You are on page 1of 183

Petroleum Geology of South Australia

Volume 3: Officer Basin

edited by: J.G.G. Morton and J.F. Drexel

MINES and ENERGY SOUTH RESOURCES AUSTRALIA

The Petroleum Geology of South Australia


Volume 3: Officer Basin
Edited by J.G.G. Morton and J.F. Drexel

1st Edition April 1997

SA Department of Mines and Energy Resources 191 Greenhill Road PO Box 151 EASTWOOD SA 5063 Telephone: National International Facsimile National International (08) 8274 7680 +61 8 8274 7680 (08) 8373 3269 +61 8 8373 3269 Prepared for publication by: Publication Section Mines and Energy Resources South Australia Printed by Rainbow Colour Copy Centre April 1997

Web Page http://www.mines.sa.gov.au/petrol/index.html Report Book 97/19


Front cover: Observatory Hill 1 being drilled in the Officer Basin in 1985. (Photo 35507)

Bibliographic reference: Entire volume: Morton, J.G.G. and Drexel, J.F., 1997. The petroleum geology of South Australia. Vol 3: Officer Basin. South Australia. Department of Mines and Energy Resources. Report Book, 97/19. Individual chapter: ONeil, B.J., 1997. History of petroleum exploration. In: Morton, J.G.G. and Drexel, J.F. (Eds), The petroleum geology of South Australia. Vol 3: Officer Basin. South Australia. Department of Mines and Energy Resources. Report Book, 97/19:7-22. Petroleum Geology of South Australia 1st ed. Includes bibliographical references ISBN 0 7308 0644 8 (set). ISBN 07308 0643 X (v.1). ISBN 07308 0762 2 (v.2). ISBN 07308 0164 0 (v.3: prepublication). 1. Petroleum Geology South Australia. 2. Petroleum Geology Otway Basin (Vic. and S. Aust). 3. Petroleum Geology Officer Basin (S. Aust and W.A.). 4. Petroleum Geology Eromanga Basin. 5. Petroleum reserves South Australia. 6. Petroleum reserves Otway Basin (Vic. and S. Aust). 7. Petroleum reserves Officer Basin (S. Aust. and W.A.). 8. Petroleum reserves Eromanga Basin. I. Morton, John George Godfrey. II. Drexel, J.F. (John F.), 1952-. III. Alexander, Elinor M. IV. Hibburt, J. V. South Australia. Dept. of Mines and Energy. Petroleum Division. (Series: Report book (South Australia. Dept. of Mines and Energy); 95/12. 553.28099423 Keywords: Petroleum geology / South Australia / Stratigraphy / Structural geology / Tectonics / Petroleum reservoir characterisation / Source rock studies / Officer Basin / Palaeozoic.

EE

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix Chapter 1: Introduction and summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Chapter 2: History of petroleum exploration


Discovery of the Officer Basin . . . . . . . Early European exploration . . . . . . . . . Weapons testing . . . . . . . . . . . . . . . Post-World War II geological exploration . . Aboriginal land rights and exploration access Petroleum industry exploration . . . . . . . OEL 8 . . . . . . . . . . . . . . . . . OEL 12 . . . . . . . . . . . . . . . . OEL 19 and 28 . . . . . . . . . . . . OEL 33 . . . . . . . . . . . . . . . . Downturn to discovery . . . . . . . . . . PEL 10, 11 and 12 . . . . . . . . . . PEL 13 . . . . . . . . . . . . . . . . PEL 23 and 30 . . . . . . . . . . . . PEL 24 . . . . . . . . . . . . . . . . PEL 29 . . . . . . . . . . . . . . . . PEL 33 . . . . . . . . . . . . . . . . Revitalisation . . . . . . . . . . . . . . . PEL 61 and 63 . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 . 7 . 9 10 11 16 16 16 16 19 19 20 20 20 21 21 21 21 22 22

Chapter 3: Natural and cultural environment


Climate . . . . . . . . . . . . . . . . . . . Landforms . . . . . . . . . . . . . . . . . . . Western sandplains region . . . . . . . Central tablelands . . . . . . . . . . . . Nullarbor Plain . . . . . . . . . . . . . Native vegetation . . . . . . . . . . . . . . . . Environmental considerations . . . . . . . . . National parks and reserves . . . . . . Summary of environmental regulation . Cultural heritage . . . . . . . . . . . . . . . . European heritage . . . . . . . . . . . Aboriginal heritage . . . . . . . . . . . Aboriginal lands . . . . . . . . . . . . . . . . History . . . . . . . . . . . . . . . . . Access to Aboriginal lands . . . . . . . AP Lands . . . . . . . . . . . . . . MT Lands . . . . . . . . . . . . . . Access for seismic surveys . . . . . Other land issues . . . . . . . . . . . . . . . . Commonwealth land . . . . . . . . . . Maralinga and Emu nuclear weapon test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 23 23 23 24 24 24 24 24 24 24 24 25 25 26 26 26 27 27 27 27

iii

Woomera Prohibited Area . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 Mintabie Precious Stones Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

Chapter 4: Infrastructure and groundwater


Infrastructure . . . . . . . . . . . . . . . Transport links . . . . . . . . . . Pipelines and production facilities Access to potential markets . . . . . . . Industries in the region . . . . . . Gas . . . . . . . . . . . . . . . . Crude oil . . . . . . . . . . . . . Groundwater . . . . . . . . . . . . . . . Surface water . . . . . . . . . . . Aquifers . . . . . . . . . . . . . . Recharge . . . . . . . . . . . . . Local recharge potential . . . . . Discharge . . . . . . . . . . . . . Potentiometric surface . . . . . . Water quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 29 29 31 31 32 32 32 32 32 32 33 33 34 34

Chapter 5: Geological setting and structural history


Introduction . . . . . . . . . . . . . . . . . Plate tectonic setting . . . . . . . . . . . . . Basement structural elements . . . . . . . . Eastern margin ---- Gawler Craton . . Tallaringa Trough and Nawa Ridge . Northern margin ---- Musgrave Block Middle Bore Ridge . . . . . . . . . . Mafic dykes . . . . . . . . . . . . . . Coompana Block . . . . . . . . . . . Structures beneath the basin . . . . . Basin architecture . . . . . . . . . . . . . . Munyarai Trough . . . . . . . . . . . Birksgate Sub-basin . . . . . . . . . Marla Overthrust Zone . . . . . . . . Ammaroodinna Ridge . . . . . . . . Manya Trough . . . . . . . . . . . . Murnaroo Platform . . . . . . . . . . Nullarbor Platform . . . . . . . . . . Structural history . . . . . . . . . . . . . . . Stage 1 . . . . . . . . . . . . . . . . Stage 2 . . . . . . . . . . . . . . . . Stage 3 . . . . . . . . . . . . . . . . Stage 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 36 37 37 37 37 39 39 40 40 41 41 41 41 43 43 43 43 44 44 44 44 44

Chapter 6: Lithostratigraphy and environments of deposition


Introduction . . . . . . . . . . . . . . . . . . . . Officer Basin . . . . . . . . . . . . . . . . . . . . Undrilled sequence below the Callanna Group Callanna Group equivalents . . . . . . . . . . Pindyin Sandstone . . . . . . . . . . . . . Alinya Formation . . . . . . . . . . . . . . Coominaree Dolomite . . . . . . . . . . . Cadlareena Volcanics . . . . . . . . . . . . Umberatana Group equivalents . . . . . . . . Chambers Bluff Tillite . . . . . . . . . . . Wantapella Volcanics . . . . . . . . . . . . Lake Maurice Group . . . . . . . . . . . . . . Tarlina Sandstone . . . . . . . . . . . . . . Meramangye Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 47 47 47 50 51 52 53 54 54 56 56 56 57

iv

Murnaroo Formation . . . . . . . . . . Ungoolya Group . . . . . . . . . . . . . . Dey Dey Mudstone . . . . . . . . . . . Karlaya Limestone . . . . . . . . . . . Tanana Formation . . . . . . . . . . . . Wilari Dolomite Member . . . . . . Munyarai Formation . . . . . . . . . . Narana Formation . . . . . . . . . . . Munta Limestone Member . . . . . Mena Mudstone Member . . . . . . Punkerri Sandstone . . . . . . . . . . . Marla Group . . . . . . . . . . . . . . . . Relief Sandstone . . . . . . . . . . . . Ouldburra Formation . . . . . . . . . . Observatory Hill Formation . . . . . . Cadney Park Member . . . . . . . . Wallatinna Member . . . . . . . . . Parakeelya Alkali Member . . . . . Moyles Chert Marker Bed . . . . . Oolarinna Member . . . . . . . . . Arcoeillinna Sandstone . . . . . . . . . Apamurra Formation . . . . . . . . . . Mount Johns Conglomerate Member Trainor Hill Sandstone . . . . . . . . . Late Cambrian Volcanics . . . . . . . . . . Kulyong Formation . . . . . . . . . . . Munda Group . . . . . . . . . . . . . . . . Mount Chandler Sandstone . . . . . . . Byilkaoora Member . . . . . . . . Indulkana Shale . . . . . . . . . . . . . Blue Hills Sandstone . . . . . . . . . . Devonian Sequence . . . . . . . . . . . . Mimili Formation . . . . . . . . . . . . Arckaringa Basin . . . . . . . . . . . . . . . . Eromanga Basin . . . . . . . . . . . . . . . . Tertiary . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

58 59 59 60 62 64 64 65 66 67 67 68 68 70 72 73 74 75 75 76 76 77 78 79 79 79 80 80 81 82 83 83 83 84 86 86

Chapter 7: Biostratigraphy and correlation


Introduction . . . . . . . . . . . . . . . Neoproterozoic stratigraphy and age . . . Stromatolites . . . . . . . . . . . Early Neoproterozoic acritarchs . Late Neoproterozoic acritarchs . . Late Neoproterozoic invertebrates Geochronology . . . . . . . . . . Chemo- and magnetostratigraphy Lithology and event stratigraphy . Cambrian stratigraphy and age . . . . . . Invertebrate fossils . . . . . . . . Acritarchs . . . . . . . . . . . . . Geochronology . . . . . . . . . . Lithology . . . . . . . . . . . . . Ordovician stratigraphy and age . . . . . Ichnology . . . . . . . . . . . . . Geochronology . . . . . . . . . . Lithology . . . . . . . . . . . . . Devonian stratigraphy and age . . . . . . Vertebrate fossils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 87 87 88 88 89 89 89 94 94 94 95 95 95 95 95 96 96 96 96

Palynomorphs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 Geochronology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

Chapter 8: Source rock distribution and quality


Introduction . . . . . . . . . . . . Formation description . . . . . . . Alinya Formation . . . . . . Coominaree Dolomite . . . Meramangye Formation . . Dey Dey Mudstone . . . . . Karlaya Limestone . . . . . Tanana Formation . . . . . Munyarai Formation . . . . Narana Formation . . . . . Ouldburra Formation . . . . Observatory Hill Formation Apamurra Formation . . . . Indulkana Shale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 . . 99 . . 99 . 101 . 101 . 102 . 102 . 103 . 103 . 103 . 104 . 106 . 108 . 108

Chapter 9: Petroleum maturation and migration


Introduction . . . . . . . . . . . . . . . . . . . . . . Oil-source correlation and maturity . . . . . . . . . . Trainor Hill Sandstone . . . . . . . . . . . . . Observatory Hill Formation . . . . . . . . . . Ouldburra Formation . . . . . . . . . . . . . . Relief Sandstone . . . . . . . . . . . . . . . . Cambrian maturity mapping . . . . . . . . . . Tanana Formation, Karlaya Limestone and Dey Murnaroo Formation . . . . . . . . . . . . . . Alinya Formation . . . . . . . . . . . . . . . . Proterozoic maturity mapping . . . . . . . . . Thermal maturity . . . . . . . . . . . . . . . . . . . . Permian vitrinite reflectance . . . . . . . . . . Fluid inclusion microthermometry . . . . . . . Apatite fission track analysis . . . . . . . . . . Geohistory modelling . . . . . . . . . . . . . . . . . Manya Trough . . . . . . . . . . . . . . . . . Ammaroodinna Ridge . . . . . . . . . . . . . Marla Overthrust Zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Dey Mudstone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 109 109 109 . 112 . 112 . 112 . 113 . 114 . 114 . 114 . 114 . 114 . 116 . 116 . 117 . 118 . 119 . 119

Chapter 10: Reservoirs and seals


Introduction . . . . . . . . . . . Formation description . . . . . . Pindyin Sandstone . . . . Tarlina Sandstone . . . . . Murnaroo Formation . . . Relief Sandstone . . . . . Ouldburra Formation . . . Arcoeillinna Sandstone . . Trainor Hill Sandstone . . Mount Chandler Sandstone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 121 121 121 122 123 125 125 126 127

Chapter 11: Potential traps and prospects


Introduction . . . . . . . . . . . . . . . . . Previous unsuccessful tests . . . . . . . . . Arcoeillinna Sandstone in Munyarai 1 Arcoeillinna Sandstone in Ungoolya 1 Arcoeillinna Sandstone in Karlaya 1 . Arcoeillinna Sandstone in Lairu 1 . . Murnaroo Sandstone in Munta 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 130 130 130 130 130 130

vi

Potential trap types . . Potential trap volumes Prospect A . . Prospect B . . Prospect C . . Prospect D . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

. . . . . .

131 131 131 131 131 131

Chapter 12: Exploration and production economics


Introduction . . . . . . . . . . . . . . Methodology . . . . . . . . . . . . . . Exploration success ratios . . . . . . . Exploration and development scenarios Case 1 . . . . . . . . . . . . . . Case 2 . . . . . . . . . . . . . . Royalty and licence fees . . . . Results . . . . . . . . . . . . . . . 20 mmbbl OOIP case study . . Discussion . . . . . . . . . . . Minimum economic field size . Upside potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141 141 141 141 142 142 143 143 143 144 145 145

Chapter 13: Undiscovered resources


Introduction . . . . . . . . . Method . . . . . . . . . . Discussion of parameters . . . Potential plays . . . . . . . . Pindyin Sandstone . . Tarlina Sandstone . . . Murnaroo Formation . Relief Sandstone . . . Ouldburra Formation . Arcoeillinna Sandstone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 147 148 150 150 150 150 150 151 151

Appendices
1.1 3.1 Abbreviations used throughout the text . . . . . . . . . . Sections of the Pitjantjatjara Land Rights Act 1981 relevant to petroleum exploration in the Officer Basin . . 3.2 Sections of the Maralinga Tjarutja Land Rights Act 1984 relevant to petroleum exploration in the Officer Basin . . 12.1 Economic model data and assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

154 159 165

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168

vii

viii

This summary of the of the petroleum geology of the South Australian portion of the Officer Basin forms the third volume of the Petroleum geology of South Australia series. It follows on from Volume 1: Otway Basin published in 1995 and Volume 2: Eromanga Basin published in 1996. The remaining volumes will be published on the following schedule: 1998 Volume 4: Cooper Basin 1999 Volume 5: Stansbury, Arrowie and Warburton Basins 2000 Volume 6: Duntroon and Bight Basins It is intended that the volumes be fit for purpose and, consequently, this has resulted in some compromise on print quality and cost. To avoid delays in publication, the volumes are not being indexed. However, this approach enables each volume to be updated and republished quickly as significant new data come to hand. Seismic contour maps and image-processed data which include the latest open file interpretations are continuously available on request from MESAs seismic interpretation digital database. Any and all comments or criticisms on this volume on the Officer Basin are welcome and will assist us in improving future publications. Please address correspondence to the Director, Petroleum Division, MESA, PO Box 151, Eastwood, SA 5063. Email: rlaws@msgate.mesa.sa.gov.au. Further information is available through our World Wide Web site: http://www.mines.sa.gov.au/petrol/index.html

EN

The idea for a series of publications on the petroleum basins of South Australia was originally conceived by Bob Laws, Director, Petroleum Division, MESA, who has enthusiastically supported the project. Thanks are due to Greg Borchers (AP Legal Service) and Andrew Collett (MT Legal Service) for their comments on text relating to Aboriginal issues. Drs John Lindsay and James Leven (AGSO) kindly provided a pre-print of their paper on the tectonic evolution of the Officer Basin. John Naylor (Hemley Exploration Pty Ltd) generously furnished information for use in the economic model. Information on road construction costs and airports was provided by officers of the South Australian Department of Road Transport. In addition to the authors, Colin Gatehouse, Paul Rogers, Reza Moussavi-Harami and Wenlong Zang read early versions of the text and made valuable comments. David Gravestock and Wenlong Zang compiled the lithology logs for the reference sections and provided photographs and assistance through many discussions on the stratigraphy of the Officer Basin (Ch. 6). John Iredale, Don Vinall and Peter Hough assisted in compiling the seismic section figures for Chapter 11. Editing was by John Morton and John Drexel, with assistance by David Gravestock. Word processing was by Jeanette Bell and Melanie Lenuzzi. The references were compiled and checked by John Drexel from the authors contributions. The figures were drafted by Elaine Appelbee, Gayle Bruggemann, Adrian Francis, Tricia Fraser, Todd McKenzie, Michael Ross, Jeff Weber and Jamie Williams under the supervision of Barry Frost. Spatial data were provided by Neil Sandercock. The publication was designed and desktop published by Rachel Froud.

$%

The Officer Basin covers an area of >300 000 km 2 in South Australia and Western Australia; this publication summarises the petroleum potential of the 100 000 km 2 in South Australia (Fig. 1.1). An industry survey conducted by Morton (1996a) indicated that the petroleum industry in general was unable to make a judgement on the prospectivity of the Officer Basin (as well as the other Cambrian Basins in South Australia). Those representatives that could comment were generally negative in their views, quoting unfavourable economics, lack of geological knowledge, difficulties with land access and poor source quality as the main deterrents to exploration. These views are not shared by all industry members however, and exploration activity in the area is at an all time high, with two licences granted and a third under application (Fig. 1.2). This publication is written specifically for a petroleum industry audience, and aims to dispel some of the myths surrounding the Officer Basins petroleum potential. The Officer Basin has close geological affinities with the productive Amadeus Basin in the Northern Territory, and with basins in the former USSR and Oman, both of which host giant oil and gas fields and have proven oil reserves in the order of billions of barrels. Numerous oil shows are known in the Officer Basin from mineral and stratigraphic drillholes, although there has been little on-structure drilling for petroleum targets. Excellent reservoir quality and source are proven. Exploration economics are favourable, even at relatively low oil prices (US$18/bbl), and the area is one where Native Title considerations are not a major issue. The Officer Basin represents one of the last remaining onshore frontier exploration areas where large petroleum discoveries may still be made. Previous exploration in the Officer Basin has been limited to a few shallow wells (mostly off-structure and drilled for mineral or stratigraphic purposes), a few regional seismic lines, and aeromagnetic and surface mapping by the Government. Petroleum licences have been issued sporadically in the area since 1954, but little modern petroleum exploration has taken place. The reasons for this have been the areas remoteness, previous access restrictions (atomic bomb testing and weapons testing), and perception that little oil will be found in Proterozoic and early Palaeozoic rocks in Australia. The State and Federal Governments have long been involved in geological investigations in the area, and have carried out most of the petroleum orientated exploration, including source rock studies, seismic acquisition, reservoir core analysis, surface mapping, aeromagnetic surveys, and stratigraphic drilling. Less than

7200 km of seismic data have been recorded and only 30 drillholes deeper than 500 m have been drilled; of these only seven had petroleum targets, and only one has subsequently been shown to have been a valid structural test. Exploration commitments for the current licences total $22 million over the next five years, and this effort should begin to validly test the potential of the area. The natural environment in the Officer Basin is regarded as harsh desert, with low and irregular rainfall, and extremely hot conditions in the summer months. Landforms comprise undulating plains and extensive dunefields (Great Victoria Desert) with vegetation of low open woodlands, deceptively lush in spite of the harsh desert climate, low tablelands with undulating plains covered in rubble, and in the south the Nullarbor Plain, a flat featureless limestone plain. Native vegetation in these latter areas is very sparse and comprises saltbush and grasses only. There are six parks and reserves in the region, all of which allow exploration access except for the Unnamed Conservation Park on the western State border, and the Nullarbor National Park on the southern coastal region. Most of the area is held as Aboriginal land (as a freehold title by a body corporate), the Anangu Pitjantjatjara (AP) in the north and the Maralinga Tjarutja (MT) in the south. In both of these areas, the Aboriginal people have the right to control entry to their lands and seek compensation for disturbance to their ways of life. Industry have successfully negotiated access to these lands for exploration and production. In the case of MT Lands, compensation at the exploration stage is limited to that which would apply to any other landowner in the State, as provided under the Petroleum Act. Some land access restrictions also apply to small areas around the atomic bomb test sites and in the Woomera Prohibited Area (access may be granted by the Department of Defence). Infrastructure in the region is relatively limited, with the fully sealed Stuart Highway the main road link to Adelaide and Darwin. The Central Australia Railway links to Adelaide, and airstrips suitable for small aircraft are available at several locations. No pipelines cross the Officer Basin, although the Alice Springs--Darwin gas pipeline is 550 km to the north, and the Moomba--Port Bonython liquids pipeline ~600 km to the east. It is likely that a transcontinental pipeline will be built in the next 10--20 years to link the North West Shelf and Timor Sea gas fields with markets in the populated southeastern parts of Australia; this pipeline would most likely cross part of the Officer Basin. Surface water supplies in the area are virtually non-existent and rely on groundwater which is mostly highly saline.

Fig. 1.1

Fig. 1.2 Well locations and seismic coverage, Officer Basin.

The basin in South Australia is bounded to the north by the Musgrave Block, to the southwest and southeast by the Coompana Block and Gawler Craton respectively, but is poorly defined to the northeast where it merges with the Warburton and Amadeus Basins and is covered by Arckaringa and Eromanga Basin sediments. Sediments are thickest in the Munyarai Trough, and are abruptly truncated by thrust faults against the southern margin of the Musgrave Block. Thrusting is most clearly seen in the Marla Overthrust Zone. Other significant depocentres are the Tallaringa Trough, Birksgate Sub-basin and Manya Trough. The southern part of the basin, designated the Nullarbor Platform, is poorly known due to Tertiary cover of the Eucla Basin, but is most likely to be Neoproterozoic sediments only. The structural history of the Officer Basin, from mid-Neoproterozoic to Late Devonian (~820360 Ma), comprises four stages. Stage 1 (~780760 Ma) was the development of a Centralian Superbasin which linked the Officer Basin with the other main central Australian Proterozoic to early Palaeozoic basins, and which were also

in close proximity to basins with similar geology in North America, Siberia and the Persian Gulf prior to the breakup of the Rodinian Supercontinent. Evaporitic units began to move soon after deposition and were responsible for some of the present structural architecture of the basin. At the end of this stage, ~100500 m of erosion may have occurred due to glaciogenic processes associated with deposition of the overlying Sturtian tillite. Stage 2 (~560550 Ma) marked the onset of compressional basin development, deposition of further Neoproterozoic sediments, and culminated in the Petermann Ranges Orogeny with extensive thrust faulting and up to 3000 m of erosion. Stage 3 (~536507 Ma) comprised Cambrian deposition in elongate troughs (possibly mildly extensional) halted by up to 2000 m of uplift associated with the Delamerian Orogeny. Some reactivation of earlier thrust faulting also occurred at this time. Stage 4 (~500360 Ma) comprised Ordovician to Devonian deposition as a thick wedge, thickening to the north against the Musgrave Block, and terminated by the Alice Springs Orogeny.

The lithostratigraphy of the basin is still relatively poorly known, due mainly to poor stratigraphic control and a multitude of sedimentary units with similar environments of deposition but of differing ages. The stratigraphy is complex, with at least 26 mappable formations and many members. The formations have been grouped to reflect the structural stages described above. The Callanna Group sediments comprise fluvial and aeolian sandstone (Pindyin Sandstone), and marginal marine evaporitic units (Alinya Formation and Coominaree Dolomite). The intervening Sturtian glacial sequence (Chambers Bluff Tillite) is found only in the northern margin of the basin. Both this and the Callanna Group sequences are terminated by volcanics, also known only from the northern margin of the basin. The overlying Lake Maurice Group comprises fluvio-deltaic to aeolian sediments (Tarlina Sandstone, Meramangye Formation and Murnaroo Formation), which is overlain by the lower Ungoolya Group of more marine origin (Dey Dey Mudstone, Karlaya Limestone, Tanana Formation and Munyarai Formation). The Narana Formation (upper Ungoolya Group) is interpreted to have been deposited as a submarine canyon fill. The basal Cambrian Marla Group comprises the aeolian to marine Relief Sandstone intercalated with the marine Ouldburra Formation. The overlying Observatory Hill Formation has been well studied and records a variety of palaeoenvironments from fluvial to alkaline playa lake. The upper Marla Group comprises shallow marine to fluvial sandstones (Arcoeillinna and Trainor Hill Sandstones) separated by the shallow marine Apamurra Formation. The shallowest sediments of the Officer Basin, separated from the Marla Group by the poorly known Kulyong Formation (volcanics and sands), are the marginal marine Mount Chandler Sandstone, Indulkana Shale and Blue Hills Sandstone of the Munda Group. The youngest unit in the basin is the Devonian non-marine Mimili Formation, restricted to the central Munyarai Trough. Previously there have been significant problems associated with a lack of biostratigraphic correlation tools for the largely non-marine and Precambrian succession in the Officer Basin. This has recently been dramatically altered by the demonstration of abundant acritarchs in the more marine parts of the sequence and establishment of five preliminary acritarch assemblages that can be used for Proterozoic well to well correlations. Other biostratigraphic tools that have been or may be used in the basin are stromatolites, marine invertebrates and trace fossils. Other techniques, such as geochronology, magneto- and chemostratigraphy, and eventand lithostratigraphy, have also proven useful for regional correlation. The Officer Basin has undoubtedly sourced oil as evidenced from the oil shows recorded, some of which have been proven from biomarkers to have derived from Neoproterozoic sources. Potential source rocks appear to be quite organically lean on average, but this is largely due to conventional sampling techniques not being suitable for the cyanobacterial mat source prevalent in the Officer Basin, where the source lithologies occur as thin laminae in a lean matrix. In the Siberian Platform, the source for the prolific oil and gas fields is believed to average only 0.3% Total Organic Carbon (TOC) and is of similar cyanobacterial origin. The main source rock formations in the Officer Basin

(Alinya Formation, Dey Dey Mudstone, Karlaya Limestone, Ouldburra Formation and Observatory Hill Formation) have average TOC between 0.2 and 0.42%, but locally may range up to 4.6%. Kerogen type is generally oil prone Type I to III, and maximum genetic potentials range from 0.91 to 7.34 kg hydrocarbon per tonne of source rock. Maturity of the Officer Basin sediments cannot be derived using conventional vitrinite reflectance measurements due to lack of land plant material in pre-Devonian rocks. Instead, the distribution of triaromatic hydrocarbons (methylphenanthrene index ---- MPI) extracted from source rocks and oil shows can be used and calculated as an equivalent vitrinite reflectance (VR calc). The source rocks in the Officer Basin are thermally mature, with large areas of both the Proterozoic and Cambrian source formations being within the present-day oil window (VR calc = 0.65--1.0), although some areas of the Munyarai Trough, Marla Overthrust Zone and Manya Trough appear to be over mature. There are difficulties in modelling maturity for the Officer Basin wells due to inadequate temperature, maturity and kinetic data. The Ouldburra Formation in the Manya Trough is calculated to have entered the oil window between 510 and 370 Ma, and entered the dry gas window at 315 Ma, where it remains at present. On the Ammaroodinna Ridge, the Precambrian succession entered the oil window at 570 Ma (and is still within it), and the Dey Dey Mudstone and Karlaya Limestone are presently just within it. In the Marla Overthrust Zone, the Dey Dey Mudstone and Karlaya Limestone entered the oil window at ~570 Ma and the wet gas window at 550 Ma. In contrast, the Observatory Hill Formation entered the oil window at 450 Ma, where it remains. At least eight reservoir horizons have been identified in the Officer Basin, nearly all of which are fluvial or aeolian arkosic sandstones with secondary porosity. Reservoir thicknesses are generally in the range 100--400 m, with the exception of multiple sandstone reservoirs in the Ouldburra Formation which average 4 m, but cumulatively may reach 100 m in individual wells. Carbonate reservoirs with vuggy porosity are also known from the Ouldburra. Core analysis of the sandstone reservoirs indicate excellent characteristics, w i t h a v e r a g e p o r o s i t i e s i n t h e r a n g e 1 0 --2 5 % a n d permeabilities up to 8000 md. The sands are generally clean with low clay content. It is considered that there is low risk associated with reservoirs or seals in the Officer Basin. There have been few valid tests of the variety of compressional structural traps that have been identified in the Officer Basin. Trap types consist of foreland and detached thrusts (including trapdoor structures), tilted fault blocks (associated with salt withdrawal) and salt walls, domes, etc. Potential also exists for subcrop unconformity traps, stratigraphic traps and palaeo-relief associated with submarine canyon cutting. Potential volumes of typical traps are very large, up to billions of barrels, and the smallest are of the order of several hundred million barrels (unrisked oil in place). Exploration and production economics have been modelled to quantify the minimum prospect size that could be targeted by potential explorers. This model accounted for both exploration expenditure prior to discovery (dry holes,

exploration seismic, etc.) and development capital and operating expenditures; a 12.5% real discount rate was used. At an oil price of ~US$25/bbl, a 5 million barrel (800 000 kL) oil in place field would be economic to produce and explore for, provided that the field was discovered within the initial five-year term of a PEL. If the oil price were US$18/bbl, the minimum economic field size would be 10 million barrels (1.59 million kilolitres) oil in place. In both these scenarios, oil would be trucked to market. If a larger field was discovered (50 million barrels (7.9 million kilolitres)), then construction of a pipeline to Port Bonython would be economic. The undiscovered resource potential of the basin has been assessed using Monte Carlo techniques, with a 50% probability of exceeding 500 million kilolitres (3 billion barrels) of recoverable oil. The range of estimates was a 90% probability of exceeding 260 million kilolitres (1.6 billion barrels) to a 10% chance of exceeding 744 million kilolitres (4.7 billion barrels). While this assessment may appear to be unrealistically large, other Precambrian and Cambrian petroleum basins (with which the Officer Basin has significant geological similarities) in the USSR (Moscow Basin and Lena-Tunguska Province), China and Oman have potential oil reserves up to 16 billion barrels. Six main reservoir plays (Pindyin Sandstone, Tarlina Sandstone, Murnaroo Formation, Relief Sandstone, sands of the Ouldburra Formation and the Arcoeillinna Sandstone) were assessed using relatively conservative input parameters, such as a wildcat success rate of 1 in 10 wells, gross reservoir thickness of 16 to 100 m, net to gross ratios of 35--87%, and reservoir porosities of 12--18%. The potentially productive area for each play is mapped using all available data on source rock distribution, maturity, reservoir distribution and probable depths to targets. Abbreviations used in this text are summarised in Appendix 1.1.

DISCOVERY OF THE OFFICER BASIN


Petroleum exploration in the South Australian sector of the Officer Basin from 1954 in essence has comprised basic surface mapping, exploratory and stratigraphic drilling, a few limited gravity and magnetic surveys, and cursory seismic traverses. The sedimentary basin underlies the Great Victoria Desert. The remoteness and desolation of the inhospitable, almost inaccessible and largely uninhabitable region hindered exploration and limited interest in its hydrocarbon prospectivity. But these were factors which led to a large part of the area being proclaimed as the North-West Aboriginal Reserve in 1921 and being used for the United Kingdom-Australia missile testing program from 1946 to 1980 and for the British nuclear tests from 1952 to 1963. One aspect of the preparation for the testing of long-range missiles and atomic weapons was the commencement of modern geological work in the region. Hydrological surveys by the then Mines Department assisted in establishing sites for these programs. As a consequence of this, and a growing interest in South Australias petroleum prospectivity, the first indications of the basin came through a Bureau of Mineral Resources (BMR) small-scale reconnaissance aeromagnetic survey in 1954, which indicated some thick sedimentary sequences (Quilty and Goodeve, 1958). Earlier announcements about the petroleum prospectivity of the far west and northwest of South Australia had not been very encouraging. For example, although there had not been any drilling for oil in the far west, in 1944 the Government Geologist, Dr Keith Ward, discounted the Precambrian rocks of the northwest because there were then no producing oilfields of Precambrian age in the world (Ward, 1944). The introduction of specialists and advisers with expertise in new exploration techniques, such as improved drilling standards and equipment (rotary and percussion drilling replacing cable tool drilling), geophysics (seismic, gravity and aeromagnetic surveys) and theoretical constructs for investigating buried structures rather than deposits (as the oil search demanded), was critical to the search after World War II (ONeil, 1982, 1995). Evidence of the new approach was seen in the efforts of the Frome-Broken Hill Co. Pty Ltd, in conjunction with the BMR, in the east of the State immediately after the war (ONeil, 1996a). Until the mid-1950s, the whole of the far west and much of the northwest was still known geologically as the Eucla 2 Basin. This comprised an area of ~388 500 km in western South Australia and southeastern Western Australia,

including 178 700 km of the Nullarbor Plain. The Mines Department commenced a geological survey, for minerals, in the northwest in 1953 under geologist Reg Sprigg, who later claimed to have referred informally to the area as the Officer Basin and that it was formally named in 1954 by F.H. Quilty and P.E. Goodeve (Sprigg, 1983). Although Quilty and Goodeve did report indications of a basin from the results of their survey across the area to as far east as Oodnadatta in May 1954, they did not refer to it as the Officer Basin. The separation of the basin into the Eucla and Officer Basins was, however, recorded in the 1959--60 Annual Report of the Mines Department. It can be assumed that the name of the basin is related to Mount Officer and Officer Creek, which extends from the Musgrave Ranges via Fregon west of Everard Ranges and into the northern part of the basin.

EARLY EUROPEAN EXPLORATION


In September 1873 during his second trip into the South Australian interior, Ernest Giles and another party member, William Tietkens, had an encounter with an estimated 200 male Aborigines. Shots were fired by the Europeans in retaliation for the spears thrown. The Europeans escaped unharmed; Giles did not record if there were any black casualties. However, he acknowledged that the usual cause of Aboriginal aggression throughout the history of land exploration by Europeans in Australia was white trespass on black land (Giles, 1889). He named the river at the scene of this confrontation The Officer; it was renamed Officer Creek six decades later. A hill to the northeast of the creek and west of Mount James-Winter was named Mount Officer ---- C.M. Officer was a contributor to the fund raised by Baron Ferdinand von Mueller in Melbourne for this expedition (Manning, 1990). Several exploration parties visited the far west and northwest of South Australia in the quest for water supplies, overland stock routes or pastoral land, and to establish lines of transport or communication. Some of these explorations were also made with an interest in the geology or mineralogy of the land. Edward John Eyres overland journey during 1840--41 to central and Western Australia attempted to link South Australia with its colonial neighbour. On the way, Eyre passed through the southern margin of the basin. In 1870, John Forrest, travelling Eyres route in reverse, was the first person to cross from west to east. Ernest Giles made five trips to the interior of Australia between 1872 and 1876, the first two (1872 and 1873--74) using horses and the next three camels. The ships of the desert were the preferred means of transport for most of the

explorations in the outlying regions. Giles was accompanied by Tietkens on the second to fourth trips: Tietkens himself continued exploring and later led the Central Australian Exploring Expedition in 1889 (Tietkens, 1890). It was in September 1873 during his second trip that Giles named Officer Creek. At that time an expedition by William Gosse was trying to find a route from the Overland Telegraph line (completed in 1872) to the west coast (Gosse, 1874). Gosse and his party reached Mount Davies in the Tomkinson Ranges in that August, followed a month later by Giles. In 1874, John Forrest travelled from the west to the Overland Telegraph line to the south of Oodnadatta across the area of the Musgrave Block. In the same year, J. Ross examined the area for fresh water supplies and explored for pastoral land west of Lake Phillipson (Ross, 1875; Jack, 1931). During Giles fourth expedition, which went from Beltana to Port Augusta to Perth during May to July 1875, he despatched Tietkens and Jess Young to examine the area north of Ooldabinna. Giles spelt this as Yooldil-beena, meaning swamp where I stood to pour out water. It was a native well northwest of Tietkens Wells, which was 70 km north of Ooldea. They went slightly north of latitude 28, between 130 and 132 longitude, to a point which was approximately the boundary of the Musgrave Block and the Officer Basin. Giles had hoped that they would find sufficient water supplies to establish a route from Fowlers Bay to the Musgrave Ranges, but this was not to be. Even Tietkens Wells, sunk in 1875, were abandoned as one was dry at 18 m and the other was very saline at 37 m. Tietkens returned to the vicinity to supervise further water drilling operations in 1879 (Gara, 1994). Ernest Giles fifth expedition, between November 1875 and April 1876, was from Geraldton in Western Australia, across the Gibson Desert to the Overland Telegraph line, and to Blinman in South Australia in an easterly traverse through the far northeastern reaches of the Officer Basin (including the Musgrave Ranges, Mount Officer, The Officer, Everard Ranges and Alberga River). Between 1883 and 1892, the South Australian Lands Department undertook triangulation surveys of the area and the colonial borders (Carruthers, 1892). Carruthers Depot was named after Jonathon Carruthers, the Departments surveyor. During 1891--92, the Elder Scientific Exploring Expedition under the sponsorship of Sir Thomas Elder and command of David Lindsay, and including geologist Victor Streich, crossed part of the area while travelling from Warrina, a railway siding on the Marree--Alice Springs railway line to Coolgardie, during an investigation into central and Western Australia (Lindsay, 1891; Streich, 1892). There were repeated requests from members of the public and through debates in Parliament for the northwest region in the vicinity of the Musgrave Ranges to be examined. This interest focused on the likely mineral (especially gold) potential of the area, the possible expansion of the railway network to the ranges and the expectation of unlocking more land for pastoralists and agriculturalists, in particular to help relieve the unemployment being endured in the 1890s depression (ONeil, 1982). Thus, Government Geologist H.Y.L. Brown was sent to Western Australia to examine the

Coolgardie region, which he visited between October and December 1895. He reported that the country north of the Nullarbor Plain adjacent to the Western Australian border was a continuation of the geological formation in Western Australia and so it could be gold prospective. But the prospecting to then had been cursory and the vast area to the Musgrave Ranges required a more thorough examination (Brown, 1896). Brown himself was an inveterate explorer and between April and June 1896 he explored the western part of South Australia from Ooldea north for ~160 km, returned to Ooldea and then went to Mount Eba and Marree (Brown, 1898--99). Other explorers or prospectors identified with the area include J. Lamb (1894), A.H. Warman (~1900) and W.J. Cockrum (~1900). The 1896--97 South Australian Stock Route Expedition, under Captain S.G. Hbbe with William Murray in the party, took nine months to travel from Oodnadatta to Kalgoorlie (Hbbe, 1897). Murray then recorded the position of The Officer more precisely (Hbbe, 1897). In 1900, a party from the North-Western Prospecting Syndicate of Western Australia reached Oodnadatta after suffering various hardships, including encountering strong resistance from Aborigines in South Australias northwest ( Advertiser , 11.4.1901). Between 1897 and 1903, Richard Maurice undertook at least eight expeditions relating to the Great Victoria Desert while living for most of this period at Pidinga, 128 km northwest of Fowlers Bay. He recorded data on the plants, animals and landscape as well as collecting mineralogical, biological and ethnographical specimens. During 1898--99 he travelled north from Fowlers Bay across the desert, through the Everard Ranges to Stuart (later Alice Springs), and west to Hermannsburg Mission Station in the Northern Territory. His 1901 expedition was accompanied by William Murray. This search for water and pastoral land included some mineral examination, and attempted to fill in the gaps between the expeditions of Giles and Lindsay. They crossed southwest to northeast from Ooldea to Tallaringa Well (an Aboriginal soak), Oolarinna Spring and the Everard Ranges, and back to Ooldea by a westerly route. Then the party went from Ooldea to Punthanna Native Well and west-northwest past Pat Aulds Vat to the western border (Jack, 1931). During this exploration they saw evidence of the visits by the Western Australian prospectors, Warman, Cockrum, and the Lindsay and Hbbe expeditions. The party camped on The Officer while travelling to Oolarinna and Ooldea; and again when they went to the Northern Territory in 1902 (George, 1904; Advertiser , 11.4.1901). In 1902, Maurice and Murray went north from Fowlers Bay to the Kimberley district in Western Australia. The mineral specimens they collected were given to the Mines Department for analysis. The South Australian House of Assembly passed a motion in December 1901 that the Government undertake a mineralogical examination of the country between Tarcoola, the Musgrave Ranges and the northwest of the State. In the following December, the Government North-West Prospecting Expedition, including Lawrence Wells, Herbert Basedow and Frank George set forth (Wells and George, 1904; Basedow, 1905, 1914). Between April and September 1903 the party journeyed northwest from Oodnadatta to Todmorden, Alberga River, Mount Mystery, Krupps Hill

and Indulkana (the far northeastern extent of the basin), and returned. The expedition helped to fill in gaps of the Elder and Horn expeditions; the Horn Scientific Expedition had gone to central Australia from Oodnadatta in 1894 (Winnecke, 1896). The North-West Expedition found little conclusive evidence of valuable minerals but H.Y.L. Brown doubted its results because of the short time available to prospect properly. Instead he relied on his own impressions of the region and of geology in suggesting that gold and other mineral discoveries would be made there. In April 1904, Brown again left Adelaide and visited the region from Lake Phillipson to the Wildingi granite (Brown, 1905; Jack, 1931). His route took him slightly west of Mount Byilkaoora and north to the Northern Territory border, Stuart Range area, Todmorden, Indulkana, Alberga and Bitchera Creek. Using Maurices camels and equipment, Frank George (then a Mines Department surveyor) led a prospecting expedition in 1904 from Fowlers Bay to northwest of Lake Dey Dey and across the State border to the Boundary Dam salt lakes in Western Australia. He named Lake Maurice, which had been discovered but not named by Ernest Giles in 1875, and reported that gold and other metallic mineral discoveries were unlikely in the region (George, 1905). Another of his Government prospecting expeditions, during 1905--06, went to the southwest of the Northern Territory; upon Georges death, Murray (then with the Mines Department) took charge and the party went to the Buxton and Davenport Ranges from Oodnadatta via Todmorden and Indulkana (George, 1907). Between 1912 and 1917, the east--west Trans Australian Railway across the Nullarbor Plain was constructed by the Commonwealth Government in the southern part of the basin. The geological and surveying work for this project was naturally directed towards ensuring the best possible route for the line. The year 1914 was one of severe drought in the State but it was at the end of this year that the Assistant Government Geologist, R. Lockhart Jack, made the hazardous trek to the Musgrave Ranges. The expedition was a geological survey and an examination for water supplies, minerals and possible pastoral land south of the ranges. The noted ornithologist, Captain Samuel White, accompanied Jack to Todmorden,

96 km northwest of Oodnadatta where the Government Astronomer, George Dodwell, joined the group (AR*, 1914; Jack, 1915). Jack was unimpressed with the mineral possibilities there. Also during that year a petroleum specialist, Arthur Wade, was appointed by the State Government, at the request of several parties interested in discovering petroleum, to investigate several supposed oil-bearing areas in the State (Wade, 1915). Wades brief examination of lower Eyre Peninsula, essentially the coastal region, extended west to Streaky Bay. He reported in passing that gas in connection with mound springs, petroleum-like streaks on Streaky Bay and a bitumen discovery at one locality were not evidence of oil or gas seeping to the surface. His expectations of the States petroleum potential were not encouraging and he dismissed the possibility of petroleum supplies being found in this region in particular. For the next 40 years there was little activity or even interest in the mineral or petroleum prospects of the basin. In 1917, Talbot and Clarke from the Western Australian Geological Survey travelled east as far as Mount Gosse and the western Musgrave Block. In 1925, Ward and Jack visited an area lying beyond the occupied pastoral holdings in the northwest and reported on prospecting for water there, and two years later they inspected areas of the far west where water supplies might be located (AR, 1925, 1927). Prior to leaving the Department in 1931, Jack investigated the geology north and northwest of Tarcoola and went to Wilgena, Commonwealth Hill, Coober Pedy, Tallaringa Well amongst other sites on the periphery of the basin (Jack, 1931). Several of these explorers are renowned for the heroic and epic nature of their journeys. Nevertheless, their endeavours rarely met their expectations: the land was not conducive to supporting large populations, it was not good enough for pastoral and agricultural pursuits, nor did it provide substantial water supplies. Just as these desired outcomes were not achieved, the mineral and petroleum potential remained to be tested more thoroughly. From the mid-1950s, however, geological, geophysical and drilling exploration were to generate limited information from sparse programs.

WEAPONS TESTING
As part of a fledgling effort to bring South Australia into the atomic age, such was the interest being generated by the States uranium supplies at Mount Painter and Radium Hill from the mid-1940s, the desert then seemed an apt site for tests on long-range weapons and atomic bombs (Morton, 1989; ONeil, 1996b). Although they originated as discrete projects, the tests overlapped in the initial period after the atomic tests were moved to the Australian mainland in 1953. In the context of limited resources, the harsh environment and the same British, Australian and South Australian political masters, it made practical sense to share equipment, facilities, communications, transport and working time (Morton, 1989). For example, the Native Patrol Officers scoured the desert in respect of both projects as did Len Beadell in establishing sites and routes, planes from the Maralinga

Lockhart Jacks party leaving for the Musgrave Ranges from Wantapella Well on Indulkana in 1914. The Government Astronomer, George Dodwell, is in the lead with his dog, Speck, not far behind.
(Photo N679)

* Annual Report, under varying titles, issued by the Mines Department.

atomic tests were decontaminated at the long-range weapons base at Woomera (named after a non-local Aboriginal spear throwing and carrying implement) and the Mines Department conducted groundwater investigations and drilling for both projects. A major missile testing range operated in South Australia from 1946 to 1980 under joint United Kingdom and Australian control. The general firing direction was from Woomera over the South Australian desert towards the northwest of the continent. Thus, part of the Officer Basin fell within the Woomera Prohibited Zone, an area under the control of the Australian Government, to which access was restricted. All people were required to be issued with a permit from the authorities at Woomera before entering this area and movement within the area was often under scrutiny, at least in theory if not in practice. In 1954, for example, when the Australian Mining and Smelting Co. was considering a work program as part of its licence commitments, Premier Tom Playford wrote to Prime Minister Robert Menzies on behalf of the company to secure access to the zone. On the mainland, two atomic bombs were exploded in 1953 at Emu Field, ~500 km west of Woomera and 250 km west of Coober Pedy, in the prohibited zone and in the firing line of the rocket range (Symonds, 1985). The first (Totem 1) was exploded at Emu on 15 October 1953. Some Aborigines have attributed a black mist passing over Wallatinna and Welbourn Hill from this test as a cause of their ill health (Lester, 1993). Emu, a remote claypan, was difficult to access and the logistics of the exercise meant that after a second test there, the atomic program was moved to a site 80 km north of Watson on the Trans Australian Railway. The new site, Maralinga (an Aboriginal word meaning thunder), is 177 km south-southwest of Emu. Seven bombs were tested there between September 1956 and October 1957. Minor trials of small nuclear weapons, which dispersed plutonium, continued at Maralinga until 1963 and the site was abandoned in 1966. Clean-up exercises in 1964 and 1967 were found by the Australian Royal Commission into British Nuclear Tests in Australia (1984--85) to have not only been ineffectual but also made the site more dangerous. As a result of this most recent inquiry, MESA has been associated with preparations for the current clean-up operation. The Mines Department undertook geological and drilling work for both atomic testing projects. This began in 1947 when Assistant Government Geologist Tom Barnes prepared a hydrological survey with some geological work in the northwest for the Long-Range Weapons Project (AR, 1947). A preliminary data search by the Department for the area of the atomic tests included checking the region west of the north-south railway line to the Western Australian border, and north from the Trans Australian Railway to the Northern Territory border (ONeil, 1995). Over the years (and even after he became the Director of Mines in 1956), Barnes continued his special hydrological investigations. Through the 1950s, percussion-drilled deep and shallow bores were sunk at Maralinga for underground water and for special purpose bores. The water work continued into the 1960s.

The Federal Governments Petroleum Search Subsidy Scheme from 1957 subsidised stratigraphic drilling but required the results of work programs to be published or else the funding was likely to be refused (Passmore, 1994). As exploration in the Officer Basin required clearances from the authorities in charge of both Woomera and Maralinga, security concerns meant that permission was also necessary before results could be published.

POST-WORLD WAR II GEOLOGICAL EXPLORATION


In 1953, a Departmental party led by Reg Sprigg and including Ron Coats went to Mount Davies on a mineralogical expedition essentially to investigate uranium deposits, but some of the samples collected showed traces of nickel. (A gold lease had been pegged in the area by the prospector Cockrum in 1900.) In 1954, Special Mining Lease 20 was pegged by Gold and Mineral Exploration NL in the Mount Davies area of the Musgrave Block adjoining the Officer Basin. The lease was taken over by Southwestern Mining Ltd in 1955. In this period, the Department sank eight bores for private hirers in the vicinity, of which three were productive (AR, 1955). Nickel mineralisation in the northwest province continued to attract attention into the next decade. In November 1965, a low-level survey of parts of the MANN and WOODROFFE 4-mile sheet areas was flown by the Department as part of an investigation into basic and ultrabasic rocks in the area, which was the prime focus of the regional mapping and nickel investigations. Regional geophysical surveys were conducted spasmodically in the Officer Basin after the first aerial survey in 1954. Aeromagnetic surveys over the far northwest, including the Tomkinson, Mann and Birksgate Ranges, were flown by BMR in June 1960. The Mines Department was mapping in the field: this focus was essentially mineralogical, with nickel in the Tomkinson Ranges being the target. Nevertheless, in 1960, the first field work on petroleum in the basin commenced. On behalf of Exoil, the Departments Geophysics Section ran a single ground traverse of gravity and magnetic observations from Ooldea north along the 131 line of longitude across the Officer Basin using a helicopter to transport the equipment. The survey, which gave further evidence of a basin, was conducted along the eastern perimeter of Maralinga because permission was refused to enter the area (Mumme, 1961). Petroleum industry exploration is reviewed in a section below. The following refers to some of the State Government w o r k . D u r i n g 1 9 6 1 --6 2 , a D e p a r t m e n t a l g e o l o g i s t accompanied a brief aerial reconnaissance of the southern Officer Basin and part of the Eucla Basin with Exoil representatives. A Departmental seismic crew was operating west of Emu Field: international seismic crews were also introduced to the basin under contract for reflection and r e f r a c t i o n s e i s m i c s u r v e y s . D u r i n g 1 9 6 4 --6 5 , t h e Departments Seismic Geophysics Section spent 3 1 2 months in the South Australian part of the Eucla Basin on reconnaissance work and detected a sedimentary trough at a much greater depth than anticipated. A seismic reconnaissance was made from Maralinga to north of Emu and then from Emu to Mabel Creek (AR, 1965--66).

10

outcrops. Reconnaissance work, including two regional helicopter surveys of the Officer and Eucla Basins during 1968--69, helped to define targets and further delineated the margins of the basins. From the early 1970s, the Department also undertook stratigraphic drilling, including Mount Willoughby 1 (November 1970), Wallira West 1 (Arckaringa Basin, March 1971), Marla 1 and Manya 1 (September 1974), and Murnaroo 1 (November 1976). Marla 1 and Manya 1 were on the far eastern perimeter of the Officer Basin. Also drilled were Marla 1A and 1B; drilling problems caused Marla 1B, ~20 m from Marla 1A, to be abandoned at 379 m (Pitt et al. , 1980).
Refuelling the helicopter during the Officer Basin gravity survey in May 1962. (Photo T3372)

Although the Department was then scaling down its large-scale seismic operations, it ran a seismic survey in the eastern Officer Basin in July 1966. The combined seismic reflection, seismic refraction and gravity survey was run as a single traverse across a marked aeromagnetic low beginning at Emu 1 and extending 122 km north to the lower reaches of the Officer Creek. This was considered as probably the most difficult seismic traverse the Department had ever undertaken as the dune-covered terrain comprised east--west trending, scrub-covered sand ridges which made access difficult even with the use of a bulldozer. The mobile camp was serviced with fresh water from either Maralinga or Everard Park, making for long hauls. The water for drilling was saline, and was obtained from shallow bores near Emu Field and bulldozed soaks 80 km north of Emu. The reflection shooting indicated a probable thick sedimentary section in the north of the basin (AR, 1966--67; Wopfner, 1969, 1970). During the 1960s, the Department was concerned to increase the quantity and quality of exploration in the basin and to expand the evolving knowledge of the basins geology. The first significant geological traverse of the Tallaringa Well area was made in 1966 by Departmental geologists Bruce Webb and Bryan Forbes, although Brown and Jack had made the first geological investigations there decades before (Benbow, 1993). In the mid-1960s, the Department re-examined all of the rock samples from the basin and used the services of Amdel for petrographic testing of new samples. Departmental petroleum and regional mapping geologists visited the basin to investigate sedimentary rock

However, the Departments basic exploration work then was often curtailed because of funding constraints. The cessation of field work in the basin came at a critical time; the similarity between the sequences in the basin and those at Palm Valley and Mereenie in the Northern Territory required further assessment. During 1971--72, the drilling of a major well on the southern shelf of the Munyarai Trough was not carried out due to lack of funds, but a ground traverse was conducted between Everard and Watson, and some experimental seismic work was recorded from Wallatinna Waterhole into the Munyarai Trough. During 1972--73, seismic profiling produced poor results in a narrow extension of the basin in the east; on the northern margin, south of the Everard Ranges, gravity, magnetic and seismic data were recorded. In 1974, the Department continued its seismic operations in the eastern Officer Basin and recorded 140 km. Mapping and drilling in the Musgrave, Everard and Tomkinson Ranges in the northwest (particularly at Mount Davies for nickel) continued during the 1960s and into the 1970s. But, by the mid-1970s, access to the area for the Department and mining and petroleum companies had become increasingly more restricted. Despite the promising indications of potential mineral wealth in the area and the vastly underexplored petroleum potential, Departmental and company work programs ground to a halt there (ONeil, 1995).

ABORIGINAL LAND RIGHTS AND EXPLORATION ACCESS


The Aboriginal occupants of the area had several different clan backgrounds but they are now labelled broadly as Pitjantjatjara and Maralinga people, though within these groups there are strong tribal and regional differences: for example, According to missionary Violet Turner, Ooldea Soak ... was visited for ceremonial purposes by Kukatas from Tarcoola, the Minnings from Eucla, the Aluridjas from the Musgraves, the Wongapitchers from the Mann Ranges. (Mattingley and Hampton, 1988, p.235). However, for much of the period, the Aborigines were treated as a monocultural race and this is evident in the promotion of reserves for their protection and improvement or, in some cases, to ease their presumed passing. Daisy Bates, at Ooldea from 1918 to 1934, was one who attempted to retain the Aboriginal peoples independence.

The Officer Basin seismic survey, looking south along the survey line from shot point EO 155, in 1966. (Photo N16912)

11

For its part, the Mines Department had elsewhere confronted the question of Aborigines, their relationship to the land and mining by Europeans. In 1905, the Department reserved land near Parachilna for Aborigines to use in their traditional way. Another instance occurred after Jacks trip in 1914 to the Musgrave Ranges; in 1919 he reported on the desirability of establishing a reserve for Aborigines in the northwest (Jack, 1919). An important development followed when in 1921 the State Government gazetted more than 56 700 km 2 of the land as the North-West Aboriginal Reserve to safeguard the Aborigines against encroachment by white people, be they settlers or transient interest groups such as miners and prospectors. Additions were made to the reserve in 1938, 1949 and 1974; by 1949, the area proclaimed in South Australia was 71 500 km 2, which with additional areas in Western Australia and Northern Territory formed the Central Aboriginal Reserve. Establishing reserves for Aborigines began soon after South Australia was colonised, but those reserves were intended to be places where Aborigines would be Christianised and civilised. The North-West Aboriginal Reserve, however, was in part more of a case of out of sight, out of mind, as was the nearby Ernabella Mission (run by the Presbyterian Board of Missions) which combined the religious theme with an encouragement of Aboriginal culture after it was established in 1937 in the Musgrave Ranges (ONeil, 1995). The United Aborigines Mission, founded at Ooldea in 1933, closed in 1952 and the people were moved to Yalata Aboriginal Reserve, ostensibly because of the impending long-range weapons and atomic tests. Amata (then Musgrave Park, 130 km west of Ernabella) was founded in 1961; later reserves were established at Fregon and Indulkana, the latter in 1968 from a small section of the Granite Downs pastoral lease. The reserves were not inviolate for all time and they were subject to European intrusion for defence, mineral and petroleum exploration, and geological mapping. However, companies wishing to work in the North-West Aboriginal Reserve had to observe rather strict guidelines in their relations with the local Aborigines. In 1954, when the Australian Mining and Smelting Co. was proposing its licence commitments, the Director of Mines wrote to the Aborigines Protection Board to request information on access conditions. The company subsequently advised him that it did not intend to enter any areas covered by the Aboriginal reserves although it did not have a map showing these (MESA file 2658/1953). The South Australian Governments Aboriginal Lands Trust Act 1966 , prepared by Don Dunstan when he was Attorney-General and Minister of Aboriginal Affairs from 1965 to 1967, set the pace for the nation on legislation and policies on Aboriginal affairs. A Government proposal to provide compensation to the Aboriginal people to ensure to them control of mineral rights in any lands held as Aboriginal lands was defeated in the Legislative Council. The Government instead signed an Indenture with the Aboriginal Lands Trust to the effect that all royalties for any mining on Aboriginal Lands Trust land would be paid to the Aboriginal Lands Trust.

