Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Archaeological Sciences 1995: Proceesings of a conference on the application of scientific techniques to the study of archaeology
Archaeological Sciences 1995: Proceesings of a conference on the application of scientific techniques to the study of archaeology
Archaeological Sciences 1995: Proceesings of a conference on the application of scientific techniques to the study of archaeology
Ebook1,360 pages15 hours

Archaeological Sciences 1995: Proceesings of a conference on the application of scientific techniques to the study of archaeology

Rating: 0 out of 5 stars

()

Read preview

About this ebook

A huge collection of papers on scientific analysis in archaeology from a conference held in Liverpool in 1995. Papers are grouped under ten headings: Scientific analysis of Petrology ( 3 papers ); of Glass ( five papers ); of Ceramics ( 7 papers ); Metallurgy ( 9 papers ); Chronological Studies ( 9 papers ); Dendrochronological Studies ( 3 papers ); The Study of Ancient Enviroments ( 16 papers ); Remote Sensing ( 2 papers ); The Analysis of Human Remains ( 5 papers ) and Perspectives on Human Evolution and Early Hominid Artefacts ( 9 papers ). Subjects studied include Stonehenge dolerites, early Egyptian glass, Greek Archaic transport amphorae, Roman brooches, South Indian bronzes, dendrochronological datings of Viking ships, phytolith analysis from the Indus Valley, Pakistan, the microflora of the Milos catacombs, Greece and Medieval glass making technology.
LanguageEnglish
PublisherOxbow Books
Release dateJul 31, 2017
ISBN9781785708060
Archaeological Sciences 1995: Proceesings of a conference on the application of scientific techniques to the study of archaeology

Related to Archaeological Sciences 1995

Related ebooks

Archaeology For You

View More

Related articles

Related categories

Reviews for Archaeological Sciences 1995

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Archaeological Sciences 1995 - Anthony Sinclair

    Implement Petrology: 60 Years of Service to the Archaeological Sciences

    R.V. Davis

    Chair of CBA Implement Petrology Committee. Cumboots Old Farmhouse, Scalby, Scarborough, North Yorks, YO13 OPO

    Abstract

    The paper briefly sets out the contribution which implement petrology has made, and continues to make, to archaeological enquiry. The Implement Petrology Committee (IPC) has its origins in pioneer work by Keiller, Piggott and Wallis (Keiller et al. 1941) in 1936 which was later incorporated into the Council for British Archaeology at its inception in 1945. IPC members have been, and still are, at the forefront of petrological research and development on prehistoric stone tools. Enduring problems are recognised, previous successes acknowledged, and future directions explored.

    IPC Mission

    The IPC is responsible for the British Implement Petrology Programme which was one of the earliest in Europe to analyze prehistoric artefacts. It demonstrated that Neolithic societies were involved in large scale extraction, manufacture and trade in stone tools. Members of the Committee have carried forward this work continuously for the past sixty years through a series of individual and collaborative research programmes involving laboratory-based investigations, replicative studies, and fieldwork at quarry sites and manufacture sites. They have also supervised research on stone axes and their petrology; it has been particularly rewarding for the Committee to invite former research students to full Membership. Justifiably, the Committee feels a strong sense of continuity and purpose, a tradition which it is eager to further in the spirit of its founders whilst remaining responsive to contemporary needs and circumstances.

    The Past

    The petrographic analysis of stone implements has been carried out since the 1930’s when the museums of South Western England combined to organise a systematic survey of the Neolithic stone axes in their collections. Five reports illustrate the origins of the work of the present Committee (Keiller, Piggott and Wallis 1941; Stone and Wallis 1947, 1951; Evens et al. 1962; Evens, Smith and Wallis 1972). More recently, the Committee has published its results in two CBA Research Reports (Clough and Cummins 1979, 1988), providing petrological identifications, distribution maps, regionally based articles and a database for over 7,600 artefacts in the United Kingdom. These represent the great bulk of axes in museum collections and the lists are a major resource for any prehistorian seeking to understand the dynamics and mechanics of production and exchange within British Neolithic society. Many petrographic analyses appear as notes in published excavation reports, others as appendices to major surveys (for example, Davis in Cherry and Cherry 1987), others within the fabric of textbooks (for example, Bradley and Edmonds 1993), others as captions in museum displays. Since 1952 the organisation of the continuing work of sectioning and identifying axes nationally, and keeping up to date with new finds, has been the chief responsibility of the IPC. In the past the Committee thin-sectioned all axes and similar implements in non-flint materials available to it, but now work concentrates more on well provenanced and dated artefacts. This work has revealed the large-scale production and widespread dispersal of these implements and led to the recognition of sophisticated exchange systems throughout Neolithic Britain.

    The Present

    With the publication of its second Research Report (Clough and Cummins 1988) the Committee’s work might be said to have been achieved. However, it would be truer to say that this publication was only a foundation for a more refined analysis and a more subtle understanding of the petrological evidence; many of the difficulties of analysis and interpretation discussed by Davis (1985) still remain. On the petrographic side there is a need to re-assess the original groupings and re-consider many of the ungrouped rocks; on the archaeological side the typology and context of axes as cultural markers (Wright 1977) need to be more closely examined. The work of the Committee, therefore, is entering a new and important phase of consolidation and discrimination.

    A database, such as the Committee has provided, gains in value as its completeness increases and its maintenance must be a matter of steady, if undramatic work. The Committee provides a network of archaeologists and petrologists throughout the country. Anyone, excavator, curator or amateur fieldworker who finds an axe or other stone implement, is asked to contact the Archaeological Recorder for their region who can then arrange the petrological analysis of the specimen. A current list of regional Archaeological Recorders and Petrologists can be obtained from the CBA. Unfortunately a small charge has to be made to over the cost of sectioning. In the case of well-dated, excavated material, which is particularly desired, this should be a charge on the post-excavation account.

    Since 1992 the activity of the Committee has divided into three spheres: petrology (routine investigations and training workshops), monitoring the database (upgrading records), and custodianship (storage of paper and thin section archives). The Committee will continue to collaborate in all these areas with other organizations, such as university departments and museums services, the Lithics Society, the Irish Stone Axe Project and the Scottish Stone Axe Project. The Committee is also looking to extend its UK network overseas, across Europe and elsewhere. The Committee’s work will continue to generate publications and valuable information for archaeologists and the general public. A major goal is the publication of a definitive Petrological Atlas of Prehistoric Stone Implements which will consist of descriptions and photomicrographs in transmitted and reflected light of the 34 axe source rocks. It is our intention to follow up this Atlas with an electronic version suitable for home access using PC software, and on the Internet. Optical microscopy still remains the central basic scientific technique used by the IPC for the initial analysis of stone implements. But the Committee has always been keen to promote appropriately the use of more sophisticated geochemical analytical techniques to help solve petrological and archaeological problems. For example, as long ago as 1981 the Committee was using XRF and XRD data to help differentiate between similar rocks from widely dispersed implements; fifteen years ago the Committee introduced the use of back-scatter SEM techniques and microprobe analysis to investigate the sourcing and distribution of axes in the North West English Highland Zone (Fell and Davis 1988). More recently, the Committee has begun to explore the use of reflected light microscopy to help characterise opaque minerals.

