You are on page 1of 16

,

,i ~ ~
. +

,I
Sedimentary Geology 90 (1994) 241-256

SEDIMENTARY GEOLOGY

ELSEVIER

Identifying provenance-specific features of detrital heavy mineral assemblages in sandstones


A n d r e w C. M o r t o n , C l a i r e H a l l s w o r t h
British GeologicalSurvey, Keyworth~Nottingham NG12 5GG, UK

(Received July 30, 1993; revised version accepted November 11, 1993)

Abstract The composition of heavy mineral assemblages in sandstones may be heavily influenced by processes operating during transport, deposition and diagenesis. As a consequence, conventional heavy mineral data may not be a reliable guide to the nature of sediment source material. Certain features of heavy mineral suites, however, are inherited directly from the source area without significant modification, such as the varietal characteristics of individual mineral species. This paper describes an alternative approach to varietal studies that concentrates on relative abundances of minerals that are stable during diagenesis and have similar hydraulic behaviour. Ratios of apatite to tourmaline, T i P 2 minerals to zircon, monazite to zircon, and chrome spinel to zircon provide a good reflection of the source rock characteristics, because they are comparatively immune to alteration during the sedimentary cycle. These ratios are described as index values (ATi, RZi, MZi and CZi, respectively). This approach avoids some of the practical problems associated with varietal studies, such as the need to make subjective decisions about "mineral properties or to use advanced analytical techniques that may not be accessible to the analyst. It also makes use of more components of the heavy mineral suite and thus provides a more balanced view of provenance characteristics. The use of these ratios is illustrated with examples from Upper Jurassic sandstones in the Outer Moray Firth area of the UK continental shelf and Triassic sandstones from onshore and offshore UK. Heavy mineral indices, notably ATi and MZi, show marked variations in Upper Jurassic Piper sandstones of the Outer Moray Firth. Main Piper sandstones have lower ATi and MZi values compared with Supra Piper sandstones, indicating significant stratigraphic evolution of provenance. The UK Triassic shows major regional variations in a number of index values, including ATi, MZi and CZi, demonstrating that sediment was supplied from several distinct source regions. This indicates a need for some modification of existing palaeogeographic models for the UK Triassic.

1. Introduction H e a v y m i n e r a l analysis is o n e o f t h e m o s t i m p o r t a n t a n d widely u s e d t e c h n i q u e s in t h e determination of the provenance of sand and sandstone. A large n u m b e r o f d e t r i t a l heavy m i n e r a l s p e c i e s have b e e n f o u n d in s a n d s t o n e s ; in t h e o r d e r of fifty a r e d e s c r i b e d by M a n g e a n d M a u r e r

(1992). The value of the technique lies not only in the number of possible minerals, but also in that many of them have restricted parageneses that positively identify the involvement of particular parent rocks. Because of this sensitivity in constraining s o u r c e r o c k lithology, heavy m i n e r a l analysis is w i d e l y a p p l i e d to m o d e m a n d a n c i e n t siliciclastic s e q u e n c e s to identify s e d i m e n t p r o v e -

0037-0738/94/$07.00 1994 Elsevier Science B.V. All rights reserved SSDI 0037-0738(93)E0129-4

242

A. ('. Morton. C. Hallsworth / Sedimentary Geology 90 (1994) 241- 25~

nance and dispersal patterns and to aid the correlation and differentiation of sand bodies.
1.1. Factors influencing heacy mineral assemblages

Despite their sensitivity to sediment provenance, heavy mineral assemblages do not only reflect the composition of the source material. Several processes have the ability to alter relative abundances of heavy minerals during the sedimentary cycle (Morton, 1985). These processes operate during weathering, transport, deposition and diagenesis (Fig. 1). Rocks at outcrop in the source terrain are subject to a weathering regime that may alter the composition of the heavy mineral assemblage prior to its incorporation into the transport system. There are many well-documented examples of mineral destruction in soil profiles (see Bateman and Catt, 1985), yet for the most part present-day rivers contain rich and diverse heavy mineral suites (e.g. Russell, 1937;

Shukri, 1949; van Andel, 1950; Morton and Johnsson, 1993) indicating that on a gross scale weathering does not significantly affect the diversity of mineral assemblages incorporated into transport systems. This may be because the vast majority of such studies have been concerned with weathering-limited denudation regimes, in which the transport processes removing weathered material from an area are potentially more rapid than the weathering processes generating the material (Johnsson et al., 1991). Studies of transport-limited regimes, in which the maximum weathering rate exceeds the ability of transport processes to remove material, may reveal a stronger weathering control on assemblages, because in such circumstances weathering products have a longer time to react with soil and groundwaters. During transport itself, minerals are subjected to abrasion processes that could affect the relative abundances of minerals with different me-

WEATHERING IN SOURCE AREA

Fig. 1. Schematicillustration of the processesthat operate during the sedimentarycycle to overprint the mineralogicalsignature of the source rocks.

A.C. Morton, C. Hallsworth/ Sedimentary Geology 90 (1994) 241-256

243

chanical stabilities. Despite experimental work that enabled the determination of relative mechanical stabilities for detrital minerals (Friese, 1931; Thiel, 1940, 1945; Dietz, 1973), there is little evidence that abrasion processes play a significantly role in natural situations (Morton and Smale, 1990). A potentially more important process during transport is the effect of alluvial storage. In fluvial systems, abundances of unstable minerals may be reduced during periods of weathering during alluvial storage on the floodplain, particularly in humid tropical environments. Johnsson et al. (1991) showed the importance of this process on bulk sand mineralogy, and subsequent work has shown that the process may also be effective in modifying aspects of the heavy mineral suite (Morton and Johnsson, 1993). The most important factors, however, are hydraulics and diagenesis. The hydrodynamic conditions at the time of deposition affect the relative abundances of minerals with different hydraulic properties. Variations in hydraulic behaviour result from differences in the density, size and shape of heavy mineral grains. Rubey (1933) was the first to conclusively demonstrate that changes in hydraulic conditions introduce marked variations to heavy mineral suites, and there has been extensive work on this topic subsequently (see review by Mange and Maurer, 1992). However, the effects of diagenesis are arguably the most profound, with post-depositional processes causing partial or complete loss of many heavy mineral species. Several case studies have shown that with increasing burial, heavy mineral suites respond to the associated increases in pore fluid temperatures by losing unstable minerals through dissolution (Morton, 1984; Milliken, 1988; Cavazza and Gandolfi, 1992). For example, Late Paleocene sandstones of the Central North Sea, which contain apatite, amphibole, epidote, garnet, kyanite, rutile, staurolite, titanite, tourmaline and zircon at shallow depths, progressively lose amphibole, then epidote, titanite, kyanite and staurolite as burial increases, resulting in an assemblage at the basin centre of apatite, corroded garnet, rutile, tourmaline and zircon (Morton, 1984). Heavy mineral suites also lose minerals through dissolution under the influence of low-temperature acidic

