Sie sind auf Seite 1von 22

Journal of Sedimentary Research, 2005, v. 75, 921942 DOI: 10.2110/jsr.2005.

071

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS IN SILICICLASTIC ROCKS AND THEIR ROLE IN FELDSPAR DISSOLUTION DURING TRANSPORT AND DIAGENESIS
IAN PARSONS,1 PAULINE THOMPSON,1 MARTIN R. LEE,2
1 2

AND

NICOLA CAYZER1

Grant Institute of Earth Science, University of Edinburgh, West Mains Road, Edinburgh EH9 3JW, U.K. Division of Earth Sciences, Centre for Geosciences, University of Glasgow, Lilybank Gardens, Glasgow G12 8QQ, U.K.

ABSTRACT: Alkali feldspars usually exhibit optical or sub-optical intergrowths of albite and K-feldspar (known generically as perthite) which have morphologies and crystallographic characteristics that are distinctive of the igneous or metamorphic environment in which they grew and cooled to surface temperatures. We review the current terminology and understanding of how the microtextures form and discuss their role in feldspar dissolution during weathering, transport, and diagenesis. We show how microtextures can be used as provenance indicators in arkosic siliciclastic rocks, using examples from the Fulmar Formation, a reservoir rock in the Upper Jurassic Humber group in the Central North Sea, in which feldspar dissolution is a major source of secondary porosity. We describe the most effective techniques for routine characterization, of which the most practical is back-scattered electron imaging in a scanning electron microscope, often coupled with cathodoluminescence. We provide an atlas of the main types of perthitic intergrowth likely to be encountered in igneous and metamorphic rocks worldwide, and show how many of the microtextures can be matched in clastic grains in the Fulmar Formation. Grains which are preferentially preserved have microtextures that make them relatively unreactive in aqueous fluids at low temperatures, and we conclude that provenance has a major impact on reservoir quality. Detrital grains in the Fulmar exhibit authigenic albitization, K-feldspar overgrowths, and replacement, and we discuss how diagenetic features can be distinguished from replacive features developed in the protolith.

INTRODUCTION

We here describe how intracrystal microtextures in detrital alkali feldspars can be used to infer the provenance of siliciclastic rocks and reliably distinguish detrital feldspars from authigenic grains and overgrowths. We discuss how these microtextures influence the rate at which detrital alkali feldspar grains degrade and dissolve during weathering (Lee and Parsons 1995; Lee et al. 1998) and undergo dissolution and replacement during diagenesis (Lee and Parsons 1998) and hence control which grains, themselves usually fragments of crystals of igneous or metamorphic origin, are likely to survive or to dissolve to create secondary porosity in potential reservoir rocks. From our previous experience of intergrowths in igneous or metamorphic rocks we can suggest a variety of ultimate provenances for detrital alkali feldspars with considerable confidence. The use of microtextures to deduce provenance, and their role in the reactivity of feldspars during diagenesis, was briefly discussed by Lee and Parsons (2003) and the present paper develops these ideas. The proportion of feldspars relative to other minerals has frequently been used as a provenance indicator (e.g., Arribas et al. 2000), but as far as we know our use of intracrystal feldspar microtexture has not been suggested before. We begin this paper with a review of the current understanding of alkali feldspars relevant to the characterization of detrital grains in clastic sedimentary rocks, and to their relative stability during burial and

diagenesis, and suggest the most effective techniques for routine work. We illustrate the method with sedimentary examples from the Fulmar Formation, an important reservoir rock in the Upper Jurassic Humber group in the Central North Sea, and with examples from our earlier work on igneous and metamorphic rocks. The Fulmar oilfield, from which the samples in the present study come, lies in a small, steep-flanked anticline on the southwestern edge of the Central Graben in the U.K. sector of the North Sea. Our observations were made on a total of 24 core fragments from 5 wells from measured depths ranging from 3273 to 5483 m. The Fulmar Formation sandstones and sandbodies are Upper Jurassic (Oxfordian to Kimmeridgian) in age (Stockbridge and Gray 1991) and comprise fine- to medium-grained, heavily bioturbated and largely structureless, shallow marine arkosic sandstones that possibly were derived from Triassic sedimentary rocks. The reservoir is overpressured and currently at a temperature of , 130uC (Saigal et al. 1992). In the Fulmar Formation feldspar dissolution is believed to be the main cause of secondary porosity and it has been suggested that the amount of detrital alkali feldspar decreases systematically with burial depth (Wilkinson and Haszeldine 1996; Wilkinson et al. 1997). Nevertheless, we have found considerable amounts of detrital alkali feldspar (. 10 vol. % of solids) in deep Fulmar wells. We emphasize that the survival of any particular feldspar grain, in pore fluids which can be expected to be at or very close to local equilibrium with the mineral assemblage in the clastic rock, is strongly dependent on intra-grain microtexture, rather than chemistry, so that provenance is directly relevant to secondary porosity development, and may have predictive potential. By comparing the populations of distinctive feldspars to those in earlier sedimentary formations in the same basin, it should be possible, in principle, to establish whether they have passed through previous sedimentary cycles, for example, in the case of the Fulmar, to establish whether grains have come from reworked Devonian or Permo-Triassic sandstones, as suggested by Stewart (1986).
UNDERLYING PRINCIPLES

Nomenclature and Polymorphism of Alkali Feldspars The terminology used here is that of Smith and Brown (1988) and as developed in reviews by Parsons and Brown (1984), Brown and Parsons (1989), Brown and Parsons (1994), and Parsons and Lee (2000). Many illustrations of perthitic textures can be found in Smith (1974, Chapter 19) and Smith and Brown (1988, Chapter 19). A convention to be noted is that the word Albite is capitalized when it refers to the twin law, whereas albite refers to the mineral phase. Similarly, Adularia refers to the characteristic {110} habit of the low-temperature variety of K-feldspar known as adularia. At igneous and metamorphic temperatures alkali feldspars crystallize as homogeneous ternary solid solutions of the three end-member components NaAlSi3O8 (albiteAb), KAlSi3O8 (orthoclaseOr) and CaAl2Si2O8 (anorthiteAn). (They may also contain significant amounts

Copyright E 2005, SEPM (Society for Sedimentary Geology)

1527-1404/05/075-921/$03.00

922

I. PARSONS ET AL.

JSR

of the component BaAl2Si2O8 [celsianCsn]). As temperature decreases, the originally homogeneous crystals usually exsolve (unmix is an alternative term) in the solid state to form intergrowths of two feldspar phases known generically as perthitic intergrowths. One phase is rich in the Or component and is best called the Or-rich phase; the other is rich in the Ab component and is best called the Ab-rich phase or the plagioclase phase. Although petrographers often refer to albite exsolution lamellae in perthitic crystals, this is not necessarily correct; they often contain significant An and may be oligoclase (Ab90An10Ab70An30) or exceptionally andesine (Ab70An30Ab50An50). The term perthite should properly be reserved for intergrowths in which Ab-rich plagioclase lamellae or patches are enclosed by Or-rich feldspar, antiperthite is used when Or-rich feldspar is enclosed by plagioclase, and mesoperthite when the two phases are present in approximately equal proportions. Although many perthites form by exsolution it is certainly the case that some form by replacement and many single grains contain both exsolved and replacive intergrowths (e.g., Lee et al. 1995). An important point is that intergrowths which have formed by exsolution must have developed initially in homogeneous single crystals at an elevated temperature and when found in sedimentary rocks must therefore be detrital. Intergrowths formed by replacement, on the other hand, can form both during cooling from igneous and metamorphic temperatures and during diagenesis. Criteria for differentiating the different types of intergrowth are given later. Sub-optical intergrowths (i.e., , 1 mm in scale) are called cryptoperthites (or cryptomesoperthites, cryptoantiperthites), optically visible intergrowths are given the prefix micro-. Modern electron microscope work has shown that many crystals contain both cryptoperthite and microperthite. Textbooks frequently assert that cryptoperthite is characteristic of alkali feldspars from volcanic rocks and is a result of rapid cooling, but this is correct only in a very general way, and it often occurs, together with coarser intergrowths, in relatively slowly cooled igneous (Brown et al. 1983; Lee et al. 1995) and metamorphic rocks (Evangelakakis et al. 1993; Cayzer 2002). This is a point to consider when identifying fragmentary clastic grains. Replacement by Na-rich feldspar is often called albitization, although the replacive phase is not always pure albite (Lee and Parsons 1997b). Albitization may occur both in cooling igneous rocks and during diagenesis. When encountered in clastic grains it is not self-evidently a product of diagenesis. Replacement by K-rich feldspar (Morad et al. 1989; Lee and Parsons 1998) has been called microclinization, but this term has also been used for the process by which orthoclase changes into microcline, and is best avoided. Furthermore, replacive K-feldspar formed in diagenesis is not usually microcline. The high-temperature monoclinic form of K-feldspar, sanidine, including fully disordered (with respect to Si and Al in the Si,AlO framework) high sanidine and partially ordered low sanidine, may grow metastably (Baskin 1956; Glover and Hoseman 1970) during diagenesis. Orthoclase has a distinctive microtexture, visible only at the transmission electron microscope (TEM) scale known as tweed. This is composed of alternating twin-like domains only a few unit cells thick (Eggleton and Buseck 1980) and develops from low sanidine by a continuous process in which alternating left-handed and right-handed triclinic domains form. Orthoclase retains average overall monoclinic symmetry and becomes stranded and unable to order further to produce fully ordered, twinned triclinic microcline. The word orthoclase will be used here in its modern sense of an Or-rich K-feldspar with a well developed tweed microtexture. The symbol Or refers to the KAlSi3O8 component only, and carries no crystal structural implications. Fluidfeldspar reactions are usually required for the transition from orthoclase to triclinic microcline via what has been called an unzipping reaction (Brown and Parsons 1989; Putnis 2002). Microcline usually exhibits the well known cross-hatched tartan twinning in the optical