    The Future

    Naturally, owners would prefer to avoid damage to their implements, indeed access to some exceptionally fine specimens has been denied in the past because of the destructive (but now effectively repairable) cutting or coring techniques used to obtain a rock sample for analysis. So the development of non-destructive techniques (Williams-Thorpe et al. 1995, and see also Kayani and McDonnell, this volume) is of particular interest to the Committee. Similarly, the Committee looks forward to future collaboration (Crompton, this volume) in the application of the GRF-ND Landmark Analysis Programme to study the technology and morphology of roughout stone tools. An illustrative list of future IPC-related research and development includes:

    –a study of axes and other artefact types using portable XRF and magnetic characterisation. This is pioneering research which aims to assist in the reinterpretation of current theories concerning the procurement, dispersal and use of these important cultural markers in prehistory

    –building on the application of the recently-developed laser microprobe argon-argon techniques to give a geological date for a rhyolitic tuff stone axe fragment from the Stonehenge environs. The method requires only milligramme-sized samples and gives dates of sufficient accuracy to aid in provenancing artefacts to sources, as well as information on the heating histories of samples

    –IPC data is held on computer by English Heritage (and the CBA). Access to the database is normally through the IPC, not because we wish to restrict its use, but because, like all specialist databases, guidance about its use is advised, especially when its data is to be interpreted by researchers not familiar with its scope and limitations. Work which aims at improving data quality and accessibility by developing a PC-compatible facility will continue, facilitating improved local access to regional and national implement petrology data-sets using standalone PCs and the Internet. Textual and numeric data will be available for individual implements and, in some cases, a photomicrograph of the rock in thin section, location map and scale drawings will be provided.

    –historically, research on prehistoric stone implements from SW England has formed the backbone of the development of implement petrology nationally. The present reinterpretation of implement petrology data in light of recent geological research on the origin and distribution of Cornish Greenstones and extrusive igneous rocks (Floyd, Exley and Styles 1993) suited to stone tool manufacture will continue.

    The Graidlwyd Axe Factory: a case-study

    The IPC poster illustrated a range of IPC activities of which excavation and petrographic research at the Graiglwyd Axe Factory, Penmaenmawr by two IPC members, D. Jenkins and J. Ll Williams, is illustrative of work associated with the IPC.

    The stone axe factories of Penmeanmawr are the source for the most important stone implement group in Wales (Group VII). They exploit the eastern chilled margin (pyroxene granophyre) of a large intrusion (hypersthene microdiorite). The Graiglwyd site was in fact one of the first of its type to be surveyed and examined in the British Isles. The discovery was made in 1919 by Hazzledine Warren and excavations were conducted by him over the next two summer seasons during which five prehistoric working floors and numerous sites with struck stone flakes were identified. Unfortunately few records of the excavations themselves remain and no substantial investigations have since taken place.

    The present investigation therefore arose partly from an acute awareness of our own ignorance of the Graiglwyd axe making complex, the realisation that one of the most important Neolithic sites in Britain had been for so long ignored and neglected, and partly to assess the claims made by some archaeologists that the greater part of the site had been destroyed by quarrying. Warren’s principal site on the northern scree slope was re-excavated in 1993 when an undisturbed baulk was located under a wall containing a roughout, abundant flakes and small amounts of charcoal; dates are awaited.

    A detailed EDM survey was recently carried out which identified several new areas of potential quarrying/axe-manufacture. Of these the most significant site is a quarry in a rock outcrop on the crest of Graiglwyd, associated with flakes and roughouts. There are also high concentrations of struck flakes in adjacent mound-like features which appear to be unique to Graiglwyd. This site, on which excavations were started in 1993, is the first of its type to be identified at Graiglwyd, since all the others investigated by Warren exploited loose scree rather than rock outcrop. Other aspects of axe-quarrying, such as variations in petrology, the weathering of the rock, and the palaeo-environmental context, are also currently under investigation. The IPC believes that such reinvestigations will lead to a cautious reinterpretation of archaeological theories taking account of new evidence properly interpreted by those with appropriate knowledge, skill and understanding of the data and how its was gathered.

    Figure 1. Possible early quarry sites at the Graiglwyd axe factory. (Repr. by permission of the Gwynedd Archaeological Trust).

    Parallel investigations by IPC members concern Group XII and the Stonehenge ‘Bluestones’, Group I and the Cornish ‘Greenstones’, the Irish Group IX porcelanite, and the Scottish Group XXII riebeckite felsite. The outcomes of these and other investigations will be communicated, in time, in the usual way.

    In conclusion, implement petrology was one of the earliest scientific material analysis programmes set up, well before the main archaeological science programme began in the 1960’s. It provided a powerful indicator of the non-parochial nature of British Neolithic societies. The British Implement Petrology Programme, together with the work of Professor Goit in Brittany, acted as the model for European-wide petrological analysis of Neolithic and Bronze Age stone artefacts. Analyses are under way in Mediterranean coast-lands (variscite, jadeite), central Europe (amphibolite, schist, basalt), north-east Europe (serpentine), the Low Countries (jadeite, phanite), Scandinavia (dolerite), France and neighbouring countries (dolerite, fibrolite); and the Scottish and Irish Stone axe surveys are each adding valuable new information to the archaeological record. Members of the IPC have provided generations of archaeologists with primary research data for more than sixty years. Although membership of the IPC is by invitation only, the Committee welcomes expressions of interest in its work from archaeologists and petrologists in UK and abroad.

    Acknowledgements

    The author wishes to gratefully acknowledge the help received from Pat Phillips, Joan Taylor and other IPC colleagues and especially to David Jenkins and John Williams who provided the Graiglwyd case study material reported in this paper, and to the Gywnedd Archaeological Trust.

    References

    Bradley, R. and Edmonds, M. 1993. Interpreting the Axe Trade: Production and Exchange in Neolithic Britain. Cambridge, Cambridge University Press.

    Clough, T.H.McK and Cummins, W.A. (eds) 1979. Stone Axe Studies, Volume 1. London, CBA Research Report No. 23.

    Clough, T.H.McK and Cummins, W.A. (eds) 1988. Stone Axe Studies Volume 2. London, CBA Research Report No. 67.

    Davis, R.V. 1985. Implement Petrology: the state of the art – some problems and possibilities, in TheArchaeologist and the Laboratory (ed. P Phillips). London, C.B.A. Research Report No. 58, pp. 33–35.

    Davis, R.V. 1987. An analysis of prehistoric lithic material fromthe limestone uplands of Cumbria between Shap and Ash Fell. In Prehistoric Habitation Sites on the Limestone Uplands of Eastern Cumbria (eds J. Cherry and P.J. Cherry). Cumberland & Westmorland Antiquarian and Archaeological Society Research Series 2, pp. 77–84

    Evens, E.D., Grinsell, L.V., Piggott, S. and Wallis, F.S. 1962. Fourth Report of the Sub Committee of the South-Western Group of Museums and Art Galleries on the Petrological Identification of Stone Axes. Proceedings of the Prehistoric Society 29: 209–266

    Evens, E.D., Smith, I.F. and Wallis, F.S. (1972). Fifth Report of the SubCommittee of the South-Western Group of Museums and Art Galleries on the Petrological Identification of Stone Axes. Proceedings of the Prehistoric Society 38: 235–275.

    Fell, C.I. and Davis, R.V. 1988. The petrological identification of stone implements from Cumbria, in Stone Axe Studies (eds T. Clough and W. Cummins). London, C.B.A. Research Report No. 67, pp. 71–77.

    Floyd, P.A., Exley, C.S. and Styles, M.T. 1993. Igneous Rocks of South-West England, Geological Conservation Review Series, Chapman and Hall.

    Keiller, A., Piggott, S. and Wallis, F.S. 1941. First Report of the SubCommittee of the South-Western Group of Museums and Art Galleries on the Petrological Identification of Stone Axes, Proceedings of the Prehistoric Society 7: 50–72

    Stone, J.F.S. and Wallis, F.S. 1947. Second Report of the SubCommittee of the South-Western Group of Museums and Art Galleries on the Petrological Identification of Stone Axes, Proceedings of the Prehistoric Society 13: 7–55

    Stone, J.F.S. and Wallis, F.S. 1951. Third Report of the SubCommittee of the South-Western Group of Museums and Art Galleries on the Petrological Identification of Stone Axes, Proceedings of the Prehistoric Society 17: 99–158.

