484
Cr 0.002 0.004
Fe 1.036 1.397 0.967 2.149 0.891 2.605
Mn 0.005 0.011 0.010 0.033 0.006 0.037
Mg 4.753 4.332 4.799 3.661 4.815 3.411
Na 0.031 0.041 0.D28 0.051 0.023 0.052
K 1.864 1.882 1.874 1.849 1.877 1.826
Ba 0.406 0.028 0.022 0.040 0.030 0.042
Ni
mg 0.821 0.756 0.832 0.630 0.844 0.567
aFeOr = total Fe expressed as FeO; n.d. = not detected; C = core; R = rim.

they have no simple genetic relationship to the geographically associated Swartruggens


orangeites.
The micas are very poor in Cr203 «0.1 wt %) and NiO «0.05 wt %). BaD ranges
from 0.25 to 1.1 wt %, and F from 0.7 to 1.6 wt %. No consistent zonation trend is evident,
and individual crystals may exhibit trends of increasing or decreasing Ba and F from core
to margin. With respect to the minor elements only the low Cr contents of these micas
distinguishes them from typical orangeite microphenocrystal micas.

2.1.6. Summary of Mica Compositional Variation


Microphenocrystal phlogopites exhibit a limited range in composition with respect
to their AI, Fe, and Ti contents. Intra- and inter-intrusion composition differences are not
significant. These observations are interpreted to suggest that all of the microphenocrystal
micas are derived from compositionally similar parental magmas.
120 CHAPTER 2

15

SWARTRUGGENS

1
14 MALE LAMPROPHYRE

~ 13
..:
• .
~
12
a rt)

I&J
~
« l-
II I&J
(!)

z
~
a:
0
10

9
2 3 4 5 6
Ti0 2 wt. %
Figure 2.23. Al203 versus Ti02 compositional variation of microphenocrystal micas from the Swartruggens
Male lamprophyre dike.

The coexistence of microphenocrysts that are normally or reversely zoned with


respect to Ti, Cr, and Fe, demonstrates that the microphenocrystal mica assemblage could
not have crystallized in situ. The complex zoning and mantling patterns documented in
this work and by Mitchell and Meyer (1989a) can be best explained by assuming that the
majority of the microphenocrystal micas crystallized from several, slightly composition-
ally different, batches of parent magma. Microphenocrysts (and some macrocrysts)
crystallizing from each batch of magma were subsequently mixed, extracted from a
common magma chamber, and emplaced as a heterogeneous hybrid assemblage.
Although crystallization in a magma chamber is invoked here to explain the complex
zonation observed in the micas, it should be realized that orangeites do not appear to form
large long-lived continuously replenished fractionating magma chambers of the type
associated with some minettes or alkaline basaltic magmas.
An important consequence of the above hypothesis is that the microphenocrysts must
be a transported assemblage and their present high modal concentrations are the result of
MINERALOGY OF ORANGEITES 121

14

13

-
12 0

II ....

1
~
10


..:
7
If)
0
~ 6
«
5

3
SWARTRUGGENS
2 MALE LAMPROPHYRE

. CORE TO RIM
ZONATION
• 0
TRENDS

.
3 4 5 6 7 8 9 10 II 12 13 14 15 16 17 18 19 20

FeOT wt. %

Figure 2.24. AI203 versus FeOT compositional variation of microphenocrystal micas from the Swartruggens
Male lamprophyre dike.

differentiation processes. It follows from this conclusion that the whole rock major
element compositions oforangeites are far removed from those of their parental magmas
(Mitchell and Meyer 1989a; see 3.2).
Macrocrysts of similar composition to microphenocrysts are merely larger crystals
derived from the same sources as the microphenocrysts. Rare AI-rich phlogopite mi-
croxenoliths may represent high-pressure cumulates. Fe-poor mantle-derived mica
xenocrysts are relatively rare. Aluminous biotite macrocrysts are cryptogenic, but the
constant association with orangeite suggests a cognate relationship may be possible.
Subsequent to mixing and emplacement, each batch of crystal-laden magma under-
went crystallization in a particular pressure-temperature-oxygen fugacity regime. Dif-
ferences in these intensive parameters resulted in different compositional trends
developing in the groundmass micas. These and micas occurring as mantles upon
microphenocrysts, macrocrysts, and xenocrysts are considered to have crystallized in situ.
Two principal trends are evident:
122 CHAPTER 2

1. A tetraJerriphlogopite trend, characterized by extreme Al depletion coupled with


Fe enrichment at relatively constant Mg contents. Ti may decrease slightly or
remain constant.
2. A biotite trend, characterized by Fe enrichment and Mg depletion accompanied
by moderate Al depletion. Ti may decrease or increase slightly.
Groundmass micas following the biotite trend may be more common in occurrences
which differentiate to felsic residua, i.e., Postmasburg, Voorspoed. However, the associa-
tion is not universal, as micas from Makganyene follow the tetraferriphlogopite trend and
those from Sover North follow a range of compositional trends between the two extremes.
The origins of the trends are undoubtedly related to the oxygen fugacity of the
magma. The tetraferriphlogopite trend may represent mica evolution under relatively
oxidizing conditions with the bulk of the iron being present as Fe3+. The biotite trend may
represent crystallization under relatively reducing conditions, with iron being distributed
between Fe2+ and Fe3+. However, other factors, such as the peralkalinity of the magma,
playa role in determining the oxidation ratio of minerals crystallizing from the magma.
Unfortunately, the relative role of each factor cannot as yet be assessed.
Differences in oxygen fugacity may be caused by extensive crystallization, contami-
nation, and/or interaction of orangeite magmas with ground waters. Unfortunately,
discussion of the origins of fugacity differences must necessarily remain speculative, as
the nature of orangeite magmas and the physical conditions prevailing during their
crystallization are unknown.

2.1.7. Solid Solutions in Orangeite Mica


The nature of the solid solutions present in orangeite micas is difficult to assess as
most compositions have been determined by electron microprobe analysis. Water and
halogens are typically not determined; consequently, structural formulae are calculated
on the basis of 22 oxygens/formula unit. The actual Fe3+lFe2+ ratios of the micas are
unknown, but in some instances may be approximately evaluated by recalculation of their
compositions on a stoichiometric basis. However, calculation of ferric and ferrous iron
contents by the methods suggested by Droop (1987) is not always successful, as the micas
are commonly partially altered and/or contain less than 14 tetrahedral and octahedral
cations/formula unit. In many instances unrealistically high ferric iron contents result
from this recalculation procedure.
The majority of primary phenocrystal and groundmass micas contain insufficient Si
and Al to occupy all of the available tetrahedral sites. Consequently, Al vi is absent and,
within the overall compositional variation Si increases as Al iv decreases. Si commonly
exceeds the ideal value of six atoms of Si/formula unit (Figure 2.25) in the most evolved
micas. In common with lamproite micas, which exhibit similar compositional trends (see
2.1.9), it is suggested that the Al deficiency is. undoubtedly a direct reflection of the
peralkalinity of the parent magma. The tetrahedral site deficiency increases with differ-
entiation (Table 2.9) and is principally remedied by the entry of Fe3+to this site. For many
micas there is insufficient Fe3+, Ti, and Cr present to fill the tetrahedral sites (Table 2.9).
In such instances the remaining deficiency may be eliminated by the entry ofMg into the
tetrahedral site. The existence of Mi v has been proposed by Robert (1981) on the basis
MINERALOGY OF ORANGElTES 123

2'0

1'0

0.5 + SWARTRUGGENS - FINSCH


- LACE - BELLS BANK
• SOVER NORTH
o BESTERSKRAAL
• VOORSPOED - + + +
POSTMASBURG - MAKGANYENE

5'5 6'0 6'5

Si •
Figure 2.25. Aliv versus Si (atomsl22 oxygens) for orangeite micas. Field of lamproite micas from Mitchell
and Bergman (1991).

of the synthesis of MgiV-mica by Tateyama et al. (1974). The presence of excess


octahedral-site cations in the recalculated structural formulae (Table 2.9) would seem to
support this hypothesis.
Figure 2.26 illustrates the variation of Al with MgI(Mg + Fer) in orangeite micas and
shows clearly that micas evolve from phlogopite-rich phenocrysts to eithertetraferriphlo-
gopites (trend 1) or AI-deficient biotites (trend 2).
124 CHAPTER 2

Table 2.9. Representative Compositions of Swartruggens Groundmass Micaso


Wt% 2 3 4 5 6
Si02 40.38 40.01 39.56 40.43 39.93 39.04
Ti02 1.83 1.59 1.49 1.27 1.23 1.93
AI 20 3 9.74 7.23 4.99 2.31 \.31 0.62
Cr203 0.10 0.04 0.08 0.05 0.08 0.08
Fe203 1.18 4.57 8.64 9.18 11.85 12.67
FeO 6.29 7.72 6.30 8.79 8.63 8.24
MnO 0.09 0.14 0.16 0.15 0.20 0.34
MgO 24.41 23.14 23.35 22.44 22.32 22.25
NiO 0.11 0.03 0.14 0.08 0.02 0.12
Na20 0,07 0.08 0.08 0.\3 0.11 0.09
K20 10.24 10.02 9.89 9.65 9.79 9.62
BaO 0.72 0.53 0.27 0.24 0.33 0.68
95.16 95.10 94.55 94.72 95.79 95.67

Structural formulae based on 22 oxygens


Si 5.888 5.941 5.927 6.149 6.073 5.990
AI 1.674 1.265 0.881 0.414 0.235 0.112
Ti 0.201 0.178 0.168 0.145 0.141 0.223
Cr 0.012 0.005 0.009 0.006 0.010 0.010
Fe 3+ 0.130 0.510 0.974 1.051 \.357 1.462
Fe 2+ 0.767 0.959 0.789 1.118 1.097 1.057
Mn 0.011 0.0\8 0.020 0.0\9 0.026 0.044
Mg 5.305 5.121 5.214 5.087 5.060 5.088
Ni 0.013 0.004 0.017 0.0\4 0.002 0.0\5
Na 0.020 0.023 0.023 0.038 0.032 0.027
K 1.905 1.898 1.890 1.872 1.819 1.883
Ba 0.041 0.031 0.016 0.014 0.020 0.041
T 7.905 7.899 7.959 7.765 7.816 7.797
0 6.096 6.102 6.040 6.238 6.185 6.204
CAT 15.967 15.953 15.928 15.927 15.872 15.952
·Pt!203 and PeO calculated by the method of Droop (1987) on the basis of 14 tetrahedral (T) and octahedral (0) cationsl22
oxygens. CAT = cation total.

Figures 2.1 and 2.26, and the above discussion demonstrate that the majority of
orangeite micas are members of a solid solution series between phlogopite
[K2Mg6Si6Ah02o(OH)4], annite [K2F~+Si602o(OHh] and tetraferriphlogopite [K2Mg6-
Si6Fe~+02o(OH)4]. Trend 1 is primarily a solid solution between phlogopite and tetrafer-
riphlogopite, while trend 2 involves a three component solid solution between phlogopite,
annite and tetraferriphlogopite.
Some phenocrysts and macrocrysts are slightly enriched in Al and contain more than
two atoms AI/formula unit (Figure 2.26). Such micas may be considered as exhibiting
limited solid solution toward a hypothetical "eastonite" molecule
[K2MgsA1SisAh02o(OH)4]. (Livi and Veblin 1987 have shown that "eastonite" micas are
actually a mixture of serpentine and phlogopite.) Significant amounts of this molecule
MINERALOGY OF ORANGElTES 125

PHENOCRYST I
CORES ~

2'0 -t-_ _ _B_IO_T_IT_E_ _ _I--P_H_L__ ~++fj


+~+ •
PHL

~ :1'.......

0'
0' ~ A-
+
.1).
.+1...
+ +' +e+.
++ ~ ••• ~.
"2' '''' +RIMS + + ~ •••
••,••"
e"!,.
~
,k"'" + + ++ •• .,.
+ +

1 - -_ _ _ _ ---1.1 ;~:
+ VOORSPOED - POSTMASBURG -=... •RIMS
O SOVER NORTH
PHENOCRYST RIMS
• z).a'-~
-_.-11 •••
·elr.o.

.,1·.
o SOVER NORTH GROUNDMASS 0 . ,~.•
4fJ BESTERSKRAAL

. .- :
1'0 W GROUNDMASS 0 0 •
•• •
1--.-S=:-W:7:A:-::R:=T~RU~G:-::G:=E~NS=---":"""LA-:-C::".:E::-I ....
- FINSCH -BELLSBANK
- MAKGANYENE -
NEW ELANDS
i:
0'5
.cD

0'4 0'5 0'6 o·g 1'0

Figure 1.26. AI (atoms/22 oxygens) versus Mg/(Mg + FeT) for orangeite micas. FeT = total Fe expressed
as Fe2+. PHL = phlogopite. See text for explanation of evolutionary trends I and 2.

are present in aluminous micas in microxenoliths (Table 2.6). These micas have no
tetrahedral site deficiency and contain Alvi.
The Ti-rich micas from Besterskraal and Sover North plot on the tetraferriphlogopite
trends in Figures 2.25 and 2.26, suggesting that they have significant Fe3+ contents.
However, their ferric and ferrous iron contents cannot be evaluated as these micas contain
less than 14 tetrahedral and octahedral cations/formula unit. This cation deficiency is
probably due to the presence of vacancies in the octahedral sites introduced by the
126 CHAPTER 2

presence of Tivi. Hence, these micas are best regarded as solid solutions between
phlogopite, tetraferriphlogopite, and Ti-octahedral site-deficient phlogopite
[K2M~oriSi602o(OH)4] .

2.1.S. Mica in Kimberlites


The compositional characteristics of primary micas in archetypal kimberlites are insuf-
ficiently known because previous studies typically did not distinguish between these rocks
and orangeites. Hence, the phlogopite-tetraferriphlogopite evolutionary trend, shown in
Section 2.1.7 to be characteristic of orangeites, was considered to be typical of kimberlites
in general (Mitchell 1986). This work demonstrates that this conclusion is incorrect.

16 0 o PHENOCRYSTS - WESSELTON
KIMBERLITE

"SECONDARY - TEXTURED
MICA- XENOLITHS
15 0

i
0
+ 0

14

~
.,; 13
~

If)
0
~ 12
«

II

10

2 3 4
Ti0 2 wt. %
Figure 2.27. Ah03 versus Ti02 compositional variation of macrocrystal and megacrystal micas from
orangeites (this work) and kimberlites (Shee 1985. Dawson and Smith 1975. this work) compared with that of
primary and secondary textured micas in mantle-derived xenoliths (Dawson and Smith 1975. 1977. Carswell
1975. Delaney et al. 1980).
MINERALOGY OF ORANGEITES 127

2.1.8.1. Macrocrysts
Many micas in archetypal kimberlites are cryptogenic macrocrysts or megacrysts
whose composition is relatively well-characterized (Mitchell 1986, Farmer and Boettcher
1981, Dawson 1980, Dawson and Smith 1975). These micas are aluminous phlogopites
(10-15 wt % Ah03) which exhibit a wide range in Ti02 (0.5-5.5 wt %) and Cr203
(0.1-2.0 wt %). Individual micas are homogeneous, although there is wide intergrain
compositional variation.
Dawson and Smith (1975) proposed that the origins of megacrystal micas in South
African kimberlites may be deduced on the basis of their Cr, Ti, and Fe contents.
Megacrysts relatively rich in Cr203 (>0.5 wt %) and poor in Ti02 «0.6 wt %) and FeO
«3.7 wt %) are considered to be xenocrysts derived from phlogopite-bearing lherzolites.
Megacrysts relatively poor in Cr203 «0.5 wt %) and rich in Ti02 (0.6-2.0 wt %) and
FeO (>3.75 wt %) are interpreted to be phenocrysts. However, note that such empirical
studies might not universally valid, as the compositions of micas in the upper mantle in
other regions might be very different from those of the control group of primary and
secondary micas, initially used by Dawson and Smith (1975) to establish the composition
of mantle-derived xenocrystal mica.

o PHENOCRYSTS-
WESSELTON MEGA I MACROCRYSTS
KIMBERLITE • ORANGEITE
16 o +

t
SOMERSET ISLAND
KIMBERLITES
• WESSELTON
KIMBERLITE

14
~
..:
~
10
0
~ 12
<{

10

2 4 6 8
FeOr Wt. % •
Figure 2.28. Ah03 versus FeOT compositional variation of macrocrystal and megacrystal micas from
orangeites and kimberlites compared with that of primary and secondary textured micas in mantle-derived
xenoliths. Data sources as in Figure 2.27.
128 CHAPfER2

MEGA!
MACROCRYSTS
o

t
SOMERSET ISLAND
2'0
~ KIMBERLITES

0
SECONDARY - TEXTURED
0 MICA - XENOLITHS 0
1'5 0
~ 0
<gc9
..: 0
~
0
If)
0
0 1'0 0
N 0
~
0 0 0

0 0
0

0'5 oOO~
0
0

2 3 4 5

Ti02 wt. %
Figure 2.29. Cn03 versus n02 compositional variation of macrocrystal and megacrystal micas from
orangeites and kimberlites compared with that of primary and secondary textured micas in mantle-derived
xenoliths. Data sources as in Figure 2.27.

Recent studies (Mitchell 1986) of the mica macrocryst population in Somerset Island
kimberlites have shown that. within a given kimberlite province. individual kimberlites
are characterized by compositionally distinctive macrocryst suites. This observation
suggests that many of the macrocrysts may indeed by cognate.
Macrocrystal micas in kimberlites are very similar in major (Figures 2.27 and 2.28)
and minor (Figure 2.29) element composition to the least-evolved macrocrystlmicrophe-
nocrysts in orangeites. Thus macrocrystal mica compositions are considered to be oflittle
use in discriminating between kimberlites and orangeites.

2.1.8.2. Primary Micas


Few studies of the compositional variation of bonafide primary micas in kimberlites
have been undertaken and a characteristic evolutionary trend has not previously been
recognized. This work summarizes existing data and incorporates new compositional data
(approximately 350 analyses) for groundmass primary micas from phlogopite-bearing
archetypal kimberlites from Somerset Island. Colorado-Wyoming. Siberia, China, and
MINERALOGY OF ORANGElTES 129

Guinea. Because of the diversity of mica populations, individual occurrences are de-
scribed separately prior to summarizing the compositional trends.

2.1.B.2.a. Somerset Island Kimberlites. The Tunraq kimberlite has previously been
termed a micaceous kimberlite by Mitchell (1979). It contains large (1-5 mm) flow-
aligned, deformed phlogopite macrocrysts (Ah03 = 12.0-13.8 wt %, Ti02 = 1.6-5.5
wt %, FeOT =4.4-6.2 wt %, mg =0.87-0.92) but lacks groundmass micas. No compo-
sitional differences exist between micas in the different facies of this kimberlite. The
micas are unevolved, with respect to their Fe and Al contents, and are interpreted (this
work) to represent phenocrysts which have crystallized at depth. These have been
concentrated prior to intrusion.
In addition to rare phlogopite macrocrysts, the Elwin Bay kimberlite (Mitchell
1978a) contains common late-stage groundmass colorless micas (this work). These are
not distributed uniformly throughout the groundmass but occur, in association with
calcite, as discrete patches or segregations of interlocking small prisms and plates. Micas

Table 2.10. Representative Compositions of Groundmass Micas from the Elwin Bay and Jos
Kimberlitesa
Wt% 1 2 3 4 5 6 7 8 9 10
SiCh 39.59 37.73 35.14 32.65 30.07 35.38 32.65 29.11 27.19 26.55
n02 0.04 0.43 0.65 0.74 1.21 1.84 1.98 0.24 0.12 0.32
Ah03 14.47 17.29 18.22 19.42 19.59 17.54 19.63 17.27 17.69 21.77
CflO3 0.07 0.03 0.05 0.02 n.d. 0.37 0.33 0.12 n.d. 0.05
FeOr 4.23 2.44 2.49 2.62 2.55 4.51 3.44 3.12 1.99 2.27
MnO 0.08 0.01 0.01 0.02 n.d. 0.04 0.01 0.10 0.05 0.03
MgO 27.89 25.36 24.18 23.23 22.74 23.46 21.96 23.22 23.50 21.96
Na20 0.07 0.05 0.06 0.13 O.ll 0.02 0.07 0.07 0.07 0.09
K20 9.64 9.78 8.75 7.39 6.09 9.20 8.44 3.67 3.19 4.72
BaO 0.08 2.54 5.82 8.73 12.63 3.58 6.15 18.98 21.22 16.23
96.16 95.66 95.37 94.93 94.99 95.94 94.66 95.90 95.02 93.99

Structural formulae based on 22 oxygens


Si 5.572 5.358 5.168 4.925 4.673 5.153 4.892 4.695 4.498 4.287
Ti 0.004 0.046 0.072 0.084 0.141 0.202 0.223 0.029 0.015 0.039
Al 2.400 2.996 3.158 3.452 3.588 3.011 3.466 3.283 3.449 4.143
Cr 0.008 0.003 0.006 0.002 0.043 0.039 0.Q15 0.006
Fe 0.498 0.290 0.306 0.331 0.331 0.549 0.431 0.421 0.275 0.307
Mn 0.009 0.001 0.012 0.003 0.005 0.001 0.014 0.007 0.004
Mg 5.850 5.368 5.301 5.222 5.268 5.093 4.904 5.582 5.795 5.286
Na 0.019 0.014 0.017 0.038 0.033 0.006 0.020 0.022 0.023 0.028
K 1.731 1.772 1.642 1.422 1.207 1.709 1.622 0.755 0.673 0.972
Ba 0.004 0.141 0.335 0.516 0.769 0.204 0.361 1.200 1.376 1.027
mg 0.922 0.949 0.945 0.940 0.941 0.903 0.919 0.930 0.955 0.945
aFeOr = total Fe expressed as FeO; n.d. = not detected. CaO and NiO not detectable by electron microprobe. Compositions
1-5 Elwin Bay, 6-10 Jos. (All data this work.)
130 CHAPTER 2

within each segregation consist of Ba-free, AI-rich cores which are strongly zoned toward
Ba- and AI-rich margins. The intra- and intergrain zonation is irregular, and the texture is
interpreted to suggest reaction between previously formed Ba-free groundmass micas and
late-stage Ba-rich fluids. The micas contain <0.1-13.4 wt % BaO, 5.8-1004 wt % K20,
<0.05-1.2 wt % Ti02, and 2-5 wt % FeOT (Table 2.10). The compositional variation is
shown in Figures 2.30 and 2.31. BaO and Ah03 (13.3-19.7 wt %) contents vary inversely
with K20 and Si02 (42.1-29.7 wt %) indicating considerable solid solution from phlo-
gopite toward the brittle mica kinoshitalite [BaMg3AhShOlO(OHh]. The coupled sub-
stitution (K+,Si4+ = Ba2+,AI 3+) is illustrated in Figure 2.32 by the 1: 1 correspondence
between K and Ba. The micas are best regarded as barian phlogopites, as the number of
Ba atomsl11 oxygens does not exceed 0.5.

21
0

. ELWIN BAY

.' ... ..
20 0 JOS

I
ZONATION TREND
19

18

ae
-~
17

10
16
0
(\J

«
15
.. LEUCITE
HILLS
14 PRIMITIVE
MICAS
r-\---
I
13

2 3 4 5

Ti0 2 wt. % •
Figure 2.30. A1203 versus Ti02 compositional variation of groundmass micas in the Elwin Bay and Jos
kimberlites. Somerset Island (this work). Field of Leucite Hills lamproite primitive micas from Mitchell and
Bergman (1991).
MlNERAWGY OF ORANGElTES 131

20
o


0",

00

19

t
18 0

17 0
~
.,; 00
0
~ 0
16
It) 0
0 <:0 0
t\I
<t 0
15
o ELWIN BAY
o JOS
14 •
ZONATION TREND

13
LEUCITE HILLS ..............
PRIMITIVE OR4NG~~'
MICAS
PRIM EirE -
M'C~~/VE
2 4 5


3

FeOT wt. %
Figure 2.31. AI203 versus FeOr compositional variation of groundmass micas in the Elwin Bay and Jos
kimberlites. Somerset Island (this work). Field of Leucite Hills lamproite primitive micas from Mitchell and
Bergman (1991).

Ba-free or Ba-poor groundmass micas are rich in Ah03 (> 13 wt %) and poor in Cr203
«0.05) and Ti02 relative to macrocrystal micas «13 wt % Ah03, >1 wt % Ti02, >0.5
wt % Cr203, Figures 2.27-2.29).
Three varieties of mica occur in the Jos kimberlite (Mitchell and Meyer 1980). Type
A micas are yellow-brown, low-Ti02 (<2.0 wt %), low-Cr203 «0.2 wt %) phlogopites
[Mg/(Mg + Fer) = 0.93-0.87]. Type B phlogopites [Mg/(Mg + Fer) = 0.93-0.84] are
darker in color and richer in Ti02 (2.5-5.6 wt %). Significant intergrain compositional
variation is found with respect to Cr and Ti contents. There is a compositional gap between
type A and B micas with respect to Ti02.
--
1/1 0'4
E
o
c
0'2 • JOS
+ ELWIN BAY

0'2 0·4 0 ·6 0·8 1'0

Ba (atoms / II oxygens ) ~
Figure 2.32. Ba versus K (atomic) for groundmass micas from the Elwin Bay and los kimberIites. Somerset
Island.

Figure 2.33. Ba-rich micas belonging to the phlogopite-kinoshitalite serie8 occurring as discrete mantles on
earlier-formed groundmass Ba-poor phlogopites. Jos kimberlite. Somerset Island. Backscattered electron image
(500x).

132
MINERALOGY OF ORANGEITES 133

Type C micas are colorless, rich in Ah03 (15.0-21.8 wt %), and poor in Ti02 «2
wt %), Cr203 «0.5 wt %), and FeOT «4.5 wt %) relative to type A and type B micas.
Type C micas exhibit continuous compositional zonation with respect to Ba (Table 2.10,
Figure 2.32) and range from Ba-bearing «0.1 Ba/afu) phlogopite through barian phlo-
gopite (0.1-0.5 Ba/afu) to potassian kinoshitalite (>0.5 Ba/afu). Ba enrichment exceeds
that found in Elwin Bay micas and is in accord with the more-evolved character of the
Jos calcite kimberlite relative to the Elwin Bay monticellite kimberlite. Increasing Ba
contents are coupled with increasing Al and decreasing Si, Fe, and Ti contents (Figures
2.30 and 2.31). Hence, Mg/(Mg+FeT) ratios increase as the compositions evolve toward
kinoshitalite. Previously, type C micas were regarded, incorrectly, as chloritized "eastoni-
tic" micas (Mitchell and Meyer 1980, Mitchell 1986).
Type A micas may be mantled by type B or C micas. Type B micas have discrete type
C mantles. Type C micas also occur as poikilitic groundmass plates and small prismatic
crystals. Type C micas, replacing earlier formed micas, may occur as diffuse irregular
patches at crystal margins and along cleavages and fractures or as discrete epitaxial
mantles (Figure 2.33). Several growth periods of Ba-poor and Ba-rich mica have

Table 2.11. Representative Compositions of Groundmass Micas from the Iron Mountain
Kimberlitesa
Wt% 2 3 4 5 6 7 8 9 10
Si02 39.98 40.53 38.88 39.40 39.03 36.70 35.57 34.83 32.64 30.59
Ti0 2 0.74 2.00 1.96 0.18 0.57 0.86 0.91 1.27 1.52 1.91
AI20 3 12.90 11.53 11.84 12.21 14.69 15.37 15.63 15.34 16.13 17.91
Cr203 n.d. 0.22 0.12 0.03 0.06 0.05 n.d. 0.06 0.04 0.08
FeOr 4.06 4.34 4.01 4.50 3.16 3.67 2.97 3.20 3.65 3.80
MnO n.d. 0.04 0.03 0.09 0.06 0.03 n.d. 0.03 0.01 0.04
MgO 26.24 25.37 24.48 29.02 26.25 25.19 26.40 24.10 23.18 22.46
Na20 0.02 0.07 0.04 0.03 0.03 0.03 0.05 0.06 0.08 0.13
K20 9.29 9.83 9.55 9.24 10.28 8.69 8.36 7.71 7.28 6.52
BaO 1.08 1.91 3.22 1.15 2.07 4.74 6.50 8.54 10.19 11.24
94.31 95.84 94.13 95.85 96.20 95.33 96.39 95.14 94.72 94.68

Structural formulae based on 22 oxygens


Si 5.758 5.817 5.739 5.623 5.569 5.383 5.220 5.259 5.041 4.767
Ti 0.080 0.216 0.218 0.193 0.061 0.095 0.100 0.144 0.177 0.224
Al 2.190 1.951 2.060 2.054 2.471 2.657 2.704 2.729 2.936 3.289
Cr 0.025 0.014 0.034 0.007 0.006 0.007 0.005 0.010
Fe 0.489 0.521 0.495 0.537 0.377 0.450 0.365 0.404 0.471 0.495
Mn 0.005 0.004 0.011 0.007 0.004 0.004 0.001 0.005
Mg 5.633 5.428 5.386 6.173 5.583 5.507 5.775 5.424 5.336 5.216
Na 0.006 0.019 0.011 0.083 0.008 0.008 0.014 0.D18 0.024 0.039
K 1.707 1.800 1.798 1.682 1.871 1.626 1.565 1.485 1.434 1.296
Ba 0.060 0.107 0.186 0.064 0.116 0.272 0.374 0.505 0.617 0.686
mg 0.920 0.912 0.916 0.920 0.937 0.925 0.941 0.931 0.919 0.913
aFeOr =total Fe expressed as FeO; n.d. =not detected; Cao and NiO not detectable by electron microprobe. Compositions
1-3 Iron Mountain #7, 4-10 Iron Mountain #24 (All data this work.)
134 CHAPrER2

Figure 2.34. Groundmass laths of phlogopite in Iron Mountain (Wyoming) kimberlite showing diffuse patchy

t "..
replacement by Ba-rich mica. Backscattered electron image (800x).

1·0
'.~

~1-'"

"'
(/) 0'8
C
Q)
01
>.
)(
,
0 0'6
"
""
.......
""
""
(/) 0'4
E
0
+-
0
""
~
0 ·2 •

m
IRON MOUNTAIN
""
CHICKEN PARK
""
0 '2 0'4 0·6
"0 '8 1·0

Ba ( atoms / II oxygens) •
Figure 2.35. Ba versus K (atomic) for groundmass micas from the Iron Mountain (Wyoming) and Chicken
Park (Colorado) kimberlites.
MINERAWGY OF ORANGElTES 135

..•. ..
18 - \
~

17
-
-
o
Oo} •• •
-:.
- .... ,
0
0

·,
• • IRON MOUNTAIN

· t·~
16 - 0 CHICKEN PARK
- •

t
• ••
15 - •• •
• • j
-
14 - • LEUCITE HILLS
PRIMITIVE MICAS
..:
~
- r----I-,
I I
13 -
l
If)
o • • I
~
« - I

.. •
1
• I
12 - \ I
I
• •• I • 0
- • •
I
II -
• ,. 0
I 0
0

-

~----_Oi
0
10 - ORANGEITE
PRIMITIVE COO 0
- MICAS

0

9 -
I I
2 4

Ti02 wt. % •

Figure 2.36. Al203 versus Ti02 compositional variation of groundmass micas from the Iron Mountain
(Wyoming) and Chicken Park (Colorado) kimberlites. Field of compositions of Leucite Hills primitive
lamproite micas from Mitchell and Bergman (199\).
136 CHAPTER 2

occurred, as inferred from the common presence of oscillatory-zoned crystals. Mica types
A and B have not crystallized in situ and are interpreted as transported hybrid microphe-
nocrystal assemblages.

2.1.B.2.h. C%rado-»yoming Kimberlites. The Iron Mountain kimberlites of


Wyoming (Smith 1977, McCallum et al. 1975) contain abundant poikilitic plates (Figure
2.34) of colorless groundmass micas that are continuously, irregularly-zoned from
Ba-poor cores to Ba-rich margins (Table 2.11, Figure 2.35). The compositional variation
is from Ba-bearing phlogopite to barian phlogopite. Increasing BaO (1.2-11.3 wt %) is
accompanied by increasing Ah03 (11.2-18.0 wt %). Ti02 and FeOT contents initially
decline then increase slightly as Al increases (Figures 2.36 and 2.37). Individual intrusions

·.tI
.
18
• • e•

0-:.•• • • IRON MOUNTAIN


17 ••• o CHICKEN PARK
• ti •
16 ..o
-,
.
• :''0 •
.. :

f
~
15
• •• •
•• •
• •
..: 14 •
~

",
r-------,

'I~'·.
o 13
~ LEUCITE

.0
c:[
HILLS
12 PRIMITIVE :\. ~
MICAS I .
I • ~ 0
1 · 0
I • 0 0
II
~
L________ : ORANGEITE
PRIMITIVE 0 0 0

MICAS o
10
00 0
o
9

2 3 4 5 6 7 8 9 10

Figure 2.37. AI203 versus FeOT compositional variation of groundmass micas from the Iron Mountain
(Wyoming) and Chicken Park (Colorado) kimberlites. Composition field of Leucite Hills primitive lamproite
micas from Mitchell and Bergman (1991).
MINERAWGY OF ORANGEITES 137

Table 2.12. Representative Compositions of Groundmass Micas from the Chicken Park
Kimberlitea
Wt% 2 3 4 5 6 7
Si02 39.29 37.38 39.91 39.19 37.99 45.32 42.42
Ti0 2 0.51 0.68 3.53 2.79 3.82 0.46 0.57
AI 20 3 16.18 15.84 9.67 10.69 11.79 0.78 0.11
Cr203 0.02 n.d. 0.02 0.02 0.06 n.d. n.d.
FeOT 3.05 4.98 10.67 8.05 8.34 11.45 14.82
MnO n.d. 0.07 0.16 0.05 0.15 0.09 n.d.
MgO 25.72 26.65 21.62 23.51 22.07 25.79 26.19
Na20 0.26 0.14 0.02 0.15 0.19 0.26 0.12
K 20 9.99 9.62 10.18 9.61 9.62 9.90 9.71
BaO n.d. n.d. 0.37 1.21 1.74 n.d. n.d.
95.02 95.36 96.15 95.22 95.77 94.05 93.94

Structural formulae based on 22 oxygens


Si 5.559 5.340 5.848 5.745 5.585 6.766 6.497
Ti 0.054 0.073 0.389 0.308 0.422 0.052 0.066
AI 2.698 2.667 1.699 1.847 2.043 0.137 0.020
Cr 0.002 0.002 0.002 0.007
Fe 0.361 0.595 1.307 0.987 1.025 1.430 1.398
Mn 0.009 0.020 0.006 0.019 0.011
Mg 5.424 5.675 4.722 5.137 4.836 5.739 5.979
Na 0.071 0.039 0.006 0.043 0.054 0.075 0.036
K 1.803 1.753 1.903 1.797 1.804 1.885 1.897
Ba 0.016 0.070 0.100
mg 0.938 0.905 0.783 0.839 0.825 0.801 0.759
"FeOT = total Fe expressed as FeO; n.d.= not detected. CaO and NiO not detectable by electron microprobe. Compositions
1-2 Chicken Park I; 3-8 Chicken Park 3; Compositions 6-7 are tetraferriphlogopites. (All data this work.)

differ with respect to the degree of Ba enrichment, e.g., Iron Mountain 7 micas contain
only 1.2-3.3 wt % BaO, while Iron Mountain 25 micas have 1.2-11.3 wt % BaO.
The Chicken Park kimberlite, Colorado (McCallum 1989), contains abundant
strongly pleochroic brown, poikilitic groundmass phlogopites. The majority of these
micas (Table 2.12) are rich in FeOT (3.0-10.1 wt %) and Ti02 (0.6-3.7 wt %) relative to
Iron Mountain (Figures 2.36 and 2.37) and other kimberlite groundmass micas. In some
instances narrow mantles of Fe-poor tetraferriphlogopite (FeOT = 10.7-14.8 wt%, Ti02
< 1.0 wt %, BaO <0.5 wt %) are developed upon these micas. The groundmass phlogopites
are BaO poor «0.2-2.0 wt %) with those richest in AI having the lowest Ba and Fe
contents (Table 2.12). Chicken Park 1 contains micas that are BaO poor «0.5 wt %)
relative to those from Chicken Park 3 (BaO> 1.0 wt %).

2.1.B.2.c. Guinea Kimberlites. The Antochka and Bounoudou kimberlites of the


Guinea kimberlite province (Bardet 1974) contain abundant colorless and pale yellow
laths, and poikilitic plates of groundmass mica, respectively.
138 CHAPTER 2

Table 2.13. Representative Comparisons of Groundmass Micas from Guinea Kimberlitesa


Wt% 2 3 4 5 6 7 8 9 10
Si02 38.52 38.69 37.35 42.83 39.82 35.49 39.89 35.43 35.75 34.42
1102 1.97 2.29 2.98 0.18 3.45 1.79 3.84 1.66 1.53 1.82
Ah0 3 13.82 14.07 13.37 2.69 13.71 17.25 13.51 18.17 16.25 18.38
Cr203 n.d. n.d. 0.04 0.13 1.58 0.08 1.74 0.07 0.11 0.16
FeOr 6.39 6.19 7.79 12.80 4.50 5.99 4.67 4.78 4.05 3.81
MnO 0.11 0.11 0.15 0.11 0.03 0.11 0.03 0.02 0.09 0.06
MgO 24.15 23.51 22.63 28.36 22.79 22.41 22.44 23.49 25.04 22.98
Na20 0.20 0.16 0.15 0.22 0.07 0.06 n.d. n.d. 0.04 0.10
K20 9.91 9.53 9.36 9.30 10.44 9.54 10.40 9.20 8.45 8.82
BaO 0.78 1.46 2.05 0.07 n.d. 2.57 n.d. 2.71 3.67 5.10
NiO 0.04 0.05 n.d. n.d. 0.23 n.d. 0.18 n.d. n.d. n.d.
--
95.89 96.06 95.87 94.63 96.62 95.29 96.70 95.53 94.98 95.65

Structural formulae based on 22 oxygens


Si 5.526 5.569 5.472 6.278 5.630 5.208 5.638 5.144 5.233 5.066
11 0.214 0.248 0.328 0.019 0.367 0.198 0.408 0.181 0.168 0.201
AI 2.356 2.387 2.309 0.465 2.284 2.983 2.250 3.109 2.803 3.188
Cr 0.005 0.Ql5 0.018 0.009 0.194 0.008 0.013 0.019
Fe 0.773 0.745 0.955 1.569 0.532 0.735 0.552 0.580 0.496 0.469
Mn 0.014 0.013 0.019 0.014 0.004 0.014 0.004 0.003 0.011 0.008
Mg 5.207 5.044 4.942 6.196 4.802 4.902 4.727 5.084 5.463 5.041
Na 0.056 0.045 0.043 0.063 0.019 0.017 0.011 0.029
K 1.829 1.749 1.749 1.739 1.883 1.786 1.875 1.704 1.578 1.656
Ba 0.044 0.082 0.118 0.004 0.148 0.154 0.2ll 0.294
Ni 0.005 0.006 0.026 0.020
mg 0.871 0.871 0.838 0.798 0.900 0.869 0.895 0.896 0.917 0.915
DFeOr= total Fe expressed as FeO; n.d. = not detected. CaO not detectable by electron microprobe. Compositions 1-4 Antochka;
5-10 Bounoudou. Compositions 5 and 7, and 6--8 are cores and rims, respectively. All data this work.

Microphenocrystal micas in the Antochka kimberlite (Table 2.13) contain 12.0-14.1


wt % A}z03, 6.4-8.5 wt % FeOT, and 0.2-2.1 wt % BaO. Rarely, they exhibit thin rims
of AI-poor tetraferriphlogopite-Iike mica (Table 2.13, anal. 4). Individual grains are
homogeneous, although considerable intergrain compositional variation is evident (Fig-
ures 2.38 and 2.39).
Groundmass plates and laths of mica in the Bounoudou kimberlites (Table 2.13,
Figures 2.38 and 2.39) are, in contrast, rich in Ah03 (15.3-18.4 wt %) and BaO (0.1-5.7
wt %) (Figure 2.40). FeOT contents (3.~.8 wt %) are relatively low. The micas are
continuously irregularly zoned with respect to their Ba contents (Figure 2.41). Although
the majority of the cores are poor in Ba and Al relative to crystal margins, the reverse
situation may also be found. The overall compositional trend is one of increasing Ba, AI,
and Fe coupled with decreasing Ti and Si (Table 2.13).

2.1.B.2.d. Chinese Kimberlites (Shandong Province). The Xi-Yu and Shengli (a.k.a.
Changma) kimberlites of the Shandong province (Dobbs et al. 1994) contain abundant
MINERAWGY OF ORANGEITES 139

- .• • • •

18 - • • •
••
- • '.'
17 - . • "• •

- •
• ••
16 -

i
- • BOUNOUDOU •
0 ANTOCHKA •
15 -
- •
ae 14 -
0
0 .
~
~ - .'
r--O--~ •
It) I OJ
0 13 -
~
« -
I
I
0
0 LEUCITE
I

....
12 - I 0 HILLS
I PRIMITIVE
I MICAS
- I
I
II - I 0
IL... _ _ _ _ _ _
-
10 - ORANGEITE
0
.... PRIMITIVE
MICAS
Figure 2.38. AI203 versus Ti02 -
compositional variation of ground·
mass micas from the Bounoudou and 9 -

..
Antochka kimberlites (Guinea). Com- I I I I
positional field of Leucite Hills primi- 2 3 4
tive lamproite micas from Mitchell
and Bergman (1991). Ti0 2 Wt. %
140 CHAPl'ER2

..
• • •
18

• •
•••• • •
• • • • ••
17 ••

o· • • • BOUNOUDOU
16
o ANTOCHKA

1 15
••

~
14 0 •
.: ••• 0
~ r------,I 0
0
If) 13
0
C\I LEUCITE
<i HILLS
12 PRIMITIVE
MICAS
ORANGEITE 0
II PRIMITIVE 0
L _______
MICAS

10
0

2 3 4 5 6 7 8

FeOT wt. %

Figure 2.39. AI203 versus FeOr compositional variation of groundmass micas from the Bounoudou and
Antochka micas (Guinea). Compositional field of Leucite Hills primitive lamproite micas from Mitchell and
Bergman (199\).

colorless poikilitic groundmass micas. These are strongly continuously or complexly


oscillatory zoned (Figure 2.42), Ah03-rich (14.9-18.0 wt %), barlan (1.0-12.20 wt %
BaO) phlogopites (Table 2.14, Figures 2.40, 2.43, 2.44). Fe and Ti contents do not vary
significantly with Al and Ba contents. Tetraferriphlogopites are absent. Zonation patterns
appear to be random and the cores of coexisting crystals mayor may not consist ofBa-rich
phlogopite (Figure 2.42).

2.1.B.2.e. Namibia. In Namibian kimberlites, ground mass phlogopite forms small


«0.1 mm) colorless equant grains which include previously formed spinels and
Figure 2.40. Ba versus K (atomic) for groundmass micas from Guinean, Chinese, and Namibian kimberlites.

perovskites (Spriggs 1988). The micas are Ah03-rich (14.0-18.4 wt %), Ti02-poor
(0.8-1.6 wt %), barian (1.3-11.5 wt % BaO) phlogopites (Figures 2.40, 2.43, 2.44).
Spriggs (1988) noted that all of the crystals were continuously-zoned from Ba-rich cores
to Ba-poor rims and suggested that the onset of groundmass mica crystallization resulted
in rapid depletion of Ba in residual liquids, due to very high crystal-liquid distribution
coefficients for Ba. However, evidence from other kimberlites (this work) suggests that
there is no such simple trend of Ba depletion with crystallization. Spriggs (1988) was the
first to document the existence of Ba-rich groundmass micas in kimberlites, although he
did not recognize the solid solution trend toward kinoshitalite.

2.1.8.2.f. Orroroo. The majority of micas found in the Orroroo kimberlite, South
Australia (Scott Smith et al. 1984) are poikilitic colorless AI-rich phlogopites mantled by
discrete thin rims of red tetraferriphlogopite. The aluminous micas are weakly zoned with
Ti02 (1-2 wt %), BaO (0.8-3.9 wt %) and Ah03 (15.4-13.2 wt %) decreasing from core
to rim. The micas contain 0.9-2.2 wt % F. Their compositional variation is illustrated in
Figure 2.41. Complex compositional zoning in groundmass micas from the Bounoudou kimberlite (Guinea).
Light gray regions are enriched in barium. Backscattered electron image (500x).

Figure 2.42. Extremely complex oscillatory zoning in groundmass micas from the Xi-Yu kimberlite (China).
Light gray regions are enriched in barium. Backscattered electron image (500x).
142
MINERALOGY OF ORANGElTES 143

Table 2.14. Representative Compositions of Groundmass Micas from the Xi-Yu and Shengli
Kimberlites Q

Wt% 2 3 4 5 6 7 8 9 10
Si02 38.31 37.85 36.36 34.26 32.81 33.24 32.64 32.58 31.69 29.57
Ti02 0.81 0.76 1.21 0.92 1.28 0.91 0.88 1.24 1.08 1.20
Ah03 15.35 14.82 14.88 16.08 16.85 17.44 17.99 18.03 18.21 17.95
Cr203 n.d. n.d. 0.09 0.05 n.d. n.d. 0.07 n.d. n.d. n.d.
FeOr 3.72 3.86 3.73 3.69 3.77 3.68 3.67 3.25 3.25 4.22
MnO 0.04 0.03 n.d. n.d. 0.06 0.06 n.d. 0.03 n.d. 0.02
MgO 25.80 25.48 24.68 24.27 24.31 24.06 23.67 23.72 23.27 22.78
Na20 0.11 0.12 0.10 0.15 0.19 n.d. 0.14 0.06 0.04 0.06
K20 10.29 9.84 8.45 7.98 7.16 7.80 7.63 7.61 7.06 5.96
BaO 0.98 2.71 5.59 7.86 9.75 7.69 8.34 9.02 10.61 12.58
95.41 95.47 95.09 95.18 96.18 94.88 95.03 95.54 95.21 94.34

Structural formulae based on 22 oxygens


Si 5.486 5.488 5.385 5.164 4.966 5.021 4.949 4.928 4.865 4.681
Ti 0.087 0.083 0.135 0.104 0.146 0.103 0.100 0.141 0.125 0.143
Al 2.591 2.533 2.598 2.856 3.006 3.105 3.215 3.215 3.295 3.349
Cr 0.011 0.006 0.008
Fe 0.446 0.468 0.462 0.465 0.477 0.465 0.465 0.411 0.417 0.559
Mn 0.005 0.004 0.008 0.008 0.004 0.003
Mg 5.507 5.507 5.449 5.452 5.485 5.417 5.349 5.348 5.325 5.375
Na 0.031 0.034 0.029 0.044 0.056 0.041 0.018 0.012 0.018
K 1.880 1.820 1.597 1.534 1.383 1.503 1.476 1.468 1.383 1.204
Ba 0.056 0.154 0.325 0.464 0.578 0.455 0.496 0.535 0.638 0.780
mg 0.925 0.922 0.922 0.921 0.920 0.921 0.920 0.929 0.927 0.906
aFeOr =total Fe expressed as FeO; n.d. =not detected. Cao not detectable by electron microprobe. Compositions 1-5, Xi-Yu;
6-10, Shengli. (All data this work.)

Figures 2.45-2.47. Tetraferriphlogopite mantles are poor in F (<0.3 wt %), Ah03 «1


wt %), and Ti02 «0.1 wt %) and rich in FeOr (15.0 wt %) relative to Al-phlogopites.

2.1.B.2.g. Benfontein. The highly evolved Benfontein calcite kimberlite (Dawson


and Hawthorne 1973) contains colorless poikilitic groundmass plates of al uminous micas
that are very rich in BaO (Mitchell 1994b; Ah03 = 16.8-21.0 wt %; BaO = 15.9-23.8
wt %). These micas contain limited amounts ofK20 (2.1-4.8 wt %) and are best described
as potassian kinoshitalite. Individual crystals are relatively uniform in composition,
although considerable intergrain compositional variation is present (Mitchell 1994b;
Figures 2.45-2.47). Macrocrystal and tetraferriphlogopite micas are not present.

2.1.B.2.h. Kirkland Lake Province. Micas in the kimberlites of the Kirkland Lake
area of Ontario (Brummer et al. 1992) have been briefly described by Smith et al. (1978)
and Arima et al. (1986). The kimberlite occurring as a dike in the Upper Canada (gold)
Mine contains dark brown xenocrystal aluminous biotites similar to those in the southern
African orangeites. These are partially resorbed and exhibit a wide range in FeOT (11-21
144 CHAPTER 2

- • CHINA
-++-
II + NAMIBIA
18 - • I)

•••
•,+

I
- +

17 -
"i
.+
• +
-
., •
-
~ ++ +
16 -
++ +
~ •
- +
•+ ++
It)
0
.,i\I
<{ + +
• •••
15 -

14 -
LEUCITE HILLS
PRIMITIVE MICAS
- ORANGEITE
PRIMITIVE
r---------.,I
I
Figure 2.43. AI203 versus Ti02
compositional variation in ground-
I

13 -
MICAS..,. l I mass micas from the Xi-Yu and
II I Shengli kimberlites (China) and Na-
mibian kimberlites (Spriggs 1988).
I I I I I
I 2 3 Compositional field of Leucite Hills

Ti0 2 wt. % • primitive iamproite micas from


Mitchell and Bergman (1991).

wt %) and Ti02 (3--6 wt %) content. These micas are mantled by phlogopites similar in
composition to macrocrysts and microphenocrysts in kimberlite and orangeite. Tetrafer-
riphlogopite and barian phlogopite groundmass micas are apparently absent.
The Nickila Lake kimberlite contains phlogopite macrocrysts exhibiting a limited
range in composition (<2 wt % BaO, <1.5 wt % Cr203. 1-4 wt % Ti02. 5-12 wt % FeOT).
Micas in volcaniclastic rocks are poorer in Ba. Fe. and TI and richer in Cr than micas in
hypabyssal facies rocks. Tetraferriphlogopites and aluminous biotites are absent.
Arirna et al. (1986) do not tabulate sufficient data to allow determination of Al-TI
and AI-Fe evolutionary trends. Although described as micaceous kimberlites by Arima
et al. (1986). the apparent absence of the tetraferriphlogopite trend indicates that the rocks
are unlikely to be derived from orangeite-type magmas. It is suggested here that the rocks
are archetypal kimberlites which have been modally enriched in rnacrocrystaVphenocrystal
MINERAWGY OF ORANGElTES 145

-
* •• •
• •••
,
-
18

- +
+
• ,• • •
•+
- •
+
+ • •+
-
+ t ••
~
16 - • +
-
• • +
+
++
• +
.."
o
t\I
- +
<{ 15
•• •
-

• CHINA

14 - + NAMIBIA

- LEUCITE HILLS
PRIMITIVE MICAS
r--------------,
- I I
13
~
I I I I
:
I I I
2 3 4 5
FeOT wt. % ..

Figure 2.44. AI203 versus FeOr compositional variation in groundmass micas from the Xi-Yu and Shengli
kimberlites (China) and Namibian kimberlites (Spriggs 1988). Compositional field of Leucite Hills primitive
lamproite micas from Mitchell and Bergman (1991).

mica. Aluminous biotites are considered by Arima et al. (1986) to be crustally-derived


xenocrysts (however, see 2.1.4).

2.1.8.2.i. Siberian Kimberlites. The Zagodochnaya kimberlite (Egorov et al. 1991)


contains large (5 mm) macrocrystal micas and smaller (2-3 mm) "ground mass micas."
Insufficient textural information is provided for the groundmass micas, although they
appear to be microphenocrysts. The macrocrysts are poor in Ti02 (<0.5 wt %) and Cr203
«0.2 wt %) relative to the "groundmass" micas (2-4 wt % Ti02, 0.8-1.2 wt % Cr203).
The Ah03 (13-14 wt %) and FeOr (3.8-4.7 wt %) contents of both types of mica are
similar (Table 2.15). The ground mass micas are weakly-zoned toward increasing Ti and
146 CHAPTER 2

21 -t + BENFONTEIN

~ M MAYENG
K KOIDU

l* • ORROROO
20
U UDACHNAYA
+ Y YUBILEINAYA
I ,., Z ZAGODOCHNAYA
Y Y
19 "1- Y

t \
-1I
18 -i+ Y

...J
+ Y
+ Y
~ I
-*
17 Y U
..: U
~ K
- K
J'
If)
0
~
16 -
<t
M~
...~'"
- U
• \K K U.}I
15 - •
- • • U K Y
M

LEUCITE HILLS K UU Z ~M
14 - PRIMITIVE· Y Y
MICAS· K U
- Z
Z
r\---U-- 1 JK J
J
• I 1 U
13 - ulUz
J
ORANGEITE U
- PRIMITIVE
K
U U
MICAS
I
1 I
2 3 4

Ti0 2 wt. %

Figure 2.45. Ah03 versus Ti02 compositional variation of groundmass micas from diverse southern African.
Australian and Russian kimberlites. Data sources Benfontein (Mitchell 1994b). Mayeng (Apter et al. 1984).
Orrorroo (Scott Smith etal. 1984). Udachnaya (Egorov etal.1991. this work). Yubileinaya (this work).
Zagodochnaya (Egorov etal. 1991). Compositional field of Leucite Hills primitive lamproite micas from
Mitchell and Bergman (1991).
MINERALOGY OF ORANGEITES 147

21 + + BENFONTEIN
M MAYENG
+ K KOIDU

20
~ • ORROROO
U UDACHNAYA
Y YUBILEINAYA
Z ZAGODOCHNAYA
+ Y
Y
19 + + + y
+ +
+ -;.

18 :tt' + y
+
+ y

-
~ ++ y

17 yU
-tI- +
~ + U
K
K
It) U U
0 16 K
C\I M KK
<{ K U
U K M
15 • U K
K

• •
M
uU
U


Z K
14 y y

r------ •
Z
y U
Y UK
ZI
ZY

13 I LEUCITE HILLS I .
U
I PRIMITIVE K
I MICAS
I

2 3 4 5 6 7

FeOT wt. % •
Figure 2.46. AI203 versus FeOT compositional variation of groundmass micas from diverse southern African,
Australian, and Russian kimberlites. Data sources as in Figure 2.45.
148 CHAPTER 2

Table 2.15. Representative Compositions of Mica in the Zagodochnaya and Yubileinaya


Kimberlites, Russiaa
Wt% 2 3 4 5 6 7 8 9 10
Si02 42.99 41.04 40.40 42.48 42.86 38.76 31.89 36.41 32.23 31.97
TI02 0.47 2.86 3.67 0.26 0.95 3.80 0.53 1.39 1.75 1.13
AI 20 3 13.57 12.94 14.18 12.85 11.97 13.95 17.68 17.07 19.09 19.25
Cr203 0.22 1.16 1.08 0.25 0.12 0.94 0.05 n.d. 0.07 0.10
FeOT 3.84 4.21 4.69 3.72 4.52 5.47 2.85 5.13 4.60 3.23
MnO n.d. 0.01 0.04 0.10 O.ll 0.08 0.06 0.03 0.04 n.d.
MgO 26.50 21.89 22.74 24.37 23.82 22.54 22.41 24.19 23.96 22.52
CaO n.d. n.d. n.d. n.d. 0.04 n.d. 0.38 0.07 n.d. 0.19
Na20 0.20 0.15 0.13 0.30 0.31 0.20 0.15 0.18 0.27 0.20
K20 8.56 10.80 10.55 10.88 10.35 10.04 3.41 9.16 7.85 6.75
BaO n.a. n.a. n.a. n.a. n.a. 0.37 16.01 2.62 4.93 9.25
NiO 0.07 0.12 0.17 n.d. n.d. 0.03 n.d. n.d. 0.05 n.d.
95.72 95.18 97.65 94.91 95.05 96.10 95.42 96.05 94.84 94.59

Structural formulae based on 22 oxygens


Si 5.935 5.866 5.639 6.025 6.092 5.365 4.985 5.250 4.799 4.871
TI 0.049 0.307 0.385 0.Q28 0.102 0.408 0.062 0.151 0.196 0.130
AI 2.208 2.179 2.333 2.148 2.005 2.349 3.258 2.901 3.350 3.457
Cr 0.024 0.131 0.119 0.Q28 0.014 0.106 0.006 0.008 0.012
Fe 0.443 0.503 0.548 0.441 0.537 0.653 0.373 0.619 0.573 0.412
Mn 0.001 0.005 0.012 0.013 0.009 0.008 0.004 0.005
Mg 5.453 4.663 4.731 5.152 5.047 4.799 5.222 5.199 5.317 5.155
Ca 0.006 0.064 0.011 0.031
Na 0.054 0.042 0.062 0.083 0.085 0.055 0.046 0.050 0.078 0.059
K 1.508 1.969 1.879 1.969 1.876 1.829 0.680 1.685 1.491 1.312
Ba 0.021 0.980 0.148 0.288 0.552
Ni 0.008 0.014 0.019 0.003 0.006
mg 0.925 0.903 0.896 0.921 0.904 0.880 0.933 0.893 0.903 0.925
aFeOr =total Fe expressed as FeO; n.d. =not detected; n.a. =not analyzed. Compositions 1-5, Zagodochnaya (Egorov el 01.
1991); 6-10. YubiJeinaya (this work); I, macrocryst; 2-3, microphenocrysts; 4-5, core and rim microphenocryst; 6-7, core
and rim groundmass micas; 8-10, groundmass micas.

Fe from core-to-margin. Figures 2.45 and 2.46 show that Zagodochnaya micas are
relatively unevolved, Micas belonging to the tetraferriphlogopite trend are absent. The
mica compositional data indicate that this kimberlite is an archetypal kimberlite which is
modally enriched in macrocrystallphenocrystal mica.
The Yubileinaya kimberlite contains small, anhedral, light brown microphenocrystal
micas that have been replaced at their margins by Ba-bearing mica (this work). BaO
contents are typically from 0.1 to 5 wt %, although higher levels may be rarely found
(Table 2.15, anals. 7 and 10). Figures 2.45 and 2.46 show that many of the micas are
relatively unevolved. Increases in AI are accompanied by a decrease in Ti and increase in
Fe. Groundmass poikilitic micas and tetraferriphlogopites are absent in the suite of
samples examined in this work.
MINERAWGY OFORANGEITES 149

~
-
1'0 + BENFONTEIN

• ORROROO

• YUBILEINAYA

~ Bo - MICAS IN
CARBONATITES,
0'6 LEUCITITES ,
NEPHELINITES
.......
CI)

--
E 0'4
o
c
0'2

0'2 0'4 0'6 0'8 1'0

Ba( atoms / II oxygens) ~

Figure 1.47. Ba versus K (atomic) for groundmass micas from the Benfontein (Mitchell 1994b), Orroroo (Scott
Smith et al. 1984), and Yubileinaya (this work) kimberlites. Compositional field of Ba-rich micas in carbona-
tites, leucitites, and nephelinites from Mitchell (1994b).

Micas in the groundmass of the East Udachnaya kimberlite (Egorov et al. 1991, this
work) are anhedral, light-brown microphenocrysts. Many of these micas are similar in
composition (Table 2.16, Figures 2.45 and 2.46) to unevol ved macrocrystal micas in other
kimberlites. They are considered to be cognate earlier-fonning phases which have not
crystallized in situ. The margins of the crystals are slightly enriched in BaO. Colorless
groundmass poikilitic plates ofbarian phlogopite are absent. Phlogopite in mica-pyroxenite
clasts from the same locality (Egorov etat. 1986) is of a similar composition to the
macrocrystal micas (Table 2.16, anal. 10).
The West Udachnaya kimberlites contain groundmass micas identical in their par-
agenesis, and of similar composition (Table 2.16, Figures 2.45 and 2.46) to those of East
Udachnaya. They differ in being relatively poor in BaO (0.3-0.9 wt %) and richer in FeOT
(5.1-1.0wt%).

2.1.8.2.j. Aries. The Aries kimberlite (Western Australia) contains abundant phe-
nocrystal and microphenocrystal mica (Edwards et al. 1992), although poikilitic ground-
mass plates appear to be absent. Phlogopite compositional zoning is weak, but complex,
and differs in each portion of the intrusion. The overall composition is similar to that of
unevolved macrocrystal and phenocrystal phlogopites from kimberlites, orangeites, and
lamproites (Figure 2.48). Complex normal and reverse zoning trends indicate that magma
mixing played a significant role in the development of the mica assemblage. Some of the
150 CHAPTER 2

Table 2.16. Representative Compositions of Micas from the Udachnaya KimberiiteO


Wt% 2 3 4 5 6 7 8 9 10
Si02 39.09 39.96 36.68 36.61 39.21 39.60 39.63 36.49 39.49 38.89
Ti0 2 3.68 2.69 3.23 3.19 2.93 3.04 2.99 1.84 2.69 2.92
AI 20 3 13.19 14.47 16.15 16.24 12.89 13.73 13.62 17.65 12.87 12.78
Cr20, 1.50 0.07 0.03 0.02 0.53 1.32 0.05 0.07 0.06 0.08
FeOT 5.08 6.26 4.87 5.17 6.35 5.12 7.25 11.00 7.74 8.20
MnO n.d. 0.08 0.08 0.05 0.11 0.05 0.08 n.d. 0.07 0.05
MgO 23.36 22.46 22.76 22.09 23.56 23.70 22.00 19.08 22.78 21.99
Na20 0.03 0.22 0.05 0.05 0.26 0.23 0.29 0.58 0.09 0.15
K20 10.18 10.07 10.19 9.81 9.93 10.10 9.97 9.04 8.03 8.88
BaO 0.17 0.38 1.14 1.59 0.32 0.25 0.61 0.50 n.a. n.a.
NiO 0.15 0.06 0.03 0.03 0.09 0.09 n.d. n.d. n.a. n.a.
96.43 96.12 95.18 94.85 95.86 97.23 96.49 96.25 93.83 93.88

Structural fonnulae based on 22 oxygens


Si 5.563 5.663 5.313 5.334 5.618 5.581 5.673 5.310 5.733 5.691
Ti 0.394 0.287 0.352 0.349 0.316 0.322 0.322 0.201 0.294 0.321
AI 2.212 2.417 2.757 2.788 2.177 2.281 2.298 3.027 2.202 2.204
Cr 0.169 0.008 0.003 0.002 0.060 0.147 0.006 0.008 0.007 0.009
Fe 0.605 0.742 0.589 0.630 0.761 0.604 0.868 1.339 0.940 1.004
Mn 0.009 0.009 0.006 0.013 0.006 0.009 0.008 0.006
Mg 4.955 4.744 4.914 4.797 5.031 4.979 4.694 4.138 4.929 4.796
Na 0.008 0.060 0.014 0.014 0.072 0.063 0.081 0.164 0.025 0.043
K 1.848 1.820 1.883 1.823 1.815 1.816 1.821 1.678 1.487 1.658
Ba 0.009 0.021 0.065 0.091 0.Gl8 0.014 0.034 0.029
Ni 0.017 0.007 0.004 0.004 0.010 0.010
mg 0.891 0.865 0.893 0.884 0.869 0.892 0.844 0.756 0.840 0.827
,'!FeOT = total Fe expressed as FeO;n.d. = not detected; n.a.= not analyzed. Compositions 1-4. East Udachnaya (this work);
5-8. West Udachnaya (this work); 9. average composition of three "microphenocrysts." East Udachnaya (Egorov et al. 1991);
10. phlogopitein mica pyroxenite clast. East Udachnaya (Egorov et al. 1986).

more evolved micas contain small quantities of tetrahedral Fe3+, indicating the presence
of minor amounts of the tetraferriphlogopite molecule. However, micas exhibiting the
extreme Al depletion, characteristic of orangeites, are not present as the majority of micas
are aluminous and plot in the Fe-Mg-AI ternary system (Figure 2.49) on the AI-rich side
of the phlogopite-biotite join. Insufficient compositional zoning is present to state
unequivocably (Edwards et al. 1992) that the Aries intrusion has mineralogical affinities
to orangeites. The micas contain only 0.6-1.5 wt % BaD, and Ba-rich micas are absent.

2.1.B.2.k. Koidu. The Koidu kimberlite, Sierre Leone (Grantham and Allen 1960),
contains reversely-pleochroic Fe-rich, Cr- and Ti-poor mica macrocrysts interpreted by
Tompkins and Haggerty (1984) to be xenocrysts. These are mantled by Ti- and Cr-bearing
phlogopites. Microphenocrystal or groundmass micas are similar in composition to the
rims of cores and are, in tum, rarely mantled by thin rims of reversely-pleochroic, low-AI
MINERALOGY OF ORANGEITES 151

NORTH EXTENSION CENTRAL LOBE


16 - "Primitive field"
K'-,
__ ~ .!2!. Mitchell (~8~ ___ ~

a: :: ~ ~2kJ~
<:(
+ ~~, rnettes
\ \', ... , ..........
10 - - - -------,........... ...
OPM \ ' .
\ Lamproltes
8-
I
KI
I I I I
~ I
" I
NORTH LOBE 0 2 4 6
16 TiOZ

+ core
c intermediate zone
• second intermediate zone
10 o rim
• groundmoss grain
8 / zoning trend

o 6

Figure 2.48. Compositional variation of micas in the Aries kimberlites, Australia (after Edwards et al. 1992).
ZI-Z4 Central Lobe; Sc, Sr, Mc, and Mr are cores and rims of grains from lithologies M and S of the North
Lobe. K = trends of "kimberlite" mica compositions from Mitchell (I 986).

tetraferriphlogopite-like micas. Unfortunately, mica compositions are not tabulated by


Tompkins and Haggerty (1984). Data obtained during the course of this work show that
microphenocrysts contain 14.0-16.6 wt % Ah03, 2.5-3.4 wt % Ti02, and 5.0-7.5 wt %
FeOT (Figures 2.45 and 2.46). The margins of the crystals rarely exhibit thin mantles of
Ah03-poor «1.0 wt %) tetraferriphlogopite.

2.1.8.2.1. Mayeng. The Mayeng sills, South Africa (Apter et al. 1984), in addition
to pale-colored phlogopite macrocrysts, contain two types of groundmass mica. All
textural varieties of the kimberlite contain an early-forming, dark-colored mica which
forms euhedral equant crystals. Following crystallization of this mica individual sills
evolved separately and produced different assemblages of late-stage micas as mantles
upon preexisting dark micas. Two of the sills thus crystallized pale-colored micas which
poikilitically enclose spinels, while a third crystallized dark-red, reversely-pleochroic
phlogopite.
The early dark micas contain 4-6 wt % Ti02, 0-1 wt % Cf203, 4.5-5.3 wt % FeOT,
and 13.3-15.6 wt % Ah03. The pale micas are also rich in Ah03 (13.8-15.8 wt %), but
are poorer in Ti02 (2.5-2.9 wt %) and Cr203 (0.04-D.12 wt %). Low analytical totals for
these micas may reflect the presence of Ba not determined by Apter et al. (1984). The
152 CHAPfER2

AI AI
A

5 15 25 15
AI
c D

,, ,, ++
,, , +

,, \
,\ +

\ , +
\ \+
\ + ANTOCHKA \
, • BOUNOUDOU \
'---r~r--r-r---7':":":''''' FeT ,"P '---r---.,,..--r--r-""';'!.!:.!:+ FeT '"P
Mg ~ 15 25 Mg 5 [5 25

AI AI
E 50 F

~o,
/ 70 .ttl % ., 70
li.OfJooo 0 8
\-~·t~~--- PHL ,----- - ---

\\ '·.'f' ,
\
\

• ARIES • UDACHNAYA WEST


• ORROROO o UDACHNAYA EAST
x BENFONTEIN YUBILEINAYA
o KOIDU \
, + ZAGODOCHNAYA
\
\ TFP \ TFP
'---r---:r---r---:r--+.:....:..... FeT
Mg 5 15 25 Mg 5 15 25

Figure 2.49. Compositions of micas from diverse kimberlites plotted in the ternary system Al-Mg-FeT
(atomsl22 oxygens). Data sources given in the text. FeT = total Fe expressed as Fe2+. EAST = "eastonite,"
PHL =phlogopite, TFP =tetraferriphlogopite.
MINERAWGY OF ORANGElTES 153

evolutionary trend appears, on the basis of the limited data tabulated, to be one of
increasing Al and Fe with decreasing Ti (Figures 2.45 and 2.46) and is thus similar to that
found in other archetypal kimberlites. Tetraferriphlogopites have low Ah03 «2 wt %),
Ti02 (0.3-0.7 wt %), and Cr203 «0.1 wt %), and moderate FeOT (10-13.3 wt %)
contents. They are relatively poor in FeOT compared to similar micas in orangeites.
Apter et al. (1984) consider the pale micas to be representative of the final stages of
crystallization of this kimberlite but provide no explanation as to why tetraferriphlogopite
forms in one textural variety. However, it is suggested here that the differences are due
to different late-stage liquid compositions and/or conditions of crystallization subsequent
to emplacement.

2.1.B.2.m. Fayette County. The Fayette County (Pennsylvania) kimberlite dike


contains microphencrystal and groundmass mica (Hunter et al. 1984). The microphe-
nocrysts are zoned toward margins whose compositions overlap those of the groundmass
micas. All micas are relatively Ah03 rich (approx. 14.0 wt %). No systematic variations
in Al content are evident, although there is a change in Cr, Ti, and Fe with paragenesis.
Hence, phenocryst cores contain 1.3-1.8 wt % Cr203, 4.5-5.0 wt % Ti02, and 4.6-5.5
wt % FeOT. The rims contain 0.1-1.0 wt % Cf203, 3.6-4.5 wt % Ti02, and 4.9-6.0 wt %
FeOT. The groundmass micas contain <0.2 wt % Cr203, 1.7-4.5 wt % Ti02, and 4.9-6.0
wt % FeOT. The overall trend of compositional evolution is one of decreasing Cr, Ti, and
increasing Fe from core, via rim, to groundmass. Tetraferriphlogopite is absent.

2.1.B.2.0. Blue Ball. The Blue Ball (Arkansas) "kimberlite" (Sal pas et al. 1986, Neal
and Taylor 1989) contains phenocrystal and groundmass phlogopites, in addition to micas
which occur as coronas about olivine phenocrysts. The micas range in composition from
phlogopite to tetraferriphlogopite. Two compositionally distinct varieties of phenocrysts
are recognized on the basis of their mg numbers (Figure 2.50) or Ti02 «1.0 wt % or 1-1.5
wt %) and BaO «0.5 or >0.5 wt %) contents. Phenocrysts are mantled by micas
exhibiting a very wide range in composition with respect to their Al and Fe contents.
Some rim compositions are identical to those of micas defined as groundmass mica.
Evolutionary trends from core to rim in the phenocrysts are of decreasing AI, Ba, and Ti
with increasing Fe and octahedral site deficiency. The composition of the groundmass
micas encompasses the entire range of phenocryst core-to-rim compositional variation
with respect to Fe but not Al (Figure 2.50). Some of the groundmass micas may be
enriched in Ti02 (up to 3 wt %) relative to the phenocrysts, yet others have compositions
identical to those of the phenocrysts. The data suggest that all of the micas are of one
generation/paragenesis, thus, many of the small micas interpreted by Neal and Taylor
(1989) as ground mass mica may be in reality microphenocrysts. The mica assemblage
has similarities to that of orangeites in being a complex transported assemblage.
Figure 2.51 shows the evolutionary trend of the Blue Ball micas is identical to that
of orangeite micas. This trend, the absence of micas belonging to the phlogopite-kinoshi-
talite series and other mineralogical features, e.g., spinels ranging in composition from
chromite through Ti-magnetite to magnetite, coupled with the absence of magnesian
ulvospinel, suggests that the Blue Ball intrusion is not an archetypal kimberlite. Samples
examined during the course of this work showed the rocks to consist of serpentinized
154 CHAPTER 2

2'0

I- 1'0
<J

0'5

40 50 60 70 80 90
Mg ~
Figure 2.50. Tetrahedral site deficiency (~T) versus mg number of micas from the Blue Ball "kimberlite,"
Arkansas (after Neal and Taylor 1989). Stars represent the earliest phenocrysts. Solid circles represent the cores
of the second group of phenocrysts. Open circles represent phenocryst rims. Tie lines join core and rims of the
same phenocryst. The inset demonstrates the evolutionary path of compositions.

t

12

10

..: ORANGEITE
~ • PRIMITIVE MICAS
It) 6
oC\I
Ci 4

2 BLUE BALL •
ARKANSAS

3 6 8 10 12 14 16 18 20 22 24
Ti02
-wt. %--.
FeOT Wt. % •
Figure 2.51. Compositional variation of micas from the Blue Ball "kimberlite," Arkansas (after Neal and Taylor
1989).
MINERALOGY OF ORANGElTES 155

olivine set in a matrix of mica, apatite, and carbonate. Also present as accessory phases
are perovskite (poor in Sr, Nb, REE), rutile, Ca-REE phosphate, Ni-pyrite, djerfisherite,
barite, and Nb-Mn ilmenite. The rocks have mineralogical affinities with both lamproites
and orangeites, but without further mineralogical study defy exact classification. Their
proximity to the Arkansas lamproite province (Mitchell and Bergman 1991) and possible
similarity in age (? mid-Cretaceous) suggests that they may be a part of this province.

2.1.8.2.n. West Greenland. The compositional variation of micas in the West Green-
land kimberlites (Scott 1981, Larsen and Rex 1992) has been insufficiently studied to
allow determination of evolutionary trends. Data given by Scott (1981) show the
macrocrystal and groundmass micas to be Ti-bearing phlogopite (Ah03 = 11.5-15.5
wt %, Ti02 =0.4-4.72 wt %, FeOr = 6.1-10.21 wt %). Many groundmass crystals
exhibit rims of tetraferriphlogopite. Barian micas appear to be absent.

2.1.8.3. Summary of Kimberlite Mica Compositional Variation


Macrocryst compositions are not significantly different from macrocrysts/microphe-
nocrysts in orangeites. Macrocrysts and the cores of microphenocrystal (and/or ground-
mass) micas are relatively low in AI, compared with crystal margins and poikilitic
groundmass micas. Two evolutionary trends may be discerned on the basis of the data
given above. The dominant and characteristic one is Al enrichment. Figure 2.49A-C
shows that mica compositions belonging to this trend plot away from the phlogopite-
biotite join toward aluminous phlogopites and the hypothetical "eastonite" composition.
Micas following this trend are commonly enriched in BaO, and late-stage poikilitic micas
are colorless members of the phlogopite-kinoshitalite series. Late-stage Ba-rich micas
are commonly depleted in FeOr relative to unevolved micas. Ti02 contents are low «4
wt %) and do not show any systematic trends. The high Ti02 (>5 wt %), Ba-rich, pink
micas characteristic of the groundmass of leucitites and melilitoids are absent.
The second, less common, evolutionary trend is toward tetraferriphlogopite. This
trend is illustrated in Figure 2.490 by mica mantles from the Antochka kimberlite.
Importantly, these AI-poor micas are typically developed as discrete thin mantles upon
cores of AI-rich groundmass micas. Tetraferriphlogopite formation does not occur in all
facies of a given kimberlite and has not occurred in most archetypal kimberlites. The
abrupt change from aluminous mica crystallization to tetraferriphlogopite indicates
sudden, drastic changes in the redox conditions prevailing during the final stages of
crystallization. Its formation may be associated with the addition of ground waters to the
magma and/or rapid carbon dioxide loss.
In conclusion, archetypal kimberlites may easily be distinguished from orangeites
on the basis of the different evolutionary trends of groundmass micas (Figures 2.49,
2.52-2.53). Although some kimberlites do contain groundmass tetraferriphlogopite, its
presence cannot be regarded as characteristic. Tetraferriphlogopite development appears
to be random and is principally confined to thin discrete mantles upon preexisting AI-rich
phlogopites. This paragenesis is in marked contrast to the ubiquitous presence oftetrafer-
riphlogopite, as either the outer margins of continuously-zoned groundmass micas or
late-stage poikilitic plates in orangeites.
156 CHAPfER2

20

18

1
16

14

12
~
.;
~ 10
If)
0
N 8
<t
6

2
Figure 2.52. Compositional (A1203 versus Ti02)
evolutionary trends of micas from orangeites (this
work).lamproites (Mitchell and Bergman 1991).
2 4 6 8 10 12
minettes (Mitchell and Bergman 1991). and kim-
Ti0 2 Wt. % • berli tes (this work).

FIELDS OF MICROPHENOCRYST
20
COMPOSITIONS
ffllI ORANGEITE [ ] LAMPROITE

t
18

16

14
ae
..: 12
~
10
II)
0
~ 8
«
6

4
Figure 2.53. Compositional (A1203
2
versus FeOr) evolutionary trends of
micas from orangeites (this work).
lamproites (Mitchell and Bergman
2 4 6 8 10 12 14 16 18
1991). minettes (Mitchell and
Bergman 1991). and kimberlites (this
work).
MINERALOGY OF ORANGElTES 157

2.1.9. Mica in Lamproites


The compositional variation shown by micas in lamproites is well characterized and
has been summarized by Mitchell and Bergman (1991).
Individuallamproite provinces contain phenocrystal micas of distinct composition
with respect to their Ti, AI, and Fe contents. Within a lamproite province phenocrysts may
exhibit relatively little compositional variation or vary considerably from vent-to-vent.
Groundmass micas and mantles upon phenocrysts are typically depleted in Cr and AI,
and enriched in Fe, Ti, Ba, and Na relative to phenocrysts, and thus may be considered
more evolved than the latter.
Mitchell and Bergman (1991) have shown that compositional and zonation trends
with respect to the AI, Ti, and Fe contents of mica may be used to assess the degree of
evolution of the magma from which they crystallized. Thus, micas with high Ah03
(10.5-\3.5 wt %), 4-10 wt % Ti02, and 2-4 wt % FeOT typically occur as phenocrysts.
Such micas are considered to be relatively primitive un evolved micas which may have
formed in the magma prior to eruption. Micas with low Ah03 (0.5-10 wt %) and high
FeOT (>4 wt %) form mantles upon phenocrysts and occur as groundmass poikilitic
plates, and are interpreted to have formed at relatively low temperatures and pressures
after eruption or emplacement of the magma.
Compositional evolutionary trends fall between two extremes (Figures 2.52-2.56).
One is of slight-to-moderate Al depletion, coupled with increasing Ti and Fe and
decreasing Mg, i.e., a trend reflecting Fe2+ increase, and representing evolution from

AI

MO TFA
5 10 15 20 30 35 40 45 50 55 60 65

F 2 + __......
eT •

Figure 2.54. ComP9sitional evolutionary trends (A-F) of micas from di verse lamproi tes plotted in the ternary
system Al-Mg-FeT2+ (atomic). A = Hatcher Mesa, Leucite Hills; B = Middle Table Mountain, Leucite Hills;
C =Pilot Butte and Badger's Teeth, Leucite Hills; D =Mount North and Rice Hill madupitic lamproites, West
Kimberley; E =Mount North phlogopite lamproites, West Kimberley; F = Mount Gytha, West Kimberley.
PHL =phlogopite, TFP =tetraferriphlogopite, TFA =tetraferriannite.
158 CHAPTER 2

FIELDS OF
20
m
MICRO PHENOCRYST COMPOSITIONS
.-.,
ORANGEITE L_J LAMPROITE

18 • MINETTE

16

14

f 12
~
..:
~
10
II)
0
01
<i 8

2 4 6 8 10 12

Figure 2.SS. A1203 versus 1102 compositional variation of micas from minettes. Data sources given in
Mitchell and Bergman (1991). Compositional trends for micas in orangeites, lamproites, and kimberlites from
Figure 2.52.

titanian phlogopite toward titanian biotite. The other is a trend of strong Al depletion
associated with increasing Ti and Fe at essentially constant Mg content, i.e., a trend
reflecting increasing Fe3+, and representing evolution from titanian phlogopite toward
titanian tetraferriphlogopite. The compositional trend exhibited by mica in a given
lamproite may lie anywhere between these two extremes and reflects the local post-
emplacement crystallization environment with respect to redox conditions, water content,
and cooling conditions.
Lamproite phlogopites as a group have insufficient Si and Al to fill the tetrahedral
sites in the mica structure, consequently Al vi is absent from phenocrystal and groundmass
MINERALOGY OF ORANGElTES 159

ae 14


..:

.., 12
oN
«

2 4 6 8 10 12 14 16 18 20

FeOT wt. %
Figure 2.56. Al203 versus FeOr compositional variation of micas from minetles. WG = Wattle Gill. England;
CEL = Celebes. Indonesia; DH = Dale Head. England; SC = Shaws Cove. Canada; NAV = Navajo. U.S.A.;
COL = Colima, Mexico; LIN =Linhaisai. Indonesia; DEV =Devonshire. England; H =Holmead Farm.
England; BOH-Aalkaline minettes. Bohemia; BOH-B otherminettes. Bohemia. Data sources given in Mitchell
and Bergman (1991). LH-P = Compositional fields of Leucite Hills lamproite phenocrystal micas. Composi-
tional trends for micas from kimberlites. orangeites. and lamproites from Figure 2.53.

micas. The tetrahedral site deficiency may be accommodated by the entry of Ti4+, Fe3+,
or Mg2+ to this site. Solid solutions present are primarily between phlogopite-bi-
otite, octahedral site-deficient titanian phlogopite [K2(Mg,Fe~TiDSi6Ah02o(OH~], and
tetraferriphogopite-tetraferriannite.
Other characteristic features of lamproite mica compositions are the high F (1-7
wt %) contents and enrichment of Na in groundmass micas. The latter contain 0.5-1.8
wt % Na20 and may be regarded as sodian titanian tetraferriphlogopites.
Orangeite micas have some compositional similarities to lamproite micas in that
similar evolutionary trends with respect to AI-Fe are present. Figures 2.53 and 2.54 show
that orangeite micas define compositional trends which overlap with those oflamproites.
The trends originate on AI-Fe diagrams at similar primitive unevolved phenocryst
compositions. By analogy with lamproites it is suggested that which trend is followed
must depend upon local post-emplacement crystallization conditions.
Significant differences exist with respect to AI-Ti compositional trends and the Ti
contents of micas. Lamproite micas are typically enriched in Ti02 (1-12 wt %) relative
to orangeite micas, and compositions evolve toward increasing Ti contents (Figure 2.52).
In contrast, the majority of orangeite micas are relatively poor in Ti02 (1-3 wt %) and
evolution toward tetraferriphlogpite is accompanied by decreasing or constant Ti contents
(Figure 2.52) with decreasing AI.
160 CHAPTER 2

However, other orangeites, which have evolved toward sanidine and richterite-bearing
residua, contain micas with Al-Ti evolutionary trends similar to those oflamproites. Thus,
micas from Postmasburg and Voorspoed follow typicallamproite mica Al-Ti evolution-
ary trends (Figures 2.6, 2.7,2.52,2.53), and AI-poor groundmass micas in the Besterskraal
and Sover North occurrences are identical in composition to groundmass micas in some
lamproites (Figures 2.5, 2.17, 2.52, 2.53).
Orangeite micas may be distinguished from lamproite micas on the basis of their low
Na (typically 0.5 wt % Na20; Table 2.5) and F «1.0 wt %) contents relative to lamproite
micas. Further, the complex mantling and reverse zoning patterns found in the mica
microphenocryst populations of orangeites are not present in lamproite micas.

2.1.10. Mica in Minettes


Micas in minettes form a complex-mantled, continuously- or reversely-zoned assem-
blage of phenocrysts (Bachinski and Simpson 1984, Jones and Smith 1983, 1985, Schulze
etal. 1985, Bergman etal. 1988, Mitchell and Bergman 1991, Meyer etal. 1994).
Phenocrysts show wide variations in composition within and between minette occur-
rences. The cores of phenocrysts are, in some instances, compositionally similar to
groundmass micas. Many crystals commonly have oxidized and/or resorbed margins.
Mitchell and Bergman (1991) have summarized the compositional variation of
minette micas and shown that the characteristic evolutionary trend is one of increasing
Fe with slightly increasing or constant Al (Figure 2.55). The Ti content may increase or
decrease slightly with respect to Al (Figure 2.56). Many minette micas are aluminous
and do not exhibit any tetrahedral site deficiency. The presence of Al vi indicates solid

AI
,
+-----.!..,-----SID

~
/ 70

leo
~ PHL

10 50 60

Figure 2.57. Compositional field of micas from minettes plotted in the ternary system AI-Mg-FIlT2+ (atomic).
Data sources as in Figure 2.56 plus Meyer et al. (1994). EAST ="eastonite," SID =siderophyllite. PHL =
phlogopite. ANN =annite. TFP =tetraferriphlogopite. TFA =tetraferriannite.
MINERALOGY OF ORANGEITES 161

solution toward "eastonite"-siderophyllite end-member molecules (Figure 2.57).


Tetraferriphlogopites have not been reported from minettes.
Figures 2.52, 2.53, and 2.57 show that unevolved minette phenocrysts overlap the
compositions of unevolved microphenocrysts from orangeites and lamproites. However,
the compositional trends diverge from these common unevolved micas, such that the
minette micas are always Al rich relative to orangeite and lamproite micas of equivalent
Fe content. The evolutionary trend is from phlogopite toward titanian aluminous biotite
and represents solid solution between phlogopite-biotite and the "eastonite"-siderophyllite
molecules.
Minor elements in minette micas exhibit very wide ranges in concentration (F =
0.2-5.0 wt %, BaO =0-2 wt %, Cr203 =0-1 wt %) and cannot be used as discriminators
between magma types.
In summary, orangeite micas may be easily distinguished from minette micas on the
basis of their Al-Ti and AI-Fe evolutionary trends.

2.1.11. Mica in Ultramafic Lamprophyres


Discussion of the composition of micas in ultramafic lamprophyres is required
because of the petrographic similarity of some olivine phlogopite lamprophyres to
orangeites. Unfortunately, there have been few detailed studies of the compositional
variation of micas in ultramafic lamprophyres, and most works tabulate only a few
random compositions. Rock (1986, 1990) reviewed these rocks and noted that micas in
them exhibit an extremely wide range in composition with respect to their AI, Fe Ti, and
Ba contents. However, no characteristic evolutionary trends were identified. Tetrafer-
riphlogopites occur in the groundmass of some examples (Rock 1990).
Table 2.17 and Figures 2.58-2.60 illustrate the compositional variation of mica in
some ultramafic lamprophyres belonging to the alnoite-polzenite (or melnoite; Mitchell
1994c) suite. These rocks are characterized by widely varying modal amounts of olivine,

Table 2.17. Representative Compositions of Mica from Ultramafic Lamprophyresa


Wt% 2 3 4 5 6 7 8 9 10
Si02 38.71 38.04 36.93 35.78 34.95 30.73 40.99 37.01 39.64 37.74
Ti02 1.57 1.62 5.37 4.04 5.11 3.09 1.55 5.55 2.02 2.24
Ah03 17.18 17.73 15.59 16.96 16.73 19.25 13.24 15.22 12.80 13.77
Cr203 0.04 0.02 0.57 n.d. 0.02 n.d. 0.15 0.16 0.09 0.17
FeOT 4.91 5.26 8.18 7.43 9.00 5.50 3.72 6.94 6.13 5.64
MnO 0.08 0.03 0.05 0.04 0.14 0.12 n.d. n.d. n.d. 0.04
MgO 22.53 22.59 18.66 20.44 18.85 20.61 25.96 20.54 23.99 23.36
Na20 0.83 1.70 n.d. n.d. 0.39 0.16 n.d. n.d. n.d. n.d.
K20 9.20 8.16 9.79 9.57 8.56 6.19 10.26 9.90 10.35 9.57
BaO n.d. n.d 0.77 2.11 2.68 11.09 n.d. n.d. 1.40 3.45
NiO 0.09 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.
95.14 95.15 95.91 96.37 96.43 96.84 95.87 95.32 96.42 95.98
"FeOT = total Fe expressed as FeO; n.d. = not detected. CompoSitions 1-2, alntiite, Como, Quebec; 3-4, alntiite, type locality
Alnti, Sweden; 5-6, polzenite, Polzen, Czech Republic; 7-8, alntiite, N. Hutson, Kentucky; 9-10, alntiite, Haystack Butte,
Montana. (All data this work.)
162 CHAPTER 2

t
20 @PI B
• ~!'

18

~ 16
.,;
~ 14
If)
0
..!I 12
«
10

2 3 4 5 2 3 4 5 6

Ti02 wt. % •
Figure 2.58. Compositional variation (AI203-Ti02) of micas in ultramafic lamprophyres. PI, P2, and P3 =
diverse polzenites, Polzen, Czech Republic; HB = Haystack Butte, Montana; A-OPM = diverse olivine
phlogopite alnoites, Alno complex, Sweden; A-TL = alnOite type locality, Alno, Sweden. All data this work,
except Vestfold (Delor and Rock (1991) and Karinya (Muller et al. 1993). LHPM and OPM =compositional
fields of microphenocrysts from Leucite Hills 1amproites (Mitchell and Bergman 1991) and orangeites (this
work), respectively. K, 0, and L are compositional trends of micas from kimberlites, orangeites, and lamproites,
respectively.

t Ki:~A~ ~I
20
"""''''''''
18 FEN
CIt! 16
..:
~

0
It) 14
------,
\ . HB

~12
<I LHPM
10

8
3 4 5 6 789 3 4 5 6 7 8 9 10 II

FeOr wt. %
Figure 2.59. Compositional variation (A1203-FeOT) of micas in ultramafic lamprophyres. Abbreviations and
data sources as in Figure 2.58.
MINERALOGY OF ORANGEITES 163

AI AI

55 \
/

Mg
Figure 2.60. Compositional fields of micas from ultramafic lamprophyres plotted in the ternary system
AI-Mg-Fer (atomic). MO = Monticellitovaya; other abbreviations as in Figure 2.58. Also shown are compo-
sitional evolutionary trends for groundmass micas in Somerset Island (SIK), Chinese (CK), Iron Mountain (1M),
and Benfontein (B) kimberlites. EAST ="eastonite," PHL =phlogopite, TFP =tetraferriphlogopite. Fer =total
Fe expressed as Fe2+.

phlogopite, spinel, perovskite, apatite, monticellite, and calcite. Melilite and clinopy-
roxene mayor may not be present. Although each occurrence is characterized by mica of
a particular composition, their overall compositional evolution is typically toward in-
creasing Ti and Fe, at constant or slightly increasing Al contents. All of the micas are rich
in AI, and solid solutions are primarily between phlogopite-annite and "eastonite"-
siderophyllite. These aluminous micas commonly coexist with aluminous pyroxenes
(2.2.7).
Compositional trends are unlike those observed in either orangeites or lamproites
(Figures 2.58-2.60) but similar to those found in micas from minettes (Figures 2.55-
2.57). The compositions of the relatively unevolved micas overlap with those of
unevolved groundmass micas in archetypal kimberlites. In common with the latter they
may in some occurrences, e.g., Haystack Butte (Wendlandt 1977), Polzen (this work),
have significant Ba contents. These micas may be distinguished from Ba-rich micas in
kimberlites on the basis of their evolutionary trends toward increasing Ti and Fe (Figures
2.58-2.60).
The diamond-bearing ultramafic lamprophyres from Bulljah Pool, Western Australia
(Hamilton and Rock 1990) contain micas which are members of the titanian phlogopite-
164 CHAPTER 2

AI AI
I
I
/ PHLOGO- I BIOTITES

WO
PITES I
t::.
MT. BUNDEY
IFI
o PHENOCRYSTS
I•• t::. GROUNDMASS
• BULLJAH
10

TFA
Mg Fe 2+
30 20 10 T
90 80 60 50 40
- - Fe2+~
T
Figure 2.61. Compositions of micas from the Mount Bundy (Sheppard and Taylor 1992) and Bulljah Pool
(Hamilton and Rock 1990) lamprophyres plotted in the ternary system AI-Mg-Fer2+ (atomic). Abbreviations
as in Figure 2.60.

titanian tetraferriphlopite series with minor solid solution toward tetraferriannite and/or
annite (Figure 2.61). They contain 6.4-11.7 wt % Ah03, 2-3 wt % Ti02 (rarely reaching
7 wt %), and <0.1 wt % Cr203. BaO contents are mostly <1 wt % but may reach 2.3 wt %.
Figure 2.62 shows that the micas are slightly evolved relative to primitive micas in
orangeites and lamproites. Unfortunately, Hamilton and Rock (1990) do not provide
detailed phenocryst-groundmass compositional trends. Figure 2.62 suggests that the
compositional trend is toward increasing Ti with decreasing AI, i.e., a lamproitic trend
rather than an orangeite trend.
Olivine-mica-sanidine lamprophyre dikes from Mt. Bundey, Australia (Sheppard
and Taylor 1992), have mineralogical affinities to evolved orangeites and lamproites. The
micas in these rocks are AI-rich members of the phlogopite--biotite-siderophyllite series
(Figure 2.61) which exhibit a wide range in Ti02 (1-8 wt %) and BaO (0.2-3.5 wt %).
Figure 2.62 shows that they are richer in Al than primitive micas from orangeites and
lamproites. Their compositional evolution toward increasing Ti (and Fe), at constant Al
contents, serves to distinguish them from the latter rocks.
Micas in the Bow Hilllamprophyre dikes (Fielding and Jaques 1989) show a wide
range in composition (Figure 2.62). Rare Ti02-rich (6 wt %) biotite cores of phenocrysts
are mantled by AI-rich phlogopites (11.5-16 wt % Ah03) of variable Ti02 (1-2 wt %)
contents. These micas are richer in F (>1 wt %) and poorer in Crz03 «0.05 wt %) than
the biotite (0.5 wt % F, >0.1 Cr203), and many are in turn mantled by narrow rims of
Ti02-poor «0.5 wt %) tetraferriphlogopite. The overall zonation trend has similarities to
that of orangeites rather than lamproites (Figure 2.62). Micas differ from those of
orangeites in being AI-rich and coexisting with andradite--melanite garnet.
~

~
~
o

MT. BUNDEY BOW HILL BULLJAH POOL


~
;d
v.>

t 16
14 + +:t • •
... tit-.
• + f .,~+
r- -IJ ·
~ 12 + +\"9-
+

~
-= 10

It) 8
0
~ 6 00
<t 0
4 -l I
+ PHENOCRYSTS ll.OLIVINE - PHLOGOPITE
• GROUNDMASS LAMPROPHYRE
2 .. ANDRADITE - PHLOGOPITE
LAMPROPHYRE
0
0 2 4 6 8 0 2 4 6 8 10 0 2 4 6 8

Ti0 2 (wt. %)
Figure 2.62. AI203 versus 1102 compositional variation of micas from the Mount Bundey (Sheppard and Taylor 1992), Bow Hill (Fielding and Jaques
1989), and Bulljah Pool lamprophyres (Hamilton and Rock 1990). 8:
-
166 CllAPTE1l2

2.2. CLINOPYROXENE

2.2.1. Paragenesis
Clinopyroxene is the only primary pyroxene found in orangeites. Modes vary widely
and may range from trace quantities to as much as 26 vol % (Wagner 1914, Skinner and
Clement 1979, Fraser 1987).
In unevolved orangeites, diopside typically forms small (mainly 0.1 to 0.5 mm and
rarely up to 1.5 mm in length) microphenocrysts of colorless prismatic crystals, some of
which are simply twinned. The diopsides in many instances are corroded and embayed
and have, evidently, undergone resorption during transportation or crystallization of the
mesostasis. The microphenocrysts are, rarely, continuously weakly-zoned to pale-yellow
rims and/or mantled by pale-green thin rims of titanian aegirine. Strong continuous
zoning, complex oscillatory zoning, sector zoning, and multiple epitaxial mantling are
not characteristic of orangeite diopsides (this work, Mitchell and Meyer 1989a, Dawson
eta/. 1977, Wagner 1914).
Clinopyroxene is most abundant in evolved orangeites. In these rocks it forms
interlocking aggregates of relatively large resorption-free blocky crystals. Unlike the
resorbed diopsides of unevolved orangeites these pyroxenes appear to have crystallized
in situ. Thin green mantles of titanian aegirine are characteristic (Figure 1.48). Similar
pyroxenes are found along fractures and cleavages. Commonly, pyroxenes are poikiliti-
cally included in groundmass mica and amphibole (this work, Tainton 1992).
Unevolved orangeite fields differ with respect to their pyroxene contents. In the
Barkly West field, diopside is not a characteristic primary groundmass mineral (Tainton
1992, Bosch 1971). Here, diopside occurs as xenocrysts derived from country rock
Ventersdorp basalts and in reaction assemblages adjacent to intrusion margins and
xenoliths (Tainton 1992). In contrast diopsides are common in orangeites from the
Boshof, Winberg. and Kroonstad fields, Swartruggens and Finsch (this work, Fraser 1987,
Dawson et al. 1977, Wagner 1914).

2.2.2. Composition
Reconnaissance studies of the composition of groundmass c1inopyroxenes in
orangeites have been undertaken by Dawson et al. (1977), Skinner and Scott (1979),
Robey (1981), and Mitchell and Meyer (1989a). These studies demonstrated that the
pyroxene is typically diopside of restricted compositional range and that significant
inter-intrusion differences are not apparent.
The summary of the compositional character of diopside in orangeites presented
below is based on over 250 new analyses of pyroxene obtained during the preparation of
this monograph. Representative compositions of microphenocrystal pyroxene are given
in Tables 2.18 and 2.19. These tables and Figures 2.63-2.65 confirm that clinopyroxenes
in orangeites typically exhibit only limited compositional variation.