1. Each member of any party entering the reserve to supply two personal references of character from reputable persons preferably Justices of the Peace, Government Officials, Ministers of Religion, the Director of Mines [or the companys Managing Director]. 2. Each member to supply a medical certificate to the effect that the person is in good general health and is not suffering from any contagious disease. 3. That the leader of any party give an assurance that he will do all in his power to prevent any members of the party from clashing with the aborigines, will not encourage or permit the aborigines to congregate near any camp and that the leader will accept the responsibility of seeing that none of the party are intimate with female aborigines. 4. That no member of the party shall remove from the reserve nor trade with the aborigines for any ethnological specimens and shall not distribute to the natives any goods or chattels by way of barter or exchange. 5. That the leader of the party shall supply the Aborigines Protection Board with regular reports covering the numbers and conditions of any aborigines encountered and places where they met, any incidents which occurred between the party and the aboriginals, and anything of general interest to this Department.

On 1 April 1954, the Secretary of the Aborigines Protection Board wrote to the Director of Mines (MESA file 2658/1953) stipulating the guidelines for persons wishing to enter an Aboriginal reserve on behalf of the Australian Mining & Smelting Co. Ltd.

Under the Act, the Aboriginal Lands Trust was authorised to hold land titles in trust on behalf of Aborigines in South Australia. The Act did not permit Aboriginal land owners to negotiate mining agreements; the M ining and Petroleum Acts applied to Trust land but could be subject to additional conditions. The Act was amended in 1973 to permit Aboriginal communities to control mining and exploration on Trust land, which was then exempted from the M ining and Petroleum Acts ; exploration and mining activities could not be prohibited, but the Trust and an individual community could stipulate conditions of access and operation, and the Governor could proclaim these special conditions. The Government could grant money to the Trust from exploration lease payments or mining royalties obtained through activity on Trust land. However, the Trust preferred to seek more titles to land rather than engage in managing the titles acquired, an activity which, in its view, was the responsibility of the Aboriginal communities concerned. Up to 30 June 1984, the Trust held 485 585 ha of land; as there had been no mining on this land, no mining royalties had been paid (Mattingley and Hampton, 1988). The North-West Aboriginal Reserve included part of the Officer Basin which, in contrast to Jacks earlier views, was now considered to have mineral and petroleum prospectivity. Since World War II, the Department had sought to promote exploration opportunities in the area by mapping and drilling, especially in the 1960s, and companies such as Southwestern Mining, Exoil and Conoco had explored in the area too. In 1972, the Governor of South Australia, Sir Mark Oliphant,

12

the land would be open under protection so that the Department and companies could go there amicably by agreement. An industry and Departmental preference for completely unrestricted access was not likely. The crews would occasionally help Aboriginal groups during Departmental field trips to the Musgrave Ranges area of the reserve. In September 1976, Departmental geologist Colin Gatehouse and other Government officials consulted with the Aboriginal people, and at Coffin Hill (on the northern perimeter of the Officer Basin) Gatehouse completed some jobs for the Aborigines. During the course of their discussions with him concerning exploration and mining in that area, the tribal elders made known their wish for the land to be transferred to them. There was increasing public interest in handing back land to the Aboriginal people (Toyne and Vachon, 1984). Based on experiences of Aboriginal communities in the Northern Territory, particularly through the impact of mining companies and their activities, the Pitjantjatjara and Yankunytjatjara communities in the northwest of the State formed the Pitjantjatjara Council in 1976. The Department had already experienced some difficulty in getting its scientific parties into Pitjantjatjara country, and this was a warning of tougher times to come. The council pursued its claim to land title without reference to the Aboriginal Land Trust, arguing that, because the Pitjantjatjaras link with the land and their lifestyle had been maintained, then their land should not be subject to other Aboriginal people who had lost their traditional ways. Premier Dunstans election policy in 1977 committed the Government to implementing land rights for the Pitjantjatjara people and, in November 1978, Parliament considered the Pitjantjatjara Land Rights Bill . Its passage was delayed by the Premiers resignation because of ill-health and the subsequent electoral defeat of his successor in September 1979. A modified version of the Bill was finally passed by the Tonkin Government on 4 March 1981, though it was not proclaimed until October of that year.

A Mayhew drilling plant at site 4 on the Indulkana Aboriginal Reserve, July 1970. Ordovician sandstone of the Indulkana Range is in the background to the north. (Photo T12568)

included the Musgrave Ranges, Ernabella and Everard Park in a tour that he made to inspect the Aboriginal situation in the State. An Australia-wide campaign for Aboriginal land rights had been stimulated and intensified by the Aboriginal tent embassy on the lawns of Parliament House in Canberra between January and July 1972. Almost immediately upon the election of the Whitlam Federal Government in December 1972, A.E. Woodward QC was commissioned to inquire into Aboriginal land rights in Federal territories. In May 1974, the Woodward Commission recommended that Aborigines should receive title where traditional land ownership in Aboriginal reserves and other unalienated Crown lands could be shown, and in alienated land where traditional land ownership could be demonstrated or if title was socially and economically desirable. The Commission considered that mining on reserves should be allowed but only with Aboriginal consent, except where mining was in the national interest. Aboriginal claims to mineral rights were rejected. Following the States pioneering legislation on Aboriginal land and in accordance with the Woodward Commissions recommendation, in May 1974 the Dunstan Government announced an important initiative. In respect to mineral exploration and mining, Aborigines living in the area concerned were to be consulted so that they would fully understand the proposed work and, upon consenting to it, could participate in the venture to the fullest extent that they were able. Only in matters of State or national importance would Cabinet consider over-riding this (AR, 1973--74). Government officials met with the Aboriginal people at Yalata early in July 1974 to discuss implementation of the Governments rules. Meanwhile, active exploration was held over in 1974 and 1975 when the Pitjantjatjara people and the Departments of Mines and Community Welfare undertook to identify and locate Aboriginal sacred sites in the reserve. In a sense this continued the concern for the local Aborigines that had led to the Reserves creation. For example, seismic lines were to be positioned so that sites of significance would be avoided, training and employment were offered to some Aboriginal people to work with the exploration parties, and

The Mines Department drilled for water in and around the NorthWest Aboriginal Reserve in 1970 and 1971 on behalf of the Department of Aboriginal Affairs. Four engine-driven and seven hand-operated pumps, four windmills and storage tanks were installed to provide water supplies for the Pitjantjatjara people. Being erected by the Department at Indulkana in 1971 were this 5000 gallon tank on a 30-foot stand and 48-foot windmill. (Photo
T21773)

13

Monster of Community Welfare Dear Sir We want this areas. So we can try to get this land for us. So we have meeting at Coffin Hill all old people want this Land this is what they want. 39 Gilpi wati have been this meeting and we been writed the Line on the map showing or country Some Old Gilpi write name on map to show they true country and after this meeting we want Mr Busbridge and Mr Nicholas to come to Indulkana for meeting so we can have meeting in Dec the 6 and we want this country for all the people with name underneath Manage Iltur Community. Punch Thompson Windlass Jimmy Stewart Mike Punch Jack Cox Jack Windlass Ray Ayaiya Paddy Charlie Tambo Tommy Micky Norman King Everard Jimmy David Taylor Con Larry Old Everard Mike Taylor Jim Pingey Willy Willy Murray Dan Murray Harry Wiland Charlie Tunkin Wilbur Brooks Mark Anderson Tommy Queama

The original Pitjantjatjara Land Rights Bill h a d established Anangu Pitjantjatjara (Pitjantjatjara people of central Australia) as a legal entity with title to nucleus lands, that is the North-West Aboriginal Reserve and any pastoral leases in the region held by Aborigines. It also proposed a right to claim non-nucleus land, such as nearby pastoral leases held by non-Aboriginal people, and a right to ban exploration and mining on their land. The Tonkin Government objected particularly to these latter proposals and, in February 1980, Premier David Tonkin announced that mineral exploration would be allowed on non-nucleus lands. In addition, the Government announced there would be a working party to register sacred sites on Pitjantjatjara land but this idea was unacceptable to the Pitjantjatjara Council and the sacred sites working party proposal lapsed without the party holding a meeting. Further negotiations resolved the claims to the non-nucleus lands, the controls over exploration and mining, the rights of opal miners at Mintabie, and the Granite Downs lease. The Government incorporated significant concessions to the mineral ownership, royalty and disturbance provisions of the Bill. The inalienable freehold title to 102 630 km 2 of arid Reserve land and vacant Crown land was vested in Anangu Pitjantjatjara. Some non-nucleus land at Granite Downs was to be included in the deal, from 2008 when the leases expired, but no further claims were to be allowed there, nor could other vacant Crown land on pastoral leases outside the Reserve be claimed. Exploration and mining were not banned, but the Pitjantjatjara people could stipulate conditions of access and operations and were entitled to compensatory payments for disturbance to their land, people and way of life. The Minister of Mines and Energy retained the right to appoint an independent arbitrator if exploration or mining was vetoed or allowed but subjected to conditions to which the mining company objected. The economic significance of a mining project to either the State or Australia would be a prime consideration for over-riding the Pitjantjatjaras interests. The opal miners were granted a 21-year lease for the township of Mintabie. The transfer of the land title to Anangu Pitjantjatjara was made at a ceremonial occasion presided over by Premier Tonkin on 4 November 1981 near Ernabella. Yami Lester (1993, pp.148-149) recalled the occasion:
Ill always remember the premiers speech that day. He said: All the world and the people of South Australia are watching you and what youre going to do with this land. So I often remind Anangu about that: Theyre watching you. If we dont do the right thing on this land, the Government is always watching to take it away. And thats the thing I cant understand. Its hard! We know this has always been our land. We got the stories, we got the culture; we got the language; we got the Law -- our own Law. So why is the white man saying were watching you, watching what youre going to do? The land was ours all the time. OK, we didnt have a piece of paper, but it was still ours.

Bob Jones Jack Baker Harry Joe Windlass Jimmy N. Killy B.

Transcript of a letter written to the Minister of Community Welfare, Ron Payne, by a group of Pitjantjatjara males (Gilpi -- tjilpi -- an old man or elder of the group; wati -- a man) at Coffin Hill in the North-West Aboriginal Reserve on 27 September 1976. Not all signatures could be transcribed accurately and the second list provides additional names of some of those who signed the map of the North-West Aboriginal Reserve.

The Ministers of Mines and Energy wanted to ensure that the Bill would not totally block that land off for any future exploration and development; a lengthy process of protest, extensive negotiation and consultation about the legislation with the Aboriginal people ensued. Although successive Governments supported the concept of land rights, they opposed sterilising huge tracts of land from any activity at all for all time. The Departments previous investigations in the Pitjantjatjara Lands, in particular into the potential for petroleum, had aroused interest from companies and, with the land rights issue apparently resolved, access to the Officer Basin was encouraged by the Department, especially because of traces of oil bleeding from Byilkaoora 1, a stratigraphic well drilled in May 1979 at Mount Byilkaoora in the northeastern reaches of the Officer Basin. Although the Minister had previously called for expressions of interest in the area from companies, it was still too early to become excited about oil because the Pitjantjatjara were concerned with winning their battle for land rights, and a proposed exploration program was continually deferred.

While the Act recognised Aboriginal interests, the subsequent closure of the land to exploration, and often to entry, indicated that in practice the mining and petroleum industries were discouraged.

14

During the process of proclaiming the Pitjantjatjara Land Rights Act , the Department had eight applications involving 16 companies under consideration for exploration in the area. However, after so long in the balance, the Governments immediate moves to encourage exploration on Pitjantjatjara land did not give the Pitjantjatjara time to adapt to the new regime. In 1980, a consortium led by Haematite Petroleum Pty Ltd (the precursor of BHP Petroleum) had already applied for a Petroleum Exploration Licence (PEL) in the Officer Basin in Pitjantjatjara land. After about a year of negotiating the process broke down when discussing the question of up-front payment of compensation for exploration work (the seismic lines, roads, airstrips, etc.) for disturbing the homelands out from Indulkana, Mimili and Fregon. The Bannon Government, elected in November 1982, maintained support for land rights but downplayed the significance of the mining sector and so indicated to companies to give the Pitjantjatjara time to think through and discuss the issues. The Act was now being interpreted in a way which had not been intended so that exploration could be frustrated or denied. But the outcome was one which really pleased neither the Government nor the Department. Another decade was to pass before real progress was made to resolve land access issues. While the Act specified compensation for disturbance, Haematite refused to budge on the issue of paying compensation up-front for disturbance to the land during the initial exploration. Claiming that arbitration would involve expensive legal fees, Haematite also did not want to establish a precedent for the mining or petroleum industries by paying compensation (McRae et al. , 1991; Advertiser , 11.7.1984; MESA confidential file SR 27/2/43). The Governments understanding was that there would be no up-front payments for exploration; compensation clauses would apply if a resource was found. In July 1984, the Haematite consortium withdrew its licence application. The Maralinga Tjarutja Land Rights Act 1984 , although similar in many respects to the Pitjantjatjara Land Rights Act 1981 , modified the conditions for mining on Aboriginal land; exploration was defined more clearly than under the Mining and Petroleum Acts, sacred sites were excluded from initial

Premier Tonkin symbolically proclaiming the Pitjantjatjara Land Rights Act at Ernabella on 4 November 1981. (Photo T23104)

tenements, the Minister of Aboriginal Affairs was to be involved in decision making, and compensation for work or disturbances in the exploration phase was limited to that allowed in the Mining and Petroleum Acts (McRae et al. , 1991). The new Act generally limited the Yalata (southern Pitjantjatjara) Aborigines influence over mining on their land. A resort to arbitration was retained at the production stage but the requirement for the Yalata Aborigines consent for access to their land was lessened. Compensatory payments to the Yalata Aborigines were maintained for exploitation agreements. The freehold title to the 76 420 km 2 of Maralinga Tjarutja land was granted on 6 December 1984; the land grant document was presented ceremoniously on 18 December 1984 near Maralinga. While it was fine in theory that companies and Aborigines might negotiate successfully, there was a substantial degree of difficulty in reaching agreement, especially as the view of the industry at that time was that petroleum and mineral potential was not high. Although relations between the Pitjantjatjara people and the Department could also become tense on occasions, the Department became more attuned to negotiating and assisting Aboriginal groups, and the mid-1980s saw remote Aboriginal communities provided with better electricity supplies and water wells being drilled by the Department on Aboriginal land. By then the higher world oil prices of the early 1980s had again spurred petroleum exploration in the Officer Basin.

South Australian Premier David Tonkin and Pantju Thompson on 2 October 1980 signing an agreement enabling the Pitjantjatjara Land Rights Bill to proceed. (Photo 32375)

15

The first PEL granted within Pitjantjatjara lands was issued in November 1985 to Amoco Australia Petroleum Co. (50%; the operator), AP Oil Pty Ltd (20%), Crusader Resources NL (15%) and Quadrant Energy Development Ltd (15%). PEL 29 covered 20 749 km 2 of the Officer Basin in the southeast of the Pitjantjatjara Lands. The Amoco group reached an agreement with a company owned by the traditional owners to become involved in a joint PEL. After the Haematite episode, the Pitjantjatjara Council had applied for a PEL in October 1984. The Department advised the Minister to refuse the request because the application failed to demonstrate that the Council had the financial resources and technical expertise to fulfil the licence conditions. The Council acted on suggestions from the Departments Oil, Gas and Coal Division and pursued expressions of interest from recognised petroleum companies. The Pitjantjatjara Council formed AP Oil Pty Ltd and then negotiated to share in a licence with the Amoco consortium and an amended application was resubmitted. The agreement with the Amoco consortium involved the group funding the Councils administration expenses and included the novel provision for AP Oil to hold a 10% vote in the management of the operation. In return, the Council exercised its statutory right not to claim compensation for disturbances to the local community during exploration and production. The agreement also provided that AP Oil would hold a 20% carried interest share in the licence during the exploration phase. This did not include any contribution to exploration costs. On the discovery of an economic petroleum deposit, AP Oil had the right to convert this carried interest to a 20% working interest on the payment of 40% of the cost to date. Alternatively, they could choose to retain a 10% net profit interest in the discoveries. This agreement was an interesting development in Australian petroleum exploration as it predicted a way forward for mining and exploration on Aboriginal land. However, the project stalled in October 1987 when the licence was cancelled due to the effects of the oil crash; one of the joint venturers collapsed following the fall in world oil prices in 1986. Anangu Pitjantjatjara remained prepared to enter joint ventures with companies which combined technical expertise with a respect for the Pitjantjatjara interests and concerns. The Department continued to signal its desire to have Aboriginal land opened up for mineral searches. For example, in 1989 it re-assessed the mineral potential of the Maralinga lands (Flint et al. , 1989) and late in 1991 a data package titled G e o l o g y a n d m i n e r a l p o t e n t i a l o f t h e Pitjantjatjara lands was released. These areas were the least examined mineral frontiers in the State. Departmental officers had continued talking with Anangu Pitjantjatjara, who seemed to favour oil exploration over mineral exploration, since petroleum companies were considered to have substantial financial resources. Furthermore, the focus for the proposed petroleum searches is well away from the area of rock outcrops in the Musgrave Ranges where Aboriginal communities and sites of significance are concentrated. More recent exploration initiatives involving Pitjantjatjara land are described below under Revitalisation.

PETROLEUM INDUSTRY EXPLORATION


OEL 8
The first Oil Exploration Licence (OEL) for the Officer Basin was granted to the Australian Mining and Smelting Co. Ltd (operating through Frome--Broken Hill with funding from Consolidated Zinc Corporation Pty Ltd) for two years from January 1954. This followed the discovery of oil elsewhere in Australia, notably at Rough Range in Western Australia in 1953 which stimulated exploration in South Australia (including the formation of Santos and Geosurveys in March 1954) and brought a revised approach to exploration 2 thinking. OEL 8 covered 125 486 km of the southwestern corner of the State (Fig. 2.1). The companys work was confined to the southern portion of the licence and extended into Western Australia. In December 1953, Frome-Broken Hill requested BMR to survey part of the area, and Quilty and Goodeve (1958) first reported on the basin after BMR had flown six flights in May 1954, including from Oodnadatta to Ceduna, Kalgoorlie to Oodnadatta, and Forrest to Mount Harriet, Cook and Ceduna. In March 1955, the company proposed sending a field party to study rocky outcrops south of the Musgrave Ranges but the companys director, Maurie Mawby, held little hope for the area. He wrote to the Director of Mines prior to surrendering the licence in May 1955 that: It would ... appear that the only area of thick sedimentary section in the Eucla Basin lies immediately south of the Warburton--Musgrave Ranges and there the sediments hold little prospect of oil generation (MESA file 354/1955).

OEL 12
After the withdrawal of Australian Mining and Smelting, Clarence River Basin Oil Exploration Co. NL lodged two applications for OEL in February 1956. The first was granted in May 1956 as OEL 12 over an area between Coober Pedy and Pimba. The second application over an area of the Eucla Basin lapsed because the company was unable to complete the geological studies necessary to formulate a proper application for the area (MESA file 244/1956). At this time, the Department suggested to the Minister of Mines that a further investigation with air and ground field work of the Nullarbor Plain over about six months was warranted to determine if test drilling was necessary. Tom Barnes wrote in March 1956 (MESA file 244/1956):
It is my personal belief that the Nullarbor Plain is an area where the Government might well undertake a comprehensive geological survey on its own accord ---- the area is very poorly understood geologically, and has considerable hydrological interest, both for W.R.E. in relation to Maralinga, and also the few landowners; in addition to the unknown oil prospects. In such an isolated area it would be impossible to exercise effective control over an exploratory company, and a licence should only be granted to a company of unquestioned integrity.

OEL 19 and 28
OEL 19 was granted over 222 160 km in the Officer and Eucla Basins to Oil Drilling and Exploration Pty Ltd in October 1958. The company drilled Eyre 1 and Gambanga 1 in the Eucla--Madura portion of the Eucla Basin. OEL 19 was
2

16

Fig. 2.1 History of Officer Basin petroleum tenements, 1954--97.

transferred to its subsidiary Exoil Pty Ltd in July 1959 and was surrendered in September 1962, but a consortium led by Exoil (Oilmin NL from July 1973) held it until 1976; by then 2 the area had been reduced progressively to 24 605 km through subsequent renewals (being reissued as OEL 28 in 2 October 1962 over 160 000 km and as PEL 10, 11 and 12). The exploration boom that permeated the Australian mining scene from the mid-1960s introduced several new players to the northwest oil and gas search. Transoil Pty Ltd,

Continental Oil Co. of Australia Ltd (Conoco) and Australian Sun Oil Co. Ltd were farm-in companies to Exoils licence between December 1964 and September 1969. The method of a licensee farming-out areas to others prepared to farm-in to a joint arrangement was legislated for by the State Governments amendment to the M ining (Petroleum) Act 1940 in 1958, which also allowed a company to chequerboard its permit area and work the new blocks over five years instead of one. The introduction of these companies, particularly Conoco, strengthened exploration prospects in the area.

17

Using a Mayhew to drill a shothole with compressed air at Giles Junction, west of Emu, for Namco International Geophysics Ltd in 1962. (Photo T6491) Personnel and no. 2 core at Continentals bore in the Officer Basin, December 1966. (Photo T7157)

Exoils Mabel Creek seismic survey from April to August 1962 did not provide good quality data but it did reveal the first definite proof of a thick, subhorizontally bedded sedimentary sequence (MESA confidential file SR 11/5/29). The Namco traverse of 470 km went west across the basin along Giles Road north of Lake Dey Dey to Serpentine Lakes on the border and into Western Australia. Exoils initial seismic and gravity surveys were followed up by the drilling of Emu 1, the first petroleum wildcat well in the basin, which the Department drilled as a shallow stratigraphic test well near the Emu Field for Exoil. It was spudded in August 1963 and abandoned two months later at 417 m after reaching unmetamorphosed shale and sandstone (Grasso, 1963). Conoco participated in an aeromagnetic survey over the eastern Officer Basin after farming into the area. The survey, which covered ~155 400 km 2 between latitudes 27 and 30S, and longitudes 129 to 134E, was conducted from October 1964 to April 1965 except for three weeks when the Maralinga aerodrome was closed. The aircrew was based at Maralinga, Forrest and Oodnadatta. The survey identified the western half of the permit as having more petroleum potential because of the average thicker sections (~ 2440 m) on the southwest shelf and trough with more foothills type of structures developed within it (Steenland, 1965).

In 1965--66, Seismograph Services Ltd carried out a Vibroseis reflection survey for Conoco near Serpentine Lakes in the northwest of OEL 28. This was followed by a helicopter-supported stratigraphic survey beginning in March 1966 over outcrops to the north of the licence and on its margins. The Munyarai structure was discovered by the Departments 1966 seismic line (the EO line). Conoco subsequently detailed the structure by an additional seismic grid and drilled stratigraphic well Officer 1 (TD 183 m; Krieg, 1967) in November 1966 at the northern end of the EO line. Two deep wildcat wells, Birksgate 1 and Munyarai 1, were then drilled. Birksgate 1, spudded in January 1967, was positioned on a minor anticline as defined from the seismic work and reached 1878 m (Henderson and Tauer, 1967). The drilling required access clearance from Commonwealth Department of Supply and the South Australian Department of Aboriginal Affairs. Constructing a road to the site was expensive and so heavy loads were carried by large trucks fitted with desert tyres; earth moving equipment was used and an airstrip was laid down at the drill site. Access was difficult because of the high sand dunes and there was a lack of water. A stratigraphic well also at Birksgate, spudded in November 1966, was drilled to 644 m, plugged back to 392 m and completed as a standby water well. Birksgate 1 passed through probable Precambrian sediments to its total depth. In February 1967, Conoco proposed leaving this and another water well at Birksgate to the Mines and/or Aboriginal Affairs Departments because the effort to find potable water was expensive and it would have been a waste not to retain them. Aboriginal Affairs was very interested and arranged with the Mines Department for one well to be capped and the other to be fitted with a windmill and tank. In the following year Namco, on behalf of Conoco, followed up the structural lead with more detailed seismic work using the thumper method. As a result, a large anticlinal structure was delineated and a site for Conoco to drill Munyarai 1 was chosen. But heavy rain prevented the well from being spudded until July 1968. Munyarai 1 terminated at 2899 m without hydrocarbon shows being recorded (Conoco, 1969).

Looking south past the leading vibrator truck along a longitudinal line east of Serpentine Lakes in the Officer Basin on 1 March 1966.
(Courtesy H. Wopfner; Photo 44362)

18

OEL 33
In 1964, licensee Al Jergins (Outback Oil Co. NL) and Departmental officials made a two-day aerial reconnaissance of the Eucla Basin from Ceduna to Forrest (Western Australia) and back to Ceduna via the coastline. Some geomorphological anomalies and joint or fault controls were noted. The company had taken up OEL 33 (126 910 km 2 of the Eucla Basin and offshore) in January 1964. The focus was still the Eucla Basin to the south, where in November 1964 Outback Oil drilled Cook 1C (A and B were unsatisfactory sites but the third attempt became Cook 1) near Cook in the northern Eucla Basin. It was abandoned as a dry well after bottoming in probable Neoproterozoic sediments at 279 m. Outback Oil undertook a helicopter gravity survey over much of its licence area west of 13045E followed by an aeromagnetic survey of its offshore area. Between February and May 1966, four shallow wells (Hughes 1, 2, 3 and Denman 1) were drilled near Hughes on the Trans Australian Railway; Tertiary and Cretaceous sediments were noted to a depth of ~229 m, and possible early Palaeozoic sediments were noted below that unconformity but no significant hydrocarbon shows were reported (AR, 1965--66). This section is now considered to be Neoproterozoic (Ch. 5). In the South Australia portion of the Eucla Basin the company undertook several photogeological studies which revealed some geomorphological anomalies. These were inspected by a ground survey team. Farminees to the licence between June 1966 and June 1968 included Union Texas (Aust.) Co., Rock Island Oil and Refining Co. Inc., Tenneco Australia Inc. and Coastal Petroleum NL. The licence expired in January 1969 and the onshore area was then issued to Outback Oil as PEL 4. From June 1969, Outback Oil drilled Mallabie 1, which proved a thicker sedimentary section than previously thought, thus improving the exploration potential though the dry well was plugged and abandoned (TD 1672 m). PEL 4 expired in January 1974.

As petroleum developments in the Cooper Basin began to unfold under their own momentum, the Department began to concentrate on collecting basic data in the areas of low or marginal prospectivity to encourage companies to explore areas such as the Arckaringa and Officer Basins. By the late 1960s, the Permian sequences east of the Officer Basin were considered as possible petroleum sources and some attention was given to them. When no hydrocarbon discoveries were made, the exploration focus shifted to the central and western Arckaringa Basin--eastern Officer Basin area where a p r e - P e r m i a n c a r b o n a t e --r e d b e d e v a p o r i t e s e q u e n c e (Observatory Hill Formation) was considered to have economic hydrocarbon potential (Hibburt, 1984). The evaporites in Cootanoorina 1 in the Boorthanna Trough of the Arckaringa Basin and in the area to the southwest were identified as being of Devonian age. The evaporite beds below the Permian north of Coober Pedy were equivalent to the Observatory Hill Formation. Observatory Hill near Maralinga was named by Len Beadell in 1955 as the hill was similar in shape to an astronomical dome. The Arckaringa Basin was another little-known basin in a high exploration risk category but information from the west and northwest of that basin was of direct importance to the understanding of the Officer Basin ... which contains at least 16 000 feet of Palaeozoic sediment-fill [and] is one of the real challenges and its investigation is a natural follow-on from the work carried out in the Arckaringa Basin (AR, 1970--71, p.8). A downturn in the petroleum industry occurred from 1973 under the imposts of a Federal Government which terminated the Petroleum Search Subsidy Scheme, abolished tax concessions, banned the export of LPG and moved against the involvement of foreign companies such as Aquitaine and Delhi. It also moved to create a Petroleum and Minerals Authority. In South Australia, the inability to secure a petrochemical plant despite strenuous efforts throughout the 1970s was also a disincentive to explorers discovering liquids-rich gas. Exploration and development drilling was all but abandoned; only one well was drilled in 1973, and none in 1974 and 1975. Few seismic surveys were conducted. Although there was little exploration by either companies or the Department in the Officer Basin from 1969 to 1974, thereafter the Department drilled seven wells to 1979 to establish how widely spread the pre-Permian sequence was. Of these wells, Byilkaoora 1 and Wilkinson 1 intersected oil-mature Cambrian source rocks (Hibburt, 1984). Between July and December 1976, the Department drilled four shallow stratigraphic wells in the southern Officer Basin ---- Murnaroo 1 (7 km south of Observatory Hill; TD 628 m), Ooldea 1 and Reid 1 and 1A ---- which were all dry and abandoned though Murnaroo 1 revealed a potential reservoir sandstone. Then Wilkinson 1 was spudded in June 1978 and completed as a dry and abandoned well in August 1978 (TD 710 m). This hole, with its excellent oil-source potential, encouraged more exploration and led the Department to form an Officer Basin Study Group (there being an Eromanga Basin Group also), which began to draw together previous work on the basin and to reassess its petroleum potential. As a result, Byilkaoora 1 was drilled on the northeastern margin of the basin in the Mount Johns Range. This well, ~300 km north-northeast of Wilkinson 1, was spudded in

DOWNTURN TO DISCOVERY
Oil and gas discoveries elsewhere in Australia had sustained hopes for the Officer Basin search (ONeil, 1995, 1996a; Wilkinson, 1988). For example, in the Cooper Basin, Innamincka 1 in 1959 revealed a Permian basin below the Great Artesian Basin and hydrocarbon shows in the Permian sediments suggested its oil prospectivity. The Gidgealpa-Merrimelia Trend was first identified on seismic lines shot by the Mines Department in August 1962 and Gidgealpa 2 was drilled on this structure; the well produced gas and condensate. This testing of commercial quantities of gas from the Permian was the first petroleum discovery in the Cooper Basin and set South Australias petroleum industry on the way. The discovery of gas at Gidgealpa on 31 December 1963 and Moomba in March 1965, and oil at Tirrawarra in 1970, in the Cooper Basin indicated that the local efforts were not in vain. Gas production from Permian Cooper Basin reservoirs in the Gidgealpa and Moomba fields to Adelaide commenced in 1969. Follow-up drilling in 1970 and 1971 through farmout arrangements with several new companies resulted in significant Cooper Basin oil and gas discoveries.

19

PEL 10, 11 and 12


There was little company exploration in the basin for almost 15 years after OEL 28 became PEL 10 and 11 in 2 August 1969. PEL 10 was held by Exoil over 24 864 km and PEL 11 was held by Conoco, Transoil NL and Australian 2 Sun Oil Co. Ltd over 24 346 km . When PEL 10 and 11 were relinquished in January 1971, PEL 12 was granted to Exoil 2 and Transoil over a portion (24 605 km ) of the old areas. The area bordered the Everard Ranges and included the southern reaches of Officer Creek. No active work followed in the short term and efforts to farm out were unsuccessful until June 1974 when Shell Development (Aust.) Pty Ltd joined for nine months. Shell recommenced the exploration phase and ran 154 line km of seismic (the Everard seismic survey) in October and November 1974. The results of this were later shown not to be as expected. PEL 12 was surrendered in June 1976.

PEL 13
PEL 13 was issued in September 1972 to Planet 2 Exploration Co. Pty Ltd over 24 518 km in the Arckaringa Basin, the far west of the Great Artesian Basin. The northern perimeter of the licence, which extended west of Coober Pedy and north of Lake Phillipson, was on the northeastern margin of the Officer Basin but PEL 13 was surrendered in December 1973 without being explored.

PEL 23 and 30
Drilling Byilkaoora 1 in 1979.
(Photo T14751)

May 1979 and completed in July 1979 to 497 m. Oil-source rock correlations in Byilkaoora 1 indicated an alkaline playa-lacustrine origin for the immature crude (McKirdy and Kantsler, 1980). Carbonates equivalent to the hydrocarbon shows found in Byilkaoora 1 were detected in Marla 1A and 1B. This drilling suggested that the Cambrian carbonate sequences in the basin correlated to the Observatory Hill Formation, which by 1980 was considered to be the major potential source of petroleum in the eastern Officer Basin (Pitt et al. , 1980, p.209). Although the signs at Wilkinson 1 gave a good reason for companies to want to resume exploration in the basin, it was the oil bleeding from Byilkaoora 1 cores which provided the first really significant oil shows, confirmed the prospects revealed by Wilkinson 1, and raised expectations even higher. Complementing the work in the northeastern Officer Basin, in 1979--80 the Department undertook a helicopter2 based geological survey over 48 000 km of South Australias western portion of the basin in order to better understand this little studied region. The exercise had three main objectives: to delineate further the rock units (particularly the Observatory Hill Formation) mapped or intersected in drillholes to the east; to recognise Officer Basin rock units mapped in eastern Western Australia; and to assess the degree of deformation in the area with a view to identifying structural leads for petroleum exploration (Pitt et al. , 1980, p.215).

Active company petroleum exploration returned to the Officer Basin with Comalco Aluminium Ltd holding PEL 23 and 30 from January 1983 and February 1985, respectively, until early 1989. Comalco became interested in the petroleum prospects of the Officer Basin after the Byilkaoora 1 discovery; its mineral exploration in the period from 1979 included 20 fully cored drillholes in its mineral tenements in the region. Comalcos search was essentially for evaporite minerals, base metals and coal. Although its extensive exploration program failed to find alkali evaporites, indications of enhanced prospectivity for the basin were noticed; oil bleeds were detected wherever Comalco drilled the Observatory Hill Formation and the underlying Rodda beds. Commencing in March 1984 with a 1200 line kilometre seismic survey, Comalcos petroleum exploration included 2613 km of reconnaissance and semi-detailed seismic, and the drilling of five cored wells (at an average depth of 2000 m). This work improved the understanding of the geology and petroleum prospectivity of the eastern Officer Basin. The company partially completed its own review of the basin stratigraphy prior to drilling the first exploration well (Giles 1) during September--October 1985 (TD 1327 m; Stainton et al. , 1988). Of Comalcos plugged and abandoned dry wells, Ungoolya 1 (November 1985) revealed encouraging oil shows over several hundred metres in low porosity, early Palaeozoic and Neoproterozoic clastics; Karlaya 1 (April 1987), Lairu 1 (July 1987) and Munta 1 (September 1987) also showed traces of oil in these sediments. PEL 23 had been applied for in 1980 but a delay in granting was caused through stalled negotiations with

20

up new permits in the Pitjantjatjara and Maralinga lands of the Officer Basin. Petroleum exploration was planned onshore in the Officer Basin in 1990 and discussions were held with the Pitjantjatjara Council about access for proposed licence holders. Three potential reservoirs ---- Murnaroo Formation, Relief Sandstone and Observatory Hill Formation ---- had been determined by the time a data package was released on 44 300 km 2 in and adjoining Aboriginal land in four areas of the basin in late 1990. Although there had not yet been any commercial hydrocarbon discoveries, the basins potential reservoirs were then estimated to contain more than 523 bcf of sales gas or more than 451 mmbl of recoverable oil (Morton, 1992).
Company oil exploration in the Officer Basin in 1987.
(Photo 36174)

another company for adjoining acreage. PEL 23, from Emu to the Marla area, was renewed for five years in January 1988 2 2 and the area reduced from 23 222 km to 15 505 km by relinquishing in the southeast and north of the licence. PEL 30 2 was granted over 434 km adjoining the northwestern perimeter of PEL 23. Both PEL were surrendered early in 1989.

PEL 24
In November 1983, CRA Exploration Pty Ltd was granted 2 PEL 24 over 21 778 km on the southern margins of the Officer and Arckaringa Basins. A $6.4 million exploration program was proposed, including the drilling of at least four wells over the five-year licence term. However, only 435 line kilometres of seismic were recorded in 1985 and 1986, and Arkeeta 1, the only well, was drilled in December 1986. In July 1987, Pacific Oil and Gas Pty Ltd became the operator but the licence was surrendered four months later.

PEL 29
Exploration by Amoco, the operator, recorded 235 km of seismic including a section between Ungoolya 1 and Munyarai 1, but no wells were drilled. This, and the work by Comalco, delineated a number of large structures capable of trapping hydrocarbons. Amoco obtained good quality seismic data in the Munyarai Trough in 1987.

An Officer Basin team of professional and technical staff was formed within the Department in August 1992 to liaise with the Pitjantjatjara and Maralinga people, to carry out water well drilling and to survey for a seismic transect. This survey had been proposed in 1989 by the Department for the Australian Geological Survey Organisation (AGSO, formerly the BMR) to undertake. AGSO was to interpret existing geological and geophysical data, and to acquire new seismic, source rock, stratigraphic and petrophysical information. This was intended to form part of a major study by the Department on the structure, stratigraphy, petroleum source and reservoir potential of the Officer Basin in South Australia. The important potential of the basin would thus be highlighted. In the new regime applying to exploration, anthropological work and work corridor clearance demonstrated the modern approach to Aboriginal liaison and environmental management. The results included establishing a better correlation with the Amadeus Basin in the Northern Territory where oil and gas was already being produced at the Palm Valley and Mereenie fields, an improved knowledge of the sandstone reservoirs, hydrogeological and structural features of the basin, and better seismic interpretation from the reprocessing of existing data. Aerial surveys were flown by AGSO and the Department as part of its South Australian Exploration Initiative (SAEI). Under the National Geoscience Mapping Accord (NGMA) and the SAEI, the Department funded petroleum exploration to acquire seismic data in areas that had been ignored by the private sector but where the Department

PEL 33
PEL 33 was issued to a consortium of small companies ---- Median Oil NL (operator), Geometals Oil Exploration Ltd, Heron Petroleum Pty Ltd, Malita Exploration Pty Ltd, Gulf Resources NL, Southern Cross Exploration NL, Forsyth Oil and Gas NL, Antarctic Petroleum Pty Ltd, Spectrum Gold 2 NL ---- in June 1985 over 23 793 km in the Eucla Basin where an experimental seismic survey of 40 km was recorded before the licence was cancelled in April 1989.

REVITALISATION
The Department continued its regional geological studies and assisted Comalco in refining the stratigraphy of the Officer Basin (Brewer et al. , 1987). From the mid-1980s, the hydrocarbon potential of the basin was promoted by the Department in data packages prepared with seismic, drilling, mapping and reports. Companies were encouraged to take

Anthropologist Scott Cane, at right, with southern Pitjantjatjara tribal elders during line scouting for the seismic survey on Maralinga Lands in November 1993. (Courtesy S. Cane)

21

Geosystems Pty Ltd Vibroseis trucks operating on the Officer Basin seismic survey in August 1993. (Photo 41474)

considered there was petroleum potential. A frontier geological province such as the Officer Basin required earlier information to be revised. This work was now subject to agreements with the Aboriginal landholders for access for water well drilling, seismic line surveying and seismic surveys. The planned NGMA transect of 600 km from the Musgrave Block in the north to the Nullarbor Plain was modified when access to the unnamed conservation park was denied, and seismic test work on the Nullarbor Plain could not penetrate the surficial cavernous limestone (Gravestock and Lindsay, 1994). Reflecting the modern regime, the AGSO--MESA transect in 1993 included environmental audits of seismic practices. In conjunction with AGSO, the Department conducted seismic surveys in unexplored Pitjantjatjara and Maralinga Lands in the eastern Officer Basin late in 1993 (Gravestock and Lindsay, 1994). AGSO undertook a 550 km regional transect to tie a northern line, including reprocessed 1966 seismic data, to a series of lines in the south. This formed part of the NGMA with Federal Government funds and South Australian logistical support on Aboriginal liaison, line surveying and water wells. (See Ch. 3 for more detail on the land access arrangements.) As well as investigations to establish water supplies for mineral and petroleum companies during 1993, especially in the desert areas outside the Great Artesian Basin, the Department examined water supply options in the Officer Basin and at Oak Valley where the Maralinga people had formed an outstation since 1984. An earlier camp had been at Dey Dey. These projects identified water of stock quality in the eastern Eucla Basin through to potable supplies in the Musgrave Ranges. The Department reviewed seismic data for the Marla area and contracted 378 km of seismic survey work along nine regional lines west of Marla. The NGMA transect was the first in the barely explored central and southern Officer Basin, while the SAEI grid linked seismic acquired in the mid-1980s by Comalco and Amoco in the eastern Officer Basin. Some 140 km of Comalco seismic data were reprocessed to the same standard. The new seismic acquired in the basin followed a reinterpretation of Comalcos 1983--85 data (Mackie and Gravestock, 1993).

The $5 million of surveys and associated geological studies indicated that considerable undiscovered hydrocarbon potential exists in the Officer Basin but that further studies would be needed to assess its potential. The investigation confirmed thrust faulting (Alice Springs Orogeny) along the structural northern basin margin, the thrusts propagating south within as well as beneath the sedimentary cover indicated 6 km or more of Neoproterozoic sediment in the Birksgate sub-basin [but that] the southern Murnaroo Platform is unlikely to contain large structures, strengthened biostratigraphic correlation with the Amadeus Basin and confirm[ed] the utility of acritarchs for Neoproterozoic zonation and indicated a potential sabkha-associated source rock near the base of the succession (Gravestock and Lindsay, 1994, p.65). The prospects for major investigations were enhanced by the discovery of sufficient ground water in the south-central Officer Basin to support shothole drilling. The new seismic investigated the southwestern extension of the Manya Trough and Marla Overthrust Zone, which have revealed most of the oil shows. As well, biostratigraphic and petrophysical studies of wells throughout the basin correlated the Neoproterozoic more effectively and improved the known reservoir characterisation. New structural data adjacent to known oil-bearing rocks in the Marla Overthrust Zone were obtained in the adjacent troughs and ridges but they are poorly delineated by seismic and are relatively undrilled. Furthermore, the Mesoproterozoic Ammaroodinna Inlier poses questions as to its origin and structural position southwest of the Marla Overthrust Zone. The correlation between strata in the Munyarai and Manya Troughs, both potentially oil-generating kitchens, requires further investigation to reveal their significance for petroleum exploration (Gravestock and Lindsay, 1994).

PEL 61 and 63
In May 1996, PEL 61 and 63 were granted over 6258 and 2 19 930 km , respectively, in the Marla area of the Officer Basin, after Hemley Exploration Pty Ltd successfully concluded access negotiations with the Aboriginal landowners. Drilling is scheduled to commence in 1997.

CONCLUSION
The search to date of the Officer Basin has been sparse and <7200 km of seismic data have been recorded and only 30 wells deeper than 500 m drilled. Only 12 of the ~70 drillholes are petroleum exploration wells of sufficient depth to enable the stratigraphy to be pieced together. Core and wireline logs from mineral, groundwater and stratigraphic drillholes, especially in the eastern part of the basin, provide useful data. There has been limited wildcat exploration but the promising potential of the region is such that it deserves more attention, especially given that the logistical problems and land access issues can be overcome. The current interest demonstrates that the Officer Basin is now regarded, at least by some parts of the industry, as having significant petroleum potential.

22

CLIMATE
Whilst extending south to the coast in the vicinity of Nullarbor National Park, the Officer Basin for the greater part lies inland and is comprised of relatively level terrain with few hills or mountains. As a result there is little variation in the climate throughout the region. The average annual rainfall is low and variable, ranging from ~250 mm/year in the north (Fig. 3.1) to 300 mm/year towards the southern coastal region. Rainfall events show only a weak seasonal pattern, with the southern margins receiving more winter rainfall whilst the northern areas may experience summer rainfall. Long sustained periods of rain are rare although large falls can occur over short periods. In the northern areas, heavy rainfall can occur during any month of the year although it is not uncommon to experience 23 very dry months. Prolonged droughts are frequent throughout the area and evaporation rates are very high, often exceeding 3800 mm/year. The summer months from December through to February tend to be the hottest time of the year. Mean maximum daily temperatures usually exceed 32C and are often over 37C, with the temperature dropping 1020C at night. During the winter months it can be mild to warm with a mean maximum temperature of 17C and a mean minimum of ~5C.

These landforms can be grouped into three major environmental regions; the western sandplains, the central tablelands and the Nullarbor Plain. Each region consists of a number of major landforms, including dunefields, undulating plains, sandplains, tablelands, clay pans and salt lakes. A number of minor landform patterns or environmental units occur within each of these major landforms. The occurrence of minor landforms can vary, depending upon the detailed morphological characteristics of a major landform based upon the local geology, soil type, topography, drainage patterns and biota.

Western sandplains region


The western sandplains region is characterised by undulating plains and extensive dunefields. Throughout the dunefields there are occasional silcrete rises, saline depressions and low gibber-covered rises with occasional low hills and ridges. This region includes the Great Victoria Desert which is a transitional zone between the northern margin of the Nullarbor Plain region and the western sandplains region. The Great Victoria Desert is characterised by longitudinal sand ridges up to 20 m high and 100 km long. The desert is so named because of the lack of modern surface drainage.

Central tablelands
The central tablelands consist of undulating plains covered with silcrete pebbles and rubble. The plains are cut by several large seasonal creek beds and associated floodplains, along with scattered groups of dissected tablelands, mesas, clay pans and salt pans.

LANDFORMS
The Officer Basin underlies a vast area of land totalling over 375 000 km2, and covers a wide range of landforms.

Fig. 3.1 Average monthly rainfall and average maximum daily temperature for Marla.

Spinifex and mallee vegetation, western sandplains, Officer Basin.


(Photo 43105)

Nullarbor Plain
The Nullarbor Plain is characteristically an undulating, featureless limestone plain with occasional sinkholes and caves, and traces of surface drainage in the form of elongated chains of dry lakes.

ENVIRONMENTAL CONSIDERATIONS
National parks and reserves
Six parks and reserves are fully or partly located within the boundaries of the Officer Basin (Fig. 3.2). These were created to preserve the best examples of vegetation and landforms within the region. Access for exploration and mining is allowed in all parks except the Unnamed Conservation Park and Nullarbor National Park. The conditions for access vary from park to park, based on the type of reserve classification (Conservation Park, National Park or Regional Reserve), the activity proposed and the impact it is likely to have on the environment.

NATIVE VEGETATION
The natural vegetation associations throughout the Officer Basin vary considerably, ranging from woodlands and tall shrublands, to hummock grasslands, chenopod shrublands, grasses, ephemeral forbs and occasionally samphire associations. Many regions within the basin area, whilst classified as desert, appear very unlike the traditional concept of what a desert should look like. For example, low open woodlands, which generally occur on dunefields, can be found throughout the basin in a variety of topographic situations. Often these woodlands are comprised of one or a number of different species such as myall, black oak, northern cypress pine or mallee. The understorey of woodlands on the dunes typically consists of tussock grasses, spinifex or sclerophyll shrubs, whereas the understorey on interdunal lows generally consists of various species of saltbush and bluebush. Throughout the plains, the woodlands often give way to mixed chenopod shrubland comprised of species such as saltbush or bluebush with a hummock or tussock grass understorey, or tall shrubland comprised of species such as cassia or mulga with a grass or chenopod understorey. Vegetation throughout the Officer Basin can vary from very dense to sparse. Generally each plant community occupies a particular habitat, with the distribution of vegetation often reflecting the landform and soil types. Although there are often large areas with only one vegetation association, many instances occur where there are complexes of two or more different vegetation associations, each occupying a specific niche in the environment. Small differences in habitat such as depressions or drainage lines may produce slightly more growth and a greater variety of species. Perennial plants throughout the region have adapted to endure long dry spells and extreme temperatures, whilst the appearance of herbaceous plants can be episodic and infrequent depending upon suitable climatic conditions.

Summary of environmental regulation


A number of environmental issues are pertinent to petroleum exploration in the Officer Basin, all of which can be resolved by proper operational planning in the initial stages. In order to ensure that activities are undertaken in a manner which minimises environmental impacts, a number of documents are required before approval to commence operations is given. A Declaration of Environmental Factors (DEF) is required from the licensee. This is the licensees assessment of the environmental impact of an activity. A Code of Environmental Practice (CEP) is also required by regulation. The code describes the procedures that the proponent will adopt during the planning, assessment, field management auditing and monitoring phases of the operation. A CEP for seismic operations within the Officer Basin has been developed by MESA and is available to licensees. This provides advice on environmental issues that need to be taken into consideration in planning a project. A company may either adopt this code or use its own CEP subject to approval by MESA. Although a number of documents are required, the approval process is not onerous. MESA is able to assist licensees by providing examples of the documentation and advising on their scope.

CULTURAL HERITAGE
European heritage
Sites of European heritage significance such as historic buildings, graves and geological monuments may be found in the Officer Basin. These are indicated on environmental sensitivity maps which can be purchased from MESA. The majority of sites are small and easily avoided by exploration and production activities.

Aboriginal heritage
In South Australia it is an offence to disturb or destroy Aboriginal sites, objects or remains. Standard procedures for determining the presence of Aboriginal heritage prior to commencement of activities have been determined. These involve consulting with the relevant Aboriginal organisation and maintaining a watch for sites, objects or remains during exploration. These sites are generally no larger than a few hundred square metres and are easily avoided. Since the
"

Meramangye Lake, Officer Basin.

(Photo 43070)

Fig. 3.2 Parks and reserves, Aboriginal lands, and defence areas covering the Officer Basin area.

inception of the Aboriginal Heritage Act 1988, there have been no conflicts between Aboriginal heritage sites and exploration or production activities. MESA can provide advice to companies on Aboriginal heritage. The Aboriginal Heritage Act also applies to Aboriginal lands held in freehold, including the Anangu Pitjantjatjara and Maralinga Tjarutja Lands. Access to Aboriginal land for site clearance must be negotiated with the traditional owners (see Aboriginal lands).
#

ABORIGINAL LANDS
History
Following the arrival of Europeans in the 1920s to the desert regions of northwestern South Australia, many of the Aboriginal people who lived in those regions moved in three main directions to Areyonga in the Northern Territory, Ernabella Mission just south of the Northern Territory border, and south to Ooldea Mission on the Trans Australian Railway.

Ooldea Mission closed in 1952 and the people were transferred to a new mission at Yalata. In 1953, the British Government established a base 35 km north of the old Ooldea Mission on what is now known as Maralinga Lands. The purpose of the base was for nuclear weapons testing. Even if Ooldea Mission hadnt closed prior to arrival of the British, it would have had to close due to the proximity of the nuclear testing. In 1981, the Pitjantjatjara Land Rights Act was passed under which a body corporate the Anangu Pitjantjatjara (AP) holds the land under inalienable freehold title. The Act also gives to AP control of entry to the lands and a share of the production of petroleum royalties earned from the land. The Aboriginal people south of the AP Lands sought similar legislation and, in 1984, the Government recognised their claim and passed the Maralinga Tjarutja Land Rights Act. Under this Act, a corporate body called the Maralinga Tjarutja (MT) was established to receive freehold title to the land. The AP Lands abut the northern border of the MT Lands and extend as far north as the Northern Territory, covering an area of ~103 000 km2. The MT Lands extend south to the Trans Australian Railway, covering ~80 764 km2. Whilst the Officer Basin incorporates all of the MT Lands, only the southern portion of the AP Lands falls within its boundaries. Traditional owners believe that their role with respect to land rights is to fulfil their responsibilities to that land. They are under social and cultural direction to ensure that the land is protected and that sites of significance are avoided. To enable them to discharge that responsibility, traditional owners have to be fully informed of any activities proposed within the lands and the impacts of those activities on the lands. To ensure that all of the people with responsibilities for the land under consideration have been consulted, it is necessary for exploration companies to submit an application with the traditional owners at the earliest possible stage of an exploration program. Maralinga Tjarutja work closely with the Pitjantjatjara Council in relation to areas of common interest such as petroleum exploration. It is important that potential explorers are prepared to work with the traditional owners to ensure that a reasonable balance is struck.

royalty, which is shared 13 to AP, 13 for all Aboriginal people in South Australia, and 13 to South Australian general revenue. AP Lands Except for a limited number of instances as set out in the Pitjantjatjara Land Rights Act, all non-Pitjantjatjara people are required to apply for permission to enter AP Lands. Exploration companies must first obtain the approval of the Minister administering the Mining and Petroleum Acts to apply to AP for permission to enter those lands. Once permission is obtained, a company may submit an application to the Executive Board of Anangu Pitjantjatjara which then has 120 days from the date of application to grant unconditional permission, permission subject to conditions or to refuse the application. If the AP refuses permission or imposes conditions unacceptable to the applicant, or the applicant has not received a notice of a decision by AP within 120 days from the date of application, the applicant may request the Minister for Mines to refer the matter to an arbitrator. A determination under this section is binding upon the AP, the applicant and the Crown. There are special provisions in Sections 20 to 24 of the Act in relation to applicants for petroleum exploration and mineral licences (Appendix 3.1). MT Lands Whilst provisions in the Maralinga Tjarutja Land Rights Act relating to access for exploration companies are similar to those for AP Lands, there are two main differences the Maralinga Tjarutja Land Rights Act limits the amount of compensation provided at the exploration stage to no greater than the amount of compensation provided for under the Mining and Petroleum Acts, and there is a slightly different procedure for appeals. Where the MT refuses permission or imposes conditions unacceptable to the applicant, or the applicant has not received a notice of a decision by MT within 120 days from the date of application, the applicant may request the Minister for Mines to attempt to resolve the matter by arbitration with the assistance of the Minister for State Aboriginal Affairs. These conditions are set out in detail in Sections 21 to 26 of the Maralinga Tjarutja Land Rights Act 1984 (Appendix 3.2).

Access to Aboriginal lands


Exploration activities have been successfully undertaken in both MT and AP Lands, and the traditional owners are well acquainted with the transient land use of exploration. Both the MT and AP are proud of the work they have done with petroleum companies and have indicated a wish to continue to work towards balancing their own sense of responsibility towards the land and the interests of those who seek to explore and produce from it. In the access agreement negotiated for PEL 61 in AP Lands, provision was made for compensation for exploration activity (expected to be ~$20 000/year) for a joint-venture partenership option for the Aboriginal owners of the land, and payment of production royalties of 13% based on a sliding scale linked to the quantity of oil or gas produced (in the event of a commercial discovery). This royalty to AP is in addition to the 10% State
$

Signing the access agreement for the Maralinga Seismic Survey, 1992. (Photo 40420)

Access for seismic surveys No petroleum exploration was carried out between 1988 and 1992 following withdrawal of Comalco and Amoco from the Aboriginal lands, but two seismic surveys have been carried out since 1992. The first, of 550 km, was recorded as part of the National Geoscience Mapping Accord (NGMA) by the Australian Geological Survey Organisation (AGSO), partly in AP Lands but mainly in MT Lands. The second survey, of 379 km, was recorded by MESA as part of the South Australian Exploration Initiative (SAEI) in eastern AP Lands. Land access was summarised by Gravestock and Lindsay (1994) as follows. Discussions with AP and MT outlining plans for the NGMA transect commenced in 1990. These involved several visits to Aboriginal communities by AGSO and MESA personnel, and the wide distribution of a leaflet illustrating seismic recording techniques and their effect on the land. In 1992, discussion was formalised and embodied in two agreements signed separately by AP, MT and the South Australian Minister for Mines and Energy. A third agreement between AP and the Minister, modelled on its predecessors, was signed in 1993 to enable the SAEI seismic program to commence. In essence, each agreement sets out the access conditions for a line scouting stage followed by the line clearing and recording stage. All personnel involved in the scouting teams, both Aboriginal and non-Aboriginal, were fully briefed as to the nature and type of work to be undertaken. In the scouting stage, groups of four men and four women responsible for safeguarding sites of significance were accompanied by anthropologists and MESA officers to mark the route of seismic lines (using GPS) and to clear a work area 200 m either side of the proposed line with due regard to cultural significance. Where necessary, seismic lines were moved or bent to avoid sites of significance without compromising the objectives of the survey. The cleared work corridors were agreed to in writing by AP, MT and MESA. All visitors to the Aboriginal lands were issued with permits stipulating the conditions of entry. Seismic line preparation followed the marked route ahead of the recording crews, at which time one or more senior Aboriginal men were employed by the AP and MT Councils to act as liaison officers. Liaison officers were chosen to act as representatives for each Council to ensure that the daily management of the seismic program proceeded in accordance with the access agreements. The men had several important duties to perform: they were responsible for the pre-work cultural briefing in the field for the seismic crews and other contractors they ensured that permit entry conditions were honoured crews were kept within the agreed work corridors, except for surveyors and water drillers who were accompanied to trig points and drillsites they were responsible for providing an account of the impacts of the seismic lines to their respective Aboriginal communities.
%

Advice from the liaison officers was passed to the crew chief or the Ministers representative. In addition, these men (and sometimes their families) took off-duty crew members on bush tucker trips. Immediately before crew departure, the liaison officers accompanied AGSO, MESA and contracted personnel on a line inspection to ensure that all rubbish was removed or buried and, where required, seismic lines were rendered impassable to future visitors. As a result of requests, some lines were left open for 4WD vehicles. The practice of carrying out a scouting stage followed by a monitoring stage for petroleum exploration is likely to become standard practice for licence holders. The cost of this exercise for the seismic surveys varied between $100 and $150 per kilometre recorded.

OTHER LAND ISSUES


Commonwealth land
Commonwealth land includes defence facilities, various railway easements, post offices, aerodromes, lighthouses, telecommunication facilities and prohibited areas such as the Maralinga test site known as Section 400. Access to Commonwealth land may be granted, subject to predetermined conditions, by the relevant Commonwealth department with the exception of sensitive defence areas.