    Williams-Thorpe, O., Potts, P.J., Webb, P.C. Tindle, A.G. and Jones, M.C. 1995. Non-destructive, field-portable methods of analysis for archaeological stone. Paper presented at the conference, on Archaeological Sciences 1995, Liverpool, July 1995.

    Wright, R.V.S. (ed.) 1977. Stone Tools as Cultural Markers: Change Evolution and Complexity. Canberra, Australian Institute of Aboriginal Studies, Prehistory and Material Culture Series No. 12.

    The Potential of Scanning Electron Microscope Techniques for Non-destructive Obsidian Characterisation and Hydration Rim Dating Research

    Peter I. Kayani and Gerry McDonnell

    Department of Archaeological Sciences, University of Bradford, Bradford BD7 1DP

    Abstract

    The scanning electron microscope (SEM) is one of the most versatile, and widely available, instruments for non-destructive microstructual analysis. This paper assesses the potential of a number of SEM techniques that can be used to analyse archaeological obsidians, with particular reference to determining the morphological, chemical and crystallographic characteristics of included microcrystalline phases present in obsidian, together with their relative and absolute abundances and sizes and also the resolution of hydrated glass (perlite) zones. Preliminary results on obsidians from known archaeological sources in the Mediterranean area are reported. These results suggest that SEM-based approaches offer a powerful alternative to existing non-destructive characterisation methods, particularly for combined obsidian characterisation and hydration rim research.

    Introduction

    In a recent paper, Williams-Thorpe (1995) identified one of main reasons for the limited progress in Old World obsidian lithic studies since the 1980s as arising from the lack of a rapid, low-cost, generally available and non-destructive sourcing method which would allow provenancing to become a routine part of post-excavation work. In addition, the lack of use of obsidian artefact hydration rim dating in the Mediterranean and the Near East compared to the U.S.A was also highlighted (Willams-Thorpe 1995: 219). This paper explores the potential of the scanning electron microscope for non-destructive obsidian characterisation and for combined characterisation and hydration rim dating research.

    The only reliable high-precision method for whole rock rare earth element (REE) analysis that can be used on as-received artefacts rather than samples is non-destructive instrumental activation analysis (INAA) (Kayani et al. 1994). However, non-destructive INAA requires highly specialised, and therefore, generally inaccessible, instrumentation (Warren 1973). The most widely available non-destructive method for whole rock REE analysis is wavelength-dispersive x-ray fluorescence (WDS-XRF) analysis (Kayani et al. 1994). However, high-precision WDS-XRF is dependant on sampling several grams of an artefact (Bouey 1991; Williams-Thorpe 1995) and is, therefore, generally inappropriate for analytical studies of archaeological obsidian (Kayani et al. 1994).

    Attempts to develop alternative non-destructive and low-cost instrumental methods for whole rock analyses have only been partially successful (Kayani et al. 1994). Work at Bradford University in the 1980s using natural obsidian β particle emission (Leach and Warren 1981) and whole rock magnetic analysis (McDougall et al. 1983) indicated that these parameters showed significantly less between-source variation than reported for REE concentrations (e.g. Aspinall et al. 1972; Hallam et al. 1976; Francaviglia 1984) resulting in relatively poor source discrimination, and in some cases, source overlap. Such methods are only likely to be useful in terms of reducing the number of whole rock REE analyses required (Williams-Thorpe 1995)

    A more promising technique, recently pioneered by Burton and Krinsley (1987) for obsidian from the South-western United States, is back-scattered electron (BSE) petrography. As BSE-petrography requires only a conventional scanning electron microscope (SEM), one of the least specialised and therefore generally available analytical instruments, it is among the most accessible and non-destructive char-acterisation methods available to geoarchaeologists.

    Background

    Although amorphous on a macro scale, obsidian typically contains a significant component of microscopic and submicron sized crystallite phases (Burton and Krinsley 1987). These crystallites comprise of minerals incorporated from surrounding rocks, primary phenocrysts and microphencrysts developed in the magma chamber, post-eruption crystallites that precipitated as a result of the rapid cooling of the magma or in the subsequent glass, and the products of sub-solidus processes such as devolatilzation and devitrification (Kirkpatrick 1975; Swanson 1977). The numbers, sizes and textures of these crystallites result from a unique set of kinetic factors such as the degree of superheating (which affects the number of seed nuclei and melt polymerisation) and supercooling (which affects chemical and thermal diffusion) (Kirkpatrick 1975; Swanson 1977) and are therefore likely to characteristic of a particular obsidian source.

    Obsidian surfaces in contact with atmospheric water vapour gradually hydrate with time. A thin hydrated-glass rim is formed at a rate mainly determined by the composition of the obsidian and average ambient temperature. The thickness of the hydration rim is therefore time-temperature dependant and so the measurement of the hydration rim (using artefact thin-sections) formed on obsidian artefacts from the same site gives a relative indication of time since last working. Using known-date and known-composition standards (i.e. artefacts which have been dated by a absolute method such as radiocarbon dating) average ambient temperatures for a site can be determined and therefore absolute dates can be obtained from known-composition obsidian artefacts (e.g. Ericson 1988).

    Whole Rock Analysis

    SEM’s have been increasingly used in recent years for the x-ray microanalysis of synthetic glasses using either WDS or EDS detectors (e.g. Henderson 1988; Verita et al. 1994). The advantages of SEM-WDS or SEM-EDS are principally the ability to distinguish and analyse, or exclude, inclusions, opacifiers and weathered areas (Verita et al. 1994). Additionally, elemental analysis using the SEM is more rapid than other conventional methods and is also non-destructive in that it allows multiple analyses to be made on the same sample (Verita et al. 1994).

    The WDS and EDS detectors have significantly different analytical performances for the analysis of glasses. It is known that EDS detectors suffer from inferior peak resolution and overlap (e.g. between aluminium and magnesium) compared to WDS detectors resulting in poorer precision (Verita et al. 1994). Minimum detection limits for EDS detectors are significantly worse than those for WDS detectors due to a poorer signal-to-noise ratio. While generally performing better than EDS detectors, WDS detectors suffer from the need to use a higher electron beam energy that results in the migration of sodium ions away from the point of analysis, a potential source of error in analyses (Henderson 1988; Verita et al. 1994).

    The poor sensitivities of both WDS and EDS detectors to REE’s (Verita et al. 1994), found to be the most effective elemental discriminators of obsidian sources (e.g. Francaviglia 1984), is probably the main reason for the limited use of SEM-based microanalysis for obsidian characterisation studies. Sample masses generally preclude obsidian trace-element analysis using the SEM. An alternative approach is to use minor and major element concentrations to distinguish between obsidian sources. Except for the element’s aluminium and magnesium, EDS and WDS detectors are broadly comparable in terms of precision and accuracy (Verita et al. 1994). However, the reported extent of obsidian between-source variation in minor and trace elements is significantly less than for the REE’s (Bird et al. 1983) and, therefore, often do not give unambiguous source discrimination (Francaviglia 1984: 330; Tykot 1994). However, major element microchemical analysis in the SEM can be useful as an initial screening procedure in distinguishing between the three broad types of obsidian, alkaline, calcalkaline and peralkaline. Also, since the rate of the development of hydration rims is dependant on obsidian composition, the major element chemistry of an artefact is required in order for an absolute date to be determined by this method (Ericson 1988). SEM-based major element analysis, therefore, has a potential role in obsidian hydration rim dating.