2.2.2.1. Diopside
The majority of the phenocrystal pyroxenes are Fe-poor diopsides exhibiting very
little solid solution toward hedenbergite or aegirine (Table 2.18). The pyroxenes are poor
MINERAWGY OF ORANGEITES 167

Table 2.1S. Representative Compositions of Iron-Poor Clinopyroxenesa


Wt% 2 3 4 5 6 7 8 9 10
Si0 2 52.92 53.36 54.04 53.50 53.14 53.54 53.81 53.46 53.91 53.74
Ti02 0.84 0.78 0.42 0.82 0.68 0.87 0.78 0.25 1.22 0.73
AI 20 3 0.39 0.88 0.48 0.27 0.30 0.19 0.37 0.19 0.21 0.37
Cr203 0.75 0.89 0.23 0.16 0.18 0.26 n.d. 0.56 n.d. 0.06
FeOr 2.23 2.54 2.62 2.96 3.01 3.07 3.56 3.60 3.88 4.04
MnO 0.04 0.06 0.06 0.11 0.08 0.05 0.11 0.15 0.11 0.14
MgO 17.22 16.72 17.31 16.86 16.73 16.80 17.23 16.73 15.81 16.76
CaO 24.07 24.20 24.64 24.89 24.75 24.68 23.59 23.46 24.09 24.05
Na20 0.29 0.34 0.25 0.43 0.25 0.32 0.48 0.67 1.05 0.30
98.75 99.77 100.05 100.00 99.12 99.78 99.93 99.07 100.28 100.19
1.23 0.53 1.40 2.55 1.80 1.51 1.98 2.87 2.60 1.49
1.12 2.07 1.36 0.67 1.39 1.71 1.78 1.02 1.54 2.70
98.87 99.82 100.19 100.26 99.30 99.93 100.13 99.36 100.54 100.34
Structural fonnula on the basis of four cations and six oxygens
Si 1.951 1.952 1.965 1.950 1.956 1.959 1.960 1.965 1.963 1.961
AI 0.017 0.038 0.021 0.012 0.013 0.008 0.016 0.008 0.009 0.016
Ti 0.023 0.021 0.011 0.022 0.019 0.024 0.021 0.007 0.033 0.020
Cr 0.022 0.026 0.007 0.005 0.005 0.008 0.016 0.002
Fe 3+ 0.034 0.015 0.038 0.070 0.050 0.042 0.054 0.079 0.071 0.041
Fe 2+ 0.035 0.063 0.041 0.020 0.043 0.052 0.054 0.031 0.047 0.082
Mn 0.001 0.002 0.002 0.003 0.002 0.002 0.003 0.005 0.003 0.004
Mg 0.946 0.911 0.938 0.916 0.918 0.916 0.936 0.917 0.858 0.912
Ca 0.951 0.948 0.960 0.972 0.976 0.%7 0.921 0.924 0.940 0.940
Na 0.021 0.024 0.018 0.030 0.018 0.023 0.034 0.048 0.074 0.021
End member molecules (mol %)
CaTiAI 20 6 0.9 1.9 0.9 0.6 0.7 0.4 0.8 0.4 0.5 0.8
CaAISiAI06 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Ae 2.4 2.1 2.0 3.0 1.8 2.3 3.4 4.8 7.4 2.1
Wo 47.2 47.0 46.8 48.0 48.2 48.1 45.5 45.9 46.9 46.5
Fs 2.2 2.9 2.8 3.0 3.7 3.6 3.7 3.1 2.2 5.1
En 47.3 46.1 47.4 45.5 45.6 45.7 46.6 45.8 43.0 45.5
</1. Sover North; 2, Besterskraal; 3, Sover Mine; 4, New Elands; 5, Swartruggens; 6, Silvery Home; 7, Blaauwbosch; 8,
Makganyene; 9, Lace; 10, Voorspoed. FeOT = total iron calculated as FeO. Fe20J and FeO calculated by the method of Droop
(1987). End member molecules calculated assuming all Na20 is present as aegirine.

in Ti02 (typically 0.1-2.0 wt%), Ah03 (0.02-1.3 wt %), and Cr203 (typically 0-1.0
wt %). Intra- and inter-intrusion compositional variation is very limited and most py-
roxenes are not zoned.
Figures 2.63 and 2.64 illustrate the compositional range of pyroxenes in repre-
sentative intrusions containing zonation-free pyroxenes. The figure demonstrates that
inter-intrusion compositional variation is not significant. The majority of pyroxenes are
diopsides containing between 2 and 6 atomic % Fe and more than 47 atomic % Ca.
Mg-rich augite is found only in pyroxenes from Blaauwbosch. Note that the older
168 CHAPTER 2

Table 2.19. Representative Compositions of Relatively Iron Rich Pyroxenesa


Wt% 2 3 4 5 6 7 8 9 10
Si02 53.38 53.18 51.57 50.35 51.09 51.26 51.45 51.54 50.97 51.38
Ti0 2 0.20 0.37 0.76 1.31 1.07 1.22 1.20 1.59 2.02 0.31
Al 20 3 0.11 0.15 0.51 0.97 0.55 0.38 0.24 0.22 0.24 0.03
Cr203 0.01 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.
FeOr 6.77 7.59 9.74 10.76 12.39 13.28 15.97 17.58 19.50 21.63
MnO 0.21 0.17 0.38 0.24 0.34 0.20 0.17 0.10 0.27 1.07
MgO 14.29 13.75 12.35 11.68 10.43 10.28 8.68 7.39 5.89 4.56
CaO 22.36 22.12 23.27 21.80 20.92 16.60 14.34 12.09 10.05 10.62
Na20 1.53 1.69 0.86 1.52 2.07 4.62 5.96 7.11 7.98 7.61
98.96 99.02 99.44 98.63 98.86 97.84 98.01 97.62 96.92 97.21

Fe203 4.65 5.02 3.99 6.02 6.30 13.45 16.69 18.06 19.32 20.02
FeO 2.59 3.07 6.15 5.35 6.72 1.18 0.95 1.33 2.12 3.61
99.33 99.52 99.84 99.23 99.49 99.19 99.68 99.43 98.86 99.22

Structural fonnula on the basis of four cations and six oxygens


Si 1.982 1.977 1.942 1.911 1.943 1.935 1.942 1.953 1.954 1.984
AI 0.006 0.010 0.022 0.037 0.031 0.035 0.034 0.045 0.011 0.001
Ti 0.005 0.007 0.023 0.043 0.025 0.017 0.011 0.010 0.058 0.009
Fe 3+ 0.130 0.141 0.113 0.172 0.180 0.382 0.474 0.515 0.557 0.582
Fe2+ 0.080 0.095 0.194 0.170 0.214 0.037 0.030 0.042 0.068 0.117
Mn 0.007 0.005 0.012 0.008 0.011 0.006 0.005 0.003 0.009 0.035
Mg 0.791 0.762 0.693 0.661 0.591 0.578 0.488 0.417 0.337 0.262
Ca 0.890 0.881 0.939 0.886 0.853 0.671 0.580 0.491 0.413 0.439
Na 0.110 0.122 0.063 0.112 0.153 0.338 0.436 0.522 0.593 0.570

End member molecules (mol %)


CaTiAl20 6 0.2 0.3 1.1 2.1 1.2 0.8 0.5 0.5 0.6 0.1
CATS 0 0 0 0 0 0 0 (l 0 0
Ae 11.0 12.2 6.2 11.1 15.2 33.6 43.3 52.4 60.1 57.8
Wo 44.3 43.8 46.1 42.8 42.0 32.9 28.5 24.4 20.7 22.3
Fs 5.0 5.7 12.1 11.4 12.1 4.0 3.4 1.7 1.6 6.5
En 39.5 38.0 34.4 32.7 29.5 28.7 24.2 20.9 17.1 13.3
al_5. Sover North; 6--9. Postmasburg; 10. Voorspoed. FeOr = total iron expressed as FeO. Fe203 and FeO calculated by the
method of Droop (1987). End-member molecules calculated assuming all Na20 is present as aegirine. CATS = CaAlSiAlO6.
n.d. = not detected.

Swartruggens intrusion does not contain pyroxenes significantly different to those in the
younger orangeite intrusions.
Figure 2.65 shows that microphenocrystal pyroxenes from different facies of the
Finsch intrusion are very similar in composition. Pyroxenes found in the different dikes
of the Swartruggens system are also identical in their composition. These observations
suggest that pyroxene compositions are of little use in determining the sequence of
intrusion of different batches of orangeite magma in complex bodies.
MINERALOGY OF ORANGEITES 169

- Fe

-
SWARTRUGGENS
~_.L. . . 'r .,. I I I I ~

- -- LACE

--",-- ,-
--I. VOORSPOED
46
-
48

Figure 2.63. Compositions of mi-


crophenocrystal clinopyroxenes in di-
verse orangeites plotted in the ternary
48
-
system Ca-Mg-Fe (atomic). All data
- SILVERY HOME
this work except Silvery Home
(Robey 1981). Fe

The low Al content of the diopsides commonly results in there being insufficient Al
to fill the tetrahedral sites in the pyroxene structure; thus, Si + Al is commonly less than
2.0 afu. The tetrahedral site deficiency of 0.0 15-0.035 afu may be remedied by the entry
of Fe3+ to this site, as typically Fe3+ is greater than, and Ti less than, the site deficiency.
The low Al and Cr contents also typically result in Na being greater than Al + Cr, i.e., the
pyroxenes contain an aegirine component rather than kosmochlor or jadeite.
Significant compositional zonation is found only in pyroxenes in the more evolved
varieties of orangeite, e.g., Sover North, Postmasburg, Makganyene, and is toward Fe
and Na enrichment at the margins of the crystals (Table 2.19, Figure 2.66). This trend
culminates in the formation of aegirine-rich mantles (see below).
-Fe ..

1
49
• • •
48 ·t'

D
Co
••
,~ ~
47 • • •
•• • • •
() 46
(j 45 ,
.. I •
/ 44 •• Mg Fe
43 •

BLAAUWBOSCH
••

NEW ELANDS
Figure 2.M. Compositions of microphenocrystal c1inopyroxenes from the Blaauwbosch (Scott Smith and
Skinner unpublished) and New Elands (this work) orangeites plotted in the ternary system Ca-Mg-Fe (atomic).

Fe •
2 345 6 7

F2

F4

••• F7

Mg Fe FINSCH
Figure 2.65. Compositions of microphenocrystal clinopyroxenes in different intrusions within the Finsch
orangeite (this work) plotted in the ternary system Ca-Mg-Fe (atomic).

170
MINERAWGY OF ORANGElTES 171

Fe



POSTMASBURG

.. .
••••
•• •
.. •
.~.

. ~
\
• MAKGANYENE
46

.... ..
l~l!
••• •
.
• SOVER NORTH

Figure 2.66. Compositions of clinopyroxenes from the Postmasburg (this work), Makganyene (this work), and
Sover North (this work, Tainton 1992) orangeites plotted in the ternary system Ca-Mg-Fe (atomic).

2.2.2.2. TItanian Aegirine


Figure 2.66 illustrates the pyroxene compositional variation found in the Postmas-
burg, Makganyene, and Sover North intrusions. Here, pyroxenes are zoned from Fe-poor
cores, identical in composition to the unzoned diopsides found in other intrusions, to
margins which are richer in Fe and Na. This compositional trend reflects aegirine
enrichment (Table 2.19) and culminates in the formation of discrete aegirine mantles
(Figure 2.67). Similar compositional trends have been reported by Tainton (1992) from
the Pniel occurrence, although aegirines in these rocks are relatively poor in Ti02 «1
wt%).
Table 2.20 gives representative compositions of mantling and groundmass aegirine-
rich pyroxenes. TheirTi02 contents range from 0.3 to 6.6 wt %. The more Ti-rich varieties
may be regarded as titanian aegirine. Recalculation of the compositions on a stoichiomet-
ric basis indicates low-to-zero Fe3+ and high Fe2+ contents. Calculation of potential
end-member pyroxene molecules suggests that they are members of an unusual quater-
nary solid solution between Na(Fe2+o.s,Tio.s}Sh06, Na(Mgo.s,Tio.s}Sh06, diopside, and
aegirine (Table 2.20). In these pyroxenes all Fe2+is considered to be combined with Na
172 CHAPTER 2

Ae

-#
-,
90
I
80 1 Ae

~!7oi ~ Hd
~ 60 Oi Hd
0\0
.
.$
~

50

/ 40 ) • POSTMASBURG
+ SOVER NORTH
co.• o MAKGANYENE
• VOORSPOEO

o
o +
:a+ .ut+
o~·..~·~
Oi Hd
10 20 30
Figure 2.67. Compositions of evolved aegirine-rich clinopyroxenes from the Posttnasburg, Sover North,
Makganyene, and Voorspoed orangeites plotted (mol %) in the ternary system diopside (Di)-hedenbergite
(Hd)-aegirine (Ae). All data this work.

in the Na(Fe2+o.5,Tio.5)Sh06 molecule. Consequently, the pyroxenes are considered not


to contain the hedenbergite molecule. This recalculation scheme is considered the most
plausible as it is the only one which results in the number of cations not assigned to a
pyroxene molecule being less than 10% of the total cations present, i.e., <0.4 afu. Other
recalculation schemes, including the hedenbergite molecule, result in a significant excess
of unassigned Na and Ti.

2.2.2.3. MilWr Elements


Table 2.21 summarizes the minor element variation exhibited by microphenocrystal
clinopyroxenes. Each intrusion is characterized by pyroxenes of similar compositional
range, and only minor inter-intrusional differences exist, e.g., Voorspoed and Lace contain
pyroxenes which are slightly lower in Ti, AI, and Cr, relative to other intrusions. The
Swartruggens pyroxenes are notably very low in AI.
MINERALOGY OFORANGEITES 173

Table 2.20. Representative Compositions of Aegirine and Titanian AegirineQ


Wt% 2 3 4 5 6 7
Si02 51.94 50.24 51.44 51.58 51.18 50.82 50.83 52.91 52.29
TI02 1.55 2.42 4.27 4.72 4.78 4.72 4.17 0.30 0.80
Ah03 n.d. 0.17 0.21 0.33 0.25 0.29 0.27 0.25 0.80
Fe203 29.55 29.55 26.47 24.49 22.39 26.07 22.23 22.21 25.61
FeO 0.0 0.0 0.0 0.0 3.71 0.83 2.61 0.0 0.41
MnO 0.09 0.24 0.16 0.34 0.34 0.53 0.21 0.19 0.10
MgO 1.85 1.24 2.51 2.01 1.53 0.94 2.28 2.45 3.58
CaO 2.08 2.46 2.09 2.72 2.47 1.72 4.64 0.39 4.84
Na20 12.68 12.48 12.48 13.09 11.98 12.89 11.15 13.08 10.82
-- --
99.74 98.80 100.99 99.29 98.63 98.81 98.39 95.75 98.63

Structural fonnula on the basis of four cations and six oxygens


Si 1.986 1.965 1.955 1.962 1.979 1.960 1.967 2.018 2.005
AI 0.008 0.009 0.015 0.011 0.013 0.012 0.009 0.023
TI 0.045 0.071 0.122 0.135 0.139 0.137 0.121 0.011 0.009
Fe3+ 0.879 0.866 0.757 0.756 0.651 0.757 0.647 0.903 0.739
Fe2+ 0.120 0.027 0.084 0.013
Mn 0.003 0.008 0.005 0.011 0.011 0.017 0.007 0.006 0.003
Mg 0.105 0.072 0.142 0.114 0.088 0.054 0.132 0.139 0.205
Ca 0.085 0.103 0.141 0.111 0.102 0.071 0.192 0.016 0.198
Na 0.940 0.947 0.919 0.965 0.898 0.964 0.837 0.967 0.805

End member molecules (mol %)


Esseneite 0.9 1.1 1.6 1.2 1.5 1.4 1.2 0.9
NaFeTIPX 19.6 4.4 14.2 2.0
NaMnTiPX 0.5 1.3 0.9 1.8 1.8 2.9 1.9 1.0 0.5
NaMgTIPX 6.4 10.1 19.6 18.6 1.3 9.0 5.1 0.4 1.1
Diopside 6.5 1.0 2.8 8.7 11.4 1.7 20.4
Aegirine 86.7 86.7 75.6 78.0 67.4 82.3 66.8 95.7 75.2
°1_2, Voorspoed; 3, Makganyene; 4-5, Sover Nonh; 6-7, Postmasburg; 8-9, Pniel. Fe:z03 and FeO calculated by the method
of Droop (1987). NaFeTIPX = NaFeo.sTIo.sSh06; NaMnTIPX = NaMno.sTIo.sSi:z06;NaMgTIPX = NaMgo.sTIo.sSh06-

Figures 2.68 and 2.69 illustrate Ti versus Al contents. Well-defined compositional


trends are absent. A weak correlation between increasing Ti and Al is evident at New
Elands and Voorspoed, but typically Ti is independent of AI. Zoning, when present, is
very weak and follows a trend of increasing Ti and AI. The Ti and Al contents are typically
<0.07 and <0.05 (commonly <0.03) afu, respectively. Figure 2.69 shows that pyroxenes
from the Finsch F4 intrusion fall into two groups. Those from the low-AI group have
significantly lower Al contents than pyroxenes in the F2 and F7 intrusions, while those
from the group with relatively high Al content are identical in composition to pyroxenes
from other Finsch intrusions. Pyroxenes from the various dikes occurring at Swartruggens
are identical in their minor element compositions.
Cr contents vary widely (Table 2.21) and show no correlation with Al or Ti. Cr is not
detectable by electron microprobe in the more-evolved relatively Fe-rich pyroxenes.
The trace element contents of microphenocrystal pyroxenes have not been determined.
174 CBAPTER2

Table 2.21. Minor Element Content of Microphenocrystal Clinopyroxenea


Location Ti02 Cr203 Ah03 N Source
B1aauwbosch 0.22-1.97 n.d. - 2.14 0.05 -0.64 48 2
Besterskraal 0.36 - 1.15 0.13 - 0.89 0.30-0.89 20
FinschF2 0.42 - 1.56 0.08-0.65 0.16 -1.62 15
FinschF4 0.45 - 2.48 0.13-0.74 0.09-0.58 14
Finschf7 0.49 - 1.32 0.04-0.85 0.03 -0.74 20
Lace 0.52 - 1.38 n.d.-0.45 0.06-0.26 21
Makganyene 0.24-0.63 n.d. - 1.72 0.15 -0.61 38 1,2
NewElands 0.10 - 1.85 n.d.-0.71 0.02 -1.53 35 1,3
Pniel 0.19 - 0.83. 0.11-0.90 0.13 -0.64 10 6
Postmasburg 0.17 -1.06 n.d.-O.72 0.03 -1.04 22 1
Roberts Victor 0.25-0.79 0.18 - 1.20 0.06 -0.38 13 4
Skietkop 0.76-1.17 0.11- 0.55 0.13 -1.03 8
Silvery Home 0.78-2.74 0.17 - 0.26 0.15 -0.97 14 1,5
SoverNorth 0.26 - 1.52 n.d.-0.86 0.06 -1.26 29 1,2
Swartruggens Main 0.36-0.88 n.d.-0.46 0.15 -0.58 14 1
Swartruggens 0.48 - 0.76 0.15 - 0.23 0.18 -0.29 8 2
Changehouse
Voorspoed 0.32-0.74 n.d.-O.40 O.ll - 0.61 26
"Data sources: I, this work; 2, Scott Smith and Skinner (pers. comm.); 3, Mitchell and Meyer (1989b); 4, Dawson et al. (1977);
= =
5, Robey (1981); 6, Tainton (1992). N number of analyses. n.d. not detected.

0'06 - • SWARTRUGGENS
- • BLAAUWBOSCH
+

t
0'05 - + NEW ELAN OS +

0'04 ~-
I- 0'03-
+•
I 0'02~-
0'01-
-
o I I I I I I I I I I I I I I I I I I I I I I
o 0'005 0'010 0'015 0'020

Figure 2.68. Al versus Ti (atomic) compositional variation of clinopyroxenes from the Swartruggens (this
work), Blaauwbosch (Scott Smith and Skinner unpublished), and New Elands (this work) orangeites.
• SOVER NORTH 0 LACE
0'07 ,,", X MAKGANYENE • BESTERSKRAAL
\
,, \ • FINSCH \7 POSTMASBURG
\
, \
0'06
. t (F2,F4,F7) • VOERSPOED
," " ..... FINSCH - F4
, FIELD OF
, \ \7 X
I \ LAMPROITE
0·05 I \
, PYROXENE~
t 0'04 I
, 0
\
\ o
\7 • • •
,(5) \7
., &
i=
, 0 '
0'03
• • ••
, • ----.. •
'. 0 r •• ·fill • • • •• \7'"\
\,. ,0 ·•
• FINSCH-F4~...... •
. . • • :,el
0'02 " 0 •• • • t;..XY \7 ••
I... e·.·
... • • •
..... •• X .~ • '\kl
\fQ--.r£./ •• ~x. t· •
0'01 ~ Xx \7 ••
\7"V.
167.· X.~
·. x. '...
\7
0
0 0'005 0'010 0'015 0'020 0'025 0'030 0'035 0'040

AI
Figure 2.69. AI versus Ti (atomic) compositional variation of c1inopyroxenes from diverse orangeites. All data this work. Compositional field of lamproite pyroxene from
Mitchell and Bergman (1991).
...~
176 CHAPTER 2

2.2.3. Pyroxenes in the Swartruggens Male Lamprophyre


This dike contains up to 40 vol %, euhedral-to-acicular (0.4 mm long, length-to-
breadth ratios 8: 1) diopsides (Dawson et al. 1977). The pyroxenes are strongly-zoned
toward Fe and Na-enriched margins (Table 2.22). Figure 2.70 shows that the relatively
unevolved pyroxenes, constituting the bulk of the population, have compositions similar
to those of unzoned pyroxenes in the Swartruggens orangeites (Figure 2.63). Figure 2.71
shows that the Male lamprophyre pyroxenes exhibit a zoning trend toward enrichment in

Table 2.22. Representative Compositions of Pyroxenes in the Swartruggens Male


Lamprophyre Dike a
Wt% 2 3 4 5 6 7 8
Si02 51.90 52.54 50.83 50.36 49.91 48.41 48.50 47.51
Ti0 2 1.13 1.35 1.61 1.50 1.99 1.86 2.64 2.46
AI 20 3 1.23 1.29 1.67 0.93 1.22 1.55 1.42 2.13
Cr203 0.54 0.04 0.0 0.0 0.0 0.0 0.0 0.0
FeOr 3.35 4.47 6.49 8.21 10.81 11.62 13.63 14.14
MnO 0.11 0.08 0.14 0.23 0.27 0.40 0.33 0.36
MgO 16.72 16.08 14.67 13.29 1l.39 10.74 9.72 10.88
CaO 23.82 24.06 23.69 24.17 22.74 22.30 20.84 20.05
Na20 0.34 0.28 0.39 0.47 1.15 0.85 1.62 2.64
99.14 100.19 99.49 99.16 99.48 97.77 98.70 100.17

Fe203 2.53 1.44 3.01 4.12 4.87 4.58 5.70 14.23


FeO 1.07 3.18 3.78 4.50 6.43 7.50 8.50 1.34
99.39 100.33 99.79 99.57 99.97 98.19 99.27 101.59

Structural fonnula on the basis of four cations and six oxygens


Si 1.911 1.925 1.890 1.896 1.889 1.897 1.869 1.779
Al 0.053 0.056 0.073 0.041 0.054 0.071 0.064 0.094
Ti 0.031 0.037 0.045 0.042 0.057 0.054 0.077 0.069
Cr 0.016 0.001
Fe 3+ 0.070 0.040 0.084 0.117 0.139 0.133 0.165 DAOI
Fe 2+ 0.033 0.097 0.118 0.142 0.204 0.243 0.274 0.042
Mn 0.003 0.002 0.004 0.007 0.009 0.013 0.011 0.011
Mg 0.918 0.878 0.813 0.746 0.643 0.620 0.558 0.607
Ca 0.910 0.944 0.944 0.975 0.922 0.926 0.860 0.804
Na 0.024 0.020 0.028 0.034 0.084 0.064 0.121 0.192

End member molecules (mol %)


CaTiAI 20 6 2.7 2.8 3.6 2.0 2.7 3.5 3.2 4.5
CATS 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Ae 2.4 2.0 2.8 3.4 8.4 6.3 12.0 18.3
Wo 45.4 45.7 44.9 46.9 44.4 44.1 41.2 36.2
Fs 3.9 5.8 8.6 11.0 12.8 15.5 15.8 12.0
En 45.6 43.7 40.2 36.7 31.8 30.7 27.8 29.0
aFeOr = total Fe calculated as FeO. Ft:203 and FeO calculated by the method of Droop (1987). End-member molecules
calculated assuming all Na20 is present as aegirine. CATS = CaAlSiAl06.
MINERALOGY OF ORANGEITES 177

Co
- - Fe •
5 10 15 20
I • Fe
• •• •

Hg Hg
L Fe

Figure 2.70. Compositions of clinopyroxenes in the Swartruggens Male lamprophyre plotted in the ternary
system Ca-Mg-Fe.

hedenbergite. The trend is similar to those found in some orangeites but is displaced from
the main trend of evolution of orangeite pyroxenes.
The Male lamprophyre pyroxenes are not significantly enriched in minor elements
(Table 2.21) and are weakly-zoned toward increasing Ti and Al from core to margin. They
contain significantly more Al than Swartruggens orangeite pyroxenes (Table 2.18).

2.2.4. Megacrystal Pyroxenes


Unlike archetypal kimberlites, orangeites do not characteristically contain a suite of
subca1cic diopside megaimacrocrysts. Chrome-poor clinopyroxene macrocrysts have
been found in heavy mineral concentrates from Wielpan South and Silvery Home
(Skinneret al. 1994) and have also been noted from Lace by Bell and Read (pers. comm.).

• Ae
~1
~
~ If!
0\0
"I
~. 20 •
~ ~~I Di Hd

,: ..
~~ / .
/ ••
I
/ 10
", /
• •

DiL----T~~,-----r----,------~ Hd
10 20 30 40
Mol. % Hd ---.
Figure 2.71. Compositions of clinopyroxenes in the Swartruggens Male lamprophyre plotted (mol %) in the
ternary system diopside (Di)-hedenbergite (Hd)-aegirine (Ae). Main orangeite trend from Figure 2.67.
178 CHAPTER 2

Many of the Cr-rich megacrysts reported from the Barldy West dikes (Hops et al. 1992)
are probably derived from eclogites (Moore and Gurney 1991). Compositional data for
megacrystal pyroxenes have not been published.

2.2.5. Comparison with Pyroxenes in Kimberlites


Bona fide primary pyroxenes are considered not to occur in archetypal hypabyssal
kimberlites. Pyroxenes commonly occur as reaction mantles around crustal xenoliths, and
such pseudoprimary pyroxenes crystallize only from magmas which have been contami-
nated by such xenolithic material. Their mode of formation is similar to that advocated
by Scott Smith et al. (1983) for pseudoprimary pectolite. The composition of the very
fine grained pyroxenes occurring in groundmass of diatreme facies rocks has not been
determined.
Pseudoprimary phenocrystal and/or groundmass pyroxenes from the Premier type 3
(Scott and Skinner 1979, six analyses) and Orroroo (Scott Smith et al. 1984, three
analyses) hypabyssal kimberlites are similar in their major element composition (Table
2.23, Figure 2.72) to the least Fe-rich pyroxenes in orangeites. They differ in being
relatively poor in AI, Ti, and Cr (Table 2.23). The few data for pyroxenes (Table 2.23,
Figure 2.72) from the Schuller pipe (Scott and Skinner 1979, three analyses) in contrast
are relatively Fe rich and variable in their minor element content relative to orangeite
pyroxenes. Diopside (mg =0.92) in the Aries kimberlite (Edwards et al. 1992) is very
poor in Ah03, Cr203, and Na20 «0.1 wt %) and Ti02 (0.15 wt %). These small (5-u0
~m) pyroxenes of irregular habit are intergrown with serpentine and talc and are probably
secondary in origin.
Microphenocrystal pyroxenes in the Zagodochnaya kimberlite (Egorov et al. 1991;
Table 2.23) are low in Ah03 (O.3-Do4 wt %), Ti02 (0.1 wt %), Cr203 (0.6-1.1 wt %), and
FeO (2.4-304 wt %). Other AI- and Cr-rich pyroxenes in this kimberlite appear to be
derived from disaggregated upper mantle xenoliths.
Low-Cr, AI-poor pyroxenes in mica-rich xenoliths occurring in the East Udachnaya
kimberlite (Egorov et al. 1986) are of similar composition to those of orangeites and

Table 2.23. Compositions of Pyroxenes in KimberJitea


Wt% 2 3 4 5 6 7 8
Si02 54.66 53.84 54.39 53.78 54.63 53.94 56.16 52.2
Ti0 2 0.19 0.88 0.22 0.57 0.25 1.08 0.08 1.14
Al 20 3 0.01 0.06 0.01 0.01 0.22 0.39 0.36 0.86
Cr203 0.12 0.13 n.d. n.d. 0.13 0.12 1.10 n.d.
FeOT 2.88 3.95 1.64 3.06 4.54 6.81 2.36 4.74
MnO 0.10 0.10 0.13 0.21 0.18 0.17 0.06 0.10
MgO 16.96 15.52 17.78 16.70 15.46 14.12 15.79 16.0
CaO 24.85 24.05 25.58 25.11 21.88 22.90 24.13 25.6
NazO 0.54 0.90 0.14 0.27 1.63 1.23 0.82 0.28
100.31 99.34 99.89 99.71 99.92 100.76 100.85 100.84
= =
"FeO,- total Fe expressed as FeO; n.d. not detected. Compositions 1-2, Premier (Scott and Skinner 1979); 3-4 ,Orroroo
(Scott Smith el al. 1984); 5-6, Schuller (Scott and Skinner 1979); 7, Zagodochnaya (Egorov el al. 1991); 8, Udachnaya
(Egorov el al. 1986).
MINERALOGY OF ORANGElTES 179

Co
-Fe~

-+-L-L-L-L-L-L-L-L--L--L--L-L-L-L-L-I~ Fe

40
LAMPROITES

Mg

Figure 2.72. Compositions of clinopyroxenes from orangeites compared with those of clinopyroxenes in
kimberlites (Orroroo, Scott Smith et al. 1984; Premier and Schuller, Scott and Skinner 1979), minettes (Agathla,
Jones and Smith 1983; Devon, Jones and Smith 1985; Colima, Luhr and Carmichael1981; Allen and Cannichael
1984; Wallace and Carmichael 1989; Linhaisai, Bergman et al. 1988), and lamproites (Mitchell and Bergman
1991).

lamproites (Table 2.23). The xenoliths are petrographically unlike orangeites and appear
to be phlogopite pyroxenite cumulates.
Pyroxenes are relatively common in the atypical, phlogopite kimberlites from West
Greenland (Scott 1981). These are very different to orangeite pyroxenes, being relatively
rich in Ti02 (0.9-2.8 wt %, Scott 1981; 0.1-4.6 wt%, this work) and FeOr (4.0--6.1 wt %,
Scott 1981; 1.7-8.2 wt %, this work). Their high Alz03 contents (0.9-4.5 wt %, Scott
1981; 0.3-10.1 wt %, this work) suggests an affinity with ultramafic lamprophyres such
as alnoites or aillikites (see 2.2.6).

2.2.6. Comparisons with Pyroxenes in Lamproites


The major element composition of microphenocrystal pyroxenes in lamproites
(Mitchell and Bergman 1991) is identical to that of orangeite pyroxenes (Table 2.24,
Figure 2.72). Lamproite pyroxenes typically are not zoned, although examples from the
Leucite Hills, in rare instances, may be zoned to Fe-rich margins and exhibit acmitic
mantles. However, Fe enrichment does not exceed 10 wt % FeOr and titanian aegirine is
not found.
Figures 2.69 and 2.73 show that most lamproite microphenocrystal pyroxenes have
very low «0.03 afu) and Ti «0.07 afu) contents. The majority of orangeite pyroxenes
are of identical composition, although the compositional range for Al extends to 0.05 afu.
180 CHAPTER 2

Table 2.24. Representative Compositions of Pyroxenes from Lamproitesa


Wt% 2 3 4 5 6 7 8 9 10
Si0 2 54.41 53.13 54.25 52.96 54.28 54.01 53.14 51.99 52.16 51.85
Ti02 0.74 2.03 0.73 1.41 1.69 2.18 1.36 2.08 1.48 1.17
AI 20 3 0.04 0.04 0.32 0.26 0.22 0.25 0.23 0.41 0.59 1.27
CrZ0 3 n.d. 0.14 0.06 0.05 0.38 n.d. 0.16 0.19 n.d. 0.92
FeOT 2.05 2.47 2.44 9.31 3.18 3.92 2.53 3.21 4.58 3.51
MnO 0.11 0.06 0.13 0.17 n.d. n.d. 0.08 0.10 0.08 0.06
MgO 17.% 17.29 17.87 13.00 17.83 17.54 17.48 16.56 16.06 16.52
CaO 24.49 23.24 24.11 17.36 22.10 22.09 24.80 24.57 23.61 23.40
NazO 0.29 0.40 0.34 3.74 n.d. n.d. 0.22 0.46 0.61 0.44
100.09 99.30 100.25 98.26 99.68 99.99 100.00 99.57 99.17 99.14
"FeOr = total Fe expressed as FeO; n.d. = not detected. Compositions 1-2, West Kimberley (Mitchell and Bergman 1991);
3-4, Leucite Hills (Mitchell and Bergman 1991); 5-6, Smoky Butte (Mitchell el al. 1987); 7-8, Prairie Creek (Scott Smith
et al. 1984); 9--10, Kapamba (Scott Smith el al. 1989).

Only the Kapamba lamproite pyroxenes exceed these Al contents (Figure 2.73). In
common with orangeite pyroxene there is insufficient Al present to occupy all of the
tetrahedral sites in the structure.

2.2.7. Comparisons with Pyroxenes in Ultramafic Lamprophyres


Ultramafic lamprophyres typically contain highly aluminous pyroxenes, e.g.,
Wauboukigou Province (Illinois-Kentucky), 1.0-2.0 wt %; Alna, 0.5-10.0 wt %; Oka,

ORANGEITES
0·10
LAMPROITES ., ,
,,

t
",

0'08
.,
., .,
",

,
",
",

0'06
......
0'04 ROMAN PROVINCE

--- -
0'02
--- --- ---
0
0 0'10 0'20 0'30 0'40
-- AI

Figure 2.73. AI versus Ti (atomic) compositional fields of clinopyroxenes from orangeites (this work), di verse
lamproites (Mitchell and Bergman 1991), Kapamba 1amproites (Scott Smith et al. 1989), Roman Province lavas
and Ugandan kamafugites (Mitchell and Bergman 1991).
MINERALOGY OF ORANGElTES 181

6.3-12.1 wt % Ah03 (Mitchell, unpublished data). Ti02 contents (wt %) are also com-
monly high, e.g., Wauboukigou, 0.8-S.2 wt %; Alno, 0.9-S.9; Como, 1-2; Oka, 2.O--S.4;
Haystack Butte, 1.6-2.5; Polzen, I.S-S.l (Mitchell unpublished data). The combination
of high Al coupled with high Ti indicates that these pyroxenes are members of the solid
solution series between CaMgSh06 (diopside )-CaAlAISi06 (Ca-Tschermak's pyroxene)
and CaTiAh06. Such pyroxenes typically contain Al vi and are thus unlike pyroxenes
found in orangeites or lamproites.
The diamond-bearing Bulljah Pool ultramafic lamprophyres contain pyroxenes with
relatively low Ah03 «1.0 wt %) and high Ti02 (1-2 wt %) contents (Hamilton and Rock
1990). Late-stage aegirine-augite and aegirine are present in "alteration assemblages" in
these rocks. Unfortunately, compositional data are not provided by Hamilton and Rock
(1990), and they cannot be compared with late-stage titanian aegirines found in some
orangeites.
Primitive olivine-mica-sanidine lamprophyre dikes from Mt. Bundey, Australia
(Sheppard and Taylor 1992), have petrographic affinities with lamproites and evolved
orangeites. Pyroxenes in these rocks differ from orangeite pyroxenes in being oscillatory-
zoned, in containing exsolved orthopyroxene and having high-Ah03 (1.3-3.2 wt %)
contents.
Primary pyroxenes in the Bow Hill lamprophyres (Fielding and Jaques 1989) are not
significantly different from those of orangeites.

2.2.S. Comparisons with Pyroxenes from Minettes


Figure 2.72 shows that pyroxenes from orangeites and minettes (Luhr and Car-
michael 1981, Jones and Smith 1983, 1985, Allen and Carmichael 1984, LeCheminant
etal. 1987, Bergman etal. 1988, O'Brien etal. 1988, Wallace and Carmichael 1989)
cannot be distinguished on the basis of their major element composition. However,
pyroxenes in minettes are commonly enriched in Al relative those of orangeites (Figure
2.73). Although some pyroxenes of low Ah03 content «1.0 wt %) may be found in
minettes, the majority contain from 1.0 to 8.0 wt % Ah03. Their compositional evolution
toward increasing AI, from core to margin, serves to distinguish them from orangeite
pyroxenes. In addition, many pyroxenes in minettes are complexly-zoned and mantled
(LeCheminant et al. 1987, O'Brien et al. 1988, MacDonald et al. 1992).

2.3. OLIVINE

2.3.1. Paragenesis
Olivine is a ubiquitous constituent of diatreme and hypabyssal facies orangeites.
Olivines in diatreme facies rocks have not been extensively studied as only macrocrysts
appear to be present. These are characteristically completely pseudomorphed by serpen-
tine.
In hypabyssal facies orangeites, olivine is typically present as discrete grains which
exhibit a wide range in size and modal abundance. The larger grains are typically rounded,
whereas the smaller crystals may be either rounded or subhedral to euhedral. These
observations suggest that, in common with archetypal kimberlites, two groups of olivine
182 CHAPrER2

are present: rounded macrocrysts which are primarily xenocrysts, and subhedraVeuhedral
phenocrysts/microphenocrysts. Rounded microcrysts are considered to be a part of the
xenocrystal suite.
It is not always possible to distinguish accurately between olivine xenocrysts and
phenocrysts. Skinner (1989) has suggested that all grains which exhibit defonnation
features (undulose, extinction, kink banding, recrystallization) or contain inciusions of
other lherzolitic-suite minerals are xenocrysts. All strain-free crystals exhibiting planar
crystal faces are believed to be phenocrysts. However, there is no a priori reason why all
xenocrysts should be strained, and Mitchell (1986) considers that deformation-free
rounded macrocrysts may be either xenocrysts or resorbed phenocrysts. Unfortunately,
there are no simple compositional parameters which might permit discrimination between
these two groups. Moore (1988) has noted that olivines in orangeites from Newlands,
Sover, and Bellsbank very rarely show any evidence of undulose extinction.
In orangeites, rounded macrocrysts may constitute up to 50 vol % of the rock (Bosch
1971, Skinner and Clement 1979, Fraser 1987). However, modal abundances vary widely
within and between intrusions as a consequence of flow differentiation (Tainton 1992).
On average some orangeites, i.e., Bellsbank, Sover, appear to be enriched in macrocrystal
olivine relative to others, e.g., New Elands, Swartruggens. Evolved orangeites are
relatively poor in olivine. Macrocrysts may be fresh, partially or completely serpenti-
nized, and/or carbonatized.
Some macrocrysts in the Bellsbank, Blaauwbosch, Roberts Victor, and Sanddrift
orangeites exhibit pronounced overgrowths of optically and compositionally distinctive
olivine which may be up to 0.5 mm in width (Skinner 1989, Tainton 1992). These
overgrowths may be nonnally- or reversely-zoned relative to the cores (see 2.3.2).
Phenocrysts are considered to be euhedral-to-subhedral crystals, commonly 0.1 to
2 mm in longest dimension, although the majority are less than 0.5 mm (Skinner 1989).
Abundances vary widely within and between intrusions, i.e., from 8 to 23 vol % at Finsch
(Fraser 1987). The ratios of xenocrysts to phenocrysts typically range between 1: 1 and
5: 1 according to Skinner (1989), although he notes that higher values may also be found.
The extreme variation in xenocryst/phenocryst ratios is exemplified by the Bellsbank
intrusions. Here, Bosch (1971) did not observe any phenocrysts, whereas Tainton (1992),
using the recognition criteria of Skinner (1989), reports that they are common. The
differences may be real or simply reflect the difficulties in distinguishing the two
populations.
Phenocrysts may be fresh, partially or completely altered to serpentine, and/or
replaced by carbonate. Chromite is the only inclusion recognized in these olivines by
Skinner (1989).
Skinnner (1989) has noted that many ofthe larger phenocrysts in the Blaauwbosch,
Roberts Victor, and Sanddrift orangeites exhibit olivine overgrowths similar to those
observed occurring on coexisting xenocrysts.
Olivines in evolved orangeites exhibit many of the characteristics described above.
They differ in containing macrocrysts and "phenocrysts" which are mantled by parallel
aggregates (dog's tooth habit) of prismatic olivine. Examples have been described from
Postmasburg, Sover North (SN1), and the Bellsbank West Fissures by Tainton (1992) and
the Prieska region by Skinner et al. (1994). Olivines of this morphology are considered
MINERALOGY OF ORANGElTES 183

to be either primary growth features resulting from the rapid crystallization of primary
olivine on macrocrystal nuclei (Mitchell and Bergman 1991) or an imposed morphology
produced during crystallographically controlled resorption of macrocrysts (Scott Smith
et al. 1989). Olivine of this habit is common in olivine lamproites (Mitchell and Bergman
1991) but has not been observed in archetypal kimberlites.
Macrocrystal olivine mantled by phlogopite, optically and compositionally similar
to groundmass phlogopite is common in evolved orangeites from Pniel, Sover North, and
Besterskraal (this work, Tainton 1992). Such reaction mantles are not observed in
unevolved orangeites.

2.3.2. Composition
Compositional data for phenocrystal (Table 2.25) and macrocrystal olivines have
been provided by Skinner (1989), Moore (1988), Fraser (1987), and Tainton (1992) and
for macrocrystal olivines by Mitchell and Meyer (1989a). Studies of olivine composi-
tional variation are hindered by the common alteration and pseudomorphing of olivines.
Commonly, small microphenocrystal (groundmass) olivines are completely pseudomor-
phed by serpentine and/or calcite. Alteration at crystal margins is a particular impediment
to determining the compositional evolution of olivines. As a consequence the majority of
the compositional data available are for the cores of crystals.
Skinner (1989) and Moore (1988), in general studies of orangeite olivines, have
shown that, although indi vidual crystals are of uniform composition, there is considerable
intergrain compositional variation. Typically, phenocrysts and xenocrysts of different
composition are juxtaposed, suggesting that the assemblage results from the mixing of
olivines derived from several sources. Figure 2.74 shows that the cores of the phenocrysts
and xenocrysts studied by Skinner (1989) range in mg from 0.87 to 0.95 [mg =Mg/(Fer +
Mg)]. However, the mode for the xenocrysts lies at a slightly more magnesian composition
(mg = 0.93) than that of the phenocrysts (mg = 0.91-0.92) and is similar to the mean
composition of olivine in peridotite xenoliths. Similar conclusions may be drawn from

Table 2.25. Representative Compositions of Olivine Phenocrystsa


2 3 4 5

Wt% C R C R C R C R C R
Si02 41.00 40.55 40.59 39.77 40.85 41.13 40.46 40.63 40.40 39.94
TiOz 0.03 n.d. 0.03 0.04 n.d. n.d. n.d. n.d. n.d. n.d.
FeOr 6.65 9.86 10.36 8.65 7.49 8.23 8.30 7.36 7.78 9.16
MnO 0.12 0.16 0.12 0.14 0.09 0.12 0.11 0.12 0.11 0.17
MgO 51.91 48.97 48.33 49.45 51.18 50.52 50.38 51.37 51.21 50.03
NiO 0.42 0.32 0.46 0.28 0.35 0.35 0.42 0.36 0.47 0.16
CaO n.d. 0.06 0.04 0.10 0.01 0.04 0.06 0.09 n.d. 0.03
100.13 99.92 99.93 98.43 99.97 100.39 99.73 99.93 99.97 99.49
mg 0.933 0.899 0.893 0.911 0.924 0.916 0.915 0.926 0.922 0.907
"FeOr = total Fe expressed as FeO; n.d. = not detectable. 1, Bellsbank (Tainton 1992); 2, Sover (Tainton 1992); 3-4, Finsch
(Scott Smith pers. comm.); 5, Roberts Victor (this work).
184 CHAPTER 2

A B c
30 30 30

N 20 20 20

10 10 10

87 91 95 87 89 91 93 95 87 89 91 93 95
- Mg / ( Mg + Fe~+) .... -Mg/(Mg +Fe~+)­ - Mg/(Mg + Fe~+)-
PHENOCRYSTS
.-,1 .-, .-,
F
1 I
0 E
I 1 1
1_, 1 L., 1 L.,
I 1 I
I I I
30 I 30 I 30 1
I I I
I 1 1
I I I
I 1
I I
N 20 20
r-~
I 20 r-~
I

I I
I
1
I
10 10 I 10
I
r'"
I '- ...
I
I
.... ...0
r---- J

87 91 95 87 91 95 87 91 95
- Mg/(Mg+Fe~+)- - Mg/(Mg +Fe~+)- - Mg/(Mg+Fe~+) ....
XENOCRYSTS / MACROCRYSTS

Figure 2.74. Histograms of olivine compositions. (A, B) Phenocrysts in diverse orangeites (data from Skinner
1989. Tainton 1992, respectively). (C) Phenocrysts in the Finsch orangeite (Fraser (1987). (0, E) Xenocrysts
and macrocrysts in diverse orangeites (data from Skinner 1989, Tainton 1992, respectively). (F) Macrocrysts
in the Finsch orangeite (Fraser 1987). Dashed histogram in D-Frepresents compositions ofolivines in peridotite
xenoliths (Clement 1982).

the limited data presented by Tainton (1992) and Moore (1988) for the Postmasburg,
Pniel. Sover, Bellsbank, and Newlands orangeites and by Fraser (1987) for olivines in
the Finsch occurrence (Figure 2.74).
Skinner (1989) has noted that overgrowths on xenocrysts or phenocrysts are of
similar composition in different orangeites. They are weakly-zoned from mg =0.91
adjacent to the core to mg =0.90 at the margin. Overgrowths are of similar character
regardless of the composition of the core. Thus, reverse or normal compositional mantling
is present (Figure 2.75).
Nickel is the only minor element present in the olivines in significant quantities
(approx. 0.25~.5 wt % NiO; Table 2.25). There are typically no differences evident in
the Ni contents of the cores of phenocrysts and xenocrystslmacrocrysts. Considerable
intergrain compositional variation is present, although individual crystals are typically of
uniform composition.
MINERALOGY OF ORANGEITES 185

BLAAUWBOSCH ROBERTS VICTOR

( Phenocryst ) ( Phenocryst)
91
94
89
Fo
87 ~,"","""~
90
CORE EDGE CORE EDGE

0 0'48 0'48 NiO


0
U
d5 0'36 0'36

0 0'24 0'24
Z
~
0
0'12 0'12 CoO
..: 0'00 o---J\ 0'00 ~~~~~~~
~

300 200 100 o }.1m 300 200 100 o}.lm

Figure 2.7S. Compositional zoning trends in orangeite olivines (after Skinner 1989).

Skinner (1989) demonstrated that significant systematic variations in NiO content


are present in overgrowths. Figure 2.75 shows that NiO contents, which may be relatively
higher than those of the cores, rapidly decrease with decreasing mg toward the overgrowth
margin. Clearly, the crystallization of overgrowth olivines, together with phenocrysts of
similar composition, rapidly depletes the parental magma in Ni. Moore (1988) has noted
that the rims of microphenocrystal olivines at Bellsbank and Newlands are depleted in
NiO relative to the cores. Calcium contents (Table 2.25) range from not detectable by
electron microprobe to as high as 0.l3 wt % CaO at the margins of phenocrysts, regardless
of the presence or absence of overgrowths. The latter may be slightly enriched in CaO
relative to the cores, but distinct core-mantle compositional discontinuities are not
present (Figure 2.75). MnO contents range from approximately 0.01 to 0.2 wt %. The
margins of crystals may be slightly enriched in Mn relative to the cores.

2.3.3. Comparisons with Olivines in Kimberlites


The paragenesis and composition of olivine in archetypal kimberlites has been
reviewed in detail by Mitchell (1986). Olivine is the most common and most characteristic
mineral of kimberlite and occurs as rounded, mantle-derived macrocrysts and euhedral-to-
subhedral microphenocrysts. The macrocrysts are of two varieties, xenocrysts derived
from disaggregated lherzolites and harzburgites, and rare relatively Fe-rich olivines of
cryptogenic origin which are members of the discrete nodule suite.
186 CHAPTER 2

Olivine macrocrysts/xenocrysts in kimberlites cover a wide range in composition


from approximately F084 to F095. Individual crystals are homogeneous and macrocrysts
of very different composition may occur adjacent to each other. Regardless of composi-
tion, macrocrysts may possess thin mantles of olivine compositionally similar to euhedral
groundmass olivine (F087-89). Macrocryst compositions are similar to those of olivines
in mantle-derived lherzolites and harzburgites. Olivines of this paragenesis are identical
in composition and origin to the macrocrystal olivine suite of orangeites.
Microphenocrystal (or groundmass) olivines in kimberlites on a worldwide basis also
exhibit a wide range in composition (F087-93). The margins of crystals commonly have
narrow rims of olivine which may be Fe-rich or poor, relative to the core. Regardless of
whether the zoning is normal or reversed, mantle compositions converge upon F087-89.
The range in core compositions and the uniform composition of the mantles has been
interpreted by MitcheU (1986) to imply that the groundmass olivine population is the
result of mixing of several batches of magma containing phenocrysts (and macrocrysts)
of slightly different composition. The normal and reversed zoned mantles found upon the
microphenocrysts (and macrocrysts) reflect attempts by the crystals to equilibrate with
the final hybrid magma precipitating olivines in the compositional range F087-90.

KIMBER LITES
40
ORANGEITES

30

20

10

87 91 95
Mg / ( Mg + Fe~+) •
Figure 2.76. Histograms comparing the compositions of phenocrystaJ oli vines in kimberlites (Shee 1986,
Clement 1982) and orangeites (Skinner 1989).
MINERAWGY OF ORANGEITES 187

Figure 2.76 shows that microphenocrystal olivines in South African kimberlites


(Clement 1982, Shee 1985) have a similar range in composition to phenocrysts in
orangeites (Skinner 1989). A significant difference is that the mode for orangeites
(F091-93) lies at more magnesian compositions than that of the majority of kimberlite
olivines (F088-90). Unfortunately, the data base is limited and the modes are not statisti-
cally significant. Nevertheless, Skinner (1989) has interpreted the data to suggest that
orangeites crystallize microphenocrystal olivines that are slightly more magnesian than
those of kimberlites. No significant differences are evident between Ni, Ca, and Mn
contents of orangeite and kimberlite phenocrysts.
The group of relatively Fe-rich (F078-88) olivine macrocrysts found in some kimber-
lites, e.g., Monastery (Gurney et al. 1979), Letseng-la-terae (Dawson et al. 1981), has
not yet been recognized in orangeites or in many other archetypal kimberlites. Many of
these macrocrysts contain inclusions of Cr-poor, Ti-pyrope or magnesian ilmenite and
are clearly a part of the megacryst or discrete nodule suite. The absence of such Fe-rich
olivines in orangeites is not unexpected given that the minerals of the megacryst suite are
not characteristic of orangeites.
In summary, orangeites and kimberlites are essentially identical with respect to the
paragenesis and composition of olivine. Orangeites share with kimberlites the charac-
teristic that, despite the ubiquity of olivine, it exhibits extremely limited compositional
variation. Primary olivines richer in iron than Foss have never been formed. The limited
range in composition is in marked contrast to that of olivine in many other ultrabasic
mantle-derived rocks, e.g., melilitites, alnoites, which display much greater ranges in
olivine composition during their crystallization. Hence, olivine compositions cannot be
used to assess the degree of evolution of different batches of orangeite magma in
composite dike systems or multiple intrusions.
Juxtaposition of macrocrystal and phenocrystal olivines of different core composi-
tions, which have attempted to equilibrate with the magma crystallizing compositionally
distinct groundmass microphenocrystal olivines, demonstrates that the olivine assem-
blage in both orangeites and kimberlites has multiple sources. This hybrid assemblage is
formed by the mixing of magmas containing compositionally distinct suites of macro-
crysts and phenocrysts. The apparently slightly more magnesian composition of mi-
crophenocrystal olivines in orangeites, relative to those in kimberlites, may imply that
their parental magmas were slightly more magnesian than kimberlite-forming magmas.
This may reflect source characteristics and/or the degree of assimilation of mantle-derived
xenocrystal material.
Recognizing that orangeites are hybrid rocks and the majority of the olivine macro-
crysts are xenocrysts has important consequences with regard to geochemical studies of
orangeites. Clearly, whole rock abundances of major elements and compatible trace
elements (Ni, Co, Sc) hosted by olivine will not reflect those of the parental magma (see
3.3.2).

2.3.4. Comparisons with Olivines in Lamproites


Although olivine is a common constituent of lamproites, it is neither ubiquitous nor
a characteristic mineral. Olivine contents are highly variable and reach maximum modal
188 CHAPTER 2

amounts (30-40 vol %) only in olivine lamproites. Two texturaly distinct varieties of
olivine are found in most olivine-bearing lamproites: anhedral macrocrysts or xenocrysts
and subhedral-to-euhedral phenocrysts. Olivines exhibiting "dog's tooth" habit are com-
mon in lamproites. The compositions of the xenocrystal olivines are identical to similar
olivines in both orangeites and kimberlite and are considered to be derived by the
disaggregation of similar mantle-derived ultramafic xenoliths. Phenocrysts in lamproites
have similar compositions to those of orangeites, consequently, distinction between these
rocks cannot be made on the basis of olivine compositions.
The textures of unevolved macrocrystal orangeites (Figures 1.26-1.44) are signifi-
cantly different from those of olivine lamproites (see illustrations of the latter in Mitchell
and Bergman 1991 and Jaques et at. 1986). Evolved orangeites, in contrast, have some
similarities to olivine lamproites in containing dog's tooth habit olivine and macrocrysts
mantled by phlogopite.

2.4. SPINEL
2.4.1. Paragenesis
Primary spinels occur as discrete euhedral-to-subhedral crystals. The majority of the
crystals are small, from 0.01 to 0.02 mm. Rocks containing very small «0.00 1-0.Q1 mm)
spinels are common, e.g., Swartruggens, New Elands. Typically, such spinels are anhedral
and highly corroded. Large primary spinel crystals are very rare.
Although spinels are Ubiquitous, their modal abundances are low. Few accurately
determined modes are available. Fraser (1987) determined that different facies of the
Finsch orangeites contain from 1.9 to 9.0 vol % opaques (presumably predominantly
spinel). Skinner and Clement (1979) give abundances from 1 to 6 vol % for examples
from Sydney-on-Vaal, Swartruggens, and Star-Burnes orangeites. Many orangeites, e.g.,
New Elands, Star, are estimated to contain <1-3 vol % spinel. In such rocks K-Ba
titanates (2.5) are common and appear to have replaced spinel as the principal late-stage
primary Ti-bearing mineral.
Commonly, spinels are single-phase grains or composite crystals consisting of
discrete cores and mantles. Cores may be deep-red and transparent or opaque. Mantles
are characteristically opaque. Single-phase spinels typically occur as euhedral-to-subhe-
dral crystals poikilitically enclosed by olivine, phlogopite, and tetraferriphlogopite and,
rarely, diopside. Spinels enclosed by primary olivines are commonly euhedral, transpar-
ent, red crystals. Spinels classified as late-stage ground mass phases are commonly
enclosed by late-stage tetraferriphlogopite, amphibole, sanidine, and carbonates.
Atoll-textured spinels (Figures 1.76, 1.79) are extremely rare. Tainton (1992) noted
that rare, relatively large (>0.01 mm) euhedral spinels exhibit an atoll texture in samples
from Sover-Doornkloof. However, such textures are not typical of the spinel population
in these rocks.
Large (up to 0.1 mm) macrocrystal spinels are rarely encountered in thin sections,
although they are common in heavy mineral concentrates. They may be mantled by thin
rinds of opaque spinels which are compositionally similar to primary groundmass spinels.
MINERALOGY OF ORANGEITES 189

2.4.2. Composition
Few detailed investigations of the compositional variation exhibited by orangeite
spinels have been undertaken. Significant compositional data have been provided only
by Boctor and Boyd (1982), Mitchell and Meyer (1989a), and Tainton (1992). The
discussion below is based on these data and new data for Lace, Bellsbank, Besterskraal,
and Sover-Doomkloof obtained during the preparation of this work.
The compositions of orangeite spinels fall within the eight-component system
MgCr204 (magnesiochromite)-FeCr204 (chromite)-MgAh04 (spinel)-FeAh04 (her-
cynite)-Mg2Ti04 (magnesian ulvospinel or qandilite)-Fe2Ti04 (ulvospinel)-MgF~04
(magnesioferrite)-Fe304. Manganese and zinc contents are typically low, and the
MD2Ti04 (manganese ulvospinel), MnFe:!04 Gacobsite), MnCr204 (manganoan chro-
mite), ZnAh04 (gahnite), and ZnCr204 (zincian chromite) end members are usually
unimportant. Vanadium contents have not been determined in this or previous studies but
are expected to be low given the reasonable weight percent oxide totals reported.
Spinel compositional variations are conventionally represented by plotting compo-
sitions in the "oxidized" and "reduced" spinel prisms (Irvine 1965, Haggerty 1976,
Mitchell 1986). In the "reduced" spinel prism, total iron is calculated as FeO (Fe?+ or
FeOT), whereas in the "oxidized" prism Fe3+ and Fe2+ are estimated from stoichiometry

Thble2.26. Representative Compositions of Spinels from the Bellsbank and Lace OrangeitesQ
Wt% 2 3 4 5 6 7 8 9 10
Ti02 1.78 4.01 9.30 7.87 6.91 2.29 3.12 4.41 6.01 8.82
AI 20 3 0.50 0.98 0.14 0.12 0.17 3.31 3.00 1.59 1.03 n.d.
Cr203 57.21 54.61 1.16 0.77 0.30 54.30 55.07 35.48 21.43 6.30
FeOr 26.58 27.98 76.57 79.18 80.86 29.16 26.27 50.14 60.63 73.83
MnO 0.58 0.50 1.45 0.85 0.86 0.61 0.88 0.81 0.95 0.63
MgO 10.82 10.32 5.46 5.54 4.46 9.34 10.64 6.08 6.58 5.68
-- --
97.47 98.40 94.08 94.33 93.56 99.01 98.98 98.51 99.63 95.26

Recalculated compositions
Fe203 11.75 9.75 52.16 55.80 57.04 10.59 9.16 26.63 38.27 48.72
FeO 16.01 19.20 29.64 28.97 29.53 19.63 18.03 26.17 26.20 30.00
98.65 99.38 99.31 99.92 99.27 100.07 99.92 101.l6 100.46 100.16

Mol % end-member molecules


MgAI 20 4 0.9 1.8 0.2 0.2 0.3 6.1 5.5 2.7 1.6
Mg21104 6.4 14.1 14.9 15.1 12.2 8.1 10.9 14.2 18.1 15.9
Mo21104 2.3 1.3 1.4 1.0
Fe2Ti04 8.7 5.4 5.8 8.0
MnCr204 1.6 1.3 1.6 2.3 2.0 2.2
MgCr20 4 41.7 27.3 26.6 29.1 4.2 0.4
FeCr204 28.4 38.5 l.l 0.8 0.3 38.9 36.1 33.9 20.1 6.2
Fe304 21.0 17.1 72.7 77.3 80.1 18.7 16.0 43.0 57.7 68.8
"FeOr = total Fe calculated as FeO; n.d. = not detected. 1-5. BeIlsbank; 6-10. Lace. (All data this work except 3 from Boctor
and Boyd 1982.)
190 CHAPTER 2

(Droop 1987) or by the recalculation procedure of Carmichael (1967b). Mitchell (1986)


noted that the reduced prism is particularly useful for illustrating spinel compositions in
that all of the major elements present in kimberlite (and orangeite) spinels are included
in the prism. A disadvantage of this projection is that the prism fails to illustrate variations
in Fe304 or MgFe204 end members which may be important in evolved spinels. Oxidized
prisms have the disadvantage of failing to illustrate the presence of inverse spinels, as Ti
is not included in the projection. Consequently, neither prism is entirely satisfactory for
illustrating the complete compositional variation of spinels, which may range from
magnesian ulvospinel-rnagnesiochromite~hromite solid solutions to ulvospinel-mag-
netite. Commonly, the reduced spinel prism, together with projections of the data onto
the front and bottom prism faces, provides the best illustrations of compositional vari-
ation.
Representative compositions of primary spinels are given in Tables 2.26 and 2.27.
These fall into two broad compositional groups:
Group A. Titanian (1-5 wt % Ti02) magnesiochrornite~hromite solid solutions
characterized by very high Cr/(Cr + AI) ratios (typically >0.9) and moderate

Table 2.27. Representative Compositions of Spinels from the Besterskraal, Sover North, and
Pniel Orangeitesa
Wt% 2 3 4 5 6 7 8 9 10
Ti02 1.03 0.95 6.11 1.34 9.63 10.81 2.20 13.39 1.31 2.16
AI 20 3 3.37 2.32 0.31 2.38 0.44 n.d. 2.89 n.d. 4.07 3.26
Cr203 64.39 61.74 32.57 59.42 14.59 11.16 56.46 10.59 57.09 51.58
FeOr 19.62 29.79 55.45 31.09 67.24 71.16 31.65 65.52 27.38 28.78
MnO 0.67 1.17 1.16 1.20 2.67 0.88 1.14 0.81 n.d. n.d.
MgD 10.41 4.03 2.39 4.07 2.44 3.01 4.10 4.53 6.81 10.03
99.49 100.00 95.60 99.50 97.01 97.02 98.44 94.84 %.66 95.81

Recalculated compositions
Fe203 3.09 4.17 25.41 5.47 36.88 39.05 5.51 33.58 5.41 11.92
FeD 16.84 26.04 32.58 26.17 34.06 36.02 26.70 35.30 22.51 18.05
-- -- -- -- -- --
99.81 100.42 100.54 100.06 100.70 100.92 98.99 98.20 97.20 97.00

Mol % end-member compositions


MgAI 20 4 6.6 4.7 0.5 4.8 0.7 5.8 8.1 6.1
Mg2Ti04 3.9 3.7 7.4 5.1 6.7 8.7 8.4 13.2 5.0 7.7
Mn2Ti04 2.2 4.5 1.5 1.4
Fe2Ti04 10.5 17.6 21.4 24.9
MnCr204 1.9 3.4 3.5 3.3
MgCr204 39.4 11.0 9.0 3.7 19.6 31.0
FeCr204 42.6 69.2 37.5 67.2 15.3 11.4 68.4 11.0 56.9 33.7
Fe304 5.7 8.1 41.8 10.5 55.2 57.0 10.5 49.6 10.4 21.4
"Fe(h =total Fe calculated as FeO. n.d. =not detected. 1--Q, Besterskraal (this work); I, euhedral spinel in olivine phenocryst; .
2-4,3-5, cores and mantles respectively; 6, euhedral groundmass spinel; 7-8, Sover North (TaintoD 1992); 9-10, Pniel
(Tainton 1992).
MINERAWGY OF ORANGEITES 191

Figure 2.77. Representative compositions of spinels from the Bellsbank, Lace (this work) and Bums (Mitchell
and Meyer 1989a) orangeites chosen to illustrate the compositional trends of orangeite spinels in the reduced
spinel prism.

Fei+/(FeT2+ + Mg) ratios (0.4-0.7). Ah03 contents are low, typically <5 wt %
and commonly <2 wt %.
Group B. Alumina-poor «1 wt % Ab03) titanian (5-10 wt % Ti02) magnesian «
10 wt % MgO) magnetites.
Red and/or opaque group A spinels typically form single-phase crystals. Opaque
group B spinels occur as discrete single crystals and as mantles upon group A spinels.
Representative spinels from Bellsbank, Lace, and Burns are plotted in the reduced
and oxidized spinel prisms, Figures 2.77 and 2.78 respectively, to illustrate the nature of
evolutionary trends of spinel compositions in orangeites.
Group A spinels plot near the base of both prisms and evolve along the prism axis
toward increasing Fer/(FeT + Mg) or Fe2+/(Fe2+ + Mg) ratios at approximately constant
Cr/(Cr + AI) ratios and Ti contents. Group B spinels plot within the prism and near to the
front rectangular face as a consequence of their high Cr/(Cr + AI) ratios, even though Cr
contents are low «10 wt % Cr203). They evolve at approximately constant
Fei+/(Fei+ + Mg) ratios (0.8-0.9) or Fe2+/(Fe2++ Mg) ratios (0.7-0.8) toward the apices
of the prisms. The trend reflects primarily increases in magnetite, magnesian-ulvospinel,
aM ulvospinel contents.
192 CHAPTER 2

........ CI'/,
O·A

0·E>
rCI'"
"4/) 0. 5
..........

Figure 2.78. Representative compositions of spinels from the Bellsbank, Lace (this work) and Bums (Mitchell
and Meyer 1989a) orangeites chosen to illustrate the compositional trends of orangeite spinels in the oxidized
spinel prism.

These data demonstrate that orangeite spinels evolve from titanian magnesio-
chromite--chromite solid solutions toward magnesian ulvospinel-ulvospinel-magnetite
solid solutions (Tables 2.26 and 2.27). This trend involves increases in Fe2+, Fe3+, and Ti
with concomitant decreases in Cr and AI.
The entire compositional trend is not present in all orangeites. Zoning trends within
individual crystals conform to the overall trend. The presence of discrete cores and
mantles implies that spinel precipitation may not be continuous during post-intrusional
crystallization. The compositional hiatus may result from a miscibility gap or be due to
bulk compositional effects.
Figures 2.79 and 2.80 illustrate the compositions of spinels from orangeites, charac-
terized by the phlogopite-tetraferriphlogopite mica composition trend (2.1.7), projected
onto the bottom and front square faces of the reduced spinel prism. These data clearly
show the differing compositions of group A (cores) and group B (mantle) spinels. Only
minor inter-intrusion differences are evident. Thus, group B spinels from Lace do not
exhibit the same degree of Ti and Fe3+ enrichment as spinels from Burns or Bellsbank.
(This may be an artifact due to a paucity of data.) Group B spinels from Newlands may
have slightly lower Fe/(Fe + Mg) ratios than other orangeite spinels.
Figures 2.81 and 2.82 illustrate the compositions of spinels from the Sover Mine
orangeite and amphibole-bearing orangeites exhibiting the phlogopite-biotite mica com-
r::::
LACE
Z
r,.J
• • LACE
~
0-1 -i + BELLSBANK 0-9 + BELLSBANK ;::~ >
0 BURNS o BURNS CA.6tg
• NEW ELANDS • NEW ELANDS
S
~
0-2 -i.6 NEWLANDS O-S +0 0<
.6 NEWLANDS 0
.6 ""l
0-3 -i \
1 0-7 0:> 0
.6 ~ 0
GROUP B !::z
0
SPINELS +~ ~
~ 0-4 0-6 r,.J
.6 U GROUP B .a.• ~
+0 r,.J
+
~ 0-5 +
__ 0-5 SPINELS o 0 I -
til
U l- •
i:> •
...... 0-6
~
0 ...... 0-4 0
u +
i=
.6
•• 0-3
a. GROUP A
o-a7 MACROCRYSTS co,,,,!::", c- ~o 0-2 J SPINELS
(V.. -
I ~ -
0-9 A .............. t"•• --ibI 0-1
r
0-2 0-3 0-4 0-5 o-a 0-9 0-5 0-6 0-7 o-s 0-9

Fe T 2 + + Mg)
2+ I( Fe T • Fe~+/(Fe~++ Mg} •
Figure 2.79. Compositions of spinels from diverse orangeites exhibiting the Figure 2.80. Compositions of spinels from diverse orangeites exhibiting
tetraferriphlogopite mica compositional trend plotted on the base of the reduced the tetraferriphlogopite mica compositional trend plotted on the front face
spinel prism_ All spinels are considered to be primary with the exception of those of the reduced spinel prism_ Data sources as in Figure 2_79_
from New Elands, which are macrocrysts_ Data from Boctor and Boyd (1982),
Mitchell and Meyer (I 989a), Tainton (1992), and this work_ AMC = composi-
tional trend of aluminous magnesian chromites from kimberlites (Mitchell 1986)_ ~
-
...~
COMPOSITIONAL
•SOVER FIELD OF SPINELS • SOVER
0·1 + SOVER NORTH IN ORANGEITES 0.9l + SOVER NORTH
o BESTERSKRAAL EXHIBITING THE o BESTERSKRAAL I .,
0·2 • PNEIL TETRAFERRIPHLOGOPITE 0.8 • PNEIL
EVOLUTIONARY TREND
0·3 \ 1 COMPOSITIONAL •
FIELD OF SPINELS
« 0·4 IN ORANGEITES
~ o. 07 EXHIBITING THE
TETRAFERRIPHLOGOPITE +
+
U 0·5 + 0·5 EVOLUTIONARY TREND

--:: 0·6
i=
...... 0-4
1
U "'cP
i=
0·3

O·S
• 0·2

0·9
t1 J Po 0·1
r' . cP-"'8o 0
,~'\....~.•+~~.
-:«:. 0
~l

0

0·2 0·3 004 0·5 0·6 0·7 O·S 0·9 0·3 0·4 0·5 0·6 0·7 O·S 0·9

Fe 2 + I(Fe 2 +
T T + Mg) • Fe~+/(Fe~++ Mg) •
("'l
Figure 2.S1. Compositions of spinels from diverse evolved orangeites Figure 2.82. Compositions of spinels from diverse evolved orangeites
(Tainton 1992, this work) plotted on the base of the reduced spinel prism. (Tainton 1992, this work) plotted on the front face of the reduced spinel
AMC = compositional trend of aluminous magnesian chromites from prism.
kimberlites (Mitchell 1986).
~....
MINERALOGY OF ORANGElTES 195

positional trend (2.1.7) from Sover North, Besterskraal, and Pniel. Mantled spinels in the
Sover Mine intrusion are identical to other orangeite spinels. Discrete spinels from Sover
North correspond mainly to group A spinels. Group B spinels are rare (two analyses;
Tainton 1992) and have high Ti02 contents (12.2-13.4 wt %). Their compositions fall
within the Sover Mine group B trend.
Besterskraal spinels occur as inclusions in primary olivine and as two-phase crystals.
Spinels included in olivine are Fe poor [Fer2+/(Fer2+ + Mg) < 0.5] relative to the cores
of discrete, mantled spinels [Fer2+/(Fer2+ + Mg) =0.55-0.75]. Mantles (group B) do not
show extreme enrichment in Ti or Fe3+, although they may have slightly higher
Fer2+/(Fer2++ Mg) ratios (>0.9) relative to other orangeite spinels. Only low-Ti group A
spinels are present at Pniel.
Figure 2.82 suggests that group A spinels from amphibole-bearing evolved
orangeites have slightly lower Ti/(Ti + Cr + AI) ratios than other orangeite group A spinels
and that magnetite-rich group B spinels are not present. Further detailed investigations
of these spinel populations are required.
The composition of macrocrystal spinels has been insufficiently documented.
Mitchell and Meyer (1989a) have shown that spinel macrocrysts from New Elands are
Ti02-poor «0.5 wt %) aluminous magnesian chromites exhibiting a wide range in their
Cr203 (43.7-62.8 wt %) and Ah03 (6.9-22.7 wt %) contents. Those with high Cr/(Cr + AI)
ratios form a continuation of the orangeite group A spinel compositional trend (Figure
2.79).
Mitchell (1986) has noted that Ti-, Fe3+-poor spinels from kimberlites, mantle-
derived lherzolites and harzburgites, and mid-ocean ridge basalts, all plot on the base of
the reduced spinel prism and have very similar compositions. These spinels have no
compositional characteristics permitting their assignment to a particular paragenesis. The
New Elands spinel macrocrysts have similar compositions to these Ti-poor spinels;
consequently their origins cannot be unambiguously determined.

2.4.3. Comparisons with Kimberlite Spinels


The compositional trends of spinels in kimberlites are well documented. Previous
studies did not distinguish between archetypal kimberlites and orangeites, although it was
recognized that mica-rich kimberlites did contain spinels different in composition from
those in monticellite kimberlites. Mitchell (1986) has shown that three main groups of
spinels, each defining a distinct compositional trend in spinel prisms (Figure 2.83) are
present:
1. The macrocrystal or aluminous magnesian chromite trend
2. Magmatic trend 1 or the magnesian ulvospinel trend
3. Magmatic trend 2 or the titanomagnetite trend
A fourth group of AI-rich ground mass spinels forming the pleonaste reaction trend
is known only from a few kimberlites. These spinels have a multiplicity of origins
(Mitchell 1986) and are not considered further in this work.
Spinels belonging to the AMC trend are Ti02 poor «2 wt %) and compositions plot
on the base of spinel prisms (Figure 2.83). They range in composition from magnesian
1% CHAPTER 2

Figure 2.83. Representative compositions of spinels from kimberlites plotted in the reduced spinel prism. See
text for an explanation of trends I and 2.

aluminous chromite (MAC) to aluminous magnesian chromite (AMC). Individual crys-


tals are typically homogeneous, consequently it is difficult to establish compositional
trends. However, Ti-bearing magnesian chromites, similar in composition to the more
Cr-rich macrocrysts, mantle the latter spinels and occur as primary groundmass phases.
This observation suggests that some AMC macrocrysts may be cognate.
MAC-AMC spinels are similar in composition to spinels occurring in a wide variety
of ultrabasic and basic rocks. They do not exhibit any compositional or textural features
that permit conclusive recognition of their origin. Thus, Mitchell (1986) suggests that
Cr-rich members of the AMC trend are cognate and that AI-rich members are xenocrysts.
In contrast, Shee (1984) regards all of the macrocrysts as xenocrysts.
Magmatic trend 1 is now recognized as the characteristic spinel compositional trend
of archetypal kimberlites (Mitchell 1986, Mitchell and Bergman 1991). In the reduced
spinel prism (Figure 2.83a) the trend is across the prism from the base near the MgCr204-
FeCr204join [commonly Cr/(Cr+ AI) =0.80-0.95, Fei+/(FeT2+ + Mg) =0.4-0.6] toward
the rear rectangular face [Le., decreasing Cr/(Cr + AI) ratios] and upward toward the
Mg2Ti04-Fe2Ti04 apex. Spinel evolution is from titanian magnesian aluminous chromite
(TIMAC) or titanian magnesian chromite (TMC) containing 1-12 wt % Ti02 toward
members of the magnesian ulvospinel-ulvospinel-magnetite (MUM) series and is a
trend of increasing Ti, Fe3+/Fe2+, and total Fe, and decreasing Cr at approximately
constant F~2+/(F~2+ + Mg) ratios. Alumina may decrease, increase, or remain constant
MINERAWGY OFORANGElTES 197

over this trend. MnO contents are typically low «1 wt %), but may increase slightly in
the more evolved spinels. The trend culminates with the formation of Ti- and Mg-free
magnetite. The presence ofTi- and Mg-rich spinels (12-23 wt % Ti02, 12-20 wt % MgO)
containing substantial proportions of the magnesian ulvospinel molecule (20-40 mol.%
Mg2Ti04) is the hallmark of this compositional trend. Archetypal kimberlites are the only
igneous rocks yet known to contain spinels rich in magnesian ulvospinel. Representative
compositions can be found in Mitchell (1986), Shee (1984), and Pasteris (1983). The
entire magmatic trend 1 spinel assemblage is not found in all kimberlites and individual
occurrences may exhibit only a portion of the trend. Numerous examples of the trend are
described in detail by Mitchell (1986).
Magmatic trend 2 is characterized by spinels ranging in composition from AMC
through TMC and titanian chromite (TC) to members of the ulvospinel-magnetite
(USP-MT) series. In the reduced spinel prism (Figure 2.83) the trend is initially along
the axis ofthe prism toward increasing Fer2+/(Fei+ + Mg) ratios at relatively constant,
but low, Ti02 contents and high Cr/(Cr + AI) ratios (>0.85), followed by a rapid increase
in Ti at high Fer2+/(Fer2++ Mg) ratios (>0.8) toward the Fe2Ti04 apex. MnO enrichment
(> 1 wt %) occurs in the more evolved spinels. The trend is characterized by rapid MgO
depletion, and spinels rich in magnesian ulvospinel (>20 mol %) are not formed. All of
the spinels from trend 2 are poor in Al relative to those from trend 1.
Magmatic trend 2 is uncommon in archetypal kimberlites and insufficiently charac-
terized. It has been recognized only from the Tunraq (Mitchell 1979), Zagodochnaya
(Rozova et al. 1982), and Koidu (Tompkins and Haggerty 1984) kimberlites. The

1'0

<t
+ 0'6
...
U

+
i= 0'4

......
i=
0'2

Figure 2.84. Compositional trends of


spinels from kimberlites (Tl and T2;
Mitchell 1986. this work). orangeites (this 0'2 0'4 0'6 0'8 1'0
work). and lamproites (L; Mitchell and
Bergman 1991). - - FeT / 2+ (I:'
leT2+ + Mg )
--.
198 CHAPTER 2

common feature of these kimberlites is that they appear to have crystalized abundant
phlogopite prior to the formation of the bulk of the ground mass spinels. In the case of the
Tunraq kimberlite, spinels belonging to both magmatic trends are present in different
facies of the intrusion. However, only the facies rich in macrocrystal phlogopite contains
spinels belonging to trend 2.
Very rare spinels belonging to this trend have also been noted in the Elwin Bay
(Mitchell 1978a), De Beers Peripheral (Pasteris 1980), and Marushkaya (Rozova et al.
1982) kimberlites.
Figures 2.77 and 2.84 show that orangeite spinels exhibit compositional trends
identical to those found in kimberlite magmatic trend 2 (Figure 2.83). Spinels belonging
to magmatic trend 1 are not present in orangeites.
Spinels in orangeites are not characterized by the presence of significant quantities
(>20 mol %) of the magnesian ulvQspinel molecule (Tables 2.26 and 2.27). Spinels from
orangeites containing 10-20 mol % Mg2Ti04 are Cr-rich members of the Mg2Ti04-
FeCr204-Fe304 series rather than the Cr-poor Mg2Ti04-Fe2Ti04-Fe304 series charac-
teristic of kimberlites.

2.4.4. Spinel Compositional Variation in Lamproites and Lamprophyres


Spinels of high Crf(Cr + AI) ratios (>0.9) very similar to trend 2 spinels, are also
found in lamproites (Mitchell and Bergman 1991). In these rocks spinels evolve from

• PRAIRIE CREE K
o ELLENDALE

o.'~

O·A
.........
C"/~ 0.6
C,.
'1-4/) 0. 8
..........

Figore 2.8S. Representative compositions of spinels from lamproiles plotted in the reduced spinel prism (after
Mitchell and Bergman 1991).
MINERALOGY OF ORANGElTES 199

Ti-poor aluminous magnesiochromites, through titanian aluminous magnesiochromites,


to AI-poor titanian magnesian chromite, and magnesian titaniferous magnetite. The
evolutionary trend is initially along the axis of the spinel prism (Figures 2.85) with
increasing Fer2+/(Fer2+ + Mg) ratios. Subsequently, the spinels evolve with increasing
Ti/(Ti + Cr + AI) at constant Fer2+/(Fer2++ Mg).
Mitchell and Bergman (1991) have noted that kimberlite spinel trend 2 and the
lamproite spinel trend are Cr-rich variants of a common spinel trend found in a wide
variety of rocks that include basalt, alnoites, melilitoids, minettes, and other lampro-

• ULTRAMAFIC LAMPROPHYRES + MINETTES


o KIMBERLITE TREND 2 • ORANGEITES
1'0

0·9

0·8 o


- - - - - - --j

0·7
o
0'6
+
~
() 0·5
+
-
~

"'-
t-
0'4

0'3

0'2

O· I

2 4 5 6 7 8 9 10
Cr I( Cr + AI)
Figure 2.86. Ti/(Ti + Cr + AI) versus Cr/(Cr + AI) for spinels from kimberlites (Mitchell 1986, Shee 1984.
Pasteris 1980),lamproites (Mitchell and Bergman 1991), orangeites (this work, Tainton 1992, Boctor and Boyd
1982), basalts (Ridley 1977), ultramafic lamprophyres (this work) from Oka, Pol zen, Como, Alno, and Haystack
Butte (HB), and minettes (Jones and Smith 1985).
200 CHAPfER2

phyres. The trends are similar in that their evolution is from Ti-poor, Cr-rich spinels
toward ulvospinel-magnetite solid solutions. Compositional trends originate at the base
of the prisms and evolve along the axes toward the ulvospinel or magnetite apices. The
trends differ principally with respect to the Cr/(Cr + AI) ratios of the least-evolved
members of the series (Mitchell 1986). Cr-rich spinels in alnoites and polzenites with
Cr/(Cr + AI) ratios <0.85 are mantled by aluminous Cr-poor or Cr-free spinels (Figure
2.86). Spinels from different petrological provinces differ with respect to the composition
of the least-evolved spinel. All follow an evolutionary trend ofCrdepletion coupled with
Ti and Fe enrichment.
Spinel compositions, plotted in Ti/(Ti + Cr + AI) versus Cr/(Cr + AI) diagrams,
clearly demonstrate these differences (Figure 2.86). Thus, spinel compositions with
Cr/(Cr + AI) > 0.85 are considered to be indicative of orangeites, lamproites, and some
varieties of kimberlite. Spinels with Cr/(Cr + AI) < 0.85 are characteristic of basaltoids,
melilitoids (including alnoites and polzenites), and minettes. Detailed studies of the
compositional variation shown by spinels in minettes have not yet been undertaken. The
majority of minettes examined to date (Jones and Smith 1985, Bergman 1987, Peterson
1994, Lange and Carmichael 1991 , Allan and Carmichael 1984) contain ulvospinel-mag-
netite solid solutions as discrete crystals and as mantles upon earlier-formed MAC-AMC
solid solutions.
Mitchell (1986) and Mitchell and Bergman (1991) have previously noted that
kimberlite magmatic trend 2 is not diagnostic of a kimberlitic or lamproitic paragenesis.
As orangeite spinels are of similar composition, it must be concluded that Cr-rich
members [Cr/(Cr + AI) > 0.85] of spinel trends similar to kimberlite magmatic trend 2
are indicative of (1) kimberlites containing phlogopite macrocrysts, (2) lamproites, and
(3) orangeites. Exact classification of rocks containing such spinels can only be made in
conjunction with other mineralogical and geochemical criteria.

2.S. POTASSIUM BARIUM TITANATES

2.S.1. Hollandite
Compounds belonging to the hollandite group have the general formula AI-2BI-2Ti6-7-
016, where A =K, Ba, Rb, Cs, Sr; B =Fe2+, Fe3+, AI, V, Ce, Ga, Sc, In, Ru; and C =Ti,
Nb, Ge, Zr, Sn. They consist of paired chains of edge-sharing (B,C)06 octahedra
extending along the crystallographic c-axis. The chains are linked by comer- and
edge-sharing to form a framework containing continuous tunnels of approximately square
cross section aligned along the c-axis. Each tunnel is bounded by four chains of paired
octahedra. The tunnels are filled by A-site cations to varying degrees; thus hollandites
typically display non-integral stoichiometries. Cation ordering within and between the
tunnels results in incommensurate superlattice ordering (Bursill and Grzinic 1980, Pring
and Jefferson 1983, Kesson and White 1986, Zandbergen et al. 1987).

2.5.1.1. Paragenesis
Hollandite group minerals are common, but not ubiquitous, late-stage groundmass
phases in some orangeites. To date they have been recognized in the Star, Lace,
MINERALOGY OF ORANGElTES 201

Besterskraal, Sover North, and New Elands orangeites (Mitchell and Haggerty 1986,
Mitchell and Meyer 1989b; this work). Hollandite does not occur in all samples from a
given intrusion and typically occurs in rocks containing the most evolved micas. The
factors controlling the presence of hollandite in some rocks but not others have not been
established. The apparent absence in many orangeites may simply be a consequence of
lack of investigation.
Hollandites typically occur as stellate clusters of subhedral prismatic crystals (Figure
2.87) enclosed within groundmass calcite segregations. Silicate-rich portions of the
groundmass typically lack hollandite. Hollandites are opaque in thin section, with reddish
internal reflections being observable only at the edges of very thin crystals. In reflected
light the mineral is gray and exhibits medium reflection pleochroism and medium-to-
strong anisotropy in tones of light-to-dark gray. Estimated white light reflectivities in air
are approximately 12-15% and in oil immersion 5% (Mitchell and Haggerty 1986). BSE
imagery shows that zoning may be weakly- or strongly-developed (Mitchell and Meyer
1989b).

2.5.1.2. Composition
Minerals belonging to the hollandite structural group form a complex series of solid
solutions, the most important of which are between the K2(Mg,Fe2+)TbOI6-
Ba(Mg,Fe2+)Ti7016 series and the K2(M 3+hTi6016-Ba(M3+hTi6016 series where M3+ =
Fe, Cr, V, Ce, AI. The existing nomenclature of these phases is unsatisfactory, as only four
of the many possible end-member molecules have been defined as mineral species. These
are: BaFe3+2Ti6016, termed "barian priderite" by Zhuravleva et al. (1978); mannardite,
(Ba,H20)V2Ti6016 (Scott andPeatfield 1986, Szymanski 1986); redledgeite BaCr2Ti6016
(Gatehouse et al. 1986); and ankangite Bao.83V2.29Cro.osTis.83016 (Xiong et al. 1989).
Mannardite, redledgeite, and ankangite have been accepted by the International
Mineralogical Association as valid mineral names. Currently, the potassian varieties of
these minerals are referred to simply as potassium analogues, e.g., potassian redledgeite,
rather than being given distinct names (Mitchell and Meyer 1989b). The status of
ankangite as a distinct species is questionable. Wu et al. (1990) have shown this mineral
to possess an incommensurately modulated structure, and, although compositionally
similar to the V-hexatitanate, tenned mannardite, it is structurally similar to septetitanates
such as priderite (see below). Hence, ankangite and mannardite may both be members of
a Ba-V hexatitanate polysomatic series for which a multiplicity of names is undesirable.
Although priderite (Norrish 1951) is accepted as a valid name, it is not actually an
end-member composition. Most priderites are intermediate members of the
K2(Mg,Fe2+)ThOI6-Ba(Mg,Fe2+)ThOI6 series (Fe2+ » Mg) with some solid solution
toward BaFe3+2Ti6016. The latter hexatitanate is an end-member molecule for which a
new name is desirable as it is not a barian priderite. Ifpriderite is to be retained as a valid
mineral species then the tenn must be redefined as an end-member molecule. Mitchell
and Meyer (1989b) have suggested that the solid solutions based upon K-Ba septati-
tanates be tenned the priderite series, but did not redefine any of the end-member
molecules as priderite.
Insufficient is known of the role of H20 in hollandites, as most natural hollandites
are not analyzed for water. Scott and Peatfield (1986) and Szymanski (1986) consider
202 CHAPTER 2

Figure 2.87. Acicular hollandite in calcite and phlogopite matrix, New Elands. (A) Transmitted light, field of
view 0.25 mm. CB) reflected light with partially crossed polars, field of view 0.1 mm.
MINERAWGY OF ORANGElTES 203

that mannardite and redledgeite contain one H20 molecule per formula unit (Le., approx.
2.1 wt % H20). However, as reasonable analytical totals are typically obtained for most
electron microprobe analyses of hollandite in lamproites, it is assumed that the amounts
of water present in most hoIlandites are very small.
Unfortunately, it is not possible to assess the FeO and Fe203 contents ofhoIlandites
analyzed by electron microprobe. This is because the common nonstoichiometry of the
compounds precludes estimation ofFe2+ and Fe3+ from the structural formula by standard
methods. It should be clearly realized that natural hoIlandites contain both Fe2+ and Fe3+
and may even contain Ti3+ (Myhra et al. 1988). In this work total Fe is expressed as Fe203,
and all structural fomulae quoted are based upon 16 oxygens in order that hoIlandites
from diverse parageneses may be compared on the same basis. The actual oxygen content
and number of cation vacancies may vary substantially, depending upon composition
(Myhra et al. 1988, Kesson and White 1986).
Representative compositions of hoIlandites from Lace, Besterskraal, Sover North,
Star, and New Elands are given in Tables 2.28 and 2.29. Individual crystals may be
strongly-zoned with respect to their Ba, K, Nb, and V contents. Figure 2.88 demonstrates

Table 2.28. Representative Compositions of Hollandites from the Lace and Besterskraal
OrangeitesO
Wt% 2 3 4 5 6 7 8 9
Nb20 s 0.95 0.76 1.13 1.99 2.68 3.96 4.97 5.64 6.84
Ti02 69.17 63.78 61.72 63.93 61.98 61.93 62.39 64.59 63.08
Cr203 n.d. 0.29 0.56 n.d. n.d. n.d. n.d. n.d. n.d.
V20 3 2.89 4.92 5.59 n.d. n.d. n.d. n.d. n.d. n.d.
Fe203 6.70 7.22 7.02 13.40 13.40 12.89 12.34 9.87 10.60
MgO n.d. n.d. n.d. n.d. n.d. n.d. 0.52 n.d. n.d.
BaO 17.58 21.42 23.32 20.04 21.52 20.71 18.63 18.00 18.42
K20 2.11 1.62 1.66 0.63 0.21 0.51 0.67 1.91 0.72
99.40 100.00 100.00 99.99 99.99 100.00 99.52 100.00 99.66

Structural formulae based on 16 oxygens


Nb 0.055 0.046 0.068 0.118 0.162 0.237 0.295 0.322 0.405
Ti 6.688 6.356 6.222 6.316 6.218 6.178 6.159 6.322 6.209
Cr 0.030 0.059
V 0.295 0.518 0.596
Fe 0.648 0.720 0.708 1.325 1.345 1.287 1.219 0.967 1.044
Mg 0.102
Ba 0.886 1.112 1.173 1.032 1.125 1.077 0.958 0.918 0.945
K 0.346 0.272 0.284 0.106 0.036 0.086 0.112 0.317 0.120

Site occupancy
A 1.232 1.385 1.456 1.137 1.161 1.163 1.070 1.235 1.065
B 0.944 1.269 1.363 1.325 1.345 1.287 1.321 0.967 1.044
C 6.743 6.402 6.291 6.430 6.379 6.416 6.454 6.322 6.613
aTotal Fe expressed as Fe203. n.d. =not detectable. Compositions 1-3, Lace; 4-9, Besterskraal. All data this work.
204 CHAPTER 2

Table 2.29. Representative Compositions of Hollandites from the Sover North, Star, and New
Elands Orangeitesa
Wt% 2 3 4 5 6 7 8 9
Nb 20 s 0.86 1.15 2.37 4.57 1.87 0.87 n.a. n.a. n.a.
Ti0 2 66.88 67.25 71.20 71.30 67.10 66.90 78.15 75.41 75.73
Cr203 n.d. n.d. n.d. 0.71 0.41 0.58 n.d. n.d. 0.50
V20 3 1.23 1.16 1.24 2.79 5.11 9.85 2.52 3.61 4.51
Fe203 10.36 10.14 9.32 6.62 6.49 4.19 3.55 3.96 1.52
Ce203 n.d. n.d. n.d. n.d. n.d. n.d. 1.76 1.74 3.52
MgO 0.74 0.33 0.21 0.21 0.10 1.37 1.80 0.94
BaO 19.21 19.63 13.72 8.36 15.60 15.80 4.72 4.47 3.92
K20 0.25 0.34 2.16 4.81 1.66 2.66 8.30 8.33 8.69
99.53 100.00 99.81 98.37 98.45 100.95 100.37 99.32 99.33

Structural formulae based on 16 oxygens


Nb 0.050 0.067 0.133 0.253 0.109 0.050
Ti 6.513 6.541 6.664 6.559 6.480 6.332 6.976 6.821 6.897
Cr 0.069 0.042 0.058 0.048
V 0.127 0.119 0.123 0.271 0.521 0.985 0.238 0.345 0.434
Fe 1.010 0.987 0.873 0.609 0.627 0.397 0.317 0.358 0.139
Ce 0.076 0.077 0.156
Mg 0.143 0.064 0.038 0.040 0.019 0.242 0.323 0.166
Ba 0.975 0.995 0.669 0.401 0.785 0.779 0.220 0.211 0.186
K 0.041 0.056 0.343 0.751 0.272 0.427 1.257 1.278 1.342

Site occupancy
A 1.016 1.051 1.012 1.151 1.057 1.206 1.476 1.489 1.528
B 1.279 1.170 0.996 0.988 1.230 1.458 0.874 1.103 0.943
C 6.563 6.608 6.797 6.812 6.589 6.381 6.976 6.821 6.897
"Total Fe expressed as Fe203; n.d. =not detected; n.a. =not analyzed. Compositions 1-3, Sover North (this work); 4-6, Star
(Mitchell and Meyer 1989b); 7-9, New Elands (Mitchell and Haggerty 1986).

that each intrusion is characterized by hollandite of a particular composition. The majority


of the data plot within or close to the quadrilateral of compositions defined by end-member
K-Ba septe- and hexatitanates, suggesting that the solid solutions present are primarily
between BaFe3+2Ti6016, BaFe2~i?OI6, and K2Fe2~i?OI6.
Figure 2.89 shows that hollandites differ with respect to their BaO and K20 contents.
Those from Lace and Besterskral are Ba rich (Table 2.28) relative to hollandites from
Sover North and Star (Table 2.29). Hollandites from New Elands are richest in K20 (Table
2.29). Figures 2.88 and 2.89 demonstrate that the solid solutions present, with the
exception of New Elands hollandites, are dominated by Ba-septe- and hexatitanates. The
New Elands hollandites are close to K2FeTh016 in composition.
Hollandites from each intrusion differ with respect to their Fe, V, and Nb contents.
Figure 2.90 and Tables 2.28 and 2.29 show that hollandites from New Elands, Star, and
Lace are enriched in V, relative to those from Sover North and Besterskraal. Samples with
MINERALOGY OF ORANGEITES 205

+ BESTERSKRAAL
• SOVER NORTH
• LACE

\
o STAR /
• NEW ELANDS
~

\
K!.TRISKAIDECATITANATES
MgO +M20 3 Ti0 2 + Nb 20 5

Figure 2.88. Compositions of hollandites. K-triskaidecatitanates. and Ba-pentatitanates from diverse


orangeites plotted in the ternary system (K20+BaO)--(Ti02+Nb20S)--(MgO+M203) (wt %). where M = Fe.
Cr. V. and AI. Data sources: this work. Mitchell and Meyer (I989b). Mitchell and Haggerty (1986).

Fe3+j(Fe3+ + V) ratios greater than 0.5 may be regarded as mannardite-potassium man-


nardite solid solutions.
Hollandites typically contain from 0.5 to 3.0 wt % Nb20s, with those from
Besterskraal being relatively enriched in NbzOs (2.0-6.8 wt %). Table 2.28 shows that as
Nb contents increase, Fe contents decrease. There is no simple negative correlation
between Ti and Nb, suggesting that Nb is accommodated at the B- and C-octahedral sites
by a complex coupled substitution involving Nbs+, Fe3+, Ti4+, and lattice vacancies. The
Nb end member of the solid solutions involved is as yet unidentified but is unlikely to be
a hollandite group compound, as the host of barium niobium titanates in the iron-free
BaO-Ti02-Nb20S system (Millet et at. 1987) does not belong to this structural group.
Mitchell and Haggerty (1986) have noted that hollandites from New Elands contain
0.7-1.7 wt % Ce203. Other hollandites analyzed during the preparation of this work were
found to contain no detectable levels of Ce203 by electron microprobe analysis.
206 CHAPTER 2

BoO

I
\

o
00 0
o
o
• LACE
+ BESTERSKRAAL
• SaVER NORTH
o STAR
• NEW ELANOS

. 'III
\
Field of K-
.. TRISKAIOECATITANATES
: . saVER NORTH -
STAR - LACE

Figure 2.89. Compositions of hollandites. K-triskaidecatitanates. and Ba-pentatitanates from diverse


orangeites plotted in the ternary system BaO-{Ti02+Nb203+M203)-K20 (wt %), where M =Fe. Cr, V, and
AI. Data sources: this work, Mitchell and Meyer (1989b). and Mitchell and Haggerty (1986).
MrnNERALOGYOFORANGEITES 207


O·g • LACE
o STAR
+ BESTERSKRAAL
•• • •: . •• •

t-
0·8 • SOVER NORTH
• NEW ELANDS

0'7 o o

c o
m 0'6 o
o o o o
+ o

-
0·5
~ o
o
0'4
o 0 o
& o
0'3 • •
o
o

0'2 •• •
0'1

0'1 0'2 0'3


BaV2 TiS°H;

Figure 2.90. Compositions ofhollandites (this work. Mitchell and Meyer 1989b, Mitchell and Haggerty 1986)
from diverse orangeites plotted as Fe3+/(Fe3+ + V) versus KI(K + Ba) (atomic). Compositional field of other
hollandites from Mitchell and Bergman (1991), Mitchell and Vladykin (1993), and Mitchell (1994c).

