Painting Amara West: The Technology and Experience of Colour in New Kingdom Nubia (British Museum Publications on Egypt and Sudan, 13) 9789042945265, 9789042945272, 9042945265

The first in a series of volumes generated by the British Museum's Amara West Research Project, this explores the u

193 64 25MB

English Pages 126 [127] Year 2022

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Cover
Title
Copyright
Table of contents
Recommend Papers

Painting Amara West: The Technology and Experience of Colour in New Kingdom Nubia (British Museum Publications on Egypt and Sudan, 13)
 9789042945265, 9789042945272, 9042945265

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

BRITISH MUSEUM PUBLICATIONS ON EGYPT AND SUDAN 13

PAINTING AMARA WEST The technology and experience of colour in New Kingdom Nubia

Kate FULCHER

PEETERS

PAINTING AMARA WEST

BRITISH

MUSEUM AMARA

PUBLICATIONS WEST

RESEARCH

ON

EGYPT

AND

PUBLICATIONS

PAINTING AMARA WEST

The technology and experience of colour in New Kingdom Nubia

Kate FULCHER

PEETERS LEUVEN – PARIS – BRISTOL, CT 2022

SUDAN 1

13

The work presented here formed the main part of my doctoral thesis at the Institute of Archaeology at University College London, submitted October 2017. The research, in collaboration with the British Museum, was funded by the Arts and Humanities Research Council as part of the Collaborative Doctoral Partnership scheme. This book has been peer reviewed.

Cover illustration: Experimental paint palettes. Photo by the author.

A catalogue record for this book is available from the Library of Congress. ISBN 978-90-429-4526-5 eISBN 978-90-429-4527-2 D/2022/0602/30 © 2022, Peeters, Bondgenotenlaan 153, B-3000 Leuven, Belgium No part of this book may be reproduced in any form or by any electronic or mechanical means, including information storage or retrieval devices or systems, without the prior written permission from the publisher, except the quotation of brief passages for review purposes.

TABLE OF CONTENTS

List of figures ....................................................................................................................................................

VII

List of tables ......................................................................................................................................................

IX

Preface................................................................................................................................................................

XI

Acknowledgements ............................................................................................................................................

XIII

INTRODUCTION.....................................................................................................................................................

1

PART I: ARCHAEOLOGICAL CONTEXT AND SCIENTIFIC ANALYSIS CHAPTER 1 AMARA WEST ..................................................................................................................................

5

Palettes and pigments at Amara West...............................................................................................................

9

Painted architecture at Amara West ..................................................................................................................

10

Colour in the cemeteries at Amara West: coffins ............................................................................................

15

Nubian and Egyptian identities at Amara West ................................................................................................

15

Natural resources around Amara West .............................................................................................................

16

CHAPTER 2 COLOUR IN

EGYPT .............................................................................................................

19

Paints and pigments in ancient Egypt ...............................................................................................................

20

Paints and pigments in ancient Nubia ...............................................................................................................

30

Binders in ancient Egypt ...................................................................................................................................

32

Ancient Egyptian painting tools ........................................................................................................................

33

CHAPTER 3 AMARA WEST PAINTS:

ANALYSIS .....................................................................................................

35

Materials analysed from Amara West ...............................................................................................................

35

Pigment and binder analysis: methodology ......................................................................................................

38

Pigment analysis: results ...................................................................................................................................

40

Binder analysis: results .....................................................................................................................................

58

ANCIENT

Part II: ETHNOARCHAEOLOGICAL AND EXPERIENTIAL STUDIES CHAPTER 4 ETHNOARCHAEOLOGY: PERSPECTIVES ON PAINTING AMARA WEST ..................................................

61

CHAPTER 5 EXPERIENTIAL ARCHAEOLOGY ..........................................................................................................

65

Collection of materials ......................................................................................................................................

65

Processing ..........................................................................................................................................................

67

Application.........................................................................................................................................................

71

VI

TABLE OF CONTENTS

CHAPTER 6 RECONSTRUCTION OF ANCIENT ACTIVITIES .......................................................................................

75

Activity 1: Collection of paint materials ..........................................................................................................

76

Activity 2: Preparation ......................................................................................................................................

80

Activity 3: Application......................................................................................................................................

82

Activity 4: Experiencing coloured surfaces......................................................................................................

84

Activity 5: Discard ............................................................................................................................................

86

Activity 6: Maintenance and refurbishment .....................................................................................................

87

CONCLUSION........................................................................................................................................................

89

APPENDIX: Experimental details ........................................................................................................................

91

BIBLIOGRAPHY.....................................................................................................................................................

95

INDEX ..................................................................................................................................................................

107

LIST OF FIGURES All images are © the Trustees of the British Museum, unless otherwise indicated.

Figure 1

Map of Upper Egypt and Nubia showing the location of Amara West and other sites mentioned in the text

Figure 2

Aerial photograph of Amara West

Figure 3

Map showing colour finds in Phases 1–2 of area E13, the northwestern part of the walled town at Amara West

Figure 4

Map showing colour finds in Phases 3, 4, and 5 of area E13

Figure 5

Map showing colour finds in the Western Suburb at Amara West

Figure 6

Grindstone with yellow pigment from E13.31.2 [5332] (F5962)

Figure 7

The east part of the mastaba in E13.7.6

Figure 8

Fragment of painted plaster F5133d from the niche in E13.7.6 [4561]

Figure 9

Traces of red and yellow paint in E13.7.5 [4566]

Figure 10 Fragment of painted plaster F5049 from E13.7.6 [4561] Figure 11 Painted wall plaster F5133c from the painted niche in E13.7.6 [4561] Figure 12 Fragments of palettes F15020, from D12.8.7 [12840] Figure 13 Yellow and red pigments F2471, from E13.20.1 [10301] Figure 14 Painted coffin plaster F9707 from G244 [9505] Figure 15 Dispersion of pale yellow pigment from ceramic palette F6079 (PS128), from E13.14.8 [5222], mostly calcite Figure 16 FTIR spectrum for white coffin paint PS307 from coffin fragment F8110, from grave G309 [8163] Figure 17 Dispersion of black pigment PS295 from coffin F9743, from G244 [9515] Figure 18 Dispersion of black pigment PS527 from coffin fragment F8110, from G309 [8163] Figure 19 Mass chromatograms for ions m/z 191 and m/z 217 for palette PS152, from E13.14.6 [5284] Figure 20 Palette F7130 (PS448) from E13.31.1 [5325]. Palette holds both light and dark yellow paint Figure 21 Dispersion of lighter yellow paint from palette F7130 (PS448), from E13.31.1 [5325] Figure 22 Dispersion of red pigment F6177 (PS329), from E13.31.1 [5356] Figure 23 Dispersion of red paint PS380 from the pit in D12.7.1 Figure 24 Dispersion of paint containing Egyptian blue from palette F12423 (PS538), from D12.10 [12211] Figure 25 FTIR spectrum for blue paint on palette F6223 (PS140) from E13.7.12 [5224] Figure 26 Ceramic palette F6147 from E13.31.1 [5325] with traces of blue pigment in voids Figure 27 Palette F2644 (PS539) from D12.7.6 [12062] with grey-blue paint Figure 28 Pale green paint on palette F6119 (PS126), from E13.31.2 [5331] Figure 29 House just outside Abri painted in yellow gir with gypsum around the doors Figure 30 Farmer’s house near Abri. Internal room decorated with bomastic Figure 31 One sofiha of gir Figure 32 Graffito of a longhorned cow on rock north of Amara West Figure 33 Group of stone tools found in deposit 2749 within alcove 2793 in house D11.2 Figure 34 Grinding experiment

VIII

LIST OF FIGURES

Figure 35 Summary of people, tools, places, and actions that may be involved in the task of collecting raw materials for paint production Figure 36 Summary of people, tools, places, and actions that may be involved in the task of processing raw materials for paint production Figure 37 Summary of people, tools, places, and actions that may be involved in the task of applying paint Figure 38 Summary of people, tools, places, and actions that may be involved in experiencing paint Figure 39 Summary of people, tools, places, and actions that may be involved in the discard of paint manufacturing materials Figure 40 Summary of people, tools, places, and actions that may be involved in the maintenance of painted surfaces

LIST OF TABLES Table 1

Chronology of Upper and Lower Nubia and the contemporaneous periods in Egypt

Table 2

Architectural phases at Amara West

Table 3

Grindstones with traces of pigment

Table 4

Hammerstones with traces of pigment

Table 5

Results of pigment analysis of white paints, pigments, and plaster

Table 6

Results of pigment analysis of black paints, pigments, and plaster

Table 7

Results of pigment analysis of yellow paints, pigments, and plaster

Table 8

Results of pigment analysis of red paints, pigments, and plaster

Table 9

Results of pigment analysis of blue paints, pigments, and plaster

Table 10

Results of pigment analysis of green paints, pigments, and plaster

Table 11

Summary of pigments identified from Amara West

Table 12

Results of grinding experiment

PREFACE The Amara West Research Project conducted archaeological fieldwork at the Pharaonic town in northern Sudan from 2008 through 2019, with post-excavation analyses and research still ongoing. The research will be published both online and in a series of edited volumes and monographs. This first volume, on the use of pigments and the experience of colour in the ancient town, embodies the approach of the wider project, in combining scientific analyses, archaeological fieldwork, and modern ethnographic perspectives, to seek nuanced perspectives on lived experience at Amara West. It also reflects the collaboration of the British Museum with university partners and the National Corporation of Antiquities & Museums in Sudan.

Neal Spencer, British Museum

ACKNOWLEDGEMENTS This research would not have been possible without the assistance of the National Corporation for Antiquities and Museums (Sudan), who generously allow the export of archaeological samples for scientific research, and provided the concession for fieldwork at Amara West. The samples were excavated as part of fieldwork of the British Museum Amara West Project, directed by Neal Spencer and funded by the Qatar-Sudan Archaeological Project, The Leverhulme Trust, the Institute of Bioarchaeology, the British Academy, and the Fondation Michela-Schiff Giorgini. Particular thanks are due to Abdelrahman Ali Mohamed, Shadia Abdu Rabo, Mohamed Saad, Mat Dalton, Anna Stevens, Manuela Lehmann, and all the researchers who have worked at Amara West. Thanks to the Arts and Humanities Research Council for funding this joint doctoral project, and to the team who run the Collaborative Doctoral Partnership scheme and support it within the non-University partners. Huge thanks are due to the staff of the Institute of Archaeology, UCL, and at the British Museum, both in the Department of Scientific Research, and the Department of Egypt and Sudan, including my supervisors Ian Freestone, Ruth Siddall, Neal Spencer, and Rebecca Stacey, and also Janet Ambers, Joanne Dyer, Duncan Hook, Harriet White, Patrick Quinn, and Tom Gregory, among others. Many thanks also to the two anonymous peer reviewers whose comments improved the text immensely.

INTRODUCTION

The town of Amara West, in Upper Nubia, was occupied by people 3,000 years ago. The world they lived in was very different from the one we inhabit today, yet they performed activities that are recognizable to us: cooking, raising children, keeping animals, burying their dead, and decorating houses and objects with coloured paint. Accessing the thought processes of these long-gone societies may seem impossible, but we have clues at our disposal: the archaeological record they left behind, and the materials they used, as well as a range of textual sources from Pharaonic Egypt that provide further context. By analysing the evidence that we have, we can learn more about those societies’ activities and priorities. Ancient people made choices in the materials they used and the way in which they used them; these choices reveal decisions and strategies that tell us something of the way in which they formed their world view and lived within it. Moreover, the material composition of a built environment is not static; changes are made, decorative schemes are altered, things are used and discarded. Following the life history of objects and buildings, tracking changes to original approaches or decisions, and following processes from instigation to completion, and then reconsideration, can tell us something about the shifting priorities in the lives of the people who created these material situations. The location of the town in Upper Nubia, a distant colony, may result in local practices divergent from the norm in Egypt. Analysing paint materials is one way in which to examine this idea. Research on paint in ancient Egypt has focused on objects now held in museums, and elite tombs and temples. The corpus of material from Amara West provides the opportunity to investigate paint and its use across several types of environment within one New Kingdom town, including domestic buildings, tombs, and a possible workshop area. The National Corporation for Antiquities and Museums (Sudan) generously permits

the export of samples for analysis, allowing a detailed scientific study of the inorganic and organic components of the paints. This has been largely absent from previous studies, owing to the restrictions on sampling museum objects and the ban on exporting samples from Egypt. This study is formed of two parts. Part I discusses the archaeological evidence for paint and painting materials at Amara West, and the scientific investigation of the paint to identify the pigments and binders. Part II uses ethnographic and experiential techniques to consider the materiality and phenomenology of the processes of paint preparation, application, and perception. This takes the form of experiments in creating and using paints and tools similar to those identified in the archaeological record, which are informed by several ethnoarchaeological interviews with people that live in the area and who still use locally available materials to make paint for their houses. This experiential practice revealed the complexity of the process of painting, and the range of materials and actors that were potentially involved. In the section that follows (‘Reconstruction of ancient activities’) all of the strands (archaeology, science, experiments, and ethnoarchaeology), are brought together to attempt to reconstruct the ancient process of making paint and how it was experienced by the inhabitants of Amara West. Short fictional passages are included to explore their experiences and how this may have entangled with the rest of their daily lives. This is the first holistic study of colour within a Pharaonic town—across domestic, workshop, and funerary environments—and the first that has combined a scientific and phenomenological approach. Using this combination of methods enriched the research, providing a more nuanced conception of the lives of the people who lived at Amara West, imbuing them with an agency and a presence that is lacking from purely scientific studies.

I ARCHAEOLOGICAL CONTEXT AND SCIENTIFIC ANALYSIS

The first part of this book discusses the archaeology of Amara West, focussing on the evidence for paint production and application, and then the evidence for the types of pigments used in ancient Egypt and Nubia. The scientific investigation of the paint materials from Amara West is then described. This approach enables the raw materials used for paint at Amara West to be identified, and compared to those known from other locations and time periods.

CHAPTER 1

AMARA WEST

The ancient town now referred to as Amara West is located in Nubia, between the Second and Third Nile Cataracts in modern Sudan (Figs 1, 2). The term ‘Nubia’ is used to designate the regions adjacent to the Nile in southern Egypt and northern Sudan where the Nubian languages are spoken, from the First to the Fourth Nile Cataract (Fernea and Rouchdy 2010). Pharaonic Egypt annexed much of Lower Nubia in the Middle Kingdom (2000–1700 BC) and built forts as far south as Semna, between the Second and Third Nile Cataracts. Control of traffic and trade is assumed to have been a prime motivation for this conquest, as Nubia offered both mineral wealth, especially gold, and a conduit towards the rich resources of Africa (S. T. Smith 1997; cf. Kemp 1997). Egypt lost control of Nubia towards the end of the Middle Kingdom, a period when Egypt itself was politically divided. The Kingdom of Kush, centred at Kerma, emerged as a dominant force in the Middle Nile region, conducting raids into southern Egypt (Davies 2003). At the beginning of the New Kingdom, around 1500 BC, pharaoh Kamose of Thebes and his son Ahmose reunified Egypt by recapturing the north that had fallen under Hyksos control. Parts of Lower Nubia were reconquered at this period, as evidenced by monuments at Buhen in the Second Cataract area and the establishment of an administrative role, ‘the King’s Son of Kush’, typically translated as Viceroy of Kush (H. S. Smith 1976, 8–9; Edwards 2004, 106). Under Thutmose I, the Pharaonic state extended its sphere of control further up the Nile into Upper Nubia, with the foundations of towns such as Sai and Sesebi, and rock inscriptions as far upstream as Kurgus (Davies 2017). The final conquest of the Kerma polity may not have been complete until the reign of Thutmose III (Edwards 2004, 101–102). Nubia was then subdivided into two regions for administrative purposes: Wawat (Lower Nubia) and Kush (Upper Nubia) (Wendrich 2010, 7), with a ‘Deputy of Wawat’ and a ‘Deputy of Kush’, respectively, overseeing these areas (Edwards 2004, 106; Török 2009, 171). Several Middle Kingdom fort sites in Lower Nubia were refurbished during this period, but New Kingdom settlements in Upper Nubia are

notable for having a less fortified appearance than the earlier foundations, perhaps reflecting an economic rather than a military focus (Török 2009, 185; N. Spencer 2019). The Ramesside Period (1295–1069 BC) was seemingly characterized by ongoing occupation of these towns, though with less evidence for major templebuilding projects beyond the series of rock-cut temples in Lower Nubia, including Abu Simbel. However, two new towns were founded in Upper Nubia in the early Ramesside era: Aksha and Amara West. The layouts of these two towns were similar, with a temple, storage facilities and housing areas (Rosenvasser 1964; N. Spencer 2017), although research at Amara West underlines how such towns changed over time, particularly through local agency (N. Spencer 2017). By the end of the New Kingdom (1069 BC), Egypt lost control over Nubia. Archaeological fieldwork in recent years has revealed how many of the Pharaonic towns continued to be occupied, though the nature of that occupation, and the transition to Napatan rule (Welsby and Welsby Sjöström 2007; Török 2009), is still poorly understood. A chronology of Nubia, including all periods mentioned in the text, is given in Table 1. The first excavations at Amara West were directed by H. W. Fairman on behalf of the Egypt Exploration Society (EES), from 1938 to 1939, with three further seasons after World War II (P. Spencer 1997). These excavations focused on the temple, two areas of the walled town, and some tombs in the adjacent cemeteries. A British Museum research project between 2008 and 2019 deployed interdisciplinary methodologies to elucidate aspects of lived experience and cultural entanglement (N. Spencer 2012; Spencer, Stevens and Binder 2014). A full publication of the excavations and research project is in preparation. Founded in the reign of Seti I (1294–1279 BC), as evidenced by bricks in the town wall stamped with his name, the walled town (108 × 108m) featured a sandstone temple, decorated in the reign of Ramses II but bearing inscriptions of kings and officials up to the reign of Ramses IX (c. 1120 BC) (P. Spencer 2016). Door jambs and lintels from a large building (E.13.2)

CHAPTER

6

1

Table 1: Chronology of Upper and Lower Nubia and the contemporaneous periods in Egypt. Upper Nubia

Lower Nubia

Contemporaneous period in Egypt

Early Khartoum Khartoum Neolithic Pre-Kerma

Approximate dates 8000–5000 BC 4500–3500 BC

Early A-Group

Naqada Ic–IIa/d

3500 BC

Pre-Kerma

Classical A-Group

Naqada III

3200 BC

Pre-Kerma

End of A-Group

Unification to mid-Dynasty 1

3000 BC

Old Kingdom

2800–2200 BC

Late Old Kingdom

2100 BC

2000–1700 BC

Population not identified Pre-Kerma

Egyptian occupation of part of lower Nubia including Buhen, and expeditions C-Group IA, IB

Early Kerma

Part occupation, Egyptian expeditions C-Group IIA, IIB

Middle Kerma

Egyptian control including Second Cataract forts

Middle Kingdom

Classic Kerma

C-Group III

Second Intermediate Period

New Kingdom (Egyptian rule) Pre / early Napatan Period

1700–1500 BC 1500–1000 BC

Third Intermediate Period

Napatan Period Dynasty 25, Nubian rule in Egypt

1000–750 BC 750–650 BC

Napatan Period

Late Period

650–332 BC

Meroitic Period

Ptolemaic and Roman Period

332 BC–AD 350

indicate that this was the residence of the ‘Deputy of Kush’, which supports the interpretation that Amara West was a new administrative centre for the control of Upper Nubia (P. Spencer 1997, 220; N. Spencer 2017, 328–32). The other areas of the walled town comprised dwellings and storage magazines, with many of the latter being transformed into further, and more densely arranged, housing (N. Spencer 2017). One of these zones, area E13 (see Fig. 2), was the subject of British Museum excavations between 2009 and 2015. A sequence of architecture from the foundation of the town to its abandonment reflects the shift from formal storage facilities to housing (N. Spencer 2015). An extra-mural area, designated the ‘Western Suburb’, was developed from late Dynasty 19/early Dynasty 20 with further housing, including large villas (e.g. E12.10, see N. Spencer 2009). Buildings in the town have been assigned to one of a series of architectural phases (1A, 2B, 2A, 2B, 3A, 3B, 4 and 5; Table 2). While this clearly simplifies complex histories of use and change, it does provide a useful framework for studying changes over time. The ancient town was located upon an island in the Nile, and two cemeteries on the north mainland served

the town’s population. Cemetery D on the desert escarpment featured tombs with pyramid chapels, while chamber tombs (with no surviving superstructure), seemingly of lower status, were built upon an alluvial terrace above a wadi to the east (Cemetery C) (Binder, Spencer, and Millet 2010; Binder 2011; Binder 2017). While the latest architecture identified in the town dates to the late New Kingdom, ceramics scattered on the surface, and burials in the associated cemeteries, suggest activity after the end of Pharaonic control of the area. Burials in the cemetery continued through the 9th/8th centuries BC (Binder, Spencer, and Millet 2010), but it is unclear if the town remained inhabited. Climate change, specifically the contraction of the Nile and resulting aridification of the landscape, may have been a principal factor in the abandonment of the site (Woodward et al. 2017). The New Kingdom environment at Amara West may have been similar to that of the island of Ernetta, just upstream of the ancient site (Spencer, Macklin, and Woodward 2012). Inhabitants would have benefited from the water channel and tree cover along each bank, which protected against strong northerly winds and aeolian sand influx. Areas of the ancient island would have been suitable for agriculture

AMARA WEST

Figure 1: Map of Upper Egypt and Nubia showing the location of Amara West and other sites mentioned in the text. Map by Claire Thorne, © Amara West Project (British Museum).

7

CHAPTER

8

1

Figure 2: Aerial view of Amara West. Area E13 is situated within the town walls, which incorporate the West Gate. The Nile is visible at the top right. © Amara West Project (British Museum).

Table 2: Architectural phases at Amara West. Phase

Period/Date

Buildings in E13

1A c. 1300–1250 BC

Early Dynasty 19 (Seti I), foundation of walled town

Enclosure wall; kiln 5029 in building E13.15; Rubbish dumping area (?) E12.6 storage facility, space E13.23

Western Suburb

1B c. 1250–1210 BC

Early–mid-Dynasty 19 (Rameses II and later)

E13.24 metallurgy working area; magazine complex E13.14, house E13.10, space E13.22

Rubbish dumping area

2A / 2B c. 1210–1180 BC

Mid-Dynasty 19–early Dynasty 20

Houses E13.3, E13.7, E13.8, buildings E13.20, E13.22. E13.17 (area with kilns and ovens) and E13.21; E13.31 workshop (?) in magazine E13.14

Rubbish dumping area; garden plots (?)

3A / 3B c. 1180–1140 BC

Early–mid-Dynasty 20 (Rameses III or later)

House E13.3 divided [E13.3-N and E13.3-S]; houses E13.4+9, E13.5, E13.16, E13.6

4 c. 1140–1100 BC

Mid–late Dynasty 20

House E13.4+9 divided [E13.4 and E13.9]

First villas (E12.10, D12.5, D11.1) constructed. Further houses built in spaces between: D11.2, D12.5, D12.6, D12.7, D12.8, D12.9, D12.12.

5 c. 1100–1000 BC

Late Dynasty 20 onwards

Small-scale adjustments to architecture and abandonment of buildings

AMARA WEST

and livestock grazing. Optically stimulated luminescence (OSL) dates from silt and sand fluvial couplets within sediments excavated from a palaeochannel adjacent to the north edge of the town indicate that by 1270 BC (± 215) this northern channel at Amara West was flowing only intermittently, eventually precipitating the joining of Amara West to the north bank of the river (Spencer, Macklin, and Woodward 2012). As the north channel dried up, the protective vegetation was lost, and the associated sand dunes eroded, exposing the town to aeolian sand brought by the northern winds. This would have made living conditions much less appealing, as can be seen from the continued attempts of the ancient inhabitants of the houses to prevent sand ingress using blocking walls around doorways, and from evidence of deteriorating health conditions, including an increased prevalence of respiratory illnesses (Woodward et al. 2017). The evidence for colour—pigment lumps, sherds used as mixing palettes, painted objects, and architecture—at Amara West was found in considerable concentrations within parts of E13, and to a lesser extent in the western suburb and cemeteries. However, much of the evidence for the use of colour at Amara West will have been lost over time. There are no upper storeys preserved in the buildings, although the presence of stairs indicates that they may have existed. Organic preservation is poor in the town, so we know little about wood, textile, or leather objects that may have borne paint or been used in the process of painting or making paint. The strong north wind has eroded many built surfaces that may originally have been painted. While wood survives better in the cemetery, much has been lost to termite damage. The grid system used by the EES to refer to buildings and rooms (P. Spencer 1997, 6, pl. 2) was extended into the areas excavated by the British Museum project. Spaces are designated by grid/building/room, thus D12.7.1. refers to room 1 in house 7 within grid D12. Each designated building would have been recognized as a distinct structure by the ancient inhabitants, though complex patterns of rebuilding, modifications and reuse are typical, including the amalgamation or subdivision of buildings. Objects and samples discussed in this book are designated by Find (F), Ceramic (C), Petrographic Sample (PS) and Archaeological Sample (AS) numbers, each associated with a specific archaeological context designated with a four- or fivedigit number.

9

Palettes and pigments at Amara West Most of the evidence for palettes and pigments in the walled town comes from area E13.14 (Figs 3, 4). House E13.7 was formed from the repurposing and conversion of parts of a Phase 1B magazine complex (E13.14), with the more easterly zones of E13.14 being subdivided into smaller units including a suite of three spaces designated as E13.31. These yielded over 300 ceramic sherds with paint on the concave surface, and many small lumps of raw pigment. The ceramic sherds with paint are thought to have been used as palettes for holding and/or mixing paint. Some of the repurposed spaces may have been working areas: the front of a magazine and the space before it (E13.31.1 and E13.31.2) yielded large grinding stones used for pigment processing, alongside a high density of palettes and pigments, and also small copper alloy objects, pieces of ceramic crucible, ostrich eggshell, faience beads, small flint blades, and worked pieces of stone. Other parts of the old magazines (E13.14.3, E13.14.5, E13.14.6) more closely resembled dumping areas for materials from adjacent houses or work areas, with colour-related finds mixed with rubble; some of the material might represent deposits used to fill in spaces prior to building the next phase of architecture, although here also the colour finds could have been part of an occupational deposit, later mixed into accumulated rubble and debris. A short distance to the northeast, area E13.17 also resembled a working area, with large amounts of pigments and palettes in occupational deposits across the floors, in association with other common domestic or working features and finds such as hearths, ceramics, small objects such as ceramic counters, and bone. Phase 1 buildings E13.23/24/25, underlying E13.17, contained kilns, and 17 pigments and 33 palettes. Beneath house E13.5, which contained only one palette, the Phase 2 spaces E13.20 featured curved walls, possibly house exteriors, and contained 16 pigment pieces and 14 palettes. Below this was E13.22, an area of rectilinear buildings with 11 pigments and 2 palettes. These areas were once again refashioned, in phases 3A and 3B, creating a new layout of houses (E13.3, E13.4, E13.5, E13.6, E13.9). It is striking how little evidence for colour preparation was found in these houses compared to earlier phases, though some of the walls were painted white (N. Spencer 2014a, 464). It is also possible that the houses were being regularly swept out, in which case evidence for pigments and processing would be found only in secondary deposits.

10

CHAPTER

Palettes and pigments were found throughout the Western Suburb, with concentrations in houses D12.8, D12.9, and D11.1 (Fig. 5). The colour finds in D11.1 include some possible windblown intrusions but mainly seem to comprise occupational debris and finds predating the house, although overlying an earlier floor. There is also a late deposit that, although lying in windblown sand, consists of aligned schist stones surrounded by 9 hammerstones and polished stones, which suggests an organized working place; 10 palettes were found here. The finds in D12.9 pre-date the house, being found within a dense accumulation of rubbish, characteristic of the area prior to houses being constructed. This rubbish most likely came from the walled town (although not necessarily from E13, the northwestern corner of the walled town). Colour-related evidence in D12.8 derives from a mixture of rubbish and occupation deposits, but many of these deposits lie in a loose, sandy layer near the top of the archaeological phases, which makes their context less secure. Trench D11.7, at the southwest of the Western Suburb, revealed rubbish layers with further examples of palettes and pigments, but these cannot be securely dated or associated with a building. Obtaining a meaningful count of palettes and pigments is extremely difficult owing to the fragmentary nature of the finds. The count of palettes is somewhat more reliable, but some have probably been broken into multiple pieces during archaeological deposition; pigment lumps have often crumbled into pieces, in which case it is nonsensical to count each minute piece as a separate find. As a compromise, pigments have been counted by find number. Where the find is particularly large this has been noted. In practice, a ‘difference’ of, for example, 3 versus 5 pigments or palettes cannot be taken to be significant, but a difference between 5 and 60, say, is perhaps significant. The high density of finds in Phase 2 of area E13 can, therefore, be contrasted with the low number of finds from subsequent phases in this zone. Of the 927 palettes found at Amara West, 78 had evidence of blue pigment (9%). Twenty of the palettes with blue were from areas E13.14 and E13.31. Fifty of the blue palettes were from the Western Suburb, including 21 from D12.8 from occupational contexts, 13 from a windblown surface layer in D12.8.7, and 16 from D11.1, also mostly from windblown surface layers. Most of the other palettes were either yellow or red, and a few retained a black pigment, or white plaster / paint. Often more than one colour was found on a palette.

1

Of the approximately 260 pigment lumps found, most were small pieces of red and yellow, except for a few very small lumps of blue pigment, which were found in E13.20 (three tiny pieces) and E13.7 (two pieces, approximately 2.5cm and 1cm across), and a larger piece about 3.5cm across in the sandy infill of D12.8.9 (context uncertain). Two vessel fragments of Egyptian blue were found in E13.14, each about 2cm across. Two large caches of ochre (including single pieces about 20cm long and 10cm wide), one red and one yellow, were excavated from room D11.1.1 in the Western Suburb. Pieces of red and yellow ochre, alongside stone slabs and grinders, were found in the town south of the temple, excavated in the 1930s, in what were described as ‘workshops and storerooms, very probably attached to the temple’ (P. Spencer 1997, 103). Evidence for the preparation of pigment was found in the form of grindstones and hammerstones from the walled town and the Western Suburb, some of which retained traces of pigment, mostly red and yellow (Tables 3, 4). Two large grindstones were found in the areas outside the magazine entrances, one covered in yellow pigment (Fig. 6), and one with traces of green and red pigment. A large grindstone with red pigment was found in a spoil heap to the south of E13. Five grindstones with evidence for pigment grinding were found in the Western Suburb, two from D12.7, two from D13.3 (all with yellow and/or red pigment), and one with a trace of greenish paint in D12.11. Many other grindstones and hammerstones have been excavated from Amara West that do not bear any traces of colour but might nonetheless have been used for grinding pigments; many such tools are likely to have been multifunctional. Painted architecture at Amara West The full plan of house E13.7 (Phase 2A–2B) has not been revealed, as it is partly covered by later architecture. Nonetheless, it is clear that it went through several phases of refurbishment. Room E13.7.6 had a mastaba (bench) against the southeastern wall (Fig. 7), in front of which lay many fragments of painted and moulded mud plaster. Their form and decoration have led the excavators to suggest there was a niche in the wall above the mastaba, perhaps to hold a stela (N. Spencer 2014a). There appear to be three phases of decoration of the mastaba and the niche, on the basis of a study by N. Spencer (unpublished).

AMARA WEST

11

Figure 3: Map showing colour finds in Phases 1–2 of area E13, the northwestern part of the walled town at Amara West, dating to mid-/late Dynasty 19 or early Dynasty 20. Note that spaces E13.14.1 and E13.14.2 were re-formatted as working spaces (E13.31.1 and E13.31.2) in phase 2A. House 13.7 (bottom left) contained a painted mastaba with a painted niche in the wall above it, and painted walls. © Amara West Project (British Museum).

Figure 4: Map showing colour finds in Phases 3, 4, and 5 of area E13, dating from early to late Dynasty 20. © Amara West Project (British Museum).

12

CHAPTER

1

Figure 5: Map showing colour finds in the Western Suburb at Amara West. Of the finds shown, 27 pigments and 87 palettes are recorded as coming from a windblown context and therefore are possibly not in their original context; 11 pigments and 26 palettes were found in contexts pre-dating the Western Suburb (most likely rubbish from the walled town). © Amara West Project (British Museum).

Figure 6: Grindstone with yellow pigment from E13.31.2 [5332] (F5962). © Amara West Project (British Museum).

1. Mud-plastered mastaba, yellow (and black) paint on wall behind; simple niche with architrave painted white. 2. Walls all whitewashed, including mastaba. Cavetto cornice added to niche; niche painted with polychrome decoration (red, yellow, black, and blue on a white background; Fig. 8). 3. Everything re-whitewashed, niche also painted over with white.

The east wall of the same room (E13.7.6) was painted white to a height of about 1m with a band of black along the top, about 2cm thick; this decorative scheme extended onto the western wall of E13.7.3 but thereafter the walls of this room have only white paint remaining. There were traces of red and yellow paint on top of white on the walls of E13.7.5, adjacent to E13.7.6 (Fig. 9). Deposits and walls from the earliest level E13.23 (Phase 1A), below E13.22, contained fragments of yellow-painted wall plaster (context 10475). It is unclear if this was a house or fulfilled a different purpose. Colour was noted on the walls of a few Phase 3 houses (see Fig. 4): traces of red and yellow were found in room E13.4.2 (part of house built on top of E13.7.6); the walls of the small contiguous houses E13.3-N and E13.3-S were extensively and repeatedly whitewashed. In the Western Suburb several houses had whitepainted walls, but the evidence for colour on walls is limited. House D12.7 had several splashes of red paint on the walls and floor (Dalton 2020, 248–52) and a small loose fragment of mud plaster with two smears of blue paint. Pink-painted ceiling fragments were

13

AMARA WEST

Table 3: Grindstones with traces of pigment (not all of these were sampled due to minute quantities).

Table 4: Hammerstones with traces of pigment (not all of these were sampled due to minute quantities).

Find Building number

Context

Phase

Colours observed

F2061

E12.10.1

2015

2

Red

F2009

E12.10.2

2002

3A

Green

F2817

D12.11

2662

3/4/5

Green

F2017

E12.10.1

2015

3A

Yellow

F3062

D13.3.1

3008

3/4/5

Red

F3061

D13.3.1

3008

4

Red

F3105

D13.3.2

3025

3/4/5

Red, yellow

F4082

E13.9.16

4061

4

White, black

F4038

E13.9.16

4061

4

Red

F4098

E13.9.16

4091

3

Red

F4111

E13.9.19

4111

3–4

Red

F4108

E13.9.19

4111

3–4

Red

F4190

E13.3.26

4275

3B

Red

F4109

E13.9.19

4111

3–4

Red

F4265

E13.14.5

4316

2A–2B

Red

F4138

E13.9.19

4110

2B or 3

Yellow

F4476

E13.4.1

4540

3–4

Yellow

F4222

E13.3.24

4298

3A

Red

F4999

E13.3.34

4679

2B

Green

F4308

E13.4

4400

4

Red

F5185

E13.9.14

4253

4

Red, yellow

F4397

E13.4.1

4475

4

Green

F5962

E13.31.2

5332

2B

Yellow

F4730

E13.3.27

4369

3A–3B

Red

F6184

E13.14.1

5361

1B

Red, green

F4858

E13.9.19

4063

3A–3B

Green, white

F6302

E13.13

4880

2B

Red

F4861

E13.9.19

4063

3A–3B

Red

F6814

E13.5.6

5420

3A–3B

Red, yellow

F4866

E13.9.19

4063

3A–3B

Red

F6827

E13.16.5

5413

4

Red

F5095

E13.4

4561

2B–3A

Red

F6837

E13.5.6

5662

spoil

Red, yellow

F5288

E13.7.7

4772

2A

Red

F12273

D12.7.1

12050

3/4/5

Yellow

F5289

E13.7.7

4772

2A

Red

F12318

D12.7.9

12076

3/4/5

Red

F5710

E13.8.1

5114

3A

Red

F13478

D11.1.2

12520

3/4/5

Red

found in D12.6 (see PS263). House D12.8 contained a small piece of plaster painted in yellow and red, three pieces of black-painted plaster on mud brick from a wall, and a stone door lintel and stone threshold painted in red and yellow. Excavations in the 1930s uncovered extensive decoration in the temple at Amara West. Fairman commented in his diary, in reference to the columns in the hypostyle hall, ‘quite a lot of colour is preserved and it is very bright, and the effect of the miniature scenes is quite charming’ (P. Spencer 1997, 40). The peristyle hall, hypostyle hall, vestibule, and the three chapels in the sanctuary were all decorated in colour, as were the staircase adjoining the sanctuary, and the external chapel to the east of the peristyle hall (P. Spencer 1997, 27–52). In addition, the walls and floors of the temple magazines were painted white, and one storage area (E.14.8, the format of this room numbering is from EES excavations) retained a band of red paint 11cm high, which ran along the wall for 1.1m (P. Spencer 1997, 56–57). The temple and the associated storerooms are currently buried in sand and cannot be accessed.

Find number

Building

Context

Phase

Colours observed

F5821

E13.6.2

5211

3B

Green, white

F5874

E13.29.1

5230

2B

Red, yellow

F5906

E13.6.3

5301

3B

Red

F5913

E13.6.2– E13.6.3

5309

3B

Red

F5925

E13.6.10

5317

4

Red

F5951

E13.31.2

5326

3A

Red

F6051

E13.7.11

5255

2B

Red, yellow

F6053

E13.7.11

5255

2B

Red

F6189

E13.14.1

5361

1B

Red, yellow, white

F6227

E13.7.12

5224

2B–3A

Red, yellow

F6799

E13.5.1

5381

3/4/5

Green

F6813

E13.16.5

5428

4

Red

F6824

E13.5.6

5445

3A–3B

Red

Fairman’s excavations in the town south of the temple uncovered a room D.14.5 with ‘fine decorated plaster’ on the walls (P. Spencer 1997, 122). The south wall was whitewashed with a 2.5cm-deep black band painted around; further west on the south wall of D.14.5 was a niche topped with moulding, painted in white and red (P. Spencer 1997, 125). Among the houses to the north of the temple magazines, a niche was cut into the wall in the southwest corner of E.12.1. The niche was painted red above bands of yellow and black, and the face of the south wall had also been painted in coloured

14

CHAPTER

1

Figure 7: The east part of the mastaba in E13.7.6, bisected by a later wall. © Amara West Project (British Museum).

Figure 8: Fragment of painted plaster F5133d from the niche in E13.7.6 [4561]. © Amara West Project (British Museum).

AMARA WEST

15

was partly recorded using visible-induced luminescence imaging (VIL). Traces of red, yellow, and black pigments were visible by eye. Colour in the cemeteries at Amara West: coffins

Figure 9: Traces of red and yellow paint in E13.7.5 [4566].

bands (P. Spencer 1997, 175). There was also evidence of a cavetto cornice, which had been painted red and black. The 1930s excavators also found fragments of painted plaster in the floor fill of E.12.3, decorated with ‘squares, rosettes and other patterns in red, blue, white and black paint’ (P. Spencer 1997, 172), which brings to mind the decorated pieces of plaster from E13.7.6 (Fig. 10). The West Gate into the walled town at Amara West was documented by the EES excavators, who made an epigraphic copy (unpublished), and it was re-excavated by the British Museum team in 2015 (N. Spencer 2017, fig. 5). There is trace evidence of extensive polychrome decoration on the interior walls of the gateway, which

Two burial areas have been excavated by the British Museum project team: Cemeteries C and D. Cemetery C is 200m northeast of the walled town, and features tombs cut into an alluvial terrace along a wadi, with simple pit burials, multi-chamber tombs (the superstructures did not survive), and tumulus tombs with burial niches. One notable tomb, G244, combined multiple burial chambers beneath a tumulus (Binder 2011; Binder 2017). Ceramics from the tombs indicate that the cemetery was in use from the late Ramesside Period until the 8th century BC (Binder 2011). Cemetery D is set upon the desert escarpment north of the town, on the other side of the now dry Nile channel. Alongside further pit burials, tumulus tombs, and burial chambers lined with vaulted brick structures, a number of tombs featured pyramids fronted by rectangular chapels (Binder, Spencer, and Millet 2010; Binder 2017). No evidence of painted decoration survives from these chapels, but parallels from other New Kingdom cemeteries in Nubia indicate such monuments were often painted (Smith and Buzon 2017). The cemetery was first used in Dynasty 19 and continued in use (and reuse) through to the 8th century BC (Binder, Spencer, and Millet 2010; Binder 2017). Coffins, badly damaged by looting and termite activity, were recovered in fragments, many of which bore painted decoration. The wood was plastered before being painted in white, black, red, yellow, and blue. Painted plaster was found as part of coffin fragments in tombs G244 (Dynasties 19–20), G201 (late New Kingdom to early Napatan Period), G222 (10th–9th centuries BC), G238 (9th–8th centuries BC), all in Cemetery C, and G301, G309, G320, and G322, all pyramid tombs in Cemetery D, Dynasties 19–20. Nubian and Egyptian identities at Amara West

Figure 10: Fragment of painted plaster F5049 from E13.7.6 [4561]. © Amara West Project (British Museum).

The cultural context of Amara West was markedly different from that of towns situated in Egypt. In all likelihood there were significant Nubian communities who had not, or only partly, adopted Egyptian culture, religion and languages (Smith 2003; N. Spencer 2010). Their indigenous traditions may have influenced the technologies and practices within the towns founded by

16

CHAPTER

the Pharaonic state. During the New Kingdom, Egyptian campaigns reached as far south as Kurgus, beyond the Fourth Cataract (Davies 2017), and Egyptian towns were established up to the Third Cataract, including Amara West. Former models of the impact of the Egyptians in the region focused on the Egyptianization and cultural domination of local populations, but more recent research has reconsidered the situation as one of interaction, entanglement, hybridization and two-way influence between the two cultures (Spencer, Stevens, and Binder 2017). Ethnicity is both self-defined and ascribed by others, and is usually based on kinship or common ancestry, a common language, and unified by constructions of the past (Emberling 1997, 304), but new ethnic groups can be formed at the edge of an expanding state when previously independent people are forced together (Emberling 1997, 308). Ethnicity is not expressed solely materially, in ways that are accessible in the archaeological record, and it may be that we have lost evidence of a more significant Nubian contribution. It is possible that people in the area were flexible as regarded their ethnicity, and affiliated themselves with whichever group it was most politic to belong to, projecting their identity depending on what they wanted to achieve (Nash 1987; Emberling 1997, 306). Ethnicity and identity are complex ideas that tend to be very fluid, and are represented and experienced in a variety of ways in life, many of which may not appear in the archaeological record. It has proved difficult in other studies to ascertain whether Egyptians or nonEgyptians inhabited towns in Nubia, and to what extent these terms might have been applied to the people that lived there, either by themselves or by other people (Buzon 2006). Burials at Amara West feature architecture, placement of the bodies and grave goods that reflect both Nubian and Egyptian cultures (Binder 2017). One Middle Kerma (c. 2050–1750 BC) tomb in Cemetery D (Spencer, Stevens, and Binder 2014) underlines how this area had been used for burials long before the period of Pharaonic control, and tumulus tomb G244 in Cemetery C, dating to Dynasty 19–20, demonstrates the continuity of Nubian funerary practices (Binder 2017). This tomb features a tumulus superstructure but an Egyptian-style substructure, with apparently Egyptian-style grave goods that, on closer inspection, may demonstrate a local interpretation of Egyptian traditions (Binder 2017, 597–604). Nubian-style niche graves and tumuli then become the favoured style of tomb at Amara West very shortly after the end of the New Kingdom (Binder 2011): indigenous funerary traditions persisted

1

through colonial occupation and then re-emerged as dominant in the aftermath of Egyptian control. A Bes amulet (F9453) from a Nubian-style niche burial G216 in Cemetery C, post-dating the New Kingdom, evokes the hybridized identities that may have emerged (Binder 2017). Typical of Egyptian depictions of Bes, a household deity, in its stance and bodily form, the face is rather un-Egyptian and probably represents a local adaptation (Binder 2011, 46, pl. 13). Pottery assemblages from the settlement comprise between 1% and 10% Nubian vessels, typically cooking pots, similar to Askut (S. T. Smith 2003, 115; Spataro, Millet, and Spencer 2015). Analysis of the Amara West ceramics shows that vessels of both Egyptian and Nubian type were most likely made locally, and others were imported from elsewhere, including Egypt and the Levant (Spataro, Millet, and Spencer 2015), and Mycenaean vessels came from the Greek mainland and Cyprus (Spataro et al. 2019). Although the town architecture is largely consistent with that found in contemporary Egyptian settlements, one oval building (E12.11) at Amara West is more consistent with Nubian architectural tradition (N. Spencer 2010). The lack of leather, basketry, and woodwork surviving in the town, most likely due to termite activity, may well mask further aspects of Nubian material culture and practices located at Amara West. Natural resources around Amara West The geology of an area determines the minerals available to the ancient population and thus is of significant relevance to the study of ancient pigments, most of which were obtained from rocks. The rocks available to the ancient people of Egypt and Nubia are considered only briefly here, but are dealt with in much more detail elsewhere (Klemm and Klemm 2008; Said 1990). In Egypt, the areas immediately adjacent to the Nile and its branches in the Delta are covered with deposits of Nile alluvial silt. Late Cretaceous to Neogene marine carbonates, predominantly limestones, make up the main exposed bedrock extending 200km east and west of the Nile from Cairo. South of Esna and into Sudan the Nile cuts through the extensive Cretaceous Nubian Sandstone Formation. The mountainous region of the Eastern Desert and the Red Sea coast are composed of Neoproterozoic igneous and metamorphic assemblages: granitoids, gneiss, greywackes, and volcanic rocks.