groundwaters (e.g. Friis, 1976; Morton, 1984, 1986), for example in fluvial sandstones or below subaerial unconformities. Therefore, although the effects of all these processes have still not been fully evaluated, there is abundant evidence that, either in isolation or in combination, they can modify heavy mineral assemblages to the extent that the nature of the source material is no longer apparent. Furthermore, they may mask original differences between heavy mineral assemblages, so that sandstones derived from different source rocks can no longer be distinguished, and they may introduce variations to assemblages of sandstones that were derived from the same source.
1.2. Varietal studies

In order to counteract the effects of the overprinting factors, it is necessary to identify and quantify those features of heavy mineral assemblages which are a true reflection of the mineralogy of the source rocks. The usual method of achieving this is to carry out varietal studies. These are studies that concentrate on the variations shown by an individual mineral or mineral group. Because this minimises the density and stability range within the population under investigation, the data obtained from varietal studies are essentially independent of variations in hydraulic and diagenetic conditions and are therefore representative of the source material. Most varietal studies have concentrated on subdividing populations on the basis of optical properties, such as colour, shape and habit. Zircon and tourmaline have both been used extensively in such studies (Krynine, 1946; Poldervaart, 1955). However, subdivision on the basis of optical properties is a subjective exercise, and boundaries between classes within the population are by necessity often arbitrarily drawn. For example, classification on the basis of grain shape involves determining categories such as angular, subangular, euhedral, subhedral, subrounded, rounded and well-rounded, and problems are inevitable in classifying some of the grains in the population. Many colour series are similarly gradational. The problems with such subjective techniques lie

244

A. C. Morton, C. Hallsworth / Sedimenta~" Geology 90 (1994) 241-256

partly in internal consistency on behalf of the analyst, but are considerably greater when comparing data from several analysts, because there can be no guarantee that the same class boundaries will be chosen by each one. Therefore, although there is still an important place for varietal studies that make use of optical properties, conclusions drawn solely on the basis of such data should be treated with caution. The recent advent of readily available microanalysis equipment, such as the electron microprobe, has enabled varietal studies to be carried out on the basis of single-grain geochemical analysis. This has considerable advantages over optical methods of varietal studies, because it provides objective numerical data that can be treated with considerably more confidence and because it provides a basis for direct comparison with source rock mineralogy. The method can be applied to a wide variety of detrital heavy minerals, both translucent (Morton, 1991) and opaque (Grigsby, 1990; Basu and Molinaroli, 1991), as many species show some degree of compositional variation. However, it may not always be practicable to undertake single-grain geochemical studies. For example, equipment may not be easily accessible, there may be constraints on time, and the mineral suites under investigation may not be suitable. In addition, because varietal studies concentrate on a single mineral type, they only provide information on the provenance of that particular mineral, causing a biased view of the provenance of the sediment as a whole. For example, the varietal study may concentrate on a mineral that is confined to one specific rock type in the hinterland. By concentrating on this one mineral, there is a danger that information from the remainder of the heavy mineral suite, which includes elements from all source rock types, may not be taken fully into account.

analysis of heavy mineral suites under the petrographic microscope) in constraining sand provenance. These criteria must involve features of the conventional heavy mineral data that are directly acquired from the source area and are not fractionated during the sedimentary cycle. One approach that meets these requirements is the determination of the relative proportions of minerals that behave in a similar way during the processes of transport, deposition and diagenesis. They should, therefore, have similar mechanical and chemical stability and similar hydraulic behaviour. However, as mechanical abrasion does not appear to play an important role in altering heavy mineral assemblages, variations in mechanical stability can probably be safely ignored. Therefore, the important considerations are the stability of heavy minerals in diagenesis and their hydraulic behaviour.

Chemical stability'
To avoid the possibility that variations in relative abundance of heavy minerals have resulted from differing diagenetic conditions, only proportions of minerals that are stable under the diagenetic regime should be considered. Therefore, in order that unstable mineral suites can be compared with their deeply buried equivalents, particular attention should be given to the stable component of the suite. Minerals known to be stable under conditions of high pore fluid temperatures (as found in deep burial) are apatite, the TiO 2 polymorphs (rutile, anatase and brookite), tourmaline and zircon (Morton, 1984; Mange and Maurer, 1992). These species are also stable in acidic groundwaters, except apatite, which is readily dissolved under such circumstances. Monazite, chloritoid and spinel are generally considered to be stable (Morton, 1984), although Mange and Maurer (1992) suggest on the basis of unsubstantiated observations that chloritoid and spinel are less stable than garnet. However, work in progress by C. Hallsworth on North Sea sandstones shows that monazite and two types of spinel, gahnite (the zinc-bearing spinel) and chrome spinel, occur in an uncorroded state in association with highly etched garnet in a variety of sandstones, including Middle

2. Heavy mineral ratios Despite the major advances made recently in varietal techniques, it is vital that criteria are established for the use of conventional heavy mineral data (that is, the data acquired by optical

A.C. Morton, C. Hallsworth/ Sedimentary Geology90 (1994) 241-256

245

Jurassic sandstones of the northern North Sea and Lower Cretaceous sandstones of the O u t e r Moray Firth. Furthermore, uncorroded monazite and chrome spinel grains are present at burial depths exceeding 3800 m in both Late Carboniferous sandstones of the southern North Sea and Triassic sandstones of the Central North Sea. Monazite and spinel can therefore be considered stable during deep burial, at least in North Sea sandstones. Work in progress has also demonstrated that chloritoid grains in association with highly corroded garnet display minor etching, and can be considered to be more stable than garnet during deep burial. Therefore, monazite, spinel minerals and chloritoid can be used to constrain provenance, although the distribution of chloritoid must be considered suspect in very deep burial situations. This is important, because unlike apatite, TiO 2 minerals, tourmaline and zircon, spinels, monazite and chloritoid have more restricted parageneses that provide greater constraints on the nature of the source lithologies. Garnet, the most stable of the common heavy minerals proven to undergo dissolution, may also be used in many circumstances, but caution is required if it displays surface textures indicative of corrosion. Garnet tends to disappear from heavy mineral assemblages in permeable North Sea sandstones between 3000 m and 3500 m because of dissolution (Morton, 1986). In zones where garnet dissolution is relatively advanced, therefore, caution is needed when ascribing variations in garnet abundance to changes in provenance. In many instances, variations in abundances of minerals such as staurolite, kyanite, titanite, epidote, amphibole and pyroxene cannot be confidently ascribed to changes in source because of their instability, although if present they are useful indicators of the nature of the source area.
Hydraulic behaviour Although there is considerable debate about the hydraulic behaviour of heavy minerals (see Mange and Maurer, 1992), there is no doubt that mineral grains with similar hydraulic properties behave in a similar fashion. The hydraulic behaviour of heavy minerals is controlled by grain