2.5.1.3. Comparison with Hollandites from Lamproites, Kimberlites, and Other


Potassic Rocks

2.5.1.3.a. Lamproites. The priderite variety of hollandite is one of the typomorphic


minerals of the lamproite clan (Mitchell and Bergman 1991). In these rocks it occurs as
euhedral prismatic crystals which form after phenocrystal phlogopite and prior to ground-
mass phlogopite. Typically, strong compositional zoning is not present. Priderites crys-
tallize contemporaneously with silicates and are not characteristically associated with
calcite. A single occurrence of priderite within calcite globules has been reported by
Jaques et al. (1989a) from the Argy~ olivine lamproite dikes. Priderites may be mantled
by late-stage ilmenite or jeppeite.
208 CHAPTER 2

Table 2.30. Representative Compositions of Hollandites from Lamproitesa


Wt% 2 3 4 5 6 7 8 9 10
Nb20s n.d. 0.06 0.08 n.d. n.d. n.d. n.d. 0.46 n.d. 0.45
Ti02 74.24 72.84 72.78 70.01 69.55 68.17 66.62 71.24 70.70 70.60
Cr203 0.06 2.20 4.08 2.03 3.25 4.33 5.16 5.60 6.64 11.03
V203 0.35 0.17 0.22 1.19 0.88 1.04 1.17 1.33 1.25 n.d.
Fe203 10.81 9.15 8.18 12.85 12.31 11.49 10.82 9.65 6.87 6.29
MgO 0.60 0.93 0.95 n.d. n.d. n.d. n.d. 1.62 1.36 0.38
BaO 6.91 6.49 6.96 10.34 10.51 10.85 12.74 4.67 7.00 5.39
K20 6.91 8.24 7.44 3.44 3.50 3.48 3.49 5.08 6.18 5.86
99.89 100.08 100.68 99.86 100.00 99.36 100.00 99.65 100.00 100.00

Structural formulae based on 16 oxygens


Nb 0.003 0.004 0.025 0.024
Ti 6.748 6.642 6.596 6.484 6.448 6.389 6.302 6.382 6.437 6.373
Cr 0.006 0.211 0.388 0.198 0.317 0.427 0.513 0.527 0.632 1.047
V 0.034 0.016 0.021 0.116 0.086 0.103 0.117 0.126 0.120
Fe3+ 0.983 0.835 0.741 1.191 1.142 1.078 1.024 0.865 0.626 0.568
Mg 0.108 0.168 0.170 0.288 0.245 0.068
Ba 0.327 0.308 0.328 0.499 0.508 0.530 0.628 0.218 0.332 0.254
K 1.065 1.275 1.142 0.540 0.550 0.553 0.560 0.772 0.954 0.897

Site occupancy
A 1.393 1.583 1.471 1.039 1.058 1.083 1.188 0.990 1.286 1.151
B 1.131 1.230 1.321 1.505 1.545 1.607 1.655 1.806 1.623 1.683
C 6.748 6.645 6.601 6.484 6.448 6.389 6.302 6.406 6.437 6.397
"Total Fe expressed as Fe203. n.d. = not detected. Compositions 1-3, Mt. North, West Australia; 4-7, Endlich Hill, Leucite
Hills, Wyoming; 8--10, Francis, Utah. (All data this work.)

Mitchell and Bergman (1991) have shown that each lamproite province is charac-
terized by hollandite of a particular KlBa ratio. The solid solutions present in lamproite
hollandites are similar to those of hollandites from orangeites. Hence, Table 2.30 and
Figure 2.91 show that lamproite and orangeite hollandites do not differ significantly in
their major element compositions.
Figure 2.92 shows that orangeite hollandites, with the exception of New Elands, are
typically richer in BaO than lamproite hollandites. BaO-rich hollandites from the Leucite
Hills have compositions which overlap those of orangeite hollandites while those from
West Kimberley are significantly enriched in K20.
Significant differences exist with respect to the V and Crcontents of orangeite (Tables
2.28 and 2.29) and lamproite hollandites (Table 2.30). Mitchell and Bergman (1991) have
noted that the V203 contents of lamproite hollandites typically do not exceed 1 wt %,
although Jaques et at. (l989a) have reported priderite with 1.3-1.7 wt % V 203 from
Argyle. Data obtained during the preparation of this work suggest that the V 203 contents
of iamproite hollandite are unlikely to exceed 2.0 wt % (Table 2.30). In marked contrast
to the very low Cr contents of orangeite hollandites, the Cr203 contents of lamproite
MINERALOGY OF ORANGElTES 209

/
\

BENFONTEIN -
WESSELTON
KIMBERLITE
o
3+
K2 Fe 2 Ti60 16
.\

LAMPROITE~"""

(MgO + M2 0 3 ) ...~--,.--~-~--,.--~---->
30 25 20 /5 /0 5 Ti0 2 +Nb 20 5

Figure 2.91. Comparison of the compositions of hollandites, K-triskaidecatitanates, and Ba-pentatitanates


from orangeites (this work), Benfontein and Wesselton kimberlites (Mitchell1994c),lamproites (Mitchell and
Bergman 1991, this work), and Murun ultrapotassic syenites (Mitchell and Vladykin 1993). P = Ba-pentati-
tanates, H = hollandites, T = K-triskaidecatitanates.

hollandite varies from <0.2-7 wt % and rarely reaches 11 wt % (Mitchell and Bergman
1991, this work).
Figure 2.93 illustrates how lamproite hollandites may be distinguished from those
of orangeites on the basis of their Cr203 and V203 contents. Note that because individual
hollandites may have low Cr or V contents it is necessary to analyze a suite ofhollandites
from a given locality before drawing any conclusions as to their magmatic affinity.
In summary, lamproite hollandites are very similar in composition to those in
orangeites. The principal differences are that orangeite hollandites show solid solutions
toward the mannardite--potassium mannardite series, while those from lamproites exhibit
solid solutions toward the redledgeite--potassium redledgeite series. The hollandites differ
significantly in their paragenesis, those in lamproite being relatively early-formed
groundmass or microphenocrystal phases, whereas those in orangeites are very late stage
groundmass phases intimately associated with carbonates and potassium triskaidecati-
tanates.
210 CHAPTER 2

BoO
\

Ba3TisO'3 /
K2 0 ~
\• I
/
85
\
BENFONTEIN -
JEPPEITE
/ WESSELTON
\ I BaTi ,3 0 27

\ /
\ I
\ /

6\
BoO \
I
/
I
K~ Ti02 \ /
\/
,/ I
15 1- 1- 1- 10 5 Ti0 2 + M2 0 3
~ ,A~-r. ~-r.
,. IP III.,
·0 ~"~,,
+ Nb 2 0,
..
~ ~. ~.
~o "'0

'-.,-J
K - HOLLANDITES
Figure 2.92. Comparison of the compositions of hollandites. K-triskaidecatitanates. Ba-pentatitanates from
orangeites. kimberlites,lamproites. and ultrapotassic syenites. Data sources and abbreviations as in Figure 2.91.
MINERALOGY OF ORANGElTES 211

t ': 1ORANGEITE/
-8-i'
'1-.
' I
I

i : j : /
1t)5~
~ 4 -1 / ,."
3 J .·/ ",,,,,,,
%~ I ", ", ", ,.LAMPROITE
2 J-'" i O 00
- O~~o 800~'b°Q
g'lJ OC% 0
234567891011

Cr203 wt. %) ~

Figure 2.93. enD3 versus V203 for hollandites from orangeites and iamproites. All data this work.

2.5.1.3.b. Kimberlites. Hollandites have only been described from the unusual.
highly differentiated. Wesselton Water Tunnels and Benfontein calcite kimberlites
(Mitchell 1994b). Here they occur in the groundmass of the rock as small (5 x 25 ~m)
euhedral crystals intergrown with calcite and dolomite. Their typical absence from other
less-evolved kimberlites. including segregation-textured hypabyssal varieties. indicates
that hollandite is not a characteristic mineral of archetypal kimberlite. Its absence may
be a consequence of the early crystallization of Ti-bearing spinels and perovskite and
rapid depletion of the magma in Ti. prior to crystallization of the calcite-rich groundmass
at temperatures which might permit K-Ba titanate crystallization. Consequently. Ba
remains in the magma until sequestration in late stage phlogopite-kinoshitalite solid
solutions.
Representative compositions of kimberlite hollandite are given in Table 2.31. Par-
ticularly noteworthy is the absence of potassium and chromium. Hollandites from
Wesselton are poor in Nb (not detected) and rich in V203 (0.5-4.0 wt %) relative to those
from Benfontein (1.0-5.7 wt % Nb205. V not detected). All are Cr203 poor. The minerals
are principally members of the solid solution series between BaFe3+2Ti6016 and BaFe2+
Tb016 with minor solid solution toward mannardite and Nb-bearing hollandite and are
thus very similar in composition to Ba-rich hollandites in orangeites (Figures 2.90-2.92).

2.5.1.3.c. Other Alkaline Rocks. Hollandites from the Kovdor carbonatite


(Zhuravleva et al. 1978) occur in association with geikileite. zirkelite. and clinohumite.
This hollandite. which probably formed by reaction of early crystallizing Ti minerals with
residual carbonatite magma, is a Ba-septetitanate. Although lacking V, it is similar in
212 CHAPTER 2

Table 2.31. Representative Compositions of Hollandites from Kimberlites, Ultrapotassic


Syenite, and Carbonatite Complexesa
Wt% 2 3 4 5 6 7 8 9 10
Nb20 5 n.d. n.d. n.d. U19 2.47 5.65 n.d n.d. n.d. 0.91
Ti02 71.72 72.23 69.39 68.41 67.55 67.96 69.71 65.90 72.00 69.02
Cr203 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.
V20 3 0.50 1.35 4.01 n.d. n.d. n.d. n.d. n.d. n.d. 0.52
Fe203 12.02 9.38 6.85 14.99 13.71 11.95 12.80 13.31 11.39 8.87
MgO n.d. n.d. n.d. n.d. n.d. n.d. 1.10 n.d. n.d. 0.26
BaO 15.08 17.60 18.40 15.51 16.09 14.43 14.12 21.10 15.00 17.79
K20 n.d. n.d. n.d. n.d. n.d. n.d. 1.13 0.33 0.40 n.d.
99.32 100.56 98.65 100.00 99.82 99.99 98.86 100.64 98.39 99.53

Structural formulae based on 16 oxygens


Nb 0.062 0.128 0.320 0.042
Ti 6.744 6.802 6.720 6.475 6.593 6.401 6.595 6.468 6.645 6.773
Cr
V 0.050 0.134 Q.411 0.054
Fe3+ 1.131 0.884 0.665 1.420 1.182 1.126 1.212 1.307 1.304 0.871
Mg 0.206 0.051
Ba 0.739 0.864 0.930 0.765 0.722 0.708 0.696 1.079 0.721 0.910
K 0.181 0.055 0.063

Site occupancies
A 0.739 0.864 0.930 0.765 0.722 0.708 0.877 1.134 0.784 0.910
B 1.181 1.018 1.036 1.420 1.182 1.126 1.418 1.307 1.304 0.975
C 6.744 6.802 6.720 6.537 6.721 6.721 6.595 6.468 6.645 6.815
"Total iron expressed as Fe203; n.d. = not detected. Compositions 1-3, Wesselton and 4-6, Benfontein Kimberlites (Mitchell
1994b); 7-8, Little Murun ultrapotassic syenite (Mitchell and Vladykin, 1993); 9, Kovdor complex (Zhuravleva et al. 1978);
10, Schryburt Lake complex, includes 1.38 WI % Ce:!03, 0.74 wI % CaO, 0.24 wI % srO (Platt 1994).

composition, Ko.06Bao.74f'eI.08Ti6.81016 (Table 2.31), but not paragenesis, to hollandites


from orangeite.
Hollandites have been described from the Schryburt Lake carbonatite complex (Platt
1994). Here they occur enclosed within perovskite in a perovskite-spinel cumulate
derived from an ultramafic lamprophyre. They are similar in composition (Table 2.31) to
the Kovdor hollandite in being a Ba-septetitanate. They differ in containing significant
amountsofCe203 (1.2-1.7 wt %), CaO (0.6-1.1 wt %), and Nb20S (0.5-0.7 wt %). They
also differ in paragenesis, and Platt (1994) considers them to be an early primary phase
and not a reaction product.
Hollandites from the Little Murun ultrapotassic complex have been described by
Mitchell and Vladykin (1993). Here they occur in aegirine potassium feldspar syenites
in association with wadeite, tausonite, sphene, Ti-magnetite, and barytolamprophyllite.
They occur as small (<10 ~m) subhedral prisms mantling earlier-formed tausonite,
Ti-magnetite, and K-triskaidecatitanate. The paragenesis is interpreted to indicate that the
MINERALOGY OF ORANGEITES 213

hollandites are the products of reaction between preexisting titanates and groundmass-
forming magma. The hollandites contain 0.2-2.0 wt % K20, 8.7-14.4 wt % Fe203, and
11.5-21.1 wt% BaO (Table 2.31) and are essentially solid solutions between BaFe2+-
TbOl6 and BaFe3+ri6016.
The Little Murun hollandites lack V, Cr, and Nb but are otherwise similar in
composition to orangeite hollandites (Figures 2.91 and 2.92). Particularly noteworthy is
their association with K-triskaidecatitanate (2.5.2)

2.5.2. Potassium Triskaidecatitanate


The Lace, Sover North, and Star orangeites contain a Ba-free potassium titanate
whose composition is regarded as K2Ti13027. The Sover North and Star occurrences have
been previously described by Mitchell and Vladykin (1993) and Mitchell and Meyer
(1989b) respectively.
At Lace this mineral occurs as stellate clusters of slender primatic crystals (Figure
2.94), at Sover North as anhedral crystals mantled by hollandite (Figure 2.95), and at Star
as anhedral isolated grains. Optically, the mineral is very similar to hollandite, being
opaque in thin section and gray in reflected light.

Figure 2.94. Prismatic potassium triskaidecatitanate (KT), Lace. Backscattered electron image. C= calcite,
S =serpentine, M =mica.
214 CHAPTER 2

Figure 2.95. Potassium triskaidecatitanate (KT) mantled by hollandite (H), Sover North. Backscattered
electron image.

Tables 2.32 and 2.33 give representative compositions of the mineral and demon-
strate there are significant inter-intrusion compositional differences. Individual crystals
may be homogeneous (Star) or strongly-zoned with respect to V and Fe (Lace). The
mineral has significant V and Nb contents but lacks, or is poor in, K and Cr. There are
negative correlations between Ti and Nb or V and Fe.
This titanate was previously regarded as potassium mannardite by Mitchell and
Meyer (1989b); however, it has greater Ti02 (approx. 80 wt %) contents and lower K20
(8-10 wt %), Fe203 (1-8 wt %), and V203 (1-14 wt %) than K2Fe3+2Ti6016-K2V2Ti6016
solid solutions (K20 = 12.8-13.0 wt %, Fe203 =21.8 wt %, V203 =20.7 wt %, Ti02 =
65.4-66.3 wt %). Structural formulae calculated on the basis of 16 oxygens are not in
accord with a hollandite structure. Figures 2.88 and 2.89 show that the mineral is
compositionally distinct from hollandite and plots in Figure 2.89 close to ideal K2Ti\3027.
Structural formulae calculated on the basis of 27 oxygens, with the exception of high-V
examples, are in reasonable agreement with this composition.
Accordingly, the mineral is believed to have the general composition A2B\3027,
where A = K, Ba and B = Ti, Nb, V, Fe, Cr. Deviations from the ideal composition are
MINERALOGY OF ORANGEITES 215

Table 2.32. Representative Compositions of K-V Triskaidecatitanates from the Lace and
Sover North Orangeitesa
Wt% 2 3 4 5 6 7 8
Nb20 s 2.15 0.49 0.79 1.76 1.76 0.95 1.66 2.29
Ti02 82.22 80.38 79.58 80.52 76.90 75.72 80.52 79.24
Cr203 0.19 n.d. n.d. n.d. n.d. n.d. 0.25 0.15
V20 3 1.55 3.66 4.98 5.23 9.61 14.28 0.20 n.d.
Fe203 4.87 7.03 5.28 3.10 3.58 1.06 8.23 8.60
MgO n.d. n.d. n.d. n.d. n.d. n.d. 0.03 n.d.
BaO n.d. n.d. n.d. n.d. n.d. n.d. 0.59 0.76
K20 9.02 8.47 9.37 9.39 8.12 7.99 9.05 9.23
100.00 100.03 100.00 100.00 99.97 99.48 100.53 100.27

Structural formulae based on 27 oxygens


Nb 0.188 0.043 0.070 0.155 0.158 0.084 0.146 0.203
Ti 11.970 11.718 11.656 11.762 11.259 11.080 11.773 11.672
Cr 0.029 0.038 0.023
V 0.238 0.564 0.721 0.807 1.487 2.208 0.031
Fe 0.709 1.026 0.774 0.453 0.525 0.155 1.204 1.268
Mg 0.009
Ba 0.045 0.058
K 2.228 2.095 2.328 2.237 2.017 1.983 2.245 2.306
A 2.228 2.095 2.328 2.237 2.017 1.983 2.290 2.365
B 13.135 13.351 13.270 13.177 13.428 13.528 13.202 13.166
"Total Fe expressed as Fe:!03; n.d. = not detected. Compositions 1-6, Lace; 7-8, Saver North. (All data this work.)

undoubtedly due to nonstoichiometry resulting from the presence of elements occurring


in two valence states and site vacancies created by complex B-site substitutional schemes.
The compound is considered to be a new mineral. As X-ray diffraction data for this phase
are unavailable, the mineral remains unnamed. In this work it is termed "potassium
triskaidecatitanate" to reflect its composition.
The only other known occurrences ofK-triskaidecatitanate are from the Little Murun
complex, Siberia (Mitchell and Vladykin 1993) and a lamproite from Sisimiut, Greenland
(Scott 198 I).
In the Little Murun tausonite syenites it occurs as small anhedral grains mantled and
replaced by Ba-rich hollandites. Prismatic crystals are absent, and it cannot be determined
if the mineral is primary phase or a reaction product. The Murun material differs from
that of orangeites in that it lacks Nb, V, and Cr and is essentially K2(Ti,Fe)13027 (Table
2.33).
The Sisimiut occurrence is described as priderite by Scott (1981); however, the Ti02
content (78.9-81.6 wt %) is too high for this mineral to be a hollandite (Table 2.33). In
common with orangeite potassium triskaidecatitanate, the Sisimiut example contains
appreciable Fez03 (9.5-9.6 wt %) and has very low BaO contents (0.3-0.8 wt %). This
is the only known occurrence of this mineral in a lamproite.
216 CHAPTER 2

Table 2.33. Representative Compositions of K-V Triskaidecatitanates from Star Orangeite,


Little Murun Syenite, and Sisimiut Lamproitea
Wt% 2 3 4 5 6 7
Nb20 S 1.61 4.82 6.52 n.d. n.a. n.a.
Ti02 80.91 75.28 74.49 86.0 81.56 78.86 91.69
Cr203 0.19 1.10 1.42 n.d 0.42 0.48
V20 3 1.43 4.42 3.06 n.d n.a. n.a.
Fe203 7.03 3.84 3.90 4.8 10.56 10.69
MgO 0.14 1.37 0.98 n.d. n.a. n.a.
BaO n.d. n.d. n.d. n.d. 0.26 0.76
K20 8.50 9.08 9.10 8.5 8.65 9.59 8.31
99.81 99.91 99.47 99.3 101.45 100.38

Structural formulae based on 27 oxygens


Nb 0.141 0.428 0.584
Ti 11.814 11.111 11.079 12.456 11.769 11.631 13.000
Cr 0.029 0.181 0.222 0.064 0.074
V 0.221 0.689 0.481
Fe 1.027 0.567 0.580 0.696 1.525 1.578
Mg 0.441 0.401 0.289
Ba 0.019 0.058
K 2.105 2.273 2.296 2.088 2.118 2.399 2.000
A 2.105 2.273 2.296 2.088 2.137 2.457
B 13.273 13.367 13.235 13.152 13.358 13.283
"Total Fe expressed as Ftl203; n.d. = not detected; n.a. = not analyzed. 1-3, Starorangeite (Mitchell and Meyer 1989b); 4, Little
Mumn syenite (Mitchell and V1adykin 1993); 5-6, Sisimiut lamproite (Scott 1981); 7, ideal composition ofK21i13027.

2.5.3. Barium Pentatitanate


The Sover North orangeite contains rare anhedral opaque crystals of an Fe-bearing
barium titanate. A similar mineral has been reported by Mitchell and Vladykin (1993)
from the Little Murun tausonite syenite. Both occur as overgrowths upon preexisting
titanates. Table 2.34 indicates that both examples have compositions in accord with the
general formula Ba(Ti,Fe)sOIl. The monoclinic compound BaTi50n has been synthe-
sized by Tillmans (1969) and Ritter et al. (1986) but has not previously been reported as
a mineral.

2.6. PEROVSKITE

2.6.1. Paragenesis
Perovskite is a common groundmass mineral in orangeite. Modal abundances vary
from a few grains per thin section (<<0.1 vol %) to approximately 5 vol % (Skinner and
Clement 1979; see also 2.6.3). Perovskite abundances range from trace amounts to 1.5
vol % (Fraser 1987) or 3.3 vol % (Clement 1982) in the Finsch orangeites. Commonly,
MINERALOGY OF ORANGEITES 217

Table 2.34. Representative Compositions of Barium


Pentatitanate from the Sover North Orangeite and the Little
Murun Ultrapotassic SyeniteQ
Wt% 2 3 4 5
Nb 20 s 0.48 0.64 n.d. n.d.
Ti02 69.57 69.78 66.74 61.76 72.26
Cr203 0.05 0.06 n.d. n.d.
Fe203 3.33 3.01 6.48 11.58
MnO n.d. n.d. n.d. 0.07
MgO n.d. 0.23 n.d. n.d.
BaO 27.09 26.28 26.34 25.02 27.74
K20 0.16 0.16 0.14 0.29
---
100.68 100.26 99.70 98.72

Structural formulae based on 11 oxygens


Nb 0.020 0.027
Ti 4.807 4.815 4.674 4.406
Cr 0.004 0.004
Fe 3+ 0.230 0.208 0.454 0.827
Mn 0.004
Mg 0.031
Ba 0.975 0.945 0.961 0.930
K 0.019 0.019 0.017 0.035

Site occupancies
A 0.994 0.964 0.978 0.965
B 5.061 5.085 5.129 5.233
"Total Fe expressed as Ff!203. Compositions 1-2. Sover North; 3-4. Little Murun; 5.
ideal BaTisOll.

perovskite is inhomogeneously distributed. Thus, some of the Swartruggens dikes appar-


ently contain only trace amounts of perovskite, yet other contemporaneous dikes contain
significant amounts (this work).
Most orangeite perovskites are very small (<0.01 mm) subhedral-to-rounded dark-
brown crystals. The smallest are difficult to identify optically, as they are opaque and thus
may be misidentified as spinels. Euhedral perovskites of cubic habit are only found as
chadacrysts in groundmass mica. At Swartruggens euhedral perovskite has been found
enclosed in wadeite (this work). The Besterskraal evolved orangeite contains deep-
red-brown groundmass plates of perovskite that poikilitically enclose altered prismatic
silicate minerals (Figure 2.96).
Most orangeite perovskites have habits which indicate that they have been resorbed
during the later stages of crystallization of the groundmass. Such perovskites lack ilmenite
or rutile mantles, although some from Lace have apatite mantles. Those from Besterskraal
are intergrown with apatite and barite.
218 CHAPTER 2

Figure 2.%. Poikilitic perovskite (P). Besterskraal. Backscattered electron image. PX =diopside, 0 =olivilJe.

2.6.2. Composition
Compositional data have previously been reported only for perovskites from
Bellsbank (Boctor and Boyd 1982), New Elands (Mitchell and Meyer 1989a), and
Sydney-on-Vaal (Mitchell and Reed 1988). Of these studies, only Boctor and Boyd (1982)
presented full analyses of the mineral. Although new data are provided in this work for
perovskites from Besterskraal, Bellsbank, Sover, and Sover North, much further work is
required to characterize fully the compositional variation of orangeite perovskite.
Semiquantitative data obtained during the preparation of this work indicate that
orangeite perovskite varies considerably in composition. Those present in the least-
evolved orangeites contain fewer REE [<6 wt % total (REEh03] and srO «1 wt %) than
those in the most evolved varieties [> 10 wt % (REEh03, >3 wt % SrO]. Perovskites from
Sover and Bellsbank are rich in REE, but poor in srO «1 wt %).
Representative compositions of perovskite are given in Table 2.35. Sr- and REE-poor
perovskites exhibit limited solid solution toward loparite [< 10 mol % (Nao.5,REEo.5)Ti03]
and tausonite «1 mol % SrTi03) and may be termed "Ce-bearing perovskite" or
Table 2.35. Representative Compositions of Perovskite in OrangeitesO
Wt% 2 3 4 5 6 7 8 9 10
Nb20S 0.32 0.63 0.43 0.37 0.60 0.65 0.81 3.09 2.92 2.18
Ti02 54.60 52.99 54.97 55.41 51.17 50.75 51.25 50.44 52.18 50.04
Th02 n.a. n.a. n.a. n.a. 0.31 0.18 0.21 1.55 0.63 n.a
La203 1.51 2.89 5.31 3.34 3.20 3.38 3.55 3.77 2.56 3.70
Ce203 3.54 3.13 8.39 6.34 7.85 8.00 7.61 6.92 4.67 7.90
Pr203 n.d. n.d. n.d. n.d. 0.72 0.74 0.68 n.d. n.d. 0.99
Nd203 0.81 1.19 1.38 2.91 2.11 2.18 2.11 1.16 1.20 1.75
Sm203 n.d. n.d. n.d. n.d. 0.15 0.10 0.18 n.d. n.d. 0.07
FeO 2.64 2.57 1.00 2.42 1.21 1.25 0.67 2.30 2.70 1.62
MnO n.a. n.a n.a. n.a. n.d. n.d. n.d. n.a. n.a. n.a
MgO n.a. n.a. n.a. n.a. n.d. 0.20 n.d. n.a n.a. n.a
CaO 35.39 35.55 22.19 22.27 27.31 26.86 24.39 27.47 30.35 27.44
srO 0.70. 0.29 4.48 4.50 3.15 3.27 5.84 0.90 0.95 n.a.
Na20 1.06 0.96 1.85 2.44 1.82 1.93 2.40 2.12 2.30 1.54
-- -- -- -- --
100.57 100.20 100.00 100.00 99.60 99.49 99.70 99.72 100.46 97.60
Mol % end-member molecules
Loparite 9.7 8.7 20.2 23.9 17.7 18.8 23.6 21.2 15.8 15.8
Lueshite 3.0
Ca2Nb20 7 0.3 0.5 0.4 0.9 0.5 0.6 0.7 2.7 0.1 1.9
Ce21h07 0.2 1.8 5.5 2.1 4.0 3.8 1.3 0.6 6.0
Tausonite 1.0 0.4 7.3 9.3 4.5 4.8 8.6 1.4 1.4
Perovskite 88.9 88.6 66.3 63.9 73.2 72.0 65.8 74.1 79.7 76.3
"Total iron expressed as FeO; n.d. =not detected; n.a. =not analyzed. Compositions 1-2, Sover Mine; 3-4, Sover North; 5-7.
Besterskraal; 8-9, Bellsbank Southern Extension; 10, Bellsbank. Data sources 1-9. this work; 10, Boctor and Boyd (1982) .

..!.
o

• BESTERSKRAAL
+ SOVER NORTH
• SOVER MINE

Figure 2.97. Compositions of perovskites (mol %) plotted in the ternary system perovskite-Ioparite-tausonite.
LH and K are compositional fields ofperovskites from Leucite Hills lamproites (Mitchell and Steele 1992) and
kimberlites (Mitchell 1986), respectively.

119
220 CHAPrER2

w
I-
a:
0
z
0
:I:
U
..... 10 3
W
I-
!II::
en
>
0
a:
w
a..
ORANGEITES
• SYDNEY - ON - VAAL
10 2 + BESTERSKRAAL
o BELLSBANK
• SaVER MINE
KIMBERLITES
x BENFONTEIN

oF:TI FIELD OF
HYPABYSSAL KIMBERLITES

10
La Ce Pr Nd SmEu Gd Tb Dy Ho Er TmYb

Figure 2.98. Chondrite normalized rare earth distribution patterns for perovskites from orangeites (this work.
Mitchell and Reed 1988), Benfontein calcite kimberlite (Jones and Wyllie 1984), and diverse hypabyssal
kimberlites (Mitchell and Reed 1988).

"perovskite" (sensu stricto). Figure 2.97 shows that Sr- and REE-rich perovskites from
evolved orangeites may be termed "cerian strontian perovskite." The solid solutions
present are primarily between perovskite (65-80 mol % CaTi03), loparite (19-27 mol %),
and tausonite (5-10 mol %). Sr- and REE-poor perovskites from unevolved orangeites
plot near the CaTi03 apex of this diagram. The evolutionary trend of composition is from
CaTi03 toward loparite with slightly increasing tausonite contents.
MINERALOGY OF ORANGEITES 221

Figure 2.98 illustrates chondrite nonnalized REE distribution patterns for REE in
orangeite perovskites. Unfortunately, the electron microprobe data given by Boctor and
Boyd (1982) have been shown by Jones and Wyllie (1984), and Mitchell and Reed (1988)
to be erroneous. Consequently, these data are not plotted in Figure 2.98. As electron beam
methods of analysis are unsatisfactory for the analysis of the heavy REE, a complete REE
distribution pattern (obtained by ion microprobe) is available only for one perovskite
from Sydney-on-Vaal (Mitchell and Reed 1988).
The available data nevertheless demonstrate the extraordinary enrichment of light
REE in these perovskites as indicated by the very steep slopes of the REE distribution
patterns. The LalYb ratio of the Sydney-on-Vaal specimen is 3145, and no Eu anomaly
is present.
Other elements present in significant amounts include FeOT (1-4 wt %), Nb20S
(0.6-3.6 wt %), Na20 (1.9-2.3 wt %), and Th02 «0.1-1.5 wt %).

2.6.3. Comparison with Perovskites from Kimberlite


Perovskite is a ubiquitous mineral in archetypal kimberlites. Modal abundances
range from trace amounts to 15 vol % (Skinner and Clement 1979, Mitchell 1986,
McCallum 1989), but can rise to major levels (>25 vol %) in occurrences such as the
Benfontein Sills, where perovskite has been locally concentrated by differentiation
processes.
The bulk of the perovskite occurs as discrete euhedral-to-subhedral or rounded
crystals ranging in size from 0.01 to 0.2 mm, with the majority of unresorbed crystals
being 0.5 to 0.1 mm. Perovskite also typically occurs as reaction mantles about magnesian
ilmenite macrocrysts, rutile, and spinels. Perovskite is itself resorbed and mantled by
rutile. For further details of perovskite paragenesis in kimberlites, see Mitchell (1986).
The principal differences in the perovskite paragenesis ofkimberlites and orangeites,
first noted by Skinner (1989), are in the size and abundance of the crystals. Orangeites
typically contain relatively few small perovskite grains, whereas kimberlites contain
perovskites which are typically 2-5 times larger and 2-10 times more abundant.
Perovskites in kimberlites are similar in composition (Table 2.36) to those from
unevolved orangeites in being essentially CaTi03 (>90mol %) and containing limited
amounts of FeOT (1-2 wt %) and Nb20s (0.5-2.0 wt %). There is little variation in the
composition of kimberlite perovskites with respect to either paragenesis (Mitchell 1986,
McCallum 1989, Jones and Wyllie 1984) or the degree of differentiation of the magma.
Sr-, Na-, and REE-rich perovskites do not occur even in the most-evolved kimberlites,
e.g., Benfontein [5--6 wt % (REEh03, <1 wt % Na20, <0.5 wt % srO; Jones and Wyllie
1984, Boctor and Boyd 1981], Wesselton Sill [4.3-5.5 wt % (REEh03; Mitchell 1986].
Hence, the majority of kimberlite perovskites so far examined exhibit only very limited
solid solution toward loparite «10 mol %) or tausonite «1 mol %) (Table 2.36, Figure
2.97). As perovskites from both orangeites and kimberlites are relatively poor in Na, there
is an excess of Ce after formation of loparite in molecular calculation schemes such as
used by Mitchell and Vladykin (1993). This excess Ce may be expressed as theCe2Th07
molecule, which typically amounts to <5 mol % (Table 2.36).
222 CHAPl'ER.2

Table 2.36. Representative Compositions of Perovskite in Kimberlites and Lamproites Q

Wt% 1 2 3 4 5 6 7 8 9
Nb 20 S 0.60 0.38 1.33 0.9 n.a. 1.09 0.69 0.47 n.d.
Ti02 53.66 55.53 52.1 55.8 54.28 50.2 50.1 55.01 48.2
Th02 1.02 n.a. n.a. n.a. n.a. n.a. n.a. 0.22 n.a.
La203 1.11 0.49 1.04 0.32 0.67 3.72 4.14 1.38 2.0
Ce203 2.41 1.49 3.05 0.80 2.41 7.62 9.26 1.05 5.1
Pr20 3 0.18 0.33 0.34 0.06 0.05 0.78 0.87 n.d 1.5
Nd20 3 1.94 0.65 1.26 0.30 1.14 2.53 2.61 0.55 1.9
Sm20 3 0.24 0.13 0.14 0.04 0.03 0.20 0.15 n.d. n.a.
FeO 0.87 1.06 1.95 1.05 1.24 0.49 0.72 0.25 11.4
MnO n.a. 0.02 0.04 0.02 n.a. n.a. n.a. n.d. n.a.
MgO n.d. 0.18 0.11 0.11 0.34 n.a. n.a. n.d. n.d.
CaO 35.89 39.38 35.7 39.5 36.54 22.54 21.1 36.62 25.1
SJO 0.23 n.a. 0.23 0.08 n.a. 7.02 5.53 2.83 3.8
BaO n.a. n.a. 0.14 0.16 n.a. 0.32 0.26 n.a. n.d.
Na20 0.08 0.38 0.33 0.36 0.79 2.94 3.01 n.d. 1.0
--
98.23 100.02 97.76 99.50 97.49 99.45 98.44 98.38 100.00

Mol % end-member molecules


Loparite 0.8 3.4 3.1 2.6 7.3 28.1 30.07 11.1
Lueshite OJ 0.7
Ca2N~07 0.5 0.3 1.1 0.5 0.4 0.6 0.4
Ce2Ti207 4.8 0.9 3.6 0.1 1.0 2.6 5.4
BaTi03 0.1 0.1 0.3 0.3
Tausonite 0.3 0.3 0.1 10.5 8.4 3.9 6.3
Perovskite 93.6 95.4 91.8 96.4 92.7 59.9 59.0 93.1 77.1
"Total Fe expressed as FeO; n.d. = not detected; n.a. =not analyzed. Compositions 1-5, kimberlites; 6-9,lamproites; I, Frank
Smith (Ibis work); 2, Chicken Park (McCallum 1989); 3, Benfontein (Jones and Wyllie 1984); 4, Premier (Jones and Wyllie
1984); 5. liqhobong (Boctor and Boyd 1980); 6-7. Middle Table Mountain. Leucite Hills (Mitchell and Steele 1992); 8.
Walgidee Hills, West Australia (Ibis work); 9, Pilot BUlle, Leucite Hills (Carmichael 1967a).

Figure 2.99 shows that although the Sr and Ce contents of perovskites from kimber-
lites and unevolved orangeites are similar, there are no counterparts to the orangeite Sr-
and REE-rich perovskites in kimberlites.
The single exception to the above observations is the Green Mountain "kimberlite"
from which Boctor and Meyer (1979) report perovskites with 6.8-11.5 wt % Nh205 and
5.9-10.3 wt % (REEh03. These perovskites lack Na; thus, solid solutions with CaNh206
or Ca2Nb207 rather than loparite must be present. The anomalous composition of these
perovskites, .relative to those from all other kimberlites, together with the unusual
mineralogy of the rocks, i.e., abundant diopside and enstatite(?), suggest that the classi-
fication of this rock as a kimberlite should be re-appraised.
In common with orangeite perovskites those from kimberlites are enriched in the
light rare earths. Chondrite normalized REE distribution patterns (Figure 2.98) have very
steep slopes and no Eu anomalies. LalYb ratios of perovskite from hypabyssal kimberlites
range from ] 828 to 3229 and are thus not significantly different from that of the
MINERALOGY OF ORANGElTES 223

7'0

tae 6'0

+
BESTERSKRAAL
SOVER NORTH
• SOVER

a
o BELLSBANK
5'0
..:
~
4'0

...
0
C/) 3'0
WALGIDEE
HILLS ••
" "PILOT BUTTE
LEUCITE HILLS
2'0 ~
/ PRAIRIE
' " CREEK OLIVINE LAMPROITE

....
_-,------------,/ ELLENDALE
1'0 LI , (/ .... , •.• ,,) 0."'" - .......... o ..... PEROVSKITE
/ ~" .' ...... '· ... 0
PYROXENITES
.-~ :...,....-.-...... "_~.",, 00
KIMB R
I I
~ANDtmOITESo
I I
......
~~_r_-._--r---r-__,_-_,.-_,---I

1'0 2'0 3'0 4'0 5'0 6'0 7'0 8·0 9'0 10'0 11'0 12'0

Ce203 wt. % •
Figure 2.99. Ce203 versus srO (wt %) compositional variation of perovskites from orangeites (this work).
lamproites (Mitchell and Steele 1992). kimberlites. alntiites. and perovskite pyroxenites (this work).

Sydney-on-Vaal orangeite perovskite. However, Figure 2.98 demonstrates that


perovskites from orangeites are enriched in the light REE (La-Nd) relative to those from
kimberlites. The figure also shows that kimberlite perovskites typically have LaN < CeN
in contrast to those from orangeites where LaN> CeN. Although the data base is extremely
limited, it appears that the heavy REE abundances are not significantly different.

2.6.4. Comparison with Lamproite Perovskite


Perovskite is not a ubiquitous mineral in lamproites. It is found only in olivine
lamproites (Ellendale, Prairie Creek) and evolved madupitic lamproites such as occur at
Pilot Butte and Middle Table Mountain (Leucite Hills). It typically forms small (<0.50
~m) subhedral-to-euhedral crystals and is commonly poikilitically enclosed by ground-
mass mica (Mitchell and Bergman 1991). Poikilitic plates of perovskite, similar in
character to those found in the Besterskraal orangeite, have been described from the
Middle Table Mountain lamproite by Mitchell and Steele (1992).
The composition (especially the REE abundances) of perovskite in lamproites is
inadequately characterized. Available data are summarized by Mitchell and Bergman
(1991). Perovskite in Ellendale (West Australia) olivine lamproites varies widely in
composition within and between intrusions and contains 3-12 wt % (REE)203, 0.2-1.5
wt % srO, 0.9-2.1 wt % FeO, and 0.6-1.3 wt % Nb20S (this work). Perovskites from the
Walgidee Hills (West Australia) lamproite pegmatite and the Prairie Creek (Arkansas)
madupitic lamproite contain 0.2-2.8 wt % srO and <2 wt % Ce203 (Mitchell and
224 CHAPTER 2

lLI
I- 10 4
a::
0
z
0
I
U
.......
lLI
I-
~
(J)
>
0
a:: 10 3
lLI
a..

LAMPROITES
: } MIDDLE TABLE MOUNTAIN

o WALGIDEE HILLS
[8l FIELD OF ORANGEITE COMPOSITIONS

La Ce Pr Nd Sm Eu Gd Tb Oy Ho Er

Figure 2.100. Chondrite normalized rare earth distribution patterns of perovskites from lamproites (Mitchell
and Steele 1992, Mitchell and Reed 1988).

Bergman 1991, Figure 2.99). Perovskite from Pilot Butte and Middle Table Mountain are
enriched in Sr and REE relative to all other lamproite perovskites. The purple poikilitic
perovskites from Middle Table Mountain are remarkably similar in composition to Sr-
and REE-rich perovskites from the Besterskraal and Sover North orangeites (Table 2.36,
Figures 2.97 and 2.99).
Chondrite normalized REE distribution patterns for lamproite perovskite are illus-
trated in Figure 2.100. Although reliable data for the heavy REE are not available, the
patterns indicate that these perovskites are strongly enriched in the light REE with LaN>
CeN. The distribution patterns and REE abundances are similar to those of orangeite
perovskite.
MINERAWGY OF ORANGEITES 225

2.7. PHOSPHATES

2.7.1. Apatite

2.7.1.1. Paragenesis
Apatite is a ubiquitous groundmass mineral with abundances ranging from trace
amounts « 1 vol %) to 10 vol %. The distribution is not homogeneous within and between
intrusions. Thus different facies of individual dikes at Swartruggens contain markedly
varied abundances of apatite, and at Bellsbank the Bobbejaan dike contains significantly
more apatite than the Main dike (Bosch 1971, Skinner and Clement 1979, Clement 1982,
Fraser 1987, this work).
Apatite occurs primarily as euhedral elongated prisms ranging from 0.05 mm in
width to 0.3 mm in length. Prisms of 0.05-0.1 mm in length are common. Apatite is
commonly resorbed and may be replaced by calcite, Sr-Ba carbonate, or barite. In the
Lace, Swartruggens, and Besterskraal orangeites, apatite occurs as large (0.1-0.3 mm)
anhedral groundmass plates which poikilitically enclose previously formed spinels and
perovskite. Inclusions of pyrite and magnetite are found in some apatites at Swartruggens.
Apatite may occur as mantles upon perovskite or be complexly intergrown with daqing-
shanite and other REE phosphates. Apatite is not preferentially associated with ground-
mass carbonates and appears to have crystallized prior to and/or contemporaneously with
late-stage mica. The radiating aggregates of acicular apatite, which are common in
kimberlites, do not appear to be characteristic of orangeites.

2.7.1.2. Composition
The composition of apatite in orangeites has not been previously studied. Repre-
sentative compositions of apatites analyzed during the preparation of this work are given in
Table 2.37. Fluorine and water contents were not determined, so it is not known whether the
mineral is fluoro- or hydroxyapatite. X-ray spectra did not indicate the presence of chlorine.
Table 2.37 shows that apatites contain significant quantities of Si replacing P and
that the principal compositional variation is with respect to their Sr, REE, and Ca contents.
Prismatic apatites have the least Sr and REE contents and may be zoned from Sr-poor
cores to Sr-rich margins. The Sr content varies between intrusions; thus, apatites from
Swartruggens contain 3-6 wt % SrO, whereas Lace, Sover Mine, and Bellsbank apatites
contain only 1-3 wt % srO. REE contents of all these early-formed apatites are low
[<2(REEh03, commonly <1 wt %].
Groundmass poikilitic plates of apatite from Besterskraal contain from 2 to 22 wt %
srO, 0.60 to 2.6 wt % Si02, <1 to 8 wt % (REEh03, and up to 1.5 wt % BaO (Table 2.38).
Typically, Sr < Ca in these apatites, and they are rarely replaced by another discrete
Sr-phosphate with Sr> Ca (see 2.7.4) and Sr-Ba carbonate. Besterskraal apatites are light
REE rich with LaN> CeN.

2.7.1.3. Comparison with Kimberlite and Lamproite Apatite

2.7.1.3.a. Kimberlite. Apatite is a ubiquitous late-crystallizing groundmass mineral


in kimberlites in quantities ranging from 1 to 10 vol%. Crystals are typically euhedral,
226 CHAPTER 2

Table 2.37. Representative Compositions of Euhedral ApatiteU


Wt% 2 3 4 5 6 7 8
CaO 54.48 52.42 54.28 55.48 54.64 55.79 53.17 54.39
srO 2.56 5.36 1.27 2.43 2.00 1.97 5.61 4.24
BaO n.d. 0.60 n.d. n.d. n.d. n.d. n.d. n.d.
FeO 0.21 n.d. n.d n.d. 0.57 0.22 0.35 0.25
Th0 2 n.d. n.d. n.d. n.d. n.d. n.d. n.a. n.a.
La203 0.33 n.d. 0.40 n.d. n.d. n.d. n.d. n.d.
Ce203 0.46 n.d. 1.02 0.35 n.d. 0.53 n.d. n.d.
Pr20 3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. n.d.
Nd20 3 n.d. n.d. 0.60 n.d. n.d. n.d. n.d. n.d.
P20 S 40.97 40.28 40.73 40.16 40.33 40.43 40.87 40.12
Si02 0.59 0.94 2.29 1.58 2.46 1.06 n.a. n.a.
99.60 99.60 100.19 100.00 100.00 100.00 100.00 99.00
aTotal iron expressed as FeO; n.d. = not detected; n.a. = not analyzed. Composition 1-2, Lace; 3-4, Bellsbank; 5-6, Sover
Mine; 7-8, Swartruggens. (All data this work.)

small (0.1-0.2 mm) prisms. Apatites are particularly abundant in calcite-rich portions of
the groundmass where they commonly occur as radiating sprays of acicular crystals which
have nucleated at the margins of calcite-serpentine segregations. The habit of such
apatites suggests they formed during rapid quenching of the magma (Mitchell 1986,
Clement 1982). Poikilitic groundmass plates of apatite have not been reported from
archetypal kimberlites. Apatite is difficult to recognize optically where it is intimately
intergrown with groundmass calcite. In many instances such apatites are completely
pseudomorphed by calcite. For further details of apatite paragenesis see Mitchell (1986).
Very little is known of the composition of kimberlite apatites (Mitchell 1986). They
are poor in SrO «1 wt %; Scott Smith et al. 1984, Exley and Smith 1982) and (REEh03
«1 wt %; Ilupin et al. 1971, Exley and Smith 1982, Mitchell 1984a). There is apparently
no difference in their Sr and REE contents with respect to the degree of evolution of the
host kimberlite. Jones and Wyllie (1984) have shown that REE abundances in perovskites
are greater by an order of magnitude than those of coexisting apatite. This relationship is
also observed in orangeites. Apatites from Benfontein and Premier have linear chondrite

Table 2.38. Representative Compositions of Poikilitic Apatite from Besterskraalu


Wt% 2 3 4 5 6 7
CaO 52.63 51.39 44.94 43.26 41.86 42.06 34.58
srO 2.53 6.11 8.04 10.00 12.15 14.06 21.66
BaO 0.21 2.25 1.43 1.06 1.32 n.d. 0.82
FeO n.d. 0.69 0.24 n.d. n.d. 0.41 0.26
La203 0.66 n.d. 1.30 1.47 1.22 n.d. 0.78
Ce203 1.50 0.38 2.50 3.79 3.41 0.80 1.90
Pr203 n.d. n.d. n.d. n.d. n.d. n.d. n.d.
Nd20 3 0.38 n.d. 0.99 1.58 1.59 n.d. n.d.
P20 S 40.74 38.09 40.13 38.23 37.47 41.80 37.55
Si02 1.35 0.68 0.45 0.61 0.98 0.59 2.44
aAll data (this work) are SEMIEDS analyses summed to 100 wt % oxides. Total Fe expressed as FeO; n.d. = not detected.
MINERAWGY OF ORANGElTES 227

normalized distribution patterns in which LaN> CeN and Eu anomalies are absent (Jones
and Wyllie 1984). Apatites in kimberlites contain significant quantities of Si02 (0.7-2.2
wt %; Scott Smith et.al. 1984, Exley and Smith 1982).
Kimberlite apatites differ from those in orangeites with respect to their paragenesis.
In orangeites, apatites are principally microphenocrysts and late-stage groundmass crys-
tals. Rapidly-quenched apatites associated with calcite segregations are not characteristic
of orangeites. Late-stage crystallization as poikilitic plates is not observed in kimberlite.
Orangeite apatites are typically richer in srO (> 1 wt %) than those from kimberlites. REE
abundances, with the exception of the Besterskraal apatites, are similar.

2.7.1.3.h. Lamproite. Apatite is a ubiquitous mineral in lamproites, occurring as


phenocrysts and microphenocrysts. Many apatites exhibit rounded and embayed habits.
Resorption is not associated with replacement by calcite. Hollow-cored apatites which
may have grown rapidly from supersaturated melts are common. Apatite crystallizes after
phenocrystal phlogopite and is poikilitically enclosed by groundmass potassium richterite
and sanidine. Crystallization is contemporaneous with priderite, wadeite, and perovskite
(Mitchell and Bergman 1991).
Lamproite apatites are fluor-apatites (2-7 wt % F) characterized by high srO
(typically 1-6 wt %, ranging up to 12 wt %, in some Leucite Hills examples). BaO
contents vary widely. Those from the Leucite Hills typically contain 0.2-0.4 wt % BaO,
with the exception of Sr-Ba apatites from Middle Table Mountain which contain up to
18 wt % Ba. The majority of apatites in the West Kimberley province contain 0.5-1.0
wt % BaO, although examples from the Walgidee Hills contain 2.1-12.3 wt % BaO
(Mitchell 1986, Edgar 1989). Apatites with high Ba and Sr contents appear to be
characteristic of lamproites. Although few reliable data exist, lamproite apatites appear
to be poor in (REEh03 (<2 wt %; Kuehner et al. 1981, Carmichael 1967a). The single
REE distribution pattern available is linear with LaN> CeN, a LalYb ratio of 90 and no
Eu anomaly (Mason 1977). Si02 contents are low (typically <1 wt %; Mitchell 1986,
Edgar 1989).
Apatites inlamproites do not differ significantly from those occurring in orangeites
either in paragenesis or composition with the possible exception of their Si02 contents.
However, further studies are required, as the composition of apatite in both rock types is
insufficiently characterized.

2.7.2. Daqingshanite
Orangeites from the Sover Mine contain anhedral 10-50 Ilm crystals of a REB-Sr
phosphate set in a matrix of calcite. Table 2.39 indicates that this mineral has a composi-
tion which corresponds to that of daqingshanite [(Sr,Ba)3REE(P04)(C03)3; Yingchen
et al. 1983]. The Sover North example is very similar to daqingshanite-(Ce) from the
Nkombwa carbonatite complex, Zamibia (Appleton et al. 1992) in being poor in BaO
relative to barium daqingshanite-(Ce) from the type locality at the Bayan Obo iron
ore-REE deposit, China (Yingchen et al. 1983). Although X-ray data are not available,
the compositional data indicate that this Sr-REE carbonate is undoubtedly daqingshan-
ite-(Ce). No other orangeites have been examined in sufficient detail to determine if
228 CHAPTER 2

Table 2.39. Representative Compositions of DaqingshaniteQ


Wt% 2 3 4 5 6
CaO 3.45 3.68 0.94 6.17
SJO 52.76 49.94 41.82 26.10 45.85 56.94
BaO 3.01 4.78 4.57 15.98
FeD 0.41 0.44 n.a. n.a.
La203 6.99 7.86 10.22 7.88
Ce203 14.84 15.79 12.24 10.16 24.21 30.06
Pr20 3 1.09 1.44 0.83 0.68
Nd20 3 3.04 3.23 1.71 1.59
P20 S 11.75 10.15 10.50 11.73 10.47 13.00
Si02 2.66 2.69 n.a. n.a
100.00 100.00 82.83 80.29 80.52
Structural fonnulae based on seven oxygens
Ca 0.322 0.356 0.122 0.723
Sr 2.669 2.614 2.694 1.655 3.000
Ba 0.103 0.169 0.199 0.685
Fe 0.030 0.033
La 0.225 0.262 0.419 0.318
Ce 0.474 0.522 0.498 0.407 1.000
Pr 0.035 0.047 0.034 0.027
Nd 0.095 0.104 ·0.068 0.062
P 0.868 0.776 0.987 1.086 1.000
Si 0.232 0.243
Site occupancies
A 3.124 3.172 3.005 3.063 3.000
B 0.829 0.935 0.864 0.814 1.000
C 1.100 1.019 0.987 1.086 1.000
"Total Feexpressedas FeO; n.a. = not analyzed. Compositions 1-2, Sover Mine (this work) expressed
as 100 wt % oxides on a C02-free basis; 3, Nkombwa (Appleton et al. 1992); 4, Bayan Obo
(Yingchen et al. 1983); 5--6, ideal composition of Sr3Ce(P04)(C03)3 expressed in terms of oxide
wt % on a C02-free basis; 6, composition recalculated to 100 wt % oxides.

daqingshanite is a common mineral in these rocks. The abundance of Sr-Ba-REE


carbonates and other Sr-REE-rich phosphates suggests that further investigations will
bring to light new occurrences of this mineral. Daqingshanite has not been found in either
lamproites or kimberlites.

2.7.3. Monazite
Orangeites from Bellsbank Southern Extension commonly contain oval 1-10 J..lm
grains of a monazite-like, complex Th-Sr-REE phosphate, AB04, where A =REE, Th,
Sr, Ca, and B =P, Si. The phase occurs as isolated discrete crystals set in an Fe-bearing
dolomite groundmass, and as irregular complex intergrowths with Sr-poor apatite and
Nb-rutile. Rarely, the phosphate is completely mantled by Nb-rutile. Isolated small grains
of the same mineral are also common in the Lace orangeites.
MINERALOGY OF ORANGElTES 229

Table 2.40. Representative Compositions of Monazitea


Wt% 2 3 4 5 6
CaD 3.23 2.57 5.68 2.60 2.75 2.74
srO 4.93 4.44 4.65 7.95 2.36 4.57
BaD 1.68 4.06 n.d. n.d. n.d. n.d.
FeD 0.70 1.49 2.96 0.64 0.20 0.66
MgO n.d. 1.44 1.41 1.18 n.d. n.d.
ThD2 3.70 3.15 n.d. 7.93 0.63 1.82
La2D3 13.91 13.07 8.27 10.61 15.68 13.27
Ce2D3 32.53 28.93 34.98 29.87 31.29 30.25
Pr2D3 n.a. n.a. n.a. n.a. 4.03 5.64
Nd2D3 6.45 8.00 9.63 8.42 12.58 13.26
P2DS 30.54 27.79 29.50 29.59 29.54 26.45
SiD2 2.33 2.41 1.35 1.20 0.54 1.35
"Total Fe expressed as FeO: n.d. = not detected: n.a. = not analyzed. All data (this work) SEMIEDS
analyses summed to 100 wt % oxides. Compositions 1-4, Bellsbank: 5-6, Lace.

The mineral shows considerable intergrain compositional variation with respect to


Th, Sr, and Ba (Table 2.40). Monazites from Lace are not as rich in Th and Ba as those
from Bellsbank. Extremely high Th contents are rarely found in Bellsbank monazite.
Monazite is not present in kimberIites or lamproites (Mitchell 1986, Mitchell and
Bergman 1991).

2.7.4. Sr-REE Phosphate


At Besterskraal, an unidentified REE-poor Sr-phosphate with srO > CaO occurs as
discrete mantles upon relatively Sr-poor apatite. This mineral is intergrown with Sr-Ba
carbonates and strontianite and may represent a reaction product formed between preex-
isting apatite and Sr-rich late-stage deuteric hydrothermal fluids. The mineral contains
26-37 wt % SrO, 20-36 wt % CaO, 1-5 wt % BaO, <4 wt % (REEh03, 31-34 wt %
P205, and 0.8-1.3 wt % Si02.

2.S. AMPHIBOLES-POTASSIUM RICHTERITE


Amphiboles with the optical characteristics of potassium richterite have been recog-
nized in evolved varieties of orangeite from Pniel (ErIank 1973, Tainton 1992), Sover
North (Tainton 1992, this work), Besterskraal (Skinner pers comm., this work), Lace (this
work), Makganyene (this work), and occurrences in the Prieska area (Skinner et al. 1994).

2.S.1. Paragenesis
Potassium richterite crystallizes as a late-stage groundmass mineral. At Pniel, and in
the SN2 intrusion at Sover North, it forms anhedral tablets which may subpoikilitically
enclose phlogopite and diopside. Pleochroism is from yellow-to-pinkish brown and
crystals are zoned toward more strongly pleochroic rims. Thin overgrowths of green-
brown pleochroic amphibole may occur upon the potassium richterite or diopside
(Tainton 1992). At Sover North, amphibole forms reaction coronas around olivine
xenocrysts. Amphibole is rarely found as strongly-yellow-pink pleochroic anhedral
230 CHAPTER 2

crystals in the SN 1intrusion at Sover North (Tainton 1992). In the Besterskraal orangeite,
amphibole occurs as pink poikilitic plates intergrown with groundmass potassium feld-
spar (this work).
Skinner and Scott (1979) have tentatively suggested that altered elongated laths of a
mineral from Swartruggens, which is regarded as an abundant (trace to 12 vol %) partially
altered primary phase, may be "anthophyllite." The composition (Si02 =52.5, Ah03 =
1.98, FeO = 5.4, MgO = 26.6, CaO = 0.6, Na20 = 0.2, K20 = 0.6, total = 88.1 wt %) is
not in accord with that of anthophyllite and may represent that of a mixture of alteration
products (serpentine, chlorite, brucite, etc.). It is considered here that anthophyllite is not
present in orangeites, and the material described by Skinner and Scott (1979) represents
pseudomorphs after an undetermined primary mineral.

2.8.2. Composition
The few compositional data available (Tainton 1992, this work) demonstrate that
there is considerable intra- and inter-intrusion compositional variation (Figures 2.101 and

"' ....
• BESTERSKRAAL
'" '" \

'"
,,''
\

0 PNIEL
'"
• SOVER NORTH
,
..
12

,,,
I

,,,
(
ZONATION TREND
,
10 ,, WEST ,,
,'KIMBERLEY
ae ,,,
-
~

I-
8 ,,,
,
0 ,
I
/
If
: /-k-"\<V I
"<
6
, .~Q~
Cj <Q~
I
,, '\
, I
,,, \ I KAPAMBA
V
4
,
,,
I

,,
2
2 3 4 5 6
Na20 wt. %
Figure 2.101. Na20 versus FeOT compositional variation ofamphiholes from orangeites. Data sources: Sover

North (this work, Tainton 1992), Besterskraal (this work), Pniel (Tainton 1992). Compositional fields of
amphiboles from lamproites from Mitchell and Bergman (1991).
MINERALOGY OF ORANGEITES 231

UCITE HILLS • BESTERSKRAAL


o PNIEL
• SOVER NORTH
0'.

0'4

0'2

MARIO

1'0 2'0 3'0 4'0 5'0 6'0


No I K
Figure l.IOl. Na/K versus Tt (atomic) compositional variation of amphiboles from orangeites (this work,
Tainton 1992). West Kimberley. Leucite Hills. Smoky Butte. and Francis lamproites (Mitchell and Bergman
1991). Kapamba lamproites (Scon Smith et al. 1989). and potassic lamprophyres from Kajan (Wagner 1986).
Pendennis (Hall 1982). Bohemia (Nemcl! 1988). Yinniugou (Zhao et al. 1993). New South Wales (NSW)
leucitites (Mitchell and Bergman 1991). and the Murun ultrapotassic complex (this work).

2.102). At Pniel, amphiboles are very low in Ti02 (0.1-1.8 wt %) ranging in composition
from potassium richterite cores to magnesioarfvedsonite mantles (Table 2.41; Tainton
1992). Amphiboles at Sover North are primarily potassium richterites which exhibit a
wide variation in Ti02 content (2.0-7.1 wt %) at approximately constant NaIK. ratios.
Those richest in Ti02 (>0.25 Ti atoms/23 oxygens; Figure 2.102) are titanian potassium
richterites and titanian potassium magnesiokatophorites (Table 2.42; Tainton 1992, this
work). These amphiboles are mantled by Ti02-poor «1.5 wt %) magnesioarfvedsonite
and arfvedsonite (Table 2.42, Tainton 1992). Amphiboles at Besterskraal show consider-
able intergrain compositional variation with respect to Ti02 (4.7-7.1 wt %) at approxi-
mately constant NaIK. ratios (Table 2.41, this work). The general evolutionary trend in all
occurrences is one of an initial increase of Ti with Fe, followed by decreasing Ti and K
and increasing Fe3+ with Na.
Amphiboles in the Pniel and Sover North orangeites are characterized by low Ah03
«0.7 wt %) contents, in contrast to those from Besterskraal which contain 0.8-1.6 wt %
Ah03 (Tables 2.41 and 2.42), Richterites from Pniel contain sufficient Si and AI to occupy
232 CHAPTER 2

Table2Al. Representative Compositions of Amphiboles from Pniel and BesterskraalQ


Wt% 2 3 4 5 6 7 8 9 10
Si02 55.21 54.06 55.48 54.97 55.25 51.87 52.05 50.53 48.99 49.17
Ti02 0.94 0.96 0.65 0.52 0.12 0.40 4.83 5.63 6.27 7.14
AI 20 3 0.41 0.71 0.50 0.34 0.17 0.15 1.25 1.65 1.49 1.22
FeOr 4.21 4.33 4.28 8.53 11.86 21.61 5.25 6.44 7.57 9.15
MnO 0.07 0.15 0.08 0.15 0.44 1.00 0.09 n.d. 0.18 0.16
MgO 20.40 16.76 20.06 17.80 15.81 8.12 18.94 18.21 17.04 15.61
CaO 5.69 5.80 5.77 2.65 1.57 0.78 5.15 5.22 5.18 3.87
Na20 4.18 4.11 4.21 5.72 6.40 6.47 4;70 4.37 4.39 5.10
K20 4.96 4.94 4.94 5.04 4.76 5.15 4.94 4.58 4.57 4.59
96.13 95.27 95.97 95.72 96.38 95.58 97.20 96.97 95.68 96.38
FeO 4.08 4.40 6.11 18.13 6.07
Fe203 0.15 4.59 6.39 3.87 0.42
96.09 95.27 95.97 96.18 97.62 96.04 97.20 96.97 95.68 96.38

Structural formulae based on 13 cations and 23 oxygens

IIS·1 T
7.941 8.132 7.986 7.980 8.034 8.094 7.506 7.361 7.286 7.339
0.059 0.014 0.020 0.212 0.283 0.261 0.214
n 0.282 0.356 0.453 0.447

All 0.010 0.126 0.071 0.038 0.029 0.028


n 0.102 0.109 0.085 0.057 0.013 0.047 0.242 0.261 0.253 0.353
Fe3+ C 0.016 0.501 0.699 0.453 0.046

:j
Mg
0.049
0.009
4.373
0.545
0.009
3.758
0.525
0.010
4.304
0.534
0.018
3.851
0.743
0.054
3.427
2.361
0.132
1.885
0.633
0.011
4.071
0.739

3.954
0.942
0.023
3.777
1.141
0.020
3.468

cal 0.877 0.936 0.850 0.412 0.245 0.130 0.796 0.815 0.825 0.618
Na B 1.123 1.064 1.110 1.588 1.755 1.870 1.204 1.185 1.175 l.382

Nal 0.042 0.135 0.065 0.022 0.049 0.084 0.110 0.049 0.091 0.092
K A 0.910 0.948 0.907 0.933 0.883 1.023 0.909 0.851 0.867 0.873
aFeO and Fez03 calculated on the basis of stoichiometry (Droop 1987); FeOr =total Fe expressed as FeO; n.d. =not detected.
Compositions 1-3, potassium richterite, Pnie1 (fainton 1992); 4-6, potassium magnesioarfvedsonite, Pniel (fainton 1992);
7-10, titanian potassium richterite and magnesiokatophorite, Besterskraa1 (this work).

all of the tetrahedral sites and thus contain octahedrally coordinated AI. Richterltes from
Besterskraal and Sover North contain insufficient Si and Al to occupy all the available
tetrahedral sites in the structure. This deficiency is probably remedied by entry ofTi to
this site, as suggested for lamproite-derived amphiboles by Mitchell and Bergman (1991)
and Thy et al. (1987). In contrast, potassium magnesioarfvedsonites characteristically
contain octahedrally coordinated AI.
In many cases it is not possible to calculate the Fe2+ and Fe3+ contents of the
amphiboles using the method of Droop (1987), as they are nonstoichiometric and contain
less than 13 cations/23 oxygens. This, in some instances, is because the presence of Ti4+
creates vacancies in the structure (Mitchell and Bergman 1991). However, in all cases
where estimation of oxidation state is possible, it appears that Fe2+lFe3+ ratios are much
greater than unity (Tables 2.41 and 2.42).
MINERALOGY OF ORANGEITES 233

Table 2.42. Representative Compositions of Amphiboles from Sover NorthQ


Wt% 2 3 4 5 6 7 8
Si0 2 55.16 54.50 53.20 51.37 51.29 52.57 52.74 53.30
Ti0 2 1.91 2.49 3.23 6.54 7.12 0.55 1.09 0.84
A120 3 0.50 0.27 0.54 0.66 0.53 0.25 0.31 0.17
FeOr 4.55 4.68 5.84 5.03 5.70 20.99 19.98 19.70
MnO 0.25 0.07 0.12 0.11 0.13 1.01 0.34 0.98
MgO 19.80 19.88 19.26 18.81 18.04 8.95 9.97 9.90
CaD 5.58 5.70 5.62 4.24 3.90 0.33 0.54 0.69
Na20 4.37 4.10 4.62 4.87 5.06 7.08 6.82 6.76
K2D 5.01 4.71 4.74 4.62 4.73 5.42 5.22 5.23
---
97.13 96.40 97.17 96.25 96.50 97.15 97.01 97.57
FeD 17.16 15.85 15.77
Fe203 4.25 4.58 4.36
97.13 96.40 97.17 96.25 96.50 97.58 97.47 98.01

Structural formulae based on 13 cations and 23 oxygens


7.879 7.847 7.625 7.464 7.462 8.037 7.993 8.039
~1
S'I T 0.080 0.046 0.092 0.113 0.091 0.007
Ti 0.041 0.107 0.283 0.423 0.447
All 0.045 0.041
Ti 0.164 0.163 0.067 0.292 0.332 0.063 0.124 0.095
Fe 3+ 0.489 0.523 0.495

~~c
0.544 0.564 0.705 0.611 0.694 2.154 2.010 1.990
0.030 0.009 0.015 0.014 0.016 0.131 0.044 0.125
Mg 4.216 4.243 4.141 4.074 3.912 2.040 2.252 2.226
Ca I 0.854 0.879 0.869 0.660 0.608 0.054 0.088 0.112
Na B 1.146 1.121 1.131 1.340 1.392 1.946 1.912 1.888

Na I 0.064 0.024 0.161 0.032 0.035 0.153 0.092


K A 0.886 0.883 0.855 0.856 0.878 1.057 1.009 1.006
"FeO and Ft!:!OJ calculated on the basis of stoichiometry (Droop 1987). FeOr = total Fe expressed as FeO. Compositions 1-3,
potassium richterite (Tainton 1992); 4-5, titanian potassium richterite (this work); 6-8, potassium magnesioarfvedsonite
(Tainton 1992).

Levels of BaO and Cr203 in amphiboles in orangeites are not detectable by standard
electron microprobe analytical methods «0.1 wt %). Fluorine contents have not been
determined.

2.8.3. Comparison with Potassium Richterite in Lamproite and Other Potassic


Rocks
Titanian potassium richterite is one of the typomorphic minerals of lamproites
(Mitchell and Bergman 1991). However, its presence is neither ubiquitous nor confined
to lamproites. Importantly, primary amphiboles do not occur in archetypal kimberlites
(Mitchell 1986).
In lamproites, potassium richterite occurs principally as groundmass poikilitic plates
and is one of the last minerals to crystallize. Potassium richterite is commonly optically
234 CHAPTER 2

zoned, from pale-yellow-pink cores, through reddish-pink regions, to dark-red-brown


margins. The amphiboles are commonly intergrown with groundmass plates of potassium
feldspar. A second, less common paragenesis, is as small euhedral prisms lining vesicles
in extrusive lamproites. These amphiboles are optically and compositionally similar to
the groundmass poikilitic variety.
The paragenesis of poikilitic potassium richterite occurring in evolved orangeites is
essentially identical to that of richterite in lamproites. Although at Pniel potassium
richterite does not coexist with potassium feldspar, intergrowths with this mineral occur
at Sover North and Besterskraal.
Mitchell and Bergman (1991) have shown that individuallamproite provinces are
characterized by the presence of titanian potassium richterite and titanian potassium
magnesiokatophorite of distinct composition with respect to their Fe and Na contents and
Na/K ratios. Amphiboles from Murcia-Almeria, Kapamba, and Francis are Na-rich
relative to those from the Leucite Hills, West Kimberley, Smoky Butte, and Prairie Creek
(Figure 2.1 01). Compositional trends in all provinces are similar and characterized by
increasing Fe and Na contents, commonly at nearly constant K contents. This trend
represents evolution from FeOT-poor (2-3 wt %) titanian potassium richterite, through
FeO~-rich (10-14 wt %) potassium richterite, to titanian potassium magnesioarfved-
sonite. Lamproite amphiboles are Ah03 poor «1.5 wt %) and Ti02 rich (2-9 wt %).
Substantial tetrahedral site deficiencies are typically present as a consequence of the low
Al contents. Ferric iron-rich amphiboles are not characteristic of lamproites, and even
rocks containing tetraferriphlogopite lack such amphiboles.
Figure 2.101 shows that amphiboles from orangeites have compositions similar to
those of amphiboles in many lamproites. Compositional zonation and evolutionary trends
are similar. Potassium richterites in orangeites have Na20 < K20 ratios greater than 0.8
(Pniel, 0.82-0.95, most <0.9; Sover North, 0.84-1.03; Besterskraal, 0.95-1.12) and in
this respect are similar to amphiboles in some slightly evolved (6-10 wt % FeOT)
amphiboles in the West Kimberley and Leucite Hills provinces. Tetrahedral site deficien-
cies are common to amphiboles from both rock types.
Figure 2.102 shows that richterites in orangeites differ principally from those in
lamproites in terms of their Ti contents. All bona fide lamproites contain more than 0.25
atoms Ti/23 oxygens. Amphiboles from Sover North exhibit a wide range in Ti content
at approximately constant Na/K ratios. Although the majority of these data do not fall
within fields defined by lamproite amphiboles, some high Ti (>0.7 afu) examples of
lamproitic character are present in examples with the highest Na/K ratios.
Richterites from Besterskraal have similar Ti contents and Na/K ratios to amphiboles
from the Leucite Hills and Smoky Butte lamproites (Figure 2.102). Although they have
Ti contents similar to amphiboles from Francis and Kampaba, they do not plot in the fields
defined by these amphiboles as the latter have significantly lower K20 contents «4
wt %).
Amphiboles from Pniel are particularly low in Ti, and compositions plot far from the
fields defined by lamproite amphiboles in Figure 2.102. This may be a consequence of
the relatively unevolved character of their host rocks, if the trend defined by the Sover
amphiboles is considered indicative of the evolutionary trend of orangeite amphiboles
with respect to Ti and Na/K ratios. In this context it is significant that the Pniel amphiboles
MINERALOGY OF ORANGEITES 23S

have the lowest Na201K20 ratios and do not coexist with potassium feldspar. Hence, they
and their host rocks may be considered as unevolved, relative to Sover North and
Besterskraal orangeites.
Ti-poor magnesioarfvedsonites occur as mantles on richterite in both orangeites and
lamproites. These amphiboles probably form as a result of reaction of preexisting
amphiboles with late-stage Na- and Fe-rich deuteric fluids and may even be of subsolidus
origin (Mitchell and Bergman 1991). Alkali amphiboles in lamproites tend to be richer
in Ti02 (1-7 wt %; Mitchell and Bergman 1991) than those in orangeites «1 wt % Ti02),
but are otherwise of similar composition.
In summary, amphiboles in orangeites and lamproites are similar in paragenesis and
composition, the principal difference being that richterites in orangeites tend to be poorer
in Ti02. On the basis of the existing data, there is no means of distinguishing richterites
in orangeites from those in lamproites on the basis of their composition (see below).
Figure 2.102 also illustrates the compositions of some richteritic and magnesioarf-
vedsonitic amphiboles from peralkaline minettes, leucitites, and potassic ultramafic
lamprophyres. Amphiboles from these occurrences typically have much higher Na/K
ratios and lower Ti contents than amphiboles in lamproite ororangeite. Many have higher
Ah03 contents (1-3 wt %) and coexist with nepheline and/or aluminous pyroxenes.
Mitchell and Bergman (1991) discussed the occurrence and paragenesis of richteritic
amphiboles and concluded that, in some instances, amphiboles of similar composition
may occur in different rock types because of a convergence of amphibole evolutionary
trends in genetically unrelated magmas. Amphibole composition alone is not regarded as
diagnostic of any particular magma type and must be considered in conjunction with the
composition and paragenesis of associated minerals, if it is used in classifying a given
rock. Thus, richterites in leucitites, while similar in some respects to those in some
lamproites or minettes, differ in being richer in Al and occurring in association with
nepheline and AI-rich pyroxenes.

2.9. POTASSIUM FELDSPAR


Potassium feldspar has been recognized as a late-stage groundmass mineral in
orangeites from Besterskraal (Skinner pers. comm., this work), Sover North (Tainton
1992), Postmasburg (this work), Makganyene (this work), Voorspoed (this work), and
the Prieska region (Brandewynkuil, Sweetput-Soutput, Droogfontein, Albertshoop,
Nauga; Skinner et al. 1994). Typically the mineral occurs as colorless groundmass
poikilitic plates and, apart from serpentine, is the last complex silicate phase to crystallize.
The feldspar is typically altered to clay minerals and/or replaced by serpentine. Potassium
feldspar forms pseudomorphs after (?) leucite in the Sover North and Postmasburg
orangeites (Tainton 1992).
Table 2.43 gives representative compositions of primary potassium feldspar from the
Sover North and Postmasburg occurrences (Tainton 1992, this work). Feldspars from
Postmasburg contain <0.26 wt % Fe203, 0.15-0.4 wt % Na20, and 0.7-2.1 wt % BaO.
Those from Sover North contain 0.3-1.7 wt % Fe203, <0.02--0.12 Na20, and 0.1-1.5
wt % BaO. Potassium feldspar pseudomorphing "leucite" (see 1.10) does not differ in
composition from primary feldspar (Table 2.43; Tainton 1992).
236 CHAPTER 2

Table 2.43. Representative Compositions of Potassium Feldspar"


Wt% 2 3 4 5 6 7
SiOl 64.01 63.63 62.62 64.37 63.06 62.44 62.68
A1 20 3 18.14 17.79 18.49 17.77 18.69 19.17 19.22
FeZ03 0.46 1.24 0.56 1.13 0.26 0.20 0.13
CaO n.d. n.d. n.d. n.d. n.d. n.d. n.d.
Nap 0.06 0.12 0.09 0.08 0.23 0.34 0.40
KzO 16.44 16.28 16.04 16.81 15.89 15.70 15.30
BaO 0.15 0.08 1.25 n.a. 0.73 1.66 2.07
---
99.26 99.14 99.05 100.16 98.86 99.51 99.80

Structural formulae based on eight oxygens


Si 2.991 2.983 2.958 2.990 2.966 2.939 2.943
Al 0.999 0.983 1.029 0.973 1.036 1.063 1.620
Fe 0.016 0.044 0.020 0.040 0.009 0.071 0.005
Ca
Na 0.005 0.011 0.008 0.007 0.021 0.031 0.037
K 0.980 0.974 0.966 0.966 0.953 0.943 0.916
Ba 0.003 0.002 0.023 0.014 0.031 0.038

Mol % end-member compositions


Cn 0.3 0.2 2.3 1.4 3.1 3.8
An
Ab 0.6 l.l 0.8 0.7 2.1 3.0 3.8
FeOr 1.6 4.4 2.0 3.9 0.9 0.7 0.5
Or 97.5 94.3 94.9 95.4 95.6 93.2 91.9
aTotal iron expressed as Fe20J; n.d. =not detected; n.a. =not analyzed. Compositions 1-3. groundmass sanidine, Sover North
(this work); 4, sanidine pseudomorphing "Ieucite," Sover North (Tainton 1992); 5-7, groundmass sanidine, Postmasburg
PK37 (this work).

Potassium feldspar is not found in archetypal kimberlites but is a common mineral


in phlogopite and madupitic lamproites (Mitchell and Bergman 1991), where it occurs as
euhedral-subhedral prismatic microphenocrysts, as groundmass poikilitic plates, and as
pseudomorphs after leucite. Only the latter two parageneses are typical of potassium
feldspar in orangeites.
Mitchell and Bergman (1991) have shown that sanidines in lamproites are charac-
terized by low Na20 contents «2.5 wt %), widely varying Fe203 contents «0.1-5 wt %,
typically >1 wt %), and 0.1-1.7 wt % BaO. Iron-rich feldspars belonging to the iron
sanidine-sanidine series containing more than 1 wt % Fe203 (>4 mol % KFeSbOs), less
than 1.0 wt % Na20, and negligible CaO «0.1 wt %) appear to occur exclusively in
lamproitic parageneses. Each lamproite province is characterized by feldspar of a distinct
composition with respect to Fe and Na content.
Figure 2.103 indicates that potassium feldspars from orangeites are similar in
composition to groundmass sanidines from the West Kimberley and Kapamba lamproite
provinces, but poor in Fe203 and Na20, relative to groundmass and microphenocrystal
sanidines from other provinces. The BaO contents of both lamproite or orangeite feldspars
MINERALOGY OFORANGElTES 237

MURCIA - ALMERIA
MADUPITIC
LAMPROITES
1'0 LEUCITE
HILLS
~
,FRANCIS

It)
02'0 'SMOKY BUTTE
N
If

I
+

'- ---
MURCIA -ALMERIA--
1'0 -+ PHLOGOPITE LAMPROITES
• POSTMASBURG - PK37
+ SOVER NORTH
PRIMARY
• SOVER NORTH-
PSEUDOMORPHS

o 1'5

Figure 2.103. Na20 versus Fe203 (wt %) compositional variation of potassium feldspars from orangeites (this
work, Tainton 1992) and lamproites (Mitchell and Bergman 1991, Scott Smith et al. 1989). WK = West
Kimberley; K = Kampamba.

have been insufficiently characterized, but on the basis of the available data do not appear
to be significantly different.
In summary, potassium feldspars in orangeites are similar in paragenesis and com-
position to some Fe- and Na-poor sanidines in lamproites. The iron-rich sanidines which
are characteristic of lamproites have not been recognized in orangeites.
Potassium feldspars in the Swartruggens Male lamprophyre are similar to those in
orangeites in having negligible Na20 contents and containing up to 1.0 wt % Fe203. They
differ in that they are irregularly, continuously-zoned to Ba-rich varieties, which may
contain up to 7 wt % BaO (this work).

2.10. ILMENITE
Macrocrystal magnesian ilmenite is not a characteristic mineral of orangeites (see
1.5). Groundmass ilmenite has been recognized in the Lace, Sover Mine, Besterskraal,
Voorspoed, and Finsch orangeites (this work). In these occurrences, ilmenite occurs as
small subhedral plates «25 /lm) and subhedral-to-euhedral (25 x 1-5 /lm) laths. The
238 CHAPTER 2

Table 2.44. Representative Compositions of llmeniteQ


Wt% 2 3 4 5 6 7 8 9
Ti0 2 53.10 52.27 53.41 SO.56 52.88 51.77 52.26 49.46 49.72
AI 20 3 n.d. n.d. n.d. n.d. n.d. n.d. n.d. 1.22 n.d.
Cr203 n.d. n.d. n.d. n.d. 0.23 0.34 0.41 0.52 0.43
FeOT 36.75 35.47 33.55 34.04 29.94 28.87 28.04 40.86 38.45
MnO 8.38 9.59 10.46 9.52 16.95 17.44 18.32 4.40 7.22
MgO n.d. n.d. n.d. n.d. n.d. 0.35 n.d. 1.67 2.98
Nb20 S 1.77 3.07 2.57 5.49 n.d. 1.22 0.90 1.86 1.20
--
100.00 100.40 99.99 l00.D1 100.00 99.99 99.93 99.99 100.00

Structural formulae based on three oxygens


Ti 1.001 0.986 1.003 0.963 1.001 0.981 0.990 0.993 0.940
AI 0.036
Cr 0.005 0.007 0.008 0.010 0.009
Fe2+ 0.770 0.736 0.701 0.715 0.628 0.608 0.591 0.857 0.808
Mn 0.178 0.204 0.221 0.203 0.362 0.372 0.391 0.098 0.154
Mg 0.013 0.062 0.112
Nb 0.020 0.D35 0.029 0.062 0.014 0.010 0.021 0.014

Mol % end-member molecules


AI 20 3 1.8
Cr203 0.2 0.3 0.4 0.5 0.4
Nb20 S 1.0 1.7 1.4 3.1 0.7 0.5 1.1 0.7
Hematite 1.9 5.3
Pyrophanite 17.6 20.3 21.7 20.4 36.0 37.6 39.1 9.5 15.3
Geikielite 1.3 6.3 11.1
Ilmenite 81.4 78.0 76.8 76.5 63.8 60.1 60.0 78.8 67.2
aFeOr = total Fe expressed as FeO; n.d. = not detected. Compositions 1-4. Sover Mine (this work); 5-9, Lace (this work).
Composition 8 contains 2.0 wt % Fe:!OJ, and 39.1 wt % FeO. and composition 9 contains 5.6 wt % Fe20J and 33.4 wt % FeO
when recalculated on a stoichiometric basis.

irregular habit of the crystals suggests that original euhedral ilmenite laths became
unstable subsequent to their crystallization and were resorbed during the later stages of
crystallization of their parent magma. I1menites are not homogeneously distributed and
not preferentially associated with any other groundmass phase. They may be found in
contact with spinel and/or perovskite and included in apatite. The majority occur as
discrete crystals in a carbonate mesostasis. Groundmass ilmenite does not occur in the
Swartruggens orangeite or lamprophyre dikes.
Although few data are available, it appears that ilmenites from orangeites exhibit
significant intra- and inter-intrusion compositional variation (Table 2.44). I1menites from
one sample from Lace contain <0.1-3 wt % MgO, 1.2-1.9 wt % Nb20S, and 2.6-7.2 wt %
MnO, whereas those in another contain <0.1-0.6 wt % MgO, 0.5-1.3 wt % Nb20S, and
17-19 wt % MnO (this work). I1menites from Sover Mine contain <0.1 wt % MgO and
8.4-10.2 wt % MnO and are relatively enriched in Nb20s (1.8-5.5 wt %; this work).
MINERALOGY OF ORANGElTES

t4;!~~t;] CARBONATITES

KIM8ERLITE
GROUNOMASS
AND
MEGACRYST

90

MnTi0 3 60 10 FeTi0 3

Figure 2.104. Compositions (mol %) of ilmenites from the Lace and Sover Mine orangeites (this work) plotted
in the ternary system geikielite (MgTi03Hlmenite (FeTi03)-pyrophanite (MnTi03). (*) = Mg-rich ilmenites
from Lace. Compositional fields for ilmenites from kimberlites (Mitchell 1986, Shee 1984), for carbonatites
from Gaspar and Wyllie (1984) and Mitchell (l978b). Field P, ilmenites from the Premier calcite kimberlite
dike (Gaspar and Wyllie 1984). Field R, outlined by dashed lines, from Wyatt (1979) for Premier kimberlites.
Points joined by arrow are Mg-rich core and Mn-rich margin of an ilmenite from the Premier calcite kimberlite
dike.

All the ilmenites contain negligible Ah03 «0.1 wt %) and have low Cr203 contents
«0.1 wt % at Sover Mine; 0.2-0.6 wt % at Lace).
Table 2.44 and Figure 2.104 indicate that the ilmenites are primarily members of the
pyrophanite (MnTi03Hlmenite (FeTi03) solid solution series. Attempts to calculate
ferric iron contents on the basis of stoichiometry (Droop 1987) demonstrate that the
majority of the ilmenites analyzed do not contain ferric iron and hence show no solid
solution toward hematite. Maximum calculated Fe203 contents reach 5.6 wt % for an
ilmenite from Lace. Because of the paucity of MgO, the majority of ilmenites examined
exhibit no solid solution toward geikielite (MgTi03). The maximum geikielite content
found was 11 mol % for a sample from Lace.
Manganese-rich ilmenites are not typically associated with either kimberlites or
lamproites (see below), although they are commonly found in carbonatites (Figure 2.104).
240 CHAPTER 2

2.10.1. Comparison with Groundmass IImenites from Kimberlites


Groundmass ilmenite typically forms anhedral crystals. These are intergrown with
spinels and/or perovskites, or occur as inclusions in olivines (Mitchell 1986, Shee 1984).
The majority of groundmass ilmenites are probably microcrysts (see 1.2). Euhedral
prisms are very rare and have been reported only from the Liqhobong (Boctor and Boyd
1980) and Lepelaneng (Haggerty 1975) kimberlites.
The few compositional data available (Shee 1985, 1984, Apter et al. 1984, Agee et al.
1982, Pasteris 1980, Boctor and Boyd 1980) indicate that kimberlite groundmass il-
menites are members of the ilmenite-geikielite series (50-90 mol % MgTi03) containing
5-10 mol % hematite (Figure 2.105). Cr203 contents are typically high (1-7 wt %), and
MnO contents are characteristically low «1 wt % MnO).
Mn-bearing ilmenites are known from some kimberlites (Figure 2.104), and Mitchell
(1986) has described a "manganese-enrichment" trend characterized by an increase in
MnO (1-5 wt %) at the margins of Mg-ilmenite megacrysts. This paragenesis is clearly
different from that of primary groundmass ilmenite and represents an attempt by
megacrysts to equilibrate with carbonate-rich late-stage groundmass fluids (Haggerty
et al. 1979, Boctor and Meyer 1979). Extreme examples of this trend are known from the
Premier kimberlite where megacryst ilmenites are strongly, continuously zoned from
MgO-rich cores to MnO-rich (up to 9 wt %) margins (Wyatt 1979, Gaspar and Wyllie
1984). Ilmenites in the late-stage carbonate-rich dikes at Premier (Gaspar and Wyllie
1984) contain similar amounts ofMnO (12-18 wt %) to those of groundmass ilmenites
in orangeites, but may be easily distinguished from the latter on the basis of their higher
MgO (1-5 wt %) and Fe203 (2-5 wt %) contents.
Pasteris (1980) has reported low MgO (1.2 wt %), MnO-rich ilmenite (21 wt %)
coexisting with other MnO-bearing (2.5-5.5 wt %) ilmenites, forming complex inter-
growths with Mg-ilmenite, spinel, and perovskite in the De Beers kimberlite. This
occurrence is not typical or characteristic of kimberlites.
The Nb contents of ground mass ilmenites in kimberlite are inadequately known.
Ilmenite in the Premier calcite kimberlite contains 1.2-2.8 wt% Nb20S (Gaspar and
Wyllie 1984).

Figure 2.10S. Compositional fields (mol %) of i1menites from orangeites [0] (this work). lamproites [L]
(Mitchell and Bergman 1991). and groundmass ilmenites from kimberlites [K] (Shee 1984). plotted in the
ternary system hematite (Fe203)-ilmenite (FeTi03)-geikielite (MgTi03).
MINERALOGY OF ORANGEITES 241

In summary, Mn-bearing ilmenites originating from kimberlites or orangeites may


be distinguished from each other on the basis of their MgO, Cr203, Fe203, and MnO
content and, to a lesser degree, on their paragenesis.

2.10.2. Comparison with I1menites in Lamproites


Ilmenite is not a characteristic mineral of most lamproites. It is known from the
Jumilla (Murcia-Almeria), Sisco (Corsica), and Oscar, Mount North, Rice Hill, and other
West Kimberley intrusions, where it occurs in the groundmass as subhedral prisms.
Ilmenite is apparently absent from extrusive lamproites (Mitchell and Bergman 1991).
Lamproite ilmenites have significant MgO (1-7.5 wt %) and relatively low MnO
(0.5-2 wt %) contents and are FeO-rich members of the ilmenite-geikilite series (>70
mol % FeTi03; Figure 2.1 05). Recalculation of the compositions shows that the ilmenites
are poor in Fe203 «0.1-9 wt %) and typically contain less than 15 mol % hematite
(Figure 2.1 05). Cr203 and Ah03 are typically less than 0.5 wt %.
The only lamproite-derived ilmenites known to contain significant amounts ofMnO
(3-8 wt %) are found in carbonate in olivine lamproites from Argyle (West Australia).
These are MgO poor «1 wt %) and contain up to 1.2 wt % Nb205 (Jaques et al. 1989a).
In summary, lamproite-derived ilmenites may be distinguished from those in
orangeites on the basis of their MgO and MnO contents (Figures 2.104 and 2.105). In
both rock types, ilmenite occurs as partially resorbed late-stage groundmass prisms.

2.11. RUTILE
Rutile is present as a trace accessory groundmass mineral in orangeites. It has been
recognized in the Lace, Sover Mine, Bellsbank Southern Extension, and Swartruggens
orangeites (this work). At Lace it occurs as small «50 !lm) anhedral crystals, which may
be present as single grains, commonly with a "doughnut-like" habit, or intimately
intergrown with anhedral ilmenite, monazite, or barite. At Bellsbank, rutile typically
occurs as irregular intergrowths with anhedral apatite and monazite. Discrete rounded
crystals occur at Sover Mine. Rutile is rare in the Swartruggens orangeites and occurs as
irregular grains and/or very thin laths «1 lim) in segregations consisting principally of
an unidentified Ca-Zr silicate, quartz, and apatite.
Table 2.45 shows that the rutiles contain significant amounts ofFe203 (0.5-1.2 wt %
at Lace, 0.9-1.6 wt % at SoverMine, 1.8-4 wt % at Bellsbank) and Nb20s (1.1-1.9 wt %

Table 2.45. Representative Compositions of Rutile a


Wt% 2 3 4 5 6 7
Nb20 S 3.17 4.57 5.35 1.09 1.44 1.69 3.09
Ti0 2 95.20 94.44 93.33 97.67 97.28 97.40 91.91
Fe203 1.29 1.09 1.20 0.92 1.17 1.01 4.39
MnO 0.48 n.d 0.25 0.41 0.23 n.d. n.d.

100.14 100.10 100.13 100.09 100.12 100.10 99.39


"Total Fe expressed as Fez03; n.d. = not detected. Compositions 1-3. Sover Mine; 4-6 Lace; 7, Bellsbank
(this work).
242 CHAPTER 2

at Lace, 3.2-5.4 wt % at Sover Mine, 3-4.9 wt % at Bellsbank) in solid solution (this


work). Note that coexisting groundmass i1menites at Sover Mine are also enriched in Nb
relative to i1menites from Lace. Chromium contents are very low «0.1 wt % Cr203).

Comparison with Rutile in Kimberlites and Lamproites


Rutile occurs in six parageneses in kimberlites as (1) associated with microphe-
nocrystal olivines, (2) a primary groundmass phase, (3) reaction mantles upon perovskite,
(4) large discrete crystals (?macrocrysts) mantled by Mg-ilmenite, (5) graphic or sym-
plectic rutile-silicate intergrowths, (6) lamellar or sigmoidal rutile-ilmenite intergrowths.
Parageneses 5 and 6 are not found in orangeites and hence not discussed further in this
work.
Needle-like and tabular crystals associated with olivine (Mitchell 1986, Skinner
1989, Pasteris 1980) are the most common paragenesis of rutile encountered in kimber-
lites. These may occur as inclusions confined to the margins of olivine microphenocrysts
or be present as numerous small crystals in the groundmass adjacent to such olivines. In
the latter case, rutile is commonly associated with spinels and/or perovskites in a
"necklace-like" texture. These rutiles are high Cr203 (up to 2.5 wt %), low Fe203 «0.5
wt %) varieties. Nb contents are unknown but may be appreciable, i.e., 1-5 wt % Nb20s
(Pasteris 1980).
Skinner (1989) initially noted that there are no counterparts in texture and composi-
tion to paragenesis 1 rutile in orangeites. Although chromite occurs as inclusions in
orangeite olivine microphenocrysts, rutile, perovskite, and ilmenite are absent.
Primary groundmass rutile is known from the lower Benfontein sill (Mitchell 1994b,
Boctor and Boyd 1981). Here it occurs as discrete twinned euhedral (1 (}-50 11m) crystals
and anhedral-to-subhedral crystals intergrown with the corroded Ti-Mg magnetite or
Ti-magnetite cores of atoll-textured spinels. These intergrowths are mantled by Mn-bearing,
Mg-poor ilmenite, and magnetite. Both varieties of rutile may be replaced by parisite,
ancylite, and ferroan dolomite. The rutile contains up to 3.3 wt % Nb20S and 1-3 wt %
Fe203 (Mitchell 1994b). Orangeites do not contain rutiles belonging to this paragenesis.
Paragenesis 3 has not yet been observed in orangeites. However, in kimberlites
complete replacement of perovskite leads to the development of discontinuous doughnut-like
rings of rutile which are ultimately disaggregated and strewn throughout the groundmass.
Such a process may account for the irregular rutile grains found at Lace. However, the
absence of precursor perovskite':"ilmenite composite grains suggests this rutile may have
other origins.
Anhedral fragments of paragenesis 4 rutile may occur in the groundmass of kimber-
lites. These may be distinguished from groundmass rutiles by their characteristic mantles
ofMg-ilmenite, perovskite, and/or spinel. Anhedral rutile crystals in orangeites lack such
mantles and are unlikely to be of similar origin.
Few data are available regarding the paragenesis and composition of rutile in
lamproites, and the majority of lamproites studied lack rutile (Mitchell and Bergman
1991). Rutile is relatively common only in the anomalous Sisimiut lamproites. From
these, Thy et at. (1987) describe rutile as a groundmass phase intergrown with ilmenite
in amphibole lamproites and as the only opaque groundmass phase in leucite lamproites.
MINERAWGY OF ORANGEITES 243

Rutile from Sisimuit contains up to 2.2 wt % Cr203 and from 0.1-11 wt % Fe203
(Thy et al. 1987, Scott 1981). The Nb content of lamproite rutile is not known.
Secondary anatase is common in some lamproites but has not yet been recognized
in orangeites.
In summary, rutile found in orangeites may be distinguished from that found in
kimberlite on the basis of its paragenesis and Cr and Nb content. Rutile is not a
characteristic or common mineral in lamproites.

2.12. ZIRCONIUM SILICATES


Zirconium silicates are relatively common as late-stage groundmass minerals in
orangeites. Unfortunately, they have not been extensively studied and few paragenetic
and compositional data are available.

2.12.1. Zircon
Zircon occurs in the Swartruggens, Lace, and Bellsbank Southern Extension
orangeites as very small «10 ~m) irregular crystals set in a carbonate mesostasis (this
work). At Lace, zircon occurs rarely as small patches (20-50 I!m) of very small «5I!m)
euhedral, flow-aligned, prismatic crystals associated with subhedral wadeite crystals.
Compositional data are not available.

2.12.2. Wadeite
A potassium zirconium silicate, provisionally considered to be wadeite (K2ZrSh09),
has been identified by BSEJEDS methods in the Swartruggens and Lace orangeites (this
work). At Swartruggens it occurs rarely as euhedral crystals (50 I!m) in orangeites from
level 6 in the mine. These and associated unidentified Ca-Ti-Fe silicate (2.14), strontian-
ite, and ancylite are set in a matrix of calcite. Small euhedral crystals of perovskite are
poikilitically enclosed by the wadeite. At Lace, wadeite forms small subhedral crystals
associated with zircon.
Table 2.46 gives representative compositions (WDS-EMPA) of Swartruggens
wadeite and demonstrates that they are not significantly different from those of wadeites
occurring in lamproites. Wadeites from both parageneses are essentially pure potassium
zirconium silicates. With the exception of Ti02, other elements are not present in
significant quantities. These data are considered to indicate that the K-Zr silicate from
Swartruggens is probably wadeite. However, without X-ray diffraction data, the presence
of kostylevite (monoclinic K2ZrSh09 . H20; Khomyakov et al. 1983a) or umbite (or-
thorhombic KzZrSh09 . H20; Khomyakov 1983b) cannot be ruled out.
Wadeite is considered by Mitchell and Bergman (1991) to be one of the characteristic
minerals of lamproites. Although wadeite is not ubiquitous in lamproites, it is easily
recognized by standard petrographic methods as it commonly forms large crystals
occurring in significant modal amounts.
Wadeite is a rare mineral that, with the exception of lamproites, has been previously
recognized only from two other parageneses, i.e., late-stage carbonatite-like rocks from
the Khibina (Tikhonenkov et al. 1960) and Kovdor (Kapustin 1980) complexes and in
244 CHAPTER 2

Table 2.46. Compositions of Zirconium Silicatesa


Wt% 2 3 4 5 6 7 8 9 10
Si02 46.41 46.92 46.08 46.36 45.91 23.44 28.42 15.8 39.1 40.7
Ti02 0.72 1.22 1.60 1.03 0.50 13.52 25.46 11.9 n.d. n.d.
Zr0 2 31.43 31.19 31.27 29.62 32.77 14.40 n.d. 24.2 29.4 28.8
Al20 3 n.d. n.d. n.d. n.d. n.d. n.d. 0.37 3.6 n.d. n.d.
Fez03 n.d. n.d. n.d. n.d. n.d. 14.14 5.98 14.8 n.d. n.d.
FeO 0.41 0.36 n.d. n.d. 0.04 n.d. n.d. n.d. 4.5 3.9
MnO 0.02 0.02 n.a. n.a. n.d. 0.10 0.83 n.a. n.d. n.d.
MgO 0.05 0.02 0.01 n.a. n.d. 2.35 3.79 n.a. 10.0 10.9
CaO n.d n.d. 0.08 0.47 n.d. 30.08 34.43 30.6 16.4 15.1
NazO 0.09 0.14 n.d. 1.28 n.d. 1.06 0.95 n.d. n.d. n.d.
KzO 21.48 20.81 21.46 20.99 21.90 n.d. n.d. n.d. n.d. n.d.
BaO n.d. 0.38 0.08 n.d. n.d. n.a. n.a. n.a. n.d. n.d.
100.61 101.06 100.58 99.75 101.12 99.09 100.23 100.09 99.4 99.4
"Total Fe expressed as Fel03 or FeO; n.d. =not detected; n.a.= not analyzed. Compositions 1-5. wadeite; 1-2. Swartruggens
orangeite (this work); 3. West Kimberley larnproite (Jacques et al. 1986); 4. Kovdor carbonatite (Kapustin 1980); 5. Little
Murun tausonite syenite (Mitchell and Vladykin 1993); 6. Kimzeyitic gamet. New Elands (Mitchell and Meyer 1989a); 7.
schorlomitic garnet. Burns (Mitchell and Meyer 1989a); 8, Kimzeyitic garnet. Wesselton (Mitchell I 994b); 9-10. unidentified
Ca-Zr silicate. Skietkop (this work).

tausonite syenites from the Little Murun uItrapotassic complex (Mitchell and Vladykin
1993).