Maralinga and Emu nuclear weapon test sites


Between 1953 and 1963, the British Government conducted a program of nuclear weapons development trials at the Maralinga and Emu sites in South Australia. During that period nine atomic bombs were detonated and 700 minor trials were undertaken. The Maralinga site covers ~3000 km2. The Emu test site, 200 km north of Maralinga, covers ~500 km2. The Maralinga site is owned by the Commonwealth whilst the Emu site is owned by the Crown. Areas within both test sites are contaminated by radioactive wastes, the principal contaminant being plutonium used during the minor trials. In 1985, the Australian Government established a Royal Commission into British nuclear tests in Australia which proposed that the MT Lands be cleaned up and decontaminated, and that the traditional Maralinga owners be compensated. The Australian Government recently initiated decontamination of the site, with the British Government agreeing to contribute to the cost. It is expected that the clean-up will be completed by the end of 1998, after which the land will be returned to the traditional owners. Public access to the Maralinga and Emu test sites is generally prohibited.

Woomera Prohibited Area


This area, covering 133 300 km2 (13.5% of SA), was established under the Defence (Special Undertakings) Act 1952. Public access is restricted to Woomera township, Stuart Highway and the Coober PedyWilliam Creek road. Access elsewhere in this area is by permit, which can be obtained from the Area Administrator at the Defence Support Centre in Woomera. There is no formal avenue for appeal if the Area Administrator refuses to issue a permit to enter.

Operators within the prohibited area are required to carry $7 million public liability insurance.

Mintabie Precious Stones Field


Opal was first discovered at Mintabie in the 1920s although it was not until 1976 that heavy earth moving machinery moved in, after which several large finds were made. In 1988, Mintabie produced an estimated $39 million dollars worth of opal, making it the largest producer of precious opal in the world at that time. Mintabie is referred to as a precious stones field and, for those wishing to prospect or mine for opal, a Precious Stones Prospecting Permit is required. Many hundreds of opal workings occur throughout the field and consist primarily of large areas ripped by bulldozers to remove the overburden and expose the shallow opal seams. The Mining Act 1971 provides for stratified titles (section 63(a)), which are proclaimed by the Governor under Section 8(1)(ba). This enables a petroleum exploration licence or production licence to be acquired over a subsurface stratum where the surface is being mined for opal. However, the nature of opal mining in the Mintabie Precious Stones Field could make access for petroleum exploration hazardous due to the number of open cuts.

&

INFRASTRUCTURE
Transport links
The main road in the region is the sealed Stuart Highway which links Adelaide and Darwin (Fig. 4.1). Adelaide (population ~1 million) is 1000 km southeast of the Marla region by road. Alice Springs (population 27 500) is 550 km to the north by road. Port Bonython (740 km), and Thevenard near Ceduna (525 km), are the nearest State commercial deep sea ports; Whyalla (740 km) is the nearest privately operated deep sea port. Important towns in the region include Marla (population 243) and Coober Pedy (the main regional administrative centre, population ~4000). The Central Australia Railway links Adelaide to Alice Springs (Fig. 4.1). The railway may eventually be extended to Darwin, where modern port facilities provide a link to southeast Asia. The nearest airport is Coober Pedy (sealed and gravel runways); a number of settlements and towns have airstrips, including Fregon (sand silt), Indulkana (silt clay), Marla (gravel), Granite Downs (silt clay) and Oodnadatta (Fig. 4.1).

naphtha feedstock from the liquids plant. The mini-refinery supplies the northern Spencer Gulf region. The nearest pipelines in South Australia to Marla are the 659 km MoombaPort Bonython Liquids Line and the 781 km MoombaAdelaide Pipeline, operated by EPIC Energy (Fig. 4.2). In the Northern Territory, ~550 km north of Marla, is the 1500 km Amadeus BasinDarwin gas pipeline, completed in 1986 and operated by NT Gas Pty Ltd. This has a maximum daily capacity of 2.5 million m3 (87 mmcf), and en route supplies Northern Territory towns as well as several gold mines and a major leadzinc mine at McArthur River. The capacity of the McArthur spur is 0.45 million m3/day (16 mmcf/day). The current Northern Territory market for gas is 15 PJ/year; a major potential market is the alumina refinery at Gove which would consume 2025 PJ/year if converted from diesel. The 270 km MereenieAlice Springs oil pipeline in the Northern Territory was constructed in 1985. It is operated by Santos Ltd and has a capacity of 1490 kL/day (9400 bbl/day); current throughput is 480 kL/day (3000 bbl/day). Oil is transported to a tankfarm and railhead at Brewer Estate, 20 km south of Alice Springs, where it is loaded into tank railcars on a spur line of the Central Australia Railway and transported 1500 km by rail to Port Adelaide. From here it is trucked to Port Stanvac refinery, 10 km south of Adelaide.

Pipelines and production facilities


The Moomba Plant, operated by Santos Ltd ~660 km east of Marla (Fig. 4.2), produces sales gas for Adelaide and Sydney, and processes 25.4 million m3 (902 mmcf) of raw gas and 6 600 kL (42 000 bbl) of condensate and crude oil per day. Condensate, LPG, crude and some ethane are transported as a cocktail via pipeline to Port Bonython where they are separated and marketed within Australia and overseas. The Port Bonython liquids plant, also operated by Santos, produces crude, naphtha, butane and propane. A mini-refinery adjacent to the Port Bonython plant produces 95 kL/day (600 bbl/day) of gasoline by refining

Airstrip in the Officer Basin.

(Photo 40271)

Aerial view of Port Bonython tanker loading facilities.

(Photo 40416)

'

A transcontinental gas pipeline, linking gas reserves in the Northwest Shelf and Timor Sea to markets in southeastern Australia, is likely to be constructed in the next 1020 years (Australian Gas Association, 1988). It is possible that such a pipeline would be routed through the Amadeus Basin to Moomba where it would link with the existing Moomba Sydney gas pipeline. The economics of Officer Basin gas could be significantly improved if sufficient reserves were discovered prior to construction, and the pipeline routed via the basin. Port Stanvac refines petroleum products mainly for the South Australian market. The refinery commenced operations in 1963 and the adjacent lubricating oil refinery came on stream in 1976. The main products are LPG, solvents, motor gasoline, jet fuel, kerosene, diesel (both automotive and industrial), lube oil basestocks for Australian and overseas markets, fuel oil and bitumen.

Fig. 4.1 Officer Basin infrastructure.

!

Fig. 4.2 Australian gas and liquids pipelines, treatment plants and refineries.

ACCESS TO POTENTIAL MARKETS


Industries in the region
The northern part of South Australia is sparsely populated and relatively undeveloped due to its remoteness and harsh climate. The main primary industry in the Officer Basin region is cattle, which are run on large pastoral leases. Tourism (including an eco-tourism venture by Anangu Pitjantjatjara) is a growing industry in the region. A large proportion of the worlds opal is mined at Coober Pedy and Mintabie. The Olympic Dam Mine, 480 km southeast of Marla, is the worlds largest copperuranium deposit. Western Mining Corporation is planning to more than double copper production from the current level of 85 000 t/year to ~200 000 t/year, together with associated uranium, gold and silver. The township of Roxby Downs supports the mine, and has a population of 2700 which will grow as the expansion
!

commences. Current energy usage is 3040 MW/day (~1.5 PJ/year). Extensive mineral exploration is currently underway on the Gawler Craton, south of the Officer Basin. High-grade gold intersections have been made at Challenger prospect and significant gold mineralisation has been discovered at other prospects (Campfire Bore and Golf Bore). The prospects for commercial gold developments in the region are rated very highly by the mineral exploration industry. Extensive subeconomic Early Permian coal deposits (~15 billion tonnes) occur in the Arckaringa Basin, in the Coober Pedy region. The South Australian Steel and Energy (SASE) Project is investigating development of iron ore and coal resources south of Coober Pedy. The project plans to apply Ausmelt technology to produce pig iron. A pilot plant is being built at Whyalla. Some of these projects may be potential users of natural gas.

Gas
Ex-field natural gas prices in South Australia are freely negotiated between buyer and seller, and the Commonwealth, States and Territories have agreed to remove impediments to across State borders trade in gas from July 1996. The rights of access to gas transmission and reticulation pipelines will be provided, and direct negotiations between consumers and producers facilitated.

spans a host time range from Tertiary palaeochannel sediments to Cambrian Observatory Hill Formation. These sediments are highly variable in composition and are impossible to divide into aquifers and aquitards on the available information. Shale of the Observatory Hill Formation contains sand sequences, while the Trainor Hill Sandstone may be impermeable in parts. A major increase in the density of geological information is needed before there is any hope of mapping individual aquifers, if such exist. Lau et al. also noted that there may well be perched aquifers in the palaeodrainage channels, hosted by Tertiary Hampton Sandstone or Pidinga Formation. A confined or semi-confined aquifer may be present in the Murnaroo Formation, which is intersected in several holes in the southeastern part of the basin. This aquifer yielded saline water in the Tallaringa Trough and near Maralinga, and is probably recharged in the area of the Nawa Ridge where the formation is closest to surface and subcrops under Tertiary sediments.

Crude oil
A free market was introduced in 1988 for all oil and condensate produced in Australia. There is no restriction on imports or exports of crude oil or refined petroleum products. A similar regime has applied since 1991 for LPG. Markets for crude oil and condensate exist in South Australia and Australia, and low sulphur light crude oils find a ready domestic and overseas market.

GROUNDWATER
The Officer Basin occurs in an area of low rainfall and high evaporation. Surface water (ephemeral and permanent) is virtually non-existent, and groundwater, where present, is usually highly saline. Although data on groundwater are very sparse, it is unlikely that more extensive searching in this environment will yield any major resources of low salinity groundwater.

Recharge
Subsurface water flow may occur through the Tertiary palaeochannels which extend southwards from the Musgrave Block over the full surface extent of the Officer Basin. While there is no firm evidence of such water movement, it must be expected on the basis of the proven water content in these features to the north, the permeability of the sediments, and the potentiometric gradient. Such waters would be expected to be saline. Below the base of the Tertiary sediments, any southward flow of groundwater into the basin is expected to be blocked by the steep, fault-controlled northern edge of the basin and the steeply dipping Adelaidean sediments. Movement of water from Western Australia eastwards into the South Australian portion of the Officer Basin is possible, but the potentiometric surface indicates that the major flow direction is to the south. Thus, apart from surface flow, most recharge into the groundwater system of the Officer Basin is expected to come from local recharge.

Surface water
There is little evidence of surface water in the Officer Basin region. The ephemeral streams of the Musgrave Block to the north and Eromanga Basin to the east vanish abruptly in the environment of higher permeability and porosity of the Palaeozoic sandstones in the Officer Basin, which result in faster penetration of water into the subsurface and a lower watertable. The ephemeral Officer Creek, which flows into the area from the north, extends into the basin ~50 km before being absorbed, but is by far the most persistent of such features. Other indications of the gathering of surface waters for recharge are very few and uncertain. The only other evidence of surface water of any kind comprises the salt lakes around the southern edge of the basin outcrop at Serpentine Lakes, at Lakes Dey Dey and Maurice, at Wyola and Wilkinson Lakes, and north of Emu Junction. These highly saline environments, which occur along the southern boundary of the north-dipping Palaeozoic sediments of the basin, are interpreted as discharge zones (Fig. 4.3; Lau et al., 1995a,b).

Aquifers
On the basis of the very sparse information that exists, Lau et al. (1995a,b) referred to all known groundwater as being interlinked in one unconfined system, with the Precambrian surface regarded as hydrogeological basement. While this may be a simplification, there is insufficient evidence to justify subdivision. The system extends from surface in the discharge zones south of the basin to considerable depth in Birksgate 1, and
!
Marla township showing groundwater recharge swamps, railway line and airstrip. (Photo T22004)

Fig. 4.3 Groundwater potentiometric surface and salinity (after Lau et al., 1995a,b).

Local recharge potential


While there are no statistics from the area of outcropping Officer Basin sediment, rainfall in the basin is expected to range from 150 mm in the south to no more than 250 mm in the ranges along the northern margin. This precipitation level is low by any standards and, when combined with high daytime temperatures and consequent high evapotranspiration rates, does not auger well for recharge. According to Jacobson et al. (1994), the monthly rainfall in the arid zone must exceed 130 mm before there is any contribution to recharge. However, using the approximations implicit in Thornthwaites equation (Thornthwaite and Mather, 1957), monthly rainfall and temperature figures for Ernabella which, while not in the Officer Basin is near enough to assist in such estimates, yield rather more optimistic figures, as shown on Figure 4.4. Combining the two estimates, and bearing in mind that the figures for potential recharge on Figure 4.4 usually require that the recharge occurs over one, or at most two months, it would appear that for this location significant recharge can be expected in at least 12 years of the 45 for which data are available, or one year in four on average. This is much better than the one year in 1520 suggested by Jacobson et al., but
!!

their figure may well be more applicable for the Officer Basin itself, which has significantly lower rainfall than Ernabella in the Musgrave Ranges. The increased elevation and greater likelihood of monsoonal rains from the north increase the precipitation at Ernabella compared to that for the Officer Basin. Whichever scenario is taken, it is clear that local recharge is a rare event in the Officer Basin and its surrounds, occurring on average every four to 15 years.

Discharge
The only evidence of surface discharge from Officer Basin aquifers is the salt lakes along the southern margin. Subsurface discharge may occur into the overlying Eucla Basin to the south of the Ooldea Range, but this is speculative. Since the Officer Basin sediments dip gently northwards, and the base of the Cambrian is not deep below the Ooldea Range, the presence of impermeable Wirrildar beds in seismic shotholes in this area may be a barrier to southward groundwater movement and may be partly responsible for the surface discharge in this area. Thus, there

continuous over this area but with little local recharge, so that there is little likelihood of lower salinity water other than very restricted, fragile supplies. It is probable that the low salinity (<1 500 mg/L) supplies encountered at Birksgate 1 and near Coffin Hill in the BIRKSGATE map area are the result of local recharge. As such, they could be fragile and of limited extent. The low recharge potential of the area precludes large supplies of potable water.

Fig. 4.4 Annual potential groundwater recharge at Ernabella. The monthly potential recharge (precipitation less potential evapotranspiration) is calculated from precipitation and temperature data. The annual potential recharge is the sum of positive monthly potential recharge figures.

may be little or no movement of water between the Officer and Eucla Basins. It is presently thought that virtually all discharge from the Officer Basin sediments is through the salt lakes near the Ooldea Range.

Potentiometric surface
The potentiometric surface reflects the smoothed topographic surface, and confirms the general southward gradient (Fig. 4.3).

Water quality
Salinity contours indicate that groundwater is usually saline to very saline (Fig. 4.3). These contours give a general idea of what can be expected, but are much biased by a few samples, such as Birksgate 1 which is the only sample over a large area and could well be a more local anomaly tapping a small area of local recharge. In the Musgrave Block to the north of the Officer Basin, it is common for groundwater to have nitrate and chlorine levels in excess of WHO approved limits. This could also apply to groundwater in the Officer Basin. In the Maralinga area, groundwater in the Tertiary palaeochannels and possibly in the underlying Officer Basin sediments tends to be very acidic (pH ~4) and have a very high iron content (>200 mg/L). Similar impurities are found in palaeochannels in the Lake Maurice area. Such problem waters tend to be associated with water that has resided for a long time in the Tertiary Pidinga Formation sediments, which often contain pyrite and lignite. An undesirably high content of radioactive minerals is also common. Impurities such as these are not anticipated in the northern part of the Officer Basin, but are expected to become more prevalent to the south, parallelling the higher salinity. The drilling of water wells in the western portion, particularly on NOORINNA and WELLS map areas, have generally been successful within 100 m of surface, with yields of 23 L/s being common. The salinity is reasonably high, ~10 000 mg/L, and becomes more saline to the south. It seems most likely that the unconfined aquifer is reasonably
!"

INTRODUCTION
The Officer Basin of southern Australia occupies an area of ~375 000 km2, with its present-day margins bounded by the Yilgarn Craton to the west, Musgrave Block to the north and Gawler Craton to the southeast. The northeastern extent of the basin is poorly known, but its strata are presumed to continue beneath Mesozoic and Permo-Carboniferous rocks into the Amadeus and Warburton Basins. The relationship of the Officer Basin to surrounding regions is illustrated on

Figure 5.1, which also shows the approximate positions of the East Antarctic and Laurentia continental blocks in Neoproterozoic time. The Tasman Line marks the approximate margin of the Australian continental plate after the separation of Laurentia and before the accretion of microcontinents in Palaeozoic time. The mid-Neoproterozoic to Late Devonian (~820 360 Ma) geological evolution of the Officer Basin can most simply be described in four stages, each terminated by an

Fig. 5.1 Outline of the Officer Basin and adjacent regions of continental Australia. The cross-section is shown on Figure 5.3.

!#

Fig. 5.2 Neoproterozoic continental reconstruction (after Hoffman, 1991), showing Neoproterozoic and Cambrian saline basins: 1. Iran Pakistan basin (Neoproterozoic); 2. Centralian Superbasin (Neoproterozoic) and component Officer Basin (Neoproterozoic plus Cambrian); 3. North Canada basins (McKenzie, Colville, Big Bear; Cambrian); 4. Siberian Platform (Cambrian). Arrows crossing the Centralian Superbasin indicate direction of compression.

orogenic episode or a major unconformity. Adopting the terminology of Hoskins and Lemon (1995), these are: Stage 1 sag development of the ~2 x 106 km2 Centralian Superbasin, culminating in uplift and erosion (~780760 Ma). Stage 2 onset of compressional basin development culminating in the Petermann Ranges Orogeny (~560550 Ma). Stage 3 Cambrian deposition halted by uplift associated with the Delamerian Orogeny (~507 Ma). Stage 4 Ordovician to Devonian deposition terminated by the Alice Springs Orogeny (~360 Ma).

(Moores, 1991; Hoffman, 1991) places Laurentia opposite East Antarctica and Australia prior to opening of the Palaeopacific Ocean in Early Cambrian time. As Figure 5.1 shows, the North American Cordillera (Laurentia) is presumed to have faced south-central Australia across the Tasman Line. The similarities of the Neoproterozoic glacial and interglacial successions of the Adelaide Fold Belt and the North American Cordillera were emphasised by Young (1992). Using Hoffmans (1991) reconstruction (Fig. 5.2), there is a pronounced alignment of the major Neoproterozoic and Cambrian salt deposits although, by the end of the Early Cambrian, palaeomagnetic data show that Australia, Laurentia and Siberia would have been separated by the Palaeopacific Ocean. Most of these salt deposits are associated with petroleum source rocks (see Ch. 8). The Hormuz Series and Punjab Saline Series of the IranPakistan salt basin are Neoproterozoic, as is the Alinya
!$

PLATE TECTONIC SETTING


The position of Australia relative to other continental blocks in Neoproterozoic time is shown on Figure 5.2. The SWEAT (Southwest USEast Antarctic) hypothesis

Formation of the eastern Officer Basin and equivalents elsewhere in the Centralian Superbasin. Ouldburra Formation salt in the Officer Basin and the Saline River Formation of the McKenzie, Colville and Big Bear Basins in Canada (Morrell, 1995) are Early Cambrian in age. Spanning both the latest Proterozoic and Early Cambrian are the halite anhydritedolomite complexes of the Danilov and Usol Formations of the Siberian Platform (Kuznetsov et al., 1992). Officer Basin evolution in the era between the deposition of Neoproterozoic salt and Early Cambrian salt (~800 535 Ma) is punctuated by compressional tectonic activity which caused the Centralian Superbasin to be foreshortened in a northsouth direction and thus form into separate component basins, of which the Officer is the most southerly (Walter and Gorter, 1994). A cross-section through these basins, and a scale comparison with the Persian Gulf and Siberian regions, is shown on Figure 5.3. Lindsay and Leven (1996) regarded the polyphase evolution of the Officer Basin to have been governed by forces acting at continental plate scale. Shaw (1991) came to the same conclusion for the Amadeus Basin. Li et al. (1996) suggested that after ~800 Ma, the history of the Rodinia supercontinent was largely extensional as different crustal blocks separated episodically during the Neoproterozoic and Palaeozoic. Whilst this may have been the case for blocks on the Tasman Line margin of the supercontinent, compression was clearly the major force operating on the interior basins. Contrary to Li et al. (1996), it is also contended here that Laurentia and AustraliaAntarctica did not separate until Early Cambrian time (as evidenced by the great volumes of within-plate basalt in the Early Cambrian of southern Australia). The separation of Siberia and Laurentia is also constrained to the Early Cambrian (Pelechaty, 1996). The direction of compressional stress, which terminated stages 2 to 4 tectonism in the Officer Basin, is illustrated by opposing arrows on Figure 5.2. These deformations are interpreted here as responses to differential rotation between the China continental blocks to the north and the AustraliaAntarctic block to the south.

separated from the Officer Basin by the Karari Fault which exhibits dip-slip displacement greater than 1000 m, and has an aeromagnetic signature which can be traced for more than 300 km. Limbs (1980) suggestion that major faulting may have produced less dense material at depth suggests that the Karari Fault is overthrust to the northwest. The magnetic anomaly associated with the Karari Fault dips 60 towards 135 (Benbow, 1993), adding further evidence for thrust faulting. A southwestnortheast curvilinear belt of magnetic anomalies on the Gawler Craton and subparallel to the Karari Fault Zone has been interpreted by Daly (1996) as a Palaeoproterozoic collision zone between the Gawler Craton and the Yilgarn Craton the west. If so, the Karari Fault Zone and adjacent Tallaringa Trough overlie a mobile belt of considerable antiquity.
Tallaringa Trough and Nawa Ridge

The Tallaringa Trough (Fig. 5.4) is almost 200 km long, 40 km wide, and contains the richest oil-prone source rocks in the Officer Basin (see Chs 8 and 9). It was initially recognised from aeromagnetic data as a syncline, barely delineated by single eastwest flight lines (Steenland, 1965). Subsequent seismic and gravity surveys have detailed small portions of the Tallaringa Trough (Townsend, 1976; Milton, 1974, 1975; Benbow, 1993) and revealed it to be structurally complex. Near its centre, the trough contains up to 600 m of Cambrian strata and possibly 1600 m of underlying Neoproterozoic sedimentary rocks above magnetic basement. The Tallaringa Trough is bounded to the northwest by the Nawa Ridge, a positive structure composed of Gawler Craton crystalline rocks covered by a thin veneer of Neoproterozoic Officer Basin strata. Basement rocks are mainly gneiss and banded iron formation (BIF) which forms linear to complex, high-amplitude magnetic anomalies. This region of shallow basement is known as the Nawa Subdomain, and its northwestern margin (shown by dots on Fig. 5.4) is regarded as the subsurface margin of the Gawler Craton (Parker and Daly, 1993).
Northern margin Musgrave Block

BASEMENT STRUCTURAL ELEMENTS


Basement beneath the Officer Basin cannot be determined directly except at rare outcrops and in a few drillholes. However, by basinward extrapolation of the surrounding basement block characteristics using aeromagnetic and deep seismic data, something can be learned of what lies beneath the basin floor. Interpretations of aeromagnetic and/or gravity data in specific areas have been made by Finlayson (1979), Ashley (1984), Womer et al. (1987) and Benbow (1993), and reviewed at basin scale by Leven and Lindsay (1992). A map of total magnetic intensity illustrates some of the features of the basins deep structure and provides a basis for delineating its major structural elements (Fig. 5.4). Further details of the eastern reaches of the basin, where structural complexity is greatest, are shown on Figure 5.5.
Eastern margin Gawler Craton

A major contribution by Steenland (1965, p.12) was his recognition based on regional aeromagnetic data that the foredeep or trough [Munyarai Trough] just south of the front [S. margin of Musgrave Block]...is broken with many local structures...which are steep, probably thrust faulted on their southern flanks. It has taken a number of years, using a combination of magnetic, gravity and seismic data, to confirm that thrust faults do indeed truncate the northern margin of the Officer Basin (Milton and Parker, 1973; Lindsay, 1995; Lindsay and Leven, 1996). Outcropping crystalline rocks of the Musgrave Block have a complex aeromagnetic signature and consist of amphibolitegranulite facies acid metamorphic rocks (mainly gneiss) intruded by felsic plutons, and basic and ultrabasic rocks. The rocks may be up to 1600 million years old, but a major thermal event (Musgravian Orogeny) was recorded at ~1200 Ma (Major and Conor, 1993). Faults in
!%

The Archaean to Palaeoproterozoic Gawler Craton has been studied in detail because of its mineral potential. It is

Fig. 5.3 Cross-section through dismembered components of the Centralian Superbasin. The Persian Gulf Basin and Siberian Platform are compared at the same scale.

!&

Fig. 5.4 Aeromagnetic map of Total Magnetic Intensity showing principal structural elements of the eastern Officer Basin.

the Musgrave Block were reactivated during the Petermann Ranges and Alice Springs Orogenies.
Middle Bore Ridge

Mafic dykes

The Middle Bore Ridge lies south of the Manya Trough and is parallel to the Ammaroodinna Ridge (Figs 5.4, 5.5). Comalco Middle Bore 1, drilled in 1985, intersected crystalline basement beneath the Early Cambrian Ouldburra Formation. Basement is composed of undifferentiated pyroxene granulite in fault contact with the Cambrian cover rocks. The pyroxene granulite is part of the Nawa Subdomain, thus the Middle Bore Ridge is on the northwestern subcropping margin of the Gawler Craton. Aeromagnetic data confirm this interpretation (Hamer, 1994).
!'

Aeromagnetic surveys have revealed a system of northwest-trending parallel mafic dykes (Fig. 5.5) which pass northwestwards from the Gawler Craton, through the Nawa Subdomain including the Middle Bore Ridge, across the floor of the Munyarai Trough and into the Musgrave Block. These are part of the Gairdner Dyke Swarm which either slightly pre-dates or is the same age as the oldest Officer Basin strata. Orientation of the dyke swarm implies northeastsouthwest extension (Ashley, 1984; Cowley and Flint, 1993). An important characteristic of the aeromagnetic image is that there is little or no lateral offset of the dykes, thus strike-slip faulting has been very minor and the Ammaroodinna and Middle Bore Ridges are major dip-slip thrust features.

Fig. 5.5 Aeromagnetic map of Total Magnetic Intensity of the northeastern Officer Basin showing selected mafic dykes, locations of two seismic lines (see Fig. 5.6), and time structure contours of base of Cambrian (after Womer et al., 1987; Mackie, 1994; Lindsay, 1995).

Coompana Block

Structures beneath the basin

The Coompana Block, situated west of the Gawler Craton (Fig. 5.1), is composed of Archaean to Mesoproterozoic gneiss and granite intruded by northwest-trending mafic dykes (?equivalent to the Gairdner Dyke Swarm), and overlain by Nullarbor Platform cover. KAr geochronology indicates that the gneiss was deformed between ~1180 and 1160 Ma, indicating a Musgravian Orogeny overprint (Flint and Daly, 1993).
"

The northward thickening Birksgate Sub-basin and Munyarai Trough are overthrust by the Musgrave Block and ramp gently up-dip onto the broad Murnaroo Platform. Deep seismic data show basement beneath these regions to be pervaded by a complex of north-dipping reflectors which Lindsay and Leven (1996) interpreted as faults. These reflectors appear to commence near the base of the crust (depth ~42 km) and are terminated by the unconformity at

the base of the Officer Basin. Lindsay and Leven (1996, p.6) also pointed out that several of these faults have been reactivated by later compressional events to form major thrust complexes. A gravity and coincident magnetic anomaly known as the Nurrai Ridge strikes northsouth beneath the eastern Birksgate Sub-basin (Fig. 5.4). The ridge has been suggested to be responsible for facies differences between the eastern and central regions of the Officer Basin (Stainton et al., 1988; Zang, 1995a). The southern part of the anomaly was investigated by seismic which revealed no structural dislocation across it (Lindsay, 1995; Lindsay and Leven, 1996; Gravestock and Lindsay, 1994). Further south, there are other anomalies similar to the Nurrai Ridge (Leven and Lindsay, 1992), and a major magnetic feature runs northeasterly beneath the Murnaroo Platform (between arrows on Fig. 5.4). These are interpreted as very old features (Ashley, 1984) which have no discernible structural expression within the basin itself. However, some superimposed magnetic features may be thrusts. A channel or graben, presumably striking eastwest, is visible on seismic beneath the Officer Basin (Fig. 5.6a). This graben, ~7 km wide and a short distance north of Ungoolya 1, underlies the Pindyin Sandstone and Alinya Formation as illustrated by Hoskins and Lemon (1995), and Zang (1995a). The graben is considered by Zang (1995a, Zang and Preiss, in prep.) to contain strata of Willouran age, based on his assumption that the basal succession in Manya 5 is also within a similar graben, or represents graben fill. An alternative seismic interpretation shown on Figure 5.6b suggests that the basal section drilled in Manya 5 represents a sediment wedge that onlaps a basement outlier of the Ammaroodinna Ridge in a northwesterly direction (Mackie and Gravestock, 1993). The succession was not restricted to grabens but was widespread prior to erosion, as evidenced by outcrops of a succession similar in all respects to that cored in Manya 5, but 200 km distant in the Peake and Denison Ranges (Ambrose et al., 1981; see Ch. 6). A graben or channel clearly exists but has not yet been drilled. An analogous succession beneath the Amadeus Basin was thought to represent an early rift sequence (Lindsay and Korsch, 1991). This sequence of fluvial clastics and volcanics is now thought to pre-date the initial development of the Amadeus Basin (Korsch et al., 1993). The existence of these grabens may be of considerable significance for petroleum exploration. A grid of very large (hundreds of kilometres long by tens of kilometres wide) rifts or transverse aulacogens underlies the Siberian Platform and contains organic matter of Early Riphean age (~1000800 Ma). Average organic matter content of the rift fill is only 0.3% (Surkov et al., 1991), but these Riphean rocks are the source of most of the oil found in the Vendian and Cambrian of the Siberian Platform (A.Yu. Rozanov, Palaeontological Institute, Russian Academy of Sciences, pers. comm., 1995). It is unlikely that the sub-Officer Basin structures reached similar dimensions, but they should not be ignored as possible hydrocarbon source areas.
"

BASIN ARCHITECTURE
The thickest Officer Basin strata are contained in an arcuate string of asymmetric sub-basins or troughs which deepen towards the overthrust southern margin of the Musgrave Block. They are the Yowalga, Lennis and Waigen Sub-basins in Western Australia, and the Birksgate Sub-basin and Munyarai Trough in South Australia. The Birksgate Sub-basin and Munyarai Trough abut the almost flat-lying Murnaroo Platform to the south and are interpreted to be underlain by Mesoproterozoic crystalline basement (Lindsay and Leven, 1996). In the more structurally complex eastern reaches of the basin, basement is in fault contact with rocks as young as Cambrian. The chief elements of basin architecture are described below.
Munyarai Trough

The Munyarai Trough (the foredeep of Steenland, 1965) is an elongate southwestnortheast orientated foreland trough which deepens asymmetrically towards the Musgrave Block (Figs 5.4, 5.5). The trough contains up to 10 km of fill based on geophysical data (Womer et al., 1987) and is ~7500 km2 in area. The youngest rocks are of Late Devonian age and are lacustrine. In contrast to the Amadeus Basin, syn-orogenic Devonian conglomerates have not been recognised. The conglomerate drilled in Officer 1 (TD 183 m) is thought to be Permian glacial outwash. The deepest well in the trough is Munyarai 1 (TD 2899 m) drilled on a large anticline.
Birksgate Sub-basin

The Birksgate Sub-basin is adjacent to the Munyarai Trough and is separated from it by a broad structural arch which does not correspond to the Nurrai Ridge but lies east of it (Fig. 5.4). As far as can be determined from limited seismic data, the arch is 50100 km wide. The Nurrai Ridge is beneath the Birksgate Sub-basin (Lindsay, 1995; Fig. 5.4). A sediment thickness of 5 km is interpreted from seismic (Lindsay, 1995). The oldest (Willouran) strata form large elongate ridges and anticlinal outcrops near the northern thrust margin, and the youngest (Ordovician) are distributed as flat-lying outcrops further south (Krieg in Lindsay, 1995; Major, 1973b). The only well in the Birksgate Sub-basin is Birksgate 1 (TD 1878 m).
Marla Overthrust Zone

The most significant oil shows in the Officer Basin are from nine mineral and stratigraphic wells drilled in the Marla Overthrust Zone, as reviewed by Hibburt et al. (1995). The Marla Overthrust Zone lies at the eastern end of the Munyarai Trough (Fig. 5.5) and is a complex series of anastomosing thrust ramp and duplex structures which affect rocks as young as Ordovician (Mackie, 1994; Gravestock and Lindsay, 1994; Lindsay, 1995; Hoskins and Lemon, 1995). This zone marks the point of closest convergence of the Musgrave Block and Gawler Craton, and the propagation of thrust faults into the Cambrian cover may result from lack of a dcollement surface at the basin floor (J.F. Lindsay, AGSO, pers. comm., 1995). This could be due to lack of evaporites in the Alinya

Aliny

a Fo

rmat

ion

Graben

Crystalline Basement

0 KILOMETRES

Crystalline Basement

0 KILOMETRES

97-0235 MESA

Fig. 5.6 Seismic sections: (a) part of Amoco Line IP-2A (vibroseis) showing graben; (b) part of Comalco lines 83-600 and 84-620 (thumper), showing onlap of the lowest strata onto crystalline basement (after Mackie, 1994).

"

Formation in this part of the basin. Ordovician rocks are preserved at the surface and the oldest unit drilled to date is the Meramangye Formation in Marla 9. The basin fill has not been fully drilled in the Marla Overthrust Zone, most wells having reached total depth in the Cambrian Observatory Hill Formation.

Sub-basin onto the Murnaroo Platform (Lindsay, 1995; Lindsay and Leven, 1996), whereas in the eastern portion the transition from the Munyarai Trough to Murnaroo Platform is more abrupt, via the Ammaroodinna Ridge (Stainton et al., 1988; Thomas, 1990). Four key wells Murnaroo 1, Observatory Hill 1, Lake Maurice West 1 and Lake Maurice East 1 have been drilled on the Murnaroo Platform, but only the last listed has penetrated to crystalline basement. In Lake Maurice East 1, high-grade gneiss of the Palaeoproterozoic Nawa Subdomain is overlain unconformably at 691 m by the Marinoan Tarlina Sandstone. The Cambrian Arcoeillinna Sandstone and Observatory Hill Formation comprise the youngest outcropping strata.

Ammaroodinna Ridge
A small outcrop of basement rocks named the Ammaroodinna Inlier occurs immediately southwest of the Marla Overthrust Zone and ~80 km south of the Musgrave Block (Krieg, 1972, 1993). It is composed of schist, gneiss and granitoid rocks and, like the Musgrave Block, was also affected by the ~1200 Ma Musgravian Orogeny. However, this sliver of basement has been emplaced tectonically. Krieg (1972) noted that faults bounding the inlier also displaced the Cambrian Observatory Hill Formation and thus were relatively young. Seismic records show that the inlier is located on a major southwest-trending thrust fault complex the Ammaroodinna Ridge that separates the Munyarai and Manya Troughs (Stainton et al., 1988; Thomas, 1990; Mackie, 1994). The Ammaroodinna Ridge is at least 140 km long (Figs 5.4, 5.5) and, though unnamed, was first recognised on aeromagnetic data by Steenland (1965). The ridge has been variously named Ammaroodinna High by Stainton et al. (1988), Ammaroodinna High Platform by Thomas (1990) and Ungoolya Hinge by Zang (1995b). All of these investigators recognised that the ridge or hinge controlled sedimentation, facies changes and structural style within the basin.

Nullarbor Platform
The Nullarbor Platform is a stable region of relatively thin (~2 km) sedimentary cover which overlies the Coompana Block. In ascending stratigraphic order, the platform comprises NeoproterozoicCambrian Officer Basin, Permo-Carboniferous Denman Basin, JurassicCretaceous Bight Basin, and Tertiary Eucla Basin. The modern surface of the platform is the Nullarbor Plain. The Nullarbor and Murnaroo Platforms are contiguous but are almost separated locally by shallow crystalline basement as drilled in Ooldea 1 (Fig. 5.7). This basement rise, named the Watson High (Rankin in Parker, 1993, fig. 2.4), also forms the southern margin of the Tallaringa Trough. The deepest well on the Nullarbor Platform is Mallabie 1 which drilled 905 m of sandstone, arkose and basic volcanics before intersecting the Coompana Block. The age of the volcanics is unknown, but they post-date the ~1200 Ma Musgravian Orogeny and are undeformed. The Cadlareena Volcanics of the Peak and Denison Inliers (Ambrose et al., 1981) and the volcanics intersected in Manya 5 may be correlatives.

Manya Trough
The Manya Trough lies between the Ammaroodinna Ridge and Middle Bore Ridge. Like the adjacent Wintinna Trough (Fig. 5.5), the Manya Trough is characterised by coincident positive gravity and negative aeromagnetic anomalies. The high gravity value results from carbonates of the Ouldburra Formation which are almost 1000 m thick. Manya Trough strata have not been fully penetrated, the deepest well being Manya 6 which reached TD 1765 m in the Relief Sandstone. It is worth noting that the numerous halite beds at the OuldburraRelief transition have not been tectonically disturbed in this well, despite its proximity to the Marla Overthrust Zone. Prior to recognition of the Middle Bore Ridge, the Manya and Wintinna Troughs were not differentiated. Here, the Wintinna Trough is considered as a structural element of the Arckaringa Basin because it contains at least 800 m of Permian strata (cf. 324 m in Manya 6).

Murnaroo Platform
The Murnaroo Platform extends south of the Birksgate Sub-basin and Munyarai Trough. It is a poorly defined region crossed diagonally by the Ammaroodinna Ridge which divides the platform into eastern and western portions (Fig. 5.4). Regional seismic data in the western portion indicate a very gentle (0.3) ramp from the Birksgate
"!

Fig. 5.7 Cross-section through the Nullarbor Platform, the most southerly structural element of the Officer Basin.

Drillholes Hughes 1, 2 and 3, located ~120 km northwest of Mallabie 1, intersected clastics and carbonates assumed by Gravestock and Hibburt (1991) to be Cambrian in age. However, seismic recorded in 1993 indicates that the Cambrian has been eroded (Lindsay, 1995; Lindsay and Leven, 1996), thus the Hughes wells are now considered to have intersected Neoproterozoic strata; lithologies are reminiscent of the Alinya Formation.

faults, the gravity driven collapse of half-graben fill and detachment on salt beds in the Alinya Formation. Badley also noted that at least two compressions and one major uplift had affected the basin since salt withdrawal in the Alinya Formation. According to Hoskins and Lemon (1995), compression was also responsible for: uplift of the Musgrave Block, culminating in the Petermann Ranges Orogeny northward tilt of the basin floor, sediment loading and salt movement submarine canyon erosion. The Petermann Ranges Orogeny (~560550 Ma) severed connection with the Centralian Superbasin by way of the Musgrave Block. Earlier workers (e.g. Gravestock and Sansome, 1994) ascribed the main phase of thrust-related deformation to the Late Devonian Alice Springs Orogeny, but Hoskins and Lemon are emphatic, and Lindsay and Leven (1996) concur, that the Petermann Ranges Orogeny was responsible for an earlier phase of thrusting. Up to 3000 m of strata are estimated to have been eroded during this orogeny (Moussavi-Harami and Gravestock, 1995). Stage 2 of Officer Basin development is shown schematically on Figure 5.8b.

STRUCTURAL HISTORY
Stage 1
Originally the Officer Basin was a component of the 2 x 106 km2 Centralian Superbasin, a giant sag basin initiated by crustal extension as evidenced by the Gairdner Dyke Swarm (Ashley, 1984). However, as Lindsay and Leven (1996, p.18) pointed out, the thick crust and north-dipping deep crustal reflectors suggest large-scale subcrustal processes rather than mechanical extension. Similarly, the sub-basin channel or graben near Ungoolya 1 implies that the Officer Basins prehistory is more complex than previously envisaged. Seismic stacking patterns and the even spacing of basal and top reflectors indicate that Stage 1 sediments represent the infilling of a broad, shallow, intra-cratonic sag basin (Hoskins and Lemon, 1995, p.398). Thomas (1990) envisaged an initial platform stage followed by creation of a topographic gradient as salt migrated within the Alinya Formation. According to Lindsay and Leven (1996), evaporitic units within the Alinya Formation began to flow soon after deposition and continued sporadically until the Devonian Alice Springs Orogeny. Stage 1 is illustrated on Figure 5.8a. There is a major hiatus between the Alinya Formation and overlying late Neoproterozoic in much of the eastern Officer Basin. An intervening Sturtian tillite is known only from the northeastern reaches of the basin, suggesting a lacuna of at least 20 million years. However, as the Alinya Formation and overlying units are structurally concordant, there is no evidence of a Stage 1 terminal orogeny. Moussavi-Harami and Gravestock (1995) have estimated uplift and erosion of 100500 m of strata from above the Alinya Formation due to glaciogenic processes, but a reason for the termination of Stage 1 remains essentially unknown.

Stage 3
During Stage 3, the thickest Cambrian sediments, chiefly carbonates, were deposited in elongate troughs between the Karari Fault Zone and Ammaroodinna Ridge. Thinner sediment packages extended into the Munyarai Trough to the northwest and onlapped the Murnaroo Platform and probably also the Gawler Craton to the southeast. Lindsay and Leven (1996) suggested that regional subsidence in a mildly extensional regime characterises the Cambrian; alluvial fan conglomerates (e.g. Wallatinna Member of the Observatory Hill Formation) attest to contemporaneous localised uplifts. Hoskins and Lemon (1995) illustrated a diapiric piercement stage (Fig. 5.8c) which Thomas (1990) suggested may have occurred somewhat earlier during the Neoproterozoic. Cambrian deposition in the Adelaide Fold Belt was halted by the Delamerian Orogeny which commenced at ~507 Ma. In the Officer Basin, the Delamerian Orogeny caused uplift and erosion, estimated to have exceeded 2000 m in the far northeast of the basin (Moussavi-Harami and Gravestock, 1995). Stage 2 thrust faults were also reactivated (Badley, 1988; Hoskins and Lemon, 1995).

Stage 2
Stage 2 spans late Neoproterozoic deposition, terminating in the Petermann Ranges Orogeny. This stage may be divided into upper and lower successions by a localised though prominent canyon incision surface which initiated on the Ammaroodinna Ridge and extended northwestwards into the Munyarai Trough, with canyons up to 700 m deep (Thomas, 1990; Sukanta et al., 1991; Lindsay and Leven, 1996). Hoskins and Lemon (1995) noted that northsouth compression was responsible for differentiation of the basin into the Munyarai Trough and Murnaroo Platform, separated by the Ammaroodinna Ridge (cf. extensional model of Gravestock and Sansome, 1994). However, Badley (1988) was the first to recognise that the seismic reflection pattern is explicable by a model involving up-slope propagation of
""

Stage 4
Relatively little is known of Stage 4 which commenced with deposition of a thick wedge of Ordovician (?and Silurian) siliciclastic sediments and culminated in the Late DevonianEarly Carboniferous Alice Springs Orogeny. About one-quarter of the sediment fill in the northern Munyarai Trough is Devonian on seismic evidence but, as the section is undrilled, the presence of a syn-orogenic molasse such as the Brewer Conglomerate of the Amadeus Basin remains speculative (Fig. 5.8d). Lindsay and Levens (1996) estimate of Devonian in the trough (almost half) appears too high. Part of this section may alternatively be

Fig. 5.8 Schematic section from the Murnaroo Platform to Musgrave Block, showing four stages of Officer Basin evolution (after Thomas, 1990; Hoskins and Lemon, 1995).

OrdovicianSilurian (Gravestock and Sansome, 1994; Hoskins and Lemon, 1995), but this is also untested. Heat flow associated with the Alice Springs Orogeny was surprisingly widespread as evidenced by fission track annealing in apatites from pre-Devonian strata (see Ch. 9). Thrust faults were reactivated and others may have been initiated; locally they propagated as ramps and duplex structures in the Marla Overthrust Zone (Mackie and

Gravestock, 1993) and Ammaroodinna Ridge (Hoskins and Lemon, 1995). Elevation of basement blocks in central Australia (e.g. Musgrave Block, Arunta Block) has been suggested as a major trigger for Carboniferous glaciation (Veevers and Powell, 1989), which initiated the Gondwanan depositional phase over the eroded remnants of the Centralian Superbasin.

"#

"$

INTRODUCTION
The Officer Basin spans 350 000 km2 of central Australia, from the Yilgarn block in Western Australia to the Gawler Craton in South Australia. It is an arcuate depression ~500 km long with six main depocentres containing up to 10 000 m (Womer et al., 1987) of gently folded Neoproterozoic and Palaeozoic sediments. In spite of its large areal extent, the stratigraphy of the basin is relatively poorly known compared to other South Australian basins, possibly due to: Wells with stratigraphic information are often shallow, and are widely spaced. Outcrops are poor and of limited stratigraphic interval because of low dips and subdued topography; this has led to many, possibly synonymous, names being introduced. The stratigraphy is complex, with at least 33 mappable stratigraphic units identified. Biostratigraphic control is poor. The early Palaeozoic sediments are largely non-marine and lithologically similar to Neoproterozoic sediments. Biozonation of the Neoproterozoic based on acritarchs has recently been achieved (Zang, 1994, 1995a,b), which has significantly improved correlation. Previous nomenclatures are summarised on Figure 6.1, and that recommended for use in the future on Figure 6.2. This chapter has been compiled using the published data and interpretations of Major and Teluk (1967), Major (1973a,b,c, 1974), Benbow (1982), Gatehouse et al. (1986), Brewer et al. (1987), Gaughan and Warren (1990), Gravestock and Hibburt (1991), Preiss (1993), Gravestock and Sansome (1994), Zang (1994, 1995a,b), Gravestock et al. (1995) and Moussavi-Harami and Gravestock (1995). Radiometric ages of the Palaeozoic section are based on those in Tucker and McKerrow (1995). Because of the poor biostratigraphic control, sequence stratigraphy has been used as an aid to correlation (Thomas, 1990; Gravestock and Hibburt, 1991; Sukanta, 1993; Gravestock and Sansome, 1994; Zang, 1995a,b; MoussaviHarami and Gravestock, 1995). Depositional sequences in open marine and deep sea environments are demonstrated to be controlled by eustatic sea-level change. However, the intracratonic Officer Basin was strongly influenced by tectonism which generated accommodation space. The major sequence boundaries in the eastern Officer Basin are correlated with tectonically generated unconformities
"%

(Moussavi-Harami and Gravestock, 1995; Zang, 1995b). Twelve sequence sets were recognised by Preiss (1993) in the Neoproterozoic in the Adelaide Fold Belt (Adelaide Geosyncline), but in the eastern Officer Basin they are incomplete. Ten sequence sets (super-sequences) from the Neoproterozoic to the Devonian can be recognised, some sequence boundaries can be observed in core and in the field, but most are interpreted from seismic data. Some have been divided into smaller (third order) sequences. Systems tracts (Moussavi-Harami and Gravestock, 1995) have also been assigned for each sequence (Fig. 6.2). These include lowstand systems tract (LST), transgressive systems tract (TST), highstand systems tract (HST), maximum marine flooding surface (MFS) and incised valley fill (IVF). The subcrop geology of the Officer Basin is shown on Figure 6.3.

OFFICER BASIN
UNDRILLED SEQUENCE BELOW THE CALLANNA GROUP
On some seismic lines (e.g. IP1-2A, Fig. 5.6), a sequence of sediments is present in valley fill or narrow grabens ~10 km wide and up to 1300 m thick. This sequence has not been drilled, but is older than the Callanna Group equivalents. Zang (1995a) suggested that the Pindyin Sandstone and Alinya Formation are younger than the Callanna Group (equivalent to the Burra Group), and that the undrilled sequence is the Younghusband Conglomerate and Coominaree Dolomite of the Peake and Denison Inliers, which are considered to be Callanna Group equivalents. The sediments may be an equivalent of the Pandurra Formation (fluvial sandstone) on the Gawler Craton and Stuart Shelf, which is of late Mesoproterozoic age (Mason et al., 1978; Preiss, 1987). The Moorilyanna Formation (Wilson, 1952; Coats, 1963; Gravestock et al., 1995) and the Levenger Formation (Major, 1973d), both of which occur in isolated grabens in the Musgrave Block, may be equivalents. In the Amadeus Basin, the Mount Harris Basalt, and the fluvial Bloods Range and Dixon Range beds, interpreted to be an early rift sequence (Lindsay and Korsch, 1991), may also be equivalents. The nature of these sediments will remain speculative until they are drilled.

CALLANNA GROUP EQUIVALENTS


The Pindyin Sandstone, Alinya Formation, Coominaree Dolomite and Cadlareena Volcanics were probably deposited in response to the initial Neoproterozoic sag during the

Fig. 6.1 Summary of previous stratigraphic nomenclature.

Willouran, and have been designated Sequence W. The Coominaree Dolomite, Cadlareena Volcanics and an earlier Pindyin Sandstone equivalent (Younghusband Conglomerate) in the Peake and Denison Inliers have been correlated with the Arkaroola Subgroup of the Callanna Group in the Adelaide Fold Belt (Preiss, 1987; Rogers and Freeman, 1996). Seismic evidence indicates that the Callanna Group is widespread subsurface in the Officer Basin. It forms the major detachment for propagation of thrust faults and is the

source of salt in halokinetic structures (Gravestock and Sansome, 1994). To the east of the Marla Overthrust Zone, the facies change from clastic to carbonates and volcanics (Coominaree Dolomite and Cadlareena Volcanics), although Zang (1995a) considered these to be older than the Pindyin SandstoneAlinya Formation succession. The lowest siltstone of the Alinya Formation was probably deposited in a TST, with seismic interpretation suggesting the possible onlap from northwest to southeast (Zang and McKirdy, 1993; Zang, 1995a). The Pindyin

"&

Fig. 6.2 Summary of current stratigraphic nomenclature (Palaeozoic dates after Tucker and McKerrow, 1995).

"'

MUSGRAVE BLOCK

Pindyin Hills Punkerri Hills

Purndu Hills

28

EROMANGA BASIN

OFFICER BASIN

29

1 Chambers Bluff 2 Mt Johns Range 3 Indulkana Range

ARCKARINGA BASIN
0
130

50 KILOMETRES
132

GAWLER CRATON
134
97-0070 MESA

MESOZOIC CRETACEOUSJURASSIC Bulldog Shale, Cadna-owie Formation, Algebuckina Sandstone PALAEOZOIC PERMIAN Mount Toondinna Formation, Stuart Range Formation, Boorthanna Formation DEVONIAN Mimili Formation ORDOVICIANSILURIAN Munda Group CAMBRIAN Kulyong Formation, Marla Group NEOPROTEROZOIC MARINOAN Ungoolya Group, Lake Maurice Group STURTIAN Wantapella Volcanics, Chambers Bluff Tillite TORRENSIANWILLOURAN Callanna Group MESOPROTEROZOIC Basement Stratigraphic well Dry exploration well Well with oil show

Pindyin Sandstone
Definition and nomenclature Thomson (1969) first used the term Pindyin Sandstone, which was later defined as Pindyin Beds by Major (1973b), and renamed Pindyin Sandstone by Zang (1995b). The Townsend Quartzite (Townson, 1985) of the western Officer Basin may be a synonym. Type section North Pindyin Hills (outcrop), as defined by Major (1973b). A subsurface reference section is defined as 12891326.8 m in Giles 1 (Fig. 6.4; TD was in Pindyin Sandstone). Lithology and distribution The formation is generally composed of fine to coarse-grained quartzose sandstone. In places, the sandstone is feldspathic and contains rare quartz pebbles and clay galls.

Fig. 6.3 Subcrop geology map (pre-Tertiary) of the Officer Basin.

Sandstone is inferred to be a LST to TST. The upper part of the Alinya Formation and Coominaree Dolomite represent a HST in a sabkha to marine palaeoenvironment; the top was extensively eroded during glaciogenic uplift in the Sturtian. The Alinya Formation is overlain by the Cadlareena Volcanics east of the Marla Overthrust Zone (MoussaviHarami and Gravestock, 1995).

Outcrop of Pindyin Sandstone, Belundinna Hill.

(Photo 44370)

#

Sedimentology and palaeoenvironment In the lower part of the type section, the presence of conglomerate along with sedimentary structures such as trough cross-bedding and some large-scale, low-angle cross-bedding suggest a fluvial palaeoenvironment. Palaeocurrent interpretation of cross-beds, the southward thinning of the conglomerate and the orientation of ripple marks imply a northern provenance. The overlying sandstone is characterised by herring bone cross-bedding, asymmetrical ripple marks and mudcracks, probably indicating a tidal or peritidal depositional setting, with finer grained transgressive shallow marine sandstone higher in the section. The Pindyin Sandstone on the Murnaroo Platform is interpreted to have a different depositional setting, possibly aeolian, due to the presence of minor anhydrite, halite, and well-rounded haematite-rimmed quartz grains.

Alinya Formation
Definition and nomenclature The formation was included in the Pindyin Beds by Major (1973b). It was originally named the Alinya beds by Stainton et al. (1988), and renamed Alinya Formation by Zang (1995b). Type section Not originally defined, but here proposed as 1233 1289 m in Giles 1 (Fig. 6.4).

Fig. 6.4 Giles 1 type section for Alinya Formation, and reference section for Pindyin Sandstone.

A basal pebble conglomerate occurs at the type section. The upper part of the unit comprises pink quartzose sandstone, and minor shale interbeds. The unit is presumed to be widespread in the deeper parts of the basin based on seismic data, and may occur on the Nullarbor Platform. Relationships and boundary criteria The Pindyin unconformably overlies crystalline basement or the older undrilled sequence, and is conformably overlain by the Alinya Formation. It may be equivalent to the basal part of the Younghusband Conglomerate of the Peake and Denison Inliers (although Zang (1995b) disputed this and considers the Younghusband Conglomerate to be older), the Heavitree Quartzite in the Amadeus Basin, and the Arkaroola Subgroup in the Adelaide Fold Belt. Thickness In outcrop the sandstone is ~200 m thick; in Giles 1 only the top 39 m were penetrated, but seismic data suggest the thickness ranges from 100 to 200 m. Age Precambrian, Neoproterozoic, Willouran to Torrensian Epochs. Zang (1995b) considered the unit to be Torrensian because of its wide distribution based on seismic evidence.
#
Herringbone cross-bedding in the Pindyin Sandstone, Belundinna Hill. (Photo 44371)

Tidal-influenced ripple marks in the upper part of the Pindyin Sandstone, North Pindyin Hills. (Photo 44372)

Alinya Formation (lower part) at 1272 m in Giles 1, showing red-brown sandstone and grey-green siltstone interbeds. Scale bar is 10 mm for each black and white grid. (Photo 44373) Pindyin Sandstone at 1324.51327 m in Giles 1; the sandstone is of probable aeolian to fluvial origin. Scale bar is 10 mm for each black and white grid. (Photo 42401a)

Thickness The thickness is 56 m in Giles 1 but may reach 500 m, derived from seismic data. Age Precambrian, Neoproterozoic, Willouran to Torrensian Epochs, acritarch assemblage AAW 1a (Giles 1) and AAW 1b (Manya 5). Zang (1995b) considered the unit to be Torrensian. Sedimentology and palaeoenvironment The lower siltstone and sandstone units were deposited on a tidal flat. The upper cyclic unit was deposited in a coastal sabkha setting. The sandstone at the top of each cycle is of aeolian origin.

Lithology and distribution Zang (1995b) divided the Alinya Formation into upper and lower units. The lower unit consists of red-brown siliciclastics (siltstone and sandstone) and evaporite (anhydrite). The upper unit comprises stacked cycles of grey siltstone, black shale, grey-green silty shale, anhydrite and red-brown siltstone and dolomite beds rich in microbial matter (cyanobacterial mats), with sand capping the cycles. The Alinya is widespread, and may occur on the Nullarbor Platform, but is interpreted to change facies in the east into the Coominaree Dolomite. Relationships and boundary criteria The formation conformably overlies the Pindyin Sandstone, and is unconformably overlain by the Tarlina Sandstone or conformably overlain by the Cadlareena Volcanics. In the Peake and Denison Inliers, the equivalent is the upper part of the Younghusband Conglomerate (Ambrose et al., 1981; cf. Zang, 1995b), and in the Amadeus Basin the equivalent is the Bitter Springs Formation (Gillen Member). The Browne and Lefroy beds (Townson, 1985) of the western Officer Basin, and the lower part of the Wright Hill beds (Major, 1973c), may be equivalents.
#

Coominaree Dolomite
Definition and nomenclature The unit was defined in the Peake and Denison Inliers by Ambrose et al. (1981). Type section Outcrop in the Peake and Denison Inliers, latitude 282635S, longitude 1355911E. A subsurface reference

Alinya Formation (upper part) at 12421245 m in Giles 1, showing grey-green shale and red-brown siltstone; anhydrite is common (Zang, 1995b). Scale bar is 10 mm for each black and white grid.
(Photo 44374)

section for the Officer Basin is here defined as 11551265 m in Manya 5 (Fig. 6.5). Lithology and distribution The formation has been subdivided into two units in outcrop. The lower is interbedded dolomite with minor sandstone and conglomerate. The upper unit consists of dolomite, which is oolitic at the base and stromatolitic at the top. The Coominaree is restricted to the eastern part of the basin, in the Manya Trough area. Relationships and boundary criteria The unit is probably a lateral facies equivalent of the upper Alinya Formation, although this is disputed by Zang (1995b), and may correlate with the Skates Hill Formation (of the western Officer Basin, Savory Sub-basin). It is conformably overlain by the Cadlareena Volcanics. Thickness Up to 77 m in outcrop, but 110 m thick in Manya 5. Age Precambrian, Neoproterozoic, Willouran to Torrensian Epochs. Zang (1995b) considered the unit to be Willouran. The stromatolites have been identified as Acaciella sp. (possibly A. australica), which is apparently restricted to the Callanna Group (Grey, 1995). Sedimentology and palaeoenvironment The Coominaree Dolomite was probably originally calcareous or aragonitic. The presence of stromatolites
#!