    Backscattered Electron Atomic Number Contrast (BSE-Z) Imaging

    BSE-petrography uses reflected (typically 10–30 keV) (e.g. Burton 1987; Meeks 1988) high-energy electrons, the number of which, collected by a BSE detector in an SEM, is proportional to the mean atomic number of the phase under the incident electron beam. This dependence on mean atomic number produces atomic number contrast images where the grey-scale of the image is a function of relative atomic number differences between phases. A relatively high mean atomic number phase such as iron oxide therefore appears ‘brighter’ than a phase of relatively low mean atomic number such as quartz in the same image (e.g. Figure 1). This compositional imaging ability of BSE allows the study of chemical and textual relationships among and within minerals on a centimetre to micrometre scale.

    The back-scatter coefficient (η) is defined as the ratio of the total BSE current received by the BSE detector to that of the incident SEM electron beam and for pure elements its relationship to atomic number (Z) for elements greater than Z=10 is given by equation 1 and is graphically illustrated in Figure 2.

    For single phase specimen’s (i.e. single crystals) η is constant and the image consists of uniform contrast. For poly phase specimens (i.e. polycrystalline materials), η is variable and the image consists of different contrasts with higher atomic number phases being brighter.

    Table 1: Typical major element compositional ranges for obsidians from different sources world-wide and the associated measurement precision of SEM-EDS and SEM-WDS (modified from Cann 1983 and Verita et al. 1994).

    Figure 1. The characeristic semi-concentric and aurora-like sub-micron iron-oxide (bright) flow-banding present in obsidian from the Gabellotto flow on the island of Lipari is clearly differentiated from the relatively low mean atomic number host-glass (dark). Note: areas of relatively less dense sub-micron iron-oxide particles appear less bright than denser areas. (BSE image 2000x magnification, accelerating voltage 20 kV, beam current 10 nA, working distance 13 mm, carbon-coated sample).

    For compounds, the BSE emission originates from a number of elements that have been combined to form the material.A weighted mean back-scatter coefficient η* is usually employed (Hall and Lloyd 1981) which is defined by equation 2.

    where C(i) is the concentration by weight of the i th element in the compound, η(i) is the elemental BSE coefficient and n the number of elements. There is also an associated mean atomic number Z* for the compound that is defined by equation 3, where N, Z and A are respectively the number of atoms, atomic number and atomic weight of each element (Σ(NA) is therefore the molecular weight).

    A computer program MEATNO, written in MicroSoft QuickBasic for IBM compatible computers, is available to calculate η, Z, Z* and η* from inputted chemical formulae (e.g. SiO2, CaCO3, etc.)

    Minerals that belong to solid-solution series (e.g. feldspars) have a range of compositions. Each composition typically has different Z* and η*, resulting in variable Z-contrast images.

    For minerals, the available BSE-Z signal is typically <20% of the total possible signal (Hall and Lloyd 1981). Thus, some form of signal and/or image processing is often necessary (i.e. gamma correction).

    The maximum atomic number resolution (Rz) obtainable depends on the contrast (δ) between adjacent phases (e.g. A and B). For η*

    A

    η*

    B

    , (δ) is defined by equation 3 (Hall and Lloyd 1981).

    In general, Rz varies between 0.1Z and 0.01Z (Hall and Lloyd 1981; Burton 1987; Meeks 1988). However, resolution can be degraded by specimen surface topography (Hall and Lloyd 1981), adjacent phases and the configuration of phase boundaries. Resolution may be improved by using lower accelerating voltages (i.e. lower beam penetration) and by better electron gun configurations (e.g. LaB6, field emission) and/or BSE detectors (e.g. Meeks 1988). However, due to potential anomalous effects introduced into image contrasts by beam penetration and consequent expansion of the BSE emission volume across mineral boundaries, care should be taken in the interpretation of BSE-Z images.

    In general, spatial resolution is determined by changes in the shape of the BSE emission volume. Scattering or divergence of incident electrons increases with penetration (i.e. increased accelerating voltage or low target atomic number), leading to an increase in the emission volume and the surface area of BSE emission, resulting in a decrease in spatial resolution. However, the emission volume will decrease as target atomic number increases, resulting in improved spatial resolution. Typically, BSE-Z spatial resolution is 0.1–0.4 microns (Burton 1987).

    BSE emission is highly directional. As a target is tilted with respect to the vertical beam, the angle of beam incidence decreases. Both electron penetration and scattering are decreased, and BSE emission becomes preferential in direction. Although, in principal, this should result in improved spatial resolution, in practice, it can have serious consequences for atomic number contrast (Hall and Lloyd 1981). Also, rough specimen surfaces produce a large effective variation in tilt angles. This results in ‘uncontrolled’ scattering and BSE emission and a decrease in the signal-to-noise ratio due to the effective suppression of the BSE signal. Thus, flat specimen surfaces, orientated normal to the beam direction are usually used in BSE-Z imaging.

    Figure 2. The relationship between the mean back-scatter coefficient and mean atomic number for common obsidian-forming minerals and for the feldspar, mica, amphibole and pyroxene solid-solution series which have variable compositions and therefore a range of back-scatter efficiencies. In a BSE image, therefore, ilmenite appears brighter than quartz and micas appear brighter than feldspars. (Modified after Burton 1987).

    Although the BSE-Z image depends on composition and shows relative variations in composition, it does not provide information on the specific elemental variations, so an energy-dispersive x-ray analyser (EDX) is used simultaneously with BSE analysis. When EDX data are combined with the BSE image, individual microcrystalline phases can be identified and variations within phases can be examined. With experience, however, crystal morphology and grey level on BSE images are often sufficient for unambiguous identification (Burton 1987).

    Backscattered Electron Channelling Pattern (BSE-ECP) Analysis

    The use of the SEM to obtain crystallographic information has not been previously reported in the geoarchaeological literature, although it is rapidly gaining popularity in mainstream geology, and may provide useful additional data concerning the identify of included mircocrystalline phases in obsidian. The simplest and most commonly used technique is electron channelling pattern (ECP) analysis (e.g. Goodhew 1975). Most (i.e. typically >95%) of the BSE signal is a result of atomic number contrast. However, when imaging crystalline phases, a contribution to the total BSE signal is made by the interaction between incident electrons and the structure of the crystal. This is known as the electron channelling effect and is a result of the range of electrons within a target specimen (i.e. <500 nm) significantly exceeding the atomic period. Incident electrons are therefore channelled between rows of atoms. The atomic packing density of crystals varies with crystal direction. Thus, the interaction between a crystalline phase and the incident electrons also varies with angle of beam incidence. For a high packing density, most phase-electron interactions occur near the surface and hence the probability of BSE emission is high. For a low packing density, near surface phase-electron interactions decrease and the probability of BSE emission decreases. Thus, BSE emission, and therefore image contrast (δ), is modulated by the crystal structure of the target phase.

    Image Analysis

    Additional information on the identity of the microcrystalline phases and on the proportions of each phase in a BSE-Z contrast image can be provided by SEM-based image analysis. The BSE-Z grey-level spectrum consists of a series of intensity peaks corresponding to the brightness (and therefore mean back-scatter coefficient of each phase in the image. If the intensity values and compositions of two phases are known (or one and the zero intensity position) then the mean back-scatter coefficient of other known intensity peaks can be calculated (Hall and Lloyd 1981). The area under each peak is proportional to the amount of that brightness level in the image and therefore to the to the concentration of phase responsible for the peak (Hall and Lloyd 1981; Owen and Day 1994). Further textual information as to shape and size variations among crystallites may also be obtained using SEM-based image analysis systems.