microscope, although twins may be visible only using TEM (see, e.g., Fig. 7 for a detrital example). Much microcline in igneous and metamorphic rocks is very irregularly twinned at the TEM scale and is usually called irregular microcline. Within single K-feldspar grains it is often mixed with volumes of tweed orthoclase on the scale of a few mm (see Fitz Gerald and McLaren 1982, their fig.11). Many authigenic Kfeldspars are the variety known as adularia, which is distinguished by its {110} habit and structurally by partial ordering and a weak departure from monoclinic symmetry (see Lee and Parsons 2003 for examples from the Fulmar Formation, and also Worden and Rushton 1992). Although many diagenetic adularias have weakly developed tweed microtexture, well developed tweed is probably diagnostic of detrital grains. Well developed tartan twinning is a reliable indicator of a detrital origin. Ab-rich feldspars also undergo a monoclinic R triclinic phase transition on cooling, from high albite and high Na-sanidine to triclinic anorthoclase. This phase transition leads to tartan twinning in anorthoclase, but the twins are usually finer-scale, straighter, and more regular than in microcline, with angular intersections. In microcline the tartan texture is visible only in sections nearly parallel to (001), on which only one cleavage, (010), is visible, whereas in anorthoclase the tartan twinning is visible on (100) and two cleavages, (010) and (001), are visible. Anorthoclase cannot grow during diagenesis, and discovery of anorthoclase in a sedimentary rock would clearly indicate a detrital origin. Solvus Relationships The compositions of pairs of feldspar phases coexisting in equilibrium, whether in perthitic intergrowth or as separate crystals, become progressively more different as temperature decreases, and are defined by a temperatureomposition surface, the ternary feldspar solvus (Fig. 1). The significance of the various curves is discussed in the caption. When An is not present the composition of coexisting feldspars is defined by the binary alkali feldspar solvus (Smith and Parsons 1974), and this is often used as an approximation of solvus relationships. However, although many alkali feldspars have low bulk An contents (12 mol%), small amounts of the third component, An, have a large effect on the temperature at which exsolution begins and exercise a strong control on the microscopic appearance of perthitic intergrowths (Fig. 2). AbOrAn feldspars with more than a few percent of the least abundant component are often loosely called ternary feldspars, although strictly speaking all natural feldspars are at least ternary. The strain-free ternary feldspar solvus (the term strain-free is explained below) defines the minimum temperature at which a feldspar phase of a given bulk composition could have grown from magma, in a metamorphic rock, or from a diagenetic fluid. Igneous rocks in which a pair of discrete feldspar phases crystallize from magma (e.g., tie line P BAF in Fig. 1) are called subsolvus rocks. Magmatic rocks in which a single feldspar with a bulk composition such as B crystallizes are called hypersolvus rocks. Small amounts of An in an alkali feldspar and of Or in a plagioclase feldspar have an extremely large effect on minimum possible crystallization temperature (Fig. 1). At diagenetic temperatures stable feldspars are nearly pure Ab and Or, and the amount of An that can dissolve in an alkali feldspar at diagenetic temperatures is very small (% 0.5 mol%). This provides a way to distinguish authigenic from detrital feldspars using back-scattered electron imaging (BSE) because of the higher mean atomic number when Ca is present. Similarly, the amount of Or that can dissolve in a plagioclase (AbAn) feldspar growing at low temperature is small. Determination of Bulk Composition When stating a chemical analysis of an alkali feldspar, it is crucial to make clear, particularly when it has been analyzed using electron-probe microanalysis (EPMA) or energy-dispersive X-ray microanalysis (EDX)

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

923

FIG. 1.Temperaturecomposition prism for the ternary feldspar system AbOrAn. The Strain Free Solvus surface (SFS, purple) defines the compositions of pairs of feldspar phases (P: plagioclase, AF: alkali feldspar) growing simultaneously from magma, during metamorphism, or during diagenesis. P and AF must lie on the ends of a tie line passing through the bulk composition B. The position of this tie line depends on temperature, and to a lesser extent on pressure. This relationship is the basis of two-feldspar geothermometry. The SFS also defines the compositions of the phases coexisting in perthitic intergrowths provided that they do not share a common Si,Al O framework (i.e., they are incoherent). When they share a common framework (are coherent), the phase compositions are depicted by the Coherent Solvus (CS, blue). If an alkali feldspar crystal of bulk composition AF (typical of a calc-alkaline granite) is cooled, together with a coexisting plagioclase P, the two phases should, in principle, react continuously with one another and maintain compositions on the SFS. In practice this usually does not occur. P and AF behave as separate systems and begin to exsolve when they encounter the CS, each unmixing into two phases. For the alkali feldspar these will have compositions N and K, on the blue surface, in proportions which reflect the bulk composition of AF, and the intergrowth is a perthite. P will also exsolve, forming an antiperthite. As cooling proceeds the compositions of N and K become progressively more different, but unless the crystals undergo dissolutionreprecipitation reactions in an aqueous fluid they will remain on the CS. When a fluid is present, all or part of the crystal my recrystallize to a strain-free assemblage on the SFS. The driving force for this recrystallization is coherency strain energy. Strain-free assemblages at low temperature are close to end-member feldspars, so feldspars growing, or actively replacing detrital feldspars during diagenesis, are nearly pure phases (N and K on the tie line passing through B 5 AF, which is drawn at the present-day temperature of the Fulmar Formation). The SFS and CS shown are for equilibrium Si,Al ordering (Brown and Parsons 1989). The 750uC isotherm is from Fuhrman and Lindsley (1988) and is for disordered feldspars.

in the scanning electron microscope (SEM), whether it is a bulk composition or a phase composition that has been obtained. Phase compositions can be obtained by point analysis of coarse perthitic intergrowths or by measurement of peak positions using X-ray diffraction, but they have little value in provenance studies because compositions of intergrown phases usually remain in equilibrium down to low temperature and therefore approach pure albite and almost pure K-feldspar (. Or90) on the ternary solvus (Fig. 1). The only exceptions are some quickly cooled volcanic rocks (e.g., MacKenzie and Smith 1956) and rare granulite facies xenoliths in volcanic rocks (Hayob et al. 1989). Bulk compositions can provide useful provenance information; for a compilation of feldspar bulk analyses (both major and trace elements) see numerous figures in Smith and Brown (1988), Chapter 14. The bulk composition of completely homogeneous crystals can be obtained by simple EPMA point analysis, but such crystals are rare and most natural alkali feldspars are perthitic. If the exsolution textures are on a , mm scale an estimate of bulk composition can be obtained using a defocussed beam of . 10 mm diameter or a rastered beam, but to obtain the bulk composition of a coarsely exsolved microperthite requires the integration of counts over a large area of a section. A combination of EPMA and BSE image analysis has recently been applied successfully (Marks and

Markl 2001). Even if this is done, it is often uncertain that the bulk composition obtained is that with which the crystal initially grew, because replacement reactions involving addition of both Ab or Or can affect alkali feldspars during cooling from igneous temperatures and during diagenesis (e.g., Lee et al. 1995; Lee and Parsons 1998). The replacive feldspars may differ in composition from those which have exsolved (Lee et al. 1995). Exsolution Microtextures Truly homogeneous alkali feldspar crystals are extremely rare, restricted to sanidines quenched very rapidly from volcanic temperatures and not subsequently subject to fluidrock interactions, and to crystals or overgrowths growing at very low temperature during diagenesis. Petrographic descriptions based on optical microscopy (OM) often describe optically featureless alkali feldspars as unexsolved, but except in the case of authigenesis the chances are high that the crystal in question is cryptoperthitic. BSE imaging in an SEM can reveal cryptoperthite, except for the finest intergrowths, which require TEM. The most convenient way to distinguish detrital alkali feldspars that are featureless using OM and BSE imaging, from authigenic crystals, is cathodolumi-

924

I. PARSONS ET AL.

JSR

FIG. 2. Schematic diagram showing coherent exsolution microtextures, as they appear viewed down the Z axis, developed as a function of composition in slowly cooled feldspars in the system AbOrAn. Vertical lines are Albite twins in plagioclase and in the Ab-rich phase in perthitic intergrowths; cross hatching at the right represents tweed orthoclase. Unornamented areas are Or-rich feldspar, low sanidine where intergrown albite lamellae are straight, microcline where zigzag or where enclosing lozenge-shaped areas of albite. Note that these textures apply to crystals that have not been externally deformed. From Brown and Parsons (1988).

nescence (CL). This is likely to be bright in the volcanic case but absent from authigenic grains. Note that in the present paper all CL images were obtained in an SEM with a detector with a spectral range of 275700 nm. Microperthitic and cryptoperthitic crystals fall into two broad categories: finer scale (usually # 1 mm), regular strain-controlled intergrowths and coarser, much more irregular, deuterically coarsened intergrowths (Parsons and Brown 1984). The former are the result of a continuous process in which Na and K ions diffuse through an unbroken SiAlO framework (Ca will also diffuse, as discussed below). The latter are the result of a discontinuous process involving dissolution

and reprecipitation in an aqueous fluid, while the crystal retains its external shape. The microtextural differences between the two types are extremely profound, and have a strong influence on the rate of dissolution of feldspars during weathering, transport, and diagenesis. The most obvious difference in OM is that in plane-polarized light crystals with strain-controlled intergrowths are glass-clear (apart perhaps from isolated fluid or mineral inclusions) whereas deuterically coarsened intergrowths are in various degrees turbid. Most feldspars from common types of igneous and metamorphic rocks are to some extent turbid.