Fig. 6.5 Manya 5 reference section for Coominaree Dolomite and Cadlareena Volcanics.

indicates a shallow marine palaeoenvironment (Ambrose et al., 1981).

Cadlareena Volcanics
Definition and nomenclature The formation was defined in the Peake and Denison Inliers by Ambrose et al. (1981). Type section Outcrop in the Peake and Denison Inliers: basal part latitude 28621S, longitude 1355329E; upper part latitude 284245S, longitude 135159E. A subsurface

Thickness The thickness is 79 m in Manya 5, and may be up to 730 m in outcrop in the Peake and Denison Inliers. Age Precambrian, Neoproterozoic, Willouran to Torrensian Epochs (780760 Ma). Zang (1995b) considered the unit to be Willouran. Radiometric dating of possible correlatives in the Adelaide Fold Belt (Boucaut Volcanics; Fanning, 1989) gave an average UPb age of 78342 Ma, and in the Polda Basin (Kilroo Formation; Flint et al., 1988) gave KAr ages on plagioclase of 7689 and 76442 Ma. Sedimentology and palaeoenvironment reference section for the Officer Basin is here defined as 10761155 m in Manya 5 (Fig. 6.5). Lithology and distribution The unit has been intersected only in Manya 5 in the northeastern Officer Basin, where it consists of basalts, redbeds and pyroclastics. The amygdaloidal basalt and redbeds in Mallabie 1 in the far south of the basin may also be attributable to this formation. In outcrop, the formation comprises altered vesicular basalt and dolerite, with minor andesite, dacite, rhyolite, and lapilli tuff beds. Lenticular red shale and quartzose sandstone may also be present. The formation is not observed in Giles 1, the only other well to penetrate the Alinya Formation, and distribution may be patchy due to erosion. Relationships and boundary criteria The lower boundary is probably conformable on the Coominaree Dolomite or Alinya Formation. In the type area (Peake and Denison Inliers), the unit either unconformably overlies crystalline basement or conformably overlies Younghusband Conglomerate. The upper boundary is unconformable, and overlain by the considerably younger Tarlina Sandstone. The Cadlareena Volcanics may be equivalent to the Wooltana and Boucaut Volcanics, which covered an area of ~210 000 km2 in the Adelaide Fold Belt, and the Kilroo Formation in the Polda Basin (Flint et al., 1988). The volcanics are subaerial lava flows, with individual flows 2030 m thick. Reworking of the volcanics into shallow water (?lacustrine) environments is evident in outcrop.

Coominaree Dolomite at 1160.11160.3 m in Manya 5; note the stromatolites (Acaciella sp.). (Photo 42400b)

UMBERATANA GROUP EQUIVALENTS


Neoproterozoic glaciogenic successions in the Adelaide Fold Belt (Umberatana Group) include lower Sturtian tillites, upper Marinoan tillites and thick (up to 4000 m) interbedded siliciclastics and carbonate sediments. Equivalent units in the Officer Basin region are informally assigned Sequence S and crop out northwest of the Marla region, but have not been intersected in the Munyarai Trough or on the Murnaroo Platform. The Chambers Bluff Tillite and Wantapella Volcanics are exposed in fault-bounded outcrops along the northern overthrust margin of the basin. The relationship of these units to the underlying Callanna Group and overlying Lake Maurice Group sequences is poorly known due to the lack of subsurface intersections, but they occur in Nicholson 2 below the Tarlina Sandstone and are presumed to occur in the deeper parts of the Manya Trough (Fig. 6.6).

Chambers Bluff Tillite


Definition and nomenclature The unit was defined by Wilson (1952). Townson (1985) defined the Lupton and Turkey Hill beds in the western Officer Basin, which are considered synonymous. Type section Outcrop 9.6 km north-northwest of Chambers Bluff on the northern overthrust margin of the basin. A subsurface reference section is defined as 384816.3 m in Nicholson 2. Lithology and distribution The tillite comprises yellow to pale green, pebbly, silty diamictite in the lower part with beds of sandstone becoming more common westwards. A 1.2 m thick bed of partly sandy, partly stromatolitic, flaggy pink-buff dolomite is present towards the top of the formation. The unit is poorly known in the subsurface, but is present east of the Manya Trough, and may be more widespread.
#"

Cadlareena Volcanics at 1118.51119.65 m in Manya 5; the arrow indicates a contact between the underlying mafic volcanics and an interbedded red sandstone. Scale bar is 10 mm for each black and white grid. (Photo 42400c)

Fig. 6.6 Chambers Bluff Tillite and Wantapella Volcanics cross-section (Chambers Bluff, Wantapella Anticline, Nicholson 2). Nicholson 2 lithology by W. Zang.

Relationships and boundary criteria The formation is unconformably(?) overlain by the Wantapella Volcanics (Krieg, 1973) or the Lake Maurice Group, and in the Peake and Denison Inliers unconformably overlies the Callanna Group. It is equivalent to a range of glacigene formations of the Adelaide Fold Belt (Preiss, 1987), the Calthorinna Tillite in the Peake and Denison Inliers (Ambrose et al., 1981) and the Areyonga Formation in the Amadeus Basin. Thickness The thickness is over 432 m in Nicholson 2. Age Precambrian, Neoproterozoic, Sturtian Epoch.
##
Glaciogenic diamictite of the Chambers Bluff Tillite at 678.0 678.15 m in Nicholson 2. Scale bar is 10 mm for each black and white grid. (Photo 42403a)

Sedimentology and palaeoenvironment Palaeoenvironments range from fluvioglacial, glaciolacustrine to supraglacial. Erratics of fresh basalt in the fluvioglacial deposits and the diamictite may be from eroded Cadlareena Volcanics, and black chert pebbles probably came from eroded remnants of Torrensian rocks which are several kilometres thick in the Peake and Denison Inliers 480 km to the southeast (Gravestock and Sansome, 1994).

LAKE MAURICE GROUP


Zang (1995a) introduced the name Maurice Group for the transgressiveregressive cycle designated Sequence M by Moussavi-Harami and Gravestock (1995). Due to prior usage of the name elsewhere, it is here modified to Lake Maurice Group. The Lake Maurice Group consists of the Tarlina Sandstone, Meramangye Formation and Murnaroo Formation, forming a complete transgressiveregressive cycle. The basal Tarlina Sandstone contains a bed of lowstand fluvial conglomerate and grades upwards into transgressive to highstand siltstone and silty mudstone of the Meramangye Formation. A regression is indicated by continuously shallowing-upwards into shoreface and estuarinefluvial sediments of the Murnaroo Formation. The Officer Basin was subject to relatively rapid subsidence during deposition of the Lake Maurice Group; this is particularly evident on the Murnaroo Platform. In Giles 1, the Lake Maurice Group (Sequence M) contains 585 m of sediment, while Sequence W and the Ungoolya Group are only 270 and 63 m thick, respectively. Rapid subsidence might also have resulted in more siliciclastics being transported into the basin.

Wantapella Volcanics
Definition and nomenclature The name was used by Krieg (1973) and later by Preiss (1993) without formal definition. Type section Here designated as the Chambers Bluff section of Preiss (1993). Lithology and distribution The volcanics consist of altered grey-green amygdaloidal tholeiitic basalt. A few gritty sandstone interbeds are present. The formation has not been intersected in the subsurface, but may be present in the Marla Overthrust Zone. Relationships and boundary criteria The lower boundary with the Chambers Bluff Tillite is inferred to be disconformable, but has not been observed in outcrop (Preiss, 1993). The overlying unit contains a conglomerate that is entirely of volcanic derivation, and Preiss (1993) suggested that the time break may not have been very long. Thickness The thickness in outcrop appears to be ~190 m, and probably increases to the northeast. Age Precambrian, Neoproterozoic, either late Sturtian or early Marinoan Epochs. Sedimentology and palaeoenvironment The volcanics were probably extruded subaerially as multiple flows.

Tarlina Sandstone
Definition and nomenclature Originally introduced as Tarlina beds by Stainton et al. (1988), and changed informally to Tarlina Sandstone by Sukanta (1993), the unit was formally defined by Zang (1995a). Type section The type section is here redefined from Zang (1995a) to 10641233 m in Giles 1 (Fig. 6.7). A reference section was defined by Zang (1995b) as 532691.3 m in Lake Maurice East 1 (TD was in Tarlina Sandstone). Lithology and distribution The lithology consists predominantly of sandstone with minor thin silty mudstone interbeds. The sandstone is brown to light brown, fine to coarse-grained, quartzose to feldspathic. The base is conglomeratic and the amount of mudstone increases up-section. The unit occurs on the Murnaroo Platform and is interpreted seismically in the

Outcrop of Chambers Bluff Tillite and Wantapella Volcanics unconformably overlain by lower Lake Maurice Group at Chambers Bluff.
(Photo T5965)

Coarse alluvial sandstone of the Tarlina Sandstone at 1227.4 1227.6 m in Giles 1. Scale bar is 10 mm for each black and white grid. (Photo 42404b)

#$

Meramangye Formation or Murnaroo Formation. The Meramangye Formation may be a lateral equivalent in the Chambers Bluff and Wantapella Anticline area. Thickness The formation is 169 m thick in Giles 1, but is up to 373 m thick in Manya 5. Age Precambrian, Neoproterozoic, early Marinoan Epoch, acritarch assemblage AAM 1. Sedimentology and palaeoenvironment The sandstone at the type section is poorly sorted, commonly massive and lacks bedding features. Graded bedding is often present when sandstone is associated with thin silty mudstone. Angular breccia or pebble beds are present in the lower part. The sandstone was probably deposited in a fluviallacustrine environment and may possibly be deltaic in the upper part. The basal conglomerate is probably fluvial.

Meramangye Formation
Definition and nomenclature Originally introduced as Meramangye beds (Stainton et al., 1988) and Giles Mudstone (Sukanta, 1993), the unit was defined as Meramangye Formation by Zang (1995a). Type section The type section was defined by Zang (1995a) and is modified here to 8691064 m in Giles 1 (Fig. 6.7). Reference sections were also defined by Zang (1995b), and for Meramangye 1 is here modified to 451690 m. The section in Lake Maurice East 1, assigned to the Meramangye Formation by Zang (1995b), is now considered more likely to be Tarlina Sandstone. Lithology and distribution The lithology consists of red-brown and minor green-grey silty mudstone interbedded with thin siltstone or fine-grained sandstone, with minor silty limestone. The formation is present on the Murnaroo Platform and is interpreted seismically in the Ungoolya Hinge Zone and Munyarai Trough. A lateral equivalent of this unit may be present in the Birksgate sub-basin.

Fig. 6.7 Giles 1 type sections for Tarlina Sandstone and Meramangye Formation, and reference section for Murnaroo Formation.

Munyarai Trough. Tarlina Sandstone was intersected on the Nawa Ridge (Sukanta, 1993) and may occur in the Tallaringa Trough and on the Nullarbor Platform. It is probably lithologically transitional upwards and laterally with the Meramangye Formation. Relationships and boundary criteria The Tarlina Sandstone unconformably overlies the Alinya Formation, Wantapella Volcanics, Chambers Bluff Tillite or metamorphic basement, and is conformably overlain by the
#%
Prodelta silty mudstone and siltstone of the Meramangye Formation at 979.2983.15 m in Giles 1. Scale bar is 10 mm for each black and white grid. (Photo 42405a)

and dark red-brown by iron oxide in groundwater. The sandstone is generally poorly sorted with occasional very well rounded, in part bimodal, fine to coarse-grained and minor granule quartz sand with rare conglomerate; lithic grains and clay clasts are variably rare to common. Feldspar and heavy minerals are present, the latter often as concentrations on bedding planes; biotite and muscovite are similarly concentrated. Shale interbeds are more common near the top but are scattered randomly throughout the section. Cement most commonly is silica, with carbonate to a lesser extent. Accessory minerals include feldspar, rare gypsum, and possibly glauconite.
Slumped prodelta siltstone of the Meramangye Formation at 873.65873.8 m in Giles 1. Scale bar is 10 mm for each black and white grid. (Photo 42405b)

Relationships and boundary criteria The Murnaroo Formation conformably to unconformably overlies Meramangye Formation, Tarlina Sandstone or Alinya Formation. The upper boundary is conformably overlain by the Dey Dey Mudstone or unconformably by the Relief Sandstone (Manya 5). The unit is equivalent to part of the Wright Hill beds of Major (1973b). Thickness Variable, from 4 m in Marla 9 to 391 m in Lake Maurice East 1.

Relationships and boundary criteria Contacts with the underlying Tarlina Sandstone and overlying Murnaroo Formation are conformable at the type section. The Meramangye Formation unconformably overlies Wantapella Volcanics in the northern outcrop areas. It is equivalent to part of the Wright Hill beds of Major (1973b). Thickness The formation is 195 m thick in Giles 1 but is over 239 m thick in the Munyarai Trough (Meramangye 1). Age Precambrian, Neoproterozoic, early Marinoan Epoch, acritarch assemblage AAM 1. Sedimentology and palaeoenvironment Sedimentary structures include thin-bedded to laminated mudstone, siltstone with planar cross-beds, ripple cross-lamination and graded beds. Slump structures are common. Upward-shallowing parasequences from silty mudstone to siltstone indicate a passage from prodelta to delta front settings.

Murnaroo Formation
Definition and nomenclature The Murnaroo Formation was defined by Gatehouse et al. (1986). Type section The type section is 316.8627.5 m in Murnaroo 1 (section not fully penetrated). A reference section was defined as Lake Maurice East 1, here modified to 117.3508.3 m, which is a complete section. However, the Murnaroo in this well overlies Tarlina Sandstone, and thus the lower boundary cannot be definitively identified. For this reason, a new reference section is here defined as 583869 m in Giles 1 (Fig. 6.7). Lithology and distribution The lithology consists predominantly of sandstone, pale grey-green to pale green when fresh, altered to dark brown
#&
Estuarine fluvial shoreface sandstone of the Murnaroo Formation (lower part) at 736.3736.45 m in Giles 1. Scale bar is 10 mm for each black and white grid. (Photo 42406a)

submarine canyon (IVF), which probably represents initial movement of the Petermann Ranges Orogeny (~575 Ma). The upper part of the Ungoolya Group (Narana Formation and equivalents) was deposited in very shallow marine to shoreline environments, TST to HST. Moussavi-Harami and Gravestock (1995) estimated that >4000 m of Ungoolya Group sediments were originally deposited in the Munyarai Trough area prior to the Petermann Ranges Orogeny, but were folded at ~560 Ma and more than 2000 m eroded.

Dey Dey Mudstone


Definition and nomenclature
Fluvial cross-bedded sandstone of the Murnaroo Formation (upper part) at 615.8615.9 m in Giles 1. (Photo 42406b)

The formation was informally introduced as Dey Dey Mudstone by Sukanta (1993) and formally defined by Zang (1995b), who subdivided it into upper and lower units. The Dey Dey has previously been termed lower Ungoolya Formation (cf. Dunster, 1987a), and lower Rodda beds (cf. Brewer et al., 1987; Thomas, 1990). Type section The type section of Zang (1995b) in Lake Maurice West 1 is here modified to 316.3479.7 m. The reference section for Munta 1 is also modified to 16751973 m (Fig. 6.9), and Murnaroo 1 is unchanged at 208.9316.8 m. Lithology and distribution Two units are recognised, separated by a bed of dolomitic intraclasts which may be correlated seismically with a bed of limestone in the Munyarai Trough. The lower unit contains mainly red-brown with some green-grey silty mudstone; the upper unit comprises dolomitic or calcareous siltstone and mudstone. The basal Dey Dey Mudstone is locally conglomeratic. The formation is widely distributed over the Murnaroo Platform and in the Munyarai Trough, and probably occurs in the Birksgate Sub-basin. Relationships and boundary criteria The Dey Dey Mudstone conformably or unconformably overlies sandstone of the Murnaroo Formation and is overlain transitionally by the Karlaya Limestone. The boundary between the lower and upper units is marked by a bed of dolomitic ribbon intraclasts on the Murnaroo Platform. The Dey Dey is correlated with unit 4 of the Wright Hill beds in the central Officer Basin, Bunyeroo Formation in the Adelaide Fold Belt, and Yarloo Shale on the Stuart Shelf. The upper unit correlates with unit 5 of the Wright Hill beds and lower part of the Wonoka Formation in the Adelaide Fold Belt. The Dey Dey Mudstone is considered to correlate with the Pertatataka Formation in the Amadeus Basin. Thickness From 86 m in Giles 1 to 298 m in Munta 1. Generally the Dey Dey Mudstone thickens towards the Munyarai Trough where it may be up to 900 m thick. Age Precambrian, Neoproterozoic, mid-Marinoan Epoch, acritarch assemblage AAM 2 to 3.
#'

Ripple cross-bedded sandstone with mudstone rip up clasts of the Murnaroo Formation (upper part) at 615.7615.8 m in Giles 1. Scale bar is 10 mm for each black and white grid. (Photo 42406c)

Age Precambrian, Neoproterozoic, early Marinoan Epoch. Sedimentology and palaeoenvironment Sukanta (1993) recognised local aeolian palaeoenvironments in the hinterland to the southwest, with distal fan, sheet outwash, channelised fluvial and tidal deposits progressing northwards.

UNGOOLYA GROUP
The name Ungoolya Group was introduced as Ungoolya Formation by Dunster (1987a), and changed informally to Ungoolya Group by Sukanta (1993) before being formally defined by Zang (1995b). The group has also previously been termed Rodda beds (Brewer et al., 1987) and in part the Wright Hill beds of Major (1973b). Sequence E of Moussavi-Harami and Gravestock (1995) is an informal equivalent. Ungoolya Group has been intersected in 10 wells across the Murnaroo Platform (Fig. 6.8), Ungoolya Hinge and Munyarai Trough, and four higher order sequences can be recognised. The lower two (Dey Dey Mudstone and Tanana Formation) were deposited in deep to shallow marine environments during TST and HST (Moussavi-Harami and Gravestock, 1995). These two sequences were cut by a

Acraman impact debris in the Dey Dey Mudstone at 440.73 440.78 m in Lake Maurice West 1. (Photo 42408a)

Observatory Hill 1 (355.0 m) and Lake Maurice West 1 (440.78 m) contain a debris layer correlated with the Acraman meteorite impact ejecta, which also occurs in the Adelaide Fold Belt and Stuart Shelf. Sedimentology and palaeoenvironment The basal conglomerate was probably deposited in a fluvial environment. The red-brown silty mudstone in the lower unit is massive to laminated and occasionally cross-bedded, grading up to rhythmite or tempestite, which is considered to have been deposited in delta front to prodelta environments. The upper unit contains massive to laminated dolomitic siltstone and mudstone with intraclasts, exhibiting marine influence, and was probably deposited in prodelta to shelf settings with reducing sediment supply (Zang, 1995b).

Karlaya Limestone
Definition and nomenclature
Dey Dey Mudstone (lower part) at 527.7527.8 m in Giles 1. Scale bar is 10 mm for each black and white grid. (Photo 42407a)

The unit was informally introduced as Karlaya Limestone by Sukanta (1993) and formally defined as a member of the Tanana Formation by Zang (1995b). As the unit is widespread and mappable seismically, it is here raised to formation status. Type section The Karlaya 1 type section is here modified to 2024 2090 m (Fig. 6.10). The reference section in Lake Maurice West 1 is here modified to 237.6316.3 m. The reference section in Munta 1 is now considered to be Tanana Formation. Lithology and distribution The Karlaya Limestone is predominantly micritic limestone with thin silty mudstone interbeds. It is distributed over the Murnaroo Platform and Munyarai Trough, and is a distinct seismic marker in the Officer Basin. The formation has been recognised in Birksgate 1.
$

Dey Dey Mudstone at 505.9506 m in Giles 1; note dolomite intraclasts. Scale bar is 10 mm for each black and white grid. (Photo
42407b)

Fig. 6.8 Neoproterozoic cross-section: Lake Maurice West 1, Murnaroo 1, Giles 1, Munta 1, Munyarai 1, Marla 9.

Relationships and boundary criteria The Karlaya Limestone conformably overlies the dolomitic and calcareous Dey Dey Mudstone, and is conformably overlain by the Tanana Formation. It unconformably overlies the Wilari Dolomite on the Murnaroo Platform. Thickness From 13 m in Giles 1 to 64 m in Munyarai 1.
$

Age Precambrian, Neoproterozoic, mid-Marinoan Epoch, acritarch assemblage AAM 2 to 3. Sedimentology and palaeoenvironment Mainly horizontal thin-bedded micritic limestone, with dark grey mudstone layers; some limestone intraclasts are occasionally present, indicating a subtidal shelf environment, probably below fair weather wave base (Zang, 1995b).

Dey Dey Mudstone (lower part) at 323.5323.6 m in Lake Maurice West 1; note graded beds (probable slope deposits). (Photo 42408b)

Dey Dey Mudstone (upper part) at 319.9320 m in Lake Maurice West 1; note dolomitic breccia. Scale bar is 10 mm for each black and white grid. (Photo 42408c)

Fig. 6.9 Munta 1 Dey Dey Mudstone reference section.

Tanana Formation
Definition and nomenclature The unit was formally defined by Zang (1995b), but was previously termed Upper Ungoolya Formation (cf. Dunster, 1987a), and upper Rodda beds (cf. Brewer et al., 1987; Stainton et al., 1988). Type section The type section in Munta 1 is here modified to 1152 1675 m (Fig. 6.11). Reference sections are also modified to 16952024 m in Karlaya 1, 21552581 m in Munyarai 1, and 421484 m in Giles 1. The reference section in Lake Maurice West 1 is now considered to be Karlaya Limestone.
$

Karlaya Limestone at 1662.11662.3 and 16511651.15 m in Munta 1. Scale bar is 10 mm for each black and white grid. (Photo
42411b)

Lithology and distribution The lithology consists of limestone, calcareous siltstone and minor sandstone. The formation is distributed over the Murnaroo Platform and Munyarai Trough, and may occur in the Birksgate Sub-basin. Relationships and boundary criteria The basal contact is conformable. Where the upper boundary is eroded by canyon deposits of the Narana Formation, the boundary is unconformable, but the overlying

Fig. 6.10 Karlaya 1 Karlaya Limestone type section.

Munyarai Formation is conformable. The formation is correlated with unit 6 of the Wright Hill beds in the western Officer Basin, and middle and upper members of the Wonoka Formation in the Adelaide Fold Belt. The shale unit in the middle part of the formation was informally named Leemurra Mudstone by Sukanta (1993), but is not recognised regionally.

Fig. 6.11 Munta 1 Tanana Formation type section.

Thickness From 118 m in Marla 9 to 665 m in Ungoolya 1. Age Precambrian, Neoproterozoic, mid-Marinoan Epoch, acritarch assemblage AAM 3.
Tanana Formation at 1196.81196.95 and 1233.051233.20 m in Munta 1; note graded bedding and erosional structure (probable slope deposits). Scale bar is 10 mm for each black and white grid.
(Photo 42411c)

Sedimentology and palaeoenvironment The formation contains a limestone unit at the top which is fractured and contains intraclasts, and may have been
$!

formed in intertidal to upper subtidal shelf settings. The siltstone or silty mudstone is massive to laminated, occasionally cross-bedded or slump cross-bedded, and was probably deposited in prodelta, distal delta front to shelf settings. The upper part of the formation (= channel units, cf. Sukanta, 1993) contains red-brown to green-grey siltstone interbedded with thin sandstone; cross-bedding and erosional surfaces are common. The formation was probably deposited in subtidal, intertidal shelf, shoreface to fluvial settings. Sequentially the sedimentary features in the Tanana Formation indicate a shallowing-upward depositional environment (Zang, 1995b). Zang interpreted the section in Ungoolya 1 as a succession of mainly siliciclastics with slope ribbon limestone, conglomerate and turbidites, indicating a submarine fan deposit. Wilari Dolomite Member Definition and nomenclature The unit was informally introduced as Wilari Dolomite by Sukanta (1993) and formally defined as a member of the Tanana Formation by Zang (1995b). Type section The type section was defined by Zang (1995b) as 266.3289.3 m in Lake Maurice West 1. This section is now considered to be a reference section for the Karlaya Limestone, and the 166.6182 m reference section in Murnaroo 1 is here defined as a new type section. A reference section is defined as 178.26239.7 m in Observatory Hill 1 (Fig. 6.12). Lithology and distribution The dolomite is vuggy and weathered, massive to occasionally thin bedded, with thin siltstone layers. The dolomite is probably secondary. The member is intersected on the southeastern Murnaroo Platform but is unknown elsewhere in the basin. Relationships and boundary criteria Both the upper and lower boundaries are erosional. The member unconformably overlies the Karlaya Limestone and is unconformably overlain by the Early Cambrian Relief Sandstone or Cadney Park Formation. Thickness Thickness ranges from 14.3 m in Murnaroo 1 to 61.6 m in Observatory Hill 1. Age Precambrian, Neoproterozoic, mid-Marinoan. Relationships and boundary criteria Sedimentology and palaeoenvironment The Wilari Dolomite Member was probably deposited on an intertidal to subtidal shelf (Zang, 1995b).
$"

Fig. 6.12 Observatory Hill 1 Wilari Dolomite Member (Tanana Formation) reference section.

Munyarai Formation
Definition and nomenclature The unit was defined by Zang (1995b). Type section The type section is 17072155 m Munyarai 1 (Fig. 6.13). The reference section defined in Karlaya 1 is now considered to be Narana Formation. Lithology and distribution The lithology consists of grey to dark grey calcareous siltstone with thin limestone interbeds in the lower part. The formation is distributed over the Munyarai Trough.

The Munyarai Formation conformably overlies the Tanana Formation and is unconformably overlain by canyon deposits of the Narana Formation. The formation is

Narana Formation
Definition and nomenclature The unit was defined by Zang (1995b). Type section The type section in Munyarai 1 is here modified to 1699 1707 m (Fig. 6.13). Reference sections are also modified to 1281.71526 m in Ungoolya 1, 13671695 m in Karlaya 1, and 9001152 m in Munta 1. Lithology and distribution The Narana comprises conglomerate (clasts of chaotic to imbricated mudstone, sandstone, and micritic and oolitic limestone), sandstone, dark grey silty mudstone and silty limestone in varying proportions in different areas. The formation is distributed over the Munyarai Trough and in channels incised into the Murnaroo Platform. Relationships and boundary criteria The Narana Formation unconformably overlies the Munyarai Formation or Tanana Formation with an erosive boundary. It is unconformably overlain by Marla Group sediments. The Narana may correlate with the Pound Subgroup in the Adelaide Fold Belt. The equivalent unit in the Amadeus Basin is the lower Arumbera Sandstone.

Fig. 6.13 Munyarai 1 Munyarai Formation and Narana Formation type sections.

correlated with unit 7 of the Wright Hill beds in the western Officer Basin, and Bonney Sandstone in the Adelaide Fold Belt. Thickness The thickness is 448.1 m in Munyarai 1. Age Precambrian, Neoproterozoic, mid- to late Marinoan Epoch, acritarch assemblage AAM 4. Sedimentology and palaeoenvironment The formation is massive to laminated with graded bedding, and is considered to have been deposited in prodelta to shelf environments (Zang, 1995b).
$#
Breccia and sandstone (submarine fan deposits) of the Narana Formation at 1297.41299.5 m in Ungoolya 1. Scale bar is 10 mm for each black and white grid. (Photo 42409b)

Unconformable contact between Tanana and Narana Formations at 1152.5 m in Munta 1. (Photo 42412b)

Narana Formation calcareous siltstone with slump structures at 987.35989.6 m in Munta 1. Scale bar is 10 mm for each black and white grid. (Photo 42412d)

Thickness From 8 m in Munyarai 1 to 328 m in Karlaya 1. Seismic data indicate thicknesses of up to 1500 m in the northeastern Munyarai Trough (seismic line IP1-8). Age Precambrian, Neoproterozoic, late Marinoan Epoch, acritarch assemblage AAM 4. Sedimentology and palaeoenvironment The formation contains two canyon-fill sequences. The lower consists of debris flow deposits grading to tidal flat deposits at the top. The upper turbidite sequence was deposited on a transgressive to highstand deeper water submarine fan, shallowing to marine shelf limestone and mudstone, which become dominant towards the top (Zang, 1995b). Munta Limestone Member Definition and nomenclature The unit was defined informally as Munta Limestone by Sukanta (1993); it is here defined as a member of the Narana Formation.
$$

Type section Here defined as 14201447 m in Karlaya 1 (Fig. 6.14). Reference sections are defined as 16501670 m in Lairu 1, and 14951526 m in Ungoolya 1. Lithology and distribution The member consists of grey to greenish grey, massive and laminated micritic limestone and marl. Limestone and marl units are rhythmically interbedded, with sharp, lower boundaries to the limestone, grading upwards into marl (Sukanta, 1993). The member is present in the Munyarai Trough. Relationships and boundary criteria The Munta Limestone Member is conformably underlain by the canyon fill sequences, and overlain either conformably by the Mena Mudstone Member, or unconformably by Marla Group sediments. Thickness The thickness varies from 20 m in Lairu 1 to 45 m in Munta 1.

Age Precambrian, Neoproterozoic, late Marinoan Epoch. Sedimentology and palaeoenvironment Sukanta (1993) reviewed the possible palaeoenvironments represented by the Munta Limestone Member. Generally, low energy subtidal marine shelf environments are represented, but deeper environments, perhaps upper slope, may be represented in the Lairu and Ungoolya area. Mena Mudstone Member Definition and nomenclature The unit was defined informally as Mena Mudstone by Sukanta (1993); it is here defined as a member of the Narana Formation. Type section Here defined as 13671420 m in Karlaya 1 (Fig. 6.14). Reference sections are defined as 14091650 m in Lairu 1, and 12821495 m in Ungoolya 1. Lithology and distribution The member consists of grey-green mudstone with thin parallel laminae and contains interbeds of grey-green medium to coarse-grained sandstone. It is similar lithologically to parts of the Cambrian Ouldburra Formation, but the latter may contain burrows (Sukanta, 1993). The member occurs in the Munyarai Trough. Relationships and boundary criteria The Mena Mudstone Member conformably overlies the Munta Limestone Member and is unconformably overlain by Marla Group sediments. The uppermost part of the member may pass laterally westwards into the Punkerri Sandstone. Thickness The thickness varies from 53 m in Karlaya 1 to 241 m in Lairu 1. Age Precambrian, Neoproterozoic, late Marinoan Epoch. Sedimentology and palaeoenvironment The mudstone is finely laminated, but the sandstone interbeds are poorly sorted with sharp erosive bases and may display graded bedding, ripple and rare trough cross-bedding. Rare mudclasts are present at the erosive base of the sandstones. The mudstone is presumed to have formed in a low energy subtidal shelf environment. Poorly sorted and coarse-grained sandstone is better developed up section, and indicates shallowing (eustatic sea-level regression) and, finally at the level of the unconformity with the overlying Marla Group, the red oxidised mudstone may indicate subaerial exposure and erosion (Sukanta, 1993).
$%
Fig. 6.14 Karlaya 1 Munta Limestone Member and Mena Mudstone Member (Narana Formation) type sections.

Punkerri Sandstone
Definition and nomenclature The unit, first mentioned as Punkerri Sandstone by Thomson (1969), was defined as Punkerri Beds by Major (1974), and is here redefined as Punkerri Sandstone. Type section The type section (Fig. 6.15) was defined by Major (1974) in outcrop, and is located across the southern flank of the Punkerri Hills at 2740 latitude and 13025 longitude. Lithology and distribution The formation was informally subdivided into two units by Major (1974). The lower is generally a purple or red-brown, medium-grained flaggy quartzose sandstone with some feldspar and biotite. The upper unit consists of red and white, medium-grained feldspathic sandstone and some quartzose sandstone with interbedded red sandstone and siltstone (Major, 1974). Outcrops of the Punkerri are restricted to the Birksgate Sub-basin, but it is presumed to occur at depth in the western part of the Officer Basin.

are presumed to have come from the upper Punkerri Sandstone. Sedimentology and palaeoenvironment Sedimentary structures in the lower Punkerri Sandstone include scour casts, ripple marks, pellets and flakes of siltstone. The upper unit shows some cross-bedding, and clasts of siltstone and quartz pebbles. The environment of deposition is tidally influenced shallow marine, similar to the Pound Subgroup of the Adelaide Fold Belt (Preiss, 1987).

MARLA GROUP
The term Marla Sequence was introduced by Krieg (1973) and defined as Marla Group by Benbow (1982). The lower part of the group, comprising aeolian and fluvial Relief Sandstone and evaporitic Ouldburra Formation, contains three third order sequences (Gravestock and Hibburt, 1991; Moussavi-Harami and Gravestock, 1995). Major regression took place at the end of ReliefOuldburra time, and the succeeding units (Wallatinna and Cadney Park Members of the Observatory Hill Formation) were formed in an alluvial fan to alluvial plain environment. The remainder of the Observatory Hill Formation was also a non-marine hinterland tract in playa lakes. The fluvial to shoreline Arcoeillinna Sandstone followed, then a major transgression began in the east and deposited the shallow marine Apamurra Formation. The Trainor Hill Sandstone was deposited in a fluvial environment in the north and sandy tidal flats in the south. Total thickness of the Marla Group was probably originally >2600 m in the Manya Trough, but only 1000 m on the Murnaroo Platform, indicating a depocentre to the northeast and major erosion (12 km) in the central part of the basin as a result of the Petermann Ranges Orogeny (Fig. 6.16). The Moorilyanna Formation (Coats, 1963; Gravestock et al., 1995) and Levenger Formation (Major, 1973d), both of which occur in isolated grabens in the Musgrave Block, may correlate with either the upper or lower Marla Group, although there is a possibility that they may be older proximal equivalents to the Pindyin Sandstone, or the earlier undrilled graben-fill sequence which may be equivalent to the Pandurra Formation of the Gawler Craton and Stuart Shelf.
Fig. 6.15 Punkerri Sandstone section (after Major, 1974).

Relief Sandstone
Definition and nomenclature The formation was defined by Brewer et al. (1987). Type section The type section in Meramangye 1 is here redefined as 359451 m (Fig. 6.17). Reference sections are 304421 m in Giles 1, and 155.2178.3 m in Observatory Hill 1. Lithology and distribution

Relationships and boundary criteria The upper and lower boundaries are not seen in outcrop, but both the base and top are presumed to be unconformable. The formation is equivalent to the Pound Subgroup in the Adelaide Fold Belt (Preiss, 1993). Thickness The formation is over 1200 m thick in the type section. Age Precambrian, Neoproterozoic, latest Marinoan Epoch. An Ediacara type fauna, with at least five genera, is known from loose rocks in the northern Punkerri Hills, which
$&

Gaughan and Warren (1990) subdivided the formation into nine stratigraphic units based on lithology and interpreted palaeoenvironment (Fig. 6.18). In general, the formation is characterised by well sorted, fine to medium-grained or poorly sorted fine to coarse-grained

Fig. 6.16 Marla Group cross-section, through Wilkinson 1, Murnaroo 1, Giles 1, Munta 1, Munyarai 1 and Marla 3.

slightly feldspathic sandstone with clay coated grains (illite, variably corroded). Quartz overgrowths and carbonate cement are found mainly in the Marla area. Relationships and boundary criteria The formation disconformably overlies either Ungoolya Group formations or earlier Murnaroo or Meramangye Formations. The unit is synchronous and intercalates with the more marine Ouldburra Formation, and is conformably or disconformably overlain by the Observatory Hill Formation. In Manya 5, the top of the unit is eroded and is unconformably overlain by Mount Chandler Sandstone. There may be an equivalent to the Relief Sandstone below the Cootanoorina Formation (Townsend and Ludbrook, 1975) in the Boorthanna Trough (Weedina 1).

Age Early Cambrian, 540524 Ma, based on stratigraphic position and synchroneity with the Ouldburra Formation. The unit is unfossiliferous. Sedimentology and palaeoenvironment Gravestock and Sansome (1994) suggested that much of the Relief Sandstone could be recycled Punkerri Sandstone.

Thickness Thickness varies from 23.1 m in Observatory Hill 1 to possibly 168 m in Manya 5.
$'
Relief Sandstone at 1671.2 m in Manya 6; core diameter is 38 mm.
(Photo 44375)

Fig. 6.17 Meramangye 1 Relief Sandstone type section.

Gaughan and Warren (1990) determined the palaeoenvironments to range from aeolian and fluvial to tidal, which are related to marine lowstandhighstand cycles, and probably synchronous with sequences in the adjacent marine Ouldburra Formation (Gravestock and Hibburt, 1991). Aeolian lithofacies are characterised by fine to mediumgrained, well to very well sorted, subangular to subrounded, low to high angle cross-bedded, pale brown to red sandstone. Fluvial lithofacies are characterised by poorer sorting, locally coarser grain size, shale intraclasts, and planar and trough cross-bedding. Tidal lithofacies are laminated with thin, fine to coarse interbeds, ripple cross-lamination, and mud drapes.

Ouldburra Formation
Definition and nomenclature The formation was defined by Brewer et al. (1987).
%
Fig. 6.18 Relief Sandstone palaeoenvironments in Giles 1 (after Gaughan, 1989; Gaughan and Warren, 1990).

Type section The type section was originally defined as 571.4 1685.4 m in Manya 6. It is here redefined in Manya 6 as 5711558 m, as the section below 1558 m is predominantly sandstone and is thus more correctly attributed to the Relief Sandstone (Fig. 6.19). Reference sections are 493607 m in Marla 3, and 199 m (top eroded) to 685.5 m in Wilkinson 1. Lithology and distribution The formation comprises mainly interbedded muddy limestone, dolostone, sandstone and evaporites. In the type section, the basal part is typified by halite and minor sandstone grading up to stacked sandsiltmudstone sets which become increasingly calcareous. These are overlain by a thick sequence of calcareous and dolomitic carbonates with sporadic clastics and gypsumanhydrite interbeds. The carbonate lithofacies are dominated by laminated and silty carbonate mudstone. The top of the formation is typically an intercalation of laminated carbonate mudstone and redbed siltstone with abundant nodular sulphate evaporite and bedded chicken wire anhydrite. The unit is present in the Manya and Tallaringa Troughs. Relationships and boundary criteria The formation disconformably overlies Ungoolya Group formations or earlier Murnaroo or Meramangye Formations. It overlies and is intercalated with the non-marine Relief Sandstone, and is conformably overlain by the Observatory Hill Formation or Wallatinna Formation in Byilkaoora 1. The Ouldburra may be equivalent to part of the Cootanoorina Formation (Townsend and Ludbrook, 1975) in the Boorthanna Trough. Thickness The thickness ranges from 114 m in Marla 3 to 987 m in Manya 6 (the type section), but could be up to 1100 m where not eroded (Gravestock and Sansome, 1994). Age Early Cambrian, 536524 Ma, acritarch assemblage AAC 1.

Fig. 6.19 Manya 6 Ouldburra Formation type section.

%

Lithology and distribution The main part of the formation consists of multicoloured micaceous siltstone and claystone, calcareous and dolomitic in part, with minor light yellow-brown, very fine-grained sandstone and light grey to dark grey limestone and dolomite. The siltstone and claystone are dominantly red-brown to brown but range through purple and greenish grey to dark grey, especially where they are interbedded with the carbonates. The formation was deposited over a wide area of the eastern Officer Basin.
Evaporitic salina; interbedded halite and dolomitic siltstone of the Ouldburra Formation at 657.7 m in Wilkinson 1. (Photo 44376)

Relationships and boundary criteria The formation was deposited conformably above the Ouldburra Formation or Relief Sandstone, and disconformably above the Ungoolya Group in the Manya and Munyarai Troughs. The upper contact with the Arcoeillinna Sandstone is conformable. The formation may be correlated with part of the Wirrildar beds of the central Officer Basin (Major, 1973a), and is equivalent to the Billy Creek Formation in the Arrowie Basin. Thickness
Ouldburra Formation feldspathic sandstone with dolomitic intraclasts at 664.1 m in Marla 6. The vuggy porosity is due to the solution of dolomite. (Photo 44377)

The thickness varies from 155.2 m in Observatory Hill 1 to 466 m in Marla 3.

The middle part of the formation in the type section contains an undescribed species of trilobite (Abadiella sp.), suggesting an Atdabanian age (Jago et al., 1994), but archaeocyaths suggest a Botomian age (Gravestock and Sansome, 1994). Acritarch assemblage AAC 1 is present. Sedimentology and palaeoenvironment The lower part of the Ouldburra Formation is a cyclic suite deposited in isolated salinas on a shallow marine to subaerially exposed sandy mudflat which extended southwestwards into the Tallaringa Trough (Dunster, 1987a). The main part of the formation is more marine, and records sawtooth epeiric sea transgressiveregressive sequences. Trilobites and archaeocyaths are found in this part of the formation. The upper part of the formation was deposited in a regressive redbed and carbonate sabkha.

Observatory Hill Formation type section.

(Photo 43062)

Observatory Hill Formation


Definition and nomenclature The name was introduced by Wopfner (1969) as Observatory Hill Beds, and designated a formation by Brewer et al. (1987). Type section Outcrop at latitude 2858.2S, longitude 13157.7E, and the subsurface section at 0155.2 m in Observatory Hill 1. This well was spudded just above the base of Wopfners outcrop type section. Reference sections are defined at 30.5335 m in Emu 1, and 8304 m in Giles 1 (Fig. 6.20).
%

Cherty dolomite of the Observatory Hill Formation; upper type section. (Photo 43061)

Contact of Observatory Hill Formation and overlying Arcoeillinna Sandstone, Mount Johns Range. (Photo 30890)

Cadney Park Member Definition and nomenclature The unit has been in informal use for many years as the Cadney Park Formation, and is defined as a member of the Observatory Hill Formation herein. Type section The type section is here defined as 116493 m in Marla 3. Reference sections are defined as 236304 m in Giles 1 (Fig. 6.20), and 254359 m in Meramangye 1. Lithology and distribution The member comprises red, sandy to calcareous siltstone with occasional conglomerate. It is widespread, and is generally found in most wells which fully penetrate the Observatory Hill Formation. In Manya 2, the Cadney Park Member overlies the Wallatinna Member, and is not easily distinguished from it. Relationships and boundary criteria The member was deposited conformably above the Ouldburra Formation or Relief Sandstone, and disconformably above the Ungoolya Group in the Manya and Munyarai Troughs. It is conformable with the main part of the Observatory Hill Formation. Thickness The thickness varies from 54 m in Lairu 1 to 377 m in Marla 3. The base is seismically mappable (Stainton et al., 1988). Age Early Cambrian, 524522 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment The member was deposited in a distal alluvial fan to ephemeral lake environment (Gravestock and Sansome, 1994).
%!

Fig. 6.20 Giles 1 Observatory Hill Formation and Cadney Park Member reference sections.

Age Early Cambrian, 524520 Ma, based on stratigraphic position.

Sedimentology and palaeoenvironment The Observatory Hill Formation has been the subject of detailed studies and is generally non-marine, with palaeoenvironments ranging from distal alluvial fan, fluvial, and ephemeral lake, to alkaline playa lake. At least three lacustrine complexes are recorded (White and Youngs, 1980; Southgate and Henry, 1984; Southgate et al., 1989; Brewer et al., 1987; Gravestock et al., 1995).

Wallatinna Member Definition and nomenclature The unit was defined by Benbow (1982) as the Wallatinna Formation. As it is restricted in distribution, and is not easily distinguished from conglomeratic facies of the Cadney Park Member, it is here redefined as a member of the Observatory Hill Formation. Type section The type section is defined from outcrop on the northern margin of the Mount Johns Range. A subsurface reference section is defined as 378486 m in Byilkaoora 1 (Fig. 6.21).

Lithology and distribution The unit comprises flat-bedded, coarse-grained to granule arkose with interbedded conglomerate and siltstone. The arkose is moderately to well sorted, with angular to sub-rounded grains of quartz, feldspar and minor biotite. Calcite cement is locally abundant. The conglomerate may contain well rounded clasts up to 0.3 m diameter of white granitoid and more rarely black chert and grey limestone. The unit is restricted to the northeastern part of the basin in the vicinity of the Mount Johns Range. Relationships and boundary criteria The basal contact is apparently conformable, but outcrop evidence suggests that it may be unconformable (Benbow,

Alluvial fan conglomerate of the Observatory Hill Formation (Wallatinna Member) in Byilkaoora 1. (Photo 44378)

Fig. 6.21 Byilkaoora 1 reference sections for Wallatinna Member, Parakeelya Alkali Member, Moyles Chert Marker Bed and Oolarinna Member (Observatory Hill Formation).

Observatory Hill Formation (Wallatinna Member), northwestern side of the Mount Johns Range. (Photo 30887)

%"

1982). The upper contact is conformable and is generally overlain by the Oolarinna Member. Thickness The thickness ranges from 50 m in Byilkaoora 2 to 260 m in outcrop, and 298 m in Byilkaoora 3. Age Early Cambrian, 524522 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment The unit was deposited in a proximal alluvial fan palaeoenvironment (Gravestock and Sansome, 1994; Benbow, 1982). Parakeelya Alkali Member Definition and nomenclature The unit was defined as a member of the Observatory Hill Formation by Brewer et al. (1987). Type section The type section is 240334.5 m in Byilkaoora 1 (Fig. 6.21). Lithology and distribution The unit comprises a sequence of laminated to thinly bedded, dominantly light brownish grey and greenish grey to dark grey limestone, dolomite, siltstone and claystone. Unique characteristics include the presence of abundant desiccation features, chert lenses, nodules and fragments, and calcite and dolomite crusts and pseudomorphs after evaporitic minerals (trona). The member is widely distributed in the Manya and Munyarai Troughs, and on the Murnaroo Platform. Relationships and boundary criteria The upper and lower boundaries are conformable within the Observatory Hill Formation. Thickness The thickness ranges from 62 m in Observatory Hill 1 to 138 m in Byilkaoora 1. Age Early Cambrian, 522521 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment The member was deposited in an alkaline playa lake environment. Detailed studies of the palaeoenvironments represented by this unit were published by Stainton et al. (1988), and Southgate et al. (1989).
%#
Laminated chert and carbonate of the Observatory Hill Formation (Parakeelya Alkali Member) in Byilkaoora 1; salina and mudflat facies. (Photo 44379)

Moyles Chert Marker Bed Definition and nomenclature The member was defined by Brewer et al. (1987) as the Moyles Chert Marker Bed. Type section The reference section of Brewer et al. (1987) is here assumed to be the type section (200213.1 m in Byilkaoora 1; Fig. 6.21). Lithology and distribution The member consists of greenish grey to dark grey limestone, dolomite and siltstone containing abundant chert nodules and lenses. The Moyles Chert Marker Bed is a distinctive seismic reflector and is widespread over much of the eastern part of the basin. Relationships and boundary criteria The upper and lower boundaries are conformable within the Observatory Hill Formation. Thickness The thickness is 13 m in Byilkaoora 1, but ranges up to 36 m in Byilkaoora 2.

Relationships and boundary criteria The lower boundary is conformable within the Observatory Hill Formation; the upper boundary with the Arcoeillinna Sandstone may be partly unconformable to the north but is also conformable to the south. Thickness The thickness ranges up to 44 m in Byilkaoora 1. Age Early Cambrian, 521520 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment The Oolarinna Member indicates more oxidising conditions relative to the Parakeelya Alkali Member, and was probably deposited in a fluvial environment.

Arcoeillinna Sandstone
Definition and nomenclature The sandstone was defined by Benbow (1982). Type section The type section is defined in outcrop on the eastern flank of the Mount Johns Range. Subsurface reference sections are defined as 10371191 m in Karlaya 1 (Fig. 6.22), and 98.5156 m in Byilkaoora 1. Lithology and distribution Age Early Cambrian, 521 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment The marker bed was deposited in an alkaline playa lake environment. Oolarinna Member Definition and nomenclature The unit was defined as a member of the Observatory Hill Formation by Benbow (1982). Type section The type section is defined as outcrop on the northern margin of the Mount Johns Range. A subsurface reference section is defined as 156200 m in Byilkaoora 1 (Fig. 6.21). Lithology and distribution The unit consists of red-brown, thinly bedded micaceous siltstone and claystone, with interbedded pebble conglomerate and sandstone. The unit is presumed to be present over most of the basin, except where removed by erosion.
%$

Calcite pseudomorphs after trona in the Observatory Hill Formation (Parakeelya Alkali Member) at 320 m in Byilkaoora 1. (Photo
42435)

The lithology is predominantly red-brown, very fine to medium-grained feldspathic sandstone. Minor interbeds consist of claystone and siltstone similar to the Oolarinna Member of the Observatory Hill Formation. Some pebbly horizons occur, composed of rounded, elongate green, red-brown and white claystone (Benbow, 1982). The formation is widely distributed in the Officer Basin. Relationships and boundary criteria The Arcoeillinna conformably overlies the Observatory Hill Formation (usually the Oolarinna Member), and is conformably overlain by Apamurra Formation or unconformably overlain by Mount Johns Conglomerate. It may be equivalent to part of the Wirrildar beds in the central Officer Basin (Major, 1973a). The interpreted equivalent unit in the Arrowie Basin is the upper part of the Billy Creek Formation (Gravestock and Hibburt, 1991). Thickness The thickness ranges from 46 m in Byilkaoora 2 to 172 m in Munta 1. Age Early Cambrian, 520518 Ma, based on stratigraphic position.

Arcoeillinna Sandstone at 137.7 m in Byilkaoora 1.

(Photo 44380)

Fig. 6.22 Karlaya 1 upper Marla Group reference section (Arcoeillinna Sandstone, Apamurra Formation and Trainor Hill Sandstone).

Sedimentology and palaeoenvironment Many of the grains have a thin film of red iron oxide which was present before compaction of the sediment, and sedimentary structures include abundant cross-bedding. Soft-sediment slump features and scour and fill structures are present in the coarser sediments. Benbow (1982) interpreted a fluvial-lacustrine palaeoenvironment, with an aeolian influence. Current directions and increasing thickness suggest a west or southwest provenance.
Cross-bedding in Arcoeillinna Sandstone, Mount Johns Range.
(Photo 30891)

here raised to formation status (and the Mount Johns Conglomerate reduced to member status) to more appropriately reflect the more widespread and finer grained facies. Type section

Apamurra Formation
Definition and nomenclature The unit was originally defined as the Apamurra Member of the Mount Johns Conglomerate by Benbow (1982). It is
%%

The type section was defined from outcrop on the southwestern margin of the Mount Johns Range (Benbow, 1982). Subsurface reference sections are defined as 9591037 m in Karlaya 1 (Fig. 6.22), and 49.698.5 m in Byilkaoora 1.

Lithology and distribution The formation consists of two units, a basal conglomerate and sandstone (the Mount Johns Conglomerate Member), and a generally thicker upper unit consisting of calcareous and dolomitic, red-brown, fine to very fine-grained sandstone and siltstone. The finer sediments are bimodal, with fine to coarse-grained angular and subangular sand grains scattered throughout. Minor lithologies include thin, light green limestone, and light green siltstone and sandstone (Benbow, 1982). The unit is widespread over the Officer Basin and is seismically mappable (Stainton et al., 1988). Relationships and boundary criteria The upper and lower contacts are conformable, although the base of the Mount Johns Conglomerate Member is probably at least locally disconformable. The unit may be equivalent to part of the Wirrildar beds (Major, 1973a) in the central Officer Basin. Interpreted equivalent units in the Arrowie Basin are the Aroona Creek and Wirrealpa Limestones (Gravestock and Hibburt, 1991). Thickness The thickness ranges from 27 m in Munyarai 1 to 78 m in Karlaya 1. Age Early Cambrian, 518516 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment The silt matrix is heavily stained with iron oxide, but minor tangential and herringbone cross-bedding is present in coarser, well-sorted sandy interbeds (Benbow, 1982). Trace fossils (Rusophycus) which are generally attributed to trilobites are common. These indicate a shallow marine to tidal flat palaeoenvironment for the finer grained unit. The lower conglomeratic unit is interpreted to have been deposited in a proximal alluvial fan environment, with fluvial and floodplain environments also possibly being represented (Benbow, 1982). Mount Johns Conglomerate Member Definition and nomenclature The unit was introduced by Krieg (1973) and defined with a type area by Benbow (1982). It is here redefined as a member of the Apamurra Formation. Type section The type section is outcrop on the northern margin of the Mount Johns Range. A subsurface reference section is here defined as 504505.6 m in Byilkaoora 2. Lithology and distribution The member is characterised by poorly sorted, red-brown conglomerate and red-brown sandstone. The conglomerate contains well-rounded clasts up to 0.2 m in size, comprising diverse lithologies including arkose, quartzite, pegmatite,
%&

siltstone and vein quartz. The sandstone is only slightly feldspathic, and individual grains may be coated with iron oxide or clay. The unit is restricted to the northern part of the basin, and is absent to the south. Relationships and boundary criteria The lower boundary with the Arcoeillinna Sandstone may be locally disconformable (Benbow, 1982), but the upper

Apamurra Formation (Mount Johns Conglomerate Member) at 97.8 m in Byilkaoora 1; the black mineral in the vugs is probably haematite. (Photo 44381)

Southerly view of the east side of the Mount Johns Range. The low-lying area to the left is Observatory Hill Formation with a low ridge of Arcoeillinna Sandstone; the area immediately flanking the range is Apamurra Formation (Mount Johns Member) and the prominent ridge is Trainor Hill Sandstone. (Photo 30894)

contact with the finer grained facies of the Apamurra Formation is gradational and conformable. Gravestock et al. (1995) suggested that the Moorilyanna Formation (Coats, 1963) and Levenger Formation (Major, 1973d), which occur in isolated grabens in the Musgrave Block, may be correlatives of the Mount Johns Conglomerate. However, as noted earlier in this chapter, these graben filling units may be as old as Mesoproterozoic. Thickness The thickness ranges from 1.6 m in Byilkaoora 2 to 78 m in outcrop. Age Early Cambrian, 518516 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment Sedimentary structures in the sandstone include tabular cross-bedding which suggests a northwesterly provenance. Rapid thickening of the member to the north also suggests a northern provenance. A proximal alluvial fan palaeoenvironment (with fluvial and floodplain environments also probably represented) was interpreted by Benbow (1982).

Trainor Hill Sandstone at 114.7 m in Marla 10; note the black cryptomelane (KMn8O16) staining. (Photo 44382)

Trainor Hill Sandstone


Definition and nomenclature The unit was introduced by Krieg (1973), with the base immediately above the Observatory Hill Formation and the top within the Mount Chandler Sandstone. The formation was defined by Benbow (1982). Type section The type section was defined from outcrop on the southwestern margin of the Mount Johns Range. Subsurface reference sections are defined as 512959 m in Karlaya 1 (Fig. 6.22), and 245460 m in Byilkaoora 2. Lithology and distribution The lithology consists of a well-sorted, medium to fine-grained, white to light grey, kaolinitic and feldspathic sandstone, with minor interbeds of red-brown micaceous siltstone and claystone, and pebbly horizons. The sandstone becomes light red-brown, feldspathic, calcareous and dolomitic towards the top (Benbow, 1982). The unit is widely distributed in the Officer Basin. Relationships and boundary criteria The base of the Formation is sharp but there is little evidence for an unconformity. The upper boundary is erosional and unconformable, with up to 500600 m of erosion (Gravestock and Sansome, 1994). The unit may be equivalent to part of the Wirrildar beds (Major, 1973a) in the Birksgate Sub-basin. The interpreted equivalent in the Arrowie Basin is the Moodlatana Formation.
%'

Thickness The thickness ranges from 87 m in Munyarai 1 to 420 m in outcrop, and 520 m in Lairu 1. As the top is eroded, it may have originally been up to 1000 m thick in the Mount Johns area (Gravestock and Sansome, 1994). Age Middle Cambrian, 516510 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment Cross-bedding is common throughout the unit , and is trough-like near the base, but steeply tangential to tabular higher in the section. Some cross-beds have been overturned. The finer grained units display herringbone cross-beds and ripple marks. Trace fossils (Skolithos ichnofacies), clay galls and desiccation cracks are also present. Facies in the Trainor Hill Sandstone have been interpreted to include tidal channels and flats, upper shoreface (beach and bar) sands, and deltaic sands, which were deposited in a transgressive barrier bar system (Rudd, 1995). Current directions generally indicate a southwesterly source, but a northwesterly source is indicated in the north where facies are more typical of fluvial palaeoenvironments (Benbow, 1982).