    Sample Preparation

    The optimal sample preparation protocols for quantitative x-ray microchemical analysis, BSE-Z contrast imaging, BSE-ECP analysis and image analysis, are identical. Obsidian specimens of <5 centimetres in diameter are embedded in cold-mounting epoxy resin using an appropriate standard mould of 2.5 or 5.0 centimetres in diameter. A small area on the specimen (typically between 0.1 and 1 square centimetres depending on the density of characteristic features) (Burton 1987; Kayani 1995; Kayani and McDonnell, forthcoming) can then be exposed by abrading the epoxy disc using silicon carbide polishing papers on an automatic polisher (e.g. MINIMET). This area can then be further polished using successively finer grades of diamond abrasive down to 0.25 micron.

    After applying a thin layer of carbon to suppress charging, the epoxy disc can then be placed in the specimen chamber for BSE-Z imaging. After the examination, the obsidian specimen may be retrieved by immersing the epoxy disc in an acetone bath. Typical immersion times are 12–24 hours depending on temperature.

    For combined obsidian characterisation and hydration rim measurement in the SEM, the procedure is virtually identical to that used for the conventional hydration rim measurement, which uses an optical microscope. Standard obsidian thin-sections are prepared from artefacts, but instead of a cover-slip, the exposed thin-section is coated with a thin layer of carbon to make it suitable for SEM examination. Existing obsidian thin-sections can be adapted for SEM examination by removing the cover-slip by xylene immersion and subsequently coating the exposed thin-section with carbon.

    If only BSE-Z contrast imaging and qualitative x-ray microanalysis are required, unembedded specimens may be examined by spot-polishing a similar area and mounting the whole specimen on a 2.5 centimetre stub (e.g. using Bluetack) (Burton 1987). The specimen should be positioned in such a way as to ensure that the polished area is in approximately the same plane as the surface of the stub. The mounted specimen is then carbon-coated and placed in the specimen chamber (Burton 1987). The orientation of the polished area can then be adjusted using the SEM controls to optimise BSE-Z contrast.

    Preliminary Results

    A preliminary BSE-Z contrast examination of twenty-four embedded diamond polished geological samples from the main archaeological sources of obsidian from the Mediterranean region (Lipari-Gabellotto, Lipari-Aqucacaldra, Sardinia-Conca Cannas, Palmarola, Pantelleria, Melos and Giali) revealed that obsidian from these sources typically contain large numbers of crystallites (typically 5–30 microns in diameter). These inclusions could be imaged and characterised (by qualitative EDX) and variations in composition, texture, and abundance, which could not be detected optically using reflected light microscopy, could be detected using back-scattered electron imaging.

    Significant differences in the type of crystallites present in geological samples from different Mediterranean sources were apparent that suggested that artefactual material from these sources may be characterised (Kayani and McDonnell, forthcoming). For example, there are two potential sources of prehistoric obsidian on the island of Lipari; the Gabellotto flow and the older Aqucaldra flow. A limited fission-track dating study of both geological samples from Gabellotto and Aquacaldra and artefactual material from southern Italy suggested that the Gabellotto flow was the exclusive source of obsidian from Lipari (Bigazzi and Bonadonna 1973). Whole rock REE analysis by NAA failed to adequately resolve these two flows (Hallam et al. 1976) and subsequent whole rock Mossbauer (Longworth and Warren 1979) and magnetic analysis (McDougall et al. 1983) gave ambiguous results. BSE-Z contrast imaging of obsidian from the two flows, however, indicates that obsidian from the Gabellotto flow is characterised by the presence of semi-concentric and aurora-like sub-micron iron-oxide flow-banding (fig. 1) which is absent in obsidian from Aqucacaldra and all other obsidian sources examined to date (Kayani and McDonnell, forthcoming). Therefore, it should be possible to readily distinguish between the two possible sources of Lipari archaeological obsidian by BSE-petrography, which is significantly less destructive of the artefact than the fission-track method.

    Preliminary results also support the potential use of BSE-petrography in combined obsidian characterisation and hydration rim dating studies. BSE-Z imaging has been shown to clearly resolve lower mean atomic number hydrated glass (perlite) zones surrounding vesicles, from the non-hydrated, higher mean atomic number, host-glass (fig. 3). The examination of obsidian thin-sections in the SEM would therefore allow the simultaneous precise measurement of the artefact’s hydration rim and characterisation using BSE-petrography.

    Figure 3. BSE-Z contrast image of a large vesicle present in obsidian from the Rocche Rosse flow on the island of Lipari. A 100-150 micron zone of hydrated glass (perlite) surrounding the vesicle is clearly visible as a relatively darker (relatively low mean atomic number) band against the lighter (relatively high mean atomic number) host-glass. The small bright phases visible are iron-rich phases. Carbon-coated sample, accelerating voltage 15 kV.

    Conclusions

    One of the main reasons for limited progress in obsidian lithic studies since the 1980s has been due to a lack of appropriate characterisation methods. The SEM offers the potential for low-cost and virtually non-destructive characterisation and therefore may significantly improve both access to museum collections and increase the number of analyses carried out. In addition the SEM examination of artefact thin-sections allows the simultaneous determination of the major element composition and petrography of the obsidian and the resolution and measurement of the hydration rim. SEM-based approaches, therefore, would seem to have considerable potential for combined obsidian characterisation and hydration rim dating programs.

    References

    Aspinall, A., Feather, S.W. and Renfrew, C. 1972. Neutron activation analysis of Aegean obsidians, Nature, 233, 242.

    Bigazzi, G. and Bonadonna F.P. 1973. Fission track dating of the obsidian of Lipari Island (Italy), Nature, 242, 322.

    Bird, J.R., Duerden, P. and Wilson D.J. 1983. Ion beam methods in archaeology and the arts, Nuclear Sience Applications: Section B, 1 (5), 357–526.

    Bouey, P.D. 1991. Recognizing the limits of archaeological applications of non-destructive energy-dispersive x-ray fluorescence analysis of obsidians, Materials Research Society Symposium Proceedings, 185, 309–320, Materials Research Society, USA.

    Burton, J.H. and Krinsley, D.H. 1987. Obsidian provenance determination by back-scattered electron imaging. Nature, 326, 585–587.

    Burton, J.H. 1987. Selected petrological applications of back-scattered electron imaging. Unpublished PhD dissertation, University of Arizona, Tempe, USA.

    Cann, J.R. 1983. Petrology of obsidian artefacts, in The Petrology of Archaeological Artefacts (eds D.R.C. Kempe & A.P. Harvey), 227–255, Oxford University Press, Oxford, UK.

    Erickson, J.E. 1988. Obsidian hydration rate development, in Material issues in art and archaeology, (eds. E. Sayre, P. Vandiver, J. Druzik and C. Stevenson). Proceedings of a Symposium of the Materials Research Society, Remo, Nevada, 1988, Materials Reseach Society, 123, Pittsburgh, Pennsylvania.

    Francaviglia, V. 1984. Characterization of Mediterranean obsidian sources by classical petrochemical methods, Preistoria Alphina, 20, 311–332.

    Goodhew, P.J. 1975. Electron Microscopy and Analysis, The Wykeham Science Series, Wykeham Pububications, London.

    Hall, M.G. and Lloyd, G.E. 1981, The SEM examination of geological samples with a semiconductor back-scattered electron detector, American Mineralogist, 66, 362–368.

    Hallam, B.R., Warren, S.E. and Renfrew, C. 1976. Obsidian in the West Mediterranean: Characterisation by neutron activation analysis and optical emission spectroscopy, Proceedings of the Prehistoric Society, 42, 85–110.

    Henderson, J. 1988. Electron probe microanalysis of mixed-alkali glasses, Archaeometry, 30, 77–91.