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

925

FIG. 3. Generalized sketch of an , 100-mm-long cleavage fragment of an alkali feldspar phenocryst from the Shap granite, Cumbria, U.K., a typical calc-alkaline granite, as it might occur as a detrital fragment. The upper surface is the (001) cleavage, the right-hand is the (010) cleavage, and the front is the Murchison plane. Thicker lamellae and subgrains of albite are filled in gray. K-feldspar, orthoclase in the areas with lamellar perthites and microcline in areas of patch perthite, is unfilled. Alkali feldspars from many granites and alkali feldsparbearing gneisses have similar microtextures, so that, on an Earth-wide basis, this is the most common type of detrital feldspar likely to be encountered in a siliciclastic rock. Note how the exsolution textures vary in appearance depending on viewing direction, and the large variety of textures that might be found in detrital grains formed from the further disintegration of such a fragment. (Modified from Parsons et al. 1999).

Strain-Controlled Intergrowths.The Si,AlO framework passes continuously, or nearly so, between the Or-rich and Ab-rich phases. Such intergrowths are said to be coherent, or, if dislocations are present on the interfaces, semicoherent. Because of the difference in cell dimensions between Na- and K-rich feldspar the framework is elastically strained close to the interface of the two phases. The chemical driving force which leads to exsolution has two opposing forces: surface and strain energy. The compositions of the exsolving phases are defined by a coherent solvus which is at lower temperature than the strain-free solvus (Fig. 1). Coherent exsolution is a continuous process which begins during cooling at or just below the coherent solvus. It is incorrect to speak of one exsolution temperature and preferable to describe this temperature as the beginning of exsolution. The coherent solvus defines the minimum temperature at which a feldspar of any particular bulk composition can persist as a homogeneous crystal during cooling unless cooling rates are extremely high, for example in ash deposits. The mechanisms by which exsolution occurs in alkali feldspars are discussed by Parsons and Brown (1991). As exsolution proceeds, surface energy is minimized by coarsening of the intergrowths and lamellae initially a few nm thick become visible

optically (. 1 mm). Structural strain is minimized by migration of the interfaces of the two phases into orientations where strain energy is least (Fig. 2). In perthites sensu stricto coherent lamellar interfaces are usually 01) and (8 01) of the approximately straight and initially between (6 monoclinic host. The overall shape of the albite lamellae is that of extremely extended flat lenses (e.g., Lee and Parsons 1995, their figs. 14) when viewed on (010) and rather shorter lenses on (001), the overall shape being like an aircraft wing (Fig. 3; also Figs. 1214). Such intergrowths are called film perthites. In cryptoantiperthites, lenses of K-feldspar occur in a plagioclase host. When the proportion of K-feldspar is low, the lenses resemble coffee beans (Brown and Parsons 1988) and the lenses bend where they intersect Albite twin lamellae (Fig. 2, area 1). The orientation and shape of coherent intergrowths formed during slow cooling, in plutonic igneous and metamorphic rocks, depend on the bulk composition of the crystal. The final products are summarized in Fig. 2. Coherent lamellae in perthites and antiperthites remain essentially planar, but in coherent mesoperthites the film lamellar interfaces bend when framework ordering and associated twinning occur (Brown and Parsons 1984), giving zigzag perthite (area 5 in Fig. 2) and for mesoperthites with low An contents a so-called braid perthite develops,

926

I. PARSONS ET AL.

JSR

FIG. 4.Micrographs illustrating the role of microtexture in alkali feldspar during weathering and diagenesis. Parts A, B, and D are secondary electron SEM images, Part C is a BSE image. A) (001) cleavage surface of an alkali feldspar phenocryst from the Shap granite that has been briefly etched (, 50 secs) in HF vapor in the laboratory. Albite has dissolved slightly less rapidly than orthoclase so that the perthitic lamellae stand out as low ridges. The black dots on thicker lamellae are dissolution etch pits on edge dislocations, showing that thicker lamellae are semicoherent. Thinner lamellae lack dislocations and are coherent. The band in the center is an igneous growth zone with slightly higher An content and contains only very fine cryptoperthite. The absence of dislocations makes such zones particularly resistant to HF etching and to weathering (modified from Lee and Parsons 1997a). B) Naturally weathered (001) surface of an alkali feldspar fragment from a peat soil overlying Shap granite, with semicoherent albite lamellae. The albite has been almost completely dissolved because adjacent etch pits have combined, leaving lines of pits resembling the perforations on tear-off labels. K-feldspar between lamellae is breaking off as thin flakes (from Lee et al. 1998). C) Detrital grain of perthitic alkali feldspar from the Fulmar Formation. In parts of the grain, mainly near the edges, albite lamellae (medium gray) have been partly or completely dissolved. Weakened by partial dissolution, the grain has broken during compaction. Subsequently, authigenic K-feldspar (light gray) has formed a partial overgrowth, and formed a lining on both surfaces of the fracture between the two largest grains. Dark grains at the extreme right are quartz and the medium gray grains forming a cement around much of the feldspar are dolomite. D) Same origin as Part B. Weathered (001) surface on which a band of patch perthite cuts across semicoherent film perthite. The irregular areas mark zones where preferential dissolution has occurred along subgrain boundaries within the patch perthite (from Lee et al. 1998).

in which lozenges of albite are separated by sheets of microcline (area 4). Braid perthites are visible under high power using OM and are easily visible using BSE in an SEM (e.g., Fig 10). These intergrowths, and the fine straight intergrowths which precede them, are fully coherent. Coherency strains are minimized by the complex shapes of the interfaces between the phases. However, in other parts of Figure 2, where lamellae are straight or nearly so (in fact the lamellae have much higher aspect ratios than shown; see Figs. 3, 1214) spaced dislocations form and the intergrowths become semicoherent (next section). The coarseness of coherent exsolution textures has been shown to be a function of cooling rate by experimental work (Yund et al. 1974; Yund and Davidson 1978) and by inference in both volcanic rocks (Yund and Chapple 1980; Snow and Yund 1988) and relatively rapidly cooled plutonic igneous rocks (Brown et al. 1983; Waldron and Parsons 1992). The coarsest regular coherent intergrowths occur in feldspars in granulite facies metamorphic rocks (Evangelakakis et al. 1993; Cayzer 2002) that have been cooled extremely slowly or maintained for long periods at sub-

solvus temperatures. However, the relationship between coarseness and cooling history in plutonic rocks is not a simple one, except in exceptional circumstances (e.g., in the Klokken syenite intrusion; Brown et al. 1983), and many individual grains of K-feldspar from granitic rocks contain coherent and semicoherent exsolution textures at a range of scales from several mm to a few nm, with variation in wavelength and scale occurring over short distances (e.g., Lee et al. 1995). The presence of the An component (and possibly Csn) is probably an important control on coarsening rate, because diffusion of divalent Ca2+ (and Ba2+) is coupled, to preserve charge balance, with that of Al3+ in the framework of the feldspar. During exsolution Ca goes into the Ab-rich phase (Mason 1982) so that the rate-limiting step in coarsening is diffusion of framework Al, which is much slower than that of the alkali ions (see Smith and Brown 1988, Sec. 16.2.4 and references therein). This is explains the presence, in many normally microperthitic granitic feldspars (e.g., Shap figs. 3 and 4a), of zones in which the lamellae are cryptoperthitic (Lee and Parsons 1997a). The zones represent chemical

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

927

FIG. 5. Optical micrograph (crossed polarizers) of perthitic microtextures in an alkali feldspar phenocryst in a sample of fresh granite from a quarry in the Shap granite. In places the feldspar has been unaffected by deuteric fluids, and is glass-clear, with nearly straight film albite lamellae. Elsewhere turbidity has begun to spread along lamellae, as dissolution and replacement take place on and around dislocation outcrops. This is the beginning of deuteric coarsening. Veins of deuterically coarsened patch perthite also cut across the crystal, probably partly following fractures.