LATE CAMBRIAN VOLCANICS


Kulyong Formation
Definition and nomenclature The unit was defined by Major and Teluk (1967) as the Kulyong Volcanics. Jackson and van de Graaff (1981) suggested that the Kulyong Volcanics and Table Hill

Volcanics (Compston, 1974) comprise a single tholeiitic basalt suite and the latter is therefore a synonym. The unit is herein changed to Kulyong Formation to include the related underlying sedimentary units. Type section Outcrop at latitude 273620S, longitude 12926E (Fig. 6.23). No subsurface reference section is designated as the unit has not been encountered in any well, but it is expected to occur at depth in the central and western Officer Basin. Lithology and distribution The lower 2 m of the volcanic unit is soft, grey and vesicular whereas the upper 1 m is hard, red-brown, finegrained rock with no recognisable layering. Petrographically it is a sub-ophitic mesh of plagioclase and augite, the brown colour being due to films of haematite. The rock is of basic igneous composition and is classified as a tholeiite. In places there occurs crystalline quartz, agate, pyrolusite and minor amounts of a dark green mineral, probably chlorophaeite, which resembles chrysocolla in hand specimen (Major and Teluk, 1967). The underlying sedimentary units comprise thinly bedded, silicified ferruginous sandstone, silicified greywacke (possibly tuffaceous) and fine-grained micaceous sandstone with a thin (50 mm) limestone interbed. The unit occurs in outcrop and at depth over >20 000 km2 in the western Officer Basin (Townson, 1985). Relationships and boundary criteria The unit in outcrop has no top, but from the relative height difference to the nearest outcrop of Mount Chandler Sandstone, and assuming a regional dip of 0.5, it is probable that the Mount Chandler Sandstone conformably overlies the

Kulyong Formation. The formation unconformably overlies the gently folded Trainor Hill Sandstone or earlier units. Thickness The volcanics are up to 3 m thick in outcrop, and the underlying sedimentary units are up to 8 m thick. The maximum drilled thickness of Table Hill Volcanics in the western Officer Basin is 118 m (Townson, 1985). Age Middle-Late Cambrian, 507505 Ma. KAr geochronology suggests an Early Ordovician minimum age of 48520 to 47520 Ma for the Kulyong Volcanics (Major and Teluk, 1967), and a poorly constrained RbSr age of ~570 Ma for the Table Hill Volcanics (Compston, 1974; Jackson and van de Graaff, 1981). Both of these determinations are questionable, and an age of no greater than late Middle Cambrian age is more likely based on stratigraphic position and the relative lack of folding compared to the underlying Trainor Hill Sandstone (Gravestock et al., 1995). Sedimentology and palaeoenvironment The vesicular nature of the volcanic unit suggests a subaerial flow, but no upper contact is seen. Possible flute casts are present in the ferruginised sandstone underlying the basalt. The thickness of the sandstone varies from 0.15 to 0.6 m. The greywacke is only weakly stratified but the basal sandstone is thin bedded. The limestone interbed has minor cross-bedding. The presumed palaeoenvironment is non-marine, to shallow marine at the base.

MUNDA GROUP
The Munda Sequence was defined by Krieg (1973) for the predominantly Ordovician sequence of formations, and is redefined here as the Munda Group. The group is bounded by two unconformities formed during the Delamerian Orogeny and Rodingan Event. The sequence of formations generally records a major transgression, from LST in the Mount Chandler Sandstone, to TST in the Indulkana Shale, which was formed during maximum transgression. The uppermost unit, the Blue Hills Sandstone, was formed in a HST. The maximum thickness of the Munda Group was probably more than 3000 m in the Marla area, increasing towards the Amadeus Basin (Moussavi-Harami and Gravestock, 1995). This would indicate that 15002000 m of erosion may have occurred.

Mount Chandler Sandstone


Definition and nomenclature The name Mount Chandler Sandstone was introduced by Coats (1963). The formation was redefined by Benbow (1982) who separated the Byilkaoora Formation, which is here redefined as a basal member of the Mount Chandler Sandstone. This sandstone should not be confused with the Chandler Formation of the Amadeus Basin, which is a carbonateevaporite unit.
&

Fig. 6.23 Kulyong Formation section (after Major and Teluk, 1967).

Type section The type section has not been defined previously, and is here defined as 40512 m in Karlaya 1. Reference sections are defined as 33245 m in Byilkaoora 2, and 11891355 m in Munyarai 1 (Fig. 6.24). Lithology and distribution The lower part of the formation comprises quartzose sandstone to friable, well-sorted, well-rounded, fine-grained white sandstone. In the upper part, the sandstone is slightly feldspathic, and is orange-brown to slightly reddish. Minor layers of rounded to subangular polished white quartz pebbles occur. The unit is widely distributed over the Officer Basin.

Relationships and boundary criteria The basal contact is unconformable, and the upper is either conformable with the Indulkana Shale or unconformable with Permian sediments of the Arckaringa Basin in the Manya Trough. The unit is probably equivalent to the Pacoota Sandstone of the Amadeus Basin. Thickness The thickness varies from 160 m in Manya 5 to 472 m in Karlaya 1, but may be up to 609 m in outcrop in the Cartu Hill area (Benbow, 1982). Age Late Ordovician, 457450 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment Thick-bedded tabular cross-bedding is common throughout the unit, and a tide-dominated fluvial deltaic palaeoenvironment is interpreted for the formation. Trace fossils are present in the unit and include Diplocraterion and Skolithos. These are characteristic of the Skolithos ichnofacies (Seilacher, 1967) which generally indicates a sandy, high-energy shore environment. Byilkaoora Member Definition and nomenclature Originally the basal part of the Mount Chandler Sandstone, the unit was defined as the Byilkaoora Formation by Benbow (1982). The unit is here redefined as a member of the Mount Chandler Sandstone as it is too thin to recognise in the subsurface. Type section The type section is defined as outcrop north of Mount Johns, and an outcrop reference section is defined on the eastern margin of Mount Byilkaoora. No subsurface sections can be defined, although it is likely to occur in Byilkaoora 2 and 3, which are near the type area.

Fig. 6.24 Munyarai 1 Munda Group reference section (Mount Chandler Sandstone, Indulkana Shale and Blue Hills Sandstone).

Skolithos, top of the Mount Chandler Sandstone, northwestern Mount Johns Range. (Photo 30902)

&

Lithology and distribution The unit is characterised in the type section by conglomerate with well-rounded clasts of grey feldspathic sandstone (?Trainor Hill Sandstone), with light blue, fine-grained quartzose sandstone, quartz, veined quartzite, and shale. A sandy facies is also present, comprising white kaolinitic, very fine-grained sandstone with minor siltstone and claystone. Relationships and boundary criteria The basal contact is sharp and unconformable; the upper is gradational and difficult to clearly identify. Thickness The unit is 1520 m thick in outcrop in the type area. Age Late Ordovician, 457455 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment Tangential cross-bedding is present in the sandstone interbeds. A predominantly fluvial palaeoenvironment was interpreted by Benbow (1982). The conglomerate interbeds thin to the south, suggesting a north to northwesterly provenance.
Indulkana Shale

Lithology and distribution The lithology consists of maroon and green shale with thin, flaggy sandstone beds bearing clay pellets near the base and top. Limestone lenses and micaceous silty sandstone occur locally. The formation is generally irregularly distributed over the basin due to removal by erosion, but is present in the Munyarai Trough. Relationships and boundary criteria Both the upper and lower contacts are conformable. The unit is probably equivalent to the Horn Valley Siltstone of the Amadeus Basin. Thickness The thickness varies from 9 m in Munyarai 1 to 11 m in Lairu 1, but may be up to 60 m in outcrop in the Indulkana Range (Krieg, 1973). Age Late Ordovician, 450445 Ma, based on stratigraphic position and radiometric data. RbSr whole-rock ages of 46015 Ma (Webb, 1978) and 43810 Ma (Womer et al., 1987) suggest a minimum Middle to Late Ordovician age.

Definition and nomenclature The name was introduced informally by Packham and Webby (1969), and used by Krieg (1973). Type section The type locality was nominated as outcrop on the northern side of the Indulkana Range. Subsurface reference sections are defined as 282293 m in Lairu 1 and 1180 1189 m in Munyarai 1 (Fig. 6.24), although the latter correlation is not certain (Womer et al., 1987).

Mount Chandler Sandstone east of Mount Byilkaoora.

(Photo 30900)

Indulkana Shale, Blue Hills.

(Photo T11207)

&

Sedimentology and palaeoenvironment Sedimentary structures recorded by Packham and Webby (1969) in the sandstone facies include ripple marks and desiccation cracks. Trace fossils in the sandstone include Diplocraterion and Skolithos. A very shallow marine, probably tidal bay to shoreface palaeoenvironment is indicated.

Thickness The thickness varies from 162 m in Munyarai 1 to 187 m in Lairu 1, but may be over 1900 m in outcrop to the east (Krieg, 1973). Age Late Ordovician to Early Silurian, 445435 Ma, based on stratigraphic position. Sedimentology and palaeoenvironment Sedimentary structures are dominated by thick sets of cross-bedding, but other features include Skolithos and probable Cruziana trace fossils, desiccation cracks, and ripples which indicate tidal to fluvial-deltaic palaeoenvironments.

Blue Hills Sandstone


Definition and nomenclature The unit was introduced informally by Packham and Webby (1969) as Policeman Well Sandstone, and the name Blue Hills Sandstone was introduced by Krieg (1973). The overlying Cartu beds and Mintabie beds (Packham and Webby, 1969; Krieg, 1973; Townsend, 1990) are here taken as synonyms. Type section The type locality was nominated in outcrop as southwest of Mount Johns, near Policeman Well. Subsurface reference sections are defined as 95282 m in Lairu 1 and 1018 1180 m in Munyarai 1 (Fig. 6.24). Lithology and distribution The lower parts of the formation most likely to be encountered in Officer Basin wells consists of fine to medium-grained quartzose sandstone, with some thin conglomerate beds. Overlying this are soft, kaolinitic sandstone and well-sorted porous sandstone. In the Mintabie area, and possibly higher in the section, the sandstone is feldspathic and micaceous. The formation is very irregularly distributed in the Officer Basin, being present in the Munyarai Trough where not removed by later erosion, but is more likely to be present to the northeast. Relationships and boundary criteria The lower contact with the Indulkana Shale is conformable, and the upper contact is unconformable with the overlying Mimili Formation. The equivalent unit in the Amadeus Basin is the Stairway Sandstone, but the upper parts of the unit may be equivalent to the Stokes Formation and Carmichael Sandstone.

DEVONIAN SEQUENCE Mimili Formation


Derivation of name The formation is named after the Mimili Aboriginal settlement. Definition and nomenclature This name is introduced here for the youngest sandstone unit in the Officer Basin. Type section Surface to 1018 m in Munyarai 1 (Fig. 6.25). Lithology and distribution Tucker (1994) recognised three units in the type section. The lowermost consists of fine to medium-grained micaceous and occasionally calcareous clean arkosic sandstone. The middle unit is fossil-bearing and composed of interbedded greenish grey micaceous, calcareous mudstone with minor siltstone and fine to medium-grained sandstone. The uppermost unit is generally ferruginised, muddy brown arkosic sandstone. Seismic data suggest that the Mimili Formation is restricted to the Munyarai Trough area (Figs 6.26, 6.27). Relationships and boundary criteria The upper contact is not known, but is presumed to be unconformably overlain by Permian, Cretaceous or younger sediments. The basal contact with the Blue Hills Sandstone is disconformable, but has not been cored. The formation may correlate with the Pertnjara Group and possibly the Mereenie Sandstone of the Amadeus Basin. Thickness The Mimili Formation is over 1018 m thick in Munyarai 1, with the top eroded; from seismic data it may be up to 2000 m thick. Age

Mount Chandler Sandstone (foreground), Indulkana Shale and Blue Hills Sandstone, northwestern margin of the Mount Johns Range.
(Photo 30904)

Late Devonian, Frasnian (375360 Ma). A vertebrate fish fauna indicates an Early to Middle Devonian (Eifelian) age, and palynology indicates a Late Devonian (Frasnian) age
&!

Mimili Formation at 923.5 m in Munyarai 1.

(Photo 44383)

in Australia in both marine and freshwater palaeoenvironments (Long et al., 1988). From a regional perspective, a non-marine palaeoenvironment is favoured (Tucker, 1994).

ARCKARINGA BASIN
Permo-Carboniferous Arckaringa Basin sediments are found overlying the eastern Officer Basin, and comprise the Boorthanna, Stuart Range, and Mount Toondina Formations. The geology of the Arckaringa Basin was summarised by Alexander and Sansome (1996). The Waitoona beds (Krieg, 1973) in Officer 1 were previously thought to be Devonian in age. However, based on the common presence of organic-walled worm tubes, they are now known to be equivalent to or younger than the Stuart Range Formation, and are considered synonymous with the Arckaringa Basin succession. Shearer (1994) recently mapped the Permian sediments from seismic uphole and stratigraphic drillhole data. Townsend (1976) and Moussavi-Harami (1994) provided lithofacies details. The total Permian thickness is generally 100400 m, and is greater in the troughs than over ridges. The Permian is very thin to absent in the Marla Overthrust Zone. Palaeoenviron&"

Fig. 6.25 Munyarai 1 Mimili Formation type section.

the latter is considered more reliable. The upper and lower parts of the sequence are barren, and thus the age spanned by the Mimili Formation is wider than indicated by fossils. Sedimentology and palaeoenvironment Planar and ripple laminations are present in the basal part of the sequence, but the vertebrate faunas are found elsewhere

Fig. 6.26 Seismic line IP1-2, through Munyarai 1 (see Fig. 6.27 for location).

Fig. 6.27 Mimili Formation subcrop map.

&#

ments range from fluvioglacial outwash in the lower parts of the succession to shallow marine in the upper parts.

EROMANGA BASIN
A thin veneer of Eromanga Basin sediments also occurs over the eastern Officer Basin, and the units found are the Algebuckina Sandstone, Cadna-owie Formation and Bulldog Shale. These are described in more detail by Alexander and Sansome (1996).

TERTIARY
The Officer Basin region is crossed by southerly directed palaeochannels which drained into the Tertiary Eucla Basin. Visible on NOAA-AVHRR imagery, the channels have incised locally to depths exceeding 100 m and are filled with Eocene to early Pliocene sediments (Benbow, 1995; Benbow et al., 1995; Statham-Lee, 1995).

&$

INTRODUCTION
In a basin with an area of 100 000 km2 and a stratigraphic succession up to 6 km thick, the problem of having too little biostratigraphic data means that there will inevitably be a range of plausible depositional models. With hindsight, some of these prove to be widely off target, as was the case in the eastern Officer Basin when Munyarai 1 was drilled by Continental Oil Company of Australia (Conoco, 1969). Prior to the drilling of Ungoolya 1 in 1986, and despite radiometric age evidence to the contrary (see below), the Munyarai Trough was thought to contain Devonian strata to at least 2399 m, the total depth of Munyarai 1. However, this interpretation did not sit comfortably with outcrops on the eastern and southern margins of the Munyarai Trough which were interpreted to be mainly Neoproterozoic to Ordovician (Krieg, 1973; Krieg et al., 1976). In fact, certain isolated outcrops mapped as Mintabie beds were thought to be Devonian because of the evidence from Munyarai 1 (Krieg, 1969, 1973). A seismic tie from Ungoolya 1 to Munyarai 1 finally settled the order of succession in the Munyarai Trough Devonian on Cambro-Ordovician on Neoproterozoic even though conflicting interpretations were still emerging from careful palynological processing of core samples (Womer et al., 1987). Since 1980, there have been several macrofossil discoveries in the Palaeozoic and major advances in acritarch biostratigraphy in the Neoproterozoic, leading to a more detailed understanding of the Officer Basins depositional sequences. This chapter is concerned chiefly with the biostratigraphic scheme as currently understood, but also reviews radiometric, lithologic and other data that assist correlation in the eastern Officer Basin.

the Amadeus Basin (Stainton et al., 1988). Now, however, biostratigraphic correlation is possible using acritarchs. Recently, Zang and Walter (1992) demonstrated the utility of acritarch studies in parts of the Neoproterozoic and Cambrian of the Amadeus Basin. The biostratigraphy of acritarch assemblages, though still at a preliminary stage, has also greatly improved correlation within the Officer Basin. The preservation of microflora in deeply weathered exposures of the Adelaide Fold Belt is too poor to permit interpretation but samples from cored drillholes on the Stuart Shelf have been tied to a biostratigraphic scheme proposed by Zang (1994). Recent work by Zang (much of it unpublished), based on 207 samples from Officer Basin wells, has added greatly to understanding the composition of acritarch assemblages in the eastern Officer Basin. In his scheme, Zang (1994; Zang and Preiss, in prep.) distinguished eight informal Neoproterozoic acritarch assemblages based on the first appearance of key species. The number has now been reduced to six, as listed in Table 7.1 which also allows for a possible new assemblage (AAC 1) from Cambrian strata. Two Ordovician assemblages (AAO 1, AAO 2) known in the Warburton Basin (Zang, 1994) have been added, based on the likelihood that acritarchs will be recovered from rocks of this age in the Officer Basin. Key elements of the Neoproterozoic microfloral assemblages can also be recognised on the Stuart Shelf. From there, lithostratigraphic correlation provides the only tie to the Neoproterozoic of the Adelaide Fold Belt. The Officer Basin correlations reviewed below are based on a variety of attributes. These include acritarchs and stromatolites in the Neoproterozoic, Cambrian invertebrates, Devonian vertebrates and palynomorphs, and magneto-, chemo- and event stratigraphy, all of which have been applied by different workers to obtain a clearer history of the Officer Basin.

NEOPROTEROZOIC STRATIGRAPHY AND AGE


The Neoproterozoic has been defined globally as beginning at 1000 Ma. In the sense used here, early Neoproterozoic refers to the first three stages of the local Adelaidean succession, namely the Willouran, Torrensian and Sturtian stages. The age of the base of the Willouran stage is ~850 Ma, which corresponds to the late Riphean in Russian terminology. Lack of macrofossils has caused emphasis to be placed on lithostratigraphic comparisons with better known regions such as the Adelaide Fold Belt (Krieg, 1973; Preiss, 1993), the Officer Basin in Western Australia (Townson, 1985) and
&%

Stromatolites
Stromatolites occur in the Coominaree Dolomite of the Peake and Denison Ranges, and two species Acaciella cf. australica and Gymnosolen cf. ramsayi were described by Preiss (1973); these were later amended to Acaciella f. indet and Gymnosolen f. indet (Preiss, 1987). Acaciella also occurs in the Coominaree Dolomite in Manya 5 (Preiss, 1993). This form is very similar to A. australica which occurs in the Skates Hill Formation of the Savory Basin (=Savory Sub-basin of the WA Officer Basin (Perincek, 1996)), and in the Bitter Springs Formation of the Amadeus

Basin. A. australica appears to be restricted to the Callanna Group at the base of the Adelaidean succession, although its full stratigraphic range is not known (Grey, 1995).

Early Neoproterozoic acritarchs


Acritarchs from the Alinya Formation in Giles 1 are interpreted by Zang (1995a; Zang and Preiss, in prep.) to be Torrensian in age. His species (Zang, 1995a, fig. 22) correlate with those from the Bitter Springs Formation (Zang and Walter, 1992), thus Zang places both units and their correlatives in the Torrensian. Zang (1995a; Zang and Preiss, in prep.) reserved a Willouran age for the Cadlareena VolcanicsCoominaree Dolomite Younghusband Conglomerate suite, and has extracted acritarchs from Manya 5 core which he interpreted to be Younghusband Conglomerate. Zang considered this suite to be representative of a graben system beneath the widespread Alinya FormationPindyin Sandstone stratal package. The CadlareenaYounghusband suite crops out in the Peake and Denison Ranges and was fully cored in Manya 5. These strata were not deposited in a graben, and they are considered more or less coeval with the AlinyaPindyin suite (Fig. 5.6), though they may be slightly older (Arkaroola Subgroup versus Curdimurka Subgroup). This work follows the more widely held view that the AlinyaPindyin of the eastern Officer Basin, the Browne LefroyTownsend [Skates hillMunjdajini Spearhole] of the Savory Basin and western Officer Basin, and the Bitter SpringsHeavitree of the Amadeus Basin, are all Willouran in age. The Manya 5 acritarch assemblage is designated AAW 1b in Table 7.1. It may be older than, or the same age as, the Giles 1 acritarch assemblage, designated AAW 1a in Table 7.1. The two assemblages contain a number of species in common, but six species occur only in Manya 5 (Zang and Preiss, in prep.). Zangs (1995a) correlation raises the question as to what does constitute the fill of the channel or graben-like features evident on seismic beneath the AlinyaPindyin stratal package. As discussed in Chapter 5, this underlying sedimentary pile is thought to be older than the Adelaidean.

In the eastern Officer Basin there is a major hiatus between the Alinya Formation and the late Neoproterozoic (Marinoan) Tarlina Sandstone. A Sturtian tillite does intervene but this is restricted to the northeast near Chambers Bluff and to Nicholson 2. There remains uncertainty as to the existence of a disconformity in the Chambers Bluff Tillite, with upper beds being possibly Marinoan in age (Preiss, 1993). No diagnostic acritarchs have been recovered from this part of the succession, although Zang (1994) reported Favosphaeridium and Comaspheridium in an inter-tillitic shale. However, both genera are long ranging, which emphasises the need to identify assemblages to species level for more reliable correlation.

Late Neoproterozoic acritarchs


Arcritarchs from late Neoproterozoic strata were first documented in Munyarai 1 (NaranaDey Dey) by van Neil (1984). He recognised a large (~350 m) form possibly referable to Micrhystridium, but this genus ranges through the Neoproterozoic and Phanerozoic and is of no value for correlation. Van Neil also recognised algal filaments and sphaeromorphs in the Tanana Formation and Dey Dey Mudstone in Murnaroo 1; again, these are long-ranging forms. Possible Trachysphaeridium laufeldi was noted by Eames (in Womer et al., 1987) from 1251.3 m in Birksgate 1, suggesting a likeness to forms known to be of Neoproterozoic age. With regard to Munyarai 1 however, Womer (in Womer et al., 1987) stated The persistent reports of Middle Devonian Ancryospora below 1550 metres tends to support the interpretation that this interval is an Indulkana Shale equivalent, although...this conclusion is very tenuous. It was not until 1992 that Jenkins et al. (1992) and Zang (unpublished report to Comalco) began to carefully extract and systematically document the fragile late Neoproterozoic acritarchs from Officer Basin wells. Jenkins et al. recovered two acritarch assemblages, an older assemblage mainly from the lower Karlaya Limestone in Murnaroo 1 (AAM 3, Table 7.1), and a younger one from the Narana Formation in Ungoolya 1 (AAM 4, Table 7.1). Seismic data (e.g. Sukanta

Table 7.1 Officer Basin acritarch assemblages (after Zang, 1994, 1995a,b).
Zang (1994, 1995a,b assemblages) New assemblage names Genus/species Giles AAW1a Manya AAW1b 1,2,3 AAM1 4 AAM2 5 AAM3 6 AAM4 AAC1 AAC2 AAO1 AAO2

Amadeusphaeridium cyathospora Ambiguaspora parvula Comasphaeridium magnum Comaspaeridium strigosum Cymatiosphaeridium kullingii k Goniosphaeridium cebrum Unispinosphaeridium n.sp. Hocosphaeridium scaberfacium Leiofusa squama Petaloferidium sp. Skiagia spp. Tongzia meitana Trachystrictosphaera aimika k Trachystrictosphaera vidalii k Unispinosphaeridium willouranum gen et sp nov.
k=Key first appearance species

k k k k k k k k k k k k

&&

et al., 1991) indicate that the beds yielding these assemblages are separated by a canyon-cutting unconformity. Jenkins et al. (1992) noted that the older assemblage consisted of diverse spinose acritarchs and few filamentous forms. The younger assemblage was dominated by filamentous types with 810 species of very large, unsculptured spheroidal forms. Jenkins et al. (1992, p.405) noted that the differences between the two microfloral assemblages may be due to chronological or biofacies differences, or both. Zang (1994, p.204) noted that both abundance and diversity of species increased during marine transgressions which he interpreted as a transfer of phytoplankton from relatively calm offshore environments to turbulent shelf environments with rising sea level necessitating adaptation. The schematic curve on Figure 7.1 (right hand column) illustrates the change in species diversity which reached a maximum of more than 80 species in the upper Dey Dey Mudstone, representing the Hocosphaeridium scaberfacium Goniosphaeridium crebrum Assemblage (Zang 1994). Many of the species in this assemblage (AAM 3, Table 7.1) had not previously been found in strata older than Palaeozoic although several, including the nominate species, have been recorded from the Pertatataka Formation in the Amadeus Basin by Zang and Walter (1992). Another significant discovery by Zang (1994) is the oldest occurrence of the trilete spore Ambiguaspora parvula in chert from the Punkerri Sandstone in the Birksgate Sub-basin (AAM 4, Table 7.1), the first record of a higher plant anywhere in the world.
Late Neoproterozoic invertebrates

trace fossil. The fauna permits correlation from the Ediacara Member of the Adelaide Fold Belt to at least part of the Punkerri Sandstone. Impressions resembling the attachment stalk of Charniodiscus have been found near Albany more than 1500 km distant (Cruse et al., 1993), attesting to the wide distribution of this fauna. Jenkins (1992) suggested that acritarchs may have constituted the food supply for some of the Ediacara metazoans, and Zang (1993) regarded the oldest semi-aquatic plant (parent of the trilete spores in the Punkerri Sandstone) as another food source.
Geochronology

As mentioned above, Compston (in Conoco, 1969) obtained an isochron of 650 Ma from the Munyarai Formation in Munyarai 1. However, Webb (in Preiss, 1993) calculated a considerably younger RbSr age of 58835 Ma from the Yarloo Shale of the Stuart Shelf which correlates with the Dey Dey Mudstone. Since the Dey Dey underlies the Munyarai Formation, an age discrepancy of at least 70 million years is evident. Greater RbSr ages have been obtained: 73154 Ma from the Rodda beds east of Chambers Bluff and 715210 Ma for the Dey Dey Mudstone in Murnaroo 1 (Webb in Womer et al., 1987), demonstrating that geochronology has not assisted Neoproterozoic correlation in the Officer Basin. Nevertheless, the age difference between the Devonian Mimili Formation and Neoproterozoic Munyarai Formation, shown on Figure 7.2, should have been evident shortly after Munyarai 1 was drilled.
Chemo- and magnetostratigraphy

Specimens of the soft-bodied Ediacara fauna have been recorded from loose blocks in the Punkerri Sandstone. Daily (in Major, 1974) recorded Charnia, ?Charniodiscus, Tribrachidium, Rangea and a double spiral, presumably a

Patterns of secular change in isotopes of carbon, sulphur and strontium can indicate rates of burial of organic carbon as well as biological events such as major extinctions, and erosional cycles. It may be possible to track parallel

Fig. 7.1 Eastern Officer Basin Neoproterozoic lithostratigraphy and acritarch assemblages. The curved line in the right-hand column represents relative acritarch species abundance (after Zang, 1994, 1995b).

&'

Acritarchs from the upper Younghusband Conglomerate, a possible equivalent of the Pindyin Sandstone, in Manya 5. A Amadeusphaeridium sp. B Acritarch with a prominent process. C Trachyhystrichosphaera aimika Hermann, 1976 emend. Butterfield, Knoll and Swett, 1994. Acritarchs from the Alinya Formation in Giles 1. D Comasphaeridium tonium Zang, 1995. E Lomentunella sp. F Trachyhystrichosphaera sp. Acritarchs from the Meramangye Formation in Meramangye 1. G ? Vandalosphaeridium sp. H Acritarch with a prominent process. I Pterospermella sp. Scale bar (in F) is 20 m for D, G, H, I; 25 m for B; 32 m for C, E, F; and 80 m for A.

'

Acritarchs from the Dey Dey Mudstone in Lake Maurice West 1. A Tanarium sp. B Comasphaeridium sp. C Unnamed species. Acritarchs from the Tanana Formation in Giles 1. D Multifronsphaeridium pelorium Zang in Zang and Walter, 1992. E Hocosphaeridium scaberfacium Zang in Zang and Walter, 1992. F Acritarchs with a prominent process and numerous conical spines. Acritarchs from the Munyarai Formation in Munyarai 1. G Goniosphaeridium sp. H Trachyhystrichosphaera sp. I Gyalosphaeridium sp. Scale bar (in I) is 20 m for B, H, I; 32 m for C, G; 50 m for F; 63 m for E; and 125 m for A, D.

'

Acritarchs from the Narana Formation in Munta 1 (A) and Lairu 1 (B, C). A Comasphaeridium sp. B Brevitrichoides sp. C Leiosphaeridia sp. Acritarchs from the Early Cambrian Ouldburra Formation in Manya 6. D Ceratophyton sp. E Ceratophyton sp. Miospores from the Devonian Mimili Formation in Munyarai 1. F Ancyrospora sp. G Ancyrospora sp. H Punctatisporites sp. I Retusotriletes sp. Scale bar (in I) is 20 m for A, B; 25 m for C, D, E; 32 m for F, G, I; and 40 m for H.

'

variations in isotopic signatures within and even between basins, so providing another correlation tool. The Neoproterozoic of the Officer Basin has few carbonate-rich units but nevertheless preliminary 13C measurements have been made and tentative correlations suggested. A slight contrast (~3) between the 13C-enriched Karlaya Limestone in Murnaroo 1 and the negative to moderately positive TananaNarana Formations in Ungoolya 1 was noted by Jenkins et al. (1992), and attributed to the different sample ages with some contribution from facies effects. Pell et al. (1993) presented a preliminary carbon isotope curve for the South Australian late Marinoan stage based on values from the Wonoka and Billy Springs Formation of the Adelaide Fold Belt, augmented by data from Observatory Hill 1 and Ungoolya 1. This curve displays consistently negative values (-6 to -8.5) through 360 m of Wonoka Formation in Bunyeroo Gorge, implying low rates of burial of organic carbon and hence poor source rocks from this interval. In contrast, total organic carbon of the slightly older Dey Dey Mudstone (=Bunyeroo Formation) in Lake Maurice West 1 and Karlaya 1 is ~0.8 to 1.5% (see Ch. 8), which is moderately rich. These wells are on the Murnaroo Platform which subsided at one quarter the rate of the Munyarai Trough (Moussavi-Harami and Gravestock, 1995), and thus even richer source rocks are anticipated there. These should equate to strongly positive 13C values. Clearly, the isotopic results from the limited data set in the Officer Basin are very preliminary and it is likely that facies effects will loom large in any interpretation. The magnetostratigraphic characteristics of cores from Observatory Hill 1 and Lake Maurice West 1 have been documented by Chamalaun et al. (1990). The remanent magnetisation of ~400 samples was measured with a spinner magnetometer through the lower Karlaya Limestone, Dey Dey Mudstone and upper Murnaroo Formation after carefully fitting the core together. Since the cores are vertical and not orientated with respect to azimuth, an arbitrary fiducial line was scribed and directions were measured relative to it. Two magnetic events are represented within the Dey Dey Mudstone in both wells. Magnetic inclination values are

more variable in the lower event which may represent an epoch of magnetic field excursions rather than a reversal. The upper event is sharp and probably a true reversal. As shown on Figure 7.3, both events are within the Comasphaeridium magnum acritarch assemblage zone (AAM 2, Table 7.1), which suggests that magnetostratigraphy holds promise for more detailed correlations within acritarch zones. The two events are 75 m apart and, assuming a subsidence rate of 15 m per million years

Fig. 7.2 RbSr isochrons of the Neoproterozoic Munyarai Formation, Ordovician Indulkana Shale and Devonian Mimili Formation. See text for details.

Fig. 7.3 Lake Maurice West 1 Neoproterozoic lithostratigraphy, acritarch assemblages, and depths of magnetic field excursions and the Acraman impact layer.

'!

(Moussavi-Harami and Gravestock, 1995), they are separated in time by about five million years.
Lithology and event stratigraphy

It should have been obvious at the time of drilling that the tracheidal wood fragments claimed to have been recovered from total depth in Munyarai 1 did not come from the hard, laminated dolomitic siltstone characteristic of the Dey Dey Mudstone intersected in that well. The lithologic succession is now sufficiently well known to enable wireline log correlations to be carried out with some confidence. It is only in the Marla Overthrust Zone where Proterozoic intersections are few, and at the Rodda beds type section which exceeds 3 km in thickness (Preiss and Krieg, 1992), that correlation remains a problem within the basin. Correlation with other basins is still a matter of debate. On a sequence stratigraphic basis, Sukanta et al. (1991) correlated the Marinoan in the Officer Basin with the Adelaide Fold Belt. A key argument for this is a set of submarine canyon incisions common to both regions which suggested correlation from the Narana Formation in the Officer Basin to the Wonoka Formation in the Adelaide Fold Belt. However, from acritarch studies, Zang (1995b) correlated the canyon surface at the base of the Narana Formation with the disconformity at the base of the Rawnsley Quartzite and not the Wonoka Formation which is stratigraphically older. One key horizon for interbasinal correlation is low in the Dey Dey Mudstone in the Officer Basin and low in the Bunyeroo Formation of the Adelaide Fold Belt. This is the Acraman impact layer, a bed of ejecta which resulted from a bolide impact on the Gawler Craton. The impact crater is modern Lake Acraman and impact debris was spread over a wide area to form a unique marker bed. The bed has been identified in cores from Observatory Hill 1 and Lake Maurice West 1, and comprises medium to coarse-sand-sized acid volcanic clasts. A thin, pale green alteration halo is evident in the surrounding red-brown mudstone. The preservation of the layer suggests that the Dey Dey Mudstone is a deep water deposit (Wallace et al., 1990). The impact layer is a few metres above the magnetic polarity excursion in both wells (Fig. 7.3) which lends weight to magnetostratigraphy as a correlation tool.

al., 1987), had been incorporated within the Observatory Hill Formation. Fragments of two small trilobites in a core from 87.85 m in Marla 1 were identified as belonging to a genus of redlichiids, and are therefore Early Cambrian in age (Jago and Youngs, 1980). One further specimen from this depth, and other trilobite remains from 333 m in the adjacent Marla 1B, remained unidentified. Well-preserved trilobites from 967.7 to 970.1 m in Manya 6 also came from the Ouldburra Formation. These belong to a new species of Abadiella and indicate an Early Cambrian (Atdabanian) age (Jago et al., 1994). Trilobite fragments have been recorded over a 394 m interval (8891263 m) in this well by Dunster (1987a). He also noted the lowest fossils (sponge spicules) at 1391 m. Detailed core logging by him also revealed trilobites between 357.8 and 385 m in Manya 3; this interval is in sequence C1.3, considered Early Cambrian (Botomian) in age (Gravestock and Hibburt, 1991). Possible archaeocyaths recorded at 1207 m in Manya 6 are calcite pseudomorphs of aragonite fan cements from a reworked bioherm. Dunster (1987a) has documented archaeocyath calcimicrobe biohermal buildups between 399 and 654 m in Marla 6, as well as wackestones with fossil hyoliths (suggesting a provenance for Gatehouses (1976) sample), sponges and ostracods. This fauna indicates an Early Cambrian age of the Ouldburra Formation even though fossil preservation is too poor to permit identification to genus level.

CAMBRIAN STRATIGRAPHY AND AGE


As no Cambrian body fossils have been found in outcrops, correlation prior to their discovery in a drillhole in 1978 was by lithological comparison with the well-known succession of the Flinders Ranges (e.g. Wopfner, 1969). A possible hyolith conch (?Biconulites) was reported from float in a modern claypan near the Observatory Hill Formation type section (Gatehouse, 1976). However, the oolitic lithology which encloses the specimen does not occur in this formation and the specimen is presumed to have been transported. Biostratigraphic and lithological correlations of Cambro-Ordovician strata are shown on Figure 7.4.
Invertebrate fossils

Marine invertebrates have been found in the Ouldburra Formation which, prior to stratigraphic revision (Brewer et
'"

Fig. 7.4 CambroOrdovician lithostratigraphy, fossils and geochronology.

The presence of chitinophosphatic shelly fossil fragments, reported from Wilkinson 1 by Kantsler (in Gatehouse, 1979), could not be verified from thin section examination.

Acritarchs
An impoverished assemblage of acritarchs was reported from between 314.5 and 512.2 m in Wilkinson 1 by Muir (in Gatehouse, 1979), an interval now correlated with the Ouldburra Formation. At that time, Muir considered the Tommotian Stage to be latest Proterozoic, whereas it is regarded now as the second stage of the Early Cambrian. In any event, the stratigraphic range of the four species reported by her can only be regarded as Neoproterozoic to Middle Cambrian and better age resolution is not possible. Re-examination of the same interval in Wilkinson 1 failed to find any microfossils, although sheet-like or granular organic matter was found to be fairly common (van Neil, 1984). Acritarchs have been suggested as possible source organisms for dinosterane which occurs in samples from the Ouldburra Formation in Karari 1 (Kamali, 1995a,b). Cambrian acritarchs recovered from the Early Cambrian of the Manya Trough and Marla Overthrust Zone are poorly preserved and of low diversity in comparison to coeval deposits in the Arrowie and Stansbury Basins (W-L. Zang, MESA, unpublished data, 1996). Only one assemblage has been assigned (AAC 1, Table 7.1), and the characteristically Cambrian genus Skiagia has not yet been recognised.

chert, calcite-after-trona pseudomorphs and green garnet. The first two correlate with the Moyles Chert Marker and Parakeelya Alkali Member respectively, and the green garnet is a common accessory in lag gravel at the disconformable base of the formation in both wells. This disconformity is currently known only in the Ungoolya area and on the Munyarai structure; no garnets have been noted where the lower boundary of the Observatory Hill Formation is conformable. Fine-grained, centimetre-thick crystal tuff bands have been recognised in the Cadney Park Member which is dominantly a gypsiferous redbed suite. The tuff bands (at 545.2 and 562.3 m in Manya 2; 436.2 m in Manya 4; 625.9 and 636.6 m in Byilkaoora 3) are buff-grey and composed of subhedral, partly sericitised feldspar crystals and relict shards in a glassy groundmass. Henry and Brewer (1984) noted this tuffaceous component and suggested a tentative correlation between the Cadney Park Member and the basaltic Cadlareena Volcanics (their Welbourn Volcanics). These basalts are now known to be Proterozoic and unrelated, but that does not lessen the correlation potential of the tuffs within the Cambrian succession. Pontifex (1984, p.190) commented that The source of this fine tuffaceous material...is uncertain, but seems likely to be deposited from high-level drifting clouds from a volcanic centre, maybe 1000 plus kilometres away. Their presence supports correlation of the Cadney Park Member with formations of the upper Hawker Group and/or Billy Creek Formation in the Arrowie Basin (sequences C1.3 and C2.1 of Gravestock and Hibburt, 1991), where tuff beds up to 1.5 m thick attest to active rift volcanism. It is possible that the Officer Basin tuffs represent fallout from the same volcanic complexes on the palaeo-Pacific margin of Gondwana.

Geochronology
An outcropping silty shale in the Chambers Bluff area, and reported as apparently conformable beneath the Mt Chandler Sandstone, had a calculated RbSr age of 57434 Ma (Webb, 1978). The shale was thought to be Observatory Hill Formation but, if this was the case, the contact would not be conformable. Given the magnitude of the associated error, this age does not improve correlation. Still less accurate ages of 52468 and 66060 Ma were obtained from the Observatory Hill Formation in Byilkaoora 1 (Webb, 1978) and Cadney Park Member in Byilkaoora 3 (Henry and Brewer, 1984). In the Ouldburra Formation in Wilkinson 1, strontium levels are too high for reliable age determination by the RbSr method (Webb, 1978), and it may be that the Observatory Hill Formation and Cadney Park Member are unsuitable for the same reason.

ORDOVICIAN STRATIGRAPHY AND AGE


Ordovician strata are interpreted in Munyarai 1 between 1018 and 1355 m depth, based on lithological similarity with outcrops of Blue Hills Sandstone, Indulkana Shale and Mount Chandler Sandstone in the EVERARD 1:250 000 map area. These outcrops have been correlated with the Ordovician succession in the Amadeus Basin, relying strongly on similarities of the lithology, sequence and tectonic setting (Krieg, 1973, p.14). No body fossils have been found and acritarchs have not been recovered from the few, mainly sandy well intersections. The Ordovician acritarch assemblages shown in Table 7.1 are in the Warburton Basin.

Lithology
Diagnostic lithologies have been employed to demonstrate or reinforce correlations in the absence of fossil data. The most widely correlated lithology is chert, which occurs abundantly as concretions, sheets and breccias at two key levels of the Observatory Hill Formation the Moyles Chert Marker Bed and the Parakeelya Alkali Member. Brewer et al. (1987) considered the chert lags and concretions found at the type section near Observatory Hill (Wopfner, 1969) to correlate with the Parakeelya Alkali Member. As part of their objective of correlating Munyarai 1 with Ungoolya 1, Womer et al. (1987) employed three key lithologies associated with the Observatory Hill Formation
'#

Ichnology
Trilobite resting traces (Rusophycus) and Skolithos and Diplocraterion burrows in the Mount Chandler Sandstone suggested equivalence with the Pacoota Sandstone, although some differences were noted (Packham and Webby, 1969; Krieg, 1973; Benbow, 1982). Trace fossils are usually considered to be good facies and palaeoenvironmental indicators, and as most types have long time ranges they are not often used for stratigraphic correlation and age dating. However, the exception to this rule are the trace fossils associated with trilobites in early Palaeozoic rocks. These trace fossil types include Rusophycus, Cruziana, and

Rusophycus sp. in Apamurra Formation siltstone, Mount Johns Range. Photo courtesy of Dr J.B. Jago. (Photo 44430)

Diplichnites. There are 19 species of Cruziana that have restricted ranges in the Cambrian and Ordovician (Crimes, 1975). In the Mount Johns Range adjacent to the Marla Overthrust Zone, trilobite trace fossils are common, and there is potential for stratigraphic correlation of the Marla and Munda Groups based on trace fossils. Although bioturbation has been recorded in drill cores, diagnostic trace fossils have not been found.

Geochronology
Six samples were collected from outcrops 2 km east-northeast of Chambers Bluff to determine the age of the Indulkana Shale. RbSr analysis produced the ratios plotted on Figure 7.2, which represents a Late Ordovician isochron of 46015 Ma (Webb, 1978). An additional surface sample was collected some years later and its age determined as 43810 Ma (Early Silurian) using the KAr method (Womer et al., 1987). Webbs (1978) age is preferred because of the consistency of data from the six samples. No fossils have been found in the Blue Hills Sandstone and the age of this unit may well range into the Silurian.

Lithology
The Mount Chandler Sandstone, besides containing trace fossils, is distinctive by virtue of its clean, well-sorted, mature quartzose lithology, which contrasts with other more feldspathic or micaceous units in the basin (Krieg, 1973). It is an attribute that can be used for correlation in the absence of other data. The Mount Chandler Sandstone in Manya 5 has been identified in this manner. The Blue Hills Sandstone is the only other lithologically similar formation. Detailed aeromagnetic maps show a distinctive magnetic marker bed which is clearly associated with continuous outcrops of Indulkana Shale on the eastern rim of the Munyarai Trough. Shallow subcrops having the same magnetic character occur northeast of the Marla Overthrust Zone and, though undrilled, have been correlated confidently with the Indulkana Shale by Hamer (1994). The magnetic source, presumably magnetite, has not been identified.
Fig. 7.5 Munyarai 1 Devonian lithostratigraphy, fossils and geochronology.

Vertebrate fossils
Cuttings between 816.9 and 823.0 m in Munyarai 1 yielded scales from thelodont, acanthodian and ganoid (osteichthyan) fish (Gilbert-Tomlinson in Conoco, 1969), indicating a maximum age of Early Devonian for this part of the section. The fossiliferous interval was extended to lie between 738 and 860 m (J. Hibburt, MESA, pers. comm., 1987) and the fossils were examined by Long et al. (1988). These authors identified scales of thelodonts, an acanthodian, teeth, plates and bone fragments, and correlated this assemblage with the early Middle Devonian (Eifelian) Wuttagoonaspis fish fauna of the Georgina and Amadeus Basins. The Munyarai fauna is not large enough to determine whether the host strata are marine or non-marine.
'$

DEVONIAN STRATIGRAPHY AND AGE


Devonian fossils occur in situ in the Mimili Formation and comprise vertebrate remains and spores (Fig. 7.5).

Palynomorphs
Palynomorphs were not recovered from the fish-bearing strata by Long et al. (1988) and, despite preparation of large quantities of material, only two spores were recovered from core 3 by Harris (in Conoco, 1969). The specimens (cf. Leiotriletes) only indicate a Devonian or younger age. Fragments of tracheidal wood reported from the bottom core and as high as core 3 are now known to be contaminants, but it was these that stood for many years as testimony to the supposed Devonian age of the Munyarai 1 succession below the fish scale zone. Further sampling was carried out by Vlierboom (1973) who was the first to report palynomorphs from core 2, at a depth of 548.6 m (189 m above the top of the fish fauna). Vlierboom carried out two preparations on this sample to improve the microfossil yield; he stated In the first sample preparation several well-preserved Triassic sporomorphs were found but as the samples were previously determined as Siluro-Devonian it was suspected that the samples had been contaminated. The first preparation yielded only seven specimens, two of which Platysaccus queenslandi and Alisporites sp. indicated a Triassic age. The second preparation yielded 34 specimens, all of which were concluded by Vlierboom to indicate a Late Devonian (probably Famennian) age. Further work is planned to examine the Triassic question. More recently, Eames and Miller (in Womer et al., 1987) identified well-preserved miospores in three samples from 640.1 to 853.5 m. The good preservation of this assemblage contrasted strongly with the very poor preservation of eight samples recovered from the much deeper interval (1652.02225.1 m) in Munyarai 1. Although the lower samples were regarded as probably Middle Devonian, they are now known to be Neoproterozoic in age. The upper assemblage of miospores was confidently assigned a Late Devonian (Frasnian) age because of similarity to a Frasnian assemblage from the Carnarvon Basin in Western Australia. The stratigraphic position of the Devonian assemblage is shown on Figure 7.5, which indicates an overlap of 116 m that may be either Middle Devonian (based on fish fossils) or Late Devonian (based on spores). The interval which may be positively correlated with Devonian strata based on fossil content comprises only 30% of the top 1000 m drilled. At least half the drilled section, from the surface to 548.6 m, may be considerably younger than Devonian, but this section is mainly sandy and lacks fossils. A sample of core 1 (249.3253.0 m) has been submitted to Geotrack in Melbourne for assessment of its zircon fission track age.

Geochronology
Whole-rock samples of core in the depth range 632.5634.3 m and 2133.62136.4 m were analysed for RbSr age information by Compston (in Conoco, 1969). Compston showed clearly that the shallower samples occupied a distinct field with a lower gradient and an equivalent age between 300 and 400 Ma, i.e. Carboniferous to Devonian (Fig. 7.2). Interpreted simply as an isochron, the slope of the well-aligned deeper samples gave an equivalent age of 650 Ma, clearly Neoproterozoic as was concluded much later.
'%

'&

INTRODUCTION
One key attribute, namely the association of Proterozoic and Cambrian source rocks and evaporites, is shared between the Officer Basin, Persian Gulf and southwest Siberian Platform. In these regions, halite and anhydrite are associated at two main stratigraphic levels with argillaceous dolomite (the chief source rock) and sandstone. These were originally interbedded sabkhalagoonal deposits but they have been extensively altered by halokinesis. Mobile salt occurs in the early Neoproterozoic (~800 Ma) Alinya Formation of the Officer Basin, Bitter Springs Formation of the Amadeus Basin, and Hormuz Series of the Persian Gulf region where salt diapirs have become major oil traps (Edgell, 1991). The evaporite association also occurs in younger strata of late Proterozoic and early Cambrian age (~650518 Ma) in southwestern Siberia (Kuznetsov et al., 1992), Oman and the ManyaTallaringa Trough region of the eastern Officer Basin. The salt has remained largely immobile in the Officer Basin and Siberia. Despite major oil and gas production from Siberia and the Persian Gulf, and numerous in situ oil bleeds (e.g. Byilkaoora 1), the ProterozoicCambrian hydrocarbon potential of the Officer Basin has been perceived as low, a perception shared with rocks of similar age in the Amadeus Basin (e.g. Summons and Powell, 1991). Current interpretations emphasise that Mereenie and Palm Valley oil and gas migrated from source facies of the Ordovician Horn Valley Siltstone, whereas oil generated in the Proterozoic or Cambrian migrated prior to trap development (e.g. Jackson et al., 1984). This assumption, based mainly on the stratigraphic range of the cyanobacterium Gloeocapsamorpha prisca, is weakened by several factors. These include the dry gas composition at Palm Valley, the presence of gas but not oil at West Walker 1 (Jackson et al., 1984), wet gas in the Bitter Springs Formation at Magee 1 (WakelinKing, 1994), and the discovery of G. prisca in the Early Cambrian Ouldburra Formation in the Officer Basin (Kamali, 1995a; Michaelsen et al., 1995). On the basis of this and other evidence, the Amadeus Basin oil and gas may have been generated from formations as old as the Bitter Springs Formation, a play that was actively explored by Pacific Oil and Gas Pty Ltd and resulted in the Magee 1 discovery (Wakelin-King, 1994). A lack of knowledge of pre-Ordovician source rocks should not be a deterrent to exploration in the Officer Basin. As McKirdy (1993) pointed out, the eastern Officer Basin is remarkably well endowed with oil shows in Neoproterozoic
''

and Cambrian rocks, the oils representing four genetically distinct families. Potential source rocks appear to be organically lean on average but this is largely due to the pyrolysis technique being not well suited to treatment of the carbonate-dominated lithologies. Most samples comprise vertical portions of core, even though the organic-rich material is horizontally laminated. Pyrolysis is carried out on powdered rock samples which thus comprise perhaps 90% matrix and must be considered minimal because samples are pre-digested in ~30% HCl, which is harsh treatment. The Total Organic Carbon (TOC) content of Officer Basin samples is usually quite low (<0.5%) but, where organic-rich layers are specifically sampled, the TOC may be >4.5% (McKirdy et al., 1983). Figure 8.1 provides a stratigraphic summary of both the source potential of the Officer Basin and oil shows.

FORMATION DESCRIPTION
Alinya Formation
Source richness Six samples of the Alinya Formation analysed from Giles 1, and nine from Manya 5, revealed that the source richness is very poor to poor (average TOC = 0.24, range 0.020.62). However, samples are from thin black shale beds at the base of upward-shallowing evaporite cycles (Zang and McKirdy, 1994) in Giles 1 and redbeds in Manya 5, and may not be representative of the more basinal parts of the unit. Given the very limited sampling from only two wells, it is premature to dismiss the unit as too organically lean for hydrocarbon generation. Available Rock-Eval data for the Alinya are summarised in Table 8.1. Kerogen type and source potential Hydrogen Index (HI) values are commonly low (0106), indicating gas-prone Type IV to III kerogen (Fig. 8.2). The samples in Manya 5 have near zero genetic potential; these are overmature and contain intertinite or sub-graphitic kerogen (McKirdy in Weste, 1984). Samples from Giles 1 indicate poor genetic potential up to 0.91 kg hydrocarbon per tonne of source rock, but an almost identical molecular biomarker assemblage to the Alinya Formation is found in the correlative Gillen Member of the Bitter Springs Formation (Amadeus Basin; McKirdy in Zang and McKirdy, 1993), which also contains an oil-prone black, pyritic shale (in Bluebush 1, Mount Charlotte 1; Jackson et al., 1984). Thus, oil-prone kerogens may be found in the Alinya Formation. Confirmation of this potential is provided by oil

Fig. 8.1 Summary stratigraphic column of the Officer Basin, showing potential and proven source units, and oil shows.



Fig. 8.2 HI versus Tmax plot, Alinya Formation.

Photomicrograph of algal mats in the Alinya Formation in Giles 1: Top: intertidal anhydrite; cross-polars, width of view is 5.7 mm. (Photo 44405) Bottom: subtidal shale; plane polarised light, width of view is 2.2 mm. (Photo 44406)

shows recorded from the Relief Sandstone in Observatory Hill 1, and from the Murnaroo Formation in Lake Maurice East and West wells. The source for this oil has been unequivocally determined as Alinya Formation based on biomarker distributions (McKirdy, 1993). Origin and organic petrology McKirdy (in Zang and McKirdy, 1994) has identified a diverse assemblage of molecular biomarkers from the Giles 1 evaporitic facies association, which indicates derivation from precursor eucaryotic algae and eubacteria. The tidal flat facies of the Alinya Formation in Giles 1 contains abundant mats of the cyanobacterium Eoentophysalis gilesis (Zang, 1995a), which may form stratiform stromatolites 18 mm thick. The thinly bedded to laminated tidal flat evaporite association is very similar to that described from Shark Bay by Ferguson and Skyring (1995, fig. 5d), despite the 800 million year age difference
Table 8.1 Rock-Eval source rock data, Alinya Formation.
Well Depth (m KB) Giles 1 1243.0 1256.0 1265.5 1266.4 1283.9 1300.6 1314.2 TOC (%) 0.62 0.40 0.62 0.26 0.28 0.26 0.46

between the two occurrences. Such cyanobacterial mats are capable of transforming into oil-prone kerogens under anoxic conditions.

Coominaree Dolomite
The Coominaree Dolomite is restricted to the eastern part of the basin, and has been intersected only in Manya 5. It is more widespread in the Peake and Denison Ranges, 200 km to the southeast. Contrary to Gravestock and Sansome (1994), no samples have been analysed for source rock quality, but the presence of stromatolites in the unit may indicate a minor source potential. However, the oxygenated environment may have resulted in the organic matter being consumed by aerobic bacteria.

Meramangye Formation
Only two samples (Marla 9) from the marine Meramangye Formation have been analysed for TOC, and

Genetic potential (kg/t) 0.91 0.26 0.47 0.00 0.01 0.01 0.01

Oxygen Index 29 8 0 158 64 50 61

Hydrogen Index 106 58 60 0 0 50 0

Tmax (C) 445 475 439 427 323 218 271

Manya 5



indicate a very poor source with TOC ranging from 0.07 to 0.09%. No Rock-Eval analyses have been performed

marine. Rock-Eval data for the Dey Dey Mudstone are summarised in Table 8.2. Kerogen type and source potential HI values are moderate (100382), indicating oil-prone Type II kerogen (Fig. 8.3), with poor to fair genetic potential up to 2.92 kg hydrocarbon per tonne of source rock. The oil potential of the upper Dey Dey Mudstone is confirmed by the presence of oil shows in thin sandstone beds in Karlaya 1 and Marla 9. This is discussed further in Chapter 9. Origin and organic petrology The potential source beds are thin but contain oil-prone lamalginite which was probably derived from acritarchs and cyanobacterial mats (McKirdy, 1993). Sterane distributions suggest a green algal source (see Tanana Formation).

Dey Dey Mudstone


Source richness Fifty-one samples of the Dey Dey Mudstone analysed from Giles 1, Karlaya 1, Munta 1, Munyarai 1, and Ungoolya 1 indicate source richness ranging from very poor to poor (average TOC = 0.11%, range 0.030.81%). The source potential improves towards the top of the formation, where the depositional environment changes from fluvial to

Karlaya Limestone
Source richness The Karlaya Limestone is widely distributed over the Murnaroo Platform and in the Munyarai Trough, and is also probably present in the Birksgate Sub-basin. Twenty-eight samples have been analysed for TOC from Karlaya 1, Munyarai 1, Murnaroo 1 and Giles 1, which indicate a poor source with an average TOC of 0.20% (range 0.020.72%); this is, however, significantly better than the overlying Tanana Formation. Rock-Eval analyses are summarised in Table 8.3. Kerogen type and source potential HI values are moderate (173194, two samples only), indicating oil-prone Type II kerogen (Fig. 8.4), with poor genetic potential up to 1.54 kg hydrocarbon per tonne of source rock. Given that the Karlaya Limestone contains isotopically heavy shelf carbonate indicative of organic matter burial (Jenkins et al., 1992; Pell et al., 1993), the source potential might be expected to be better than the data would suggest. Oil shows, which may have been sourced from the underlying Dey Dey Mudstone, have been recorded from Karlaya 1.
Fig. 8.3 HI versus Tmax plot, Dey Dey Mudstone. Table 8.2 Rock-Eval source rock data, Dey Dey Mudstone.
Well Depth (m KB) Karlaya 1 2093.7 2094.9 2096.2 2099.2 2101.0 TOC (%) 0.81 0.40 0.21 0.23 0.31 Genetic potential (kg/t) 2.92 2.10 0.32 0.30 0.48 Oxygen Index 34 50 119 126 61 Hydrogen Index 285 382 109 100 119 Tmax (C) 432 431 422 424 426

Table 8.3 Rock-Eval source rock data, Karlaya Limestone.


Well Depth (m KB) Murnaroo 1 182.3 182.4 185.4 TOC (%) 0.40 0.72 0.60 Genetic potential (kg/t) 0.22 1.54 1.18 Oxygen Index 98 116 Hydrogen Index 194 173 T max (C) 417 426 424



0.021.11%). The richest samples occur in Marla 9, although the highest value was from Lairu 1. Rock-Eval data for the Tanana Formation are summarised in Table 8.4. Kerogen type and source potential HI values are low (3100), indicating gas-prone Type III kerogen (Fig. 8.4), with a poor genetic potential of up to 0.67 kg hydrocarbon per tonne of source rock. However, oil bleeds from fractures in Marla 9 suggest that oil has been generated. Origin and organic petrology The Marla 9 (Tanana Formation) and Karlaya 1 (Dey Dey Mudstone) oils have distinctive sterane distributions similar to those in Neoproterozoic oils and source rocks from Oman and Siberia. Primitive green algae appear to be the precursors (McKirdy, 1993).

Munyarai Formation
Four samples were analysed from the Munyarai Formation in Munyarai 1, with TOC ranging from 0.08 to 0.20%. No Rock-Eval analyses were performed. The unit is predominantly fine grained and the environment is interpreted to be marine prodelta and outer shelf. This formation may have source potential but further drilling is required to obtain samples for analysis.
Fig. 8.4 HI versus Tmax plot, Karlaya Limestone and Tanana Formation.