    Kayani, P.I., Pollard, M. and McDonnell, G. 1994. Non-destructive sourcing of Old World obsidians, in Abstracts: Science and archaeology; Towards an interdisciplinary approach to studying the past, October 14–16, Harvard University, Cambridge, Massachusetts.

    Kayani, P.I. 1995. Obsidian characterisation by back-scattered electron (BSE)-petrography, Unpublished Msc Disseration, University of Bradford.

    Kayani, P.I. and McDonnell, G. forthcoming. An assessment of back-scattered electron petrography as a method for distinguishing Mediterranean obsidians, Archaeometry.

    Kirkpatrick, R.J. 1975. Crystal growth from the melt: a review, American Mineralogist, 60, 798–814.

    McDougall, J.M., Tarling, D.H. and Warren, S.E. 1983. The Magnetic Sourcing of Obsidian Samples from Mediterranean and Near Eastern Sources, Journal of Archaeological Science 10, 441–452.

    Meeks, N. 1988. Back-scattered electron imaging of archaeological materials, in Scanning Electron Microscopy in Archaeology, (ed. S.L. Olsen) British Archaeological Reports International Series, 452, 23–44.

    Leach, B.F. and Warren, S.E. 1981. Neutron activation of New Zealand and Oceanic obsidians: towards a simple screening technique, in Studies of Pacific stone resources (eds. F. Leach and J. Davidson), 151–66, British Archaeological Reports International Series, 104, Oxford.

    Longworth, G. and Warren, S.E. 1979. The Application of Mossbauer Spectroscopy to the Characterisation of Western Mediterranean Obsidian, Journal of Archaeological Science, 6, 179–193.

    Owen, J.V. and Day, T.E. 1994. Estimation of the bulk composition of fine-grained media from microchemical and back-scatter-image analysis: application to biscuit wasters from the Bow factory site, London, Archaeometry, 36, 217–226.

    Swanson, S.E. 1977. Relation of nucleation and crystal-growth role to the development of granitic textures, American Mineralogist, 62, 966–978.

    Tykot, R.H. and Hartshorn, K. 1994. The source of corse obsidian: Provenience analysis based on major element chemistry, in Abstracts: Science and archaeology; Towards an interdisciplinary approach to studying the past, October 14–16, Harvard University, Cambridge, Massachusetts.

    Verita, M., Basso, R., Wypyski, M.T. and Koestler, R.J. 1994. X-ray microanalysis of ancient glassy materials: a comparative study of wavelength dispersive and energy dispersive techniques, Archaeometry, 36, 241–252.

    Warren, S.E. 1973. Geometrical factors in the neutron activation analysis of archaeological specimens, Archaeometry, 15, 115–122.

    Williams-Thorpe, O. 1995. Obsidian in the Mediterranean and the Near East, Archaeometry, 37, 217–248.

    Detailed Provenancing of the Stonehenge Dolerites using Reflected Light Petrography: A Return to the Light

    R.A. Ixer

    School of Earth Sciences, University of Birmingham, Edgbaston, Birmingham, B15 2TT, U.K

    Abstract

    Recent work, largely based upon lithogeochemistry, has provenanced eleven Stonehenge bluestone dolerites to named outcrops on Mynydd Preseli – namely Stonehenge monoliths SH33, 37, 43, 49, 61, 62, 65 and 67 to outcrops within the Carn Menyn – Carn Gyfrwy or Cerrig Marchogion – Carn Goedog areas; SH44 and 45 to the Carn Ddafad-las outcrop and SH42 to Carn Breseb. New independent provenancing based upon reflected light petrography suggests that SH33, 37, 43, 49, 65, 67 and 61 originate from within the Carn Menyn outcrop or from Carn Goedog; that SH44 and 45 come from the Carn Ddafad-las outcrop and SH42 from Carn Bica or Carn Ddafad-las. However, no good petrographical match was obtained for SH62.

    Comparison of the results of the two provenancing methods shows that together they demonstrate that SH33, 37, 43, 49, 65, 67 and 61 are a very tight lithological-petrographical group from outcrops in Carn Menyn or Carn Goedog rather than those at Carn Gyfrwy or Cerrig Marchogion. Both methods are in agreement for the source of SH44 and 45. The mismatch for SH42 and no match for SH62 may be due to the greater sensitivity shown by petrography compared with geochemistry to small scale inhomogeneities within an outcrop. The results strongly support the belief that sampling of the so far unsampled Stonehenge monoliths, integrated with intensive further collecting of the Preseli dolerites, will delineate an area or areas of extraction of the bluestones and so indicate the mechanism of extraction, be it natural or anthropogenic.

    Introduction

    Provenancing lithic materials and especially axe-heads using only petrography is now regarded by many workers as discredited (Berridge, 1994). There are good reasons for this view, especially with regard to ambiguities in the Implement Petrology Committee’s axe-head groups as defined and discussed in the two volumes of Stone Axe Studies (Clough and Cummins 1979, 1988).

    The original axe-head groups were defined by professional petrographers who even in their day (1940–1975) were anachronistic but who had the experience and practice to recognise their groups and to impose self-consistent petrographical limits to them. These same people had sufficient status that they were kept abreast of developments within petrology and so were able to update the groupings. Today axe-head petrography often is the part-time occupation of non-professional petrographers who are no longer able to see all the available material belonging to a group, visit potential outcrop/factory sites or keep up-to-date with the geological and/or archaeological literature. For example, the work of Bevins and co-workers (Bevins, 1978; 1982; Bevins and Rowbotham, 1983) on the mineralogical and petrographical effects of low-grade metamorphism in Wales and their potential use in helping to discriminate between Welsh and English fine-grained volcanic and volcanoclastic rocks has not been recognised.

    Further causes for concern are the sectioning and grouping of poorly archaeologically-contexted material for ill-defined or diffuse archaeological reasons and the continued use of terms such as near Group X, close to Group VI. The definition of many groups is becoming lost.

    However, underlying all these arguments is the belief that petrography is an out-dated skill of little value that is unable to match the precision or sophistication of geochemistry, especially expensive geochemistry. But, as with very many techniques, the value of petrography lies in its being used alongside complementary methods. ‘Total petrography’, the use of reflected light methods, as well as those of transmitted light, should be the norm when using a combination of lithogeochemistry and petrography for lithic studies. This combination is still the cheapest and most effective way forward.

    Reflected light petrography and its potential use in the characterisation of lithic materials has been discussed by Ixer (1994). Here, it was suggested that for igneous rocks, the combination of magma type, cooling history, auto-metasomatism/alteration and weathering should produce an opaque mineral assemblage that, in terms of mineral species and their intergrowths, is unique, a petrographical fingerprint. Since reflected light petrography (which for most lithics is the study of iron-titanium oxides, iron-chromium oxides, and iron, copper, nickel ± zinc ± lead sulphides) describes the mineralogy of a significant number of the elements that are used in lithogeochemistry to discriminate between lithologies, Ixer further proposed that the technique could be used to distinguish between rocks of ‘identical’ chemistries.

    As a worked example, millstones manufactured from basalts from the island of Pantelleria were provenanced petrographically and the results compared to those of Williams-Thorpe and Thorpe (1990), who used lithogeochemistry as their provenancing tool. A comparison of the relative sensitivities of total petrography (transmitted and reflected light techniques) and geochemistry showed that individually both were capable of broad scale discrimination (millstones made from basalts that did not come from Pantelleria were easily identified by either method) but that for very small scale provenancing they were best used together, to act as checks upon each other. Although for the vast majority of samples there was very good agreement between the methods, for a significant, if numerically low, minority there was little or poor correspondence. This was inferred to be due to small scale heterogeneities in the rocks that have a profound petrographical but little or no geochemical effect (Ixer 1994).