AMARA WEST

Gypsum quarries in Egypt are known in the Fayum at Umm el-Sawan and near Qasr el-Sagha, dating to the Early Dynastic Period and Old Kingdom, and one is located on the Red Sea coast near Wadi el-Anbaut, dating to the Roman Period (Harrell and Storemyr 2009; Harrell 2017). Recently Harrell (2017) has identified extensive gypsum deposits in a thin soil on the limestone plateau to the north and east of Amarna, with quarries dating to the Old to New Kingdoms. These deposits are calcite rich, and so were not the ones used for the Amara West buildings. In northern Sudan, Nubian Sandstone is the predominant lithology in both outcrop and sub-cropping beneath recent windblown sand. Around the Nile south of Wadi Halfa (close to the present-day Egyptian– Sudanese border) there are schists with occasional small outcrops of volcanic and intrusive igneous rocks. The local mineral resources available to the people living in north Sudan were thus quite different from those available to people living in the centre of Egypt. There is little information available on the distribution of earth pigments, but one might assume that they were widely available in Egypt and Nubia, including yellow and red ochre, and green earth. Certainly in the vicinity of Amara West there are visible outcrops of

17

red and yellow soft rock, some of which have visible evidence of small-scale mining. Local people in the areas around Abri describe collecting rock from these outcrops to use in painting their houses, and mention other sources that were close by but not visible from the road, including a source of white gir (a word used for soft rock that can be used as plaster or paint) near to Amara West itself (see Chapter 4: Ethnoarchaeology). Modern local people claimed that they purchased all of their gypsum in the market, and prior to that it was not used. However, one interviewee described digging for gypsum in the desert, so it is possible that a source was known and exploited in the area. Gypsum can occur as veins in sedimentary rocks (Harrell 2017) and nodules of the material are plentiful in the sand around Amara West and could conceivably have been gathered and heated to make plaster. Calcite can form in spars in fissures within rocks, and thus can occur in small quantities in a wide variety of environments (Kerr 1959, 212), and may have been available around Amara West in this form. A largescale source of calcite can be found in the form of marble at the Third Nile Cataract (Feneuille, Letourneu, and Bouchar 2014).

CHAPTER 2

COLOUR IN ANCIENT EGYPT

The way in which people conceive colour is strongly influenced by their culture and history; in the West our concept of colour has been shaped by the development of scientific techniques, technology, and medicine (Saunders and Brakel 2002; Young 2006). Sciencebased explanations of colour use physics to describe how humans experience colour: visible light is part of the electromagnetic spectrum and this is made up of waves of different wavelengths. The human eye has two types of light receptor, called rods and cones. Rods can detect a single photon; they allow humans to see in dim light (Baylor 1995). There are three types of cones, which are receptive to long-, middle- and shortwave light, and by detecting differences between the relative intensities of these light waves, the cones can allow the brain to distinguish a huge variety of colours (Stockman and Sharp 1999). Colours are measured within a colour space, in terms of saturation, hue, and brightness. Modern science thus dematerializes colours into wavelengths of light, quantifiable, and also universally received and translated by the human cognitive system (Saunders and Brakel 2002; Young 2006). Coming from this scientific tradition, Berlin and Kay conducted a now famous experiment in San Francisco among people of different linguistic backgrounds, with the aim of identifying universal ‘Basic Colour Terms’ (BCTs) across languages (Berlin and Kay 1969). Based on this, and many subsequent revisions (Kay 1975; Kay and McDaniel 1978; Kay and Kempton 1984; Kay, Berlin, and Merrifield 1991), they concluded that BCTs existed as a ‘natural law’, building blocks of language, which develop in stages, from simple black and white to the complete description and understanding of the colour space (Kay and McDaniel 1978). The BCTs are said not only to be elemental units of language, but to be ‘derived directly from the neural response patterns that underlie the perception of color’ (Kay and McDaniel 1978, 630). There have been many criticisms of the BCT hypothesis. The method is criticized for assuming a static, universal system that will only manage to test its own assumptions, including that all colour terms can be described using English terms (Saunders and

Brakel 2002). Other features that have been criticized include: giving greater status to abstraction, for example the use of the word ‘blue’ rather than ‘like the sky’ (Gellatly 1995); using Munsell colour chips as a test set, a stimulus that is relevant to English and the way our culture thinks about and references colour, but may have little relevance for other cultures (Levinson 2000); failing to consider surface properties other than colour, for example texture or wetness (Gellatly 1995; Levinson 2000). Those who criticize the Berlin and Kay hypothesis argue instead for an anthropological approach, which discusses colours not in the abstract, but in context (Saunders and Brakel 2002; Young 2006). Their approach does not assume that colour is a universal natural thing, but makes colour a creation of the person, situation, and socio-historical moment. An anthropocentric, or phenomenological, perspective on colour would ask: ‘why does the object have the colours it does?’; ‘what do those colours do for the object?’; ‘what kind of effect do they have?’ (Young 2006). In the Egyptian written language there were four colour terms: black (km), white (ḥḏ), red (dšrt) and green (wꜢḏ), and a term used for animal skins and feathers that might be translated as ‘variegated’ (sꜢb) (Baines 1985). However, all of these terms had other meanings and it is not clear from texts that the words refer abstractly to colour. Although the colour spectrum can be described by physics, cultural divisions of this quality are far more problematic, being embedded in cultural meanings (Quirke 2001; Young 2006). The word for white also had the meaning ‘silver’ and ‘bright’; black had associations with fertility (black soil) and Osiris, was the colour of underworld demons, and the favoured colour for magical statues (Wilkinson 1994, 109–10; Taylor 2001, 166). Green was also connected to Osiris, and with meanings that are sometimes difficult to translate but seem to revolve around ideas of freshness, moistness, living, and growing (Wilkinson 1994, 108; Pinch 2001). The word for red was also used for ‘desert’, and was associated with things that were considered to be powerful and dangerous (Pinch 2001).

20

CHAPTER

Much more certain are the Egyptian terms for pigments and minerals: lapis lazuli (ḫsbḏ), turquoise (mfkꜢt), orpiment (ḳnỉ), realgar (Ꜣw-ỉb), red minerals (mnš/prš), red ochre (ṯmḥy), and ochre (sty) (Quirke 2001; Lesko and Lesko 2002). There was no basic colour term for blue, but the word for lapis lazuli was used to describe other substances made of blue, the term ‘true lapis’ (ḫsbḏ mꜢꜤ ) being reserved for the mineral itself; there is no specific word recorded for Egyptian blue pigment (Harris 1961). It has been suggested that colour was not as important an identifying feature for Egyptians (and other societies) as were size, shape, location, and brightness and texture (Lyons 1995; Quirke 2001). In the modern world we are surrounded by shiny surfaces, such as metals and glass, but in ancient times these materials were much more scarce, and therefore the shine or sparkle effect they produced was probably more impactful (Hurcombe 2007). Faience was the most widely available luxury material; small objects made of faience have been found in the domestic areas of Amarna, indicating that most people here had access to this material to some degree (Shortland, Nicholson, and Jackson 2001). Similarly, small faience beads are found throughout the town of Amara West, suggesting that access to this material was widespread, if limited in scale. The Egyptian word for faience was ṯḥnt, meaning ‘brilliant’, ‘gleaming’ or ‘dazzling’ (Friedman 1998), words which refer not to its colour but to the effect of light across its surface, indicating that the defining feature of faience was this surface effect. The wider use of the word used for green (wꜢḏ) to describe plants, and amulets made from shiny and semi-transparent materials (Pinch 2001, 183), reinforces this notion. The introduction of a golden micaceous slip to Egyptian pottery forms in Nubia in the Middle Kingdom evidences a similar interest in glimmering, shimmering surfaces (Knoblauch 2011). Paints and pigments in ancient Egypt The ancient Egyptians used a variety of naturally occurring and manufactured pigments to paint their tombs, both royal and private, coffins, temples, palaces, houses, textiles, and objects of wood, stone and ceramic (although in general, ancient Egyptian pottery is not heavily decorated), and to write on and illustrate papyri. There are also coloured objects that use minerals to impart colour, for example glass and faience vessels, jewellery, and inlays. Examples of plastered and

2

decorated walls are known from many sites in Egypt. In some cases the evidence is extensive, such as in the large houses at Amarna, and in other cases only a few fragments of plaster remain. The latter situation is often the case for humbler private houses, where the evidence for the use of paint relies on good excavation records. In some cases pieces of pigment and sherds used as palettes have also been excavated, but these are not always mentioned in the archaeological reports, and even then are sometimes summarized very briefly. Very few of the paints and pigments from domestic areas have been scientifically analysed. Most scientific studies of pigments are on elite funerary structures, temples, and objects in museums, most of which also originate from the elite of Egyptian society. In order to organize the evidence to compare it to the material from Amara West, the discussion below presents the results of previous studies on Egyptian pigments on various types of objects by colour, which is often how they are reported, then presents the evidence for the use of paints and pigments in houses in Egypt (with pigment identification where it exists). Pigments identified from ancient Egypt The analysis of Egyptian pigments has received considerable attention. The first summary was published in 1934 by Lucas (Lucas 1962) but this early work lacked modern methods of analysis. Probably the most comprehensive analysis project was undertaken by the Max Planck Institut in the 1980s. They analysed 1,350 samples of pigment from stone surfaces (tombs, sarcophagi, and temples), mostly from Thebes, dating from the Old Kingdom to the Roman Period (El Goresy et al. 1986; Blom-Böer 1994). This research led them to develop a theoretical palette used by the ancient Egyptians, varying over time and with status (royal versus non-royal). However, many of their conclusions have since been challenged by other researchers (PagèsCamagna and Colinart 2003; 2006; Schiegl and El Goresy 2006), and the palette they identified is gradually expanding as modern research methods, such as Raman spectroscopy, allow a wider range of pigments to be more precisely identified (Heywood 2001; Ambers 2004). The Centre de Recherche et de Restauration des Musées de France (C2RMF) published analysis of pigments from 300 objects in French museum collections spanning from the Old Kingdom to the Ptolemaic Period on substrates including limestone, wood, and mud plaster (Colinart, Delange, and Pagès

COLOUR IN ANCIENT EGYPT

1996; Pagès-Camagna and Guichard 2010). The British Museum has done a great deal of work on pigment analysis, from papyri (Lee and Quirke 2000; Daniels and Leach 2004), tomb paintings (Ambers 2004; 2008) and coffins (Middleton and Humphrey 2001). There have also been many studies of individual tombs, temples and objects (or groups of objects). A summary of the range of pigments analysed and the contexts and collections from which they were selected is given below.

21

white (lead carbonate hydroxide, 2PbCO3·Pb(OH)2) was used by the Greeks and Romans (Caley 1946; Varone and Béarat 1997; Mazzocchin et al. 2003). It has been identified from ancient Egypt on cartonnage dating to the Graeco-Roman Period, but not before (Scott et al. 2003; Scott et al. 2009; Rowe, Siddall, and Stacey 2010). Different white pigments have been found within one painting, suggesting that the whites were used for different purposes, for example calcium sulphate for the background and huntite for details (Heywood 2001; Ambers 2004; Ambers 2008).

White White is widely found and is most often identified as gypsum or anhydrite (hydrated and anhydrous calcium sulphate, CaSO4, respectively) or calcite (calcium carbonate, CaCO3). Calcite is the main component of limestone, commonly found in the Nile valley in micritic form (Hussein 1990); most studies do not describe the morphology of the calcite. Both micritic limestone and gypsum are very soft and easily crushed to make a pigment. Often white plaster from Egypt is identified in the literature as ‘lime’ or ‘gypsum’, with no scientific basis. Lime (calcium oxide, CaO) is produced by calcining (burning) calcium carbonate (for example, limestone) above 750°C (Regev et al. 2010; Feneuille, Letourneu, and Bouchar 2014). To make lime plaster, water is added to the lime to produce slaked lime (calcium hydroxide). As it dries, this reacts with carbon dioxide in the air to form lime plaster (calcium carbonate). Lime and gypsum plaster are difficult to distinguish without scientific analysis, although gypsum is softer, and lime plaster will ‘fizz’ in mild acids such as lemon juice. For many years it was thought that calcium sulphate and calcium carbonate were the only two white pigments used, but since its first identification in 1974 on Nubian pottery from c. 1600 BC (Riederer 1974), huntite (a calcium magnesium carbonate, Mg3Ca(CO3)4) has been found on objects and paintings from the Old, Middle and New Kingdoms (Noll 1978; El Goresy et al. 1986; Middleton 1999; Heywood 2001; Uda et al. 2000; Middleton and Humphrey 2001; Ambers 2004). Several sources of huntite have been found within Egypt and it is thought that it would have been quite widely available (Heywood 2001). Huntite is a soft, fine powder that gives an intense pure white colour, and seems to have been used either alone as a bright white, or under other colours to increase their vibrancy (Lee and Quirke 2000; Uda et al. 2000; McCarthy 2001; Ambers 2004; Ambers 2008). Lead

Black Nearly all ancient Egyptian black pigments that have been examined have been identified as carbon (El Goresy 2000; Lee and Quirke 2000). Morphological and structural studies make it possible to determine whether the carbon is a result of burning plant materials, soot, or crushed charcoal (Pagès-Camagna, Colinart, and Guicharnaud 2004). The presence of phosphate indicates that the carbon was obtained by burning bone or ivory, which contain apatite (Winter 1983; Pagès-Camagna and Guichard 2010, 26–27). Carbon is, of course, very readily available, being the result of burning any organic matter. In a few cases, manganese pigments have been identified on Egyptian artefacts. Manganese-based pigments were found in several areas of relief from the Dynasty 12 tomb of Djehutynakht at Deir el-Bersheh (Middleton 1999, 42), and in the wall paintings from the palace at Malqata (Lacovara and Winkels 2018), and used as an ink, along with carbon black, on a Dynasty 18 shroud (Pullan et al. 2012). The analysis methods used did not allow a more precise identification, although the Raman spectra for the ink on the shroud were a close match to published reference spectra for manganite (MnO(OH)) (Pullan et al. 2012, 20). Manganese was extensively used for the black lines decorating faience objects (Tite, Manti, and Shortland 2007) and for the purple colouring of faience and glass from Dynasty 18 (Shortland 2002, 518). Manganese deposits are known in the Sinai, the eastern desert, to the north of Hurghada, and the southeastern desert near the modern border with Sudan (Hussein 1990). There are also significant manganese deposits around the copper mines at Timna in the Negev (modern Israel), which were heavily exploited by Egyptians in the Ramesside Period (Weisgerber 2006).

22

CHAPTER

Bitumen was used extensively in ancient cultures for both practical and decorative purposes (Forbes 1936a, 80; Connan 1999), but its use in Egypt is largely unknown or undocumented outside of mummification (Newman and Serpico 2000). Ancient bitumen deposits are known from the Middle East and Egypt (Forbes 1936a; Connan 1999; Serpico 2000), important sources for Egypt being the Dead Sea (Connan, Nissenbaum, and Dessort 1992), and seeps along the coast of the Red Sea (Harrell and Lewan 2002; Barakat et al. 2005). Bitumen is solid or semisolid fossilized organic matter that has not migrated from its source rock (Peters, Walters, and Moldowan 2005, 991). Its physical and chemical properties depend on the organic matter that formed it; these variations can be used to identify and provenance bituminous materials (Peters, Walters, and Moldowan 2005). The terms asphalt and bitumen are often used interchangeably in the archaeological and scientific literature (Forbes 1936b; Chilingarian and Yen 1994); since the term ‘bitumen’ can be used as an umbrella term for these substances, it is the one employed here. References to the ancient use of ground bitumen as a pigment are rare (Connan et al. 2004). Three instances of the use of powdered bitumen as a pigment in ancient Egypt have been identified by Ruth Siddall (2011) on objects from the Petrie Museum (UCL): two on Roman Period mummy masks and one on a model boat from a Dynasty 19 Tomb (611) at Gurob. Yellow The most commonly identified yellow pigment in Egypt is yellow ochre (Bikiaris et al. 2000; Lee and Quirke 2000; Pagès-Camagna and Guichard 2010). The exact chemical composition of ochres varies, but they are usually mixtures of iron oxide hydroxides, mainly goethite (α-FeOOH), plus clays and quartz, according to the natural environment in which the ochre has formed (Hradil et al. 2003; Helwig 2007). Yellow and red iron oxides are distributed throughout the Nile valley (Hussein 1990; El Goresy 2000). Because natural iron ochres are so widely available and have low mineralogical variability, the geographical source of the pigment is very difficult to identify (Hradil et al. 2003). Two other yellow pigments have been identified on ancient Egypt objects and paintings: jarosite (and natrojarosite), and orpiment. These pigments can be

2

initially distinguished from ochre visually, although analysis is required to confirm observations. Jarosite has a lemony colour (Colinart 2001) as compared to the dull yellow given by ochre, and orpiment is laminar and therefore sparkles (Lee and Quirke 2000). It has been claimed that pigments identified as jarosite (KFe 3(OH) 6(SO 4) 2) or natrojarosite (NaFe3(OH)6(SO4)2) are the degradation products of another pigment, perhaps brown-green in colour (Schiegl, Weiner, and El Goresy 1989; Schiegl, Weiner, and El Goresy 1992; Lee and Quirke 2000), but further recent identifications confirm jarosite as a pigment in its own right (Le Fur 1994; Colinart 2001; Middleton and Humphrey 2001; Ambers 2004; Ambers, Stacey, and Taylor 2007; Pagès-Camagna and Guichard 2010). Sources for jarosite have been identified within Egypt, around the Dakhleh Oasis (Berry 1999) and in the Western Desert near Aswan (Hussein 1990). Jarosite has been identified mostly on material from the Old and Middle Kingdoms. Orpiment (an arsenic sulphide, As2S3) has a bright glittery effect due to its laminated structure. A Max Planck Institut project in the 1980s identified orpiment as a pure pigment only on royal sarcophagi and royal tomb walls in Dynasty 18; they posited that it was subsequently adopted by non-royals for limited use on coffins in the late New Kingdom (El Goresy et al. 1986; El Goresy 2000). However, orpiment is now definitely known from the Middle Kingdom, and very probably also from the Old Kingdom (Lee and Quirke 2000; Colinart 2001; Middleton and Humphrey 2001; Pagès-Camagna and Guichard 2010), as well as on non-royal tomb walls, shrouds and papyri of Dynasty 18 (Lee and Quirke 2000). Many researchers have found ochre and orpiment present together, usually in layers (Green 1995; Colinart 2001; McCarthy 2001). Orpiment is thought to have been imported into Egypt, there being no local source (El Goresy 2000), but Le Fur (1994) states that it occurs alongside the gold ores in Nubia and the copper ores in the Sinai, so it is possible the Egyptians could have mined it themselves. Orpiment pigment was found on the Uluburun shipwreck (14th century BC), which suggests that it was being traded around the Mediterranean region (Bass 1986). Different yellow pigments were sometimes used within one composition, indicating that each was chosen for its specific shade and effect (McCarthy 2001; Pagès-Camagna and Guichard 2010).

COLOUR IN ANCIENT EGYPT

Red Yellow and red pigments are often discussed together because the most common of each is a form of iron oxide (Lucas 1962; Lee and Quirke 2000). Red iron oxides are the anhydrous form. The main colourant is haematite (αFe2O3), although since the pigments used almost always contain other materials, such as quartz and clays, the pigment tends to be referred to by the less specific name of red ochre or red earth (Helwig 2007). Pure haematite occurs as a mineral and was used as a pigment in ancient Egypt (Froment, Tournié, and Colomban 2008; Pagès-Camagna and Guichard 2010, 28). Iron oxide minerals are widely available, and frequently local sources are used for pigments (Helwig 2007, 65). However, some particularly high-quality earth pigments were traded over long distances; Greek and Roman sources document known sources of good earth pigments (Helwig 2007, 65–66). It is likely that ancient Egyptians would have used local sources of red ochre, but possible that it was also traded within the country or internationally. A less commonly used red/orange pigment was realgar (an arsenic sulphide, AsS), which is very unstable and decomposes to pararealgar, which is yellow, and has the same chemical formula as realgar but a different structure (Douglass, Shing, and Wang 1992; Green 2001; Daniels and Leach 2004). Thus far it appears that the use of realgar was very limited compared to the widespread use of ochre (Pagès-Camagna and Guichard 2010). Realgar has been found in Egypt on New Kingdom objects: on papyri (Daniels and Leach 2004); in tomb paintings (Blom-Böer 1994; Vandenabeele et al. 2009); as raw pigment from Amarna (Saleh et al. 1974; David et al. 2001); and on wood, stone, and cartonnage objects from the New Kingdom onwards (Pagès-Camagna and Guichard 2010, 28). Realgar is found in geological association with orpiment and therefore may have been obtained from the same source (Fitzhugh 1997; Daniels and Leach 2004). Two other red pigments have been reported from Egypt, an association of iron and arsenic (possibly realgar with ochre?), mixed with huntite and used for skin colour in the Dynasty 18 Theban tomb of Neberhebef (Pagès-Camagna and Guichard 2010, 28), and cinnabar (mercuric sulphide, HgS). Cinnabar is the name given to the natural mineral, and vermilion is the name for synthetic forms (Gettens, Feller, and Chase 1972). Cinnabar has been found on two items now in the

23

Louvre, one dating to the Ramesside Period and one to the Third Intermediate Period (Pagès-Camagna and Guichard 2010), on a Dynasty 26 coffin (Bonizzoni et al. 2011), and on a funerary papyrus of the late Ptolemaic or early Roman Period (Lee and Quirke 2000, 113). Cinnabar was used as a pigment in Classical Greek, Hellenistic, and Roman times, although its use was probably limited by its high cost (Gettens, Feller, and Chase 1972; Kakoulli 2002; Mazzocchin, Baraldi, and Barbante 2008). It was often adulterated by red lead (Pb3O4, also known as minium), a more common red pigment manufactured from the heating of white lead, a process written about by Vitruvius and Pliny (Fitzhugh 1986, 110). Red lead has been found on Egyptian cartonnage dating to the Graeco-Roman Period (Scott et al. 2003; Rowe, Siddall, and Stacey 2010). Madder is an organic dyestuff derived from the roots of the Rubiaceae family, notably the Rubia genus (Eastaugh et al. 2004a, 250; Daniels et al. 2014). It can be precipitated or adsorbed onto an inorganic substrate, commonly aluminium hydroxide or calcium sulphate, to form a pinkish red colourant, known as a lake pigment (Daniels et al. 2014). Madder was used in Egypt from Dynasty 18 for dyeing textiles, and numerous examples are known from elite and non-elite contexts (Wouters, Maes, and Germer 1990; Germer 1992). Madder lake pigments in Egypt have been identified on Roman mummy portraits (Cartwright and Middleton 2008), Roman mummy cartonnage (Scott et al. 2009; Rowe et al. 2010), and in a pot from a Roman tomb at Hawara (Russell 1892). A pink pigment consisting of an organic substance on a gypsum ground was found on a Third Intermediate Period papyrus and tentatively identified as madder (Lee and Quirke 2000). Madder was widely used in the ancient world and has been documented in Greek, Roman, Mesopotamian, and Parthian contexts (Daniels et al. 2014). Blue In the vast majority of cases the blue pigment found in ancient Egypt, from the Early Dynastic Period onwards, is a synthetic pigment known to modern researchers as Egyptian blue, or blue frit (Lee and Quirke 2000; Hatton, Shortland, and Tite 2008; Moussa and Ali 2013; Corcoran 2016). It is a manufactured pigment, made by heating together a source of each of calcium, silica, copper and a small amount of

24

CHAPTER

alkali flux at 900–1000ᵒC (Tite, Bimson, and Cowell 1984; Tite, Bimson, and Cowell 1987; Hatton, Shortland, and Tite 2008). A multiphase material is formed, in which the blue colour is given by calcium copper silicate crystals (CaCuSi4O10), the synthetic analogue of the rare mineral cuprorivaite (Warner 2011, 43), the other components being unreacted quartz and a glassy phase that holds it all together (Tite, Bimson, and Cowell 1987, 40). The source of the silica used to manufacture Egyptian blue was most likely sand (Jaksch et al. 1983, 533; Pagès-Camagna and Colinart 2003, 643; Hatton, Shortland, and Tite 2008, 1598), and the calcium was probably introduced from calcium carbonate in the sand (Tite, Bimson, and Cowell 1987, 40; Hatton, Shortland, and Tite 2008, 1598). The copper was introduced as a copper oxide probably obtained from roasting scrap metal (Hatton 2005, 55). Some Egyptian blue has been found to contain tin, arsenic, and lead in the ratios represented in contemporary copper alloys, indicating that these were the source of the copper used to manufacture the pigment (Tite, Bimson, and Cowell 1987, 40). Analysis of sodium oxide (Na2O) to potassium oxide (K2O) ratios in samples of Egyptian blue suggest that either natron (sodium carbonate) or plant ash (containing potassium) may have been used as a flux, perhaps depending on the availability of the materials to the manufacturer (Jaksch et al. 1983; Pagès-Camagna and Colinart 2003; Hatton, Shortland, and Tite 2008, 1599). The use of plant ash as a flux allows for the localized production of synthesized pigments, plant ash being more easily obtainable than natron. During Pharaonic times, Egyptian blue was formed into flat cylindrical cakes that could then be transported and ground for use; such cakes have been found in a Middle Kingdom temple at Karnak, and New Kingdom examples come from Amarna in a ‘factory area’ (Dynasty 18), the craftsmen’s village at Karnak, the courtyard of the tomb of Keruef at Thebes (Dynasty 18), and the workmen’s village of Deir el-Medina (Saleh et al. 1974; Weatherhead and Buckley 1989; Rouchon et al. 1990; Pagès-Camagna, Colinart, and Coupry 1999; Pagès-Camagna and Colinart 2003; Hatton, Shortland, and Tite 2008, 1593). Evidence for the production of Egyptian blue comes from Amarna and Pi-Ramesse (modern Qantir), both royal cities of the New Kingdom. Workshop areas QI and QIV at the Ramesside city of Pi-Ramesse were at some point levelled to create military workshops, stables, and a parade ground, but within the floor debris

2

was evidence for bronze-working, glass production, faience production, and Egyptian blue in the form of fragments of slabs, cakes, bags, and half-finished objects (Rehren and Pusch 1997; Pusch and Rehren 2007). A cream-coloured plaster-like material that the excavators describe as ‘Schamotte’, consisting of calcium carbonate and silica, frequently occurs at Qantir (Pusch 1999). A crucible formed from ‘Schamotte’ with adhering pieces of Egyptian blue was excavated from area QIV, and points to the primary production of Egyptian blue at the site (Pusch 1999; Rehren et al. 2001). This plaster-like material is also found at Amarna, as are cakes of Egyptian blue (Weatherhead and Buckley 1989; Nicholson 2007). Evidence from Amarna also indicates the existence of industrial high-temperature works: glass-making and -working, faience manufacture and Egyptian blue manufacture (Nicholson 2007; Nicholson 2008; Smirniou and Rehren 2011). It has been suggested that some hightemperature industries, notably glass production, and possibly the manufacture of Egyptian blue, were centralized in state workshops (Rehren 1995; Shortland, Nicholson, and Jackson 2001; Smirniou and Rehren 2011). However, the manufacture of Egyptian blue and faience appears to be at a smaller scale within the industrial facilities investigated at Qantir (Rehren, Pusch, and Herold 2001), and evidence for faience manufacture at Amarna (in the form of moulds) is found not only in the larger industrial ‘factory’ areas such as O45.1, but also within residential areas, in workshops that may have operated partly as private producers of faience (Shortland 2000; Shortland, Hope, and Tite 2001). The wide distribution of faience objects has led to the consensus that faience manufacture was a cottage industry, taking place within houses (Shortland, Hope, and Tite 2001). The glass industry, in contrast, is argued to be state controlled based on textual evidence (association with cartouches), its use as royal tribute, the evidence from Qantir, and the distribution of glass at Amarna, where it is primarily found in palace dumps rather than houses (Shortland, Hope, and Tite 2001). It is unclear where the production of Egyptian blue falls on this spectrum of state to domestic production. To produce a successful result, a reasonable level of knowledge and control over the process is required, which makes it likely that the production of Egyptian blue was a specialist skill, but the level of state control over the process is not known. Different shades of Egyptian blue could be deliberately manufactured (Tite, Bimson, and Cowell 1987;

COLOUR IN ANCIENT EGYPT

Pagès-Camagna and Guichard 2010, 30). A dark blue can be achieved by producing a coarse-textured pigment, in which there are clumps of blue crystals, increasing the density of the blue colour. The light blue has a fine texture in which the blue crystals are more evenly distributed in the glassy phase, which masks the blue colour (Tite, Bimson, and Cowell 1987). Cobalt blue (cobalt aluminate, CoAl2O4) was used as a pigment on ceramic during Dynasties 18–19, most notably at Amarna; this use of the pigment on ceramic is the only use attested for cobalt as a paint in Egypt (Shortland, Hope, and Tite 2006), though it was employed a colourant in blue glass and glazes in the New Kingdom (Rehren 2001; Tite and Shortland 2003). Early attestations of azurite (a basic copper (II) carbonate, Cu3(CO3)2(OH)2) used as a blue pigment in ancient Egypt are unreliable (Lucas 1962, 343; Lee and Quirke 2000); however, a pigment lump from Amarna was identified as azurite, as was the blue pigment on a painted leather fragment dated to Dynasty 18 (David et al. 2001; Heywood 2010). Azurite is unstable and may degrade to malachite, which might explain the lack of further identifications (Scott 2010b, 35), or it may have been very rarely used. Ground lapis lazuli (natural ultramarine; the blue component is lazurite which is a complex silicate material with approximate formula (Na,Ca)4-8Al6 Si6O24(S,SO4)1-2) has also recently been found as a pigment, for the first time in an Egyptian context, on a stone bust dated to the Second Intermediate Period (Heywood 2010). It is perhaps surprising that it has not been previously identified, since it is known that the Egyptians used carved pieces of lapis lazuli decoratively. The likely reason for this is that simply grinding lapis lazuli results in a dull grey-blue pigment; to achieve a bright blue, improved extraction methods are required to increase the lazurite yield, and these appear to have been introduced after AD 1200 (Eastaugh et al. 2004a). An ‘optical’ blue is known from First Intermediate Period tomb chapels. The blue effect is made by applying black charcoal over a white base, which creates an impression of a bluish-grey colour (Gettens and Stout 1958; Blom-Böer 1994; Howard 2003; Eastaugh et al. 2004a). Blue-dyed textiles are known from the New Kingdom; the source of the blue colour has been identified as indigotin, which is found in indigo and woad (Vogelsang-Eastwood 2000, 278). Indigo can also be used as

25

a pigment, and has been identified from the GraecoRoman Period in Egypt, but not from earlier contexts (Rowe, Siddall, and Stacey 2010; Scott 2010b). Green The most commonly reported green pigment from ancient Egypt is Egyptian green, also known as green frit. Its colour varies from pale green to turquoise to pale blue. It was manufactured in a similar manner to Egyptian blue, using the same ingredients, but in different ratios; in the manufacture of Egyptian green the calcium content exceeds the copper and the excess calcium is precipitated as a copper rich calcium silicate, also known as cupro-wollastonite ((Ca,Cu)SiO3), which gives the pigment its green colouring (Eastaugh et al. 2004a, 403). It is also observed that the soda and silica contents are higher in Egyptian green than in Egyptian blue (Pagès-Camagna and Colinart 2003; Hatton, Shortland, and Tite 2008). The dates during which Egyptian green was in use are not agreed upon. Pagès-Camagna and Guichard (2010, 29) state that it has been found on artefacts from the Middle and New Kingdoms. Jaksch et al. (1983, 532) say that they found it in several tombs of Dynasties 5–6. El Goresy (2000, 61–62) strongly claims that Egyptian green was used only from Dynasty 18, and that its invention was a new technological development. He refutes all earlier identifications of Egyptian green as mistaken identifications of deterioration products of other pigments. He suggests there was a range of earlier green pigments, which have now all degraded, but this is not a consensus point of view (Colinart 2001; Ambers 2004; Pagès-Camagna and Colinart 2006). Some researchers believe that the green copper compounds atacamite and paratacamite (polymorphous copper chlorides, Cu2Cl(OH)3) and/or malachite (copper carbonate hydroxide, Cu2CO3(OH)2) were used as pigments in ancient Egypt (Riederer 1974; PagèsCamagna and Guichard 2010; Scott 2010b), although others believe they are solely deterioration products of the synthetic green and blue pigments (Schiegl, Weiner, and El Goresy 1989; El Goresy 2000, 57). The latter seems unlikely given close examination of artefacts, the observed stability of Egyptian blue in many contexts (Green 2001; Daniels, Stacey, and Middleton 2004; cf. Lau et al. 2008), and the identification of atacamite and malachite alongside Egyptian blue (Green 1995; Lee and Quirke 2000, 112; Middleton

26

CHAPTER

and Humphrey 2001). Copper chlorides can be formed as deterioration products of Egyptian blue and faience where chloride salts are present (for example on tomb walls), but they were also used as green pigments in their own right (Giménez 2015). Atacamite occurs around copper ores in arid regions or as a corrosion product of copper; given its rarity as a naturally occurring mineral it is likely that it was synthesized for use as a pigment, in which case it should be referred to as copper chloride hydroxide, atacamite type (Eastaugh et al. 2004a, 33). Theophilus provides a recipe for a green pigment viride salsum, which produces atacamite, using copper, salt, and vinegar (Scott 2002, 281). Malachite was found as a colourant on a green inlay from a Dynasty 4 tomb (Ambers, Stacey, and Taylor 2007), and on a wooden coffin from Dynasty 6, with no evidence of a previous glassy pigment, for instance silicon or alkali metals (Green 2001, 44). Malachite was used to paint hieroglyphs on a First Intermediate Period or early Middle Kingdom coffin in the British Museum, on which Egyptian blue was also used for wedjat eyes, the two pigments being distinct (Lee and Quirke 2000, 112). Malachite pigment was also found to have been used to paint the face of a wooden coffin of Dynasty 26 (Scott 2010a). PagèsCamagna and Guichard (2010, 29) suggest that copper chlorides and malachite were used as pigments in the Old Kingdom and from the Third Intermediate Period, with Egyptian green being the green pigment of choice in the intervening period. Chrysocolla, a hydrated copper silicate mineral (variable formula Cu2-xAlx(H2-xSi2O5)(OH)4·nH2O), has been reported as a pigment from Amarna, from Karnak, and on the walls of Theban temples, the latter alongside Egyptian green (Riederer 1974; Weatherhead and Buckley 1989; Rouchon et al. 1990; Scott 2010b). Again, the identification of this pigment is refuted by some scholars (Schiegl, Weiner, and El Goresy 1989). A range of organo-copper complexes has been identified from New Kingdom, Late Period and GraecoRoman objects (Scott 2010b). These consist of copper or a copper-containing mineral mixed with an organic substance, with which the mineral reacts to produce a bright emerald green. A copper-proteinate pigment, probably manufactured from verdigris (a loose term used for copper corrosion products, usually copper chlorides or copper acetate (Cu(O2CCH3)2·2 Cu(OH)2) (Eastaugh et al. 2004a, 391)) and animal glue, was found on a Graeco-Roman cartonnage from the University of Southern California (Scott et al. 2003). A green

2

pigment from a second cartonnage fragment was found to be a copper-carbohydrate, originally made using a plant gum (Scott et al. 2004). Four samples of green waxy paint from museum objects studied at the British Museum were found to consist of beeswax and copper. Experimentation showed that a waxy paint could be achieved by heating these materials together (Daniels 2007). Two of the objects dated to the Late Period (a coffin and a corn mummy), and two to Dynasty 20 (both shabtis), indicating that organo-copper pigments were used during the New Kingdom. Two further examples of waxy green paint from Dynasty 19 jar fragments were also found to consist of beeswax and copper; experiments by the authors showed that a similar paint could be created by heating copper with malachite, chrysocolla, verdigris, or copper acetate, or by heating beeswax in a copper crucible (Liang and Scott 2014). Green earth is a naturally occurring pigment containing iron and consisting of a variety of siliceous clay minerals, depending on its geographical origin (Grissom 1986). The iron-rich micas, celadonite (K(Mg,Fe2+)Fe3+(Si4O10)(OH)2), and glauconite ((K,Na) (Fe3+,Al,Mg)2(Si,Al)4O10(OH)2) are usually the primary coloured components of green earth, but other greenish minerals have also been found, including the chlorites, serpentines, and smectites (Grissom 1986; Hradil et al. 2003; Ospitali et al. 2008). Green earth was widely used in the ancient world (Grissom 1986; Hradil et al. 2003). In Egypt it is attested on cartonnage from the Third Intermediate Period and the Graeco-Roman Period (Scott et al. 2003; Scott et al. 2009). The recent identification of green earth from the Dynasty 18 palace of Amenhotep III at Malqata (Lacovara and Winkels 2018) suggests that green earth may have been underidentified either owing to the subject matter that has formed the main part of the analyses to date, or owing to the comparative ease with which Egyptian green can be identified, because it is crystalline and contains copper. At Malkata the green earth occurs in conjunction with Egyptian green and Egyptian blue; a less careful analysis may have led to the green earth being overlooked. Greens were also achieved by mixing blue and yellow pigments. Egyptian blue and orpiment were used in the Dynasty 18 tomb of Menna (TT69) (Vandenabeele et al. 2009), and yellow ochre with Egyptian blue on the walls of the tomb of Amenhotep III, also Dynasty 18 (Uda 2004). Egyptian blue mixed with orpiment, with yellow ochre, and with green earth have

COLOUR IN ANCIENT EGYPT

all been found on Graeco-Roman cartonnage fragments (Scott et al. 2003; Scott et al. 2009), the mixing of blue and yellow to achieve green was frequently used by the Greeks (Kakoulli 2009). Paints and pigments in ancient Egyptian houses Very little domestic architecture has survived from the Old Kingdom (Correas-Amador 2013, 150). Plaster recovered from the Khentakawes town at Giza indicates that houses were usually plastered in white, sometimes painted red, which was usually in trace amounts, with the exception of room O in which there were ‘sheets of painted red wall plaster’ (Lehner, Kamel, and Tavares 2009, 62). There were also remains of black-painted plaster in the western town (Lehner, Kamel, and Tavares 2006, 70). More evidence for paint in domestic spaces is known from the Middle Kingdom. At Kahun, important rooms were commonly painted with a dado. Black was painted up to 90cm–1.5m from the ground, then an area of red lines on white; the upper part was yellow (CorreasAmador 2013, 399). At Lisht, which dates from the Middle Kingdom into the Second Intermediate Period, the walls of room h of house A.1.3., and the entrance chamber doorway and the antechamber, were plastered and painted 85cm high in black and yellow, separated by a white line. In house A.3.3. there were similar fragments of painted plaster in the living room: black and yellow sections separated by black, white, red and brown stripes (Correas-Amador 2013, 407). Paints and pigments from a few sites have been studied in more detail, and these are summarized below. Pigments at Elephantine There are towns and temples on the island of Elephantine dating from the Predynastic Period to the Graeco-Roman Period. From the Old Kingdom domestic architecture on Elephantine, red and yellow panels or stripes were found in the central room of a building in the ‘Oststadt’ (Ziermann 1988), and red, white and yellow fragments of painted mud plaster were found in area XXX (building of unknown function) (ForstnerMüller and Raue 2008, 133). Elephantine has only a painted fragment dating to the Middle Kingdom (von Pilgrim 1996), but from the New Kingdom has more evidence of colour. House 55 has painted scenes of boats and crew dating to Dynasty 17 (von Pilgrim 2016), and the Ramesside house H61 has fragments of

27

coloured plaster showing that the walls were painted in red, yellow and white (von Pilgrim and von Pilgrim 2007), and blue (D. Raue, pers. comm., August 2014). A large number of ceramic palettes were found in houses from across all the time periods at Elephantine, containing paint coloured white, yellow, blue and red (Pagès-Camagna and Raue 2016); House 55 alone contained more than 20 palettes, and more than 50 objects with pigments adhering, mostly stone tools (J. Budka, pers. comm, June 2020). Lumps of pigments were also found, from around the temples and in the city, but the authors are not specific about the dates or exact locations of the find spots (Pagès-Camagna and Raue 2016). In situ scientific analysis of the pigments was conducted on over 400 samples. From houses in the western part of the city, from near the Satet temple, south of the Khnum temple and south of the port, came 200 red fragments; the reds were all identified as haematite and dated from Dynasty 2 to the Second Intermediate Period. Greens were infrequent and came from the western part of the city and the Satet temple: 6 were identified as malachite, all from Dynasty 2; one Dynasty 6 piece was reported as sampleite; and one Dynasty 19 example as amazonite. Yellows came from near the Satet temple and houses; 65 pieces were all identified as goethite, dating from the Old Kingdom to the Late Period. Twenty samples of orange pigment from near the Satet temple, dating from Dynasty 2 to the First Intermediate Period, were reported to be ochre. Of the 11 blue examples analysed, 10 were Egyptian blue, dating from the New Kingdom to the Roman Period, and 1 of uncertain date was lapis lazuli. The authors included gemstones in the study; the lapis lazuli was polished on one side and therefore appears to be an inlay rather than a lump of raw pigment. Ten white samples were mainly found to be calcite, from the Old Kingdom to the Roman Period, with 1 example of anhydrite from the Roman Period. The 5 pinks were a mixture of calcite/huntite and haematite, all New Kingdom, and the 3 greys were calcite mixed with carbon black, also all New Kingdom. Pigments at Amarna Amarna, given its short occupation (around 15 years) and the sustained archaeological fieldwork undertaken from the late 19th century onwards, provides a good opportunity to understand the range of environments in which colour was deployed, and the types of pigments used.

28

CHAPTER

Elite houses at Amarna were elaborately decorated in bright colours, and small areas of painted walls from private houses are known from several areas. Large houses were extensively decorated in polychrome, including blue ceilings, and polychrome garlands, friezes, and false doors using yellow, red, blue, green, black, and white, e.g. V37.1, K50.1, N49.18 (Frankfort and Pendlebury 1933; Peet and Woolley 1923; Kemp and Stevens 2010a). Smaller houses also had polychrome decoration, e.g. M50.16, and finds of raw pigments, e.g. P46.13 (Peet and Woolley 1923). Of the 37 houses excavated in the workmen’s village in the 1920s, 7 had painted wall plaster, monochrome and polychrome, including dancing Bes figures in front of the goddess Taweret in Main Street House 3, and a line of human figures, possibly dancers, in Long Wall Street House 10 (Weatherhead 1995a; Kemp 1979). Nearly 70 pigment specimens were recovered from this area in the late 1970s and early 1980s, with samples being analysed using X-ray diffraction (XRD) and electron microprobe (Weatherhead and Buckley 1989). The blues were found to be Egyptian blue, and the turquoise pigments, and one green pigment, were Egyptian green. Two other green specimens were shown to be chrysocolla, a red proved to be haematite, and yellow ochre was also identified. Weatherhead’s later publication discusses the distribution of pigment finds across the workmen’s village (Weatherhead 1995a). Of the approximately 70 pieces of pigment from in and around the workmen’s village, 61% were green or turquoise, 23% were blue, 10% were yellow, and just four pieces of red were found (6%). The pigments were not all analysed but a visual examination suggested that the green and turquoise pigments were Egyptian green, the blues were Egyptian blue, and the reds and yellow were ochre, with the exception of specimen K18 (1), which appeared to be orpiment (Weatherhead 1995a, 388). A bright crystalline yellow pigment in a grinding bowl, found in a well near Q48.4, was also described as orpiment (Weatherhead 1995a, 394–95), as was a yellow pigment from Room 4 in House O49.17 (Weatherhead 1995a, 395). The residents of the Amarna workmen’s village built a series of painted chapels in their village. The painted plaster fragments from the main chapel have been excavated and reconstructions made of the designs (Weatherhead and Kemp 2007). The outer areas of the chapel appear to have been painted in white, but the Inner Hall and Sanctuary were more highly decorated. There appears to have been a white ‘skirting’ to a

2

height of 40–80cm, above which were painted coloured designs. In some cases these are stripes of white, blue and red, or the red and black palace façade design, and striped cavetto cornice fragments are also present in the rubble. In other cases fragments can be pieced together to reconstruct vultures, banquet scenes, and winged sun discs. The colours used are bright and include the usual white, black, yellow, red, blue, and green. The excavations at Amarna of Ranefer’s house (N49.18) and grid 12 in the Main City uncovered pigment lumps and ceramic sherds with paint adhering, similar to the palettes found at Amara West. Almost as many were found in a construction phase of Ranefer’s house as were found in the rest of the grid, indicating a greater investment in colour in the large house (Kemp and Stevens 2010b, 533). In grid 12, fourteen lumps of blue pigment, 6 lumps of green/turquoise, and 3 red were found. Among the sherds with pigment adhering, 9 had blue pigment, 6 had red, 12 had yellow, and 7 had green (Kemp and Stevens 2010b, 533–44). The Amarna Stone Village Survey, which focused on a small settlement southeast of the workmen’s village, likely to have housed a community of labourers and/or a provisioning place for desert-based workforces, uncovered seven pieces of pigment and two ceramic sherds with pigment adhering. Two of the pigments were red, resembling ochre or haematite, three were blue, visually identified as Egyptian blue, one yellow was thought to be ochre and the other yellow piece resembled orpiment (Stevens 2012, 339–41). The two ceramic sherds contained a yellow pigment described as matte and powdery, and a grindstone was found in trench 4 which had been used to grind a yellow pigment resembling the one identified as orpiment. Approximately 40 pigment fragments were recovered from industrial area Q48.4 in the Main City at Amarna. The vast majority were red and yellow, with very few blue or green (Weatherhead 1995a, 389). It is thought that area Q48.4 contained a workshop related to the production and decoration of pottery, and that these red and yellow pigments were used to create slip (Weatherhead 1995a, 389). Nicholson (2007) describes colour-related finds from site O45.1, a workshop area at Amarna related to the production of faience and glass objects and Egyptian blue and green pigment on the west side of the Royal Road, south of the Small Aten Temple. The finds included 28 specimens of Egyptian green and blue, 24 pieces of red ochre, 12 pieces of yellow

COLOUR IN ANCIENT EGYPT

ochre, and 1 piece of orpiment (Nicholson 2007; Anna Stevens, pers. comm., December 2013). Hatton examined the chemical composition and microstructures of samples of blue and green frit cakes, powder residues (adhering to ceramic), and frit vessel and bead fragments from O45.1, alongside other ancient samples from Thebes (Dynasty 18), Karnak (Dynasty 18) and Zawiyet Umm el-Rakham (Dynasty 19), and from 15th-century BC Mesopotamia (Nicholson 2007; Hatton, Shortland, and Tite 2008). A comparison of the microstructure of the ancient blue frit samples with modern replicas suggested that Egyptian blue cakes were the primary product, which were then ground to produce a pigment (Hatton, Shortland, and Tite 2008, 1596). The source of the quartz in the Egyptian samples was identified as sand, based on the rounded appearance of the grains of silica and the presence of alumina and iron oxides, typically present in quartz sands (1598, 1601). Sand from Amarna was analysed as part of this study, and found to contain about 17% calcium carbonate, which would have been sufficient to produce Egyptian blue and green without an addition as a separate ingredient (1598). The Na2O/ K2O ratios for the glass phase in all the Egyptian blue frits indicated that the flux used was plant ashes (1599). The presence of tin in the frits suggested that bronze was the source of the copper component (1598). Eight pigments excavated from Amarna by Petrie in 1891–92, and now in the Manchester Museum, were analysed using Raman spectroscopy and microscopy. They were identified as yellow and red ochre, realgar, Egyptian blue, azurite and malachite (David et al. 2001). The palace buildings at Amarna—the Great Palace, the King’s House, the North Palace, the North Riverside Palace, and the Meru-Aten—have significant remains of painted walls and pavements (Weatherhead 1992, 1994, 1995b, 2007). They were painted in a naturalistic way, with motifs taken from the natural world alongside some of the more traditional geometric designs. Analysis of painted plaster fragments from the Maru-Aten in Liverpool Museum and from the Great Palace in the Petrie Museum (UCL) identified red pigments as haematite, yellow pigments as goethite, blues as Egyptian blue, and a mixture of Egyptian blue and yellow ochre to create green pigment (Weatherhead 2007, appendix 2). The plaster was identified as lime, used both fresco and fresco-secco. The blue and green paint was applied secco because the large pigment grains from the Egyptian blue require a gum binder.