FIG. 7. Bright-field TEM image of a detrital microcline from the Fulmar Formation with fine (sub-optical) tartan twinning and cryptoperthitic albite lamellae (lower right). The grain has a diagenetic overgrowth of adularia in the form of tightly packed subgrains (upper left). Despite the wealth of microtexture in this image, the grain and overgrowth would look clear and featureless in OM. However, the overgrowth does not exhibit CL, in contrast with the detrital core. (From Lee and Parsons 2003).

zoning in An and Csn inherited from magmatic growth and are less susceptible to weathering and diagenetic dissolution than the bulk of the crystal (see Lee et al. 1998 and below). Semicoherent Strain-controlled Intergrowths.As film perthites and some nearly straight film mesoperthites and antiperthites coarsen, intergrowths often become semicoherent by developing spaced edge dislocations which can be made visible by HF etching (Fig. 4A; see Waldron et al. 1994 for practical details). These dislocations are common in feldspars from subsolvus granites (Fig. 4) and have a crucial role in their low-temperature reactions during igneous cooling (Lee and Parsons 1997b, 1998), weathering (Fig. 4B; see Lee and Parsons 1995; Lee et al. 1998), and diagenesis. They also contribute to mechanical degradation during compaction (Fig. 4C). Each edge dislocation is a line representing

the end of a single 020 lattice plane in the Ab-rich phase where it terminates at the lamellar interface. The dislocations develop as lamellar coarsening occurs because it is impossible for the Si,AlO framework of both phases to remain continuous while sustaining the strains which develop because of the difference in cell dimensions. The strains are minimized by the formation of regularly spaced dislocations on the 01) and (8 01), one set surface of the perthite lamellae, which is between (6 parallel to the crystal b axis, the other at right angles in the ac plane. The dislocations form very extended lens-shaped loops around lamellae (Fig. 3) and therefore appear in pairs straddling lamellae on fragment surfaces. A simplified diagram illustrating the relationship is given as Figure 2 in Lee and Parsons (1998). This crystallographically irrational plane was known as the Murchison plane of easy fracture by early mineralogists and, with the perfect (010) and (001) cleavages, defines the shape of many detrital alkali feldspar grains. The angles between (010) 01)(8 01) are 90u in monoclinic feldspars, and and (001), and (010) and (6 01) and (8 01) and (001) are 95u and 101u, the angles between (6 respectively. Thus broken fragments of feldspar tend to be approximately

FIG. 6.Detrital alkali feldspar fragment from the Fulmar Formation showing ambiguous late albitization. A) BSE image. The lenticular lamellae of albite (dark gray) indicate an igneous origin, but the irregular patches of albite are of uncertain origin. B) CL image. The K-feldspar is brilliantly luminescent, and the regular albite lamellae show weak luminescence. Albite patches are completely non-luminescent, but their origin is ambiguous. Their presence in the fragment interior suggests that they originated by replacement while in the igneous rock. Had they formed a rim or partial rim to the detrital grain, an authigenic origin would have been inferred.

928

I. PARSONS ET AL.

JSR

differ most when the aqueous solution is close to chemical equilibrium with the solids. This situation is likely to pertain in confined spaces in igneous rocks, in sedimentary basins, or in established soil waters. Figure 4B shows the naturally weathered surface of an alkali feldspar from a peat soil, in which etch pits on dislocation outcrops have combined near the grain surface, leading to dissolution of albite and mechanical degradation (see Lee et al. 1998 for a detailed discussion). Deuteric Intergrowths.These form subsequent to coherent and semicoherent exsolution and replace strain-controlled microtextures. They are up to 103 times coarser than coherent intergrowths and are easily visible using OM. The microtexture is profoundly different to strain-controlled intergrowths (e.g., Figs. 4D, 5, 8, 10 (right), 11, and 12) and is composed of a composite of subgrains on a range of scales from a few tens of nm to several mm. TEM micrographs of such intergrowths can be found in Worden et al. (1990), Guthrie and Veblen (1991), Waldron and Parsons (1992), Walker et al. (1995), and Lee et al. (1995). Coherency between the Ab- and Or-rich phases has been lost (they are incoherent), the intergrowths are essentially strain-free, and phase compositions will be defined by the strain-free solvus (SFS in Fig. 1). The driving force for this recrystallization is loss of coherency strain energy (Brown and Parsons 1993) during reactions between the feldspar and deuteric or hydrothermal fluids. Alkali feldspars which have escaped such reactions are rare in plutonic rocks, although volcanic sanidines are more commonly unaffected. In some rocks the alkali feldspars have been recrystallized in their entirety (e.g., Fig. 11), but in many cases (such as the Shap feldspars illustrated in Figs. 3, 4, and 12) parts of the original crystal have strain-controlled microtexture while elsewhere they are deuterically coarsened. Deuteric intergrowths are turbid in plane-polarized light. It is usually, although not invariably, the Or-rich phase which is the most turbid. Although frequently ascribed to clay minerals, turbidity is overwhelmingly caused by myriads of sub-mm micropores (Worden et al. 1990; Walker et al. 1995), usually developed at the junctions between subgrains. Very turbid detrital alkali feldspars in sedimentary rocks probably owe their appearance to enlargement of preexisting pores and growth of clay minerals, although some clay particles are imported from the protolith igneous or metamorphic rock (see e.g., Worden et al. 1990, their fig.8). The intergrowths are irregular, and various descriptive terms are useful: patch perthite, when each phase forms shapeless patches (Fig. 10, right); and vein perthite, when the lamellae form sinuous bands (Figs. 11, 12). Note, however, that at a scale smaller than the patches and veins

FIG. 8.BSE image of Fulmar sandstone showing four detrital feldspar grains with three distinct provenances. The two bright featureless grains at the left are both volcanic sanidines. The right hand grain has regular lenticular albite lamellae, and is likely to have come from a granitic protolith, and the central grain is a deuterically coarsened perthite, with a few more regular lamellae, also from a granitic protolith. In this case the grain has been subject to fluidrock reactions below , 450uC.

rectangular from whichever crystallographic axial direction they are viewed. Dislocation outcrops on grain surfaces or internal fractures are extremely reactive in aqueous fluids, and the role of various defects in feldspar weathering has been recognized for many years (e.g., Berner and Holdren 1977, 1979). Dislocations are strongly reactive because of Gibbs free energy contributed by undirected bonds in the core and because of elastic strain energy in the disturbed crystal structure surrounding the core. Blum and Lasaga (1987) used Monte Carlo simulations to demonstrate the importance of dislocations in feldspar and quartz dissolution. TEM work currently in progress (Fitz Gerald and Parsons 2003) on alkali feldspar phenocrysts from the Shap granite shows that even in what appear optically to be the freshest parts of grains from fresh igneous rock the dislocation cores have been dissolved away, leading to tiny (, 2 nm) tubes. A most important concept is that of chemical affinity (Velbel 1989), which implies that the relative dissolution rates of structure close to dislocations and normal crystal structure

FIG. 9. A) BSE and B) CL image of a detrital sanidine (left) from the Fulmar Formation, with strong BSE and CL signals. A strongly resorbed authigenic K-feldspar overgrowth has a strong BSE signal but no CL, and a larger area of authigenic feldspar occurs at the right. Although authigenic feldspars may appear optically defectfree, TEM shows that they can be composed of fine-scale subgrain mosaics (Fig. 7; see also Worden and Rushton 1992, Lee and Parsons 2003). Dissolution has affected the reactive authigenic overgrowth, leaving the comparatively defect-poor sanidine unaffected, even though they differ little in chemical composition.

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

929

Deuteric coarsening can occur in a number of ways which reflect the underlying strain-controlled microtexture and/or external deformation. Changes in the bulk composition of the crystal (giving replacement perthite) may also occur (next section). Because replacement perthite may form during diagenesis, it is important to be aware of the variety which can arise from processes occurring before transport and sedimentation. All types of deuteric coarsening involve interaction between a fluid film and the feldspar which leaves a microporous microtexture behind an advancing recrystallization front. Putnis (2002) discusses such processes in minerals in general. In alkali feldspars three broad types of process can be inferred: (i) Fully coherent lamellar and braid mesoperthites recrystallize pervasively along a sharp interface (Fig. 10). A fluid film moves forward along a broad front into the crystal, releasing (unzipping) the elastic strain in the coherent intergrowth and precipitating a strain-free, coarsened and microporous mosaic in its wake. The exact details are complex (Lee et al. 1997; Brown et al. 1997). In originally semicoherent intergrowths deuteric coarsening may be guided by dislocations on lamellae. Figure 5 shows a typical film microperthite from the Shap granite. Areas of fresh film perthite are glass-clear, and the lamellae can be seen because of the difference in refractive index or birefringence. Deuteric alteration leads at first to faint turbidity which extends along lamellae. The tiny dusty particles can be shown using SEM to be pores (Walker et al. 1995) or minute grains of new albite (Lee and Parsons 1997b, their figs. 24). Both features nucleate on dislocations. These small albites increase in size and coalesce leading to areas of patch perthite. Deuteric coarsening also occurs along fractures through crystals (Figs. 5, 12); this feature, where coarse irregular perthite forms

FIG. 10. BSE image of braid cryptoperthite (bottom left) that has partly recrystallized to a patch perthite during deuteric reactions in the cooling igneous rock. From the Klokken syenite, South Greenland. The surface is viewed in a direction approximately parallel to the Z crystallographic axis; X is horizontal and Y vertical. The low albite phase is dark gray, the low microcline phase is light gray. Black dots are micropores, rare in areas with the braid texture. The patch perthite areas are irregular mosaics of Ab- and Or-rich subgrains at a scale considerably smaller than the patches themselves.

(ii)

themselves the deuteric perthite is actually a mosaic of subgrains. Like the dislocations discussed above, the subgrain boundaries have a strong influence on the dissolution of alkali feldspars, and weathered surfaces may be very irregular where preferential dissolution has occurred along them (Fig. 4D).

(iii)

FIG. 11. A) BSE image of crystals of coarse vein perthite in a quartz syenite sheet cutting the Klokken syenite (compare Fig. 10). Hypersolvus alkali granites usually have feldspars with similar microtextures. Note the low magnification. Black crystals are quartz; brightly emitting grains are mafic silicates. B) Higher-magnification image of small area of Part A, showing a relic of braid texture in vein perthite. The black dots are micropores, rare in braid areas where they are aligned and follow healed cracks. (From Lee et al. 1997).