Narana Formation
Source richness Fifty-one samples of the Narana Formation analysed from Byilkaoora 1, Karlaya 1, Lairu 1, Munta 1 and Ungoolya 1 indicate source richness ranging from very poor to moderate (average TOC = 0.16, range 0.040.57%). Rock-Eval data for the Narana Formation are summarised in Table 8.5.

Tanana Formation
Source richness One hundred and seven samples of the Tanana Formation, analysed from Giles 1, Karlaya 1, Lairu 1, Marla 9, Munta 1, Munyarai 1 and Ungoolya 1, indicate source richness ranging from very poor to moderate (average TOC = 0.14%, range
Table 8.4 Rock-Eval source rock data, Tanana Formation.
Well Marla 9 Ungoolya 1 Depth (m KB) 218.9 269.3 1623.1 1930.6 TOC (%) 0.92 0.53 0.20 0.20

Genetic potential (kg/t) 0.05 0.67 0.16 0.12

Oxygen Index 11 28 140 75

Hydrogen Index 3 100 75 60

Tmax (C) 438 350 362

Table 8.5 Rock-Eval Source rock data, Narana Formation


Well Depth (m KB) Karlaya 1 1424.6 1435.7 1594.57 1287.2 1287.6 1299.9 1325.1 1345.1 1450.1 1499.8 1523.6 1525.0 TOC (%) 0.21 0.32 0.21 0.57 0.30 0.20 0.20 0.34 0.21 0.21 0.22 0.21 Genetic potential (kg/t) 0.08 0.29 0.10 0.09 0.06 0.08 0.27 0.21 0.20 0.42 0.38 0.33 Oxygen Index 347 262 319 26 40 140 180 91 171 171 209 186 Hydrogen Index 38 81 47 16 20 40 120 56 86 171 145 129 Tmax (C) 421 418 423 327 276 335 421 423 419 421 421 421

Ungoolya 1

!

Kerogen type and source potential HI values are low to moderate (16171), indicating gas prone Type III kerogen (Fig. 8.5), with poor genetic potential of no more than 0.42 kg hydrocarbon per tonne of source rock. Origin and organic petrology No organic petrology has been carried out. The Narana Formation comprises a lower canyon fill succession which is organically lean and an upper limestonemudstone succession which contains the richest source rocks. The latter are presumed to derive from algalcyanobacterial precursors.

Ouldburra Formation
Source richness More than 260 TOC measurements have been made on samples from the Ouldburra Formation, including Marla 1A, 1B, 3, 6, 7 and Manya 3, 6 in the Marla Overthrust Zone and Manya Trough, and Wallira West 1, Wilkinson 1 and Karari 1, 2A in the Tallaringa Trough. The average TOC of all samples is 0.29% but may be much higher (up to a very good 4.56% in Marla 1A) in thin, organic-rich layers that are only reliably detected by close sample spacing. A profile of part of the Ouldburra Formation in Manya 6 (Fig. 8.6) shows the downhole variation in TOC (Kamali, 1995a). Between 680 and 850 m depth there are frequent high TOC spikes against a low background, while from 1050 to 1450 m, spikes are more sporadic but the background TOC level rises to 0.3%.

Fluorescing Gloeocapsamorpha prisca and lamalginite in the Ouldburra Formation at 263.35 m in Karari 1, Tallaringa Trough (from Kamali, 1995a). (Photo 44404)

Fig. 8.5 HI versus Tmax plot, Narana Formation.

Fig. 8.6 Manya 6 TOC and copper profile (after Kamali, 1995b).

"

Kamali (1995a) related the organic matter distribution in Manya 6 to five organic matter preservation cycles separated by exposure surfaces. In the two lower cycles (10351500 m), organic matter is in laminae of algal cyanobacterial origin interbedded with halite which indicates a hypersaline environment. Organic matter in the upper three cycles (6351035 m) is associated with limestone or dolomite interbedded with anhydrite and thin silty layers. The high frequency of organic-rich versus organic-poor alternations is suggested by Kamali (1995a) to stem from periodic flooding events and lowstands. This cyclicity is best developed in transgressive and highstand tracts of sequence C1.2 (703849 m; Gravestock and Hibburt, 1991), which may correlate with a global relative sea-level maximum in Early Cambrian (Botomian) time. Rock-Eval data for the Ouldburra Formation are summarised in Table 8.6.
Table 8.6 Rock-Eval source rock data, Ouldburra Formation.
Well Manya Trough Manya 3 Depth (m KB) 246.7 360.4 418.0 657.9 691.7 698.0 698.6 732.7 773.8 780.0 785.8 807.9 810.5 826.2 899.8 999.4 1057.0 1127.0 1229.7 1231.9 1247.9 1267.5 1277.1 1279.1 1279.8 1302.1 1312.7 1332.3 1386.6 1392.4 1419.3 1459.2 136.7 138.6 416.0 671.2 700.2 392.8 263.3 275.4 211.8 285.5 298.0 344.4 344.6 462.1 480.9 TOC (%) 0.37 0.45 0.37 0.36 0.97 0.68 0.82 0.93 0.34 0.48 0.34 0.63 0.43 0.69 0.91 0.34 0.31 0.63 0.64 0.22 0.32 0.43 0.42 0.52 0.64 0.38 0.61 0.33 0.37 0.31 0.36 0.32 0.64 0.48 1.34 1.13 0.47 0.53 1.18 0.20 0.34 0.73 0.59 0.72 0.78 0.58 1.10

Kerogen type and source potential There is a problem with Rock-Eval pyrolysis data for samples from the Manya Trough. The very low HI (091) suggests the presence of poor quality gas prone Type III or IV kerogen and an associated mineral matrix effect in argillaceous samples, even though their n-alkane profiles indicate an algal source (McKirdy et al., 1984). These authors also pointed out the contrast between organic-rich laminae (TOC = 4.56%) and the dolomitic host rock (TOC = 0.18%). Ill-defined S2 peaks also result in anomalously low Tmax values, and oxygen indices (OI) are variable. Samples from the Ouldburra Formation in the Tallaringa Trough are in complete contrast, even though the organic mattercarbonateevaporite association is the same. This contrast is clearly demonstrated on Figure 8.7 which shows

Genetic potential (kg/t) 0.02 0.02 0.02 0.03 0.75 0.57 0.81 0.13 0.04 0.10 0.04 0.10 0.04 0.09 0.07 0.01 0.00 0.01 0.03 0.01 0.00 0.00 0.21 0.15 0.01 0.00 0.00 0.03 0.04 0.01 0.17 0.08 0.15 0.13 1.44 0.27 0.08 0.43 5.18 1.27 1.12 3.69 2.90 1.91 1.64 1.74 5.35

Oxygen Index 75 75 64 139 69 56 52 41 126 48 115 11 49 57 11 27 39 24 30 49 44 86 214 133 92 121 80 312 86 155 86 88 87 137 30 15 87 47 455 150 91 652 567 79 69 63 50

Hydrogen Index 2 2 2 0 41 38 61 9 0 6 0 5 2 6 3 0 0 0 2 0 0 0 36 12 0 0 0 0 3 0 31 19 12 10 91 11 0 60 388 510 285 438 372 243 193 227 409

Tmax (C) 419 440 221 249 307 436 476 359 279 333 279 317 279 317 279 279 279 279 241 266 241 266 466 304 241 205 221 281 241 232 241 241 440 432 422 341 237 403 427 435 430 427 425 422 424 431 433

Manya 6

Marla 1B Marla 6

Marla 7 Tallaringa Trough Karari 1 Karari 2A

Wilkinson 1

#

the consistently high HI of samples from the Tallaringa Trough. Kamali (1995a) ascribed this contrast in the organic matter to a lower degree of oxidation, less bacterial degradation and less compaction compared to the Manya Trough. Michaelsen et al. (1995) suggested that the Ouldburra Formation in the Manya Trough was deposited under dysoxic conditions (pristanephytane ratio = 1.02.6), compared to anoxic conditions in the Tallaringa Trough (pristanephytane ratio 1.2). It is here suggested that these differences are due to the over maturity of the Manya Trough samples (supported by burial history modelling and MPI maturity data, see Chapter 9), and that in other areas not presently sampled, suitable Ouldburra Formation facies (transgressive tracts, evaporite associations) are likely to have fair to good source potential. Genetic potential of the Tallaringa Trough samples is moderately good, with values up to 5.18 kg hydrocarbon per tonne of source rock. Origin and organic petrology Examination of the base-metal assays routinely carried out on Comalco cores from mineral drillholes (Brewer, 1984) showed that copper and zinc display elevated values at depths comparable with those corresponding to some of the organically rich layers. A profile for copper illustrates this effect and suggests reducing conditions in the depositional environment (Fig. 8.6). Elevated base-metal values are commonly associated with organically rich sediments including sabkhas rich in cyanobacterial mats (Demaison and Moore, 1980; Ferguson and Skyring, 1995).

New molecular biomarkers found in the Tallaringa Trough include 24-Isopropylcholestanes, possibly from fossil sponges (McCaffrey et al., 1993), and dinosterane, a dinoflagellate indicator (Michaelsen et al., 1995). Of perhaps the greatest importance is the discovery of telalginite composed of the cyanophyte G. prisca in the Early Cambrian Ouldburra Formation (Kamali, 1995a,b; Michaelsen et al., 1995). This organism had hitherto been identified only in Ordovician oil-prone source rocks (e.g. Summons and Powell, 1991). In this context it is interesting to note the observations of Dow (in Womer et al., 1987) on a whole rock extract from the Observatory Hill Formation in Ungoolya 1. Dow stated:
The whole extract gas chromatogram reveals a distinct odd carbon predominance in the C 14-C20 carbon number range .... ascribed ... to a particular organism Gloeocapsamorpha prisca. The high pristane and phytane content of the subject sample (from 1208.3 m) and the numerous extraneous peaks, however, are uncharacteristic of Ordovician oils and may be due to contamination, possibly from the drilling fluid.

As an alternative to contamination, this extract is suggested to represent oil which has migrated from the Ouldburra Formation and may have mixed with oil from a Neoproterozoic source sharing the same migration pathway. This implies that the Ouldburra Formation occurs downdip of Ungoolya 1, as indicated by Moussavi-Harami and Gravestock (1995) on their restored isopach map.

Observatory Hill Formation


Source richness Nearly 150 samples have been analysed from the Observatory Hill Formation from Byilkaoora 1, 2, 3, Emu 1, Giles 1, Karlaya 1, Munyarai 1, Murnaroo 1 and Ungoolya 1. The TOC of the Observatory Hill Formation is higher in the Marla Overthrust Zone ( average 0.42%, maximum 2.29%) than on the Murnaroo Platform (average 0.20%, maximum 0.49%) where the alkaline playa facies was either not well developed or deep weathering has leached the shallow to exposed beds. There is no record of the Observatory Hill Formation in the Tallaringa Trough; strata originally ascribed to this formation in wells such as Wallira West 1, Wilkinson 1, and Karari 1 and 2A have since been assigned to the Ouldburra Formation (e.g. Stainton et al., 1988). A TOC profile of the Observatory Hill Formation in Byilkaoora 1 is shown on Figure 8.8. Most of the elevated values are from the Moyles Chert Marker bed and Parakeelya Alkali Member, although values are variable due to staining by migrated oil (McKirdy et al., 1984). Copper assays are elevated and these also partly correspond to the organic-rich levels (data in Brewer, 1984). Rock-Eval data for the Observatory Hill Formation are summarised in Table 8.7. Kerogen type and source potential A plot of HI versus Tmax (Fig. 8.9) shows that the Observatory Hill Formation has moderate to good oil source potential, and suggests Type II kerogen, but high H/C and C/N atomic ratios are consistent with an origin from Type I kerogen (McKirdy and Kantsler, 1980; McKirdy et al., 1983).
$

Fig. 8.7 HI versus Tmax plot, Ouldburra Formation.

bleeding from vugs in core from Byilkaoora 1) confirms the oil potential of this formation. Origin and organic petrology The discovery of oil bleeding from vugs and fractures in core from Byilkaoora 1 first drew attention to the hydrocarbon potential of the Officer Basin. The oil was generated in situ from halotolerant algae and bacteria in the non-marine playa lake facies of the Parakeelya Alkali Member of the Observatory Hill Formation (McKirdy and Kantsler, 1980; Pitt et al., 1980; Brewer et al., 1987), an argillaceous dolomitic mudrock with pseudomorphs of the evaporites trona and shortite. Organic petrographic examination has revealed abundant lamalginite which forms a web of organic matter such as might be expected from an algal mat (McKirdy and Kantsler, 1980, p.84).
Fig. 8.8 Byilkaoora 1 TOC and copper profile.

Genetic potential is best in the Marla Overthrust Zone with potential yield up to 7.34 kg hydrocarbon per tonne of source rock, and the abundant oil shows (including oil

The importance of cyanobacterial mats as oil-prone kerogen precursors has been discussed by Zhmur et al. (1994) for settings ranging from lagoonal to marine. Zhmur and colleagues from the Paleontological Institute, Russian Academy of Sciences, are currently studying cyanobacterial mats from the Officer Basin, including those from the

Table 8.7 Rock-Eval source rock data, Observatory Hill Formation.


Well Byilkaoora 1 Depth (m KB) 200.7 202.3 204.3 254.9 259.4 286.5 660.9 670.0 677.4 680.5 688.5 699.6 703.3 705.7 711.9 720.1 720.3 727.4 728.0 728.5 729.1 729.7 738.0 343.0 346.2 346.8 348.4 365.2 368.7 373.2 212.6 213.3 224.4 Munyarai 1 Ungoolya 1 1692.9 1203.6 1203.8 1207.5 1243.8 1245.6 TOC (%) 0.76 0.67 0.58 0.83 0.68 0.58 0.44 0.50 0.53 0.77 0.39 2.29 0.20 0.13 0.62 0.28 0.39 0.31 0.39 0.34 0.68 0.27 0.43 0.73 0.61 0.56 0.35 0.55 0.50 0.81 0.47 0.42 0.56 0.12 0.43 0.50 0.26 0.21 0.27 Genetic potential (kg/t) 1.79 2.43 1.80 2.88 2.51 1.90 0.65 1.41 2.16 3.43 1.22 7.34 0.37 0.02 1.99 0.62 1.30 1.05 1.80 1.00 2.64 0.79 1.41 3.59 2.66 2.47 1.09 3.64 1.98 3.69 0.50 0.54 0.99 0.02 1.51 1.48 0.74 0.07 0.14 Oxygen Index 28 35 50 27 52 55 77 100 94 42 64 27 120 146 46 86 146 139 110 91 57 156 105 18 18 54 123 13 80 86 57 40 14 200 33 50 50 33 170 Hydrogen Index 211 298 281 310 301 260 134 240 362 412 254 257 160 15 283 200 303 284 395 262 350 241 263 386 364 355 289 600 370 417 98 124 173 0 333 276 262 29 44 Tmax (C) 428 427 420 425 420 419 434 417 416 422 415 439 420 404 416 418 423 410 415 414 412 415 431 423 413 414 416 426 420 423 424 426 430 433 422 418 426 432 345

Byilkaoora 2

Byilkaoora 3

Giles 1

%

Fig. 8.9 HI versus Tmax plot, Observatory Hill Formation.

alkaline playa facies of the Observatory Hill Formation. Results of this study will be available in 1998.

Apamurra Formation
Four samples from Byilkaoora 1 were analysed for TOC, and results ranged from 0.33 to 0.64%. No Rock-Eval analyses have been performed. The unit is widespread over the eastern Officer Basin and is of shallow marine origin; it warrants further analysis to determine the source potential.

Indulkana Shale
No source rock analyses have been performed on the Ordovician Indulkana Shale because the unit has only been intersected in the subsurface in a few wells and has not been cored. The formation may have source potential as it is a probable correlative of the marine, organic-rich Horn Valley Siltstone in the Amadeus Basin. A distinctive aeromagnetic signature is associated with the Horn Valley Siltstone (Hamer, 1994) which may be caused by an oxidised pyrite-rich layer indicative of formerly reducing conditions.

&

INTRODUCTION

Observatory Hill Formation


Oil in the Early Cambrian Observatory Hill Formation has been described in detail by McKirdy and Kantsler (1980) and McKirdy et al. (1983, 1984), and a review of all eastern Officer Basin oils is provided in McKirdy (1993). The Observatory Hill Formation oil is mainly in the Parakeelya Alkali Member and Moyles Chert Marker Bed in the Marla Overthrust Zone; it is non-marine, algal-sourced and partly biodegraded. In Byilkaoora 1, the Observatory Hill source rock maturity, expressed as calculated vitrinite reflectance (VRcalc) from the methylphenanthrene index (MPI), is VRcalc = 0.991.09%, which is at or just past peak maturity for Type I kerogen (McKirdy and Michaelsen, 1994). However, sterane and hopane distributions in Byilkaoora 1 oil suggest expulsion from a marginally mature source, and VRcalc of the oil is 0.49% (McKirdy et al., 1984; McKirdy, 1993). The Byilkaoora 2 mud pit oil is biodegraded as is the EOM in the host rock, and shows evidence of short distance migration (Woodhouse in Weste, 1984). There is thus sufficient information to show that present day maturity of the Observatory Hill Formation in the Marla Overthrust Zone is in the middle oil window. However, the indigenous oil was generated at a much lower maturity level and has migrated short distances within the formation. This can only have been achieved via a fracture network. In Byilkaoora 1 and 2, bitumen is also present in the conformably underlying Wallatinna Member. Organic matter was extracted from the Observatory Hill Formation in stratigraphic well Observatory Hill 1, drilled at the type section on the Murnaroo Platform. The aromatic maturity of the sample yielded a VRcalc of 0.94% (McKirdy et al. in Gatehouse and Hibburt, 1987). The maturity of this extract is comparable to that of the source rocks in the Marla Overthrust Zone and suggests that most of the Observatory Hill Formation is now at peak oil generation. The identification of ?G. prisca in Ungoolya 1 by Dow (in Womer et al., 1987) has been mentioned in Chapter 8. The organism was identified by its C14C20 odd carbon number predominance. G. prisca, or a precursor, has also been identified from petrographic examination of telalginite from the Early Cambrian Ouldburra Formation (Kamali, 1995a). Assuming the Ouldburra Formation to be the only G. prisca rich source rock in the eastern Officer Basin, the Ungoolya 1 occurrence is particularly significant because the extract was taken from the Observatory Hill Formation in an interval (12011215 m) with poor shows in vugs and very small vertical fractures in chert laminae displaying light brown to
'

Excitement followed the discovery of oil in fractures and vugs in core from Byilkaoora 1 (Benbow and Pitt, 1979; Pitt et al., 1980; McKirdy and Kantsler, 1980), but it was the presence of calcite pseudomorphs of trona and shortite that led initially to extensive mineral exploration in the northeastern Officer Basin. Comalco drilled 20 cored holes in the Marla area, nine of which displayed oil shows. Most of these were from the Observatory Hill Formation but oil was also found bleeding from calcite veins in the upper Rodda beds (?Tanana Formation) in Marla 9, and oil was recovered from the mud pit in Byilkaoora 2. A review of oil shows from mineral wells is provided by Hibburt et al. (1995). Woodhouse (in Weste, 1984) suggested the mud pit oil, though similar to extracts from oil-impregnated core, had migrated much further. These are clear, macroscopic indications of oil generation and migration in Neoproterozoic and Cambrian rocks. There are also microscopic indications of oil migration from both Cambrian and Neoproterozoic strata. These are reviewed below with an assessment of the thermal maturity of source rocks and the timing of hydrocarbon migration.
OIL-SOURCE CORRELATION AND MATURITY

The thermal maturity of hydrocarbon source rocks is generally assessed from reflectance measurements of vitrinite phytoclasts. However, in pre-Devonian rocks devoid of higher plant material, the distribution of triaromatic hydrocarbons can be exploited as a measure of the maturity of oils and source rock extracts. The aromatic maturity of Officer Basin source rocks has been reviewed by McKirdy and Michaelsen (1994), and available data are presented in Table 9.1. Oil shows and oil-source correlations are discussed briefly below, from the stratigraphically youngest to the oldest occurrences.

Trainor Hill Sandstone


The youngest Extractable Organic Matter (EOM) is in the Trainor Hill Sandstone, only metres below its eroded top in Marla 10. Cryptomelane (KMn8O16) was identified as the source of a black stain in the sandstone. A trace of organic matter was extracted but has not been identified (Watson, 1994a). The EOMcryptomelane association suggests secondary hydrocarbon migration due to hot groundwater movement and may be a Tertiary to Recent (500 Ma) phenomenon (see Apatite fission track analysis).

Table 9.1 Aromatic maturity data for the Officer Basin (after McKirdy and Michaelsen ,1994; Kamali, 1995a).
Formation Well depth (m) Sample type MPI MPR VRcalc % (a) VRcalc % (a1) VRcalc % (b)

Devonian Mimili Formation Cambrian Observatory Hill Formation

Munyarai 1 633.00 Byilkaoora 1 262.95 285.00 293.17 Observatory Hill 1 34.14

extract

1.43

2.78

1.26

1.22

0.98

extract extract oil extract extract* extract extract* extract extract extract extract* extract extract extract extract extract extract* oil

0.98 1.16 0.38 0.90 0.58 0.83 0.52 0.38 0.46 0.30 0.34 1.00 1.43 2.13 1.30 1.11 0.85 1.32

1.15 0.87 0.25 0.88 0.64 0.45 0.55 0.46 0.60 0.47 0.38 1.24 1.76 4.98 1.92 2.34 1.33 1.46

0.99 1.09 0.63 0.94 0.75 0.90 0.71 0.63 0.68 0.58 0.60 1.00 1.26 1.68 1.18 1.07 0.91 1.19

0.91 1.03 0.49 0.85 0.63 0.80 0.58 0.49 0.46 0.92 1.22 1.71 1.13 1.00 0.82 1.15

1.00 0.88 0.34 0.89 0.75 0.60 0.68 0.60 0.72 0.62 0.52 1.03 1.18 1.63 1.22 1.31 1.06 1.10

Ouldburra Formation (Tallaringa Trough)

Wilkinson 1 333.07 390.08 461.87 Karari 1 263.35 Karari 2A 285.50 298.0 298.13

(Manya Trough) Marla 3 619.60

Marla 6 416.0 671.25 Marla 7 392.85 Manya 6 698.60 1279.15


Relief Sandstone

Observatory Hill 1 155.35

Neoproterozoic Tanana Munyarai 1 Formation 2289.81 Marla 9 209.75 234.83 269.93


Karlaya Limestone

extract extract extract extract extract extract extract extract extract extract extract oil oil extract

1.63 0.66 0.68 0.71 1.67 0.64 0.70 0.66 0.37 0.28 0.29 1.03 1.40 0.43

0.90 0.73 0.94 0.85 1.0 0.75 0.90 0.63 0.82 1.27

1.38 0.80 0.81 0.83 1.40 0.78 0.82 0.79 0.62 0.57 0.57 1.02 1.24 0.66

1.36 0.70 1.39 0.94 1.20 0.52

0.89 0.81 0.92 0.87 0.94 0.82 0.89 0.74 0.86 1.04

Munyarai 1 2611.83 Murnaroo 1 183.90 190.95

Dey Dey Mudstone

Karlaya 1 2093.73 2345.15 Lake Maurice West 1 417.7 418.20

Murnaroo Formation

Lake Maurice West 1 534.14 Lake Maurice East 1 540.83

Alinya Formation

Giles 1 1237

* stained by oil (a) VRcalc = 0.60 MPI+0.40 (0.65 VR 1.35%) VRcalc = -0.60 MPI+2.30 (VR >1.35%) (a1) VRcalc = 0.7 MPI + 0.22 (0.5 VR 1.7%) (b) VRcalc = 0.99 log10 MPR + 0.94 (0.4 VR 1.7%)



Oil bleeds from the Observatory Hill Formation in Byilkaoora 1; core width is 41 mm: (a) alkaline playa sequence, 219.75 m (Photo 44388) (b) 277.0 m (Photo 44390) (c) 293.0 m (Photo 44391) (d) 295.5 m (Photo 44392) (e) 278.9 m (Photo T15624) (f) 219.45 m (Photo 44389).



yellow-gold fluorescence, pale straw cut and slow streaming bright yellow cut (Henry, 1986). Thus at Ungoolya 1 on the edge of the Murnaroo Platform, the Observatory Hill Formation contains traces of live oil, some of which has migrated from the Early Cambrian Ouldburra Formation, or some other G. prisca-bearing (?Ordovician) source rock.

direction into the Manya and Munyarai Troughs (Fig. 9.1). New seismic data indicate that the Cambrian is eroded on the southern Murnaroo Platform (Lindsay, 1995; Lindsay and Leven, 1996), and the formations intersected in the Hughes wells are Neoproterozoic.

Ouldburra Formation
No shows have been recorded from the Ouldburra Formation during drilling operations. Kamali (1995b) reported live oil from Marla 6 and Manya 6 in the Manya Trough, and Karari 1 and 2A in the Tallaringa Trough. These are microscale shows from organic petrographic studies. Dead oil and bitumen have been recorded in Marla 3 (Hibburt et al., 1995) and illustrated by Kamali et al. (1995) who noted that such material indicates oil migration paths. Calculated vitrinite reflectances show that the Ouldburra Formation in the Manya Trough is late mature to overmature and generally gas-generative (VRcalc = 1.01.68%; McKirdy and Michaelsen, 1994). However, one sample from 1279 m in Manya 6 is stained by migrated hydrocarbons and has an MPI-derived VRcalc of 0.820.91% (Kamali et al., 1993; Kamali, 1995b; Table 9.1). Interestingly, the stain is in limestone ~4 m above the top salt bed in this well, thus the oil can only have migrated laterally. MPI-derived maturities of Ouldburra Formation samples from the Wilkinson and Karari wells in the Tallaringa Trough are significantly lower (VRcalc = 0.580.68%, Table 9.1) and the unit is mature for oil generation. As a result of their lower maturity, organic macerals are more readily distinguished, in particular lamalginite and talalginite, which display evidence of active oil expulsion into microfractures and veinlets (Kamali, 1995a,b). These samples are from present depths <300 m, whereas samples from the Ouldburra Formation in Wilkinson 1 (65 km northeast) at similar depths are slightly more mature (VRcalc = 0.710.90%; Table 9.1). McKirdy and Michaelsen (1994) have suggested that regional maturity of the Ouldburra Formation increases southwest towards Hughes 2 on the Murnaroo Platform (VRcalc = 0.99% at 243 m in Hughes 2). However, it is more likely that the maturity of the Ouldburra Formation increases in the opposite direction in the Tallaringa Trough (i.e. towards the northeast), and also increases in a generally northwards

Relief Sandstone
No shows have been recorded while drilling the Early Cambrian Relief Sandstone. Giles 1, when drilled, was thought to be located on an anticline with the Relief Sandstone inside closure. Re-mapping by Mackie (1994) suggested, however, that the southeastern flank of the Giles structure is faulted and the anticline is open to the east. Consequently there have been no valid structural tests of the Relief or any other sandstone reservoir under seal in the eastern Officer Basin. Bitumen was identified in thin section from the Relief Sandstone in Observatory Hill 1 (Gatehouse and Hibburt, 1987), with VRcalc from an oil extract of 1.15% (McKirdy and Watson, 1989). This occurrence is highly significant. The oil is in a basal Cambrian sandstone but molecular biomarkers identify the oil with a specific Proterozoic source, namely the Alinya Formation (see below). The oil is clearly a product of post-Petermann Ranges Orogeny hydrocarbon migration from a source rock at peak levels of oil generation.

Cambrian maturity mapping


A preliminary iso-reflectance map of Cambrian aromatic maturity is shown on Figure 9.1. Potential source rocks in the Ouldburra and Observatory Hill Formations are restricted to the eastern part of the basin. Based on available data, two areas, the Munyarai and Manya Troughs, are overmature and are therefore in the gas zone. A region south of the Watson High on the Murnaroo Platform is designated as immature and seismic data show that Cambrian strata have not been preserved to the southwest (Lindsay, 1995). Sparse data suggest that the Birksgate Sub-basin, west of the Nurrai Ridge, is relatively immature (McKirdy and Kantsler, 1980, p.83, fig. 15). The bulk of the eastern Officer Basin ranges from initially mature in the southern Tallaringa Trough to mature on the southwestern margin of the Munyarai Trough. Although there is not much evidence, the region between Emu 1 and

Oil staining in Relief Sandstone at 161.5 m in Observatory Hill 1: Left: UV fluorescence (Photo 44386) Right: transmitted white light (Photo 44387). Field of view for both images is 0.5 mm.



Giles 1 appears more mature than the Ammaroodinna Ridge and its extension into the Marla Overthrust Zone. Faults in the latter region were reactivated by the Devonian Alice Springs Orogeny, whereas the Giles structure is a product of the older Petermann Ranges Orogeny. Whether this maturity contrast is related to the age difference between these structures, or due to greater uplift and erosion of the Marla Overthrust Zone after the Alice Springs Orogeny, is not clear.
Tanana Formation, Karlaya Limestone and Dey Dey Mudstone

KarlayaDey Dey boundary, both formations are regarded as part of the same source rock package. Bleeds from small, disconnected vugs with patchy to pin point bright yellow fluorescence were recorded in Karlaya Limestone core from Karlaya 1 (Dunster, 1987b). Fluorescence was also noted in very thin sandstone beds in the upper Dey Dey; extracts from the upper Dey Dey (depth 20942345 m) gave VRcalc = 0.820.89%. In Lake Maurice West 1 (depth 418 m), the Dey Dey Mudstone yielded VRcalc = 0.740.86%, and the Karlaya Limestone in Murnaroo 1 (depth 184191 m) gave VRcalc = 0.870.94% (Table 9.1). These values indicate that the KarlayaDey Dey source rock package is early mature on the Murnaroo Platform. Maturity increases into the Munyarai Trough, since VRcalc in

Oil shows have been recorded in the Tanana Formation, Karlaya Limestone and upper Dey Dey Mudstone immediately beneath. Since most of these shows span the

Fig. 9.1 Eastern Officer Basin Cambrian aromatic maturity. Contours represent vitrinite reflectance calculated from methylphenanthrene index.

!

the Devonian Mimili Formation is 0.98% and in the Karlaya Limestone (depth 2612 m) is 1.4% in Munyarai 1 (McKirdy and Michaelsen, 1994; Table 9.1). A Proterozoic oil show was recorded while drilling Marla 9 in the Marla Overthrust Zone over an interval tentatively correlated with the Tanana Formation (depth 245287 m). The show comprises oil bleeds from fractures partly cemented by calcite. VRcalc of extracts from 210 to 270 m depth is 0.810.92%. Weak bleeds were also reported from the Dey Dey Mudstone (depth 328 m). Marla 9 is not in the Manya Trough as McKirdy and Michaelsen had assumed, hence its maturity is not anomalously low, although it is somewhat lower than nearby Cambrian values. Byilkaoora 1 is the only other well to have intersected the Neoproterozoic in the Marla Overthrust Zone; no shows were recorded in this part of the section. The KarlayaDey Dey oil is marine and its distinctive sterane distribution links it to green algal precursors. Similar sterane distributions are found in Neoproterozoic oils from Oman and Siberia (McKirdy, 1993).

Proterozoic maturity mapping


A preliminary iso-reflectance map of Neoproterozoic aromatic maturity is shown on Figure 9.2. No source rocks exist in the easternmost parts of the basin where Mesoproterozoic basement underlies Neoproterozoic or Cambrian sedimentary rocks at shallow depths. Like the Cambrian map, it is a hybrid of VRcalc values from two distinct oil families in samples taken over a wide depth range. Both maps are thus highly interpretive. Overmature, gas-prone conditions are indicated for the Manya and Munyarai Troughs and the Ammaroodinna and Middle Bore Ridges. Iso-reflectance contours are displaced to the south and west compared to Cambrian contours, thus the northern Murnaroo Platform is at a peak stage of oil generation. No maturity data exist for the Birksgate Sub-basin or Tallaringa Trough. Basement from refraction seismic investigation could be 800 m below the drilled Cambrian section in the Tallaringa Trough (Milton, 1975). The Cambrian aromatic maturity values suggest that any underlying Proterozoic source rocks are likely to be within the oil window. The Neoproterozoic samples from Marla 9 indicate that the Marla Overthrust Zone is within the oil window. Similarly, results from Hughes 2 suggest that the southwestern Murnaroo Platform is oil mature.

Murnaroo Formation
The Proterozoic Murnaroo Formation is a potential reservoir, not a source rock. As with the Cambrian Relief Sandstone, thin sections from Lake Maurice West 1 (depth 534576 m) were oil stained and small quantities of oil were subsequently extracted from drillcore. VRcalc of the EOM is 0.941.20% and it has a similar maturity and composition to residual hydrocarbon from the Relief Sandstone in Observatory Hill 1 (McKirdy and Watson, 1989). Thus, two wells 60 km apart on the Murnaroo Platform contain oil from a common source in sandstone reservoirs of Proterozoic and Early Cambrian age. The Murnaroo and Relief shows are perhaps the most significant in the Officer Basin.

THERMAL MATURITY
There are few usable bottom-hole temperature records from eastern Officer Basin wells. Drillstem test temperatures from Lairu 1 (40.6C at 1090 m) and Munta 1 (73.3C at 1981 m) indicate a low to average present-day geothermal gradient ranging from 14 to 24C/km. However, this basin has had a complex history of subsidence and uplift, and present-day geothermal gradients may not apply to conditions prior to the Neogene. To better understand the thermal history of the eastern Officer Basin, apatite fission track analysis (AFTA) and fluid inclusion geothermometry have been carried out on a small sample population. In addition, vitrinite reflectances were measured on Permian phytoclasts from several wells in order to provide some constraints on Palaeozoic thermal history. The results presented below have far-reaching consequences for geohistory modelling.

Alinya Formation
It was McKirdy (1993) who recognised that biomarker distributions (hopanesterane and diasteranesterane ratios; McKirdy and Watson, 1989) point unequivocally to the Alinya Formation as the source of the MurnarooRelief oil shows. This constitutes a second family of Neoproterozoic oil, as confirmed from extracts of the Alinya Formation in Giles 1. VRcalc from core in this well (depth 12371266 m) yields a value of 1.04% (McKirdy and Michaelsen, 1994; Table 9.1). The calculated maturity is comparable to maturities calculated for extracts from the Relief Sandstone (1.15%) and Murnaroo Formation (0.941.20%), suggesting that the Alinya Formation was the source of this oil, which must have migrated after deposition of the Relief Sandstone. Moussavi-Harami and Gravestock (1995) suggested that oil migration took place after burial beneath thick Ordovician sediments. Because the same biomarker assemblage has been found in the Bitter Springs Formation of the Amadeus Basin (Zang and McKirdy, 1994), the oil source facies was probably very widespread and was certainly far richer organically than the meagre Giles 1 data would suggest.
"

Permian vitrinite reflectance


In their search for economic coal deposits, Comalco drilled eight holes in the Permo-Carboniferous Arckaringa Basin near the northwestern margin of the Boorthanna Trough (Bourke and Senapati, 1983). Cuttings from two of these holes, 42 km southeast of Manya 2 (Fig. 9.1), were sampled for coal from the Permian Mount Toondina Formation at depths ranging from 187 to 216 m. Measured vitrinite reflectance (VRmeas), organic petrology and VRcalc from MPI were recorded (Watson, 1994b). Further measured VR and petrographic descriptions were obtained from core samples of the Mount Toondina and Boorthanna Formations (255497 m) in Manya 2, and the Cretaceous Bulldog Shale and Mount Toondina Formation (62 and 360 m, respectively) in Mount Willoughby 1 (Tingate, 1994).

In Manya 2, measured VR varies from 0.35 to 0.51%, which is inconsistent with the present depth of the samples. The degree of gelification of telovitrinite in samples from the Mount Toondina Formation suggests a cover of not less than 1.21.5 km since Permian deposition (Keiraville Konsultants in Tingate, 1994), pointing to a greater depth of burial in Mesozoic time. Measured vitrinite reflectance in the two Comalco coal drillholes ranges from 0.29 to 0.50%. In these holes and in Manya 2, the Mount Toondina Formation contains fluorescing telalginite composed of Reinschia and Botryococcus-related genera, as well as Tasmanites. Oil drops are also present. Of greatest interest, however, is the calculated vitrinite reflectance of 0.85%, corresponding to a

methylphenanthrene index of 0.75. This is considerably higher than the measured vitrinite reflectance of 0.330.50% on adjacent samples, pointing to suppression of vitrinite reflectance due to the presence of hydrogen-rich macerals (Hutton and Cook, 1980; Keiraville Konsultants in Tingate, 1994). Reflectance suppression of fluorescent lamalginite by as much as 0.40.5%, compared to VRcalc from MPI, is evident in source rocks from the Proterozoic McArthur Basin (Crick, 1992), and a similar problem is encountered in the Triassic of the North West Shelf (Beardsmore and OSullivan, 1995). As a result of this phenomenon, the Permian vitrinite reflectances should be regarded as minimum values and post-Early Permian burial may have been as high as 2 km in

Fig. 9.2 Eastern Officer Basin Neoproterozoic aromatic maturity. Contours represent vitrinite reflectance calculated from methylphenanthrene index.

#

the vicinity of these wells. It would also appear that the maturity of Permian source rocks in the Boorthanna Trough has been underestimated. A technique developed by CSIRO Division of Petroleum Resources which examines fluorescence alteration of multiple macerals (FAMM; Wilkins et al., 1992) will be applied in this region. Preliminary results indicate VR suppression of 0.1% (N. Sherwood, CSIRO, pers. comm., 1997). It is noteworthy that van Neil (1984) first documented fluorescence alteration in a sample from the Ouldburra Formation in Wilkinson 1.

window for oil generation, oil was not noted in these inclusions either, oil in vugs and fractures notwithstanding. It is possible that homogenisation temperatures in the Byilkaoora samples are related to a much younger (100 Ma) thermal event revealed by apatite fission track data. All the fluid inclusions analysed to date are from the structurally complex Marla Overthrust Zone (Byilkaoora wells) and the adjacent Manya Trough (Manya 6), and these interpretations should thus be regarded as preliminary.

Fluid inclusion microthermometry


Fluid inclusions were studied in three samples from the Ouldburra Formation in Manya 6 in the Manya Trough (Kamali, 1995b). Microthermometry was performed on inclusions from early calcite cement, late calcite cement, early dolomite and saddle dolomite in an attempt to determine palaeotemperature history. The sampled depths were 889.5, 896.5 and 956.8 m, with the deepest yielding the majority of inclusions in late calcite and saddle dolomite. Saddle dolomite from the deepest sample contained fluid inclusions with four or more homogenisation temperatures ranging from 163 to 332C, which significantly exceed the temperature of dolomite crystal formation, and thus have re-equilibrated. Such high temperatures suggest the passage of hydrothermal brines, but calculated vitrinite reflectance values of 1.07% (from 698.6 m depth) and 0.91% (from 1279.15 m; oil stained) argue against this hypothesis and also negate a widespread thermal event (Kamali, 1995b). Furthermore, salt in this well has not been mobilised, thus the high temperature data must be treated with caution. Homogenisation temperatures for early calcite (3150C) and early dolomite 6068C are consistent with the entrapment of primary fluid inclusions with increasing burial. Late calcite (homogenisation temperature 132170C) may have been precipitated at a depth exceeding 4 km from a more saline brine which is consistent with burial beneath a thick OrdovicianDevonian section prior to the Alice Springs Orogeny. Gravestock and Sansome (1994) suggested that the Relief Sandstone in Manya 6 may have been buried to a depth approaching 6 km beneath an advancing thrust front to account for the degree of compaction of the sandstone. Modelling (see below) suggests that a depth of 4.6 km would be sufficient to account for present-day maturity values. The absence of hydrocarbons from Ouldburra Formation fluid inclusions in Manya 6 does not downgrade the source potential of this unit. Instead, it points to their migration at some time between the relatively early low-temperature cements and the relatively late high-temperature cement. Two-phase fluid inclusions were also examined by the Baas Becking Geobiological Laboratory (1983) from the Parakeelya Alkali Member of the Observatory Hill Formation in the Byilkaoora wells. These inclusions are in carbonate pseudomorphs of shortite (an indicator of alkaline playa environments), and have homogenisation temperatures in the range 60108C. Even though this is in the thermal
$

Apatite fission track analysis


Apatite fission track analysis was carried out by Geotrack (1994) on sandstone samples from five wells in three structurally contrasting regions. Wells in the Marla Overthrust Zone and Manya Trough were predicted to have thermal histories dominated by the Alice Springs Orogeny, whereas an older thermal history was expected from cooler, shallower wells on the Murnaroo Platform. Sandstones in the Tallaringa Trough were not analysed. Apatite yield was excellent from all but the oldest and youngest samples. The yield from the oldest (Pindyin Sandstone in Giles 1) was very poor and the Pindyin was therefore not analysed. Yield was also poor from the Late Jurassic Algebuckina Sandstone in Manya 2, but good information was obtained nevertheless. Analytical and palaeotemperature data (Table 9.2) were interpreted by Tingate (1994) based on a geothermal gradient of 25C/km, which was assumed in the absence of measured temperature data. One unexpected outcome of this analysis was the total thermal annealing of tracks associated with the Alice Springs Orogeny, regardless of location. Maximum palaeotemperatures in excess of 110C were experienced by the Relief Sandstone and Murnaroo Formation in Lake Maurice West 1 during the Late Devonian and Carboniferous, yet this well is 200 km south of the Musgrave Block. Similarly, apatites analysed from early Palaeozoic strata beneath the Pedirka Basin, up to 160 km east of the Musgrave Block, have experienced the same effect (Tingate in Alexander et al., 1996). The uraniferous granites beneath the Cooper Basin were intruded during the Early to Middle Carboniferous and other thermal events have been recorded as far south as the Flinders Ranges (Gatehouse et al., 1995). The Alice Springs Orogeny was evidently associated with high heat flow over a wide area of northern South Australia, and this is taken into account for geohistory modelling (see below). Permian and Jurassic apatites from Manya 2, and older apatites from other wells (Table 9.2), were heated to 90C prior to cooling during the Cretaceous (11070 Ma). Tingate (1994) correctly associated this higher temperature with Late Permian to Cretaceous burial, and suggested a cover of 1.5 km would be enough to provide conditions suited to the elevated temperature. Based on measured vitrinite reflectance suppression, and assuming a lower geothermal gradient, the interpreted depth of burial is now closer to 2 km, at least for the Manya Trough and Marla Overthrust Zone. There is also consistent evidence that palaeotemperature was elevated during the last 50 million years, partly due to

Table 9.2 Sample analytical and palaeotemperature data (after Geotrack, 1994; Tingate, 1994).
Formation Well depth (m) Algebuckina Sandstone Boorthanna Formation Cadney Park Formation Cadney Park Formation Relief Sandstone Strat. age (Ma) 150130 290280 525518 Fission track age (Ma) 201.217.6 27916.2 237.216.8 Mean track length (m) 11.590.34 11.760.15 11.350.25 Standard deviation (m) 1.47 1.54 2.57 Present1 temp. (C) 31 37 38 110 90 90 90 60 60 70 Maximum palaeotemperature2 360-300 Ma 50-0 Ma 110-70 Ma

Manya 2 245.5247.7
492.9494.3 510516

Manya 6 448.2448.8
1699.61701.2

525518 540520

236.211.8 231.015.9 218.826.9

11.430.19 11.210.64

1.70 2.92

36 68

110 >110

90 90

60 70

Relief Sandstone Murnaroo Formation Tarlina Sandstone Relief Sandstone Murnaroo Formation Relief Sandstone Tanana Formation Tarlina Sandstone
1 2

Manya 5 455.1455.5
459459.6

540-520 615-600

216.614.5 192.113.1 195.620.8

11.920.15 12.220.28

1.57 2.00

36 37

110 >110

90 90

60 60

1054.51055.8

650-640

238.418.7

11.230.21

2.12

51

>110

90

70

Lake Maurice West 1 217.7218.7 540520


488.2488.8 615600

323.619.9 240.214.2

12.120.22 11.980.16

2.32 1.68

30 37

100 110

90 90

60 60

Giles 1 416.6416.9
422.3422.6

540520 585575

299.317.4 251.615.5 233.522.3

11.950.18 12.100.18

1.79 1.74

35 36

100 110

90 90

60 60

1063.41063.8

640615

148.212.6 139.622.8

11.160.19

1.87

51

>110

100

90

Calculated assuming a geothermal gradient of 25C/km. Palaeotemperature estimates have an error of approximately 10C. Underlined ages are central ages used when sample single grain age data have chi squared probabilities of <50% (see Geotrack (1994) for further details).

thicker Tertiary cover, since eroded. Tingate (1994) considered the main reason to be hot water flow. Support for the existence of heated aquifers in Tertiary time comes from the presence of abundant kaolinite in the Algebuckina Sandstone (which probably caused the poor apatite yield). The presence of Liesegang rings in Permian strata also points to circulation of warm water and suggests the aquifers were not confined solely to the Jurassic sandstones. The apatite fission track data are consistent and uniform enough to constrain palaeotemperatures in the eastern Officer Basin over the past 300 million years. However, measurement of thermal events in the preceding 800 million years of the basins history is elusive, and analysis of this period is a best guess.

a modular software program developed by Paltech Pty Ltd. Modelling was undertaken on three wells, each of which has undergone a differing tectonic history and is located in a distinct structural setting: Manya 6 (Manya Trough) Giles 1 (Ammaroodinna Ridge) Byilkaoora 1 (Marla Overthrust Zone). Considerable difficulty was experienced in constraining maturity models for all wells, with the possible exception of Manya 6. This has arisen as a result of inadequate temperature, maturity and kinetic data. Whilst estimates of the amount and timing of erosion have been addressed by Moussavi-Harami (1994) and summarised in Table 9.3, an accurate estimate of the original Rock-Eval properties for principal source rocks, in particular TOC and HI, has been hampered largely as a consequence of the complex structural history and great depths of burial.
%

GEOHISTORY MODELLING
Geohistory and maturity modelling of the eastern Officer , Basin has been undertaken using version 2.05 of Winbury

Table 9.3 Interpreted amount of sediment lost, selected wells, Officer Basin (m).
Tectonic event Manya 6 Giles 1 200 100 255 625 2400 500 Byilkaoora 1 1000 500 1735 450 1675 1000

Continental uplift 1300 Alice Springs Orogeny 2750 Rodingan Event 450 Delamerian Orogeny 1250 Petermann Ranges Orogeny 770 Sturtian uplift not reached

log. One attribute of the latter is temperature suppression through the salt zone in the lower Ouldburra Formation and upper Relief Sandstone. For the purposes of more accurately modelling the richest source rock units, the 1 km thick Ouldburra Formation has been divided into three informal units, of which the richest source quality is found in the upper unit (average TOC = 0.82%). Present day maturity zones are summarised on Figure 9.3 and Table 9.4. Formations below the Relief Sandstone (in which the well reached total depth) are reconstructions based on data from Manya 5. Source rocks of the upper and middle Ouldburra appear to have entered the oil window (VRcalc = 0.65%) at ~370 Ma, just prior to the Alice Springs Orogeny, and the wet gas window shortly thereafter (~360 Ma). They have remained in the wet gas window to the present day. Structures which were in place as a result of the Petermann Ranges and Delamerian Orogenies could be expected to be charged. By comparison, the lower Ouldburra Formation source unit entered the oil window just prior to the Delamerian Orogeny at ~510 Ma and remainded there until 460 Ma before entering the wet gas window (VRcalc = 1.0%) just prior to the Alice Springs Orogeny (~370 Ma). This unit has been in the dry gas window since the Carboniferous (~315 Ma).

Source rock packages modelled for the three wells are the Cambrian Observatory Hill and Ouldburra Formations, and the Neoproterozoic Karlaya Limestone and Dey Dey Mudstone. The Alinya Formation was not considered due to a lack of data. All Rock-Eval parameters for the Early Cambrian Ouldburra Formation, the most prospective source unit, are extrapolated from Karari 1 and 2A wells in the Tallaringa Trough, which still display active oil expulsion.

Manya Trough
The thermal and burial history of the Manya Trough is represented by Manya 6. Modelling of this well is fairly well constrained by fission track data and a detailed temperature

Fig. 9.3 Geohistory plot, Manya 6.

&

Table 9.4 Hydrocarbon maturity, Manya 6, Manya Trough.


Formation VRcalc (%) Maturity window Immature (oil) Wet gas Dry gas Depth (m subsea) Surface ~175 ~944

Table 9.5 Hydrocarbon maturity, Giles 1, Ammaroodinna Ridge.


Formation VRcalc (%) Maturity window Immature (oil) Mature (oil) Depth (m subsea) Surface ~123

Post-Devonian <0.65 Cadney Park Member, upper and middle 1.0 Ouldburra Formation Lower Ouldburra Formation, 1.6 Relief Sandstone, Murnaroo Formation*, Tarlina Sandstone* * Not penetrated.

Ammaroodinna Ridge
The thermal and burial history of the Ammaroodinna Ridge is represented by Giles 1. Modelling of this well is poorly constrained and draws heavily on heatflow constructed for Manya 6. Present-day maturity zones are summarised on Figure 9.4 and Table 9.5. The entire Precambrian succession at the Giles 1 location entered the oil window at ~570 Ma, prior to the Petermann Ranges Orogeny, and has stayed there until the present day. The Dey Dey MudstoneKarlaya Limestone source rock package is presently just within the oil window (VRcalc = 0.7%). By comparison, the Cambrian succession remains

Observatory Hill Formation, <0.65 Cadney Park Member, Relief Sandstone, 0.65 Tanana Formation, Karlaya Limestone Member, Dey Dey Mudstone, Murnaroo Formation, Meramangye Formation, Tarlina Sandstone, Alinya Formation, Pindyin Sandstone, Tarlina Sandstone

immature largely as a result of significantly reduced depositional rates on the Ammaroodinna Ridge.

Marla Overthrust Zone


The thermal and burial history of the Marla Overthrust Zone is represented by Byilkaoora 1. There is no bottom hole temperature data for this well and a regional geothermal gradient of 25C/km is assumed. Only one VRcalc value was available (1.0% at 200 m) and heatflow modelling has been drawn from Manya 6; the pre-Narana Formation stratigraphy

Fig. 9.4 Geohistory plot, Giles 1.

'

is a reconstruction below total depth. Present-day maturity zones are summarised on Figure 9.5 and Table 9.6.
Table 9.6 Hydrocarbon maturity, Byilkaoora 1, Marla Overthrust Zone.
Formation VRcalc (%) Maturity window Mature (oil) Depth (m subsea) Surface

Trainor Hill Sandstone, <0.65 Apamurra Formation, Arcoeillinna Sandstone, Observatory Hill Formation, Cadney Park Formation upper Narana Formation 1.0 lower Narana Formation, 1.6 Tanana Formation*, Karlaya Limestone*, Dey Dey Mudstone*, Murnaroo Formation*, Meramangye Formation*, Tarlina Sandstone* * Not penetrated.

The Dey Dey MudstoneKarlaya Limestone source rock package entered the oil window at ~570 Ma, and the wet gas window at ~550 Ma before passing into the dry gas window between ~490 and 475 Ma during the Delamerian Orogeny. More significantly, source rocks of the Observatory Hill Formation appear to have remained within the oil window from ~450 Ma (i.e. post Delamerian Orogeny) until the present day which may account for the highly degraded nature of the oil recovered from bleeds in the core.

Wet gas Dry gas

~100 ~700

Fig. 9.5 Geohistory plot, Byilkaoora 1.

 

INTRODUCTION
The Officer Basin contains a number of reservoirs with excellent porosity and permeability. These are generally sandstones of fluvial or aeolian origin, and were originally feldspathic, but dissolution of the feldspars has led to extensive secondary porosity development, in some cases with permeabilities over 8 darcys (8000 md). Reservoir potential also exists in carbonates (vuggy porosity). The only previous core analysis for reservoir properties carried out by industry was for Comalco in the early 1980s, which concentrated on the Ungoolya Group reservoirs (Weste, 1984), but MESA has since obtained data from all available cored sandstone reservoirs as part of the South Australian Exploration Initiative (106 samples; Sansome and Gravestock, 1993). In addition, some analytical work was initiated by university student projects (Gaughan, 1989; Kamali, 1995b).

FORMATION DESCRIPTION
Pindyin Sandstone
Distribution Pindyin Sandstone crops out in the Birksgate Sub-basin but in the subsurface it has only been intersected by Giles 1. The formation is, however, presumed to be widespread in the deeper, undrilled parts of the basin. The thickness ranges from 100 to 200 m. Petrophysics Twelve samples from Giles 1 range from 3.8 to 22.5% porosity with an average of 11.8%, and permeability values reach 1538 md. The porositylog permeability (phi log k) plot is linear (Fig. 10.1). Calculated shale volume (Vshale)

Aeolian sandstone of the Pindyin Sandstone at 1291.36 m in Giles 1. Quartz grains are evenly coated with haematite. Porosity is coloured blue; core porosity from a nearby sample was 22.6% and permeability was 1538 md. Plane polarised light; field of view is 1.6 mm. (Photo 44393)

rarely exceeds 5% and the Gamma Ray is generally between 20 and 40 API units (Fig. 10.2). Wireline log porosity is readily calculated from the density log using a quartz matrix density of 2.65 g/cm3. The Pindyin Sandstone aeolian facies is the cleanest potential reservoir in the basin. Seal Siltstone and evaporites of the Alinya Formation may act as a semi-regional seal for the Pindyin Sandstone on the northern margin of the Murnaroo Platform.

Tarlina Sandstone
Distribution Tarlina Sandstone is distributed over the Murnaroo Platform, Manya Trough and may possibly extend to the

Fig. 10.1 Porositylog permeability plot, Pindyin Sandstone; Murnaroo Platform depth range 12911326 m.

Fig. 10.2 Vshale versus Gamma Ray for Giles 1, Pindyin Sandstone.

 

Murnaroo Formation
Distribution The Murnaroo Formation is a key petroleum reservoir target. The formation is widespread on the Murnaroo Platform (maximum thickness 391 m) and extends to the Manya Trough where it was fully cored in Manya 5 (246 m thick). In Marla 9, however, the formation is represented by a thin condensed section. It underlies the Dey Dey Mudstone (Murnaroo Platform and Marla Overthrust Zone) and Cambrian Relief Sandstone (Manya Trough), and overlies the Meramangye Formation or Tarlina Sandstone.
Pindyin Sandstone aeolianite at 445.4 m in Watson Siding 1a. Plane polarised light; field of view is 3.2 mm. (Photo 44429)

Tallaringa Trough. The formation is up to 183 m thick to the south of the basin on the Murnaroo Platform, but may be up to 373 m in Manya 5. It may not be present on the northern margin of the basin due to facies changes. A section corresponding to the Murnaroo Formation in Manya 5, as originally interpreted by Gravestock and Sansome (1994), has been revised. Moussavi-Harami (1994) and Moussavi-Harami and Gravestock (1995) now consider that the Murnaroo Formation is underlain by the Tarlina Sandstone. This interpretation better fits the burial history of this well. Petrophysics Twenty-one samples from Giles 1 and Manya 5 have porosities ranging from 9 to 19.6% with low permeabilities which average 1.0 md. The porositylog permeability plot is clustered and there is no discernible difference between Murnaroo Platform and Manya Trough samples of the Tarlina Sandstone, supporting the reinterpretation of Manya 5 stratigraphy as discussed above (Fig. 10.3). Calculated Vshale in Giles 1 is mainly 1020% with few values exceeding 30%. Gamma Ray (Fig. 10.4) is generally 6090 API units in Giles 1, reflecting the feldspar content. Scattered high values to 150 API are attributed to slumped mudclast horizons. Porosity from the wireline density log (quartz matrix) correlates reasonably well with measured values (Fig. 10.5). The relatively low permeability values, despite good porosity, suggest that pore-bridging clays may be responsible. Seal Mudstone of the Meramangye Formation may act as a seal for the Tarlina Sandstone in the northern Murnaroo Platform where it reaches a thickness of 195 m in Giles 1. The Meramangye Formation may disappear towards the basin margins, resulting in stacked reservoirs. The Tarlina Sandstone is overlain by Murnaroo Formation in Lake Maurice East (southern Murnaroo Platform) and Manya 5 (Manya Trough).


Fig. 10.3 Porositylog permeability plot, Tarlina Sandstone; Murnaroo Platform depth range 10641230 m; Manya Trough depth range 7151053 m.

Fig. 10.4 Vshale versus Gamma Ray for Giles 1, Tarlina Sandstone.

Fig. 10.5 Core porosity versus porosity calculated from the density log, Tarlina Sandstone.