    Figure 1. Sampling localities for dolerite samples on Mynydd Preseli, Pembroke.

    As a further worked example to test the relative sensitivities of petrography and lithogeochemistry and to compare this to their combined use, the altered ophitic dolerites of Implement Petrology axe-head Group XIII are chosen.

    This group comprises a small number of archaeological artefacts – the Stonehenge bluestones being the most famous, and a numerically significant assemblage within the group. Indeed it is unlikely that Group XIII would be a separate group but for them. Most examples of the group have been found within good archaeological contexts. Outcrops related to Group XIII are well-defined and geographically compact; they all crop out within a few square kilometres. The rocks have been subjected to a number of different processes and so have been the object of intense geological and petrological study (Bevins et al., 1989) based upon a very complete mineralogical and geochemical database. Only reflected light studies are missing from these data. Finally there is a well-defined archaeological reason for choosing this example – the existence of a petrographically-lithogeochemically tight group within the monoliths that can be provenanced to a single or very few outcrops might suggest human quarrying, whereas multiple-sourced monoliths hint at more natural agencies being responsible for their removal.

    Provenance Studies of the Stonehenge Monolith Bluestones

    Introduction and Background

    Thomas (1923) first proposed that the bluestones of Stonehenge were the same as rocks cropping out within the Preseli Hills of Dyfed, southwest Wales. Later, Stone and Wallis (1951) proposed that spotted dolerites (preselites) be named Group XIII, of the British Stone Implement Petrology Groups. This group is now a subgroup within Implement Petrology Group XXIII, an eclectic set of lithologies ranging from granodiorites to quartz dolerites, but all from the St David’s to Preseli area of South Wales. Group XIII is a small group of 42, mainly perforated implements, summarised in Clough and Cummins (1988), that make up approximately one percent of all grouped axes (Clough 1988).

    Most recently, Thorpe et al. (1991) have returned to the problem of the provenance of the Stonehenge monoliths. Drawing upon and extending the definitive geochemical and petrographical data and petrological discussions of Bevins (1978, 1982), and based upon new extensive sampling of South Wales’ rocks, Thorpe et al. (1991) were able to reconfirm that outcrops on the Preseli Mountains were the source of the dolerite bluestones. Furthermore, by comparing the geochemistry of the archaeological artefacts with that of naturally occurring lithologies, they were able to suggest that individual monoliths could be assigned to named outcrops. In Table 11, (p. 139) Thorpe et al. (1991) assigned monoliths 33, 37, 43, 49, 61, 62, 65 and 67 to East Preseli, in particular an area bounded by Carn Menyn-Carn Gyfrwy or Cerrig Marchogion-Carn Geodog (see Figure 8, p. 126); monoliths 44 and 45 were assigned to East Preseli Carn Ddafad-las and monolith 42 to East Preseli, Carn Breseb. Stonehenge monolith numbering is after Atkinson (1979), outcrop named after Thorpe et al. (1991).

    Although both transmitted light and preliminary reflected light data (Ixer in Thorpe et al., 1991: 150–152) were reported in the paper, the matching of Stonehenge monoliths to outcrop was essentially done geochemically.

    Sample Preparation and Method of Study

    To determine whether detailed petrography and especially reflected light petrography could further refine the provenancing of the monolith dolerite bluestones the samples obtained by Thorpe et al. (1991) were used. One polished thin section was prepared in the standard way (Ixer 1990) for each of the eleven dolerite monoliths and for the outcrop material shown in Figures 1 and 2. Two additional polished thin sections were prepared from SH45, CM2 and CM12.

    Figure 2. Outcrops samples at Preseli.

    The prepared sections were chosen at random and examined using standard reflected light techniques under X8, X16 and X40 oil immersion lenses. All phases larger than 5pm were visually identified. The presence or absence of mineral species, their intergrowths and associations were noted. Detailed modal and grain size analyses were not attempted nor was detailed transmitted light petrography other than to note gross mineralogical differences. The results were tabulated and used to provide an empirically derived set of qualitative mineralogical and textural criteria that could be used to assess comparability between sections. All the sections were re-examined using these criteria and correlations made in order to provenance the bluestones.

    Results

    Stonehenge Monoliths

    Macroscopically SH44, 45 and 62 are unspotted dolerites, whereas all the others show 1–2 cm diameter pale-coloured spots that comprise plagioclase crystals altering to fine-grained assemblages of metamorphic minerals (Bevins et al., 1989).

    In reflected light all the monoliths have very similar petrographies, sufficiently so that a generalised description based on SH33 can be given. Variations from the common petrography (SH33) were used as the basis for grouping the monoliths as follows:

    Group SH33

    This group includes SH33, 37, 43, 49, 65, 67 and SH61.

    Ilmenite and titanomagnetite were the main primary oxides forming crystals and intergrowths, up to millimetres in diameter, accompanied by trace amounts of chromite. Although some sulphides may be magmatic most pyrrhotite, chalcopyrite, and pyrite are later phases associated with metamorphism of the rocks. Titanite and TiO2 minerals, including rutile, are secondary minerals mainly replacing the primary iron-titanium oxides – they are the most abundant ‘opaque’ phases. Sulphides have altered to limonite and trace amounts of secondary copper minerals.

    Ilmenite forms discrete, lobate laths, skeletal, equant crystals or coarse-grained graphic intergrowths with pyroxene and altered plagioclase. Although relict ilmenite is present in all the sections it shows an extensive but characteristic alteration sequence of ilmenite to ‘speckled ilmenite’ to pale-coloured, twinned rutile to orange-brown TiO2. A thin white titanite rim overgrows altered ilmenite. Characteristically, an orange-coloured TiO2 mineral (? rutile) forms thin, acicular crystals forming a crystallographically orientated, triangular, trellis-like, open texture.

    Titanomagnetite forms equant to skeletal grains, up to several millimetres in diameter, that are intergrown with plagioclase and pyroxene and enclose altered plagioclase laths. Many crystals are intergrown with ilmenite to form internal and external composites (the ‘sandwich texture’ of Haggerty 1976a) and many display thin ilmenite oxidation-exsolution lamellae. These lamellae have two thicknesses, the thicker set >5 μm has altered to orange-brown TiO2 minerals, whereas the thinner >1–5 pm is now pseudomorphed by white TiO2 minerals. Relict titanomagnetite is absent, it being totally pseudomorphed by 4–20 μm diameter crystals of titanite with white internal reflections.

    Trace amounts of chromite form rounded to euhedral crystals and show slight alteration to pale-coloured, poorly characterised ferritchromit rims. Sulphides, in minor amounts, are widely distributed, either as discrete, coarse-grained crystals or more often as smaller disseminated grains, forming sulphide-rich patches, Pyrrhotite ± pentlandite exsolution flames (SH43, 65, 61) and chalcopyrite are more abundant than pyrite and marcasite which may be alteration products after pyrrhotite. The majority of the sulphides, especially the sulphide-rich patches, are associated with white mica, epidote and quartz which themselves are part of the metamorphic suite of minerals. Euhedral rhombic titanite is present in the matrix of the rocks.

    The petrography of the SH33 group is remarkably consistent, i.e. it defines a very ‘tight’ group, with only SH61 showing minor differences, for example in having very coarse-grained chromite 300 pm, and pyrite up to 200 pm in diameter.

    SH42, SH44 and SH45

    SH42 differs from group SH33 in one significant respect, namely the absence of thin ilmenite oxidation-exsolution lamellae in altered magnetite/titanomagnetite. In all other respects it is similar.

    Macroscopically both SH44 and 45 are unspotted dolerites and under the microscope they form a loose group with some similarities to SH42. Whilst much of the petrography of SH44 and SH45 is similar to that of group SH33, the rocks display the following significant differences.