29

Pigments at Deir el-Medina The inhabitants of Deir el-Medina excavated, decorated, and furnished tombs for themselves in the cliffs and rock around their village, some of which they decorated in lavish detail with bright colours. Others were painted in a palette of yellow and black on a white background (Bruyère 1952). Their houses were also painted, as were objects found in the houses, including anthropoid busts and figured ostraca (Keith 2011; Cooney 2012). There were also various types of evidence for the practice of painting: brushes, palettes and pigments (Andreu and Barbotin 2002). Most paint materials are not described by the excavators, with the exception of a group of blue and turquoise pigments found in house S.E. V, which may have been ground and mixed with a binder before solidifying in dishes (Bruyère 1939, 221–22). Eleven pigment cakes (three green, eight blue) from Deir el-Medina, now in the Louvre, were examined by scanning electron microscopy with energy dispersive X-ray spectroscopy (SEM-EDS), Raman spectroscopy, optical microscopy, X-ray diffraction (XRD), transmission electron microscopy (TEM–EDS), UV-visible spectrometry and colorimetry to determine their structure and morphology (Pagès-Camagna and Colinart 2003). In agreement with Hatton et al. (2008), they concluded that the source of the silica was sand, the flux was most likely plant ashes, and the presence of tin suggested the use of bronze scraps as a source of the copper (PagèsCamagna and Colinart 2003). Pigments at Pi-Ramesse Pi-Ramesse (modern Qantir) was the Ramesside capital and royal residence, located in the Eastern Delta, and contemporaneous with Amara West (Prell 2011, 17). As mentioned above (see section on Blue, p. 24), there is evidence of Egyptian blue production, in combination with other high-temperature industries, from workshop areas in the city. Grindstones (Prell 2011, 77, 86) and ceramic palettes retaining pigment have also been found; the total number of objects designated in the Qantir finds database as ‘Farbpalette’ (‘colour palette’) is 82 (Henning Franzmeier, pers. comm., March 2017). The excavators estimate that 22 of these are complete, while the remainder are fragments. All but 6 are ceramic: 4 are shell and 2 schist stone. The pigment colours are blue, yellow, red and black (Prell 2011, 98–99; Henning Franzmeier, pers. comm., March 2017), and a few of the

30

CHAPTER

palettes had been used for more than one colour. The ceramic forms used for the palettes were from open and closed vessels (Prell 2011, 98). Lumps of raw pigment, 10 red, 14 yellow (both most likely ochres), and 4 blue (probably Egyptian blue), were also found in area QI (Prell 2011, 100–101). Pigments in royal palaces Several painted royal palaces in Egypt are known. The palace buildings at Amarna are discussed above. Two other examples of palaces with significant remains of painted decoration remaining are Tell el-Daba‘a and Malqata. Palatial buildings stood at Tell el-Dab‘a from the Hyksos Period to the Ramesside Period (1295–1069 BC). A mansion of the late Hyksos was painted with horizontal stripes in blue, red and white along the lower parts of the walls (Bietak 2013). The upper parts of the walls were painted yellow, and blue-painted fragments of wall plaster seem to have fallen from the ceiling. In a room described as a workshop (L1421), pre-dating the Dynasty 15 palace and destroyed by fire, stood two amphorae filled with Egyptian blue pigment (Bietak et al. 2009, 109). None of these have been analysed. The famous paintings from Palaces F and G at Tell el-Daba‘a were executed in Minoan colour conventions, although those in Palace G appear to have been Egyptian-style motifs whereas those in Palace F were Minoan in style (Bietak and Forstner-Müller 2003, 45; Bietak 2005; Morgan 2006). The paintings were most likely completed in the reigns of Hatshepsut and Thutmose III (Bietak 2005; Bietak et al. 2012/13). An analysis of the plaster at Tell el-Dab‘a from the Thutmoside Palace district found all the plaster to be lime with a very low gypsum content from natural impurities (Winkel 2007). Specks of blue pigment in the plaster were identified as Egyptian blue using polarized light microscopy (PLM) (Winkels 2007, 286). A detailed analysis of the Hunt Frieze from Palace F identified lime plaster, painted with yellow and red ochres, Egyptian blue, and carbon black (Morgan 2018). The painting was done in secco and used organic binders (Seeber 2000). Colours were mixed to achieve different shades, for example red ochre with calcite to make pink, and red and yellow ochre to make orange (Morgan 2018, 240). Green was made by mixing together Egyptian blue and yellow ochre (Morgan 2018, 242). It has been suggested that the painting was done by Minoan artists using local materials (Morgan 2018).

2

Amenhotep III’s palace complex at Malqata has many fragments of painted plaster preserved from the walls, floors, and ceilings. A microscopic analysis of the paintings reveals that the paint on the floors and daises was applied in a fresco-secco technique onto lime plaster (Lacovara and Winkels 2018). The paintings on the walls and ceilings were applied in a secco technique directly onto a brown clay plaster surface (Lacovara and Winkels 2018). Three uses of blue/green were documented—Egyptian blue, Egyptian blue mixed with Egyptian green, and a green earth pigment containing glauconite, which was sometimes mixed with Egyptian blue and green (Lacovara and Winkels 2018). Also identified were different shades of yellow ochre, red iron oxide, manganese and charcoal black, calcium carbonate used as a fine-grained white pigment, and a bright yellow pigment that is probably orpiment (Uda et al. 2000; Lacovara and Winkels 2018). Paints and pigments in ancient Nubia Most of the evidence relating to colour in Nubia is from sites dating to later periods than Amara West, although evidence from the New Kingdom is growing, and paintings from the Classic Kerma Period are preserved at Kerma. Scientific analyses have been conducted at a number of sites. All sites mentioned here, from north to south, are shown on the map in Fig. 1. Reiderer (1974) identified huntite as the white pigment on two late Nubian C-Group (c. 1600 BC) bowls and several ceramic sherds from Koshtamna, just south of Aswan. The site of Qasr Ibrim in Lower Nubia, now in Egypt, preserved fragments of wall paintings from the Meroitic temple (1st century AD) that depict divine and human figures in black, white, yellow, red, blue, and green (Pyke 2007). Painted stone blocks are also known from Qasr Ibrim and may have originated in the temple (Rose 2007). No materials analysis has been undertaken. The Archaeological Survey of Sudanese Nubia (ASSN) was conducted between 1963 and 1969, before the building of the Aswan Dam, over c. 130km of the Nile valley between Gemai, just south of the Second Cataract, and Dal (Edwards and Mills 2020). A few instances of paints and pigments were recorded, none were analysed. A four room house (or possibly working area) at Semna, probably dating to Dynasty 18, contained several small pieces of red and yellow ochre, small grinding stones with red and black pigment,

COLOUR IN ANCIENT EGYPT

possibly for use as make-up, and a grindstone with blue pigment (197–211). A working area in the Saras region, around Askut, had fragments of red ochre, and pieces of what appeared to be bitumen (145). A Dynasty 18 grave from the same region contained painted gypsum fragments, of unclear function (67). A New Kingdom grave in the Ukma region, 50km south of Semna, contained a polychrome wooden mummiform coffin (314–15). Sai Island, just a few kilometres upstream of Amara West, was chosen as the location for a Pharaonic settlement during the conquest of Upper Nubia in early Dynasty 18. Excavations in the early to mid-Dynasty 18 buildings in the town yielded palettes bearing colour (yellow, red, white and blue), alongside grindstones and hammerstones. Many came from the northeast sector (SAV 1N) in a space resembling a workshop with small-scale mud-brick buildings (Budka 2020, 420; Fulcher and Budka 2020). The important Pharaonic site of Tombos, at the Third Cataract, features cemeteries in use from midDynasty 18 to the Napatan Period (S. T. Smith 2003). Excavations in the cemeteries recovered painted coffins and sarcophagi, and tomb chapels (S. T. Smith 2007). The chapels were originally highly decorated, although now only fragments of painted plaster remain (Smith and Buzon 2014, 199). Kerma, centre of a Nubian polity that would threaten parts of Egypt during the Second Intermediate Period, was founded as early as 3000 BC and remained a centre of religious and political power through the Egyptian occupation and into Meroitic times. The religious centre is based around the enormous Western Deffufa, whereas a large circular building 50m to the southeast is thought to be the kingly seat. Palettes containing red and yellow pigment, assumed to be ochre, were excavated from tombs and funerary chapels in the eastern necropolis (Minor 2012; Charles Bonnet, pers. comm., May 2017). Chapels KXI and KII, and the burial chamber of Tumulus KIII, from the Classic Kerma Period, c. 1700–1500 BC, were painted in red, black and yellow (Reisner 1923, 136, 263–64; Lacovara 1986; Bonnet and Valbelle 2000; Minor 2012). The painted chapels depicted ships on the east walls and processions of animals on the west (Lacovara 1986). Kawa (ancient Gematon) is located between the Third and Fourth Cataracts, near the modern town of Dongola, and was occupied from the mid-14th century BC to the 3rd century AD (Welsby 2000). A mud-brick three-room shrine at the south end of the town, dating

31

to the reign of the Napatan king Taharqo, Dynasty 25, c. 650 BC, preserved Egyptian-style polychrome decoration (Welsby 2000; 2013; 2014). Store building F1 contained some broken pieces of mud brick bearing painted plaster, including one with hieroglyphic signs (Welsby 2014, 24). The pigments from the shrine at Kawa have been analysed by the author and consist of red and yellow ochres, Egyptian blue, and carbon black on a gypsum plaster background. One sample of red paint was identified as madder (Fulcher 2017). The tombs in the Kushite royal cemetery at el-Kurru, within the Gebel Barkal area, were decorated in red, yellow, blue, white, black and grey (Therkildsen 2015). Egyptian blue was confirmed by the use of visibleinduced luminescence imaging, and red and yellow oxides were strongly suggested by XRF identifying the presence of iron; XRF also identified calcium and sulphur in the white background (Therkildsen 2015), which indicates the use of gypsum, although for portable XRF analysis in air (rather than in a vacuum), sulphur is at the limits of detection of the instrument. The mountain of Gebel Barkal forms the focus for a series of temples, palaces, tombs, and town sites, which date from the New Kingdom and the Napatan, Meroitic and post-Meroitic Periods (Kendall and Mohammed 2016). Many of the monuments were painted in bright colours, and the pigments are being studied by Ca’ Foscari University of Venice and the Chemistry Department at Khartoum University (Ciampini 2018). Fragmentary wall paintings were found on columns in the 1st-century AD Amun temple at the Meroitic site of Dangeil (Anderson and Ahmed 2011). The pigments have been identified using Raman spectroscopy as yellow and red ochres, and Egyptian blue, on lime plaster (Sweek et al. 2014). A large number of painted plaster fragments depicting deities were excavated from three small temples at Meroe, from a precinct leading up to the large Amun temple, probably all dating to the 1st century AD (Shinnie and Bradley 1980, 62–65; Bradley 2003). The fragments are now all in the Sudan National Museum in Khartoum, the most complete being on display (Bradley 2003). Red, yellow, black, and blue pigments are preserved, but no materials analysis has been conducted. Building M.292 at Meroe, known as the ‘Augustus Temple’, was also originally decorated, although all of these paintings are now destroyed, and the only records of them are negatives held at Liverpool University and watercolours in the Museum of Fine Arts, Boston (Shinnie and Bradley 1981). They

32

CHAPTER

depict enthroned figures with bound captives on the footrests. The proportion of the figures and the iconography suggest a date of 1st century BC to 1st century AD (Shinnie and Bradley 1981). The Meroitic site of Naqa includes temples, a necropolis, and a town area. Remnants of polychrome wall paintings were found in the Amun temple, dating to AD 1–50, depicting Nile gods and the royal family (Hesse 2006). Analysis of the pigments using XRF identified lead as one of the constituent elements, but ancient fire damage to the paintings meant that the original colour of the lead pigments could not be determined (Hesse 2006). Analyses of mortars have been conducted from several sites in Nubia, over a range of time periods. Samples of mortar were examined from Dynasty 18 and Ramesside phases of the Amun temple at Dokki Gel (adjacent to Kerma), and various Nubian monuments of the 1st century AD from Dokki Gel, Gebel Barkal, Kawa, Dangeil, el-Hassa, Damboya and Naqa (Feneuille, Letourneu, and Bouchar 2014). They concluded that plasters used during the Egyptian New Kingdom were composed of gypsum imported from Egypt, and those of the later period were lime, possibly from marble in the region of the Third Cataract (Feneuille, Letourneu, and Bouchar 2014). However, the argument for the gypsum source being Egypt is based solely on the lack of knowledge of sources in Sudan; should a more local source be identified, this conclusion could be revisited. Moreover, the sources of Egyptian gypsum that are known to have been exploited in ancient times contain high levels of calcite, which are not reported here, although it is possible to clean the calcite out of the gypsum by flotation or simple chemical means (Harrell 2017). During both periods local Nile alluvial clays were also used as mortars (Feneuille, Letourneu, and Bouchar 2014). Mortars from the Dynasty 18 town and cemetery on Sai Island have also been analysed and were identified as softburned (rather than calcined) lime, with gypsum inclusions (Budka 2020, 269–73). Binders in ancient Egypt Paints are made up of the colourant, or pigment, and a substance that holds the pigment in place after the paint dries, called a binder. Binders for paint can be inorganic, for example lime or gypsum dissolved in water, or they can be organic. Organic binding media are composed of a film-forming material, such as gum,

2

glue or oil, and often a solvent or diluent, such as water or an organic solvent. Lucas (1962) hypothesizes that the organic binders most probably used by the ancient Egyptians were animal glue, beeswax, egg white, and plant gum or resin—all materials that would have been locally available. Proteinaceous glues and carbohydrates (e.g. sugars such as plant gums) have been identified as binders in ancient Egyptian paints using gas chromatography-mass spectrometry (GC-MS). Both of these materials are water soluble, which makes them especially useful as paint binders. Plant gum has been identified in paint on tomb and temple walls, coffins, and mummy masks from Dynasty 18 onwards (Stulik, Porta, and Palet 1993; Le Fur 1994; Newman and Serpico 2000; McCarthy 2001; Daniels, Stacey, and Middleton 2004; Stacey 2008). The binder used at Karnak temple was identified as Acacia nilotica for samples from all periods, Dynasty 18 to Late Antiquity (Le Fur 1994, 62–66), which seems remarkably consistent and might perhaps be an over-confident identification of a specific plant gum; the constituents used to identify plant gums can be fairly variable within a species (Anderson and Karamalla 1966). Wall paintings from the tomb of Nefertari were found to contain an Acacia gum (Stulik, Porta, and Palet 1993). Gum tragacanth was identified in a white material from a Dynasty 21 coffin (MasscheleinKleiner, Heylen, and Tricot-Marckx 1968), and possibly in samples of the Third Intermediate Period from the Museum of Fine Arts, Boston in a mixture with other gums (Newman and Halpine 2001). Animal glues have been identified in paint from Old and New Kingdom tomb walls, wooden figures from the Middle Kingdom, and Graeco-Roman Period and Late Period to Graeco-Roman cartonnage (Jaksch 1985; Newman and Serpico 2000; Scott et al. 2003; Stacey 2008). In the case of the tomb of Nebamun, animal glue was used as the binder only for the blue and green pigments; the other pigments used gum as the main binder (Stacey 2008). This could be explained by the relatively large grain size of the synthesized copper pigments, which have a tendency to fall off the substrate, and thus require a stronger binder (Stacey 2008, 58). Egg white was found as a varnish in the tomb of Nefertari, but the authors question whether this could be a modern addition (Stulik, Porta, and Palet 1993). Egg was also found on cartonnage fragments of the Late, Ptolemaic and Roman Periods and on Fayum mummy portraits from the 4th century AD (Ramer

COLOUR IN ANCIENT EGYPT

1979; Scott et al. 2009). Egg with casein and egg with Acacia gum were tentatively identified as binders in the Minoan paintings from Tell el-Dab‘a (Seeber 2000). Beeswax has not been confidently identified as a binder, but was used in painting schemes to enhance small areas (Stacey 2008), and also seems to have been used as a varnish (Daniels, Stacey, and Middleton 2004; Bonizzoni et al. 2011), although again it is impossible to determine if the varnish was applied in antiquity or in modern times. Petrie, for example, is known to have coated his more fragile findings in wax (Petrie 1904, 90). Plant resin, often identified as Pistacia spp., was used as a varnish (Newman and Serpico 2000; Serpico and White 2001; Daniels, Stacey, and Middleton 2004; de Vartavan 2007; Bonizzoni et al. 2011), but has not been confirmed as a binder. Fresh Pistacia resin is viscous and quickly solidifies, requiring a solvent or direct heat to reliquify it, although other resins may be stored as liquids. Resins are not water soluble, but most are soluble in alcohol, which was available. It is also possible that resins were distilled to make essential oils, in which resin could be dissolved, although direct evidence for this is lacking. With an appropriate solvent resins could have acted as paint binders, but they do not appear to have been used in this way. In addition to these more common substances, there are a range of unidentified sugary substances used as binders, which might possibly include honey (Masschelein-Kleiner, Heylen, and Tricot-Marckx 1968; Stacey 2008). A Dynasty 6 box of paints at the

33

British Museum was found to use a drying oil as a binder, a substance not previously found in this context (Ambers, Stacey, and Taylor 2007). Animal fat was found to be one of the constituents of a black fill applied to hieroglyphs on a Dynasty 18 sarcophagus (Newman 1993). Ancient Egyptian painting tools Paintbrushes are known from ancient Egyptian sites, though many are poorly provenanced. The paintbrushes now in museum collections are stated as being of ‘palm fibre’ (British Museum (BM) EA 36889; Metropolitan Museum of Art (Met) 17.190.1967), ‘reed(?)’ (BM EA 36895), ‘vegetal material (not identified)’ (BM EA 41186); ‘cordage’ (Penn Museum 2003-34-378), and ‘halfa grass’ (Met 28.3.1 and other brushes from the tomb of Meketre). It appears that in most cases the materials used to manufacture the paint brushes have not been investigated scientifically. A few other plant fibre objects, such as baskets and shoes, have been analysed, with halfa grass and palm leaf identified as the most important materials for use in basketry (Wendrich 2000), while shoes were made with various plant materials including halfa grass, reed, palm leaf and the fruit-bearing stalks of the date palm (Veldmeijer 2010, 145–47). Crude paintbrushes excavated at Shalfak, a Middle Kingdom fortress at the Second Cataract, were simply fashioned from sticks with the ends pounded to create bristles (Claudia Näser, pers. comm., May 2017).

CHAPTER 3

AMARA WEST PAINTS: ANALYSIS

A substantial number of samples of pigment from Amara West were subjected to materials analysis, both in the field and in British Museum laboratories. This provides confirmation of the inorganic and organic substances from which the pigments were made and provides a foundation for understanding the sources and processing methods. Materials analysed from Amara West In this analysis I refer to paints, pigments, and binders. Pigments are colourants, which in this study are usually produced by grinding rocks or from manufactured pigment cakes. Binders are the organic substances added to the pigments to hold the pigment in place; when these are mixed with pigments, paint is created for application to objects or buildings. I consider the material on the palettes to be paint, made up of pigment(s) and binder, whereas coloured material on grindstones is pigment. ‘Plaster’ is used to refer to white gypsum, or pigment mixed with gypsum, applied in large areas across a wall, or applied as a ground on a wooden coffin before painting. Mud plaster is also referred to; this is a mud-based slurry that is applied to walls and other architectural elements (such as mastabas) to give a smooth finish over mud brick.

also Figs 20, 26–28); as stated above, individual F-numbers might comprise a group of palettes found in the same context. The majority of samples (97) come from the large assemblage of palettes in E13.14 and E13.31. Samples from 21 palettes were taken from the Western Suburb. Blue was sampled in every case where it was present in sufficient quantities. All green examples were sampled. Palettes holding red and yellow paint were numerous; those retaining the most paint were taken to London for further analysis. Polychrome objects were sampled several times, for each colour present. In total, 130 palettes were analysed: 16 white, 11 black, 37 yellow, 42 red and pink, 20 blue, and 4 green. Pigments Lumps of red and yellow raw pigment were numerous and very similar in appearance. Only a selection was exported for analyses; a further 11 red and 13 yellow pigments were analysed using portable XRF (pXRF) at the project house (Fig. 13). All excavated lumps of blue pigment were taken to London for analysis. No lumps of white or black pigment were found (although charcoal is abundant at Amara West). In total 17 yellow pigments were analysed, along with 15 red, 7 blue, and 1 green.

Walls, floors, and ceilings Samples of plaster from walls which retained colour were infrequent at the site, and come from only a few houses (Fig. 11; see also Figs 7–10). All white and coloured wall plasters and paints on walls were sampled. In cases where there were several layers of white plaster, the top layer was sampled, excepting the painted niche in E13.7.6, which had been stripped back by conservators so that earlier layers were also sampled. In total, 77 samples were taken from walls, floors, and ceilings: 26 white, 7 black, 18 yellow, 21 red, and 5 blue. Palettes Over 900 ceramic palettes, or fragments thereof, have been excavated from Amara West (Fig. 12; see

Grindstones Samples were taken from all grindstones that retained enough pigment (Fig. 6). Most grindstones and hammerstones were not sampled because the pigment survived only in trace amounts, or was not visible at all. Five samples were taken from grindstones: 1 yellow, 3 red, and 1 green. Coffins Painted plaster was found as part of coffin fragments in tombs G201, G222, G238, G244 (Fig. 14), G301, G309, G320, and G322. Forty-two samples were taken from coffins from five tombs, selected on the basis of whether they had sufficient paint remaining to sample,

36

CHAPTER

3

Figure 11: Painted wall plaster F5133c from the painted niche in E13.7.6 [4561].

AMARA WEST PAINTS: ANALYSIS

Figure 12: Fragments of palettes F15020, from D12.8.7 [12840], a windblown deposit of sand.

Figure 13: Yellow and red pigments F2471, from E13.20.1 [10301].

37

CHAPTER

38

3

Figure 14: Painted coffin plaster F9707 from tomb G244 [9505].

and the extent to which each fragment had been consolidated, as conservation materials can obfuscate analytical results. A range of colours were examined, along with the plaster ground to which the paint had been applied: 11 white coffin plasters, 1 white paint, 11 black, 7 yellow, 9 red, and 3 blue.

fragments to determine if bitumen was present. For binding media analysis, 17 samples were analysed by GC-MS to determine if a polysaccharide-based binding medium was present: 15 samples of paint from palettes (3 black, 3 white, 2 blue, 4 yellow, and 3 red) and 2 white wall plasters. Details of the instruments and methods used are given in the Appendix.

Pigment and binder analysis: methodology A range of analytical methods was employed to identify the inorganic components of the paints. In the first instance, a visual inspection provided an indication of the type of pigment that was expected; this was mainly influenced by colour and crystallinity. Portable XRF (pXRF) was used on site to analyse palettes that had been excavated by February 2014, and the presence and extent of Egyptian blue at Amara West was checked on site (West Gate) and in the excavation house (coffin fragments and architectural elements) using visible-induced luminescence (VIL) photography. Samples that were exported to London were analysed initially by polarized light microscopy (PLM), and then some examples by infrared spectroscopy. In the cases where the results were inconclusive, or further investigation was deemed necessary, elemental analysis was conducted using a scanning electron microscope coupled with an energy-dispersive X-ray spectrometer (SEM-EDS). One green pigment and one blue pigment were analysed using X-ray diffraction (XRD). In the case of the black pigments, gas chromatography-mass spectrometry (GC-MS) was undertaken on black paint from 10 palettes and 4 painted coffin

Portable X-ray fluorescence spectrometry (pXRF) XRF is an elemental analysis technique. The elemental composition of a sample can be determined by measuring the fluorescent (secondary) X-rays emitted from a sample when it is excited by a primary X-ray source. Each of the elements present in the sample produces a set of characteristic fluorescent X-rays that is unique for that specific element (Shackley 2011). The benefits of XRF are that it is quick and gives an immediate result. The limitations of the XRF technique are that it has fairly high limits of detection, and that the beam penetrates the analysed surface, and so may penetrate very thin layers of pigment to include the substrate in the analysis. If the analysis is conducted in air then X-rays from the lighter elements will be absorbed by the air; thus nothing lighter than potassium can be measured, and the analysis cannot be quantified (Tite 2000). Furthermore, the portable XRF instrument used on site has quite a wide beam, which can be problematic when trying to focus on a small area of pigment. It is also fairly heavy and therefore not suitable to hold over very fragile objects, such as coffin fragments.

AMARA WEST PAINTS: ANALYSIS

Polarized light microscopy (PLM) PLM is a widely used technique for the identification of historical pigments (Kakoulli 2002; Eastaugh et al. 2004b). A small sample of pigment was dispersed on a microscope slide in MeltmountTM, refractive index of 1.662, and covered with a cover slip. Each sample was then observed in plane polarized light and between crossed polars. More detail on the method is given by Easthaugh et al. (2004b). The technique allows the full assemblage of grains in the sample to be observed if a representative sample has been taken. The percentage of each mineral present in the sample was estimated by eye by counting grains within a quarter segment of the visible dispersion under ×400 magnification. Infrared spectroscopy Fourier-transform infrared spectroscopy (FTIR) has frequently been used for the analysis of ancient pigments (Edreira, Feliu, and Mart 2001; McCarthy 2001; Genestar and Pons 2005; Aliatis et al. 2009; Kakoulli 2009; Vahur et al. 2016). FTIR measures the energy absorbed by the sample from a beam of infrared radiation. Covalent bonds within molecules absorb the infrared energy and vibrate, giving each molecular structure a unique absorption spectrum (Stuart 2004). By comparing the spectra of an unknown substance with a spectra library, inorganic and organic substances can be identified. Samples were analysed using attenuated total reflection Fourier-transform infrared spectroscopy (ATRFTIR). The ATR method produces reflection spectra using an internally reflected infrared beam. An infrared beam is directed onto a diamond window (in this case) at a certain angle. The diamond has a high refractive index, which causes internal reflectance of the beam in the window. This internal reflectance creates an evanescent wave, which extends into the sample held in contact with the window. The sample absorbs energy in certain regions of the infrared spectrum, which reduces the amplitude of (attenuates) the evanescent wave. The attenuated beam is directed to the detector in the spectrometer, which generates an IR spectrum. The ATR method is useful because it requires very little sample preparation and is quick to run. Peaks can be in slightly different positions to those from spectra run by transmission due to differences in the methods, which should be considered when comparing spectra from libraries (Derrick, Landry, and Stulik 1999, 60; Koulis, Reffner, and Bibby 2001).

39

Scanning electron microscopy with energy-dispersive X-ray spectrometry (SEM-EDS) The scanning electron microscope (SEM) can be used to take high magnification and resolution images, and to determine elemental composition when coupled with an energy-dispersive X-ray spectrometer (EDS). The SEM can form two types of image, using either secondary electrons or backscattered electrons (BSE). Secondary electrons are emitted when the incident electron beam causes electrons to be ejected from the K-shells of the specimen atoms. The number of secondary electrons reaching the detector from each point on the specimen determines the brightness of the image; the shape of the surface affects the number of secondary electrons that ‘escape’ and thus can reach the detector, so secondary electrons can be used to build images showing topography and morphology. BSE are those electrons that originated in the incident beam and have been deflected by the sample; larger atoms have a greater chance of creating an elastic collision due to their greater cross-sectional area. The number of BSE reaching the detector is proportional to the atomic number of the atom, therefore an image can be built showing phases with higher (brighter) and lower (darker) average atomic number (Frahm 2014). EDS is the detection and measurement of X-rays emitted as a result of electron shell transitions in the sample caused by the incident electron beam. It allows elemental analysis to be undertaken on the sample across areas or in point locations (Goldstein 2003; Egerton 2005, 125–76). Unless the sample is polished flat, the EDS results are not fully quantitative; some analyses were conducted on powders or directly onto palettes and some samples were prepared in polished blocks. SEM was used to take magnified pictures of individual grains of pigments, and for EDS analysis. X-ray diffraction (XRD) XRD provides phase and structural information of crystalline materials. The crystalline structure of the material to be analysed causes a beam of incident X-rays to diffract and the following interference results in a pattern of higher and lower intensities, which are picked up by a detector moving around the sample. The output is an X-ray diffraction pattern or diffractogram, which is a plot of the intensity of the X-rays scattered at different angles. Each phase (combination of

40

CHAPTER

chemistry and crystal structure) produces a unique diffraction pattern (Schreiner et al. 2004; Garrison 2014). Visible-induced luminescence imaging (VIL) When excited by visible light, Egyptian blue luminesces in the infrared spectrum in the 800–1000nm range (Accorsi et al. 2009). Using a camera fitted with a filter that admits only infrared light, this luminescence can be directly captured, which is a quick and portable method to identify Egyptian blue. Because the luminescence is so strong, even minute amounts of Egyptian blue can be detected in this way, which can indicate areas that were originally painted blue but are no longer visibly blue to the naked eye (Verri et al. 2010). Gas chromatography-mass spectrometry (GC-MS) Gas chromatography-mass spectrometry (GC-MS) analysis was undertaken on paint samples from Amara West to seek and identify polysaccharides (plant gums, honey). Black samples were analysed for bitumen also using GC-MS. In each case, reference material from the British Museum Reference Collection was analysed alongside the samples. GC-MS is an analytical technique that allows the molecules in a substance to be separated and identified (Evershed 2000). The material to be analysed is vaporized, and then swept through the column by an inert gas, referred to as the mobile phase. The column is lined with a liquid stationary phase; the molecules in the sample are separated as they pass along the column due to differences in partitioning behaviour between the two phases. The column is connected to a mass spectrometer that ionizes the incoming molecules and measures their mass to charge ratio. The method followed for the analysis of Amara West paints for plant gums was the standard operating procedure used at the British Museum for the preparation of polysaccharide samples for GC-MS analysis of neutral sugars and uronic acids, based on a published method (Bleton et al. 1996). GC-MS analysis for polysaccharides was undertaken on 17 samples, which were selected for their range in colour and volume of sample available. All of the samples were taken from paint on ceramic palettes or pieces of wall plaster because these provided the largest volumes of material. Large volumes were needed because the likelihood of organic material surviving was thought to be low. Large paint samples provide a potentially larger amount of organic

3

residue to be analysed. Reference samples of Acacia (gum arabic), Prunus (plum), and gum tragacanth were taken from the British Museum Reference Collection. For bitumen analysis, samples were solvent extracted and then hexane was added to precipitate the asphaltene fraction and obtain the maltene fraction. Each maltene fraction was then fractionated using column chromatography. The elutes were collected and the first fraction was analysed by GC-MS (see Fulcher, Stacey, and Spencer 2020). Pigment analysis: results Pigments were identified from paints using a range of techniques. The methods used complemented one another to provide secure identifications. In total the mineral components of 296 paints or pigments from Amara West were identified: 130 paints from palettes; 42 paints from coffins; 77 paints and plasters from walls, floors, and ceilings; 40 raw pigments; 5 pigments from grindstones; 1 pigment from a small clay bowl; 1 pigment from a stone sherd. Details are given in the Appendix. White White minerals were widely used at Amara West. They were used alone for plastering walls, and also combined with other pigments. The white pigments were analysed using FTIR, PLM, and SEM-EDS (Table 5). All solely white samples on palettes were found to be gypsum or anhydrite. Gypsum (calcium sulphate dihydrate CaSO4·2H2O) is clear and colourless under plane polarized light, with moderate relief and frequently has inclusions; under crossed polars it exhibits first order whites and greys, and inclined extinction (Eastaugh et al. 2004b, 295). Anhydrite (calcium sulphate CaSO4) is the dehydrated form of gypsum, which can occur naturally, or be produced by heating gypsum (Lacovara and Winkels 2018). Under plane polarized light it has clear crystals with no inclusions, with moderate and variable relief (Eastaugh et al. 2004b, 291); often the crystals are rhombic or rectangular due to cleavage at high angles to each other. Under crossed polars, this pigment shows bright, second order interference colours. Both calcite (calcium carbonate CaCO3) and gypsum were found mixed with red, yellow, and blue pigments on palettes (see following sections). In plane polarized

AMARA WEST PAINTS: ANALYSIS

light calcite (calcium carbonate) is clear and colourless with variable relief, and in crossed polars it exhibits fourth order or above birefringence (Eastaugh et al. 2004b, 279). The rhombic shape of the calcite particles observed in PLM (Fig. 15) and the lack of fossils within it indicate that the source may have been a local vein calcite, and not chalk, which is micritic (very fine) and usually contains coccoliths (Kerr 1959, 212; Gettens, Fitzhugh, and Feller 1974; Deer, Howie, and Zussman 1992, 630; Eastaugh et al. 2004b, 279). Calcite frequently occurs in association with ochres (Eastaugh et al. 2004b, 365), which may explain its presence to some extent, but the absence of calcite from the raw pigments examined from Amara West, and its inclusion in blue paints made from manufactured Egyptian blue (which does not itself contain calcite) suggest that it was deliberately added as a white pigment to other pigments in certain circumstances. In addition, there are examples of palettes holding two shades of paint, one a darker shade, and the other a lighter paint made up of the coloured pigment mixed with calcite (e.g. F7130; see Fig. 19). Gypsum was used for wall plaster, with occasional very minor quantities of calcite inclusions of 1% or less. The gypsum was usually ground quite fine, down to 20μm and below. Calcite can occur as a mineral association with gypsum, which probably explains the inclusions in the wall plaster, rather than an intentional addition. The samples of white pigment from coffins were in most cases from the underlying plaster, used to create

41

Figure 15: Dispersion of pale yellow pigment from ceramic palette F6079 (PS128), from E13.14.8 [5222], mostly calcite; rhombic crystals are visible. Crossed polars ×1000.

a ground on which to paint colours. However, the plaster was also intended to show through in some designs, therefore acting as a white decorative element. Of the 11 plaster samples taken from coffin fragments, 8 were solely composed of gypsum, and 3 were a gypsum:calcite mix in the ratios 50:50, 60:40 and 80:20. Calcite was found in the coloured paints used to decorate the coffins and in some cases appears to have been deliberately added to the paint: for example, PS502 had a 100% gypsum ground, but 20% calcite in the red paint; PS307 had 20% calcite in the ground but 40% calcite in the blue paint. Calcite was also found in other coffin paints but could have been a contaminant from the ground.

Figure 16: FTIR spectrum for white coffin paint PS307 from coffin fragment F8110, from tomb G309 [8163]. Peaks indicate huntite (1505, 1432, 890, 885), and gypsum from the coffin plaster (1111, 671, 599).

CHAPTER

42

3

Table 5: Results of pigment analysis of white paints, pigments, and plaster. Wall plaster samples do not have Find (F) numbers because they are samples taken directly from walls on site and are not finds. Results given here are for pigments / paints / plasters that appeared white to the eye. The majority of calcite identifications were made within mixtures with colours and so are not presented in this table. Find number

Sample number

Type of object

Building

Context

Phase

Analysis

Mineral identification

F4881

PS285

Wall plaster

E13.7.6

4561

2B–3A

PLM

Gypsum

F4881

PS286

Wall plaster

E13.7.6

4561

2B–3A

PLM

Gypsum

F5003g

PS264

Wall plaster

E13.7.6

4561

2B–3A

PLM, FTIR

Gypsum

F5003i

PS266

Wall plaster

E13.7.6

4561

2B–3A

PLM, FTIR

Gypsum

F6147

PS131

Palette

E13.31.1

5325

3A

PLM, FTIR

Gypsum, calcite

F6147

PS133

Palette

E13.31.1

5325

3A

PLM, FTIR

Anhydrite

F6169

PS436

Palette

E13.31.2

5346

2B

PLM

Gypsum

F6170

PS419

Palette

E13.6.4

5341

3B

PLM, FTIR

Gypsum

F6190

PS122

Palette

E13.31.2

5332

2B

PLM, FTIR

Gypsum, calcite

F6223

PS140

Palette

E13.7.12

5224

3A

PLM, FTIR

Gypsum

F6264

PS417

Palette

E13.7.12

5261

2A

PLM, FTIR

Gypsum

F6423

PS451

Palette

E13.14.10

5339

2B

PLM

Gypsum

F6446

PS445

Palette

E13.14.10

5339

2B

PLM

Gypsum

F6463

PS155

Palette

E13.14.10

5346

2B

PLM, FTIR

Gypsum

F6463

PS248

Palette

E13.14.10

5346

2B

PLM, FTIR

Gypsum, calcite

F6467

PS249

Palette

E13.31.2

5334

2B

PLM, FTIR

Gypsum

F6493

PS440

Palette

E13.14.1

5361

1B

PLM, FTIR

Gypsum

F6499

PS246

Wall plaster

E13.14.10

5339

2B

PLM, FTIR

Gypsum

F7794

PS473

Palette

E13.17

10092

2A–2B

PLM

Calcite

F8110

PS307

Coffin plaster

G309

8163

3/4/5

PLM, FTIR

Gypsum, calcite

F8110

PS307

Coffin paint

G309

8163

3/4/5

PLM, FTIR

Huntite

F8110

PS527

Coffin plaster

G309

8163

3/4/5

PLM, FTIR

Gypsum, calcite

F9079

PS531

Coffin plaster

G201

9018

3/4/5

PLM, FTIR

Gypsum

F9485

PS529

Coffin plaster

G222

9485

3/4/5

PLM, FTIR

Gypsum

F9535

PS528

Coffin plaster

G238

9195

3/4/5

PLM, FTIR

Gypsum

F9707

PS503

Coffin plaster

G244

9505

3/4/5

PLM

Gypsum, calcite

F9741

PS296

Coffin plaster

G244

9511

3/4/5

FTIR

Gypsum

F9741

PS298

Coffin plaster

G244

9511

3/4/5

FTIR

Gypsum

F9742

PS301

Coffin plaster

G244

9519

3/4/5

FTIR

Gypsum

F9761

PS502

Coffin plaster

G244.1

9244

3/4/5

PLM

Gypsum

F9761

PS504

Coffin plaster

G244.1

9244

3/4/5

PLM, FTIR

Gypsum

F14043

PS567

Stone lintel

D12.8.3

12831

3/4/5

PLM

Gypsum

F15014

PS532

Palette

D12.12.2

12301

3/4/5

PLM, FTIR

Anhydrite

F15020

PS534

Palette

D12.8.7

12840

3/4/5

PLM

Gypsum, quartz, calcite, red ochre

PS255

Wall plaster

E13.7.5

4566

2

PLM

Gypsum

PLM, FTIR

PS256

Wall plaster

E13.7.5

4566

2

PS257

Wall plaster

E13.3.1

4133

2A–2B–3A–3B–4 PLM, FTIR

Gypsum Anhydrite

PS259

Wall plaster

E13.3.24

4068

2A–2B–3A–3B–4 PLM, FTIR

Gypsum

PS273

Wall plaster

E13.7.6

4727

2B

FTIR

Gypsum

PS279

Wall plaster

E13.7.6

4727

2B

FTIR

Gypsum

PS485

Wall plaster

E13.4.2

4471

4

PLM

Gypsum

PS491

Wall plaster

E13.7.3

4728

2A

PLM

Gypsum

PS492

Wall plaster

E13.3.11

4568

3A

PLM

Gypsum

PS493

Wall plaster

E13.7.6

4735

2B

PLM

Gypsum

AMARA WEST PAINTS: ANALYSIS

Find number

43

Sample number

Type of object

Building

Context

Phase

Analysis

Mineral identification

PS496

Wall plaster

E13.7.6

4727

2B

PLM

Gypsum

PS497

Wall plaster

E13.7.6

4727

2B

PLM

Gypsum

PS499

Wall plaster

D12.7.1

12013

3/4/5

PLM

Gypsum

PS500

Wall plaster

D12.7.1

12152

3/4/5

PLM

Gypsum

PS544

Wall plaster

D12.8.1

12887

3/4/5

PLM

Gypsum

PS551

Wall plaster

E13.7.6

4727

2B

PLM

Gypsum

PS553

Wall plaster

E13.7.6

4727

2B

PLM

Gypsum

PS554

Wall plaster

E13.7.6

4727

2B

PLM

Gypsum

PS558

Wall plaster

E13.7.6

4566

2B

PLM

Gypsum

PS559

Wall plaster

E13.7.6

4566

2B

PLM

Gypsum

PS590

Modern gir

Pit near Amara West

3/4/5

PLM, FTIR, SEM-EDS

Dolomite

Thus while gypsum is ubiquitous, calcite appears to have been reserved for specific purposes, since it is only found in significant quantities on coffins and palettes. This pattern of use at Amara West is consistent with results from studies of objects from Egypt. An analysis of many ancient Egyptian items in the Louvre found that gypsum and anhydrite were usually employed for coatings or underlayers, whereas white carbonate pigments were used as pigments in decorative applications (Pagès-Camagna and Guichard 2010, 26). Huntite (magnesium calcium carbonate Mg3Ca(CO3)4) was found on one coffin from Amara West (PS307, F8110), from tomb G309. It was applied as a white paint of a much brighter hue than the plaster. The plaster was found to be composed of gypsum, and the white paint was identified as huntite using PLM and FTIR (Fig. 16). Huntite is composed of very fine particles in aggregates that resemble pom-poms (Eastaugh et al. 2004b, 289). In crossed polars it exhibits moderate birefringence with high first order and second order interference colours. This was the only coffin fragment with remains of this bright white paint. A sample of white powder from a source close to Amara West that is used by the current inhabitants of the area was identified as dolomite (calcium magnesium carbonate CaMg(CO3)2). Dolomite is very similar in appearance to calcite and can be distinguished using FTIR and SEM-EDS but not PLM (Deer, Howie, and Zussman 1992, 641; Appendix Tables 1–3). Black All instances of black paint or pigment found at Amara West were sampled, if sufficient material was

present. A combination of PLM, FTIR and GC-MS was used to distinguish the black materials (Table 6). Bitumen and carbon are not easy to distinguish, both essentially being formed of carbon, rendering elemental analyses unhelpful. Under polarized light, bitumen is more likely to present tobacco-coloured flakes with conchoidal fracture (Eastaugh et al. 2004b, 240; Fig. 17), but occasionally charcoal has been seen to present similar particles, so that sometimes the two can be confused (Ruth Siddall, pers. comm., November 2016). If bitumen is present the FTIR spectrum will show peaks at 2924 and 2854, the asymmetric and symmetric stretching of C-H in the methylene and methyl groups, and bands from the methyl and methylene bending vibrations 1456, 1376 cm-1 (Tomasini, Siracusano, and Maier 2012). Analysis using GC-MS can provide conclusive evidence for the presence of bitumen if the biomarkers for hopanes and steranes can be identified (Peters, Walters, and Moldowan 2005; Fulcher, Stacey, and Spencer 2020). The majority of the black pigments were found to be carbon. Sometimes the remains of burnt plant matter, still retaining the cellular structure, can be observed by PLM, and this confirms an origin for the pigment as organic carbon (Fig. 18). Nine samples of black paint from palettes were found to contain bitumen. Only one (PS119) was found to be entirely composed of bitumen, the other eight were a mixture of bitumen and carbon, as could be seen from the presence of both isotropic black particles and tobacco flakes in PLM. Four of these palettes with mixtures of bitumen and carbon had only traces of bitumen (Table 6; Fig. 19) (Fulcher, Stacey, and Spencer 2020). Several coffin fragments from G244 and G309 were found to be painted using bitumen as a pigment, in most cases mixed with

44

CHAPTER

Figure 17: Dispersion of black pigment PS295 from coffin F9743, from G244 [9515] in plane polarized light ×400. The tobacco colour and conchoidal fracture of the particles strongly indicate bitumen. Calcite particles also present. Scale bar at bottom right represents 30μm.

3

Figure 18: Dispersion of black pigment (PS527) from coffin fragment F8110, from G309 [8163] in plane polarized light ×400. Remains of burnt plant matter. Scale bar at bottom right represents 30μm.

Figure 19: Above: Mass chromatogram for ion m/z 191 (terpanes and hopanes) for palette PS152 from E13.14.6 [5284], showing positions of terpanes (20/3 to 30/3), hopanes (29αβH to 34αβH; hopanes 31–34 are split into S and R), Ts, Tm, and gammacerane (GCR). Below: Mass chromatogram for ion m/z 217 for palette PS152, showing positions of steranes (see Fulcher, Stacey, and Spencer 2020).