Table 1 Density and typical detrital grain habits of the more stable heavy minerals Mineral Habit Density Mid-range range value Tourmaline Equant/prismatic 3.03-3.25 3.14 Apatite Equant/prismatic 3.10-3.35 3.23 Chloritoid Platy 3.51-3.80 3.66 Anatase Equant 3.82-3.97 3.90 Garnet Equant 3.58-4.32 4.10 a Brookite Equant/prismatic 4.08-4.18 4.13 Rutile Equant/prismatic 4.23-5.50 b 4.23 b Gahnite Equant 4.62 4.62 (spinel) Zircon Equant/prismatic 4.60-4.70 4.65 Chrome Equant 4.43-5.09 4.76 spinel Monazite Equant 5.00-5.30 5.15 Density values are taken from Deer et al. (1982) and Mange and Maurer (1992). a The mid-range value for garnet is taken assuming a composition of 70% almandine and/or spessartine and 30% grossular and/or pyrope. b The 5.50 value refers to rutile with high Nb and/or Ta contents. These elements are only rarely major components of rutile and thus the low density value of 4.23 has been cited as the mid-range value in this case. size, density and shape. The shape factor controls the relative abundance of minerals with different habits, so that minerals with equant or prismatic habits behave differently from those with bladed or platy habits. Minerals with bladed or platy habits tend to be hydraulically equivalent to smaller quartz grains compared with their equant or prismatic counterparts (Flores and Shideler, 1978). This behaviour is taken to extremes with mica; sand-size mica particles, although denser than quartz, are hydraulically equivalent to quartz grains of silt size (Doyle et al., 1983). Thus, for provenance evaluation, best results will be achieved by comparing proportions of minerals with similar habits. In this circumstance, the relative abundance of heavy mineral grains with similar size and density will be unaffected by changing hydraulic conditions and will, therefore, provide a good reflection of the source area characteristics. Examination of the range and means of the densities of the stable heavy minerals (Table 1) shows that there are several pairs of commonly

246

A,C. Morton, C. Hallsworth / Sedimentary Geology 90 (1994) 241-256

Table 2 Density differentials of mineral pairs Mineral ratio Zircon/chrome spinel Tourmaline/apatite Anatase/rutile Rutile/zircon Zircon/monazite Garnet/zircon Tourmaline/monazite Tourmaline/zircon Density ratio 0.98 0.97 0.92 0.91 I).90 0.88 0.61 0.67 Density differential (%) 2 3 8 9 10 12 39 33

The ratios of the densities of several of the mineral pairs are given in Table 2. The density differentials of these ratios are small, ranging from 2% (chrome spinel and zircon) and 3% (apatite and tourmaline) to a maximum of 12% (zircon and garnet). These values are significantly less than the density differential across the entire heavy mineral spectrum (tourmaline to zircon, 33%; tourmaline to monazite, 39%), and can therefore be considered to provide a better reflection of source area characteristics.

Data calculated using mid-range density values from Table 1.

2.1. Determination of ratio values


occurring minerals that fulfil the criteria for similar hydraulic behaviour. These are apatitetourmaline, TiP2 minerals-zircon, monazitezircon and chrome spinel-zircon. The T i P 2 polymorphs anatase, rutile and brookite are also likely to behave in a similar hydraulic fashion to one another. Gahnite-zircon also fulfils the criteria, but the general scarcity of gahnite suggests that the ratio of these two minerals may have limited application. Chloritoid has a density that falls midway between the "heavier" heavy minerals (TiP 2 group, zircon, spinel group and monazite) and the "lighter" heavy minerals, but its platy habit is likely to cause it to behave more like the "lighter" group than the "heavier" group. If abundances of this mineral are used for provenance evaluation, they should therefore be used in comparison with either apatite or tourmaline; however, because of its higher density and its platy habit, the reliability of this ratio as a provenance indicator is open to doubt, and should be used with caution.
Table 3 Determination of provenance-sensitive index values Index ATi GZi RZi CZi MZi Mineral pair Apatite, tourmaline Garnet, zircon TiP 2 group, zircon Chrome spinel, zircon Monazite, zircon Index determination 100 x 100 100 100 x 100 apatite count/(total apatite plus tourmaline) garnet count/(total garnet plus tourmaline) TiO a group count/(total TiO 2 group plus zircon) chrome spinel count/(total chrome spinel plus zircon) monazite count/(total monazite plus zircon)

The above discussion shows that the aspects of conventional heavy mineral data that most closely reflect the nature of the source area are ratios of stable minerals with similar hydraulic behaviour. Conventional heavy mineral data are generally calculated by undertaking a count of 200 to 300 detrital non-opaque grains from the residues (Morton, 1985; Mange and Maurer, 1992). Although this approach gives a good characterisation of the entire mineral suite, the errors associated with the abundance of individual minerals are significant, particularly for minerals present in small amounts. Consequently, if mineral ratios are calculated using these data, the errors associated with each value are likely to be considerable. In order to produce more accurate data that can be used with confidence to constrain provenance, each individual ratio must be determined on the basis of separate counts; a minimum of 100 grains per mineral pair is required, and a total of 200 should be the aim. To avoid the problems of

RZi is determined here using the total of the TiP 2 polymorphs, because many analysts do not distinguish the three minerals during the counting procedure, particularly if identification is made using energy-dispersive X-ray analysis rather than by optical methods! If separate counts of rutile, anatase and brookite are made, the ratio could be determined using rutile alone. This would have some advantages over the use of the entire TiP 2 group (see text).

A.C. Morton, C. Hallsworth / Sedimentary Geology 90 (1994) 241-256

247

infinite values that would be obtained from a normal ratio determination should the second mineral of the pair be absent from the assemblage, the ratios are given in the form of index values scaled from 0 to 100, as shown in Table 3.