    Titanomagnetite/magnetite has altered to white or pale, peach-coloured titanite devoid of ilmenite oxidation-exsolution lamellae (SH44) or locally showing very densely packed lamellae (SH45). Chromite is absent from SH44 or small and very rare in SH45 and titanite forms small, up to 10μmm diameter, anhedral grains rather than euhedral rhombic crystals. Ilmenite, altered ilmenite and sulphides have the same mineralogy and textures and approximate grain size as those seen in group SH33, other than a relative paucity of pyrrhotite in SH44 and 45 and the presence of cubanite exsolution lamellae in chalcopyrite in SH44.

    SH62

    Petrographically SH62 shows the greatest number of differences from group SH33. Titanomagnetite has altered to dense, white titanite that carries very rare, altered ilmenite oxidation-exsolution lamellae. Relict ilmenite is absent and ilmenite has altered to orange- and white-coloured TiO2 minerals and titanite.

    Dolerites from Preseli Outcrops

    Dolerites taken from the Preseli Mountains outcrops (figs 1 and 2) show a wider range in their detailed petrographical characteristics than do the Stonehenge monolith dolerites. However, it must be re-emphasised that both monoliths and outcrop samples share enough petrographical characteristics to group them all as Group XIII.

    In hand specimen Carn Ddafad-las (CM 12), and Craig Talfynydd (TF2, 3) are unspotted dolerites, Carn Marchogion (CMA1) is ‘less spotted’ and Carn Menyn (CM2) is a weakly spotted/unspotted dolerite. At all other outcrops spotted dolerites were sampled (Thorpe et al., 1991).

    Carn Menyn (CM1, CM2, CM11) and Carn Gyfrwy (CM10)

    Although the outcrops lie within one kilometre of each other, there are significant petrographical variations between these dolerites especially in the textures and alteration of their iron-titanium oxides.

    Millimetre-sized titanomagnetites carry two generations of ilmenite oxidation-exsolution lamellae along (111), (CM 10, CM11). The thinner lamellae (>2 pm) have altered to white TiO2 minerals, whilst the thicker, 5–20 pm but up to 40pmm in CM 10, are replaced by brown-coloured TiO2. In CM1, however, the oxidation-exsolution lamellae are thin and rarely present and are absent from CM2. The titanomagnetite host is altered to loosely-packed white titanite except for CM2 where fine-grained ilmenite is present in darker cores within pale titanite pseudomorphs after titanomagnetite and may be relict ilmenite from internal composite crystals or late-stage neomorphic ilmenite.

    In all the sections magnetite and tabular ilmenite form internal and external composite crystals. The alteration of ilmenite varies between the outcrops but in all cases relict ilmenite is present. Relict ilmenite is surrounded by blue- coloured TiO2 (rutile); that in turn is altered to titanite (CM1, CM2); or is accompanied by a highly anisotropic? carbonate phase and then alters to orange-coloured TiO2 (CM11); or is accompanied by fine-grained haematite (CM 10). The presence of haematite as part of the alteration of ilmenite is unique to CM 10 in the present study despite, elsewhere, it being a very common and widespread mineral replacing ilmenite (Ixer 1990). The fine-grained nature of the haematite-TiO2 intergrowths suggest that they may be a more altered form of ‘speckled ilmenite’.

    All the rocks, except for CM1, carry large, rounded to euhedral chromite grains that show dissolution features and ferritchromit alteration rims. All carry up to 300 pm diameter disseminated sulphide patches comprising chalcopyrite or chalcopyrite and pyrite, these patches are associated with secondary silicates, including quartz and titanite. Small chalcopyrite or hexagonal pyrrhotite grains ± pentlandite exsolution flames cement acicular box-work TiO2 crystals within altered ilmenite or form discrete inclusions in epidote. Chalcopyrite alters to digenite, covelline and ‘idaite’ and pyrrhotite to pyrite/marcasite along (0001); all the sulphides are oxidised to limonite. Discrete rhombic crystals of titanite are present in all the dolerites.

    Carn Goedog (CGD 1, 2), Carn Breseb (CBR1) and Carn Bica (CB1)

    These three outcrops, each approximately one kilometre apart (fig. 1), share many petrographical features.

    Millimetre diameter titanomagnetite has altered to white or pale, lime-green coloured titanite although the cores of some altered titanomagnetites are darker and carry small magnetite/ilmenite grains (CBR1). Ilmenite oxidation- exsolution lamellae, 1–2 pm in thickness and altered to white TiO2 and accompanied by thicker lamellae that have altered to pale orange TiO2 minerals, are common except for Carn Bica (CB1). Here thin lamellae are rare to absent.

    All the rocks carry the same internal and external composite ilmenite-titanomagnetite intergrowths and most show the same type and degree of ilmenite alteration. Relict ilmenite, altering to ‘speckled ilmenite’, is surrounded by pale-coloured TiO2 minerals accompanied by a highly anisotropic mineral and these in turn alter to a triangular mesh- work of pale orange, acicular TiO2 crystals, all this is enclosed within a titanite rim. In CDG2, relict ilmenite is absent, it is totally pseudomorphed by orange-coloured TiO2 minerals and titanite. Chromite, up to 100 pm in diameter, is present in CB 1 but is smaller in CBR 1 and CDG1/2, but in all cases shows oxidation to ferritchromit.

    Disseminated chalcopyrite or hexagonal pyrrhotite grains form sulphide patches associated with secondary silicates and both minerals cement acicular TiO2 crystals within altered ilmenite. Chalcopyrite alters to spionkopite and pyrrhotite to limonite along (0001) planes. Euhedral, rhombic crystals of titanite are present in all the rocks and trace amounts of pyrite are present in CBR1 and CGD1.

    Carn Marchogion (CMA2), and Craig Talfynydd (FT 2, 3)

    Despite the distance between the outcrops of about two kilometres, their rocks share many characteristics.

    In all sections magnetite, now altered to white titanite, carries rare to very rare 1–2 μm oxidation-exsolution lamellae now altered to white TiO2 minerals; a second generation of very rare, thicker lamellae, some of which retain relict ilmenite, is present in CMA2 and TF2.

    Ilmenite and titanomagnetite form internal and external composite mixed crystals. Relict ilmenite alters to TiO2 minerals and then to titanite in CMA2 and TF2 but the more extensive alteration sequence of ilmenite to ‘speckled ilmenite’ to pale TiO2 to orange TiO2 plus a titanite rim is seen in the rocks from Craig Talfynydd (TF3). Chromite is present in these same rocks but not seen in CMA2.

    All the sections show abundant sulphides notably hexagonal pyrrhotite often intergrown with chalcopyrite. Although disseminated sulphide patches associated with later metamorphic minerals are present, the majority of the sulphides are discrete and coarse-grained. Pyrrhotite ± pentlandite exsolution flames is intergrown with lesser amounts of chalcopyrite and pyrite and all are oxidised to limonite. Titanite is present as small, poorly crystalline aggregates and as euhedral rhombs (TF2, 3).

    Carn Ddafad-las (CM 12)

    Three polished thin sections were made and they show some petrographical variation.

    Titanomagnetite has altered to dense, white titanite, some with darker coloured cores. White TiO2 pseudomorphs after thin ilmenite oxidation-exsolution lamellae are locally present but more commonly they are very rare or absent.

    Ilmenite shows the typical alteration sequence of relict ilmenite, altering to TiO2 minerals and a highly anisotropic phase, themselves altering to orange-coloured TiO2 and enclosed within a titanite rim. Rare, partially dissolved,

    Enjoying the preview?
    Page 1 of 1