2.12.3. Zirconium-Bearing Gamet


Zirconium-rich garnets occur as brown, small (l00-500 J.lm) euhedral isotropic
crystals in the groundmass of the New Elands orangeites (Mitchell and Meyer 1989a).
The garnets lack reaction rims and resorption features and are thus considered to be
primary phases which have crystallized in eqUilibrium with the calcite-rich groundmass.
Table 2.46 shows the garnets to have high Ti02 and Fe203 and low Ah03 contents,
suggesting they are members of a solid solution series between schorlornite and kimzey-
ite. Ti-rich garnets with very low Zr02 contents found in the Bums orangeite (Mitchell
and Meyer 1989a) represent the schorlornitic end member (Table 2.46). Schorlomitic
garnets containing up to 32 wt % Ti02 have been reported in calcite-serpentine segrega-
tions in a Postmasburg area orangeite (Tainton 1992). Schorlomitic garnets are rarely
found at Newlands (Tainton 1992).
Kimzeyitic garnets are not characteristic minerals of archetypal kimberlites (Mitchell
1986) and have never been found in lamproites (Mitchell and Bergman 1991). The only
reported occurrence of Ti-Zr-rich garnet in a bona fide kimberlite is from a highly
differentiated calcite kimberlite siIl at WesseIton (Mitchell 1994b). This garnet occurs as
small (30-50 J.lm) anhedral groundmass crystals in samples lacking baddeleyite and
calcium zirconate. The garnet is richer in Zr02 than the Zr-bearing garnets from the New
Elands orangeite (Table 2.46).
Zr- and Ti-rich garnets are common in alnoites and ultramafic lamprophyres associ-
ated with carbonatite complexes (Platt and Mitchell 1979, Mitchell 1983). A very wide
range of compositions has been reported within the series andradite-schorlornite-
MINERALOGY OF ORANGEITES 245

kirnzeyite. These encompass the compositions described above. Kirnzeyitic garnets from
this paragenesis commonly coexist with monticellite and melilite.

2.12.4. Calcium Zirconium Silicate


An unidentified calcium zirconium silicate occurs rarely in the groundmass of the
Swartruggens and Skietkop orangeites. At Swartruggens it forms ragged, irregular, small
(1-5 !lm) crystals intergrown with rutile and apatite, set in a matrix of quartz. Composi-
tional data are not available.
In the Skietkop occurrence the mineral forms granular aggregates of small (10-50
!lm) anhedral crystals within a calcite matrix. Semiquantitative analyses indicate 39.1-
40.7 wt % Si02, 28.8-30.3 wt % Zr02, 15.1-16.4 wt % CaO, 9.1-10.9 wt % MgO, and
3.7-4.5 wt % FeOT. The mineral contains insufficient CaO and Si02 to be either Zr-Ti
garnet or calcium catapleite, respectively.

2.13. CARBONATES
Carbonates are common late-stage groundmass minerals in orangeites with modes
ranging from <1-13 vol % (Skinner and Clement 1979, Fraser 1987). Detailed investi-
gations of their composition and paragenesis have not yet been undertaken.

2.13.1. Calcite
Calcite is the commonest carbonate found in orangeite. It typically occurs as anhedral
patches of interlocking crystals intimately intergrown with groundmass serpentine. Less
commonly, it forms the margins of calcite-serpentine segregations. Calcite may replace
serpentinized olivine macrocrysts and phlogopite. Despite the ubiquity of calcite, com-
positional data have not been reported in previous studies of orangeite. Limited data
obtained during this study show that it is typically pure CaC03 containing less than 0.2
wt % MgO, FeO, SrO, and MnO.

2.13.2. Dolomite
Dolomite is a relatively common groundmass constituent of the Bellsbank and
Newlands orangeites (this work, Tainton 1992). The groundmass of the Main and
Bobbejaan dikes consists of a serpentine-dolomite intergrowth. The dolomite forms
poikilitic grains (0.6 mm). With increasing degrees of deuteric alteration, dolomite/ser-
pentine ratios increase until all of the serpentine is replaced (Tainton 1992). In the
Bobbejaan dike the matrix consists entirely of dolomite. Similar dolomite/serpentine
relationships have been observed at Newlands, where fresh samples contain more
serpentine than dolomite. The abundance of dolomite increases with increasing degrees
of deuteric alteration (Tainton 1992). At Newlands, dolomite crystallization occurs before
calcite (Tainton 1992).
The dolomites of the Bellsbank Southern Extension orangeite contain rounded
patches of (?)exsolved norsethite (this work, see below). The dolomite is an Sr-Mn-bear-
ing ferroan dolomite (Table 2.4 7). No other compositional data for dolomite in orangeites
have been published.
246 CHAPTER 2

Table 2.47. Representative Compositions of Carbonatesa


Wt% 2 3 4 5 6 7 8 9
CaO 1.46 0.69 54.9 52.5 52.8 0.3 6.8 8.9 35.3
srO 1.01 0.71 1.5 2.5 4.8 33.3 16.6 89.9 63.1
FeOr 0.86 0.79 5.3 8.0 7.1 n.d. n.d. 1.2 1.0
MgO 14.21 15.39 36.9 34.4 34.9 n.d. n.d. n.d. n.d.
BaO 56.50 57.00 0.8 2.2 n.d. n.d. n.d. n.d. n.d.
MnO n.d. n.d. 0.7 0.5 0.4 n.d. n.d. n.d. 0.6
La203 n.a. n.a n.a. n.a. n.a. 28.6 22.1 n.d. n.d.
CeZ0 3 n.a. n.a. n.a n.a. n.a. 29.1 41.4 n.d. n.d.
Pr20 3 n.a. n.a. n.a. n.a. n.a. 2.9 2.8 n.d. n.d.
Nd20 3 n.a. n.a. n.a. n.a. n.a. 5.8 10.3 n.d. n.d.
74.04 74.58 100.0 100.0 100.0 100.0 100.0 100.0 100.0
aFeOr =total Fe expressed as FeO; n.d. =not detected. n.a. =not analyzed. Compositions 1-2 obtained by EMPAlWDS
analysis, 3-9 by SEMIEDS methods and expressed as oxides summed to 100 wt % on a C02-free basis. Compositions 1-2,
norsethite, Bellsbank; 3-5, ferroan dolomite, Bellsbank; 6-7, ancylite, Bellsbank, and Swartruggens, respectively; 8-9,
strontianite and calcian strontianite, Swartruggens. (All data this work.)

2.13.3. Other Carbonates


Other carbonates present in orangeites include strontianite (SrC03), barian strontian-
ite, witherite (BaC03), unidentified Ba-Ca carbonate (?barytocalcite), an Sr-rare earth
carbonate (Table 2.47), considered to be ancylite [Sr(REE)(C03h(OH)· H20], and
norsethite [BaMg(C03h].
The majority of these carbonates occur as irregular crystals within the groundmass
and are intimately intergrown with calcite, dolomite, serpentine, and barite. Commonly,
they replace earlier crystallizing minerals, e.g., barian strontianite replaces apatite at Lace;
witherite replaces sanidine at Sover North and calcite at Sover Mine; ancylite replaces
perovskite at Makganyene and Swartruggens. This petrographic evidence suggests that
the majority of these carbonates are late-stage hydrothermalldeuteric minerals. Norsethite
(Table 2.47) is common in the Bellsbank Southern Extension orangeite as rounded
droplets within a ferroan dolomite matrix. The norsethite is considered to have exsolved
from a complex dolomitic carbonate precursor as it is only found as inclusions within
ferro an dolomite (this work).
Carbonates are major components of archetypal kimberlites. The parageneses of
calcite and dolomite in these rocks (Mitchell 1986) are similar to those of orangeite. The
principle differences are that kimberlites contain more carbonate, and calcite-serpentine
segregations are common. Insufficient compositional data are available to permit useful
comparisons of carbonate compositions in the two rock types. Surprisingly, detailed
studies of the carbonate assemblages in kimberlites have not been undertaken; thus, it is
not known whether or not Sr-, Ba-, and REE-rich carbonates are common. Ancylite and
Ca-Ba carbonate have been reported in highly-evolved calcite kimberlites from the
Benfontein sills (Mitchell 1994b). Preliminary studies of Somerset Island kimberlites
have not revealed the presence of any Sr-Ba-REE carbonates.
MINERALOGY OF ORANGElTES 247

Mi tchell and Bergman (1991) have noted that carbonates are not characteristic
groundmass minerals of lamproites. Groundmass calcite, when present, appears to result
from post-emplacement secondary carbonatization. The majority of lamproitic carbon-
ates occur in vesicles. Calcite, witherite, calcian strontianite, and strontian calcite have
been noted in vesicles (Mitchell and Bergman 1991, Wagner and Velde 1986).

2.14. OTHER MINERALS


Barite is a common accessory mineral in most orangeites and occurs as irregularly
shaped aggregates and veins of anhedral crystals throughout the groundmass or replacing
earlier crystallizing minerals. Intergrowths of calcite and barite are common. Barite is
considered to be a late-stage groundmass and deuteric hydrothermal mineral. Qualitative
analysis indicates that many barites contain significant quantities of Sr.
The groundmass of the Swartruggens orangeites contains an unidentified euhedral
purplish brown mineral occurring in matrices which include barite, calcite, and quartz.
The habit is hexagonal, and the mineral exhibits first-order interference colors. The
average (wt %) of four WDS-EMP analyses is 30.3% CaO, 26.2% Si02, 21.4% Ti02,
14.0% FeOr, 1.6% MgO, and 0.65% Na20 (94.5% total oxides). AI, Zr, and Ba were not
detected.
Chalcopyrite and galena have been identified in the Besterskraal, Sover North, Lace,
Bellsbank, and Swartruggens orangeites. Pyrite is common as interstitial late-crystal-
lizing laths and plates at Voorspoed and Swartruggens. Fe-Ni sulfides are present in the
groundmass at Postmasburg, Sover Mine, and Bellsbank. Ni-sulfides are common in
serpentinized olivine macrocrysts in most orangeites. Small «111m) rounded crystals of
FeAs2 (?loellingite) occur rarely in the groundmass of the Sover Mi ne orangeite.
Quartz is a common component of the mesostasis of the Swartruggens and the
Sweetput-Souput orangeites. It is unclear whether the quartz is a late-stage primary phase
or a secondary mineral. Ba- and Na-zeolites, replacing feldspars, occur in the groundmass
of the Besterskraal and Swartruggens rocks.

2.15. SUMMARY
The above mineralogical studies demonstrate conclusively that the rocks here
referred to as orangeites, previously termed "micaceous or group 2 kimberlites," have
little mineralogical affinity with archetypal or "group 1" kimberlites. Notable differences
exist with respect to the nature of the composition and evolutionary trends of mica
compositions. Kimberlite micas are characterized by trends of Al enrichment, leading to
the formation of barian aluminous phlogopite-:kinoshitalite solid solutions. In contrast,
orangeite micas exhibit trends of Al depletion, leading to the formation of tetraferriphlo-
gopite. Orangeites and kimberlites differ with respect to the nature of the primary silicate
phases. Kimberlites crystallize monticellite and serpentine but not diopside, potassic
amphibole, and potassium feldspar. The assemblages and compositional trends of iron
and titanium oxides are different in the two rock types. Spinels rich in the magnesian
ulvospinel molecule appear to be restricted to kimberlites. Spinels in orangeites are very
248 CHAPTER 2

Cr-rich and AI-poor relative to those in kimberlites. Perovskites in kimberlites are poor
in REE and Sr relative to those in lamproites.
Although orangeites have few mineralogical characteristics in common with kim-
berlites, they have much closer affinities to lamproites. Micas follow similar evolutionary
trends toward tetraferriphlogopite. The assemblages differ in that Ti-rich micas are not
preSent in orangeites. Primary diopside, potassic amphibole, and spinel group minerals
haVie similar compositions and evolutionary trends in both rock types. Orangeites and
larriproites share the characteristic of containing primary AI-poor silicates which com-
monly contain insufficient Al to occupy all of the available tetrahedral lattice sites in the
crystal structure. This evidence suggests that orangeites also crystallized from peralkaline
parental magmas. Lamproites and orangeites differ primarily in the nature and composi-
tion of the accessory mineral assemblage. Hollandites in lamproites are principally
Cr-bearing K-Ba septetitanates, whereas those in orangeites are V-bearing Ba-rich
hexatitanates. K-triskaidecatitanates do not occur in lamproites. Orangeites contain a
greater variety of Sr-, Ba-, and REE-bearing phosphates than most lamproites. The
affinities with lamproites are discussed further in Section 4.7.2.
Get your facts first, and then you can distort 'em as much
as you please.
Mark Twain

GEOCHEMISTRY OF ORANGEITES

Most previous studies of the geochemistry of orangeites have been undertaken and
interpreted on the premise that they are merely mica-rich varieties of kimberlite. Thus,
compilations of the average trace element abundances in kimberlites, made by Wedepohl
and Muramatsu (1979), Dawson (1980), Muramatsu (1983), and Mitchell (1986), are
based upon combined data for both kimberlites and orangeites. Given the conclusion that
kimberlites and orangeites are members of different petrological clans, such averages
must now be seen as incorrect. Similarly, comparative discussions of the geochemistry
of ''micaceous kimberlites" and "mica-poor kimberlites," as presented by Mitchell (1986),
Fesq et al. (1975), and Kable et al. (1975), must now be viewed in a different light.
The earliest studies of the geochemistry of kimberlites and orangeites were restricted
to establishing differences in their major element composition. These studies detennined
that "the only apparent difference in composition between micaceous and basaltic
kimberlites is that the former are somewhat richer in potassium oxide and alumina, and
somewhat poorer in magnesia, as one would expect from their mineralogical constitution"
(Wagner 1914, pp. 110-111). This conclusion has been reiterate~ in many subsequent
studies, e.g., Dawson (1971, 1980), which have also emphasized the volatile-rich nature
of both varieties of rocks and the high content of Cr and Ni coupled with high REB, Sr,
and Ba contents. Tainton (1992) has provided the most recent significant discussion of
the major element geochemistry ofkimberlites with respect to inter-element variation and
the problems of alteration and contamination.
Detailed investigations of the trace element geochemistry of orangeites were not
undertaken until the early 1970s. The studies by Fesq et al. (1975), Kable et al. (1975),
Mitchell and Brunfelt (1975) and Gurney and Berg (1969) established that, compared to
kimberlites, orangeites were richer in many incompatible elements and exhibited very
highly fractionated REE distribution patterns. One important conclusion of the work of
Kable et al. (1975), based upon NblP ratios, was the first suggestion that kimberlites and
orangeites must be derived from mineralogically different mantle sources. A major
conclusion ofthe initial geochemical studies was that orangeites and kimberlites represent
mixtures of an incompatible element-rich melt with mantle-derived xenocrysts.
Subsequent to the isotopic studies (Smith 1983) which established "micaceous
kimberlites" as a group of rocks with different origins from other "kimberlites," Smith et
al. (1 985b ) proposed a new geochemical classification of "kimberlites" based upon
249
250 CHAPTER 3

isotopic criteria. The major and trace element compositions of the isotopically defined
groups were shown by multivariate statistical methods to be significantly different.
Isotopic group II kimberlites (orangeites) were noted to have higher abundances of Si02,
K20, Pb, Rb, Ba, and light REE and lower abundances ofTi02 and Nb than isotopic group
I rocks (archetypal kimberlites). This conclusion was predictable from general geochemi-
cal principles and the observed mineralogy. Smith et al.'s (1985b) study was especially
valuable in that a wide spectrum of trace element abundances was obtained on a suite of
hypabyssal facies rocks, carefully selected as being contamination-free and minimally
altered.
Smith et al. (1985b) further divided isotopic group I kimberlites into "on craton"
(IA) and "off-craton" (IB) subgroups; the latter were considered to be relatively enriched
in Ti, P, Nb, Zr, Y. Unfortunately, only 33 samples were analyzed and the value of
undertaking extensive multivariate statistical analysis on such a small sample is question-
able. Not surprisingly, the IA-IB subdivision is not supported by the more extensive data
set now available (see below).
The most significant advance resulting from the Smith (1983) and Smith et al.
(1985b) studies was the conclusion that isotopic group II rocks originated from li-
thospheric sources, whereas isotopic group I rocks were derived from asthenospheric-like
sources. Subsequent isotopic studies ofkimberlites and orangeites have confirmed these
original observations (see 3.8).
Recently, the detailed studies of the Finsch (Fraser 1987, Fraser and Hawkesworth
1992) and Barkly West orangeites (Tainton 1992) have substantially increased the number
of major, trace, and isotopic analyses of well-characterized suites of samples. These data
sets are extremely valuable for assessing the geochemical variation within and between
orangeite intrusions.
Despite the advances noted, our knowledge of the geochemistry of orangeites is still
limited. Although new data are presented in this work for orangeites from Swartruggens,
New Elands, and Star, there remains a paucity of data for orangeites from the Swartrug-
gens, Kroonstad, Winburg, Boshof, and Prieksa areas. Determinations of the trace element
and isotopic composition of the minerals comprising orangeites have not yet been
undertaken.
This chapter summarizes the basic features of the major and trace element and isotope
geochemistry of orangeites and provides comparisons with that ofkimberlites and olivine
lamproites. The geochemistry of orangeites is not, compared with that of the mica-rich
rocks occurring in West Greenland, termed "micaceous kimberlites" (Scott 1981, Larsen
and Rex 1992). This omission is intentional, as the rocks are not orangeites and are more
probably related to ultramafic lamprophyres than kimberlites.

3.1. CONTAMINATION AND ALTERATION


Investigation of the geochemistry of orangeites is subject to all of the contamination
and alteration problems discussed by Mitchell (1986) and Clement (1982) with regard to
kimberlites.
Because of the extremely high contents of Mg, Ni, Cr, Sr, Ba, REE, Zr, and Nb
relative to their abundance in common crustal rocks, it is now considered that crustal
GEOCHEMISTRY OF ORANGEITES 251

contamination cannot be responsible for either the observed enrichment in these elements
or the compositional variation within and between intrusions (Tainton 1992, Fraser 1987,
Mitchell 1986). Tainton (1992) has also noted that the ZrlY (>10) and NblY (>4) ratios
of orangeites are far in excess of those of possible crustal contaminants (ZrlY < 10, NblY
< 3). Hence, incompatible element ratios may reflect those of their parental magma rather
than mixing with crustal contaminants.
Xenoliths of all types of crustal rocks found in unevolved orangeites either show no
effects of incorporation in the magma or are slightly thermally metamorphosed. The latter
process leads, in limestones and shales, to the formation of concentric zones of recrystal-
lization and slight metasomatism, recognizable by color banding. Shale and basalt
xenoliths immersed in evolved orangeites commonly undergo extensive reaction and are
replaced by phlogopite, amphibole, and sanidine. The formation of these minerals
represents the response of the magma to xenolith assimilation by the precipitation of only
the current or possible liquidus phases.
Various contamination indices have been devised to estimate the degree of contami-
nation (or alteration) ofkimberlites (Ilupin and Lutz 1971, Clement 1982) and may be
directly applied to orangeites. Of these, the most widely used is Clement's contamination
index (CI), which is defined as the ratio of (Si02 + Ah03 + Na20)/(MgO + K20) wt%.
For kimberlites, contamination indices close to unity are considered to indicate uncon-
taminated or fresh kimberlite. Clement (1982) has noted that many apparently contami-
nation-free orangeites have CIs of 1-5. These indices are not a result of contamination
but a consequence of high modal contents of phlogopite. CIs for contaminated orangeites
are considered more likely to be greater than 1.5. However, CIs for apparently uncon-
taminated orangeites analyzed by Dawson (1987) range from 1.5 to 2.6.
Alteration of orangeites leads to the formation of chlorite and clay minerals from
micas, serpentine from olivine, and the replacement of pyroxene and apatite by carbon-
ates. Spinels and titanates may remain as residual minerals. The geochemical effects of
alteration include loss of mobile alkali elements (K, Rb, Cs), decrease in the Mg/Si ratio,
and increase in Ca content. The overall result is an increase in the contamination index.
Although altered rocks may have undergone significant changes in major element
composition, the trace element signature of the more immobile elements may not have
been substantially altered. Thus, strongly-altered rocks with high Ni and Cr contents
coupled with high Ba, Sr, and REE might still be recognized as orangeite, kimberlite, or
lamproite.
Geochemical characterization of altered rocks is an important component of many
exploration programs for diamond-bearing rocks. Commonly, the only material available
for the identification of a particular exploration target is highly-altered rock obtained by
core drilling. A combination of trace element analysis and identification of relict resistant
accessory minerals might be the only means of classifying such samples (Mitchell 1995).
Diatreme facies rocks are obviously particularly prone to contamination and altera-
tion, and analysis of such material should be avoided. Studies of diatreme facies
kimberlites have clearly shown that it is impossible to rid samples of microscopic crustal
clasts (Dawson 1980). The bulk compositions of all diatreme facies rocks should be
regarded as contaminated.
252 CHAPTER 3

3.2. PRIMARY MAGMA COMPOSITIONS


Fortunately, many orangeites occur as relatively-fresh hypabyssal intrusions, con-
taining few crustal xenoliths. Hence, whole rock compositions can be expected to be
representative of the crystal-phyric magma composition at the time of consolidation.
Petrographic studies (1.10) and major element compositional data (3.3) indicate that
all orangeites have been contaminated with mantle-derived olivine xenocrysts. Rocks
which are relatively xenocryst-free contain transported assemblages of macro crystal and
microphenocrystal phlogopite (2.1.7) in addition to primary groundmass minerals which
have crystallized in situ. The ratios of primary phlogopite to apatite, diopside, and
carbonates vary widely, suggesting that pre- and post-intrusion crystal fractionation (flow
differentiation, filter pressing, crystal settling) plays a significant role in controlling the
final modal assemblage. It is this assemblage plus the amount ofxenocrystal olivine which
controls the bulk composition of the rock.
The observation that the modal proportions of the primary groundmass minerals are
not constant within a given intrusion demonstrates that this matrix cannot represent the
bulk composition of the orangeite parental magma. If the matrix represented a relatively
rapidly-quenched liquid, one would expect to find approximately constant proportions of
primary minerals in all samples. This conclusion implies that if major and trace elements
are subtracted from the whole rock composition in proportion to the amount of xenocrys-
tal olivine present, the result will not approximate the composition of the orangeite
parental magma. The "corrected" compositions can only represent those of random
mixtures of microphenocrysts, ground mass, and mesostasis minerals.
It is concluded from the above observations that the bulk compositions of orangeites
cannot represent those of the parental magma or of magma plus olivine xenocrysts.
Unfortunately, glassy or aphyric margins to hypabyssal intrusions, which would provide
a better estimate of magma compositions, are not found. Consequently, it is suggested
here that we do not know the true compositions of primary or derivative orangeite
magmas. This conclusion has very important consequences regarding attempts to model
orangeite petrogenesis using techniques devised for rocks whose compositions approxi-
mate those ofliquids (Fraser and Hawkesworth 1992, Tainton 1992, Tainton and McKen-
zie 1994, Mitchell and Brunfelt 1975). The problem is discussed further in Sections 3.3.2
and 4.1.

3.3. MAJOR ELEMENT GEOCHEMISTRY


The principal features of the major element geochemistry of orangeite were estab-
lished by Wagner (1914) and reiterated by Dawson (1980,1987). Subsequently, Smith et
al. (1985b) demonstrated that orangeites have compositions statistically distinct from
those of archetypal kimberlites. Detailed geochemical studies of the compositional
variation within individual intrusions have been undertaken by Fraser (1987), Fraser and
Hawkesworth (1992), and Tainton (1992). The latter study, together with that of Skinner
et at. (1994), documented for the first time the geochemistry of relatively silica-rich
orangeites from the BarkIy West and Prieska areas respectively.
GEOCHEMISTRY OF ORANGEITES 253

The rocks referred to in this chapter as "unevolved orangeites," are composed


primarily of phlogopite, apatite, and carbonate; late-stage richterite and sanidine are
absent, and olivine macrocrysts mayor may not be present. Representative examples
include orangeites from Bellsbank, Sover, Swartruggens, New Elands, and Star. "Evolved
orangeites" contain richterite and sanidine; examples are known from Pniel, Postmasburg,
Kroonstad, and the Prieska areas (see 1.10).

3.3.1. Unevolved Orangeites


Tables 3.1 and 3.2 present the average major element composition of unevolved
orangeites from several of the main occurrences. Notable aspects of their composition
are the wide ranges in the concentration of many elements (MgO, CaO, P20S, K20), low
Si02 and Na20 contents, and high MgO, K20, P20S, and volatile contents. As a
consequence of high K20 coupled with low Ah03 and Na20 contents, the majority of the
rocks are peralkaline [molar (K20 + Na20)/Ah03 > 1], ultrapotassic (molar K20/Na20
> 3), perpotassic (molar K20/Ah03 > 1), and ultrabasic.
The majority of the Fe contents have been detennined by X-ray fluorescence
methods, and Fe203/FeO ratios are not well-established. However, these can be expected
to be high given the common occurrence oftetraferriphlogopite in the orangeite ground-
mass. Given the prevalence of late-stage deuteric alteration in many orangeites, it is
extremely unlikely that measured oxidation ratios will reflect the redox conditions of the
parent magmas. Data given by Smith et al. (1985b) indicate that FeO and Fe203 contents
of 16 diverse orangeites range from 2.88 to 5.10 wt% and 2.50 to 5.25 wt%, respectively.
The majority of the samples have Fe203/FeO ratios> 1 (range 0.46-1.71, average 1.16).
Orangeites from Finsch (Clement 1982) contain 3.52-4.66 wt% FeO and 3.18-7.68 wt%
Fe203 and have Fe203/FeO ratios of 0.76-1.81 (average of 7 samples = 1.17). Five
randomly selected orangeites analyzed by Dawson (1987) contain 3.11-4.58 wt% FeO
and 1.94-5.85 wt% Fe203, with Fe203/FeO ratios of 0.50 to 1.88 and averaging 1.14.
Tables 3.1 and 3.2 demonstrate that most orangeites have high and extremely
wide-ranging values ofloss on ignition (LOI), representing primarily the sum of H20 and
C02. Increasing LOI commonly correlates positively with increasing CaO, reflecting the
very high modal calcite and/or dolomite contents of some samples. Actual H20 and C02
contents quoted by Clement (1992), Dawson (1987), and Smith et al. (1985b) vary widely
and depend upon the relative proportions of phyllosilicates to carbonates. Thus, different
rocks from a given intrusion may be rich in H20 or C02. In all instances neither the
volatile content (LOI) nor the H20/C02 ratio reflects that of the parent magma.
The sulfur and fluorine contents have been insufficiently investigated, and no data
have been reported for chlorine or bromine. Dawson (1987) reports the F content of five
orangeites from not detectable to 0.03 wt% F.
Seven determinations of S content (0.Q1-0.12 wt% S, average 0.05 wt% S) for Finsch
orangeites have been given by Clement (1982). Five determinations of S03 reported by
Dawson (1987) range from 0.03 to 0.74 wt% and average 0.37 wt%. The high modal
contents of barite in many orangeites indicates that the parental magmas must contain
significant amounts of S. The paucity of sulfides and the dominance of sulfate suggests
that orangeite magmas are highly oxidizing relative to kimberlites and basaltic magmas.
~

Table 3.1. Average and Range of Major Element Compositions of Representative Orangeites
Wt% Swartruggens Finsch Bellsbank Sover
Si02 36.44 ± 2.98 30.00-40.75 37.53 ±3.13 27.6~1.93 33.02 ± 1.89 27.59-36.00 35.09± 1.72 32.8~.41
Ti02 1.58±0.30 1.28-2.52 0.88 ± 0.20 0.60-1.57 0.74±0.13 0.43-0.97 1.06±0.29 0.48-1.64
AI203 4.02±0.87 2.76-6.03 3.34±0.90 1.62-5.69 1.64 ± 0.38 0.91-2.29 2.55±0.70 1.30-3.96
Fe203 8.15±0.85 6.17-10.30 7.99±0.75 5.68-8.84 7.77±0.37 6.83-8.56 7.78±0.47 7.07-9.21
MnO 0.16 ± 0.03 0.12-022 O.17±0.07 0.13-0.46 0.16 ± 0.24 0.12-0.24 0.15 ±0.03 0.09-0.22
MgO 21.25±4.25 14.80-27.30 28.18±5.09 10.44-33.38 31.40±4.27 20.97-39.49 29.02±4.53 20.94-36.47
Cao 8.39 ± 3.75 2.94-15.26 6.54±4.48 3.29-24.48 6.61 ±2.07 3.49-13.41 6.49 ± 2.46 3.50-11.48
Na20 0.24±0.10 0.03-0.58 0.21 ±O.l8 0.03-0.74 0.12±0.06 0.01-0.25 0.18±0.1O 0.03-0.52
K20 4.65± 1.00 3.4~.72 3.14 ± 0.76 0.81-4.43 1.72±0.55 0.67-3.16 2.91 ± 1.32 1.16-5.79
P20S 1.34 ± 0.44 0.74-2.25 0.61 ±0.19 0.30-1.18 1.41 ±0.52 0.72-3.31 0.68 ± 0.42 0.10-1.62
LOI 1O.45±2.26 6.31-15.00 9.90±3.71 5.28-21.47 12.99±2.14 9.80-18.40 1l.76± 1.42 8.58-14.50
Total 96.67 98.49 97.58 97.67
PI 1.35 Ll2 1.26 1.35
UPf 12.75 9.84 9.43 10.64
PPI 1.25 1.02 1.14 1.24
(N) 18 30 35 31
LOI = loss on ignition. PI. UPI. and PPI are the peraIkalinity, ultrapotassic, and perpotassic indices, respectively. Total Fe is expressed as F~3. (N) = number of samples. Data sources: Swartruggens
(this work, Smith et al. 1985a,b); Finsch'(Fraser 1987, this work); Bellsbank (fainton 1992, Smith et al. 1985a,b, this work); Sover (fainton 1992, this work).

~
~
GEOCHEMISTRY OF ORANGElTES 255

Table 3.2. Average and Range of Major Element Compositions of Representative Orangeites
Wt% Newlands New Elands Star
Si02 33.52 ± 1.85 29.80-36.31 36.64 ± 0.09 35.83-37.90 34.01 ± 1.09 32.10-35.30
TI02 0.62±0.13 0.47-0.91 l.32±0.16 1.08-1.43 1.27+0.18 1.01-1.61
Ah03 1.71 ±0.32 1.29-2.32 4.23±0.27 3.84-4.44 2.79 ± 0.47 2.13-3.66
Fe203 7.36±0.25 6.86-7.71 7.30± 1.45 5.13-8.22 8.59 ± 0.60 7.30-9.13
MnO 0.14±0.02 0.11-0.16 0.31 ±0.20 0.19-0.60 0.26 ±0.14 0.16-0.57
MgO 34.08 ±3.21 28.40-39.84 20.86 ± 5.79 12.30-25.04 25.34±5.66 18.80-31.70
CaO 6.12± 1.64 3.11-8.58 10.18 ± 5.71 6.15-18.65 7.92±2.65 4.90-12.30
Na20 0.11 ±0.05 0.03-0.23 0.19 ±0.05 0.13-0.25 0.17 ±0.1O 0.05-0.35
K20 1.02 ±0.36 0.52-1.65 4.73±0.36 4.27-5.12 2.95±0.27 2.51-3.48
P20S 1.13±0.24 0.65-1.52 1.22±0.18 1.00-1.41 0.82±0.23 0.39-1.16
LOI 12.42 ±2.19 8.71-16.80 11.43±2.86 9.48-15.68 12.89 ± 2.99 8.75-17.20
Total 98.23 98.98 97.01
PI 0.75 1.28 1.24
UPI 6.10 16.38 11.42
PPI 0.65 1.21 1.14
(N) 19 4 8
LO! =loss on ignition. PI, UPI, and PPI are the peralkalinity, ultrapotassic, and perpotassic indices, respectively. Total Fe
expressed as Fe.!03. (N) =number of samples. Data sources: Newlands (Tainton 1992, Smith et al. 1985a, this work); New
Elands (Smith et al. 1985b); Star (this work).

3.3.2. Mineralogical Controls on the Major Element Geochemistry


Petrographic examination reveals that unevolved orangeites are essentially mixtures
of olivine, phlogopite, carbonate (calcite and/or dolomite), and apatite. The majority of
the olivines are macrocrysts derived by the fragmentation of mantle-derived harzburgite
or lherzolite xenoliths. Thus, they may be regarded as contaminants in the magma,
whereas the other minerals are primary phases. Consequently, the bulk major element
compositions of orangeites represent mixing lines between the composition of forsteritic
olivine and orangeite primary phases. The relatively small amounts of primary olivine in
many orangeites (LlO, 4.5.4) do not contribute significantly to the bulk composition.
Tables 3.1 and 3.2 indicate that the P205 contents of orangeites are relatively low; hence,
the bulk compositions of orangeites may be regarded as reflecting modal variations in
olivine macrocrystal, phlogopite, and carbonate content.
Compositional data for unevol ved olivine macrocryst-rich orangeites from Sover and
BelIsbank (Tainton 1992), when plotted in the ternary system MgO-K20-CaO (Figure
3.1), demonstrate clearly that bulk compositions are controlled by mixing of the assem-
blage phlogopite--carbonate with macrocrystal olivine. Different olivine control lines for
each intrusion reflect their differing phlogopitelcarbonate ratios. Figure 3.1 does not
reflect the variations in the calciteldolomite ratio known to occur in these suites of samples
(Tainton 1992). However, the figure indicates that the presence of dolomite will move
bulk compositions to relatively CaO-poor compositions for a given phlogopitelolivine
ratio.
256 CHAPTER 3

MgO

/ 5
+ SOVER
• BELLSBANK
• SWARTRUGGENS
o NEW ELANDS

o
"0
35

40 \

45\

CoO
CT
CoO
CT
Figure 3.1. Compositions (wt%) of orangeites plotted in the ternary system MgO-K20-CaO. The diagram
also shows compositional tie lines for mixtures of olivine (OL). phlogopite (PHL). calcite (Cf). or dolomite
(DOL). Isocompositionallines show the ternary percentages of these minerals. e.g .• the 850L line shows the
varying composition of ternary mixtures containing 85% forsteritic olivine with respect to changing cal-
citelphlogopite ratios. Compositions of olivine and phlogopite used in the calculations are from Table 2.25
(anal. 4C) and Table 2.1 (anal. 10). respectively. Data for Sover and Bellsbank from Tainton (1992);
Swartruggens and New EIands from this work and Smith et al. (1985b).
GEOCHEMISTRY OF ORANGEITES 257

Similar relationships are evident (Figure 3.1) for relatively olivine-poor Swartrug-
gens and New Elands orangeites. Note that several Swartruggens samples have bulk
compositions that plot parallel to the phlogopite-calcite join. This agrees with pet-
rographic observations that the rocks are essentially mixtures of phlogopite and calcite
(see 1.10).
Figure 3.1 is interpreted to show that the bulk compositions of orangeites from Sover
and Bellsbank cannot represent liquid compositions. The compositions found are the
result of mixing of the crystal-laden magma, which formed the groundmass, with
xenocrystal olivines. Addition of xenocrystal olivine implies that lherzolite-derived
orthopyroxene, together with minor clinopyroxene and garnet, must have been added to
the magma, unless the xenocrystal contaminants are entirely derived from dunites.
Orthopyroxene xenocrysts have not been recognized in orangeites, implying that any
lherwlite-derived enstatite must be completely assimilated by the magma either at its
source or during transport. Assimilation of orthopyroxene will raise the silica content of
the hybrid magma. The addition of the amounts of orthopyroxene typically found in
lherwlite xenoliths (20-40 vol%) may lead to the formation of relatively siliceous
evolved orangeites. However, lacking knowledge of the composition, volume, and
temperature of the primary magmas involved, calculation of the potential compositions
of hybrids is fraught with uncertainty.
Orangeites relatively-poorin olivine, such as occur at New Elands and Swartruggens,
might have lost the majority of their load of xenocrystal olivines during periods of
stagnation in the ascent of the magma. Alternatively, the magmas might not have been
extensively contaminated at their sources. Regardless, the complex mica assemblage
present in these rocks and the bulk compositional variation attributable to varying modal
phlogopite/carbonate ratios, demonstrate that they were intruded as crystal-charged
slurries with a minor fluid content.
The above observations suggest that the composition of the parental orangeite
magma cannot be determined by the simple subtraction of olivine from the measured bulk
compositions, as the proportions of phlogopite, apatite, and carbonate vary widely
because of pre- and post-intrusion crystal fractionation. This conclusion has important
implications regarding the interpretation of the trace element geochemistry of orangeite.
For example, Fraser (1987) and Fraser and Hawkesworth (1992) have concluded that the
variations in incompatible element abundances in the Finsch orangeites merely reflect
the results of mixing varying amounts of peridotite contaminant with an unfractionated
trace element-rich magma. However, early crystallizing apatite and diopside are common
in many orangeites, and fractional crystallization processes involving these phases may
playa role in determining the abundances of Sr and the REE abundances.

3.3.3. Evolved Orangeites


Table 3.3 presents the averages and ranges of composition of evolved orangeites from
Sover North and Postmasburg (Tainton 1992), together with representative compositions
of similar rocks from Pniel (Tainton 1992) and the Prieska area (Skinner et al. 1994).
Compared to unevolved orangeites, the rocks are relatively rich in silica as a consequence
of the presence of potassium feldspar and richterite. Other significant differences are their
258 CHAPTER 3

Table 3.3. Average and Range of Major Element Compositions of Representative Evolved
Orangeites
Wt% 2 3 4 5
Si0 2 45.75 45.20-46.92 44.87 44.13-45.57 42.51 41.01 44.27
Ti0 2 1.87 1.79-2.03 1.29 1.23-1.39 1.18 1.24 1.91
AI 20 3 5.75 5.53-6.10 7.41 7.12-7.87 4.08 6.43 6.26
FeZ03 8.20 7.92-8.40 8.80 8.71-8.98 8.26 7.67 6.26
MnO 0.12 0.11-0.13 0.17 0.15-0.20 0.13 0.11 0.14
MgO 20.92 19.74-21.96 14.89 14.54-15.47 27.99 20.40 18.23
CaO 4.86 4.42-5.16 10.36 10.12-10.79 4.04 6.15 7.02
Na20 1.12 0.82-1.56 0.83 0.61-1.06 0.60 0.36 0.68
K20 4.07 3.35-4.84 5.52 5.38-5.70 4.31 3.17 3.70
P20 S 0.78 0.55-1.07 0.71 0.41-1.09 0.43 1.06 1.09
LOI 4.25 3.12-5.09 3.29 2.64-3.95 4.26 8.42 5.51
Total 97.64 98.14 97.79 96.02 97.19
PI 1.09 0.99 1.39 0.63 0.82
UPI 2.39 4.38 4.73 5.79 3.58
PPJ 0.77 0.81 1.14 0.53 0.64
(N) 9 3 1
LOI =loss on ignition. PI, UPI, and PPJ are the peralkalinity, ultrapotassic, and perpotassic indices respectively. Total Fe is
expressed as Fez03. (N) =number of samples. I =Sover North; 2=Postmasburg 24/PK37; 3=Pniel; 4 =Brandewynskuil; 5
=Slypsteen. Data sources: 1-3 (Tainton 1992); 4-5 (Skinner et QI. 1994).

relative enrichment in Ah03 and depletion in MgO and volatiles. Rocks from Sover North
and Postmasburg have high Na20 contents which are considered by Tainton (1992) to
result from low-temperature alteration ofleucite and sanidine to Na-zeolites. Data on the
abundances of the individual volatiles constituting the LOI are not available. The paucity of
carbonate in evolved orangeites (1.10) suggests that the major volatile component is H20.
The limited data available indicate that evolved orangeites are typically miascitic,
not perpotassic, and only weakly ultrapotassic compared with unevolved orangeites
(Table 3.3). The Pniel orangeite is anomalous in that it is agpaitic and perpotassic as a
consequence of the high modal abundance of potassic richterite (Tainton 1992).
Tainton (1992) considers that the compositions of evolved orangeites, i.e., Si02-rich
rocks, do not lie on extensions of the linear arrays defined by unevolved orangeites on
plots of Si02, Ah03, CaO, and Na20 versus MgO (Figure 3.2). However, it should be
realized that these plots mainly reflect variations in macrocrystal olivine content and not
the compositions of evolving liquids. Hence, there is no a priori reason why any simple
compositional relationship should exist between diverse orangeites in these bivariate
plots. Moreover, Figures 3.2B and 3.2F appear to contradict Tainton's (1992) assertions.

3.3.4. Comparison with Kimberlites


Kimberlites show a remarkably wide range in their major element composition
(Mitchell 1986, Smith et at. I985b, Gurney and Ebrahim 1973) as a consequence of
differentiation and modal variations in their macrocrystaI and primary mineral contents
(MitcheII 1986). Average compositions (Table 3.4) are unlikely to have any real geo-
chemical significance but are useful for comparative purposes. MitcheII (1986) has noted
GEOCHEMISTRY OF ORANGEITES 259

8
50 - A B
-
D 6
~ 40-
CJ) - :~.~ 0
••
OJ
30 - 2
o
I I I I I I

2-5 -
c 9-5 -
o
- .i4 -
o •
o(\11-5 - • • If) 8-5 -
AD ~Q 0
0

•• • • 0(\1 - •• ~v~'o
i= - D~. If
.. o ..~
'·0 0
7-5-
0-5 - Ql~O -
DO ""

I I I I I 6-5~-~1-r-1-r-1-~1~1--~1

1-6 6 -
E
-
1-2
o 4-
o
(\I
0-8 -
Z II • D
0-4 • •
0-0 +---,--,----,

12 o 2-5 -
G
-
oo 8 010 1-5-
U ~
4 0-5 -

I I I I I I I I I I I I
10 20 30 40 10 20 30 40
MgO MgO
Figure 3.2. Major element compositional variation of orangeites from the Barldy West (Bellsbank, Sover,
Sover North, Pniel) and Postmasburg (PK35-37) regions (after Tainton 1992)_ 0 Be\lshank and Newlands;.
Sover; 0 Pniel and PK35; • Sover North and PK36; II PK37_

that kimberlites may be considered to be undersaturated ultrabasic rocks (Si02 = 25-35


wt%) with low Ah03 contents «5 wt%) and low Na201K20 ratios «0.5)_ Calcite and
dolomite are major minerals in most kimberlites; consequently volatile contents are high
(> 10 wt%) and dominated by C02_ Kimberlites are typically potassic but not agpaitic_
Thus, the major geochemistry of kimberlites is, in many respects similar to that of
unevolved orangeites_ Hence, major element compositions do not provide any simple
means of distinguishing the two rock types. Smith et aI_ (l985b) have shown that K20
and Ti02 may provide the only effective discriminant when both elements exceed I wt%.
;J

Table3A. Representative Compositions of Kimberlites and Olivine/Madupitic Lamproites


Wt% 2 3 4 5 6 7 8 9
Si<h 30.00 31.99 34.37 37.48 33.86 33.92 42.31 ±2.21 45.47 ± 1.16 39.91
n02 1.72 2.32 0.74 0.38 1.77 1.46 3.75 ±0.82 2.34 ± 0.32 2.89
Ah03 1.99 2.68 1.04 2.31 3.88 4.26 3.92±0.87 8.89±0.67 3.88
Fe203 5.23 5.64 4.12 3.88 10.48 8.27 8.71
FeO 3.32 3.24 3.56 3.40 8.27±0.54 5.99 ± 0.27
MnO 0.16 0.16 0.13 0.11 0.17 0.16 0.13
MgO 32.49 32.44 38.55 34.43 30.67 21.93 24.42±3.56 11.15±0.94 27.17
Cao 10.90 6.71 7.03 2.13 8.64 14.60 5.00±0.95 11.84 ± 1.79 5.16
Na20 0.19 0.05 0.19 0.03 0.24 1.00 0.50 ± 0.25 0.83±0.15 0.32
K20 0.70 1.11 0.80 0.65 0.86 2.92 4.01 ± 1.09 7.75 ± 1.49 2.69
P205 1.89 1.51 1.70 0.21 0.80 1.43 1.59 ± 0.48 2.08±0.66 0.35
LO} 10.71 11.51 7.42 13.87 8.94 9.56 6.07± 1.88 3.49± 1.19 8.63
99.30 99.36 99.65 98.88 100.31 99.51 99.84 99.83 99.78
On-craton kimberJites: I De Beers; 2 Wesselton; 3 Dutoitspan; 4 Jagersfontein (1-3; Clement 1982; 4, Smith et al. I 985a). Off-eraton kimberlites: 5 Berseba Reserve #2; 6 Anis Kubub (5-6 Spriggs
1988); 7 average of 105 Ellendale olivine Jamproites based on the data of Jaques et al. (1986); 8 average of 6 Leucite Hi1Is madupitic lamproites based on the data of Carmichael (l967a); 9
olivine-madupitic lamproite, Prairie Creek (Fraser 1987). LOI loss on ignition, this is C02 and H20 in kimberlites but mainly H20 in lamproites.
=

to>
~
GEOCHEMISTRY OF ORANGElTES 261

t 5·0 • •
KIMBERLITES
ON -CRATON
• • OFF-CRATON
••
....•••••

.
4'0
• ORANGEITES

. ..
-t. ~
..: 3'0 •• •
~
N • •
Q 2'0
~

I 1·0

0
o 1'0 2·0 3·0 4·0 5·0 6·0 7'0

- - K20 Wt.% •

Figure 3.3. Ti02 versus K20 for on- and off-craton kimberlites and orangeites (after Smith et al. 1985b).

Thus, kimberlites are typically characterized by low K20 and high Ti contents, whereas
orangeites exhibit the inverse relationship (Figure 3.3). The high Ti contents of kimber-
lites are attributable to the characteristically high modal abundances of groundmass
Ti-rich spinels and perovskite. The low K contents reflect a paucity in phlogopite and the
common presence ofkinoshitalite-rich groundmass micas.
As the mineralogical differences between evolved orangeites and kimberlites are so
distinctive, it is not surprising that evolved orangeites are easily distinguishable from
kimberlites on the basis of their higher Si02 and Ah03 and lower MgO, CaO, and C02
(LOI) contents.

3.3.5. Comparison with Lamproites


Lamproites show an exceedingly wide range in composition because of the numerous
possible primary minerals coupled with extensive differentiation within the clan (Mitchell
and Bergman 1991). Phlogopite and sanidine lamproites are mineralogically so different
from all orangeites that comparison of bulk rock compositions is unnecessary. However,
olivine and madupitic lamproites are low-silica, high-K20 rocks with some mineralogical
and compositional affinities with evolved orangeites.
Dawson (1987) has noted that the compositions of unevolved orangeites are similar
to those of olivine lamproites when the former are expressed on a CaC03-free basis (Table
3.4). Although this procedure does indeed result in bulk compositions resembling those
of olivine lamproites, it should be realized that there are no petrogenetic grounds for
undertaking this recalculation procedure. Excluding the constituents of one of the major
primary groundmass minerals from the bulk composition is petrologically unsound. The
inappropriateness of Dawson's (1987) approach may be realized by considering the
analogous deduction of an amount of Si02 equivalent to the quartz in a granite and then
claiming that granites are compositionally similar to syenites. Recalculation procedures
262 CHAPTER 3

of the type utilized by Dawson (1987) are valid only if the subtracted components are the
constituents of secondary minerals introduced into a rock subsequent to consolidation.
Thus, it is suggested here that unevol ved orangeites are compositionally distinct from
olivine lamproites, being poorer in Si02 and richer in CaO and C02. In contrast, evolved
orangeites have bulk compositions closely resembling those of olivine lamproites (Table
3.4).

3.4. FIRST-PERIOD TRANSITION ELEMENTS


First-period transitional elements may be considered compatible trace elements in
orangeites (and kimberlites) as they substitute for Fe and Mg in the principal early
crystallizing primary phases: olivine (Sc, Ni, Co), phlogopite (Sc, Cr, Cu), spinel (Sc, V,
Cr, Co, Zn), and pyroxene (Sc, Cr). Chromium occurs as a major element in primary
groundmass magnesiochromites. Thus, Cr abundances are strongly controlled by the
presence or absence of this mineral, e.g., the spinel-free Swartruggens orangeites are
relatively poor in Cr compared to spinel-rich rocks from Finsch. Primary sulfides are very
rare in orangeites and play no significant role in controlling the distribution of chaJcophiJe
transition elements. Nickel sulfides are common in setpentinized olivine macrocrysts.
However, the nickel forming these sulfides was originally present in solid solution as the
liebenbergite molecule in olivine and has been merely redistributed during serpentiniza-
tion.
Tables 3.5 and 3.6 show the abundance of first-period transition elements to be
similar within and between orangeites. The only significant difference between unevolved
and evol ved orangeites is with respect to their Ni contents. The lower Ni contents of the
latter correlate to the relative paucity of olivine in these rocks.
Data presented by Tainton (1992) indicate no systematic correlations between Sc, V,
Cr, Zn, and Cu with MgO. Ni correlates positively with Mg as expected, as olivine is the
major host for Ni. Absolute abundances for Cr and Ni in orangeites are high compared to
other mantle-derived basic magmas (Tainton 1992, Fraser 1987) and significantly higher
than predicted for primary melts (Ni =300-400, Cr =400-500 ppm) from peridotite
sources that have not been metasomatized (Fraser 1987).
Abundances of Sc, V, Co, Cu, and Zn are not very different from levels found in a
wide variety of mantle-derived magmas, including kimberlite and olivine lamproite
(Table 3.5; Mitchell 1986, Mitchell and Bergman 1991). Smith et aZ. (1985b) have
suggested that kimberlites have, on average, lower Cr contents (1000 ppm) than
orangeites (1800 ppm), a conclusion not supported by the data of Tainton (1992) or this
work.
In summary, first-period transitional element abundances are of little use in distin-
guishing orangeites, kimberlites, and olivine lamproites. Ni abundances, being related to
the presence of macrocrystal olivine, might permit estimation of the amount of contami-
nation of orangeite magma with xenocrystal components (see 4.1.2). Abundances and/or
ratios of abundances of first period transition elements, with the exception of Ti and Ni,
are oflittle use in geochemical modeling of orangeite petrogenesis.
Table 3.5. Average and Range of First-Period Transition Element Abundances (ppm) in Unevolved Orangeites
Swartruggens Finsch Bellsbank Sover Newlands New Elands
Sc
V
20: 16-28
131: 91-152
17: 12-23
132: 6-285
22: 11-39
72: 41-102
16:2-26
82:26-180
22: 16-32
48:30--77
23: 19-25
105:83-148
i
~
~
Cr 1207:315-1424 1765: 1100--2190 1670: 1130--2251 1852:975-2865 1891: 1616-2861 1514: 1430--1641
Ni 1034: 470--1742 1214:21-1544 1396: 573-2022 1253:648-920 1450:812-1749 1036: 902-1348
Co 73:54-96 71:61-93 96: 87-1121 80:83-92 71: 65-84 69:62-76
Cu 29:25-34 36 21: 7-50 21: 2-49 22:8-38 38:32-42
Zn 84:79-88 53 53:46-62 71: 41-409 43:40--48 83:75-87
(n) 4-21 1-30 6-48 3-31 5-19 4

Star Kimberlite WK-01-lamproite PC-OI-Iamproite


I
Sc 23: 18-35 14:6-38 21: 9-39 15: 14-16
V 100:21-760 82:20--267 46:27--68
Cr 2156: 1410--2620 893:430--2554 1014: 379-1703 1447: 1391-1500
Ni 1207: 895-1570 965:471-1800 968: 401-1500 1356: 1285-1443
Co 78:56-89 65: 9-125 69:31-92 96:95-97
Cu 93: 6-1320 55:39-93 52:47-57
Zn 69: 10--287 73:58-107 73:71-74
(n) 8
Data sources: Swartruggens (this work, Smith et al. 1985b); Finsch (this work, Fraser 1987); Bellsbank (this work, Tainton 1992); Sover (this work, Tainton 1992); Newlands (Tainton 1992); New
Elands (Smith et aJ. 1985b); Star (this work). Average kimberlite (Mitchell 1986); average West Kimberley olivine lamproite (WK-OI-Iamproite), Jaques et al. (1986); average Prairie Creek
lamproite (PC-Ol-Iamproite), Fraser (1987).

~
264 CHAPTER 3

Table 3.6. Abundances (ppm) of First-Period Transition Elements in Evolved Orangeites


2 3 4 5
Sc 11: 10--13 12:9-15 9
v 161: 139-182 104: 80--146 159 13I 231
Cr 1239: 1082-1399 1155: 970--1219 1965 1412 1267
Ni 495:470--524 960: 811-1146 1266 972 676
Co 77 69
Cu 25:24--26 31:23-36 31 28 54
Zn 60:51--65 77:72-92 60 76 86
(11) 3 9
1 ~ Postmasburg 24/P37; 2 ~ Sover North; 3 ~ Pniel; 4 ~ Brandewynskuil; 5 ~ Slypsteen. Data sources: 1-3 Tainton (1992);
4-5 (Skinner et al. 1994).

3.5. INCOMPATIBLE ELEMENTS


The incompatible trace elements (Sr, Ba, Zr, Nb, REE, Rb, Th, etc.) are usually
defined as elements having solid/liquid distribution coefficients for common rock-
forming silicates of approximately zero. They are strongly partitioned into the liquid
phase during partial melting of Iherzolitic sources and preferentially concentrated in
derivative liquids during crystal-liquid fractionation processes. Their abundances and
inter-element ratios are commonly used to infer the nature and degree of partial melting
of magma sources. Much of the geochemical lore pertaining to incompatible trace
elements has been deri ved from studies of basalts and related rocks. Parental magmas to
these rocks are thought to be derived from simple lherzolitic mantle source rocks.
Metasomatic phases enriched in incompatible elements have been postulated as being
present in lherzolitic sources of magmas of more extreme compositions, e.g., melilitite,
kimberlite. In such cases the incompatible element-rich phases are usually considered to
be completely consumed during the partial melting episode that gave rise to the incom-
patible element-rich magma.
As a caveat to the above, it should be realized that if incompatible element-rich phases
are not consumed during melting, then the elements in question are no longer incompat-
ible. In these instances, extensions of hypotheses derived to explain the trace element
distributions of basaltoid rocks to kimberlites and orangeites (Fraser 1987, Fraser and
Hawkesworth 1992, Tainton 1992, Tainton and McKenzie 1994) are unlikely to be
realistic.
In common magmas the elements may behave as incompatible elements throughout
most of their crystallization interval. In the case of magmas of extreme composition
early- forming liquidus phases include many minerals which have trace element solid/liq-
uid distribution coefficients greater than zero, e.g., REE and Sr in apatite or Rb, Ba, and
Cs in phlogopite. In these instances the elements in question are compatible and their
abundances may be significantly affected by crystal-liquid fractionation and/or crystal
accumulation. The latter process is particularly important regarding orangeites composed
principally of phlogopite.
GEOCHEMISTRY OF ORANGElTES 265

In some instances incompatible elements might not be removed from the liquid until
the later stages of crystallization of the groundmass. Therefore whole rock analysis may
provide reasonable estimates of their abundances and ratios. The mixing of significant
amounts of xenocrystal olivine with orangeite and kimberlite magmas implies that
absolute abundances of incompatible elements are of little geochemical significance
unless the magnitude of intra- and inter-intrusion differences are expressed against some
normalizing index or are extremely large (Le., differ by a factor of 10 or more).
Importantly, studies oflamproites and kimberlites have suggested that simple garnet
Iherzolites are not suitable source rocks for their parental magmas (Foley 1990, 1992a,
Edgar 1987, Mitchell and Bergman 1991, Mitchell 1986, Smith et al. 1985b), and their
sources may retain a residual mineralogy in which many so-called incompatible elements
are actually compatible. Recognition of this aspect of magma genesis and the possibilities
of compatibility during much of the crystallization interval have important ramifications
regarding the interpretation of the trace element geochemistry of kimberlites and
orangeites.
Tables 3.7 and 3.8 give abundances of incompatible elements in orangeites and
provide some comparisons with kimberlites and lamproites. In some instances the
difference in abundance is significant and useful in discriminating between rock types.
However, all of these geochemical data should be regarded in the context of the above
comments, realizing that abundances measured do not reflect those of the actual magma,
as they have been reduced or "diluted" by mixing with macrocrystal olivine.

3.5.1. Alkaline Earths


In unevolved orangeites, Ba is hosted primarily by phlogopite, and also occurs as a
major element in Ba-carbonates, hollandites, and barite. Sr is hosted primarily by apatite,
perovskite, and late-stage carbonates. Ba and Sr exhibit wide ranges in their abundances
within and between orangeites (Table 3.7). High and widely ranging Ba abundances, e.g.,
Swartruggens, reflect the presence of abundant late-stage, irregularly distributed barite.
High Sr abundances, e.g., Sover, reflect the presence of Sr-carbonates and phosphates.
BalSr ratios range from 0.6 to 4.8 (Table 3.9). There is no correlation between Ba and Sr
abundances, and both elements are regarded as not being strictly incompatible.
Evolvedorangeites have Sr and Bacontents (Table 3.8) and BalSrratios (1-2) similar
to unevolved types.
Unevolved orangeites are richer in Ba and Sr than kimberlites and poorer in Ba than
olivine and madupitic lamproites (Table 3.7). Evolved orangeites are poor in both Ba and
Sr compared to the latter.

3.5.2. Second- and Third-Period Transition Elements

3.5.2.1. Zirconium and Hafnium


Zr and Hf are concentrated in the groundmass of orangeites and hosted primarily by
late-crystallizing zirconium silicates. There is a wide range of Zr and Hf abundances
within and between orangeites (Table 3.7), and the highest contents are found in rocks in
which groundmass Zr minerals are most abundant, e.g., Swartruggens and New Elands.
Kable et al. (1975) report low Zr and Hf abundances in rocks of the Bellsbank Water
~

Thble 3.7. Average and Range ofBa, Sr, Zr, Hf, Nb, Ta, Th, and U Abundances (ppm) in Unevolved Orangeites
Swartruggens Finsch BeIlsbank Sover Newlands New Elands
Ba 5183: 1390-16300 1467:290-2590 3439:644-5942 2442:792-5480 3351:557-8536 1895: 1606--2167
Sr 1139: 51(}"2880 738:416-1423 1414: 601-2484 1127: 48(}..6591 1261:545-2630 1356: 1184-1487
Zr 401:256-1060 184: 11(}"359 291: 120-598 214:53-496 193: 129-395 400: 319-472
Hf 10:6-26 5: 3-10 8:4-15 5:3-8 4:3-6 12:9-15
Nb 143: 85-258 51: 2-93 168: 108-289 97:46-161 139: 95-169 111:94-120
Ta 8: 5-13 3: 2--6 14:7-26 9:4-14 9:7-12 7:5-8
Th 26: 16-42 9:5-16 45:20-74 30:6-51 33:21-44 27:24-30
U 7:5-17 3:2-4 7: 3-14 3: ]-5 5: 3-9 5:4--6
(n) 17 21-30 25-48 15-31 15-19 4

Star SAK ZK KIMBK West Kimberley Leucite Hills Prairie Creek


Ba 4370: 312(}"57oo 850 973: 195-1652 833:164-2292 10584:3007-41378 9831:4319-14400 1971: 1624-250
Sr 1808: 136(}..2550 1020 468: 186--829 1054:346-2428 1325:725-1889 3860: 2581-4904 1094:972-1284
Zr 194: 165-277 385 125:73-164 279: 12(}"717 1167: 564-3740 1302: 1152-1557 718:662-745
Hf 7: 4-13 3: 2.5-3.7 6:3-15 38:32-49 42:38-43 17: 16-18
Nb 134: 119-173 210 153: 132-193 150:37-346 186: 104-309 99: 79-137 101: 87-112
Ta 10:5-18 14:11-19 9:2-23 10: 6-14 6:5-7 6: 5-7
Th 28: 14-50 27 15:14-18 19:9-5 57: 18-98 37:33-45 12: 11-14
U 6 3:3-5 2: 1-0 9: 8-9 3:2-3
(n) 4-8 10 23 18-35 9-109 3-10 3-5
Data sources for orangeites as in Table 3.5. SAK South African kimberlite (Smith et al. I 985b); ZK:= Zaire kimberlites (Kampata 1993); KIMBK
= =Kimberley group kimberlites (Gement 1982,
this work); averages for the West Kimberley, Leucite Hills, and Prairie Creek lamproites from Mitchell and Bergman (1991).

c.>
~
GEOCHEMISTRY OF ORANGElTES 267

Table 3.8. Abundances (ppm) ofIncompatible Elements in Evolved Orangeites


2 3 4 5
Ba 2234: 1080--29292002: 1902-2077 1354 1929 1479
Sr 1293: 1404-1500 795: 624-1120 624 1445 1072
'Ix 554:516-550 305:233-425 171 297 191
Hf 19
Nb 63: 62-70 54:30--78 30 67 75
Ta 5:5-6 4:3-5 2
Th 17: 12-17 8: 7-10 4
U 3:2-5 1: 1-2 2
(n) 1-9 3
1 =Sover North; 2 =Postmasburg 24/PK37; 3 =Pniel; 4 =Brandewynskuil; 5 = Slypsteen. Data sources: 1-3. Tainton (1992);
4-5. Skinner et al. 1994.

Table 3.9. Averages and Ranges of Ratios of Incompatible Elements in Orangeites


Swartruggens Fiosch Bellsbank Sover
Rb/Sr 0.21: 0.05-0.36 0.20: 0.10--0.33 0.08: 0.03-0.20 0.18: 0.08-{).40
ZrlNb 2.75: 2.23-4.11 3.70: 2.27-5.69 1.71: 0.76-2.85 2.16: 0.13-3.75
BalSr 4.76: 1.55-9.67 1.97: 0.39-3.06 2.62: 0.39-5.53 2.85: 0.65-7.46
PblCe 0.08: 0.05-0.09 0.09: 0.04-0.14 0.06: 0.04-0.11 0.05: 0.02-0.08
NbIU 20.9: 15.2-28.15 18.7: 13.3-27.3 24.9: 11.7-38.9 29.3: 16.8--58.1
NbIY 7.49: 5.19-11.58 5.28: 3.78-8.86 10.3: 6.1-18.6 6.18: 4.08-8.82
KlRb 203: 169-255 197: 126-342 134:83-167 149: 101-166
ZrlHf 40.9: 33.8-43.9 38.2: 34.1-41.3 39.2: 20.4-50.1 34.1: 24.5-43.0
Nb/Ta 18.4: 15.6-21.6 14.4: 12.1-17.5 14.0: 7.8-23.3 11.6: 6.2-18.9
KlNb 283: 177-439 565:243-937 81:44-129 243: 142-334
KlBa 10.4: 2.3-23.5 23.0: 6.5-76.2 6.3: 1.2-33.5 12.9: 3.2-35.8
La/Yb 116: 55-192 88: 54-115 243:92-483 179:83-267
LalNb 1.52: 1.17-2.11 1.27: 0.99-1.59 1.46: 0.91-1.86 1.82: 1.16-2.69
CeIY 22.1: 15.6-41.4 14.0: 11.4-19.1 27.6: 16.7-42.8 21.5: 14.5-30.0
(n) 4-17 21-30 27-35 28--31

Newlands NewElands Star SoverNorth


Rb/Sr 0.05: 0.02--{).OS 0.12: 0.10--{).13 O.OS: 0.05--{).1O 0.11: O.OS--{).13
'IxlNb 1.41: 0.79-2.00 3.61: 2.80--3.93 1.45: 1.12-1.65 8.06: 7.4S-8.91
BalSr 2.67: 0.67-4.79 1.29: 1.26-1.50 2.53: 1.22-3.68 1.54: 0.67-2.08
PblCe 0.05: @I-O.09 0.09: 0.08-0.11 0.06: 0.05-0.07
NbIU 28.7: 15.1-38.2 22.8: 17.7-29.0 23.8: 15.0--31.3
NbIY 11.7: 9.5-16.0 5.65: 5.37-6.34 7.76: 5.67-10.2 3.47: 3.19-4.14
KlRb 147: 110--174 243:232-262 187: 169-210 213: 158--271
'IxlHf 46.4: 35.4-58.9 33.5: 29.1-43.S 32.3
Nb/Ta 15.1: 12.8-21.1 14.7: 9.5--26.1 13.3: 11.2-16.7
KINb 59.7: 30.9-100.3 359:295-433 184: 167-212 496:334-603
KlBa 3.95: 0.72-13.1 21.0: 18.8--23.5 5.85: 3.88--7.72 17.1: 10.9-34.6
La/Yb 266: 185-382 167: 144-188 263: 185-338 129: 108-138
LalNb 1.48: 1.15-1.83 1.87: 1.78-2.00 1.40: 0.99-1.84 2.17: 1.86-2.50
CeIY 30.9: 25.3-37.3 19.3: 18.7-20.9 20.8: 17.8-27.1 14.0: 10.3-15.5
(n) 15-19 4 4-8 1-9
(n) =numberofsamples. Data sources: Swartruggens (Smithet al. I 985b. this work); Finsch (Fraser 1987. this work); Bellsbank
(Tainton 1992. Kable et al. 1975. this work); Sover (Tainton 1992. this work); Newlands (Tainton 1992); New Elands (Smith
et al. 1985b. this work); Star (this work); SoverNorth (Tainton 1992).
268 CHAPTER 3

t 1000-

E 500-

-
Q.
Q.

I I I
5 10 15
- - Hf (ppm) •

Figure 3.4. Zr versus Hf for orangeites. Data sources: Bellsbank (Tainton 1992, this work); Sover (Tainton
1992); Swartruggens (this work); Star (this work); Finsch (Fraser 1987).

Fissure relative to the Main and Bobbejaan dikes; this observation is significant if the
Water Fissure rocks are evolved orangeites (Tainton 1992). Zr/Hf ratios of 24-50 are
similar to those of a wide variety of mantle-derived rocks, including lamproite and
kimberlite (Mitchell 1986, Mitchell and Bergman 1991). The coherent behavior ofZr and
Hf on logarithmic plots of abundances (Figure 3.4) indicates that these elements are highly
incompatible in orangeites and their ratios are unaffected by fractional crystallization or
hybridization. Thus, Zr/Hf ratios may reflect those of the magma sources. Similar ratios
in all orangeites suggest that these elements are hosted by only one phase in the mantle
sources of the parental magmas.
Zr abundances of all orangeites are, on average, significantly less than those of
olivine lamproites (Tables 3.7, 3.8). In particular, the low Zr contents of evolved
orangeites provide a means of discriminating between these rock types and olivine
lamproites of similar major element composition. There are no significant differences in
the Zr (and Hf) contents of orangeites and kimberlites (Table 3.7).

3.5.2.2. Niobium and Tantalum


Nb and Ta are concentrated in the groundmass of orangeites where they substitute
for Ti in late-stage rutile, Mn-ilmenite, and hollandite. Abundances are widely variable
within and between intrusions. Finsch appears to be depleted in both elements relative to
other unevolved orangeites. Average NblTa ratios of 11-14 are similar in all intrusions,
although wide ranges are apparent within individual intrusions. Evolved orangeites are
poor in Nb and Ta relative to unevolved types (Table 3.8, Figure 3.5). Significantly, the
Water Fissure is low in Nb and Ta relative to unevolved Bellsbank orangeites (Kable et
al. 1975), although Nb/Ta ratios (approximately 20) are similar (Figure 3.5).
GEOCHEMISTRY OF ORANGElTES 269

1992

-
E

-
c.
c.
,g
Z
100
• BELLSBANK
+ SOVER
• STAR
o FINSCH
• SWARTRUGGENS

5 10 15 20
To (ppm) •
Figure 3.5. Nb versus Ta for orangeites. Data sources: Bellsbank (Kable et al. 1975, Tainton 1992); Sover,
Sover North, Pniel, and Postmasburg (Tainton 1992); Star and Swartruggens (this work); Finsch (Fraser 1987).

Figure 3.5 shows significant discrepancies exist between the data of Tainton (1992)
and Kable et al. (1975) for the Bellsbank unevolved orangeites. Ta abundances given by
Tainton (1992) and obtained by ICP-MS are much higher than the INAA data of Kable
et al. (1975). Nb and Ta show only a weak correlation, and NblTa ratios vary considerably
(Table 3.9). Similar trends are seen in Tainton's (1992) data for Sover. In contrast,
orangeites from Finsch and Swartruggens, whose Ta content was determined by INNA,
show greater coherence in their NblTa ratios, and they plot in Figure 3.5 close to the Nbrra
ratio defined by the Bellsbank samples analyzed by Kable et al. (1975). Star orangeites
resemble Sover in their Nb and Ta (lNAA) abundances. Further study is required to
determine whether the differences noted above are real or due to inaccurate determination
ofTa.
The majority of unevolved orangeites does not differ significantly in Nb and Ti
contents relative to those ofkimberlites or olivine lamproites (Table 3.7, Figure 3.6). The
contention of Smith et al. (1985b) that orangeites are poorer in Nb (120 ppm) than are
kimberlites (210 ppm) is not supported by the larger data base now available (Table 3.7,
Figure 3.6). Orangeites from Bellsbank are much richer in Nb than most low-Ti02
kimberlites. Only "off-craton" Namibian kimberlites (Spriggs 1988) have high Nb and
low Ti02 contents. The spread of data in Figure 3.6 for Bellsbank and Sover is related to
mixing with macrocrystal olivine (Tainton 1992).
Mineralogical observations suggest that the controls on Ti and Nb distribution are
different in orangeites and kimberlites. Much of the Ti in kimberlites is bound in the
abundant groundmass perovskite and spinels. Nb in kimberlites is carried primarily by
270 CHAPTER 3



BELLSBA~"'''''7 •

., ., /
",.
ISWR
I •
••

n:. ,. :.
E 200 I I
a. ••• •
a.

Y . "-...
•1 . . ;I ,

l
l 'I.
.tl •
Z
100 • ?
SOVER ..... '" .~ SOVER NORTH
/F ,.. ~

• I EO PNIELI
W POSTMASBURG
o
o 1·0 2'0 3·0 4'0
Ti0 2 wt. % •

Figure 3.6. Nb versus Ti02 for orangeites and kimberlites. Data sources for orangeites as in Figure 3.5. Data
for kimberlites from South Africa (Smith et al. 1985b, Clement 1982) and Namibia (Spriggs 1988). SWR =
Swartruggens; F = Finsch; EO = evolved orangeites.

perovskites, although kimberlites rich in macrocrystal magnesian ilmenite may exhibit


enhanced Nb contents, e.g., Premier (Kable et al. 1975, Mitchell 1986). Thus, increasing
Nb is correlated with increasing Ti and reflects modal increases in groundmass titanium-
rich oxides.
Figure 3.7 demonstrates that evolved orangeites cannot always be distinguished from
unevolved types on the basis of Zr and Nb abundances, e.g., orangeites from Finsch and

t
300

--. 200
E
a.
a.
.tl
Z
100
SOV:*,~

o
o 100 200 300 400 500 600
Zr (ppm) •

Figure 3.7. Nb versus Zr for orangeites. Data sources as in Figure 3.5.


GEOCHEMISTRY OF ORANGEITES 271

t 300
• KIMBERLITE

E 200
~ ...
.J:l
Z 100

o
o 100 200 300 400 500 600 700 800 900 1000

Zr (ppm) ~

Figure 3.8. Nb versus Zr for orangeites, kimberlites. and olivine lamproites. Data sources for orangeites as in
Figure 3.5. Data for Ellendale and Prairie Creek lamproites from Jaques et al. (1986) and Fraser (1987).
Kimberlite data from Clement (1982). Smith et al. (1985b). Spriggs (1988). and Kampata (1993).

Sover have low Nb contents. However, evolved varieties are characterized by low Nb
abundances «100 pmm) and widely varying Zr contents with relatively high (>3) ZrlNb
ratios. Unevolved orangeites display a positive correlation with respect to increasing Nb
and Zr contents. Individual localities have distinct, but overlapping, Nb and Zr contents
and similar ZrINb ratios (Table 3.9).
Figure 3.8 indicates that kimberlites cannot be distinguished from orangeites on the
basis of their Nb and Zr contents. Olivine lamproites may be easily distinguished from
both kimberlites and orangeites on the basis oftheir much higher Zr contents. Sover North
has Zr and Nb contents similar to the most Zr-poor olivine lamproites. Figure 3.8 indicates
that petrographically similar olivine-rich rocks, with ZrINb > 3 and Zr > 500 ppm, are
more likely to be olivine lamproites than unevolved orangeites.

3.5.3. Thorium and Uranium


Thorium and uranium are probably primarily concentrated in groundmass apatite
and perovskite. Table 3.7 and Figure 3.9 show that Th and U abundances vary widely
within and between intrusions. Extremely high Th (630, 920 ppm) and U (15.0, 22.9ppm,
Th/U >40) abundances have been reported by Fesq et al. (1975) for apatite-rich orangeites
from the Bellsbank Main Fissure. On the basis of the limited data available, with the
exception of Finsch, evolved orangeites appear to be poor in Th and U relative to
unevolved types. Th/U ratios range from 3 to 9 (Figure 3.8). There is a notable lack of
correlation between Th and U on logarithmic plots of abundances for Bellsbank, Sover,
and Sover North orangeites. Coherence is better for Swartruggens and Finsch. Whether
the data spread is real and due to removal of U as the highly soluble uranyl ion during
weathering (Paul et al. 1977, Fesq et al. 1975) or results from the compatible behavior
of U is unresolved.
Th and U abundances and Th/U ratios of kimberlites are insufficiently well estab-
lished to state conclusively that the Th/U ratios of orangeites (6-11) are greater than those
272 CHAPTER 3

t 100

-
50

E
c..
c..
.£:.
I-- 10

• SWARTRUGGENS
• SOVER
+ BELLSBANK
o FINSCH
(ff} SOVER NORTH

5 10
U (ppm) ---.

Figure 3.9. Th versus U for orangeites. Data sources as in Figure 3.5.

of kimberlites (3-7) as suggested by Gurney and Hobbs (1973). Paul et al. (1977) have
shown that kimberlite Th/U ratios may exceed 10.
Olivine lamproites may have very high Th/u ratios (24) with no correlation evident
between Th and U abundances (Jaques et al. 1986). It is not known whether the Th/U
ratios of kimberlites and olivine lamproites reflect primary variations in Th and U
abundances or are elevated because of U leaching.

3.5.4. Rare Earth Elements


Rare earth elements are concentrated in apatite, perovskite, and REE-bearing car-
bonates during the later stages of crystallization of orangeites (this work, Fesq et al. 1975,
Mitchell and Reed 1988, Mitchell and Steele 1992). Orangeites typically have high total
REE abundances (Table 3.10) and are characterized by extreme fractionation of the light
REE relative to the heavy REE (Mitchell and Brunfelt 1975, Fesq et al. 1975).
Tables 3.9 and 3.10 indicate that LalYb ratios and REE abundances vary widely
within and between intrusions. Total REE abundances typically exceed 500 ppm. Fesq et
al. (1975) have reported atypical, extraordinarily high REE abundances (870 and 1120
ppm La, 1910 and 2080 ppm Ce) in apatite-rich orangeites from the BeIIsbank Main
Fissure.
Representative chondrite normalized REE distribution patterns are given in Figures
3.10 to 3.12. Within a given intrusion, light REE are not fractionated relative to each other
GEOCHEMISTRY OF ORANGEITES 273

Table 3.10. Averages and Range of Rare Earth Element Abundances (ppm) in Orangeites
Swartruggens Finsch Bellsbank Sover
La 218: 145-316 62:41-100 252: 126-504 168:80-250
Ce 429:276-719 132: 82-235 464:236-871 324: 147-544
Pr 55:28-91 36: 17-55
Nd 158: 106-218 55:32-94 163:83-286 115: 53-178
Sm 20.0: 14.1-32.3 7.7: 4.8-12.8 17.0: 8.5-28.3 12.6: 5.9-18.0
Eu 5.03: 3.58-10.5 1.80: 1.14-3.10 3.93: 1.97-6.33 3.00: 1.50-4.62
Od 13.2: 6.7-19.3 9.51: 4.81-15.2
Tb 1.65: 0.86-3.31 0.52: 0.32-0.98 1.05: 0.56-1.59 0.82: 0.43-1.28
Dy 4.09: 1.57-6.33 3.09: 1.39-5.13
Ho 0.62: 0.25-1.00 0.52: 0.23-0.89
Er 1.51: 0.47-2.98 1.20: 0.48-2.33
Tm 0.17: 0.05-0.38 0.19: 0.43-1.90
Yb 2.07: 0.95-5.77 0.72: 0.37-1.11 1.13: 0.25-2.57 0.99: 0.43-1.90
Lu 0.14: 0.02-0.36 0.13: 0.05;0.34
(n) 15-21 22 21-35 28-30

New1ands New Elands Star Sover North


La 203: 140-265 247: 175-334 192: 164-220 149: 142-168
Ce 346:253-480 439:319-647 332:238-271 288:263-316
Pr 38:27-51 35:32-38
Nd 120: 84-156 174: 138-215 113: 107-132
Sm 11.3: 8.0-14.3 21.4: 17.3-24.8 15.0: 12.7-17.9 15.1: 14.1-16.2
Eu 2.69: 1.96-4.15 5.95: 4.33-6.78 3.16: 2.19-4.28 3.58: 3.04-4.14
Od 8.43: 6.08-14.1 10.5: 9.73-12.0
Tb 0.63: 0.45-0.87 1.36: 1.19-1.62 0.69: 0.06-1.08 1.07: 0.94-1.1 7
Dy 2.73: 2.08-3.63 4.08: 3.04-5.08
Ho 0.42: 0.40-0.51 0.66: 0.61-0.81
Er 0.97: 0.79-1.25 1.51: 1.34-1.94
Tm 0.13: 0.10-0.20 0.18: 0.15-0.23
Yb 0.76: 0.49-0.97 1. 72: 1.39-2.02 0.72: 0.26-1.19 1.17: 1.00-1.55
Lu 0.11: 0.08-0.13 0.14: 0.12-0.16
(n) 14-18 4-8 8 9
(n) = number of samples. Data sources: Swartruggens (Mitchell and Brunfelt 1975, this work); Finsch (Fraser 1987); Bellsbank
(Tainton 1992, this work); Sover (Tainton 1992); New1ands (Tainton 1992, this work); New E1ands (Smith et al. 1985b, this
work); Star (this work); Sover North (Tainton 1992).

and the patterns are subparallel. Divergence of the patterns, represented by upward
inflections, are observed only with respect to the heavy REE. The divergences are
attributable to either experimental errors (Tainton 1992) or to contamination of the rock
with crustal material (Mitchell 1986).
Figures 3.10 and 3.11 show that unevolved orangeites have very similar REE
distribution patterns and high LalYb ratios (typically 100-350). In contrast, evolved
orangeites (Figure 3.12) are not as enriched in their absolute REE abundances and have
lower LalYb ratios (80-150). Interestingly, the Finsch orangeites have very low LalYb ratios
(20-115) and REE distribution patterns (Figure 3.12), similar to those of evolved orangeites.
No distribution pattern exhibits significant Eu or Ce anomalies. Fesq et al. (1975)
have reported large negative Eu anomalies in samples from Bellsbank. Recent studies of
274 CHAPTER 3

• SOVER
o BELLSBANK
+ NEWLANDS

IJ.J
.....
a:o 10 2
Z
o
::z::
()
.......
IJ.J
.....
jjj
(!)
z
<[ 10
a::
o

La Pr Eu Tb Ho Tm Lu
Ce Nd Sm Gd Oy Er Yb

Figure 3.10. Chondrite nonnalized REE distribution patterns for unevolved orangeites from Sover, Bellsbank,
and Newlands. All data from Tainton (1992).

Bellsbank orangeites have not supported the existence of this anomaly (Tainton 1992,
this work). Tainton (1992) considers that Fesq et al. (1975) analyzed strongly weathered
material and the Eu anomaly resulted from preferential leaching of Eu2+ from the rocks
by reducing ground waters. Mitchell and Brunfelt (1975) reported weak negative Eu
anomalies in the Swartruggens orangeites. This anomaly may also may be related to
weathering, as less-altered samples from Swartruggens analyzed in this work were found
not to exhibit any Eu anomaly (Figure 3.11). It is concluded in this work and by Tainton
(1992) that the REE distribution patterns of orangeites do not characteristically exhibit
negative or positive Eu anomalies.
Surprisingly, the REE contents and distribution patterns of kimberlites are not well
established. Although the basic features of the REE geochemistry are known, Mitchell
(1986) has commented that many of the published data are subject to serious analytical
errors, or analyses have been undertaken upon contaminated diatreme facies material
and/or altered rocks. There have been no systematic studies of REE distributions in
well-characterized suites of fresh uncontaminated kimberlite from a single intrusion.
Data for fresh kimberlites (Mitchell and Brunfelt 1975, Mitchell and Clement
unpublished, Kampata 1993, Spriggs 1988) indicate that kimberlites are light REE
GEOCHEMISTRY OF ORANGEITES 275

103 0 .....
~" • SWARTRUGGENS
+~,,>, 0 NEW ELANDS
~.\
0, + STAR
\'
\\ • ROBERTS VICTOR

~+~\
IJJ
l-
ll:: 10 2
0

\
z
0
::c ~,
\('
u
.....
IJJ "
I- '-': "
-.;::',
IJJ -...;::',
<.!)

+~
Z
~ 10
Il::
0 "0

+-+

La Pr Eu Tb Ho Tm Lu
Ce Nd Sm Gd Oy Er Yb
Figure 3.11. Chondrite nonnalized REE distribution patterns for unevolved orangeites from Swartruggens,
New Elands, Star, and Roberts Victor. All data this work.

enriched, REE distribution patterns are linear, Eu anomalies are absent, and REE
abundances and LalYb ratios vary widely between intrusions.
Figure 3.13 indicates that kimberlites cannot be distinguished from orangeites on the
basis of their Sm content and LalYb ratios. The claim by Mitchell and Brunfelt (1975)
and Mitchell (1986) that micaceous kimberlites (i.e., orangeites) have higher LalYb ratios
than kimberlites is not supported by the data in Figure 3.13.
Mitchell and Bergman (1991) have previously concluded that olivine iamproites
cannot be distinguished from orangeites on the basis of their REE geochemistry, as LalYb
ratios and REE distribution patterns are similar. This conclusion is supported by Figure
3.13.
Using a very limited data base, Mitchell (1986) suggested that kimberlites and
orangeites could be distinguished on the basis of their SmiTh ratios. This conclusion is
not supported by the more extensive data now available.
Smith et al. (1985b) have noted that a plot of P205 versus Ce content may serve to
distinguish kimberlites from orangeites, although Figure 3.14, incorporating recent data,
does not support this observation. Although most on-craton South African kimberlites
have high P205 and low Ce contents relative to orangeites, there are no significant
• SOVER NORTH
o PNIEL
+ POSTMASBURG-
PK37
• FINSCH
w
~
a:: 10 2
0
Z
0
::I:
U
......
W
~
LiJ
(!)
z
«
a:
10
0

La Pr Eu Tb Ho Tm Lu
Ce Nd Sm Gd Dy Er Yb

Figure 3.12. Chondrite normalized REE distribution patterns for a Finsch orangeite (Fraser 1987) and evolved
orangeites from Sover North. Pniel. and Postmasburg (Tainton 1992) .

• KIMBERLITE

t"
E

Ii 20

..
E
en POST;
10 M4SBURG
~ ~-
~-
FINSCH
0
0 50 100 150 200 250 300
La / Yb ~

Figure 3.13. Sm versus La/Yb ratio for orangeites. kimberlites. and olivine lamproites. Data sources for
orangeites as in Figure 3.5. Kimberlite data from Mitchell and Clement (unpublished). Mitchell and Brunfelt
(1975). Spriggs (1988). and Kampata (1993). Olivine lamproite data for Ellendale. Argyle. and Prairie Creek
from Jaques et al. (1986. 1989) and Fraser (l987).

276
GEOCHEMISTRY OF ORANGElTES 277

3'0 • ON - CRATON KIMBERLITES

t
o OFF -CRATON KIMBERLITES
o
+ UNEVOLVED ORANGEITES
• SOVER NORTH • 0
~ FINSCH
+

-
~
It)

• + +

o
rr 1'0


100 200 300 400 500 600

Ce (ppm) •
Figure 3.14. P205 versus Ce for orangeites and kimberlites. Data for orangeites from Tainton (1992), Fraser
(1987), Smith et al. (1985b) and this work. Data for kimberlites from Smith et al. (l985b), Spriggs (1988),
Kampala (1993), and Mitchell and Clement (unpub.). SN = Sover North; KK = Kundelungu kimberlites.

differences for samples with low P and Ce contents. On-craton kimberlites from Kun-
de)ungu, and some Ce- and mica-rich South African kimberlites plot well within the
"orangeite field," i.e., P205/Ce ratios <0.05, as do off-craton kimberlites from South
Africa and Namibia.

3.5.5. Alkali Elements


Rubidium is primarily hosted by phlogopite in all orangeites, and by amphibole and
potassium feldspar in evolved varieties. Wide variations exist in Rb contents within and
between orangeites, e.g., Swartruggens (mean 191, range 146-273 ppm; Mitchell and
Brunfelt 1975, Smith et al. 1985b, this work), Finsch (mean 137, range 43-187 ppm;
Fraser 1987, this work), Bellsbank (mean 101, range 56-185 ppm; Tainton 1992, this
work), Sover (mean 159, range 64-305 ppm; Tainton 1992), Newlands (mean 60, range
27-89 ppm; Tainton 1992), New Elands (mean 162, range 153-176 ppm; Smith et al.
1985a, this work), Star (mean 132, range 110-159 ppm; this work), Sover North (mean
60, range 134-202 ppm; Tainton 1992), Postmasburg (mean 234, range 212-256 ppm;
Tainton 1992). Much of the variation reflects mixing with Rb-free olivine macrocrysts
and variations in phlogopite to carbonate plus apatite ratios. The data indicate no
significant differences in the Rb contents of unevolved and evolved orangeites. There is
a strong positive correlation (Figure 3.15) between K and Rb, reflecting increases in
modal phlogopite contents.
278 CHAPI'ER3

• SaVER
o BELLSBANK

-
• SWARTRUGGENS
+ FINSCH
E

-
Q.
Q.

10000

50 100 500
Rb (ppm) •
Figure 3.1S. K versus Rb for orangeites and olivine lamproites. Data sources for orangeites as in Figure 3.5.
Ellendale and Prairie Creek data from Jaques et al. (1986) and Fraser (1987). respectively.

KlRb ratios (Table 3.9) also vary widely within and between intrusions. Some of the
higher ratios. e.g., those greater than 300 at Finsch, may reflect preferential removal of
Rb during weathering and/or deuteric alteration (Mitchell and Crocket 1971, Barrett and
Berg 1975). There are no significant differences in the KlRb ratios of unevolved and
evolved orangeites. KlRb ratios in the majority of orangeites lie between 100 and 250
and are thus typically lower than the crustal average of 250 (Figure 3.15).
Orangeites, because of their higher modal abundances of phlogopite, are enriched in
Rb relative to kimberlites: e.g., Bultfontein, 72-111 ppm; Wesselton, 88-107 ppm; De
Beers, 26-91 ppm; Jagersfontein, 51-107 ppm (Gurney and Berg 1969). Orangeites
cannot be distinguished from kimberlites on the basis of their KlRb ratios (Gurney and
Berg 1969, this work). KlRb ratios of kimberlites in the Kimberley group range from 90
to 238 (Gurney and Berg 1969).
Olivine lamproites from Ellendale (Jaques et al. 1986) and Prairie Creek (Fraser
1987) have higher Rb contents than orangeites of similar K content (Figure 3.15);
consequently, KlRb ratios are less than those of orangeites. Note, that unlike orangeites,
K and Rb in the Ellendale lamproites are not positively correlated (Figure 3.15).
Cesium is hosted primarily by phlogopite in all orangeites, but interpretation of Cs
data is rendered difficult by the extreme mobility of this element during weathering and/or
deuteric alteration (Fesq et al. 1975).
The few data available indicate that, in the Barkly West region, Cs contents of
unevolved orangeites (Tainton 1992) are similar within and between intrusions, e.g.,
Sover (mean 9.51, range 3.29-19.96 ppm), Bellsbank (mean 9.53, range 4.79-18.33
ppm), and Newlands (mean 9.51, 3.29-19.96 ppm). Cs abundances show no correlation
with K contents, suggesting that some of the variation may be due to alteration. Signifi-
cantly lower Cs contents (3.0-4.4 ppm) for Bellsbank, which may also result from
alteration, have been reported by Fesq et al. (1975). The Finsch orangeites have signifi-
GEOCHEMISTRY OF ORANGElTES 279

cantly lower Cs contents (mean 2.39, range 0.9-3.27 ppm; Fraser and Hawkesworth
1992) than the Barkly West orangeites. Roberts Victor orangeites have a mean Cs content
of 6 ppm (range 5-8; Gurney and Berg 1969). Evolved orangeites from Sover North and
Postmasburg average 4.09 ppm (1.21-8.91) and 4.07 ppm (3.01-4.64 ppm Cs), respec-
tively. The wide range in Cs abundances found at Sover North shows no correlation with
K content.
KlCs ratios of unevolved orangeites (Tainton 1992), e.g., Sover (mean 3357, range
1001-104140), Bellsbank (mean 1612, range 403-2734), Newlands (mean 3986, range
1892-64260) are lower than those of evolved orangeites, e.g., Sover North (mean 12867,
range 5060-240130, Postmasburg (mean 11678, range 9998-14838). Finsch orangeites
are anomalous with respect to other unevolved orangeites in having high KlCs ratios
(mean 11000, range 6159-15400, Fraser and Hawkesworth 1992).
Insufficient data are available to state conclusively whether or not the Cs contents of
kimberlites differ significantly from orangeites. Older spectrographic data, which are
subject to serious analytical errors resulting in overestimation ofCs abundances, suggest
7-76 ppm Cs (KlCs =118-4500) for Kimberley area kimberlites (Gurney and Berg 1969).
Recent NAA data for the same kimberlites (Mitchell and Clement unpublished) give
lower abundances (1.68-44.9 ppm Cs) and low K/Cs ratios (495-4440).
Few data exist for Cs in olivine lamproites. Cs abundances (0.7-44 ppm) and hence
KlCs ratios (1900-25000) vary enormously for Ellendale olivine lamproites (Jaques et
al.1986)
In summary, orangeites, olivine lamproites and kimberlites cannot be distinguished
from each other on the basis of their Cs contents or KlCs ratios. Orangeites have greater
Rb contents than kimberlites, although KlRb ratios are similar. Olivine lamproites are
relatively rich in Rb and appear to have lower KlRb ratios than either kimberlites or
orangeites.

3.5.6. Lead
Lead in orangeites is probably hosted primarily by phlogopite. Galena is present only
as an insignificant accessory phase. The Pb contents of unevolved and evolved orangeites
do not differ significantly, e.g., Sover (mean 17, range 6-31 ppm), Bellsbank (mean 27.8,
range 14-50 ppm), Newlands (mean 18.8, range 4-32 ppm), New Elands (mean 35.5,
range 30-41 ppm), Swartruggens (mean 35.7, range 22-40), Sover North (mean 13.8,
range 14-23 ppm; Tainton 1992, Smith et al. 1985h). Finsch is anomalous in having
relatively low Ph (mean 12.62, range 4.55-24.96) contents (Fraser and Hawkesworth
1992).
The limited data available indicate that kimherlites have widely varying Pb contents,
e.g., Kimberley group, 2.3-21.9 ppm (Smith et al. 1985b); 10-37 ppm (Clement 1982);
Namibia, 5.1-8.5 ppm (Spriggs 1988). Despite the wide range, Pb contents are typically
less than 17.5 ppm (Figure 3.16). High Pb contents are found primarily in varieties
modally enriched in mica.
Smith (1983) has noted that the Pb contents of orangeites are higher than those of
kimberlites. A plot of Pb versus Ce (Figure 3.16) demonstrates that orangeites typically
contain more than 15 ppm Pb and 200 ppm Ce, suggesting that this diagram, in most
280 CHAPTER 3

t 50 -

40 -
+. KIMBERLITES
ORANGEITES
• SOVER NORTH
.. FINSCH
+ + +
+
+

E +
Q. 30 - ++ +t. -i:t-+ + + + +
Q. + SN + ~ ++ ~ ..

if 20 - .. + _~~ tr +++ tr / • +
~- -.-; - ~."'-.-:~1!~-*-.t-:-:++++- +; - - - - -
1e..... .
10-
· ..~
.,... .;
.... "'"+
lr . . .;. •
1 1 1 III 1 1 1 1 1 1
100 200 300 400 500 600

Ce (ppm) ..

Figure 3.16. Pb versus Ce for orangeites and kimberlites. Data sources as in Figure 3.14. SN =Sover North.

instances, may be used to discriminate between the two rock types. Orangeites from
Finsch are anomalous with respect to their Pb and Ce contents relative to other orangeites
(Figure 3.16).
It is noted above that kimberlites are poor in Rb relative to orangeites; thus, plots of
Pb versus Rb should also serve to discriminate between the two rock types. However,
such plots will merely reflect the high modal abundances of phlogopite in orangeites
relative to kimberlites and are thus unnecessary.
Olivine lamproites contain widely varying amounts of Pb (mean 51, range 17-124
ppm; Jaques et al. 1986) and do not appear to be significantly different from orangeites
in their Pb content.

3.6. INTER-ELEMENT RELATIONSHIPS


3.6.1. Extended Incompatible Element Distribution Diagrams
Extended incompatible element distribution diagrams or "spidergrams" have become
a sine qua non of geochemical studies of basaltic and related rocks. While such diagrams
are entirely appropriate for interelement comparisons in rocks which approximate liquid
compositions, their applicability to rocks whose compositions are controlled by crystal
accumulation is open to question.
The relative order of elements in these diagrams reflects decreasing liquid--crystal
distribution coefficients for a "normal" mantle source. This sequence gives smooth
interelement distribution curves for oceanic basalts when trace element abundances are
normalized to a primitive mantle composition. Clearly, magmas which are not derived
from mineralogically similar sources will not give smooth distribution patterns on such
plots. Thus, distribution patterns for rocks such as potassic lavas and melilitites charac-
teristically exhibit negative anomalies for several trace elements when they are normal-
GEOCHEMISTRY OF ORANGEITES 281

ized to primitive mantle compositions. Interpretation of the geological meaning of such


anomalies is subjective. Negative anomalies are commonly considered to reflect the
presence of residual phases in the mantle which sequester the elements in question.
However, the presence of these residual phases implies that the source mantle is unlikely
to be primitive, therefore, the use of primitive mantle as a normalizing index must be
inappropriate. Alternatively, the anomalies may be simply an indication that a given
magma was derived from a mantle source very different from that of oceanic basalts. In
the case of lamproites such a conclusion is not surprising, given that radiogenic isotopic
studies indicate that lamproite parental magmas might be derived from an evolved mantle
that has undergone several episodes of metasomatism.
Several studies of orangeites and related rocks (Tainton and McKenzie 1994, Fraser
and Hawkesworth 1992, Tainton 1992, Sheppard and Taylor 1992, Jaques et al. 1989,
Rock 1990) have used such extended incompatible element distribution diagrams in
attempts to deduce the geochemical characteristics of the sources of parental magmas. In
all of these studies it is claimed that incompatible element abundances, normalized to the
composition of the primitive mantle, reflect the composition of the parent magma and
not the effects of extended fractional crystallization and/or hybridization. Negative
anomalies are interpreted as indicating the presence of residual phases in the mantle
sources of the magmas. These sources are postulated to range in character from metaso-
matized evolved mantle to subducted oceanic crust; however, in all cases they are
certainly not equivalent to the "normal" primitive peridotite mantle proposed as the source
of oceanic basalts.
Figures 3.17-3.24 present incompatible trace element abundances for orangeites and
related rocks normalized to the composition of primitive mantle (Sun and McDonough
1989). Bearing in mind the above comments, it is considered that inferences about the
nature of the mantle sources from these diagrams mayor may not be realistic. Their
principal value is for comparative purposes, as differences in the distribution patterns may
reflect differing processes or sources involved in the generation of these magmas.
Figure 3.17 shows that unevolved orangeites from the Barkly West District give
distribution patterns with prominent negative Rb, K, and Sr anomalies. Wide fluctuations
in Cs abundances may reflect alteration as noted above. The figure clearly shows the
enrichment of Bellsbank orangeites in many incompatible elements relative to those from
Sover. Orangeites from Swartruggens, New Elands, and Star exhibit similar distribution
patterns (Figure 3.18) to those of the Barkly West orangeites, but differ in that the
magnitude of the Rb and K anomalies is much less. The absence of Hf anomalies in these
orangeites suggests that the minor Hf anomalies present in the Barkly West distribution
patterns are an artifact of the analytical technique used to determine Hf. This conclusion
is supported by the lack of a corresponding Zr anomaly in the Barkly West samples.
Figure 3.19 gives distribution patterns for evolved orangeites from Sover North and
Postmasburg. The Sover North pattern is very irregular and differs from unevolved
orangeites in having no significant negative Rb and K anomalies. Negative Sr and P and
positive Zr and Hf anomalies are present. Postmasburg orangeites exhibit positive K and
negative Sr, Th, and U anomalies. Figure 3.19 also shows the distribution pattern of the
Finsch orangeites is similar to those of evolved orangeites in that they both lack Rb and
K anomalies and have negative Sr and Th anomalies.
282 CHAPTER 3

0
w 10 3 ORANGEITES
...J
I-
Z
<t
:!:
w
>
~ 10 2
:!:
a::
a..
.....
w
I-
W

-
(!) 10
z 0--0 BELLSBANK
<t
a:: SOVER
0
*--- ... NEWLANDS

Pb Rb Th K Nb Ce Nd Sm Zr Ti Yb
Cs 80 U To La Sr P Hf Eu Y Lu

Figure 3.17. Incompatible element distribution diagrams for unevolved orangeites from Bellsbank, Sover, and
Newlands. All data from Tainton (1992).