AMARA WEST PAINTS: ANALYSIS

45

Table 6: Results of pigment analysis of black paints, pigments, and plaster. Wall plaster samples do not have Find (F) numbers because they are samples taken directly from walls on site and are not finds. Find number

Sample number

Type of object

Building

Context

Phase

Analysis

Mineral identification

F5003a

PS290

Wall plaster

E13.7.6

4561

2B–3A

PLM

Gypsum, yellow ochre, carbon

F5003h

PS326

Wall plaster

E13.7.6

4561

2B–3A

PLM, FTIR

Gypsum, carbon, yellow ochre

F5003h

PS456

Wall plaster

E13.7.6

4561

2B–3A

PLM

Gypsum, carbon, yellow ochre

F5094a

PS292

Wall plaster

E13.7.6

4561

2B–3A

PLM, FTIR

Carbon, gypsum

F6167

PS415

Palette

E13.31.1

5348

2A

PLM, FTIR, GC-MS

Bitumen / carbon

F6281

PS139

Palette

E13.7.11

5246

3A

PLM, FTIR, GC-MS

Carbon

F8110

PS307

Coffin

G309

8163

3/4/5

PLM, FTIR

Gypsum, calcite, bitumen / carbon, yellow ochre

F8110

PS527

Coffin

G309

8163

3/4/5

PLM, FTIR

Carbon, calcite, gypsum

F8110

PS899

Coffin

G309

8163

3/4/5

PLM, FTIR, GC-MS

Carbon

F8115

PS165

Coffin

G309

8163

3/4/5

PLM, FTIR

Carbon

F9079

PS531

Coffin

G201

9018

3/4/5

PLM, FTIR

Carbon, gypsum

F9485

PS529

Coffin

G222

9485

3/4/5

PLM, FTIR

Gypsum, bitumen / carbon

F9535

PS528

Coffin

G238

9195

3/4/5

PLM, FTIR

Carbon, gypsum

F9715

PS901

Coffin

G244.1

9510

3/4/5

PLM, FTIR, GC-MS

Carbon, trace bitumen

F9736

PS297

Coffin

G244

9511

3/4/5

PLM, FTIR, GC-MS

Trace bitumen, yellow ochre

F9742

PS301

Coffin

G244

9519

3/4/5

PLM, FTIR

Quartz, gypsum, bitumen / carbon

F9743

PS295

Coffin

G244

9515

3/4/5

PLM, FTIR, GC-MS

Bitumen, calcite

F15020

PS534

Palette

D12.8.7

12840

3/4/5

PLM, FTIR

Carbon, yellow ochre

F15132

PS879

Palette

D12.7.7

12111

3/4/5

GC-MS

Carbon, trace bitumen

F15137

PS877

Palette

D12.7.6

12062

3/4/5

PLM, FTIR, GC-MS

Carbon, trace bitumen

F15279

PS861

Palette

D11.2.5

2772

3/4/5

PLM, FTIR, GC-MS

Carbon, trace bitumen

F15666

PS873

Palette

D12.12.3

12346

3/4/5

PLM, FTIR, GC-MS

Carbon, trace bitumen

F15670

PS864

Palette

D11.2.6

2738

3/4/5

GC-MS

Carbon, trace bitumen

F17310

PS119

Palette

E13.29.1

5230

2B

PLM, FTIR, GC-MS

Bitumen

F17311

PS121

Palette

E13.14.7

5243

3A

PLM, FTIR, GC-MS

Bitumen / carbon

F17312

PS152

Palette

E13.14.6

5284

1B

PLM, FTIR, GC-MS

Bitumen / carbon

PS384

Wall plaster

E13.7.3

4709

2

PLM, FTIR

Gypsum, carbon, yellow ochre

PS494

Wall plaster

E13.7.6

4709

2

PLM, FTIR

Gypsum, carbon

PS545

Wall plaster

D12.8.1

12887

3/4/5

PLM

Yellow ochre, carbon

46

CHAPTER

carbon. The black paint on coffin F9743 (PS295) from G244 was bitumen mixed with calcite, which is unlikely to be contamination from the ground, as no gypsum was observed in the sample (see Fig. 17). In most cases the pattern of hopanes and steranes indicated that the bitumen had been sourced from the Dead Sea (see Fig. 19). There was one exception, palette PS121; the source of the bitumen on this palette has not yet been identified (Fulcher, Stacey, and Spencer 2020). Bitumen has only previously been identified as a ground pigment in Egypt by PLM on one object from the New Kingdom, and on Roman cartonnage (Siddall 2011). The identification of bitumen in several paint palettes at Amara West is a significant addition to our knowledge of Egyptian paints. Given the difficulties of distinguishing bitumen and carbon without molecular analysis, it is possible that bitumen has been underidentified as a pigment in previous analyses, and in future studies bitumen should be considered as a possible black pigment. Yellow Yellow paint on palettes and yellow lumps of pigment were frequently found at Amara West. The yellow pigments were identified using PLM, FTIR, SEM-EDS, and pXRF (Table 7). All the yellow pigments at Amara West have been identified as yellow ochre. As discussed above, yellow ochre is an iron oxide hydroxide (FeOOH), usually goethite, plus clays and quartz. Under plane polarized light, yellow iron oxide has yellow crystals,

3

Figure 20: Palette F7130 (PS448) from E13.31.1 [5325]. Palette holds both light and dark yellow paint.

varying in size and appearance from fibrous acicular crystals to large aggregates of fine crystals (Eastaugh et al. 2004b, 365; Helwig 2007). Under crossed polars they are birefringent but this is masked by their body colour so the grains appear bright yellow. Quartz grains in plane polarized light are clear and colourless with few internal features, conchoidal fracture, and low relief (Eastaugh et al. 2004b, 331). In crossed polars quartz has first order birefringence, increasing to second order in large crystals. Peaks for goethite

Figure 21: Dispersion of lighter yellow paint from palette F7130 (PS448), from E13.31.1 [5325], in plane polarized light (left) and crossed polars (right) ×400. Yellow iron oxide and calcite. Scale bars represent 30μm.

AMARA WEST PAINTS: ANALYSIS

47

Table 7: Results of pigment analysis of yellow paints, pigments, and plaster. Objects only analysed by portable XRF (pXRF) do not have a sample (PS) number because they were analysed directly and no sample was taken. Wall plaster samples do not have Find (F) numbers because they are samples taken directly from walls on site and are not finds. Find number

Sample number

Type of object

Building

Context

Phase

Analysis

Mineral identification

F2470

PS321

Palette

E13.20.1

10301

2B–3

PLM

Yellow ochre, calcite

Raw pigment

E13.20.1

10301

2B–3

pXRF

Yellow ochre

F2471 F2482

PS322 dark Palette

E13.20.1

10306

2B

PLM, pXRF

Yellow ochre

F2482

PS322 light Palette

E13.20.1

10306

2B

PLM, pXRF

Yellow ochre, calcite

F3136

PS407

Raw pigment

D13.4.6

3127

3/4/5

PLM, FTIR

Yellow ochre

F4630

PS426

Palette

E13.14.5

4356

1B–2A

PLM, FTIR, pXRF

Yellow ochre

F4881g

PS282

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Yellow ochre, gypsum

F4913

PS278

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Yellow ochre, gypsum

F5049g

PS291

Wall plaster

E13.7.6

4561

2B–3A

PLM

Yellow ochre, gypsum

F5101a

PS277

Wall plaster

E13.7.6

4561

2B–3A

PLM, FTIR, pXRF

Yellow ochre, gypsum

Raw pigment

E13.13

4745

2B–3A

pXRF

Yellow ochre

F5114 F5133c

PS293

Wall plaster

E13.7.6

4561

2B–3A

PLM

Yellow ochre, gypsum

F5133d

PS455

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Yellow ochre, gypsum

F5962

PS306

Grindstone

E13.31.2

5332

2B–3A

PLM, pXRF

Yellow ochre

F6044

Raw pigment

E13.7.11

5255

2A

pXRF

Yellow ochre

F6048

PS420

Palette

E13.29.1

5219

2B–3A

PLM, pXRF

Yellow ochre, calcite

F6079

PS128

Palette

E13.14.8

5222

3A

PLM, FTIR, pXRF, SEM-EDS

Yellow ochre, calcite

F6079

PS129

Palette

E13.14.8

5222

3A

PLM, FTIR, pXRF, SEM-EDS

Yellow ochre, calcite

F6079

PS422

Palette

E13.14.8

5222

3A

PLM, pXRF

Yellow ochre, red ochre

F6119

PS423

Palette

E13.31.2

5346

2B

PLM, FTIR, pXRF

Yellow ochre, gypsum, calcite

F6124

Palette

E13.31.2

5346

2B

pXRF

Yellow ochre

F6142

PS421 dark Palette

E13.31.1

5336

2B

PLM, pXRF

Yellow ochre, calcite

F6142

PS421 light Palette

E13.31.1

5336

2B

PLM, pXRF

Yellow ochre, calcite

F6147

PS132

Palette

E13.31.1

5325

3A

PLM, FTIR, pXRF

Yellow ochre, anhydrite

F6147

PS427

Palette

E13.31.1

5325

3A

PLM, pXRF

Yellow ochre, gypsum

F6147

PS429

Palette

E13.31.1

5325

3A

PLM, pXRF

Yellow ochre, calcite

F6170

PS419

Palette

E13.6.4

5341

3A–3B

PLM, pXRF

Yellow ochre

F6190

PS433

Palette

E13.31.2

5332

2B

PLM, pXRF

Yellow ochre, calcite

Raw pigment

E13.7.13

5283

2A

pXRF

Yellow ochre

F6201 F6219

PS434 dark Palette

E13.14.7

5243

3A

PLM

Yellow ochre

F6219

PS434 light Palette

E13.14.7

5243

3A

PLM

Yellow ochre, calcite

F6223

PS439

Palette

E13.7.12

5224

3A

PLM, pXRF

Yellow ochre, gypsum

F6264

PS435

Palette

E13.7.12

5261

2A

PLM, pXRF

Yellow ochre, calcite

Raw pigment

E13.15

5039

1

pXRF

Yellow ochre

Raw pigment

F6356 F6411

E13.31.2

5334

2B

pXRF

Yellow ochre

F6419

PS454 dark Palette

E13.31.2

5334

2B

PLM

Yellow ochre, red ochre

F6419

PS454 light Palette

E13.31.2

5334

2B

PLM

Yellow ochre, calcite

F6423

PS452

Palette

E13.6.3

5301

3B

PLM

Yellow ochre, gypsum

Raw pigment

E13.14.1

5365

1B

pXRF

Yellow ochre

F6433 F6440

Raw pigment

E13.14.10

5346

2B

pXRF

Yellow ochre

F6446

PS445

Palette

E13.14.1

5402

1B

PLM

Yellow ochre

F6467

PS430

Palette

E13.31.2

5334

2B

PLM, FTIR, pXRF

Yellow ochre, red ochre

Raw pigment

E13.14.2

5376

1B

pXRF

Yellow ochre

F6489

CHAPTER

48

3

Find number

Sample number

Type of object

Building

Context

Phase

Analysis

Mineral identification

F6493

PS440

Palette

E13.14.1

5361

2A

PLM

Yellow ochre

F6499

PS451

Palette

E13.14.10

5339

2B

PLM

Yellow ochre, gypsum

F6915

PS376

Raw pigment

E13.16.1

5495

3B

FTIR

Yellow ochre

F7124

PS444

Palette

E13.14.1

5402

1B

PLM

Yellow ochre

F7130

PS448 dark Palette

E13.31.1

5325

3A

PLM, FTIR

Yellow ochre

F7130

PS448 light Palette

E13.31.1

5325

3A

PLM, FTIR

Yellow ochre, calcite

F7139

PS441

Palette

E13.14.1

5297

1B

PLM

Yellow ochre

F7278

PS323

Palette

D12.5.12

2538

3/4/5

PLM, pXRF, SEM-EDS

Yellow ochre

F7452

Raw pigment

E13.5

10300

3/4/5

pXRF

Yellow ochre

F7557

Raw pigment

E13.20.5

10327

2–3

pXRF

Yellow ochre

F7577

Raw pigment

E13.20.5

10339

2B

pXRF

Yellow ochre

F7684

PS537

Palette

E13.20.5

10434

2B

PLM

Yellow ochre, gypsum, calcite

F7794

PS473

Palette

E13.17

10092

2A–2B

PLM

Yellow ochre, calcite

F8110

PS307

Coffin

G309

8163

3/4/5

PLM, FTIR, SEM-EDS

Yellow ochre, gypsum

F8110

PS527

Coffin

G309

8163

3/4/5

PLM

Yellow ochre, gypsum

F9485

PS529

Coffin

G222

9485

3/4/5

PLM

Yellow ochre, gypsum

F9535

PS528

Coffin

G238

9195

3/4/5

PLM

Yellow ochre, gypsum, carbon

F9714

PS300

Coffin

G244

9505

3/4/5

PLM, SEM-EDS

Yellow ochre, gypsum

F9735

PS297

Coffin

G244

9511

3/4/5

PLM

Yellow ochre, gypsum

F9741

PS296

Coffin

G244

9511

3/4/5

PLM, SEM-EDS

Yellow ochre, gypsum

F12204

PS310

Raw pigment

D12.7.1

12000

3/4/5

SEM-EDS, pXRF, FTIR

Yellow ochre

Raw pigment

D12.7.3

12048

3/4/5

pXRF

Yellow ochre

F12219 F14018

PS542

Clay bowl

D12.8.7

12840

3/4/5

PLM

Yellow ochre

F15020

PS534

Palette

D12.8.7

12840

3/4/5

PLM

Yellow ochre

F15029

PS543

Palette

D11.1.5

12706

3/4/5

PLM

Yellow ochre

F17313

PS114

Raw pigment

E13.15

5037

1

FTIR

Yellow ochre

PS255

Wall plaster

E13.7.5

4566

2B

PLM

Yellow ochre, gypsum

PS267

Wall plaster

E13.4.2

4471

4

PLM, FTIR

Yellow ochre

PS269

Wall plaster

E13.4.1

4425

3A

PLM, FTIR

Yellow ochre, calcite

PS272

Wall plaster

E13.7.6

4727

2B

PLM, FTIR

Yellow ochre, calcite

PS273

Wall plaster

E13.7.6

4727

2B

PLM

Yellow ochre, gypsum

PS279

Wall plaster

E13.7.6

4727

2B

PLM

Yellow ochre, gypsum

PS482

Wall plaster

E13.23

10475

1A

PLM

Yellow ochre, gypsum

PS488

Door fill

E13.4.2

4471

4

PLM

Yellow ochre, gypsum

PS552

Wall plaster

D12.8.2 or D12.9.14?

found out of context

3/4/5

PLM

Yellow ochre

PS559

Wall plaster

E13.7.5

4566

2B

PLM

Yellow ochre

PS564

D12.7.1 pit

D12.7.1

12068

3/4/5

PLM

Yellow ochre, calcite

PS566

Stone threshold D12.8.2

12894

3/4/5

PLM

Yellow ochre, gypsum

AMARA WEST PAINTS: ANALYSIS

were only rarely identified using FTIR, but clays and quartz were identifiable; kaolinite bands were most prominent and presumably masked those for goethite. Those samples analysed by SEM-EDS and pXRF contained no elements outside of the range expected for earth pigments, namely iron, aluminium, and silicon (Helwig 2007). Often the yellow pigments are found mixed with a white pigment, either gypsum or calcite. Calcite mixed with yellow is found mainly on ceramic palettes (Figs 20, 21). There are only two instances of yellow/ calcite on walls, one in E13.4.1 (30% calcite, PS269) and one from the mastaba in E13.7.6 (5% calcite, PS272). The calcite in these mixtures has probably been introduced as an inclusion in the ochre. All other yellow samples from walls consisted of yellow ochre mixed with gypsum. Many of the yellow samples had a significant quantity of quartz present. Quartz can be added to pigments to assist with the grinding process, especially if the raw material is soft or clay-like (Eastaugh et al. 2004a, 321). That quartz was being used in this way at Amara West is suggested by the high percentage of quartz (50%) in the yellow pigment sample taken from the large grindstone found in E13.14.2 (F5962, PS306, see Fig. 6), and the high quantities of quartz in many of the yellow and red samples. However, it is also possible that there was a high level of quartz naturally occurring in the yellow ochre that was being used to create some of the pigments. Lumps of raw pigment that differ in appearance found at Amara West suggest that different sources of ochre were being used, some providing fine-grained ochre with a strong yellow colour and others a paler, coarser yellow sandstone, which is yellow iron oxide particles adhering to large rounded grains of silica. If yellow sandstone were ground sufficiently finely, it would be indistinguishable from fine-grained yellow ochre with quartz. The two are chemically identical enough (iron oxide and quartz) that they give the same results from elemental analyses and FTIR. Seven samples of yellow paint were taken from coffin fragments. The pigments were all identified as yellow iron oxide. Red Similarly to yellow, red paint on palettes and red pigments were frequently found in excavations at Amara West. The red pigments were identified using PLM, FTIR, SEM-EDS, and pXRF (Table 8). All the

49

red pigments at Amara West have been identified as red ochre, coloured by haematite, an anhydrous iron oxide (αFe2O3), with quartz and clays. Pinks and oranges are included here, because they comprise red ochre combined with other minerals. Under plane polarized light, red iron oxide has bright red to brown crystals, varying in size and appearance from minute round ruby red crystals to large dark aggregates of fibrous crystals (Eastaugh et al. 2004b, 363). Under crossed polars they are birefringent but this is masked by their body colour so the grains appear bright red (Fig. 22). Peaks for haematite could not be identified using FTIR, but clays and quartz were identifiable. In common with the yellow pigments, the samples analysed by SEM-EDS and pXRF contained no elements outside of the range expected for earth pigments. The red pigments are often found mixed with a white pigment, creating a pink paint, the white component being either gypsum or calcite. Of the 39 red/pink samples from palettes analysed, 13 contained gypsum, and 21 contained calcite. Calcite mixed with red is found mainly on ceramic palettes and coffins. Pink paint on palettes can contain up to 99% calcite. Other samples of red mixed with a white pigment contained gypsum. All the samples of red paint from plastered walls were found to be red iron oxide in combination with gypsum, with the exception of one sample from the mastaba in E13.7.6 [4727], which contained a small amount of calcite (10%), probably a mineral inclusion in the ochre rather than a deliberate addition. Of the 9 red samples taken from painted coffin plaster, 4 contained calcite (PS307 1%, PS308 10%, PS502 20%, PS531 1%). The plaster on one coffin (F9761, PS502) was composed of 100% gypsum, meaning that the calcite in the paint was not contamination from the ground, and could have been added deliberately, or have been present in the ochre. Gypsum was also found in association with red paint on the coffins, although it was difficult to determine if this was the result of depositional contamination. Red sediments were found on the floors of several rooms at Amara West, and often there were several layers visible in thin section, from multiple applications of red material. Matthew Dalton has studied the floor deposits at Amara West and concluded that redcoloured rocks were being crushed and mixed with water to use as a red slurry (Dalton 2020, 241–72). A range of rocks were used for this purpose, as demonstrated by a sequence of six micromorphologically analysed red sediment layers in oven room

50

CHAPTER

3

Figure 22: Dispersion of red pigment F6177 (PS329) from E13.31.1 [5356] in plane polarized light (left) and crossed polars (right) ×400. Scale bars represent 30μm.

Figure 23: Dispersion of red paint PS380 from the pit in D12.7.1, in plane polarized light (left) and crossed polars (right) ×400. Red iron oxide and quartz. Scale bars represent 30μm.

D12.8.8: sediments derived from red sandstone, ochre, and schist were all identified. Similar red slurry was splashed on the walls in several rooms, including the walls around the pit in D12.7.1. The splashed red consists of large (approx. 30μm diameter) rounded transparent grains that have first order birefringence in crossed polars, coated in tiny yellow or red grains (Fig. 23). This indicates that the material is composed primarily of quartz grains covered in red iron oxide, and thus appears to derive from sandstone. Such red splashes, with large quartz grains, are identifiable on the walls of E13.4.2 (PS268, PS487, PS490), and the walls and floor of D12.7.1 (PS250, PS252, PS380,

PS377, PS379, PS525). However, it also occurs in some of the palettes from E13.14 (PS124 from F6142 and PS451 from F6499), so it seems that this material was used across a range of contexts at Amara West. There appear to have been two main types of red mineral in use at Amara West: fine-grained ochre and red sandstone (schist is much less used). The uses of red ochre and red sandstone seem to have overlapped. Both types of red are found on palettes, on walls, and over floors. If finely ground, red sandstone would be difficult to distinguish from ochre with a high quartz component, so the two may have been used interchangeably.

AMARA WEST PAINTS: ANALYSIS

51

Table 8: Results of pigment analysis of red paints, pigments, and plaster. Objects only analysed by portable XRF (pXRF) do not have a sample (PS) number because they were analysed directly and no sample was taken. Wall plaster samples do not have Find (F) numbers because they are samples taken directly from walls on site and are not finds, with the exception of F5133 which was found on the floor and was highly decorated. Find Sample number number F2439

PS325

F2471

Type of object

Building

Context

Phase

Analysis

Mineral identification

Palette

E13.16.2

5585

3B

PLM, FTIR, pXRF

Red ochre

Raw pigment

E13.20.1

10301

2B-3

pXRF

Red ochre

F2493

PS289

Raw pigment

E13.20.1

10323

2B

PLM

Red ochre

F4190

PS390

Grindstone

E13.3.26

4275

3B

PLM

Red ochre, gypsum, calcite

F4265

PS370

Grindstone

E13.14.5

4316

3A

PLM

Red ochre

F4881

PS280

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Red ochre, yellow ochre, gypsum

F4881

PS285

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Red ochre, gypsum

F5003a

PS290

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Red ochre, yellow ochre, gypsum

F5014

PS404

Raw pigment

E13.9.13

4713

4

PLM, FTIR

Red ochre

F5133d

PS455

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Red ochre, calcite

F5833

Raw pigment

E13.29.1

5219

2B–3A

pXRF

Red ochre

F6077

Raw pigment

E13.14.8

5222

3A

pXRF

Red ochre

F6079

PS129 red

F6119

Palette

E13.14.8

5222

3A

PLM, FTIR, SEM-EDS Red ochre

Palette

E13.31.2

5331

2B

pXRF

Red ochre Red ochre, yellow ochre, gypsum, calcite

F6124

PS423

Palette

E13.31.2

5346

2B

PLM, pXRF

F6124

PS424

Palette

E13.31.2

5331

2B

PLM, pXRF

Red ochre, calcite

F6142

PS124

Palette

E13.31.1

5336

2B

PLM

Red sandstone

F6142

PS125

Palette

E13.31.1

5336

2B

PLM, FTIR, SEM-EDS Red ochre, gypsum, anhydrite, calcite

F6142

PS421

Palette

E13.31.1

5336

2B

PLM

Red ochre

F6144

Raw pigment

E13.31.1

5336

2B

pXRF

Red ochre

F6146

Palette

E13.31.1

5336

2B

pXRF

Red ochre Red ochre

F6147

PS130

Palette

E13.31.1

5325

3A

PLM, FTIR, pXRF

F6147

PS131

Palette

E13.31.1

5325

3A

PLM, FTIR, SEM-EDS, Red ochre, gypsum, calcite pXRF

F6147

PS428 pink Palette

E13.31.1

5325

3A

PLM, pXRF

F6147

PS428 red

Palette

E13.31.1

5325

3A

PLM, pXRF

Red ochre, gypsum, calcite

F6156

PS416

Palette

E13.31.1

5354

2B

PLM

Red ochre

F6165

PS425 pink Palette

E13.31.1

5348

2A

PLM, pXRF

Red ochre, calcite

F6165

PS425 red

Palette

E13.31.1

5348

2A

PLM, pXRF

Red ochre, calcite

F6170

PS419

Palette

E13.6.4

5341

3B

PLM, pXRF

Red ochre, gypsum, calcite

F6177

PS329

Raw pigment

E13.31.1

5356

2A

PLM, FTIR, pXRF

Red ochre

F6184

PS274

Grindstone

E13.14.1

5361

1B

PLM

Red ochre

F6190

PS123 pink Palette

E13.31.2

5332

2B

PLM, FTIR

Red ochre, gypsum

F6190

PS123 red

Palette

E13.31.2

5332

2B

PLM

Red ochre, gypsum

F6190

PS432

Palette

E13.31.2

5332

2B

PLM, FTIR, pXRF

Red ochre, gypsum, calcite

F6194

PS136

F6198

Red ochre, calcite

Palette

E13.31.2

5352

3A

PLM

Red ochre

Raw pigment

E13.31.2

5352

3A

pXRF

Red ochre

Palette

E13.14.7

5253

3A

PLM, SEM-EDS

Red ochre

F6210

PS446

F6219

PS434 pink Palette

E13.14.7

5243

3A

PLM

Red ochre, calcite

F6219

PS434 red

Palette

E13.14.7

5243

3A

PLM

Red ochre, gypsum, calcite

F6223

PS439

Palette

E13.7.12

5224

3A

PLM

Red ochre, calcite

F6264

PS435

Palette

E13.7.12

5261

2B

PLM, SEM-EDS

Red ochre, calcite

F6289

PS453

Palette

E13.7.11

5255

2B

PLM

Red ochre

Palette

E13.31.1

5345

3A

pXRF

Red ochre

F6403

CHAPTER

52

3

Find Sample number number

Type of object

Building

Context

Phase

Analysis

Mineral identification

F6407

Raw pigment

E13.31.2

5334

2B

pXRF

Red ochre

F6408

PS141

Palette

E13.31.2

5334

2B

PLM, FTIR

Red ochre, calcite

F6408

PS247

Palette

E13.31.2

5334

2B

PLM, FTIR

Red ochre, calcite

F6435

Raw pigment

E13.31.2

5371

2A

pXRF

Red ochre

F6441

Raw pigment

E13.31.2

5372

2A

pXRF

Red ochre

Palette

E13.14.2

5346

2B

PLM

Red ochre, calcite

F6463

PS436

F6467

PS430 pink Palette

E13.31.2

5334

2B

PLM, FTIR, pXRF

Red ochre, calcite

F6467

PS430 red

Palette

E13.31.2

5334

2B

PLM, FTIR, pXRF

Red ochre, gypsum, calcite Red ochre

F6468

Raw pigment

E13.31.2

5334

2B

pXRF

F6499

PS451

Palette

E13.14.10

5339

2B

PLM, SEM-EDS

Red sandstone

F6519

PS442

Palette

E13.7.12

5224

2B–3A

PLM, FTIR

Red ochre, calcite

F7120

Raw pigment

E13.14.1

5402

1B

pXRF

Red ochre

F7124

PS444

Palette

E13.14.1

5402

1B

PLM

Red ochre

F7128

PS449

Palette

E13.14.7

5274

3A

PLM, SEM-EDS

Red ochre, gypsum

F7153

PS447 dark Palette

E13.14.3

5281

1B

PLM, SEM-EDS

Red ochre

F7153

PS447 light Palette

E13.14.3

5281

IB

PLM, SEM-EDS

Red ochre

F7794

PS473

Palette

E13.17

10092

2A–2B

PLM

Red ochre, gypsum

F8100

PS308

Coffin

G309

8137

3/4/5

PLM, FTIR

Red ochre, calcite

F8110

PS307

Coffin

G309

8163

3/4/5

PLM

Red ochre, calcite

F9079

PS531

Coffin

G201

9018

3/4/5

PLM

Red ochre, gypsum, calcite

F9485

PS529

Coffin

G222

9485

3/4/5

PLM

Red ochre, calcite

F9658

PS302

Coffin

G244

9505

3/4/5

PLM

Red ochre, carbon

F9714

PS300

Coffin

G244

9505

3/4/5

PLM

Red ochre, gypsum

F9741

PS296

Coffin

G244

9511

3/4/5

PLM

Red ochre

F9742

PS301

Coffin

G244

9519

3/4/5

PLM

Red ochre, gypsum

F9761

PS502

Coffin

G244.1

9244

3/4/5

PLM

Red ochre, calcite

F12018

PS288

Raw pigment

D12.6.1

2803

3/4/5

PLM, FTIR, pXRF

Red ochre

F12233

PS260

Wall plaster

D12.7.3

12001

3/4/5

pXRF

Iron oxide

F15020

PS534

Palette

D12.8.7

12840

3/4/5

PLM

Red ochre, gypsum, calcite

F15026

PS536

Palette

D12.8.9

12836

3/4/5

PLM

Red ochre

F17314

PS101

Raw pigment

E13.8.3

4673

2B–5

PLM

Red ochre

PS250

D12.7.1 N wall

D12.7.1

12017

3/4/5

PLM

Red sandstone

PS252

D12.7.1 W wall

D12.7.1

12018

3/4/5

PLM, FTIR

Red sandstone, gypsum

PS254

Wall plaster

E13.7.6

4566

2B

PLM, FTIR

Red ochre, gypsum

PS258

Wall plaster

E13.3.24

4068

3A– 3B–4

PLM

Red ochre, yellow ochre, gypsum

PS263

Ceiling fragment D12.6.2

2812

3/4/5

PLM, FTIR, pXRF

Red sandstone

PS268

Wall plaster

E13.4.2

4471

3A– 3B–4

PLM

Red sandstone

PS377

D12.7.1 pit

D12.7.1

12068

3/4/5

PLM

Red sandstone

PS379

D12.7.1 pit

D12.7.1

12068

3/4/5

PLM

Red sandstone

PS380

D12.7.1 E wall

D12.7.1

12012

3/4/5

PLM, FTIR

Red sandstone

PS385

Plaster on mastaba

E13.7.6

4727

2B

PLM

Red ochre, gypsum, calcite

PS487

Wall plaster

E13.4.2

4426

3A

PLM

Red sandstone, calcite

PS490

Wall plaster

E13.4.2

4425

3A

PLM

Red sandstone

PS525

D12.7.1 pit

D12.7.1

12068

3/4/5

PLM

Red sandstone, gypsum, calcite

PS526

D11 from floor

D11.1.2

12518

3/4/5

PLM

Red ochre, calcite

PS558

Wall plaster

E13.7.5

4566

2B

PLM

Red ochre, gypsum

PS565

Stone threshold

D12.8.2

12894

3/4/5

PLM

Red ochre, gypsum

AMARA WEST PAINTS: ANALYSIS

Blue Blue pigment was less common at Amara West than red and yellow. All samples of blue that have been excavated were analysed, and identified using PLM, pXRF, SEM-EDS, and FTIR (Table 9). Egyptian blue was identified in 32 of the 35 blue samples from Amara West. In plane polarized light Egyptian blue particles are light blue translucent to transparent crystals; large particles are typical and there is often a large distribution of particle sizes (Fig. 24) (Eastaugh et al. 2004b, 27). They are pleochroic from blue to colourless, and red through a

53

Chelsea filter. In crossed polars they display third order and above birefringence colours but this may be masked by body colour. pXRF results revealed a high incidence of copper compared to the substrate. FTIR spectra for these samples were a good match with reference materials of Egyptian blue and spectra in reference libraries (Fig. 25). In the scanning electron microscope the BSE image showed dark grey, pale grey, and bright white areas, which were analysed using EDS. The ratio of the constituent elements in the pale grey areas was Ca:Cu:Si:O 1:1:4:10, consistent with calcium copper silicate (cuprorivaite CaCuSi4O10), the blue component of the Egyptian blue pigment (Appendix Table 4).

Figure 24: Dispersion of paint containing Egyptian blue from palette F12423 (PS538), from D12.10 [12211], seen under plane polarized light (left) and crossed polars (right), both at ×400. Scale bar shows 30μm.

Figure 25: FTIR spectrum for blue paint on palette F6223 (PS140) from E13.7.12 [5224]. Peaks indicate Egyptian blue.

54

CHAPTER

Bright white areas were found to contain a high percentage of tin, and pXRF also identified some tin in the blue pigment, suggesting that the pigment was manufactured using a copper-tin alloy. Blue paint on palettes is mostly present in only small quantities, especially compared to palettes with yellow and red paint. It appears that the blue paint may have been scraped or washed off the palettes after use, sometimes leaving only traces behind in the voids of the ceramic (Fig. 26). Washing a palette, retaining the water, and evaporating it would allow the pigment to be reclaimed for further use. This suggests that blue paint was scarce and valued at Amara West. Egyptian blue was identified on 16 palettes. On 4 of these, blue was mixed with gypsum. A very low amount of calcite (2%) was found in association with Egyptian blue on one palette (PS538 from F12423). Samples of blue paint were taken from painted coffin fragments from G244 (PS503 from F9707 and PS302 from F9658) and G309 (PS307 from F8110); these were all identified as Egyptian blue. Only one clear example of blue being used to paint domestic walls was excavated at Amara West. A fragment of wall plaster from the niche in E13.7.6 retained a decorative scheme on an earlier layer, after the upper layers of plaster had been removed by a conservator (see Fig. 8). Another, less certain, use of blue for decoration can be seen on a mud plaster fragment F12233 from D12.7.3 [12001], which retains a streak of light blue and a streak of dark blue. That these paints occurred alongside one another suggests that the painter

Figure 26: Ceramic palette F6147 from E13.31.1 [5325] with traces of blue pigment in voids.

3

was aiming to achieve a lighter and a darker blue. These examples were all Egyptian blue. The composition of blue paint from Amara West, observed using PLM, shows that only a little Egyptian blue pigment is required in the paint to achieve a blue colour. In no instance was the percentage of calcium copper silicate in the paint above 50%; the blue paint on F5133d was 5% calcium copper silicate, and the blues from F12233 contained only about 1% calcium copper silicate, although in both cases this is a difficult number to estimate because the substrate may be accidentally sampled along with the paint where it is very thin. VIL analysis was undertaken on site on the wall plaster from the mastaba and the niche above the mastaba from E13.7.6, and on the West Gate into the walled town. The images showed that the use of Egyptian blue at Amara West was more extensive than can be seen by the naked eye. The West Gate scenes retained traces of Egyptian blue in the spear held by a figure of a king, the wheels of his chariot, the hieroglyphic inscriptions, and register lines. Traces of Egyptian blue were also identified using VIL on many of the excavated pieces of mud plaster from E13.7.6, most of which were no longer visible. No evidence for the manufacture of Egyptian blue has been found at Amara West, for example crucibles with blue crystals adhering to the inner surface, or cakes of Egyptian blue at various stages in the production cycle, as have been found at other sites (Rehren, Pusch, and Herold 1998; Nicholson 2007). Fragments of plaster trays from Amara West (for example F15198) may be indicative of faience production, given the association of these object types with glass- and faience-production areas at Amarna (Nicholson 2007, 139–41). Many fragments of such trays were found in area E13.17, along with metal-working detritus such as crucibles. The evidence for these high-temperature activities at Amara West suggests that the skills to manufacture Egyptian blue were available, at least during this phase of the town (Phase 1A–2B), and it is possible that Egyptian blue manufacture was taking place in or near the town in an area that has not yet been excavated. Three examples of a grey-blue paint were found on palettes in the western suburb (Fig. 27), and one in the rubbish layers in D11.7. Sample PS539 from palette F2644 was analysed using SEM-EDS; it contained no copper or cobalt, the elements present were primarily iron, magnesium, silicon, and aluminium. All four were

AMARA WEST PAINTS: ANALYSIS

55

Table 9: Results of pigment analysis of blue paints, pigments, and plaster. Portable XRF is abbreviated to pXRF. Find number

Sample number

Type of object

Building

Context

Phase

Analysis

Mineral identification

F2301

PS313

Limestone fragment

D12.5.11

2320

3/4/5

PLM, pXRF

Egyptian blue

F2600

PS540

Palette

E13.20.1

10324

2B

PLM

Egyptian blue

F2644

PS539

Palette

D12.7.6

12062

3/4/5

PLM, FTIR, SEM-EDS, XRD

Blue earth

F4881a

PS281

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Egyptian blue, gypsum, yellow ochre

F5133d

PS455

Wall plaster

E13.7.6

4561

2B–3A

PLM, pXRF

Egyptian blue, gypsum

F5273

PS394

Raw pigment

E13.7.8

4762

2B

SEM-EDS

Egyptian blue

F6045

PS315

Raw pigment

E13.29.1

5219

2B–3A

SEM-EDS

Egyptian blue

F6047

PS418

Palette

E13.29.1

5219

2B–3A

PLM, pXRF

Egyptian blue

F6147

PS132

Palette

E13.31.1

5325

3A

PLM, FTIR, SEM-EDS, pXRF

Egyptian blue

F6147

PS427

Palette

E13.31.1

5325

3A

PLM, FTIR, pXRF Egyptian blue, gypsum

F6169

PS437

Palette

E13.31.2

5352

2B

PLM

F6170

PS127

Palette

E13.6.4

5341

3B

PLM, FTIR

Egyptian blue, gypsum

F6190

PS431

Palette

E13.31.2

5332

2B

PLM, pXRF

Egyptian blue

F6223

PS140

Palette

E13.7.12

5224

3A

PLM, FTIR, SEM-EDS, pXRF

Egyptian blue

F6223

PS438

Palette

E13.7.12

5224

3A

PLM, pXRF

Egyptian blue, gypsum

F6255

PS316

Raw pigment

E13.7.11

5248

2B

SEM-EDS

Egyptian blue

F6264

PS417

Palette

E13.7.12

5261

2A

PLM, pXRF

Egyptian blue

F6474

PS443

Palette

E13.31.1

5336

2B

PLM, pXRF

Egyptian blue

F7353

PS317

Raw pigment

E13.7.9

5702

2A–2B

SEM-EDS

Egyptian blue

F7530

PS305

Raw pigment

E13.16.2

5583

4

pXRF, FTIR, SEM-EDS

Egyptian blue

F7537

PS304

Raw pigment

E13.16.2

5583

4

SEM-EDS, pXRF

Egyptian blue

F7569

PS287

Palette

E13.20.5

10331

2B–2C

PLM, pXRF

Egyptian blue

F7569

PS324

Palette

E13.20.5

10331

2B–2C

PLM

Egyptian blue

F7684

PS537

Palette

E13.20.5

10434

2B

PLM

Egyptian blue, gypsum

F8110

PS307

Coffin

G309

8163

3/4/5

PLM

Egyptian blue, gypsum, calcite

F9658

PS302

Coffin

G244

9505

3/4/5

PLM

Egyptian blue, gypsum, calcite, carbon, yellow ochre

F9707

PS503

Coffin

G244

9505

3/4/5

PLM

Egyptian blue, gypsum, calcite, yellow ochre

F12233

PS260

Wall plaster

D12.7.3

12001

3/4/5

PLM, pXRF

Egyptian blue, carbon, yellow ochre

F12233

PS261

Wall plaster

D12.7.3

12001

3/4/5

PLM, pXRF

Egyptian blue, carbon, yellow ochre

F12423

PS538

Palette

D12.10

12211

3/4/5

PLM

Egyptian blue, calcite

F15020

PS534

Palette

D12.8.7

12840

3/4/5

PLM

Egyptian blue

F15193

PS893

Palette

D12.8.8

12891

3/4/5

FTIR

Egyptian blue

F15656

PS860

Palette

D11.2.4

2716

3/4/5

PLM

Blue earth

F16667

PS865

Palette

D11.7

13569

3/4/5

PLM, FTIR

Blue earth

F17315

PS103

Raw pigment

E13.6.3

5301

3B

FTIR

Egyptian blue

Egyptian blue

56

CHAPTER

Figure 27: Palette F2644 (PS539) from D12.7.6 [12062], with grey-blue paint.

analysed under PLM, but no observable blue particles could be seen. FTIR analysis was inconclusive, giving only the peaks expected for clays. A sample of the grey-blue pigment from palette PS539 was sent to Dr Trevor Emmett at Anglia Ruskin University for XRD analysis. It was found to contain major phases of clinochlore (magnesium rich chlorite) and amphiboles. Chlorites are mixed crystals with complex chemistry, and the metal ions present within them determine their colour (Lauf 2010; Haldar and Tišljar 2014). Chlorite has been described as having a ‘distinctly bluish’ colour (Grissom 1986, 37). The colour may also be influenced by the amphiboles (NaCa2(Mg,Fe,Al)5(Al,Si)8O 22(OH)2; the Fe, Mg, and Al ions substitute freely for one another), some of which have a blue colour (Eastaugh et al. 2004a, 17). The presence of this mineral on a palette so closely resembling those hundreds from the site which have been used for paint suggests this was a ‘paint’ of some sort. If so, it may represent an attempt to find a blue paint that was locally available and did not require access to Egyptian blue.

3

the green paint from the palettes was partially obscured by the peaks for calcite, but included peaks in the 980 region and at 756 and 645, suggesting a clayey green earth (Grissom 1986; Moretto, Orsega, and Mazzocchin 2011). Alongside the green particles in the pale green paints was a large proportion (25–45%) of colourless micas and feldspars. Feldspars are extremely common and occur in earth pigments and clays (Deer et al. 1992, 391–95); micas are sheet silicate minerals that are compositionally closely related to chlorites and occur alongside (Deer et al. 1992, 332). A grindstone (F6184) from E13.14 retained traces of a soft bright green pigment (PS506), and small lumps of a visually similar material were found in the sand nearby (PS118). Both samples were analysed by PLM, FTIR, and SEM-EDS (see Table 10). In PLM the pigments were very similar, and had a vivid green colour in plane polarized light, with moderate birefringence and some anomalous blue in crossed polars, which suggests atacamite or malachite (Eastaugh et al. 2004b, 54–55, 60–61). The FTIR spectra of the two samples were also very similar, and a good match for atacamite. SEM-EDS analysis showed that the majority of both pigments were made up of copper and chlorine. Therefore it can be concluded that these green pigments, from the grindstone and the floor next to it, were copper chloride hydroxide, atacamite type. One palette with a smear of green paint (PS869, F16767) was excavated from D11.7 [13568], a trench in the Western Suburb that contained rubbish. A sample

Green Very few examples of green pigment were found at Amara West. Three palettes held pale green paint, and all were analysed by PLM and FTIR (Table 10; Fig. 28). The pale green paint from palette F6119 (PS126) was also analysed in the scanning electron microscope by EDS and a secondary electron image captured, and by XRD by Trevor Emmett at Anglia Ruskin. In PLM the green crystals in the paint from palettes exhibited typical chlorite attributes; in plane polarized light the particles appeared translucent green with pleochroism to yellowish green, and in crossed polars the particles showed anomalous blue (Kerr 1959, 399; Eastaugh et al. 2004b, 103). The FTIR analysis of

Figure 28: Pale green paint on palette F6119 (PS126), from E13.31.2 [5331].

AMARA WEST PAINTS: ANALYSIS

57

Table 10: Results of pigment analysis of green paints, pigments, and plaster. Find number

Sample number

Type of object

Building

Context

Phase

Analysis

F6119

PS126

Palette

E13.31.2

5331

2B

PLM, FTIR, Chlorite, micas & feldspars, quartz, calcite SEM-EDS, XRD

F6184

PS506

Grindstone

E13.14.1

5361

1B

PLM, FTIR, SEM-EDS

Copper chloride (atacamite type)

F6463

PS248

Palette

E13.14.10

5346

2B

PLM, FTIR

Chlorite, micas & feldspars, quartz, calcite

F6467

PS249

Palette

E13.31.2

5334

2B

PLM, FTIR

Chlorite, micas & feldspars, quartz, calcite

F16767

PS869

Palette

D11.7

13568

3/4/5

PLM, FTIR

Egyptian blue, yellow ochre, calcite

F17309

PS118

Raw pigment

E13.14.1

5365

1B

PLM, FTIR, SEM-EDS

Copper chloride hydroxide (atacamite type)

was analysed by PLM and FTIR and identified as Egyptian blue mixed with yellow iron oxide. This mixture used to make a green pigment is known from other Egyptian contexts. Summary of pigment analysis results Most of the pigments identified were those commonly identified from Egypt itself: yellow and red ochres, gypsum, calcite, carbon black, and Egyptian blue (Table 11). The low frequency of blue finds can be compared to the much higher frequency at Amarna, which as an Egyptian royal city presumably had better access to skills and materials. Particular

Mineral identification

white pigments appear to be reserved for specific roles. A few examples of non-copper-based blues and greens evidence the use of naturally occurring earths, which may be a local adaptation to a lack of access to the more usual glassy manufactured blue and green pigments. It is also possible that blue and green earths were more widely used but have been under-identified due to in situ analysis methods and lack of evidence from domestic buildings. Another green was identified as copper chloride hydroxide (atacamite type), which was probably manufactured, given its rarity as a mineral. Bitumen was tentatively identified using FTIR, and confirmed using molecular analysis.

Table 11: Summary of pigments identified from Amara West. Colour

Pigments identified

Context

Gypsum

All walls, plasters, mixed with colours in palettes

Calcite

Mixed with colours in palettes, coffin paint and plaster

Huntite

1 coffin

Carbon

Palettes, walls, coffins

Bitumen

Coffin paint, palettes

Yellow

Ochre

Palettes, walls, coffins, grindstones, pigments

Red

Ochre

Palettes, walls, coffins, grindstones, pigments

Egyptian blue

Palettes, 2 walls, coffins, pigments

Blue earth

4 palettes

Chlorite

3 palettes

Copper chloride hydroxide (atacamite type)

1 grindstone, 1 pigment

Egyptian blue + yellow ochre

1 palette

White

Black

Blue

Green

58

CHAPTER

Binder analysis: results Of the 17 Amara West samples analysed, 8 gave results containing identifiable sugars, and 4 had a good range of monosaccharides preserved. The most complete ranges were found for PS445 (white palette, gypsum), PS435 (red palette, iron oxide), PS129 (yellow palette, iron oxide) and PS256 (white wall plaster, gypsum). This indicates that a plant gum was being used as a binder at Amara West. Both fucose and mannose were identified in two samples, which suggests a mixture of Acacia sp. and Astragalus sp. gums (Bleton et al. 1996; Vallance et al. 1998). Acacia is known to have been available at Amara West from plant remains in the archaeological record (Ryan, Cartwright, and Spencer 2012; Cartwright and Ryan 2017). Gum from the legume genus Astragalus is usually referred to as tragacanth gum and would have been imported into the region from the Middle East (Newman and Serpico

3

2000, 477). Mannose was present in eight paint samples and indicates the use of a fruit gum. Several fruits have been identified at Amara West, including the sycomore fig, doum palm, white cross berry, Christ’s thorn, Cucumis sp. (a genus that includes melons), colocynth, and watermelon (Ryan, Cartwright, and Spencer 2012). The presence of plant gums in the palettes is further evidence that pigments and binders were being mixed to use as a paint. Plant gums are known to have been used as binders in ancient Egypt, and a range of monosaccharides similar to those identified here was identified from one New Kingdom object and two Third Intermediate Period objects from the Museum of Fine Arts, Boston (Newman and Halpine 2001). The paints were not analysed for proteinaceous binders such as egg and animal glue, so the use of these binders cannot be ruled out.