930

I. PARSONS ET AL.

JSR

individual crystals the process is isochemical. In contrast, irregular intergrowths introduce the possibility that they form non-isochemically, by a replacement process. Non-isochemical replacement is well known to occur during diagenesis, replacement by albite (often called albitization) being most common (Ogunyomi et al. 1981; Walker 1984; Morad 1986; Saigal et al. 1988) although introduction of K-feldspar also occurs (Milliken 1989; Morad et al. 1989; Lee and Parsons 1998; Lee and Parsons 2003). However, both these processes have also been shown to occur in cooling igneous (Ferry 1985; Aldahan et al. 1987; Ramsayer et al. 1992; Lee and Parsons 1997b) and metamorphic rocks (Pryer and Robin 1996; Cayzer 2002). Unambiguous criteria for distinguishing diagenetic replacement from non-isochemical replacement in an igneous or metamorphic protolith are not easy to establish. In an igneous rock deuteric microtextures are likely to be isochemical if a large number of individual alkali feldspar crystals have the same bulk composition (see Brown et al. 1983). However, this approach is unworkable in a siliciclastic rock. Areas or veins of Ab-rich feldspar which are not pure Ab but which contain a few percent of An are likely to be due to non-isochemical replacement of late igneous origin. In a study of low-temperature replacement in the Shap granite, Lee and Parsons (1997b) found that replacive plagioclase forming veins and showing strong blue cathodoluminescence was Or1.5Ab89.8An8.7. Within , 1 mm of the crystal surface, turbid, non-cathodoluminescent replacive albite Or0.7Ab99.1An0.2 also occurred, in part penetrating along preexisting semicoherent perthitic albite lamellae. However, it would be extremely difficult to distinguish such albite from diagenetic albite once fragmented and transported into a sedimentary rock. Figure 6 shows a detrital grain with patches of non-cathodoluminescent albite, which is similar to the patches of pure replacive albite seen in the Shap granite and therefore could be of igneous origin. A continuous or semicontinuous rim of such albite would unambiguously indicate a diagenetic origin. It is possible that trace-element or isotopic criteria might exist to distinguish these possibilities.
PREPARATION, ANAYTICAL, AND IMAGING TECHNIQUES

FIG. 12. Collage of BSE images of a typical area of film, vein, and patch perthite in a phenocryst from the Shap granite, viewed approximately parallel to the Z axis. Similar microtextures are found in most subsolvus granites, with variation imposed by bulk composition, in this case , Ab25Or75, and the degree of deuteric alteration. A 50 mm detrital grain might sample regions with strikingly different microtextures. Albite is dark gray, K-feldspar is medium gray, black dots are micropores. Regular lamellae near MP are semicoherent films of albite in orthoclase, as seen on the etched surfaces in Figure 4. Elsewhere, ragged veins of albite, Ab, correlate with regions rich in micropores. Featureless K-feldspar in the vicinity of the albite veins is intermediate microcline, Im. This material and the albite veins and patches are an incoherent mosaic of subgrains, and the irregular texture, with its micropores, marks feldspar which has pervasively recrystallized in a deuteric fluid. (From Lee and Parsons 1997b).

veins across grains, is common in alkali feldspars from granites, and usually coexists with, and grades into, more pervasive coarsening of type (ii). It perhaps forms because of fluid flow along fractures which develop in response to external stresses. In the case of the Shap feldspar illustrated, EPMA analysis showed that the albite in these veins is an oligoclase of composition different from the exsolved Ab-rich phase, so this is strictly a replacement perthite (Lee and Parsons 1997b). Replacement Textures.In the case of regular coherent and semicoherent intergrowths we can be confident that the microtexture formed by exsolution which occurred during cooling following crystal growth at high igneous or metamorphic temperatures. Such intergrowths cannot be produced by any known low-temperature process. At the scale of

The techniques adopted in the present study were designed to provide a rapid survey of as many samples as possible by SEM, including EDX, which then could, if required, be followed by OM and TEM of selected grains. Optical microscopy has inadequate resolution to distinguish most of the types of feldspar intergrowth encountered in siliciclastic rocks, and the general turbidity of many detrital feldspar grains makes identification difficult. In our study of the Fulmar Formation samples were initially made into resin-impregnated polished blocks for SEM, which could then be turned over and made into double polished sections for OM and TEM. While TEM can be very informative in diagenetic and provenance work (Worden and Rushton 1992; Lee and Parsons 1998; Lee and Parsons 2003; Lee et al. 2003) it is time consuming and unlikely to become a routine method. The polished blocks were characterized petrographically (i.e., grain size, shape, mineralogy, and porosity) using BSE with a Philips XL30CP SEM at Edinburgh University operated at 20 kV. All varieties of feldspar observed were recorded. Detrital grains were classified using the criteria provided and illustrated in the next section. Authigenic K- and Nafeldspar overgrowths and replacement, separate authigenic grains, and evidence of corrosion of authigenic overgrowths and detrital grains were recorded. The general type of each feldspar grain encountered along a regular stepped traverse in the SEM was estimated by qualitative EDX using an Oxford ISIS or PGT Spirit EDX system. This point-counting did not include all feldspar microtextural types in a sample, some of which occurred as only a single grain, but it provided a reasonably quantitative guide to the main varieties of feldspar present. An advantage of the EDX point-counting technique is that it allows An-bearing plagioclase to be

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

931

distinguished from An-poor plagioclase by the presence of a Ca peak in EDX spectra. Lee et al. (2003) suggest a detection limit of , 1.5 mol% An. A disadvantage of this technique is that K-feldspar and quartz overgrowths could not be confidently distinguished from their detrital cores in the BSE images alone during the point-counting procedure. However, corresponding CL images obtained in the SEM could be used to identify overgrowths from their low luminescence intensities. It was not practical to use CL when point counting because of the time taken to acquire an image. An advantage of the SEM/EDX technique is that it distinguishes clear, unexsolved, and untwinned grains of feldspar from quartz, which is difficult during OM point counting. We have detected significant amounts of feldspar in samples logged commercially as feldspar-free, presumably because turbidity was the main criterion used to differentiate feldspar. For each sample a total of over 450 points were counted to ensure that abundances as low as 5% were meaningful, with a relative error of 20% at the 95% confidence level. We found that porosities estimated by the SEM method were larger, by a factor of two in many cases, than those obtained using OM on the same core by commercial logging. For point counting by OM the rock samples are normally impregnated with blue resin prior to making the thin section so that holes produced subsequently by plucking will be obvious. Possibly porosities may be overestimated in our point-counting statistics because of plucking of grains during sample preparation for the SEM. Conversely it is likely that in the SEM it is easier to detect smaller pores. Where necessary the blocks were made into thin sections and feldspars were analyzed quantitatively using a Cameca SX50 electron probe in the Division of Earth Science, University of Glasgow and a Cameca Camebax electron probe in the Grant Institute of Earth Science, University of Edinburgh. Data were acquired using an accelerating voltage of 20 kV and a beam current of 10 nA, and the beam was defocused to 20 mm diameter. Where required, following SEM and electron probe work, it is possible to do TEM using the polished thin sections, provided that they have been mounted using low-meltingtemperature resin. For this purpose 3.05 mm diameter copper discs were attached to the thin section with their central 0.4 mm diameter hole positioned over a grain of interest. The copper discs, with grains attached, were then extracted from the thin section and Ar ion beam thinned until electron-transparent using an Atom Tec 700 series thinner operated at , 6 kV (Edinburgh University) and a Gatan Duomill operated at 5 kV (Glasgow University). High-resolution TEM images and electron diffraction patterns were acquired from the thinned foils using a Philips CM120 Biotwin TEM operated at 120 kV (Edinburgh University) and a Jeol 200FX TEM operated at 200 kV (Glasgow University). A useful auxiliary SEM technique is to etch feldspar cleavage surfaces in HF vapor for short periods (typically , 50 s) followed by secondary electron (SE) imaging, as described by Waldron et al. (1994). An example is given as Figure 4A. The etching provides pronounced surface topography caused by rapid dissolution on dislocation outcrops, in regions of high coherency strain (along coherent or semicoherent perthite lamellar boundaries) and on subgrain boundaries in deuteric intergrowths. Less pronounced relief is produced by the different dissolution rates of Ab- and Or-rich feldspar. The method has higher resolution than BSE imaging (see e.g., fig. 2B in Waldron et al. 1994, which shows features , 50 nm across) but requires that natural fresh cleavage surfaces are available. An example of this technique applied to a detrital grain is given by Lee and Parsons (1998, their fig. 6c).
RECOGNITION OF ALKALI FELDSPAR MICROTEXTURES