MINERAL A
f I

MINERAL B J;--I,~ME~M;FINE
-I
I I I I
i

r-i i i

I I

Selection of grain size fraction


Although heavy mineral analysis is a widely practised technique, there is no standardised procedure with regard to which fraction of the sand should be analysed. Different analysts recommend widely differing approaches (see discussion by Mange and Maurer, 1992). These fall into two basic categories: the analysis of a relatively narrow size fraction, and the analysis of the entire residue or a wide size range. It is beyond the scope of this paper to discuss the relative merits of the two methods for conventional heavy mineral analysis and interpretation, but it is important to consider which approach is more suitable for the determination of provenance-sensitive heavy mineral ratios. For this purpose, it is crucial that differences in sediment grain size, which reflect variations in hydrodynamic conditions at the time of deposition, are not reflected by differences in the ratio values. If the hydrodynamic behaviour of heavy minerals was the only control on their grain size distribution, neither approach would have any particular merit in terms of calculating the ratio values, because minerals with similar hydraulic behaviour would have the same grain size distribution and would respond in the same way to varying hydraulic conditions. In such circumstances, therefore, the ratios of hydraulically equivalent heavy minerals would not vary according to sediment grain size. However, hydrodynamic behaviour is not the only control, because heavy mineral species inherit a grain size distribution from their parent rocks. Certain heavy minerals, notably rutile and zircon, tend to occur as relatively small grains, whereas others, such as garnet and tourmaline, are frequently found as large crystals (Milner, 1962; Mange and Maurer, 1992). Consequently, there is a potential for changing hydraulic conditions to cause variations in the ratios of hydraulically equivalent minerals if a wide size range is employed for the ratio determination. "
-

t r-t
I

I
r

i l

E "6 == ,so

2,,

34

II
34

11 f
E
0

II

34

34

Grain size (0 units)

Fig. 2. Idealised example to show the possible problems caused by determination of ratio values from the entire sand fraction. Mineral A and mineral B have identical hydraulic behaviour, but mineral A is available in all size grades in the source, whereas mineral B only occurs as grains less than 125 /~m in diameter. The histograms show the grain size distribution of the two minerals in a medium-fine sand and a very fine sand. The ABi value calculated from the entire sand fraction is 80.0 for the coarser-grained case and 66.6 for the finergrained case, but is 50.0 if the calculation uses only the 63-125 ~m fraction.

This potential variability is minimised by the analysis of a narrow range, as illustrated by the theoretical example in Fig. 2. This example considers the behaviour of two minerals, A and B, which have similar hydraulic properties. If both minerals had the same size range, their relative proportions would be unaffected by changes in grain size. However, although mineral A is available as large grains in the source terrain, producing a wide range of grain sizes in the sediment, mineral B has a more restricted size distribution in the parent rock, with a considerably smaller maximum size. This is likely to be the case with, for example, tourmaline and apatite. Obviously, therefore, although mineral B has the same hydraulic behaviour as mineral A, it cannot be deposited as large grains hydraulically equivalent to those of mineral B because they are simply not

248

A.C. Morton, C. ttallsworth / Sedimentary Geology 90 (1994) 241-256

present in the transport system. Coarse-grained sediment and fine-grained sediment from the same source cannot, therefore, have the same ratios of the two minerals if a wide size fraction is analysed. If, however, analysis is carried out on a narrow size interval across which grains of both minerals are available in the parent rocks, variations in grain size will have no effect on the ratio. It is, therefore, recommended that for the determination of provenance-sensitive heavy mineral ratios, analysis is carried out on the very fine sand fraction (63-125 ~m), as this fraction is finer than the maximum size of the typically finegrained heavy minerals such as zircon, which rarely exceeds 250 izm (Poldervaart, 1955) and rutile, which rarely exceeds 200/~m (Mange and Maurer, 1992). It is also important to stress that ratio values are likely to vary considerably with the chosen grain size range, because of the grain size distributions inherited from the parent rocks. Therefore, comparisons of ratios determined on different size fractions may not be valid, and a standardised approach is advisable.

2.2. Possible problems with determination and interpretation of mineral ratios Identification
A prerequisite to accurate determination of the ratio values is clearly the identification process. The recent publication of the colour heavy mineral identification manual by Mange and Maurer (1992) is a major step towards correct mineral identification, and accurate determination of apatite, the TiO 2 minerals, tourmaline and zircon should not prove difficult. However, monazite and chrome spinel are more difficult to identify and require a greater degree of expertise. Monazite and zircon are commonly confused, despite the former having lower relief and birefringence than zircon, together with a typical pale yellow to green colour. Use of a spectroscopic attachment (Hering and Zimmerle, 1963) or use of energy-dispersive X-ray analytical equipment should be considered where doubt exists. The difficulty in identifying chrome spinel lies in its tendency toward opacity, with many grains being identifiable only by the observation of its typical

red-brown isotropic characteristics around the thin grain margins. The tendency for grains to be opaque is obviously greater for larger grains, providing another good reason for restricting the analysis to the very fine sand fraction. It is recommended that for accurate quantification of chrome spinel, the heavy mineral residue is examined under the optical microscope with the condenser attachment in place and racked fully up, to increase the amount of light through the grains. If optical analysis is carried out without the condenser in place, there is a significant chance of failure to identify some or all of the chrome spinels. As with monazite, use of energy-dispersive X-ray analysis will provide an unequivocal basis for identification. There may also be problems in the determination of the abundance of the TiO 2 minerals, because they are commonly present as secondary phases as well as detrital grains. The ratio of TiO 2 minerals to zircon is only valuable for provenance purposes if the grain count does not include secondary grains. Anatase is the most common secondary TiO 2 mineral in sandstones, but secondary rutile may also occur. Detrital and secondary grains can be reasonably easily distinguished on the basis of optical observations of grain morphology. Because TiO 2 minerals occur both in detrital and secondary forms, use of energy-dispersive procedures to quantify their abundance is not recommended.