II ETHNOARCHAEOLOGICAL AND EXPERIENTIAL STUDIES

The second part of this book uses interviews with residents of the areas around Amara West, experiments with materials, and experiences of the landscape, to build a description of the activities around paint collection, production, and application, and how these may have been experienced in the ancient town.

CHAPTER 4

ETHNOARCHAEOLOGY: PERSPECTIVES ON PAINTING AMARA WEST

Ethnoarchaeology is the process of seeking perspectives from modern populations and applying them to those who lived in the distant past. Such information cannot provide absolute answers about ancient people and practices, but it is a ‘flexible interpretative tool’ (Wendrich 2002, 10) that can be used to suggest activities that might otherwise have been opaque for a nonlocal archaeologist. However culturally different the modern and ancient inhabitants may be, they experienced and exploited a broadly similar environment, changed slightly in terms of climate (Woodward et al. 2017) but not geology. Both groups have used the land to find materials for painting homes largely built in mud brick, mud, and plaster with organic roofing. The current inhabitants of the area can also reveal local sources of materials and the practices used to obtain them, which can inform the researcher on ancient experiences and practices. The scientific analysis revealed several pigments (ochres and earths) that were most probably gathered locally from the land around Amara West; red and yellow ochres commonly occur in the area. Similar pigments are still used today by the inhabitants of the areas around Amara West to paint their houses. As part of the investigation into how people may have gathered and processed pigments, ethnoarchaeological interviews were conducted with people who had experience of this in the modern context. Interviews with residents Twelve interviews were conducted in January 2015 with residents in areas near to Amara West, which is itself uninhabited. Interviews took place in three locations: (1) Ernetta, an island in the Nile and location of the a project house, between Amara West and Abri, the local town on the mainland; (2) Amara East, a village east / downstream of Abri, on the opposite bank of the Nile from Amara West; and (3) the large island of Sai, about 11km upstream of Amara West, which has archaeological evidence of inhabitation broadly concurrent with Amara West (Budka 2017). In all three locations, most modern house complexes consisted of an outside wall, within which stood the main house and

several outbuildings, usually built of courses of mud brick (jalus), with some walls painted (Wenzel 1972; Dalton 2017), and a large open courtyard area. The interviews were conducted through a translator. Respondents were approached in the street, or by knocking on the door of their house, and asked if they wished to participate. A brief introduction was given that included the information that I was involved in the project at Amara West, with which most respondents were familiar, that I was a student, and that I wished to talk to them about the way they painted their house. The interviews were, in the main, conducted inside or just outside the interviewees’ homes and varied in length, depending on the availability of the interviewee and the translator. The questions put to the respondents followed a general structure, but varied slightly according to the direction the conversation was taking and the willingness of the respondent to engage in each topic. This interview process was not intended as an in-depth study of modern practices, but rather to inform my understanding of possible practices at ancient Amara West. Interviewees The interviewees have been anonymized in accordance with UCL Research Ethics Committee guidelines (UCL 2015), but are summarized in the below list. – Builder on Ernetta (male, 45) – Older woman, Ernetta (60–70) – Carpenter, Abri (male, 40–50) – Husband and wife, Abri (both about 50) – Farmer and wife, Ernetta (male, 65–70, and female, 50ish) – Farmer’s wife, east of Abri (40ish) – Two women, Amara East (45 and 39) – Young woman, Ernetta (23) – Farmer and wife, Ernetta (male, 35 and female, 30) – Woman, Sai Island, Morkot village (40ish) – Mother and daughter, Sai Island, Morkot village (60 and 25) – Man, Sai Island, Morkot village (50)

CHAPTER

62 Interview questions – – – – – – – – – –

What colours do you use to paint your house? From where do you obtain these colours? Who is responsible for collecting gir? How is the gir prepared? Do you use gum arabic, or another gum? Who is responsible for preparing it? Who does the painting? What tools are used? Do you use gypsum? Do you mix gypsum in with colours? Where are different paints or colours used in the house? Is it possible to employ a decorator? If you need help, who do you ask? How often is the house repainted? What is the learning process for young people?

Colour-related terminology The following Arabic words were frequently used in the interviews: – Bomastic—modern acrylic paint purchased in the market. – Gir—soft rock (clay) that can be gathered and mixed with water for use as paint (not relating to a specific mineral). – Gyps—pure gypsum, purchased at the market, rather than a more general term for white paint. A clear distinction was made between white gir and gypsum. – Sofiha—a large can (Fig. 31). Summary of findings Attitudes towards the collection and use of paint varied, but there were some fairly consistent themes. Gir is still collected from the desert, various sources are known, and either men or women may do this. However, it is the women who plaster and paint the house with gir, and they choose the colours. In contrast, men paint with bomastic purchased from the market. Gir is used to plaster and paint the outer areas of the house (courtyard and walls) (Fig. 29), whereas modern paints (bomastic) are more often used to paint interior walls (Fig. 30). There was some disagreement over whether it was possible or acceptable to pay others to paint; paying a man to paint with bomastic was not controversial, but paying a woman to paint or plaster was deemed by some to be either ridiculous or impossible. If a woman needs help with plastering or painting she can ask neighbours or relatives, and this can be

4

turned into a social event. Girls learn to plaster and paint by gradually helping their mothers in the process from early teens onwards. Repainting is done as required, for example after heavy rain, or every two years as the surface starts to deteriorate. A few people mentioned that they might repaint for a special occasion, such as for the Eid feast. An ethnoarchaeological study of plastering and floor-laying practices in the area around Amara West found that, similarly to decorative processes, housebuilding techniques are widely known and practised by individuals, often in a system of reciprocal labour exchange with friends and neighbours, although there is an increasing trend towards payment of semi-professionals to undertake building work (Dalton 2017). Respondents indicated that, ideally, mud plaster would be replaced for significant religious events, but in reality it seems that the plaster is more likely to be replaced when it begins to fail, which is usually in the time frame of one to two years (Dalton 2017). This finding is in agreement with the responses to my questions on colour. Gir is prepared simply by soaking in water; all interviewees agreed on this. If the gir contains grit it can be sieved. The gir is not ground but it can be encouraged to break apart by stirring. The gir is measured by can (sofiha); four sofiha of gir are needed to paint one house (Fig. 31). Some people added gum arabic, which is now purchased in the market, but it is known that it can be picked off acacia trees also. Modern paintbrushes are now used, but many women remembered their mothers or grandmothers using either a sheep’s tail or a piece of animal skin with the wool still attached. In one house in Morkot village on Sai Island the residents said that if a large area needed painting then the paint was poured over the wall from a teapot. These interviews provided an insight into how a community might perform the tasks of collecting, processing, and applying paints in the landscape around Amara West, and suggest ways in which communities might organize these tasks that would be very difficult to identify in the archaeological record, such as gender divisions, and the use of organic materials for tools. Various methods were identified for acquiring gir (collected by the women, a woman’s husband, purchased from someone who collected too much, collected in a group), which is interesting to think about for the ancient population; their ways of collecting materials may also have been varied, with each procedure involving varying levels of input from different people, and

ETHNOARCHAEOLOGY: PERSPECTIVES ON PAINTING AMARA WEST

Figure 29: House just outside Abri painted in yellow gir with gypsum around the doors. Author’s photograph.

Figure 30: Farmer’s house near Abri. Internal room decorated with bomastic. Author’s photograph.

63

64

CHAPTER

4

Figure 31: One sofiha of gir. The gir is measured by can (sofiha). 1 sofiha = SDG 10 (around GBP 0.20 at January 2020 rates). Four sofiha of gir are needed to paint one modern house. Author’s photograph.

possibly therefore different associations attached to the materials. The disagreement on whether a woman could be paid to paint with gir shows that although societies often have an ‘ideal behaviour’, there are many ways to

subvert this. We are reminded that humans will rarely adhere to strict guidelines when performing tasks: rules are challenged, and practices alter over time.

CHAPTER 5

EXPERIENTIAL ARCHAEOLOGY

During the 2017 season at Amara West, an experiential study of ancient paint was undertaken, guided by the archaeological and scientific analyses, and the information gathered from ethnoarchaeological interviews. The experiences included walking out into the desert to a white rock source that is located close to Amara West, gathering tools for grinding the paint, finding plant gum to use as a binder, and collecting plant materials to make paintbrushes. These materials were then used to make and apply paint. The aim of this study was to think about the holistic process of paint manufacture, from beginning to end, and to try to understand all the stages that might be involved, including intangible elements that cannot be accessed from the archaeological record. Collection of materials Pigments The first material to obtain in order to make paint is the raw pigment itself. Landscapes are lived-in spaces that are known to their inhabitants; humans alter the landscapes around them, so the land is not just imbued with meaning and memories, but is used to create them, and therefore retains them (Edmonds 1999, 9; Hamilakis 2013, 103). These memories can be recent, but can also stretch into the past and be passed down from elder to younger (Fulcher 2019). Group and personal memories combine to give an environment a character and voice of its own. It is part of the story of the people who lived here. It is impossible exactly to replicate the ancient experience of collecting paint materials from the local environment. I do not belong in this desert landscape; my life experience is of growing up in a middle-class family in southeast England, far removed in both time and geography from ancient Upper Nubia. My health is good, my teeth are in good condition, I am well fed; these are not conditions that we can assume for the ancient inhabitants of Amara West (Hummler 2008; Binder and Spencer 2014). In my day-to-day life I personally gather nothing directly from nature to fashion into something useful, not even

firewood, which is delivered. My clothes and shoes are completely different from those of ancient Nubians and Egyptians. My eyesight is corrected by glasses or contact lenses, and I have sunglasses to protect my eyes. When I set out from Amara West across the desert to fetch rock to make pigment, my outlook is vastly different from that of an ancient inhabitant. But having made clear these differences (Hamilakis 2013, 100), I can seek to explain the experience and try to imagine this experience for someone else. The white rock source I visited is 1.1km to the northeast of Amara West (20°49ʹ49 N, 30°23ʹ28 E). It is one of the sources of white rock used today by local inhabitants for painting their houses, although there is no evidence that it was used in ancient times. It took 25 minutes to walk there from Amara West, across the desert. To navigate I used GPS coordinates but originally the location was shown to Matthew Dalton (one of the Amara West team) by a local worker employed as excavator by the project. The rock is soft and can be dug using a small tool such as a piece of schist rock or a sherd of pottery; I used a small trowel. Digging the rock results in a pile of fairly fine powder. The mineral identity of the rock was identified as dolomite using polarized light microscopy (PLM) and Fourier-transform infrared spectroscopy (FTIR). This was not the source of the rock used in the New Kingdom to paint Amara West; this source has not been identified. Facing north from Amara West the land dips into the palaeochannel, which early in the history of the town would have been flowing with water and would have required a boat to cross, but which within one generation or so had become dry for much of the year (Woodward et al. 2017). Beyond this, the land rises to Cemetery D. A person or group heading north would have had to walk either through or around the cemetery. Either way, they see from the town side, and then pass by, three pyramids each standing to a height of 10m. Perhaps more importantly to these people, there would also be the graves of their own ancestors, which they may have visited regularly to lay offerings and seek favours (Binder 2017, 604). A cemetery is a

66

CHAPTER

large modification to the landscape, and almost certainly a place to which strong feelings would have been attached. It must have been more common in ancient Amara West than in modern times to witness the death of a family member, and trips to the cemetery may have been fairly frequent; the cemetery was a familiar place (Stevens 2017). People had to prepare their tombs in advance, and so there may well have been workers in the cemetery; perhaps they called out greetings or shared a joke. Slightly further on there is an outcrop of black volcanic rock which has been inscribed with the name of the scribe of the temple Hatiay (it is unknown if he lived at Amara West) (Spencer, Stevens, and Binder 2014, 22) and longhorned cattle (Fig. 32), the latter carved before the period of Egyptian occupation, and perhaps even prehistoric. Through time, people passed by these rocks and felt the need to decorate them—they are prominent in an otherwise fairly flat landscape. After the cemetery and rocks, there is flat desert. As the people moved away from the town and the cemetery, the noises of life, of animals, of work and play, would have faded away, and been replaced by the quiet of the desert. The travellers may have set out early on their journey, to avoid the heat of the day. In that case, the sun would have only just been rising as they made their way to the top of the escarpment that then drops away into the desert. The air is cool at this time of day, and the wind is low. There are almost no insects. The modern desert from this point on is crisscrossed by tracks made by motor vehicles, and areas formed into little fields that people have tried to farm. In modern times there is almost no vegetation visible, but optically stimulated luminescence (OSL) dates showed that during the New Kingdom some channels of water were flowing seasonally, and where there is water, there are plants. At the time when Amara West was inhabited the landscape would have been less dry and desolate than it is today. The desert is scattered with archaeological remains, including Neolithic and Kerma tombs, rubbish mounds and settlement remains belonging to Dynasty 18 and pre-New Kingdom populations (Spencer, Stevens, and Binder 2014; Stevens and Garnett 2017). The inhabitants of Amara West would have been familiar with this evidence of previous inhabitants; this was not a virgin site, and the area already had a history, with which some of the inhabitants may have identified. During the New Kingdom some of the historic buildings may have still have been standing, providing oases of shelter.

5

The desert bedrock is of a grey-blue schist, which frequently lies in a plane in which the sharp edges are pointing up at an angle off the vertical. Sand lies between these ridges of rock, and piles up against them in small soft dunes. Walking across this landscape is hard on the feet. The rock can be very sharp, which would hurt an unshod foot, or damage a sandal, and the sand can be very soft, which means that two steps forward involve one step back as feet sink into the sand. The sand varies in both texture and colour. It is made up of various types of rock and the eye sometimes picks up on small pieces of bright red and bright green in the otherwise fairly monotonous colours of black and yellow. As the sun climbs in the sky, the colours of the desert become brighter and more assertive: the vivid blue of the sky, often without a cloud in sight, the very dark black of the rock, and the warm yellow of the sand lying between. At around 9am the desert wind starts up. It varies in intensity but certainly buffets clothing and scarves around the body and prevents easy communication. Perhaps the casual conversation that has until now been exchanged among people in the party dies off as they wind scarves around their ears, noses and mouths to prevent the ingress of sand, and to deaden the howling noise. As the wind dies off, the insects arrive. In modern times, in February and March, there is a plague of tiny blackflies called nimiti who come to life for about six weeks and make everybody’s life unpleasant; the nimiti prefer wetter environments, so they may have also have been around in ancient times when the climate was slightly different. Also present would have been the persistent flies, who spy an interesting target in the otherwise relatively barren desert, and come to investigate. At about 11am there is a sudden realization that it is very hot and the sunlight very intense. There is nowhere to hide; there are few trees or buildings, and the rocks are too small to cast a shadow. Perhaps the group followed a flowing channel in order to shelter under the available foliage, or took a rest in the shade of a donkey; or perhaps they did not consider it to be uncomfortably hot, especially if they chose a good winter day and left early. The route to the rock sources would have been known, and younger people would have learnt the route by accompanying more experienced people on the journey. What seems to me to be a featureless desert is, of course, nothing of the sort. There are rock outcrops, and mountains on the horizon to navigate by, and the course of the sun through the sky. On arrival at the source they would have collected what they needed.

EXPERIENTIAL ARCHAEOLOGY

67

Binders

Figure 32: Graffito of a longhorned cow on a rock north of Amara West. Author’s photograph.

Ceramic containers are heavy, so perhaps they used bags, baskets, or scarves. Modern people use large metal cans, thobes (a woman’s wrap-around garment) or scarves. To dig they may have mainly used their hands, or small metal or stone tools. A piece of flat stone picked up in the desert would make an effective shovel and obviate the need to carry a tool from the town and back. The journey home would have been hot. If there was no donkey, the rock would be heavy, and walking uncomfortable. They would have been glad to get home, unload their spoils, and have a cool drink and some food. The experience of fetching the pigment is, then, part of the pigment. And not only this one experience, but the cumulative experience of the community, and each person’s combined individual experiences. A pigment is not just a coloured rock, but a material from a specific place, gathered by a group of people in a social setting, who are repeating a task that has been done many times previously, and has been learnt from those who went before.

The gas chromatography-mass spectrometry (GCMS) analysis of the binders in the Amara West paints indicates the use of a plant gum, and plant gums are the most commonly identified binders in other studies of Egyptian paints. For the experiential study, I used two organic binders. Acacia gum was sourced from Vachellia nilotica trees on Ernetta Island (Royal Botanic Gardens Kew 2016). The gum seeps out where the tree has been damaged. In the trees I sampled this appeared to be from natural causes, or from branches being removed, possibly for firewood. The tree can also be damaged intentionally to encourage the gum to exude. The amount of gum varied hugely, from a large single piece about 5cm across, to thin dribbles on branches which had to be scraped off with a fingernail. Remains of ancient charcoal and seeds at Amara West indicate that Acacia was also available to the inhabitants (Ryan, Cartwright, and Spencer 2012; Cartwright and Ryan 2017), in which case plant gum would have been readily available. However, the gum analysis suggests the use of tragacanth gum, which grows in Turkey, Syria, Iraq and Iran, and would have to have been imported (Newman and Serpico 2000, 477). A chicken egg was also used in the experiential study; chickens were not commonplace until Dynasty 30 at the earliest, but the Egyptians did eat goose, duck, and other birds, from which eggs could have been obtained (MacDonald and Edwards 1993). A small number of bird remains, not yet identified in detail, were excavated from E12.10 and E13.3 at Amara West. An important final component was water. This could have been taken from a storage jar or collected directly from the Nile. Jar-stands and stand-impressions in the mud floor are found in houses at Amara West, and storage jars for water are likely to have been common. Water was mixed with ground pigments to make a paint, and potentially used for washing palettes and grinding equipment. If washing were to take place, the heat of the sun is useful for drying out the grindstones and hammerstones. Processing Tools Alongside the raw materials, a set of tools is also required. A grinding stone of some sort is needed, and this means either sourcing a schist rock from the desert or finding one that has been previously used. The desert

68

CHAPTER

5

Figure 33: Group of stone tools found in deposit 2749 within alcove 2793 in house D11.2.

floor is covered in potentially useful lumps of schist, some of which might be used as grinding stones without working, but most of which would require a least a little alteration. Much of the schist, due to its weathered nature and propensity to laminate, is not suitable for grinding. The wear on some of the grinding stones from Amara West has shaped the rock, which indicates either deliberate working or the result of a long period of use, most likely a combination of the two: an initial forming of the stone and then a gradual smoothing from use. A small grindstone would be useful only for a small amount of pigment, and thus a small painting project. Larger grindstones would be much more difficult to move around or store, and since no workshop has been found and there was unlikely to be a daily requirement for pigment, it is unlikely that a large permanent grindstone just for pigments would have been considered a useful item. Large grindstones with pigment from Amara West have only been excavated from Phase 2B in E13.31, and are part of the paint-related assemblage that suggests an early phase large-scale painting project. However, large grindstones might have had multiple uses, including an occasional use for pigment grinding. Many large grindstones have been found at Amara West that do not retain evidence of pigment processing, but could have been used in this way at some point in their use lives.

Another tool required is a hammerstone. This was probably more easily found; hammerstones should fit the hand and are, therefore, smallish. A great many have been found at Amara West; I have recorded only those which retain evidence of pigment (Table 4). The Nile bank at the local town across the river (Abri) is a shingle beach, from where it is a simple task to pick up various smooth, hand-sized rocks. When Amara West was an island such river pebbles may have been even closer, on the Nile bank next to the town. The small size of the hammerstones means that a person may even have had a favourite that they carried with them or kept at home, such hammerstones presumably having a great number of potential uses. A collection of stone tools (deposit 2749), including 28 handheld stones in various materials and a quartzite quernstone, was found placed inside the front door of house D11.2, possibly a cache of useful tools (Fig. 33). Grinding experiment An experiment was conducted to compare the grinding time required for different pigments. The pigments used in the experiment are listed in Table 12, along with their sources. Fine-grained ochres were found on the surface at Amara West, and correspond well to those from archaeologically secure

69

EXPERIENTIAL ARCHAEOLOGY

contexts. Red sandstone is available locally, and bears a close resemblance to sandstone excavated from Amara West, for example the large lump F13534 from D11.1.1. The sources of gypsum and calcite used by the people of Amara West have not been identified. For this experiment, walnut-sized gypsum nodules were collected from the desert floor, where they occur naturally, and dolomite was dug from a rock source close to Amara West that is exploited by the local people to the present day. Dolomite is very similar to calcite in both appearance and characteristics (Deer, Howie, and Zussman 1992, 641). Two samples of Egyptian blue were used, one a modern pigment manufactured by and purchased from Kremer Pigmente in 2017 and the other an ancient sample from Amara West. A grindstone and a hammerstone were acquired from the 1930s spoil heaps at Amara West. The working area on the grindstone was 15cm by 7cm, comparable with the smaller grindstones from archaeological contexts at Amara West. Approximately equal-size lumps of pigment were selected by eye, each about 3cm across. The pigments in question had very different densities, so weight was not used as a comparative measure. Each pigment was ground in turn on the grindstone until a fine, even powder was created (Fig. 34). The time taken to achieve this was recorded (Table 12). Ochres and sandstones are highly variable in hardness, and this is reflected in the range of grinding times for these rocks. The modern Egyptian blue was very hard and difficult to grind, but the ancient one was extremely easy to crush. The glass component in the ancient sample is completely degraded, which would make it much softer; I did not have a large enough sample of this to get a comparative grinding time. The dolomite was dug as a powder and so required very little grinding.

This was a difficult experiment to control and measure. As I have mentioned, weight is not a good comparative measure for the pigments, but selecting pieces by eye is not an accurate measure either. The effort put into the grinding would depend on my physical condition (including the tiredness of my arm) and other variables such as how hot I felt. And the end point was a purely subjective decision. However, the experiment did highlight the relative efforts of grinding pigments from different sources. Fourteen minutes of grinding (red ochre) is quite a long time given the small amount of pigment that was obtained. I have written in my diary, after grinding six samples, ‘my arm hurts and I’m shaking. More effort than I thought. Imagine doing this for a whole wall!’ The hand holding the hammerstone also aches, and the fingers get knocked and scraped. The position in which one performs this task is close to the ground: sitting, kneeling, or squatting. In any of these positions, the back is bent and the head is bowed over the grindstone. This task is hard work, though may have been more familiar to the ancient inhabitants, for whom a ‘low horizon of activity’ was the norm (Kemp and Stevens 2010a, 507). To have ground pigment in large quantities on the big grindstones in E13.14 would have been a large undertaking, requiring commitment of time and effort. Grinding experience Grinding the raw pigments to create pigment powders initially seems to be a simple task, but there are some important considerations. The location chosen for the grinding is important. There are strong winds across the desert at Amara West; these would have been lesser when the town was still located on an island and the town protected by vegetation, but for most of its history, the desert winds blew straight into the town. In

Table 12: Results of grinding experiment. Pigment

Source

Time taken to grind

Red ochre

Surface find from Amara West

14m 8s

Yellow ochre

Surface find from Amara West

7m 56s

Red sandstone

Dug from nearby rock source

8m 26s

Yellow sandstone

Dug from nearby rock source

12m 4s

Charcoal

Purchased from market

5m 33s

Gypsum nodules

Collected from desert floor

14m 22s

Dolomite (similar to calcite)

Dug from nearby rock source

2m 14s

Egyptian blue (modern)

Purchased from Kremer Pigmente

18m 16s

Egyptian blue (ancient)

Sample from excavation at Amara West (PS481 from D12.8.9)

13s

70

CHAPTER

5

Figure 34: Grinding experiment. Left: Piece of red ochre to be ground on my grindstone. Right: after grinding; pigment has spilled onto the floor and there is no working space left on the grindstone.

even a light breeze the pigment powders will blow away, wasting all the effort that has gone into collecting and grinding the pigment. A sheltered area is best, or a very still day. On a small grindstone, only a small amount of pigment can be ground at one time. If more pigment is added the powder starts falling off the edge of the grindstone, and it becomes impossible to achieve a good particle size because the build-up of powder prevents the particles from being crushed between the grindstone and the hammerstone. Therefore the powdered pigment has to be regularly decanted to another vessel if any sizeable amount is to be ground. The maximum amount that could be ground at one time on my grindstone was a piece of pigment with a diameter of 2–3cm (see Fig. 34), approximately 10g of fine-grained ochre. There is a strong sensory aspect to grinding. The choice of hammerstone is based on how the stone fits in the hand, how the fingers curl around it, and how the weight feels when it is brought down. After only two days of intermittently grinding pigments I had a favourite hammerstone, one that was comfortable to hold. Some are heavier than others; weight can be a good thing because it brings more force down onto the grinding surface, but the rock must be lifted again in order to continue, and this is the work of the arm muscle, which quickly begins to ache. I do not have the muscle built up from years of carrying water, animal husbandry, and tending crops, but if grinding pigment was an occasional task at Amara West, rather than a regular

one, which seems to have been the case, then the muscles used for grinding may not have had the chance to have been developed, and the process may have been hard for ancient people too. Small injuries to the grinding hand would also have been hard to avoid, especially if no callouses had been built up by regular practice. On a hot day, of which there are many at Amara West, the person grinding would have worked up quite a sweat and continued in some discomfort. The sound of grinding is distinctive. Initially there are percussive sounds as the rock is broken down into small pieces, then intermittent scraping, as the rock is dragged across the grindstone, and more percussion, as larger pieces are hammered again and again. Eventually the main sound is a regular scraping; on the large grindstones this would have been continuous, in circles; on my small grindstone it was backwards and forwards in a line. Sounds are very evocative, as are many senses conjuring up memories with other sensory associations (Hamilakis 2013). If there were other activities associated with grinding pigment, such as singing, the action and sounds of the task would have invited in thoughts of these other things, maybe causing the grinder to break into song. It is probably unrealistic to think of one person grinding pigment alone and isolated (Brück 2005). A small town is a busy space, especially within the narrow streets and cramped houses of the walled town. In the Phase 2 pigment production area of E13.31, or in a nearby space if this was a rubbish dump, it is likely that several people were working together, as suggested by

EXPERIENTIAL ARCHAEOLOGY

the presence of more than one large grindstone, in which case they may have been conversing, or exchanging insults, or singing. In later phases there may have been only one person grinding at a time, but they would not have been alone. There would have been children running around, animals bleating, and neighbours passing by. The sound of the grinding may have attracted attention, and joined the other sounds of activities: stone-working, cereal grinding, etc. If the sound of the task evoked memories for the person performing it, it would have done the same for those passing by and hearing it. Some may have passed without comment but continued on their way thinking of the events that had led to the need for pigment, or the last time their family painted their house. Some may have been prompted to make a comment to the person grinding, perhaps a welcome interruption. If pigment grinding was a not a regular task, this is even more likely. People enjoy a break in routine. During a mini ‘workshop’ at the project house at Amara West I showed the other archaeologists staying at the house how I was making paint from raw pigment, and encouraged them to have a go. The local people who work in the house (or who generally help us out) do not usually join in research tasks, but in this case they all came over to give their opinions on the paint, and to use the home-made paint to graffiti the dig house. A task that was out of the ordinary attracted the attention of people who were not obliged to be present. Samir, the boatman, began talking about how all the local people used this paint; he offered opinions and information on the colours, and where they could be found. The tasks initiated comment, advice, and exchange, as they may have also done in ancient times. The grinding of the pigment is potentially a very important part of the development of the paint, not just in terms of materials, but also in terms of the creation of a substance that is applicable for its destined task. The actions performed during the making of the paint are essential to its becoming paint. The creation is not a side issue of little import, but rather the process has its own value; indeed it is the human engagement with the materials that imbues them with their value (Dobres 2001; Pfaffenberger 2001). Thus the identity of the grinder of the pigment must be important. The persons performing the various tasks, including collecting and grinding, must be appropriate in terms of gender, age, social status, and status in relation to the final product (for example the owner of the house if it is wall paint). The ancient Egyptians gave great weight to the spoken and written

71

word, especially in relation to efficacy (Ritner 1993); it is possible that words were spoken or sung to give further potency to the process. Many Egyptian magical and religious ceremonies used props such as amulets or figurines (Ritner 1993; Szpakowska 2003); the combination of speech and action was meaningful to ancient Egyptians on an everyday basis (Meskell 2004), and it is not unreasonable to consider the use of both during the processing of paint to imbue the materials with increased effect, particularly as painted objects such as coffins had important ritual functions. Application The most numerous paint-related finds from Amara West are ceramic palettes that hold paint. These palettes are also known from Egyptian sites, and thus it seems that this was common practice (Kemp and Stevens 2010a; Pagès-Camagna and Raue 2016). Ceramic sherds would have been easy to obtain, and may even have been created for the purpose by deliberate breakage. I searched in the spoil heaps from the 1930s and 1940s EES excavations and found many useful-sized sherds that could be held in the hand, but that retained enough curvature of the original pot to hold a liquid paint. Many of the ceramic palettes holding paint found at Amara West are small, and therefore quite flat; from my experience of mixing paint, these would have been impractical palettes at their current size, and must have been larger when in use, and subsequently broken during discard and burial. The palettes function better when damp because it prevents the water soaking straight into them when it is added to the pigment powder, so they may have been soaked in water before use. Both the dolomite rock from the pit near Amara West and the gypsum from the desert floor needed filtering before they made a good liquid paint. I suspended the ground rock in water, left the heavier particles to settle, poured off the suspension of water and fine particles, and left the fine particles to dry in the sun, creating a fine white pigment. Modern local people told me that they used a similar method or sieved the suspension using scarves or home-made sieves. After this each pigment mixed well with water to make a good white paint. The dry pigments were similar in appearance, but in sunlight the gypsum was more matt and the dolomite slightly more sparkly. For large-scale projects it would have made sense to grind the pigments and decant them into a container

72

CHAPTER

before mixing at the point of use. Binders might also be pre-prepared, for example plant gum ground and stored in a container. This would result in the painter having a vessel for each colour required (possibly six: red, yellow, white, black, green, blue), plus one for water, plus one for binder, plus palette(s). A set of small bags may have held the required pigments, which would have been much easier to carry than a set of ceramic vessels. Plant gum can be ground much like a pigment, and a pinch added as the pigment is mixed. When I asked modern people about binder, they mimed adding plant gum using the pinch of a thumb and forefinger. I did not find any heat was required to dissolve the gum into the water as long as it was not added in large lumps. Moreover, the heat from the sun at Amara West, even in the winter months, is usually sufficient to heat an entire 20 litre plastic bucket of water to about 40°C over the course of a day, so to obtain warm water it would be sensible in this wood-poor environment to leave a pot of water in the sun rather than waste fuel heating it. Egg can simply be stirred together to mix yolk and white, and added directly, with or without water, or the white and yolk separated and used similarly. GC-MS analysis of the Amara West paints from palettes in E13.14 and E13.31 found plant gum in gypsum plaster, black carbon paint, yellow and red ochre paint, and Egyptian blue paint. This suggests that the people of Amara West were adding binders to all colours, although not necessarily in every circumstance; some of the samples tested had no result for sugars, although they could have been present at levels below the detection limit of the instrument. I did not find gum was needed for making paints with whites, ochres, or carbon, as the ground pigment mixed well with water alone and was easy and smooth to paint with, and adhered well to the wall when dry. The only pigment that required a binder was Egyptian blue. The blue crystals in Egyptian blue pigment are large and heavy, and require a binder to hold them in suspension in the paint, and to hold them on the painted surface. Without a binder the particles sink to the bottom of the palette and cannot be picked up by the brush. Both plant gum and egg worked well for this. Gum does not seem to have been added to the ancient paints solely for its ability to carry pigment during painting, because it is used both for blue, which requires a binder, and the finer pigments, which do not. However, there may have been other reasons for adding gum, such as surface finish, or the longevity of paint in a high-use area (such a use for

5

binder was mentioned by modern interviewees). Therefore, we might conclude that gum might have been added for various reasons, including use of the area being painted, and perhaps also personal preference or habit. The availability of gum must also have been a factor, especially if the gum was imported. Experimentation with the application of paint has led me to conclude that a paintbrush would have been a necessary item. Thick mud plaster and white plaster may have been applied by hand, as is done today, but in E13.7.6 and on coffins the paint has been applied in thin stripes and decorative schemes that would have necessitated a more delicate application device. I attempted to paint stripes onto a gypsum plastered mud-brick wall using my finger and found it to be impossible. The water that holds the pigment in suspension is absorbed immediately into the plaster, leaving dry pigment on the finger and preventing the sweeping action that would create a line of paint. In contrast, the bristles of a paintbrush hold the whole suspension and allow the paint to be applied across the surface. No paintbrushes have been found at Amara West, but paintbrushes would have been made from organic materials and there is very little organic preservation at the site, other than degraded wood and basketry in the cemetery. I experimented with various local plant resources to attempt to manufacture a paintbrush. Date palm trees on Ernetta, where the project house is located, are felled and left to dry where they have fallen until there is a requirement for the materials; branches are locally used as fuel for cooking fires. Palm leaf fibres are far too wide and flat to be used as bristles, and the fibres from the palm trunk could be bundled together to create a brush but have very little stiffness and at the same time are too thick to act well as bristles. The most success was had with the fruitbearing branches of the date palm. These can be snapped into shorter lengths, and are made up of many thin strands that naturally hold together well, but at the broken end can be beaten with a hammerstone to create bristles. If necessary, the bristles at the end can be held together using a thin strand pulled off the branch and tied around. It is necessary to soak the strand in water before doing this to increase its flexibility. A brush manufactured in this way holds paint well and is easy to manipulate. Many brushes can be made from one branch, and this is one way in which similar paintbrushes may have been made by the people of Amara West. Archaeobotanical studies of Amara West have shown that doum palm products were used by the

EXPERIENTIAL ARCHAEOLOGY

ancient population (Ryan, Cartwright, and Spencer 2012; Ryan 2016). Even if their brushes were made from other fibres, the process highlights the importance of preparation of the brushes before the use of paint. A tree needs to be felled, or branches removed, a few weeks in advance so that the plant material can dry out. Then the strands need to be snapped apart and soaked in water before the bristles are hammered and the end tied. Although full-time painters would probably have had brushes to hand, there is nothing at Amara West to suggest such a full-time job existed, and brushes may have had to be made on demand. My interviews with modern residents suggested that brushes based on animal parts might also have been an option; goat remains are abundant in the faunal record at Amara West (N. Spencer 2010, 22). Again, these would have taken some preparation, for example the removal of a tail, or the acquisition of a piece of skin with wool attached after an animal had been slaughtered. Soft tissue animal products would be very unlikely to survive in the archaeological record. For the task of painting I found the quality of light to be important. Early in the morning the light is soft and diffuse with little or no shadow. By midday the sun is overhead and the light is very intense, rendering colours difficult to distinguish and casting harsh, sharp shadows. If the area to be painted is outside and requires colours to be mixed and applied in a particular scheme (rather than, for example, a plain white wall), then the early morning would be a much better time to work. During the process of gathering tools, mixing, and painting, the actors would have been unavailable to perform their normal tasks, so the whole process may have been accomplished in a group, with one or more people painting and others forming a support network,

73

preparing food, looking after children and animals, and perhaps taking part in the painting as a social event. Painting may have taken place at particular life events or particular times of year and the painting process could have been integrated with these celebrations. The ethnoarchaeological interviews indicated that it was desirable to schedule redecoration of the house around important events, even if this did not always transpire in reality. The experiential study has demonstrated that the preparation of paints and the execution of painting was not a simple process. Many materials had to be gathered, traded, manufactured, and processed, taking time, effort, and planning. There would have been many people involved, both directly and tangentially, and therefore social interactions, gestures, performances, and actions. The performance of all these actions would have been culturally regulated, including gestures, songs, timings, and the status of the actors within the society. Objects and tasks peripheral to the one focused on here probably included cooking (requiring food, pots, fire, utensils), travelling by donkey or boat, making bags or baskets, producing items to trade, meeting and trading with other people, collecting water, minding animals and children, and cleaning. The task of painting was part of a much wider interconnected taskscape (Ingold 1993) enacted within the habitus of the people of Amara West, situated within their own landscape and society, and the wider cultural world (Fulcher 2018). A ‘taskscape’ can be described as the interlocking ensemble of activities performed by people (Ingold 2011, 195), a concept that discourages the separation of landscape, person, technical task, and socio-cultural life. These things do not exist independently, and should be considered as one organism, each affecting the other.

CHAPTER 6

RECONSTRUCTION OF ANCIENT ACTIVITIES

In this chapter the evidence from Amara West is considered step by step in a series of activities. The evidence presented comes from four strands of research: examination of the archaeological context; scientific analysis; ethnoarchaeological interviews; and the experiential study. The scientific analysis of the paints and study of the archaeological assemblage of colour-related finds at Amara West provided evidence for technical choices made by the ancient inhabitants, such as the selection of minerals and binders, and their uses in specific contexts. The analysis also indicated from where the raw minerals might have been sourced, and thus what journeys might have been made to transport them to Amara West. Archaeological sites provide a huge amount of data about the tangible remains but it can be difficult to interpret these in terms of the human beings who created and used them. Interviews were conducted with the populations local to Amara West, focusing on the collection of painting materials, the method of preparation and application, and the people who performed these actions. The information gleaned from these interviews has fed into the re-creation of the paint production process. The interviews were intended to provide an insight into human considerations, actions and decisions in relation to the painting of houses in the area. A piece of experiential archaeology was conducted to provide further information on actions and technical choices. By going through the process of creating and applying the paint, the range of options from which the technical choices were taken has become clearer. The choices made appear to be guided mainly by sociocultural considerations rather than the materials themselves. The taskscape within which the manufacture and application of paint existed is fleshed out as a range of embedded activities and actions. For each activity the elements of the associated taskscape are shown in a diagram which indicates the possible actors, tools, places, and micro-actions (within the activity) that could have been involved in performing the activity, and lists the technical choices, performance

characteristics, and guidance that may have been required by the actors. ‘Performance characteristics’ are the features an artefact must possess in order to perform its function. The makers start with what they know they want the performance characteristics of the artefact to be (for example, colour, adherence to wall) and make technical choices (for example, choice of pigment, choice of binder) and perform actions (for example, grind) to engineer them into being. ‘Guidance’ consists of the knowledge and skills that people carry around with them that allow them to negotiate their physical and socio-cultural world. One way in which archaeologists have tried to insert people into their archaeological interpretation is by the use of fictional narratives (Spector 1993; Edmonds 1999; Skeates 2010; Tringham 2016), which combine archaeological knowledge with imagined actors and scenarios that are in keeping with historical or archaeological contexts. The fictional narratives are explicitly works of an individual’s imagination and thus declare their position from the outset. To a certain extent, all archaeology is storytelling, since the ‘truth’ can never be known for certain; all archaeological reporting uses a narrative form of some sort, some of which have become institutionalized and are therefore hardly recognized as storytelling (Pluciennik 1999; Majewski 2000; Joyce 2002). Sometimes these narratives are constructed from the basis of a single object or event (Spector 1993; Tringham 2016). These ‘microhistories’ usually have the aim of trying to access an empathetic multi-sensory experience for past people as opposed to the detached, objective approach of scientific description and theory building (Spector 1993; Tringham 2015). Other narratives take a larger or longer view, such as an entire landscape, and try to describe the experience of living in that environment or time period (Edmonds 1999; Skeates 2010). Since I am interested in the sensory and emotional impact of the interactions of people with materials to create paint, I have included short fictional passages (in italics) alongside the steps outlined below to try to illustrate some of the experiential findings, and how these might have played out in ancient Amara West. The aim is to demonstrate how

76

CHAPTER

important the intangible aspects of the production of paint may have been to its creation. The narratives are based on small-scale events, individual actions within the process of making a paint, but it is hoped that they will help to envision the process with all of its wider associations. Critics of archaeological imaginative narratives have pointed out that the device can be taken too far and all sense of the original archaeology lost, or at least the line between fact and fiction irretrievably blurred (Fleming 2006; Van Dyke and Bernbeck 2015). In the hope of avoiding these pitfalls, I place my narratives at the end of sections of archaeologically and scientifically supported results and discussions. This should make clear what there is evidence for, and what has been fabricated by my imagination. I also hope that my stated aim in using these narratives limits the scope for over-interpretation. In addition, I keep them short so that they do not overshadow the results for which there is material evidence. Activity 1: Collection of paint materials To collect the various materials needed to make paint involves travel, acquisition (trade, dig) and transportation for both people and pigment. Each stage requires technical choices to be made (Fig. 35), which include the location selected (distance, ease of access, availability of required material), the method of travel (speed, cost, availability), and the method of acquisition (dig, gather, trade). Yet each of these choices will be heavily influenced by the required performance characteristics of the pigment (see Fig. 35), including the performance of its manufacture. The people who collect the raw materials will be determined by a socio-cultural norm, as might the time of day/year and location of the collecting. There are many sources of iron oxides in the area around Amara West; large outcrops of red and yellow soft rock can be seen from the roadside. For about a generation after it was founded, Amara West sat on an island in the Nile, and no source of coloured rocks are known from the island area, so any journey to collect materials from elsewhere would have necessitated a boat trip. Local sources of yellow and white rock are used by modern people to paint their homes and can be gathered by anyone without any special tools. There are specific areas from where the gir is collected; not all outcrops are equal, and there are gir destinations. Finds of red and yellow rock, sometimes in large lumps

6

(e.g. F13535), are common at Amara West, across all phases, which indicates that this material was not scarce or controlled. However, the rocks vary considerably in quality, from coarse-grained sandstone to finegrained ochre. It is likely that some of the yellow and red coarse-grained sandstones were collected locally, but the fine-grained ochres may have been traded from other areas with better-quality sources. There may have been special locations, both locally and in a wider geographical area, where the rock was thought to be superior, and the origin would have been one of the performance characteristics of that pigment. The meaning of an object is created through its making (Dobres 2001; Pfaffenberger 2001), including its acquisition; the origin of the materials is an important characteristic of the final product. The white pigments from Amara West were used in specific contexts: gypsum was used for plastering walls and coffins, and calcite was found mixed with colours on palettes (whose intended use is unknown) and with gypsum in coffin plaster and paints. Calcite and gypsum can occur naturally together (Mermut and Khademi 2006) but the differential use of the two minerals at Amara West suggests that the sources they were using contained very little co-occurrence. These separate uses strongly indicate that the two minerals were purposefully applied to different contexts. During the experiential work, in which I ground and painted with a dolomite mineral (indistinguishable from calcite to the eye) and a gypsum mineral, I noted that the visible differences between the gypsum and the dolomite, once ground, were very minor to my eyes. Either the inhabitants of Amara West were better able to distinguish the ground pigments visually, or the important difference to them was not in the ground appearance, but rather the mineral itself, or the geographical source. One mineral may have been considered more appropriate for some tasks than the other. However, it may have been the location of the sources that defined their roles, with all the potentially associated experiences: the journey (time, distance, difficulty, people), the position in the landscape, or somewhere decreed ‘correct’. The performance characteristics for the white pigment might have been: white colour; sparkle; competence as plastering material; location and method of procurement, including persons responsible (indicating both feasibility of gathering and specialness); processing required. There is scope here for local traditions to be developed. If the early inhabitants found a calcite source but used it only for special applications, while

RECONSTRUCTION OF ANCIENT ACTIVITIES

77

Figure 35: Summary of people, tools, places, and actions that may be involved in the task of collecting raw materials for paint production. Blue circles list (possible) micro-actions involved; green circles list (possible) environments; grey circles list (possible) actors involved; orange circles list (possible) objects involved. Circles ringed in a thicker line are actions / locations / objects for which evidence exists in the archaeological record.

using the gypsum ubiquitously, this could have been passed down as the norm to the next generation, quickly becoming a standard behaviour. Black pigments at Amara West are either carbon or bitumen. The source of the carbon was most likely from household fires, although it is possible that fires were lit specifically for the creation of black pigment, in which case there would have been a performance element to this stage of acquisition. The majority of the bitumen has been identified as coming from the Dead Sea geological formation, approximately 1,500km north of Amara West. This would have been traded down through Egypt and eventually to Amara West. Bitumen was also being used in the burial process, as evidenced by bitumen in tomb G321 (Fulcher, Stacey, and Spencer 2020). Thus bitumen was acquired not just to use as a black pigment but for wider use in the funerary process. It is significant that bitumen was used as a pigment, because carbon would have been much more easily available. Carbon and bitumen as ground pigments are indistinguishable to my eye, so the performance characteristics that affected their choice were more likely to have been related to the origin, preparation, and associations of the materials. Bitumen was not identified outside of the palettes and graves; it appears that there were strong associations between bitumen and funerary objects and procedures, although the destination of the bitumen paint on the palettes is unknown (but may have included coffins and other funerary equipment). Carbon was identified in both funerary and

non-funerary contexts, so its use was less specialized. Carbon seems to have been a more neutral pigment, for use in any situation, and perhaps this was related partly to the ease of its acquisition. If a material is complex, time-consuming, and costly to acquire, it would probably have a more restricted use and a higher status than an easy-to-find material. This is an example of how the acquisition history of a material could become one of its most important performance characteristics. Blue pigment is found quite frequently at Amara West but in small amounts. Nearly all of it is Egyptian blue, a manufactured pigment, and by far the most commonly identified blue pigment from sites in Egypt across all time periods. There is no evidence for the manufacture of any vitreous pigments at Amara West, green or blue, despite the fact that there was some degree of high-temperature industry at the site in the form of metal-working. The melting points of copper and bronze are around the same as the temperature required to manufacture Egyptian green and Egyptian blue, and the ingredients of Egyptian blue were available (sand, copper, plant ash). This suggests that the part of the process that was unavailable to the people of Amara West, and therefore prevented them from making the synthetic pigments, was an understanding of the manufacturing process. If the manufacture of Egyptian blue was a state-controlled process, as it seems was glass (Shortland 2000; Pusch and Rehren 2007), the people skilled in its creation would have been located at central sites (such as Pi-Ramesse) and