10 3 ORANGEITES
w
...J
I-
z
<t
:!:
w
> 10 2
i=
~
a::
a..
.....
w

-
I-
w 10
SWARTRUGGENS
(!)
Z
<t 0--0 NEW ELANDS
a::
0 *---.. STAR

Pb Rb Th K Nb Ce Nd Sm Zr Ti Yb
Cs 80 U To La Sr P Hf Eu Y Lu

Figure 3.18. Incompatible element distribution diagrams for orangeites from Swartruggens, New Elands, and
Star. All data this work.
GEOCHEMISTRY OF ORANGElTES l83

10 3 EVOLVED ORANGEITES
LLI
...J
I-
Z
<I
~
LLI
> 10 2
~
~
it:
a..
......
LLI
I-

-
i:ij 10
(!) 0--0 SOVER NORTH
Z
<I FINSCH
0::
0 *---... POSTMASBURG

Pb Rb Th K Nb Ce Nd Sm Zr Ti Yb
Cs So U To La Sr P Hf Eu Y Lu

Figure 3.19. Incompatible element distribution diagrams for a Finsch orangeite (Fraser 1987) and evolved
orangeites from Sover North and Postmasburg (Tainton 1992).

Figure 3.20 gives distribution patterns for "on-craton" kimberlites from the Kimber-
ley area. These exhibit negative Rb and K and positive P and Zr anomalies but show no
significant depletion in Sr. In contrast, "off-craton" kimberlites from Namibia exhibit
significant negative Sr anomalies (Figure 3.21) in addition to Rb and K depletion. Figure
3.21 also shows that olivine melilitites from Namaqualand (South Africa) have virtually
identical distribution patterns to those of Namibian kimberlites. This suggests that the
processes controlling the distribution of trace elements during the partial melting epi-
sodes, which gave rise to the Namibian kimberlites and South African melilitites, were
similar (see 4.8).
Figure 3.22 shows that distribution patterns for olivine lamproites from Ellendale
have well-defined negative K, Rb, Sr, and P, and positive Zr and Hf anomalies. Similar
patterns are evident for the Prairie Creek olivine lamproites. Note that the distribution
patterns are very different from those of silica-rich lamproites (Figure 3.23), which have
negative Th, D, Nb, Ta, and Sr anomalies, but lack K anomalies. The diagrams suggest
that different processes have been involved in the formation of olivine and phlogopite
lamproites, which reinforces the proposition of Mitchell and Bergman (1991) that olivine
lamproites cannot be parental to phlogopite lamproites. Figure 3.24 demonstrates that the
distribution patterns of other ultrapotassic volcanic rocks are similar to those of phlo-
gopite lamproites but contain additional negative Ti anomalies.
In summary, incompatible element distribution patterns for orangeites suggest they
formed neither by the same processes nor from the same sources as generate phlogopite
284 CHAPTER 3

LLJ 10
3 KIMBER LITES
...J
~
Z
<t
~
LLJ
> 10 2
E
~
a:
,
11..

LLJ
t:
...J 10
0:: 0--0 DE BEERS

-
LLJ
CD
~
*----* DUTOITSPAN
WESSELTON
~

Pb Rb Th K Nb Ce Nd sm Zr Ti Yb
Cs Ba U To La sr P Hf Eu Y Lu
Figure 3.20. Incompatible element distribution diagrams for kimberlites from the Kimberley area. All data
from Clement (1982) and Mitchell and Clement (unpublished).

- K35 BERSEBA RESERVE 2


0--0 GAMOEP MELILITITE

Pb Rb Th K Nb Ce Nd sm Zr Ti Yb
Cs Ba U To La sr P Hf Eu Y Lu

Figure 3.21. Incompatible element distribution diagrams for an off-craton Namibian kimberlite (Spriggs 1988)
and a South African melilitite (Rogers et al.1992).
GEOCHEMISTRY OF ORANGElTES 285

w
>
E
,I
I I
I
~
'*,
I I
~ II
~ 10 2
Q. *
......
w
t-
O
a::

-
Q.
~ 10
<t ELLENDALE 4
..J
W 0--0 ELLENDALE 9
z
:> *---",* PRAIRIE CREEK
::J
0

Pb Rb Th K Nb Ce Nd Sm Zr Ti Yb
Cs Ba U Ta La Sr P Hf Eu Y Lu
Figure 3.22. Incompatible element distribution diagrams for Ellendale (Jaques et aJ. 1986) and Prairie Creek
(Fraser 1987) olivine lamproites.

W
..J
I- 10 3
Z
<t
~
w
~ 2
!:: 10
~
iE
Q.
.......
w
I- 10
0
a::
Q.
~
<t
..J
Cs Rb Th K Nb Ce P F Zr Ti Yb
Pb So U To La Sr Nd Hf Sm Y Lu

Figure 3.23. Composite incompatible element distribution diagram for phlogopite lamproites (after Mitchell
and Bergman 1991, Figure 7.29).
286 CHAPTER 3

kI
-oJ
~ 10 3
Z
<
:t
kI
>
~ 10 2
i
ii:
....a..
en
~ 10
<
-oJ
~
a..
0:

Ba K Nb Ce Nd Hf Sm Tb
Rb Th To La Sr P Zr Ti

Figure 3.24. Composite incompatible element distribution diagram for high-potassium Roman Province type
(RPT) lavas from Monti Ernici and Vulsini. Italy (after PeccerilJo et al. 1988).

lamproites and other continental ultrapotassic volcanic rocks. The common presence of
negative Ta-Nb-Ti anomalies in the latter indicate that their mantle sources have very
different mineralogical characteristics from those of orangeites.
Orangeites, olivine lamproites, and kimberlites are all similar in having negative Rb
and K anomalies, suggesting that similar minerals are retained in their mantle sources.
The negative Sr anomaly found in many of these diverse rocks cannot have a common
origin because of marked differences in the isotopic composition ofSr in each petrological
clan (3.8.1).
The conventional explanation of the Rb and K anomalies is that residual phlogopite
must remain in the mantle sources. The absence of K anomalies in evolved orangeites
might imply they are formed by a greater degree of partial melting, resulting in the
elimination of phlogopite as a residual phase. The Sr anomalies may be due either to the
presence of a residual Sr-bearing phase, such as a phosphate, or to intrinsic depletion of
the mantle in clinopyroxene, and hence Sr, by previous episodes of basaltic magma
formation. Potassium richterite is another possible residual phase which might retain Sr,
Rb, and K. Although Sr anomalies are present, there are no corresponding REE anomalies,
suggesting that Sr and REE are hosted by different phases in the mantle sources of these
magmas. The absence of Ta-Nb-Ti anomalies in orangeites, kimberlites, and olivine
lamproites suggests that their mantle sources do not contain residual titanates.

3.6.2. Ce/Y and LafYb versus Zr/Nh


Plots of Ce/Y versus ZrlNb are commonly used to infer the degree of melting
involved in the production of basaltic rocks from peridotitic sources. Figure 3.25
illustrates Ce/Y versus ZrlNb ratios for orangeites relative to the compositions of magmas
GEOCHEMISTRY OF ORANGEITES 287

40

30

>-
"'-
Q) 20
U
• SOVER NORTH
10 PNIEL / POSTMASBURG

2 4 6 8 10 12 14 16
Zr I Nb
Figure 3.25. Ce/Y ratio versus ZrlNb ratio for orangeites. Data sources as in Figure 3.5. B-N = Bellsbank-
Newlands; ST =Star; SWR =Swartruggens. Solid curved line from (Tainton 1992) indicates the compositions
of melts formed by various degrees (%) of equilibrium partial melting of a peridotite containing 1.4 ppm Ce,
3.45 ppm Zr, 8.51 ppm Zr, and 0.54 ppm Nb, i.e., a bulk earth composition.

formed by the partial melting of a peridotitic source. The plot shows quite clearly that
orangeites cannot be derived by single-stage partial melting of such a source (Tainton
1992). Note that evolved orangeites plot with higher ZrlNb and lower Ce/Y ratios than
unevolved orangeites, suggesting that the former are produced by greater degrees of
partial melting of a source mantle enriched in incompatible elements relative to "normal"
peridotite. Thus, Tainton (1992) interprets the trend of decreasing Ce/Y and increasing
ZrlNb from Bellsbank to Pniel as a partial melting trend. On this basis, Swartruggens
orangeites appear to represent the lowest degrees of partial melting of the orangeite source
mantle.
Figure 3.26 shows that orangeites from different intrusions define a broad hyperbolic
trend of increasing LalYb ratio with decreasing Zr/Nb ratio, which, within a given
intrusion, do not show any corresponding correlation. The conventional interpretation of
the data is that the increasing LalYb ratios represent decreasing amounts of partial
melting. Thus, evolved orangeites could be formed by greater degrees of partial melting
than unevolved varieties. Finsch is anomalous in having the geochemical characteristics
of an evolved orangeite. Note that the degree of partial melting suggested by this plot is
not the same as that deduced from Figure 3.25, e.g., Swartruggens on the basis of LalYb
ratios appears to represent a greater degree of melting than the Star and Barkly West
orangeites. An alternative explanation of this contradiction is that the Swartruggens and
Star orangeites are derived from sources with compositions different from those of the
Barkly West and Finsch orangeites. The LalYb and ZrlNb ratios are not compatible with
derivation of the parental magmas from a simple peridotite mantle source unless the
degree of partial melting is vanishingly small («<1 %).
288 CHAPTER 3

r:
.Q
Z 5
PNIEL - POSTMASBURG

.....
...
N

50 100 150 200 250 300


La / Yb •

Figure 3.26. ZrlNb ratio versus LalYb ratio for orangeites. Data sources as in Figure 3.5.

3.7. PERIDOTITE MIXING AND ASSIMILATION


The magnesium content and compatible trace element geochemistry of orangeites
(3.3, 3.4) is dominated by the presence of ubiquitous macrocrystal olivines, considered
to represent disaggregated mantle harzburgite or peridotite. Other phases derived from
these contaminants appear to have been completely dissolved (Cr-diopside, enstatite) or
mainly fractionated out (garnet, Cr-spinel). It is possible that the assimilated xenoliths
also contained phlogopite, although the effects of phlogopite assimilation are extremely
difficult to decipher from those of crystal fractionation in these mica-rich rocks (Fraser
and Hawkesworth 1992); however, they are probably insignificant as this mineral is not
abundant in common mantle peridotites «1 vol%). This conclusion will not apply if
phlogopite is a major component of the mantle sources of orangeites and residual
phlogopite (restite) is incorporated into partial melts.
Assuming orangeites are derived by small degrees of melting from a peridotite
source, Fraser and Hawksworth (1992) calculated thatthe Ni and Cr contents of the Finsch
orangeites are consistent with 5~60% entrainment of a peridotite containing 2000 ppm
Ni and 2800 ppm Cr. A major problem with this conclusion is that it is based on the
assumption of primary orangeite magmas having the same Ni and Cr contents (20~500
GEOCHEMISTRY OF ORANGEITES 289

2500 <> BELLSBANK - NEWLANDS


• SOVER
AJE25

w
99%
-- 2000
E
Q.
Q.
1500
z
1000 • SOVER POSTMASBURG
NORTH
500 -+---....,...-----r---r----r--.I
0'075 0'100 0'125 0'150 0·175

Sm/Nd
Figure 3.27. Ni versus SmlNd ratio for orangeites. Solid curve is a hypothetical mixing curve calculated for
the addition of a peridotite (AJE25) analyzed by Erlank et al. (1987) to a Bellsbank orangeite (after Tainton
1992).

ppm) as many other limited partial melts of normal mantle. The assumption is unlikely
to be correct (see 4.2.3).
Using Ni versus SmlNd ratios (Figure 3.27), Tainton (1992) has shown that incom-
patible element ratios are insensitive to mixing of garnet lherzolite with orangeite
provided that the amount of peridotite is less than 80 wt%. Orangeites from the Barkly
West region are considered to have undergone 10-80% contamination, an amount in
accord with estimates deduced in this work from the major element composition of these
rocks (3.3.1).
The effects of peridotite mixing on incompatible element abundances and ratios are
easier to estimate as normal peridotites have very low contents of these elements (<5 ppm
Rb, <30 ppm Sr, <10 ppm Zr) even relative to the amounts found in contaminated
orangeites. Fraser and Hawkesworth (1992) have concluded that peridotite entrainment
principally dilutes the incompatible element contents of the initial melt. Hence, such melts
must have had extremely high incompatible element contents and might have contained
at least 100 ppm Nd, 1500 ppm Sr, and 400 ppm Zr, with a SmlNd ratio of 0.09 (Fraser
and Hawkesworth 1992). However, these estimates are based upon the assumption that
the initial melts, derived from normal mantle, contained 300-500 ppmNi and were mixed
with peridotite containing about 2000 ppm Ni (Figure 3.28). Regardless of the veracity
of the assumptions underlying these calculations, the conclusion has the merit ofshowing
that even extremely small «0.5%) degrees of partial melting are incapable of producing
orangeites from primitive mantle compositions (Fraser and Hawkesworth 1992).
Figure 3.29 shows that the LalNd versus SmlNd array formed by Finsch orangeites
can be modeled as a mixture of a REE-enriched melt and entrained peridotite. Fraser and
Hawkesworth (1992) consider that for small degrees of melting the SmlNd ratio of the
melt is less than that of the source, and LalNd is greater. They conclude that the melt
component must have been derived from peridotite with SmlNd ratios (0.25) different
from those of the proposed entrained peridotite (SmlNd =0.2).
290 CHAPTER 3

. . 7----/
--
,'" ....... _-- ;',
3000 - PERIDOTITES

I I ',I /'
-
E 2000
a.
a. v,,'''
I I 'I I

". " , . I
• FINSCH

KlM'r UTES
.
1000 -
/
o
MELT

I I
,I

I I
005 010 015 020 025
Sm/ Nd

Figure 3.28. Ni versus SmlNd for Finsch orangeites. Solid line represents mixtures between peridotite and
melts formed from primary mantle (after Fraser 1987).

Orangeites from Barkly West form an array plotting at a high angle to the mixing
line defined for Finsch (Figure 3.29) by Fraser and Hawkesworth (1992), and are
considered by Tainton (1992) to have been formed from melts with distinct incompatible
element ratios. Orangeites from Star, New Elands, and Swartruggens define discrete fields
different from those of Barkly West and Finsch on Figure 3.29. These data might suggest
that different orangeites are derived from sources of different incompatible element
content. Tainton (1992) considers that the correlation between La/Nd and Sm/Nd ob-
served for Finsch (Figure 3.29) is due to the relatively low abundances of these elements
in the Finsch parental magma. Hence, ratios were more susceptible to mixing processes,
and initial La/Nd ratios were considerably higher than shown in Figure 3.29.
All of the above conclusions stem from the following assumptions: the composition
of the assimilate is known; there is no significant crystal fractionation of the primary
melts; these magmas are produced by small degrees of partial melting. The latter point is
addressed further in Section 4.4.2.
As the magmas ascended from a minimum depth of 150 lan, there is ample
opportunity for sampling a very wide range of mantle material, and contribution to the
olivine macrocryst suite from many sources must occur. These may range from dunites
to di verse metasomatized peridotites. Modeling of the amounts of incompatible elements
added to the melt may not be as simple as proposed by Fraser and Hawkesworth (1992)
and Tainton (1992), as such peridotites may have been cryptically or patently metasoma-
tized (Dawson 1984). Thus, the mineralogical sites of incompatible elements in the
assimilate may vary. Preferential extraction of intergranular constituents is clearly more
probable than complete assimilation of refractory single crystals, and significant amounts
of incompatible elements may be added to the melt without concomitant Ni addition.
Regarding crystal fractionation, some important physical and thermodynamic as-
pects of assimilation are not considered in geochemical models. If peridotite xenoliths
are dissolved in small-volume magmas, heat is required to assimilate large quantities of
GEOCHEMISTRY OF ORANGEITES 291

2·5 A SOVER NORTH


• FINSCH
D POSTMASBURG

2·0

"C Peridotite
Z
" 1'5
~

1'0

Melt
0'5 -f----...----.---------r---'
0·05 0'10 0'15 0'20

Sm/Nd
Figure 3.29. La/Nd ratio versus SmlNd ratio for orangeites. Data sources as in Figure 3.5. Solid line (after
Fraser 1987) represents mixtures between peridotite and melts formed from primary mantle.

orthopyroxene. This energy must be provided by crystallization from the magma of the
primary liquidus phases stable at the P-T of assimilation. These phases may be sub-
sequently removed (or concentrated) by flow differentiation and/or gravitational frac-
tionation. The complete absence of orthopyroxene xenocrysts in orangeites indicates that
the assimilation process must be complete, with the implication that magma ascent rate
through the mantle must be relatively slow in order for this process to occur. Slow ascent
rates will provide ample opportunity for crystal fractionation of both xenocrysts and
primary minerals.
Finally, the small amounts of orangeite emplaced at high crustal levels suggests that
if the magmas have indeed assimilated significant quantities of orthopyroxene, then initial
magma volumes must have been large in order to provide the heat required for the partial
assimilation of the postulated amounts of peridotite contaminant. Much of the initial
magma, derivative melts, and many of the crystals precipitated during assimilation must
be retained in the mantle.
If the above conclusions are correct, then orangeites are unlikely to represent
unmodified primary melts. Alternatively, the orthopyroxene assimilation problem may
be avoided if orangeite magmas are simply primary melts mechanically contaminated
only by dunites. The presence of diamond and subcalcic chrome pyrope xenocrysts in
orangeites is definitive evidence that disaggregation and/or assimilation of garnet-bearing
ultramafic xenoliths has occurred (see 4.5.3).
292 CHAPTER 3

3.8. RADIOGENIC ISOTOPES


3.8.1. Strontium and Neodymium
Studies of the composition of radiogenic Sr and Nd by Smith (1983) provided the
initial impetus for the confirmation of orangeites as a distinct magma type (1.1). This and
all subsequent studies (Fraser et al. 1985, Fraser 1987, Fraser and Hawkesworth 1992,
Tainton 1992, Clarke et al. 1991, Skinner et al. 1994) consider that the measured isotopic
compositions of the bulk rocks must be very similar to those of their sources. This
assumption is based upon the very high Sr and Nd contents of the rocks and on calculations
showing that either peridotite mixing or crustal contamination will have insignificant
effects on Sr and Nd isotopic compositions. These conclusions are in accord with current
opinion (McCulloch et al. 1983, Vollmer et al. 1984, Fraser et al. 1985, Nelson et al.
1986), which holds that the isotopic compositions of mantle-derived potassic magmas of
high Sr and REE content reflect those of their mantle sources.
Interestingly, the first Sr isotopic studies of rocks from Swartruggens (Mitchell and
Crocket 1971, Allsopp and Barrett 1975) established that they possessed high 87Sr/86Sr
initial ratios. Unfortunately, the significance of these observations was not realized, as
similar high initial ratios were contemporaneously determined in contaminated and
altered archetypal kimberlites.
Despite the importance of Sr and Nd isotopic studies there are few published data
for the Swartruggens (Smith 1983) and none for the Boshof, Winburg, and Kroonstad
occurrences.
Measured and calculated initial 143Ndll44Nd ratios obtained in separate laboratories
are commonly not directly comparable owing to the use of different standards and
analytical techniques (Hawkes worth and van Calsteren 1984). In an attempt to overcome
this problem, initial Nd isotopic ratios are commonly expressed as the deviation from the
bulk earth chondritic value of 143Ndll44Nd at the time of formation (1) of the samples by
the relation

143Ndll44Nd sample initial ratio (1)


[
ENd = 143Ndll44Nd CHUR (1)

where CHUR is the isotopic composition at time (1) of a chondri tic uniform reservoir
used to represent the SmlNd ratio and isotopic compositions of the bulk earth (DePaolo
and Wasserberg 1976, O'Nions et al. 1979). Values of ENd of zero or near zero in
mantle-derived rocks indicate undifferentiated primitive mantle sources in terms oftheir
SmlNd ratios. Positive or negative values require that at least one episode of fractionation
has increased or decreased the source SmlNd relative to the chondritic ratio. The ENd
values are particularly useful when attempting to compare the isotopic compositions of
kimberlites, orangeites, and lamproites.
Table 3.11 lists the range of Sr and Nd isotopic compositions found in orangeites.
Fortunately, all Nd isotopic determinations have been corrected for instrumental isotopic
fractionation using the same 146Ndll44Nd ratio (0.7219) and are, thus, directly compara-
ble.
GEOCHEMISTRY OF ORANGEITES 293

Table 3.11. Initial Sr, Nd, and Pb Isotopic Compositions of Orangeites


87Sr/86Sr 143 Ndl l44Nd
ENd (N)
Finsch 0.70777-0.70983 0.51202-0.51217 -6.2 to-9.7 (23)
Posbnasburg/Pniel 0.70718-0.70886 0.51190-0.51216 -7.6 to-l2.5 (7)
SoverNorth 0.70713-0.70721 0.51180-0.51181 -13.1 to-l3.4 (3)
Sover 0.70741-0.70776 0.51195-0.51196 -10.1 to-lO.4 (3)
Bellsbank 0.70847-0.70896 0.51204-0.51206 -9.4 to-9.6 (3)
Newlands 0.70764-0.70773 0.51202-0.51206 -9.3 to-10.3 (2)
Prieska 0.70755-0.70868 0.51190-0.51208 (II)
NewElands 0.7074-0.7076 0.51208 (1-4)
Swartruggens 0.7090-0.7109 (4)
206 pbP04 Pb 207PbP04Pb 208 pbP04Pb (N)
Finsch 17.69-18.24 15.44-15.58 37.48-38.23 (23)
Posbnasburg/Pniel 17.298-17.539 15.453-15.517 35.967-37.681 (7)
Sover North 17.062-17.310 15.441-15.498 37.445-37.602 (3)
Sover 17.365-17.428 15.465-15.529 37.442-37.658 (4)
Bellsbank 17.440-17.542 15.487-15.537 37.540-37.736 (6)
Newlands 17.479-17.550 15.514-15.538 37.657-37.674 (3)
New Elands 17.21-17.26 15.47-15.48 (4)
Swartruggens 17.63 15.62 (1)
Data sources: Finsch (Fraser 1987, Fraser and Hawkesworth 1992); PostmasburglPnie1, Sover North, Sover, Bellsbank,
New1ands (Tainton 1992); Prieska (Skinner et al. 1994); New E1ands, Swartruggens (Smith 1983). (N) = number of samples.
All Nd isotopic data are corrected for fractionation to l~dll''Nd = 0.7219.

Table 3.1 and Figure 3.30 show that within individual orangeite intrusions there is a
limited but significant range in Sr and Nd isotopic composition not attributable to
experimental error. Isotopic differences are also recognizable between groups. Thus, the
Finsch orangeites contain relatively radiogenic Sr and Nd and appear to have been derived
from sources that had slightly greater time-averaged Sm/Nd and Rb/Sr ratios than those
of the Barkly West area. There are no correlations between the Sr and Nd isotopic
compositions of the Finsch orangeites. Within the Barkly West group the evolved Sover
North orangeites have the least enrichment in radiogenic Sr and greatest depletion in
radiogenic Nd. In contrast, evolved rocks in the Postmasburg area are slightly enriched
in radiogenic Nd relative to the Barkly West samples. The data suggest no simple
fractional crystallization or partial melting relationship exists between evolved and
unevolved orangeites. As a group, orangeites from the Barkly West-Postmasburg-Finsch
areas define a weak correlation of decreasing 87Sr/86Sr with decreasing 143Ndll44Nd.
Bellsbank orangeites show no intra-intrusion variation in 143Ndll44Nd, although Sr
isotopic compositions vary widely. Tainton (1992) has suggested that this variation may
be attributable to the introduction of ground water and carbonate country rock. Bonafide
orangeites from the Prieska area have isotopic compositions overlapping (Figure 3.30)
both the Finsch and Barkly West orangeites.
Figure 3.31 illustrates the Sr and Nd isotopic composition of orangeites relative to
those of bulk earth and archetypal kimberlites, and shows clearly that each magma type
must originate from compositionally distinct sources. Thus, Smith (1983) and subsequent
294 CHAPTER 3

0-5122 -
rpRlESKA7 - ;- - II
I • •
• rI. • •1
'"0
0-5121 -
I • .... 0.. • - -7-5
Z 1 1 0
'it
! 0-5120- I 0 ~ ·0. 1 •
1
....... .1 o. • 0 o 0 f- -10-0
'"0 IL• _ _ _ _ _ _ _ ....JI eNd
f()Z 0-5119 -
'it
o BELLSBANK a NEWLANDS t- -12-5
• SOVER
0-5118 - o PNIEL a PK35
... SOVER NORTH a PK 36
PK37 t::.
_ l-_----r_ _-._---L...,..._F_I_NrSC_H_--._ _- ._ _,....t- -15-0
0-5117 I I I I I I I
0-7065 0-7075 0-7095

Figure 3.30. 143Ndll44Nd versus 87 Sr;86Sr for orangeites from the Finsch (Fraser 1987) and Barldy West-
Postmasburg region (Tainton 1992)_ Field for Prieska orangeites after Skinner et ai_ (1994).

workers have concluded that orangeites are derived from ancient sources with greater
Rb/Sr and lesser SmlNd ratios than those of bulk earth. Kimberlites, on the other hand,
appear to be derived from sources with compositions close to that of bulk earth or which
have higher SmlNd and lower Rb/Sr ratios.

- o o KIMBERLITE
0-5128 - • ORANGEITE
- o ANOMALOUS
~ 0,5126- PRIESKA
'it -
/+~
! 0'5124- o
"~ -
BULK EARTH 0

.:......~"\ ..
0'5122- n.......n .'
'-1...J:T""

.'...••:..:.:,.'.
•• ' .
f() -

'it 0'5120-
- .
0'5118 - '" •
I I I I I I I I
0'704 0'706 0'708 0'710

87Sr / 86 Sr

Figure 3.31. 143Ndll44Nd versus 87Sr;86Sr for kimberlites (Smith 1983, Skinner et al. 1994),orangeites
(Fraser 1987, Tainton 1992, Skinner et al. 1994), and anomalous Prieska samples (Skinner et al. 1994).
GEOCHEMISTRY OF ORANGElTES 295

The conventional interpretation of the isotopic data (Smith 1983) is that the orangeite
sources were isolated for 1-2 billion years prior to the partial melting events which led
to the formation of orangeite magmas. These sources are considered to be located in the
nonconvecting continental lithospheric mantle. In contrast, kimberlites are believed to be
derived from convecting asthenospheric mantle. Note that the data set in Figure 3.31
includes "on"- and "off'-craton kimberlites. These cannot be distinguished from each
other on an isotopic basis. This observation indicates derivation of the magma from
similar sources and the absence of contamination of "on" -craton kimberlites by ancient
radiogenic Sr-enriched cratonic crust during emplacement.
Figure 3.31 also shows that some rocks from the Prieska area have isotopic signatures
which are intermediate between those ofkimberlites and orangeites. Clarke et al. (1991)
and Skinner et al. (1994) refer to these rocks as "transitional kimberlites." They occur
primarily in domain V of the Prieska region (Figure 1.12). Petrographically, they differ
from bona fide orangeites in containing relatively abundant, coarse-grained, primary
spinels, and perovskite. Skinner et al. (1994) and Skinner (1989) consider these features
place the rocks as petrographically transitional between kimberlites and orangeites.
However, the rocks have other characteristics, such as the presence of amphibole and
sanidine, linking them to orangeites. Lacking detailed mineralogical investigations of
these rocks, it would seem unwise and premature to recognize a new category of
"transitional kimberlites" based upon the grain size of the ground mass minerals and only
six determinations of isotopic composition. Further studies of these rocks are desirable.

PRAIRIE
CO;
0
MADUPITIC
LAMPROITE(
LEUCITE I'
-16 HILLS ~'\

G:J,G:J~ ,ORANGEITES
-24 O PHLOGOPITE LAMPROITE
_ SMOKY LEUCITE HILLS
BUTTE

0·700 0·705 0·710 0·715 0·720

875r/865r
Figure 3.32. Isotopic composition of Nd and Sr in orangeites. lamproites. kimberlites. and potassic volcanic
rocks. Data sources as Figure 3.31 and Mitchell and Bergman (1991).
296 CHAPTER 3

As a group, orangeites exhibit limited 87Sr/86Sr (0.707-0.710) and 143Nd/I44Nd


(0.5118-0.5122) compared to many other mantle-derived potassic volcanic rocks (Figure
3.32). They share with these magmas the trait of derivation from ancient sources, having
lower Sm/Nd ratios and greater Rb/Sr ratios than bulk earth. Figure 3.32 shows that the
majority of orangeite isotopic compositions plot on the West Kimberley/Murcia-Almeria
isotopic trend, close to the composition of Ellendale olivine lamproites. Others (Sover
North, Postmasburg, Sover, Bellsbank) plot between the West Kimberley trend and that
defined by North American lamproites. These data suggest derivation of orangeites from
sources of varying Sm/Nd and Rb/Sr ratios, intermediate in character between those of
the low Rb/Sr and Sm/Nd North American lamproites and the high Rb/Sr, low Sm/Nd of
the Australian lamproites.
Isotopic data may be intetpreted to suggest that intra- and inter-intrusion variations
represent derivation from isotopically heterogeneous sources (Bergman 1987, Tainton
1992). In this case discrete domains of diverse Rb/Sr and Sm/Nd would have existed for
long time periods (1-2 Ga). Partial melting of these regions, without mixing of the
derivative magmas, is then required to explain the observed isotopic variation. While this
process may be applicable to large lithospheric domains and regional inter-intrusion
isotopic differences, it would seem less so to intra-intrusion variations, as considerable
magma mixing must occur during their generation and emplacement.
An alternative explanation of the isotopic variations proposes that they are the
products of mixing of two (or more) components of radically different isotopic compo-
sition (McCulloch et al. 1983, Vollmer et al. 1984, Mitchell et al. 1987). In this case the
isotopic compositions do not necessarily represent those of their source rocks, as observed
compositions are intetpreted as mixtures between depleted and enriched end-member
components. The sources of these components may be found in the asthenosphere and
the lithosphere, respectively. Hence, it is possible that the addition of asthenospheric
material to the lithosphere triggers partial melting of ancient Rb- and light REE-enriched
zones, leading to the formation of diverse potassic magmas containing hybridized Sr and
Nd (see 4.5.2).
Isotopic variation within the Finsch orangeites is attributed by Fraser (1987) and
Fraser and Hawkesworth (1992) to mixing of melts derived from enriched source regions
with overlying depleted mantle. However, neither details of the isotopic mixing processes
nor the cause of partial melting are specified.

3.8.2. Lead
The isotopic composition of Pb in orangeites (Table 3.11) is unradiogenic with
respect to 206PbP04Pb ratios and plots to the left of the geochron and slightly below the
Stacey and Kramers (1975) two-stage Pb growth curve in Figure 3.33. Each intrusion
differs with respect to 206PbP04Pb-207PbP04Pb and 208PbP04Pb-206pb/204Pb ratios, and
samples define distinct linear arrays on Figure 3.33. These arrays do not represent
errorchrons, isochrons, or chords (anomalous Pb lines) to the growth curve.
Figures 3.33 and 3.34 also show that the isotopic composition of Pb in archetypal
kimberlites is distinct from that of orangeites, implying that their parental magmas
GEOCHEMISTRY OF ORANGEITES 297

15'8
W. KIMBERLEY 0
GAUSS BERG -0
..c 15·6 LEUCITE HILLS
a..
v
oC\J ORANGEITES

-,
_.--
/
" 15-4 PRAIRIE CREEK (f
.
15·60
..c
,....a..
o
C\J 15.2 fit
N.~
,\)~
I....~ C;
c;;? 15'50
.,...,..
.,. .,. 4.l,. ·
<J;-~ SMOKY • ••
o BUTTE 15-45
• 6.

15-40
15'0

14 15 16 17 18 19 20

Figure 3.33. 207 pbP04 pb versus 206 pbP04 pb for orangeites (Smith 1983, Fraser 1987, Tainton 1992, Skinner
et al. 1994), southern African kimberlites (Smith 1983), and W. Kimberley, Murcia-Almeria, Gaussberg,
Leucite Hills, Prairie Creek, Smoky Butte, and Sisimiut lamproites (Nelson 1989, Alibert and Albarede 1988,
Fraser 1987, Nelson et al. 1986, Fraser et al. 1985). Growth curve is from Stacey and Kramers (1975). Field
of compositions of basalts from mid-oceanic ridges (MORB) and oceanic islands (OIB) is from Fraser (1987).
Inset shows isotopic arrays relative to the growth curve (dashed line) for suites of orangeites from Finsch and
the Barkly West region (after Tainton 1992). Symbols as in Figure 3.30.

originated from sources with very different U/Pb and Th/Pb ratios and Pb evolutionary
histories.
Figure 3.33 shows that the only other potassic magmas which have Pb isotopic
compositions similar to those of orangeites are lamproites from the Leucite Hills. The
major problem in interpreting these Pb isotopic data is explaining the combination of
unradiogenic 206PbP04Pb ratios with relatively radiogenic 207PbP04 Pb ratios. Unfortu-
nately, interpretation ofPb isotopic data is not unambiguous and subject to the prejudices
of the interpreter. Hence, some geochemists favor mixing models over multistage growth
models, and vice versa. Common to most interpretations is the conclusion that orangeite
and lamproite Pb that plot to the left of the geochron have undergone evolution in a region
of low UlPb for a significant length of time.
Tainton (1992), Fraser (1987), and Fraser and Hawkesworth (1992) consider that the
intra-intrusion Pb isotopic arrays do not result from mixing of mantle Pb with lower
crustal Pb derived from granulites, as the latter does not have sufficiently high Pb contents
and is highly radiogenic.
298 CHAPTER 3

40~---------------------------,

39
.Q
a.
~
o
N
....... 38
.Q
a. ORANGEITES
CD
o
N
37
1,0 Go
SMOKY ~ PRAIRIE CREEK
BUTTEU /

15 16 17 18 19 20

Figure 3.34. 208pbl204pb versus 206pbl204pb for orangeites. kimberlites. and lamproites. Data sources as in
Figure 3.33.

Fraser (1987) and Fraser and Hawkesworth (1992) have interpreted the Finsch Pb
array as resulting from mixing of the magma with entrained peridotite having unradio-
genic Pb isotopic compositions similar to those of diopsides separated from peridotites
by Kramers (1977, 1979). However, several isotopic varieties ofPb exist in peridotites,
and the mixing process cannot be as simple as that envisioned by Fraser (1987). Thus,
bulk rock Pb isotopic compositions of peridotites cannot equal those of diopside, as
coexisting garnet and enstatite may contain radiogenic Pb not in equilibrium with diopside
(Gunther and Jagoutz 1991). Tainton (1992) has noted that because of very low contents
of Pb in diopside «0.2-2 ppm), any bulk mixing model requires the addition of
inordinately large amounts of peridotite to generate the observed Finsch Pb isotopic array.
A further argument against this mixing model is that the Finsch Pb isotopic compositions
do not correlate with the ratios of other incompatible elements.
Tainton (1992) also considers that the Pb isotopic arrays within individual orangeites
represent mixing lines. To explain the Pb isotopic compositions a two-stage growth model
is proposed, in which differences in the UlPb ratios of the magma sources during the
second stage are required to explain the inter-intrusion isotopic differences. The Pb
isotopic arrays are then generated by the mixing of these isotopically distinct melts with
unradiogenic Pb derived from depleted peridotite. This process is not one of simple
mixing, and Tainton (1992) suggests that it is "a more complex process involving partial
re-equilibration of the lead isotopic composition of the melt with peridotite wall-rock"
(Tainton 1992, p. 147). Specific details of the physicochemical nature of this process are
not provided.

3.9. STABLE ISOTOPES


There have been few studies of the composition of hydrogen, carbon, and oxygen in
orangeites, and most have concentrated upon carbonates while the stable isotopic com-
GEOCHEMISTRY OF ORANGEITES 299

position of other minerals is unknown. Initial studies were undertaken on a few isolated
samples as part of general investigations ofthe isotopic compositions of kimberlite sensu
lato and other mantle-derived carbonates. Sheppard and Dawson (1975) give 018 0 =
10.06%0 and 013C = -7.06%0 for calcite from New Elands and 0180 = 11.43%0 and o\3C
= -5.97%0 for dolomite from Sover. The isotopic composition ofH in the carbonate-free
matrix of the Sover sample was oD = -98%0. Kobelski et af. (1979) give 018 0 of 11.04%0
and 12.23%0 and o\3C of -7.61%0 and -6.81%0 for the Star and Roberts Victor orangeites,
respectively. Too few samples were analyzed in these studies to determine whether any
significant differences exist between the stable isotope composition of archetypal kim-
berlites and orangeites.
The work of Kirkley et al. (1989) represents the only detailed investigation of the
stable isotopic composition of orangeites, and was limited to the determination of the
whole rock isotopic composition of carbon and oxygen. Unfortunately, no attempt was
made to analyze individual carbonates, thus creating a major difficulty in the interpreta-
tion of the data, as several varieties of carbonate are known to occur in the orangeites
(this work, Tainton 1992) and kimberlites (Mitchell 1994b) investigated. The problem is
especially significant for samples containing major modal amounts of dolomite and
calcite. Thus, whole rock isotopic compositions are those of mixtures which reflect the
relative proportions of these minerals. Contrary to the assertions of Kirkley et al. (1989),
the data may have absolutely no petrogenetic significance, especially if the calcite and
dolomite are not in isotopic equilibrium. This situation might arise if either mineral has
formed during post-intrusion dolomitization or serpentinization. Unfortunately, Kirkley
et al. (1989) did not undertake a complementary investigation of the carbonate mineral-
ogy of the samples analyzed.
Figure 3.35 shows that though there is wide variation in the C isotopic compositions
of orangeites, most samples fall within the limits established for primary mantle-derived
carbon. In contrast, 0 isotopic compositions exhibit significant variation and are enriched
in 180 relative to mantle carbonate values. Kirkley et al. (1989) note that the Finsch and
Bellsbank orangeites which are emplaced in dolomitic country rocks are enriched in 18 0
and depleted in \3C relative to orangeites emplaced in other terrains (Figure 3.35).
Orangeites from Swartruggens have similar isotopic compositions but are emplaced in
lavas of the Pretoria series. However, the intrusions may have passed through dolomitic
horizons at depth.
The data suggest that none of the 0 isotopic compositions represent those of the
primary magmas. This conclusion is based upon H, C, and 0 isotopic studies of
carbonatites and kimberlites (Deines 1989, Kobelski et aZ. 1979, Sheppard and Dawson
1975), which show that enrichment in 180 results from influx of meteoric water or
re-equilibration of the carbonates with magmatic waters at low temperatures.
The above interpretation of the isotopic data is in direct contrast to the hypothesis
presented by Kirkley et al. (1989)-that the isotopic compositions of Finsch, Bellsbank,
and Swartruggens are due to the assimilation of dolomite enriched in 12C and 18 0. They
suggest that the dissociation of dolomite produces C02 that mixes with mantle-derived
C02, resulting in the crystallization of the 12C_ and 180-enriched ground mass carbonates.
There is no petrographic or geochemical evidence to support the operation of this process,
especially at Swartruggens, and the temperatures of the orangeite magmas suggested by
300 CHAPTER 3

o CARBONATITES +
+
- 2 ---h~T"'7'''
I
co I
~ -4 I
U o
I
If)
CO
-6
o
o
co<?o
0 00
o
o
•00 •• 0 0
0
:,
FIELD OF
-8 KIMBERLITES

-10

5 10 15 20 25

Sl80SMOW

Figure 3.3S. a13c versus a l80 for orangeites (after Kirkley et al. 1989). Solid points are for orangeites
emplaced in dolomitic country rocks. Open circles are orangeites emplaced in non-dolomitic terrains. Crosses
are the isotopic composition of Chuniespoort dolomite. Compositional fields of carbonatites and kimberlites
are from Deines and Gold (1973). Kobelski ef al. (1979). and Kirkley ef al. (1989).

Kirkley et al. (1989) are unrealistically high (1000-125()QC). Tainton (1992) has further
noted that increasing amounts of deuteric dolomite at Bellsbank are accompanied by
decreasing amounts of serpentine. These minerals are unlikely to be in isotopic equilib-
rium with each other or primary carbonates.
Kirkley et al. (1989) further claim that orangeites emplaced in dolomitic rocks have
higher initial 87Sr/86Sr ratios than those emplaced in other rocks. While their observation
appears to be correct, it does not follow that the high ratios are a consequence of the bulk
assimilation of dolomite, as proposed to explain the e and 0 isotopic variations. Tainton
(1992) has suggested that assimilation of the amounts of dolomite required by Kirkley et
al. (1989) are petrologically unreasonable given the very low Sr contents of dolomite
relative to those of orangeites. Ground-water alteration as proposed by Barrett and Berg
(1975) provides a simpler and more realistic explanation of the high Sr isotopic ratios. In
conclusion. the observed 0 isotopic compositions of these orangeites are considered to
be due to exchange with l80-enriched ground waters derived from the dolomitic terrains
rather than bulk assimilation of dolomite.
The e and 0 isotopic compositions of orangeites are not very different from those
of archetypal kimberlites (Figure 3.35). The stable isotope composition of kimberlites
has been reviewed by Deines (1989) and Mitchell (1986). Most investigators (Deines
1989, Kobelski et al. 1979, Sheppard and Dawson 1975) have concluded that the carbon
in kimberlites is typical mantle-deri ved carbon and that the oxygen isotopic compositions
have been modified by interactions with ground water.
There are, on the basis of the existing data, absolutely no grounds to support the
hypothesis of Kirkley et al. (1989) that the mantle sources of orangeites are depleted in
I3e relative to those of kimberlites. This conclusion is based upon the wide range in the
isotopic composition of kimberlites, which completely overlap those of orangeites
GEOCHEMISTRY OF ORANGEITES 301

(Figure 3.35), and the above observation that there is no reason to believe that the whole
rock C isotopic composition of orangeites represents that of the primary magma. Further,
Kirkley et al. (1989) do not provide either calculations showing the distribution of carbon
between carbon-bearing compounds during formation and evolution of the magma as a
function of redox conditions, or information on the mineralogical control on whole rock
isotopic compositions. Provision of these data is essential if the origins of the variations
in the C and 0 isotopic compositions of orangeites are to be understood.

3.10. SUMMARY
Orangeites may be divided into unevolved and evolved types on the basis of their
bulk rock major element geochemistry. Unevolved orangeites are ultrapotassic, peralka-
line, perpotassic, ultrabasic rocks characterized by high and variable C02 and H20
contents. Bulk rock compositions are controlled by accumulation and fractionation of
primary minerals and contamination with mantle-derived peridotites and cannot represent
those of parental liquids. Evolved orangeites are richer in Si02 (41-47 wt%) than
unevolved orangeites «40 wt% Si02) and are not characteristically peralkaline. Bulk
rock compositions do not represent those of parental liquids as a consequence of
hybridization with mantle-derived peridotites and crystal fractionation.
Orangeites are enriched in REE, Zr, Nb, Sr, Ba, Rb, and other incompatible elements,
and Ni relative to common mantle-derived magmas. Enrichment in Ni (and MgO) reflects
the presence of large amounts of xenocrystal olivine. Incompatible element abundances
vary widely as a result of variations in the modal amounts of primary phlogopite, apatite,
and carbonate and/or xenocrystal olivine. Thus, high phlogopite contents result in whole
rock compositions characterized by high abundances of Rb and Pb, whereas apatite- and
carbonate-rich rocks are relatively richer in Sr and REE. Sequestration of these trace
elements in early crystallizing primary phases implies that they are not strictly incompat-
ible elements during the crystallization of orangeites. The presence of xenocrystal olivine
merely serves to dilute incompatible element abundances and has no effect upon their
ratios or isotopic composition. All orangeites are strongly enriched in the light REE
(typically La/Yb > 100). Chondrite normalized REE distribution patterns do not exhibit
any Eu or Ce anomalies. Primitive mantle normalized extended incompatible element
distribution diagrams display significant negative K, Rb, and Sr anomalies, but lack the
negative Ta, Nb, and Ti anomalies characteristic of many lamproites and potassic rocks.
Sr and Nd isotopic studies suggest that orangeites have been derived from ancient sources
with lower SmlNd and higher Rb/Sr ratios than those of bulk earth. These sources are
considered to be located in nonconvecting lithospheric mantle. Pb in orangeites is notably
unradiogenic, suggesting evolution in a regime of low U/Pb ratio for an extended time.
No simple relationships between evolved and unevolved orangeites can be recognized on
the basis of trace element and isotopic studies. However, evolved orangeites appear to
have lower REE and Nb contents, La/Yb ratios, Pb/Ce ratios, and higher ZrlNb ratios
than unevolved types.
Unevolved C02-rich orangeites are geochemically unlike kimberlites and olivine
lamproites with respect to their major and trace element geochemistry. Evolved orangeites
are similar in their major element geochemistry to olivine and/or madupitic lamproites,
302 CHAPTER 3

but may be distinguished from these on the basis of the higher Ba and Zr contents of the
latter.
Orangeites, in terms of their isotopic compositions, differ greatly from archetypal
kimberlites and most other asthenosphere-derived basaltoid magmas. Isotopically they
are similar to lamproites and other continental potassic volcanic rocks, implying deriva-
tion from similar metasomatically enriched lithospheric mantle sources.
The Finsch orangeites are anomalous in that they are petrographically similar to
unevolved orangeites, whereas their geochemical signature is that of an evolved
orangeite.
Whether you are really right or not doesn't matter; it sthe
belief that coullts.
Robertson Davies

Omnia mutantur, nos et mutamur in illis.


(Anon)

PETROGENESIS OF ORANGEITES
AND KIMBERLITES

Many years of study have not resulted in general agreement concerning the petrogenesis
of continental alkaline rocks. Unfortunately, petrologists, acting from a stance of promoting
their own particular bias, be it geophysical, experimental, geochemical, or tectonic, have
devised contradictory models for a given magma type. Once introduced, such models are
commonly promoted by their originators without impartial reflection on all aspects of the
problem or consideration of competing hypotheses.
This unsatisfactory situation arises from the absence of modern analogues of many
varieties of alkaline magmatism and the realization that the sources of alkaline magmas
may be developed by a wide range of physicochemical processes operating at inaccessible
depths in the mantle. Thus, there are few constraints upon the imagination of petrologists
seeking to explain the genesis of alkaline rocks and the subject might be regarded as being
closer to "petromancy" than an exact science.
Kimberlites have remained challenging objects for petrogenetic speculation since
their discovery over 100 years ago. Hypotheses for their genesis have ranged from the
bizarre, such as meteorite-generated electrical discharges (Khazanovich-Vulf 1991) and
mobilized sedimentary breccias (Mikheyenko 1977), to the commonplace, such as partial
melts of asthenospheric (Canil and Scarfe 1990) or lithospheric mantle (Foley 1988,
Bailey 1993). Consequently, a comprehensive petrogenetic hypothesis for kimberlites has
not yet been devised.
Orangeites have not previously been considered as a distinct magma type; conse-
quently their genesis has been discussed as though they are merely varieties of kimberlite
(sensu lato).
This chapter reviews hypotheses and evidence pertaining to the genesis of kimber-
lites and orangeites and presents some petrogenetic speculations for both magma types,
based on the conclusions of the review. It is unrealistic to expect these speculations to
represent the final word on the topic, and, undoubtedly, the petrogenetic schemes for both
magma types will be revised in the light of further studies.
303
304 CHAPI'ER4

4.1. GEOCHEMICAL MODELS OF ORANGEITE PETROGENESIS


INVOLVING LIMITED PARTIAL MELTING OF LHERZOLITIC
SOURCES
During the 1970s geochemical studies of a variety of mantle-derived alkaline rocks,
coupled with the recognition of metasomatized mantle-derived xenoliths, demonstrated
that partial melting of a simple four-phase garnet lherzolite source rock was inadequate
as a means of explaining their incompatible element contents. Consequently, older
hypotheses invoking high degrees (15-30%) of partial melting of mantle lherzolite
followed by extensive fractional crystallization (O'Hara and Yoder 1967, MacGregor
1970, Anderson 1984) were overthrown in favor of models favoring small «10%)
amounts of partial melting of fi ve- or six -phase lherzolite sources, e.g., phlogopite garnet
lherzolite (Dawson 1972). The presence of minerals such as phlogopite or amphibole was
required in the sources to support the geochemical claims that incompatible elements
must be sequestered in hydrous phases. The various hypotheses differ with respect to the
mineralogy of the source rocks, the extent of partial melting, and the origins of the
incompatible element-bearing phases. Fractional crystallization is not considered to play
a significant role in the evolution of these magmas. Thus, it is widely believed that
observed, whole rock, incompatible element abundances and distribution patterns can be
related to the degree of melting and type of source rocks by geochemical modeling, using
the forward or inverse approaches.
Most of the partial-melting models are variants of hypotheses developed to explain
the geochemistry of oceanic basaltic rocks, and their relevance to exotic continental
alkaline magmas has not been seriously questioned by their proponents. Typically, the
hypotheses are rooted in geochemical models based on trace element abundances,
especially the rare earth elements, and mineralogical and major element constraints are
rarely considered.

4.1.1. Earlier Hypotheses


Earlier hypotheses, suggesting that "kimberlitic" (sensu lato) magmas are produced
by small amounts of partial melting of lherzolitic sources in the mantle, have been
reviewed by Mitchell (1986). Most of these hypotheses were presented prior to the
mineralogical and isotopic studies, summarized in this monograph, which have demon-
strated that "kimberlites" (sensu lato) are not a single group of cogenetic rocks. Conse-
quently, common to all of these hypotheses is the assumption that kimberlites and
orangeites are derived from the same source rocks by differing degrees of partial melting.
Clearly, in the light of current studies this assumption is no longer valid.
Hypotheses presented by Mitchell and Brunfelt (1975), Paul et al. (1975), and Cullers
et al. (1982) were devised to explain the potassium contents and high LalYb ratios of
"kimberlites" (sensu lato). These hypotheses demonstrated that the requisite LalYb ratios
could be produced by extremely small (<l %) degrees of partial melting of phlogopite
garnet lherzolite sources with chondri tic-like REE abundances. The generally higher
LalYb ratios of "micaceous kimberlites," i.e., orangeites, were considered to require
lesser amounts of partial melting (0.3-0.4%) than mica-poor kimberlites (0.7-0.9%).
PETROGENESIS OF ORANGEITES AND KIMBERLITES 305

Apart from the absence of mineralogical and major element constraints, all of the
above hypotheses were unable to model the observed abundances of trace elements in the
rocks, and differentiation or other processes had to be appealed to, to remedy this
deficiency. Subsequently, enrichment of the source in REE above chondri tic abundances
by cryptic metasomatism (Dawson 1984, Mitchell and Carswell 1976) was considered
as a possible means of producing melts at the lower limits of the spectrum of "kimberlite"
REE abundances (Mitchell 1986). However, this model still required very small (0.5-
0.2%) amounts of partial melting of the source.
The very low volumes of postulated melts have typically been regarded as a major
disadvantage of the above models. Consequently, there have been doubts that such melts
could ever segregate and escape from their mantle source regions. However, recent
theoretical and experimental studies (McKenzie 1989, Hunter and McKenzie 1989) of
the migration of small volume melts in the mantle indicate that these concerns may be
groundless.
A significant conclusion of the earlier studies was that, in addition to phlogopite, a
phosphate and/or titanate must be present in the source rocks of kimberlitic magmas to
account for their P, Sr, REE, Th, U, Zr, Nb, and Ta contents (Fesq et al. 1975, Kable et
al. 1975). This observation led Mitchell (1986) to propose that "kimberlitic" (sensu lato)
magmas could be derived by 1-8% partial melting of a patently metasomatized mantle
source (Dawson 1984) consisting of an apatite- and K-Ti-richterite-bearing garnet lher-
zolite. In this model the high LalYb ratios of the partial melts were believed to reflect
those of their source, rather than resulting from element fractionation during melting of
a low LalYb source. The relatively high volumes of the melts produced, compared to
those required by melting of cryptically metasomatized garnet lherzolite, alleviated the
supposed melt extraction problems.
In summary, although most of the earlier partial melting models are now primarily
of historical interest, they were important in that they indicated kimberlites (and other
alkaline rocks) must be produced by the melting of metasomatized mantle.

4.1.2. Melting of Enriched Mantle and Peridotite Entrainment


The first model concerned specifically with the petrogenesis of orangeites (referred
to as "group 2 ultrapotassic kimberlites") was initially presented by Fraser (1987) and,
ultimately, by Fraser and Hawkesworth (1992). The model is based on geochemical and
isotopic studies of the Finsch intrusion undertaken by Fraser (1987) and Fraser et al.
(1985). These data have been reviewed in Sections 3.7 and 3.8.
Fraser and Hawkesworth (1992) consider that the geochemical characteristics of the
Finsch rocks result from the mixing of a melt component with 50-60% of entrained garnet
lherzolite. The high Nd and Sr contents, low Sm/Nd ratios, and high TalYb ratios relative
to continental crust and local country rocks are interpreted as indicating that contamina-
tion of the magmas by crustal material is not significant.
The melt component is considered to have very high incompatible element contents
and be formed by very small « 1%) degrees of melting of a source enriched in these
elements relative to depleted or primitive mantle. Fraser and Hawkesworth (1992)
demonstrated that the melt component dominates the incompatible trace element content
306 CBAPTER4

of the hybrid magma. Therefore, the Nd, Sr, and Pb isotopic ratios are considered to reflect
those of the source. The isotopic data (see 3.8) are interpreted to suggest that the melt was
derived from ancient trace-element-enriched portions of the upper mantle, most probably
situated within the subcontinental mantle lithosphere. Fraser and Hawkesworth (1992),
in agreement with Wyllie (1980) and Mitchell (1986), regard this source as a phlogopite
magnesite peridotite.
The disaggregation and incorporation of old depleted lithospheric mantle into the
melt, during ascent, generates the high compatible-element contents of the hybrid magma
and the olivine macrocrysts. In contrast to Tainton and McKenzie (1994), Fraser (1987)
noted that xenoliths which might represent source material are not present in the magma.
Fraser et al. (1985) suggested that enrichment of the source in incompatible elements
was due to introduction of H20-rich fluids. However, Fraser (1987) stated that the origins
of the trace element enrichment are unresolved and compatible with derivation from either
recycled pelagic sediments or intra-mantle processes involving the migration and crys-
tallization of small volume silicate melts. Fraser and Hawkesworth (1992) do not discuss
the nature of the enrichment process.
Fraser (1987) considered that the Finsch intrusion and other orangeites were em-
placed in response to the opening of the South Atlantic (see4.5.2.1). Thus, the propagation
of the oceanic extensional tectonic regime into the subcontinental mantle resulted in small
degrees of decompressional melting of the ancient enriched lithospheric sources. Fraser
(1987) proposed a two-stage model of melt migration. An initial slow ascent of melt along
grain boundaries into overlying depleted lherzolitic mantle, followed by rapid ascent of
melt plus entrained lherzolite along major cracks produced during uplift of the litho-
sphere.
The major contributions to orangeite petrology by Fraser (1987) and Fraser and
Hawkesworth (1992) were the realization that orangeites are hybrid rocks, whole rock
compositions do not represent those of the magma, and parental melts are most probably
derived from ancient lithospheric sources. However, the partial-melting model suggested
provided no advances as it merely reiterated the earlier models, discussed above, in
requiring very small degrees of melting of a cryptically metasomatized five-phase
lherzolitic source.
Fraser and Hawkesworth (1992) also concluded that their model of hybridization by
peridotite entrainment is directly applicable to archetypal kimberlites. Following Smith
(1983), Fraser and Hawkesworth (1992) suggestthat the melt component is derived from
asthenospheric sources.

4.1.3. Three-Stage Processes-Depletion, Enrichment, and Melting


Tainton (1992) and Tainton and McKenzie (1994) have presented a general model
to explain the geochemistry of "kimberlites and lamproites." Note, in the latter paper
unevolved orangeites are termed "group II kimberlites," the Sover North evolved
orangeite is termed "lamproite," and the only bona fide lamproite considered is the
atypical, altered, and contaminated (Jaques et al. 1989b) Argyle lamproite. In their model,
Tainton and McKenzie (1994) make no real distinction between archetypal kimberlites
and orangeites.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 307

The approach used is a variant of the inversion model of McKenzie and O'Nions
(1991), devised to explain the trace element characteristics of basaltic rocks formed by
partial melting of a primitive garnet lherzolite source. The model primarily uses REE
distribution patterns, in conjunction with trace element crystal-liquid distribution coef-
ficients, to estimate the degree of partial melting, depth of melting, and the modal
mineralogy of the residue. The model was modified by Tainton and McKenzie (1994) by
the addition of phlogopite, apatite, and chrome spinel to the source.
In agreement with Fraser (1987), Tainton (1992) considers orangeites and kimberlites
as hybrid rocks and that initial partial melts have been contaminated by entrainment of
mantle-derived peridotite (see 3.7). Tainton and McKenzie (1994) assume that the
observed incompatible element contents of orangeites and kimberlites may be corrected
for hybridization, and the corrected composition represents that of the initial partial melts.
The amount of contamination is estimated by plotting the Ni contents of bulkrocks against
their macrocrystal olivine content and by assuming that the contaminant contained 2500
ppm Ni (Tainton 1992). The calculated Ni concentration of the melt for zero contamina-
tion is then used to devise a factor which is applied to the concentrations of incompatible
elements to correct for the dilution by "olivine."
Using these corrected data, Tainton and McKenzie (1994) show that the trace element
geochemistry of orangeites and kimberlites cannot be modeled by simple partial melting
(0.3-0.4%) of a primitive mantle or MORB-type garnet lherzolite source. Consequently,
they propose a three-stage model in which an initial stage of extensive partial melting,
resulting in the depletion and fractionation of incompatible elements, is followed by
"metasomatic enrichment" involving addition of an incompatible element-rich melt.
Orangeites and kimberlites are then produced by the partial melting of this derivative-
enriched source.
As an example of their model, Tainton and McKenzie (1994) calculate that the trace
element geochemistry of the Bellsbank orangeites can be reproduced by 24% melting of
garnet-bearing primitive mantle, followed by adding to this depleted material 7% of a
metasomatic melt generated by 0.5-0.3% melting of a MORB-type source, and sub-
sequent removal of 0.4-0.3% melt from this second-stage enriched source. The mineral-
ogy of the latter is calculated to consist of 72% olivine, 24% orthopyroxene, 1.3%
clinopyroxene, and 1.8% garnet. Broadly similar results are obtained for other orangeites,
kimberlites, and Argyle lamproite (see Table 1 of Tainton and MacKenzie 1994, Tainton
1992). Each magma-forming event differs only in the degree of initial depletion and
amount of secondary enrichment. The initial depletion event, involving liquids in equi-
librium with garnet, is required to explain the heavy REE abundances of the rocks. Small
amounts of apatite in the source are considered not likely to have an important effect on
REE contents of melts, although it is recognized that phlogopite plays a significant role
in controlling alkali element abundances. In all cases examined, with the exception of
Sover North and Finsch, Tainton and McKenzie (1994) calculate that the observed content
of potassium is less than calculated. Hence, phlogopite is required in the source rocks and
may remain in the residua.
Tainton and McKenzie (1994) propose that the source of the evolved Sover North
orangeites underwent 20% melting at 90-km depth, followed by the addition of 10%
enriched melt, whereas the source of the unevolved Sover orangeites experienced 20%
308 CHAPTER 4

melting at 83-km depth and only 6% secondary enrichment. Both magmas were sub-
sequently generated by 0.3-0.4% melting of these enriched sources. It follows from these
conclusions that this scheme does not admit the possibility that Sover North rocks might
be formed from a differentiate of an unevolved orangeite magma.
The initial extensive partial melting which generated the depleted mantle is believed
by Tainton and McKenzie (1994) to represent extraction ofkomatiite during the Archean.
Isotopic constraints for orangeites (and lamproites) suggest that secondary enrichment of
the depleted material also occurred in Archean or Proterozoic times, and the enriched
source was subsequently isolated from convecting mantle for a long time. The origins of
the secondary enriching melts from MORB-type sources are ultimately placed in the
asthenosphere, and Tainton and McKenzie (1994, p.813) state that "although the total
amount of melt added during enrichment is well-constrained, there are no thermal or
geochemical restrictions on the number of events involved."
A surprising conclusion of Tainton and McKenzie (1994) is that incompatible
element-enriched alkaline rocks could be derived from depleted harzburgitic sources.
Tainton and McKenzie (1994) further claim that the common, granular-depleted garnet
phlogopite peridotite xenoliths present in kimberlites are examples of the source rocks of
kimberlite and orangeite magmas. This conclusion is at variance with the commonly held
view that these xenoliths are merely mantle-derived xenoliths unrelated to their host
magmas. Arguments presented by Tainton and McKenzie (1994) for their conclusion rely
upon the proposed small volumes of the melts involved. Thus, in their model, because
melting is limited to grain boundaries, source rock entrainment must be a necessary
consequence of melt separation from these residua. Tainton and McKenzie (1994)
consider that such garnet lherwlite xenoliths have modes and major and trace element
compositions similar to those calculated for their postulated sources.
An important consequence of this model is that the difference in isotopic composi-
tions of kimberlites and orangeites requires source enrichment to have occurred at
different times but not by different processes. Thus, Tainton and McKenzie (1994) believe
that both magmas are generated by the same processes from the same source rocks and
that archetypal kimberlites could also be lithospheric melts. This conclusion is not in
agreement with hypotheses advanced by other petrologists regarding the depth of kim-
berlite generation (see Smith 1983, Canil and Scarfe 1990, Ringwood et al. 1992, Edgar
and Charbonneau 1993, Haggerty 1994, this work).
Tainton and McKenzie's (1994) hypothesis follows an approach devised for common
basaltic rocks and is therefore unlikely to be appropriate for alkaline rocks of extreme
composition, such as orangeites and lamproites. It may be relevant to the generation of
archetypal kimberlites, but not in the tectonic setting visualized by the authors (see 4.6.1).
Mineralogical and petrological evidence suggests the hypothesis is based upon incorrect
premises. The major problem being the lack of serious consideration given to sources
other than phlogopite garnet lherzolite. This a priori assumption regarding the nature of
the source necessarily predetermines the outcome of the mathematical modeling. Hence,
the latter, which is correct and internally consistent, gives the hypothesis an unwarranted
verisimilitude as a mathematically appropriate solution can always be found.
A further aspect of the model not discussed in detail by Tainton and McKenzie (1994)
is the relevance of the crystal-liquid distribution coefficients, derived from basaltic
PETROGENESIS OF ORANGEITES AND KIMBERLITES 309

magmas, to the alkaline magmas considered. The latter differ significantly in composition
from basaltic magmas, so distribution coefficients will differ correspondingly. In fact,
Tainton and McKenzie (1994, pp. 808, 811) note that the distribution coefficients
appropriate to these magmas are very poorly determined.
The hypothesis proposes extraction of a variety of distinct magma types from the
same source. Tainton and McKenzie (1994) do not provide any realistic rationale for their
choice of a lherzolitic source and do not consider any of the recent experimental and
geochemical evidence pertaining to the petrogenesis of lamproites and kimberlites (see
4.5,4.6), which indicate very different sources. For example, Foley (l992a) has noted
that models for the origin of ultrapotassic melts by partial melting of phlogopite-bearing
lherzolites are inconsistent with the array ofliquidus experimental results on ultrapotassic
compositions (see 4.2.2). This discrepancy between partial melting models of the type
advanced by Tainton and McKenzie (1994) or Fraser and Hawskesworth (1992) lies in
the invalid assumption that incompatible elements are homogeneously distributed in the
mantle source rocks.
Unfortunately, the type of geochemical modeling employed by Tainton and McKen-
zie (1994) is nondiscriminatory and may be applied with equal success to many other
alkaline magmas if no other constraints are imposed. For example, their model could be
used for such light REE-enriched rocks as kamafugites, minettes, carbonatites, or melili-
tites with similar predictable conclusions, and the genesis of these magmas could thus be
incorporated into their general model. This undesirable conclusion arises from the failure
of the geochemical models to consider other factors, such as major element compositions
of the magmas, or mineralogical and experimental evidence pertaining to their genesis
(Foley 1992b).
Finally, the hypothesis is not supported by the conclusions of this work that kimber-
lites and orangeites are distinct magma types derived from mineralogically different
sources; orangeite bulk compositions do not, even for incompatible elements, reflect those
of primary melts as a consequence of high- and low-pressure crystal-liquid fractionation.
A merit of the Tainton and McKenzie (1994) hypothesis is its focus on the Archean
events leading to the large-scale depletion of continental lithospheric mantle in incom-
patible elements and the generation of a substrate for the subsequent addition of metaso-
matic (sensu lato) material.
Tainton and McKenzie (1994) do not discuss in any detail the nature of the partial
melting processes leading to the formation of orangeite, kimberlite, and lamproite melts.
However, they do indicate that all magmas are formed in the continental lithosphere above
the mechanical boundary layer and within the garnet stability field. Partial melting is
believed to take place over a depth range of 125-175 km; hence, melts may be produced
in either the diamond or graphite stability fields.
Tainton (1992) suggests that the metasomatically-enriched sources of orangeite are
held in the lithospheric mantle at temperatures close to their solidii. The small increase
in temperature required to initiate melting is provided by advection of heat from the
fringes of the large asthenospheric mantle plume responsible for the opening of the South
Atlantic. Subsequent segregation and ascent of the melt follows the small-volume
melt-movement model of McKenzie (1989).
310 CHAPTER 4

4.2. EXPERIMENTAL EVIDENCE PERTAINING TO ORANGEITE


PETROGENESIS
There have been no experimental investigations of the phase relationships of
orangeite compositions at low pressure and only one at high pressure. Chapters 2 and 3
emphasize the mineralogical and geochemical similarities of orangeites to lamproites;
thus, some of the experimental studies directed toward understanding lamproite petro-
genesis have a direct bearing on the origin of orangeites, because of their similar
ultrapotassic peralkaline character and proposed derivation from similar sources.

4.2.1. Liquidus Experiments on Orangeite Compositions


To date, the only melting experiments on orangeite compositions have been under-
taken by Yamashita et al. (1992) and Arima et al. (1993a), using as starting material
macrocrystal and aphanitic orangeites from Makganyene. The former is, not surprisingly,
greatly enriched in MgO (30.40 wt %) relative to the latter (18.44 wt %). Makganyene
orangeites are evolved, so the samples used in the experiments are high in silica (39.27
and 40.90 wt %), relative to unevolved orangeites (see Tables 3.1 and 3.2).
For the aphanitic orangeite, the liquidus temperature is about 1470°C at 6 GPa and
1520°C at 8 GPa. Suprasolidus phase assemblages vary with increasing pressure as
follows: from phlogopite plus liquid through phlogopite plus clinopyroxene and liquid
to phlogopite, clinopyroxene, orthopyroxene, and liquid to clinopyroxene plus garnet and
liquid at 1400°C. Phlogopite breakdown occurs between 1300-1400°C and 6-7 GPa by
the reaction of phlogopite with clinopyroxene to form garnet and liquid. No stable
K-bearing phase is observed in runs above 7 GPa, suggesting that the liquid at these
pressures is extremely rich in K and volatiles.
In the experiments on the macrocrystal orangeite, olivine is stable near-liquidus
phase up to 8 GPa. It coexists with phlogopite and clinopyroxene below 4 GPa and with
clinopyroxene, orthopyroxene, and garnet above 6 GPa at 1400°C. The persistence of
olivine as a liquidus phase to high pressures is not surprising, given the enrichment of
this sample in MgO due to the presence of macrocrystal olivine.
Arima et al. (l993a) consider that the aphanitic orangeite represents a liquid com-
position and propose, on the basis of their experiments, that orangeites can be generated
by the partial melting of phlogopite-bearing lithosphere (sic) at pressures below 6.5 GPa.
As the liquids are enriched in K, those generated at higher pressures might migrate toward
the lithosphere-asthenosphere boundary and cause metasomatism, including phlogopite
formation. The near-liquidus assemblage of clinopyroxene and garnet is interpreted to
suggest that orangeites can be equilibrated, at pressures above 6.5 GPa, with an "eclogitic
source."
The value of these experiments is questionable, given the very low probability that
any bulk rock composition can represent the composition of the parental orangeite magma
(see 3.3); moreover the compositions investigated might not even be primary. However,
the phase relationships for the aphanitic sample do indicate that olivine is unlikely to be
a stable suprasolidus phase. This observation is in agreement with experimental studies
of other uItrapotassic compositions (see 4.2.3), suggesting that these magmas, at their
sources, cannot be in equilibrium with olivine-bearing mantle (Foley 1992a). Without
PETROGENESIS OF ORANGEITES AND KIMBERLITES 311

any doubt, the macrocrystal orangeite investigated is a hybrid rock containing mantle-
derived olivine xenocrysts; thus, the liquidus relationships have no direct bearing on the
genesis of orangeite.

4.2.2. Liquidus Experiments on Lamproite Compositions


These studies are important because of the similarities in mineralogy and composi-
tion of lamproites to orangeites. However, it should be realized that lamproites are rich
in F, H20, and K20 relative to orangeites, and these differences in volatile compositions
might have significant effects on liquidus phase relations (see below). As it is beyond the
scope of this work to summarize the many studies of synthetic and naturallamproites,
the reader is referred to Foley (1990, 1992a) and Mitchell and Bergman (1991) for
comprehensive reviews of the topic. The study by Luth (1967) of the system KAISi04-
Mg2Si04-Si02-H20 is particularly relevant to the postemplacement crystallization of
C02-poor evolved orangeites (see 4.5.4). This system is also important with regard to
crystallization of ultrapotassic magmas at high pressures and has been investigated at
20-28 kbpressure, with H20, C02, or F as the volatile phase, by Sekine and Wyllie (1982),
Foley et al. (1986), and Gupta and Green (1988).
The most important conclusions of the studies of synthetic systems and natural
phlogopite and leucite lamproite compositions, relevant to orangeite crystallization at low
and high pressure, are:
• Primary olivine is a low- or high-pressure, high-temperature phenocryst. Reac-
tion of primary or xenocrystal olivine with the liquid to form phlogopite may
occur during post-emplacement crystallization.
• Phlogopite is stable at high or low pressures and may crystallize as a phenocryst
during ascent of the magma. Under some conditions early-formed phlogopite
may react with liquid to form pyroxene and leucite.
• Orthopyroxene may be a high-pressure phenocryst.
• Leucite, sanidine, and amphibole are low-pressure phases.
• The near-liquidus phase assemblage of lamproite melts suggests they may be
derived from phlogopite harzhurgite or phlogopite wherlite sources, under H20-
or F-rich conditions.
However, other interpretations of these data are possible (Foley 1992b, see 4.5.2.2).
Foley (1992a), summarizing experimental data for lamproites and kamafugites, has
stressed that models for the origins of these magmas by partial melting of phlogopite
lherzolite are inconsistent with the results of liquidus experiments on ultrapotassic
compositions. This conclusion is based on the absence of olivine, and in many instances
garnet, as a near-liquidus phase. Consequently, non-peridotitic assemblages, rich in mica
and pyroxene, which in some instances may be completely free of olivine (i.e., mica
clinopyroxenite, mica websterite) are considered to be the sources of potassic rocks.
LampJ:oites are currently believed to originate from a depleted source which was
strongly re-enriched in incompatible elements at a later stage, thus producing mica
harzburgites, and subsequently melted under H20-rich reducing conditions (Foley 1988,
1989, Mitchell and Bergman 1991). Note that this three-stage model is similar in some
312 CHAPTER 4

respects to Tainton and McKenzie's (1994) model for orangeite genesis (4.1.3), the major
difference being in the composition of the second-stage source.
Foley (l992a) has reviewed the effects of volatile composition on liquidus phase
relationships and noted that, for a leucite lamproite at 2 GPa and 1150°C, decreasing
H20/C02 ratios lead, to the addition of orthopyroxene, to mica plus clinopyroxene
assemblages. For the same bulk composition in a C02-free system, with a CIWlhO ratio
of 0.22, mica is the only liquidus phase. These data suggest that oxidation and addition
of C02 result in clinopyroxene formation. In the case of kamafugites, low values of
H20/C02 lead to the formation of clinopyroxene, whereas high values result in orthopy-
roxene plus clinopyroxene and reduced olivine stability. These data demonstrate that
volatile compositions play an important role in determining the liquidus-phase assem-
blages of ultrapotassic magmas. Hence, it is suggested that one possible reason for the
existence of diopside orangeites and diopside-free orangeites may be simply one of
crystallization under differing H20/C02 ratios.
Olivine lamproites are commonly believed to represent primary magmas (Jaques et
al. 1986, Foley 1993), although Mitchell and Bergman (1991) have argued on minera-
logical and geochemical grounds that they are hybrid derivative magmas. These rocks
have some compositional and mineralogical similarities to evolved orangeites (see 3.3.5).
Regardless of their origin, liquidus experimental studies by Foley (1993) are important
in demonstrating that olivine, orthopyroxene, and phlogopite occur together at the
liquidus at pressures in excess of 5 GPa for magmas of this composition. These data are
consistent with derivation by partial melting of phlogopite harzburgite under C02-free
reducing conditions. To explain the apparent contradiction to Foley's (1992a) conclusion
that the source rocks of ultrapotassic magmas should not contain olivine, Foley (1993)
notes that the phase relationships are also consistent with melting of a veined source in
which neither vein nor wall rock consists ofphlogopite harzburgite (see 4.5.2.2). Thus,
Foley (1993) prefers an interpretation in which the magma results from preferential
melting of phlogopite-amphibole-clinopyroxene veins within a garnet lherzolite host.
The latter, to a lesser degree, is also involved in the melting process. The liquidus-phase
assemblage found in the experiments is interpreted to be that of liquids in equilibrium
with a residuum of melted vein and wall rock. Such a vein-plus-wall rock melting process
(Foley 1993, 1992b) is directly applicable to the genesis of orangeites and is discussed
further in Section 4.5.2.2. Alternatively, the persistence of olivine to high pressures in the
liquidus experiments may be a consequence of investigating a hybrid magma composi-
tion. Contamination of a MgO-poor magma at its source with xenocrystal olivine will
undoubtedly move the bulk composition of the hybrid into an olivine stability field.

4.2.3. Melting of Mica Pyroxenites


Lloyd et al. (1985) determined the compositions of liquids produced by the partial
melting of a phlogopite pyroxenite (Phh5Cpx52Ml7Sph4Ap2) at 2 and 3 GPa. Liquids
produced by 20-30% partial melting range widely in composition, but are K20 rich (4-6
wt %), Ti02rich (3-8wt %),andpoorinAh03 (6-9wt %) and Si02 (35-39 wt %).Lloyd
et al. (1985) consider that the compositions of the melts support the hypothesis that the
PETROGENESIS OFORANGElTES AND KIMBERLITES 313

potassic rocks of southwest Uganda might be formed by partial melting of phlogopite-


clinopyroxene sources.
Importantly, the study supports the hypothesis of Foley (1992a) that partial melts of
olivine-free ultramafic rocks are ultrapotassic and compositionally different from those
produced by partial melting oflherzolitic sources. Obviously, by varying the composition
of the starting material and melting conditions, partial melts equivalent to many ultrapo-
tassic magmas may be generated. Note the large degree ofmeIting in Lloyd et al.'s (1985)
study eliminates the extraction problems perceived to be associated with small degrees
of melting of lherzolitic sources.
Foley (1992a), on the basis of Lloyd et al. 's (1985) experiments, draws the important
conclusion that the mg numbers of primary magmas derived from pyroxenitic sources
may be less (60) than the values of 65-75, which are generally taken to indicate a primary
mantle-derived melt. Similarly, high Ni and Cr (200-500 ppm) contents do not necessarily

®
1500
L
""
t
"
""
1400
,," Ap + L
,,-------
--
1300

U
0 1200
PH+Ap
La.J
0: 40 50
::> 10 20 30
~
0: 1600
lLJ
® /
£l.. 2L+ Fo ./

~Fo+\
~ /
La.J
1500 L /
I-- /
: '1,L\ /
1400 \.1 \ /
/
Fo+ /
Ap+L
1300 PH + L /

10 20 30 40 50 60
- wt. % APATITE ----.
Figure 4.1. (A) Phase relations in the system hydroxy-phlogopite-hydroxy-apatite at 2 GPa. (B) Phase
relations in the system fluor-phlogopite-fluor-apatite at 2 GPa. Fo = forsterite, Ap = apatite, PH = phlogopite,
L = liquid (after Vukadinovic and Edgar 1993),
314 CHAPTER 4

imply a primitive character, as these may result primarily from contamination with
peridotitic material (see 3.7). Foley (1992a) emphasizes that criteria used to identify
primary basaltic magmas derived from Iherzolitic sources, may be misleading for exotic
alkaline magmas.

4.2.4. Phase Relations in the System: Phlogopite-Potassium Richterite-Apatite


Vukadinovic and Edgar (1993) have investigated the pseudobinary systems:
hydroxy-phlogopite-hydroxy-apatite and hydroxyfluor-phlogopite-hydroxyfluor-
phlogopite at 2 GPa pressure. In the F-free system, apatite, phlogopite, and forsterite
coexist with a K-P-rich liquid at a pseudobinary minimum at 1225°C and PhlssAplS. In
theF-bearing system this minimum occurs at 1260° at Ph166Ap34, and liquids are relatively
poorer in K and richer in P. For phlogopite-rich compositions, forsterite and phlogopite
coexist with liquid over a wide range of temperatures above those of these minima (Figure
4.1) due to the incongruent melting behavior of phlogopite.
Figure 4.2 illustrates pseudobinary phase relationships in the systems: apatite-fluor-
potassium richterite and fluor-phlogopite-fluor-potassiumrichterite (Edgar and Pizzolato
1995). Both systems contain pseudobinary minima with K-rich liquids being present
above 1150°C and I OOO°C, respectively. Orthopyroxene appears as a supraliquidus phase
due to the breakdown of richterite. The F-free analogues ofthe richterite-bearing systems
were not investigated by Edgar and Pizzolato (1995), but the data of Vukadinovic and
Edgar (1993) suggest that minimum melting temperatures will be lower in these systems.
These data show that the presence of potassium richterite lowers the melting
temperature more effectively than either phlogopite or apatite. It is to be expected that

t 1500
® LIQUID 1300
® L + PH + OPX 7 + OL?

-
- 1400 1200
U LIQUID
o +
r- 1300 LIQUID ORTHO-
PYROXENE
1100
+ +
1200 APATITE K-RICHTERITE 1000 PH + K-RICH +OPX +L
1100 +-------.l..-------i 900
PH + K-RICH
APATITE + K- RICHTERITE

50 60 70 80 90 20 40 60 80

- wt. % K - RICHTERITE ~ - Wt. % PHLOGOPITE"


Figure 4.2. (A) Phase relationships in the system fluor-apatite-fluor-potassium richterite at 2 GPa. (B) Phase
relationships in the system fluor-phlogopite-fluor-potassium richterite at 2 GPa. K-RICH = potassium
richterite, OL =olivine, OPX =orthopyroxene, PHL =phlogopite, L =liquid (after Edgar and Pizzolato 1995).
PETROGENESIS OF ORANGEITES AND KIMBERLITES 315

minimum melting temperatures in the F-free pseudoternary system phlogopite-potas-


sium richterite-apatite will be lower than lOoo°C at 2 GPa.
These studies are important with regard to the melting of mica-rich veins in the upper
mantle if the hypotheses of Foley (1 992a,b ) are correct. The effects of increasing pressure
or the addition of other components, i.e., diopside, K-Ba titanates, are as yet unknown.
However, these data indicate that partial melting of mixtures of phlogopite, amphibole,
and apatite will produce a wide range of K-rich liquids. Clearly, experimental investiga-
tion of these systems at high pressures is critical to advancing our understanding of the
formation of potassic melts in the mantle.
The primary mineral assemblage of some orangeites consists principally of phlo-
gopite, apatite, and carbonate. Their mineralogy can be in part explained if these rocks
represent the products of partial melting of phlogopite (>85 wt % )-apatite veins in a
dunitic or lherzolitic mantle (see 4.5.2).

4.3. PETROGENESIS OF ARCHETYPALKIMBERLITES-RECENT


HYPOTHESES
Orangeite magmatism is geographically and temporally associated with archetypal
kimberlite magmatism in southern Africa. To explain this relationship, some discussion
of current hypotheses of kimberlite petrogenesis is required. Earlier hypotheses have been
summarized by Mitchell (1986). Recent hypotheses have been concerned more with the
site of partial melting rather than the details of the process.

4.3.1. Carbonated Lherzolite Sources


Recent discussion of the origin of kimberlite has centered around partial melting of
carbonated lherzolite, since the discovery by Wyllie and Huang (1 975a,b ) and Eggler
(1974) that dolomite and magnesite can be stable in the mantle at solidus temperatures.
The numerous experimental studies of natural and synthetic peridotite-C02-H20 have
been reviewed by Mitchell (1986), Wyllie (1980, 1987, I 989a,b ), Eggler (1987), and
Wyllie et al. (1990). For reasons noted below, most of these studies have been devoted
to determining the topology of the peridotite-C02-H20 solidus at 1-4 GPa.
Figure 4.3 illustrates phase relationships in the system peridotite-C02. The effect of
C02 on the solidus curve is to produce a significant reduction in the solidus temperature
at pressures at which dolomite becomes stable. This occurs where the subsolidus univari-
ant carbonation reaction (reaction 6 of Wyllie and Huang 1976)

2Mg2Si04 + CaMgSh06 + 2C02 = 4MgSi03 + MgCa(C03h


fosterite diopside enstatite dolomite

intersects the solidus. At this intersection the invariant point 16 marks the change from the
solidus for peridotite-C02 to the solidus for dolomite peridotite. At higher pressures a
second subsolidus univariant reaction (/0) results in the appearance of magnesite in place
of dolomite (Figure 4.3). Melt compositions at the solidus have been inferred to be
C02-rich undersaturated silicate melts or carbonatites at pressures greater than 16, but
relatively silica rich (basanite, alkali basalt) at lower pressures. Thus, 16 is a petrogeneti-
316 CHAPTER 4

A Magnesite garnet lherzolite


B Dolomite garnet lherzolite

- OL+OPX + CPX
+GNT + MAG
@
C
D
E
Dolomite spinel lherzolite
Garnet lherzolite + C02
Spinel lherzolite + C02

-
~ 30
F Plagioclase lherzolite

::J
f/)
f/)
LLI 20
a::
a.. FIELD OF LHERZOLITE
10
+ VAPOR PHASE
VAPOR ABSENT
SUBSOLIDUS ASSEMBLAGE

800 1000 1200 1400

- TEMPERATURE (Oe) .....


Figure 4.3. Phase relationships in the system peridotite-C02 (0.1 %) based on experimental data and deduc-
tions from the system CaO-MgO-Si02-C02 (Wyl1ie and Huang 1976). Note that the carbonation reaction
terminating at 16 on the peridotite solidus divides the diagram into vapor-absent and vapor-bearing assemblages.
as all C02 is consumed in the carbonation reaction. Solidus for anhydrous peridotite (PER) is shown as a dashed
line. OL = olivine. OPX = orthopyroxene. CPX =clinopyroxene. GNT = garnet, MAG = magnesite. DOL =
dolomite. SP = spinel. PL =plagioclase. L = liquid. V = vapor.

cally important point as it marks the depth below which C02 plays a key role in generating
primary carbonatitic to low-silica C02-rich alkaline magmas such as melilitite and
kimberlite. Consequently, much effort has been devoted to locating the position of 16. The
initial experimental studies undertaken in separate laboratories were not in agreement on
the exact P-T of 16, in part due to the different experimental techniques utilized. Recent
studies are more in agreement, although it is realized that the location also depends on
other factors such as oxygen fugacity (Wyllie et al. 1990).
It is important to recognize that the location of 16 also varies with C02ilhO ratio.
The solidus depression is at a maximum in the H20-free system and the carbonation-
related minimum is, of course, absent in C02-free systems. Further, complexities in the
topology of the solidus of peridotite-H20 and peridotite-C02-H20 are introduced by the
stabilization of amphibole (see Wyllie 1987, Falloon and Green 1989, Wyllie et al. 1990).
Fortunately, for the purposes of discussion of the genesis of kimberlite, the exact
position of 16 and topology of the peridotite-C02-H20 solidus is not critical. What is of
significance is the presence of the minimum and the associated "solidus ledge" (Figure
PETROGENESIS OF ORANGEITES AND KIMBERLITES 317

4.3) which is created at lower pressures. The solidus ledge is believed by Wyllie (1980)
to play a key role in arresting the ascent of melts derived from the asthenosphere.
Crystallization of melts at this ledge results in the release of vapors, which Wyllie (1980)
sees as the principal agent of metasomatism in the upper lithospheric mantle.
Hypotheses of kimberlite genesis are based upon the existence of the solidus ledge
(Wyllie 1980, Skinner 1989, Tainton 1992) and the observation that partial melts of
carbonated lherzolite are C02-rich (Wyllie 1980, Brey et al. 1983, Canil and Scarfe 1990).
Most hypotheses based upon peridotite-C02 equilibria are models of the genesis of
"haplokimberlites," i.e., K-, Ti-, Fe-free melts approximating to postulated primitive
magma compositions in terms of their Si02, CaO, MgO, and volatile contents. Phlogopite
and other incompatible element-rich phases are not present in the experiments. It is
assumed that their presence will not seriously change the melting relationships, and
incompatible elements held in these phases will enter the melt along with carbonate.
Determination of the actual composition of the small-volume near-solidus melts produced
in these experiments is extremely difficult, as liquids quench to cryptocrystalline inter-
growths of minerals rather than glass. Compositions, estimated by a variety of indirect
methods, are all C02 rich and undersaturated and believed to range from carbonatite to
kimberlite.

4.3.1.1. Volatile Fluxing-Diapiric Model


Wyllie (1980) presented a model for the genesis and ascent of kimberlite magmas
based on volatile-induced partial melting of lherzolite. Revisions to the model have been
presented by Wyllie (1989a,b) and Wyllie et al. (1990).
The Wyllie (1980) version of the model assumes that C02, H20, and the components
of phlogopite are present as the constituents of trace magmas within a persistent melt
zone extending from 120 to 260 km depth. This zone of melting results from the
intersection of the shield geotherm, over this depth range, with the solidus of magnesite
peridotite. In the presence of volatiles above 120 km the mantIe may be composed of
hornblende or phlogopite dolomite peridotite and vapor, and below 260 km, of magnesite
peridotite and H20-rich vapor. In the absence of volatiles the mantle would be composed
of garnet lherzolite.
The starting point in Wyllie's (1980) model is the introduction of C02 and H20 to
volatile-free mantle at depths below 260 km. These volatiles, of unknown origin, are
considered to escape from deep within the asthenosphere. Partial melting of the astheno-
spheric mantle is induced at point A on the carbonated peridotite solidus (Figure 4.4).
Melting results in a density inversion, followed by the uprise of a partially melted diapir
along the adiabiatic path A-B (Wy lIie 1980) or along the geotherm A-C (Wy Ilie 1989a,b).
Within each diapir the percentage of melting increases during ascent.
At point B (Figure 4.4) the rising diapir intersects the thermal maximum on the
carbonated lherzolite solidus. If equilibrium is maintained, the magma will evolve vapor,
and crystallize completely as phlogopite dolomite peridotite plus vapor. The hot solid
diapir may continue to rise and will, upon recrossing the solidus at lower pressures, melt
to a C02-poor liquid, possibly of alkali basaltic-nephelinitic composition.
Alternatively, if the lithosphere is under tension, the evolution of vapor at B (Figure
4.4) could result in crack formation and propagation into the overlying lithosphere, as
318 CHAPl'ER4

Explosive eruption
Gas CO2-H2O
A
50
Magma crystallize
E100 B i.'

-
~

- 150 G
~
a.
Q) 200
0

250
Rising
300
500 1000 1500

B
50

-E
100

-
~

-
Graphite
150
~ Diamond
a. E
Q) 200
0

250
Rising
300
500 1000 1500
Temperature (OC)
Figure4.4. Diapiric models for the origin and eruption of kimberlite magmas (after Wyllie 1980). (A) Partially
melted diapirs rise through the mantle along the adiabatic path A-B and crystallize at the thermal maximum
on the carbonated peridotite solidus. Crystallization is accompanied by evolution of C02 vapor which initiates
crack propagation through the overlying lithosphere. followed by explosive eruption of diamond-free kimberlite
magma from point B at about 90-km depth. (B) The system as it appears once a conduit as been established to
the surface. Magmas separating from partially melted diapirs do not crystallize at the thermal maximum and
are derived from progressively deeper levels in the mantle. Eventually magmas segregate from within the
diamond stability field at point E and give rise to diamond-bearing kimberlites. See text for further explanation.
PETROGENESIS OF ORANGElTES AND KIMBERLITES 319

suggested by Anderson (1979). Kimberlites would erupt explosively as vesiculating


partially-crystallized melts, from depths of about 90 km. The initial batches of magma
erupted from 90 km will not contain diamond as they have equilibrated at depths above
the stability field of diamond. Wyllie (1980) believes that once a conduit has been
established by rising diapirs, further escape of volatiles will trigger diapiric ascent from
successively deeper levels and, ultimately, from within the diamond stability field (Figure
4.4). Thus, Wyllie (1980) suggests that successive individual intrusions within a single
kimberlite body should contain xenoliths which record increasing depths of origin.
In this model, the maximum depth of kimberlite generation must be the deepest point
at which the geotherm intersects the carbonated peridotite solidus. However, batches of
magma can ascend from any higher level above this depth once a conduit has been
established. These magmas will ascend rapidly and will not be affected by the solidus
shelf. Kimberlite magmatism in this model is analogous to a stream of bubbles rising
from an ever-deepening source (Figure 4.4).
Other batches of magma equilibrate along geotherms as they ascend and neverreach
the solidus ledge. These magmas will crystallize at the solidus at point F (Figure 4.4), and
evolution of vapors will result in metasomatism of the adjacent mantle (Figure 4.4). In
this scenario, evolution of vapor as a metasomatic agent is a consequence of kimberlite
magmatism rather than a precursor. Wyllie (1989a,b), in the revised versions of his model,
suggests such melts may accumulate at, and invade, the base of the lithosphere (Figure
4.5). Some of these kimberlites may escape via lithospheric cracks, whereas others remain
trapped in the vicinity of the lithosphere-asthenosphere boundary (Figure 4.5). Partial
melting of previously metasomatized lithosphere during subsequent thermal perturba-
tions may generate kimberlite-like magmas, i.e., orangeites.
Though Wyllie's model has been widely accepted, it has several serious drawbacks
discussed by Mitchell (1986) and Eggler (1989). The principal problems are with respect
to the absence of a zone of persistent melt beneath continental shields, doubts whether
rising magmas would intersect the solidus shelf and initiate crack propagation and conduit
formation, and lack of geological evidence to support direct eruption of several batches
of magma from deep in the mantle along a single conduit.
With regard to the latter point, WylIie's model completely ignores the known
hypabyssal infrastructure of kimberlite fields which indicate that at deep crustal levels
kimberlite magmas behave rheologically like many other basic magmas (Eggler 1989).
Diatremes are developed above dike systems only at high levels and do not extend into
the mantle. Skinner (1989) addresses these crustal emplacement problems in stage 4 of
his model (see 4.4.4).
Wyllie's model is infinitely flexible with regard to the extent of the persistent melt
zone. The magnitude of this may vary considerably, depending upon the slope of the
carbonated mantle solidus and that of the local geotherm.
Wyllie (1980) assumes that magmatism is triggered by the ascent of volatiles into a
mantle at equilibrium; if vapor is not present there is no melting. He does not discuss the
alternative that a preexisting carbonated mantle diapir rising by convection from depth
might cross the solidus and undergo decompressional melting.
Wyllie (1989a,b) introduced asthenospheric plumes into his model and noted that
kimberlites may be generated ahead of the plume during the initial stages of melting
320 CHAYfER4

o
CRUST
MOHO

- E 100

-.¥

-.s:::.

o
~ 200
.b====i>__
. -
",-
~e\\
,,'-.. \....",. -- - -- ~t\eo
" r r €,t\\tO\
--- - ~\,l((t------
" 'I ' ,,(, Solidus - - - ......
300
H~fl
"/' I I 1.1 Magma
.. , .
r:::::t Vapor
I.l:.:LI

Figure 4.5. Cross section of a craton showing lithospheric thinning resulting from upwelling of an astheno-
spheric plume. Note that partial melts are generated below the lithosphere-asthenosphere boundary. During the
initial stages of upwelling small degrees of partial melting may give rise to kimberlitic melts which might escape
through the lithosphere or remain trapped at its base. As the trapped magmas, or failed kimberlites, crystallize,
they have the potential to metasomatize the adjacent lithosphere. Subsequently, further upweIling and rifting
leads to the formation of nephelinitic magmas at shallow depths. See text for further explanation (after WyIlie
1989a,b).

(Figure 4.5). A consequence of the plume variant ofthe model is that initial emplacement
of kimberlites is followed by lithospheric thinning, rifting, and generation of alkaline
magmas above l00-km depth.

4.3.1.2. Partial Melting of Magnesite Peridotite


Brey et al. (1983), on the basis of extrapolation of liquidus experimental studies of
olivine melilitite in the presence of C02 at 3-3.5 GPa, suggested that kimberlites might
be generated by very small degrees of partial melting of a magnesite peridotite. Sub-
sequent experimental studies of near-solidus phase relationships for synthetic peridotite-
C02 systems to 12 GPa by Canit and Scarfe (1990) provided verification of the
hypothesis.
The studies demonstrate that at all pressures from 4 to 11 GPa, the stable mantle
assemblage for the system peridotite-{:02 is magnesite peridotite or magnesite garnet
peridotite (Figure 4.6).
Canit and Scarfe (1990) noted that the composition, calculated by mass balance, of
near-solidus partial melts (23-39%) in equilibrium with magnesite at 5-7 GPa, have lower
Ca/Mg ratios than melts in equilibrium with dolomite at 3 GPa (Figure 4.6). They interpret
the former to be kimberlitic and the latter melilititic in agreement with the conclusions
PETROGENESIS OFORANGEITES AND KIMBERLITES 321

CMS - CO2 CMAS -CO2

t
12 12

10 10
C
a..
-
(!) OL OL +OPX +CPX
OL+CPX +GNT
8 + +OPX 8

,,
CPX +MAG+L + MAG
L&.J
a:: + OL
::::> MAG
en 6
I + 6 OL+OPX
en I.
I
CPX
+ +CPX +GNT
L&.J
a:: I L +L
a.. OL+CPX
4
4 +OPX
+L

1000 1200 1400 1600 1000 1200 1400 1600

TEMPERATURE (OC) ~

Figure 4.6. (Left) Phase relationships in the system CaO-MgO-Si02-C02. (Right) Phase relationships in the
system CaO-MgO-AI203-Si02-C02. OL =olivine, CPX =clinopyroxene, OPX =orthopyroxene, GNT =
gamet, MAG =magnesite, DOL =dolomite, L =liquid (after Canil and Scarfe 1990).

of Brey et al. (1983). Melts at 9 GPa are too rich in Mg to correspond with any proposed
primary kimberlite magma (Figure 4.7). Canil and Scarfe (1990) suggest that kimberlites
are generated by partial melting of magnesite-bearing peridotite at pressures of 5-7 GPa
(150-250 km depth) in the upper mantle. Kimberlite cannot be generated at pressures
below 3 GPa.
Melting and magma segregation, followed by ascent of kimberlite magma, com-
mences when rising asthenospheric diapirs of magnesite peridotite become trapped at the
lithosphere-asthenosphere boundary. In their model, Canil and Scarfe (1990) do not
discuss Wyllie's (1980, 1989a,b) models of kimberlite genesis and emplacement.
Canil and Scarfe (1990) also suggest that magmas termed "protokimberlites" could
originate at pressures above 7 GPa. Such magmas may, for tectonic reasons, remain
trapped in the lower mantle. Alternatively, fractionation of olivine at high pressure may
drive their bulk compositions toward those of the 5-7 GPa melts prior to eruption as
kimberlites.
Clinopyroxenes crystallizing in equilibrium with melts become more subcalcic with
increasing pressures. Their compositions are similar to those of subcalcic pyroxenes
belonging to the megacryst suite. Thus, Canil and Scarfe (1990) suggest that the latter
may represent high-pressure phenocrysts in kimberlite magmas.
Canil and Scarfe's (1990) work provides important experimental evidence that
haplokimberlitic magmas may be generated by partial melting of carbonated lherwlitic
sources at high pressures, consequent upon decompressional melting.
322 CHAPTER 4

PARTIAL MELTS M
o 5 GPo
() 7 GPo
• 9 GPo

Cc
C
'-------------~~----------------------~
Ln s
Figure 4.7. Compositions of "uncontaminated" natural kimberlites. lamproites. melilitites. and partial melts
formed in the system CaO-MgO-Si02-C02 projected into the ternary system CaO-MgO-Si02 from A1203.
The arrows for the partial melts indicate their changing compositions with respect to increasing degrees of
melting (after Canil and Scarfe 1990).