Secondary overgrowths
The tendency for TiO 2 minerals to occur as secondary phases causes a further problem with the determination of the RZi value, because detrital TiO~ grains commonly develop secondary overgrowths. Overgrowth increases the diameters of such grains; consequently, many of those that fall in the analysed 63-125/~m fraction are likely to have been deposited in the < 63 tzm fraction, and grains that originally had diameters between 63 ~ m and 125/xm may now fall into the coarser size fraction. There is little that can be done to counteract this problem, although it might be partially alleviated by counting rutile alone, because anatase and brookite appear to be more susceptible to secondary overgrowth. In cases

A.C. Morton, C Hallsworth / Sedimentary Geology 90 (1994) 241-256

249

where secondary TiO 2 development is common, it is probably prudent to avoid placing great importance on fluctuating RZi values, because they may not represent changes in sediment source. The same problem may be found with the ATi value on rare occasions, because apatite is also known to develop overgrowths, although much less commonly and generally at considerable burial depths (Mange and Maurer, 1992).

fluvial or fluvio-deltaic sandstones deposited by large river systems in humid tropical environments, however, changes in ATi value may not reflect changes in provenance but rather the degree of alteration on the floodplain. In such circumstances, the ATi value cannot be used to prove changes in provenance; additional evidence, from other mineral ratios or varietal characteristics, is required.

Apatite dissolution
Although apatite is stable in deep burial conditions, it is susceptible to dissolution by acidic groundwaters. There is, therefore, a possibility that the ATi values may not accurately reflect source area characteristics, as the value may have been modified during weathering. Loss of apatite during near-surface weathering conditions could take place in five different settings: (i) in the source terrain, (ii) during periods of alluvial storage, (iii) at the depositional site, in fluvial or fluvio-deltaic systems, (iv) at subaerial unconformities, and (v) at outcrop. Careful selection of samples should eliminate the last factor, and zones of apatite depletion below subaerial unconformities and in fluvial or fluvio-deltaic sandstones can be readily identified by undertaking regular detailed heavy mineral analyses through the stratigraphic section. Weathering at source presumably reduces the proportion of apatite to tourmaline, but in weathering-limited denudation regimes the degree to which this occurs is likely to be small. There is evidence, however, that significant apatite loss may occur during alluvial storage on floodplains of major rivers in humid tropical climates. This is particularly likely in fluvial systems with low-lying extensive floodplains, because this gives rise to repeated periods of prolonged storage. A modern analogue is the Llanos area of Colombia and Venezuela, the floodplain of the Apure and Orinoco rivers, where ATi values are markedly reduced once transport distances exceed 200 km (Morton and Johnsson, 1993) Therefore, in marine sediments and in terrestrial sediments from arid or semi-arid climates, changes in ATi value can be ascribed to changes in source with some degree of confidence In

3. Case examples
The use of heavy mineral ratios as indicators of changing provenance is illustrated with two
~ l ~ /

"'3 -~i'!!iiii:i::::::::" .e:':" . . . . . . : : : : : . ' : , ' "


-, . . . . . : ................

- 58 40'N

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

===========:::::::::=1== ::::::::::::::::::::::::

~::::-::::::::;:: . . . . . . . :::

QUARTZ ARENITIC SOURCE

IZZZZZZ.2~
..... ---...............

~5i?~:::::::::::'
i::::::::::::::::. . . ...........
::::::::::::::::::::::

:::::::::::::::::::::

~:i:i:i:i:]iii!!iii
I-

ii i ii::,LZZZZ_ ....
t------

! =========================. . . . . ~...~ . . . . . . . . . . . . . . . . ,,
::::::::::::::::::::::::::::
, ................

58 o O0'N

0 o 24~W

0 D 48'E

QUARTZ ARENITE ~ SUBARKOSE ~

ARKOSE DISTAL SILTSTONE/MUDSTONE

Fig. 3. Distribution of petrographicaily distinct sandstone types in Supra Piper and equivalent sandstones of the Outer Moray Firth and their proposed source regions, adapted from O'Driscoll et al. (1990). Location of Well 15/21a-33 is shown.

250

A.C. Morton, C. Hallsworth/ Sedimentary Geology 90 (1994) 241-256

examples, one demonstrating stratigraphic changes in provenance and the other demonstrating regional variations.
3.1. Stratigraphic changes in provenance: Upper Jurassic sandstones of the Outer Moray Firth, UK sector, North Sea

The Upper Jurassic Piper Formation in the I v a n h o e / R o b Roy Field area, U K Block 15/21 (Fig. 3), contains two sandstone developments, the Main Piper and the Supra Piper (Parker, 1991). These sandstones comprise lower to upper shoreface sands with subsidiary distributary channel and washover sands, and are believed to have been deposited by a northward-prograding wavedominated delta (Boote and Gustav, 1987; Parker, 1991). O'Driscoll et al. (1990), however, indicate that another source, lying to the northeast, may also have been involved (Fig. 3), because petrographic analysis shows the presence of both quartz arenites and arkoses. The quartz arenites are interpreted as having a source to the northeast, whereas the arkoses are interpreted as having a southwesterly source (O'Driscoll et al., 1990). The sequence cored in Well 15/21a-33 recovered subarkosic Supra Piper sandstones and quartz arenitic Main Piper sandstones, and was analysed to assess whether there is any heavy mineral evidence for a change in source material. The sandstones range in grain size from very fine to

very coarse or pebbly, indicating that hydraulic conditions varied considerably. Furthermore, because the sandstones have been buried to over 2300 m, heavy mineral suites are likely to have suffered depletion of many unstable species. The combination of these factors introduced a high degree of variability to the conventional heavy mineral data. The assemblages contain five main species, which show wide variations in relative abundance: apatite, which ranges from 0.0% to 14.5%; garnet, which ranges from 3.5% to 62.0%; TiO 2 minerals, which range from 8.0% to 36.0%; tourmaline, which ranges from 1.0% to 25.0%; and zircon, which ranges from 10.0% to 64.5%. In addition, there are four subsidiary species (chrome spinel, gahnite spinel, monazite and staurolite), present sporadically and in minor amounts. Because of the scatter in mineral abundances, the conventional heavy mineral data provide little firm evidence for a difference between the Supra and Main Piper sandstones. In contrast, the provenance-sensitive ratio data show that the two sandstones had distinctly different sources (Fig. 4), with three of the four indices (ATi, MZi and RZi) showing consistent differences between the two units. In the Supra Piper sand, the ATi value exceeds 40 in 7 of the 9 samples, whereas 18 of the 19 Main Piper samples have ATi below 40. However, eight of the samples in the Main Piper have

CZi

!
I i

RZi

MZi

5o='
o o
%00

10 oe

[ ~ i o o

"
o o o

oo~
o e

o
o o I ,50 ATi

o o

o
o

0 0

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

~ 5O ATi

0 0

.......