78

CHAPTER

the production process largely unknown outside of these locations. The Egyptian blue is likely to have been imported from elsewhere; unfortunately the glass phase in the Egyptian blue from Amara West was too degraded to analyse for trace elements, which might have indicated its geographical origin. A second blue pigment was found on three ceramic sherds in the Western Suburb. This pigment has been identified as a blue earth, and was probably found locally. There was Egyptian blue still available at this time in Amara West’s history, so it was not a sudden lack of blue at Amara West that led to this blue earth being used. Perhaps access to Egyptian blue was restricted at Amara West, by cost, socio-cultural norms, or the cessation of a previously active trade route, and the blue earth was an attempt to bypass the restriction. In order to obtain certain performance characteristics, others must be compromised, depending on which are more important to the user. In this case it may have been that the desire for a blue pigment led to an attempt to overcome the importance of Egyptian blue and the performance characteristics associated with it, and replace it with a local blue. Alternatively, the blue earth may have been adopted as a second blue pigment, with new performance characteristics that rivalled Egyptian blue, such as its local origin and its lack of connection with Egypt and the Egyptian elite. The later stages of Amara West, after the formal end of the Egyptian occupation of Nubia, saw an increase in the material expression of Nubian identity through funerary practices and material culture (N. Spencer 2014b; Binder 2017). The palettes holding the blue earth are from surface deposits and windblown sand layers that belong to the final use phases of houses D12.7, D11.2 and D11.7. It is possible that the adoption of this pigment could be connected to the re-assertion of Nubian practices at Amara West, although not enough is known about Nubian use of pigments to be able to link use of the blue earth with Nubian material traditions. Green paint is scarce at Amara West, but two green pigments have been identified, on palettes, a grindstone, and loose on the ground (none were found on walls or coffins). These have been identified as a naturally occurring chlorite pigment, probably available in the local area, and a copper chloride that was probably deliberately manufactured, given its rarity in nature (Eastaugh et al. 2004a, 33). The green pigment most widely used in Egypt at the time concurrent with Amara West was the vitreous manufactured pigment Egyptian green, although other green pigments, including copper

6

chlorides, have also been identified in Egypt. No Egyptian green has been found at Amara West. It seems that in the absence of the ‘normal’ or ‘expected’ green pigment, people at Amara West improvised and sourced an alternative. If the source of chlorite was local, much more would have been available than there is evidence for at Amara West, so either the chlorite pigment was ultimately undesirable or its use was restricted. Availability of a pigment does not guarantee its adoption; it may lack certain performance characteristics such the correct colour, sparkle or texture, or it may not have the appropriate associations with place and process that other pigments hold. The manufacture of copper chloride (atacamite type) green pigment may have taken place at Amara West, or elsewhere and the pigment imported to the site. Only small amounts have been found, but its presence on a large grindstone suggests that it was, at one point, used in larger quantities than current evidence suggests. The ingredients for the manufacture of copper chloride pigment are copper, salt, and vinegar, which would not necessarily leave any trace in the archaeological record if used on a small scale. The analysis of the binders in the paints points to a mix of tragacanth gum and fruit gum. Plants currently grown in the area near Amara West (for example on the island of Ernetta) are modern imports and bear very little relation to what would have been present in ancient times, which makes collecting comparable gum samples very difficult. Archaeobotanical studies of Amara West have shown that they had fruit in the form of figs and doum palm (Ryan, Cartwright, and Spencer 2012; Ryan 2016). Acacia wood has been identified from charcoal at Amara West (Cartwright and Ryan 2017), so the gum from Acacia trees (also known as gum arabic) would have been available to the ancient inhabitants. However, tragacanth gum is an exudate of several plants of the genus Astragalus in southwest Asia and the Middle East, and would have to have been imported (Nussinovitch 2010, 52). Modern interviewees stated that they purchased gum arabic in the market, or just used water with no other binder. Two interviewees said that gum could be picked off the Acacia tree, which was probably general knowledge before the arrival of gum in the market. Although it seems unlikely that plant gum would have been imported, we have seen that, despite easy access to charcoal, bitumen was imported from 1,500km away, and resins are known to have been imported in large quantities from the eastern Mediterranean in

RECONSTRUCTION OF ANCIENT ACTIVITIES

Dynasty 18, travelling as far as Buhen in Nubia (Serpico and White 2000; Serpico 2011). If there was a specific requirement for a plant gum that was not available locally, it could have been imported into Amara West. This would suggest that the binder was not merely a practical addition to the paint, but had its own role in creating the meaning of the paint through its making. My experiential work indicated that binder was not really necessary for the finely ground pigments (whites, ochre and charcoal), as these were easily suspended in water and adhered well to the wall without gum. This again suggests that any addition of plant gum into ochres was not for purely practical reasons but that gum had a place in the performance of the creation of paint, including the acquisition of the ingredients. It is much more difficult to find evidence for the way in which the materials were collected. The modern local people travel by truck, donkey, boat, or foot to find gir. The latter three methods were available to the inhabitants of Amara West. It was also mentioned in the interviews that it was possible to pay someone to fetch the gir, and some people made a living collecting gir and selling it on. When the families collected the gir themselves, most of the interviewees identified the man as the one who collected it, but a few said that that either men or women could do this. Some women described how they collected the gir either alone, holding it in their scarf, or in a group to pool resources. Modern residents of the areas around Amara West collect and measure gir in a can called a sofiha. One sofiha can hold about 20kg of gir (see Fig. 31), and four of these are needed to paint one house (the houses are considerably bigger than the ancient houses at Amara West). It would take significant effort to collect and carry 80kg of gir, probably requiring a donkey. Tools used to dig rocks for pigment in ancient times would probably have been pieces of either rock or pottery. Pottery sherds used as tools, as evidenced by their worn edges, are known from Amara West, often in the vicinity of tombs. Pigments that have been identified as imported to Amara West would have been obtained from other people, possibly at markets close to the area, or possibly further away; the journey to market itself may have been an effortful process. A market trade takes place within a taskscape (Ingold 1993). For example, grain is grown, which requires seed, water (irrigation), weeding, harvesting, and processing. The grain is kept in a sack or basket, which has to be manufactured from

79

plant materials, which themselves have to be gathered and processed. The grain is then carried to a market, which is known from established habits, or from word of mouth, both of which are enmeshed in social interactions with a network of people wider than one town. The grain is then exchanged in a socially acceptable manner, which comes from the people being in the same habitus (Bourdieu 1990), at least during the exchange, and a pigment is received, which has its own entire backstory and life history. Even the seemingly simple task of acquiring a pigment at a market has much wider implications for the interconnectivity of tasks and people. One requirement leads to another, almost indefinitely, so that these layers of actions, actors, and materials are completely entangled; pigments are one element of a taskscape that stretches much further than Amara West. The paint materials have performance characteristics at the time of acquisition as much as at any other time in their life history. The source of the materials, the person who collects them, the method of travel and carrying, the time, the weather, could all play a role in the meaningful creation of the paint. It seems that some pigments were widely available, both in terms of having abundant local sources, and being unrestricted to all, whereas others were scarce either through societally enforced restrictions, availability, cost, or preference. The green earth and blue earth pigments demonstrate a desire to bypass these restrictions, or to create a paint with different performance characteristics from those used by the Egyptian elite. It was not just pigments that went into the paint; binders too may have been an important meaningful, as well as practical, component. Paint does not merely have colour; it has meanings embedded in its source and manufacture. I got up earlier than usual this morning because today we are going to fetch paint-rock from the pit in the desert and we need to leave before the sun rises. The younger children are staying with my sister, but I’m taking the older two with me, and it’s Menet’s first time. She’s excited and gets up without complaining, despite the early morning chill. We each take a cloak and then head to the edge of the village where we are meeting Waset with the donkey and her children. Yesterday we filled a waterskin and put some bread aside, and now we tie all this plus several empty baskets onto the donkey, and as the light touches the sky we set off towards the cemetery. In the cemetery we stop for a while to visit our ancestors and leave small food offerings. The children run around

CHAPTER

80

chasing each other. We pass the boulders that mark the edge of our daily environs and look out into the desert. It’s important to start the journey facing the right direction so we take a couple of minutes to discuss the landmarks we need to situate, and then we descend into the desert. As we walk on, the wind picks up and we all wrap our heads in our cloaks to keep the sand out of our eyes and ears. I think about the previous journeys I have made to fetch paint-rock and the events that have led to these painting episodes: marriage, a new house, a new baby. This journey is to collect paint-rock for Waset. She is pregnant and she wishes to repaint the main room in her house. We also thought it was a good time to take the older children to the desert pit to show them the paintrock and how they will have to collect their own painting materials when their turn comes. As we reach the place the wind dies down and we can talk again. The children look around for rocks to use as shovels, and we find some that were used last time and left in a pile at the side of the pit. We sit to eat a little and drink some water while we tell stories of the last trip, who was there, and where we dug. Then we point to where we need to dig. The youngest digs first and fills a basket with the chunks of coloured rock. There are songs to sing while we dig, which remind us of the other times, and of the people who were here with us. Now it is getting hot and digging is hard work, we chide the children and the songs become rhythmic to help us finish the work. When we have filled all of the containers, we sit again, and now the flies have come out and they buzz around us. We head back with the sun getting hotter and hotter and pull our cloaks over our heads. There is no wind at all and we can all talk easily. We point out landmarks we need to know to find our way across the desert. Eventually we can see the pyramids in the cemetery. They grow larger on the horizon and then I can hear the workmen in the cemetery swearing and joking and swinging their hammers. We wearily head into the village, and people greet us on our return. They call out to us or stop for a chat. Some people who couldn’t come on the journey ask if they can have some of the paint-rock in return for a contribution to the party.

Activity 2: Preparation The preparation of pigments and binders had large potential for display and creating meaning, because it took place within the town and would have involved movement, smell, and noise, so that others in the vicinity could have been part of the sensory process even if they were not actively included in the processing (Fig. 36). Grindstones and hammerstones sometimes preserve traces of pigment, and are found throughout

6

the town and suburb. Although grindstones with pigment are infrequently found, it is possible that more were used that no longer have traces of pigment. Modern interviewees prepared gir by soaking in water, and this may have formed part of the process of creating the thin mud plasters on the walls at Amara West, but the evidence of pigment on grindstones indicates that coloured rocks at Amara West were ground. The white minerals used at Amara West are different from the pale white clay used in the area today and must also have been ground because both gypsum and calcite pigments are obtained from rocks, although there is very little direct evidence for grinding white pigments in the archaeological record, and it is possible they were imported pre-ground. The location of the grinding activity would have been important. The wind is strong at Amara West, and any breath of wind blows the pigment away. This would likely have been an indoor activity, done while crouched in a corner. Grinding pigments is an arduous task: the hand and arm ache and the fingers sustain small but cumulative injuries. Excepting the high level of painting activity from Phase 2 in area E13, it appears from the discard of palettes and pigments that the preparation of paint was ad hoc and individual. It is unlikely that any one person had the sole job of painter or pigment grinder. Grinding pigment was not a frequent activity, and so the person undertaking the task would not have developed particular muscles or callouses relating to it. In my grinding experiment I found that to grind a fine-grained 10g piece of red ochre took me 14 minutes. I estimate that 10g of fine-grained ochre, plus water, would be enough to cover a 0.04m2 section of wall. To cover a 4m2 section of wall, for example behind the mastaba in E13.7.6, would therefore require 1kg of ochre, which would take me 23 hours to grind. Although it is, of course, possible (or likely) that a resident of Amara West would have been quicker, this is an indication of the time investment that might be required to paint a wall. In addition, there may have been other elements of pigment preparation, such as floating the pigment to remove coarse impurities, as is done by people who live in the area today, and which would have extended the preparation time further. Some degree of colour mixing took place on the palettes. There are several examples where both ochre, and ochre mixed with a white pigment, are present on the palette, with the mixed paint lying in the centre, indicating that a pinch of white was added to the

RECONSTRUCTION OF ANCIENT ACTIVITIES

81

Figure 36: Summary of people, tools, places, and actions that may be involved in the task of processing raw materials for paint production. Blue circles list (possible) micro-actions involved; green circles list (possible) environments; grey circles list (possible) actors involved; orange circles list (possible) objects involved. Circles ringed in a thicker line are actions / locations / objects for which evidence exists in the archaeological record.

paint and stirred in (see Fig. 20). This suggests that pigment was pre-ground and held in containers before the painting process took place. Gum may also have been pre-ground; I found that lumps of gum took a very long time to dissolve into the water, whereas gum was very easy to grind, and then dissolved much more satisfactorily. The grindstone I used to grind pigment was of a typical size found at Amara West, excepting the large grindstones from Phase 2 in E13. I found only a small amount of pigment could be ground at one time; I needed to regularly decant pigment in order to be able to grind more. The maximum I was able to grind at one time was approximately 10g of ochre. As previously stated, to cover a 2 × 2m wall would require approximately 1kg of ochre. It seems very likely, therefore, that pigments were ground, decanted, and stored in containers before use. The fragments of bright green pigment (PS118) found next to the grindstone in E13.14.1 were in a small pile on the floor, suggesting that they were originally held in a container made of organic material that decayed over time. It is possible that baskets or bags were used to store ground pigment before it was mixed into paint. If this was the case, all of these containers would need to be manufactured prior to the paint-grinding activity, which would involve the collection and processing of plant materials and the construction of the container, or the purchase of the container from a manufacturer. This is another resource that would have to be considered in the

process of making paint, linking paint preparation to the wider taskscapes taking place at Amara West. Given the effort required to grind pigment, and the preparation involved, the grinding process might be viewed as a performance in a social arena. The gestures used to move the hammerstone, the position of the body, and the location of the grinding, would all be learnt and culturally specific. If the task was shared, more than one person would be present, possibly many people. It may have been seen as the opportunity for a social gathering, a chance to share food and drink, to tell stories, or to sing songs. Area E13 in the walled town has evidence for communal arrangements for food preparation, and households may have shared certain activities, and even been interconnected at times (N. Spencer 2015). It is not a great leap, then, to suggest that neighbours may have shared paint preparation. Modern interviewees indicated that girls learnt to plaster and paint from their mothers by joining in with the plastering and painting process from their early teens. We could imagine the grinding of pigments to form part of a learning process for younger members of the family or neighbourhood, perhaps taking turns at grinding, perhaps just watching and listening. This may have been the time that memories were passed on, both collective memories that describe the ways in which paint must be prepared, and why, and more personal ones of previous times that paint has been made, the people that made it, and what happened to them.

82

CHAPTER

The processing of bitumen may have involved particular techniques and possibly tools, all of which would have marked it out as special. Even though when used as a pigment bitumen would have been ground rather than heated, the associations from other uses would be connected to the bitumen, and pass from one sphere (preparation of the deceased) to another (preparation of paint). The efficacy of a material stems from its properties, treatment, and connotations, all of which are intertwined. The potency and effect of bitumen are related to its origin, its handling, its associations with ritual and people who perform those rituals, its ability to melt when heated, its noncarbon-ness, its very materiality. The concept of bitumen as the bodily fluid of Osiris (see Activity 3 below) would have been physically experienced as the bitumen was heated and turned into a liquid and was applied to the body; the materiality of the bitumen helped to activate the metaphor. The behaviour of people around this material was probably very different to the way they used carbon. It is possible that the preparation and application of such a substance would be enacted by a person of appropriate standing, accompanied by ritual sounds and gestures. I have gathered the rocks and tree gum, and the day has come when the rock must be ground to a powder. It will take a whole day, so I have asked the neighbours on both sides to join me, and my sister is here too. We will take turns grinding the powder, watching the children, and cooking. Because we have gathered we are making some special food. I remember my mother making the paint for my sister’s house and the smells of the stew she made, and the taste of the special breads. I suggest we make these foods and my sister agrees but the others have their own special foods that they want to make, so we end up with a lot of food! We find a quiet, sheltered corner of the house to grind the rock, where the wind won’t reach it, and as we finish each grind, we decant the powder into the basket my mother gave to me for holding the powder. We each grind until the ache in our hand becomes unbearable, then the next person takes a turn. As we work alongside each other we chat and reminisce and sing songs that remind us of other times when we have been grinding paint and the events that led to those times and the people involved. People passing the doorway can smell our food and come in to say hello and talk about the reason for the paint and to taste the food. We give each visitor a bread to take with them. Some of them share their memories of special food with us and some give us advice on the grinding, and how much plant gum to add. Menet sits by me as I grind and asks questions. After she has watched

6 for a while she asks to take a turn. I show her how to hold the hammerstone, the amount of rock to use, and the correct way to pound the rock. It quickly starts to hurt her hand so I send her to take water to the animals. When the basket is full we can stop grinding. We sing a song of celebration and relief, cover the basket, and put the ground rock aside for another day.

Activity 3: Application The stage of application is the point at which the paint was probably mixed from the raw materials (Fig. 37). Storing a liquid paint would be space consuming, and difficult in a hot climate owing to evaporation. Before beginning, the raw materials of ground pigment, binder, palettes, water, and paintbrush(es) would need to be gathered. To transport liquid paint from elsewhere would be impractical, especially if the paint were being mixed in shallow ceramic sherd palettes. The selection of the palettes appears to be a utilitarian choice. The palette needs to be of a size to hold in the hand and to support the amount of paint required. Different shapes of pot were used, made of a variety of clays common at Amara West; it seems likely that the palettes would be selected from abundant piles of broken pots, as found in the rubbish deposits excavated beneath house D12.9. As mentioned above, it appears that colours were mixed with white pigments in the palettes, which suggests that the painter would have had pre-ground pigments, and possibly gum, available before beginning to paint. A pinch of coloured pigment would be placed on the palette, and water added to make it into a liquid paint. Gum might be added at this stage, and a pinch of white pigment if required, then the paint quickly mixed and applied to the wall (or other destination). The heat in the middle of the day at Amara West gives quite a short working time because the small amount of water used evaporates; either the person worked quickly, or they chose to work early or late in the day to avoid the heat and use the better light. The ceramic palettes used cannot hold much paint at one time, so this process would have to be repeated often in order to be able to paint a large area. One way to extend the working time of the paint is to pre-soak the palettes before use. This ensures that the water added to the pigment does not immediately soak into the ceramic body of the sherd. My experimentation with painting tools showed that it was not possible to paint with one’s fingers. A paintbrush made from dried and soaked palm fibres worked

RECONSTRUCTION OF ANCIENT ACTIVITIES

83

Figure 37: Summary of people, tools, places, and actions that may be involved in the task of applying paint. Blue circles list (possible) micro-actions involved; green circles list (possible) environments; grey circles list (possible) actors involved; orange circles list (possible) objects involved. Circles ringed in a thicker line are actions / locations / objects for which evidence exists in the archaeological record.

very well and similar paintbrushes are known from Egypt. A paintbrush would probably have formed part of the application toolkit. As the walls were painted, the people painting them would be considering the role of the paint they were using. The colour, the specific pigment used (a choice of whites, for example), the sparkle, the design, the binder, the texture. Some of these choices would have been utilitarian—for example, choice of binder would have been at least partly practical, as it had to hold the paint onto the wall—but many of these choices would have been directed by the socio-cultural performance characteristics of the paint, both during application and with knowledge of its role when in situ. As the paint was applied the space being painted would change. The colour of the walls would obviously change, and the reflection of the light would be altered. The walls would be damp from the fresh paint, which would create the scent of wet mud plaster and moisten the air. These sensations would be reminiscent of other times when painting was done, with associated memories of events, people, and situations. The question of who painted the walls is a difficult one to address. Modern interviewees stated that women painted with gir, although men could help with hardto-reach places; only men painted with modern paints. All of the interviewees concurred with these roles, which were clearly delineated. When repainting the whole house, every few years, the neighbours could help with the mud plastering and the painting. It is not possible to reconstruct such an event at Amara West

from archaeological evidence, but given the amount of work required to paint a whole house (including acquisition of raw materials, grinding, painting, and continuing with everyday tasks), it is not unreasonable to imagine a social event during which several people come together to accomplish the painting as a group, perhaps some painting and others preparing food and performing other daily tasks. Texts from Deir el-Medina mention parties, to which people are listed as bringing an enormous amount of food and drink (Donker van Heel 2016, 137–48). Perhaps a group painting event would have provided an opportunity for such a party. Special circumstances may have required a special type of person to do (or oversee) the painting, for example someone of advanced age, or high social standing, or connected with the temple. Ritual actions and words may have accompanied the application of paint, either religious or non-religious. Texts from the Graeco-Roman Period in Egypt outline the mummification process in terms of rituals performed on the body, and some of the materials used, although these can be difficult to translate with any certainty (Töpfer 2015a). Bitumen (mnnn) is mentioned is in these contexts, forming part of the spoken, and presumably acted, ritual. The texts specify that bitumen comes from the body of Osiris and mention body parts that are to be anointed. It comes to you, who come to you, Osiris NN: It comes to you from Osiris Coptos, the great God in Coptos, in front of the ‘gold house’. He brings you the discharge that comes forth from him, the bitumen (mnnn) that emerges

84

CHAPTER

from his body. He brings you the divine material from Bat / Edfu, as he did for himself Min. Your skin is to be anointed in the Duat (?). Your face is opened in the ways of darkness. (Papyri Boulaq III and Louvre E 5158; Töpfer 2014, 213) Anoint his fingers to the ends: Ꜥnḫ-ỉmỉ-seed, natron, bitumen (mnnn), together with the divine snb-papyrus tied in 36 knots, place into his left hand. (Papyrus Boulaq III; Töpfer 2015b, 249)

Two New Kingdom Theban tombs preserve vignettes of the mummification procedure showing men applying a liquid to the body while a priest stands to the side holding a papyrus and making ceremonial gestures (Dawson 1927, 46–47). Bitumen had strong divine connotations and ritual actions and words associated with its use. Although we have written evidence for this only from later periods, the use of bitumen in purely funerary contexts from earlier periods suggests that this was an association with a long history. It seems likely that the application of a bitumen pigment would have been accompanied by some kind of ritualized actions and possibly words. This task may have been performed by a member of the community with a special status. To accomplish the task of applying paint to the wall the person painting would need to have had the various ingredients and tools to hand. All of these items required pre-preparation and forethought, and were part of the taskscape across Amara West and beyond. The performance of painting would have been culturally learnt: the shape of sherd used, the amount of paint used, the method of adding water. Each action may have varied between individuals, but the habitus in which they lived would have determined the overall method and desired result. I start the painting by applying a layer of white plaster to the whole wall. It is thick so I use my hand. I mix the plaster with some water and plant gum in a large pot and stir it with my hand, then I take a lump and smear it onto the mud-brick wall. I continue until the whole wall is covered, and there are splashes of white plaster over the floor and over me too. My oldest daughter helps me with the lower sections of the wall; it’s her first time so she makes a mess but I quickly smooth it over. This way of applying the plaster takes practice to get it smooth enough to paint on. It doesn’t take long to cover the whole wall. Then we leave it to dry while going about our normal tasks. The whole house smells damp while it is drying and I light the fire earlier than normal in the evening to get

6 rid of the damp feeling. After two days I am sure it is dry and we can paint the colour. I have prepared red and yellow ground rock, and some charcoal that I ground to make black. Blue is not for us. So I’ll stick with the normal colours. My husband mixes the ground charcoal with a little water in a piece of broken pot and paints a thin line of black around the wall. I mix red and water and paint the area above the line. He uses a brush we made from palm last time we dried the palm branches; mine is a bigger brush, made with grasses from the river bank that I gathered when the moon was small and then dried on the roof until the moon was full again. When we are done we invite the family over to admire our work and we share food that they have brought with them. We sing together and play music. We go to bed late, but before we go to sleep we speak to our ancestors and ask that all will be well in this house now we have painted the walls.

Activity 4: Experiencing coloured surfaces The sandblasting effects of the wind in the desert at Amara West have removed a large portion of the paint that once would have been on the walls of the houses in the walled town and Western Suburb. The patches of paint that survive suggest that large areas were at least painted white, and traces of coloured paint indicate that both red and yellow were used, although the extent of their use is difficult to determine. That upper storeys and ceilings do not survive further limits our understanding; ceilings were painted at Amara West. The presence of palettes and pigments across all levels suggests that coloured painting did continue on the site for most of its existence. Due to the low retention of wall paint it is difficult to comment on the function of the colour in the rooms. The varying performance characteristics of paint when it is in situ (Fig. 38) are familiar even to modern people: there are appropriate and inappropriate colours depending on the cultural situation and the function of the room, personal preferences, and also utilitarian considerations such as white paint making a room brighter. One modern interviewee commented that white paint made rooms feel cooler. In ancient Egypt paint was used to indicate the status of the house owner. Elite houses at Amarna were richly decorated with a variety of colours and expensive pigments (Peet and Woolley 1923; Kemp and Stevens 2010a). Workmen’s houses at Amarna were also decorated, but with a more limited palette and in a simpler style, sometimes with line drawings (Kemp 1979). However, the chapel at the Amarna workmen’s village

RECONSTRUCTION OF ANCIENT ACTIVITIES

85

Figure 38: Summary of people, tools, places, and actions that may be involved in experiencing paint. Blue circles list (possible) micro-actions involved; green circles list (possible) environments; grey circles list (possible) actors involved; orange circles list (possible) objects involved. Circles ringed in a thicker line are actions / locations / objects for which evidence exists in the archaeological record.

was highly decorated (Weatherhead and Kemp 2007), indicating the importance of the building and showing the investment of time and resources appropriate to a religious edifice, rather than a domestic one, including the use of a lot of colour. Thus, we could conclude that a house or room with a more intricate and colourful decorative scheme was in some way more important than an undecorated or whitewashed room, and that a decorated feature within that room might hold a particular significance. The niche above the mastaba in E13.7.6 is such a feature at Amara West. The performance characteristics of paint on a painted surface consist of its ability to stay on the surface it has been applied to, and the message it communicates to observers. This message may depend not only on the individual colour of the paint, but also its relation to other paints or architectural features around it, its sparkle, its texture, and communal and personal memories that are connected to the minerals used and their sources, and the people who gathered and prepared the materials, and who painted the wall. The history of the paint, including its origin, who collected it, the processing actions, and the people who performed them, the application procedure, and the final design, all imbue the final painted surface with a potency that has an effect on the observer who inhabits the habitus of the

culture who created the paint. The collective and personal memories connected not just to the life of the paint, but also to its specific final location, determine the effect it has. This effect would have been both shared by the community, and also individual to each person, based on their own history and position within the society and their physical characteristics: the literal and metaphorical angles from which they were viewing the painted surface. The ‘effect’ refers not just to the visual appearance in the standard Western sense, the ‘retinal journey’ (Pallasmaa 1996, 12), but the overall impact on the senses of the audience, which may consist of a variety of clues such as (but not limited to) colour, texture, sparkle, juxtaposition of colours, how much is hidden and revealed, where daylight hits it, and where the owner of the house positions themselves in relation to it. In addition there is knowledge which overlays these clues, which is culturally learnt, such as the ‘right’ pigment to use in a context, and group memories attached to certain sources of pigments, or to people who were associated with the process of creating and using paint. For the first few days after the painting, every time I enter the room I am again surprised at the change in colour and the way it makes the room feel different from before. Then after a while I get used to it. Menet says it makes

CHAPTER

86

her shy of her elders, the room feels more formal. When neighbours come into the room they behave differently from before, they do not sit so casually on the floor but stay by the doorway and wait to be invited in. But I remember how this goes; the paint is fresh and this will last for a little while, yet soon they will be back to their normal selves and gradually the paint will crack and the room will still feel different from before, but not so newly painted. My husband is pleased with the effect the paint has had. Soon it will be somebody else’s turn and we will have the chance to help them and share their food.

Activity 5: Discard The discard of used items associated with painting is the main part of the evidence for paint at Amara West (Fig. 39). Over 900 palette pieces have been excavated in the town and suburb. The high number suggests that palettes were a disposable item. Many palettes are heavily encrusted with paint, several millimetres thick. However, most of the palettes preserving traces of blue in voids in the ceramic body appear to have been scraped or washed clean (see Fig. 26). This indicates that Egyptian blue pigment was far less disposable than yellow, red, and white, and that steps were taken to conserve it. This agrees with the small amount of blue pigment found at the site compared to red and yellow. The sherd with the blue earth pigment was thickly encrusted, which may indicate a relative lack of value ascribed to this pigment. Many small, and some larger, lumps of red and yellow ochre have been found at Amara West. Those in

6

the rubbish fill that makes up the accumulation of painting materials in area E13 during Phase 2 appear to have been purposefully discarded, but some larger lumps of red and yellow, for example in D11.1.1, may have been stored for later use (although finally discarded when the site was abandoned). Although white pigment is frequently found on palettes, both alone (gypsum) and mixed with colours (gypsum and calcite), no discarded lumps of white pigment have been found at Amara West. This suggests that the collection and preparation of white pigments was different to that of red and yellow ochre. If gypsum was imported in chunks, some of these would probably have remained in the archaeological record, as do the red and yellow ochres. It is possible that gypsum was stored in an area that has not been excavated, or it was brought in as pre-ground pigment which has now been lost archaeologically. There is also no evidence of calcite in its raw form at Amara West. Given its restricted use, it may have had a symbolic value higher than that of yellow and red, and so was deemed too special to discard. Alternatively, if it was used only for limited types of painting, not a great deal of it would have been required, and it may all have been processed and used, leaving nothing to be discarded. There also appears to be no discarded bitumen. Bitumen was probably imported over a large distance, and so would have been valuable, as well as having funerary associations that might have made discarding it around the domestic area inappropriate.

Figure 39: Summary of people, tools, places, and actions that may be involved in the discard of paint manufacturing materials. Blue circles list (possible) micro-actions involved; green circles list (possible) environments; grey circles list (possible) actors involved; orange circles list (possible) objects involved. Circles ringed in a thicker line are actions / locations / objects for which evidence exists in the archaeological record.

RECONSTRUCTION OF ANCIENT ACTIVITIES

When we have finished the painting job we are looking forward to relaxing with some food, but first we need to tidy up all the pieces of pigment, palettes, and brushes we have used. When there’s an available dusty corridor we would normally just throw it all into the corner, but in this case the house is new and therefore clean. So we carry what we can to the outskirts of the town and drop it all into the general refuse area. Then we return to sweep the smaller pieces into the street, and wash away what remains with a jug of water.

Activity 6: Maintenance and refurbishment The pieces of painted wall and moulded niche that were found in E13.7.6 show several layers of plaster and paint, evidencing a refurbishment cycle (Fig. 40). The first scheme was simple yellow, black, and white; then there was a polychrome stage; and then everything was painted white. What might these changes represent? Pigments continued to be used to decorate houses and coffins at Amara West, so it seems unlikely that the importance of the paints or colours themselves had diminished. It seems more likely that the status of the room changed, and the decor had to change appropriately. Reasons for the room to change status might be that the space was repurposed, from a main room to a back room, for example, or the status of the owner changed.

87

Maintenance of painted walls can also be observed in other houses, for example E13.3-N and E13.3-S, with successive coats of whitewash. In some cases the choice has been taken to apply a layer of mud plaster over a gypsum plastered wall, and then reapply gypsum; in others, a new layer of gypsum is applied directly on top of the old one. It appears that maintenance was ongoing, although the method was not consistent, which may reflect the importance of the rooms or their inhabitants. The death of my husband has brought many changes into my life, but the most visible one is the changes we have had to make to the house. The brightly coloured mastaba and niche had to be painted white, to display our change in status, and remind visitors that a man of importance no longer lives here. Every time I walk into this room now the impact of recent events becomes more real. The room feels so different, so bare. I have noticed that people behave very differently in this room than they did before: they are subdued and stay for only a short time. There is no focal point, the room is not distinguished from the other rooms in the house. I allow the children to sit in here now and eat sitting on the floor, all formality is gone.

Figure 40: Summary of people, tools, places, and actions that may be involved in the maintenance of painted surfaces. Blue circles list (possible) micro-actions involved; green circles list (possible) environments; grey circles list (possible) actors involved; orange circles list (possible) objects involved. Circles ringed in a thicker line are actions / locations / objects for which evidence exists in the archaeological record.

CONCLUSION

The aim of this research was to use a multi-stranded approach to investigate the use and experience of paint at Amara West. The three strands—materials analysis, ethnoarchaeological interviews, and an experiential approach—combined with the archaeological data from Amara West, were used to create a series of activities by which paint was produced. Some of the tasks, objects, and materials used for these activities could be identified archaeologically or from the materials analysis, and others were suggested by ethnoarchaeological and experiential research. Each part of the production process would have added a layer of significance and value to the paint: the pigment, the provenance, the collection journey, the maker, the performance of grinding, the effort required, the binder (and its own origin), the house in which it was prepared and painted, and the final effect. The combination of these layers may have been manipulated to communicate specific messages about power and identity by the people of Amara West. The value of a painted wall can also be withdrawn, as is evidenced by the whitewashing of the painted niche in E13.7.6. The removal of colour from a room probably represents a change of function or status of the room or owner, a demotion of distinctiveness. This study, of often unimpressive and fragmentary remains, has provided a rich, experiential perspective on how paint was interwoven into ancient Nile valley life, a view that could not be achieved through the analytically focused study of high-elite painted objects in museums. The multi-stranded nature of this research, combining scientific analysis with a phenomenological approach, has led to an in-depth yet nuanced consideration of the paints from the ancient site of Amara West. The actions undertaken to make paint were not detached from other elements of the ancient population’s lives, but sat within a wide taskscape in which all activities,

people, and goals were interlinked. The importance of some specific materials has been proposed, where they have been applied in only certain contexts, suggesting the use of ritual behaviours and performances in their use. Ritual is a habituated behaviour and may have come into play for the preparation of all the paints, a culturally learnt act. Significance and value can be built up using materials chosen from a special place or obtained via a particular journey; ritualized actions then imbue the material with meaning while altering it in a specific way, and people bring their status, memories, and knowledge to the material and create a product that retains an association with their input, recognizable to those who are within the society and can read the paint. The technological choices made by the people at Amara West were led by cultural forces and the personal desire, formed within their habitus, to embellish living spaces and funerary equipment, and possibly other things that have not survived in the archaeological record, such as their own skin. The format of painted decoration on house walls, and elsewhere, would have been regulated by cultural norms, learnt through the experience of living in the community. That community would have encompassed inhabitants with varying affiliations, to Egyptian but also to Nubian culture. The processing and use of paint would have formed part of their habitus and taskscape, with their ingrained and learnt knowledge of correct behaviour, use of the environment, manipulation of space, gestures, performance, treatment of materials, time management, resource allocation, and personal interactions, both in their local environment, and the wider ancient world. It is hoped that further research of this type could be conducted at other urban sites in the Nile valley; only then would emerge potential distinctions between different communities and their approach to colour.

Appendix

EXPERIMENTAL DETAILS

X-ray fluorescence (XRF) XRF analysis was conducted at the project house in Sudan, using an Olympus Innov-X Delta Premium portable handheld XRF (pXRF). The Delta Premium has a 4W tube and rhodium anode. The mode used was 2 Beam Mining, has a live time of 32 seconds per beam and a beam size of 25 mm2 at the port. The elements searched for using this mode were: Mg, Al, Si, P, S, Cl, K, Ca, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn, W, As, Pb, Bi, Zr, Mo, Ag, Cd, Sn, Sb. The software used was the InnovX data acquisition and processing software supplied with the instrument. The operating manual for the instrument can be found here: https://usenvironmental.com/download/manuals/ Olympus - Delta User Manual.pdf Polarized light microscopy (PLM) The sample was mounted on a glass microscope slide using Meltmount, a material of known refractive index (1.662), and was then examined in transmitted light using an Olympus BH-2 polarizing light microscope, at up to 400× magnification. This method allowed the various components present in a pigment to be observed. Particles were assessed in both plane polarized light and crossed-polarized light. Features observed in plane polarized light included colour, shape, size, refractive index, relief, and pleochroism (Eastaugh et al. 2004b). Those observed in crossedpolarized light include isotropism, birefringence, and extinction (Eastaugh et al. 2004b). Fourier-transform infrared spectroscopy (FTIR) Attenuated total reflectance spectra were taken using the Smart iTR diamond accessory on the Nicolet 6700 main bench with a DLaGTS detector. Spectra were acquired over a range of 4000–550 cm-1 using 32 scans at a resolution of 4cm-1 and automatic gain, and were identified by comparison with in-built databases and British Museum reference materials. Before each sample was analysed, a background analysis was run to

remove the effects on the final spectrum of atmospheric water vapour, CO2 and the internal components of the instrument. The instrument used for this analysis operates only in the mid-IR range (4000–550 cm-1). Thus, it was unable to detect distinctive peaks for haematite, which occur below 550cm-1 (Genestar and Pons 2005). Peaks for clay-type minerals, such as kaolinite, were taken as an indication of ochres, in combination with a lack of peaks for other red pigments. For this research various libraries of FTIR spectra were consulted: a library of spectra of conservation and paint materials compiled by the University of Tartu (Estonia) analysed in ATR-FTIR (Vahur et al. 2016); the spectra available from the Infrared and Raman Users Group’s Spectral Database (Infrared & Raman Users Group 2016), submitted by various institutions; and the open source RRUFF™ Project (Lafuente et al. 2015); as well as reference spectra run for this study from the British Museum Reference Collection. Scanning electron microscope (SEM) The instrument used for all SEM analysis was a Hitachi S-3400N with an Oxford Instruments Inca Spectrometer and associated software. The instrument was calibrated with standards. EDS analyses were performed at 20 kV, 10 mm working distance and 120 seconds counting time. All samples were analysed for all elements. Some samples were prepared in polished resin blocks and carbon coated to obtain quantitative results; some samples were too small to do this and these were analysed as powders adhering to a sticky pad, also carbon coated; some paints were analysed in situ on their ceramic palettes, and these could not be carbon coated. The latter samples were analysed in variable pressure mode (some air) to avoid charge collecting on the objects. Two white samples PS125 and PS128 were prepared in polished resin blocks and carbon coated thickness 15.0 nm, density 2.25 g/cm3. Three points were analysed for each sample. These results are quantitative.

APPENDIX

92

These two samples were chosen for SEM-EDS analysis because they contained a high proportion of calcite (calcium carbonate CaCO3); this analysis was used to confirm whether they were dolomite (calcium magnesium carbonate, CaMg(CO3)2), which is indistinguishable from calcite in PLM but can be detected by elemental analysis because it contains magnesium. On the palettes PS125 and PS128, magnesium was not found in sufficient quantities to indicate the presence of dolomite

rather than calcite. The sample of gir (PS590) that was collected in 2015 from the gir pit near Amara West was not homogeneous and gave differing results depending on the site analysed. Some areas had high levels of calcium and sulphur, in a 1:1 atomic % ratio indicating calcium sulphate CaSO4 (gypsum or anhydrite), whereas other areas had high levels of calcium and magnesium in a 1:1 atomic % ratio, indicating dolomite.

Appendix Table 1: SEM-EDS results for PS125, pink palette, 98% calcite. Element

Weight%

Atomic%

Element

Weight%

Atomic%

Element

Weight%

Atomic%

Na

0.6

0.7

Mg

0.6

0.7

Na

0.7

0.6

Mg

0.5

0.6

Si

1.2

1.2

Al

9.2

7.2

Al

1.3

1.3

S

0.2

0.1

Si

31.8

23.9

Si

2.7

2.6

Cl

0.3

0.2

Cl

0.1

0.1

Cl

0.4

0.3

Ca

68.0

46.9

K

11.2

6.1

K

0.3

0.2

Fe

0.4

0.2

Fe

0.2

0.1

Ca

63.0

42.6

O

29.3

50.6

O

46.9

62.1

Fe

0.9

0.4

O

30.3

51.3

Appendix Table 2: SEM-EDS results for PS128, yellow palette 98% calcite. Element

Weight%

Atomic%

Element

Weight%

Atomic%

Element

Weight%

Atomic%

Mg

0.3

0.3

Mg

0.4

0.5

Mg

0.5

0.6

Ca

71.2

49.7

Ca

71.0

49.5

Si

0.3

0.3

O

28.6

50.0

O

28.6

50.0

Ca

70.3

48.9

O

28.8

50.2

Appendix Table 3: SEM-EDS results for PS590, gir from a site near Amara West (still in use by modern residents of the area). Element

Weight %

Atomic %

Element

Weight %

Atomic %

Element

Weight %

Atomic %

Na

0.6

0.6

Si

0.2

0.1

Na

2.3

2.4

Al

0.2

0.2

S

23.1

16.4

Mg

22.0

21.8

Si

0.5

0.4

Cl

0.4

0.2

Si

0.8

0.7

S

22.8

16.2

Ca

29.7

16.9

S

0.2

0.1

66.4

Ca

41.6

25.0

O

33.1

49.9

Cl

0.4

0.2

Ca

28.8

16.3

O

46.7

66.1

O

46.7

93

EXPERIMENTAL DETAILS

The yellow and red paints analysed by SEM-EDS were on palettes. The paints were analysed on the palettes using variable pressure, uncoated. This analysis is not quantitative and was used to check for presence and absence of elements. Three points were analysed for each sample. The yellow paints contained a high amount of iron, supporting the identification of a yellow iron oxide, and no arsenic, the presence of which would indicate orpiment. The SEM-EDS analysis found no arsenic, lead, or mercury, which would have suggested realgar, read lead and cinnabar, respectively. The main elements present were iron, silicon and aluminium, in the approximate atomic % ratio 1:5:2 (PS449, PS129, PS447 (orange-red)) or 1:3:1 (PS446, PS451, PS435). Sulphur was detected where gypsum was present, and calcium where gypsum and/or calcite

were present (e.g. PS131). PS447 (dark red) had a much higher level of iron than the other samples, and was noticeably a darker red – this was a purer iron oxide. Egyptian blue samples were prepared in polished resin blocks and carbon coated thickness 15.0 nm, density 2.25 g/cm3. These results are quantitative. Three points were analysed for each sample. When observed in back scattered electron (BSE) mode the Egyptian blue samples had three distinct areas: pale grey areas (calcium copper silicate), dark grey (glass / Si), and bright spots (high tin content). A typical result is shown in Appendix Table 4. The blue from palette PS539 was analysed as a powder on a sticky pad, carbon coated. Five areas were analysed. These results are not quantitative.

Appendix Table 4: SEM-EDS results for PS315 (F6045), raw pigment. pale grey area Element

Atomic%

Si

25.2

Ca

5.9

pale grey area Ratio to Cu

Element

Atomic%

4.2

Si

25.2

1.0

Ca

6.1

pale grey area Ratio to Cu

Element

Atomic%

Ratio to Cu

4.2

Si

25.3

4.1

1.0

Ca

6.0

1.0

Cu

6.0

1.0

Cu

6.1

1.0

Cu

6.1

1.0

O

62.7

10.5

O

62.6

10.3

O

62.7

10.2

Green samples were mounted on sticky pads and carbon coated. Five points were analysed for PS118 and five points for PS506. Two areas were analysed for PS126. These results are not quantitative.

and the step size (2θ) was 0.02°, with a nominal wavelength of 1.5402Å. Scan range and integration time varied and are stated in the results for each sample. Interpretation was accomplished using the Bruker EVA package coupled with the PDF-2 diffraction pattern database.

X-ray diffraction (XRD) The samples were hand-ground to a fine powder in an agate pestle and mortar. The powdered sample was then loaded into a Bruker ‘zero height zero background’ sample holder (a polished silcon wafer) for analysis. XRD analysis was performed using a Bruker D2 Phaser compact X-ray diffractometer located in the Forensic Laboratories of Anglia Ruskin University (Cambridge, UK). This instrument was fitted with a water-cooled 2.2 kW Cu anode X-ray tube operated at 300 W (30 kV, 10 mA), a Ni β-filter and a Bruker LynxEye™ silicon strip position-sensitive detector (operated in θ–2θ scan mode). The sample rotation was 15 r.p.m.

Gas chromatography-mass spectrometry (GC-MS) Bitumen analysis Samples were dissolved in 1ml dichloromethane (DCM), and heated at 60°C for 1 hour, after which the solution was decanted, and dried under a stream of nitrogen. 20μl DCM and 1ml hexane were added to the soluble fraction, the asphaltene fraction precipitated out, and this was left overnight to settle. The solution was then decanted and dried under a stream of nitrogen to obtain the maltene fraction. The addition of 20μl DCM and 1ml hexane was repeated twice more,

94

APPENDIX

each time leaving overnight and then decanting and drying in nitrogen. Each maltene fraction was then fractionated using column chromatography. 100μl hexane was added to the samples. Each was decanted into a glass pipette held upright and plugged with glass wool and half filled with dried silica (chromatography grade 60–120μm, pre-extracted with DCM/methanol 97:3, followed by hexane, then oven dried) to which hexane was added to exclude moisture. The first fraction was extracted using 3ml hexane washed through the pipette; the second using 3ml DCM:hexane 1:3; the third using 3ml DCM:methanol 2:1. Each fraction was collected at the base of the pipette and dried under nitrogen. For analysis, 50μl of hexane was added to the first elute. The GC-MS analysis was carried out with an Agilent HP5-MS column (30m × 0.25mm, 0.25μm film thickness) with splitless injection, coupled to an Agilent 5973 MSD. The mass spectrometer was operating in the electron impact (EI) mode at 70 eV and scanning m/z 50 to 550 amu. The oven was set at 60°C to 290°C at 4°C/min with the final temperature held for 30.5 mins. GC-MS analysis was run in two modes: scan and Selective Ion Monitoring (SIM). Acquisition in SIM mode targeted ions: 177, 191, 217, 218, 259.

Gums analysis Samples were hydrolysed by the addition of 500μl of 0.5M hydrochloric acid and heated at 80°C for 20 hours. The solution was decanted and dried under nitrogen. Samples were derivatized by the addition of 300μl Sigma-Sil A (1:3:9 ratio of trimethylchlorosilane (TMCS), hexamethyldisilazane (HMDS) and pyridine), and heated at 80°C for 2 hours. Samples were dried under nitrogen and dissolved in 100 μl hexane in preparation for injection into the GC-MS instrument. A blank and three reference samples were prepared alongside the samples using the same method. Samples were analysed using an Agilent HP5-MS column (30m × 0.25mm, 0.25μm film thickness with 1m × 0.53mm retention gap) with splitless injection at 300°C and 10.1 psi and a purge time of 0.5 mins. The carrier gas was helium with a flow at 1.5ml/min. The oven was set at 40°C to 130°C at 9°C/min, then to 290°C at 2°C/min, with the final temperature held for 10 mins. The mass spectrometer zone temperatures were 230°C at source and 280°C at interface. The solvent delay was set at 5 mins. Acquisition was in scan mode over 29–650 amu.