Two factors must be borne in mind when applying our approach: firstly, the strong effect of orientation on the appearance of a perthitic intergrowth when sliced, and secondly that the relatively small fragments likely to be found in many siliciclastic rocks may be untypical of the

original grain as a whole. In general, strain-controlled microtextures and estimates of grain bulk composition are the most informative for provenance purposes (see Fig. 2 and photomicrographs below). Fragments of deuterically coarsened perthite are relatively uninformative except in cases where they are large enough to estimate bulk composition, at least roughly. While some of the strain-controlled intergrowths are sufficiently distinctive that provenance can be suggested from a single grain, in other cases it is best to deduce provenance from families of grains within the sedimentary rock. The appearance of a perthitic intergrowth varies with viewing direction (Fig. 3). Regular lamellar intergrowths appear diffuse and irregular or even patchy if viewed from a direction nearly normal to the lamellar interface. As a routine method we recommend looking for the most regular exsolution features visible using BSE imaging. Experiment with magnification and with the BSE contrast. The presence of any sort of regular lamellar intergrowth is secure evidence that a grain is detrital. The morphology of the intergrowths, the relative proportions of the Ab- and Or-rich phases, a CL image, and, if necessary, a qualitative chemical analysis can all be obtained in the SEM. Appreciation of the orientation of exsolution lamellae relative to the two perfect cleavages of feldspar is of considerable help in interpreting the random sections of grains encountered in a sedimentary rock. If OM is used, extinction parallel to the (010) cleavage and the (010) composition planes of Albite twins (which often can be seen in the Ab-rich phase of perthites) are useful guides to orientation. Albite twins are sometimes faintly visible using SE imaging in the SEM. Cleavages are often visible in BSE images (see e.g., Figs. 13 and 14), and there will be a tendency for detrital fragments to be bounded by either the (001) and (010) cleavages 01) and (8 01) (Fig. 3). or the less regular Murchison plane between (6 Regular lamellar intergrowths imaged approximately normal to the Murchison plane will appear irregular and patchy but will show two welldeveloped cleavages at right angles. True patch perthites are less informative than strain-controlled intergrowths, although their bulk composition is some guide to potential source rocks (Fig. 2). They often coexist with strain-controlled textures within individual original crystals or detrital fragments (e.g., Figs. 10, 11, and 12). However, patch perthites contain large concentrations of subgrain boundaries and micropores, are relatively easily weathered (Fig. 4D; Lee et al. 1998), and are likely to dissolve relatively rapidly during burial. There is a natural tendency for patch perthites to disappear relatively quickly from the inventory of clastic feldspars in a given sequence of weathering, transport, and burial. Using BSE imaging alone, alkali feldspar grains free of visible microtextures may be difficult to distinguish from discrete authigenic crystals, unless the latter have grown into pore spaces and developed the Adularia habit. The most likely source for detrital fragments of this type are acid and alkaline volcanic igneous rocks, and bulk composition can provide evidence of provenance (see diagrams in Chapter 14 of Smith and Brown 1988). Lack of an exsolution microtexture makes such fragments resistant to dissolution because of the absence of either dislocations or subgrain boundaries (Lee et al. 1998), and this probably accounts for the relative abundance of such grains in the Fulmar Formation (see below). Absence of cathodoluminescence and compositions close to end-member Ab and Or (Ab.99 or Or.99) are the best guides to an authigenic origin. If TEM is used, the presence of cryptoperthite would rule out an authigenic origin, as would the regular well-developed tartan AlbitePericline twinning characteristic of microcline or, less reliably, the tweed texture of orthoclase. It is sometimes possible to see tartan twinning in microcline using SEM-CL, because the small variation in the orientation of the twins affects luminescence intensity; this observation would rule out an authigenic origin. However, Figure 7 shows an example of a detrital grain and its overgrowth in which all microtextures are suboptical and which would be indistinguishable using BSE or CL. At the TEM scale the authigenic K-feldspar overgrowth exhibits the diffuse

932

I. PARSONS ET AL.

JSR

FIG. 13.BSE images of fine film perthite grains from the Fulmar Formation, with largely coherent lamellar microperthitic intergrowths. They are most likely to have originated in subsolvus granitic protoliths, or possibly from granulite facies metamorphic rocks (see Fig. 19). If quantitative analysis indicated bulk An contents . 3% the latter would be strongly suggested. Further examples are given in Figures 6A and 8. A) Fragment with irregularly distributed lamellae of variable size and well defined (001) cleavage (compare Fig. 3). The detrital grain has a partially resorbed K-feldspar overgrowth. B) Similar grain with variable texture such that intergrowths appear to be absent at the top center. This may be because the intergrowths in this part of the grain are of very fine scale, rather than of variable bulk composition (as in the case of the zone of fine intergrowth shown in Fig. 3). Straight edge at the left is the (001) cleavage. The grain has a strongly resorbed K-feldspar overgrowth. C) Grain with mainly straight film lamellae. A few lamellae are kinked, perhaps because of deformation. At the right the grain has fractured along the plane of the exsolution lamellae, the Murchison plane. A partially resorbed K-feldspar overgrowth is mainly developed at the bottom left of the grain. A second alkali feldspar grain on the left of the image has been partially dissolved, leaving a skeletal relic (see also Fig. 13). D) Grain with fine, relatively evenly developed texture, resembling the platelet zones shown in Figures 3 and 4A, but on a coarser scale. This grain probably has a more albitic bulk composition than those shown in Parts AC. The pale lines (Or-rich) with micropores could be healed fractures along which deuteric coarsening has occurred, or a diagenetic feature. A K-feldspar overgrowth has been almost completely resorbed.

modulated microtexture characteristic of adularia and the contrast between the detrital K-feldspar core and overgrowth is obvious (Fig. 7; see also Lee and Parsons 2003, their fig. 7a, for examples of adularia veining detrital K-feldspar and replacing perthitic albite). Detrital fragments that may be untypical of a larger original grain as a whole pose problems best overcome by considering families of grains in any one slice. The problem is most acute in the case of grains which originally had coarse deuteric microtextures but which have been fragmented. Thus a single grain of Ab-rich plagioclase in a sedimentary rock might have at least six sources: 1. 2. It originally formed as an Ab-rich rich patch in an alkali feldspar by exsolution It originally formed during relatively high-temperature replacement (albitization) reactions with an alkali feldspar (as shown by Lee and Parsons 1997b). It originally formed during low-temperature igneous albitization in such a parent (Lee and Parsons 1997b). It grew as a discrete plagioclase crystal in a sub-solvus granite or metamorphic rock

5. 6.

It formed by albitization of basaltic calcic plagioclase in an ocean floor environment (spilitization). It grew as an authigenic grain or overgrowth (since fragmented) during diagenesis.

3. 4.

Fragments of types (1), (2), and (4) will luminesce, but types (3) and (6) will not. Type (5) is uncertain. In general An contents increase from type (1) to type (3), although there is much overlap. Thus Ab-rich phase compositions in coarse microperthites in the Shap granite (Lee and Parsons 1997b) are , An1, in cross-cutting patch perthites , An10, and in coexisting discrete plagioclase , An27. Replacive albite of type (5) is , An0.2, potentially the same as potential authigenic grains (type 6). There are no obvious criteria for distinguishing types (3) and (6). In the Humber Group sandstones of the Fulmar Formation, Lee et al. (2003) showed that detrital plagioclase grains ranged up to An18, and that many were lamellar peristerite intergrowths. The presence of peristerite intergrowths rules out an authigenic origin, and a source in low-grade metamorphic rocks is implied. The Or-rich phase in patch perthites from granitic source rocks usually has an appreciable Ab content; in Shap the range is from Or90 to Or97.

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

933

FIG. 14.Detrital coarse film perthite grains from the Fulmar Formation, probably from granitic protoliths. These intergrowths are probably largely semicoherent. Parts A, B, and D are BSE images; Part C is a CL image. All the grains have authigenic K-feldspar overgrowths, but those in Parts AC appear to be stable while that in Part D is strongly resorbed. A) Lenticular microperthite with a euhedral adularia overgrowth. The (001) cleavage is visible in the detrital core. Coarse albite lamellae (dark gray) have dissolved, in some cases leaving open pores (black), but neither the detrital K-feldspar nor the overgrowth have dissolved (From Lee and Parsons 2003). B) Similar grain in which some albite lamellae have dissolved while others have survived, with a nearly continuous K-feldspar overgrowth. C) Same grain as Part B. The image shows the complete lack of CL from the authigenic overgrowth, which appears as dark as the voids left by the dissolution of albite lamellae. Undissolved albite lamellae show weak CL. D) Lamellae in this fragment have partially dissolved, particularly around the thicker parts of albite lenses, where dislocations are concentrated in semicoherent intergrowths (see Fig. 4A). Areas where the grain is particularly corroded may correspond with patch perthite. The trace of the (010) cleavage can be seen across the center of the grain.