T....... 5O ATi

O~

~ao--

-. . . .

o Main Piper

Supra Piper

Fig. 4. Differentiation of Main Piper and Supra Piper sands in Well 15/21a-33 from the Ivanhoe/Rob Roy area, UK sector, North Sea, using provenance-sensitiveratios. The two units do not overlap on the ATi-RZi and ATi-MZi crossplots, but CZi values do not provide a good basis for discrimination.

A.C. Morton, C, Hallsworth / Sedimentary Geology 90 (1994) 241-256

251

particularly low ATi values. This group includes samples associated with a tightly silica-cemented zone, interpreted as a silcrete formed in the vadose zone (Parker, 1991). It is therefore considered that the very low ATi values are the result of apatite dissolution during acidic groundwater flux, and do not represent the transported sediment characteristics. Nevertheless, even discarding the samples affected by in-situ apatite dissolution, the Main Piper can still be distinguished from the Supra Piper on the basis of its lower ATi values. A possible explanation for this is that the lower values in the Main Piper could have resulted from greater apatite loss during alluvial storage in the fluvial system, as described from the Llanos region of Venezuela by Morton and Johnsson
P

(1993). However, this is unlikely, because palaeogeographic reconstructions indicate that the transport distances must have been relatively short (Boote and Gustav, 1987; O'Driscoll et al., 1991), placing severe constraints on the opportunity for sediment storage and recycling. Furthermore, Hurst (1985) considers it unlikely that the Jurassic was a time of intensive weathering under tropical conditions in the Moray Firth area, and suggests that the climate was sub-tropical, humid and possibly seasonal. Therefore, the difference in ATi values between the Main and Supra Piper sandstones most probably reflects a change in provenance. Variations in MZi values between the two units are also clearly defined. In the Supra Piper,
P P

SUPRA

/d."
20%

'\
\

pyrope

/IT'dF~'

I ---'

p
MAIN

p ~ ~

* <5%~sartine o >5%sPessartine

/olo'o
"o "

"

\
~ G

AS

0%

Fig. 5. Garnet compositionsin Main Piper and Supra Piper sands in Well 15/21a-33 from the Ivanhoe/Rob Roy area, UK sector, North Sea, expressedin terms of the relative abundancesof the almandinc plus spcssartin (AS), pyropc (P) and grossular(G) end-members. Each ternary plot represents a single garnet population determined by electron microprobe analysisof 50 garnet grains. The majority of garnets in the Supra Piper sandshave20-30% pyropc, whereasmost of the garnetsin the Main Piper sands have pyrope over 3 0 % .

252

A.C, Morton. C. ttallsworth ./Sedimentary Geolo,g9' 90 (1994) 241- 256

7 of the 9 samples have MZi > 5, whereas in the Main Piper 15 of the 18 samples have MZi < 5. Additionally, the Supra Piper tends to have higher RZi values than the Main Piper, although there is greater overlap; 8 out 9 Supra Piper samples have RZi > 35, compared with 6 out 19 from the Main Piper. The least conclusive evidence comes from the CZi values: 5 of the 9 Supra Piper samples have CZi > 1.5, compared with 6 of the 19 Main Piper samples. Therefore, even if the ATi evidence is not considered conclusive in isolation, the variations in MZi and RZi provide strong support for a difference in provenance between the two. Further evidence for a difference between the two sandstones is provided by studies of detrital garnet geochemistry. Both the Supra and Main Piper sandstones have garnet suites dominated by almandine-pyropes with low grossular contents (Fig. 5), but in the Main Piper pyrope contents commonly exceed 30%, whereas in the Supra Piper pyrope generally lies between 20% and 30%. The change in provenance identified by the provenance-sensitive heavy mineral ratios and by the garnet compositions could have either resulted from a change in source area (possibly the change from the northeasterly source to the southwesterly source suggested by O'Driscoll et al., 1990), or alternatively could have resulted from unroofing within the same general source. In the absence of a regional coverage of heavy mineral data, it is not possible to come to a firm conclusion regarding these two possibilities.

Fig. 6. Palaeogeographic reconstruction of the Early Triassic, adapted from Audley-Charles (1970) and Cope et al. (1992), together with locations of wells used in the Triassic sandstone case example. Stippled areas show proposed highland areas. Transport directions indicated by arrows.

3. 2. Regional changes m provenance: the Triassic of onshore and offshore Britain


The use of heavy mineral ratio data to identify regional differences in provenance is illustrated here with three Triassic sandstone sections from different basins onshore and offshore Britain: Wytch Farm Borehole B22, from the Wessex Basin; British Gas Well 110/2-6, from the South Morecambe Field in the East Irish Sea Basin; and BP Well 22/24A-1, from the Central North Sea Basin (Fig. 6). These sandstones have widely differing burial histories. The top of the Wytch Farm sequence, which comprises arkosic sand-

stones (Colter and Havard, 1981) is presently at about 1640 m, and has probably not been buried much deeper than that. In contrast, the Triassic sandstones of the Irish Sea, which tend to be subarkosic (Stuart, 1993), are presently relatively shallow, but have been buried to considerably greater depths in the past. In Well 110/2-6, the top of the Triassic sandstone section is at about 940 m (Ebbern, 1981), but Jackson et al. (1987) estimated that the maximum palaeodepth to the top of the sandstone could have exceeded 3500 m prior to uplift. There is little detailed published information on Triassic sandstone petrography in

A.C. Morton, C. Hallsworth / Sedimentary Geology 90 (1994) 241-256

253

the Central North Sea, although Stewart (1986) alludes to the generally arkosic nature of Triassic sandstones in the area. The Triassic sequence in Well 22/24A-1 is probably presently at its maximum burial depth, with the top at about 3540 m. Therefore, much of the variation between the heavy mineral suites of the three areas is likely to be diagenetically induced; garnet, for example is common at Wytch Farm, rare and highly corroded in 22/24A-1, and absent in 110/2-6. Apart

from differences in the abundance of garnet, there is little difference in mineral diversity between the three areas; apatite, TiO 2 minerals, tourmaline and zircon are the only abundant heavy minerals common to all three areas, with chrome spinel and monazite present sporadically and in minor amounts. Determination of the provenance-sensitive mineral ratios produces a much clearer picture of major differences in nature of the source mate-

100RZi

50-

MZi

50-

25,-

%0
0 0 0 0 ~0 0

|
O

%"1~
NO

@ @

0
I

50 10 CZi ATi

IC 100 0

50 ATi

100

&
NI

lO_
Q. (/)

--

~J ~J ,,j

E ~sd z 0 0 T 50 ATi --~----' 1 100 0 0


I I

20

40

60

80

100

% Rutile / Total TiO~

Wytch Farm B22 Well 1 1 0 / 2 - 6 Well 22/24A-1

Fig. 7. Use of ATi, RZi, MZi, CZi and variations in TiO2 po|ymorphs to iUustratc differences in provenance of Triassic sands in the Wessex Basin (Wytch Farm B22 Borchole), East Irish Sea Basin (Wcl| 110/2-6) and the Central North Sea (Well 22/24A-1).