BIBLIOGRAPHY

Accorsi, G., G. Verri, M. Bolognesi, N. Armaroli, C. Clementi, C. Miliani, and A. Romani. 2009. The exceptional near-infrared luminescence properties of cuprorivaite (Egyptian blue). Chemical Communications 23: 3392–94. doi: 10.1039/B902563D Aliatis, I., D. Bersani, E. Campani, A. Casoli, P. P. Lottici, S. Mantovan, I.-G. Marino, and F. Ospitali. 2009. Green pigments of the Pompeian artists’ palette. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 73 (3): 532–38. doi: 10.1016/j.saa.2008.11.009 Ambers, J., 2004. Raman analysis of pigments from the Egyptian Old Kingdom. Journal of Raman Spectroscopy 35 (8–9): 768–73. doi: 10.1002/jrs.1187 Ambers, J. 2008. Pigments. In: A. Middleton and K. Uprichard (eds), The Nebamun wall paintings: Conservation, scientific analysis and display at the British Museum. London: Archetype in association with the British Museum, 31–40. Ambers, J., R. Stacey and J. H. Taylor. 2007. Writing that cannot be erased: Investigations of a box of pigmented inlays from the tomb chapel of an Old Kingdom noble. British Museum Technical Research Bulletin 1: 49–54. Anderson, J. and S. M. Ahmed. 2011. Dangeil 2010: Meroitic wall paintings unearthed and conservation strategies considered. Sudan & Nubia 15: 80–89. Anderson, D. M. W. and K. A. Karamalla. 1966. Studies on uronic acid material. Part XVI. Inter-nodule variation and the acidic components in Acacia nilotica gum. Carbohydrate Research 2 (5): 403–10. doi: 10.1016/S00086215(00)80334-1 Andreu, G. and C. Barbotin. 2002. Les artistes de Pharaon: Deir el-Médineh et la Vallée des Rois. Paris: Réunion des musées nationaux. Baines, J. 1985. Color terminology and color classification: Ancient Egyptian color terminology and polychromy. American Anthropologist 87 (2): 282–97. doi: 10.1525/ aa.1985.87.2.02a00030 Barakat, A. O., A. Mostafa, Y. Qian, M. Kim, and M. C. Kennicutt. 2005. Organic geochemistry indicates Gebel El Zeit, Gulf of Suez, is a source of bitumen used in some Egyptian mummies. Geoarchaeology 20 (3): 211– 28. doi: 10.1002/gea.20044 Bass, G. F. 1986. A Bronze Age shipwreck at Ulu Burun (Kaş): 1984 campaign. American Journal of Archaeology 90: 269–96. doi: 10.2307/505687

Baylor, D. 1995. Color mechanisms of the eye. In: T. Lamb and J. Bourriau (eds), Colour: Art and science. Cambridge: Cambridge University Press, 103–26. Berlin, B. and P. Kay. 1969. Basic color terms: Their universality and evolution. Berkeley, CA; London: University of California Press. Berry, M. 1999. A study of pigments from a Roman Egyptian shrine. Australian Institute for the Conservation of Cultural Material (AICCM) Bulletin 24 (1): 1–9. doi: 10.1179/bac.1999.24.1.002 Bietak, M. 2005. Egypt and the Aegean: Cultural convergence in a Thutmoside palace at Avaris. In: C. H. Roehrig, R. Dreyfus, and C. A. Keller (eds), Hatshepsut: From queen to pharaoh. New York; New Haven, CT: Metropolitan Museum of Art; Yale University Press, 75–81. Bietak, M. 2013. The palatial precinct at the Nile branch (Area H). http://www.auaris.at/html/ez_helmi_en. html#8 [accessed January 2020]. Bietak, M. and I. Forstner-Müller. 2003. Ausgrabungen im Palastbezirk von Avaris. Vorbericht Tell el-Dab‘a/ ‘Ezbet Helmi Frühjahr 2003. Ägypten und Levante 13: 39–50. Bietak, M., I. Forstner-Müller, F. van Koppen, and K. Radner. 2009. Der Hyksospalast bei Tell el-Dab‘a. Zweite und Dritte Grabungskampagne (Frühling 2008 und Frühling 2009). Ägypten und Levante 19: 91–119. Bietak, M., N. Marinatos, and C. Palivou. 2007. Taureador scenes in Tell el-Dab‘a (Avaris) and Knossos. Vienna: Verlag der Österreichischen Akademie der Wissenschaften. Bietak, M., C. von Rüden, J. Becker, J. Jungfleisch, L. Morgan, and E. Peintner. 2012/13. Preliminary report on the Tell el-Dab‘a Wall Painting Project: Season 2011/12. Ägypten und Levante 22/23: 131–47. Bikiaris, D., Sister Daniilia, S. Sotiropoulou, O. Katsimbiri, E. Pavlidou, A. P. Moutsatsou, and Y. Chryssoulakis. 2000. Ochre-differentiation through micro-Raman and micro-FTIR spectroscopies: Application on wall paintings at Meteora and Mount Athos, Greece. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 56 (1): 3–18. doi: 10.1016/S1386-1425(99)00134-1 Binder, M., 2011. The 10th–9th century BC—new evidence from Cemetery C of Amara West. Sudan & Nubia 15: 39–53.

96

BIBLIOGRAPHY

Binder, M. 2017. The New Kingdom tombs at Amara West: Funerary perspectives on Nubian–Egyptian interactions. In: Spencer, Stevens, and Binder 2017, 589–611. Binder, M. and N. Spencer. 2014. The bioarchaeology of Amara West in Nubia: Investigating the impacts of political, cultural and environmental change on health and diet. In: A. Fletcher, D. Antoine, and J. D. Hill (eds), Regarding the dead: Human remains in the British Museum. London: British Museum, 123–36. Binder, M., N. Spencer, and M. Millet. 2010. Cemetery D at Amara West: The Ramesside Period and its aftermath. Sudan & Nubia 14: 25–44. Bleton, J., P. Mejanelle, J. Sansoulet, S. Goursaud, and A. Tchapla. 1996. Characterization of neutral sugars and uronic acids after methanolysis and trimethylsilylation for recognition of plant gums. Journal of Chromatography A 720 (1–2): 27–49. doi: 10.1016/00219673(95)00308-8 Blom-Böer, I., 1994. Zusammensetzung altägyptischer Farbpigmente und ihre Herkunftslagerstatten in Zeit und Raum. Oudheidkundige mededelingen uit het Rijksmuseum van Oudheden te Leiden 74: 55–107. Bonizzoni, L., S. Bruni, V. Guglielmi, M. Milazzo, and O. Neri. 2011. Field and laboratory multi-technique analysis of pigments and organic painting media from an Egyptian coffin (26th Dynasty). Archaeometry 53 (6): 1212–30. doi: 10.1111/j.1475-4754.2011. 00592.x Bonnet, C. and D. Valbelle. 2000. Édifices et rites funéraires à Kerma. Paris: Errance. Bourdieu, P. 1990. The logic of practice. Cambridge: Polity. Bradley, R. 2003. Painted plaster murals from Meroe townsite. Sudan & Nubia 7: 66–70. Brück, J. 2005. Experiencing the past? The development of a phenomenological archaeology in British prehistory. Archaeological Dialogues 12 (1): 45–72. doi: 10.1017/ S1380203805001583 Bruyère, B. 1939. Rapport sur les fouilles de Deir el-Médineh (1934–1935) 3me partie, Le village, les décharges publiques, la station de repos du col de la Vallée des Rois. Cairo: l’Institut français d’archéologie orientale. Bruyère, B. 1952. Tombes thébaines de Deir el-Medineh à décoration monochrome. Cairo: Institut français d’archéologie orientale. Budka, J. 2017. Life in the New Kingdom town of Sai Island: Some new perspectives. In: Spencer, Stevens, and Binder 2017, 429–48. Budka, J. 2020. AcrossBorders II: Living in New Kingdom Sai. With contributions by Johannes Auenmüller, Annette M. Hansen, Frits B. J. Heinrich, Veronica Hinterhuber, Ptolemaios Paxinos, Nadja Pöllath, Helmut Sattmann, Sara Schnedl and Martina Ullmann. Vienna: Austrian Academy of Science Press.

Buzon, M. R. 2006. Biological and ethnic identity in New Kingdom Nubia: A case study from Tombos. Current Anthropology 47 (4): 683–95. doi: 10.1086/506288 Caley, E. R. 1946. Ancient Greek pigments. Journal of Chemical Education 23 (7): 314–16. doi: 10.1021/ ed023p314 Cartwright, C. and A. Middleton. 2008. Scientific aspects of ancient faces: Mummy portraits from Egypt. British Museum Technical Research Bulletin 2: 59–66. Cartwright, C. R. and P. Ryan. 2017. Archaeobotanical research at Amara West in New Kingdom Nubia. In: Spencer, Stevens, and Binder 2017, 271–286. Ciampini, E. M., 2018. The Italian Archaeological Mission in Sudan—Jebel Barkal, Season 2017–2018. The Royal District of Natakamani at Napata. https://unive.academia. edu/MissionJebelBarkal [accessed September 2020]. Colinart, S. 2001. Analysis of inorganic yellow colour in ancient Egyptian painting. In: Davies 2001, 1–4. Colinart, S., E. Delange, and S. Pagès. 1996. Couleurs et pigments de la peinture de l’Égypte ancienne. Techne 4: 29–45. Connan, J. 1999. Use and trade of bitumen in antiquity and prehistory: Molecular archaeology reveals secrets of past civilizations. Philosophical Transactions of the Royal Society Series B: Biological Sciences 354 (1379): 33–50. doi: 10.1098/rstb.1999.0358 Connan, J., O. P. Nieuwenhuyse, A. Van As, and L. Jacobs. 2004. Bitumen in early ceramic art: Bitumen-painted ceramics from Late Neolithic Tell Sabi Abyad (Syria). Archaeometry 46 (1): 115–24. doi: 10.1111/ j.14754754.2004.00147.x Connan, J., A. Nissenbaum, and D. Dessort. 1992. Molecular archaeology: Export of Dead Sea asphalt to Canaan and Egypt in the Chalcolithic–Early Bronze Age (4th–3rd millennium BC). Geochimica et Cosmochimica Acta 56 (7): 2743–59. doi: 10.1016/0016-7037(92)90357-O Cooney, K. M. 2012. Apprenticeship and figured ostraca from the ancient Egyptian village of Deir el-Medina. In: W. Wendrich (ed.), Archaeology and apprenticeship: Body knowledge, identity, and communities of practice. Tucson, AZ: University of Arizona Press, 145–70. Corcoran, L. 2016. The color blue as an animator in ancient Egyptian art. In: R. Goldman (ed.), Essays in global color history: Interpreting the ancient spectrum. Piscataway, NJ: Gorgias Press, 59-82. Correas-Amador, M., 2013. Ethnoarchaeology of Egyptian mudbrick houses: towards a holistic understanding of ancient Egyptian domestic architecture. PhD Thesis. Durham University. http://etheses.dur.ac.uk/6916/ Dalton, M. 2017. Reconstructing lived experiences of domestic space at Amara West: Some preliminary interpretations of ancient floor deposits using ethnoarchaeological and micromorphological analyses. In: Spencer, Stevens, and Binder 2017, 357–88.

BIBLIOGRAPHY

Dalton, M. 2020. Reconstructing the use and conception of pharaonic domestic space in Nubia: Geoarchaeological investigations at Amara West (~1300–1070 BC). PhD thesis. University of Cambridge. Daniels, V. 2007. Analyses of copper- and beeswaxcontaining green paint on Egyptian antiquities. Studies in Conservation 52 (1): 13–18. doi: 10.1179/sic.2007. 52.1.13 Daniels, V., T. Devièse, M. Hacke, and C. Higgitt. 2014. Technological insights into madder pigment production in antiquity. British Museum Technical Research Bulletin 8: 13–28. Daniels, V. and B. Leach. 2004. The occurrence and alteration of realgar on ancient Egyptian papyri. Studies in Conservation 49 (2): 73–84. doi: 10.1179/sic.2004. 49.2.73 Daniels, V., R. Stacey, and A. Middleton. 2004. The blackening of paint containing Egyptian blue. Studies in Conservation 49 (4): 217–30. doi: 10.1179/ sic.2004.49.4.217 David, R., H. G. M. Edwards, D. W. Farwell, and D. L. A. De Faria. 2001. Raman spectroscopic analysis of ancient Egyptian pigments. Archaeometry 43 (4): 461–73. doi: 10.1111/1475-4754.00029 Davies, W. V., ed. 2001. Colour and painting in ancient Egypt. London: British Museum Press. Davies, W. V. 2003. Kouch en Égypte: Une nouvelle inscription historique à El-Kab. Bulletin de la Société française d’égyptologie 157: 38–44. Davies, W. V. 2017. Nubia in the New Kingdom: The Egyptians at Kurgus. In: Spencer, Stevens, and Binder 2017, 65–106. Dawson, W. R. 1927. Making a mummy. The Journal of Egyptian Archaeology 13 (1/2): 40–49. Deer, W. A., R. A. Howie, and J. Zussman. 1992. An introduction to the rock forming minerals. 2nd ed. Harlow: Longman. Derrick, M. R., J. M. Landry, and D. Stulik. 1999. Infrared spectroscopy in conservation science. Los Angeles, CA: Getty Conservation Institute. Dobres, M.-A. 2001. Meaning in the making: Agency and the social embodiment of technology and art. In: M. B. Schiffer (ed.), Anthropological perspectives on technology. Albuquerque, NM: University of New Mexico Press, 47–76. Donker van Heel, K. 2016. Mrs. Naunakhte and family. Cairo: American University in Cairo Press. Douglass, D. L., C. Shing, and G. Wang. 1992. The lightinduced alteration of realgar to pararealgar. American Mineralogist 77: 1266–74. Eastaugh, N., V. Walsh, T. Chaplin, and R. Siddall. 2004a. Pigment compendium Part 1: A dictionary of historical pigments. Amsterdam; London: Elsevier ButterworthHeinemann.

97

Eastaugh, N., V. Walsh, T. Chaplin, and R. Siddall. 2004b. Pigment compendium Part 2: Optical microscopy of historical pigments. Amsterdam; London: Elsevier Butterworth-Heinemann. Edmonds, M. 1999. Ancestral geographies of the Neolithic: Landscapes, monuments and memory. London: Routledge. Edreira, M. C., M. J. Feliu, and J. Mart. 2001. Roman wall paintings characterization from Cripta del Museo and Alcazaba in Mérida (Spain): Chromatic, energy dispersive X-ray flurescence spectroscopic, X-ray diffraction and Fourier transform infrared spectroscopic analysis. Analytica Chimica Acta 434 (2): 331–45. doi: 10.1016/ S0003-2670(01)00847-9 Edwards, D. N. 2004. The Nubian past. An archaeology of the Sudan. London: Routledge. Edwards, D. N. and Mills, A. J. 2020. The Archaeological Survey of Sudanese Nubia 1963–69. Sudan Archaeological Research Society Publication Number 23. Oxford: Archaeopress. Egerton, R. F. 2005. Physical principles of electron microscopy: An introduction to TEM, SEM, and AEM. New York: Springer. El Goresy, A. 2000. Polychromatic wall painting decorations in monuments of Pharaonic Egypt: Compositions, chronology and painting techniques. In: S. Sherratt (ed.), The wall paintings of Thera. Proceedings of the First International Symposium Thera, Hellas, 30 August– 4 September 1997. Athens: Thera Foundation, 49–70. El Goresy, A., H. Jaksch, M. Abdel Razek, and K. L. Weiner. 1986. Ancient pigments in wall paintings of Egyptian tombs and temples: An archaeometric project. Preprint Vol. 12. Heidelberg: Max-Planck Institut für Kernphysik. Emberling, G. 1997. Ethnicity in complex societies: Archaeological perspectives. Journal of Archaeological Research 5 (4): 295–344. doi: 10.1007/BF02229256 Evershed, R. P. 2000. Biomolecular analysis by organic mass spectrometry. In: E. Ciliberto and G. Spoto (eds), Modern analytical methods in art and archaeology. New York; Chichester: Wiley, 177–240. Feneuille, S., J. Letourneu, and M. Bouchar. 2014. Archaeological information extracted from a comparative study of samples of mortar collected on various ancient monuments in the Nile valley between the Third and the Sixth Cataracts. In: J. R. Anderson and D. A. Welsby (eds), The Fourth Cataract and beyond: Proceedings of the 12th International Conference for Nubian Studies. Leuven: Peeters, 827–32. Fernea, R. A. and A. Rouchdy. 2010. Nubian culture and ethnicity. In: N. S. Hopkins and S. R. Mehanna (eds), Nubian encounters: The story of the Nubian Ethnological Survey 1961–1964. Cairo; New York: The American University in Cairo Press, 289–300.

98

BIBLIOGRAPHY

Fitzhugh, E. W. 1986. Red lead and minium. In: R. L. Feller (ed.), Artists’ pigments: A handbook of their history and characteristics. Vol. 1. Washington, DC: National Gallery of Art, 109–39. Fitzhugh, E. W. 1997. Orpiment and realgar. In: E. W. Fitzhugh (ed.), Artists’ pigments: A handbook of their history and characteristics. Vol. 3. Washington, DC: National Gallery of Art, 47–79. Fleming, A. 2006. Post-processual landscape archaeology: A critique. Cambridge Archaeological Journal 16 (3): 267–80. doi: 10.1017/S0959774306000163 Forbes, R. J. 1936a. Bitumen and petroleum in antiquity. Leiden: Brill. Forbes, R. J. 1936b. The nomenclature of bitumen petroleum tar and allied products in antiquity. Mnemosyne 4 (1): 67–77. Forstner-Müller, I. and D. Raue. 2008. Elephantine and the Levant. In: E.-M. Engel, V. Müller and U. Hartung (eds), Zeichen aus dem Sand: Streiflichter aus Ägyptens Geschichte zu Ehren von Günter Dreyer. Wiesbaden: Harrassowitz, 127–48. Frahm, E. 2014. Scanning electron microscopy (SEM): Applications in archaeology. In: C. Smith (ed.), Encyclopedia of global archaeology. New York: Springer. doi: 10.1007/978-1-4419-0465-2_341 Frankfort, H. and J. D. S. Pendlebury. 1933. The city of Akhenaten II. The north suburb and the desert alters. London; Oxford: Oxford University Press for the Egypt Exploration Society. Friedman, F. D. 1998. Faience: The brilliance of eternity. In: F. D. Friedman (ed.), Gifts of the Nile: Ancient Egyptian faience. London: Thames & Hudson, 15–21. Froment, F., A. Tournié, and P. Colomban. 2008. Raman identification of natural red to yellow pigments: Ochre and iron-containing ores. Journal of Raman Spectroscopy 39: 560–68. doi: 10.1002/jrs.1858 Fulcher, K. 2017. Evidence for the use of madder as a pigment in Nubia. Sudan & Nubia 21: 113–16. Fulcher, K. 2018. Colour taskscapes in ancient Sudan. In: S. Tipper and G. Tully (eds), Current research in Nubian archaeology. Piscataway, NJ: Gorgias Press, 23–36. Fulcher, K. 2019. Practising craft and producing memories in ancient Nubia. In: L. Killroe (ed.), Invisible archaeologies: Hidden aspects of daily life in ancient Egypt and Nubia. Oxford: Archaeopress, 56–63. Fulcher, K. and J. Budka. 2020. Pigments, incense, and bitumen from the New Kingdom town and cemetery on Sai Island in Nubia. Journal of Archaeological Science: Reports 33 (102550). doi: 10.1016/j.jasrep.2020. 102550 Fulcher, K., R. Stacey, and N. Spencer. 2020. Bitumen from the Dead Sea in early Iron Age Nubia. Nature Scientific Reports 10 (8309). doi: 10.1038/s41598-020-64209-8

Garrison, E. 2014. X-ray diffraction (XRD): Applications in archaeology. In: C. Smith (ed.), Encyclopedia of global archaeology. New York: Springer. doi: 10.1007/978-14419-0465-2_339 Gellatly, A. 1995. Colourful Whorfian ideas: Linguistic and cultural influences on the perception and cognition of colour, and on the investigation of them. Mind & Language 10 (3): 199–225. doi: 10.1111/j.1468-0017.1995. tb00011.x Genestar, C. and C. Pons. 2005. Earth pigments in painting: Characterisation and differentiation by means of FTIR spectroscopy and SEM-EDS microanalysis. Analytical and Bioanalytical Chemistry 382 (2): 269–74. doi: 10.1007/s00216-005-3085-8 Germer, R. 1992. Die Textilfärberei und die Verwendung gefärbter Textilien im alten Ägypten. Wiesbaden: Harrassowitz. Gettens, R. J., R. L. Feller, and W. Chase. 1972. Vermilion and cinnabar. Studies in Conservation 17 (2): 45–69. doi: 10.1179/sic.1972.006 Gettens, R. J., E. W. Fitzhugh, and R. L. Feller. 1974. Calcium carbonate whites. Studies in Conservation 19 (3): 157–84. doi: 10.1179/sic.1974.014 Gettens, R. J. and G. L. Stout. 1958. A monument of Byzantine wall painting: The method of construction. Studies in Conservation 3 (3): 107–19. doi: 10.1179/ sic.1958.016 Giménez, J. 2015. Finding hidden chemistry in ancient Egyptian artifacts: Pigment degradation taught in a chemical engineering course. Journal of Chemical Education 92 (3): 456–62. doi: 10.1021/ed500327j Goldstein, J., 2003. Scanning electron microscopy and X-ray microanalysis. New York; London: Kluwer Academic; Plenum. Green, L. 1995. Recent analysis of pigments from ancient Egyptian artefacts. In: C. E. Brown, F. Macalister, and M. M. Wright (eds), Conservation in ancient Egyptian collections. London: Archetype, 85–91. Green, L. 2001. Colour transformations of ancient Egyptian pigments. In: Davies 2001, 43–48. Grissom, C. A. 1986. Green earth. In: R. L. Feller (ed.), Artists’ pigments: A handbook of their history and characteristics. Vol. 1. Washington: National Gallery of Art, 141–67. Haldar, S. K. and J. Tišljar. 2014. Introduction to mineralogy and petrology. Amsterdam: Elsevier. Hamilakis, Y. 2013. Archaeology and the senses: Human experience, memory, and affect. New York: Cambridge University Press. Harrell, J. A., 2017. Amarna gypsite: A new source of gypsum for ancient Egypt. Journal of Archaeological Science: Reports 11, 536–45. doi: 10.1016/j.jasrep. 2016.12.031 Harrell, J. A. and M. D. Lewan. 2002. Sources of mummy bitumen in ancient Egypt and Palestine. Archaeometry 44 (2): 285–93. doi: 10.1111/1475-4754.t01-1-00060

BIBLIOGRAPHY

Harris, J. R. 1961. Lexicographical studies in ancient Egyptian minerals. Berlin: Akademie Verlag. Hatton, G. D. 2005. The technology of Egyptian blue. PhD thesis. University of Oxford. Hatton, G. D., A. J. Shortland, and M. S. Tite. 2008. The production technology of Egyptian blue and green frits from second millennium BC Egypt and Mesopotamia. Journal of Archaeological Science 35 (6): 1591–604. doi: 10.1016/j.jas.2007.11.008 Helwig, K. 2007. Iron oxide pigments: Natural and synthetic. In: B. H. Berrie (ed.), Artists’ pigments: A handbook of their history and characteristics. Vol. 4. Washington, DC: National Gallery of Art, 39–109. Hesse, A. 2006. Wall painting fragments from the Amun temple in the ancient city of Naga in Sudan. In: D. Saunders, J. Townsend and S. Woodcock (eds), The object in context: Crossing conservation boundaries: contributions to the Munich Congress 28 August– 1 September 2006. London: The International Institute for Conservation, 333. Heywood, A. 2001. The use of huntite as a white pigment in ancient Egypt. In: Davies 2001, 5–9. Heywood, A. 2010. Evidence for the use of azurite and natural ultramarine pigments in ancient Egypt. Metropolitan Museum Studies in Art, Science, and Technology 1: 73–81. Howard, H. 2003. Pigments of English medieval wall painting. London: Archetype. Hradil, D., T. Grygar, J. Hradilová, and P. Bezdička. 2003. Clay and iron oxide pigments in the history of painting. Applied Clay Science 22 (5): 223–36. doi: 10.1016/ S0169-1317(03)00076-0 Hummler, M. 2008. Review of Stone worlds: Narrative and reflexivity in landscape archaeology. Antiquity 82 (318): 1156–57. doi: 10.1017/S0003598X00120745 Hurcombe, L. 2007. A sense of materials and sensory perception in concepts of materiality. World Archaeology 39 (4): 532–45. doi: 10.1080/00438240701679346 Hussein, A. 1990. Mineral deposits. In: R. Said (ed.), The geology of Egypt. Rotterdam; Brookfield, VT: Published for the Egyptian General Petroleum Corp., Conoco Hurghada Inc. and Repsol Exploracion, S.A. by A. A. Balkema, 511–66. Infrared & Raman Users Group. 2016. Infrared & Raman Users Group Spectral Database. http://www.irug.org/ [accessed September 2020]. Ingold, T. 1993. The temporality of the landscape. World Archaeology 25 (2): 152–74. doi: 10.1080/00438243. 1993.9980235 Ingold, T. 2011. The perception of the environment: essays of livelihood, dwelling and skill. 2nd ed. London: Routledge. Jaksch, H. 1985. Farbpigmente aus Wandmalereien altägyptischer Gräber und Tempel: Technologien der Herstel-

99

lung und mögliche Herkunftsbeziehungen. PhD thesis. University of Heidelberg. Jaksch, H., W. Seipel, K. L. Weiner, and A. el Goresy. 1983. Egyptian blue – cuprorivaite. A window to ancient Egyptian technology. Die Naturwissenschaften 70 (11): 525–35. doi: 10.1007/BF00376668 Joyce, R. A. 2002. Introducing the First Voice. In: R. A. Joyce (ed.), The languages of archaeology: Dialogue, narrative, and writing. Oxford; Malden, MA: Blackwell, 4–17. Kakoulli, I. 2002. Late Classical and Hellenistic painting techniques and materials: A review of the technical literature. Studies in Conservation 47: 56–67. doi: 10.1179/sic.2002.47.Supplement-1.56 Kakoulli, I. 2009. Greek painting techniques and materials from the fourth to the first century BC. London: Archetype. Kay, P. 1975. Synchronic variability and diachronic change in basic color terms. Language in Society 4 (3): 257–70. doi: 10.1017/S0047404500006667 Kay, P., B. Berlin, and W. Merrifield. 1991. Biocultural implications of systems of color naming. Journal of Linguistic Anthropology 1 (1): 12–25. doi: 10.1525/jlin. 1991.1.1.12 Kay, P. and W. Kempton. 1984. What is the Sapir-Whorf hypothesis? American Anthropologist 86: 65–79. doi: 10.1525/aa.1984.86.1.02a00050 Kay, P. and C. K. McDaniel. 1978. The linguistic significance of the meanings of basic color terms. Language 54 (3): 610–46. doi: 10.2307/412789 Keith, J. L. 2011. Anthropoid busts of Deir el Medineh and other sites and collections. Cairo: Institut français d’archéologie orientale. Kemp, B. J. 1979. Wall paintings from the workmen’s village at el-Amarna. The Journal of Egyptian Archaeology 65: 47–53. Kemp, B. J. 1997. Why empires rise. Cambridge Archaeological Journal 7 (1): 125–31. doi: 10.1017/ S0959774300001505 Kemp, B. J. and A. Stevens. 2010a. Busy lives at Amarna: Excavations in the main city (Grid 12 and the House of Ranefer, N49.18). Vol. I: The excavations, architecture and environmental remains. EES Excavation Memoir 90. London: Egypt Exploration Society and Amarna Trust. Kemp, B. J. and A. Stevens. 2010b. Busy lives at Amarna: Excavations in the main city (Grid 12 and the House of Ranefer, N49.18). Vol. II: The objects. EES Excavation Memoir 91. London: Egypt Exploration Society and Amarna Trust. Kendall, T. and Mohamed, E. A., 2016. The Jebel Barkal Temples. The NCAM Jebel Barkal Mission. http:// www.jebelbarkal.org/frames/VisGuide.pdf [accessed September 2020].

100

BIBLIOGRAPHY

Kerr, P. F. 1959. Optical mineralogy. New York; London: McGraw-Hill. Klemm, R. and D. D. Klemm. 2008. Stones and quarries in ancient Egypt. London: British Museum Press. Knoblauch, C. 2011. Not all that glitters: A case study of regional aspects of Egyptian Middle Kingdom pottery production in Lower Nubia and the Second Cataract. Cahiers de la céramique égyptienne 9: 167–84. Koulis, C. V., J. A. Reffner, and A. M. Bibby. 2001. Comparison of transmission and internal reflection infrared spectra of cocaine. Journal of Forensic Sciences 46 (4): 822–29. doi: 10.1520/JFS15053J Lacovara, P. 1986. The funerary chapels at Kerma. Cahiers de recherches de l’Institut de papyrologie et d’égyptologie de Lille 8: 49–58. Lacovara, P. and A. Winkels. 2018. Malqata: The painted palace. In: J. Becker, J. Jungfleisch, and C. von Rüden (eds), Tracing technoscapes: The production of Bronze Age wall paintings in the Eastern Mediterranean. Leiden: Sidestone, 149–72. Lafuente, B., R. T. Downs, H. Yang, and N. Stone. 2015. The power of databases: The RRUFF project. In: T. Armbruster and R. M. Danisi (eds), Highlights in mineralogical crystallography. Berlin: W. de Gruyter, 1–30. Lau, D., P. Kappen, M. Strohschnieder, N. Brack, and P. J. Pigram. 2008. Characterization of green copper phase pigments in Egyptian artifacts with X-ray absorption spectroscopy and principal components analysis. Spectrochimica Acta Part B: Atomic Spectroscopy 63 (11): 1283–89. doi: 10.1016/j.sab.2008.09.013 Lauf, R. J. 2010. Collector’s guide to the chlorite group. Rocks & Minerals 85 (4): 318–25. doi: 10.1080/ 00357521003727272 Le Fur, D. 1994. La conservation des peintures murales des temples de Karnak. Paris: Éditions Recherche sur les Civilisations. Lee, L. and S. Quirke. 2000. Painting materials. In: Nicholson and Shaw 2000, 104–20. Lehner, M., M. Kamel, and A. Tavares. 2006. Giza plateau mapping project. Season 2005 preliminary report. Giza Occasional Papers 2. Boston, MA: Ancient Egypt Research Associates. Lehner, M., M. Kamel, and A. Tavares. 2009. Giza plateau mapping project. Seasons 2006–2007 preliminary report. Giza occasional papers 3. Boston, MA: Ancient Egypt Research Associates. Lesko, L. H. and B. S. Lesko. 2002. A dictionary of late Egyptian. 2nd ed. Providence, RI: B.C. Scribe Publications. Levinson, S. C. 2000. Yélî Dnye and the theory of Basic Color Terms. Journal of Linguistic Anthropology 10 (1): 3–55. doi: 10.1525/jlin.2000.10.1.3 Liang, J. and D. A. Scott. 2014. Green-copper containing waxy paint on two Egyptian polychrome artifacts:

A technical study. Studies in Conservation 59 (6): 391– 403. doi: 10.1179/2047058413Y.0000000106 Lucas, A. 1962. Ancient Egyptian materials and industries. 4th ed. (1st ed. 1934). London: E. Arnold. Lyons, J. 1995. Colour in language. In: T. Lamb and J. Bourriau (eds), Colour: Art and science. Cambridge: Cambridge University Press, 194–224. MacDonald, K. C. and D. N. Edwards. 1993. Chickens in Africa: The importance of Qasr Ibrim. Antiquity 67 (256): 584–90. doi: 10.1017/S0003598X00045786 Majewski, T. 2000. We are all storytellers: Comments on storytelling, science, and historical archaeology. Society for Historical Archaeology 34 (2): 17–19. doi: 10.1007/ BF03374309 Masschelein-Kleiner, L., J. Heylen, and F. Tricot-Marckx. 1968. Contribution à l’analyse des liants, adhésifs et vernis anciens. Studies in Conservation 13 (3): 105–21. doi: 10.1179/sic.1968.009 Mazzocchin, G. A., F. Agnoli, S. Mazzocchin, and I. Colpo. 2003. Analysis of pigments from Roman wall paintings found in Vicenza. Talanta 61 (4): 565–72. doi: 10.1016/ S0039-9140(03)00323-0 Mazzocchin, G. A., P. Baraldi, and C. Barbante. 2008. Isotopic analysis of lead present in the cinnabar of Roman wall paintings from the Xth Regio “(Venetia et Histria)” by ICP-MS. Talanta 74 (4): 690–93. doi: 10.1016/j.talanta.2007.06.048 McCarthy, B., 2001. Technical analysis of reds and yellows in the tomb of Suemniwet, Theban Tomb 92. In: Davies 2001, 17–21. Mermut, A. R. and H. Khademi. 2006. Gypsum formation in gypsic soils. In: R. Lal (ed.), Encyclopedia of soil science. Boca Raton, FL: CRC Press, 800–803. Meskell, L. 2004. Object worlds in ancient Egypt: Material biographies past and present. Oxford: Berg. Middleton, A. 1999. Polychromy of some fragments of painted relief from El-Bersheh. In: W. V. Davies (ed.), Studies in Egyptian antiquities: A tribute to T.G.H. James. British Museum Occasional Paper 123. London: British Museum, 37–44. Middleton, A. and S. Humphrey. 2001. Pigments on some Middle Kingdom coffins. In: Davies 2001, 10–16. Minor, E. J. 2012. The use of Egyptian and Egyptianizing material culture in Nubian burials of the Classic Kerma Period. Berkeley, CA: University of California. Moretto, L. M., E. F. Orsega, and G. A. Mazzocchin. 2011. Spectroscopic methods for the analysis of celadonite and glauconite in Roman green wall paintings. Journal of Cultural Heritage 12 (4): 384–91. doi: 10.1016/j. culher.2011.04.003 Morgan, L. 2006. Art and international relations: The hunt frieze at Tell el-Dab‘a. In: E. Czerny, I. Hein, H. Hunger, D. Melman and A. Schwab (eds.), Timelines: Studies in honour of Manfred Bietak, Vol. II. Leuven: Peeters, 249–58.

BIBLIOGRAPHY

Morgan, L. 2018. Forming the image: Approaches to painting at Ayia Irini, Kea and Tell el-Dab‘a. In: C. von Rüden, J. Jungfleisch, and J. Becker (eds), Tracing technoscapes: The production of Bronze Age wall paintings in the eastern Mediterranean. Leiden: Sidestone, 235–51. Moussa, A. and M. F. Ali. 2013. Color alteration of ancient Egyptian blue faience. International Journal of Architectural Heritage 7 (3): 261–74. doi: 10.1080/ 15583058.2011.634960 Nash, M. 1987. Ethnicity in peninsular Malaysia: The idiom of communalism, confrontation and co-operation. In: P. Hockings (ed.), Dimensions of social life: Essays in honor of David G. Mandelbaum. Berlin: Mouton de Gruyter, 559–71. Newman, R., 1993. Analysis of red paint and filling material from the sarcophagus of Queen Hatshepsut and King Thutmose I. Journal of the Museum of Fine Arts, Boston 5: 62–65. Newman, R. and S. M. Halpine. 2001. The binding media of ancient Egyptian painting. In: Davies 2001, 22–32. Newman, R. and M. Serpico. 2000. Adhesives and binders. In: Nicholson and Shaw 2000, 475–94. Nicholson, P. T. 2007. Brilliant things for Akhenaten: The production of glass, vitreous materials and pottery at Amarna Site O45.1. London: Egypt Exploration Society. Nicholson, P. T. 2008. Glass and faience production sites in New Kingdom Egypt: A review of the evidence. In: C. M. Jackson and E. C. Wager (eds), Vitreous materials in the late Bronze Age Aegean. Oxford: Oxbow Books, 1–13. Nicholson, P. T. and I. Shaw, eds. 2000. Ancient Egyptian materials and technology. Cambridge: Cambridge University Press. Noll, W. 1978. Mineralogie und Technik der bemalten Keramiken Altägyptens. Neues Jahrbuch für Mineralogie Abhandlungen 133: 227–90. Nussinovitch, A. 2010. Plant gum exudates of the world: Sources, distribution, properties, and applications. Boca Raton, FL: CRC Press. Ospitali, F., D. Bersani, G. Di Lonardo, and P. P. Lottici. 2008. ‘Green earths’: Vibrational and elemental characterization of glauconites, celadonites and historical pigments. Journal of Raman Spectroscopy 39 (8): 1066–73. doi: 10.1002/jrs.1983 Pagès-Camagna, S. and S. Colinart. 2003. The Egyptian green pigment: Its manufacturing process and links to Egyptian blue. Archaeometry 45 (4): 637–58. doi: 10.1046/j.1475-4754.2003.00134.x Pagès-Camagna, S. and S. Colinart. 2006. Authors’ reply. Archaeometry 48 (4): 707–13. doi: 10.1111/j.14754754.2006.282_2.x Pagès-Camagna, S., S. Colinart, and C. Coupry. 1999. Fabrication processes of archaeological Egyptian blue and

101

green pigments enlightened by Raman microscopy and scanning electron microscopy. Journal of Raman Spectroscopy 30: 313–17. doi: 10.1002/(SICI)10974555(199904)30:43.0.CO;2-B Pagès-Camagna, S., A. Duval, and H. Guicharnaud. 2004. Study of Gustave Moreau’s black drawings: Identification of the graphic materials by Raman microspectrometry and PIXE. Journal of Raman Spectroscopy 35 (89): 628–32. doi: 10.1002/jrs.1215 Pagès-Camagna, S. and H. Guichard. 2010. Egyptian colours and pigments in French collections: Physicochemical analyses on 300 objects. In: J. Dawson, C. Rozeik, and M. M. Wright (eds), Decorated surfaces on ancient Egyptian objects: Technology, deterioration and conservation. Proceedings of a conference held in Cambridge, UK on 7–8 September 2007. London: Archetype in association with the Fitzwilliam Museum and Icon Archaeology Group, 25–31. Pagès-Camagna, S. and D. Raue. 2016. Coloured materials used in Elephantine: Evolution and continuity from the Old Kingdom to the Roman Period. Journal of Archaeological Science: Reports 7: 662–67. doi: 10.1016/j. jasrep.2016.02.002 Pallasmaa, J. 1996. The eyes of the skin: Architecture and the senses. Chichester: Wiley. Peet, T. E. and C. L. Woolley. 1923. The city of Akhenaten I. Excavations of 1921 and 1922 at El-’Amarneh. London: Egypt Exploration Society. Peters, K. E., C. C. Walters, and J. M. Moldowan. 2005. The Biomarker Guide. Cambridge: Cambridge University Press. Petrie, W. M. F. 1904. Methods and aims in archaeology. London: Macmillan. Pfaffenberger, B. 2001. Symbols do not create meanings— activities do: Or, why symbolic anthropology needs the anthropology of technology. In: M. B. Schiffer (ed.), Anthropological perspectives on technology. Albuquerque, NM: University of New Mexico Press, 77–86. Pilgrim, C. von. 1996. Untersuchungen in der Stadt des Mittleren Reiches und der Zweiten Zwischenzeit. Mainz: P. von Zabern. Pilgrim, C. von. 2016. Excavation of House 55 (18th Dynasty). In: J. Seidlmayer (ed.), Report on the excavations at Elephantine by the German Archaeological Institute and the Swiss Institute from autumn 2015 to summer 2016. Berlin: Deutsches Archäologisches Institut and Cairo: Swiss Institute for Architectural and Archaeological Research on Ancient Egypt, 22–25. Pilgrim, B. von. and C. von Pilgrim. 2007. Area II: Domestic quarters of the New Kingdom. In: D. Raue (ed.), Report on the 36th season of excavation and restoration on the island of Elephantine. Berlin: Deutsches Archäologisches Institut, 7.

102

BIBLIOGRAPHY

Pinch, G. 2001. Red things: The symbolism of colour in magic. In: Davies 2001, 182–85. Pluciennik, M. 1999. Archaeological narratives and other ways of telling. Current Anthropology 40 (5): 653–78. doi: 10.1086/300085 Prell, S. 2011. Einblicke in die Werkstätten der Residenz: die Stein- und Metallwerkzeuge des Grabungsplatzes Q I. Hildesheim: Gerstenberg. Pullan, M., J. Ambers, C. Cartwright, J. H. Taylor, and F. Kalloniatis. 2012. The Norwich shroud: Conservation and investigation of a rare Eighteenth Dynasty shroud. British Museum Technical Research Bulletin 6: 13–24. Pusch, E. B. 1999. Glassproduktion in Qantir. Ägypten und Levante 9: 111–20. Pusch, E. B. and T. Rehren. 2007. Hochtemperatur-Technologie in der Ramses-Stadt: Rubinglas für den Pharao. Hildesheim: Gerstenberg. Pyke, G. 2007. The fragmentary wall plaster. In: P. Rose (ed.), The Meroitic temple complex at Qasr Ibrim. London: Egypt Exploration Society, 44–68. Quirke, S. 2001. Colour vocabularies in ancient Egyptian. In: Davies 2001, 186–92. Ramer, B. 1979. The technology, examination and conservation of the Fayum portraits in the Petrie Museum. Studies in Conservation 24 (1): 1–13. doi: 10.2307/ 1505918 Regev, L., K. M. Poduska, L. Addadi, S. Weiner, and E. Boaretto. 2010. Distinguishing between calcites formed by different mechanisms using infrared spectrometry: Archaeological applications. Journal of Archaeological Science 37 (12): 3022–29. doi: 10.1016/j.jas.2010. 06.027 Rehren, T. 1995. High temperature industries in the Late Bronze Age capital Piramesse-(Qantir): I - Bronze and glass production and processing. In: F. A. Esmael (ed.), Proceedings of the first International Conference on Ancient Egyptian Mining & Metallurgy and Conservation of Metallic Artifacts, Cairo, Egypt, 10–12 April 1995. Cairo: Ministry of Culture, Supreme Council of Antiquities, 101–19. Rehren, T. 2001. Aspects of the production of cobalt-blue glass in Egypt. Archaeometry 43 (4): 483–89. doi: 10.1111/1475-4754.00031 Rehren, T. and E. B. Pusch. 1997. New Kingdom glassmelting crucibles from Qantir-Piramesses. The Journal of Egyptian Archaeology 83: 127–41. Rehren, T., E. B. Pusch, and A. Herold. 1998. Glass coloring works within a copper-centered industrial complex in late bronze age Egypt. In: P. McCray and W. D. Kingery (eds), The prehistory and history of glassmaking technology. Westerville, OH: American Ceramic Society, 227–50. Rehren, T., E. B. Pusch, and A. Herold. 2001. QantirPiramesses and the organisation of the Egyptian glass

industry. In: A. J. Shortland (ed.), The social context of technological change: Egypt and the Near East, 1650–1550 B.C. Oxford: Oxbow, 223–38. Reisner, G. A. 1923. Excavations at Kerma I–V. Cambridge, MA: Peabody Museum of Harvard University. Riederer, J. 1974. Recently identified Egyptian pigments. Archaeometry 16 (1): 102–109. doi: 10.1111/j.14754754.1974.tb01098.x Ritner, R. K. 1993. The mechanics of ancient Egyptian magical practice. Chicago, IL: Oriental Institute of the University of Chicago. Rose, P. 2007. Painting on reused stone blocks. In: P. Rose (ed.), The Meroitic temple complex at Qasr Ibrim. London: Egypt Exploration Society, 69–73. Rosenvasser, A. 1964. Preliminary report of the excavations at Aksha by the Franco-Argentine archaeological expedition, 1962–63. Kush 12: 96–101. Rouchon, O., J. Fabre, M.-P. Etcheverry, and M. Schvoerer. 1990. Pigments d’Égypte: étude physique de matières colorantes bleue, rouge, blanche, verte et jaune, provenant de Karnak. Revue d’Archéométrie 14: 87–97. Rowe, S., R. Siddall, and R. Stacey. 2010. Roman Egyptian gilded cartonnage: Technical study and conservation of a mummy mask from Hawara. In: J. Dawson, C. Rozeik, and M. M. Wright (eds), Decorated surfaces on ancient Egyptian objects: Technology, deterioration and conservation: proceedings of a conference held in Cambridge, UK on 7–8 September 2007. London: Archetype in association with the Fitzwilliam Museum and Icon Archaeology Group, 106–21. Royal Botanic Gardens Kew. 2016. Vachellia nilotica. Plants of the world online. http://powo.science.kew.org/ taxon/urn:lsid:ipni.org:names:77089275-1 [accessed September 2020]. Russell, W. J. 1892. Egyptian colours. In: W. M. F. Petrie (ed.), Medum. London: David Nutt, 44–48. Ryan, P. 2016. From raw resources to food processing. Archaeobotanical and ethnographic insights from New Kingdom Amara West and present-day Ernetta Island in Northern Sudan. In: L. Steel and K. Zinn (eds), Exploring the materiality of food “stuffs”: Archaeological and anthropological perspectives. London: Routledge, 15–38. Ryan, P., C. Cartwright, and N. Spencer. 2012. Archaeobotanical research in a Pharaonic town in ancient Nubia. British Museum Technical Research Bulletin 6: 97–107. Said, R. 1990. The geology of Egypt. Rotterdam; Brookfield, VT: Published for the Egyptian General Petroleum Corp., Conoco Hurghada Inc. and Repsol Exploracion, S.A. by A. A. Balkema. Saleh, S. A., Z. Iskander, A. A. El-Masry, and F. M. Helmi. 1974. Some ancient Egyptian pigments. In: A. Bishay (ed.), Recent advances in science and technology of

BIBLIOGRAPHY

materials. Proceedings of the 2nd Cairo Solid State Conference, 1973. Vol. 3. New York: Plenum Press, 141–55. Saunders, B. and J. van Brakel. 2002. The trajectory of color. Perspectives on Science 10 (3): 302–55. doi: 10.1162/106361402321899078 Schiegl, S. and A. el Goresy. 2006. Comments on S. PagèsCamagna and S. Colinart, “The Egyptian green pigment: Its manufacturing process and links to Egyptian blue.” Archaeometry 48 (4): 707–13. doi: 10.1111/j.14754754.2006.282_1.x Schiegl, S., K. L. Weiner, and A. el Goresy. 1989. Discovery of copper chloride cancer in ancient Egyptian polychromic wall paintings and faience—a developing archaeological disaster. Naturwissenschaften 76 (9): 393–400. doi: 10.1007/BF00366160 Schiegl, S., K. L. Weiner, and A. el Goresy. 1992. The diversity of newly discovered deterioration patterns in ancient Egyptian pigments: Consequences to entirely new restoration strategies and to the Egyptological colour symbolism. MRS Proceedings Library Archive 267: 831–58. doi: 10.1557/PROC-267-831 Schreiner, M., B. Frühmann, D. Jembrih-Simbürger, and R. Linke. 2004. X-rays in art and archaeology: An overview. Powder Diffraction 19 (1): 3–11. doi: 10.1154/ 1.1649963 Scott, D. 2002. Copper and bronze in art: Corrosion, colorants, conservation. Los Angeles, CA: Getty Conservation Institute. Scott, D. 2010a. Ancient Egyptian pigments: The examination of some coffins from the San Diego Museum of Man. MRS Bulletin 35: 390–96. doi: 10.1557/ mrs2010.572 Scott, D. 2010b. Greener shades of pale: A review of advances in the characterisation of ancient Egyptian green pigments. In: J. Dawson, C. Rozeik, and M. M. Wright (eds), Decorated surfaces on ancient Egyptian objects: Technology, deterioration and conservation. Proceedings of a conference held in Cambridge, UK on 7–8 September 2007. London: Archetype in association with the Fitzwilliam Museum and Icon Archaeology Group, 32–45. Scott, D., M. Dennis, N. Khandekar, J. Keeney, D. Carson, and L. Swartz Dodd. 2003. An Egyptian cartonnage of the Graeco-Roman Period: Examination and discoveries. Studies in Conservation 48 (1): 41–56. doi: 10.1179/sic.2003.48.1.41 Scott, D., L. S. Dodd, J. Furihata, S. Tanimoto, J. Keeney, M. R. Schilling, and E. Cowan. 2004. An ancient Egyptian cartonnage broad collar: Technical examination of pigments and binding media. Studies in Conservation 49 (3): 177–92. doi: 10.1179/sic.2004.49.3.177 Scott, D., S. Warmlander, J. Mazurek, and S. Quirke. 2009. Examination of some pigments, grounds and media

103

from Egyptian cartonnage fragments in the Petrie Museum, University College London. Journal of Archaeological Science 36 (3): 923–32. doi: 10.1016/j. jas.2008.12.011 Seeber, R. 2000. The technique of plaster preparation for the Minoan wall paintings at Tell el-Dab‘a, Egypt: Preliminary results. In: S. Sherratt (ed.), The wall paintings of Thera. Proceedings of the First International Symposium Thera, Hellas, 30 August–4 September 1997. Athens: Thera Foundation, 91–102. Serpico, M. 2000. Resins, amber and bitumen. In: Nicholson and Shaw 2000, 430–74. Serpico, M. 2011. The contents of jars in Hatshepsut’s foundation deposit at Deir el-Bahri and their significance for trade. In: D. Aston, B. Bader, C. Gallorini, P. Nicholson, and S. Buckingham (eds), Under the potter’s tree: Studies on ancient Egypt presented to Janine Bourriau on the occasion of her 70th birthday. Orientalia Lovaniensia Analecta 204. Leuven: Peeters, 843–84. Serpico, M. and R. White. 2000. The botanical identity and transport of incense during the Egyptian New Kingdom. Antiquity 74 (March): 884–97. doi: 10.1017/ S0003598X00060531 Serpico, M. and R. White. 2001. The use and identification of varnish on New Kingdom funerary equipment. In: Davies 2001, 33–38. Shackley, M. S. 2011. An introduction to X-ray fluorescence (XRF) analysis in archaeology. In: M. S. Shackley (ed.), X-ray fluorescence spectrometry (XRF) in geoarchaeology. New York: Springer, 7–44. Shinnie, P. L. and R. J. Bradley. 1980. The capital of Kush I: Meroe excavations, 1965–1972. Berlin: Akademie-Verlag. Shinnie, P. L. and R. J. Bradley. 1981. Murals from the Augustus Temple, Meroe. In: W. K. Simpson and W. M. Davis (eds), Studies in ancient Egypt, the Aegean, and the Sudan. Boston: Museum of Fine Arts, 167–72. Shortland, A. J. 2000. The number, extent and distribution of the vitreous materials workshops at Amarna. Oxford Journal of Archaeology 19 (2): 115–34. doi: 10.1111/ 1468-0092.00104 Shortland, A. J. 2002. The use and origin of antimonate colorants in early Egyptian glass. Archaeometry 44 (4): 517–30. doi: 10.1111/1475-4754.t01-1-00083 Shortland, A. J., C. A. Hope, and M. S. Tite. 2006. Cobalt blue painted pottery from 18th Dynasty Egypt. In: M. Maggetti and B. Messiga (eds), Geomaterials in cultural heritage. London: Geological Society, 91–99. Shortland, A. J., P. T. Nicholson, and C. M. Jackson. 2001. Glass and faience at Amarna: Different methods of both supply for production, and subsequent distribution. In: A. J. Shortland (ed.), The social context of technological change: Egypt and the Near East, 1650–1550 B.C. Oxford: Oxbow, 147–60.