Canil and Scarfe (1990) also demonstrate that lamproites are unlikely to represent
partial melts of carbonated lherzolite as their bulk compositions are richer in silica than
those of haplokimberlitic liquids encountered in their study (Figure 4.7).

4.3.1.3. Partial Melting o/Carbonated Phlogopite Lherzolite


Based on studies of the melting of carbonated phlogopite lherzolite, Thibault et al.
(1992) have proposed a model of cyclic metasomatism occurring beneath continental
rifts. This model has similarities with one presented by Haggerty (1989a,b) in proposing
the existence of distinct C02- or H20-rich metasomatically enriched horizons in the
lithosphere (see 4.4.2). A major difference between the models is that Thibault et al.
(1992) propose a dynamic model of metasomatism, whereas that of Haggerty is essen-
tially static.
Melting of a carbonated phlogopite lherzolite at 3 GPa and 1060-1150°C results in
the formation of alkaline dolomitic melts. The composition (wt %) of the 4% partial melt
at 1l00°C is 2.57 Si02, 4.54 FeO, 15.12 MgO, 21.60 CaO, 4.93 Na20, 7.01 K20,2.66
H20, 40.31 C02. Gamet, orthopyroxene, and olivine were found in equilibrium with the
melts at all temperatures, whereas phlogopite was consumed between 1125 and 1150°C.
Melting of a phlogopite lherzolite at 3 GPa from 1175 to 1250°C results in the
formation of hydrous potassic calcic melts. At 1225°C, the composition (wt %) of a 7%
partial melt is 47.96 Si02, 10.97 Ah03, 5.66 FeO, 13.21 MgO, 11.59 CaO, 1.04 Na20,
5.40 K20, and 2.05 H20.
Thibault et al. (1992) further showed that the alkaline dolomitic melts can metaso-
matize harzhurgite to olivine-rich phlogopite wherlite and calcite-bearing phlogopite
PETROGENESIS OF ORANGElTES AND KIMBERLITES 323

dunite. The interaction of harzburgite with the hydrous silicate melt results in enrichment
in clinopyroxene and phlogopite.
On the basis of these data Thibault et al. (1992) propose a cyclic model of metaso-
matism active beneath a continental rift. The cycle commences with the formation of a
horizon of carbonated phlogopite lherzolite at the base of the lithosphere by interaction
of barren harzburgite with dense alkaline fluids derived from a hot mantle plume.
Lithospheric thinning associated with plume uprise results in the uplift of the lithosphere-
asthenosphere boundary and melting of the carbonated phlogopite lherzolite. The low-
temperature melting components are then remobilized as small-volume volatile-rich
melts which, as they migrate upward, resolidify and react with the overlying colder
harzburgite. Through such cycles of melting, migration, and solidification the metaso-
matic front of carbonated phlogopite lherzolites eventually reaches a depth of 100 km (3
GPa).
Subsequent increase in the temperature of dolomite phlogopite lherzolite to 1150°C
results in the formation of an alkaline dolomitic melt which infiltrates overlying litho-
sphere leaving a residual C02-free garnet phlogopite lherzolite which returns to sub-
solidus conditions. As the alkaline dolomitic melt reaches 65 km (2 GPa, 950°C) it
solidifies and reacts with harzburgite to produce dolomite phlogopite harzburgite. The
metasomatic front is now decoupled and forms two distinct horizons: a dolomite
phlogopite harzburgite at 65 km and a C02-free garnet phlogopite lherzolite at 100
km.
With progressive rifting, the temperature at 65 km depth reaches lOOO°C and the
dolomitic phlogopite harzburgite is transformed to olivine-rich phlogopite lherzolite and
wehrlite with the release of C02-rich low-density fluids. At l00-km depth, temperatures
exceed the solidus of gamet phlogopite lherzolite. Partial melts from this source infiltrate
the metasome at 65 km resulting in an enrichment in phlogopite and clinopyroxene. The
final result of shallow decoupled metasomatism is a wide variety of metasomatized
ultramafic rocks comparable to the xenolith suites found in alkaline basaltic rocks.
Importantly, the studies indicate that carbonate-bearing and potassic magmas can be
generated in the lithosphere and that carbonated phlogopite lherzolite has a very low
solidus temperature at 3 GPa. Thibault et al. (1992) do not discuss the possibility of these
potassic melts escaping from the lithosphere. Interestingly, the low solidus temperature
suggests that melting of such sources at the fringes of large thermal plumes, in nonrifting
environments, might give rise to melts which would subsequently solidify in the litho-
sphere as veins of complex mineralogy. The study demonstrates the complexity of
metasomatic processes in the lithosphere and is thus relevant to the formation of
metasomatic and veined non-Iherzolitic sources for ultrapotassic rocks.

4.3.1.4. Carbonates in the Mantle?


The partial melting models of carbonated lherzolites described above are currently
the favored mechanism for generating asthenospheric kimberlites and other C02-rich
alkaline magmas. Despite the popularity of such sources it should be realized that they
are hypothetical concepts based entirely on experimental studies. Samples of dolomite-
or magnesite lherzolite have never been reported from the xenolith suites of kimberlites.
The report of the presence of a few calcite and dolomite inclusions in pyrope xenocrysts
324 CHAPTER 4

from minette diatremes in Utah (McGetchin et al. 1973, McGetchin and Besancon 1973)
is hardly overwhelming evidence in favor of widespread carbonation of the upper mantIe.
There is no doubt that C02 exists in the mantle and that the experimentally verified
carbonation reactions must occur. Consequently, the absence of carbonate-bearing xeno-
liths is ascribed to either the decomposition of all carbonate during transport or failure to
recognize, petrographically, small volumes of primary carbonate in altered xenoliths.
Conveniently, carbonate decomposition is always considered to result in complete
disaggregation of xenoliths (Wyllie et al. 1983); thus, any evidence pertaining to its
former presence is always destroyed.
Canil (1990) has attempted to resolve the problem by subjecting a synthetic carbon-
ated mantIe to various rates of isothermal decompression. A mixture of dolomite, enstatite,
and diopside, held initially at 2.5 GPa and l000°C, was shown to decompose rapidly into
forsterite, diopside, enstatite, and vapor at decompression rates of 1-3 GPa/hr (45-90
kmlhr). Decarbonated, decompressed experimental charges were found to be disaggre-
gated and granular. Fluid inclusions present in newly created olivine demonstrate that
fluid (C02-vapor) was present during decarbonation. Canil (1990) concludes that car-
bonate coexisting with silicates in mantIe-derived xenoliths could not survive entrain-
ment, even in the fastest ascending magmas, due to rapid decarbonation upon
decompression. Interestingly, decarbonation reactions may be one means of producing
some macrocrystal olivines.
Another approach to the problem is to seek textural evidence for decarbonation
reactions among the mineral constituents of xenoliths, but xenoliths have not been
extensively studied with this object in mind. Textural relationships between the presumed
products of decarbonation of dolomite, i.e., olivine and diopside, might be examined to
some avail in fresh xenoliths.
Berg (1986) has suggested that rare brucite-calcite intergrowths in mantle xenoliths
form when dolomite breaks down to calcite, periclase, and vapor. Periclase, being unstable
in a hydrous environment, is subsequently converted to brucite. Canil's (1990) experi-
ments do not support this reaction as being of importance in dolomite breakdown relative
to carbonate-silicate equilibria. A further problem with Berg's (1986) hypothesis is the
difficulty in distinguishing, on a petrographic basis, between periclase-derived brucite
and that formed by serpentinization (Berg 1989).
The apparent absence of mantle carbonates in xenolith suites can also be explained
if the lithospheric mantle is regarded as a barren residua purged of its volatiles by previous
melting episodes (Hess 1989). Hence, the lithospheric xenolith suite sampled by kimber-
lites cannot be expected to contain carbonates. Any C02 entering this environment results
in partial melting rather than carbonate formation, due to the depression of the mantle
solidus upon carbonation (Wyllie 1980).
This conclusion does not negate the possible existence of solid carbonates at greater
depth in the asthenosphere. Pristine samples of this material are unlikely ever to be present
in xenolith suites because, even during slow diapiric upwelling, decompressional partial
melting will inevitably occur and remove all traces of carbonate. Small modal amounts
of carbonate will be consumed in the initial stages of melting. Separation of melt along
grain boundaries, followed by segregation from the restite, may result in melts that
migrate relatively rapidly toward the lithosphere, leaving a restite containing no trace of
PETROGENESIS OF ORANGEITES AND KIMBERLITES 325

carbonate. Restites are unlikely ever to be entrained by their daughter melts, although
long-term mantle convection may transport restite to higher levels, where it may be
sampled by subsequent batches of magma. However, such material, because of continued
reequilibration, will contain no traces of its former character or depth of origin.
On the basis of the existing experimental data, melting of carbonated lherzolite
appears to be suitable for the production of C02-rich archetypal kimberlites but not
relatively C02-poor magmas such as orangeites or lamproites. However, detailed experi-
mental studies on H20- and alkali-bearing systems which might confirm or deny this
conclusion have not yet been undertaken. Ifindeed orangeites (and lamproites) are formed
in the lithosphere, then their origins might be clarified by further experimental studies of
non-Iherzolitic sources (see 4.2) with mixed volatiles.

4.3.2. Liquidus Experimental Studies at High Pressures


Inferences concerning the nature of the sources of magma, based on liquidus
experimental studies of kimberlite compositions, are severely hampered by our lack of
knowledge of the composition of primitive kimberlite magmas (Mitchell 1986). Further,
Mitchell (1986) has concluded that kimberlite whole rock compositions do not represent
those of the magmas from which they formed, as a consequence of volatile loss and crystal
fractionation during differentiation. Regarding volatile loss, Brey et al. (1991) have noted
that the solubility of C02 in kimberlitic melts may change dramatically at 4-5 GPa, and
large amounts of C02 may be released from melts in this pressure range during ascent
into the crust. Canil (1993) has observed that melting experiments on compositions with
such reduced C02 contents may produce points of multiphase saturation which are not
representative of the depth of origin of the magma or residual mineralogy of the source.
Macrocrystal kimberlites are particularly unsuitable material for experimental studies as
they are clearly contaminated rocks.
The few liquidus experimental studies that have been undertaken have utilized either
natural kimberlites (Edgar et al. 1988, Edgar and Charbonneau 1993) Or cobalt-substituted
synthetic analogues (Eggler and Wendlandt 1978, Ringwood et al. 1992, Kesson et al.
1994).

4.3.2.1. Liquidus Studies of Natural Kimberlite


Recognizing the above problems of sample selection, Edgar et al. (1988) and Edgar
and Charbonneau (1993) have suggested that liquidus studies of an aphanitic kimberlite
from the Wesselton Mine (South Africa) are meaningful as this material has a composition
close to those of uncontaminated kimberlite (Mitchell 1986). Edgar and Charbonneau
(1993) suggest that the absence ofxenocrystal olivine, high mg number (83.9), Ni (810
ppm), and Cr (2410 ppm) contents indicates that the parental magma of this kimberlite
has not undergone appreciable fractionation. In contrast, Mitchell (1986) and Shee
(1984), on mineralogical and petrographic grounds, regard the Wesselton aphanitic
kimberlite as a differentiate of a less-evol ved kimberlite magma. Edgar and Charbonneau
(1993) admit that some olivine and pyroxene fractionation may have occurred from the
parental magma, but nevertheless state that, for the purposes of their experimental studies,
"the actual origin of the aphanitic kimberlite is not as important as the similarity in its
326 CHAPTER 4

4 A 8

-...
\ OL "

-
5 \ + \ 10
\
L
C 6
a.
-
C
..0
(!) 7 L ~ o
_20
lLJ 8
a::: OL
lLJ
a::: \
~ 9 l LIQUID I
en + ~ I
en GNT ~ 30
A OL+L \
lLJ 10 B OL+SP+L I
a::: + lLJ
a::: C OL + SP tcPX +L \
a. II L a. o OL+SP+CPX+CT \

t 40
E OL+SP+MO+L +LI
F OL+SP+MO+CT+LI
I
I

1300 1400 1500 1600 1700 1000 1200 1400

TEMPERATURE (OC) •
Figure 4.8. Phase relationships of Wesselton aphanitic kimberlite at high (A) and low (B) pressures. OL =
olivine. GNT = garnet, L = liquid. MO = monticellite. CPX = clinopyroxene. CT = calcite. SP = spinel (after
Edgar and Charbonneau 1993. Edgar et al. 1988).

composition to that of other supposedly primitive kimberlites" (Edgar and Charbonneau


1993. p. 133). It is concluded here that, lacking samples of indisputably primitive
composition, the Wessel ton aphanitic kimberlite remains the best available sample of a
relatively unfractionated kimberlite.
Figure 4.8 illustrates the near-liquidus-phase-relationships of the Wesselton aphani-
tic kimberlite at 5-10 GPa, as determined by Edgar and Charbonneau (1993). The
experiments were undertaken with only the volatiles present in the kimberlite (Xeo =
0.24), at an oxygen fugacity estimated as slightly above the EMOG-EMOD buffer
(Eggler and Baker 1982), using presoaked Pt capsules to avoid Fe loss by alloying.
The data show that orthopyroxene and primary clinopyroxene are absent at near-
liquidus phases, although quench clinopyroxene occurs in all fields. Extrapolation of the
field boundaries suggests that the minimum pressure at which garnet and olivine might
coexist is 12 GPa. Liquidus olivines have mg = 92.4-94.4 and are thus similar to olivine
phenocrysts in the natural rock. Garnets are pyrope grossular almandines similar to those
common in eclogites.
Edgar and Charbonneau (1993) interpret the data to suggest that the liquid could
never have been in equilibrium with a garnet lherzolite source. At first sight, liquidus-
phase relations suggest garnet dunite as a potential source, although it is clearly too
refractory. Consequently, Edgar and Charbonneau (1993) note that the ubiquitous pres-
ence of quench clinopyroxene up to near-liquidus temperatures is evidence that the source
PETROGENESIS OFORANGEITES AND KIMBERLITES 327

material may contain clinopyroxene. The presence of olivine on the liquidus up to high
pressure is believed to be a consequence of the expansion of the olivine plus liquid field,
at the expense of clinopyroxene, due to C02 loss. Hence, at a higher C02 content a
multi saturation point involving clinopyroxene, olivine, and garnet would be expected in
experiments at lower pressures (10 GPa).
Experiments undertaken on the same starting composition at 1-4 GPa (Figure 4.8)
also suggest that orthopyroxene is absent from the source (Edgar et al. 1988). The absence
of Mg-carbonates at low pressure or as quench phases at high pressure is interpreted by
Edgar and Charbonneau (1993) to indicate that kimberlites are not derived from magnesite-
bearing sources. In this context, Foley (1990) notes that calcite is a possible mineral in
the source, as the reactions which limit the formation of magnesite all involve orthopy-
roxene.
Clinopyroxene is an abundant near-liquidus phase in low-pressure experiments at
1.5-3 GPa. To explain the absence of primary low-pressure clinopyroxene phenocrysts
in kimberlites, Edgar et al. (1988) suggest the magma temperatures remained high
(1250-1300°C) during rapid ascent from about 150-km depth. Under these constraints,
olivine and spinel will be the only liquidus phases during ascent until they are joined by
monticellite at pressures of less than 1 GPa.
In summary, Edgar and Charbonneau (1993) suggest that kimberlites are generated
by the partial melting of a gametite (olivine--clinopyroxene-garnet) source at 10-13 GPa
(300-330 km), which may represent deeply subducted material of basaltic composition.

4.3.2.2. Liquidus Studies of Synthetic Kimberlite


Liquidus studies of synthetic kimberlites have been undertaken by Eggler and
Wendlandt (1978, 1979), Ringwood et al. (1992), and Kesson et al. (1994). All experi-
ments use CoO as a proxy for FeO to eliminate alloying with the Pt capsules used to
contain the samples in the experiments. Co-substitution results in liquidus temperatures
about 50°C higher than in the analogous Fe-bearing systems. The results of, and
conclusions derived from, these experiments must be regarded as preliminary as detailed
investigations of the phase relationships were not undertaken.

4.3.2.2.a. Average Lesotho Kimberlite. Eggler and Wendlandt (1978, 1979) studied
a bulk composition equivalent to that of the average anhydrous Lesotho kimberlite
(Gurney and Ebrahim 1973) at 3 and 5.5 GPa with various C02IH20 ratios. In considering
the significance of their results, it should be borne in mind that Mitchell (1986 p.279) has
noted that this average composition includes macrocrystal, altered, and contaminated
kimberlites and is biased toward relatively high silica and alumina contents. Thus, the
composition is unlikely to represent primitive kimberlite magma.
Figure 4.9 shows this composition has olivine, orthopyroxene, and clinopyroxene
near the liquidus at both pressures studied. Extrapolation of the gamet-in curve indicates
that all four phases occur near the liquidus at 6 GPa. Dolomite and magnesite were found
at temperatures 300°C below the liquidus. Eggler and Wendlandt (1979) intetpret the data
to indicate that kimberlite could be a partial melt of a phlogopite magnesite lherzolite at
about 6 Pa. During melting, phlogopite and magnesite are consumed during the initial
stages and do not remain as residual phases.
328 CHAPrER4

60 LHZ +
Cm +Ph
~
+V
0

-
.c
~ 50

La.J
a:: L
=>
en 40 Figure 4.9. Phase relationships of a co-
en baltian synthetic kimberlite composition
La.J
a:: containing 5 wt % C02 and 5 wt % H20
a.. extrapolated from phase equilibria at 3.0
and 5.5 GPa. The heavy line is the univari-
30 ant zone of invariant vapor composition
(ZIVC) solidus for haploperidotite-C02.
LHZ = lherzolite, Cm = magnesite, Ph =
1000 1200 1400 1600 phlogopite, Dol = dolomite, L = liquid, V
= vapor (after Eggler and Wendlandt
TEMPERATURE (OC) 1979).

Eggler and Wendlandt (1979) suggest that kimberlites are generated by melting of
carbonated peridotite at 5-6 GPa (160-200 km) as a rising asthenospheric diapir reaches
a level at which the solidus intersects the local geotherm.

4.3.2.2.h. Ultra-high-Pressure Experiments. Ringwood et al. (1992) and Kesson et


al. (1994) investigated a Co-analogue of a composition similar to that of the average group
lAkimberlite of Smith et at. (1985b). This average composition is based upon the analysis
of a carefully selected group of samples and provides a reasonable estimate of the bulk
composition of an uncontaminated, but not necessarily primitive, kimberlite.
Experiments were undertaken at the very high pressures of 10 and 16 GPa because
the discovery of majorite garnets (Moore and Gurney 1985) and xenoliths derived from
depths greater than 300 km (Haggerty and Sautter 1990) suggested that kimberlites might
be generated within the transition zone of the mantle. Majorite garnets are formed when
the components of pyroxenes are taken into solid solution by preexisting garnets at high
pressures. Pyroxenes are not stable below 500-km depth in the transition zone. Majorite
garnets are solid solutions between normal garnets (Mg,Fe,CahAhSbOJ2 and an
alumina-free end member (Mg,Fe,Cah[(Mg,Fe,Ca)(Si)]SbOJ2 containing octahedrally
coordinated Si (Ringwood 1967).
Experiments at 16 GPa indicate that the liquidus is slightly below 1700°C, and
near-liquidus phases at 1650°C are ~-Mg2Si04 and majorite garnet. Below 1600°C,
clinoenstatite joins the assemblage. The solidus is located at about 1525°C with the
subsolidus assemblage consisting of ~-Mg2Si04, majorite gamet, clinoenstatite, and
dolomitic carbonate.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 329

Initial experiments by Ringwood et al. (1992) at 10 GPa placed the liquidus below
1650°C. The near-liquidus assemblage consisted of olivine, clinoenstatite, and perovskite.
These are joined by gamet (21% majorite) at 1500°C and by primary magnesite and
clinopyroxene at 1400°C. The solidus is located between 1200 and 1300°C.
Kesson et al. (1994) conducted a second series of experiments in recognition of the
discrepancy between their initial results and those of Edgar and Charbonneau (1993) at
10 GPa. The principal differences were the appearance of garnet well below the liquidus
and the presence of primary perovskite.
The Wesselton aphanitic kimberlite studied by Edgar and Chabonneau (1993) is
depleted in Si02 and enriched CaO relative to Ringwood et al.'s (1992) starting compo-
sition. Kesson et al. (1994) suggest that these compositional features result in the
expansion of the garnet stability field toward the liquidus in Edgar and Charbonneau's
experiments. However, the garnets in the latter are not pyrope rich and are completely
unlike those crystallized in either of the experiments by Ringwood et al. (1992) and
Kesson et al. (1994). In the latter experiments, olivine was shown to be the liquidus phase
at 10 GPa and 1650°C, followed closely by pyrope-rich garnet. Perovskite was considered
to be a near-solidus primary phase rather than a near-liquidus phase.
Ringwood et al. (1992) conclude that the experimental phase relationships preclude
kimberlite being formed by the partial melting of a garnet lherzolite at depths ofless than
300 km. The basis for this conclusion is the absence of garnet and orthopyroxene and the
presence of perovskite as near-liquidus phases. Accordingly, their preferred model is
based upon their 16-GPa experiments and involves small degrees of partial melting of
"refertilized" former harzburgites in the transition zone (>400 km). This source is
composed of majorite garnet and ~- or y-(Mg,Fe)2Si04. Based upon these few experi-
mental data, Ringwood et al. (1992) erected a complex hypothesis for the origin ofthe
source involving interactions between subducted oceanic crust and upwelling astheno-
sphere at the 650-km discontinuity. This model is discussed further in Section 4.4.1.

4.3.2.3. Summary-A Cautionary Note


The differences between the three sets ofliquidus experimental studies are undoubt-
edly mainly related to compositional differences between the starting materials. These
studies provide an excellent illustration of the problems inherent in choosing different
compositions as "primitive kimberlite," as the experimental data cannot be reconciled
and lead to very different conclusions regarding the composition of the source and depth
of origin of kimberlites.
Importantly, none of the liquidus or partial melting experimental studies summarized
above indicate that kimberlites and orangeites are formed by melting of similar source
rocks.

4.4. GEODYNAMIC MODELS OF KIMBERLITE AND ORANGEITE


GENESIS
Geodynamic models of kimberlite and orangeite genesis attempt to place their origins
in the context of our current understanding of plate tectonics and mantle dynamics. Given
our inadequate knowledge of many deep mantle processes and the long-term evolution
330 CHAPTER 4

of the Earth, many of these models are, of necessity, highly speculative; however, they
are important in that they serve to direct us to new avenues of investigation which might
provide useful evidence for rejecting or verifying these hypotheses.
Some recent geodynamic models of kimberlite genesis (Ringwood et al. 1992,
Haggerty 1994) have placed their sources at ever greater depths in the mantle, even as
deep as the core-mantle boundary. The stimulus for these hypotheses was the discovery
that some diamonds in the Monastery kimberlites contain syngenetic inclusions of
majorite garnet (Moore and Gurney 1985, 1989, Moore et al. 1991). On the basis of
experimental studies of garnet-pyroxene solid solutions at high pressures (Irifune et al.
1989), it was suggested by Moore et al. (1991) that the inclusions formed at depths of at
least 480 km, i.e., within the transition zone of the mantle.
Subsequently, Wilding et al. (1994) described inclusions within Brazilian alluvial
diamonds, supposedly of kimberlitic provenance, consisting of majoritic garnets whose
compositions indicated derivation from depths of 180--400 km. Other inclusions present
are CaSi03, MgO-FeO solid solutions (periclase-wustite), Si02 (? stishovite), diopside,
CazAbSi07, moissanite, and olivine (F08s). At depths greater than 650 km, experimental
evidence indicates that mantle olivine and pyroxene are replaced by perovskite-structured
"pyroxene" and ferripericlase. Therefore, Wilding et al. (1994) suggest that some mem-
bers of this inclusion suite (CaSi03, MgO-FeO, Si02) may represent minerals originating
from below these depths.
A deep origin for kimberlites was also considered by Haggerty and Sautter (1990)
and Sautter et al. (1991), who reported the occurrence of orientated incl usions of pyroxene
in garnet in xenoliths from the Jagersfontein kimberlite. These were interpreted to
represent majorite garnets which crystallized at 300--400 km depth and were then
transported to shallower depths where exsolution of pyroxene occurred.
Initially, the majorite garnet-bearing diamonds and xenoliths were believed to have
crystallized in the transition zone, and been subsequently carried to shallower depths by
convecting mantle, where they were eventually sampled by kimberlites (Moore and
Gurney 1985, 1989). In this scenario, kimberlites are not formed in the transition zone.
Subsequently, Moore et al. (1991) considered the possibility that the kimberlite itself was
generated at depths exceeding 450 km and transported the majorite-bearing diamonds
directly to the surface. Ringwood et al. (1992) commenced their high-pressure liquidus
experimental studies of kimberlite (4.3.2.2) on the basis of this suggestion. Haggerty
(1994) has suggested that diamonds bearing strongly-reduced mineral inclusions (SiC,
silicate perovskite, wustite-periclase), might be formed in the lower mantle, and further
implied that their transporting hosts might commence their ascent from these depths (see
4.4.2).
In considering hypotheses based upon the above observations, it should be kept in
mind that occurrences of majoritic garnets in diamonds and xenoliths are the exception
rather than the rule. Similarly, the metasomatized peridotite xenolith suite found in some
southern African kimberlites (Erlank et al. 1987) may not be characteristic of all cratons.
Consequently, general petrogenetic schemes erected upon the recognition of these few
atypical occurrences should be regarded with due caution as to their universal, and even
local, veracity.
PETROGENESIS OF ORANGElTES AND KIMBERLITES 331

4.4.1. Transition Zone Melting


Ringwood et al. (1992) have suggested that kimberlites are formed by partial melting
of a majorite plus y- Mg2Si04 assemblage in the transition zone (400-650 km). The origins
of this source are sought in a complex model, which is a variant of Ringwood's (1989)
hypothesis of subduction of oceanic crust to the 650-km discontinuity, as a means of
explaining the composition and evolution of the mantle.
In Ringwood et al.'s (1992) model, oceanic crust (basalt plus sediments) and
lithospheric harzburgite are subducted to the 650-km discontinuity. The subducted slab,
or megalith, remains at the discontinuity because of the density contrast between it and
the underlying lower mantle. Here, oceanic crust recrystallizes to a volatile-bearing
mixture of gametite (>85% majorite garnet + <15% stishovite) and recrystallized former
sediments. At these pressures and temperatures harzburgite recrystallizes to spinel-structured
y- Mg2Si04 (>80%) plus minor «20%) majorite garnet, ilmenite, and stishovite. Former
harzburgite may be mixed with the garnetite layer (Ringwood 1989) or accumulate above
it (Ringwood et al. 1992). Partial melting in the gametite layer results in the addition of
incompatible element-rich melts to overlying harzburgitic material. These melts have
higher LalYb, U/Pb, and Th/u ratios than their parental MORB-derived gametite.
Ringwood et al. (1992) suggestthat the persistence of the "refertilized" lithology for times
of up to 1 Ga, without further melting, may allow sufficient time for isotopic heteroge-
neities to develop.
Subsequently, the residual garnetite and the refertilized former harzburgite were
heated by convection currents rising from the lower mantle. The gravitationally stable,
and highly viscous, garnetite layer forms a barrier which deflects convection currents,
causing the layer to be elevated by as much as 50 km, forming a dome-like structure
several thousand kilometers in diameter (Figure 4.10). Previous melt extraction results
in the gametite layer having a higher solidus temperature than overlying refertilized
former harzburgite. Consequently, heat from the convection current causes remelting of
only the latter, forming kimberlitic (and lamproitic) magmas. These ascend to the upper
mantle, collecting xenoliths, primarily from subcontinental lithosphere, and rarely from
the transition zone. The hypothesis explains the "highly fractionated" geochemical
signature of kimberlite as resulting from a combination of two separate partial melting
events, both occurring in the presence of garnet.
Ringwood et al. (1992) derive archetypal kimberlites, orangeites, and olivine lam-
proites by the same process at the same depths. Compositional and isotopic differences
between archetypal kimberlites and orangeites are considered to originate from the
derivation of the first partial melts in the gametite layer from regions relatively poor, or
rich in, subducted pelagic sediments, respectively. Significant mixing of second-stage
melts does not occur and magmas derived from isotopically distinct portions of the
refertilized layer segregate and rise as separate batches of magma.
Apart from the implausibility of deriving three distinct magma types from a single
source by the same process, Ringwood et al. 's (1992) hypothesis has many problems of
a general geophysical-geodynamic nature related to the thermal structure and rheological
character of the mantle. Discussion of these is entirely beyond the scope of this work, and
comments on the hypothesis are confined to aspects specific to kimberlite genesis.
332 CHAPTER 4

Figure 4.10. Model for the generation of kimberlites by partial melting of refertilized former harzburgites at
the 650-km discontinuity. Melting and uplift of the garnetite layer occur in response to temperature increases
resulting from the upwelling of a lower mantle convection current. See text for further explanation (after
Ringwood et al. 1992).

Ringwood et al. (1992) consider it impossible to derive kimberlites and orangeites


from different sources because of constraints imposed by isotopic and major element
compositions. The apparent paradox between this observation and arguments in favor of
separate lithospheric and asthenospheric sources for orangeite and kimberlites, respec-
tively, noted by Ringwood et al. (1992, p. 532), stem from considering both magmas to
be genetically related. The problem disappears if the magmas are not genetically related
and derived from different sources. Requiring orangeites to be derived from an astheno-
spheric source is not in accord with the occurrence of orangeites only in the Kaapvaal
craton. Thus, the temporal and geological relationships of kimberlite, orangeite, and
lamproite required by the model are not supported by observation.
For the above reasons, coupled with the mineralogical and geochemical observations
given in this work and Mitchell and Bergman (1991), it is considered that Ringwood et
al. 's (1992) model may have applicability to archetypal kimberlites, but is not relevant
to either orangeite or lamproite genesis.
Unfortunately, even with regard to archetypal kimberlites, the model lacks specific
details regarding the extent of partial melting in both stages, and the mineralogy of the
refertilized source. In addition, surely discussion of the fate of subducted carbonate during
PETROGENESIS OF ORANGEITES AND KIMBERLITES 333

each melting episode should be a prerequisite to this model of kimberlite genesis.


Interestingly, a deep asthenospheric source for kimberlites suggests that they need not be
confined to continental regions.
It is not possible to reject or accept the novel hypothesis that kimberlite generation
is in some manner connected with partial melting of a majorite- ~- or y-spinel source
until further experiments are undertaken on the potential source material. The complexity
of Ringwood et al. 's (1992) model results from the bias held-that subduction of oceanic
crust is required to provide a source of incompatible elements. The simpler alternative
model of decompressional melting of primitive or metasomatized mantle at these depths
is not explored.

4.4.2. Metasome Melting and Mantle Plumes


Haggerty (1989a,b) has suggested that a necessary prerequisite for the genesis of
kimberlitic and other varieties of continental alkaline magmatism is the formation of
metasomatic mineral assemblages in depleted harzburgitic lithosphere. These assem-
blages are envisioned to form layers, termed "metasomes," in the continental lithospheric
mantle over a depth range of 60-100 km. Xenoliths of metasomatized peridotite contain-
ing phlogopite, K-richterite, lindsleyite-mathiasite, hawthorneite-yimengite, and other
titanates (Erlank et al. 1987, Haggerty 1983) found in some kimberlites are considered
to be fragments of these metasomes. Metasome formation takes place when astheno-
spheric melts solidify in the lithosphere as they cool below the solidus of carbonated
lherzolite. Their crystallization releases fluids containing dissolved incompatible ele-
ments which migrate into the surrounding lithosphere. The formation of metasomes is a
variant of Wyllie's (1980) model of melt-lithosphere interaction (4.3.1.1). It differs in
that Haggerty (1989a,b) requires metasomatism to be a precursor, and not a consequence,
of kimberlite magmatism.
The metasomatic zone is divided into two overlapping units, a lower (100-75 km)
zone representing the phlogopite-K-richterite peridotite (PKP) assemblage and an upper
carbonate-rich horizon <65 km) containing zircon, ilmenite, and calcite. Both zones
contain phlogopite, K-richterite, and alkali titanates. Evidence for the depths and charac-
ter of the zonation is weak, as the distributions of purported assemblages are not
constrained by accurate equilibration pressures and temperatures, paragenetic, or mineral
stability data.
Haggerty (1989,a,b) proposes that a rapidly ascending asthenospheric "protomelt"
will induce preferential partial melting of the metasome, as the latter is a region with a
lower solidus temperature than the surrounding mantle. The protomelt is considered to
be komatiitic for no obvious reasons other than Haggerty's (1989,a,b) requirement that
this magma have higher MgO contents and temperatures than those magmas (unspecified
but presumably basaltic) which generate the metasomes.
Incorporation of metasome and included volatiles into the hot komatiitic picrite then
leads to "flash melting" (an undefined process) and the explosive eruption of alkaline
melts. Kimberlites and lamproites are purported to be generated in the PKP zone and
carbonatites in the calcite-rich zone.
334 CHAPTER 4

Haggerty (1994) has suggested that events at the core-mantle boundary are ulti-
mately responsible for the generation of the protomelts which subsequently interact with
metasomatized lithosphere to generate kimberlites. The hypothesis is based on the
presence, in some kimberlites, of xenoliths and diamonds derived from, or below, the
transition zone and the assumptions that there is a supposed correlation of kimberlite
magmatic activity with times when the Earth's geomagnetic field remained stationary for
long periods (40 Ma) of time; kimberlite magmatism is considered to be global rather
than local; and there is a direct relationship between kimberlite magmatism and deep
mantle plumes.
According to this model, instabilities at the D" layer, caused by crystaJIization and/or
convection events in the core, result in the formation of mantle plumes which travel
rapidly (25-50 Ma) from the D" layer to the lithosphere. Komatiitic melts derived from
this plume are arrested in the lithosphere and cause partial melting of the PKP metasomes
to form kimberlites.
Detailed criticism of many aspects of Haggerty's (1989a,b 1994) model, i.e., plume
ascent, origin of the D" layer, relationship of magmatism to the cyclicity of the geomag-
netic field, etc., is beyond the scope of this work and best left to the geophysical
community. Other parts of the model are difficult to critique as ideas are expressed as ex
cathedra pronouncements not supported by fact or reasoned argument. Further, parts of
the model are simply incorrect, such as the statement that carbonatites and lamproites are
closely-related (Foley 1992a,b, Mitchell and Bergman 1991).
Haggerty's concept that the genesis of kimberlite hinges upon the supposed assimi-
lation of metasomes in a komatiitic-picritic magma is implausible. The principal argu-
ment against this process is that komatiitic magma is confined to Archean times
characterized by unusually high heat flows (Hess 1989). Haggerty does not explain either
how this magma is generated in post-Archean times or its relationships to the voluminous
continental flood basalts which are also ascribed to the melting of mantle plumes. The
details of the assimilative process are glossed over, and the concept of "flash melting" is
never put on a firm physicochemical basis.
Importantly, Haggerty does not explain why flood basalts in Siberia occur in the
midst of an extended period of kimberlite magmatism, kimberlites in southern Africa are
preceded by basalts, and kimberlites in Colorado, Tanzania, China, Canada (Saskatche-
wan, Somerset Island, Lac De Gras), Zaire, and Arkhangelsk are not associated at all with
flood basalts. Explanation of some of these observations is critical, given that the solidus
of metasomatized mantle is undoubtedly well below the temperature of basaltic magmas
passing through the lithosphere. The experimental data of Thibault et al. (1992) support
this observation as they show that, at 3 GPa, the solidus of carbonated phlogopite
lherzolite is as low as 1060°C. Thus, melting of metasomes by precursor episodes of
basaltic magmatism is highly probable (see 4.5.2.1). Based upon these observations,
Thibault et al. (1992) have formulated a model, similar to Haggerty's, in which decoupled
metasomatic zones rich in either C02 or H20 are generated in the lithospheric mantle by
successive cycles of melting, migration, and solidification of incompatible-element-rich
melt/fluids (see 4.3.1.3). Interestingly, neither Haggerty nor Thibault et al. (1992) refer
to each other's work. Finally, remobilization or assimilation of enriched, i.e., metasoma-
tized, lithosphere during emplacement of magma has recently emerged as a fashionable
PETROGENESIS OF ORANGEITES AND KIMBERLITES 335

concept to explain the subtleties of flood basalt geochemistry (Cox 1992, Ellam and Cox
1991, Hawkesworth et al. 1986).
Coincidence of epochs of stable magnetic polarity with episodes of kimberlite
magmatism is not surprising, given the length of the epochs relative to the time scale of
any type of frequently occurring magmatism; consequently, this aspect of the model has
little petrogenetic significance. (Haggerty does not discuss the errors attached to the time
scales of geomagnetic epochs.)
Finally, a major omission in Haggerty's (1 989a,b, 1994) papers is the failure to
discuss opposing hypotheses suggesting that plumes have absolutely nothing to do with
continental alkaline magma generation-i.e., lithospheric fracture propagation (Turcotte
and Oxburgh (1978) or lithospheric volatile fluxing (Bailey 1977, 1983, 1992).
The principal merit of Haggerty's model is that it draws our attention to the
complexity of the metasomatic events in the upper mantle. The proposed metasomatic
"stratigraphy" is innovative but, unfortunately, difficult to prove on a quantitative basis.
Although the model is entirely inappropriate for generating kimberlites, it does support
the concept that the melting of enriched regions of the lithosphere might playa role in
the genesis oflamproites (and orangeites).

4.4.3. Hot-Spot Melting


Plate tectonic reconstructions of southern Africa, combined with geochemical data,
have been used by le Roex (1986) to suggest a relationship between kimberlites and
orangeite magmatism to hot-spot activity. In this interpretation, le Roex (1986) notes that
the geochemical characteristics of orangeites are similar to those of oceanic island basaltic
(OIB) rocks occurring within the Dupal isotopic anomaly (Zindler and Hart 1986). This
anomaly encircles the Earth between the equator and 600 S latitude and is characterized
by enrichment in incompatible trace elements, high 87Sr/86Sr, 206pbP04 Pb ratios and low
143Nd/l44Nd ratios. The source of the Dupal anomaly has been interpreted as ancient
subducted-recycled oceanic lithosphere including pelagic sediments, delaminated sub-
continental lithosphere, or upwelling primordial mantle (Allegre and Turcotte 1985,
Zindler and Hart 1986). Upwelling of this anomalous asthenospheric material results in
hot-spot volcanism with an apparent lithospheric signature, whereas upwelling outside
the Dupal anomaly region leads to volcanism with the normal depleted asthenospheric
isotopic signature. Extending these observations to orangeites implies that they are
asthenospheric melts derived from Dupal-type sources (Ie Roex 1986). Orangeite trace
element and isotopic compositions suggest derivation from a source more enriched than
those of Dupal-OIB basalts. As such they could also be considered as mixtures between
MORB-type sources and an even more enriched source.
Le Roex (1986) has noted orangeites lie at the termination of compositional trends
defined by Dupal-type OIBs, whereas archetypal kimberlites terminate trends defined by
normal OIBs (Figure 4.11). Note that these trends are not well defined for all isotopic
ratios or elemental ratios. Regardless, Ie Roex (1986) suggests that both archetypal
kimberlites and orangeites are derived from asthenospheric sources.
Le Roex (1986) has attempted to correlate kimberlite and orangeite asthenospheric
magmatism with hot-spot tracks in the South Atlantic (Figure 4.12) and believes that the
336 CHAPTER 4

I I I I I I I I I
0·5122 0·5126 0'5130
143 Nd / 144 Nd

20 -

1-
ORANGEITE
I I I I I I I I I
0'704 0'706 0'708 0'710
87 Sr / 86 Sr

Figure 4.11. Correlations between Zr/Nb ratios and Sr or Nd isotopic compositions of South Atlantic basalts,
kimberlites, and orangeites (after Ie Roex 1986).

distribution of orangeites is consistent with the Shona hot-spot paleotrack. In contrast,


kimberlites show no simple correlation with either the Bouvet or Discovery paleotracks.
Similar conclusions have been drawn by Skinner (1989), who has noted that the
almost linear progression of orangeite ages in southern Africa suggest that hot-spot
activity may have been important with respect to their formation. In contrast to Ie Roex
(1986), Skinner (1989) regards hot-spot activity to transfer only heat and volatiles to the
lithosphere, where kimberlites and orangeites are generated by melting of lithospheric
sources. Thus, both authors use the same data to arrive at completely different conclusions
regarding sources of these magmas. In agreement with le Roex, Skinner (1989) finds no
correlation between kimberlite magmatism and hot-spot tracks.
Mitchell (1986) and Bailey (1983, 1993) have discussed hot-spot models of conti-
nental alkaline volcanism and found them to be unconvincing. The principal arguments
against the model are that it cannot satisfactorily explain the worldwide distribution of
PETROGENESIS OF ORANGEITES AND KIMBERLITES 337

o!

AFRICA KIMBERLITES

Figure 4.12. Location and age ofkimberlites and orangeites (from Skinner 1989) relative to postulated hot-spot
palaeotracks during the Upper Jurassic and Lower Cretaceous in southern Africa. Numbers on tracks show ages
in Ma (after Ie Roex 1986).

kimberlites in particular and continental alkaline magmas in general; most kimberlites


within a province have been emplaced at the same time over a wide region and not
sequentially over curvilinear trends; hot spots fail to explain the repeated eruption of
similar magmas over a long period of time in a relatively restricted area. Hot spots are
usually conceived of as narrowly focussed heat sources (Morgan 1971) in marked contrast
to the current interpretation of mantle plumes as large (1000-2000 km at their heads)
upwellings of mantle.
Section 1.7 suggests that the southern African orangeite province may be better
regarded as reflecting two distinct periods of magmatism rather than a single hot-spot
track. The width of the province, distribution of intrusions of different age within it (see
1.7), and especially the concentration of activity at the southeastern end preclude any
simple, single, small hot-spot model for its formation. Application of the concept to this
province would lead to an improbable proliferation of multiple hot-spot tracks. The
apparent absence of orangeites between Swartruggens and Barkly West is also difficult
to reconcile with a single hot-spot track.
338 CHAPTER 4

Bailey (1993) rejects all plume-hot-spot-related models of continental alkaline


igneous activity and suggests that the apparent age progression of orangeites may reflect
progressive structural readjustments into the craton following continental breakup and
the development of the Lebombo monocline. In this model, orangeites (and kimberlites)
are generated in response to passive reactivation of the continental plate by external forces
(see 4.5.2.1).
With regard to Ie Roex's (1986) geochemical arguments, there are simpler ways of
explaining the isotopic data than by appealing to unknown asthenospheric sources. For
example, apart from the conventional interpretation of distinct asthenospheric and li-
thospheric sources, kimberlites and orangeites could be derived from young- and ancient-
enriched lithospheric sources, respectively. In either instance trace element correlation
diagrams involving asthenospheric Dupal-type oms and lithospheric orangeites have no
petrological/geological foundation and are irrelevant, as they simply compare apples and
orange(i te)s.

4.4.4. Partial Melting of Heterogeneous Lithosphere


Skinner (1986, 1989) proposed the first hypothesis attempting to provide a general
model for the origin and emplacement of archetypal kimberlites (referred to as group I
kimberlites) and orangeites (group II kimberlites). The model relies upon small degrees
of partial melting of a phlogopite garnet lherzolite as source for the magmas. The
hypothesis is based upon Wyllie's (1980) model of kimberlite genesis (see 4.3.1.1)
together with an unconventional interpretation of Sr and Nd isotopic data (3.8.1).
Skinner (1989) considers that the broad similarities in the incompatible-element
abundance patterns of kimberlite and orangeites (see 3.6.1), recognized by Smith et al.
(1985b), indicate that both magmas are derived from "broadly similar sources" (Skinner
1989, p. 539). On this basis alone, it is assumed that both magmas are derived from
compositionally different domains in the continental lithosphere, regardless of the con-
ventional interpretation of the isotopic composition of kimberlite as indicative of asthe-
nospheric sources.
Skinner (1989) further assumes that magmas are generated by small degrees of
melting of a phlogopite-bearing source as a result of the introduction of volatile elements
(C02, H20) and heat from the asthenosphere into the lithosphere (Wyllie 1980). The
lithosphere is considered to be heterogeneous, as the source rocks of orangeites and
kimberlites are considered to be geochemically and isotopically distinct. The sources of
orangeites were enriched in K, Pb, Rb, Ba, LREE, Si02, and H20, and depleted in Ti02,
Nb, and C02, relative to those of kimberlites. Enrichment of orangeite sources occurred
at least 500 Ma-1A Ga ago and were isolated from those of kimberlites. Skinner (1989)
does not discuss the time of formation of the kimberlite sources, except to state that they
represent depleted mantle modified by older periods of metasomatism.
The crux of Skinner's (1989, p. 540) hypothesis lies in his statement that
sectors of the lithosphere that were relati vely more highly depleted with basaltic elements (i.e.,
essentially harzburgiticl may have acted as preferential sites for later incompatible element
enrichment. Melting caused by the sti1llater introduction of H20 and C02, would be accelerated
in the highly LREE enriched zones but would be retarded within those zones less enriched in
LREE. The fact that Group II kimberlites predate Group I kimberlite occurrences, where both
PETROGENESIS OF ORANGEITES AND KIMBERLITES 339

occur in the same area, indicates that if the genesis of both magma types is initiated by the same
or similar processes, then Group II magmas are generated earlier, and hence emplaced earlier
than Group I kimberlites.
Hence, Skinner (1989) believes that orangeites and kimberlites are produced by the same
process, but the rate of melting of the source rocks differs as a consequence of their
differing composition. The existence of a previously metasomatized mantle is a prereq-
uisite for Skinner's model.
Figure 4.13 summarizes Skinner's (1989) four-stage hypothesis for the genesis,
ascent, and emplacement of kimberlite and orangeite. In stage 1 the introduction of
asthenosphere-derived volatiles results in melt development at point A on the shield
geotherm. In the LREE-, Rb-, and Ba-enriched regions of the lithosphere, slow percola-
tion of melts derived from this source, i.e., orangeites, coalesce into small magma pockets
at B. Formation of kimberlite magma takes place over the region A-C and involves a
larger vertical section of lithosphere, less enriched in REE, i.e., kimberlite magma
generation is slower than that of orangeites for the same volatile introduction event and
the melts have isotopic signatures characteristic of depleted mantle. The evidence for
magma development over an extended section of lithosphere is the apparent restriction
of metasomatized lherzolitic xenoliths, and the Ti-rich macrocryst suite, to kimberlites.
Skinner (1989), following Erlank et al. (1987) and Haggerty (1986), considers that the
former are all generated at relatively high levels in the lithosphere by diverse metasomatic
processes and the latter are xenocrysts derived from magmas unrelated to kimberlites.

ORANGEITE KIMBERLITE

A .aJ Crust : : ~
50

~ 100

~
Q. 150
Q)
o Buffered
200 Peridotite

250

500 1000 1500


Temperature (OC)
Figure 4.13. Model of kimberlite (K) and orangeite (0) magmatism according to Skinner (1989). See text for
explanation.
340 CHAPTER 4

From the paucity of xenoliths found in orangeites it follows that they must be derived
from smaller volumes of mantle relative to kimberlites.
Stage 2 of the model covers the slow ascent of the magmas up to a level at which the
rate of movement rapidly increases. During this stage the magmas are believed to
incorporate the bulk of their xenolith suites. Orangeites following path A-D on Figure
4.13 do not sample the upper portions of the lithosphere, whereas kimberlites following
path A-E sample a wide variety of xenoliths. Skinner (1989) believes that some crystal-
lization of primary olivine, ilmenite, spinel, and phlogopite may occur during this stage.
During stage 3 it is envisaged that discrete batches of magma accelerate rapidly and
move upward by means of crack propagation. The rapid ascent is initiated by an increase
in the volatile content of the melt resulting from crystallization of later generations of
olivine and other primary minerals. For orangeites and kimberlites the rapid ascent paths
on Figure 4.13 are from D-F and E-F respectively. Relatively few xenoliths are incorpo-
rated during this stage of the ascent. Both magmas are believed to decelerate rapidly as
they approach point F as a consequence of degassing into open fractures.
Stage 4 from point F-G (Figure 4.13) involves the slow intrusion of the magmas as
sills and dikes, together with the formation of diatremes. Late-stage ground mass minerals
and segregationary textures are formed during an extended period of post-emplacement
crystallization.
Skinner (1989) does not discuss in any detail the origins of the asthenospheric
volatiles which initiate the partial melting process. Allusion is made to the linear age
progression of 145-110 Ma orangeites in southern Africa as being suggestive of transfer
of volatiles and heat from a stationary hot-spot into a moving lithospheric plate. However,
the lack of an apparent correlation between age and distribution of 95-80 Ma kimberlites
is not consistent with this hypothesis. The absence of a correlation may be related to the
extended time proposed by Skinner (1989) for development of kimberlites versus
orangeites. Note that Skinner's model was developed primarily to explain the age and
geochemical relationships between southern African J urassic-Cretaceous kimberlites and
orangeites. Skinner (1989) does not discuss why the emplacement of the older kimberlite
provinces of southern Africa, and elsewhere in the world, are not preceded by orangeite
magmatism.
Skinner's (1989) hypothesis is appealing in that it attempts to draw together diverse
aspects of both kimberlite and orangeite magmatism which must be considered in any
general model of continental alkaline rock petrogenesis. Unfortunately, the model is
ultimately unsatisfactory as it is based on numerous unfounded assumptions. The model
stands and falls upon the initial premises of stages 1 and 2. The claim that kimberlites are
lithospheric melts is simply an assumption and not supported by critical examination of
the geochemical data. Similarly, the explanation of how depleted isotopic signatures
might develop in the lithospheric mantle is not convincing, as no arguments are presented
against the alternative hypotheses of Smith (1983).
Skinner (1989) does not present any evidence to explain why the partial melting
processes leading to the generation of kimberlite and orangeite must be initiated by the
contemporaneous introduction of asthenospheric volatiles. Further, there is no discussion
as to why the model does not appear to be valid for other kimberlite provinces, a
prerequisite for any general petrogenetic model.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 341

Unfortunately, there is no experimental evidence to support the concept that the rate
of melting is related to the composition of the mantle. Skinner (1989) does not address
the question of the modal mineralogy of his heterogeneous mantle sources, but does imply
that both are formed by similar metasomatic processes; hence, although their contents of
incompatible elements might differ, they may be mineralogically similar. Both would
therefore have very similar solidus temperatures in contradiction to Skinner's model.
Other problematic aspects ofthe model hinge upon the validity of Wyllie's (1980)
model of kimberlite genesis and especially with regard to Skinner's uncritical application
of this model to orangeites.
In conclusion, Skinner's (1989) model, at first sight, appears to provide a catholic
interpretation of kimberlite and orangeite petrogenesis; however, upon examination it is
found wanting, as it is based upon novel unsupported interpretations of geochemical and
isotopic data combined with numerous poorly-defined petrogenetic processes. It is open
to speculation as to why Tainton (1992), Tainton and McKenzie (1994), and Fraser and
Hawkesworth (1992) have refrained from any discussion of Skinner's (1989) model.

4.4.5. Redox Melting


Green et al. (1987), Taylor and Green (1989), and Foley (1988, 1989) have proposed
a mechanism for magma production in the upper mantle, termed "redox melting." In this
scheme, melting of upper mantle peridotites is triggered by the oxidation of reduced
C-O-H fluids (CIt! + H2) to carbon (graphite or diamond), C02, and H20. Production
of water results in a reduction of solidus temperatures and initiates melting. In this
hypothesis, melting is caused by volatile introduction rather than temperature increase.
The reduced volatiles are considered to originate in the lower mantle as a consequence
of continued degassing of the primordial earth. Operation of redox melting depends upon
the existence of a reduced asthenosphere if02 < iron-wustite buffer + 1-2 log units) and
an oxidized lithosphere if02 > magnetite-wustite buffer or quartz-fayalite-magnetite
buffer). Oxidation of CIt! and H2 will occur at the interface between the reduced and
oxidized domains (Figure 4.14), and carbon may be precipitated at or below the litho-
sphere-asthenosphere boundary as microdiamonds. C02 produced during oxidation may
carbonate lithospheric peridotites or remain dissolved in the partial melt.
The initial stage of redox melting involves oxidation of CIt!-bearing fluids to
H20-rich fluids. Migration of these fluids and reaction with preexisting minerals will lead
to formation of hydrous phases at all levels in the lithosphere. With the introduction of
more fluid, increased activities of water will eventually cause partial melting. As the water
produced at the oxidation zone front migrates upward, melting will occur in the overlying
oxidized zone (Figure 4.14). The nature of the melts produced will depend upon the
lithosphere composition and the rate of oxidation of the fluid.
Foley (1988) suggests that the oxygen fugacity of the lithosphere is such that
carbonates are present as subsolidus phases. As a H20- and CIt!-rich fluid moves slowly
into the lithosphere and oxidizes to H20-rich compositions, melting should occur in the
presence of H20- and C02-rich fluids. Melts will be restricted to silica-poor compositions
which may be kimberlitic or melilititic at high and low pressures, respectively. If the fluid
moves rapidly and equilibrium is not maintained, it will not oxidize completely and
342 CHAPTER 4

KIMBERLITIC
MELT

OXIDISED
LITHOSPHERE

rn MELT W~\d ZONE OF PARTIAL MELTING

Figure 4.14. Contrasting origins for kimberlites and lamproites according to the redox melting model of Foley
(1988). (A) Kimberlites are formed during the initial stage of melting where oxidation fronts are relatively well
defmed and the oxidized zone is fertile. Partial melting is induced by the oxidation of hydrogen and methane
to water. (B) Lamproites are formed in subsequent stages of melting from mantle which is depleted in major
elements but enriched in incompatible elements. The reduced nature of this region results in H20-rich, C02-free
melting at 1.5-2.5 GPa. C02-bearing olivine lamproites (and perhaps orangeites) may originate at 4.5-5.5 GPa,
where oxidized blocks persist in regions which are primarily reduced. See text for further explanation (after
Foley 1988).

methane-bearing fluids will encounter carbonated peridotites. Carbonate should be


consumed by decomposition to graphite or diamond and more water will be produced.
High activities of water may cause melting to occur while carbonates persist. In this
disequilibrium process, restricted amounts of low-silica melt (kimberlite) may be pro-
duced before continued introduction of methane causes all of the carbonate to be reacted
out. Subsequent to exhaustion of carbonate, more voluminous Si02- and H20-rich melts
will be produced. These might be basanitic to nephelinitic in character.
Foley (1988) notes that lamproitic melts may form as a consequence of renewed
volatile introduction when Cli4-rich fluids infiltrate a mantle region which has been
previously depleted in Ca, Na, and AI. Melts produced from these regions under reducing
conditions will have a depleted major element signature and resemble lamproites provided
that metasomatic enrichment in K, Ti, Zr, and REE, etc., has occurred prior to the second
period of melting. Foley (1988) considers that the degree of geochemical depletion and
intrinsic oxidation state in the region of melting may explain the difference between
kimberlitic and lamproitic melts. Olivine lamproites may originate at similar depths to
PETROGENESIS OF ORANGEITES AND KIMBERLITES 343

kimberlites but from a depleted and reduced mantle formed during an earlier period of
melting. Leucite lamproites may form at lower pressures from similar sources in H20-
and F-rich regimes (Figure 4.14).
Foley's (1988) redox melting mechanism is an extension of the volatile-induced
melting hypotheses of Wyllie (1980, 1989a,b) and Bailey (1983, 1993). It provides a
physicochemical explanation for melting processes which link diverse continental alka-
line magmas to a common source. The model may be applicable to lamproites and, by
analogy, to orangeites, if the genesis of these lithospheric magmas is initiated by volatile
uprise. The model is less attractive for the generation of kimberlites, as their parental
magmas are probably formed by decompressional melting of diverse carbonated-ultrabasic
assemblages in the asthenosphere (4.3.1, 4.3.2).
In common with all volatile-induced melting hypotheses, the model relies on the
introduction of volatiles from unknown sources and might be considered untenable on
this basis alone.

4.5. PETROGENESIS OF THE ORANGEITE CLAN


The petrogenesis of the orangeite clan comprises four principal stages: development
of the mantle source; melting of the source; contamination and evolution of the magma
during ascent through the mantle into the crust; post-emplacement crystallization. Any
hypotheses regarding the genesis of the clan must account for the following observations:
1. Orangeites are derived from sources with time-integrated Rb/Sr and SmlNd
ratios which are higher and lower, respectively, than those of bulk earth. These
sources have been isolated from convecting mantle for 0.5-2 Ga.
2. Individual orangeites differ in their isotopic composition within and between
intrusions.
3. Multistage and/or multi component models are required to explain the Sr, Nd,
and Pb isotopic compositions.
4. The bulk rock major and trace element compositions of orangeites do not
represent those of their parent magmas.
5. Orangeite magmas have been extensively contaminated by xenocrysts derived
from disaggregated mantle-derived ultramafic rocks.
6. The composition of the parental magma of the clan is unknown, although
orangeites appear to have formed from ultrapotassic peralkaline magmas.
7. Orangeites are strongly enriched in REE, Sr, Ba, Rb, and other incompatible
elements.
8. Orangeites exhibit significant negative K, Rb, and Sr anomalies in primitive
mantle normalized extended incompatible-element diagrams. The negative Ta,
Nb, and Ti anomalies, which are characteristic of lamproites, are not present.
9. Individual fields differ in their petrographic character. Phlogopite, spinel, apa-
tite, and carbonate are the only principal primary liquidus phases in some
orangeite magmas, whereas, in others, these minerals are accompanied by
diopside. Some of the latter crystallize late-stage sanidine and potassium
richterite.
344 CHAPTER 4

10. The crystallization of primary olivine probably does not playa significant role
in orangeite evolution, although small amounts of primary microphenocrystal
olivine may crystallize prior to emplacement. In contrast to archetypal kimber-
lites second-generation subhedral primary groundmass olivines are rare. Many
orangeites are characterized by extensive pre-emplacement crystal accumula-
tion of phlogopite.
11. Orangeites have closer mineralogical and geochemical affinities to lamproites
than to archetypal kimberlites. Thus, experimental studies oflamproite genesis
might have a bearing on orangeite petrogenesis.
12. Partial melting of H20- and carbonate-bearing, non-Iherzolitic sources might be
involved in the formation of orangeites.
13. To this date orangeites have been found only in southern Africa where they are
confined to particular regions of the Kaapvaal craton, the majority occurring in
the southwestern area.
14. Although the Kaapvaal craton has been the site of several periods of kimberlite
magmatism, orangeite magmatism is restricted to Upper Jurassic and Lower
Cretaceous times.
15. Emplacement of orangeites was preceded by extensive continental flood basaltic
magmatism which apparently did not interact with the sources of orangeites.
There are no simple unambiguous explanations for the above observations. Given
that orangeites have only recently been subjected to detailed study, it is unreasonable to
expect that, on the basis of our current knowledge, we should be in a position to devise
a distinctive comprehensive petrogenetic hypothesis which will stand the test of time.
Consequently, the speculations concerning their genesis presented below must be re-
garded as preliminary and subject to change in the light of future observations and
experimental studies.

4.5.1. Development of the Source


The restriction of orangeites to the Kaapvaal craton of South Africa, coupled with
the isotopic data indicating derivation from ancient enriched material, suggests that their
sources are located in the nonconvecting lithosphere. There is no convincing evidence
for an asthenospheric source, because, if there were, orangeitic magmatism should be
widespread throughout the world. Consequently, the initial stages in the formation of the
orangeite source are believed to be related to the development of the Kaapvaal craton.
Detailed discussion of the latter is beyond the scope of this work, but some general points
concerning the structure and evolution of the craton are required to appreciate where and
how orangeite magmas are generated. Source formation is considered to have taken place
in two steps: development of a deep continental root of depleted harzburgite, and
production of incompatible element-rich domains in this root by metasomatism (sensu
lato). The depth to which the continental root extends is extremely important in placing
constraints on the maximum depths of origin of lithosphere-derived melts; therefore,
evidence pertaining to the location of the lithosphere-asthenosphere boundary is dis-
cussed here at some length.
PETROGENESIS OFORANGEITES AND KIMBERLITES 345

4.5.1.1. Continental Roots


Currently, continental lithospheric plates are considered to consist of the crust and
uppermost part of the mantle. The crust consists of a stable ancient nucleus, termed a
"craton," which is surrounded by mobile belts and may be covered with younger platform
sedimentary rocks. Cratons consist of late Archean or early Proterozoic rocks which
behave as rigid units and react to external tectonic forces by epeirogenic faulting and
uplift rather than by folding.
Most geophysicists and petrologists agree that the lithospheric mantle consists
predominantly of low-density peridotite depleted of its low-melting "basaltic" fraction.
Therefore the lithospheric mantle is less dense and substantially more refractory than the
underlying asthenospheric mantle. Lithospheric mantle is rigid, and heat transfer occurs
mainly by conduction. The chemical boundary layer, represented by depleted subconti-
nental mantle, stabilizes the continental root against thermal disruption by the convecting
asthenospheric mantle. Cratons are attached to the lithospheric mantle root and move with
it during plate motion. Depleted lithospheric mantle is subjected to metasomatism and
intrusion by asthenospheric and lithospheric melts (Jordan 1978, Jeanloz 1987, James
1989).
Although there is agreement concerning the general character of the continental
lithosphere, there is none regarding the depth of the lithosphere-asthenosphere boundary.
Models based upon seismological, mechanical, rheological, thermal, or compositional
grounds typically are not in agreement (James 1989). However, the received opinion
holds that continental roots extend no deeper than 200--250 km (Anderson 1987, Boyd
and Gurney 1986). In contrast, Jordan (1978) has presented persuasive arguments holding
that continental roots, termed the "tectosphere," extend to depths of 300 km and perhaps
even deeper. Resolution of this question by the geophysical community is important as
the depth of the boundary constrains the maximum depth of generation of lithospheric-
derived magmas and has important ramifications regarding the fate of underplated
subducted oceanic material. Boyd and Gurney (1986) have suggested that the low
geothermal gradients and deep roots, characteristic of cratons today, were initially
established more than 3 Ga ago. Although the deep root may have been initiated in the
early Archean (see below), current models of komatiite (Hess 1989) and lithosphere
genesis (Herzberg 1993) require high heat flows at that time, in contradiction to Boyd
and Gurney's (1986) hypothesis.
Cratonic root formation has been linked with the initial stages of continental crust
separation and the extraction oflarge volumes ofkomatiitic (Nixon et at. 1987, Boyd and
Gurney 1986, Canil 1991) or basaltic magmas (Jordan 1978). The origins of the li-
thospheric peridotites comprising the Kaapvaal craton root have been reviewed by
Herzberg (1993), who concludes that they originated as harzburgitic rocks formed as
either residues or cumulates from a source material enriched in Si02, relative to normal
mantle peridotite or its pyrolite analogue. These harzburgites could have crystallized as
cumulates from an ultrabasic magma richer in Si02 than most komatiites. Formation of
these magmas would have required extensive (>50%) melting of the primordial mantle.
Thus, Herzberg (1993) sees cratonic root formation as being intimately linked with early
Archean differentiation processes and confined to these times of high heat flow. In
346 CHAPTER 4

agreement with Jordan (1978) and Pollack (1986), Herzberg (1993) suggests that the
FeO-poor Kaapvaal peridotites, having a reduced density compared to prirriltive mantle,
would lead to the establishment of a "buoyant raft of lithosphere" isolated from the rest
of the mantle and acting as a foundation for the Archean continental crust.
Continental roots are thus considered to originate in the early Archean and consist
of depleted harzburgite. Formation of garnet harzburgite and garnet peridotite from this
material subsequently occurs as a result of ex solution of garnet and clinopyroxene from
orthopyroxene followed by recrystallization (Canil 1991, Cox et al. 1987).
Depleted harzburgite cratonic roots subsequently form a substrate which may be
impregnated by metasomatic fluids or small-volume incompatible-element-enriched
melts derived from the underlying asthenosphere. These may modify the composition of
the lithosphere without introducing melting. Over the course of aeons this influx of
incompatible element-rich matter produces geochemically and isotopically heterogene-
ous domains in the non-convecting lithosphere (4.5.1.3).

4.5.1.2. Depth oj Origin ojOrangeite Magmas


The depth and topography of the lithosphere-asthenosphere boundary of the
Kaapvaal craton is not well-established. The current structural model of the craton is based
primarily on the depth of origin of the peridotite xenolith suite found in kimberlites.
Boyd and Nixon (1975) initially claimed that equilibration parameters for these
xenoliths may be used to devise an upper mantle stratigraphy. Studies of xenoliths derived
from northern Lesothan kimberlites suggested the existence of two suites of xenoliths: a
relatively cool suite of granular depleted garnet peridotites and a relatively hot suite of
Fe-rich or "fertile" sheared or deformed peridotites. Boyd and Nixon (1975) and Finnerty
and Boyd (1987) consider that the xenolith suite defines a paleogeotherm with an
inflection marking the depth of the lithosphere-asthenosphere boundary at 150-160 km.
Granular and sheared xenoliths are considered to represent depleted lithospheric and
asthenospheric mantle, respectively. MacGregor and Manton (1986), Shervais et al.
(1988), and Taylor and Neal (1989) have suggested that the paleogeotherm inflection is
due to underplating of the craton with eclogitic material derived from subducted oceanic
crust.
Other studies of peridotite xenoliths have shown no general correlation between
degree of deformation, equilibration temperature, and depth of origin. Carswell and Gibb
(1987, and references therein), in particular, have suggested that inflected paleogeotherms
are an artifact of a particular choice of geothermobarometers and have no geological
significance. The "fertile" xenoliths are also considered to be formed in thermal-metasomatic
aureoles associated with kimberlite intrusions rather than being samples of undepleted
asthenosphere (Mitchell 1984b, Harte 1983).
Equilibration temperatures and pressures of peridotite xenoliths found in kimberlites,
regardless of the geothermobarometer chosen, suggest that none have originated at depths
in excess of 220-250 km. Whether the xenoliths originating below about 160-km depth
are asthenospheric or lithospheric remains a moot point.
In contrast to kimberlite, peridotite xenoliths are relatively rare in orangeites, having
been found only in the Finsch diatreme (Shee et al. 1982). The presence of olivine and
chrome garnet macrocrysts attests to the contarrilnation of orangeite magmas by ul-
PETROGENESIS OF ORANGEITES AND KIMBERLITES 347

tramafic material, but it is unclear whether the disaggregated protolith is peridotite,


harzburgite, dunite, or wehrlite. Garnets derived from a wehrlite paragenesis have been
reported from Newlands by Schulze (1989). Orangeites also contain xenocrystal subcal-
cic chrome pyropes considered to be derived from disaggregated garnet harzburgites
(Pokhilenko et at. 1977), yet examples of their parental xenoliths are exceedingly rare in
the Kaapvaal craton (Nixon et al. 1987).
The xenolithic suite of orangeites appears to be dominated by eclogites (MacGregor
and Carter 1970, Hatton 1978, Taylor and Neal 1989, Gurney 1989). It is known that
xenolith suites from geographically closely-related kimberlites are significantly different
as a result of random sampling of the upper mantle during their ascent (Nixon 1987,
Carswell and Gibb 1987, Mitchell et al. 1980). However, it is not known whether the
paucity of peridotite xenoliths in orangeites, relative to kimberlites, is the result of such
random sampling of the mantle or a reflection of some fundamental difference in the
rheology and ascent rates of orangeite magmas relative to kimberlites. On the basis of
present evidence it appears that orangeites do not typically sample the upper lithospheric
mantle postulated to be the source of the granular depleted lherzolite suite (see also 4.5.3).
Estimation of the depth of origin of eclogites is more difficult than for lherzolites
because geothermobarometry is based upon garnet-clinopyroxene Fe-Mg exchange
equilibria which are not pressure and temperature independent (Ellis and Green 1979).
Hence, equilibration temperatures may be calculated only by assuming a particular
pressure; nevertheless, eclogites hosted by orangeites have been shown to have equili-
brated over a wide range of temperatures (800-1050 DC) at pressures of 3.0--5.0 GPa
(100--160 km) by Smyth and Caporuscio (1984), MacGregor and Manton (1986), and
Taylor and Neal (1989).
The range in equilibration parameters is a reflection of the diverse origins postulated
for mantle-derived eclogites. Currently, the favored viewpoint is that eclogites may
represent either subducted underplated oceanic crust or "basaltic" magmas crystallizing
at high pressures. Eclogites belonging to both parageneses are represented in the
Bellsbank and Roberts Victor eclogite suites. Diamond-bearing eclogites clearly originate
from within the diamond stability field, and Hatton and Gurney (1979) have calculated
equilibration over a range of 1020-1140DC at 4.2-4.5 GPa (137-147 km) for an example
from Roberts Victor. A coesite grospydite from Roberts Victor is determined by Wohletz
and Smyth (1984) to have originated at 160-km depth (4.9 GPa) at 1060DC.
Xenoliths, supposedly originating in the transition zone (Moore et al. 1991, Sautter
et al. 1991), have not been reported from orangeites. MARID-suite and metasomatized
peridotites (Erlank et at. 1987) are also absent.
It has long been known that diamondiferous kimberlites and orangeites are found
only in regions underlain by Archean cratonic basement rocks (Clifford 1966). It is now
believed that most diamonds are xenocrysts in their transporting magmas and are derived
from lithospheric mantle roots where they have been preserved since Archean or Protero-
zoic times (Richardson et al. 1984, Meyer 1987).
Consequently, a further constraint on the depth of origin of orangeite is provided by
inclusions in diamonds. Meyer (1987) and Gurney (1989) reviewed this topic and noted
that diamonds contain two suites of silicate inclusions, termed the "peridotitic" (P-type)
and "eclogitic" (E-type) suites because of their similarity to minerals comprising peri-
348 CHAPTER 4

dotite and eclogite xenoliths, respectively. However, the inclusions are compositionally
distinct from the minerals of these xenoliths. There is no correlation between xenolith
suite and type of inclusions in diamonds within a given kimberlite or orangeite. Both
suites of inclusions may occur in the total diamond population, although one or the other
may be dominant at a particular locality. E-type diamonds may occur in intrusions that
are dominated by peridotite xenoliths, and vice versa. Thus, at Roberts Victor, diamonds
with P-type inclusions are the most common, yet the xenolith suite is dominated by
eclogites. At Finsch, E-type diamonds become more abundant relative to P-type with
increasing diamond size.
Currently it is believed that most P-type diamonds are not derived by disaggregation
of coarse granular garnet lherzolite protoliths, although E-type diamonds may be derived
from material similar to eclogite suite xenoliths. The source of P-type diamonds is now
believed to be garnet harzburgites derived from the highly depleted cratonic roots of the
continent (Pokhilenko et al. 1977, Nixon et al. 1987).
Equilibration temperatures and pressures calculated for both suites of inclusions have
been summarized by Meyer (1987). The most reliable data for P-type inclusions range
from about 900-1300°C at 4.5--6.5 GPa (147-212 km). At an assumed pressure of 5 Gpa
most E-type inclusions fall in the stability field of diamond. It is important to note that
the data imply that both suites of diamonds appear to originate at the base of the
lithosphere (Meyer 1987, Gurney 1989).
In summary, xenoliths and diamond inclusion studies suggest that orangeites sample
material derived from near the base of the lithosphere within the diamond stability field
at depths of 150-200 km. Clearly, the magmas must be derived either from these depths
or deeper. The absence of asthenospheric xenoliths precludes the latter. Consequently, it
is suggested that the sources of orangeites are located in the cratonic root of the Kaapvaal
craton at, or slightly above (50 km), the lithosphere-asthenosphere boundary. This region
is considered to consist of ancient depleted garnet harzburgite and garnet dunite, together
with eclogites representing ancient recrystallized subducted underplated oceanic crust
and ponded crystallized asthenospheric melts.

4.5.1.3. Compositional Heterogeneities-Veined Harzburgites


The development of compositional and isotopic heterogeneities in the lithospheric
mantle is currently considered to result from the introduction of incompatible-e1ement-
rich hydrous fluids or small-volume silicate melts (Hawkes worth et al. 1985, Menzies et
al. 1987, Dawson 1984, Haggerty 1989a,b). Commonly, compositional changes resulting
from both processes are referred to as metasomatism. However, Hawkesworth et al.
(1985) correctly note that infiltration of silicate melts is not in itself metasomatism,
although such melts may cause localized metasomatism in adjacent rocks.
The extensive mineralogical and geochemical evidence supporting the existence of
metasomatism and melt infiltration processes in the lithospheric mantle has been sum-
marized by Erlank et al. (1987), and Menzies et al. (1987). There is general agreement
that the end result of these processes is the production of a wide variety of veins with or
without metasomatic aureoles. The veins are believed to form a stockwork within the
depleted harzburgitic or peridotitic mantle (Menzies et al. 1987, O'Brien et al. 1991,
Foley 1992b).
PETROGENESIS OF ORANGEITES AND KIMBERLITES 349

The mineralogy of the veins is believed to consist primarily of widely varying modal
amounts of amphibole (kaersutite or potassium richterite), Ti-phlogopite, apatite, diop-
side, K-Ba titanates, ilmenite, and rutile. Menzies et al. (1987) suggest that veins
generated from hydrous metasomatic fluids consist of glimmerites and MARID-suite
material with aureoles of phlogopite- and/or phlogopite-K-richterite peridotite. In con-
trast, small volume melts are believed to be essentially of basanitic compositions which
crystallize to kaersuite, apatite, and mica. Peridotitic wall rocks may be subjected to
Fe-Ti-rich metasomatism with the formation of pargasite/kaersutite and high Ti-mica.
Carbonates may (Thibault et al. 1992) or may not (Wyllie 1980) be present.
Experimental studies have indicated that most ofthe minerals postulated to be present
in the veins are stable at pressures up to and exceeding those encountered at the base of
the lithosphere, i.e., 5-6 GPa. Thus, K-richterite is stable up to 14 GPa (Tr0nnes et al.
1988, Sudo and Tatsumi 1990, Foley 1991), phlogopite to 9 GPa and phlogopite plus
diopside to 7 GPa at 1200°C (Luth et al. 1993), and K-Ba titanates (hollandites,
crichtonites) to 5-6 GPa (Foley et al. 1994). While these experimental studies demon-
strate the stability of the incompatible-element-rich minerals of interest at the base of the
lithosphere, there are unfortunately few studies of the melting of vein-type assemblages
at these pressures (see 4.2.3,4.2.4).
It should be stressed that there is very little quantitative evidence regarding the
mineralogy and modal composition of the veins, in particular those postulated at the base
of the lithosphere. Vein mineralogy certainly will vary as a function of depth and the
source of the infiltrating melts and fluids. Experimental (Thibault et al. 1992) and
theoretical (Haggerty 1989a,b) studies suggest that mineral zoning of metasomes will be
commonplace. One undesirable, but unavoidable, aspect of this lacuna in our knowledge
is that petrologists commonly use the compositions of supposed partial melts to indulge
in speculation concerning the source mineralogy. This is unsatisfactory as hypotheses not
constrained by experimental data are infinitely flexible with regard to the hypothetical
source mineralogy.
All of the veins are considered to have high Rb/Sr and U/Pb ratios and low SmlNd,
relative to bulk earth. Undoubtedly, metasomatism and infiltration is not confined to
Archean times and must persist to the present day. Consequently, the Sm-Nd model ages
of the enrichments range from 0.9 to 2 Ga (Menzies et al. 1987). The ages of the MARID
suite and phlogopite K-richterite peridotite have been shown to be 84-87 Ma (Erlank et
at. 1987, Kinny and Dawson 1992) and contemporaneous with emplacement of arche-
typal kimberlites in the Kaapvaal craton.
To explain the isotopic signatures of ultrapotassic lithospheric melts, some of the
enriched veins are required to persist in isolation from convecting mantle for periods on
the order of 1-2 Ga. This constraint is required to allow mantle domains to develop
distinctive enriched isotopic signatures. Note that although Pb isotopic studies (3.8.2)
suggest the initial stages of vein formation may have involved two or more episodes of
infiltration and metasomatism, they have subsequently remained as closed systems
apparently unaffected by Phanerozoic magmatic events.
The antiquity and suitability of the veins considered as a source of ultrapotassic melts
have been demonstrated by O'Brien et al. (1991), who determined that a glimmerite vein
in a mantle-derived harzburgite from the Wyoming craton was enriched in Ba, Rb, and
350 CHAPTER 4

REE, with an ENd of -33 at 52 Ma and a Nd model age of 2.57 Ga, relative to depleted
mantle.
While there is general agreement among petrologists regarding the existence of a
stockwork of veins in the lithospheric mantle, there is no agreement as to the ultimate
source ofthe fluids and melts from which they are derived. Hypotheses include: devola-
tilization of underplated subducted material (Sekine and Wyllie 1982, O'Brien et ai,
1991); volatile release from the asthenosphere (Wyllie 1980, Foley 1988); crystallization
of asthenospheric basaltic (Hawkes worth et al. 1985, Menzies et al. 1987, Thibault et al.
1992); kimberlitic (Menzies et al. 1987, Kinney and Dawson 1992); or carbonatitic
(Haggerty 1989a,b) magmas trapped in the lithosphere or at the lithosphere-asthenosphere
boundary.
It is to be expected that metasomatism and melt infiltration will not necessarily have
the same character or take place to the same extent everywhere within the lithosphere.
Locations may be controlled by pre-exisitng fractures or zones of weakness. The topog-
raphy of the Kaapvaal lithosphere-asthenosphere boundary is not known. Most repre-
sentations (Mitchell 1991b, Haggerty 1989a,b, Boyd and Gurney 1986) depict the
boundary and craton root as a smooth convex-down keel; however, given the known
complexity of geological processes, it is entirely possible that this representation is
incorrect. Immense corrugations may exist at the lithosphere-asthenosphere boundary
resulting from asthenospheric upwelling, pileup and/or sinking of underplated subducted
slabs, or the presence of ancient welded mobile zones oflower-density material, e.g., the
Limpopo belt. Such features may channel fluids and melts to particular portions of the
lithosphere-asthenosphere boundary.
With regard to orangeites, it is evident that their primary expression is in the
southwestern part of the craton. This may result from either preferential modification of
the lithosphere underlying the region or selective melting of a portion of an extensively
modified craton. With regard to the former it is instructive to note that Tainton (1992) and
Smith (1983) consider that, on the basis ofNd isotopic compositions, the last enrichment
and closure of the sources of orangeites occurred about 0.7-1 Ga ago and was possibly
related to the later stages of the accretion of the Namaqua Province to the Kaapvaal craton
(Tainton 1992).
Hence, devolatization of underthrust Namaqua material, according to the models of
Sekine and Wyllie (1982), could have restricted the sources of orangeites to the southern
margins of the Kaapvaal craton root. Such a model could explain the paucity of orangeites
in the northern parts of the craton and their absence in the Zimbabwe craton. However,
other explanations of the distribution are possible (4.5.2.1).
In summary, the depleted continental lithosphere is undoubtedly extensively veined
by ancient and modern incompatible-element-enriched material. Partial or complete
(4.5.2.2) melting of such veins provides suitable sources for incompatible-element-
enriched lithospheric magmas. The sources of these veins are, as yet, undetermined and
may be related to either asthenospheric or subducted material.
It is concluded that the stages in the production of the sources of orangeite magmas
are:
PETROGENESIS OF ORANGEITES AND KIMBERLITES 351

1. Fonnation of the depleted harzburgitic roots of the Kaapvaal craton in the early
Archean (Herzberg 1993, Jordan 1978). This buoyant depleted root acts as
substrate for subsequent metasomatism and melt infiltration.
2. Emplacement of incompatible element-enriched veins during the interval ca. 3
Ga to 1 Ga. The exact modal mineralogy of the veins is unknown, but they
undoubtedly contain potassium richterite, Ti-phlogopite, and apatite. The source
of the vein material is unknown, but may be derived from asthenospheric melts
contributing to further growth of the cratonic roots and/or subducted oceanic
material. Oceanic crust may have been subducted under the Kaapvaal craton
during collision of the Kaapvaal and Zimbabwe cratons (Helmstaedt and
Schulze 1989) and subsequent to the accretion of the Namaqua Province. Thus,
eclogites of two ages may be underplated and incorporated with the cratonic
root. Therefore, eclogite xenoliths found in orangeites could conceivably have
two provenances (however see 4.8), although only Archean isotopic signatures
have so far been found for eclogites from Roberts Victor (Kramers 1979) and
Bellsbank (Neal et al. 1990).
3. Stabilization of the craton subsequent to the Namaqua orogeny and geochemical
isolation of orangeite sources for 0.5-1 Ga. There were apparently no thennal
effects on, or compositional changes to, the sources during the Pan-African
orogeny and the assembly of Gondwanaland or its subsequent dismemberment
in Jurassic times.

4.5.2. Melting of the Source


Aspects of the composition and mineralogical character of orangeite magma sources
may be inferred from experimental studies of orangeites and lamproites (see 4.2), in
conjunction with compositional (Chapter 3) and mineralogical (Chapter 2) constraints
derived from the study of rocks of the orangeite clan. However, it must be emphasized
that the actual composition of any orangeite magma is not known, and thus we have no
definitive knowledge of the composition of the initial partial melt or primitive magma.
Consequently, hypotheses concerning the source mineralogy or extent of partial melting
of that source must be regarded as speculative and preliminary.
Bearing the above in mind, the source should contain C02 and H20 and minerals
which sequester alkali, alkaline earth, rare earth, and second-period transitional elements.
Derivative melts must be rich in potassium and able to crystallize abundant micas,
significant amounts of apatite and carbonate, and accessory titanates and spinels. Genetic
models must explain the presence of diopside in some orangeites and the formation of
evolved orangeites.
The simplest source having the above attributes is one composed of Ti-phlogopite
and a phosphate. The latter could be apatite, a dense polymorph of apatite (Murayama et
al. 1986), or a complex phosphate such the K-Ba phosphate (16-20 wt % K20, 19-21
wt % BaO, 6 wt % CaO, 9 wt % MgO), which has recently been recognized by Mitchell
and Edgar (unpublished data) in 5-7 GPa near-solidus experimental studies oflamproites.
Whether the presence of other minerals is necessary, is unknown. One probable
candidate which may playa particular role in the fonnation of evolved orangeites is
352 CHAPTER 4

potassium richterite, as it may contribute Ti and silica-rich melt to the magma. Foley
(1991) and Trl1lnnes et al.(l988) have shown this amphibole to be stable throughout the
lithosphere. Luth et al. (1993) have noted that potassium richterite is stable to greater
depths than phlogopite, and, during subduction may incorporate all of the H, K, and
large-ion lithophile elements released by phlogopite breakdown at about 200-km depth.
Subsequent richterite breakdown, perhaps near the top of the transition zone, may cause
partial melting and release of fluids and volatiles into the craton and cause reprecipitation
of potassium richterite and/or phlogopite. The common occurrence of potassium
richterite, together with phlogopite, in metasomatized peridotite and MARID-suite
xenoliths supports the contention that this amphibole is a typical component of the vein
assemblage.
Experimental studies of the systems phlogopite-apatite, K-richterite-apatite, and
K-richterite-phlogopite (see 4.2.4), clearly demonstrate that veins composed of Ti-K
richterite, Ti-phlogopite, and apatite will have low solidus temperatures. Melting of these
assemblages will produce K~rich melts with the general compositional characteristics
ascribed to orangeite magmas. Unfortunately, experimental studies of this ternary system
at 5-7 GPa have not been undertaken.
The liquidus studies of lamproites and the melting of phlogopite clinopyroxenite
suggest that olivine is not present in the sources of orangeites. Whether diopside occurs
in the veins is unknown, as the breakdown of richterite will supply all of the components
required for the crystallization of diopside from the melt at lower pressures.
The presence of carbonate in orangeites requires a source of C02. It is unknown
whether this exists as solid carbonate in the veins or is subsequently introduced during
the events which initiated partial melting of the veins. Both sources are, of course,
possible.
It is concluded here that the incompatible element-rich veins proposed as the sources
of orangeite magmas consist of varying modal amounts of Ti-K richterite, Ti-phlogopite,
and apatite. Diopside and carbonate mayor may not be present. Ilmenite, rutile, hol-
landites, crichtonites, and wadeite are not required in the source as contributors ofTi, Ba.
and K to the melt as they are unlikely to take part in near-solidus melting reactions. The
absence of negative Ti-Nb-Ta anomalies in extended incompatible-element distribution
diagrams for orangeites is evidence that these minerals are not present as modally
significant residual phases and. hence. were unlikely to have been major components of
the initial assemblage.
Production of melts from veins in a harzburgitic substrate may be considered with
respect to the causes and mechanism of melting.

4.5.2.1. Causes of Melting


Partial melting of the source may be initiated by any of the following processes
(Mitchell and Bergman 1991):
1. Simple thermally induced melting as ambient temperatures exceed those of the
solidus. No new components are introduced. and the veins melt directly at the
fringes of large mantle plumes.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 353

2. Decompressional melting resulting from uplift, with no new components being


introduced. Uplift may be associated with mantle plumes or permissive plate
tectonics.
3. Introduction of volatiles, with melting resulting from reduction of the solidus
temperatures. The volatile influx may result from degassing of the Earth and
ascent of reduced fluids (Foley 1988, 1989, Taylor and Green 1989) or the
thermally induced decomposition of subducted hydrous phases at great depth in
the mantle (Ringwood 1989, Wyllie 1980).
4. Introduction of volatiles and other components. In this case the volatile flux
contains alkali and other incompatible elements. These may be C02- and
H20-rich limited partial melts derived from asthenospheric sources, i.e., they
are similar to the melts which caused the original enrichment of harzburgitic
mantle. This case may be considered as a further enrichment event.

Any combination of these processes may initiate partial melting. Processes 2 and 3
may be interrelated, as Bailey (1983) maintains that the introduction of volatiles will lead
to the formation of low-density minerals and ultimately to uplift of the lithosphere.
Process 4 has very important consequences given the possibility of modification of the
isotopic composition of the source (see below).
Processes I and 2 suggest melting may occur at the fringes ofthe large mantle plumes
which are currently believed to cause continental breakup and generation of continental
flood basalt provinces. The characteristics of such plumes and how they interact with the
lithosphere are reviewed by White and McKenzie (1989), Saunders et al. (1992), and
Anderson et ai. (1992).
Large plumes are believed to result from the upwelling of hot asthenospheric material
from the lower mantle. As the plume impinges upon the lithosphere it causes uplift,
thinning, and, ultimately, rifting. Igneous rocks are generated by decompressional melting
as the mantle rises passively beneath the stretched and thinned lithosphere. Plumes at the
lithosphere-asthenosphere boundary are believed to spread out as very large mushroom-
shaped bodies, typically 2000 km in diameter (White and McKenzie 1989). In this region
temperatures are raised 100-200°C above normal. The combined effects of uplift and
enhanced temperature are believed to be the principal cause of the formation of the vast
quantities of flood basalt associated with continental rifting.
With regard to continental lithospheric magmatism, interest in plumes is related to
the elevated temperatures at, and above, the lithosphere-asthenosphere boundary result-
ing from the mushrooming of the plume head. Thus, partial melting may be induced in
the lithospheric mantle above the head, or at the fringes, of a plume if ambient tempera-
tures are raised above those of material with low solidus temperatures. Melting of
lithosphere above plume heads is currently a popular means of generating the enriched
geochemical signatures of continental flood basalts (Ellam and Cox 1991, Mahoney et
at. 1989, Hawskesworth et at. 1986).
Melting at the fringes of plumes might be important for the generation of small
volume magmas such as orangeites, as only small increases in temperature, coupled with
slight uplift, may be sufficient to achieve suprasolidus conditions. Thus, with regard to
orangeites, melting of the source, in response to the uprise of large mantle plumes, is
354 CHAPTER 4

plausible, although it is shown in Section 4.4.3 that orangeites are not related to small hot
spots. Two plumes are required to explain the distribution and range in age of orangeites.
The thermal effects associated with one of these plumes may also explain the absence of
orangeites in many parts of southeastern Africa.
The genesis of the Late Jurassic orangeites (145-165 Ma) may be related to the
breakup of Gondwana which was initiated in the mid-Jurassic (170 Ma), consequent on
impingement of a large mantle plume on the eastern margin of the Kaapvaal-Zimbabwe
craton (Cox 1992, White and Mckenzie 1989). This plume is believed to be the source of
the vast quantities of Karroo basalts erupted in southeastern Africa. Recent geochemical
studies have concluded that many of the basalts are contaminated with lithospheric mantle
(Cox 1992, Ellam and Cox 1991). The contaminant has been identified by Ellam and Cox
(1991) as having isotopic and trace element characteristics similar to those oflamproites,
and is believed to be derived from metasomatized lithospheric mantle sources. Clearly,
this component could also be identified with the sources of orangeites.
Consequently, it is suggested that the absence of eitherorangeites or lamproites from
the northern Kaapvaal and Zimbabwe cratons is due to the destruction of their potential
sources ahead of the rising Karroo plume. The small amounts of partial melt produced
could easily be hybridized in voluminous basalt magmas without significant effects upon
their major element composition, while simultaneously changing their incompatible
element and isotopic abundances.
The late Jurassic orangeite magmas may be formed in the mid-Jurassic at the fringes
of the plume and remain as melts trapped in the mantle or late in the breakup process of
Gondwanaland as a consequence of slow conduction of heat to the outer fringes of the
plume. The former seems geologically unreasonable as it requires melts to exist for 25
Ma in the mantle and some further process to initiate intrusion. However, melt persistence
may be possible if solidus temperatures are low and the region is held at higher
temperatures during the existence of the plume.
Alternatively, the late Jurassic orangeites may have no direct relationship with the
Karroo plume and represent melts derived from remnants of original enriched mantle
which survived destruction because of their location at the fringes of the plume. In this
scenario, melting occurred subsequent to the decay of the Karroo plume and was initiated
by other processes. The latter must have been localized, otherwise orangeites of this age
would have been formed elsewhere in the craton.
It has been previously noted (Fraser 1987, Skinner 1989) that the main focus of
orangeite magmatism at the southwestern margin of the Kaapvaal craton is contempora-
neous with the opening of the South Atlantic and the eruption of the Parana and Etendeka
flood basalts. Figure 4.15 shows that orangeites are emplaced near the outer fringes of
the abnormally hot mantle associated with the Walvis (White and McKenzie 1989) or
Parana plume (Cox 1992). This association and the coincidence in ages cannot be
fortuitous and implies some connection between the onset of melting of the orangeite
sources and uprise of the plume. Partial melting of orangeite sources could be in response
to a combination of thermal effects and uplift at the fringes of the Wavis plume or be
entirely decompressional and in response to plume-related uplift outside the region of
enhanced temperatures. This scenario of plume-related magmatism does not involve the
addition of asthenospheric material to orangeite sources at the time of partial melting.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 355

I
South

o! 2000 km
! • ORANGEITES

Figure 4.15. Reconstruction of the South Atlantic at 120 Ma, shortly after the onset of seafloor spreading.
showing the distribution of Lower Cretaceous orangeites (after Skinner 1989) relative to the region of
abnormally hot mantle (dotted) above the Walvis or Parana plume (after White and McKenzie 1989). Solid
black areas show areas of extrusive basalts.

This constraint is apparently not violated for the Postmasburg, Barkly West, Winburg,
Kronstad, and Boshof orangeites, but may not be satisfactory in accounting for anomalous
Preiska area orangeites (see below).
In direct opposition to the above, Anderson et al. (1992) have concluded from seismic
tomographic studies that there is no evidence for the involvement of plume heads or
lithospheric modification in the genesis of the Karroo and Parana-Etendeka basalts.
Instead, these flood basalt provinces are considered to be related to plate tectonic
356 CHAPTER 4

reorganization over broad deep regions of hot, probably partially-molten mantle. Pieces
of thick continental lithosphere are believed to separate along pre-existing lines of
weakness without substantial rifting or thinning. Buoyant upwelling of the underlying
hot regions leads to further decompressional melting and production of voluminous
basaltic magmas. Clearly, if Anderson et al. (1992) are correct, plumes cannot play any
role in the genesis of orangeites.
Determining whether orangeites contain an asthenospheric component (process 4)
is extremely difficult, and hypotheses in favor of, or against, commonly hinge upon
particular subjective interpretations of isotopic compositions (see 3.8). Thus, the isotopic
composition of orangeites may be interpreted to represent:
1. Those of the sources with no mixing of individual batches of lithospheric melt
and no asthenospheric contributions.
2. Mixing of individual batches of lithospheric melt derived from veins with
slightly different time-integrated isotopic compositions and no asthenospheric
component.
3. Mixtures of asthenospheric and lithospheric melts. In this case the isotopic
composition of the melts does not represent that of the source. Potential end
members might include depleted asthenospheric MORB-type sources and a
highly enriched lithospheric end member with lower 143Ndll44Nd ratios and
higher 87Sr/86 ratios than observed in orangeites.
Unfortunately, all of the above can account for the intra- and inter-intrusion isotopic
variations found in orangeites, although there is no compelling geological or mineralogi-
cal evidence for the existence of an asthenospheric component. Contemporaneous asthe-
nospheric magmatism is apparently absent or restricted to younger Upper Cretaceous
kimberlites and melilitites. The presence of mineralogically anomalous orangeites in the
Prieska area (1.8.8) in itself does not indicate the involvement of an asthenospheric
component in their genesis, as the ilmenite, gamet, and pyroxene macrocrysts could well
be derived from an earlier period of kimberlitic or other type of magmatism. Note in
particular that the Sr and Nd isotopic compositions of these orangeites are similar to those
of other orangeites lacking macrocrysts (Figures 3.30, 3.31). Hence, even if contamina-
tion/mixing has occurred, it cannot have played a significant role in determining the
isotopic compositions of the Prieska orangeites.
The isotopically and mineralogically anomalous rocks of domain V of the Preiska
region (1.8.8) may represent archetypal kimberlite magmas contaminated by lithospheric
material whose genesis is totally unrelated to the geographically associated younger
orangeites and kimberlites. They certainly do not represent rocks derived from magmas
which are "transitional" between orangeites and kimberlite magmas, as suggested by
Skinner et at. (1994).
Addition of volatiles to lithospheric sources is a possible means of inducing partial
melting (Wyllie 1980, Bailey 1983, Foley 1988). However, this process commonly begs
the question as to the ultimate source of the volatiles, and appeals must be made to further
processes, which are impossible to verify, occurring in the transition zone, the lower
mantle, or at the lithosphere-asthenosphere boundary above subducted oceanic crust. It
is possible, but not highly probable, that the asthenospheric events giving rise to
PETROGENESIS OFORANGEITES AND KIMBERLITES 357

kimberlites may be preceded by the uprise of volatiles which add H20 and C02 to
orangeite sources and initiate partial melting. This scenario requires that asthenospheric
uprise and kimberlite segregation lag 20-25 Ma behind the ascent of the volatiles. The
hypothesis fails to explain the general absence of a temporal association between
ultrapotassic and kimberlite magmatism.
Bailey (1983, 1984, 1992, 1993) has consistently argued that continental lithospheric
magmatism occurs in response to plate tectonic events such as collisions and changes in
plate motion patterns. Thus, episodic intraplate magmatism is a consequence of external
forces acting laterally across the lithosphere rather than being initiated by mantle plumes.
Melt generation occurs in response to stress release in the lithosphere and/or volatile
fluxing.
The volatile fluxing hypothesis (Bailey 1984, 1992, 1993) requires the addition to
the lithosphere of volatiles derived from the deep mantle, a condition seemingly at
variance with Bailey's stance that asthenospheric activity plays no role in intraplate
magmatism. Regardless, Bailey's (1984, 1993) model assumes that for long periods the
lithosphere acts as a lid which permits only the slow leakage of volatiles from the
asthenosphere. Fractures developed or reactivated by tectonic processes external to the
craton may penetrate to the base of the lithosphere and provide channels for the escape
of the volatiles. These channels act as a focus for drawing volatiles and heat to one region
of the lithosphere-asthenosphere boundary. The process results in the build up, by
metasomatism, of the levels of volatile-bearing phases in regions adjacent to the channels.
Ultimately, melting of these regions will occur due to decompression consequent upon
uplift (Bailey 1984). The type of magma erupted depends on the local geothermal gradient
(Bailey 1993). Thus, kimberlites, lamproites, and carbonatites lacking associated silicate
melts would characterize cooler older lithosphere, and provinces with larger volumes of
alkaline silicate melts would be expected to occur in regions with higher geothermal
gradients, i.e., old mobile belts and rift zones. An important aspect of Bailey's (1993)
version of this model is that magmatism must be contemporaneous across the whole of
a tectonic plate. Accordingly, Bailey considers that Cretaceous kimberlitic and other
magmatism in Africa can be correlated with collision of the African plate with Europe.
Orangeites are believed by Bailey (1993) to be contemporaneous with other varieties of
platewide magmatic activity, and the apparent.age trend reported by Skinner (1989) to be
due to progressive structural readjustment of the craton as it drifts northward subsequent
to the breakup of Gondwanaland.
Bailey's hypothesis is important in that it attempts to place all late Phanerozic
magmatism in a single model; however, it fails as it requires that kimberlites, orangeites,
lamproites, melilitites, and nephelinites all be generated from the same source in the
lithosphere. Unfortunately, Bailey consistently ignores all the accumulated mineralogical,
geochemical, and isotopic evidence pointing to the derivation of these diverse magmas
from different sources and depths. Moreover, the model is specific to late Phanerozoic
African magmatism, and even the proponent has not investigated whether the model has
any general worldwide applicability.
With regard to orangeites, the model is unsatisfactory as, if Bailey (1993) is correct,
it would be expected that the low geothermal gradients found in other cratonic regions
of the African plate should result in the formation, in them, of contemporaneous
358 CHAYfER4

orangeites. However, orangeites are absent from these cratons, and other examples of
African ultrapotassic magmatism, such as the Quaternary-Recent kamafugites of the
Western Rift and the 220 Ma Kapamba lamproites of the Luangwa graben (Scott Smith
et al. 1989), are not contemporaneous.
In conclusion, it is apparent that the sources of orangeites are located in the
lithosphere, but whether melting is in response to passive or active tectonism is unknown.
The contemporaneity of the magmatism with continental breakup and the eruption of
flood basalt provinces is suggestive of some relationship with large asthenospheric
plumes. The latter may cause uplift and heating of orangeite source regions, but do not
apparently contribute asthenospheric material to the melts. Of course, there is no a priori
reason why all orangeite magmas should be generated by a single process, and attempts
to seek a definitive relationship to a particular tectonic mechanism may be inevitably
doomed to failure.