254

A. C Morton, C. ttallsworth / Sedimentary Geology 90 (1994) 241-250

rial. Each of the three sequences plot in discrete areas on the ATi-RZi, ATi-MZi and ATi-CZi crossplots, as well as showing different relative proportions of TiO 2 minerals (Fig. 7). ATi values are low to moderate in Well 110/2-6 (9-50), moderate to high in Wyteh Farm (45-92), and high in 22/24A-1 (90-98). The RZi values in 110/2-6 and Wytch Farm are comparable (20-37 and 23-40, respectively), but are higher in 22/24A-1 (35-67). The proportion of rutile within the TiO z group is low in 110/2-6 (14-38%), high in Wytch Farm (64-82%) and very high in 22/24A-1 (77-97%). MZi values are very low in both 110/2-6 and 22/24A-1, but are remarkably high in Wytch Farm, where they range from 14 to 52. The highest MZi value indicates that in one sample monazite contents exceed those of zircon. The high proportions of monazite in the Triassic of Wytch Farm indicates that granites or granitic pegmatites were major contributors of sediment to the Wessex Basin, but not to the East Irish Sea or Central North Sea areas. CZi values are very low in 110/2-6 and Wytch Farm, but are between 3 and 9 in 22/24A-1. This indicates that there was a significant contribution from ultramafic rocks, probably ophiolites, to the Central North Sea area, but not to the Wessex or East Irish Sea basins. Thus, not only do the heavy mineral ratio values provide a clear discrimination of the three sequences and indicate that they are each supplied from different source materials, but also provide important information regarding the nature of the source rocks.

4. Conclusions Heavy mineral abundances in sandstones do not simply reflect the composition of the source material, because they are heavily influenced by processes that act on the sediment during transport, deposition and diagenesis. This limits the use of conventional heavy mineral data to constrain sediment provenance, particularly where diagenesis has been severe. Nevertheless, heavy mineral assemblages contain some features that directly reflect properties of the source material. Previous work has shown that varietal studies,

which concentrate on variations in the properties of one or more individual mineral type, produce data that are inherited from the source terrain. An alternative approach is to determine ratios of stable heavy minerals with similar hydraulic behaviour. These parameters provide a good reflection of source area characteristics because they are comparatively immune to alteration during the sedimentary cycle. Because they are not affected by processes during and after sedimentation, they could be used to directly match sediment with source material, even for suites of first-cycle origin that have been heavily affected by diagenetic processes. Suitable ratios include apatite : tourmaline (ATi), TiO 2 minerals : zircon (RZi), monazite:zircon (MZi), chrome spinel: zircon (CZi) and the relative proportions of the TiO 2 minerals. These ratios should be determined on a narrow size band (63 ~m to 125 gm is recommended here) to avoid potential problems that arise from inheritance of grain size distributions from the parent rocks. The method has several advantages over varietal studies; it does not involve subjective discrimination of mineral types on the basis of optical properties, it avoids using electron microprobe equipment that may not be always available to the researcher, and it makes use of a wider number of species in the assemblage, thus avoiding a possibly biased view of provenance. There are a number of possible problems involved with the determination and interpretation of these mineral ratios. Identification of monazite and chrome spinel may prove difficult for the inexperienced analyst. These two minerals have specific parageneses that provide important constraints on the nature of the source rocks, and thus accurate quantification of MZi and CZi values is crucial. If there is doubt over the determination of these species during optical analysis, use of energy-dispersive X-ray analysis is recommended. The RZi value may be affected by the tendency of some of the TiO 2 minerals, notably anatase, but also futile, to occur in both detrital and diagenetic forms, a frequent occurrence being the development of secondary overgrowths on detrital grains. This problem cannot be easily overcome, and variations in RZi in sequences

A.C. Morton, C. Hallsworth / Sedimentary Geology 90 (1994) 241-256

255

with extensive TiO 2 diagenesis should be carefully evaluated before assigning them to changes in source. Alternatively, the ratio could be determined using rutile alone, rather than including anatase and brookite, because rutile is considerably less susceptible to secondary development than the other two polymorphs. Variations in the ATi value potentially reflect the extent to which sediment has been exposed to weathering; thus, this ratio should be used with care in fluvial or fluvio-deltaic sediments deposited in humid tropical climates. Objectively derived varietal data (such as garnet or tourmaline geochemical information) should be used to supplement and support the ratio data in cases of doubt.

Acknowledgements
We are grateful to Amerada Hess Ltd. and partners for permission to publish information gathered from Well 15/21a-33, and to Total UK for their financial support of the study of Triassic sand provenance. The constructive comments made by Dr. J.F. Hubert and Dr. N.H. Trewin helped greatly in production of the final manuscript. This paper is published with the approval of the Director, British Geological Survey (N.E.R.C.).

References
Audley-Charles, M.G., 1970. Triassic palaeogeography of the British Isles. Q. J. Geol. Soc. London, 126: 48-89. Basu, A. and Molinaroli, E., 1991. Reliability and application of detrital opaque Fe-Ti oxide minerals in provenance determination. In: A.C. Morton, S.P. Todd and P.D.W. Haughton (Editors), Developments in Sedimentary Provenance Studies. Geol. Soc. London, Spec. Publ., 57: 55-65. Bateman, R.M. and Catt, J.A., 1985. Modification of heavy mineral assemblages in English coversands by acid pedochemical weathering. Catena, 12: 1-21. Boote, D.R.D. and Gustav, S.H., 1987. Evolving depositional systems within an active rift, Witch Ground Graben, North Sea. In: J. Brooks and K.W. Glennie (Editors), Petroleum Geology of North-West Europe. Graham and Trotman, London, pp. 819-833. Cavazza, W. and Gandolfi, G., 1992. Diagenetic processes along a basin-wide marker bed as a function of burial depth. J. Sediment. Petrol., 62: 261-272.