Petrophysics The reservoir quality is variable, ranging from 3 to 20% in the three sampled wells (Giles 1, Munta 1, Manya 5; n = 28). The porositylog permeability plot shows that permeability can reach 200 md and is generally greater than 1 md (Fig. 10.6). Slightly higher porosity but lower permeability values in Manya 5 near the Marla Overthrust Zone may be related to early cementation by illitic clay prior to burial. On average, however, there is no great difference in porosity distribution between Manya 5 (Manya Trough) and the other two wells (Murnaroo Platform; Fig. 10.7). In Giles 1, the Gamma Ray varies from 20 to 200 API probably due to the feldspar, mica, heavy mineral and glauconite composition. Vshale ranges up to 40% but is predominantly 20% or less (Fig. 10.8). Water saturation Three samples from Manya 5 were submitted for airmercury capillary pressure curve analysis. Two curves from the Murnaroo are very similar, whilst the one Tarlina Sandstone sample indicates that a higher injection pressure is required for the same saturation. Calculations from capillary pressure data (assuming a 100 000 ppm brine) indicate that, for the Murnaroo Formation, a 120 m vertical height above an oilwater contact yields an irreducible water saturation (Swirr) of 22%. In contrast, the Tarlina Sandstone would yield an Swirr of 43% under similar conditions. This

result is consistent with the low permeability measurements for the Tarlina (Fig. 10.3). Seal On the Murnaroo Platform and in the Marla Overthrust Zone, the Murnaroo Formation is sealed by the Dey Dey Mudstone. Thickness of the Dey Dey Mudstone ranges from 86 m to possibly 900 m, in the Munyarai Trough. In the Manya Trough, seals are absent, resulting in stacked reservoirs, with the Cambrian Relief Sandstone overlying the Murnaroo Formation.

Relief Sandstone
Distribution The Relief Sandstone has been intersected on the Murnaroo Platform and in the Manya Trough (thickness ~100 m). It is overlain conformably by, and intertongues with, the Ouldburra Formation (transitional, with halite interbeds), and is overlain conformably to disconformably by the Observatory Hill Formation. In Manya 5, however, the Relief Sandstone is overlain by the Ordovician Mount Chandler Sandstone due to Delamerian erosion. Petrophysics Porosity is secondary; Gaughan and Warren (1990) cited one sample from Observatory Hill 1 with a permeability of 4839 md. Gravestock and Sansome (1994) reported a sample from Giles 1 with a permeability of 8033 md. Five samples from Giles 1 (40 m interval) exceed 1400 md and five samples from Meramangye 1 (17 m interval) are in the 26076297 md range. In contrast, the Relief Sandstone in Manya 6 rarely exceeds 0.08 md (Gravestock and Sansome, 1994). The porosity histogram shown on Figure 10.9 (three wells from the Marla Overthrust Zone and Manya Trough, n =19; two wells from the Murnaroo Platform, n =25) clearly illustrates the bimodal porosity distribution pointed out by Gaughan and Warren (1990). The porositylog permeability plot illustrates the marked difference between the two regions (Fig. 10.10). Two trends are evident one related to high secondary porosity and low compaction in the Marla Overthrust Zone and Manya Trough, the other related to low secondary porosity and high compaction on the Murnaroo

Fig. 10.6 Porositylog permeability plot, Murnaroo Formation; Murnaroo Platform depth range 5931985 m; Manya Trough depth range 458686 m.

Fig. 10.7 Porosity histogram of pore distribution, Murnaroo Formation.

Fig. 10.8 Vshale versus Gamma Ray for Giles 1, Murnaroo Formation.

 !

Platform and related to Carboniferous depth of burial (Gravestock and Sansome, 1994). The Gamma Ray (Giles 1) ranges from 20 to 80 API with scattered higher readings to 200 API. Vshale remains generally between 10 and 30% (Fig. 10.11). Vshale from neutron logs run in Marla, Manya and Byilkaoora wells are unreliable because of uncalibrated substandard readings. Core to log correlation is poor (Fig. 10.12). The Relief Sandstone reservoir quality is superb on the Murnaroo Platform due to high dissolution and low compaction effects. Porosity in the Marla Overthrust Zone

and Manya Trough should not be written off as it is only poor in deeply buried footwall situations. In hanging wall structures, porosity reaches 13% and permeability reaches 124 md. In Manya 5 (Manya Trough), porosity reaches 10.9% with a permeability of 0.44 md. Commercial reservoirs may be found in all areas (Gravestock and Sansome, 1994). Seal The Ouldburra and Observatory Hill Formations may act as seals for the Relief Sandstone. The Relief Sandstone intertongues with the Ouldburra Formation, possibly due to relative sea-level changes. At low relative sea level, the Relief Sandstone progressed basinwards over the Ouldburra

Fig. 10.9 Porosity histogram of pore distribution, Relief Sandstone.

Relief Sandstone at 416 m in Emu 1.

(Photo 44384)

Fig. 10.10 Porositylog permeability plot, Relief Sandstone; Murnaroo Platform depth range 304447 m; Marla Overthrust Zone and Manya Trough depth range 4051755 m. Fig. 10.11 Vshale versus Gamma Ray for Giles 1, Relief Sandstone.

Relief Sandstone at 178.9 m in Observatory Hill 1. Porosity, coloured blue, is ~20%. Plane polarised light; field of view is 6.8 mm.
(Photo 44385)

Fig. 10.12 Core porosity versus porosity calculated from the density log, Relief Sandstone.

124

Formation, while high relative sea level led to flooding of the hinterland, causing carbonate buildup (Ouldburra Formation) thus forming a seal to the Relief Sandstone (Gravestock and Hibburt, 1991). In the Marla Overthrust Zone, the Cadney Park Member of the Observatory Hill Formation might provide a seal to the Relief Sandstone. Ouldburra Formation Distribution The Ouldburra Formation has been intersected in the Marla Overthrust Zone and Manya and Tallaringa Troughs. The maximum thickness is in the Manya Trough (987 m in Manya 6). However, the Ouldburra is absent on the eastern margin of the Manya Trough in Manya 5. The maximum thickness in the Tallaringa Trough is 486 m in Wilkinson 1. Magnetic data suggest that the formation in the Tallaringa Trough thickens towards the Karari Fault, which is interpreted as a reverse or thrust fault (Milton, 1974). The Ouldburra is predominantly carbonate but it does contain sandstone reservoir potential. As an example, in Manya 3 (Middle Bore Ridge) there are 30 stacked sands (Dunster, 1987a) with an average thickness of 3.8 m comprising 100 m of sandstone in 639 m of section (16%). The average separation between the stacked sands is 17.8 m (Gravestock and Hibburt, 1991). The Ouldburra Formation is overlain conformably by the Observatory Hill Formation. It overlies and intertongues with the Relief Sandstone (Gravestock and Hibburt, 1991). In Wilkinson 1, the Ouldburra is underlain by Relief Sandstone and unconformably overlain by Permian Stuart Range Formation. Hence, the true thickness of the Ouldburra is unknown in the Tallaringa Trough.

Seal Halite is an effective stratigraphic seal across the OuldburraRelief interface, while intraformational seals are provided by micritic carbonates. In Manya 3, for example, these carbonates average 17.8 m in thickness, while the thinnest carbonate beds are 1 m thick. Locally, thin breccia beds may lower the seal efficiency of some carbonates but are not considered a major risk. Arcoeillinna Sandstone Distribution The Arcoeillinna Sandstone is a southwest-thickening unit (60172 m) which extends through the Manya and Munyarai Troughs. It reaches a maximum thickness in Munta 1 on the Murnaroo Platform. The Arcoeillinna Sandstone occurs between the Observatory Hill Formation (source rock) and the dolomitic mudstone seal of the Apamurra Formation. Petrophysics As with the older Relief Sandstone, the porosity distribution between the Marla Overthrust ZoneManya Trough areas and the Murnaroo Platform is bimodal, the latter having a very high average porosity value of 21%. However, the Arcoeillinna in the Marla Overthrust Zone and Manya Trough has quite high porosity values averaging 13.7% and ranging up to 19% (Fig. 10.14).

Petrophysics Reservoir quality in the Ouldburra Formation is variable with porosity ranging from 3 to 27% and permeability ranging from 0.005 to 1640 md. Dolomitisation has resulted in substantial secondary porosity in the carbonates (Kamali et al., 1995). Carbonate reservoir porosity averages 15.0% while the clastic reservoirs average 13.8%. The porositylog permeability plot has a poor correlation, with significant scatter for carbonate reservoirs, whilst the sandstones are in good agreement and plot on a semi-log trend with minor scatter (Fig. 10.13). In this respect, the intra-Ouldburra sandstones behave like other siliciclastic reservoirs in the Officer Basin. Kamali (1995b) identified that the better carbonate reservoirs are composed of sucrosic dolomite with intercrystalline porosity. Subaerial exposure in the early stages of diagenesis of the carbonates has largely contributed to the Ouldburras good reservoir quality. The extent and distribution of these higher quality carbonate reservoirs is yet to be determined. While no samples have been taken from Wilkinson 1, a 3 m sandstone described as very porous and vuggy was intersected at a depth of 460 m. Another ~1 m thick sandstone with a visual porosity of 10% was intersected at 704 m. The carbonates have been described as mainly micritic and have no visible porosity (Gatehouse, 1979).

Fig. 10.13 Porositylog permeability plot, Ouldburra Formation; Marla Overthrust Zone and Manya Trough depth range 187 1471 m (raw data from Kamali, 1995b).

Fig. 10.14 Porosity histogram of pore distribution, Arcoeillinna Sandstone.

125

The porositylog permeability plot (Fig. 10.15) is remarkably linear, with Marla Overthrust Zone and Manya Trough permeability values ranging between 0.1 and 50 md, and Murnaroo Platform permeabilities averaging 291 md with a maximum exceeding 1700 md. Gamma Ray and Vshale values are very high (Fig. 10.16) but two fields can be distinguished on the porositylog permeability plot. One, which comprises mostly sandstone (with muddy laminae), has a low Gamma Ray and very high Vshale. Abundant mica may be responsible. The other, which comprises mostly mudstone, has a very high Gamma Ray and low to moderate Vshale. These observations must be taken with caution because of the poor quality logging in the basin. Because of its abundant thin muddy interbeds, the Arcoeillinna Sandstone is not a high quality reservoir but could be considered a potential secondary target. The formation is a fine to medium-grained, immature micaceous arkose with numerous muddy laminae and mudstone interbeds. Benbow (1982) provided the following sandstone composition: quartz (50%), K-feldspar (35%), minor plagioclase, lithic grains (10%), and muscovite and biotite (5%). His interbedded siltstone and claystone composition is: quartz and feldspar (35%), muscovite and biotite (35%), and chlorite (30%). This heterogeneous composition is confirmed by XRD data (Gravestock and Sansome, 1994).

Seal The widespread, seismically mappable Apamurra Formation (Benbow, 1982; Stainton et al., 1988) is a potential regional seal above the Arcoeillinna Sandstone or older reservoirs. Trainor Hill Sandstone Distribution The Trainor Hill Sandstone was originally widespread but was thinned and locally removed in the Marla Overthrust Zone by Delamerian erosion. Maximum thickness in Marla 10 is 316 m but exceeds 440 m on the Murnaroo Platform, reaching a maximum preserved thickness of 520 m in Lairu 1. Where preservation is more complete it thus rivals the Murnaroo Formation in thickness. Petrophysics Like the preceding Arcoeillinna and Relief Sandstones, porosity distribution of the Trainor Hill is bimodal (Fig. 10.17). Mean porosity values are very good 15% in the Marla Overthrust Zone and Manya Trough and 22% in the Murnaroo Platform. Permeability values on the porositylog permeability plot (Fig. 10.18) are high, usually tens to hundreds of millidarcies in the Marla Overthrust Zone and up to 5249 md on the Murnaroo Platform (Ungoolya 1, depth 667 m). Core-log porosity correlation (calcite matrix density) is good (Fig. 10.19). Gamma Ray values reach 300

Fig. 10.15 Porositylog permeability plot, Arcoeillinna Sandstone; Murnaroo Platform depth range 9591200 m; Marla Overthrust Zone depth range 106566 m.

Fig. 10.17 Porosity histogram of pore distribution, Trainor Hill Sandstone.

Fig. 10.16 Vshale versus Gamma Ray for Marla 4, Arcoeillinna Sandstone.

Fig. 10.18 Porositylog permeability plot, Trainor Hill Sandstone; Murnaroo Platform depth range 472907 m; Marla Overthrust Zone depth range 105459 m.

126

API and Vshale is predominantly in the range 5 to 80%. Values >100% on Figure 10.20 are spurious and related to a correction factor for poor logs. Upper levels of the Trainor Hill Sandstone in outcrop are calcareous and dolomitic, but the abundant kaolin reported from outcrops (Benbow, 1982) is not matched by the XRD data from cores (Gravestock and Sansome, 1994). The near-surface kaolin results from a Tertiary weathering event which affected sandstones as young as Jurassic (Algebuckina Sandstone) and is evident in most of the upholes drilled for velocity data in the region. Feldspar content is usually low but in outcrop it ranges up to 35%; biotite is lacking in contrast to the Arcoeillinna (Benbow, 1982). Seal The Trainor Hill Sandstone is usually overlain disconformably by the Ordovician Mount Chandler Sandstone. The Delamerian unconformity at the top of the Trainor Hill Sandstone is a moderate reflector but due to muting and near-surface noise, coupled with the weight-drop technique, picking the reflector in the basin is quite difficult (Mackie, 1994; Rudd, 1995). The overlying Mount Chandler Formation is sandy and does not provide a stratigraphic seal. However, in Devonian thrust zones, there is a good chance of fault seal and of juxtaposition against potential Cambrian source rocks. In this scenario, the Trainor Hill Sandstone would be in the footwall and sealed by rocks in the hanging wall.

Mount Chandler Sandstone


Distribution This Ordovician sandstone is widespread and once thickened northwards over the Musgrave Block before being eroded during the Alice Springs Orogeny. Maximum drilled thickness ranges from 212 m in the Marla Overthrust Zone (Byilkaoora 2) to 472 m on the Murnaroo Platform (Karlaya 1) but may be >600 m thick in outcrop (Benbow, 1982). Petrophysics There are too few samples from the Murnaroo Platform to compare porosity distribution but the porositylog permeability plot (Fig. 10.21) indicates consistently high porosity (up to 25.4%) and permeability (up to 238 md). Log porosity data agree well with core data (Sansome and Gravestock, 1993) using a quartz matrix density of 2.65 g/cm3 (not calcite this was an error in the above abstract). Gamma Ray and Vshale values are correspondingly low.

Fig. 10.21 Porositylog permeability plot, Mount Chandler Sandstone; Murnaroo Platform depth range 323400 m; Marla Overthrust Zone and Manya Trough depth range 63192 m.

Seal The Mount Chandler Sandstone is locally sealed in the northern Munyarai TroughMount Johns region by the Indulkana Shale. The shale reaches a thickness of 60 m in the Indulkana Range (Krieg, 1973) and may be quite widespread east of the Marla Overthrust Zone based on aeromagnetic evidence (Hamer, 1994). However, in the Marla Overthrust Zone, the sandstone requires hanging wall fault structures for seal. The Mount Chandler and stratigraphically younger Blue Hills Sandstone (not studied) also run the risk of being breached by Permian erosion. Despite its excellent reservoir qualities, the Mount Chandler Sandstone is thus unlikely to be a major target for petroleum.

Fig. 10.19 Core porosity versus porosity calculated from the density log, Trainor Hill Sandstone.

Fig. 10.20 Vshale versus gamma ray for Lairu 1, Trainor Hill Sandstone.

 %

 &

INTRODUCTION
An understanding of the range of structural styles and their causes increases with the density of seismic and drilling data. In the 100 000 km2 Officer Basin in South Australia, ~6250 km of seismic have been recorded and 40 wells drilled deeper than 400 m. Understandably there have been as many structural interpretations as interpreters, with models ranging from extensional to compressional. A compressional style is now clearly evident and is particularly well expressed on the 1993 seismic surveys
Table 11.1 Eastern Officer Basin seismic surveys.
Survey code
53OF01 62OF01 66OF01

conducted by MESA and AGSO (Mackie, 1994; Gravestock and Lindsay, 1994; Lindsay, 1995; Lindsay and Leven, 1996; Hoskins and Lemon, 1995). Seismic recorded on the southeastern flank of the Munyarai Trough has also revealed structures associated with salt movement (Badley, 1988; Stainton et al., 1988; Thomas, 1990) and palaeoreliefs associated with a canyon-cutting event (Thomas, 1990; Sukanta et al., 1991). Seismic surveys are listed in Table 11.1, and seismic line coverage of the basin is shown on Figure 1.2 (Ch. 1).

Survey name
1953 atomic test deep refraction survey Mabel Creek area seismic survey Serpentine Lakes reconnaissance seismic survey

Operator
British Atomic Weapons Research Establishment Exoil Pty Ltd Continental Oil Co. (Aust.) Ltd

Contractor
BMR Namco International Inc. Seismograph Services Ltd

Year
1953 1962 1966

Line coverage (km) *refraction


137, 27* 944

MESA reference
BMR Record 1954/64 Env. 224, 225 (1, AC), 214 Env. 501 (122), 603 (12) Env. 697, 692; Report 64/27

66OF02

1966 Eastern Officer Continental Oil Co. Basin seismic reflection, (Aust.) Ltd refraction and gravity survey Eastern Officer Basin seismic and gravity survey Northern margin of the eastern Officer Basin and Wintinna Trough seismic survey Everard seismic survey OF78 seismic survey Mini-sosie seismic reflection survey in EL 699 (Manya) 1983 PEL 23 seismic survey (Marla) Marla seismic survey 1983 PEL 23 1984 seismic survey PEL 23 1985 seismic survey PEL 23, 30 Comalco 1986 seismic survey 1987 Amoco Officer Basin seismic survey NGMA transect Wallatinna seismic survey Continental Oil Co. (Aust.) Ltd SADME

SADME

1966

106, 58*

67OF01

Namco Geophysical Co.

1967

169

Env. 747, 795, 829

74OF01

SADME

1974

475 86*

Report 73/31, 73/181

74OF02 78OF01 80OF01

Shell Development (Aust.) GES Pty Ltd SADME Comalco Aluminium Ltd SADME Velocity Data Pty Ltd

1974 1978 1980

168 94 53

Env. 2509 (13) Env. 6259 (1920)

83OF01 83OF02 84OF01 85OF01 86OF01 87OF01 AGS93 OF93

Comalco Aluminium Ltd Comalco Aluminium Ltd Comalco Aluminium Ltd Comalco Aluminium Ltd Comalco Aluminium Ltd Amoco Australia Petroleum Co. AGSO MESA

Petty Ray Geophysical Geosystems Pty Ltd Petty Ray Geophysical Petty Ray Geophysical Petty Ray Geophysical Petty Ray Geophysical AGSO Geosystems

1983 1983 1984 1985 1986 1987 1993 1993

257 36 1278 706 496 235 550 379 Total coverage 6252 km

Env. 5805 (13, 56) Env. 5805 (13) Env. 5923, 6165 Env. 6993 Env. 6902 Env. 6766 Lindsay (1995) Env. 8812

 '

PREVIOUS UNSUCCESSFUL TESTS


Despite identification of an array of trap types and multiple sandstone reservoirs under seal, there have been few on-structure tests of effectively trapped sands. Table 11.2 lists the six key wildcat wells and five sealed sandstone reservoirs (Birksgate 1 is omitted due to its remoteness). Not one well in Table 11.1 is from the Marla Overthrust Zone which remains untested despite oil bleeds in nine mineral exploration drillholes (Fig. 11.8). This area is now under licence. Each of the six wells listed in Table 11.1 had four chances of hitting a sealed reservoir, totalling 24 chances. The Relief Sandstone is not included because it is present only in Giles 1 and was not drilled in closure according to Mackies (1994) interpretation. Of the 24 chances, 14 missed the opportunity to reach the target. There were only five intersections of a sandstone reservoir under seal, and four of these comprise the Arcoeillinna Sandstone. Each of the five intersections is discussed below.

dolomitic siltstone with disturbed anhydrite laminae and abundant dewatering structures. The common dewatering structures suggest an ineffective seal despite the presence of anhydrite. Minor gas shows recorded in the Apamurra and near the base of the overlying Trainor Hill support the lack of seal rather than unfavourable location with respect to migrating hydrocarbons as the well was in closure at the Apamurra level (Henry, 1986). However, Mackies (1994) mapping suggests that the structure may have been breached by faulting.

Arcoeillinna Sandstone in Karlaya 1


The Arcoeillinna Sandstone in Karlaya 1 was intersected at a depth of 1037 m beneath 78 m of Apamurra Formation (Fig. 6.22). The Apamurra is described as red-brown siltstone, sandy near the top, with 0.15 m of dolomite near the base (Dunster, 1987b). Recent seismic mapping (Mackie, 1994) suggests that Karlaya 1 was off-structure at this level. No Cambrian shows were recorded in Karlaya 1; the highest recorded show (1430 m) was from the Narana Formation.

Arcoeillinna Sandstone in Munyarai 1


In Munyarai 1, the Arcoeillinna Sandstone was intersected at a depth of 1484 m (4869 ft KB) beneath the Apamurra Formation which forms the seal. The Apamurra is represented by core 7 (2.8 m recovery), a brick-red mudrock with granules of quartz, feldspar and heavy minerals in bands or floating in a dolomitic matrix. From this description the Apamurra Formation is probably an effective seal. The lack of hydrocarbon shows suggests that the Arcoeillinna was isolated from Cambrian and Proterozoic hydrocarbon migration pathways. The only oil show recorded in Munyarai 1 was trace yellow and rare blue fluorescence in the Karlaya Limestone.

Arcoeillinna Sandstone in Lairu 1


The Arcoeillinna Sandstone in Lairu 1 was intersected at 1091 m beneath 36 m of Apamurra Formation, the latter lithologically similar to the Karlaya 1 occurrence. The well was deliberately sited off-structure (Dunster, 1987c: Stainton et al., 1988) and no significant shows were recorded.

Murnaroo sandstone in Munta 1


Sandstone of the Murnaroo Formation was intersected at a depth of 1973 m beneath nearly 300 m of Dey Dey Mudstone. The basal 50 m of the latter unit are composed of slightly calcareous and dolomitic siltstone grading to claystone, which suggests that the Dey Dey is an effective seal. Traces of gas were recorded in the Tanana Formation and Karlaya Limestone but not in the Murnaroo. Although the well was sited to test structural closure in the MurnarooGilesTarlina package (Cucuzza, 1987), detailed mapping (Thomas, 1990) suggests that salt withdrawal has

Arcoeillinna Sandstone in Ungoolya 1


In Ungoolya 1, the Arcoeillinna Sandstone was intersected at 956 m depth beneath 39 m of Apamurra Formation. The latter is described as a red-brown, slightly

Table 11.2 Opportunities to test sandstone reservoirs and reasons for failure. Mount Chandler, Blue Hills and Trainor Hill Sandstones a re omitted due to lack of seal or erosion. Numbers 1 to 5 refer to sandstones intersected under seal.
Well Year drilled Trap
Munyarai 1 1968 Anticline Giles 1985 Anticline Ungoolya 1 1985 Anticline Karlaya 1 1987 Faulted anticline Lairu 1 1987 Anticlinal flank

Reservoir sandstone Pindyin


Not reached

Tarlina
Not reached

Murnaroo
Not reached

Relief
Not present

Arcoeillinna
1

No closure

No closure

No closure

No closure

At surface

Not reached

Not reached

Not reached

Not present

Not reached

Not reached

Not reached

Not present

Not reached

Not reached

Not reached

Not present

Munta 1 Not reached 1987 Sub-unconformity tilted fault block

Not reached

Not reached

No closure

!

deformed the Murnaroo and underlying formations into a series of tilted fault blocks and, as a result, Munta 1 is off-structure at this level (Fig. 11.5). Mackie (1994) also mapped this location outside closure (Fig. 11.3).

POTENTIAL TRAP TYPES


A great variety of structural traps exists in the Officer Basin, most of which have been identified but not drilled. They include: foreland basement thrusts (Steenland, 1965) detached thrusts with hanging wall and footwall rollovers (Stainton et al., 1988) simple anticlines with at least two generations of folding (Petermann Ranges and Alice Springs Orogenies; Stainton et al., 1988) tilted fault blocks associated with salt withdrawal (Badley, 1988; Stainton et al., 1988; Thomas, 1990) salt walls and diapirs (Badley, 1988; Thomas, 1990) palaeoreliefs (buried hills) which originated with the canyon-cutting event (Stainton et al., 1988; Thomas, 1990). Subcrop unconformities also have a stratigraphic component. Purely stratigraphic traps, for example porous sandstone or dolomite beds sealed within the Ouldburra Formation (Dunster, 1987a; Gravestock and Hibburt, 1991; Kamali, 1995b) or vuggy Moyles Chert Marker Bed in the Observatory Hill Formation, also have hydrocarbon potential.

reservoir horizons (Tarlina and Pindyin Sandstones). Intense faulting is probable at the Tarlina and Murnaroo horizons due to the salt flowage (Fig. 11.5). The structure map on Figure 11.3 shows an area of closure of 65 km2 (16 000 acres) and a maximum vertical closure of 69 m (226 feet). The unrisked potential oil in place of the structure in the Pindyin Sandstone is ~67 million kilolitres (420 million barrels), and similar potential would also exist in the Arcoeillinna Sandstone, Murnaroo Formation and Tarlina Sandstone.

Prospect B
Prospect B (Fig. 11.6) is a combination stratigraphic and structural trap. It comprises a subcrop pinchout on a compressive fault-bounded anticline (Fig. 11.7). Seismic quality is reasonable, but there is insufficient seismic coverage to the north to prove closure. The Arcoeillinna Sandstone is not likely to be sealed at this location. The structure map on Figure 11.6a shows an area of closure for a trap in the Cambrian (Ouldburra Formation or Relief Sandstone ) of 54 km2 (13 350 acres) and a maximum vertical closure of 48 m (157 feet). The unrisked potential oil in place of the structure in the Relief Sandstone is ~45 million kilolitres (290 million barrels), but with the possibility of stacked sand reservoirs, the potential for the Ouldburra is over 1 billion barrels (165 million kilolitres). The structure map on Figure 11.6b shows an area of closure for a trap near basement (Pindyin Sandstone) of 130 km2 (32 000 acres) and a maximum vertical closure of 103 m (338 feet). The unrisked potential oil in place of the structure in the Pindyin Sandstone is ~120 million kilolitres (750 million barrels). Significant potential would also exist in the Tarlina Sandstone as an unconformity play beneath the erosional base of the Narana Formation, but this has not been estimated.

POTENTIAL TRAP VOLUMES


Many potential leads and prospects were identified by seismic mapping in the Munta and Marla areas by Thomas (1990) and Mackie (1994). Some additions and/or modifications were made in the Marla area Figures 11.1 and 11.2 illustrate all closures mapped at the base Cambrian and top basement horizons. Four prospects were selected to illustrate the potential volumes of oil that the various types of Officer Basin traps could contain. Average porosities, water saturations, net to gross ratios and formation volume factors used to calculate these estimates are the average values used for the various plays summarised in Chapter 13 (Undiscovered resources). All potential reserve estimates are unrisked to maximum closure and are of oil in place. Source richness and maturity will determine the extent to which these structures are filled; this is likely to be considerably less than 100%, given the large trap sizes.

Prospect C
Prospect C (Fig. 11.8) is a probable detached fold-thrust trap (Fig. 11.9). Seismic quality is poor but there is reasonable seismic coverage. Ten wells have been drilled in the vicinity of the prospect, most of which had numerous oil shows in the Observatory Hill Formation and, in the case of Marla 9, in the Tanana Formation and Dey Dey Mudstone (Hibburt et al., 1995). Most of these wells did not reach the Ouldburra Formation and Relief Sandstone, with the exception of Marla 3 which may have been drilled close to the closure for the prospect, but considerable up-dip potential exists. The Proterozoic sequence has been interpreted to be thin or absent (Mackie, 1994), but the section intersected in Marla 9 to the west of the prospect suggests that at least the Murnaroo target may be present. The main Cambrian target horizons are the Arcoeillinna Sandstone, Ouldburra Formation and Relief Sandstone. The structure map on Figure 11.8 shows an area of closure for a trap in the Cambrian of 146 km2 (38 550 acres) and a maximum vertical closure of 241 m (790 feet). The unrisked potential oil in place of the structure in the Relief Sandstone up dip from Marla 3 is ~900 million kilolitres (5.7 billion barrels).

Prospect A
Prospect A (Fig. 11.3) was identified by Mackie (1994) in the vicinity of Munta 1. The trap is a probable salt wall piercement structure with reservoir horizons abutting the sides of the salt wall (Fig. 11.4). The Relief Sandstone and Ouldburra Formation were not deposited in this area, although potential may exist in the Arcoeillinna Sandstone. Seismic quality is not good, but Munta 1 appears to have been drilled outside closure and did not penetrate the two lower
!

Prospect D
Prospect D (Fig. 11.10) is a foreland basement upthrust trap (Fig. 11.11). Seismic quality is reasonable but further

!
Fig. 11.1 Cambrian (Arcoeillinna Sandstone, Ouldburra Formation, Relief Sandstone) prospects and leads (after Mackie, 1994; base Cambrian time horizon). Contour interval 50 ms.

!!
Fig. 11.2 Proterozoic (Pindyin and Tarlina Sandstones, Murnaroo Formation) prospects and leads (after Mackie, 1994; crystalline basement time horizon). Contour interval 100 ms.

97-0205 MESA KILOMETRES

Fig. 11.3 Time structure map, crystalline basement, Prospect A (after Mackie, 1994). Contour interval 100 ms.

seismic coverage is required in the southern area. Manya 2 and 3 were drilled just outside closure. Middle Bore 1 was possibly drilled on closure just to the north of Prospect D. None of these wells reached the Relief Sandstone and in Middle Bore 1 it was faulted out. The main Cambrian target horizons are the Arcoeillinna Sandstone, Ouldburra Formation and Relief Sandstone. Potential also exists in Proterozoic reservoirs, although these and some of the other Cambrian targets may not be present due to faulting. The structure map on Figure 11.10 shows an area of closure for a trap in the Cambrian of 299 km2 (73 900 acres) and a maximum vertical closure of 207 m (680 feet). The unrisked potential oil in place of the structure in the Relief Sandstone is ~2200 million kilolitres (13.7 billion barrels).

BASE CAMBRIAN

INTRA UNGOOLYA GP UNCONFORMITY


SALT

BASE TARLINA UNCONFORMITY

BASEMENT

0 KILOMETRES

2
97-0255 MESA

Fig. 11.4 Seismic line 86-106 through Prospect A, showing probable abutment of reservoirs against a salt pillar.

!"

Fig. 11.5 Time structure map, top Murnaroo Formation, Prospect A (after Thomas, 1990). Contour interval 20 ms.

Fig. 11.6 (a) Time structure map, base Cambrian, Prospect B. Contour interval 50 ms. (b) Time structure map, crystalline basement, Prospect B. Contour interval 100 ms. (after Mackie, 1994).

!#

BASE CAMBRIAN

INTRA U NGOOL YA G UNCON FORMIT P Y

BASE TARLINA UNCONFORMITY

BASEMENT

0 KILOMETRES

2
97-0253 MESA

Fig. 11.7 Seismic line 86-0044 through Prospect B.

!$

Fig. 11.8 Time structure map, base Cambrian, Prospect C (light shading; after Mackie, 1994); additional closures are shaded darker.

!%

BAS
A EC MB RIA N

E PE

RMIA

BA

BA

SE

ME

NT

BASE

CA

E DN

AR YP

K
FM

BASE

CAMBRIAN

BASEMENT

0 KILOMETRES

2
97-0254 MESA

Fig. 11.9 Seismic line 85-0096 through Prospect C.

!&

Fig. 11.10 Time structure map, base Cambrian, Prospect D (light shading; after Mackie, 1994); additional closures are shaded in darker.

!'

BA

SE

CAMBRIAN

E AS

ME

NT

KILOMETRES

97-0256 MESA

BASE CAMBRIAN

BASEMENT

KILOMETRES

97-0257 MESA

Fig. 11.11 (a) Seismic line 84-0060 through the northern part of Prospect D. (b) Seismic line 85-0080 through the southern part of Prospect D.

"

INTRODUCTION
MESA has developed an economic model for oil exploration and development in the Marla region as a guide to potential explorers in the Officer Basin region. Although the Officer Basin has gas potential, gas discovery economics have not yet been modelled as development is dependent on future links between gas reserves in the northwest of Australia and markets on the eastern seaboard (Fig. 4.1). If sufficient reserves are discovered, a pipeline to either the Adelaide or eastern Australian markets (via Moomba) may be economic. Development of oil reserves is generally less capital intensive than gas. Two scenarios were examined: The 30 mmbbl OOIP (original oil in place) case, in which oil is trucked from Marla and the infrastructure required is a trunkline from the field to the Stuart Highway, storage tank, yard, office and workshop at Marla, basic surface facilities and truck loading facilities. The 100200 mmbbl OOIP case, in which oil is piped to Port Bonython and the infrastructure required is the same apart from a larger storage tank and trunkline. In both cases, a single oil pool in a single field has been modelled on a stand-alone basis it is likely that more than one oil pool and field would be discovered. Marla has diesel generated power, water and communication infrastructure available, and the fully sealed Stuart Highway runs through the area. It is currently uneconomic for oil from a new discovery in Central Australia to be transported via rail to Adelaide even though the railway runs through the region. Trucking oil to the Port Stanvac refinery near Adelaide using two or three tank road trains is the alternative. Privatisation of the rail freight network may improve rail transport economics in the future.

conservative and that Officer Basin reservoirs may have better reservoir characteristics. The second spreadsheet determines the net present value of the project by calculating exploration, capital and operational costs, royalty and revenue for a range of field sizes, enabling minimum economic field size in the region to be determined. A high oil price of US$25/bbl and a low oil price of US$18/bbl were used. This covers variations in oil price forecasts for 1997, which range from US$19 to US$25/bbl (Bell, 1997). Appendix 12.1 is a summary of information used in the model. Other sources of information used were GPA Engineering Pty Ltd (GPA, 1996) and Bureau of Resource Sciences (BRS, 1996).

EXPLORATION SUCCESS RATIOS


The Australian historical base case success ratio for a wildcat well to discover an economic field is ~1:12.5; the success ratio over the period 197887 was 1:8.3, reflecting modern techniques and knowledge (MacKenzie and Cai, 1993). A total of eight petroleum exploration wells have been drilled in the entire Officer Basin, but none of these are in the Marla area. In the economic model, oil is discovered by the fifth exploration well of the hypothetical drilling program, or the thirteenth petroleum exploration well in the Officer Basin, a 1:13 success ratio comparable to the historical Australian case and more conservative than the modern Australian case. A pessimistic case where oil is discovered in the eighth exploration well was examined as part of the case study below.

EXPLORATION AND DEVELOPMENT SCENARIOS


An exploration program, based on the combined program for PEL 61 and 63, was used up to the discovery of a single oil field in Year 3. The development program was then applied and run until either production ceased due to depletion or until one PPL term (21 years) had elapsed. A single Murnaroo Formation oil pool was modelled but there is a good possibility of multiple stacked hydrocarbon pools in the Officer Basin. In the Marla area, stacked pools in the Arcoeillinna, intra-Ouldburra, Relief and Murnaroo are possible. In the Munta area, stacked pools in the Arcoeillinna, Relief, Murnaroo, Tarlina and Pindyin are possible. It is also likely that more than one field would be discovered in a particular area (Mackie, 1994). Table 12.1 shows the field sizes that were modelled using US$18/bbl
"

METHODOLOGY
Two Microsoft Excel spreadsheets were used to develop the economic model. The first spread sheet modelled oil production for a range of field sizes and permeability thickness kh (which determines the flow capacity of a well; in millidarcy feet (md.ft)), and results were placed into the economic model adapted from the Cooper Basin gas model (McDonough, 1996, in press). Reservoir properties were based on the early Palaeozoic P3 Pacoota Sandstone reservoir in the Mereenie Oilfield as this information is very limited in the Officer Basin. It is likely that the Mereenie case is more

Table 12.1 Field sizes used for the economic model, in imperial and metric units.
Field size (million barrels) CASE 1 (<30 million barrels) 5 10 15 20 30 CASE 2 (>30 million barrels) 50 100 200 7 949 380 15 898 760 31 797 510 794 940 1 589 875 2 384 810 3 179 750 4 769 630 Field size (kilolitres)

(52 m3/day); commence trucking oil to Port Stanvac near Adelaide. Pay petroleum royalty. Year 4 development Conduct geological and geophysical data review, estimate oil reserves, and plan development drilling. A total of five wells will be used to produce from the field. The development drilling program is determined from the production profiles. The maximum field production rate is set at 1000 bopd (158.9 m3/day), with a limit of five wells; higher volumes of oil could not be handled by trucking to Adelaide (Appendix 12.1). For the purpose of modelling flowlines, the field is assumed to be circular (radius = 5 km; area = 75 km2); the discovery well is assumed to be near the field centre and development wells are sited halfway between the edge and centre of the field. As the reservoir is assumed to be a solution gas drive, it is assumed that there will be no water production and therefore no associated disposal requirements. Site a 10 000 bbl (1589 m3) storage tank at the field. Year 5 onwards development Construct a 50 km pipeline to the Stuart Highway, establish a truck loading facility on the highway and truck oil to Adelaide. Acquire and interpret 100 km of development seismic. Drill development wells as required by the production model.

and US$25/bbl for the oil price and for a range of reservoir properties. Case 1 was used to determine the minimum economic field size; Case 2 was used to determine a conservative upside potential. Much larger field sizes are possible in the Officer Basin (see Ch. 11). Facilities required for Case 1 and 2 are shown on Figure 12.1.

+=IA
The second exploration and development scenario, for fields in the 50200 mmbbl OOIP range, was as follows: Year 0
Fig. 12.1 Case 1 and 2 schematic of facilities.

Apply for PEL, and negotiate access agreement with Anangu Pitjantjatjara. Year 1 exploration Construct access roads and, using the existing seismic grid, drill two exploration wells. Year 2 exploration Construct an access road and drill one exploration well; acquire and interpret 320 km of seismic. Year 3 exploration and discovery of oil Construct an access road and drill two exploration wells; acquire and interpret 250 km of seismic. The second exploration well discovers a large oil field and is completed. Apply for a PPL; initially limit oil production to 3300 bopd (520 m3/day), and commence trucking to Port Stanvac. Site a 100 000 bbl (15 898 m3) storage tank at the field. Commence a feasibility study and construction of a 740 km trunkline to Port Bonython (the MoombaPort Bonython trunkline was used as model; McDonough, 1996). Year 4 development Conduct geological and geophysical data review, estimate oil reserves, plan development drilling with a
"

+=IA 
The first exploration and development scenario, for fields 30 mmbbl OOIP, was as follows: Year 0 Apply for PEL, negotiate access agreement with Anangu Pitjantjatjara. Year 1 exploration Construct access roads and, using existing seismic grid, drill two exploration wells. Year 2 exploration Construct access road and drill one exploration well; acquire and interpret 320 km of seismic. Year 3 exploration and discovery of oil Construct access road and drill two exploration wells; acquire and interpret 250 km of seismic. The second exploration well discovers oil and is completed. Apply for a PPL, and initially limit oil production to 330 bopd

maximum of 10 wells and limit field production to 10 000 bbl/day (1589 m3/day). Complete the construction of a trunkline to Port Bonython. The development model is not entirely realistic as the amount of oil produced daily during Years 3 and 4 while the trunkline is being constructed could not be handled by trucking alone. However, this is somewhat offset by the capacity of the trunkline (33 000 bbl/day or 5250 m3/day). Year 5 onwards development Acquire, process and interpret 200 km of development seismic. Drill development wells as required by the production model.

Royalty and licence fees


Guidelines for the payment of petroleum royalty were applied to these scenarios. Licensees pay royalty at a rate of 10% of the value of the petroleum at the wellhead minus all expenses (capital and operating costs) actually incurred in treating, processing or refining the petroleum downstream of the wellhead. PEL application fees were included for the first five years of the model (one PEL term). PPL application, rental and renewal fees were applied from the year that production commenced and continued for 21 years (one PPL term) or until the field ceased production.
Fig. 12.2 Net present value analysis over the life of the project; revenue versus costs (20 mmbbl OOIP field, kh = 9200 md.ft).

expenditure are relatively minor components. A reduction in the price of oil from US$25 to US$18/bbl results in a major reduction of profit and hence royalty (Table 12.2).
Table 12.2 Net present value of profit and expenses for a 20 mmbbl OOIP field (kh = 9200 md.ft).
Cash flow US$18/bbl 5.3 3.5 22.2 5.8 1.9 21.9 NPV ($ million) US$25/bbl 5.3 3.5 22.2 18.5 3.4 7.27

RESULTS
For each field size, the following kh values were used for a 60 m thick pay zone within the Murnaroo Formation: 250 md.ft the most pessimistic case where permeability = 0.787 md (based on average permeability data for the early Palaeozoic Pacoota Sandstone P3 reservoir, Mereenie oil field, Northern Territory). 1000 md.ft calculated using arithmetic average permeability (5.371 md) from the Murnaroo Formation, Marla region (Ch. 10). 9200 md.ft geometric average permeability from the Murnaroo Formation (46 md), Marla region (Ch. 10). 17 700 md.ft geometric average permeability from the Murnaroo Formation (86 md), Munta region (Ch. 10).

Exploration CAPEX OPEX Profit before tax and royalty Royalty Revenue

A breakdown of OPEX (Fig. 12.3) indicates that transport costs (i.e. trucking oil from Marla to Port Stanvac) are the most significant operating expense. PEL application fees and rent averaged under $5000/year and were applied over the first five years of the model (one PEL term). PPL application, rental and renewal fees were ~$20 000/year for case 1 and $94 000/year for case 2. Other land access costs are: payment of compensation of a minimum of $20 000 per annum to AP when exploration is carried out and,

20 mmbbl OOIP case study


Results for a 20 mmbbl OOIP field (kh = 9200 md.ft) are summarised on Figure 12.2 as an example of how each OOIP and kh scenario was modelled. Oil production and the timing of development drilling was calculated for this field size and kh, and input into the economic model. Exploration expenditure, capital expenditure (CAPEX), operating expenditure (OPEX), revenue and petroleum royalty were modelled over a 20-year period and the net present value (NPV) of the project calculated at a discount rate of 12.5%. All calculations are before tax. A breakdown of the net present value (Fig. 12.2) reveals that OPEX is the most significant component of the NPV, profit is next, while royalty, CAPEX and exploration
"!
Fig. 12.3 Net present value analysis of operating expenditure (20 mmbbl OOIP field, kh = 9200 md.ft).

payment of production royalties to AP of 13% on a sliding scale based on the quantity of oil and gas produced, following an economic discovery. CAPEX has been divided into road construction and vehicles, development drilling and seismic, and surface equipment (trunkline, flowlines, pumps) on Figure 12.4. Surface equipment makes up nearly half of the CAPEX due to construction costs of a trunkline from the field to the Stuart Highway. Seismic and drilling costs are the other significant component of CAPEX. The net present value of the before tax cash flows for the project is $21.9 million for an oil price of US$25/bbl, representing an internal rate of return of 44.5% (a real discount rate of 12.5% was used in all cases). The cashflows over the life of the project are shown on Figure 12.5. The impact of exploration and capital expenditure during the first three years of the project are obvious. Once the pipeline is commissioned, production generates revenue and positive cashflows commence. Operating costs total ~$4.9 million/year. Wells were drilled in Years 1 (discovery well), 2, 14 and 15 (two wells). The discovery of oil was delayed until Year 10 for this case study to see what effect lack of initial exploration success would have on overall project economics. A delay means that a more expensive exploration program is required and that generation of revenue from production occurs later. If the discovery of oil is delayed until Year 10 of the exploration licence after the drilling of eight exploration wells, a 20 mmbbl OOIP sized field is not economic. This

implies that a successful exploration strategy is to drill wells within the first five years to bring forward possible discoveries.

Discussion
Tables 12.312.6 summarise the NPVs and IRR (internal rate of return) of all scenarios modelled with a real discount rate of 12.5 %.
Table 12.3 NPV for all field sizes and kh; oil price = US$25/bbl.
Field size (million barrels) Case 1 5 10 15 20 30 Case 2 50 100 200 -14.9 28.3 120 35.6 110.4 199.4 80.2 129.2 219.6 135.4 223.4 -2.1 5.5 11.0 14.9 19.7 -1.9 6.7 12.2 16.8 22.1 6.5 12.5 18.1 21.9 23.2 6.5 13.4 17.6 22.1 23.2 250 kh (md.ft) 1000 9200 17 700

Table 12.4 NPV for all field sizes and kh (oil price = US$18/bbl).
Field size (million barrels) Case 1 5 10 15 20 30 Case 2 50 100 200 -57.5 -33.1 14.8 -29.6 11.5 60.2 80.2 25.68 74.63 29.43 76.30 -7.2 -3.1 -0.2 2.2 5.1 -6.9 -2.3 0.6 3.4 6.9 -1.6 1.9 5.1 7.3 7.6 -1.6 2.8 5.2 7.6 7.6 250 kh (md.ft) 1000 9200 17 700

Table 12.5 IRR as a percentage, for all field sizes and kh (oil price = US$25/bbl).
Field size (million barrels) Case 1 IRR (%) 5 10 15 20 30 -1.5 29.8 35.9 39.0 41.0 5.7 21.91 70.50 -4.7 32.2 37.0 41.1 42.0 66.1 93.67 99.93 43.9 42.1 44.4 44.4 42.1 118.9 100.46 101.76 43.9 42.8 44.4 45.0 42.1 101.42 101.76 250 kh (md.ft) 1000 9200 17 700

Fig. 12.4 Net present value analysis of capital expenditure (20 mmbbl OOIP field, kh = 9200 md.ft).

Case 2 IRR (%) 50 100 200

Table 12.6 IRR as a percentage, for all field sizes and kh (oil price = US$18/bbl).
Field size (million barrels) Case 1 IRR (%) 5 10 15 20 30 Case 2 IRR (%) 50 100 200 -1.6 11.9 17.0 21.0 -0.6 16.2 3.5 13.9 19.8 23.1 18.4 31.0 2.0 18.4 23.5 24.9 23.4 10.9 25.8 33.7 2.0 20.3 23.7 25.5 23.4 27.3 33.7 250 kh (md.ft) 1000 9200 17 700

Fig. 12.5 Net present values (before tax cash flow) for 20 mmbbl OOIP field over the life of the project (kh = 9200 md.ft, US$25/bbl oil price, discount rate = 12.5%).

""

Minimum economic field size


NPV results for high (US$25/bbl) and low (US$18/bbl) oil prices at a real discount rate of 12.5% have been plotted for Case 1 field sizes versus NPV for each individual kh (Figs 12.69). The 250 md.ft case (Fig. 12.6) represents the most conservative scenario with the poorest reservoir quality. The worst case minimum economic field size (note: original oil in place, not recoverable oil) ranges from 15 mmbbl for US$18/bbl oil down to 6 mmbbl for US$25/bbl oil. Better reservoir quality in the 1000 md.ft case means that the minimum economic field size drops slightly to 14 mmbbl for US$18/bbl oil; there is little change in the US$25/bbl case (Fig. 12.7). The profitability increases with the improvement in reservoir quality. Increases in kh extend the number of

years that maximum production is achieved and influences the timing of development wells. The 9200 and 17 700 md.ft cases show a dramatic increase in profitability from the low end cases, and the minimum economic field size drops to 7 mmbbl for US$18/bbl and <5 mmbbl OOIP for US$25/bbl (Figs 12.8, 12.9). There is little difference between the two, as field production is limited to 1000 bbl/day (158.9 m3/day) by oil transport constraints, and both models achieve the production limit for a similar period of time. The only variation between the two is in the timing and number of development wells, which produces minor variations in NPV. Further increases in kh or field size above 30 mmbbl OOIP would necessitate modification of the development scenario as alternative oil transport scenarios would become economic and the production limit could be removed or increased.

Upside potential
Three large field sizes (50, 100 and 200 mmbbl OOIP) were modelled as Case 2 to explore the upside economic potential of Officer Basin oil and the feasibility of a trunkline to Port Bonython. The production limit was increased to 10 000 bbl/day (15 898 m3/day) and NPV calculated for the standard kh values used in Case 1. Figure 12.10 shows NPVs for all scenarios modelled at an oil price of US$25/bbl. The minimum economic field size for Case 2 (kh = 250 md.ft) is ~70 mmbbl OOIP for an oil price of US$25/bbl.
Fig. 12.6 Net present value of before tax cash flows (discount rate = 12.5%) versus Case 1 field sizes for kh = 250 md.ft.

Fig. 12.7 Net present value of before tax cash flows (discount rate = 12.5%) versus Case 1 field sizes for kh = 1000 md.ft.

Fig. 12.9 Net present value of before tax cash flows (discount rate = 12.5%) versus Case 1 field sizes for kh = 17 700 md.ft.

Fig. 12.8 Net present value of before tax cash flows (discount rate = 12.5%) versus Case 1 field sizes for kh = 9200 md.ft.

Fig. 12.10 Net present value of before tax cash flows (discount rate = 12.5%) for each kh versus Case 1 and Case 2 field sizes.

"#

"$

INTRODUCTION
Estimating undiscovered petroleum reserves of the Officer Basin in South Australia provides some quantitative expression of the potential, and a basis for comparison with other basins. As the basin is clearly oil prone, only undiscovered oil resources are calculated; however, gas discoveries may also be possible although it is unlikely that small gas discoveries would be economic. As the Officer Basin has had only minimal exploration effort, with no economic discoveries so far, estimates of undiscovered resources are best carried out by a method that uses available geological data and Monte Carlo type statistical techniques to estimate, as a probability distribution, the undiscovered potential for each play (Morton, 1992, 1995, 1996b). This chapter presents a revised estimate of the undiscovered potential of the basin that reflects new knowledge of potential reservoirs and source rocks acquired since 1992. In total, the average estimate of the potential of the Officer Basin plays in South Australia is ~400 million kilolitres (2.5 billion barrels) of recoverable oil. The increase from the 1992 estimate of ~300 million kilolitres is due mainly to the recognition of additional source rocks and reservoirs, particularly two new Neoproterozoic plays (Pindyin and Tarlina Sandstones), and the potential for stacked reservoirs in the Ouldburra Formation. These estimates may appear to be large in comparison to other, geologically younger, Australian petroleum basins, but are comparable to geologically more analogous (ProterozoicCambrian) petroleum provinces elsewhere in the world. The LenaTunguska province in the Siberian Platform has a predicted potential of 318 million kilolitres (2 billion barrels) of oil and gas liquids, with 2417 billion cubic metres (85 trillion cubic feet) of gas (Meyerhoff, 1982), the Moscow Basin has a potential of 2353 million kilolitres (16 billion barrels) of liquid hydrocarbons (V. Gorbachev, NEDRA, pers. comm., 1997) and, from geochemical evidence in oils from Oman, Proterozoic sediments are now believed to be a very significant source for the prolific oil and gas fields of the entire Persian Gulf area (Edgell, 1991). The proven oil reserves in Oman alone are 795 million kilolitres (nearly 5 billion barrels; Feld, 1997). Potential (undiscovered) resources should not be compared to traditional Proved, Probable and Possible reserves in known discoveries. Undiscovered resources are calculated to give a quantitative indication of the potential of the basin, and require considerable exploration to establish their existence.
"%

METHOD
For a commercial petroleum field to exist in the eastern Officer Basin, four essential components are required: A mature source; a rock unit that contains sufficient organic matter and which has been subjected to sufficient heat and pressure over time to have produced significant quantities of hydrocarbons but not to have destroyed them through excessive heat and pressure. A reservoir horizon; a rock unit that accumulates the generated oil or gas. A reservoir rock must be porous and have sufficient permeability to produce fluids economically. A seal horizon; a rock unit that traps petroleum in the reservoir and prevents further migration. A structure over the reservoir horizon that will concentrate the petroleum in economic quantities and that was present at the time of petroleum expulsion from the source rock. This is usually an anticline, but stratigraphic traps can also be important, e.g. the Ouldburra Formation. When all four of these occur together, a petroleum play or a potential target for exploration exists. The method of estimating undiscovered resources consists of identifying all of the plays that may exist, either by discoveries made so far or by analysis of the available data (e.g. drillhole, geophysical, or outcrop). The oil potential for each play is then calculated using the following formula: Pt = Ap*AB*h*NG*FF*Por*Sh*FVF*SR*RF Where:
Pt Ap Total potential recoverable oil reserves of the play Prospective area of the basin AB Anticline to total basin area ratio h Average gross reservoir thickness. NG Net to gross pay ratio FF Anticline fill factor Por Porosity (fraction) Sh Hydrocarbon saturation (1 - water saturation) FVF Formation volume factor SR Exploration drilling success ratio RF Recovery factor.

None of the above parameters is known with certainty but most can be estimated from available data to within at least

broad limits. The most common method of combining and expressing the uncertainty associated with this type of equation is to use Monte Carlo simulation techniques (White and Gehman, 1979). A frequency distribution for each parameter is assumed, which is converted to a cumulative probability distribution, and a random number between 0 and 1 (corresponding to 0 to 100% probability) is used to sample each of the distributions; these are combined as in the equation above to give one estimate of the potential of the play. The process is repeated many times (in this case at least 1000 times) to produce multiple estimates of the potential of each play. These are then used to produce a probability versus petroleum potential distribution for each play and for the basin as a whole. Because this is computationally intensive, the calculation is carried out by computer using a commercially available simulator (@RISK, a Microsoft EXCEL spreadsheet add in). This uses a more advanced stratified sampling technique called Latin Hypercube that will converge in fewer iterations than with the traditional Monte Carlo technique.

Anticline to basin area ratio )* This is the proportion of the prospective area that is within an anticlinal trap. It was extrapolated from seismic structure mapping in the Marla and Munta areas (Mackie and Gravestock, 1993; Mackie, 1994). The top crystalline basement depth horizon was used for the Pindyin, Tarlina and Murnaroo plays, and the basal Cambrian depth horizon for the Relief, Ouldburra and Arcoeillinna plays. Although the seismic coverage in this area is the best available in the Officer Basin, it is still poor compared to that required to

DISCUSSION OF PARAMETERS
Prospective area )F This is the area of the basin that is believed to contain the three essential components of source, reservoir and seal, and where the reservoir is at an economically drillable depth (assumed to be <4500 m). This is a critical factor in determining the potential of the basin but can be mapped with reasonable accuracy from the available drillhole, source rock and regional depth to basement seismic mapping data (Lindsay, 1995). It is entered as a uniform distribution (equal probability between minimum and maximum limits). Maps for each play, taking into account distribution of source, seal and reservoir, are shown on Figures 13.1 to 13.4.

Fig. 13.2 Relief Sandstone prospectivity, Officer Basin.

Fig. 13.1 Pindyin, Tarlina Sandstone and Murnaroo Formation prospectivity, Officer Basin.

Fig. 13.3 Ouldburra Formation prospectivity, Officer Basin.

"&

Hydrocarbon saturation (SD) The average hydrocarbon saturation is partly dependent on the average porosity and the pay thickness, and the distributions are linked in the Monte Carlo simulator so that when a low value of porosity and/or pay thickness is chosen a low hydrocarbon saturation is also chosen. The range has been determined from capillary pressure data from the Murnaroo Formation, but modified for the better permeability reservoirs. A truncated normal distribution is used. Formation volume factor (FVF) The volume of oil in a reservoir decreases when brought to the surface due to the drop in pressure, and consequent loss of volatiles. The value for this factor was estimated from Eromanga Basin data (Morton, 1996b). Success ratio (SR) This is an estimate of the proportion of prospects to be drilled that will contain oil (i.e. the drilling success ratio). Like the fill factor (FF) above, this ratio is related in part to the richness of the source rocks, but other factors such as the degree of structural complexity, trap integrity and quality of seismic data are also important. A value was estimated from Australian and world average drilling results, and a truncated normal distribution is used. Recovery factor (RF) Gross reservoir thickness (h) This is the maximum vertical closure of the trap. All reservoirs are modelled as a cone (h is reduced to one-third the volume of a cone = 1/3 x area x height) with the exception of the Ouldburra reservoirs, which are modelled as a slab and h is not reduced. The parameter is modelled as a truncated lognormal distribution. Net to gross pay ratio (NG) The net to gross ratio reduces the maximum reservoir thickness to the anticipated pay (permeable reservoir) thickness. This has been estimated using core data. A truncated normal distribution is used. Anticline fill factor (FF) In oil or gas basins with commercial fields, anticlines can range from filled to spill to near 0% fill (0% = dry wells). The average fill is therefore less than one, and it is assumed that the richer the source rock the greater the average fill. This critical parameter is subjective and has been assumed to average 50% for the Officer Basin, although it would be expected to be greater for plays with rich source potential. A triangular distribution is used. Porosity (Por) The average porosity of the reservoir was estimated from available routine core analysis data (Ch. 10). A truncated normal distribution is used.
"'

Fig. 13.4 Arcoeillinna Sandstone prospectivity, Officer Basin.

identify all of the smaller prospects, and may not be typical of trap styles elsewhere in the basin, hence the values used in this assessment may be conservative. This parameter is entered as a truncated lognormal distribution.

The recovery factor converts petroleum in-place reserves to recoverable oil, and is mostly dependent on the degree of mobility of the underlying aquifer and height of the oil column. Estimates were derived from averages in the Cooper Basin. A truncated normal distribution is used.