The non-endmember composition and luminescence would serve to distinguish fragments from this source from authigenic grains. Furthermore, as noted above, optically visible tartan or, when TEM is used, suboptical tartan and tweed textures are almost universal in detrital Kfeldspar from plutonic sources (Fig. 7). Or-rich patches in patch perthites are usually microcline, often in the irregularly twinned variant known as irregular microcline.
ATLAS OF MICROTEXTURES

terranes. Deuteric coarsening is illustrated in these contexts. Some images show grains with several types of microtexture and are cross referenced from the relevant subsections. In addition, we comment in separate subsections on the role of the microtextures in the chemical reactivity and preservation of the grains, and we show how microtextures facilitate dissolution and replacement reactions. General points are made in the main text; details of interpretation of the micrographs are given in extended captions. Acid and Alkaline Volcanic Rocks.Provided that grains have escaped deuteric or hydrothermal reactions, volcanic Or-rich feldspars will be featureless in OM or show very fine-scale, approximately straight exsolution lamellae. Ab-rich volcanic alkali feldspars (anorthoclase) may in addition exhibit regular cross-hatched tartan twinning, similar to microcline (for making the distinction using OM, see nomenclature section). Volcanic alkali feldspars may likewise be featureless using BSE in the SEM (Figs. 8, 9A), or have faint straight perthitic lamellae. In the absence of visible microtexture at the SEM scale, the best way to distinguish such grains from authigenic feldspars is CL (Fig. 9B), because the latter will not luminesce. Alternatively, qualitative analysis in the SEM, showing appreciable Ab content, is a good guide to a detrital origin as is a resolvable Ca peak in the EDX spectrum. A high Ab content

In this section we illustrate our method with BSE and CL images (in the SEM) of microtextures in alkali feldspars from the Fulmar Formation, comparing them with examples from our work on igneous and metamorphic rocks. We have included a few images illustrating distinctive microtextures which we have not yet seen in the Fulmar. The worldwide range of chemical variation in igneous rocks is very large, so we have concentrated on the feldspar types that are most likely to be encountered as detrital grains. Figure 2 can be used as a guide to coherent and semicoherent microtextures in alkali feldspars from most plutonic igneous rocks. The images begin with feldspars that originated in volcanic rocks and lead through relatively rapidly cooled plutonic igneous rocks to deepseated igneous rocks and finally to very slowly cooled metamorphic

934

I. PARSONS ET AL.

JSR

(Ab.60Or,40), combined with tartan twinning, indicates volcanic anorthoclase. At the TEM scale even rapidly chilled volcanic crystals are usually exsolved (see Parsons and Brown 1984, their fig. 5, for an example), particularly if the bulk composition is in the middle region of the feldspar solvus. Presence of exsolution textures is unequivocal proof that a grain is detrital rather than authigenic. However, observing fine exsolution microtextures using TEM in randomly oriented grains is difficult, particularly when bulk compositions are close to the Or end member. A more practical technique in the case of volcanic K-feldspars is careful EPMA analysis for major and trace elements. The compositional range of volcanic alkali feldspars can extend continuously from nearly pure sanidine through ternary compositions to the plagioclase field, and the bulk composition can be a guide to possible parental rocks. We recommend comparing analyses with the compilations given by Smith and Brown (1988, Chapter 14). However, it should be noted that phenocryst and groundmass assemblages in individual volcanic rocks may show a considerable range in composition and strong chemical zoning (see Brown and Parsons 1994 for a brief treatment, and Brown 1993 for details of this complex subject), so that a varied population of volcanic feldspar fragments in a sedimentary rock need not imply a varied provenance. Reactivity and Preservation.The Fulmar Formation contains a high proportion of grains of sanidine which are devoid of microtextures at the SEM scale (Figs. 8, 9), but which show rare fine Albite-twinned albite lamellae at the TEM scale. The bulk composition of these grains is unusually Or-rich (typically Or88 to Or96), and some have high Ba contents (up to Csn4.5) (Mark Wilkinson, personal communication 2003). The most likely source is ultrapotassic volcanic or hypabyssal rocks, although where these extremely rare rock types might be located in the North Sea catchment is not known. In principle, a perfect crystal of sanidine, the structurally disordered form of K-feldspar stable at high temperatures, should dissolve slightly more rapidly in aqueous solutions at reservoir temperatures than a perfect crystal of microcline, the ordered form stable at low temperatures. However, the effect of microtexture on dissolution kinetics far outweighs the effect of structural state, and the lack of microtexture in the sanidine grains probably accounts for their relative abundance throughout the Fulmar. We would expect fresh volcanic feldspars to be considerably more resistant to degradation and dissolution during weathering, transport, and diagenesis than potentially much more common feldspars from granitic rocks (see Lee and Parsons 1995, 1998; Lee et al. 1998), because of the absence of dislocations on perthite lamellae and of boundaries to subgrains formed by deuteric coarsening in the latter (see Fig. 4). Milliken et al. (1989) showed that in the Frio Formation of South Texas, where the feldspar provenance is known, Na-bearing volcanic feldspars dissolved more rapidly with burial than more Or-rich grains. This might be accounted for by the abundance of exsolution textures in the former. Figure 9 shows that in the Fulmar the volcanic grains are less reactive than chemically similar authigenic Kfeldspar overgrowths, because the latter are made of subgrains and are microporous (Fig. 7, top left). Hypersolvus Acid and Alkaline Plutonic Igneous Rocks.Hypersolvus igneous rocks crystallize at temperatures above the strain-free ternary feldspar solvus (Fig. 1) and therefore contain a single feldspar phase prior to exsolution. They characteristically form relatively small, anorogenic plutons of alkali granite, syenite or nepheline syenite. The alkali feldspars show a wide range of coherent and semicoherent exsolution microtextures depending on bulk composition (Fig. 2) and cooling rate, modified in various degrees by deuteric coarsening. For phase-equilibrium reasons (Tuttle and Bowen 1958) the bulk compositions of the alkali feldspars are strongly concentrated in restricted ranges, , (Ab6555Or3545)10097An03 in syenites and nepheline syenites, and

(Ab6050Or4050)10098An02 in alkali granites, around 4 in Figure 2. The anorthite content is low (, An3) in most examples, but in syenitic magmas formed by fractionation, feldspar bulk compositions or chemical zonation within crystals vary continuously from the plagioclase join along curved paths roughly parallel to the path 1234 in Figure 2 (see Parsons and Brown 1988, and Brown and Parsons 1988, for examples). However, detrital fragments representing a sample of hypersolvus igneous protoliths would overwhelmingly lie in the restricted ranges given above. These are comparatively uncommon rocks, so that the feldspars have considerable potential as provenance indicators. In syenites, provided that deuteric feldsparfluid reactions have not occurred, braid perthites (Fig. 2 around 4, and Fig. 10) are common. The average repeat distance of the threads of microcline ranges from , 40 nm to . 500 nm (Brown et al. 1983) and depends on the cooling rate of the igneous body and An content (Brown et al. 1983; Brown and Parsons 1988; Waldron and Parsons 1992). Grains may have zigzag, wavy, or straight lamellae (Fig. 2, areas 5, 6, 7), but these types are less common than the braid intergrowth. The coarsest textures are just visible in OM, but all but the finest are readily imaged using BSE (Fig. 10, left). The finest intergrowths require TEM or use of HF etching and SE SEM (see, e.g., Waldron et al. 1994, fig. 2a). Deuteric reactions lead to coarsening of intergrowths by up 103, giving irregular patch perthites (Fig. 10, right). The coarsened areas are characterized by micropores (black dots in Fig. 10) at the boundaries of sub-mm subgrains (see Worden et al. 1990 for detail of the coarsening process). Although the feldspars in hypersolvus alkali granites have bulk compositions similar to those in syenites, they are usually characterized by vein perthites much coarser than braid textures (Fig. 11A). Rare relics of braid texture can sometimes be found (Fig. 11B). The example given is from a sheet of quartz syenite, an evolved late member of the Klokken intrusion with a higher water content than the syenite magmas. Lee et al. (1997) concluded that the vein perthites formed by almost completely pervasive deuteric recrystallization of braid perthites, reflecting the high water content. Reactivity and Preservation.We have not seen feldspars of these types in the Fulmar, although the Permian alkaline rocks of the Oslo graben, which are potential provenance rocks for the Fulmar, contain braid perthites (Muir and Smith 1956). Because braid perthites are fully coherent we would expect them to be considerably more resistant to dissolution than associated patch and film perthites, in which micropores and subgrain boundaries will be rapidly attacked by fluids (e.g., Fig. 4D). Subsolvus Granitic Rocks Including Acid Gneisses.Alkali feldspars in sub-solvus rocks crystallize in equilibrium with a plagioclase feldspar on an isothermal tieline such as PBAF in Figure 1. The majority of alkali feldspars in subsolvus igneous granites and granitic gneisses are in the range Ab25Or75Ab10Or90, usually with , 3% An (region 8 in Fig. 2) although a few may lie outside this range. Figure 12 shows a typical area of a phenocryst from the Shap granite. Regular strain-controlled perthites characteristically form nearly straight film lamellae which are initially fully coherent but which develop edge dislocations (become semicoherent) as they coarsen (Fig. 4A). When viewed parallel to Z the exsolution lamellae form short lenses; they are straighter and less markedly lenticular viewed parallel to the Y axis. In more Or-rich bulk compositions lamellae are less conspicuous than in relatively Ab-rich feldspars like those from Shap (near Ab25Or75). The thickest lamellae are a few mm thick (easily visible using OM and BSE in an SEM), but lamellae only a few nm thick may occur (Lee et al. 1995; also Fig. 4), below the resolution of the SEM. Images of perthitic feldspars found in the Fulmar Formation which are most probably from granitic protoliths are shown in Figures 13 to 15. Deuteric coarsening, giving either ragged vein perthites or patch perthites (Figs. 12 and 3), is common in granitic K-feldspars. If the bulk

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

935

FIG. 15.BSE images of detrital patch perthite grains from the Fulmar Formation, Part A largely undissolved, Parts BD showing different stages in dissolution. A) Patch perthite grain. The sub-regular distribution of albite (darker gray) throughout the entire grain, the edges of the albite patches parallel to faint relics of film lamellae, and the bulk composition of the grain all suggest that this is an original patch perthite from a granitic source. The grain has a partially resorbed authigenic K-feldspar overgrowth along its lower margin. B) Corroded grain in which much of the left-hand end is patch perthite (dark gray), much of which has dissolved, giving elongate pores (black). The right-hand end is devoid of visible microtexture and has not dissolved. The grain has a partially resorbed overgrowth (right lower corner). C) Highly corroded patch perthite in which only a few relics of perthitic albite have survived so that the grain is largely skeletal. A large patch of albite right of center is probably the result of diagenetic albitization. D) A feldspar grain at the right is very heavily corroded and unrecognizable. The grain at center left, a film perthite that has been albitized, has survived.