4.5.2.2. Melting of Veined Lithosphere


Foley (1992b) has reviewed the physical aspects of melt generation from a veined
lithospheric mantle and developed an important model of the melting process, with
particular relevance to the genesis of ultrapotassic rocks. The model is directly applicable
to the genesis of orangeites if they are formed by the partial melting of richterite-phlogopite-
apatite veins in a harzburgitic substrate.
Foley (1992b) introduced his vein-plus-wall rock melting mechanism to explain the
paradox of the failure ofliquidus experiments on ultrapotassic rocks to locate a peridotite-
like multiphase saturation point (Foley 1992a), and the conclusion derived from geo-
chemical (Mitchell and Bergman 1991) and other experimental studies (Edgar 1987,
Foley 1990) that these magmas might be derived from an incompatible element-rich mica
harzburgite.
Foley (1992b) notes that veins within a peridotitic or harzburgite mantle will have a
lower solidus temperature than their host, due to the high concentration of volatiles and
incompatible elements. Despite this difference, during progressive melting the veins and
host do not behave as distinct systems because an initial strongly alkaline melt, originating
exclusively from the vein assemblage, becomes progressively diluted by a peridotitic
component derived from the wall rocks. Hence, partial melts will be hybrids with
compositions dependent on the relative proportions of melts derived from the vein and
wall rock.
When partial melting begins, it will be centered on the vein assemblage. Some
accessory phases may be quickly eliminated, but the major minerals will remain stable
due to solid solution melting reactions. These reactions occur wherever a mineral is an
intermediate member of a solid solution series between two end members with very
different melting points. In this context, Foley (1991) has shown that the breakdown
temperatures of hydroxy- and fluor-amphibole differ by about 200°C (Figure 4.16).
Similar differences in stability may be expected for apatite and mica as a function of their
F/OH ratios. Hence, as minerals belonging to complex solid solution series melt, the
compositions of the derivative liquids and residual minerals will continuously change.
Further, the melting interval of a particular mineral may extend to temperatures above the
solidus of the surrounding wall rocks. Consequently, this phase will remain stable while
PETROGENESIS OFORANGEITES AND KIMBERLITES 359

K - richterite F/OH
60
...
C
.Q
~ 50

IJ..I
a::
:::> 40
(f)
(f)
IJ..I
a::
a. 30

1000 1200 1400


TEMPERATURE (OC)

Figure 4.16. Relative stabilities of hydroxy- and fluor-potassiumrichterite at high pressures (after Foley 1991).

the vein-derived melt is hybridized by melt derived from the low-melting components of
the wall rocks. Figure 4 .17 graphically illustrates Foley's (1992b) vein-wall rock melting
process at various stages of melting. An important conclusion of this model is that, at
advanced stages of melting, only fluor-phlogopite persists from the original vein assem-
blage, although olivine and orthopyroxene remain in the wall rock. Hence, melting
experiments on a rock derived from a melt produced at this stage of the melting process
would find only mica, olivine, and orthopyroxene as near-liquidus phases. Hence, the
experimentalist might incorrectly conclude that the melt was derived by the partial
melting of a homogeneous mica harzburgite. Foley (1992b) stresses that in this example
neither the vein nor the wall rock consist of mica harzburgite.
Foley's (1992b) model also postulates that melt infiltrates the wall rock and begins
to dissolve minerals such as olivine and orthopyroxene. Assimilation of these minerals
will add a refractory wall rock component to the melt, which is compositionally very
different from that derived from solid solution melting of pyroxenes and garnets. The
extent of assimilation is very difficult to predict, and while Foley (1992b) does not discuss
the thermodynamics of assimilation, he does note that the incorporation of refractory
components and dissolution in vein-melt may commence at temperatures below the
peridotite solidus. It is suggested here that small volumes of melt will be unable to
assimilate much olivine or orthopyroxene without causing complete crystallization of the
vein-melt. However, minerals separated from the wall rock by infiltration and porous
flow of vein-melt may be mechanically incorporated into the melt during segregation
and be transported.
In summary, the compositions of alkaline melts produced by vein-wall rock melting
are determined by the relative proportions of:

• Melts derived from the vein assemblage (V-component)


• Melts derived from the low-melting components of the wall rocks (W-component)
VEINED GARNET LHERZOLITE: LOW - DEGREE MELTING :
UNMELTED VEIN ONLY

INTERMEDIATE STAGE OF ADVANCED MELTING : ONLY MICA


MELTING: WALL - ROCK MELTING REMAINS FROM THE VEIN
BEGINS ASSEMBLAGE

..... OLIVINE ~ ORTHOPYROXENE ~ CARBONATE

~ GARNET ~ AMPHIBOLE j:.....


:/?/::.'I APATITE
',.

_MELT PHLOGOPITE ~~;~~


, "
CLINOPYROXENE

D RUTILE, ALKALI- TITANATE


PETROGENESIS OF ORANGEITES AND KIMBERLITES 361

• A component derived from assimilation of refractory wall rock minerals (WX-


component)

The partial melting process results in a series of hybridized melts. Initial melts will
be dominated by the vein component and have high V/(W+WX) ratios, whereas later
melts will have low V/(W+WX) ratios with W>>>WX.
Foley (1992b) correctly emphasizes that the concepts of low degree and high degree
of partial melting are a matter of scale. Thus, a strongly alkaline melt may be produced
primarily from the vein material [very high V/(W +WX) ratios]. The scale of melting with
respect to the vein is obviously very high (>50%), but low when compared to a system
consisting of a few veins (1-5%) in a substrate. This observation reinforces the conclu-
sion, advanced in this work, that concepts developed for the partial melting of homoge-
neous sources, and appropriate for basaltic magmatism, are inapplicable to the genesis of
exotic alkaline rocks.
Foley's (l992b) model is directly applicable to orangeites. Partial melting of veins
of potassium richterite, phlogopite, and apatite (± clinopyroxene) in a harzburgitic or
peridotitic substrate will melt according to the vein-plus-wall rock melting mechanism.
It is suggested here that all orangeite melts have high V/(W+WX) ratios; i.e., they are
high-degree partial melts with respect to the vein systems. Differing modal assemblages,
mineral compositions, and V/(W+WX) ratios may result in melts which differ slightly in
bulk composition. Incorporation of refractory wall rock minerals might account for the
presence of macrocrystal olivine in many orangeites. Differences in their macrocryst
contents could result from the extent to which vein-melts infiltrate wall rocks. However,
other processes, such as flow differentiation and/or gravity fractionation, can also explain
these variations (see 4.5.3).
While vein-plus-wall rock melting is an attractive process for generating orangeites,
it does have some limitations, especially with respect to the generation of diopside
orangeites and sanidine richterite orangeites. This is because the model is not constrained
and is infinitely flexible with regard to the mode of the veins. Thus, diopside orangeites
may be generated either by preferential melting of richterite or veins containing diopside.
Mineralogical evidence (Chapter 2) suggests that sanidine richterite orangeites are
unlikely to be derived from primitive melts. However, melting of veins rich in richterite
and diopside might conceivably give rise to relatively silica-rich magmas which may
differentiate at low pressures to evolved sanidine richterite orangeites.
One further major problem concerns the role of orthopyroxene. Macrocrysts of
orthopyroxene would be expected to be common in orangeites if olivine macrocrysts are
derived from harzburgitic protoliths. Their complete absence requires that orthopy-
roxenes be completely dissolved, the protolith be dunite, or wall rock orthopyroxenes not
be incorporated into the melt. Dissolution of orthopyroxene would generate silica-rich
liquids, but the extent of assimilation would require large volumes of magma and/or
superheat and is considered unlikely. A dunite protolith provides a satisfactory solution
to the problem, but requires that the diamond-bearing rocks sampled by orangeites during
their ascent to the crust, also be garnet dunites (see 4.8).
Further options are that either V/(W+WX) ratios are extremely high (>10,000) (i.e.,
essentially only the vein melts and macrocrysts are not derived from the residue of partial
362 CHAPTER 4

melting) or VIW is high and WX=O (Le., garnet and clinopyroxene in the wall rock melt
with no addition of a refractory component). However, the absence of orthopyroxene still
requires an explanation if macrocrystal olivine is derived from harzburgitic or peridotitic
contaminants during ascent of the magma (see 4.5.3).
In conclusion, orangeite primary magmas of unknown composition are probably
derived by high degrees of, or even complete, melting of richterite-phlogopite-apatite
veins in a dunitic substrate. Note that the modal amounts of these minerals in the veins
are also unknown.
The vein-wall rock melting mechanism of Foley (1992b) also provides an explana-
tion of the significant negative K, Rb, and Sr anomalies found in extended incompatible-
element distribution diagrams for orangeites (3.6.1) in that, because of solid solution
melting reactions, residual apatite and richterite may remain in the mantle sources.

4.5.3. Melt Segregation, Contamination, and Ascent


This stage of orangeite evolution is one of the least understood as there is very little
experimental evidence pertaining to the physical processes involved in magma segrega-
tion and transit through the lithosphere into the crust. Consequently, much of the
discussion which follows is highly speculative.
The formation and segregation of partial melts in the mantle has been reviewed by
McKenzie (1985, 1989), Hunter and McKenzie (1989), McKenzie and Bickle (1988),
Sleep (1988), Watson and Brennan (1987), and Foley (1992b).1t is concluded that melts
may exist in either porous or channeled flow regimes. In the former, melts are intercon-
nected along grain edges and form a three-dimensional network on an intergranular scale.
In the channeled flow regime, melt exists as a discrete phase, filling cracks and veins.
Sleep (1988) and Foley (1992b) note that the transition between the two flow regimes
is not abrupt and depends upon the scale at which a process is considered. Channeled
flow dominates during the solidification of mantle veins. Porous flow may predominate
during the initial stages of remelting, whereas channeled flow may prevail during later
stages of melting and ascent.
It is believed that the lithosphere is dominantly a regime of channeled flow and the
asthenosphere one of porous flow (McKenzie 1989, Foley 1992b). One consequence of
this assumption is that the underside of the channeled flow regime may be intensively
veined by asthenosphere-derived melts. This is because small volumes of melt, emanating
from the asthenospheric porous flow regime, will be unable to progress very far into the
lithosphere before solidifying due to lack of heat (Spera 1984).
One interesting conclusion of McKenzie's (1989) and Hunter and McKenzie's (1989)
analysis of melting is that melts formed by extremely small degrees of partial melting
«2%) may be extracted from the mantle. This observation is considered here to be
relevant to the extraction of kimberlitic melts from asthenospheric sources, but not to
orangeites from lithospheric sources (see 4.6.1).
High degrees (>50%) of melting of mantle veins in a dunitic substrate will give rise
to liquids which ascend by channeled flow through the brittle lithosphere. At their source,
porous flow may lead to infiltration of melt into the wall rocks, followed by disaggrega-
PETROGENESIS OFORANGEITES AND KIMBERLITES 363

tion and incorporation of xenocrystal olivine into the melt. Because of the nature of the
melting process, xenoliths of restite are not extracted from the source regions of the melt.
Melts derived from isotopically and modally distinct individual veins may coalesce
into larger batches of magma capable of transporting xenoliths of eclogite, diamond-bearing
eclogite, and diamond-bearing garnet dunite (or harzburgitic) rocks. Disaggregation of
the latter will add, in addition to subcalcic chrome pyrope and diamond, a further suite
of olivine xenocrysts to the melt. Thus, the macrocrystal olivine suite may consist of
olivine derived from at least two sources. The reasons for the characteristic, apparently
near-total, disaggregation of diamond-bearing ultramafic xenoliths in both orangeites and
kimberlites remains unknown. The presence in orangeites of diamonds containing eclogite-
suite inclusions clearly demonstrates that some eclogites are also disaggregated; however,
the mechanism of disintegration and the fate of the silicate components in the orangeite
magma are unknown.
The paucity of granular garnet peridotites and lower crustal xenoliths in orangeites
suggests that there is no significant interaction between the magma and wall rocks of the
conduit in the upper lithosphere and lower crust. This may reflect differences in the
rheology and ascent rates of the magma in the upper lithosphere relative to those
prevailing at the source. Unfortunately, we have no knowledge of these parameters. It is
unlikely that upper lithosphere material was sampled and subsequently fractionated from
the magma, as such a process should also remove diamond and xenocrystal pyrope. On
the basis of the foregoing it is provisionally concluded that all types of xenolithic material
found in orangeites are derived from locations close to the lithosphere-asthenosphere
boundary and in the immediate vicinity of the sources of the magma.
Orangeite magmas will lose heat during ascent and begin to crystallize. We do not
know any of the actual liquidus phases, but may surmise that the first and most important
is phlogopite. Evidence for this conclusion is found in the presence of macrocrystal mica
and overall high modal contents of orangeites. Olivine is unlikely to be an important
liquidus phase at high-to-moderate pressures because bulk compositions of the liquids
probably fall outside the primary phase volumes of olivine in ultrapotassic systems. At
lower pressures, the reaction of olivine with liquid to form phlogopite suggests that
olivine is more likely to be assimilated than to crystallize (see below).
Finally, nothing is known of the xenolithlxenocryst content, extent of crystallization,
viscosity, volatile content, or oxygen fugacity of the magma during this stage of its
evolution. Without knowing these physical parameters it is impossible to predict ascent
rates and how the magma might interact with its surroundings.
The most important geochemical aspect of this stage of evolution is contamination
of the primary magma by the addition of the olivine xenocrysts which compose the
macrocryst suite. It is envisioned that, during ascent, the orangeite magma is a slurry
consisting of xenocrysts and phlogopite macrocrysts-phenocrysts suspended in a vola-
tile-rich ultrapotassic fluid. The xenocrystlphenocryst ratio will undoubtedly vary con-
siderably during ascent due to flow differentiation, elutriation, and increasing amounts
of phlogopite crystallization in the upper lithosphere and lower crust.
Because the magma is a slurry, it is unlikely that ascent rates will be extraordinarily
high. Skinner (1989) suggests ascent times of several hours to a few days. Emplacement
of orangeites as dike swarms indicates that these magmas are rheologically very similar
364 CHAPTER 4

to other ultrabasic magmas and lamprophyric magmas. Magma is thus expected to


advance through the mantle by crack propagation (Anderson 1979) and to seek out zones
of weakness and/or preexisting fracture systems in the lower and upper crust. Magma
may flow considerable distances laterally at crustal levels in dike systems. Orangeites
(and kimberlites) certainly do not ascend directly from their mantle sources in cylindrical
conduits as fluidized high-velocity intrusions, as suggested by Wyllie (1980) and
McGetchin (1968).

4.5.4. Low-Pressure and Post-emplacement Crystallization


At the current level of erosion the majority of orangeites occur as hypabyssal dike
swarms and "blows," the latter being the lower parts of the root zones. Diatremes and the
upper levels of root zones are preserved only in the Finsch-Postmasburg area. The style
of intrusion is identical to that observed for kimberlites, and diatremes are believed to
form above precursor dikes and root zones (see 1.9.1). Thus, it is highly probable that
diatremes and craters were originally present above the dike swarms.
In areas which have undergone similar amounts of postemplacement erosion, e.g.,
the Barkly West-Kimberley-Boshof region, there are major differences in the facies of
geographically associated kimberlites and orangeites exposed at the current level of
erosion. Kimberlites are predominantly diatreme and upper root-zone facies, whereas
orangeites are hypabyssal dikes and blows. This observation implies that, for these two
magma types, there are significant differences in the depths at which diatremes originate.
If diatremes are phreatomagmatic, this may indicate significant changes in the hydrology
of the region between the Lower and Upper Cretaceous. Alternatively, the depth of
diatreme initiation may be related to the level at which devolatization of the magma
occurs. Mitchell and Bergman (1991) have noted that C02 is much less soluble than H20
in silicate melts, therefore, exsolution of a C02-rich fluid phase is expected to occur at
relatively greater depths than for a H20-dominated melt. Thus, it is suggested that dikes
of orangeite magma penetrate to much higher levels in the crust than kimberlite dikes
before volatile exsolution occurs, as a consequence of the higher H20 content of the
former.
Petrographic studies of orangeites indicate that the majority are emplaced as slurries
of crystal-rich magmas. Typically, they consist of olivine macrocrysts and closely-packed
aggregates of phlogopite set in groundmass which represents the minor former fluid
component. The complex zoning and inter-crystal compositional variation found in
macrocrystal and phenocrystal mica popUlations demonstrates that most of these micas
have not crystallized in situ. The simplest interpretation of these observations is that the
micas represent the products of crystallization of several batches of orangeite magma of
broadly similar composition. Incorporation of crystals from one batch of magma into
another will result in the development of epitaxial mantles, which represent the compo-
sition of the current liquidus phlogopite in the hybrid magma. Concentration of crystals
from different batches of magma at slightly different stages of crystallization, together
with batch mixing and hybridization ofthe magmas results, in the observed heterogeneous
mica popUlation.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 365

Support for this process is found in the observation that many orangeite dikes are
composite (see 1.8), each phase being composed of modally different assemblages of
similar macrocrysts and phenocrysts. Such composite dikes are particularly common at
Star and Swartruggens. It is probable that dike systems remained as open conduits in
which mixing of many batches of partially-crystallized magma has led to a multiplicity
of hybrid magmas of broadly similar composition. Flow differentiation may result in
further modal variations as mica and olivine are concentrated in different portions of the
dikes. It is unlikely that flow rates will be constant, and periods of stagnation may permit
the flotation of mica and the settling of olivine. The resulting mass of mica may be
subsequently swept into another part of the dike system by a subsequent pulse of incoming
magma, where it crystallizes as an olivine-free orangeite. From the above discussion it is
concluded that the petrographic differences between orangeites and macrocrystal
orangeites result entirely from flow differentiation processes.
Although some orangeites (e.g., Roberts Victor, Star) contain minor amounts of
apparently primary olivine, it is considered in this work that olivine is not a significant
liquidus phase during the later stages of ascent and post-emplacement crystallization of
orangeites (2.3.1, 4.5.3). This is in marked contrast to kimberlites, in which primary
olivine is a major component of the groundmass.
While Skinner (1989) has described mantled and zoned olivines from orangeites in
the Boshof area, these do not appear to be characteristic of all orangeite fields. The
mantling and zoning demonstrate that these olivines are not in equilibrium with their host
magmas. This olivine population probably formed prior to emplacement, the mantles
representing overgrowths upon xenocrystal substrates. Precipitation of olivine may be a
consequence of the bulk composition of these particular hybrid melts being driven into
the primary phase volume of olivine for this system. This may result from the assimilation
of xenocrystal olivine. Alternatively, the low-pressure incongruent dissolution of or-
thopyroxene may cause olivine precipitation. However, given the absence of orthopy-
roxene xenocrysts, this latter process is considered unlikely.
The restriction of diopside-bearing orangeites to particular fields suggests that the
compositional characteristics of the melt leading to diopside precipitation are imposed at
the source of the magma. The presence in some orangeites of resorbed diopside crystals
indicates instability in the transporting melt, and such diopsides must have formed prior
to the emplacement of their current hosts.
Subsequent to intrusion of the slurry, the melt fraction crystallized phlogopite as
microphenocrysts, phlogopite-tetraferriphlogopite solid solutions as mantles on all pre-
existing micas, and tetraferriphlogopite as a ground mass phase. Apatite, spinel, titanates,
perovskite, and other accessory minerals crystallized contemporaneously, and the residual
dregs of liquid formed carbonates and serpentine in the mesostasis. Intra- and inter-intrusion
mineralogical variations reflect minor differences in the bulk composition of the melt and
intensive parameters during cooling.
Evolved orangeites have no simple relationship to orangeites, although the absence
of diamond, paucity in xenocrystal olivine, assemblage of primary minerals, and their
evolutionary trends of composition suggest they are differentiates of less-evolved
orangeite magmas. Unfortunately, where and how differentiation occurred cannot be
determined as there is no geological evidence for the existence of magma chambers. An
366 CHAPfER4

equally unsatisfactory alternative is to consider their sources as being mineralogically


different from those of orangeites, and silica enrichment is not the result of differentiation.
In this scenario, evolved orangeites are melts whose silica-rich, C02-poor character is
imposed at the source, perhaps by partial melting of richterite--diopside--apatite veins
rather than phlogopite-apatite--carbonate veins. The answers to these problems will only
be resolved by further study.
As a final point note that olivine xenocrysts are unstable in evolved orangeite
magmas, as demonstrated by the presence of "dog's tooth" resorption morphologies and
reaction rims of phlogopite around xenocrysts. Such features are not observed in
orangeites, although olivine macrocrysts are commonly rounded and embayed in a
manner suggestive of resorption. This might have occurred during transportation in the
upper mantle. The differences in olivine morphology may be related to cooling rates,
implying that crystallization of orangeite magma is too rapid to permit extensive reaction
of olivine with the melt to occur. The formation of phlogopite reaction rims around olivine
is predicted by Luth's (1967) study of the system kalsiite--forsterite-silica-H20 and is
also observed in some lamproites (Carmichael 1967a, Mitchell and Bergman 1991).

4.5.5. Summary
Orangeites might be derived from phlogopite--potassium richterite--apatite--carbonate,
incompatible element-rich veins within a dunitic substrate located at or above the
lithosphere-asthenosphere boundary. The veins persist, in isolation from convecting
lithosphere from about 1 Ga until the late Jurassic and early Cretaceous. At this time
partial melting is initiated by undetermined processes. Ultrapotassic melts are produced
by a vein-plus-wall rock melting mechanism and are extensively contaminated by
mantle-derived olivine xenocrysts. The contaminated magma ascends to the crust as a
slurry. Crystallization of phenocrystal phlogopite occurs in the lower crust. Magma
mixing, flow differentiation, and low-pressure crystallization of abundant microphe-
nocrystal phlogopite gives rise to a spectrum of hybrid crystal-rich melts which are
emplaced as a system of dikes. Diatremes form above these dikes, at high levels in the
crust, by undetermined processes. Evolved sanidine richterite orangeites may be either
differentiates of orangeites or derived from mineralogically different sources.
Thus, outstanding problems remain with regard to the origin of the veins .and nature
of the processes which initiate the partial melting events leading to the generation of
orangeite magmas.

4.6. PETROGENESIS OF THE KIMBERLITE CLAN


Mitchell (1986) reviewed earlier hypotheses for the genesis of kimberlites and
concluded that none was entirely satisfactory in explaining all features of the geochem-
istry, mineralogy, and petrology of the clan. Further, it is considered that many of the
recent petrogenetic hypotheses, summarized in Sections 4.3 and 4.4 of this work, have
serious flaws. Consequently, it is concluded that, although some advances have been
made, many years of study have not resolved outstanding problems, such as the nature
of the source, depth of melting, genesis of the megacryst suite, and origins of the
PETROGENESIS OF O~ANGEITES AND KIMBERLITES 367

characteristic hybridization. In common with orangeites, a particular impediment to


understanding kimberlite genesis and evolution is our complete ignorance of the compo-
sition of primitive and derivative kimberlite magmas. Although aspects of the above
problems are considered below, any attempts to devise a comprehensive model of
kimberlite genesis are premature.
The low-pressure evolution of kimberlite magmas (i.e., post-emplacement crystal-
lization and diatreme formation) is not considered in detail in this work as few major
advances have been made in this area during the past decade (see 4.6.4). Of these, the
recognition of phlogopite-kinoshitalite solid solutions as a common groundmass mineral
(2.1.9.2) is perhaps the most significant. For reviews of aspects of this topic the reader
should consult Mitchell (1986) and Scott Smith (1992). Recent specific descriptions of
the mineralogy and petrology of a wide variety of kimberlites may be found in the
proceedings of the Fourth and Fifth International Kimberlite Conferences published as
Ross et al. (1989) and Meyer and Leonardos (1994), respectively.

4.6.1. Nature of the Source and Depth of Melting


For many years a major unresolved problem of kimberlite petrogenesis has been the
question of whether the parental magma is derived from lithospheric or asthenospheric
sources. Recent experimental studies (4.3) and geodynamic models (4.4) of kimberlite
genesis indicate that kimberlites might be produced by the partial melting of carbonated
garnet lherzolite or carbonated garnetite sources in the lower parts of the asthenospheric
mantle or the transition zone. Although there is no agreement as to the nature or depth of
melting of these sources, it is agreed that kimberlites are unlikely to be lithospheric
magmas. This latter view, held by Skinner (1989), Foley (1988), and Wyllie (1980,
1989a,b) and extensively promoted by Bailey (1980, 1983, 1992, 1993), is based
primarily upon the geographic restriction ofkimberlites to continental lithospheric plates.
Particularly strong evidence against a lithospheric source is the observation that
isotopic signatures of archetypal kimberlites (Smith 1983) are similar to those of a wide
variety of oceanic magmas (Zindler and Hart 1986), undoubtedly derived from astheno-
spheric sources. Kimberlites exhibiting an asthenospheric isotopic signature can only be
derived from lithospheric sources by petrologically implausible processes. For example,
these would require crystallization of an asthenospheric melt in the lithosphere, followed
by the partial melting of that material within a very short time to ensure that enriched
isotopic signatures are not developed. This may indeed be possible, but the initial part of
this process is identical to that suggested for the origin of veined lithospheric mantle, and
these sources cannot simultaneously give rise to kimberlites, orangeites, and other
ultrapotassic magmas. Of course, virtually untestable, complex models involving zoning
of metasomes (Haggerty 1989a,b, Thibault et at. 1992) may be considered as a means of
overcoming this paradox.
A simpler answer is to place the source of all kimberlites in the asthenosphere and
consider lithospheric metasomatism as a consequence of the interaction of kimberlites
(and other magmas) with the lithosphere rather than being a prerequisite for their genesis.
Metasomatism of the lithosphere by kimberlites is supported by the contemporaneity of
368 CHAPTER 4

the MARID and phlogopite-richterite peridotite xenolith suites and Upper Cretaceous
South African kimberlites (Kinny and Dawson 1992, Erlank et al. 1987).
An asthenospheric source is further supported by the world-wide similarity in the
mineralogy and geochemistry of kimberlites, which indicates that their parental magmas
are produced by the same process occurring repeatedly in space and time. This observation
applies to other asthenospheric magmas (MORB, tholeiitic basalt), but is in marked
contrast to the unique character of the sources of orangeites and some other ultrapotassic
magmas.
Accordingly, it is suggested that kimberlite magmas are generated by partial melting
of asthenospheric material; however, determination of the source mineralogy and resolu-
tion of the questions as to where and how the magmas are produced is more problematic.
One constraint is that the source must contain REE, Sr, K, Rb, Nb, and Ba, but never be
isolated long enough to develop unusual isotopic signatures. This implies that the bulk
composition of the sources has long-term integrated low Rb/Sr and high SmlNd ratios.
Alkali, alkaline earth, and second-period transition elements are thus concentrated in
kimberlitic partial melts derived from protoliths whose bulk compositions are relatively
poor in these elements.
Kimberlites are not especially enriched in potassium «2 wt% K20) compared to
lamproites, orangeites, or minettes (2-12 wt% K20; Mitchell 1986, Mitchell and
Bergman 1991), but the presence in them of abundant carbonate, perovskite, and Ti-rich
spinels, together with phlogopite macrocrysts and groundmass phlogopite-kinoshitalite
requires that their sources contain C02, Ti, K, and Ba. Titanian phlogopite and magnesite
are usually considered to host these elements, and the source is commonly believed to be
a magnesite phlogopite garnet lherzolite (Mitchell 1986, Brey et al. 1983).
Whether this conclusion is correct or not remains problematical. The question of the
presence of carbonate in the mantle has been discussed in Section 4.3.1.4, and it is
concluded that magnesite may be present in carbonated peridotites. although the ultimate
origin of the C02 is ambiguous. Phlogopite is commonly believed to be of "metasomatic"
(sensu lata) origin. Whether it forms by such processes in the asthenosphere is unknown,
but there is no a priori reason why metasomatism should be purely a lithospheric process.
Consequently. portions of the asthenosphere may be enriched in alkali and alkali earth
elements by melt migration and dehydration reactions.
At deep levels in the mantle, phlogopite is unlikely to be stable and K and Ti may be
hosted by richterite (Luth et al. 1993). An alternative is to sequester K in clinopyroxene,
as Edgar and Vukadinovic (1993) and Harlow and Veblen (1991) have shown that this
mineral may contain significant amounts of K20 (> 1 wt %) at pressures above 6 OPa.
This may provide a suitable source of potassium for relatively K-poor, C02-rich magmas
such as kimberlites if magnesite and K-bearing clinopyroxene are consumed together
during the initial stages of melting. At these depths Ba may be present in solid solution
with carbonates, or as complex K-Ba titanates (Foley et al. 1994), rather than silicates.
The above comments indicate that the mineralogy of the mantle which might act as
a source for kimberlites is likely to vary with depth. In the upper asthenosphere, the source
might be magnesite phlogopite garnet lherzolite, whereas in the lower asthenosphere the
same bulk composition might exist as a Ba-Sr-Ca-Mg-carbonate-bearing, K-Ti-richterite
peridotite, or a carbonate-, K-Ba-titanate-bearing peridotite containing potassic diopside.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 369

In the transition zone, potassium will probably be held in majoritic garnets and the source
assemblage would be K-bearing majorite, y-Mg2Si04, and ilmenite or hollandite-struc-
tured titanates. In an oxidized mantle, magnesite would continue to host C02 at all depths
as it is stable to 40 GPa and 2000°C (Katsura and Ito 1990). Limited partial melting of
any of these sources could generate melts having the compositional characteristics
postulated for primitive kimberlite magmas, i.e., C02-, K-, Ti-rich alkaline ultrabasic
melts.
Assuming, for the purposes of this discussion, that melting of the source takes place
at or above the 650-km discontinuity, i.e., the "uppermantle" of Anderson et al. (1992),
then the mantle source may be juvenile primitive material upwelling from the lower
mantle, metasomatized enriched "uppermantle," or a mixture of both. Formation of the
enriched "uppermantle" could involve recycling of subducted material (Ringwood 1989,
Ringwood et al. 1992, Anderson et al. 1992) or the circulation and crystallization of melts
derived from the lower mantle and/or the "uppermantle." Unfortunately, as yet it is
impossible to choose between the multitude of possible scenarios for the choice of a
source, as none are constrained by experimental data.
Asthenospheric mantle is commonly considered to undergo decompressional melt-
ing as it rises toward the lithosphere as large plumes (Campbell and Griffiths 1990, Sleep
1990, Richards et al. 1989, White and McKenzie 1989) or diapirs (Green and Guegen
1974). Interestingly, recent tomographic seismic studies of the mantle appear to contradict
the premises of these models. Thus, as an alternative, Anderson et at. (1992) have
proposed the existence of deep (200-400 km) irregular regions of hot, perhaps partially
molten mantle, termed "hot cells" or "hot regions." Discussion of the relative merits of
these contradictory models is beyond the scope of this work. However, an interesting
possibility arising from Anderson et at. 's (1992) hypothesis is that persistent partially
molten hot cells could be the site in which kimberlites and the megacryst suite are
"brewed."
Melting in the asthenosphere is probably confined to the porous flow regime,
although channeled flow in the vicinity of the lithosphere-asthenosphere boundary
cannot be discounted. The models of McKenzie (1989) and Hunter and McKenzie (1989)
indicate that it should be possible to extract very small amounts of carbonate-rich melt
from sources in the porous flow regime, a conclusion clearly applicable to asthenospheric
kimberlite genesis.
In summary, partial models, such as advocated by Canil and Scarfe (1990), Edgar
and Charbonneau (1993), Ringwood et at. (1992), appear to be viable for generating
kimberlites in the asthenospheric mantle or transition zone. The major differences
between the models lie primarily in the depth at which melting is believed to occur. The
homogeneous source model of Tainton and McKenzie (1994) may also be useful if
considered in an asthenospheric setting. Unfortunately, it is not yet possible to decide
which of these models is preferable, and experimental studies of the problem by the
inverse and forward approaches are highly desirable. Determination of the depth of
melting is critical, and any new genetic hypothesis must be based on an increased
understanding of the thermal and compositional character of the "uppermantle."
370 CHAPTER 4

4.6.2. The Megacryst Problem


Mitchell (1986, 1987) and Schulze (1987) have reviewed the characteristics of the
Cr-poor megacrystJmacrocryst suite (1.3) and summarized hypotheses regarding their
presumed cognate or xenocrystal relationship to kimberlites. Regardless of origin, the
typical presence of this suite of minerals in kimberlites cannot be accidental and requires
explanation. One significant conclusion emerging from studies by Mitchell (1986, 1987),
Hunter and Taylor (1984), Davies etal. (1991), and Hops etal. (1992) is that themegacryst
suite is itself a mixed assemblage, as significant intra- and inter-intrusion compositional
and textural differences exist between megacrystal diopside, garnet, and ilmenite.
Megacrysts within a given kimberlite appear to represent cumulates derived from the
crystallization of several batches of their parental magma at high pressures. Disaggrega-
tion of these cumulates, and the random mixing, of the megacryst minerals within the
magma which ultimately transports them, produces the hybrid heterogeneous assemblage
of megacrysts found in kimberlites. Thus, megacrysts have no simple relationship to their
current hosts. This observation is very important if megacrysts have a genetic relationship
to kimberlites, as cumulates may crystallize from one batch of kimberlite magma and be
transported by a second. Strictly speaking, the megacrysts should be considered as
xenocrysts in the latter; unfortunately this designation obscures their genetic relationship
to kimberlites. This observation suggests that much of the discussion concerning the
cognate or xenocrystal origins of megacrysts is semantic and depends on whether they
are regarded as being remotely (Hops et al. 1992, Jones 1987) or closely related to
kimberlite magmas (Mitchell 1986).
Currently, megacrysts are believed to crystallize from a fractionating magma at
depths of 150-200 km (Nixon and Boyd 1973, Eggler et al. 1979, Gurney et al. 1979b,
Harte and Gurney 1981, Mitchell 1986, Schulze 1987, Hops et al. 1992), i.e., at or just
below the lithosphere-asthenosphere boundary. The crux of the megacryst problem lies
in the identification of this magma. There is a dichotomy of opinion between those who
consider it to be kimberlite or protokimberlite (Hunter and Taylor 1984, Mitchell 1987,
1986, Canil and Scarfe 1990) and those suggesting it is basaltic (Harte 1983, Jones 1987,
Davies et al. 1991, Hops et al. 1992).
Recent studies have not resolved the megacryst problem (Davies et al. 1991, Hops
et al. 1992) as interpretations of the isotopic and geochemical data remain subjective. For
example, isotopic disequilibrium between megacrysts and host kimberlites does not
necessarily imply they are unrelated, given that the host is a hybrid-contaminated rock
and certainly does not represent the original magma from which the megacrysts might
have crystallized.
A xenocrystal origin for megacrysts has been suggested on the basis of calculation
of the trace element content of their parental magma from that of the megacrysts (Kramers
et al. 1981, Jones 1987). However, such hypotheses rest upon the dubious assumption
that crystal-liquid distribution coefficients derived for basaltic systems at low pressures
are relevant at high pressures to kimberlitic magmas.
Determination of their age is perhaps the only way in which the megacryst problem
might be resolved. Hops et al. (1992) and Jones (1987) have determined that the Sm-Nd
ages for garnet and diopside megacrysts from Jagersfontein and Premier are similar to
PETROGENESIS OF ORANGEITES AND KIMBERLITES 371

those of the host kimberlite. However, these ages might be unreliable given the possibili-
ties of interaction of the megacrysts with the light REE-rich host and the significant errors
associated with two-point isochrons. The former problem may be minimized by determi-
nation of megacryst ages by Lu-Hf geochronology using coexisting garnet and ilmenite.
The latter problem will persist, given the apparent paucity of three-phase megacrysts.
Origin of the megacryst-forming magma from asthenospheric sources is now con-
sidered to be supported by the similarity of the Sr, Nd, and Pb isotopic compositions of
the megacrysts to those of oceanic island basalts (Kramers et al. 1981, Jones 1987, Smith
et al. 1987, Davies et al. 1991, Hops et ai. 1992).
Accordingly, Hops et al. (1992) and Jones (1987) propose that both kimberlites and
megacrysts are derived from the same plume-related magmas. In these models a rising
plume undergoes decompressional melting to produce alkali basaltic melts which are
trapped as pools of magma at the base of the lithosphere. Crystallization of these pools,
which exist in various stages of differentiation, gives rise to the Cr-poor megacryst suite.
Kimberlite magmas are believed by Hops et ai. (1992) and Jones (1987) to be too
magnesian and enriched in incompatible elements to be formed by extended fractionation
of these magmas. Consequently, kimberlitic melts are considered to form when the
residual magma infiltrates and equilibrates with the wall rock peridotites and assimilates
other melts (unspecified) generated in the thermal boundary layer and at the base of the
lithosphere. These processes generate hybrid kimberlitic melts which, on eruption, disrupt
the megacryst cumulates and entrain them in the ascending magmas. Thus, a genetic
relationship, albeit distant, between kimberlite and megacrysts is implied, although the
megacrysts are correctly considered to be xenocrysts in their transporting kimberlites.
These hypotheses are attractive in explaining the heterogeneity of the macrocryst
suites and the common derivation of these and kimberlites in an asthenospheric setting.
Crystallization of megacrysts as high-pressure minerals from an "alkali basaltic magma"
is in accord with the well-known occurrence of similar megacrysts in basanites (Green
and Sobolov 1975, Parfenoff 1982, Le Blanc et al. 1982).
The most serious problem of the Hops et al. (1992) and Jones (1987) models is the
assumption that kimberlites are generated by unspecified wall rock equilibration and
assimilation processes. This hypothesis denies the existence of any uncontaminated
primitive kimberlite magma. It is extremely unlikely that the suggested equilibration-as-
similation process will be everywhere identical, and thus it cannot explain the worldwide
petrological similarity of kimberlites. Hops et al. (1992) require wall rock equilibration
to explain the high MgO contents ofkimberlites. However, the MgO content of primitive
kimberlite magma is unknown. The observed high MgO contents of kimberlite rocks
results not from an intrinsic character of the parent melt but from contamination of this
magma with mantle-derived xenocrystal olivine (see 4.6.3). The relationship of kimber-
lites and the megacryst suite to asthenospheric basaltic magmas is considered further in
Section 4.6.4.

4.6.3. Contamination of Kimberlites in the Mantle


Although Section 2.3 concludes that kimberlites and macrocrystal orangeites contain
similar suites of macrocrystal olivine, a significant difference is that kimberlites appear
372 CHAPTER 4

to contain a greater proportion of macrocrysts which can be identified, on textural


evidence, as phenocrysts. This observation is not surprising as kimberlites, unlike
orangeites, crystallize olivine as an abundant primary mineral at low pressures. In
addition, experimental studies (4.3.1.2, 4.3.1.3) indicate that olivine is undoubtedly a
primary liquidus phase during passage of the magma through the upper lithosphere and
may even be a primary phase in protokimberlite magmas in the asthenosphere.
It is considered in this work (1.3, 2.3, 3.7) that kimberlites are hybrid magmas whose
high Ni and MgO contents result primarily from the addition of copious quantities of
xenocrystal olivine to an incompatible element-rich mantle-derived magma. The problem
of the identification of a source for these xenocrysts is identical to that described above
with respect to orangeites (4.5.3).
Textural and compositional studies (2.3) indicate that many of the xenocrysts are
derived from disaggregated ultramafic rocks. The sources of the xenoliths are placed in
the upper mantle, and disaggregation is believed to occur during transport in ascending
kimberlite. The mechanics of the processes which result in disaggregation are unknown.
Some kimberlites contain xenocrystal diamond and subca1cic chrome pyrope; therefore,
in these, some of the macrocrysts must be derived from diamond-bearing garnet dunite.
There is no difference between the xenocrystal olivine suites present in kimberlites that
have passed through diamond-bearing deep cratonic roots, and those emplaced in ancient
mobile zones, i.e., "on-craton" and "off-craton" kimberlites. The latter lack subcalcic
chrome pyrope and diamond, and the kimberlites could not have encountered diamond-
bearing horizons during their ascent (see 4.8). However, passage through chrome pyrope-
bearing dunites is still required to explain the presence of olivine and apparent absence
of orthopyroxene macrocrysts. Appealing to derivation of the olivine from lherzolite
followed by the complete assimilation of orthopyroxene is unsatisfactory, as noted in
Section 4.5.3. Thus, it is considered that the common granular and sheared peridotite
xenoliths found in kimberlites are not the major source of the xenocryst suite. Clearly,
kimberlites sample such peridotites during their ascent, and the preferential occurrence
of these xenoliths in kimberlites, relative to orangeites, must be due to a greater extent of
magma-conduit wall interaction. This may be related to the higher volatile contents,
differing viscosities and ascent rates ofkimberlites relative to orangeites (and lamproites).
Unfortunately, virtually nothing is known of the values of these physicochemical parame-
ters, and further speculation on this topic is not justified.
In summary, it is suggested that asthenospheric kimberlites are contaminated, pri-
marily near the base of the lithosphere, by garnet dunites (or extremely-orthopyroxene-
poor harzburgites), some of which contain diamond. For unknown reasons deep
lithospheric dunites appear to be particularly prone to disaggregation, relative to more
competent granular garnet harzburgites and lherzolites encountered in the upper litho-
sphere. Thus, xenocrystal olivine is believed to be added primarily to kimberlite (and
orangeite) magmas in the initial stages of their passage through the lithosphere, at or above
the lithosphere-asthenosphere boundary. Olivines may be derived from many different
dunitic protoliths, but the limited compositional variation exhibited by mantle-derived
olivine precludes identification of their relative proportions.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 373

4.6.4. Post-emplacement Crystallization


During magma ascent olivine and minor phlogopite crystallize as primary liquidus
phases, and the hybrid assemblage of xenocrysts and phenocrysts is emplaced in the upper
crust as a series of dikes (Clement 1982, Mitchell 1986). During post-emplacement
crystallization all early-formed olivines attempt to equilibrate with their current host
magma by reacting to develop mantles of olivine, whose composition corresponds to that
of the final groundmass olivines (Mitchell 1986). Phenocrystal micas may be concen-
trated by flow differentiation during emplacement and it is this process which results in
the formation of the archetypal kimberlites of "lamprophyric" appearance which are
superficially similar to some orangeites.
Post-emplacement crystallization in the hypabyssal environment results in the crys-
tallization of rutile, spinel, and perovskite followed by monticellite, apatite, and
phlogopite-kinoshitalite as primary ground mass minerals (this work, Mitchell 1986).
The mesostasis, in which these crystals are set, develops by the crystallization of calcite
followed by primary serpophitic serpentine (Mitchell and Putnis 1988). These minerals
may form a uniform groundmass or discrete segregations, depending on the local cooling
rate (Mitchell 1986). Reaction of late-stage mesostasis-forming fluids with the preexist-
ing mineral assemblage may lead to prograde carbonatization and serpentinization (Jago
and Mitchell 1985).
Diatremes and volcanic craters are developed above the hypabyssal infrastructure of
precursor dikes. Reviews of the diverse hypotheses proposed to explain the origin of
diatremes are given by Clement (1982), Clement and Reid (1989), and Mitchell (1986).
These hypotheses are not discussed further in this work as no major advances in this field
have been made since publication of these reviews.

4.6.5. Summary
Kimberlites and their associated megacrysts are derived from asthenospheric sources
whose composition is probably close to that of a REE-, Ti-, K-, Ba-bearing carbonated
ultrabasic rock. The mineral assemblage represented by this bulk composition will vary
with depth. Until the depth at which melting occurs is known the actual mineralogy of
the kimberlite source cannot be determined. Regardless of the modal mineralogy of the
source, small degrees of partial melting will lead to the formation of incompatible
element-rich, C02-bearing, undersaturated, ultrabasic liquids. These segregate from their
sources in the porous flow regime and pass into the lithosphere. The megacryst suite may
be incorporated in the ascending magma either in the asthenosphere or at the lithosphere-
asthenosphere boundary (see below). During passage through the lithospheric mantle, the
megacryst-bearing melts are extensively contaminated by xenocrystal olivines derived
from the disaggregation of garnet dunites. Olivine also crystallizes as a primary phase at
low pressures in the upper lithosphere. The final hybrid assemblage of phenocrysts,
megacrysts, and xenocrysts is emplaced as a crystal-rich magma in a series of dikes in
the upper crust.
The above summary of the processes involved in magma formation and evolution
cannot be regarded as the definitive model of kimberlite genesis, as there remain many
outstanding problems with regard to the nature of the source, depth of melting, and origin
374 CHAPTER 4

of the megacryst suite. Some speculations on the latter and on the relationship of
kimberlites to other asthenospheric magmas are considered below.
Currently, the majority of megacrysts are considered to form at the lithosphere-
asthenosphere boundary. The discovery of megacrysts derived from greater depths
(Sautter et at. 1991) and experimental data of Canil and Scarfe (1990) suggest that this
conclusion is not entirely correct. Most current hypotheses also imply that megacrysts
formed in a single event; this assumption is unlikely to be valid within any uniformitari-
anist models of magmatism. In the interest of generating further discussion of the
megacryst problem, the following, highly speculative, scenario is presented as an alter-
native to current models of megacryst genesis.
Assuming that plumes or hot cells exist in the asthenosphere, it is suggested that the
mantle within these may be at temperatures close to the solidus. Consequently, decom-
pressional partial melting may be easily induced by small amounts of upwelling. This
may be initiated by processes occurring in the transition zone or even triggered by the
upwelling of lower mantle material. Partial melting will produce a variety of magmas,
depending upon the depth and degree of melting. These melts may segregate from their
sources, ascend toward the lithosphere, and commence crystallization.
It is further assumed that melts are continually produced over geological time, and
all may crystallize minerals of the megacryst suite. However, because near-solidus initial
melting would preferentially consume carbonate, Ti-bearing phases, clinopyroxene, and
garnet, the megacryst suite may be preferentially produced from initial limited partial
melts of kimberlitic composition. This is because bulk compositions of the melts would
not lie in the primary phase volume of olivine for this system. This constraint does not
preclude some precipitation of the megacryst suite from more advanced partial melts
produced from the same source. However, crystallization of megacrysts from these may
be limited due to the appearance of olivine as the dominant primary liquidus phase as the
bulk composition of the partial melt changes. Thus, in such systems, the initial precipi-
tation of megacrysts and/or increasing amounts of melting may drive liquid compositions
very quickly into the primary phase volume of olivine. It might be expected that the
potential for megacryst formation would decrease as partial melts changed composition
in the sequence kimberlite, melilitite, basanite, as the degree of partial melting increases.
Note, all of these magmas are known from geological evidence or experimental studies
to have the potential to crystallize megacrysts (Green and Sobolov 1975, Nixon and Boyd
1979, Nixon et at. 1980, LeBlanc et at. 1982, Parfenoff 1982, Mitchell 1986).
Melts may crystallize completely in the asthenosphere and/or become trapped at the
lithosphere-asthenosphere boundary, as it is very unlikely that all partial melts produced
in the mantle will be erupted. There may be a preferential concentration of crystallizing
melts at the underside of the lithosphere, where magmas are channeled into topographic
irregularities at the base of the cratonic root. Some melts might enter the lithosphere and
contribute to the veined metasomatic domains which have been proposed to exist at the
base of the lithosphere. The end result of these processes may be pockets of megacrysts,
scattered over a range of depths in the upper asthenosphere, which are derived from a
variety of magmas of different ages. Fragmentation of these cumulates by subsequent
batches of magma followed by entrainment and mixing will give rise to the observed
hybrid assemblages of megacrysts. Note that magmas causing fragmentation of cumulates
PETROGENESIS OF ORANGEITES AND KIMBERLITES 375

may themselves be simultaneously crystallizing megacrysts. This model incorporates


elements of hypotheses advanced by Harte and Gurney (1981), Mitchell (1986), Davies
et al. (1991), and Hops et al. (1992), but differs in that the megacryst assemblage in
kimberlites is considered to originate from several sources of different ages.
The above scenario suggests that kimberlite magmatism is but one facet of the
broader magmatic events related to the upwelling and melting of asthenospheric mantle.
In this context it is suggested that, depending upon the depth and extent of partial melting,
the same source may give rise to a spectrum of partial melts. One possibility is that a
mantle hot cell or rising plume may initially give rise to kimberlites which are followed
by melilitites and other alkaline magmas as the extent of partial melting increases. The
latter magmas may be generated only in the regimes of extensive lithospheric thinning
and rifting resulting from the continued upwelling of the plume or hot cell. An alternative
is that mantle hot cells may be thermally and compositionally asymmetric (laterally and
vertically). Melting in one part of the cell may give rise to kimberlites, and in another to
melilitites. If plumes or hot cells are not particularly buoyant, or if ascent is terminated,
lithospheric thinning may not occur. Partial melts migrating ahead of such plumes or hot
cells may intrude the continental lithosphere without associated rifting and nephelinite-
carbonatite volcanism. Note that different partial melts will not necessarily ascend at the
same rate or segregate and be emplaced at the same time. Such aborted plumes, or hot
cells, might be the cause of the geographic association of kimberlite and melilititic
magmatism found in southern Africa (Namibian kimberlites-Bushmanland melilitites),
the Anabar Shield (Ukukitskoye-Olenyok kimberlites-northeastern Anabar melilitites),
and the Hudson Bay Lowlands of Canada (Attawapiskat kimberlites-Lowlands melili-
tites). None of these igneous provinces are associated with rifts. This hypothesis suggests
that melilitites are not lithospheric melts, and the absence of diamond in the majority of
them is not simply due to genesis at depths above the stability field of diamond (see 4.8).
Kimberlites are not found in oceanic settings as upwelling of asthenospheric mantle
is too rapid and partial melting too extensive to enable the initial limited partial melts to
escape before being overwhelmed by basaltic magmas. Consequently, the components of
kimberlite are simply diluted by and hybridized with the more voluminous basaltic
magmas.

4.7. RELATIONSIDPS OFORANGEITES TO KIMBERLITES,


LAMPROITES, AND OTHER ULTRAPOTASSIC MAGMAS

4.7.1. Kimberlites
The central thesis of this monograph is that orangeites and archetypal kimberlites
represent distinct magma types derived from lithospheric and asthenospheric sources,
respectively. Consequently, they have no genetic relationship. The only possible relation-
ships are indirect and within the broader context of the overall evolution of the li-
thospheric and asthenospheric mantle. Hence, asthenospheric processes, which
eventually result in kimberlite or basaltoid magma generation, may passively affect the
lithospheric sources of orangeites and induce partial melting. This may result from uplift
above upwelling plumes or hot cells and/or release of juvenile volatiles. Of course,
376 CHAPTER 4

asthenospheric magmas might ultimately be the parents of the ancient veined lithospheric
sources of orangeites in the roots of cratons.

4.7.2. Lamproites
Orangeites have closer mineralogical, geochemical, and isotopic similarities to
lamproites than archetypal kimberlites, and, in terms of their C02 and H20 contents,
might be thought of as being intermediate between C02-rich kimberlites and H20-rich
lamproites. In contrast to lamproites, which only rarely contain carbonates, orangeites
characteristically contain primary carbonates. In addition, orangeites are poor in fluorine,
relative to lamproites.
Although orangeites and macrocrystal orangeites contain some accessory minerals
suggestive of a lamproitic affinity, only evolved diopside and sanidine richterite
orangeites consist of a mineral assemblage similar to that of lamproites. There are
significant differences between the clans with respect to the compositional variation and
relative abundances of typomorphic minerals, such as phlogopite, potassium richterite,
potassium feldspar, and hollandite. These differences are discussed in detail in Chapter 2.
Mitchell and Bergman (1991) have concluded that the emplacement of lamproites in
a wide variety of geological environments precludes development of a universal model
of their temporal and tectonic settings. Lamproites commonly occur along the margins
of cratons or in accreted ancient mobile belts in regions of thick crust and lithosphere.
They are not associated with modem active subduction zones, although potassic magmas
formed in this environment share some of the geochemical characteristics of lamproites,
i.e., Sr and Nd isotopic signatures suggestive of derivation from enriched sources, and
negative Ti, Nb, and Ta anomalies. Mitchell and Bergman (1991) suggest that lamproite
tectonic settings and compositional traits indicate they might represent partial melts of
ancient subduction zones.
Orangeites are restricted to the Kaapvaal craton and considered to be derived from
a lithospheric source located near the deepest parts of the cratonic root. It cannot be
unambiguously determined whether subducted oceanic lithosphere played any role in the
development of this source. In this context, extended incompatible-element distribution
patterns for orangeites do not exhibit the characteristic negative Ti-Nb--Ta anomalies
found for lamproitic and other subduction-related magmas.
Currently, lamproite sources are believed to be veined or metasomatized lithosphere
(Fraser et al. 1985, Mitchell and Bergman 1991, Foley I 992a,b ). Lamproite primary
magmas are considered by Mitchell and Bergman (1991) to be peralkaline ultrapotassic
melts of intermediate silica content, which is the reason why they crystallize abundant
potassium feldspar, leucite, and richterite and are mica-poor, relative to orangeites. The
magma compositions must reflect the presence of greater amounts of relatively silica-rich
minerals in their sources, i.e., richterite and diopside. Thus, geochemical and isotopic
studies suggest that the sources of both magma types are broadly similar in mineralogical
character but not mode. Those of orangeite must be richer in phlogopite and apatite and
contain substantial quantities of C02 and be poor in F. Only the parental magmas of
evolved orangeites might be derived from veins whose modal mineralogy overlaps that
of lamproites. Unfortunately, the exact mineralogy of the source cannot be specified. In
PETROGENESIS OF ORANGEITES AND KIMBERLITES 377

summary, the compositions of lamproites suggest that Ti-potassium richterite, diopside,


and K-Ba titanates are dominant in their sources in contrast to the phlogopite-apatite
sources suggested in this work for orangeites. Partial melting of lamproite sources
probably takes place by the vein-plus-wall rock mechanism of Foley (l992b).
In conclusion, because of the mineralogical and geochemical differences noted
above, orangeites are not regarded as members of the lamproite clan. However, their
similarities with lamproites and some other potassic rocks cannot be disregarded, and it
is suggested that all of these magmas represent different expressions of a more general
process of continental potassic magmatism (see 4.7.3).

4.7.3. Other Ultrapotassic Magmas


Lamproites are absent from the Kaapvaal craton, and the nearest occurrence of
lamproitic rocks is at Kapamba in the Luangwa valley of Zambia (Scott Smith et al. 1989).
There is also a notable absence of other types of markedly potassic magmatism in the
Kaapvaal-Zimbabwe craton, in marked contrast to the situation in other cratons. It is
suggested here that the absence of such rocks is directly related to the long-term
metasomatic history of this craton.
In the Wyoming (U.S.A.) craton, lamproites (Leucite Hills, Smoky Butte; Car-
michael 1967a, Mitchell et al. 1987, Mitchell and Bergman 1991) occur along with a
wide variety of potassic rocks belonging to the shonkinite-fergusite-missourite-minette-
suite (Highwood, Crazy, and Bearpaw Mountains; Larson 1940, Hearn 1989, O'Brien et
al. 1991, MacDonald et al. 1992). In the Aldan (Russia) craton, many rocks with
mineralogical affinities to lamproites are found in the Murun ultrapotassic complex
(Shadenkov et al. 1989, Orlova 1988, Vladykin 1985), in addition to a host of geochemi-
cally defined lamproites (Bogatikov et al. 1985, 1991) and many plutonic complexes
consisting of extremely undersaturated potassic rocks belonging to the kalsilite-Ieucite-
biotite-<>rthoclase pyroxenite suite (Kostyuk et at. 1990). In other cratons potassic
magmatism is represented primarily by lamproites-e.g., West Kimberley (Australia)
craton (Jaques et al. 1986), Bhandara-Singhbum (India) craton (Mitchell and Bergman
1991), while the minette, shonkinite, and/or kalsilite pyroxenite suites are absent.
Currently, some form oflithospheric mantle metasomatism (sensu lato) is considered
to play a role in the development of the sources of all of the above potassic rocks
(McCulloch et al. 1983, Vollmer et al. 1984, Mitchell et al. 1987, 1994, Dudas et al. 1987,
O'Brien et al. 1991, Mitchell and Bergman 1991), although opinions differ as to the
relative contribution of asthenospheric or recycled subducted components (Nelson et al.
1986) to these sources during partial melting. The differing expressions of potassic
magmatism in each craton suggest that, although they were all affected by a similar
metasomatic (sensu lato) process, the duration, style, and extent of this process was
different in each craton.
These differences are clearly expressed in the isotopic signatures of the potassic
rocks. Thus, the ancient REE and Ba enrichment of the Wyoming and Aldan cratons
appears not to have been accompanied by Rb enrichment, i.e., the lithosphere had
long-term low Rb/Sr and Sm/Nd ratios (O'Brien et al. 1991, Mitchell and Bergman 1991,
Mitchell et al. 1994). In contrast, similar REE enrichment of the sources of the Spanish
378 CHAPTER 4

and West Kimberley lamproites were associated with Rb enrichment and their sources
reflect long-term high Rb/Sr and low SrnlNd ratios.
Mantle metasomatism (sensu lato) undoubtedly prepares cratons for the generation
of potassic magmas. On a time scale of 1-2 Ga, different modal mineralogies will develop
in veins and metasomatic horizons in different cratons. For example, one craton may be
enriched in mica or potassium richterite relative to another, the differing styles perhaps
reflecting the relative proportions of small-volume melt infiltration to fluid-related
metasomatism. Some cratons may experience long-term persistent metasomatism (sensu
lato), whereas others may undergo relatively little metasomatism. Persistent metasoma-
tism may destroy the diamond-bearing roots of a craton if the melts and fluids involved
are sufficiently oxidizing. Although metasomatism (sensu lato) as a physicochemical
process will always operate in the same manner, its geological effects are likely to be
unpredictable and random. Hence, the accumulated results of metasomatism (sensu lato)
will never be the same everywhere.
It is concluded from the above observations that the metasomatic (sensu lato) history
and evolution of a particular craton must be unique. Partial melting of these diverse, yet
similar, lithospheric metasomatic assemblages will result in each craton exhibiting a
distinctive style of potassic magmatism.
Magmas will be produced in different cratons, which in some instances may be
broadly similar in bulk composition, but significantly different in trace element content
and isotopic signatures, e.g., the lamproites of the Leucite Hills and West Kimberley
provinces. Within a given craton individual lamproite fields may have similar isotopic
signatures, but significantly different mineralogical character, e.g., Francis, Leucite Hills,
Smoky Butte.
Different styles of metasomatism explain why extremely undersaturated potassic
rocks are found in the Aldan shield, but not in the Wyoming craton. Both of these cratons
appear to have experienced persistent extensive metasomatism, judging by the plethora
of potassic alkaline rocks present in them compared with other cratons. Note thatthe 1-35
Ma lamproites of the Wyoming craton do not contain diamond; Devonian Wyoming-
Colorado kimberlites have very low diamond grades; and lamproite-like rocks of the
Aldan Shield apparently lack diamond. These observations suggest that oxidation of any
diamond originally present in the craton roots occurs during extensive and extended
metasomatism.
As one might expect, there exist suites of potassic rocks which have some minera-
logical or geochemical affinities to potassic rocks known from other cratons, but are
significantly different in other respects. The Mata de Corda (Brazil) and Baker Lake
(Canada) suites of ultrapotassic volcanism fall into this category. The Mata de Corda
Formation has lamproitic and kamafugitic mineralogical and geochemical characteristics
and cannot be unambiguously assigned to either ofthese clans (Sgarbi and Valenca 1991,
Ulbrich and Leonardos 1991). Volcanic rocks of the Baker Lake Formation (Peterson
1994) are similarly ambiguous in their petrological character in having mineral assem-
blages typical of minettes, coupled with geochemical affinities to lamproites.
The existence of these apparently "transitional" rocks does not imply that kamafugite
or minette magmas differentiate to lamproites, or vice versa. The occurrences are
interpreted as indicating that the sources of these rocks may be unique, relative to the
PETROGENESIS OFORANGEITES AND KIMBERLITES 379

sources of potassic rocks in other cratons. This conclusion is supported by recent Sr-Nd
isotopic studies of kimberlites and other alkaline rocks from Brazil (Bizzi et at. 1994),
which demonstrate that the Sao Francisco craton and associated accreted mobile belts
have undergone a significantly different evolutionary history to the Kaapvaal craton.
These observations raise an interesting question: does each suite of apparently unique
potassic rocks require a separate clan name? Clearly, if this approach is taken, it could
lead to a proliferation of new magma types. Many petrologists are resistant to the
introduction of new names (including the writer) and are, in the case of alkaline rocks,
actively attempting to rationalize the archaic nomenclature by eliminating otiose unin-
formative names (Scott Smith et at. 1984, Scott Smith 1992, Mitchell and Bergman 1991,
Mitchell 1994c, Woolley et at. 1995). However, the introduction of new names is entirely
legitimate if a clan or suite of rocks is demonstrably genetically different to rocks
belonging to other clans. The recognition ofkomatiites and boninites illustrates the value
of this approach. Modal or contaminated/hybrid variants of existing clans do not warrant
new names. In this context, the Mata de Corda Formation has not yet been sufficiently
characterized with respect to its origin and evolution. It may be that it merely represents
a kamafugitic magma that has been contaminated with lithospheric components rather
than a distinctive magma type which warrants a new name.
Do orangeites require a new name, or are they merely extreme variants of the
lamproite clan? The conclusion of this work, based on mineralogical and geochemical
evidence, presented here and in Chapters 2 and 3, demonstrates that they are not
lamproites and a new name is justified. Evolved orangeites considered in isolation from
the less-evolved antecedents might indeed be described as lamproites, e.g., Tainton and
Browning (1991). Note, however, that evolved amphibole-bearing minettes (Hall 1982,
Nemec 1988) also have similar mineralogical and geochemical characteristics to both
evolved orangeites and lamproites and could be assigned to either clan. Alternatively,
evolved orangeites could be termed "minettes"! These observations do not imply that all
of these rocks are related or the names synonymous, but they do illustrate the problems
that arise in attempting to classify individual rocks in isolation from consanguineous
antecedents and descendents (see 1.4). These apparently similar rocks reflect a minera-
logical convergence resulting from differentiation of genetically diverse saturated-to-
oversaturated peralkaline magmas. The crystallization of coexisting groundmass
richterite-arfvedsonite and potassium feldspar as the major modal phases in all cases is
in response to the similar bulk major element composition of the residua. Mineralogical-
genetic classifications may be used to identify the parental clans of such petrographically
similar rocks (1.4.4).
In the Kaapvaal-Zimbabwe craton, the restriction of potassic magmatism to
orangeites suggests that this craton has not been subjected to the persistent and extensive
metasomatism that has afflicted the Wyoming, Aldan, and Sao Francisco cratons. This
may be a consequence of unusually deep, extremely durable cratonic roots acting as a
near-impermeable barrier to thermal and chemical assaults from asthenospheric magmas.
Similar conclusions might be drawn for the Anabar craton. It cannot be coincidental that
the world's major sources of diamonds originate from the roots of these cratons and not
from those of strongly metasomatized cratons characterized by extensive potassic mag-
matism.
380 CHAPTER 4

4.7.4. Summary
In summary, orangeites, or rocks which might be considered as mineralogical
variants of the clan, do not occur outside the Kaapvaal craton. They are genetically and
mineralogically distinct from lamproites and thus represent a distinct magma type. They
are believed to represent the sole expression of potassic magmatism in the Kaapvaal-
Zi mbabwe craton. As a consequence of their different metasomatic histories other cratons
vary with respect to the style of lamproite clan or plutonic-to-volcanic potassic and
ultrapotassic magamatism.
As a final point, it may be that eventually we will be able to consider all of the diverse
forms of potassic lithospheric magmatism as belonging to a family of magmas whose
origins must be sought in metasomatic modifications of the lithospheric mantle. Follow-
ing a suggestion by Barbara Scott Smith (pers. comm. 1994) these might be collectively
referred to as the Three M-family, i.e., metasomatized mantle magmas!

4.8. PRIMARY DIAMOND DEPOSITS


As a finale to a work concerned with diamond-bearing rocks some comments are
required regarding aspects of primary diamond deposits. "Primary" is used in the sense
that these diamond deposits have not been formed by surficial weathering process.
Strictly, the deposits should be called "pseudoprimary," as, if all diamonds are xenocrysts,
the actual primary source lies at inaccessible depths in the mantle.
Detailed discussion of the genesis of diamond is beyond the scope of this work.
Reviews of this topic may be found in Meyer (1985, 1987), Harris (1987), Gurney (1989),
and Kirkley et al. (1991). Current models of diamond formation differ primarily with
respect to the sources of carbon. One group suggests that carbon is juvenile, and
deposition of it as diamond occurs as methane or other hydrocarbons are oxidized during
their ascent through the upper mantle (Taylor and Green 1989) or at the lithosphere-
asthenosphere boundary (Haggerty 1986). These hypotheses are favored for the genera-
tion of diamonds containing the peridotitic suite of inclusions. Other possibilities for the
origin of this suite of diamonds include crystallization from kimberlitic liquids (Harte et
al. 1980, Arima et al. 1993b) or from ultrabasic melts during the formation of cratonic
roots (this work).
A second group of hypotheses suggests that carbon is introduced into the mantle by
subduction processes (Schulze 1986, Kesson and Ringwood 1989). The carbon is not
juvenile and may be of biogenic origin (Milledge et al. 1983, Nisbet et al. 1994). A
subduction origin is favored for the genesis of isotopically light diamonds containing the
eclogitic suite of inclusions.
Although it is now evident that several diamond-forming processes exist, many
important questions remain unanswered: the origin of mega- and microdiamonds;
whether or not diamonds form in the lower asthenosphere transition zone; and whether
diamond-forming processes are active today. The growth of diamond crystals and the
mechanism of trapping silicate inclusions within them are not understood, as it is
uncertain whether diamonds grow in the solid state as porphyroblasts or from liquids.
PETROGENESIS OF ORANGEITES AND KIMBERLITES 381

Regardless of origin, most diamonds are now believed to be xenocrysts in their


transporting magmas. Current hypotheses postulate that the roots of cratons contain
diamond-bearing horizons consisting of garnet dunite, garnet harzburgite, and eclogite.
Diamonds may also exist just below the lithosphere-asthenosphere boundary, where they
are believed by Haggerty (1986) to be formed by methane oxidation. Other diamonds
may occur in subducted oceanic material underplating the flanks of a craton. The vertical
and lateral disposition of all of these diamond-bearing horizons is completely unknown.
Disruption and disaggregation of diamond-bearing zones by the passage of magmas
ascending from greater depths results in the incorporation of diamonds as xenocrysts in
the magma. The subsequent fate of entrained xenocrysts is dependent upon the oxygen
fugacity and rate of ascent of the magma toward, and through, the crust. Slow transport
in highly oxidized hot magmas, e.g., lamproites, may result in the complete resorption of
all diamond originally present. From the foregoing it is apparent that only magmas
derived from depths below the diamond-bearing zones will contain diamonds.
Given the multiplicity of sources and the myriad of possible ascent paths, it is not
surprising that individual primary diamond deposits differ greatly in character. A given
deposit may contain diamonds derived from several protoliths, as the diamond suite
present depends upon which diamond-bearing horizons are intersected.
This study has suggested that orangeites originate from, or just above, the litho-
sphere-asthenosphere boundary. Thus, there is a high probability of these magmas
interacting with diamond-bearing zones. The diamond suite present in orangeites indi-
cates disruption of both eclogitic and harzburgitic diamond-bearing horizons. Orangeites
and macro crystal orangeites typically contain diamonds, whereas evolved orangeites
appear to be diamond free. If evolved orangeites are low-pressure differentiates of
orangeites, this may be a consequence of diamond resorption or gravitational fractiona-
tion of diamond from the magma. Of course, the parental magmas of evolved orangeites
simply may not have intersected diamond-bearing horizons, although this option seems
improbable as all occurrences of these rocks appear to be diamond free.
It is well known that only those kimberlites which are erupted through the diamond-
bearing roots of cratons contain diamonds, whereas those emplaced through accreted
mobile belts lack diamonds (Clifford 1966). Importantly, there are no mineralogical
differences between diamond-bearing and diamond-free kimberlites (Mitchell 1986).
Similarly, the Cr-poor megacryst suite is identical in both varieties (Mitchell 1987 , Hops
et at. 1992).
Boyd and Gurney (1986) have suggested that all kimberlites are derived from the
lithosphere-asthenosphere boundary, which they consider to intersect the stability field
of diamond only where the base of the craton underlies Archean shields (Figure 4.18). In
mobile belts the lithosphere-asthenosphere boundary is located at lesser depths and
within the stability field of graphite. Consequently, diamond-bearing kimberlites are
confined to cratonic settings and kimberlites erupted through mobile belts are diamond
free. Boyd and Gurney (1986) do not address the question of why most "on-craton"
kimberlites are diamond poor or diamond free.
Boyd and Gurney's (1986) hypothesis is based on a particular interpretation of the
equilibration temperatures and pressures of xenolith suites found in "on-craton" and
"off-craton" kimberlites. They believe that the inflections found in the xenolith-derived
382 CHAPTER 4

NW
•• • SE NW
•• • SE

km • ••
ASTHENOSPHERE ASTHENOSPHERE

CD DIAMOND - BEARING ASTHENOSPHERE m DIAMOND-BEARING GARNET HARZBURGITE

+ BARREN KIMBERLITES • DIAMOND-BEARING KIMBERLITES

(0) (b) +

Figure 4.18. Contrasting models illustrating why diamond-bearing kimberlites are restricted to within the
bounds of the Kaapvaal craton and barren kimberlites are confined to adjacent mobile belts. (A) In the Haggerty
(1986) and Mitchell (1986. 1987) model. kimberlites are derived from similar depths within the asthenosphere
as a result of partial melting of upwelling material. Asthenospheric diamonds are formed by methane
dissociation at the lithosphere-asthenosphere boundary (LAB) in the vicinity of the deepest parts of the craton
root (Haggerty 1986). Lithospheric diamonds occur only within the harzburgitic root of the craton. Only
kimberlites which pass through this region can entrain xenocrystal diamond. (B) In the Boyd and Gurney (1986)
model, kimberlites are derived from different depths at the lithosphere- asthenosphere boundary, the location
of which is defined by the equilibration parameters of garnet lherzolite xenoliths found in kimberlites. In the
mobile belts the boundary is considered to lie within the graphite stability field. Diamonds are believed to be
stable only within the deepest parts of the craton root. In this model all diamonds are of lithospheric origin.

paleogeothenns define the lithosphere-asthenosphere boundary. The maximum pressure


calculated for anyone suite is believed to record the maximum depth of origin of the host
kimberlite.
Contrasting interpretations of the equilibration data (Carswell and Gibb 1987)
suggest that the Boyd and Gurney (1986) model is not plausible. Further, it is suggested
in this work and by others (Canil and Scarfe 1990, Ringwood et al. 1992, Haggerty 1986,
1994) that kimberlites are of asthenospheric origin. If this hypothesis is correct, the Boyd
and Gurney (1986) model has no relevance in explaining the absence of diamonds in
"off-craton" kimberlites.
A simpler model, which makes no assumptions regarding the depth of origin of
xenolith suites, requires that kimberlites are all derived from similar depths in the
asthenosphere (Figure 4. 18). In this model only kimberlites which pass through the craton
roots traverse diamond-bearing horizons (Haggerty 1986, Mitchell 1986, 1987, 1991b).
A refined version of the model presented by Mitchell (1991 b) is presented as Fig. 4.19 to
illustrate how different kimberlites might develop contrasting suites of diamonds.
It is considered that the asthenospheric source model best explains the fonnation of
diamond-bearing and diamond-free kimberlites. Intra- and inter-kimberlite diamond
grades differ because of the heterogeneous distribution of potential sources, sorting of
PETROGENESIS OFORANGEITES AND KIMBERLITES 383
RIFTING CRATON MOBILE BELT

N o

- 50

- 150

DIAMOND - BEARING FACIES


- 250
[""' .... '] GARNET
:;f::l HARZBURGITE

- 350 km
lEI
L.:...J
LITHOSPHERIC
ECLOGITE
t t ~ SUBDUCTED
K K K
~ ECLOGITE
ASTHENOSPHERE IGLtl GA~[;.EJOS~~~~~OLlTE
Figure 4.19. Hypothetical cross section of an Archean craton, ancient accreted mobile belt, and younger
incipient rift, showing the location of the lithosphere-asthenosphere boundary (LAB, relative to the stability
fields of diamond and graphite. The diagram illustrates why different kimberlites (K) differ with respect to
sources of xenocrystal diamond. KJ may contain lithospheric and asthenospheric garnet lherzolite diamonds
together with garnet harzburgiteldunite-derived diamonds. K2 contains diamonds from the aforementioned
sources plus diamonds derived from lithospheric and subducted eclogites, i.e., five distinct sources. K3 contains
only lithospheric and asthenospheric garnet lherzolite-derived diamonds. K4 does not pass through any
diamond-bearing zones and is barren. Orangeites (0) are shown originating at the LAB and contain diamonds
derived from garnet harzburgite/dunites and subducted eclogites. Lamproite (L) contains diamonds derived
only from subducted eclogite and lithospheric garnet lherzolite sources. Melilititic magmas are considered to
be derived by greater degrees of melting from the same asthenospheric sources as kimberlites. Depending upon
the depth of segregation of the magma they may (Mj) or may not (M2) contain diamonds. High degrees of
partial melting at shallow depths in the rift zone produces nephelinite-suite magmas (N) which are always
diamond free (modified from MitchellI99Ib).

diamonds during entrainment, flow and mixing of different batches of kimberlite, and
varying degrees of resorption of diamond in the ascending magma (Mitchell 1991 b).
Lamproite diamond deposits (Argyle, Ellendale, Prairie Creek, Majhgawan) do not
contain subcaIcic chrome pyrope xenocrysts. Thus, derivation of their diamond suites
from highly depleted garnet dunitic-harzburgitic sources is unlikely. Studies of mantle-
derived lherzolites, xenocrysts, and inclusions in diamonds from the Australian lampro-
ites (Jaques 1989) suggest that their lithospheric mantle sources are less refractory than
the roots of the Kaapvaal craton. The inclusion suite indicates a predominantly eclogitic
source, which may represent subducted oceanic lithosphere which was eventually cra-
tonized. This diamond-bearing material now constitutes the deeper parts of the mobile
belt surrounding the Kimberley craton. A similar tectonic setting is evident for the
Mahjgawan lamproite, whereas that of the Prairie Creek lamproite is less well-defined
(Mitchell and Bergman 1991). Diamonds in the latter appear to be derived from both
peridotitic and eclogitic sources (McCandless et ai. 1994).
The presence of diamonds derived from cratonized mobile belts in Australia, India,
and the United States is in marked contrast to the situation in the Kaapvaal craton. Mobile
384 CHAPTER 4

belts accreted to the latter appear to be diamond free, as kimberlites erupted through them
are barren of diamonds. This observation has two implications: eclogite suite diamonds
in Kaapvaal kimberlites and orangeites might be derived entirely from the Archean
eclogites incorporated into the roots of the craton during the fusion of the Kaapvaal and
Zimbabwe cratons (Helmstaedt and Schulze 1989); it confirms that the sources of
orangeites are not in subducted mobile belts and, therefore, are not tectonically or
geochemically equivalent to those of lamproites.
It follows from the above discussion that any asthenosphere-derived magma which
passes through diamond-bearing horizons in the craton root has the potential to carry
diamonds. This conclusion may explain the rare occurrences of diamonds reported from
alkali basaltoid and lamprophyric rocks (Kaminskii 1984, Janse 1994) and the existence
of diamond-bearing melilitites in the Anabar craton (this work). The latter occur on the
northeastern flank of the exposed craton and constitute the Mesozoic West Ukukit,
Luchakanskoye, Dukenskoye, Kuranakhskoye, and Ari-Mastakhskoye "kimberlite"
fields (BrakhfogeI1984, Kovalskii et al. 1969). Examination of many samples from these
fields by the author indicates that they contain diatreme and hypabyssal facies monticellite
melilitites and alnoites rather than kimberlites (this work). Most of the intrusions are not
diamondiferous and those that are have very low grades (V, Kornilova pers. comm.). The
province is petrologically very similar to the Hudson Bay Lowlands melilitite province
(Janse et al. 1989, Reed and Sinclair 1991).
The northeast Anabar melilitites are unique among melilitite provinces in containing
diamonds, a fact that must be related to the unusual stability and very high diamond
content of the Anabar craton. Melilitites erupted through the mobile belts surrounding the
Kaapvaal-Zimbabwe craton do not contain diamonds for the same reasons that kimber-
lites in that tectonic setting are diamond free (see above). Other melilitite provinces
(Balcones, Bathurst Island, Rhine Graben) occur in regions of lithospheric thinning and
rifting and do not contain diamonds. This may be simply due to formation by partial
melting at shallow depths in regions distant from any diamond-bearing horizons (Figure
4.19). Melilitites have been insufficiently studied to determine whether the Bushmanland-
Sutherland, Hudson Bay Lowlands, and Anabar provinces are different in character and/or
derived from different depths and sources than the rift-related suites.
I maintain that every explanation may be funher explained,
by a theory or conjecture of a higher degree of universality.
There can be no explanation which is not in need offurther
explanation, for none can be a self-explanatory description
of an essence.
Karl Popper (1957)

POSTSCRIPT

During the preparation of this work it became evident that very little is known about many
aspects of orangeite petrology. Investigations of these rocks have been neglected in favor
of those ofthe more abundant and easily accessible archetypal kimberlites. Consequently,
studies of orangeites must be regarded as being in their infancy relative to those of
kimberlites (and lamproites). A long-term program of research is desirable to investigate,
in detail, the geology of the occurrences, together with their mineralogy and geochemistry.
Particular attention must be given to integrating the field, mineralogical, and geochemical
studies. The hypotheses for the genesis of continental potassic magmas summarized in
Chapter 4 immediately suggest several potentially fruitful lines of high-pressure experi-
mental research which would complement these studies. It is only by such comprehensive
investigations that we shall gain a deeper understanding of the origin and evolution of
the parental magmas.
Some specific programs of research on orangeites might include the following:
1. Detailed mineralogical investigations of specific orangeite occurrences: Curi-
ously, even the better-known localities such as Swartruggens and Bellsbank have
not been subjected to such scrutiny.
2. Determination of the mineralogical and geochemical characteristics of various
fields: This work would lead to a better understanding of the regional differences
in orangeite magmatism.
3. Characterization of evolved orangeites: These rocks are the least understood
members of the orangeite clan; hence, any new data would represent a significant
increase in our knowledge.
4. Detailed geological and petrological studies of pelletal-textured orangeites:
Such studies should be paralleled by investigation of pelletal-textured kimber-
lites, alnoites, and melilitites. The ultimate goal of this work should be toward
devising a universal model of the diatreme-forming process.
5. Studies of stable and radiogenic isotopic compositions and trace element con-
tents of minerals found in orangeites: This work would have a bearing upon the
sources and evolution of orangeite magmas.
385
386 POSTSCRIPT

6. Experimental studies of melting at high pressures of the mineral assemblages


which have been proposed as the source rocks of orangeite magmas in the lower
lithosphere, e.g., phlogopite pyroxenites, phlogopite-potassium richterite-
apatite veins.
These and other studies, which will no doubt spring to the mind of the reader, could
form the basis of numerous graduate theses and provide many years of challenging
research.
Similar research proposals could be advocated for kimberlites. Interestingly, most of
the projects suggested by Mitchell (1986) have not yet been addressed. In particular,
studies of whole fields of kimberlites, volcanological studies of crater facies rocks, and
Lu-Hf geochronology of the megacryst suite have not been undertaken. Some progress
has been made in experimental studies directed at reproducing the differentiation and
formation of kimberlite magmas at high and low pressures, but many more studies are
required to meet our objective of achieving a greater understanding of kimberlite
petrogenesis. The recent trend of placing the sources of these magmas at ever-deeper
depths in the mantle necessitates an extensive program of experimental studies at very
high pressures in order to confirm or reject these hypotheses.
In conclusion. there are opportunities for a wide variety of studies of the mineralogy,
geochemistry, and petrology of orangeites, kimberlites, and lamproites, which should lead
to an improved understanding of their genesis. Consequently, it is highly probable that
studies of these exotic rocks will continue to flourish during the next decade, althOUgh.
according to the philosophy of Karl Popper, it is unlikely that we shall ever really
understand their origins. Undoubtedly, we will continue to devise ever more elaborate
petrogenetic schemes, but should never abandon the quest for a deeper understanding of
their character. The excitement and challenge of discovery will continue to motivate
further investigation-and perhaps even prove the pessimistic Popper to be incorrect!
REFERENCES

Agee, J. 1., Garrison, J. R., Taylor, L. A. 1982. Petrogenesis of oxide minerals in kimberlite, Elliot County,
Kentucky. Am. Mineral. 67, 28-42.
Alibert, C., Albarede, F. 1988. Relationships between mineralogical, chemical, and isotopic properties of some
North American kimberlites. J. Geophys. Res. 93, 7643-7671.
Allegre, C. J., Thrcotte, D. L. 1985. Geodynamic mixing in the mesosphere boundary layer and the origin of
oceanic islands. Geophys. Res. Lett. 12,207-210.
Allen, J. F., Carmichael, I. S. E. 1984. Lamprophyric lavas in the Colima graben, SW Mexico. Contrib. Mineral.
Petrol. 88,203-216.
Allen, R. L. 1991. Nomenclature of Volcaniclastics: Discrimination of syn-eruptive and post-eruptive volcani-
clastics. IAVCEI Commission on Volcanogenic Sediments, Newsletter 5.
Allsopp, H. L., Barrett, D. R. 1975. Rb-Sr age determinations of South African kimberlites. Phys. Chem. Earth
9,605-617.
Anderson, D. L. 1984. Kimberlites and the evolution of the mantle. In Kornprobst (1984) q.v., Vol. I, pp.
395-403.
Anderson, D. L. 1987. The depths of mantle reservoirs. In Mysen, B. O. (ed.), Magmatic Processes:
Physicochemical Principles. Special Publication No. I, The Geochemical Society, pp. 3-12.
Anderson, D. L., Zhang, Y., Tanimoto, 1992. Plume heads, continental lithosphere, flood basalts and tomogra-
phy. In Storey et al. (1992) q.v., pp. 99-124.
Anderson, O. L. 1979. The role offracture dynamics in kimberlite pipe formation. In Boyd and Meyer (1979)
q.v., Vol. I, pp. 344-353.
Appleton, J. D., Bland, D. J., Nancarrow, P. H., Styles, M. T., Mambwe, S., Zambezi, P. 1992. The occurrence
of daqingshanite-(Ce) in the Nkombwa Hill carbonatite, Zambia. Mineral. Mag. 56, 419-422.
Apter, D. B., Harper, F. J., Wyatt, B. A., Scott Smith, B. H. 1984. The geology of the Mayeng kimberlite sill
complex, South A{rica. In Kornprobst(l984) q.v., Vol. I, pp. 43-57.
Arima, M., Barnett, R. L., Kerrich, R. 1986. Chemical and textural variations of mica in the Nickila Lake and
Upper Canada Mine kimberlites, Ontario, Canada. Fourth Int. Kimberlite Conf., perth, Western Australia,
Extended Abstracts, pp. 15-17.
Arima, M., Nakayama, K., Akaishi, M., Yamoaka, S., Kanda H. 1993b. Crystallization of diamond from a
silicate melt of kimberlite composition in high pressure and high temperature experiments. Geology 21,
968-970.
Arima, M., Yamasita, H., Ohtani, E. 1993a. Petrogenesis of kimberlite magma and its bearing on mantle
metasomatism. Japan Earth and Planetary Science Joint Meeting, Tokyo, Japan (abstract).
Artsybasheva, T. F., Blagulkina, V. A., Rovsha, V. S., Sarsadskikh, N. N. 1964. The problem of the classification
of the Yakutia kimberlites based on those of the Alakit-Daldynsk diamantiferous regions. Int. Geol. Rev.
6, 1173-1781.
Bachinski, S. W., Simpson, E. L. 1984. TI-phlogopites ofthe Shaw's Cove minette: A comparison with micas
of other lamprophyres, potassic rocks, kimberlites and mantle xenoliths. Am. Mineral. 69, 41-56.
Bailey, D. K. 1977. Lithospheric control of continental rift magmatism. J. Geol. Soc. London 133, 103-106.

387
388 REFERENCES

Bailey, D. K. 1980. Volatile flux, geothenns and the generation of the kimberlite-carbonatite-a1kaline magma
spectrum. Mineral. Mag. 43, 695-699.
Bailey, D. K. 1983. The chemical and thermal evolution ofrifts. Tectonophysics 94,585-597.
Bailey, D. K. 1984. Kimberlite "The Mantle Sample" formed by ultrametasomatism. In Kornprobst (1984) q. v.,
Vol. 1, pp. 323-333.
Bailey, D. K. 1992. Episodic alkaline igneous activity across Africa: Implications for the causes of continental
break-up. In Storey et al. (1992) q.v., pp. 91-98.
Bailey, D. K. 1993. Petrogenetic implications of the timing of alkaline, carbonatite and kimberlite igneous
activity in Africa. S. Afr. 1. Geol. 96, 67-74.
Bardet, M. G. 1974. Geologie du Diamant, Vol. II. Gisements de diamant d' Afrique. Bureau de Recherches
Geologiques et Minieres Memoir 83, Editions B.R.G.M., Paris, 223 pp.
Barrett, D. R, Berg, G. W. 1975. Complementary petrographic and strontium-isotope ratio studies of South
African kimberlite. Phys. Chern. Earth 9, 619-635.
Barton, M., van Bergen, M. J. 1981. Green c1inopyroxenes and associated phases in potassium-rich lava from
the Leucite Hills, Wyoming. COll1rib. Mineral. Petrol. 77, 101-114.
Bedard, 1. H., Francis, D. M., Ludden, J. N. 1988. Petrology and pyroxene chemistry of Monteregian dykes:
The origin of concentric zoning and green cores in clinopyroxenes from alkali basalts and lamprophyres.
Can. 1. Earth Sci. 25, 2041-2058.
Bell, D. R, Rossman, G. R 1992. The distribution of hydroxyl in garnets from the subcontinental mantle of
southern Africa. COl11rib. Mineral. Petrol. 111, 161-178.
Berg, G. W. 1986. Evidence for carbonate in the mantIe. Nature 324,50--51.
Berg, G. W. 1989. The significance of brucite in South African kimberlites. In Ross et al. (1989) q.v., Vol. 1,
pp. 282-296.
Bergman, S. C. 1987. Lamproites and other potassium-rich igneous rocks: A review of their occurrence,
mineralogy and geochemistry. In Fitton,J. G., and Upton, B. G. (eds.), Alkaline igneous Rocks. Geol. Soc.
London Spec. Publ. 30, pp. 95-101.
Bergman, S. C., Dunn, D. P., Krol, L. G. 1988. Rock and mineral chemistry of the Linhaisai minette, central
Kalimantan, Indonesia, and the origin of Borneo diamonds. Can. Mineral. 26, 24-43.
Bizzi, L. A., Smith, C. B., DeWitt, M. 1., Annstrong, R, Meyer, H. O. A. 1994. Mesozoic kimberlites and related
alkaline rocks in southwestern Sao Francisco craton, Brazil: A case for local mantle reservoirs and their
interaction. In Meyer and Leonardos (1994) q.v., Vol. I, pp. 156--171.
Bobrievich, A. P., Bondarenko, M. N., Gnevushev, M. A., Kind, M. A., Korshkov, B. Y., Kuryleva, N. A., Nefoca,
Z. D., Popugaeva, L. A., Popova. E. Z.. Skuiskii. V. D.. Smirnov. G. I., Yurkevtski, R. K., Fayunshteyeyuh,
G. H.• Shukin, V. I. 1959a. Diamonds of Siberia. Gozgeolteknizdat, Moscow (Russian).
Bobrievich, A. P., Bondarenko, M. N., Gnevushev, M. A., Kraslov, L. M., Smirnov, 5.1., Yurkevich, R K. 1959b.
The Diamond Deposits of Yakutia. Izdat. Nedra, Moscow (Russian).
Bobrievich, A. P., I1upin, I. P., Kozlov, I. T., Lebedev, L. I., Pankratov, A. A., Smirnov, G. I., Kharkiv, A. D.
1964. Petrography and Mineralogy of the Kimberlite Rocks ofYakutia. Izdat. Nedra, Moscow (Russian).
Boctor. N. Z., Boyd, F. R 1980. Oxide minerals in the Liqhobong kimberlite, Lesotho. Am. Mineral. 65,
631-638.
Boctor, N. Z., Boyd, F. R 1981. Oxide minerals in a layered kimberlite-carbonatite sill from Benfontein, South
Africa. Collfrib. Mineral. Petrol. 76. 253-259.
Boctor, N. Z., Boyd, F. R 1982. Petrology ofkimberlites from the DeBruyn and Martin Mine, Bellsbank, South
Africa. Am. Mineral. 67, 917-925.
Boctor, N. Z., Meyer, H. O. A. 1979. Oxide and sulphide minerals in kimberlite from Green Mountain, Colorado.
In Boyd and Meyer (1979) q.v., Vol. I, pp. 217-228.
Bogatikov, V. A., Makhotkin, I. L., Kononova, V. A. 1985. Lamproites and their place in the systematics of high
magnesium potassic rocks. izv. Akad. Nauk SSSR, 1985( 12), 3-10.
Bogatikov, V. A., Ryabchikov, I. D., Kononova, V. A., Makhotkin, I., Novgorodova, M. I., Solovova, I., Ga1uskin,
E. V., Ganeyev, N. I., Girnis, A. V., Eremeyev, N. V., Kogarko, L. N., Kudryatseva, G. P., Mikhailichenko,
O. A., Naumov, V. B., Sapozhnikova, E. N. 1991. Lamproites. Izdat. Nauka, Moscow (Russian).
Bonney, T. G. 1899. The parent rock of diamond. R. Soc. London Proc. 65, 235-236.
Bosch,1. L. 1971. The petrology of some kimberlite occurrences in the Barkly West District. Cape Province.
Trans. Geol. Soc. S. Afr. 74, 75-101.
REFERENCES 389

Boyd, F. R., Gurney, J. J. 1986. Diamonds and the African lithosphere. Science 132, 472-477.
Boyd, F. R., Meyer, H. O. A. 1979. Proc. Second Int. Kimberlite Conf. Vol. 1. Kimberlites, Diatremes and
Diamonds: Their Geology, Petrology and Geochemistry. Vol. 2. The Mantle Sample: Inclusions in
Kimberlites and Other Volcanics. Am. Geophys. Union, Washington, D.C.
Boyd, F. R., Nixon, P. H. 1975. Origins of the ultramafic nodules from some kimberlites in northern Lesotho
and the Monastery Mine, South Africa. Phys. Chem. Earth 9, 431-454.
Brakhfogel, F. F. 1984. Geological Aspects of Kimberlitic Magmatism of the Nonh-eastern Siberian Platform.
Yakutian Division of Siberian Branch of the Academy of Science SSSR, Yakutsk.
Brey, G., Brice, W. R., EJlis, D. J., Green, D. H., Harris, K. L. Ryabchikov, I. D. 1983. Pyroxene-carbonate
reactions in the upper mantle. Earth Planet. Sci. Lett. 62, 63-74.
Brey, G. Kogarko, L. N., Ryabchikov, I. D. 1991. Carbon dioxide solubility in kimberlitic melts. Neues Jahr.
Mineral. Monatsh. 1991, 159-168.
Brummer, J. J., MacFadyen, D. A., Pegg, C. C. 1992. Discovery of kimberlites in the Kikland Lake area,
Northern Ontario, Canada. II: Kimberlite discoveries, sampling, diamond content, ages and emplacement.
Explor. Mining Geol. 1,351-370.
Bursill, L. A., Grzinic, G. 1980. Incommensurate superlattice ordering in the hollandites BaxTiS-xMgxOl6 and
BaxTis-xGa2.\OI6. Acta Crystallogr. B 36,2902-2913.
Campbell, I., Griffiths, R. W. 1990. Implications of mantle plume structure for the origin of flood basalts. Earth
Planet. Sci. Lett. 99, 79-93.
Canil, D. 1990. Experimental studies bearing on the absence of carbonate in mantle-derived xenoliths. Geology
18,101\-1013.
Canil, D. 1991. Experimental evidence for the exsolution of cratonic peridotite from high temperature
harzburgite. Earth Planet. Sci. Lett. 106, 64-72.
Canil, D. 1993. Phase equilibria at high pressure applied to the Earth's mantle. In Luth, R. W. (ed.), Experiments
at High Pressure and Applications to the Earth s Mantle. Mineralogical Association of Canada Short
Course Handbook 21, pp. 197-245.
Canil, D., Scarfe, C. M. 1990. Phase relations in peridotite + C02 systems to 12 GPa: Implications for the origin
of kimberlite and carbonate stability in the Earth's upper mantle. J. Geoph)'s. Res. 958,15805-15816.
Carmichael, I. S. E. 1967a. The mineralogy and petrology of the volcanic rocks from the Leucite Hills,
Wyoming. Contrib. Mineral. Petrol. 15, 24-66.
Carmichael, I. S. E. I967b. The iron-titanium oxides of salic volcanic rocks and their associated ferromagnesian
silicates. Contrib. Mineral. Petrol. 14, 36-64.
Carswell, D. A. 1975. Primary and secondary phlogopites in garnet lherzolite xenoliths. Phys. Chem. Earth 9,
417-430.
Carswell, D. A., Gibb, F. G. F. 1987. Gamet lherzolite xenoliths in the kimberlites of northern Lesotho: Revised
P- T equilibration conditions and upper mantle paleogeotherm. Contrib. Mineral. Petrol. 97, 473-487.
Cas, R. 1991. Nomenclature of Volcaniclastics. IAVCEI Commission on Volcanogenic Sediments, Newsletter
4.
Cas, R., Wright, 1. V. 1987. Volcanic Succession: Modem and Anciellt. Allen & Unwin, London.
Clarke, D. B., Carswell, D. A. 1977. Green garnets from the Newlands kimberlite, Cape Province, South Africa.
Eanh Planet. Sci. Lett. 34, 30-38.
Clarke, T. C., Smith, C. B., Bristow, J. W., Skinner, E. M. w., Viljoen, K. S. 1991. Isotopic and geochemical
variation in kimberlites from the south western craton margin, Prieska area, South Africa. Fifth Int.
Kimberlite Conf., Araxa, Brazil. Extended Abstracts. Companhia de Pesquisa de Recursos Minerais,
Special Publication 2/91, pp. 46-47.
Clement, C. R. 1973. Kimberlites from the Kao pipe, Lesotho. In Nixon, P. H. (ed.), Lesotho Kimberlites.
Lesotho National Development Corporation, Maseru, Lesotho, pp. 110-121.
Clement, C. R. 1982. A comparati ve geological study of some major kimberlite pipes in the Northern Cape and
Orange Free State. Ph.D. thesis (2 vols), University of Cape Town, South Africa.
Clement, C. R., Dawson, J. B., Geringer, G. J., Gurney, J. J., Hawthorne, J. B., Krol, L., Kleinjan, L., van Zyl,
A. A. 1973. Guide for the first field excursion. Int. Conf. on Kimberlites, Cape Town, South Africa.
Clement, C. R., Reid, A. M. 1989. The origin of kimberlite pipes: An interpretation based on a synthesis of
geological features displayed by southern African occurrences. In Ross et al. (1989) q.v., Vol. I, pp.
632-646.
390 REFERENCES

Clement, C. R, Skinner, E. M. W. 1979. A textural-genetic classification of kimberlite rocks. Kimberlite Symp.