104

BIBLIOGRAPHY

Siddall, R. 2011. Asphaltite pigments in ancient Egypt. Traditional Paint News 3: 7–15. Skeates, R. 2010. An archaeology of the senses: Prehistoric Malta. Oxford: Oxford University Press. Smirniou, M. and T. Rehren. 2011. Direct evidence of primary glass production in Late Bronze Age Amarna, Egypt. Archaeometry 53 (1): 58–80. doi: 10.1111/ j.1475-4754.2010.00521.x Smith, H. S. 1976. The fortress of Buhen: The inscriptions. London: Egypt Exploration Society. Smith, S. T. 1997. Ancient Egyptian imperialism: Ideological vision or economic exploitation? Reply to critics of Askut in Nubia. Cambridge Archaeological Journal 7 (2): 301–307. doi: 10.1017/S0959774300002006 Smith, S. T. 2003. Wretched Kush: Ethnic identity in Egypt’s Nubian empire. London: Routledge. Smith, S. T. 2007. Death at Tombos: Pyramids, iron and the rise of the Napatan Dynasty. Sudan & Nubia 11: 2–14. Smith, S. T. and M. R. Buzon. 2014. Identity, commemoration, and remembrance in colonial encounters: Burials at Tombos during the Egyptian New Kingdom Nubian empire and its aftermath. In: B. W. Porter and A. T. Boutin (eds), Remembering the dead in the ancient Near East: Recent contributions from bioarchaeology and mortuary archaeology. Boulder, CO: University Press of Colorado, 185–216. Smith, S. T. and M. R. Buzon. 2017. Colonial encounters at New Kingdom Tombos: Cultural entanglements and hybrid identity. In: Spencer, Stevens, and Binder 2017, 615–30. Spataro, M., A. Garnett, A. Shapland, N. Spencer, and H. Mommsen. 2019. Mycenaean pottery from Amara West (Nubia, Sudan). Archaeological and Anthropological Sciences 11: 683–97. doi: 10.1007/s12520-017-0552-z Spataro, M., M. Millet, and N. Spencer. 2015. The New Kingdom settlement of Amara West (Nubia, Sudan): Mineralogical and chemical investigation of the ceramics. Archaeological and Anthropological Sciences 7 (4): 399–421. doi: 10.1007/s12520-014-0199-y Spector, J. 1993. What this awl means: Feminist archaeology at a Wahpeton Dakota village. St. Paul, MN: Minnesota Historical Society Press. Spencer, N. 2009. Cemeteries and a late Ramesside suburb at Amara West. Sudan & Nubia 13: 47–61. Spencer, N. 2010. Nubian architecture in an Egyptian town? Building E12.11 at Amara West. Sudan & Nubia 14: 15–24. Spencer, N., 2012. Amara West: life in Egypt’s Nubian empire. Current World Archaeology 54: 26–34. Spencer, N. 2014a. Amara West: Considerations on urban life in colonial Kush. In: J. R. Anderson and D. A. Welsby (eds), The Fourth Cataract and beyond: Proceedings of the 12th International Conference for Nubian Studies. Leuven: Peeters, 457–86.

Spencer, N. 2014b. Creating and re-shaping Egypt in Kush: Responses at Amara West. Journal of Ancient Egyptian Interconnections 6 (1): 42–61. Spencer, N. 2015. Creating a neighborhood within a changing town: Household and other agencies at Amara West in Nubia. In: M. Müller (ed.), Household studies in complex societies: (Micro) archaeological and textual approaches. Chicago, IL: Oriental Institute of the University of Chicago, 167–208. Spencer, N. 2017. Building on new ground: The foundation of a colonial town at Amara West. In: Spencer, Stevens, and Binder 2017, 322–53. Spencer, N. 2019. Settlements of the Second Intermediate Period and New Kingdom. In: D. Raue (ed.), Handbook of ancient Nubia. Berlin: De Gruyter, 433–64. Spencer, N., M. Macklin, and J. Woodward. 2012. Reassessing the abandonment of Amara West: The impact of a changing Nile? Sudan & Nubia 16: 37–43. Spencer, N., A. Stevens, and M. Binder. 2014. Amara West: Living in Egyptian Nubia. London: British Museum. Spencer, N., A. Stevens, and M. Binder. 2017. Introduction. In: Spencer, Stevens, and Binder 2017, 1–60. Spencer, N., A. Stevens, and M. Binder, eds. 2017. Nubia in the New Kingdom: Lived experience, Pharaonic control and indigenous traditions. Leuven: Peeters. Spencer, P. 1997. Amara West I: The architectural report. EES Excavation Memoir 63. London: Egypt Exploration Society. Spencer, P. 2016. Amara West III: The scenes and texts of the Ramesside temple. EES Excavation Memoir 114. London: The Egypt Exploration Society. Stacey, R. 2008. Paint media and varnishes. In: A. Middleton and K. Uprichard (eds), The Nebamun wall paintings: Conservation, scientific analysis and display at the British Museum. London: Archetype in association with the British Museum, 51–60. Stevens, A. 2012. Akhenaten’s workers: The Amarna Stone Village Survey, 2005–2009. Vol. II: The faunal and botanical remains, and objects. EES Excavation Memoir 101. London: Egypt Exploration Society and Amarna Trust. Stevens, A. 2017. Death and the city: The cemeteries of Amarna in their urban context. Cambridge Archaeological Journal 28 (1): 103–26. doi: 10.1017/ S0959774317000592 Stevens, A. and A. Garnett. 2017. Surveying the Pharaonic desert hinterland of Amara West. In: Spencer, Stevens, and Binder 2017, 287–306. Stockman, A. and L. T. Sharp. 1999. Cone spectral sensitivities and color matching. In: K. R. Gegenfurtner and L. T. Sharpe (eds), Color vision: from genes to perception. Cambridge: Cambridge University Press, 53–88. Stuart, B. 2004. Infrared spectroscopy: Fundamentals and applications. Chichester; Hoboken, NJ: Wiley.

BIBLIOGRAPHY

Stulik, D., E. Porta, and A. Palet. 1993. Analyses of pigments, binding media and varnishes. In: M. A. Corzo and M. Afshar (eds), Art and eternity: The Nefertari Wall Paintings Conservation Project 1986–1992. Santa Monica, CA: Getty Conservation Institute, 55–65. Sweek, T., J. R. Anderson, S. M. Ahmed, and S. Tanimoto. 2014. Conservation of an Amun temple in the Sudan. In: J. R. Anderson and D. A. Welsby (eds), The Fourth Cataract and beyond: Proceedings of the 12th International Conference for Nubian Studies. Leuven: Peeters, 703–10. Szpakowska, K. 2003. Playing with fire: Initial observations on the religious uses of clay cobras from Amarna. Journal of the American Research Center in Egypt 40: 113–22. Taylor, J. H. 2001. Patterns of colouring on ancient Egyptian coffins from the New Kingdom to the Twenty-sixth Dynasty: A overview. In: Davies 2001, 164–81. Therkildsen, R. 2015. In a royal cemetery of Kush: Archaeological investigations at el-Kurru, Northern Sudan, 2014–15. Documentation and conservation of the painted tombs: Progress report. VIL and XRF analysis of the painted tombs. Sudan & Nubia 19: 57. Tite, M. 2000. X-ray fluorescence (XRF) analysis. In: L. Ellis (ed.), Archaeological method and theory: An encyclopedia. New York; London: Garland, 672–74. Tite, M., M. Bimson, and M. R. Cowell. 1984. Technological examination of Egyptian blue. In: J. B. Lambert (ed.), Archaeological chemistry III. Washington DC: American Chemical Society, 215–42. Tite, M., M. Bimson, and M. R. Cowell. 1987. The technology of Egyptian blue. In: M. Bimson and I. C. Freestone (eds), Early vitreous materials. British Museum Occasional Paper 56. London: British Museum, 39–46. Tite, M., P. Manti, and A. J. Shortland. 2007. A technological study of ancient faience from Egypt. Journal of Archaeological Science 34 (10): 1568–83. doi: 10.1016/j.jas.2006.11.010 Tite, M. and A. J. Shortland. 2003. Production technology for copper- and cobalt-blue vitreous materials from the New Kingdom site of Amarna—a reappraisal. Archaeometry 45 (2): 285–312. doi: 10.1111/1475-4754. 00109 Tomasini, E., G. Siracusano, and M. S. Maier. 2012. Spectroscopic, morphological and chemical characterization of historic pigments based on carbon. Paths for the identification of an artistic pigment. Microchemical Journal 102: 28–37. doi: 10.1016/j.microc.2011.11.005 Töpfer, S. 2014. Bemerkungen zum Balsamierungsritual nach den Papyri Boulaq III und Louvre E 5158. In: J. F. Quack (ed.), Ägyptische Rituale der griechischrömischen Zeit. Tübingen: Mohr Siebeck, 201–21. Töpfer, S. 2015a. Das Balsamierungsritual: eine (Neu-) Edition der Textkomposition Balsamierungsritual (pBoulaq 3, pLouvre 5158, pDurham 1983.11 + pSt. Petersburg 18128). Studien zur spätägyptischen Religion 13. Wiesbaden: Harrassowitz.

105

Töpfer, S. 2015b. Funktion, Verwendung und Entstehung der Textkomposition “Balsamierungsritual.” In: B. Backes and J. Dielemann (eds), Liturgical texts for Osiris and the deceased in Late Period Egypt. Proceedings of the colloquiums at New York (ISAW), 6 May 2011, and Freudenstadt, 18–21 July 2012. Wiesbaden: Harrassowitz, 245–58. Török, L. 2009. Between two worlds: The frontier region between ancient Nubia and Egypt, 3700 BC–500 AD. Leiden: Brill. Tringham, R. 2015. Creating narratives of the past as recombinant histories. In: R. M. Van Dyke and R. Bernbeck (eds), Subjects and narratives in archaeology. Denver, CO: University Press of Colorado, 27–54. Tringham, R. 2016. Dido and the basket: Fragments towards a non-linear history. In: S. Brown, A. Clarke, and U. Frederick (eds), Object stories: Artifacts and archaeologists. London: Routledge, 161–68. UCL 2015. Research Ethics at UCL. https://ethics.grad.ucl. ac.uk/ [accessed September 2020]. Uda, M. 2004. In situ characterization of ancient plaster and pigments on tomb walls in Egypt using energy dispersive X-ray diffraction and fluorescence. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 226 (1–2): 75–82. doi: 10.1016/j.nimb.2004.04.187 Uda, M., S. Sassa, K. Taniguchi, S. Nomura, S. Yoshimura, J. Kondo, N. Iskander, and B. Zaghloul. 2000. Touch-free in situ investigation of ancient Egyptian pigments. Naturwissenschaften 87 (6): 260–63. doi: 10.1007/s001140050716 Uda, M., S. Sassa, S. Yoshimura, J. Kondo, M. Nakamura, Y. Ban, and H. Adachi. 2000. Yellow, red and blue pigments from ancient Egyptian palace painted walls. Nuclear instruments and methods in physics research, Section B: Beam interactions with materials and atoms 161: 758–61. doi: 10.1016/S0168-583X(99)00969-6 Vahur, S., A. Teearu, P. Peets, L. Joosu, and I. Leito. 2016. ATR-FT-IR spectral collection of conservation materials in the extended region of 4000-80 cm-1. Analytical and Bioanalytical Chemistry 408: 3373–79. doi: 10.1007/s00216-016-9411-5 Vallance, S. L., B. W. Singer, S. M. Hitchen, and J. H. Townsend. 1998. The development and initial application of a gas chromatographic method for the characterization of gum media. Journal of the American Institute for Conservation 37 (3): 294–311. doi: 10.1179/ 019713698806082750 Van Dyke, R. M. and R. Bernbeck. 2015. Alternative narratives and the ethics of representation: An introduction. In: R. M. Van Dyke and R. Bernbeck (eds), Subjects and narratives in archaeology. Denver, CO: University Press of Colorado, 1–26. Vandenabeele, P., R. Garcia-Moreno, F. Mathis, K. Leterme, E. Van Elslande, F.-P. Hocquet, S. Rakkaa, D. Laboury,

106

BIBLIOGRAPHY

L. Moens, D. Strivay, and M. Hartwig. 2009. Multidisciplinary investigation of the tomb of Menna (TT69), Theban Necropolis, Egypt. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy 73 (3): 546– 52. doi: 10.1016/j.saa.2008.07.028 Varone, A. and H. Béarat. 1997. Pittori romani al lavoro. Materiali, strumenti, techniche: evidenze archeologiche e dati analitici de un recente scavo pompeiano lungo via dell’Abbondanza (reg. IX ins. 12). In: H. Béarat et al. (eds), Roman wall painting: Materials, techniques, analysis and conservation: Proceedings of the International Workshop, Fribourg 7–9 March 1996. Fribourg: Institute of Minerology and Petrography, 199–214. Vartavan, C. T. de. 2007. Pistacia species in relation to their use as varnish and ‘incense’ (snṯr) in Pharaonic Egypt. Bulletin of Parthian and Mixed Oriental Studies 2: 63–92. Veldmeijer, A. J. 2010. Tutankhamun’s footwear: Studies of ancient Egyptian footwear. Norg, Netherlands: Drukware. Verri, G., D. Saunders, J. Ambers, and T. Sweek. 2010. Digital mapping of Egyptian blue: Conservation implications. Studies in Conservation 55 (Supplement 2): 220–24. doi: 10.1179/sic.2010.55.Supplement-2.220 Vogelsang-Eastwood, G. 2000. Textiles. In: Nicholson and Shaw 2000, 268–98. Warner, T. E. 2011. Synthesis, properties and minerology of important inorganic materials. Chichester: Wiley. Weatherhead, F. 1992. Painted pavements in the Great Palace at Amarna. The Journal of Egyptian Archaeology 78: 179–94. Weatherhead, F. 1994. Wall-paintings from the North Harim in the Great Palace at Amarna. The Journal of Egyptian Archaeology 80: 198–201. Weatherhead, F. 1995a. Two studies on Amarna pigments. In: B. J. Kemp (ed.), Amarna Reports VI. EES Occasional Publications 10. London: Egypt Exploration Society, 384–98. Weatherhead, F. 1995b. Wall-paintings from the King’s House at Amarna. The Journal of Egyptian Archaeology 81: 95–113. Weatherhead, F. 2007. Amarna palace paintings. London: Egypt Exploration Society. Weatherhead, F. and A. Buckley. 1989. Artists’ pigments from Amarna. In: B. J. Kemp (ed.), Amarna Reports V. EES Occasional Publications 9. London: Egypt Exploration Society, 202–39. Weatherhead, F. and B. J. Kemp. 2007. The main chapel at the Amarna workmen’s village and its wall paintings. London: Egypt Exploration Society. Weisgerber, G. 2006. The mineral wealth of ancient Arabia and its use I: Copper mining and smelting at Feinan and Timna—comparison and evaluation of techniques, production, and strategies. Arabian Archaeology and Epigraphy 17: 1–30. doi: 10.1111/j.1600-0471.2006.00253.x

Welsby, D. A. 2000. The Kawa excavation project. Sudan & Nubia 4: 5–10. Welsby, D. A. 2013. Excavations of the painted shrine and other buildings in area A (draft). London. Welsby, D. A. 2014. Kawa: The Pharaonic and Kushite town of Gematon. History and archaeology of the site. Qatar-Sudan Archaeological Project. Welsby, D. A. and I. Welsby Sjöström. 2007. The Dongola Reach and the Fourth Cataract: Continuity and change during the 2nd and 1st millenia BC. In: B. Gratien (ed.), Mélanges offerts à Francis Geus. Lille: Université Charles-de-Gaulle, 379–98. Wendrich, W. 2000. Basketry. In: Nicholson and Shaw 2000, 254–67. Wendrich, W. 2002. Moving matters, an introduction. In: W. Wendrich and G. van der Kooij (eds), Moving matters: Ethnoarchaeology in the Near East. Proceedings of the international seminar held at Cairo, 7–10 December 1998. Leiden: Research School of Asian, African, and Amerindian Studies, Universiteit Leiden, 7–12. Wendrich, W. 2010. Egyptian archaeology: From text to context. In: W. Wendrich (ed.), Egyptian archaeology. Chichester: Wiley-Blackwell, 1–14. Wenzel, M. 1972. House decoration in Nubia. London: Duckworth. Wilkinson, R. H. 1994. Symbol and magic in Egyptian art. London: Thames & Hudson. Winkels, A. 2007. Restauratorisch-naturwissenschaftliche Untersuchung von tuthmosidischen Putzen aus ‘Ezbet Helmi / Tell el-Dab‘a: Ein Beitrag zur Erforschung altägyptischer Kalkputztechnik. Ägypten und Levante 17: 273–93. Winter, J. 1983. The characterization of pigments based on carbon. Studies in Conservation 28 (2): 49–66. doi: 10.1179/sic.1983.28.2.49 Woodward, J., M. Macklin, N. Spencer, M. Binder, M. Dalton, S. Hay, and A. Hardy. 2017. Living with a changing river and desert landscape at Amara West. In: Spencer, Stevens, and Binder 2017, 225–55. Wouters, J., L. Maes, and R. Germer. 1990. The identification of haematite as a red colorant on an Egyptian textile from the second millenium B.C. Studies in Conservation 35 (2): 89–92. doi: 10.1179/sic.1990.35.2.89 Chilingarian, G. V. and T. F. Yen. 1994. Asphaltenes and asphalts 1. Developments in Petroleum Science 40A. Amsterdam: Elsevier. Young, D. 2006. The colours of things. In: C. Tilley et al. (eds), Handbook of material culture. London: Sage, 173–85. Ziermann, M. 1988. Stadt und Tempel von Elephantine: 15./16. Grabungsbericht. Mitteilungen des Deutschen Archäologischen Instituts Kairo 44: 135–41.

INDEX

Abri 17, 61, 63, 68 Acacia gum (gum arabic) 32, 33, 40, 58, 62, 67, 78 aeolian sand influx 6–9 agriculture 6–9, 66 Aksha 5 Amara East 61 Amara West architecture 5–6, 8t, 9, 10–15, 16; see also housing; tombs/tomb architecture cemeteries 5, 6, 9, 15, 16, 65–66 chronology 1, 5–6, 6t geology/environment 1, 6–9, 16–17, 65–67, 67, 69–70, 76–77, 79 housing areas 5, 6, 9–15, 11, 14, 15, 27; see also E13 (area); and Western Suburb location 1, 5–9, 7, 8 temples 5, 10, 13 West Gate 8, 15, 38, 54 Amarna 20, 23, 24, 25, 26, 27–29, 54, 57, 84–85 Amarna Stone Village Survey 28 amazonite 27 Amenhotep III 21, 26, 30 analysis, methods of (technology) 20–21, 38–40, 91–94; see also under individual entries anhydrite 21, 27, 40, 42t, 43, 47t, 48t animal fat 33 animal glues 26, 32, 58 Archaeological Survey of Sudanese Nubia (ASSN) 30–31 architecture 5–6, 8t, 9, 10–15, 16; see also housing; tombs/ tomb architecture arsenic 23, 24, 93; see also orpiment; realgar Askut 16, 31 asphalt see bitumen atacamite 25, 26, 56, 57t ATR-FTIR (attenuated total reflection Fourier-transform infrared spectroscopy) 39, 91 azurite 25, 29 backscattered electrons (BSE) images 39, 53, 93 ‘Basic Colour Terms’ (BCT) 19 baskets/basketry 16, 33, 67, 72, 73, 79, 81, 82 beeswax 26, 32, 33 Bes (deity) 16, 28 binders 32–33, 35, 58, 79, 83 *

analysis methods used for 38, 58, 67, 78 preparation of 67, 72 and trade 58, 67, 72, 78–79 bitumen 22, 31 identification/analysis of 22, 38, 40, 43–46, 44, 45t, 57t processing of 82 sources of 22, 46, 77, 86 and symbolism 82, 83–84 black (colour) 19, 66 black paint 45t, 72 on coffin fragments 15, 38, 43–46, 44, 45t on palettes 10, 29, 35, 43, 44, 45t, 46 on wall decoration schemes 12, 13, 15, 21, 27, 28, 29, 30, 31, 35, 45t, 84, 87 black pigments 10, 21–22, 29, 35, 38, 40, 43–46, 44, 45t, 57t, 77 and colour-mixing 27 on grindstones/hammerstones 13t, 29, 30 see also bitumen; carbon blue (colour) 19, 20, 24–25, 84 blue frit see Egyptian blue blue paint calcite present in 41, 54 on coffin fragments 38, 38, 54, 55t on palettes 10, 27, 31, 35, 38, 40, 53–56, 53, 54, 55t, 56, 78, 86, 93 on walls/ceilings 12, 15, 27, 28, 30, 32, 35, 38, 54, 55t blue pigments 20, 23–25, 40, 53–56, 55t cakes/lumps of 10, 24, 25, 28, 35 and colour-mixing 26–27, 30 in earths 55t, 57, 57t, 78, 79, 86 on grindstones 31 manufacture/production of 23–24, 29, 30, 54, 77–78 sources of 77–78 on wall/ceiling decoration schemes 55t see also Egyptian blue bomastic 62, 63 bronze 24, 29, 77 Buhen 5, 7, 79 building techniques (modern) 62 cakes see pigment cakes calcite 17, 21, 27, 32, 41–43, 42t, 44, 54, 80, 86, 92 and colour mixing 30, 40–41, 46, 46, 49, 69, 76

Page numbers in italics refer to figures. The letter t following a page number denotes a table.

108

INDEX

calcium carbonate 21, 24, 29, 30; see also calcite calcium copper silicate 24, 53, 54, 93 calcium sulphate 21, 23, 40, 92; see also anhydrite; gypsum carbon 21, 27, 31, 43–46, 44, 45t, 55t, 57t, 77 ceilings, painted 12–13, 28, 30, 35, 40, 52t, 84 celadonite 26 cemeteries 5, 6, 9, 15, 16, 31, 65–66 Centre de Recherche et de Restauration des Musées de France (C2RMF) 20–21 ceramics 9, 11, 12, 16, 20, 25, 28 Nubian 16, 21 see also palettes charcoal 21, 25, 30, 35, 43, 67, 78 chlorites 26, 56, 57t, 78 chrysocolla 26, 28 cinnabar 23 clays/clay minerals 26, 49, 56, 80, 82, 91; see also gir climate change 6 cobalt blue 25 coffins 15, 21, 22, 32, 71, 77 polychrome/painted 30, 31, 35–38, 38, 41, 57t, 72, 87; see also under individual paint colours colour 9, 10–15, 17, 19–20, 84–85 and reflection of status 20, 77, 84–85, 87, 89 see also polychrome decoration; and under individual colours colour mixing 26–27, 29, 30, 46, 49, 57, 80–81 copper 23, 24, 25, 26, 53, 56, 77 sources of 21, 22, 24, 26, 29 copper alloys 9, 24, 54 copper chloride hydroxide (atacamite type) 26, 56, 57, 57t, 78 copper chlorides 25–26, 56, 57t, 78 cupro-wollastonite 25 cuprorivaite 24, 53 Damboya 32 Dangeil 31, 32 date palms 33, 72 Dead Sea 22, 46, 77 decorative schemes on coffins 15, 72 and palaces 21, 26, 29, 30 and religious buildings 84–85 on walls 10–15, 14, 20, 27–30, 31, 54, 84–85 see also polychrome decoration Deir el-Bersheh 21 Deir el-Medina 24, 29, 83 deities, depictions of 16, 28, 31–32 desert environment 6–9, 65–67, 69–70 diffractograms 39–40 Djehutynakht, tomb of 21 Dokki Gel 32 dolomite 43, 43t, 69, 71, 76, 92

doum palm 58, 72–73, 78 dumping areas see spoil heaps dyeing (textiles) 23, 25 E13 (area) 6, 8, 8t, 9, 10–12, 11, 15 E12.10 6, 8t, 13t, 67 E13.3 8t, 9, 12, 13t, 42t, 51t, 52t E13.4 8t, 9, 12, 13t, 42t, 47t, 48t, 49, 50, 51t, 52t E13.5 8t, 9, 13t, 47t, 48t E13.6 8t, 9, 13t, 47t, 48t, 51t, 52t, 55t E13.7 8t, 9, 10–12, 11, 13t, 14, 15, 15, 36, 42t, 43t, 45t, 47t, 48t, 49, 51t, 52t, 54, 55t, 72, 87, 89 E13.9 8t, 9, 13t, 51t, 52t E13.14 8t, 9, 10, 11, 13t, 35, 41, 44, 45t, 47t, 48t, 49, 50, 51t, 52t, 56, 57t, 69, 72, 81 E13.17 8t, 9, 47t, 48t, 54 E13.20 8t, 9, 10, 37, 47t, 48t, 51t, 52t, 55t E13.22 8t, 9 E13.23 8t, 9, 12, 47t, 48t E13.24 8t, 9 E13.31 8t, 9, 10, 11, 13t, 45t, 46, 47t, 48t, 50, 51t, 52t, 54, 55t, 56, 57t, 68, 70–71, 72 earth pigments 17, 23, 49 blue 55t, 57, 57t, 78, 79, 86 green 30, 56, 79 red see red ochre eggs/egg whites 32–33, 58, 72 Egypt Exploration Society (EES) 5, 13–15, 71 excavation grid system 9 Egyptian blue 40, 53–54, 53, 55t, 57t, 93 and binders 29, 72 on coffin fragments 55t, 57t and colour-mixing 26–27, 29, 30, 57, 57t deterioration products of 25–26 grinding 69, 69t manufacture/production of 23–25, 28, 29, 30, 54, 56, 77–78 modern 69 in paint 10, 26–27, 29, 41, 53–56, 53, 55t, 72 on palettes 53–54, 53, 55t, 57t, 86 pigment cakes/lumps 24, 27, 28–29, 30, 54, 55t and wall decoration schemes 26, 29, 30, 31, 54, 55t, 57t Egyptian green 25, 26, 28, 29, 77, 78 el-Hassa 32 el-Kurru 31 Elephantine 27 elite, the housing of 28, 84 funerary structures/tombs 20 pigments used by 23, 78, 79 see also royalty energy-dispersive X-ray (EDS) 39; see also SEM-EDS Ernetta 6, 61, 67, 72 ethnoarchaeology 61–62, 65, 73, 75, 89

109

INDEX

faience 9, 20, 21, 24, 26, 28, 54 Fairman, H. W. 5, 13 Fayum mummy portraits 32 feldspars 56, 57t fictional/imaginative narratives, use of 75–76, 79–80, 82, 84, 85–86, 87 fresco technique 29, 30 FTIR (Fourier-transform infrared spectroscopy) 39, 40, 41, 42–43, 43t, 45t, 47t, 48t, 49, 51t, 52t, 53, 55t, 56, 57t, 65, 91; see also ATR-FTIR funerary practices/rituals 16, 31, 66, 77, 78, 83–84; see also cemeteries funerary structures see coffins; tombs/tomb architecture furniture 10–12, 14, 35, 49 GC-MS (gas chromatography-mass spectrometry) 32, 38, 40, 43, 45t, 67, 72, 93–94 Gebel Barkal 31, 32 gir 17, 43t, 62, 63, 64, 76, 79, 92 Giza 27 glass/glass-making 20, 21, 24, 25, 28 glauconite 26, 30 gneiss 16 gods/goddesses 16, 28, 82, 83–84; see also deities, depictions of goethite 22, 27, 29, 46–49 gold 5, 22 granitoids 16 grasses (use for paintbrushes) 33, 84 graves 15, 16; see also cemeteries; funerary practices/rituals Greece, ancient 16, 21, 23, 27 green (colour) 19, 20 green earth 17, 26, 30, 56, 57, 79 green frit see Egyptian green green paint 10, 28, 29, 30, 35, 56–57, 56, 57t, 78 green pigments 10, 13t, 25–26, 27, 28–29, 32, 35, 38, 56, 57t, 78, 81 and colour-mixing 26–27, 29, 30, 57 on coffins 26 on grindstones 56, 57t, 78 see also Egyptian green greywackes 16 grinding, process of 68–71, 70, 69t, 80–81, 82 and performance 80, 81, 82, 89 tools for see grindstones grindstones 10, 11, 12, 13t, 28, 35, 57t, 65, 67–71, 68, 70, 80–81; see also under individual pigment colours gum arabic see Acacia gum gyps 62 gypsum 21, 40, 62, 63, 76–77, 86 use as a binder 32, 69 and colour-mixing 40–41, 43, 49, 51, 54, 71, 93 sources of 17, 32, 69, 71, 81 see also plaster

haematite 23, 27, 28, 29, 49 hammerstones 10, 11, 12, 13t, 31, 35, 68, 69, 70, 80, 81 Hatshepsut 30 hieroglyphs 26, 31, 33 honey 33, 40 huntite 21, 23, 30, 41, 42t, 43, 57t Hyksos 5, 30 imports see trade indigo 25 indigotin 25 industries faience 24, 28, 54 glass 24, 28 pigment manufacture 23–25, 28, 29, 30, 54, 56, 77–78 plaster 17 infrared spectroscopy 38, 39; see also ATR-FTIR; FTIR interviews/interview process 61–62, 73, 75, 89 iron 26, 31, 49, 54, 93 iron oxide hydroxides 22, 46–49; see also goethite; yellow ochre iron oxides 22, 23, 51t, 52t, 58, 76, 93 red 22, 23, 30, 49, 50, 50 yellow 22, 23, 46, 46, 49, 57, 93 jarosite

22

Kahun 27 Kamose of Thebes 5 Karnak 24, 26, 32 Kawa 31, 32 Kerma 5, 7, 30, 31, 66 Keruef, tomb of 24 Khentakawes, Giza 27 Koshtamna 30 Kurgus 5, 7, 16 Kush/Kushites 5, 6, 31 labour, division of 62–64, 79–80, 83 lake pigments 23 landscape/environment 6–9, 65–67, 69–70 lapis lazuli 20, 25, 27 lazurite 25 lead white (lead carbonate hydroxide) 21 leather 9, 25 lime 21, 32 lime mortar 32 lime plaster 21, 29, 30, 31, 32 limestones 16, 21 Lisht 27 Lower Nubia 5, 6t, 30; see also Nubia madder 23, 31 malachite 25, 26, 27, 29, 56

110

INDEX

Malqata: Amenhotep III palace complex 21, 26, 30 manganese 21, 30 marble 17, 32 mastabas (benches) 10–12, 14, 35, 49 Max Planck Institut 20, 22 Meketre, tomb of 33 memory/memories 65, 70, 71, 81, 82, 83, 85, 89; see also performance Menna, tomb of 26 mercuric sulphide see cinnabar Meroe: ‘Augustus Temple’ (M.292) 31–32 metal-working 24, 54, 77 micas 26, 56, 57t mining/mines 17, 21 minium 23 Minoan paintings 30, 33 mortars 32 mud plaster 10, 12, 20, 27, 35, 54, 62, 72, 80, 83, 87 mummification 22, 83–84 Naqa 32 narratives see fictional/imaginative narratives, use of natrojarosite 22 Nebamun, tomb of 32 Neberhebef, tomb of 23 Nefertari, tomb of 32 Nile, the 8 and climate change 6 and geology 16–17, 32 and trade 5 as water source 67 Nubia 5, 6t, 7, 30–32, 78 architecture 16 ceramics 16, 21 and communities/identity 15–16, 78 funerary practices/rituals 15, 16, 31, 78 ochres 10, 20, 22, 27, 76, 79; see also red ochre; yellow ochre orange paints/pigments 23, 27, 30, 49, 93 orpiment 20, 22, 23, 26, 28, 29, 30 Osiris 19, 82, 83–84 paint production collecting materials for 65–67, 75, 76–80, 77, 81 discard from 86–87, 86 processing materials for 67–71, 75–76, 80–82, 81 paintbrushes 33, 62, 65, 72–73, 82–83, 84 painting practices 1, 72–73, 75–76, 82–84, 83 and discard 86–87, 86 see also fictional/imaginative narratives, use of paints see under individual paint colours palaces 21, 26, 29, 30 palettes 9–10, 11, 12, 27, 35, 37, 67, 71

analysis of 38, 40, 42t, 45t, 47t, 48t, 51t, 52t, 55t, 57t, 91–93 and colour-mixing 9, 80–81 as disposable items 80, 86, 87 paint samples found on see under individual paint colours plant gums found in 58, 72 shapes/materials 29–30, 35, 82 palm fibres 33, 58, 72–73, 78, 82–83, 84 pararealgar 23 paratacamite 25 performance and paint application 62–64, 73, 75, 83, 84, 85, 89 and paint production 62–64, 71, 73, 75–77, 79, 81–82, 85, 89 Petrie, William Flinders 33 Pi-Ramesse (Qantir) 24, 29–30 pigment cakes 10, 24, 28–29, 35, 54 pigments 9–10, 11, 12, 20, 29 analysis of 20–21, 38–40, 91–94 earth 10, 17, 23, 65, 66–67; see also ochres original sources of 16, 20–27, 50–51, 65–67, 75, 76–78, 85 preparation/processing of 9, 10, 12, 13t, 68–71, 80–82, 81; see also grinding, process of storage of 72, 81 synthetic 23, 24, 25, 26, 29, 77–78; see also Egyptian blue; Egyptian green see also under individual pigment colours pink paints/pigments 12–13, 23, 27, 30, 35, 49, 51t, 52t, 92t; see also red pigments plant gums 32, 40, 58, 67, 72, 81, 94 sources of 65, 67 transporting/trade in 78–79 plant resins 32, 33 plaster application process 72 gypsum 21, 31, 35, 41–43, 42t, 43t, 45t, 47t, 49, 72, 76, 87 lime 21, 29, 30, 31, 32 manufacture of 17 mud 10, 12, 20, 27, 35, 54, 62, 72, 80, 83, 87 painted 13–15, 14, 15, 27; see also under individual paint colours polarized light microscopy (PLM) 38, 39, 40, 41, 41, 42t, 43t, 43, 45t, 46, 47t, 48t, 49, 51t, 52t, 53, 54–56, 55t, 56, 57t, 65, 91 polychrome decoration 35 on architecture 12, 14, 15, 28, 31–32, 87 on coffins 30, 31, 35–38, 38, 41, 57t, 72, 87 polysaccharides 32, 38, 40 pottery see ceramics pXRF (portable X-ray fluorescence) 31, 35, 38, 46, 47t, 48t, 49, 51t, 52t, 53–54, 55t, 91 pyramids see tombs/tomb architecture

INDEX

Qantir see Pi-Ramesse Qasr Ibrim 30 quarries 17 quartz 22, 23, 24, 29, 42t, 45t, 46, 49, 50, 50, 57t Raman spectroscopy 20, 21, 29, 31 Ramses II 5 realgar 20, 23, 29 red (colour) 19 red lead 23 red ochre 10, 20, 23, 28, 29, 31, 42t, 47t, 48t, 50, 51t, 52t, 57t and colour mixing 30, 49 discard 86 grinding 69, 70, 80 origins/sources of 17, 23, 61 red paints 10, 12–13, 15, 15, 27, 28, 30, 49, 50, 51t, 52t, 93 red pigments 10, 13t, 23, 35, 49, 50, 51t, 52t red rock, sources of 17, 76 Red Sea 16, 17, 22 resins see plant resins ritual 83, 89 and funerary practices 16, 31, 66, 77, 78, 83–84 see also performance rock art 66, 67 Romans, ancient 21, 23 royalty 32 palaces of 21, 26, 29, 30 tombs of 20, 22, 31 rubbish/rubble see spoil heaps Sai Island 7, 31, 32, 61, 62 sampleite 27 sand 24, 29, 77; see also desert environment sandstones 16, 50 Nubian 17 red 50, 51t, 52t, 76 yellow 49, 76 scanning electron microscopes 39; see also SEM-EDS ‘Schamotte’ 24 schist rock 29, 67–68 secco technique 29, 30 secondary electrons images 39, 56 SEM-EDS (scanning electron microscopy with energy-dispersive X-ray spectrometry) 29, 38, 39, 40, 49, 43, 43t, 47t, 48t, 49, 51t, 52t, 53, 54, 55t, 56, 91–94, 92t, 93t Semna 5, 30–31 serpentines 26 Seti I 5 Shalfak 33 silica 23, 24, 25, 29, 49, 94 Sinai desert 21 smectites 26 sofiha 62, 64, 79 spoil heaps 9, 10, 86

111

status (social) 6, 20, 28, 71, 73, 84–85, 87, 89; see also elite, the stripes (decorative) 27, 28, 30, 72 Sudan 5, 32; see also Nubia surfaces/surface effects 20 Taweret (goddess) 28 Tell el-Dab‘a 30, 33 temples 5, 10, 13, 20, 26, 30, 31–32; see also under individual sites textiles 9, 23, 25 Thebes 20, 26 tombs 23, 24, 26, 32, 33 Theophilus 26 Thutmose I 5 Thutmose III 5, 30 Timna, Negev 21 tin 29, 54 Tombos 31 tombs/tomb architecture 6, 15, 20, 21, 22, 23, 24, 25, 26, 31, 66 G201 15, 35, 42t, 45t, 51t, 52t G216 16 G244 15, 16, 35, 38, 42t, 44, 45t, 46, 47t, 48t, 51t, 52t, 54, 55t G222 15, 35, 42t, 45t, 51t, 52t G238 15, 35, 42t, 45t, 47t, 48t G301 15, 35 G309 15, 35, 41, 42t, 45t, 47t, 48t, 51t, 52t, 54, 55t G320 15, 35 G321 77 G322 15, 35, 47t, 48t see also coffins tools for collection of raw materials 67, 79 for painting 33, 62, 65, 72–73, 82–83, 84; see also paintbrushes for pigment extraction/processing 65, 67–71, 68, 79, 82; see also grindstones; hammerstones trade 16, 73, 76, 79 in binders 58, 67, 72, 78–79 in minerals/pigments 22, 23, 32, 76, 77, 78 tragacanth gum 32, 40, 58, 67, 78 travel, methods of 65–66, 73, 76, 79–80 ‘true lapis’ see lapis lazuli turquoise pigments 20, 25, 28, 29 Ukma region 31 ultramarine 25; see also lapis lazuli Upper Nubia 1, 5, 6t, 31; see also Nubia varnishes 32, 33 verdigris 26 vermilion 23

112

INDEX

viride salsum 26 visible-induced luminescence imaging (VIL) 15, 38, 40, 54 volcanic rocks 16 walls, decoration of 10–15, 14, 27–30, 31, 54, 84–85 modern methods of 61, 62 and reflection of status 84–85, 87 and refurbishment/maintenance 87, 87 water as a solvent 21, 32, 49, 62, 67, 71–72, 78, 79, 80, 81, 84 sources of 65, 66, 67, 73, 82 for washing 54, 67 Wawat 5 West Gate 8, 15, 38, 54 Western Suburb 6, 9, 10, 12–13, 12, 35, 56–57, 78 D11.1 10, 13t, 47t, 48t, 51t, 52t, 86 D11.2 45t, 55t, 68, 68, 78 D11.7 10, 54, 55t, 56–57, 57t, 78 D12.6 12–13, 51t, 52t D12.7 10, 12, 13t, 43t, 45t, 47t, 48t, 50, 50, 51t, 52t, 54, 55t, 56, 78 D12.8 10, 13, 37, 42t, 43t, 45t, 47t, 48t, 49–50, 51t, 52t, 55t D12.9 10, 82 D12.10 53, 55t D12.11 10, 13t D13.3 10, 13t white (colour) 19

white paints on coffin fragments 15, 41, 42t, 43, 76 on palettes 10, 40, 42t, 43, 80–81 on walls/floors 9, 12, 13, 15, 27, 30, 40, 42t, 43t, 84, 87, 89 white pigments 13t, 21, 40–43, 41, 42t, 43t, 79, 80, 86 and colour-mixing 80–81, 82 white rock, sources of 65, 66–67, 76 X-ray spectroscopy see SEM-EDS XRD (X-ray diffraction) 28, 38, 39–40, 55t, 56, 57t, 93 XRF (X-ray fluorescence) 31, 32, 38; see also pXRF (portable X-ray fluorescence) yellow ochre 17, 22, 28–29, 31, 45t, 46, 47t, 48t, 49, 51t, 52t, 55t, 57t and colour mixing 26–27, 30 discard 86 sources of 17, 61 yellow paints on coffin fragments 22, 48t, 49 on palettes 10, 31, 35, 38, 41, 46, 46, 47t, 48t, 49, 93 on walls 12, 13, 15, 15, 27–28, 30, 35, 47t, 48t yellow pigments 12, 22, 23, 28, 37, 46–49, 47t, 48t and colour mixing 26–27, 30, 40, 49 on grindstones 12, 13t, 28, 35, 47t, 49 on walls 26 see also yellow ochre yellow rock, sources of 17, 76