POTENTIAL PLAYS
There are six major plays that have potential for discoveries:
1. Pindyin Sandstone

Reservoir: Pindyin Sandstone. Seal: Alinya Formation. Source: overlying Alinya Formation or the underlying, undrilled, possibly pre-Adelaidean sequence. Summary of Monte Carlo input parameters:

Minimum
Prospective area of the basin (km ) Anticline to total basin area ratio Average gross reservoir thickness (m) Net to gross pay ratio Anticline fill factor Porosity (fraction) Water saturation Formation volume factor Exploration drilling success ratio Recovery factor
2. Tarlina Sandstone

Mean
65 874 0.06 69 0.60 0.5 0.12 0.3 0.89 0.10 0.25

Maximum
116 070 0.15 192 0.91 0.99 0.14 0.45 0.91 0.17 0.32

15 620 0.02 13 0.30 0.01 0.10 0.2 0.85 0.04 0.18

Reservoir: Tarlina Sandstone. Seal: Meramangye Formation. Source: Alinya Formation Summary of Monte Carlo input parameters:

Minimum
Prospective area of the basin (km ) Anticline to total basin area ratio Average gross reservoir thickness (m) Net to gross pay ratio Anticline fill factor Porosity (fraction) Water saturation Formation volume factor Exploration drilling success ratio Recovery factor
3. Murnaroo Formation

Mean
65 874 0.06 69 0.55 0.5 0.15 0.41 0.89 0.10 0.18

Maximum
116 108 0.17 194 0.84 0.99 0.17 0.50 0.91 0.18 0.30

15 667 0.02 17 0.30 0 0.13 0.35 0.85 0.04 0.01

Reservoir: Murnaroo Formation sand. Seal: Dey Dey Mudstone. Source: Alinya Formation and Dey Dey Mudstone Summary of Monte Carlo input parameters:

Minimum
Prospective area of the basin (km ) Anticline to total basin area ratio Average gross reservoir thickness (m) Net to gross pay ratio Anticline fill factor Porosity (fraction) Water saturation Formation volume factor Exploration drilling success ratio Recovery factor
4. Relief Sandstone

Mean
65 874 0.06 69 0.87 0.5 0.15 0.35 0.89 0.10 0.25

Maximum
116 059 0.16 194 1.00 0.98 0.18 0.45 0.91 0.17 0.32

15 696 0.02 14 0.55 0.01 0.13 0.25 0.85 0.03 0.18

Reservoir: Relief Sandstone. Seal: Ouldburra Formation or Observatory Hill Formation. Source: Alinya Formation (as shown from oil extracts), Dey Dey Mudstone, Karlaya Limestone, Ouldburra Formation and Observatory Hill Formation. Summary of Monte Carlo input parameters:

Minimum
Prospective area of the basin (km2) Anticline to total basin area ratio Average gross reservoir thickness (m) Net to gross pay ratio Anticline fill factor Porosity (fraction) Water saturation Formation volume factor Exploration drilling success ratio Recovery factor 3459 0.01 15 0.28 0.02 0.09 0.15 0.85 0.04 0.18

Mean
25 711 0.05 69 0.60 0.5 0.18 0.25 0.89 0.11 0.25

Maximum
47 963 0.16 195 0.94 0.98 0.25 0.44 0.91 0.16 0.32

#

5. Ouldburra Formation

Reservoir: Ouldburra Formation sand. Seal: intra-Ouldburra Formation shale and micrite. Source: Ouldburra Formation, with possible contribution from Dey Dey Mudstone, Karlaya Limestone. Summary of Monte Carlo input parameters:

Minimum
Prospective area of the basin (km ) Anticline to total basin area ratio Average gross reservoir thickness (m) Net to gross pay ratio Anticline fill factor Porosity (fraction) Water saturation Formation volume factor Exploration drilling success ratio Recovery factor
6. Arcoeillinna Sandstone

Mean
19 626 0.05 99 0.35 0.50 0.14 0.30 0.89 0.10 0.18

Maximum
28 389 0.21 171 0.64 0.98 0.16 0.40 0.91 0.17 0.3

10 860 0.01 45 0.25 0.01 0.12 0.20 0.85 0.03 0.01

Reservoir: Arcoeillinna Sandstone. Seal: Apamurra Formation. Source: Observatory Hill Formation and Ouldburra Formation. Summary of Monte Carlo input parameters:

Minimum
Prospective area of the basin (km ) Anticline to total basin area ratio Average gross reservoir thickness (m) Net to gross pay ratio Anticline fill factor Porosity (fraction) Water saturation Formation volume factor Exploration drilling success ratio Recovery factor
2

Mean
13 134 0.05 50 0.65 0.5 0.175 0.35 0.89 0.10 0.18

Maximum
20 893 0.16 92 0.96 0.99 0.19 0.50 0.91 0.17 0.3

5387 0.01 25 0.31 0.02 0.16 0.25 0.85 0.03 0.01

Additional potential may exist in carbonate reservoirs of the Ouldburra Formation (Kamali et al., 1995b), and sand of the Tanana and Narana Formations. The Trainor Hill and Mount Chandler Sandstones (sealed by Indulkana Shale) are also a possible play, but seal and source are significant risks. The table below summarises the assessment of the undiscovered recoverable oil potential of the Officer Basin in South Australia at various probability levels.
Play Probability that the ultimate potential will exceed the stated value million kilolitres (million barrels) 90% 50% 10% 10.8 (68.1) 85.1 (535.5) 30.3 (190.7) 69.1 (434.6) 63.7 (400.6) 62.6 (394.1) 399.0 (2510.0) 27.8 209.9 95.0 210.4 186.1 172.9 674.0 (174.9) (1320.4) (597.8) (1323.7) (1170.7) (1087.4) (4239.6)

Arcoeillinna 3.3 (20.6) Ouldburra 29.4 (185.1) Relief 6.3 (39.8) Murnaroo 21.7 (136.7) Tarlina 18.6 (117.2) Pindyin 18.0 (113.2) Total 236.4 (1486.9)

151

#

Appendix 1.1 Abbreviations used throughout the text


Organisations and projects Conversions

AGSO AP BMR GPA MESA MT NGMA SAEI SASE


General

Australian Geological Survey Organisation Anangu Pitjantjatjara Bureau of Mineral Resources (now AGSO) Glen Parkinson and Associates Mines and Energy Resources South Australia Maralinga Tjarutja National Geoscience Mapping Accord South Australian Exploration Initiative South Australian steel and energy project

C = ((F - 32).5)/9 1 Petajoule (PJ) = 9.4781 1011 BTU US$1 = A$0.75 1 cubic metre (m3) = 1 kilolitre (kL) 1 standard cubic metre of gas (m3) = 5.6154 standard cubic feet of gas 1 kilolitre (kL) = 6.29 US barrels 1 kilopascal (kPa) = 0.1450 pound-force per square inch (psi)
Radiometric dating

BIF CEF DEF LPG OEL OOIP OPL PEL PPL TD XRD ZOCA

banded iron formation Code of Environmental Practice Declaration of Environmental Factors Liquefied Petroleum Gas Oil Exploration Licence original oil in place Oil Production Licence Petroleum Exploration Licence Petroleum Production Licence total depth X-ray diffraction mineral analysis zone of cooperation (Timor Sea)

KAr RbSr UPb

Potassium 40Argon 40 Rubidium 87Strontium 87 Uranium 235, 238Lead 207, 206

Source rock and maturity parameters

EOM HI MPI MPR OI VRcalc Tmax TOC

extractable organic matter Hydrogen Index methylphenanthrene index methylphenanthrene ratio Oxygen Index calculated equivalent vitrinite reflectance temperature of maximum generation of S2 hydrocarbons total organic carbon

Reservoir, engineering and financial parameters Measurement

C API BOPD ha kL kPa L/s Ma md mmbbl ms /ft ppm rb SCF stb

degrees Celsius (temperature) gamma ray log units barrels of oil per day hectare (area; = 104 m2) kilolitre (volume; = 1 m3) kilopascal (pressure; = 1 kg/m.s2) Litres per second million years before present millidarcies million barrels milliseconds microseconds per foot parts per million (= milligrams per litre) reservoir barrels standard cubic feet (gas) stock tank barrel (oil)

CAPEX FVF k kh NPV OPEX OWC Swirr

capital expenditure formation volume factor (stb/rb) permeability (md) reservoir flow capacity net present value operating expenditure oil-water contact irreducible water saturation

Sequence stratigraphy

HST IVF LST MFS TST

high stand system tract incised valley fill low stand system tract maximum marine flooding surface transgressive system tract

#!

Appendix 3.1 Sections of the

Pitjantjatjara Land Rights Act 1981 relevant to petroleum

exploration in the Officer Basin

#"

##

#$

#%

#&

Appendix 3.2 Sections of the

Maralinga Tjarutja Land Rights Act 1984 relevant to

petroleum exploration in the Officer Basin

#'

$

$

$

$!

$"

APPENDIX 12.1 Economic model data and assumptions

Costs

The following assumptions have been made (all cash values below are Australian dollars unless indicated): Conversion from US$ to A$: Crude oil price: Real discount rate: $0.79 (December 1996January 1997). High case of US$25/bbl (January 1997) and low case of US$18/bbl were selected. 12.5% (before tax). 4297 km2. $1.015 million. $3500/km. Existing regional seismic grid = 1808 km; New exploration seismic = 570 km. 475 km.

Exploration
PEL area in model: Cost of exploration well: Cost of seismic: Seismic data: Seismic per exploration well:

Capital expenditure
Cost of development well: Road construction: $0.832 million. The Department of Transport provided rough estimates for road construction costs in outback South Australia (1997). Total cost = $21 500/km which includes: $20 000 forming and sheeting road $25 000 per water bore with four bores per 100 km of road $50 000 per 100 km Aboriginal and environmental clearances. For the model, wells are located on or within 20 km of the existing MESA seismic grid, which was left as tracks at the request of the landholder. 30 m3/tank (190 bbl/tank; GPA, 1996), up to a maximum of three tanks. $70/km (GPA 1996). $94/m (GPA, 1996); trunkline capacity = 790 bbl/day (125.6 m3/day). $187/m (capital costs for MoombaPort Bonython trunkline; McDonough, 1996; Section 5.0); trunkline capacity = 33 000 bbl/day (5250 m3/day). $350 000 (BRS, 1996). $102 000 (GPA, 1996). $150 000 (GPA, 1996). US$18.75/bbl (BRS, 1996). $50 000 (typical 4WD), replaced every three years.

Road tanker capacity: Flowlines (100 mm): Trunkline (150 mm): Trunkline (300 mm): Completion cost per well: Pumps: Wellhead: Oil storage tanks: Vehicle:

Operating costs
Trucking cost: Downhole well maintenance: Pump maintenance: Pump fuel cost: Flowline maintenance: Trunkline maintenance: Marla operation: Head office: Road maintenance: $10.80/bbl (McDonough, 1996). $45 000/well/year (R. McDonough, MESA, pers. comm., 1996). 8% CAPEX/year (GPA, 1996). $1500/hp/year (GPA, 1996). 3% CAPEX (GPA, 1996). 2.5% CAPEX (GPA, 1996). $50 000/year. $150 000/year. $1100 average cost of patrol grading to suitable standard for heavy vehicles (Department of Transport, 1997); the range is $8001400/km/year, depending on weather conditions.

$#

Reservoir engineering data


The spreadsheet Oil_prod.xls calculates total production rate, annual and cumulative production, wells needed and the recovery factor. Max. wells: Max. field rate: Field area: Pay thickness: Original Oil in Place: Recovery factor: Reservoir depth: Reservoir pressure: Bottomhole pressure: Reservoir kh: Water gradient: Temperature gradient: Surface temperature: Reservoir temperature: Oil API gravity: GOR: Gas gravity: Initial Oil FVF: Rock compressibility Water compressibility Water saturation Five (discovery well and four development wells). 1000 bopd (158.9 m3/day). Case 1: 75 km2 , Case 2: 150 km2 (the average area of prospects and leads delineated by Mackie (1994) is 97 km2; area ranges from 16 to 299 km2). The field is assumed to be circular. 60 m (average closure mapped by Mackie (1994) is 173 m, closure ranges from 125 to 750 m). Two cases were modelled: Case 1 = 530 mmbbl, Case 2 = 50200 mmbbl. Note that original oil in place is referred to throughout the text, not recoverable oil. Oil recoveries did not exceed 25%. 1750 m (5743 ft), primary reservoir Murnaroo Formation (D. Gravestock, MESA, pers. comm., 1996). 17 427 kPa (2527 psia), based on hydrostatic gradient. 3250 kPa (500 psia), assumed. 250, 1000, 9200 and 17 700 md.ft were selected. 0.44 psi/ft (assumed). 20C/km. 25C. 60C (140F, 600 R). 50 (based on Mereenie Pacoota P3). 800 scf/stb (based on Mereenie Pacoota P3). 0.76 (based on Mereenie Pacoota P3). 1.42 rb/stb (calculated from correlation). The Marla region is overmature and oil here is likely to be more gas-rich than other parts of the Officer Basin. 4x10-6.psi-1 (assumed). 3x10-6.psi-1 (assumed). 50% (assumed).

Oil production above the bubble point is calculated using the material balance equation. Oil production below the bubble point is calculated using the Tracy Material Balance Method (Smith and Tracy, 1986). For a given pressure decrement, the oil production volume is calculated. Individual well flow rates are determined based on reservoir pressure and flowing bottomhole pressure. The number of wells required (up to a maximum of five) to produce at the target field flow rate is then calculated. Based on the total rate, the time to produce the oil volume for the pressure decrement is calculated to develop a oil production profile with time. Oil properties are based on Mereenie Field data, (oil gravity, gas gravity, GOR). Correlations are then used to calculate bubble point pressure, oil formation volume factor, oil viscosity, etc.

$$

Alexander, E.M., Pegum, D., Tingate, P., Staples, C.J., Michaelsen, B.H. and McKirdy, D.M., 1996. Petroleum potential of the Eringa Trough in SA and the NT. APPEA Journal, 36:322-348. Alexander, E.M. and Sansome, A., 1996. Lithostratigraphy and environments of deposition. In: Alexander, E.M. and Hibburt, J.E. (Ed.), The petroleum geology of South Australia. Vol. 2: Eromanga Basin. South Australia. Department of Mines and Energy. Report Book, 96/20, pp.49-86. Ambrose, G.J., Flint, R.B. and Webb, A.W., 1981. Precambrian and Palaeozoic geology of the Peake and Denison Ranges. South Australia. Geological Survey. Bulletin, 50:5-71. Ashley, J., 1984. Northeast Officer Basin SA. Interpretation of aeromagnetic data. South Australia. Department of Mines and Energy Resources. Open file Envelope, 5073:347-364 (unpublished). Australian Gas Association, 1988. Gas supply and demand study to 2030. A public report. Pirie Printers Sales Pty Ltd, Canberra. Baas Becking Geobiological Laboratory, 1983. Isotopic studies of palaeoenvironments, Officer Basin, SA. Quarterly report to Comalco Aluminium Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 5073:582-584 (unpublished). Badley, M., 1988. Officer Basin. A summary report on a brief review seismic interpretation. Report for Comalco Aluminium Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 5073:1228-1242 (unpublished). Basedow, H., 1905. Geological report on the country traversed by the South Australian Government North-West Prospecting Expedition, 1903. Royal Society of South Australia. Transactions, 29:57-102. Basedow, H., 1914. Journal of the Government North West Expedition. Royal Geographical Society of Australasia. South Australian Branch. Proceedings, 15:57-242. Beardsmore, G.R. and OSullivan, P.B., 1995. Uplift and erosion on the Ashmore Platform, North West Shelf: conflicting evidence from maturation indicators. APEA Journal, 35:333-343. Bell, S., 1997. Analysts see oil prices holding above $20 in 1997. Oil and Gas Gazette, December/January 1996/97:33-37. Benbow, M.C., 1982. Stratigraphy of the Cambrian?Early Ordovician Mount Johns Range, NE Officer Basin, South Australia. Royal Society of South Australia. Transactions, 106:191-211. Benbow, M.C., 1993. TALLARINGA, South Australia, sheet SH53-5. South Australia. Geological Survey. 1:250 000 Series Explanatory Notes. Benbow, M.C., 1995. Near-surface geological cross sections (1:200 000). In: Lindsay, J.F. (Ed.), Geological atlas of the Officer Basin, South Australia. Australian Geological Survey Organisation and Department of Mines and Energy, South Australia. Benbow, M.C., Lindsay, J.M. and Alley, N.F., 1995. Eucla Basin and palaeodrainage. In: Drexel, J.F. and Preiss, W.V. (Eds), The geology of South Australia. Vol. 2, The Phanerozoic. South Australia. Geological Survey. Bulletin, 54:178-186. Benbow, M.C. and Pitt, G.M., 1979. Byilkaoora No. 1 well completion report. South Australia. Department of Mines and Energy. Report Book, 79/115. Bourke, D.J. and Senapati, N., 1983. Technical report on the drilling for the 1983 coal exploration programme. Report for Comalco. South Australia. Department of Mines and Energy Resources. Open file Envelope, 6259:1108-1154 (unpublished). Brewer, A.M., 1984. Base metal geochemical data from eastern Officer Basin drillholes. Report for Comalco Aluminium Ltd. South Australia.

Department of Mines and Energy Resources. Open file Envelope, 3938:760-773 (unpublished). Brewer, A.M., Dunster, J.N., Gatehouse, C.G., Henry, R.L. and Weste, G., 1987. A revision of the stratigraphy of the eastern Officer Basin. South Australia. Geological Survey. Quarterly Geological Notes, 102:2-15. Brown, H.Y.L., 1896. Auriferous deposits of Western Australia. South Australia. Parliamentary Papers, 26. Brown, H.Y.L., 189899. Government Geologists report on exploration in western part of South Australia. South Australia. Parliamentary Papers, 46. Brown, H.Y.L., 1905. Report on geological explorations in the west and north-west of South Australia. South Australia. Parliamentary Papers, 71. Bureau of Resource Sciences, 1996. Development economics of onshore petroleum prospects in the Northern Territory. Bureau of Resource Sciences. Report (unpublished). Carruthers, J., 1892. Triangulation of N.W. portion of South Australia. South Australia. Parliamentary Papers, 179. Chamalaun, F.H., von der Borch, C.C. and Gatehouse, C.G., 1990. Magnetostratigraphic study of late Adelaidean sediments in the eastern Officer Basin, South Australia. Mines and Energy Review, South Australia, 157:36-39. Coats, R.P., 1963. The geology of the Alberga 4-mile military sheet. South Australia. Geological Survey. Report of Investigations, 22. Compston, W., 1974. The Table Hill Volcanics of the Officer Basin Precambrian or Palaeozoic? Geological Society of Australia. Journal, 21(4):403-411. Conoco, 1969. Munyarai No. 1 Well, South Australia, stratigraphic drilling project well completion report. Continental Oil Co. of Australia Pty Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 979. Cowley, W.M. and Flint, R.B., 1993. Epicratonic igneous rocks and sediments. In: Drexel, J.F., Preiss, W.V. and Parker, A.J. (Eds), The geology of South Australia. Vol. 1, The Precambrian. South Australia. Geological Survey. Bulletin, 54:142-147. Crick, I.H., 1992. Petrological and maturation characteristics of organic matter from the Middle Proterozoic McArthur Basin, Australia. Australian Journal of Earth Sciences, 39:501-519. Crimes, T.P., 1975. The stratigraphical significance of trace fossils. In: Frey, R.W. (Ed.), The study of trace fossils. Springer-Verlag, New York, pp.109-130. Cruse, T., Harris, L.B. and Rasmussen, B., 1993. Geological note: the discovery of Ediacaran trace and body fossils in the Stirling Range Formation, Western Australia: implications for sedimentation and deformation during the Pan-African orogenic cycle. Australian Journal of Earth Sciences, 40(3):293-296. Cucuzza, J., 1987. Munta 1 well completion report, PEL 23, northeastern Officer Basin. Report for Comalco Aluminium Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 7090 (unpublished). Daly, S., 1996. The Thompson Nickel Belt and other new exploration models for the west Gawler Craton. In: Resources 96 Convention Abstracts. South Australia. Department of Mines and Energy, pp.49-51. Demaison, G.J. and Moore, G.T., 1980. Anoxic environments and oil source bed genesis. AAPG Bulletin, 64(8):1179-1209. Dunster, J.N., 1987a. Sedimentology of the Ouldburra Formation (Early Cambrian) northeastern Officer Basin. University of Adelaide. M.Sc. thesis (unpublished).

$&

Dunster, J.N., 1987b. Karlaya No. 1 well completion report for Comalco Aluminium Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 7040 (unpublished). Dunster, J.N., 1987c. Lairu 1 well completion report, PEL 23, northeastern Officer Basin. Report for Comalco Aluminium Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 7041 (unpublished). Edgell, H.S., 1991. Proterozoic salt basins of the Persian Gulf area and their role in hydrocarbon generation. Precambrian Research, 54:1-14. Fanning, C.M., 1989. UPb dating of the Boucaut Volcanics and Bendigo Granite. Amdel report, G7948/89 (unpublished). Feld, L., 1997. Oman DOE/EIA. (WWW site http://www.eia.doe.gov/ emeu/cabs/oman.html). Ferguson, J. and Skyring, G.W., 1995. Redbed-associated sabkhas and tidal flats at Shark Bay, Western Australia: their significance for genetic models of stratiform Cu-(Pb-Zn) deposits. Australian Journal of Earth Sciences, 42:321-333. Finlayson, B., 1979. A geophysical interpretation with depths to magnetic basement TALLARINGA, GILES, MURLOOCOPPIE 1:250 000 sheets. South Australia. Department of Mines and Energy. Report Book, 79/90. Flint, R.B. and Daly, S.J., 1993. Coompana Block. In: Drexel, J.F., Preiss, W.V. and Parker, A.J. (Eds), The geology of South Australia. Vol. 1, The Precambrian. South Australia. Geological Survey. Bulletin, 54:168-169. Flint, R.B., Fanning, C.M., and Rankin, L.R., 1988. The late Proterozoic Kilroo Formation of the Polda Basin. South Australia. Geological Survey. Quarterly Geological Notes, 106:16-23. Flint, D.J., Gravestock, D.I. and Newton, A.W., 1989. Maralinga Lands: assessment of mineral potential. South Australia. Department of Mines and Energy. Report Book, 89/14. Gara, T., 1994. Tackling the back country: Richard Maurices expeditions in the Great Victoria Desert, 18971903. South Australian Geographical Journal, 93:42-60. Gatehouse, C., 1976. A fossil in the Observatory Hill Beds, South Australia. South Australia. Geological Survey. Quarterly Geological Notes, 60:5-8. Gatehouse, C.G., 1979. Well completion report. Wilkinson No. 1. South Australia. Department of Mines and Energy. Report Book, 79/88. Gatehouse, C.G., Benbow, M.C. and Major, R.B., 1986. The Murnaroo Formation of the Officer Basin. South Australia. Geological Survey. Quarterly Geological Notes, 97:17-20. Gatehouse, C.G., Fanning, C.M. and Flint, R.B., 1995. Geochronology of the Big Lake Suite, Warburton Basin, northeastern South Australia. South Australia. Geological Survey. Quarterly Geological Notes, 128:8-16. Gatehouse, C.G. and Hibburt, J.E., 1987. Observatory Hill 1 well completion report and geology of adjacent Observatory Hill Formation. South Australia. Department of Mines and Energy. Report Book, 87/58. Gaughan, C.J., 1989. Early Cambrian Relief Sandstone, Officer Basin, South Australia: subdivision, diagenesis and porositypermeability distribution. University of Adelaide. B.Sc. (Hons) thesis (unpublished). Gaughan, C.J. and Warren, J.K., 1990. Lower Cambrian Relief Sandstone, eastern Officer Basin, South Australia: an example of secondary porosity development. APEA Journal, 30(1):184-195. George, F.R., 1904. Extracts from journals of explorations by R.T. Maurice, Fowlers Bay to Rawlinson Ranges and Fowlers Bay to Cambridge Gulf. With plans. South Australia. Parliamentary Papers, 43. George, F.R., 1905. Prospecting expedition north of Nullarbor Plains. South Australia. Parliamentary Papers, 60. George, F.R., 1907. Journal (with plans) of the Government Prospecting Expedition to the south-western portions of the Northern Territory, by F.R. George; and to the Buxton and Davenport Ranges by W.R. Murray. South Australia. Parliamentary Papers, 50. Geotrack, 1994. Apatite fission track analysis of fourteen well samples. Geotrack report 531. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Giles, E., 1889. Australia twice traversed: the romance of exploration, being a narrative compiled from the journals of five exploring expeditions into and through Central South Australia, and Western Australia, from 1872 to 1876. Sampson Low, Marston, Searle and Rivington, London, (2 vols).

Gosse, W.C., 1874. W.C. Gosses explorations, 1873. South Australia. Parliamentary Papers, 48. GPA Engineering Pty Ltd, 1996. Cooper Basin oilfield development study. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8693 (unpublished). Grasso, R., 1963. Final well report of Emu No. 1 well. South Australia. Department of Mines and Energy Resources. Open file Envelope, 362 (unpublished). Gravestock, D.I., Benbow, M.C., Gatehouse, C.G. and Krieg, G.W., 1995. Eastern Officer Basin. In: Drexel, J.F. and Preiss, W.V. (Eds), The geology of South Australia. Vol. 2, The Phanerozoic. South Australia. Geological Survey. Bulletin, 54:35-41. Gravestock, D.I. and Hibburt, J.E., 1991. Sequence stratigraphy of the eastern Officer and Arrowie Basins: a framework for Cambrian oil search. APEA Journal, 31(1):177-190. Gravestock, D.I. and Lindsay, J.F., 1994. Summary of 1993 seismic exploration in the Officer Basin, South Australia. PESA Journal, 22:65-75. Gravestock, D.I. and Sansome, A., 1994. Eastern Officer Basin geology and hydrocarbon potential. South Australia. Department of Mines and Energy. Open file Envelope, 8591 (unpublished). Grey, K., 1995. Neoproterozoic stromatolites from the Skates Hills Formation, Savory Basin, Western Australia, and a review of the distribution of Acaciella australica. Australian Journal of Earth Sciences, 42:123-132. Hamer, K.L., 1994. Aeromagnetic interpretation of east Officer Basin (1:250 000 scale). Report for World Geoscience Corporation (unpublished). Henderson, S.W. and Tauer, R.W., 1967. Birksgate No. 1 well, South Australia, stratigraphic drilling project, well completion report. South Australia. Department of Mines and Energy Resources. Open file Envelope, 768 (unpublished). Henry, R.L., 1986. Ungoolya No. 1 well completion report for Comalco Aluminium Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 6764 (unpublished). Henry, R.L. and Brewer, A.M., 1984. Revisions to the Cambrian stratigraphy of the northeastern Officer Basin, South Australia. South Australia. Department of Mines and Energy Resources. Open file Envelope, 3938:848-860 (unpublished). Hibburt, J.E., 1984. Review of exploration activity in the Arckaringa Basin region, 18581983. South Australia. Department of Mines and Energy. Report Book, 84/1. Hibburt, J.E., Hough, L.P. and Tucker, L.R., 1995. Hydrocarbon shows in Comalco mineral drillholes, Officer Basin. South Australia. Department of Mines and Energy. Report Book, 95/11. Hoffman, P.F., 1991. Did the breakout of Laurentia turn Gondwanaland inside-out? Science, 252:1409-1412. Hoskins, D. and Lemon, N., 1995. Tectonic development of the eastern Officer Basin, central Australia. Exploration Geophysics, 26:395-402. Hbbe, S.G., 1897. Stock Route Expedition from South to West Australia. South Australia. Parliamentary Papers, 51. Hutton, A.C. and Cook, A.C., 1980. Influence of alginite on the reflectance of vitrinite from Joadja, NSW, and some other coals and oil shales containing alginite. Fuel, 59(10):711-714. Jack, R.L., 1915. The geology and prospects of the region to the south of the Musgrave Ranges, and the geology of the western portion of the Great Artesian Basin. South Australia. Geological Survey. Bulletin, 5. Jack, R.L., 1919. Report on proposed area for reserve. Commissioner of Public Works correspondence. South Australia. State Records file, GRG23/1/1917/643 (unpublished). Jack, R.L., 1931. Report on the geology of the region to the north and north-west of Tarcoola. South Australia. Geological Survey. Bulletin, 15. Jackson, K.S., McKirdy, D.M. and Deckelman, J.A., 1984. Hydrocarbon generation in the Amadeus Basin, central Australia. APEA Journal, 24(1):42-65. Jackson, M.J. and van de Graaff, W.J.E., 1981. Geology of the Officer Basin. Bureau of Mineral Resources, Geology and Geophysics, Australia. Bulletin, 206. Jacobson, G., Barnes, C.J., Fifield, L.K. and Cresswell, R.G., 1994. The time factor in arid-zone groundwater recharge. In: Water down under 94. The

$'

Insitution of Engineers Australia and International Association of Hydrogeologists, Vol. 2(B), pp.471-478. Jago, J.B., Tian-rui Lin and Dunster, J.N., 1994. Early Cambrian trilobites from the Ouldburra Formation, Manya 6, eastern Officer Basin. PESA Journal, 22:87. Jago, J.B. and Youngs, B.C., 1980. Early Cambrian trilobites from the Officer Basin, South Australia. Royal Society of South Australia. Transactions, 104(6):197-199. Jenkins, R.J.F., 1992. Functional and ecological aspects of Ediacaran assemblages. In: Lipps, J.H. and Signor, P.W. (Eds), Origins and early evolution of the Metazoa. Plenum Press, New York, pp.131-176. Jenkins, R.J.F., McKirdy, D.M., Foster, C.B., OLeary, T. and Pell, S.D., 1992. The record and stratigraphic implications of organic-walled microfossils from the Ediacaran (terminal Proterozoic) of South Australia. Geological Magazine, 129(4):401-410. Kamali, M.R., 1995a. Petroleum geochemistry of Early Cambrian carbonate source rocks, Ouldburra Formation, eastern Officer Basin. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Kamali, M.R., 1995b. Sedimentology and petroleum geochemistry of the Ouldburra Formation, eastern Officer Basin, Australia. Ph.D. thesis. National Centre for Petroleum Geology and Geophysics. The University of Adelaide. South Australia Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Kamali, M.R., Lemon, N.M. and Apak, S.N., 1995. Porosity generation and reservoir potential of Ouldburra Formation carbonates, Officer Basin, South Australia. APEA Journal, 35(1):106-120. Kamali, M.R., Lemon, N.M. and McKirdy, D.M. 1993. Ouldburra Formation as a potential source and reservoir for petroleum in the Manya Trough, eastern Officer Basin. In: Alexander, E.M. and Gravestock, D.I. (Eds), Central Australian Basins Workshop, Alice Springs, 13-14 September 1993. Programme and Abstracts, pp.65-66. Korsch, R.J., Shaw, R.D. and Goleby, B.R., 1993. Structural and tectonic framework of the Amadeus Basin, with some deep seismic examples. In: Alexander, E.M. and Gravestock, D.I. (Eds), Central Australian Basins Workshop, Alice Springs, 1993. Programme and Abstracts, pp.22-28. Krieg, G.W., 1967. Continental stratigraphic well Officer No. 1, well completion report. South Australia. Department of Mines. Report Book, 744. Krieg, G.W., 1969. Geological developments in the eastern Officer Basin of South Australia. APEA Journal, 9:8-13. Krieg, G.W., 1972. The Ammaroodinna Inlier. South Australia. Geological Survey. Quarterly Geological Notes, 41:3-7. Krieg, G.W., 1973. EVERARD, South Australia, sheet SH53-13. South Australia. Geological Survey. 1:250 000 Series Explanatory Notes. Krieg, G.W., 1993. Basement inliers southeast of the Musgrave Block. In: Drexel, J.F., Preiss, W.V. and Parker, A.J. (Eds), The geology of South Australia. Vol. 1, The Precambrian. South Australia. Geological Survey. Bulletin, 54:168. Krieg, G.W., Jackson, M.J. and van de Graaff, W.J.E., 1976. Officer Basin. In: Leslie, R.B., Evans, H.J. and Knight, C.L. (Eds), Economic geology of Australia and Papua New Guinea, 3, Petroleum. Australasian Institute of Mining and Metallurgy. Monograph Series, 7:247-253. Kuznetsov, V.G., Ilyukhin, L.N., Miller, S.A., Moskovkina, E.Yu. and Postnikova, O.V., 1992. Paleogeography of boundary sediments of Vendian and Cambrian of the south Siberian Platform. Izvestie Akademiya Nauk, Rossiya Seriya Geologiya, 5:68-83. English translation in Clarke, J., 1996. Petroleum Geology, 30(1):21-36. Lau, J.E., Dodds, A.R., Tewkesbury, P. and Jacobson, G., 1995a. Groundwater systems (Scale 1:1 000 000). In: Lindsay, J.F. (Ed.), Geological atlas of the Officer Basin, South Australia. Australian Geological Survey Organisation, Canberra, and South Australian Department of Mines and Energy, Adelaide, Plate 9. Lau, J.E., Dodds, A.R., Tewkesbury, P. and Jacobson, G., 1995b. Groundwater salinity (Scale 1:1 000 000). In: Lindsay, J.F. (Ed.), Geological atlas of the Officer Basin, South Australia. Australian Geological Survey Organisation, Canberra, and South Australian Department of Mines and Energy, Adelaide, Plate 10. Lester, Y., 1993. Yami. The autobiography of Yami Lester. Institute for Aboriginal Development Publications, Alice Springs.

Leven, J.H. and Lindsay, J.F., 1992. Morphology of the Late Proterozoic to Early Palaeozoic Officer Basin, South Australia. Exploration Geophysics, 23:191-196. Li, Z.X., Zhang, L. and Powell, C.McA., 1996. Positions of the East Asian cratons in the Neoproterozoic supercontinent Rodinia. Australian Journal of Earth Sciences, 43(6):593-604. Limb, N., 1980. A geophysical interpretation of the Tallaringa Trough and Karari Fault Zone. South Australia. Department of Mines and Energy. Report Book, 80/123. Lindsay, D., 1891. Journal of the Elder Exploring Expedition, 1891. South Australia. Parliamentary Papers, 45. Lindsay, J.F. (Ed.), 1995. Geological atlas of the Officer Basin, South Australia. Australian Geological Survey Organisation and Department of Mines and Energy, South Australia. Lindsay, J.F. and Korsch, R.J., 1991. The evolution of the Amadeus Basin, central Australia. In: Korsch, R.J. and Kennard, J.M. (Eds), Geological and geophysical studies of the Amadeus Basin, central Australia. Bureau of Mineral Resources, Geology and Geophysics, Australia. Bulletin, 236:7-32. Lindsay, J.F. and Leven, J.H., 1996. Evolution of a Neoproterozoic to Palaeozoic intracratonic setting, Officer Basin, South Australia. Basin Research, 8:403-424. Long, J.A., Turner, S. and Young, G.C., 1988. A Devonian fish fauna from subsurface sediments in the eastern Officer Basin, South Australia. Alcheringa, 12:61-78. MacKenzie, B.W. and Cai, Z., 1993. Economics of petroleum exploration in Australia. Centre for Resource Studies, Ontario, and Australian Mineral Foundation, Adelaide. Mackie, S.M., 1994. Summary of seismic interpretation, Marla and Munta areas, Officer Basin. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Mackie, S.M. and Gravestock, D.I., 1993. Summary of seismic interpretation, Marla area Officer Basin. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Major, R.B., 1973a. The Wirrildar Beds. South Australia. Geological Survey. Quarterly Geological Notes, 45:8-11. Major, R.B., 1973b. The Pindyin Beds. South Australia. Geological Survey. Quarterly Geological Notes, 46:1-5. Major, R.B., 1973c. The Wright Hill Beds. South Australia. Geological Survey. Quarterly Geological Notes, 46:6-10. Major, R.B., 1973d. WOODROFFE, South Australia, sheet SG52-12. South Australia. Geological Survey. 1:250 000 Series Explanatory Notes. Major, R.B., 1974. The Punkerri Beds. South Australia. Geological Survey. Quarterly Geological Notes, 51:1-5. Major, R.B. and Conor, C.H.H., 1993. Musgrave Block. In: Drexel, J.F., Preiss, W.V. and Parker, A.J. (Eds), The geology of South Australia. Vol. 1, The Precambrian. South Australia. Geological Survey. Bulletin, 54:156-167. Major, R.B. and Teluk, J.A., 1967. The Kulyong Volcanics. South Australia. Geological Survey. Quarterly Geological Notes, 22:8-11. Manning, G.H., 1990. Mannings place names of South Australia. G.H. Manning, Adelaide. Mason, M.G., Thomson, B.P. and Tonkin, D.G., 1978. Regional stratigraphy of the Beda Volcanics, Backy Point Beds and Pandurra Formation on the southern Stuart Shelf, South Australia. South Australia. Geological Survey. Quarterly Geological Notes, 66:2-9. Mattingley, C. and Hampton, K., 1988. Survival in our own land: Aboriginal experiences in South Australia since 1836. ALDAA/ Hodder and Stoughton, Rydalmere, NSW. McCaffrey, M.A., Moldowan, J.M., Lipton, P.A., Summons, R.E., Peters, K.E., Jeganathan, A. and Watt, D.S., 1993. 24-Isopropylcholestanes: possible sponge biomarkers in sediments and petroleum. In: Alexander, E.M. and Gravestock, D.I. (Eds), Central Australian Basins Workshop, Alice Springs, 13-14 September 1993. Programme and Abstracts, pp.37-38. McDonough, R.C., 1996. Economics of gas gathering and processing in the Cooper Basin. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8693 (unpublished). McDonough, R.C., in press. Economics of gas field developments in the Cooper Basin after 1999. APPEA Journal, 37(1).

%

McKirdy, D.M., 1993. Oil shows and source rocks of the eastern Officer Basin a review. In: Alexander, E.M. and Gravestock, D.I. (Eds), Central Australian Basins Workshop, Alice Springs, 13-14 September 1993. Programme and Abstracts, pp.68-70. McKirdy, D.M., Aldridge, A.K. and Ypma, P.J.M., 1983. A geochemical comparison of some crude oil from pre-Ordovician carbonate rocks. In: Bjory, M. et al. (Eds), Advances in inorganic geochemistry 1981. John Wiley and Sons Ltd, Chichester. pp.99-107. McKirdy, D.M. and Kantsler, A.J., 1980. Oil geochemistry and potential source rocks of the Officer Basin, South Australia. APEA Journal, 20(1):68-86. McKirdy, D.M., Kantsler, A.J., Emmett, J.K. and Aldridge, A.K., 1984. Hydrocarbon genesis and organic facies in Cambrian carbonates of the eastern Officer Basin, South Australia. In: Palacas, J.G. (Ed.), Petroleum geochemistry and source rock potential of carbonate rocks. AAPG Studies in Geology, 18:13-31. McKirdy, D.M. and Michaelsen, B.H., 1994. Geochemical measurement of thermal maturity in Neoproterozoic and Cambrian sediments, eastern Officer Basin. Report for South Australian Department of Mines and Energy. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). McKirdy, D.M. and Watson, B.L., 1989. Biomarker geochemistry of oil shows in the Late Precambrian Murnaroo Formation, Lake Maurice West 1, Officer Basin, SA. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8488:65-96 (unpublished). McRae, H., Nettheim, G. and Beacroft, L., 1991. Aboriginal legal issues: commentary and materials. The Law Book Company, Sydney. Meyerhoff, A.A., 1982. Petroleum basins of the Union of Socialist Soviet Republics and the Peoples Republic of China. PESA Distinguished Lecture Series. Michaelsen, B.H., Kamali, M.R. and McKirdy, D.M., 1995. Unexpected molecular fossils from Early Cambrian carbonates. Proceedings of the 1995 Organic Geochemistry Conference. The University of Adelaide, p.46. Milton, B.E., 1974. Seismic experiments in the northern Eucla Basin, South Australia. South Australia. Geological Survey. Quarterly Geological Notes, 51:5-9. Milton, B.E., 1975. Reconnaissance seismic exploration, southwest Arckaringa Basin, South Australia, 1974. South Australia. Geological Survey. Quarterly Geological Notes, 53:3-9. Milton, B.E. and Parker, A.J., 1973. An interpretation of geophysical observations on the northern margin of the eastern Officer Basin. South Australia. Geological Survey. Quarterly Geological Notes, 46:10-14. Moores, E.M., 1991. Southwest U.S.-East Antarctic (SWEAT) connection: a hypothesis. Geology, 19:425-428. Morrell, G.R. (Ed.), 1995. Petroleum exploration in northern Canada. A guide to oil and gas exploration and potential. Northern Oil and Gas Directorate. Indian and Northern Affairs, Canada. Morton, J.G.G., 1992. Undiscovered petroleum assessments of the potential of the Officer and Otway Basins in South Australia. Mines and Energy Review, South Australia, 158:40-47. Morton, J.G.G., 1995. Undiscovered reserves. In: Morton, J.G.G. and Drexel, J.F. (Eds), Petroleum geology of South Australia, Vol. 1: Otway Basin. South Australia. Department of Mines and Energy Report Book, 95/12. Morton, J.G.G., 1996a. 1996 MESA Petroleum Industry Survey. South Australia. Department of Mines and Energy. Report Book, 96/34. Morton, J.G.G., 1996b. Undiscovered petroleum resources. In: Alexander, E.M. and Hibburt, J. (Eds), Petroleum geology of South Australia, Vol. 2, Eromanga Basin. South Australia. Department of Mines and Energy. Report Book, 96/20. Morton, P., 1989. Fire across the desert: Woomera and the Anglo-Australian Joint Project 19451980. AGPS, Canberra. Moussavi-Harami, R., 1994. Burial history analysis of the east Officer Basin, South Australia. A preliminary study. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Moussavi-Harami, R. and Gravestock, D.I., 1995. Burial history of the eastern Officer Basin, South Australia. APEA Journal, 35(1):307-320. Mumme, I.A., 1961. Geophysical survey of the Officer Basin, South Australia. South Australia. Department of Mines. Unpublished file, SR11/5/48.

ONeil, B., 1982. In search of mineral wealth: the South Australian Geological Survey and Department of Mines to 1944. Department of Mines and Energy, Adelaide. ONeil, B., 1995. Above and below: the South Australian Department of Mines and Energy from 1944 to 1994. Department of Mines and Energy, Adelaide. ONeil, B., 1996a. History of petroleum exploration and development. In: Alexander, E.M. and Hibburt, J.E. (Eds). The petroleum geology of South Australia. Vol. 2: Eromanga Basin. South Australia. Department of Mines and Energy. Report Book, 96/20:11-31. ONeil, B., 1996b. National heroes, not national villains: South Australia and the atomic age. In: ONeil, B., Raftery, J. and Round, K. (Eds), Playfords South Australia: essays on the history of South Australia 19331968. Association of Professional Historians Inc., Adelaide. Packham, G.H. and Webby, B.D., 1969. The geology of the Officer Basin in the EVERARD 1:250 000 map area. Report for Marrumba Oil NL (unpublished). Parker, A.J., 1993. Geological framework. In: Drexel, J.F., Preiss, W.V. and Parker, A.J. (Eds), The geology of South Australia. Vol. 1, The Precambrian. South Australia. Geological Survey. Bulletin, 54:9-31. Parker, A.J. and Daly, S.J., 1993. Nawa Subdomain. In: Drexel, J.F., Preiss, W.V. and Parker, A.J. (Eds), The geology of South Australia. Vol. 1, The Precambrian. South Australia. Geological Survey. Bulletin, 54:70-71. Passmore, V., 1994. Impact of the Petroleum Search Subsidy Acts on the exploration for petroleum in Australia. In: Branagan, D.F. and McNally, G.H. (Eds), Useful and curious geological enquiries beyond the world: PacificAsia historical themes. INHIGEO Symposium, Sydney. Pelechaty, S.M., 1996. Stratigraphic evidence for the SiberiaLaurentia connection and Early Cambrian rifting. Geology, 24(8):719-722. Pell, S.D., McKirdy, D.M., Jansyn, J. and Jenkins, R.J.F., 1993. Ediacaran carbon isotope stratigraphy of South Australia an initial study. Royal Society of South Australia. Transactions, 117(4):153-161. Perincek, D., 1996. The age of NeoproterozoicPalaeozoic sediments within the Officer Basin of the Centralian Super-basin can be constrained by major sequence-bounding unconformities. APPEA Journal, 36:350-367. Pitt, G.M., Benbow, M.C. and Youngs, B.C., 1980. A review of recent geological work in the Officer Basin, South Australia. APEA Journal, 20(1):209-220. Pontifex, I.R., 1984. Mineralogical report No. 4263, for Comalco Aluminium Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 3938:188-208 (unpublished). Preiss, W.V., 1973. Early Willouran stromatolites from the Peake and Denison Ranges and their stratigraphic significance. South Australia. Department of Mines and Energy. Report Book, 73/208. Preiss, W.V. (Compiler), 1987. The Adelaide Geosyncline Late Proterozoic stratigraphy, sedimentation, palaeontology and tectonics. South Australia. Geological Survey. Bulletin, 53. Preiss, W.V., 1993. Neoproterozoic. In: Drexel, J.F., Preiss, W.V. and Parker, A.J. (Eds), The geology of South Australia. Vol. 1, The Precambrian. South Australia. Geological Survey. Bulletin, 54:171-203. Preiss, W.V. and Krieg, G.W., 1992. Stratigraphic drilling in the northeastern Officer Basin: Rodda 2 well. Mines and Energy Review, South Australia, 158:48-51. Quilty, F.H. and Goodeve, P.E., 1958. Reconnaissance airborne magnetometer survey of the Eucla Basin, Southern Australia. Bureau of Mineral Resources, Geology and Geophysics, Australia. Record, 1958/87. Rogers, P.A. and Freeman, 1996. WARRINA map sheet. South Australia. Geological Survey. Geological Atlas 1:250 000 Series, sheet SH53-3. Ross, J., 1875. Mr J. Rosss explorations, 1874. South Australia. Parliamentary Papers, 67. Rudd, D.S., 1995. Sedimentology and reservoir potential of the mid-Cambrian Trainor Hill Sandstone, eastern Officer Basin. University of Adelaide. B.Sc. (Hons) thesis (unpublished). Sansome, A and Gravestock, D., 1993. Sandstone reservoirs in the eastern Officer Basin. In: Alexander, E.M. and Gravestock, D.I. (Eds), Central Australian Basins Workshop, Alice Springs, 13-14 September 1993. Programme and Abstracts, p.60. Seilacher, A., 1967. Bathymetry of trace fossils. Marine Geology, 5:413-428.

%

Shaw, R.D., 1991. The tectonic development of the Amadeus Basin, central Australia. Bureau of Mineral Resources, Geology and Geophysics, Australia. Bulletin, 236:429-462. Shearer, A., 1994. Distribution of Permian sediments in the eastern Officer Basin, SA. South Australia. Department of Mines and Energy Resources. Open file envelope, 8591 (unpublished). Smith, R.C. and Tracy, G.W., 1986. Applied Reservoir Engineering Course Notes. Australian Mineral Foundation, 21 July to 1 August, 1986. Oil and Gas Consultants International Inc. (unpublished). Southgate, P.N. and Henry, R., 1984. Lithofacies and cyclic sequences in an alkaline lake sequence of Lower to Middle Cambrian age from the Officer Basin. Geological Society of Australia. Abstracts, 15:180. Southgate, P.N., Lambert, I.P., Donnelly, T.H., Henry, R., Etminan, H. and Weste, G., 1989. Depositional environments and diagenesis in Lake Parakeelya: a Cambrian alkaline playa from the Officer Basin, South Australia. Sedimentology, 36:1091-1112. Sprigg, R.C., 1983. The search for commercial oil and gas in South Australia. Australian Institute of Petroleum (SA Branch), From the Coorong to Port Bonython: the history of the oil industry in South Australia. Proceedings of the 1983 conference of the AIP (SA Branch), Adelaide, pp.C1-119. Stainton, P.W., Weste, G. and Cucuzza, G., 1988. Exploration of PEL 23 and PEL 30, eastern Officer Basin, South Australia, 1983-1988. South Australia. Department of Mines and Energy Resources. Open file Envelope, 5073:1243-1321 (unpublished). Statham-Lee, L., 1995. Palaeodrainage (1:1 000 000). In: Lindsay, J.F. (Ed.), Geological atlas of the Officer Basin, South Australia. Australian Geological Survey Organisation and Department of Mines and Energy, South Australia. Steenland, N.C., 1965. Eastern Officer Basin aeromagnetic survey OEL 28, SA. Adastra Hunting Geophysics Pty Ltd. Report for Exoil Pty Ltd. South Australia. Department of Mines. Open file Envelope, 527:3-64 (unpublished). Streich, V., 1892. Geological observations taken on the Sir Thomas Elder Expedition 1891. Royal Society of South Australia. Transactions, 16:74-115. Sukanta, U., 1993. Sedimentology, sequence stratigraphy and palaeogeography of Marinoan sediments in the eastern Officer Basin, South Australia. Flinders University (South Australia). Ph.D. thesis (unpublished). Sukanta, U., Thomas, B., von der Borch, C.C. and Gatehouse, C.G., 1991. Sequence stratigraphic studies and canyon formation, South Australia. PESA Journal, 19:68-73. Summons, R.E. and Powell, T.G., 1991. Petroleum source rocks of the Amadeus Basin. Bureau of Mineral Resources, Geology and Geophysics, Australia. Bulletin, 236:511-524. Surkov, V.S., Grishin, M.P., Larichev, A.I., Lotyshev, V.I., Melnikov, N.V., Kontorovich, A.Eh., Trofimuk, A.A. and Zolotov, A.N., 1991. The Riphean sedimentary basins of the Eastern Siberia Province and their petroleum potential. Precambrian Research, 54:37-44. Symonds, J.L., 1985. A history of British atomic tests in Australia. Department of Resources and AGPS, Canberra. Thomas, B., 1990. Summary of seismic interpretation in the eastern Officer Basin. South Australia. Department of Mines and Energy. Report Book, 90/58. Thomson, B.P., 1969. Precambrian basement cover: the Adelaide System. In: Parkin, L.W. (Ed.), Handbook of South Australian geology. Geological Survey of South Australia, pp.49-83. Thornthwaite, C.W. and Mather, J.R., 1957. Instructions and tables for computing potential evapotranspiration and the water balance. Drexel Institute of Technology. Climatology Laboratory. Publication, 10(3):185-311. Tietkens, W.H., 1890. Journal of Mr W.H. Tietkens Central Australian Exploring Expedition. South Australia. Parliamentary Papers, 111. Tingate, P., 1994. Interpretation of apatite fission track data from the Officer Basin. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Townsend, I.J. and Ludbrook, N.H., 1975. Revision of Permian and Devonian nomenclature of four formations in and below the Arckaringa Basin. South Australia. Geological Survey. Quarterly Geological Notes, 54:2-5.

Townsend, I.J., 1976. Stratigraphic drilling in the Arckaringa Basin, 1969-1971. South Australia. Geological Survey. Report of Investigations, 45:7-30. Townsend, I.J., 1990. Mintabie Opalfield mining and geology. South Australia. Department of Mines and Energy. Report Book, 90/73. Townson, W.G., 1985. The subsurface geology of the western Officer Basin results of Shells 1980-1984 petroleum exploration campaign. APEA Journal, 25(1):34-51. Toyne, P. and Vachon, D., 1984. Growing up the country: the Pitjantjatjara struggle for their land. McPhee Gribble and Penguin Books, Fitzroy (Victoria). Tucker, L.R., 1994. Devonian geology of the Munyarai Trough. South Australia. Department of Mines and Energy. Report Book, 94/10. Tucker, R.D. and McKerrow, W.S., 1995. Early Palaeozoic chronology; a review in light of new UPb zircon ages from Newfoundland and Britain. Canadian Journal of Earth Sciences, 32:368-379. van Neil, J., 1984. Palynology of Upper Proterozoic and Cambrian samples from 11 wells in South Australia. Report for Shell Development (Aust.) Pty Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 6846:3-17 (unpublished). Veevers, J.J. and Powell, C.McA., 1989. Phanerozoic tectonic regimes of Australia, Gondwanaland and the globe. Geological Society of Australia. Abstracts, 24:159-160. Vlierboom, F.W., 1973. Palynology and source rock potential of core samples from the Conoco exploration well Munyarai 1, Officer Basin, South Australia. Report for Shell Development (Aust.) Pty Ltd. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8590:37-48 (unpublished). Wade, A., 1915. The supposed oil-bearing areas of South Australia. South Australia. Geological Survey. Bulletin, 4. Wakelin-King, G., 1994. Proterozoic play challenges Amadeus Basin explorers. Oil and Gas Journal, February 28, 1994: 52-55. Wallace, M.W., Williams, G.E., Gostin, V.A. and Keays, R.R., 1990. The Late Proterozoic Acraman impact towards an understanding of impact events in the sedimentary record. Mines and Energy Review, South Australia, 157:29-35. Walter, M.R. and Gorter, J.D., 1994. The Neoproterozoic Centralian Basin in Western Australia: the Savory and Officer Basins. In: Purcell, P.G. and Purcell, R.R. (Eds), The sedimentary basins of Western Australia. Proceedings of the West Australian Basins Symposium, Perth. Petroleum Exploration Society of Australia (Western Australia Branch), pp.851-864. Ward, L.K., 1944. The search for oil in South Australia. South Australia. Geological Survey. Bulletin, 22. Watson, B.L., 1994a. Marla 10 core sample. Identification of staining. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Watson, B.L., 1994b. Cuttings samples (from Comalco coal exploration holes). Petrographic and geochemical analyses. South Australia. Department of Mines and Energy Resources. Open file Envelope, 8591 (unpublished). Webb, A.W., 1978. Geochronology of the Officer Basin. Amdel progress reports 1 to 6. South Australia. Department of Mines and Energy Resources. Open file Envelope, 3325:3-33 (unpublished). Wells, L.A. and George, F.R., 1904. Reports on prospecting operations in the Musgrave, Mann and Tomkinson Ranges (with plans). South Australia. Parliamentary Papers, 54. Weste, G., 1984. PEL 23. 1983 source rock geochemistry and porosity/permeability studies. South Australia. Department of Mines and Energy Resources. Open file Envelope, 5073:365-551 (unpublished). White, A.H. and Youngs, B.C., 1980. Cambrian alkali playa-lacustrine sequence in the northeastern Officer Basin, South Australia. Journal of Sedimentary Petrology, 50:1279-1286. White, D.A. and Gehman, H.M., 1979. Methods of estimating oil and gas resources. AAPG Bulletin, 63:2183-2192. Wilkins, R.W.T., Wilmshurst, J.R., Russell, N.J., Hladky, G., Ellacott, M.V. and Buckingham, C.B., 1992. Fluorescence alteration and the suppression of vitrinite reflectance. Organic Geochemistry, 18:629-640. Wilkinson, R., 1988. A thirst for burning: the story of Australias oil industry. David Ell Press, Sydney.

%

Wilson, A.F., 1952. Precambrian tillites east of the Everard Ranges, north-western South Australia. Royal Society of South Australia. Transactions, 75:160-163. Winnecke, C., 1896. Journal, etc., of the Horn Scientific Exploring Expedition to Central Australia (with plates and plans). 1894. South Australia. Parliamentary Papers, 19. Womer, M.B., Baker, R.N., Newman, E.J. and van Nieuwenhuise, R., 1987. Technical evaluation of PEL 29, east Officer Basin, Australia. Report for Amoco Australia Production Company. South Australia. Department of Mines and Energy Resources. Open file Envelope, 6843 (unpublished). Wopfner, H., 1969. Lithology and distribution of the Observatory Hill beds, eastern Officer Basin. Royal Society of South Australia. Transactions, 93:169-185. Wopfner, H., 1970. Depositional history and tectonics of South Australian sedimentary basins. Mineral Resources Review, South Australia, 133:32-50. Young, G.M., 1992. Late Proterozoic stratigraphy and the CanadaAustralia connection. Geology, 20:215-218. Zang, W-L., 1993. Review of Neoproterozoic and Early Palaeozoic acritarch biostratigraphy. In: Alexander, E.M. and Gravestock, D.I. (Eds), Central Australian Basins Workshop, Alice Springs, 13-14 September 1993. Programme and Abstracts, pp.12-14. Zang, W-L., 1994. Review of Neoproterozoic and Early Palaeozoic acritarch biostratigraphy in Australia. PESA Journal, 22:101-106. Zang, W-L., 1995a. Early Neoproterozoic sequence stratigraphy and acritarch biostratigraphy, eastern Officer Basin, South Australia. Precambrian Research, 74:119-175. Zang, W-L., 1995b. Neoproterozoic depositional sequences and tectonics, eastern Officer Basin, South Australia. South Australia. Department of Mines and Energy. Report Book, 95/21. Zang, W-L. and McKirdy, D.M., 1993. Microfossils and molecular fossils from the Neoproterozoic Alinya Formation a possible new source rock in the eastern Officer Basin. In: Alexander, E.M. and Gravestock, D.I. (Eds), Central Australian Basins Workshop, Alice Springs, 13-14 September 1993. Programme and Abstracts, pp.62-63. Zang, W-L. and McKirdy, D.M., 1994. Microfossils and molecular fossils from the Neoproterozoic Alinya Formation a possible new source rock in the eastern Officer Basin. PESA Journal, 22:89-90. Zang, W-L. and Preiss, W.V. (in prep.). Neoproterozoic Callanna Group in South Australia: rift-valley tectonism, sedimentation, mineralisation and biostratigraphy. South Australia. Department of Mines and Energy. Report Book. Zang, W-L. and Walter, M.R., 1992. Late Proterozoic and Cambrian microfossils and biostratigraphy, Amadeus Basin, central Australia. Association of Australasian Palaeontologists. Memoir, 12:1-132. Zhmur, S.I., Bazhenova, O.K., Burzin, M.B., Gorlenko, V.M., Rozanov, A.Yu. and Sokolov, B.A., 1994. The role of benthic microbial associations in oil source rock potential. In: Sokolov, B.A. (Ed.). History of oil in sedimentary basins. Interprint, Moscow, pp.35-48 (in Russian).

%!

You might also like