Colter, V.S. and Havard, D.J., 1981. The Wytch Farm Oil Field, Dorset. In: L.V. Illing and G.D. Hobson (Editors), Petroleum Geology of the Continental Shelf of North-West Europe. Heyden and Son, London, pp. 494-503. Cope, J.C.W., Ingham, J.]C and Rawson, P.F. (Editors), 1992. Atlas of Palaeogeography and Lithofacies. Geol. Soc. London, Mem., 13, 153 pp. Deer, W.A., Howie, R.A. and Zussman, J., 1982. Rock-Forming Minerals, 1A. Orthosilicates. Longman, London, 919 pp. Dietz, V., 1973. Experiments on the influence of transport on shape and roundness of heavy minerals. Contrib. Sedimentol., 1: 103-125. Doyle, L.J., Carder, K.L. and Steward, R.G., 1983. The hydraulic equivalence of mica. J. Sediment. Petrol., 53: 643648. Ebbern, J., 1981. The geology of the Morecambe Gas Field. In: L.V. Illing and G.D. Hobson (Editors), Petroleum Geology of the Continental Shelf of North-West Europe. Heyden and Son, London, pp. 485-493. Flores, R.M. and Shideler, G.L., 1978. Factors controlling heavy mineral variations on the south Texas outer continental shelf. J. Sediment. Petrol., 48: 269-280. Friese, F.W., 1931. Untersuchung von Mineralen auf Abnutzbarkeit bei Verfrachtung im wasser. Tschermaks Mineral. Petrogr. Mitt., 41: 1-7. Friis, H., 1976. Weathering of a Neogene fluviatile fining upwards sequence at Voervadsbro, Denmark. Bull. Geol. Soc. Den., 25: 99-105. Grigsby, J.D., 1990. Detrital magnetite as a provenance indicator. J. Sediment. Petrol., 60: 940-951. Hering, O.H. and Zimmerle, W., 1963. Simple method of distinguishing zircon, monazite and xenotime. J. Sediment. Petrol., 33: 472-473. Hurst, A., 1985. The implications of clay mineralogy to palaeoclimate and provenance during the Jurassic in NE Scotland. Scott. J. Geol., 21: 143-160. Jackson, D.I., Mulholland, P., Jones, S.M. and Warrington, G., 1987. The geological framework of the East Irish Sea Basin. In: J. Brooks and K.W. Glennie (Editors), Petroleum Geology of North-West Europe. Graham and Trotman, London, pp. 191-203. Johnsson, M.J., Stallard, R.F. and Lundberg, N., 1991. Controls on the composition of fluvial sands from a tropical weathering environment: sands of the Orinoco drainage basin, Venezuela and Colombia. Bull. Geol. Soc. Am., 103: 1622-1647. Krynine, P.D., 1946. The tourmaline group in sediments. J. Geol., 54: 65-87. Mange, M.A. and Maurer, H.F.W., 1992. Heavy Minerals in Coiour. Chapman and Hall, London, 147 pp. Milliken, K.L., 1988. Loss of provenance information through subsurface diagenesis in Plio-Pleistocene sediments, northern Gulf of Mexico. J. Sediment. Petrol., 58: 9921002. Milner, H.B., 1962. Sedimentary Petrography, 4th ed. George Allen and Unwin, London, 715 pp.

25~

A. C. Morton, C. Hallsworth / Sedimentary Geology 90 (1994) 241--256 15/21a-b, UK North Sea. In: I.L. Abbotts (Editor), United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume. Geol. Soc. London, Mem., 14: 331-338. Poldervaart, A., 1955. Zircon in rocks. 1. Sedimentary rocks. Am. J. Sci., 253: 433-461. Rubey, W.W., 1933. The size distribution of heavy minerals within a water-lain sandstone. J. Sediment. Petrol., 3: 3-29. Russell, R.D., 1937. Mineral composition of Mississippi river sands. Bull. Geol. Soc. Am., 48: 1308-1348. Shukri, N.M., 1949. The mineralogy of Nile sediments. Q.J. Geol. Soc. London, 105: 511-529. Stewart, D.J., 1986. Diagenesis of the shallow marine Fulmar Formation in the central North Sea. Clay Miner., 21: 537-564. Stuart, I.A., 1993. The geology of the North Morecamble Gas Field, East Irish Sea Basin. In: J.R. Parker (Editor), Petroleum Geology of Northwest Europe, Proceedings of the 4th Conference. The Geological Society, London, pp. 883-895. Thiel, G.A., 1940. The relative resistance to abrasion of mineral grains of sand size. J. Sediment. Petrol., 10: 103124. Thiel, G.A., 1945. Mechanical effects of stream transportation in mineral grains of sand size. Bull. Geol. Soc. Am., 56: 1207. Van Andel, T.H., 1950. Provenance, Transport and Deposition of Rhine Sediments. Veenman en Zonen, Wageningen, 129 pp.

Morton, A.C., 1984. Stability of detrital heavy minerals in Tertiary sandstones of the North Sea Basin. Clay Miner.. I c): 287-308. Morton, A.C., 1985. Heavy minerals in provenance studies. In: G.G. Zuffa (Editor), Provenance of Arenites. Reidel, Dordrecht, pp. 249-277. Morton, A.C., 1986. Dissolution of apatite in North Sea Jurassic sandstones: implications for the generation of secondary porosity. Clay Miner., 21: 711-733. Morton, A.C., 1991. Geochemical studies of detrital heavy minerals and their application to provenance studies. In: A.C. Morton, S.P. Todd and P.D.W. Haughton (Editors), Developments in Sedimentary Provenance Studies. Geol. Soc. London, Spec. Publ., 57: 31-45. Morton, A.C. and Smale, D., 1990. The effects of transport and weathering on heavy minerals from the Cascade River, New Zealand. Sediment. Geol., 68: 117-123. Morton, A.C. and Johnsson, M.J., 1993. Factors influencing the composition of detrital heavy mineral suites in ttolocene sands of the Apure River drainage basin, Venezuela. In: A. Basu and M.J. Johnsson (Editors), Processes Controlling the Composition of Siliciclastic Sediments. Geol. Soc. Am., Spec. Pap., 284: 171-185. O'Driscoll, D., Hindle, A.D. and Long, D.C., 1990. The structural controls on Upper Jurassic and Lower Cretaceous reservoir sandstones in the Witch Ground Graben, UK North Sea. In: R.F.P. Hardman and J. Brooks (Editors), Tectonic Events Responsible for Britain's Oil and Gas Reserves. Geol. Soc. London, Spec. Publ., 55: 299-323. Parker, R.H., 1991. The lvanhoe and Rob Roy Fields, Blocks

You might also like