II, Cambridge, Extended Abstract.
Clement, C. R., Skinner, E. M. W. 1985. A textural-genetic classification of kimberlites. Trans. Geol. Soc. S.
Afr. 88, 403-409.
Clement, C. R, Skinner, E. M. W., Scott Smith, B. H. 1984. Kimberlite re-defined. J. Geol. 32, 223-228.
Clifford, T. N. 1966. Tectono-metallogenic units and metallogenic provinces of Africa. Eanh Planet. Sci. Leu.
1,421-434.
Cloos, H. 1941. Bau und Tatigkeit von Thffschloten. Untersuchungen an den Schwabischen Vulkanen. Geol.
Rllndsch. 32, 709-800.
Coopersmith, H. G. 1993. Diamondiferous kimberlite at Kelsey Lake, Southern Wyoming Archean Province.
In Dunne, K. P. E., and Grant, B. (eds.), Mid-Continental Diamonds. GAC-MAC Symp. Vol., Edmonton,
Alberta, pp. 85-88.
Cox, K. G. 1992. Karoo igneous activity, and the early stages of the break-up of Gondwanaland. In Storey et
al. (1992), pp. 137-148.
Cox, K. G., Smith, M. R., Beswetherick, S. 1987. Textural studies of garnet Iherzolites: Evidence of exsolution
origin from high temperature Iherzolites. In Nixon (1987) q.v., pp. 537-550.
Cullers, L. R., Mullenax, J., Dimarco, M. J., Nordeng, S. 1982. The trace element content and petrogenesis of
kimberlites in Riley County, Kansas, U.S.A. Am. Mineral. 67, 223-233.
Currie, K. L., Ferguson, J. 1970. The mechanism of intrusion of lamprophyre dikes as indicated by "offsetting"
of dikes. Tectonophysics 9, 525-535.
Daly, R. A. 1914.lgneolls Rocks and Their Origin. Hafner, New York.
Danchin, R V., Ferguson, J., McIver, J. R., Nixon, P. H. 1975. The composition oflate-stage kimberlite liquids
as revealed by nucleated autoliths. Phys. Chem. Earth 9, 235-245.
Davies, G. R, Spriggs, A. J., Nixon, P. H., Rex, D. C. 1991. A non-cognate origin for the Gibeon kimberlite
megacryst suite. Fifth Int. Kimberlite Conf., Araxa, Brazil, Extended Abstracts. Companhia de Pesquisa
de Recursos Minerais Special Publication 2/91, pp. 63...Q5.
Dawson, J. B. 1967. A review of the geology of kimberlite. In Wyllie, P. J. (ed.), Utramaficand Related Rocks.
Wiley, New York, pp. 241-251.
Dawson, J. B. 1971. Advances in kimberlite geology. Earth Sci. Rev., 7, 187-214.
Dawson, J. B. 1972. Kimberlites and their relationship to the upper mantle. R. Soc. London Phil. Trans. 271A,
297-311.
Dawson, 1. B. 1980. Kimberlites and Their Xenoliths. Springer-Verlag, New York.
Dawson, 1. B. 1984. Contrasting types of upper mantle metasomatism. In Kornprobst (1984) q.v., Vol. 2, pp.
289-294.
Dawson, 1. B. 1987. The kimberlite clan: Relationship with olivine and leucite lamproite, and inferences for
uppermantJe metasomatism. In Fitton, 1. G., and Upton, B. G. (eds.), Alkaline Igneolls Rocks. Geol. Soci.
London Spec. Publ. 30, pp. 95-101.
Dawson, J. B. 1989. Geographic and time distribution ofkimberlites and lamproites: Relationships to tectonic
processes. In Ross et al. (1989) q.v., Vol. I, pp. 323-342.
Dawson, J. B. 1994. Quaternary kimberlitic volcanism on the Tanzania craton. Contrib. Mineral. Petrol. 116,
473-485.
Dawson, J. B., Hawthorne, J. B. 1973. Magmatic sedimentation and carbonatitic differentiation in kimberlite
sills at Benfontein, South Africa. J. Geol. Soc. London 129, 61-85.
Dawson,1. B., Hervig, R. L., Smith, J. V. 1981. Fertile iron-rich dunite xenoliths from the BuItfontein kimberlite,
South Africa. Fortsch. Mineral. 59, 303-324.
Dawson, J. B., Smith, 1. V. 1975. Chemistry and origin of phlogopite megacrysts in kimberlite. Natllre 253,
336-338.
Dawson, 1. B., Smith, 1. V. 1977. The MARIO (mica-amphibole-rutile-ilmenite-diopside) suite of xenoliths in
kimberlite. Geochim. Cosmochim. Acta 41, 309-332.
Dawson,1. B., Smith, 1. V., Hervig, R L. 1977. Late-stage diopside in kimberlite groundmass. Nelles Jahr.
Mineral. Monatsh. 1977, 529-553.
Deines, P. 1989. Stable isotope variations in carbonatites. In Bell, K. (ed.), Carbonatites: Genesis and
Evoilition. Unwin Hyman, London, pp. 301-359.
REFERENCES 391

Deines, P., Gold. D. P. 1973. The isotopic composition of carbonatite and kimberlite carbonates and their bearing
on the isotopic composition of deep-seated carbon. Geochim. Cosmochim. Acta 37, 1709-1733.
Delaney, J. S., Smith, J. V., Carswell, D. A., Dawson, 1. B. 1980. Chemistry of micas from kimberlites and
xenoliths. II: Primary- and secondary-textured micas from peridotite xenoliths. Geochim. Cosmochim.
Acta 44, 857-872.
Delor, C. P., Rock, N. M. S. 1991. Alkaline-ultramafic lamprophyre dykes from the Vestfold Hills, Princess
Elizabeth Land (East Antarctica): Primitive magmas of deep mantle origin. Antarctic Sci. 3, 419-432.
De Paolo, D. J., Wasserburg, G. 1. 1976. Nd isotopic variation and petrogenetic models. Geophys. Res. Lett. 3,
249-252.
Dobbs, P. N., Duncan, D. J., Hu, S., Shee, S. R., Cogan, E. A., Brown,M. A., Smith, C. B., Allsopp, H. L. 1994.
The geology of the Mengyin kimberlites, Shandong China. In Meyer and Leonardos (1994) q.v., Vol. 1,
pp.40-61.
Dobosi, G., Fodor, R. V. 1992. Magma fractionation, replenishment, and mixing as inferred from green-core
c1inopyroxenes in Pliocene basanite, southern Slovakia. Lithos 28, 133-150.
Donaldson, C. H., Reid, A. M. 1982. Multiple intrusion ofakimberlite dike. Trans. Geol. Soc. S.Afr. 85,1-12.
Donaldson, G. 1974. Scotland-The Making of a Nation. David & Charles Press, Newton Abbot, Devon, UK.
Droop, G. T. R. 1987. A general equation for estimating Fe 3+ concentrations in ferromagnesian silicates and
oxides from microprobe analyses using stoichiometric criteria. Mineral. Mag. 51, 431-435.
Duda, A., Schminke, H. U. 1985. Polybaric differentiation of alkali basaltic magmas: Evidence from green-core
clinopyroxenes (Eifel, FRG). Contrib. Mineral. Petrol. 91, 340-353.
Dudas, F. 0., Carlson, R. W., Eggler, D. H. 1987. Regional Middle Proterozoic enrichment of subcontinental
mantle source of igneous rocks from central Montana. Geology 15, 2-25.
Edgar, A. D. 1987. The genesis of alkaline rocks with emphasis on their source regions: Inferences from
experimental studies. In Fitton, 1. G., and Upton, B. G. (eds.), Alkaline Igneous Rocks. Geo!. Soc. London
Spec. Pub. 30, pp. 29-52
Edgar, A. D. 1989. Barium- and strontium-enriched apatites in lamproites from West Kimberley, Western
Australia. Am. Mineral. 74, 889-895.
Edgar, A. D., Arima, M., Baldwin, D. K., Bell, D. R., Shee, S. R., Skinner, E. M. w., Walker, E. C. 1988. High
pressure-high temperature melting experiments on a Si02-poor aphanitic kimberlite from the Wesselton
mine, Kimberley, South Africa. Am. Mineral. 73, 524-533.
Edgar, A. D., Charbonneau, H. E. 1993. Melting experiments on a Si02-poor, CaO-rich aphanitic kimberlite
from 5-10 GPa and their bearing on sources of kimberlite magmas. Am. Mineral. 78,132-142.
Edgar, A. D., Pizzolato, L. A. 1995. Distribution of F in the model mantle systems K-richterite-apatite and
phlogopite-K -richterite at 20kb: consequences for F as a compatible element in ultrapotassic magma
genesis. Milleral. Petrol. 57, in press.
Edgar, A. D., Vukadinovic, D. 1993. Potassium-rich clinopyroxene in the mantle: An experimental investigation
of a K-rich lamproite up to 60 kbar. Geochim. Cosmochim. Acta 57, 5063-5072.
Edwards, C. B., Howkins, J. B. 1966. Kimberlites in Tanganyika with special reference to the Mwadui
occurrence. £Con. Geol. 62, 537-554.
Edwards, D., Rock, N. M. S., Taylor, W. R., Griffin, B. J., Ramsay, R. R. 1992. Mineralogy and petrology of
the Aries diamondiferous kimberlite pipe, Central Kimberley Block, Western Australia. J. Petrol. 3,
1157-1191.
Eggler, D. H. 1974. Peridotite-carbonate relations in the system CaO-MgO-Si02-C02. Carnegie Irzst.
Washington Yearb. 74, 468-474.
Eggler, D. H. 1987. Discussion of recent papers on carbonated peridotite, bearing on metasomatism and
magmatism: An alternative. Earth Planet. Sci. Lett. 82, 398-400.
Eggler, D. H. 1989. Kimberlites: How do they form? In Ross et al. (1989) q.v., Vol. I, pp. 489-504.
Eggler, D. H., Baker, D. R. 1982. Reduced volatiles in the system C-O-H: Implications to melting, fluid
formation and diamond genesis. In Akimoto, S., and Manghnanai, M. H. (eds.), High Pressure Research
in Geophysics. Center for Academic Pub!., Tokyo, pp. 237-250.
Eggler, D. H., McCallum, M. E., Smith, C. B. 1979. Megacryst assemblages in kimberlite from northern
Colorado and southern Wyoming: Petrology, geothermometry-barometry and areal distribution. In Boyd
and Meyer (1979) q.v., Vol. 2, pp. 213-226.
392 REFERENCES

Eggler, D. H., Wendlandt, R. F. 1978. Phase relations of a kimberlite composition. Carnegie Inst. Washington
Yearb. 77, 751-756.
Eggler, D. H., Wendlandt, R. F. 1979. Experimental studies on the relationship between kimberlite magmas and
partial melting of peridotite. In Boyd and Meyer (1979) q.v., Vol. I, pp. 330--338.
Egorov, K. N., Bogdanov, G. V., Zavyalova, L. L. 1991. New data on the composition of kimberlite from the
Zagodachnaya pipe, Yakutia. /zv. Akad. Nauk SSSR 11,98-108 (Russian).
Egorov, K. N., Kornilova, V. P., Saphranov, A. P., Philipov, N. D. 1986. Micaceous kimberlite from the East
Udachnaya pipe. Dokl. Akad. Nauk 291, 199-202 (Russian).
Ellam, R. M., Cox, K. G. 1991. An interpretation of Karoo picrite basalts in terms of interaction between
asthenospheric magmas and the mantle lithosphere. Earth Planet. Sci. Lett. 105, 2330--342.
Ellis, D. 1., Green, D. H. An experimental study ofthe effect of Ca upon garnet-clinopyroxene Fe-Mg exchange
equilibria. Contrib. Mineral. Petrol. 71, 13-22.
Erlank, A. J. 1973. Kimberlite potassium richterite and the distribution of potassium in the upper mantle. First
Int. Kimberlite Conf., Cape Town, South Africa, Extended Abstracts, pp. 236-238.
Erlank, A. J., Waters, F. G., Hawkesworth, C. 1., Haggerty, S. E., Allsopp, H. L., Rickard, R. S., Menzies, M.
A. 1987. Evidence for mantle metasomatism in peridotite nodules from the Kimberley pipes, South Africa.
In Menzies, M. A., and Hawkesworth, C. J. (eds.), Mantle Metasomatism. Academic Press, London, pp.
221-311.
Exley, R. A., Smith, J. V. 1982. The role of apatite in mantle enrichment processes and in the petrogenesis of
some alkali basalt suites. Geochim. Cosmochim. Acta 46, 1375-1384.
Falloon, T., Green, D. H. 1989. The solidus of carbonated fertile peridotite. Earth Planet. Sci. Lett. 94, 364-370.
Farmer, G. L., Boettcher, A. L. 1981. Petrologic and crystal-chemical significance of some deep-seated
phlogopites. Am. Mineral. 66, 1154-1163.
Ferguson, 1., Darchin, J., Nixon, P. H. 1973. Petrochemistry of kimberlite autoliths. In Nixon, P. H. (ed.), Lesotho
Kimberlites. Lesotho National Development Corporation Maseru, Lesotho, pp. 285-293.
Fesq, H. w., Kable, E. 1. D., Gurney, 1. J. 1975. Aspects of the geochemistry ofkimberlites from the Premier
Mine and other South African occurrences, with particular reference to the rare earth elements. Phys.
Chem. Earth 9, 686-707.
Fielding, D. c., Jaques, A. L. 1989. Geology, petrology and geochemistry of the Bow Hilliamprophyre dikes,
Western Australia. In Ross et al. (1989) q.v., Vol. I, pp. 206-219.
Finnerty, A. A., Boyd, F. R. 1987. Thermobarometry for garnet peridotite xenoliths: A basis for upper mantle
stratigraphy. In Nixon (1987) q.v., pp. 381-402.
Fisher, R. V. 1961. Proposed classification of volcaniclastic sediments and rocks. Geol. Soc. Am. Bull. 72,
1409-1413.
Fisher, R. v., Schminke, H. U. 1984. Pyroclastic Rocks. Springer-Verlag, New York.
Foley, S. F. 1988. The genesis of continental basic alkaline magmas-an interpretation in terms of redox melting.
J. Petrol., Special Lithosphere Issue, pp. 138-161.
Foley, S. F. 1989. The genesis of lamproitic magmas in a reduced fluorine-rich mantle. In Ross et al. (1989)
q.v., Vol. I, pp. 616-630.
Foley, S. F. 1990. A review and assessment of experiments on kimberlites, lamproites and lamprophyres as a
guide to their origin. Proc.Indian Acad. Sci. Earth Planet. Sci. 99, 57-80.
Foley, S. F. 1991. High pressure stability of the fluor- and hydroxy end-members of pargasite and K-richterite.
Geochim. Cosmochim. Acta 55, 2689-2694.
Foley, S. F. 1992a. Petrological characterization of the source components of potassic magmas: Geochemical
and experimental constraints. Lithos 28, 187-204.
Foley, S. F. 1992b. Vein-plus-wail-rockmeIting mechanisms in the lithosphere and the origin of potassic alkaline
magmas. Lithos 28, 435-453.
Foley, S. F. 1993. An experimental study of olivine lamproite: First results from the diamond stability field.
Geochim. Cosmochim. Acta 57, 483-489.
Foley, S. F., Hofer, H., Brey, G. 1994. High pressure synthesis of priderite and members of the lindsleyite-mathiasite
and hawthomeite-yimengite series. Contrib. Mineral. Petrol. 117, 164-174.
Foley, S. F., Taylor, W. R., Green, D. H. 1986. The effect of fluorine on phase relationships in the system
KAISi04-Mg2Si04-Si02 at 28 kbar and the solution mechanism of fluorine in silicate melts. Contrib.
Mineral. Petrol. 93, 46-55.
REFERENCES 393

Fourie, G. P. 1958. Die diamantvoorkomste in die omgewing van Swartruggens, Transvaal. Geol. Survey S. AI
Bull. 26, 1-16.
Fraser, K. 1. 1987. Petrogenesis ofkimberlites from South Africa and lamproites from Western Australia and
North America. Ph.D. thesis, The Open University, Milton Keynes, UK.
Fraser, K. 1., Hawkesworth, C. J. 1992. The petrogenesis of group 2 ultrapotassic kimberlites from the Finsch
Mine, South Africa. Lithos 28, 327-345.
Fraser, K. 1., Hawkesworth, C. 1., Erlank, A. J., Mitchell, R. H., Scott Smith, B. H. 1985. Sr, Nd, and Pb isotope
and minor element geochemistry of lamproites and kimberlites. Earth Planet. Sc. Lett. 76, 57-70.
Frantsseson, E. V. 1968. The Petrology of Kimberlites. Izdat. Nedra, Moscow (Russian).
Frantsseson, E. V. 1970. The Peflvlogy of Kimberlites. (Translated by D. A. Brown). Publication No. 150,
Department of Geology, Australian National University, Canberra, Australia.
Fritz, W. 1. 1993. Tuff or Sandstone?-Usage of Size and Compositional Terms in the Classification of
Pyroclastic Rocks. IAVCEI Commission on Volcanogenic Sediments, Newsletter 8.
Gaspar, J. C., Wyllie, P. 1. 1984. The alleged kimberlite-carbonatite relationship: Evidence from ilmenite and
spinel from Premier and WesseltonMines and the Benfontein Sill, South Africa. Contrib. Mineral. Petrol.
85, 133-140.
Gatehouse, B. M., Jones, G. C., Pring, A. 1986. The chemistry and structure ofredledgeite. Mineral. Mag. 50,
709-715.
Grantham, D. R., Allen, J. B. 1960. Kimberlite in Sierra Leone. Overseas Geol. Mineral Res. 8, 5-25.
Green, D. H., Falloon, T. J., Taylor, W. R. 1987. Mantle-derived magmas-roles of variable source peridotite
and variable C-H-O fluid compositions. In Mysen, B. O. (ed.), Magmatic Processes: Physicochemical
Principles. Geochem. Soc. Spec. Publ. No.1, pp. 139-154.
Green, D. H., Sobolev, N. V. 1975. Co-existing garnets and ilmenites synthesized at high pressures from pyrolite
and olivine basanite and their significance for kimberlitic assemblages. COllfrib. Mineral. Petrol. 50,
217-229.
Green, H. W., Guegen, Y. 1974. Origin of kimberlite pipes by diapiric upwelling in the upper mantle. Nature
249,617--620.
Gupta, A. K., Green, D. H. 1988. The liquidus surface of the system forsterite-kalsilite-quartz at 28 kbar under
dry conditions, in the presence of H20 and C02. Mineral. Petrol. 39, 163-174.
Gurney, J. J. 1989. Diamonds. In Ross et al. (1989) q.v., Vol. 2, pp. 935-965.
Gurney, 1. J., Berg, G. W. 1969. Potassium, rubidium and cesium in South African kimberlites and their peridotite
xenoliths. Geol. Soc. S. Afr. Spec. Publ. 2, 417-427.
Gurney, 1. 1., Ebrahim, S. 1973. Chemical composition of Lesotho kimberlites. In Nixon, P. H. (ed.), Lesotho
Kimberlites. Lesotho National Development Corporation, Maseru, Lesotho, pp. 280-294.
Gurney, 1. J., Harris, 1. w., Rickard, R. S. 1979a. Silicate and oxide inclusions in diamonds from the Finsch
kimberlite pipe. In Boyd and Meyer (1979) q.v., Vol. I, pp. 1-15.
Gurney, 1. J., Hobbs, J. B. M. 1973. Potassium, thorium and uranium in some kimberlites from South Africa.
First Int. Kimberlite Conf., Cape Town, South Africa, Extended Abstracts, pp. 143-146.
Gurney, J. 1., Jakob, W. R. 0., Dawson, J. B. 1979b. Megacrysts from the Monastery kimberlite pipe, South
Africa. In Boyd and Meyer (1979) q.v., Vol. 2, pp. 222-243.
Gurney, 1. 1., Switzer, G. S. 1973. The discovery of garnets closely related to diamonds in the Finsch pipe, South
Africa. Collfrib. Mineral. Pet/vi. 39, 103-116.
Gunther, M., lagoutz, E. 1991. Systematics of isotope disequilibria between minerals of low temperature garnet
Iherzolites. Fifth Int. Kimberlite Conf., Araxa, Brazil, Extended Abstracts, Companhia de Pesquisa de
Recursos Minerais, Spec. Publ. 2191, pp. 122-124.
Haggerty, S. E. 1975. The chemistry and genesis of opaque minerals in kimberlites. Phys. Chem. Eanh 9,
295-307.
Haggerty, S. E. 1976. Opaque mineral oxides in terrestrial igneous rocks. In Rumble, D. (Ed.), Oxide Minerals.
Mineral. Soc. Am., Rev. Miner. 3, pp. HglOI-Hg300.
Haggerty, S. E. 1983. The mineral chemistry of new titanates from the lagersfontein kimberlite, South Africa:
Implications for metasomatism in the upper mantle. Geoc/lim. Cosmoschim. Acta 47, 1833-1854.
Haggerty, S. E. 1986. Diamond genesis in a multiply-constrained model. Nature 320, 34-38.
Haggerty, S. E. 1989a. Upper mantle opaque mineral stratigraphy and the genesis ofmetasomites and alkali-rich
melts. In Ross et al. (1989) q.v., Vol. 2, pp. 687--699.
394 REFERENCES

Haggerty, S. E. 1989b. Mantle metasomes and the kinship between carbonatites and kimberlites. In Bell, K.
(ed.), Carbonatites. Unwin Hyman, London, pp. 546-560.
Haggerty, S. E. 1994. Superkimberlites: A geodynamic window to the Earth's core. Earth Planet. Sci. Lett. 122,
57-69.
Haggerty, S. E., Hardie, R. B., McMahon, R. M. 1979. The mineral chemistry of ilmenite nodule associations
from the Monastery diatreme. In Boyd and Meyer (1979) q.v., Vol. 2, pp. 249-256.
Haggerty, S. E., Sautter, V. 1990. Ultra-deep (>300 km) ultramafic xenoliths: New petrologic evidence from
the transition zone. Science 252, 827-830.
Hall, A. 1982. The Pendennis peralkaline minette. Mineral. Mag. 45, 257-266.
Hamilton, R., Rock, N. M. S. 1990. Geochemistry, mineralogy and petrology of a new find of ultramafic
lamprophyres from Bulljah Pool, Nabberu Basin, Yilgarn Craton, Western Australia. Lithos 24, 275-290.
Harlow, G. E., Veblen, B. R. 1991. Potassium in clinopyroxene: Inclusions in diamonds. Science 251, 652-655.
Harris, 1. W. 1987. Recent physical, chemical and isotopic research of diamond. In Nixon (1987) q.v., pp.
477-500.
Harte, B. 1983. Mantle peridotites and processes-the kimberlite sample. In Hawkesworth, C. J., and Norry,
M. J. (eds.), Continental Basalt and Mantle Xenoliths. Shiva, Nantwich, pp. 46-91.
Harte, B., Gurney, J. 1. 1981. The mode of formation of chromium-poor megacryst suites from kimberlites. 1.
Geol. 89, 749-753.
Harte, B., Gurney, 1. 1., Harris, J. W. 1980. The formation of peridotite suite inclusions in diamonds. Contrib.
Mineral. Petrol. 72, 181-190.
Hatton, C. 1. 1978. The geochemistry and origin of xenoliths from the Roberts Victor Mine. Ph.D. thesis.
University of Cape Town, South Africa.
Hatton, C. 1., Gurney, 1. J. 1979. A diamond-graphite eclogite from the Roberts Victor Mine. In Boyd and Meyer
(1979) q.v., Vol. 2, pp. 29-36.
Haughton, S. H. 1935. The kimberlite fissures near Swartruggens. Open File Report Geological Survey Union
of South Africa, Pretoria.
Hawkesworth, C. 1., van Calsteren, P. W. C. 1984. Radiogenic isotopes, some geological applications. In
Henderson, P. (ed.), Rare Earth Element Geochemistry. Elsevier, New York, pp. 375-421.
Hawkesworth, C. J., Fraser, K. 1., Rogers, N. W. 1985. Kimberlites and lamproites: Extreme products of mantle
enrichment processes. Trans. Geol. Soc. S. Afr., 88, 439-447.
Hawkesworth, C. J., Mantovani, M. S. M., Taylor, P. N., Palacz, Z. 1986. Evidence from the Parana of South
Brazil for a continental contribution to DUPAL basalts. Nature 322,356-359.
Hawthorne, J. B. 1975. Model of a kimberlite pipe. Phys. Chern. Earth 9, 1-15.
Hawthorne, J. B., Carrington, A. J., Clement, C. R., Skinner, E. M. W. 1979. Geology of the Dokolwayo
kimberlite and associated paleoalluvial diamond deposits. In Boyd and Meyer (1979) q.v., Vol. I, pp.
59-70.
Hearn, B. C. 1968. Diatremes with kimberlitic affinities in north central Montana. Science 159, 622-625.
Hearn, B. C. 1989. Alkalic ultramafic magmas in north central Montana, USA: Genetic connections of alnoite,
kimberlite and carbonatite. In Ross et at. (1989) q.v., Vol. I, 109-139.
Helmstaedt, H., Schulze, D. J. 1989. Southern African kimberlites and their mantle sample-implications for
Archean tectonics and lithosphere evolution. In Ross et al. (1989) q. v., Vol. I, pp. 358-368.
Herzberg, C. T. 1993. Lithospheric peridotites of the Kaapvaal craton. Earth Planet. Sci. Lett. 120, 13-29.
Hess, P. C. 1989. Origins of Igneous Rocks. Harvard University Press, Cambridge, MA.
Hops, 1. J., Gurney, J. J., Harte, B. 1992. The Jagersfontein Cr-poor megacryst suite-towards a model for
megacryst petrogenesis. 1. Volcallol. Geothenll. Res. 50, 143-160.
Houghton, B. E, Smith, R. T. 1993. Recycling of magmatic clasts during explosive eruptions: Estimating the
true juvenile content of phreatomagmatic volcanic deposits. Bull. Volcanol. 55, 414-420.
Hunter, R. H., McKenzie, D. 1989. The equilibrium geometry of carbonate melts in rocks of mantle composition.
Earth Planet. Sci. Lett. 92, 347-356.
Hunter, R. H., Kissling, R. D., Taylor, L. A. 1984. Mid to late-stage kimberlitic melt evolution: Phlogopites and
oxides from the Fayette County kimberlite, Pennsylvania. Am. Mineral. 69, 30--40.
Hunter, R. H., Taylor, L. A. 1984. Magma mixing in the low velocity zone: Kimberlite megacrysts from Fayette
County, Pennsylvania. Am. Mineral. 69,16-29.
REFERENCES 395

Ilupin, I. P., Lutz, B. G. 1971. The chemical composition of kimberlite and questions on the origin of kimberlite
magmas. Soviet. Geol. 6, 61-73 (Russian).
Ilupin, I. P., Khomyakov, A. P., Balashov, Y. A. 1971. Rare earths in accessory minerals of Yakutian kimberlites.
Dokl. AkLld. Nauk SSSR 201,1214-1217 (Russian).
Irifune, T., Hibberson, W. A., Ringwood, A. E. 1989. Eclogite-gametite transformation at high pressure and its
bearing on the occurrence of gamet inclusions in diamond. In Ross et al. (1989) q. v., Vol. 2, pp. 877-882.
Irvine, T. N., 1965. Chromium spinel as a petrogenetic indicator. Part 1. Theory. Can.l. Earth Sci. 2, 648-672.
Jago, B. C., Mitchell, R. H., 1985. Mineralogy and petrology of the Ham kimberlite, Somerset Island, Northwest
Territories, Canada. Can. Mineral. 23, 619-634.
James, D. E. 1989. Continental lithosphere. In James, D. E. (ed.), The Encyclopedia o/Solid Earth Geophysics.
Van Nostrand Reinhold, New York, pp. 92-97.
Janse, A. J. A. 1994. Review of supposedly non-kimberlitic and non-Iamproitic diamond host rocks. In Meyer
and Leonardos (1994) q.v., Vol. 2, pp. 144-159.
Janse, A.J. A., Downie, I. E, Reed, L. E., Sinclair, G. L. 1989. Alkaline intrusions in the Hudson Bay Lowlands,
Canada: Exploration methods, petrology and geochemistry. In Ross et al. (1989) q. v., pp. 1192-1203.
Jaques, A. L. 1989. Lamproite diamonds and their inclusions: new insights from the West Australian deposits.
Workshop on Diamonds, 28th Int. Geol. Congress, Washington, Extended Abstracts, pp. 36-39.
Jaques, A. L., Haggerty, S. E., Lucas, H., Boxer, G. L. 1989a. Mineralogy and petrology of the Argyle (AK1)
lamproite pipe, Western Australia. In Ross et al. (1989) q.v., Vol. I, pp. 153-169.
Jaques, A. L., Lewis, J. D., Smith, C. B. 1986. The kimberlites and lamproites of Western Australia. Geol. Surv.
W. Austral. Bull. 132, 268 pp.
Jaques, A. L., Sun, S. S., Chappell, B. W. 1989b. Geochemistry of the Argyle (AKl) lamproite pipe, Western
Australia. In Ross et al. (1989) q. v., Vol. I, pp. 170-188.
Jeanloz, R. 1987. Earth's mantle. Encycl. ihys. Sci. Technol. 4, 486-502.
Jones, A. P., Smith, 1. V. 1983. Petrological significance of mineral chemistry in the Agathla Peak and the Thumb
minettes, Navajo volcanic field. l. Geol. 91, 643-656.
Jones, A. P., Smith J. V. 1985. Phlogopite and associated minerals from the Permian minettes in Devon, South
England. Geol. Soc. Fililand Bull. 57, 89-102.
Jones, A. P., Wylie, P. J. 1984. Minor elements in perovskite from kimberlites and the distribution of rare earth
elements: An electron microprobe study. Earth Planet. Sci. Lerr. 69, 128-140.
Jones, R. A. 1987. Strontium and neodymium isotopic and rare earth element evidence for the genesis of
megacrysts in kimberlites of southern Africa. In Nixon (1987) q. v., pp. 711-724.
Jordan, T. H. 1978. Composition and development ofthe continental tectosphere. Nature 274,544-548.
Kable, E. J. D., Fesq, H. W., Gurney, 1. 1. 1975. The significance ofinterelement relationships of some minor
and trace elements in South African kimberlites. Phys. Chem. Earth 9, 709-734.
Kaminsksii, E V. 1984. Diamond-Bearing Non-kimberlitic Eruptive Rocks. Nedra, Moscow (Russian).
Kampata, M. D. 1993. Mineralogie et geochimie des kimberlites du haut plateau de Kundelungu. Ph.D. thesis,
Universite Catholique de Louvain, Louvain-Ia-Neuve, Belgium.
Kapustin, Y. L. 1980. Mineralogy 0/ Carbonatites. Amerind, New Delhi.
Katsura, T., Ito, E. 1990. Melting and subsolidus phase relations in the MgSi03-MgC03 system at high
pressures: Implications to evolution of the Earth's atmosphere. Earth Planet. Sci. Lerr. 99, 110-117.
Kesson, S. E., Ringwood, A. E. 1989. Slab mantle interactions 2. The formation of diamonds. Chem. Geol. 78,
97-118.
Kesson, S. E., Ringwood, A. E., Hibberson, W. O. 1994. Kimberlite melting relations revisited. Earth Planet.
Sci. Lerr.121, 261-262.
Kesson, S. E., White, T. J. 1986. [Bax<:sy][(Ti,Al)i;+yTit2x_ylOI6 synroc-type hollandites: I. Phase chemistry.
Proc. R. Soc. London 405A, 73-101.
Kharkiv, A. D. 1990. Structure and composition of some slightly eroded kimberlite pipes. lilt. Geol. Rev. 32,
404-414.
Kharkiv, A. D. 1992. Similarities and differences between kimberlitic rocks from the northern Russian platform
and other regions. Russ. Geol. Geophys. 33, 75-8\.
Khazanovich-Vulf, K. K. 1991. Meteorite-generated electrical discharges as a possible factor governing the
occurrence of diatremes and the metallogeny of kimberlites. Dokl. Akad. Nauk SSSR 319, 127-131.
396 REFERENCES

Khomyakov, A., Polezhaeva, 1. I., Smolyaninova, N. N. 1983b. Kostylevite, K4ZrzSi60IS·H20, a new mineral.
Zap. Vsesoy. Mineral. Obshcll. 112,469-474 (Russian).
Khomyakov, A. P., Voronkov, A. A., Kobyashev, Y S., Polezhaeva, 1. I. 1983a. Umbite and parumbite, new
potassium zirconosilicates from the Khibina alkaline massif. Zap. Vsesoy. Mineral. Obshch. 112,461-469
(Russian).
Kinny, P. D., Dawson, 1. B. 1992. A mantle metasomatic injection event linked to late Cretaceous kimberlite
magmatism. Nature 360, 726-728.
Kirkely, M. B., Gurney, 1. J., Levinson, A. A. 1991. Age, origin and emplacement of diamonds. Gems Gemol.
27,2-25.
Kirkley, M. B., Smith, H. B., Gurney, J. J. 1989. Kimberlite carbonates-a carbon and oxygen stable isotope
study, In Ross et 01. (1989) q.v., Vol. I, pp. 264-281.
Kobelski, B. J., Gold, D. P., Deines, P. 1979. Variations in stable isotope compositions for carbon and oxygen
in some South African and Lesothan kimberlites. In Boyd and Meyer (1979) q.v., Vol. I, pp. 252-271.
Kornilova. V. P., Nikishov, K. N., Kovalskii, V. V. 1983. Atlas of the Textures and Structures of Kimberlitic
Rocks. lzdat. Nauka, Moscow (Russian).
Kornprobst, 1. (Ed.) 1984. Proc. Third Int. Kimberlite Conf. Vol. I. Kimberlites I: Kimberlites and Related
Rocks. Vol. 2. Kimberlites /I: The Mantle and Crust-Mantle Relationships. Developments in Petrology
11 A and lIB, Elsevier, New York.
Kostyuk, V. P., Panina, 1. I., Zhidkoy, A. y, Orlova, M. P., Bazarova, T. Y. 1990. Potassic Alkaline Magmatism
of tile Baikal-Stanovoy Rifting System. Nauka, Novosibirsk (Russian).
Kovalskii, V. V. 1963. Kimberlitic Rocks of Yak uti a and Some Principles ofTlleir Petrographic Classification
lzdat. Nauka, Moscow (Russian).
Kovalskii, V. v., Nikishov, K. N., Egorov, O. S. 1969. Kimberlitic and Carbonatitic Occllrrences in the Eastern
and Southeastern Parts of the AnabarAlltec/ise. Nauka, Moscow (Russian).
Kramers, J. D. 1977. Lead and strontium isotopes in Cretaceous kimberlites and mantle-deri ved xenoli ths from
southern Africa. Earth Planet. Sci. Lett. 34, 419-431.
Kramers, 1. D. 1979. Pb, U, Sr, K, and Rb in inclusion-bearing diamonds and mantle-derived xenoliths from
southern Africa. Earth Planet. Sci. Lett. 42, 58-70.
Kramers, 1. D., Smith, C. B., Lock, N., Harmon, R. S., Boyd, F. R. 1981. Can kimberlites be generated from
"ordinary mantle"? A trace element point of view. Nature 291, 53-56.
Kresten, P., Dempster, A. N. 1973. The geology of Pipe 200 and the Malibamatso dike swarm. In Nixon, P. H.
(ed.), Lesotho Kimberlites. Lesotho National Development Corporation, Maseru, Lesotho, pp. 172-179.
Kriuchkov, A. I., Leliukh, M. I., Krasinets, S. S., Afanasiev, V. P. 1994. Two unusual Paleozoic kimberlite
diatremes in the Daldyn-Alkait region of the Siberian platform. In Meyer and Leonardos (1994) q.v., Vol.
1, pp. 34-39.
Kudrjavtseva, G. P., Busheva, E. B., Vasiljeva, E. R., Verichev, E. M., Garanin, V. K., Grib, V. P., Laverova, T.
N., Mikhailchenko, O. A., Posukhova, T. v., Schepina, N. A. 1991. Geological structure and mineralogy
ofkimberlites of the Archangelskkimberlite province. Fifth Int. Kimberlite Conf., Araxa, Brazil. Extended
Abstracts, Companhia de Pesquisa de Recursos Minerais, Spec. Pub!. 2/91, pp. 530-531.
Kuehner, M., Edgar, A. D., Arima, M. 1981. Petrogenesis of the ultrapotassic rocks from the Leucite Hills,
Wyoming. Am. Mineral. 66, 663-677.
Lange, R. A., Carmichael, I. S. E. 1991. A potassic volcanic front in western Mexico: the lamprophyric and
related lavas of San Sebastian. Geol. Soc. Am. Bull. 103, 928-940.
Larsen, E. S. 1940. Petrographic province of central Montana. Geol. Soc. Am. Bull. 51, 887-948
Larsen, L. M., Rex, D. C. 1992. A review of the 2500 Ma span of alkaline-ultramafic, potassic and carbonatitic
magmatism in West Greenland. Lit/lOS 28, 367-402.
Leblanc, M., Dautria, J., Girod, M. 1982. Magnesian ilmenite xenoliths in a basanite from Tahalra, Ahaggar
(Southern Algeria). Comrib. Mineral. Petrol. 79, 347-354.
LeCheminant, A. N., Miller, A. R., LeCheminant, G. M. 1987. Early Proterozoic alkaline igneous rocks, District
of Keewatin, Canada: Petrogenesis and mineralization. In Pharaoh, T. C., Beckinsale, R. D, and Rickard,
D. (eds), Geochemistry and Mineralization of Proterozoic Volcanic Suites. Geo!. Soc. London Spec. Pub!.
No. 33, pp. 219-240.
Lehnert-Thiel, K., Loewer, R., Orr, R. G., Robertshaw, P. 1992. Diamond-bearing kimberlites in Saskatchewan,
Canada: The Fort a la Corne case history. Explor. Mining Geol. 1,391-403.
REFERENCE'S 397

Lenzen, G. 1980. South Africa and the birth of a great industry. In J. Legrand (ed.), Diamonds: Myth, Magic
and Reality. Crown, New York, pp. 66-85.
Ie Roex, A. P. 1986. Geochemical correlations between southernAfricankimberlites and south Atlantic hotspots.
Nature 324, 243-245.
Lewis, H. C. 1887. On diamantiferous peridotite and the genesis of the diamond. Geol. Mag. 4, 22-24.
Lewis, H. C. 1888. The matrix of diamond. Geol. Mag. 5,129-131.
Livi, K. J. T., Veblen, D. R. 1987. "Eastonite" from Easton, Pennsylvania: A mixture of phlogopite and a new
form of serpentine. Am. Mineral. 72, 113-125.
Lloyd, F. E., Arima, M., Edgar, A. D. 1985. Partial melting of a phlogopite-clinopyroxenite nodule from
south-west Uganda: An experimental study bearing on the origin of highly potassic continental rift
volcanics. Comrib. Mineral. Petrol. 91, 321-329.
Lorenz, V. 1975. Formation of phreatomagmatic maar-diatreme volcanoes and its relevance to kimberlite
diatremes. Phys. Chem. Earth. 9,17-27.
Lorenz, V. 1979. Phreatomagmatic origin of the olivine melilitite diatremes in the Swabian Alb, Germany. In
Boyd and Meyer (1979) q.v., Vol. 1, pp. 354-363.
Lorenz, V. 1984. Explosive volcanism of the West Eifel volcanic field, Germany. In Kornprobst (1984) q.v.,
Vol. 1, pp. 299-307.
Luhr, J. F., Carmichael, I. S. E. 1981. The Colima volcanic complex, Mexico: Part n. Late-Quaternary cinder
cones. Comrib. Mineral. Petrol. 76, 127-147.
Luth, R. W., Trl'lnnes, R. G., Canil, D. 1993. Volatile phases in the Earth's mantle. In Luth, R.W. (ed.),
Experimems at High Pressure and Applications to the Earth's Mamie. Mineral. Assoc. Canada Short
Course Handbook 21, pp. 445-485.
Luth, W. C. 1967. Studies in the system KAISi04-Mg2Si04-Si02-H20 I. Inferred phase relations and petrologic
applications. J. Petrol. 8, 372-416.
MacDonald, R., Upton, B. G., Collerson, K. D., Hearn, B. c., James, D. 1992. Potassic mafic lavas of the
Bearpaw Mountains, Montana: mineralogy, chemistry and origin. J. Petrol. 33, 305-346.
MacGregor, I. D. 1970. A hypothesis for the origin of kimberlite. Mineral. Soc. Am. Spec. Paper 3,51-62.
MacGregor, I. D., Carter, J. L. 1970. The chemistry of clinopyroxenes and garnets of eclogite and peridotite
xenoliths from the Roberts Victor mine, South Africa. Phys. Earth Planet. Imer. 3, 391-397.
MacGregor, I. D., Manton, w.1. 1986. The Roberts Victor eclogites: Ancient oceanic crust. J. Geophy. Res. 91,
14063-14079.
Mahoney, J. J., Natland, 1. H., White, W. M., Poreda, R., Bloomer, S. M., Fisher, R. L., Baxter, A. N. 1989.
Isotopic and geochemical provinces of the western Indian Ocean spreading centers. J. Geophys. Res. 94,
4033-4052.
Mannard, G. W. 1962. The Singida kimberlite pipes, Tanganyika. Ph.D. thesis, McGill University, Montreal,
Canada.
Mason, B. 1977. Elemental distribution in minerals from the Wolgidee Hills intrusion, Western Australia. N.
Z. Dept. Sci. Ind. Res. Bull. 218, 114-120.
McCallum, M. E. 1989. Oxide minerals in the Chicken Park kimberlite, northern Colorado. In Ross et al. (1989)
q.v., Vol. I, pp. 242-263.
McCallum, M. E., Eggler, D. H., Burns, L. K. 1975. Kimberlite diatremes in northern Colorado and southern
Wyoming. Ph),. Chern. Earth 9, 149-162.
McCandless, T. E., Waldman, M. A., Gurney, J. 1. 1994. Macrodiamonds, microdiamonds from Murfreesboro
lamproites: morphology, inclusions, carbon isotope geochemistry. In Meyer and Leonardos (1994) q.v.,
Vol. 2. pp. 78-97.
McCulloch. M. T., Jaques, A. L., Nelson, D. R., Lewis, 1. D. 1983. Nd and Sr isotopes in kimberiites and
lamproites from Western Australia: An enriched mantle origin. Nature 302, 400-403.
McGetchin, T. R. 1968. The Moses Rock Dike: Geology, petrology and mode of emplacement of kimberlite-
bearing breccia dike, San Juan County, Utah. Ph.D. thesis, Califomia Institute of Technology, Pasadena,
CA.
McGetchin, T. R., Besancon, J. R. 1973. Carbonate inclusions in mantle-derived pyropes. &lrth Planet. Sci.
Lett. 18,408-410.
McGetchin, T. R., Nikhanj, Y. S., Chodos, A. A. 1973. Carbonatite-kimberlite relations in the Cane Valley
diatreme, San Juan County, Utah. J. Geoph),s. Res. 78, 1854-1869.
398 REFERENCES

McIver, J. R. 1981. Aspects of ultrabasic and basic alkaline intrusive rocks from Bittefontein, South Africa.
Contrib. Mineral. Petrol. 78,1-11.
McIver, J. R., Ferguson, J. 1979. Kimberlitic, melilititic, trachytic and carbonatite eruptives at Saltpetre Kop,
Sutherland, South Africa. In Boyd and Meyer (1979) q.v., Vol. \, pp. 1\1-128.
McKenzie, D. 1985. The extraction of magma from the crust and mantle. Earth Planet. Sci. Lett. 74, 81-91.
McKenzie, D. 1989. Some remarks on the movement of small melt fractions in the mantle. Earth Planet. Sci.
Lett. 95, 53-72.
McKenzie, D., Bickle, M. J. 1988. The volume and composition of melt generated by extension of the
lithosphere. J. Petrol. 29, 625-679.
McKenzie, D., O'Nions, R. K. 1991. Partial melt distributions from inversion of rare earth element concentra-
tions. J. Petrol. 32,1021-1091.
McPhie, 1., Doyle, M., Allen, R. 1993. Volcanic Textures. Key Centre for Ore Deposit and Exploration Studies,
University of Tasmania, Hobart, Tasmania, 196 pp.
Menzies, M. A., Rogers, N., Tindle, A., Hawkesworth, C. 1. 1987. Metasomatic and enrichment processes in
lithospheric peridotites, an effect of asthenosphere-lithosphere interaction. In Menzies, M. A., and
Hawkesworth, C. J. (eds.), Mantle Metasomatism. Academic Press, London, pp. 313-361.
Meyer, H. O. A. 1985. Genesis of diamond: A mantle saga. Am. Mineral. 70, 344-355.
Meyer, H. O. A. 1987. Inclusions in diamonds. In Nixon (1987) q.v., pp. 501-522.
Meyer, H. O. A., Leonardos, O. H. 1994. Proc. Fifth Int. Kimberlite Conf. Vol. 1: Kimberlites alld Related Rocks
and Mantle Xenoliths. Vol. 2: Diamonds: Characterization, Genesis, and Exploration. Companhia de
Pesquisa de Recursos Minerais, Spec. Publ. IIA and lIB, Brasilia, Brazil.
Meyer, H. O. A., Mitchell, R. H., Jayaganapathy, S. 1994. Phlogopite in calc-alkaline lamprophyres of Northem
England. Mineral. Petrol. 51, 227-237.
Middlemost, E. A. K., Paul, D. K., Fletcher, I. R. 1988. Geochemistry of the minette-Iamproite association from
the Indian Gondwanas. Lithos 22, 31-42.
Mikheyenko, V. I. 1977. Evidence for the sedimentary origin of kimberlite. Dokl. Akad. Nauk SSSR 237,
198-201.
Milashev, V. A. 1963. The term kimberlite and the classification of kimberli tic rocks. Geol. Geophys. 4, 42-52
(Russian).
Milashev, V. A. 1984. Explosion Pipes. Springer-Verlag, New York.
Milashev, V. A., Krutoyarskii, M. A., Rabkin, M. I., Erlikh, E. N. 1963. Kimberlitic Rocks and Picrite Porphyries
o/the N0/1i1-eastern Part o/the Siberian Platform. Gosgeoltekhizdat, Moscow, 216 pp. (Russian)
Milledge, H. Joo Mendelssohn, M. J., Seal, Moo Rouse, J. Eoo Swart, K., Pillinger, C. T. 1983. Carbon isotopic
variations in spectral type II diamonds. Nature 303, 791-792.
Millett, J. M., Roth, R. S., Ettlinger, L. D., Parker, H. S. 1987. Phase equilibria and crystal chemistry in the
ternary system BaO-Ti02-Nb20S. J. Solid State Chem. 67, 259-270.
Mitchell, R. H. 1970. Kimberlites and related rocks-A critical reappraisal. J. Geol. 78, 686-704.
Mitchell, R. H. 1978a. Mineralogy of the Elwin Bay kimberlite, Somerset Island, N.W.T, Canada. Am. Mineral.
63,47-57.
Mitchell, R. H. 1978b. Manganoan magnesian ilmenite and titanian clinohumite from the Jacupiranga carbona-
tite, Sao Paulo, Brazil. Am. Mineral. 63, 544-547.
Mitchell, R. H. 1979. Mineralogy of the Tunraq kimberlite, Somerset Island, N.W.T., Canada. In Boyd and
Meyer (1979) q.v., Vol. 1, pp. 161-171.
Mitchell, R. H. 1983. The lie Bizard intrusion, Montreal, Quebec-kimberlite or lamprophyre?: Discussion.
Can. J. Earth Sci. 20, 1493-1496.
Mitchell, R. H. 1984a. Mineralogy and origin of carbonate-rich segregations in a composite kimberlite sill.
Neues Jahr. Mineral. Abh. 150, 185-197.
Mitchell, R. H. 1984b. Gamet Iherzolites from the Hanaus-I and Louwrensia kimber1ites of Namibia. Contrib.
Mineral. Petrol. 86,178-188.
Mitchell, R. H. 1985. A review of the mineralogy of lamproites. Trans. Geol. Soc. S.A/r. 88, 411-437.
Mitchell, R. H. 1986. Kimberlites: Mineralogy, Geochemistry, and Petrology. Plenum Press, New York.
Mitchell, R. H. 1987. Megacrysts in kimberlites from the Gibeon field, Namibia. Neues Jahr. Mineral. Abh.
157, 2657-283.·
REFERENCES 399

Mitchell. R. H. 1989. Aspects of the petrology ofkimberlites and lamproites: Some definitions and distinctions.
In Ross et al. (1989) q.v.• Vol. I. pp. 7-45.
Mitchell. R. H. 1991a. What's in a name? Suggestions for revisions to the terminology of kimberlites and
lamprophyres from a genetic viewpoint. Fifth Int. Kimberlite Conf.• Araxa. Brazil. Extended Abstracts.
Companhia de Pesquisa de Recursos Minerais. Spec. Publ. 2/91. 295-297.
Mitchell. R. H. 1991b. Kimberlites and lamproites: primary sources of diamond. Geosci. Can. 18. 1-16.
Mitchell. R. H. 1994a. Suggestions for revisions to the terminology of kimberlites and lamprophyres from a
genetic viewpoint. In Meyer and Leonardos (1994) q.v.• Vol. 1. pp. 15-26.
Mitchell. R. H. 1994b. Accessory rare earth. strontium. barium and zirconium minerals in the Benfontein and
Wesselton calcite kimberlites. South Africa. In Meyer and Leonardos (1994) q.v.• Vol. 1. pp. 115-128.
Mitchell. R. H. 1994c. The lamprophyre facies. Mineral. Petrol. 51. 137-146.
Mitchell. R. H. 1995. The role of petrography and lithogeochemistry in exploration for diamondiferous rocks.
J. Geochem. Explor.• in press.
Mitchell. R. H.• Bergman. S. C. 1991. Petrology of Lamproites. Plenum Press. New York.
Mitchell. R. H.• Brunfel!, A. 0.1975. Rare earth geochemistry of kimberlite. Phys. Chem. Earth 9. 671-686.
Mitchell. R. H.• Carswell. D. A. 1976. Lanthanum. samarium and ytterbium abundances in some southern
African garnet Iherzolites. Eanh Planet. Sci. Lerr. 31. 175-178.
Mitchell. R.. Carswell. D. A.• Clarke. D. B. 1980. Geological implications and validity of calculated equilibra-
tion conditions for ultramafic xenoliths from the Pipe 200 kimberlite. Northern Lesotho. Contrib. Mineral.
Petrol. 72. 205-217.
Mitchell. R. H.. Crocket. J. H. 1971. The isotopic composition of strontium in some South African kimberlites.
Contrib. Mineral. Petrol. 30. 277-290.
Mitchell. R. H.. Haggerty. S. E. 1986. Anew K-V-Ba titanate related toprideritefrom the New Elands kimberlite.
South Africa. Neues Jarh. Mineral. Monatsil. 1986. 376-384.
Mitchell. R. H.• Meyer. H. O. A. 1980. Mineralogy of micaceous kimberlite from the Jos dike. Somerset Island.
N.W.T.• Canada. Can. Mineral. 18. 241-250.
Mitchell. R. H.• Meyer. H. O. A. 1989a. Mineralogy of micaceous kimberlites from the New Elands and Star
Mines. Orange Free State. South Africa. In Ross et al. (1989) q.v.• Vol. I. pp. 83-96.
Mitchell, R. H.• Meyer. H. O. A. I 989b. Niobian K-Ba-V titanates from micaceous kimberlite. Star Mine.
Orange Free State. South Africa. Mineral. Mag. 53. 451-456.
Mitchell. R. H.• Platt. R. G.• Downey. M. 1987. Petrology oflamproites from Smoky Butte. Montana. J. Petrol.
28.645-677.
Mitchell. R. H.• Putnis. A. 1988. Polygonal serpentine in segregation-textured kimberlite. Can. Mineral. 26.
991-997.
Mitchell. R. H.• Reed. S. 1. B. 1988. Ion microprobe determinations of rare earth elements in perovskites from
kimberlites and alnoites. Mineral. Mag. 52. 331-339.
Mitchell, R. H., Smith, C. B., V1adykin, N. V. 1994. Isotopic composition of strontium and neodymium in
potassic rocks of the Little Murun complex. Aldan Shield. Siberia. Lithos 32. 243-248.
Mitchell. R. H.• Steele. 1. 1992. Potassian zirconium and titanium silicates and strontian cerian perovskite in
lamproites from the Leucite Hills. Wyoming. Can. Mineral. 30.1153-1159.
Mitchell. R. H.• Vladykin. N. V. 1993. Rare earth element-bearing tausonite and potassium barium titanates
from the Little Murun potassic alkaline complex. Yakutia. Russia. Mineral. Mag. 57. 651-664.
Moore. A. E. 1988. Olivine: A monitor of magma evolutionary paths in kimberlites and olivine melilitites.
Contrib. Mineral. Petrol. 99. 238-248.
Moore. A. E.• Erlank. J. 1979. Unusual oli vine zoning-evidence for complex physico-chemical changes during
the evolution of olivine melilitite and kimberlite magmas. Contrib. Mineral. Petrol. 70. 391-405.
Moore. R. 0 .• Gurney. 1. 1. 1985. Pyroxene solid solution in garnets included in diamond. Nature 318.553-555.
Moore. R. 0 .• Gurney. J. J. 1989. Mineral inclusions in diamonds from the Monastery kimberlite. South Africa.
In Ross et al. (1989) q.v.• Vol. 2. pp. 1029-1041.
Moore. R. 0 .• Gurney. J. J. 1991. Gamet megacrysts from Group II kimberlites in southern Africa. Fifth Int.
Kimberlite Conf.• Araxa. Brazil. Extended Abstracts. Companhiade Pesquisade Recursos Minerais. Spec.
Pub\. 2191. pp. 298-300.
Moore, R. 0., Gurney, J. 1., Griffin, W. L.. Shimizu, N. 1991. Ultra-high pressure garnet inclusions in Monastery
diamonds: Trace element abundance patterns and conditions of origin. Eur. J. Mineral. 3, 213-230.
400 REFERENCES

Morgan, W. J. 1971. Convection plumes in the lower mantle. Nature 230, 42-43.
MUller, D., Morris, B. J., Farrand, M. G. 1993. Potassic alkaline lamprophyres with affinities to lamproites from
the Karinya Syncline, South Australia. Lithos 30,123-137.
Muramatsu, Y. 1983. Geochemical investigations of kimberlites from the Kimberley area, South Africa.
Geochem. J. 17,71-86.
Murayama, J. K., Nakai, S., Kato, M., Kumazawa, M. 1986. A dense polymorph of Ca3(P04)2: A high pressure
phase of apatite decomposition and its geochemical significance. Phys. Eanh Planet. Inter. 44, 293-303.
Murthy, D. S. N., Dayal, A. M., Natarajan, R. 1994. Mineralogy and geochemistry ofChigicherla kimberlite
and its xenoliths, Anantapur district, South India. J. Geol. Soc. India 43, 329-341.
Myhra, S., White, T. J., Kesson, S. E., Riviere, C. J. 1988. X-ray photoelectron spectroscopy for direct
identification ofTi valence in [BaxCsyl[(Ti,AI)t+)'Ti~\\,_yl016. Am. Mineral. 73,161-167.
Neal, C. R., Taylor, L. A. 1989. The petrography and composition of phlogopite micas from the Blue Ball
kimberlite, Arkansas: A record of chemical evolution during crystallization. Mineral. Petrol. 40, 207-224.
Neal, C. R., Taylor, L. A., Davidson, J. P., Holden, P., Halliday, A. N., Nixon, P. H., Paces, J. B., Clayton,
R. N., Mayeda, T. K. 1990. Eclogites with oceanic crustal and mantle signatures from the Bellsbank
kimberlite, South Africa, Part 2: Sr, Nd, and 0 isotope geochemistry. Earth Planet. Sci. Lett. 99, 362-379.
Nelson, D. R. 1989. Isotopic characteristics and petrogenesis of the lamproites and kimberlites of central West
Greenland. Lithos 22, 265-274.
Nelson, D. R., McCulloch, M. T., Sun, S. S. 1986. The origins of ultrapotassic rocks as inferred from Sr, Nd,
and Pb isotopes. Geochim. Cosmochim. Acta 50, 231-245.
Nemec, D. 1988. The amphiboles of potassium-rich dyke rocks of the southeastern border of the Bohemian
massif. Call. Mineral. 26, 89-95.
Nisbet, E. G., Malley, D. P., Lowry, D. 1994. Can diamonds be dead bacteria? Nature 367,694.
Nixon, P. H. 1987. Mantle Xenoliths. Wiley, New York.
Nixon, P. H., Boyd, F. 1973. The discrete nodule association in kimberlites from northern Lesotho. In Nixon,
P.H. (ed.), Lesotho Kimberlites. Lesotho National Development Corporation, Maseru, Lesotho, pp. 67-75.
Nixon, P. H., Boyd, F. R. 1979. Gamet-bearing Iherzolites and discrete nodule suites from the Malaita alntiite,
Solomon Islands, S.W. Pacific, and their bearing on oceanic composition and geotherm. In Boyd and
Meyer (1979) q.v., Vol. 2, pp. 40-423.
Nixon, P. H., van Calsteren, P. W. C., Boyd, F. R., Hawkesworth, C. J. 1987. Harzburgites of the diamond facies
from southern African kimberlites. In Nixon (1987) q.v., pp. 523-533.
Nixon, P. H., Mitchell, R. H., Rogers, N. W. 1980. Petrogenesis of alnoitic rocks from Malaita, Solomon Islands.
Melanesia. Mineral. Mag. 43, 587-596.
Norrish, K. 1951. Priderite, a new mineral from the leucite lamproites of the West Kimberley area, Western
Australia. Mineral. Mag. 24, 496-501.
O'Brien, H. E., Irving, A. J., McCallum, I. S. 1988. Complex zoning and resorption of phenocrysts in mixed
potassic mafic magmas of the Highwood Mountains, Montana. Am. Mineral. 73, 1007-1024.
O'Brien, H. E., Irving, A. J., McCallum, I. S. 1991. J. Geophys. Res. 96B, 13237-13260.
O'Hara, M. J .• Yoder, H. S. 1967. Formation and fractionation of basic magmas at high pressures. Scot. J. Geol.
3,67-117.
O'Nions. R. K., Carter, S., Evensen, N. H.• Hamilton, P. J. 1979. Geochemical and cosmochemical applications
of Nd isotope analysis. Ann. Rev. Earth Planet. Sci. 7, 11-38.
Orlova, M. P. 1988. Recent findings on the geology of the Malo Murun alkaline pluton. Int. Geol. Rev. 30,
945-953.
Parfenoff, A. 1982. Une minerale traceur pour la prospection alluvionaire: L'ilmenite. Relations entre ilmenites
magnesieanes, basaltes alcalins, kimberlites et diamant. BRGM. Documents Burueau Recherche Geologie
et Minieres No. 37.
Pasteris, J. D. 1980. Opaque oxide phases of the De Beers Pipe kimberlite (Kimberley, South Africa) and their
petrologic significance. Ph.D. thesis, Yale University.
Pasteris. J. D. 1983. Spinel zonation in the De Beers kimberlite, South Africa: Possible role of phlogopite. Call.
Mineral. 21, 41-58.
Pasteris, J. D., Boyd, F. R., Nixon, P. H. 1979. The ilmenite association at the Frank Smith Mine, R.S.A. In
Boyd and Meyer (1979) q.v., Vol. 2, pp. 257-264.
REFERENCES 401

Paul, D. K., Gale, N. H., Harris, P. G. 1977. Uranium and thorium abundances in Indian kimberlites. Geochim.
Cosmochim. Acta 41, 335-339.
Paul, D. K., Potts, P. 1., Gibson, G. L., Harris, P. G. 1975. Rare earth abundance in Indian kimberlites. Earth
Planet. Sci. Lett. 15, 151-158.
Peccerillo, A., Poli, G., Serri, G. 1988. Petrogenesis of orenditic and kamafugitic rocks from central Italy. Can.
Mineral. 26, 45--65.
Peterson, T. 1994. Early Proterozoic ultrapotassic volcanism of the Keewatin hinterland, Canada. In Meyer and
Leonardos (1994) q.v., Vol. I, pp. 221-235.
Platt, R. G. 1994. Perovskite, loparite and Ba-Fe hollandite from the Schryburt Lake carbonatite complex,
northwestern Ontario, Canada. Mineral. Mag. 58,49-57.
Platt, R. G., Mitchell, R. H. 1979. The Marathon dikes. I: Zirconium-rich titanium garnets and manganoan
magnesian ulvospinel magnetite spinels. Am. Mineral. 64, 546-550.
Pokhilenko, N., Sobolev, N. v., Lavrentyev, Y. G. 1977. Xenoliths of diamondiferous ultramafic rocks from
Yakutian kimberlites. Proc. Second Int. Kimberlite Conf., Santa Fe, New Mexico, Extended Abstract.
Pollack, H. N. 1986. Cratonization and thermal evolution of the mantle. Earth Planet. Sci. Lett. 80,175-182.
Pring, A., Jefferson, D. A. 1983. Incommensurate superlattice ordering in priderite. Mineral. Mag. 47, 65--68.
Rabkin, M. I., Krutoyarskii, M. A., Milashev, V. A. 1961. Classification of Kimberlitic Rocks of Yakutia and
their Nomenclature. Trans. Inst. Arctic Geol., Yakutsk No. 121,215 pp. (Russian).
Reed, L. E., Sinclair, I. G. L. 1991. The search for kimberlite in the James Bay Lowlands of Ontario. Can. Inst.
Mining Bull. 84, 132-139.
Richards, M. A., Duncan, R. A., Courtillot, V. E. 1989. Flood basalts and hot-spot tracks: Plume heads and tails.
Science 246, 103-107.
Richardson, S. H., Gurney, 1. J., Ehrlank, A. J., Harris, J. W. 1984. Origin of diamonds in old enriched mantle.
Nature 310, 198-202.
Ridley, W. I. 1977. Crystallization trends of spinels in Tertiary basalts from Rhum and Muck, and their
petrogenetic significance. Contrib. Mineral. Petrol. 64, 243-255.
Ringwood, A. E. 1967. Synthesis of rnajorite and other high pressure gamets. Earth Planet. Sci. Lett. 12,
411-418.
Ringwood, A. E. 1989. Constitution and evolution of the mantle. In Rossetal. (1989) q.v., Vol. I, pp. 457-485.
Ringwood, A. E., Kesson, S. E., Hibberson, w., Ware, N. 1992. Origin ofkimberlites and related magmas. Earth
Planet. Sci. Lett. 113, 521-538.
Ritter, J. 1., Roth, R. S., Blendell, J. E. 1986. Alkoxide precursor synthesis and characterization of phases in the
barium-titanium oxide system. 1. Am. Ceram. Soc. 69, 155-162.
Robert, 1. L. 1981. Etudes cristallochimiques sur les micas et les amphiboles. Applications a la petrographie et
ala geochimie. These de Doctorate d'Etat, Univ. Paris-Sud, Orsay.
Roberts, B. 1976 Kimberley: Turbulent City. David Philip, Cape Town, South Africa.
Robey, J. V. A. 1981. Kimberlites of the Central Cape Province, R.S.A., Ph.D. thesis (2 vols), University of
Cape Town, South Africa.
Rock, N. M. S. 1986. The nature and origin of lamprophyres: Alniiites and allied rocks (ultramafic lampro-
phyres). 1. Petrol. 27, 193-227.
Rock, N. M. S. 1990. Lamprophyres. Blackie & Son, Edinburgh.
Rogers, N. w., Hawkesworth, C. J., Palacz, Z. A. 1992. Phlogopite in the generation of olivine-melilitites from
Namaqualand, South Africa and implications for element fractionation processes in the upper mantle.
Lithos 28, 347-365.
Rolf, D. G. 1973. The geology of the Kao pipes. In Nixon, P. H. (ed.), Lesotho Kimberlites. Lesotho National
Development Corporation, Maseru, Lesotho, pp. 101-106.
Ross, J., Jaques, A. L., Ferguson, 1., Green, D. H., O'Reilly, S. Y., Danchin, R. V., Janse, A. J. A. (Eds.) 1989.
Proc. Fourth Int. Kimberlite Conf. Kimber/ites and Related Rocks. Vol 1. Their Composition, Occurrence,
Origin and Emplacement. Vol 2. Their Mantle/Crust Setting, Diamonds and Diamond Exploration. Geol.
Soc. Austral. Spec. Publ. No. 14.
Rozova, Y. V., Frantsesson, E. v., Pleshakov, A. P., Botova, M. M., Filipova. L. P. 1982. High-iron chrome
spinels in kimberlites ofYakutia. Int. Geol. Rev. 24, 1417-1425.
Ruotsala, A. P. 1975. Alteration of the Finsch kimberlite pipe, South Africa. £Con. Geol. 70, 587-590.
Rust, G. W. 1937. Preliminary notes on explosive volcanism in Missouri. 1. Geol. 45, 48-75.
402 REFERENCES

Salpas, P. A., Taylor, L. A., Shervais, J. W. 1986. The Blue Ball, Arkansas kimberlite: Mineralogy, petrology,
and geochemistry. J. Geol. 94,891-901.
Saunders, A. D., Storey, M., Kent, R w., Norry, M. 1. 1992. Consequences of plume-lithosphere interactions.
In Storey et al. (1992) q.v., pp. 41--{j0.
Sautter, V. Haggerty, S. E., Field, S. 1991. Ultra-deep (>300 km) ultramafic xenoliths: New petrologic evidence
from the transition zone. Science 252, 827-830.
Schulze, D. J. 1986. Calcium anomalies in the mantle and a subducted metaserpentinite origin for diamonds.
Nature 319, 483-485.
Schulze, D. J. 1987. Megacrysts from alkaline rocks. In Nixon (1987) q.v., pp. 433-451.
Schulze, D. J. 1989. Green garnets from South African kimberlites and their relationship to wehrlites and crustal
uvarovites. In Ross et al. (1989) q.v., Vol. 2, pp. 820-826.
Schulze, D. J., Smith, 1. v., Nemec, D. 1985. Mica chemistry of lamprophyres from the Bohemian Massif,
Czechoslovakia. Neues Jahr. Mineral. Monat. Abh. 152321-334.
Scott, B. H. 1981. Kimberlite and lamproite dykes from Holsteinsborg, West Greenland. Meddelelser Gronland
Geosci. 4, 3-24.
Scott Smith, B. H. 1989. Lamproites and kimberlites in India. Neues Jahr. Mineral. Abh. 161, 193-225.
Scott Smith, B. H. 1992. Contrasting kimberlites and lamproites. Explor. Mining Geol. I, 371-382.
Scott Smith, B. H., Danchin, R V., Harris, J. W., Stracke, K. J. 1984. Kimberlites near Orroroo, South Australia.
In Kornprobst (1984) q.v., Vol. 1, pp. 121-142.
Scott, B. H., Skinner, E. M. W. 1979. The Premier kimberlite pipe, Transvaal, South Africa. Kimberlite Symp.
II Cambridge, Extended Abstract.
Scott Smith, B. H., Skinner, E. M. w., Clement, C. R 1983. Further data on the occurrence of pectolite in
kimberlite. Geol. Mag. 47, 75-78.
Scott Smith, B. H., Skinner, E. M. W. 1984. A new look at Prairie Creek, Arkansas. In Kornprobst (1984) q.v.,
Vol. 1, pp. 255-283.
Scott Smith, B. H., Skinner, E. M. W., Loney, P. E. 1989. The Kapambalamproites of the Luanga valley, Eastern
Zambia. In Ross et al. (1989) q.v., Vol. I, pp. 189-205.
Scott, 1. D., Peatfield, G. R. 1986. [Ba, H201Ti6V23+016, a new mineral species and new data on redledgeite.
Can. Mineral. 24, 55--{j6.
Sekine, T., Wyllie, P. 1. 1982. Phase relationships in the system KAISi04-Mg2Si04-Si02-H20 as a model for
hybridization between hydrous siliceous melts and peridotite. COllfrib. Mineral. Petrol. 79, 368-374.
Sgarbi, P. B. A., Valenca, 1. G. 1991. Petrography and general chemical features of potassic to ultramafic alkaline
volcanic rocks of Mat ada Corda fonnation, Minas Gerais State, Brazil. Fifth In!. Kimberlite Conf., Araxa,
Brazil, Extended Abstracts. Companhia de Pesquisa de Recursos Minerais, Spec. Publ. 2/91, pp. 359-360.
Shadenkov, Y. M., Orlova, M. P., Borisov, A. B. 1989. Pyroxenites and shonkinites of the Little Murun
massif-intrusive analogues of lamproite. lilt. Geol. Rev. 32, 61-69.
Shee, S. R 1984. The oxide minerals of the Wesselton Mine kimberlite, Kimberley, South Africa. In Kornprobst
(1984) q.v., Vol. 1, pp. 59-73.
Shee, S. R 1985. The petrogenesis of the Wesselton Mine kimberlites, Kimberley, Cape Province, RS.A. Ph.D.
thesis, University of Cape Town, South Africa.
Shee, S. R, Bristow, 1. w., Bell, D. R., Smith, C. B., Allsopp, H. L., Shee, P. B. 1989. The petrology ofkimberlites
related rocks and associated mantle xenoliths from the Kuruman province, South Africa. In Ross et al.
(1989) q.v., Vol. I, pp. 60-82.
Shee, S. R., Gurney, 1. J., Robinson, D. N. 1982. Two diamond-bearing peridotite xenoliths from the Finsch
kimberlite. Collfrib. Mineral. Petrol. 81, 79-87.
Sheppard, S. M. F., Dawson, 1. B. 1975. Hydrogen, carbon and oxygen isotope studies of megacryst and matrix
minerals from Lesothan and South African kimberlites. Phys. Chern. Earth 9, 747-763.
Sheppard, S., Taylor, W. R 1992. Barium and LREE-rich, olivine-mica-Iamprophyre with affinities to
lamproites, Mt. Bundey, Northern Territory, Australia. Lithos 28,303-325.
Sheridan, M. F., Wohletz, K. H. 1983. Hydrovo\canism: Basic considerations and review. J. Volcanol. Geotherm.
Res. 17, 1-29.
Shervais, J. S., Taylor, L. A., Lugmair, G., Clayton, R., Mayeda. T.• Korotev. R 1988. Archean oceanic crust
and the evolution of subcontinental mantle: Eclogites from southern Africa. Geol. Soc. Am. Bull. 100,
411-423.
REFERENCES 403

Singewald, J. T., Milton, C. 1930. An alnoite pipe, its contact phenomena and ore deposition near Avon Missouri.
J. Geol. 38, 54-66.
Sinitsyn, A. v., Dauev, Y. M., Gribb, V. P. 1992. Structural setting and productivity of the kimberlites of the
Arkhangelsk province. Russ. Geol. Geophys. 33, 61-70.
Skinner, E. M. W. 1986. Contrasting Group 1 and 2 kimberlite petrology: Towards a genetic model for
kimberlites. Fourth Int. Kimberlite Conr., Perth, Australia. Extended Abstracts, pp. 202-204.
Skinner, E. M. W. 1989. Contrasting group I and II kimberlite petrology: Towards a genetic model for
kimberlites. In Ross et al. (1989) q.v., Vol. I, pp. 528-544.
Skinner, E. M. W., Clement, C. R. 1979. Mineralogical classification of southern African kimberlites. In Boyd
and Meyer (1979) q.v., Vol. I, 129-139.
Skinner E. M. W., Clement, C. R., Gurney, J. J., Apter, D. B., Hatton, C. J. 1992. The distribution and tectonic
setting of South African kimberlites. Russ. Geol. Geoph)'s. 33, 26-31.
Skinner, E. M. W., Scott, B. H. 1979. Petrography, mineralogy and geochemistry of kimberlite and associated
lamprophyre dykes near Swartruggens, W. Transvaal, R. S. A. Kimberlite Symp. II Cambridge, Extended
Abstract.
Skinner E. M. W., Viljoen, K. S., Clark, T. C., Smith C. B. 1994. The petrography, tectonic setting and
emplacement ages ofkimberlites in the south western border region of the Kaapvaal craton, Prieska area.
In Meyer and Leonardos (1994) q.v., Vol. I, pp. 80-97.
Sleep, N. H. 1988. Tapping of melt by veins and dykes. J. Geoph)'s. Res. 93,10255-10272.
Sleep, N. H. 1990. Hot-spots and mantle plumes: Some phenomenology. J. Geoph)'s. Res. 958, 6715-6736.
Smith, C. B. 1977. Kimberlite and mantle-derived xenoliths at Iron Mountain, Wyoming. M.Sc. thesis, Colorado
State University, Fort Collins, CO.
Smith, C. B. 1983. Pb, Sr, and Nd isotopic evidence for sources of African Cretaceous kimberlites. Nature 304,
51-54.
Smith, C. B., Allsopp, H. L., Kramers, J. D., Hutchinson, G., Roddick, J. C. 1985a. Emplacement ages of
Jurassic-Cretaceous South African kimberlites by the Rb-Sr method on phlogopite and whole rock
samples. Trans. Geol. Soc. S. Afr. 88, 249-266.
Smith, C. B., Clark, T. c., Barton, E. S., Bristow, J. W. 1994. Emplacement ages of kimberlite occurrences in
the Prieska region, southwest border of the Kaapvaal craton, South Africa. Chem. Geol. (Isotope Geosci.
Sec.) 113,149-169.
Smith, C. B., Gurney, J. J., Skinner, E. M. w., Clement, C. R., Ebrahim, N. 1985b. Geochemical character of
southern African kimberlites: a new approach based on isotopic constraints. Trans. Geol. Soc. S. Afr. 88,
267-280.
Smith, C. B., Kramers, J. D., Jagovtz, E. 1987. Subcalcic megacrysts in kimberlite: Deep lithosphere or
asthenosphere origins? Terra Cognita 7, 620-621.
Smith, C. B., McCallum, M. E., Coopersmith, H. G., Eggler, D. H. 1979. Petrochemistry and structure of
kimberlites in the Front Range and Laramie Ranges, Colorado-Wyomimg. In Boyd and Meyer (1979)
q.v., Vol. I, 186-189.
Smith, J. V., Brennesholtz, R., Dawson, J. B. 1978. Chemistry of micas from kimberlites and xenoliths. I.
Micaceous kimberlites. Geochim. Cosmochim. Acta 42,959-971.
Smith, J. v., Hervig, R. L., Ackermand, D., Dawson, J. B. 1979. K, Rb, and Ba in micas from kimberlite and
peridotite xenoliths and implication for the origin of basaltic rocks. In Boyd and Meyer (1979) q.v., Vol.
I, pp. 241-251.
Smyth, J. R., Caporuscio, F. A. 1984. Petrology of a suite of eclogite inclusions from the Bobbejaan kimberlite.
II. Primary phase compositions and origin. In Kornprobst (1984) q.v., Vol. 2, pp. 121-132.
Sobolev, N. V. 1977. Deep-Seated Illelusions ill Kimberlites and the Problem of the Composition of tile Upper
Mantle. Am. Geophys. Union, Washington, DC.
Spera. F. 1984. Carbon dioxide in petrogenesis. HI: Role of volatiles in the ascent of alkaline magma with
special reference to xenolith-bearing mafic lavas. Contrib. Mil/eral. Petrol. 88, 217-232.
Spriggs, A. J. 1988. An isotopic and geochemical study of kimberlites and associated alkaline rocks from
Namibia. Ph.D. thesis, University of Leeds, UK.
Stacey, J. S., Kramers, J. D. 1975, Approximation of terrestrial lead isotope evolution by a two stage model.
Eanh Planet. Sci. Len. 26, 207-221.
404 REFERENCES

Storey. B. C.• Alabaster. T.• Pankhurst, R. J. 1992. Magmatism and the Causes o/Continental Break-up. Geo!.
Soc. London. Spec. Pub!. No. 68.
Sudo. A .• Tatsumi. Y. 1990. Phlogopite and K-arnphibole in the upper mantle: Implication for magma genesis
in subduction zones. Geophys. Res. Lett. 17.29-32.
Sun. S. S.• McDonough. W. F. 1989. Chemical and isotopic systematics of oceanic basalts: Implications for
mantle composition and processes. In Saunders. A. D. and Norry. M. 1. (eds.). Magmatism in Ocean Basins.
Geo. Soc. London Spec. Pub!'. pp. 313-354.
Szymanski. 1. T. 1986. The crystal structure of mannardite. a new hydrated cryptomelane-group (hollandite)
mineral with a double short axis. Can. Mineral. 24. 67-78.
Tainton. K. M. 1992. The petrogenesis of group-2 kimberlites and lamproites from the northern Cape Province.
South Africa. Ph.D. thesis. University of Cambridge. UK.
Tainton. K. M.• Browning. P. 1991. The group-2 kimberlite-Iarnproite connection: Some constraints from the
Barkly West District, northern Cape Province. South Africa. Fifth Int. Kimberlite Conf.• Araxa, Brazi!.
Extended Abstracts. Companhia de Pesquisa de Recursos Minerais. Spec. Publ. 2/91. pp. 405-407.
Tainton. K. M.• McKenzie. D. 1994. The generation ofkimberlites. lamproites and their source rocks. J. Petrol.
35.787-817.
Tateyama. H.• Shimoda. S.• Sudo. T. 1974. The crystal structure of synthetic Mgiv mica. Z. Kristallogr. 139.
196-206.
Taylor. L. A .• Neal. C. R. 1989. Eclogites with oceanic crustal and mantle signatures from the Bellsbank
kimberlite. South Africa. Part 1: Mineralogy. petrography and whole rock chemistry. J. Geol. 97. 551-567.
Taylor. W. R.• Green. D. H. 1989. The role ofC-O-H fluids in mantle partial melting. In Ross etal. (1989) q.v.•
Vol. 1. pp. 263-266.
Thibault. Y.• Edgar. A. D.• Lloyd. F. E. 1992. Experimental investigation of melts from a carbonated phlogopite
lherzolite: Implications for metasomatism in the continental lithospheric mantle. Am. Mineral. 77.
784-794.
Thy. P.• Stecher. 0 .• Korstgard. 1. A. 1987. Mineral chemistry and crystallization sequences in kimberlite and
larnproite dikes from the Sisimiut area, central west Greenland. Lithos 20.391-417.
TIkhonenkov. I. P.• Kokharchik. M. v.. Pyatenko. Y. A.1960. Wadeite from the Khibiny massif and the conditions
of its formation. Dokl. Akad. Nauk SSSR 134. 920-923 (Russian).
Tillmans. E. 1969. The crystal structure of BaTIsOl1. Acta Crystal/ogr. 2SD. 1444-1452.
Tompkins. L. A .• Haggerty. S. E. 1984. The Koidu kimberlite complex. Sierra Leone: Geological setting.
petrology and mineral chemistry. In Komprobst (1984) q.v.• Vol. 1. pp. 81-105.
Trl'lnnes. R. G.• Takahashi. E.• Scarfe. C. M. 1988. Stability of K-richterite and phlogopite to 14 GPa. EOS 69.
1510-1511.
Turcotte. D. L.. Oxburgh. E. R. 1978. Intra-plate volcanism. R. Soc. London Phil. Trans. 288A. 561-579.
Ulbrich. M. N. C.• Leonardos. O. H. 1991. The ultrabasic rocks of Presidente Olegario. Serra da Mata da Corda.
Minas Gerais. Brazil. Fifth Int. Kimberlite Conf.. Araxa. Brazil. Extended Abstracts. Companhia de
Pesquia de Recursos Minerais. Spec. Publ. 2/91. pp. 437-439.
Vladimirov. B. M.• Kostrovitskii. S.I.. Soloveva, L. v.. Botkunov. A. I.• Fiveiskaya, L. V.• Egorov. K. N. 1981.
Classification 0/ Kimberlites and the Internal Structure 0/ Kimberlite Pipes. Izdat. Nauka. Moscow
(Russian).
Vladimirov. B. M.• Soloveva. L. V.• Kiselev. A.I.. Egorov. K. N.• Maslovskaya. M. N.• Dneprovskaya. L. V.•
Brandt. S. B.• Semenova. V. G. 1990. Kimberlites and Kimberlite-Like Rocks. Izdat. Nauka. Moscow
(Russian).
Vladykin. N. V. 1985. First find oflarnproites in the USSR. Dokl. Akad. Nauk SSSR 280.718-722 (Russian).
Vollmer. R.. Ogden. P. R.• Schilling. J. G.• Kingsley. R. H.• Waggoner. D. G. 1984. Nd and Sr isotopes in
ultrapotassic volcanic rocks from the Leucite Hills. Wyoming. Contrib. Mineral. Petrol. 87. 359-368.
Vukadinovic. D.• Edgar. A. D. 1993. Phase relations in the phlogopite-apatite system at 20 kbar: Implications
for the role of fluorine in mantle melting. Contrib. Mineral. Petrol. 114. 247-254.
Wagner. C.• Velde. D. 1986. Davanite. K2Ti6015 in the Smoky Butte (Montana) larnproites. Am. Mineral. 71.
1473-1475.
Wagner. P. A. 1914. The Diamond Fields of South Africa. Transvaal Leader. Johannesburg. South Africa.
Wagner. P. A. 1928. The evidence of the kimberlite pipes on the constitution of the outer part of the Earth. S.
Afr. J. Sci. 25.127-148.
REFERENCES 405

Wallace, P., Carmichael, r. S. E. 1989. Minette lavas and associated leucitites from the Western Front of the
Mexican Volcanic Belt: petrology, chemistry and origin. Contrib. Mineral. Petrol. 103,470-492.
Watson, E. B., Brennan, J. M. 1987. Fluids in the lithosphere 1: Experimentally-determined wetting charac-
teristics of C02-H20 fluids and their implications for fluid transport, host-rock physical properties, and
fluid inclusion formation. Earth Planet. Sci. Lett. 85, 497-515.
Wedepohl, K. H., Muramatsu, Y. 1979. The chemical composition of kimberlites compared with the average
composition of three basalt magma types. In Boyd and Meyer (1979) q.v., Vol. I, pp. 300-312.
Wendlandt, R F. 1977. Barium phlogopite from Haystack Butte, Highwood Mountains, Montana. Carnegie
Inst. Washington Yearb. 76, 534-539.
Wentworth, C. K. 1922. A scale of grade and class terms for clastic sediments. J. Geol. 30,377-392.
White, R., McKenzie, D. 1989. Magmatism at rift zones: The generation of volcanic continental margins and
flood basalts. J. Geophys. Res. 94, 7685-7729.
Wilding, M. c., Harte, B., Fallick, A. E., Harris, 1. W. 1994. Inclusion chemistry, carbon isotopes and nitrogen
distribution in diamonds from the Bultfontein mine. In Meyer and Leonardos (1994) q.v., Vol. 2, pp.
116-126.
Williams, A. F. 1932. The Genesis of Diamond. Ernest Benn, London, UK (2 vols.).
Wilson, A. N. 1982. Diamondsfrom Birth to Eternity. Gemological Institute of America, Santa Monica, CA.
Wohletz, K. H., Smyth, J. R 1984. Origin of a Roberts Victor sanidine-coesite grospydite: Thermodynamic
considerations. In Komprobst (1984) q.v., Vol. 2, pp. 33-42.
Woolley, A. R, Bergman, S., Edgar, A. D., Le Bas, M. J., Mitchell, R. H., Rock, N. M. S., Scott Smith, B. H.
1995. Classification of lamprophyres, lamproites, kimberlites, and the kalsilite-, melilite- and leucite-
bearing rocks. (Recommendations of the lUGS Subcommission on the Systematics of Igneous Rocks).
Can. Mineral., in press.
Wu, X. J., Li, F. H., Hasimoto, H. 1990. Electron microscopy study of the incommensurately modulated
structure of ankangite. Acta Crystallogr. 846, 111-117.
Wyatt, B. A. 1979. Manganoan ilmenite from the Premier kimberlite. Kimberlite Symp. II. Cambridge,
Extended Abstracts.
Wyllie, P. 1. 1980. The origin of kimberlite. J. Geop/zys. Res. 85, 6902-6910.
Wyllie, P. 1. 1987. Discussion of recent papers on carbonated peridotite, bearing on mantle metasomatism and
magmatism. Earth Planet. Sci. Lett. 82, 391-397.
Wyllie, P. 1. I 989a. The genesis of kimberlites and some low-Si02, high-alkali magmas. In Ross et al. (1989)
q.v., Vol. I, pp. 603-615.
Wyllie, P. J. 1989b. Origin of carbonatites: evidence from phase equilibrium studies. In Bell, K. (ed.),
Carbonatites. Unwin Hyman, London, UK, pp. 500-545.
Wyllie, P. J., Baker, M. B., White, B. S. 1990. Experimental boundaries for the origin and evolution of
carbonatites. Lithos 26, 3-19.
Wyllie, P. 1., Huang, W. L. I 975a. Influence of mantle C02 in the generation of carbonatites and kimberlites.
Nature 257,297-299.
Wyllie, P. 1., Huang, W. L. 1975b. Peridotite, kimberlite and carbonatite explained in the system CaO-MgO-
Si02-C02. Geology 3, 621-624.
Wyllie, P. J., Huang, W. L. 1976. Carbonation and melting reactions in the system CaO-MgO-Si02-C02 at
mantle pressures with geophysical and petrological applications. Contrib. Mineral. Petrol. 54, 79-107.
Wyllie, P. 1., Huang, W. L., Otto, J., Byrnes, A. P. 1983. Carbonation of peridotites and decarbonation of siliceous
dolomites represented in the system CaO-MgO-Si02-C02 to 30 kbar. Tectonophysics 100, 359-388.
Xiong, M., Ma, Z., Peng, Z. 1989. A new mineral-ankangite. Chinese Sci. BlIll. 34, 592-596.
Yamashita, H., Arima, M., Ohtani, E. 1992. Melting experiments of group II kimberlites up to 10 GPa:
Petrogenesis of kimberlite magma.Int. Geol. Congress, Kyoto, Japan, Abstract, p. 538.
Yingchen, R, Lulu, X., Zhizhong, P. 1983. Daqingshanite-a new mineral recently discovered in China.
Geochem. (China) 2,180-184.
Zandbergen, H. w., Everstijn, P. L. A., Mijlhoff, F. c., Renes, G. H., Ijdo, D. 1. W. 1987. Composition,
constitution and stability of the synthetic hollandites AxM4·2xN2tDg, M =Ti, Ge, Ru, Zr, Sn and N =AI,
Sc, Cr, Ga, Ru, In and the system (A,BahTiyAI;:Og with A =Rb, Cs, Sr. Mater. Res. Bull. 22, 431-438.
Zhao, D., Smith, D. G. w., Zhou, M., Yang, J., Deng, c., Huang, Y. 1993. Yinniugoulamproites in Datong,
Northern Shanxi Province, China: First occurrence in the North China craton. In Dunn, P. E. K., and Grant,
406 REFERENCES

B. (eds.), Mid-Continent Diamonds. Geol. Assoc. Canada-Mineral. Assoc. Canada Symp. Vol., pp.
139-144.
Zindler, A., Hart, S. 1986. Chemical geodynamics. Ann. Rev. Eanh Planet. Sci 14, 493-571.
Zuffa, G. G. 1991. More comments on the use of "epiclastic" and about the distinction of volcanic and
sedimentary surface processes. IAVCEI Commission on Volcanogenic Sediments Newsletter 4, 6--8.
Zhuravleva, L. N., Yukina, N. v., Ryabeva, Y. G. 1978. Priderite, first find in the USSR. Dokl.Akad. Nauk SSSR
239, 141-143. (Russian).
INDEX

Amphiboles: see potassium richterite Composition (cont.)


Ancylite, 246 trace elements
Apatite alkali, 277-280
composition, 15,225--227 alkaline earths, 265
paragenesis, 76, 225 averages, 263-264, 266--267, 273
Autolithic kimberlite, 46--50 compatible, 262-264
incompatible, 264-265
inter-element relationships. 280-287
Barite, 247 rare earth elements, 272-277
Barium, III, 114-115,265--267 transition, 265--271
see also individual elements
volatile elements, III, 253
Calcite, 245 Copper, 262-263
Cesium, 278-279, 281
Chromium, 111,262-263,288
Clinopyroxene: see Diopside; Titanian aegirine Daqingshanite, 227-228
Cobalt, 262-263 Diamond, 1-2,4,20-21,31,91,319,347-348,
Composition 363,365,372,380-384
isotopic Diopside
carbon, 299-301 kimberlite, II, 15, 178-179
hydrogen, 299 lamproite, 179-180
kimberlite, 3-4, 293-295 lamprophyre, 176-177, 180-181
lamproite, 295--297 megacrystal, 177-178
lead, 296--298 minette, 181
neodymium, 292-296 orangeites
orangeite, 292-301,350,356 composition, 166--176
oxygen, 299-301 paragenesis, 72, 166
strontium, 292-296, 300 Dolomite, 24, 245, 299-230, 324
major elements
alteration, 251
averages, 249,254-255,258 Epiclastic rocks: see Resedimented volcaniclastic
contamination, crustal, 250-251 rocks; Volcanogenic sedimentary rocks
contamination, mantle, 252, 288-291, 298, Experimental studies on
306, 363, 372 kimberlite
evolved orangeite, 257 natural, 325--327
kimberlite, 258-261 synthetic, 327-326
lamproite, 261-262 ultra-high pressure, 328-329
mineralogical controls on, 255--257 lamproites, 311-312
primary magma, 252, 257 mica pyroxenites, 312-314
unevolved orangeite, 253-257 orangeites, 310-311

407
408 INDEX

Experimental studies on (conI.) Kimberlite (conI.)


synthetic systems petrogenesis (conI.)
CaO-MgO-AI 20 3-Si0 2-C02, 320-321 metasome melting, 322-323, 333-335
carbonated lherzolite, 315--317 partial melting lherzolite, 304-309
phlogopite-richterite-apatite, 314-315 redox melting, 341-343
source mantle, 327, 321, 329-331, 344-351
Fluorine, III, 114 transition zone melting, 331-333
volatile fluxing, 317-320, 335, 357-358
Genetic classifications, 7-8, 10 recognition, 11-15,74-89
Geodynamic models of petrogenesis, 329-343 tectonic setting, 17,335--337, 340,357-358,
372,381-384
Hafnium, 265--268, 281 Kinoshitalite, 15,74, \11, 130, 132-133, 141, 143,
Hollandite 155
alkaline rocks, 211-213
kimberlite, 211 Lamproite,8, 10-\1,26-27,79,89-90,157-160,
lamproite, 207-209 179-180, 187-188, 198--199,207-211,
orangeite 223-224,227,233-237,241-243,261-
composition, 201-207 262,268,271,285,306,308-309,322,
paragenesis, 200-201 331,342-343,376-377
Lamprophyre
Ilmenite facies, 9-10
kimberlite, 240-241 micas, 118-\19,161-165
lamproite, 241 pyroxenes, 176-177, 180-181
orangeite, 237-239 spinels, 198--200
Lead,279-280, 301
Kimberlite Leucite, 74, 235--236
classification, 3-5, 10
crater facies Macrocrysts
lavas, 37 definition, 5
pyroclastic rocks, 38-39 kimberlite,6, 15,74-77,79, 127-128, 144
resedimented volcaniclastic rocks, 38-40 orangeite, 6, 23-24, 35, 60-62, 93, 109-110,
volcanogenic sedimentary rocks, 40-41, 52, 117-\18,121,181-183,195,356
54-57 Mantle
definition, 14-15 carbonates, 315, 320-322
diatreme facies eclogite, 6, 27, 347-348
autoliths, 46-48 harzburgite, 6, 288, 308, 323, 345--351, 381, 383
definition, 41-42 lherzolite, 6, 21, 288--291, 309, 312, 326, 346-
pelletallapilli, 43-47,50-51 347
volcaniclastic kimberlites, 41-42 majorite garnet, 330-331
etymology, 2-4 MARIO-suite rocks, 6, 35, 116, 349
hypabyssal facies metasomatic veins, 312, 315, 348--352, 376
autoliths, 49-50 metasomatism, 307, 319, 322-323, 333-334,
definition, 48-49 339,348--351,376-379
globular segregations, 50-51 plumes, 320, 334, 353-356, 371, 375
nomenclature, 5, 10-11,37-50 source of orangeites, 343-352
nucleated globules, 50 Megacrysts
relation to diatreme facies, 35--36, 50-58 alkaline rocks, 371, 374
segregation textures, 49-50, 76, 83, 85-87 definition, 5
petrogenesis genesis, 370-371, 374-375
carbonated lherzolite, 315--325 kimberlite, >-6, \1,49,321
carbonated phlogopite lherzolite, 322-325 orangeite, >-6, \1,21,34,60,62,177-178
depth of melting, 367-369
heterogeneous lithosphere melting, 338-341 Melilite, 15--16
hot-spot magmatism, 335--338, 340 Melilitite, 2, 43, 51, 58, 320, 374-375, 384
magnesite peridotite, 320-322, 306 Minette, 7-8, 31,156,159-161,181,200,379
mantle plumes, 333-335 Modal classifications, 7-8
INDEX 409

Monazite, 228--229 Petrologic clans


Monticellite, II, 74 definitions, 8--9, 379
kimberlite, 10
Nickel, 262-263, 288--289, 301 lamproite, 9-10, 377, 379
Niobium, 266-271, 283, 286, 301 lamprophyre,9-IO
Norsethite,246 melilitite, 58
orangeite, 13-14,343,379
Olivine Phlogopite
kimberlite kimberlite
composition, 186-187 composition, 128--156
paragenesis, 74, 78--80, 185 paragenesis, 11-13,74-79,126-155
lamproite, 89, 62,187-188 lamproite, 156-160
orangeite lamprophyres, 118--119, 161-165
composition, 183-187 minette, 156, 159-161
paragenesis, 26, 89, 181-183, 366 orangeite
Orangeite aluminous macrocrysts, 96, 115-117, 121
age, 16-17,34 composition, 94-122,156
definition, 14 paragenesis, 11-13,61-74,91-93
diatreme facies, 19-20, 35, 43-44, 58, 60, solid solutions in, 122-126
364 Phosphates
etymology, 2-4 apatite, 225-227
evolved, 60, 73,257-258 daqingshanite, 227-228
hypabyssal facies, 21-22, 26-30, 35, 50, 58, monazite, 228--229
364-365 Sr-REE phosphate, 229
nomenclature, 60--62, 72-74 Potassium barium titanates
occurrence, 18--35 barium pentatitanate, 216
petrogenesis hollandite, 200-213
heterogeneous lithosphere melting, 338- K-triskaidecatitanate, 213-215
341 Potassium feldspar
hot-spot magmatism, 335-338 lamproite, 236-237
mantle contamination, 305-306, 359-362 minette, 237
partial melting lherzolite, 304-309 orangeite, 26-27, 34, 72, 235-237
redox melting, 341-343 Potassium richterite
three stage models, 306-309 lamproites, 233-235
veined lithosphere, 358--362 orangeites
volatile fluxing, 357-358 composition, 230-233
petrography, 60-74 paragenesis, 26-27, 34, 72-73, 229-230
recognition of, 11-14,74-77,79,89-90
relationships to
Quartz, 31,247
kimberlites, 11-13,258--261,304-305,308,
331-332,335-337,339-341,375-376
lamproites, 261-262, 376-377 Rare earth elements, 272-277, 286, 289-291,
ultrapotassic magmas, 377-378 301
tectonic setting, 16-18, 306, 340, 350, 354- Resedimented volcaniclastic rocks, 38-41
356 Rubidium, 114-115,277-278,281,286,301
unevolved,253,261-262 Rutile
kimberlite, 242
Pelletallapilli, 43-46,50-51,58 orangeites, 241-242
Peridotite contamination, 252, 288--291, 298, 305-
307,362-363,372 Scandium, 262-263
Perovskite Spinels
kimberlite, 13, 15,76,221-223 kimberlites
lamproite, 13, 223-224 composition, 15, 195-198
orangeite paragenesis, 11,13,61,76-77,89
composition, 15, 218--221 lamproites, 198--199
paragenesis, 13, 34, 216-217 lamprophyres, 199-200
410 INDEX

Spinels (conI.) Wadeite, 243


orangeites Witherite, 246
composition, 189-195
paragenesis, II, 13,34,61,77,188 Ultrapotassic rocks, 377-380
Strontium, 265-267, 281, 283, 286, 301 Uranium, 266-267, 271-272, 281, 283
Sulfides, 247
Sulfur, 253 Vanadium, 262-263
Volcaniclastic kimberlites, 42, 46-48
Tantalum, 265-269, 283, 286, 301 Volcanogenic sedimentary rocks, 40-41, 52, 54-
Tetraferriphlogopite 57
kimberlites, 137-138, 140-141, 150-153, 155
lamproites, 157-159
Xenocrysts, 5-6, 91
orangeites, 61, 72-73, 92,106,109,114,122-126
Textural-genetic classifications
kimberlite, 37-58 Zinc, 262-263
melilitites, 58 Zirconium, 265-268, 270-271, 281
orangeites, 58 Zirconium silicates
principles, 35-37 calcium zirconium silicate, 245
Thorium, 266-267, 271-272, 281, 283 wadeite, 243
Titanian aegirine, 171-172 zircon, 243
Tuffisitic kimberlite, 41-42 zirconium-bearing gamet, 244-245

Das könnte Ihnen auch gefallen