composition of the fragment is in the Ab25Or75Ab10Or90 range it is reasonable to assume that these textures formed isochemically during igneous cooling of a granitic protolith. Figure 14A shows a grain of this type. However, as we noted above (Fig. 6), non-isochemical replacive albitization can occur during both igneous cooling and diagenesis. Thus the regular film intergrowths and grain bulk composition are the most reliable provenance indicators. Reactivity and Preservation.Figures 13 and 14 illustrate the effect of dislocations on the reactivity of perthitic feldspars. The lamellae in Figure 13 are relatively thin, and the grains are relatively Or-rich. From the lack of dissolution features we infer that they are largely coherent intergrowths. There is no sign of dissolution preferentially affecting exsolution lamellae or their boundaries. Lamellae in Figure 14 are coarser, with thicker lenticular cross sections, and some of have been preferentially dissolved while others have been preserved. We infer that these thicker lamellae have dislocations along their boundaries and those that were exposed at the surface of the grains have dissolved leaving lenticular voids, while those not reaching the grain surface have been preserved. Figure 14D shows an early stage in this process. Similar behavior was noted in a Carboniferous conglomerate at Shap (Lee and Parsons 1998).

Coarse patch and vein perthites in subsolvus granitic feldspars have complex subgrain microtextures and are very reactive during weathering (Fig. 4D) and are easily corroded during diagenesis (Fig. 15BD). They are likely to be the first grains to be lost from the alkali feldspar inventory of a siliciclastic sediment, and a preponderance of fine intergrowths (Fig. 13) is likely to be encountered in suites of detrital alkali feldspars from a granitic protolith. Low-Grade Metamorphism of Granitic Rocks.Some alkali feldspars exhibit a distinctive texture know as flame perthite. Detrital examples are shown in Figure 16. The Ab-rich phase in flame perthites adopts distinctive sinuous shapes, which sometimes bifurcate. The lamellae are rather coarse and are irregularly developed throughout the K-feldspar host and are often concentrated around the margins of grains. The genesis of flame perthite has been discussed by Pryer et al. (1995) and Pryer and Robin (1995, 1996). They suggest that the flames represent a coherent replacement perthite although they do not provide electron-microscope evidence of the character of the interfaces. Flame perthite occurs in granitic rocks sensu lato which have been subject to brittle deformation in the greenschist facies, much less commonly in the lower amphibolite facies. The texture does not occur in cataclastic rocks below the

936

I. PARSONS ET AL.

JSR

FIG. 16. BSE images of detrital flame perthite grains from the Fulmar Formation. There is no sign of corrosion guided by microtextures, suggesting that these intergrowths are coherent. A) The large grain is a well preserved fragment of coarse flame perthite. Note the distinctive shape of the albite lamellae, their tendency to bifurcate, and their heterogeneous distribution. The two smaller feldspar grains are sanidines, devoid of microtexture at this scale. That on the right has an authigenic Kfeldspar overgrowth. B) Flame perthite fragment with areas of film perthite, particularly at the lower left corner of the grain.

greenschist facies. Thus flame perthite seems to have a relatively limited metamorphic provenance. Reactivity and Preservation.Flame perthites are uncommon and have a limited paragenesis. At first sight it is surprising that they have been found in the Fulmar in abundance similar to film and patch perthites from common granitic rocks. However, the two examples shown from the Fulmar (Fig. 16) are devoid of the selective dissolution seen in other coarse perthites (Figs. 14, 15), which supports the contention of Pryer and coworkers that the flames are coherent. Thus the relative abundance of the flame perthites in the Fulmar reflects microtexture that is resistant to degradation rather than a widespread source. Granulite Facies Metamorphic Rocks and Charnockites.High-grade metamorphism in the granulite facies may involve crystallization of alkali feldspar at very high temperature and pressure, sometimes . 1050uC and 1.2 GPa. At these conditions, because of the shape of the ternary feldspar solvus (Fig. 1) it is possible for ternary alkali feldspars with high An contents to grow provided that water activities are low enough to preclude melting. The exsolution microtextures are varied and here we concentrate on the commonest and most distinctive types using examples from Antarctic granulite facies suites studied by Cayzer (2002). Other illustrations or discussions of granulite feldspar microtextures can be found in Hubbard (1965), Carstens (1967), Machado (1970), Kay (1977), Yund et al. (1980), Hayob et al. (1989), Waldron et al. (1993), Evangelakakis et al. (1993), and Raase (1998, 2000). Charnockites are granitic rocks emplaced under dry granulite-facies crustal conditions; their feldspar microtextures are similar to those in metamorphic granulites. The range of ternary feldspar bulk compositions in granulites is large, varying almost continuously from antiperthites in the vicinity of 1 (Fig. 2) through mesoperthites near 3 and 5 and perthites near 8 and extending to near the AbOr join, essentially in the field of granites (Fig. 12). Feldspars with large amounts of all three feldspar components are often said to be strongly ternary and this characteristic of granulitic feldspars is matched only by those in some uncommon syenitic rocks (see hypersolvus igneous rocks, above; Parsons and Brown 1988). Four microtextural features are distinctive of granulite-facies alkali feldspars: (1) the microtextures have become relatively coarse (. or & 1 mm) while remaining coherent or semicoherent; (2) there is often evidence for at least two stages of coherent or semicoherent exsolution; (3) micropores and

therefore turbidity are rare or absent; (4) patch perthites are uncommon. All these features are consistent with growth at high temperature followed by relatively slow cooling under dry conditions. Micromesoperthites (Fig. 17) are one of the most distinctive granulite facies feldspar types; they are common in the Fulmar Formation (Fig. 18). The film lamellae are sinuous and the intergrowths are coarse compared with the lamellae in syenites, yet dislocations are rare or absent (Fig.17A, B). There are superficial similarities to the vein mesoperthites found in hypersolvus granites (compare with Fig. 11), but the lamellae in Figure 11 are ragged and pervaded by micropores (Fig. 11B), which are almost completely absent from the granulite feldspars. Granulite mesoperthites with coarse, stubby lenses of oligoclase (Fig.17C) and complex intergrowths (Fig. 17D), with two lamellar orientations, are also encountered. The origin of the latter texture is uncertain but is probably the result of reorganization of an original mesoperthite under the influence of deformation (Cayzer 2002). A similar texture was illustrated by Hokada (2001, his fig. 2c). As far as we are aware such textures are distinctive of the granulite facies. A special case of granulite facies mesoperthite was reported by Hayob et al. (1989) in xenoliths in Quaternary volcanic rocks from Mexico. Unusually, they found that the intergrown phases in these mesoperthites were far from the Ab and Or end-members, and they deduced that the cooling history had been abruptly terminated at 800950uC, when they ascended rapidly from hot lower crust. Discovery of coarse mesoperthitic intergrowths with non-end-member phase compositions in detrital grains would be an extremely distinctive provenance indicator. Perthites (sensu stricto) in the granulite facies (Fig. 19A) may form straight lamellae very similar to those in feldspars from subsolvus granites (Figs. 12, 13), and it may be difficult to assign a granulite-facies origin unequivocally to a detrital fragment of such a feldspar. However, very regular lamellae, with few dislocations and micropores, without deuteric coarsening, and relatively high An + Csn, would strongly suggest a granulite source. Two-stage perthites (Fig. 19BD) are very distinctive of the granulite facies and can also be matched in the Fulmar (Fig. 20A, B), although care must be taken to differentiate possible patchy diagenetic albitization (Fig. 20C), using CL (Fig. 20D). Two-stage perthites contain both fine (usually sub-optical) lamellae and much coarser rounded blebs which often have coarse Albite twinning (Fig. 19C). The blebs are semicoherent when small, incoherent when large, and often form trails across grains (Fig. 19D). Two-stage perthites have been described by many authors (Eskola 1952; Machado 1970; Yund and Ackermand 1979; Yund

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

937

FIG. 17.Granulite facies mesoperthites, from the Napier Complex, Antarctica (Cayzer 2002). A) SE image of an etched (001) cleavage surface of a film mesoperthite, bulk composition Ab35.3An3.2Cn1Or60.6. The Ab-rich lamellae, which stand out in relief, are relatively coarse but almost fully coherent, the whole image being nearly devoid of etch pits on dislocations. There are three pairs of etch pits, one at top right and two halfway up at the left. This is a true mesoperthite in which both phases are continuous in three dimensions (contrast the more Or-rich granitic feldspars in Figs. 1214, where there is a clear Or-rich host phase and the albite is discontinuous). B) BSE image of a granulite-facies gneiss showing many mesoperthitic feldspar crystals. Pale gray is the Or-rich phase, mid gray is the Ab-rich phase. The feldspar bulk composition is strongly ternary, Ab49.2An19.5Cn0.1Or30.8. The high An content, the smoothly sinuous character of the lamellae, and the absence of micropores are diagnostic of the granulite facies. Some of the grains have continuous plagioclase rims, another distinctively granulite feature. C) SE image of an etched (001) cleavage surface of lenticular mesoperthite, bulk composition Ab32.0An9.5Cn0.2Or58.4. There are some extended lamellae, as in Part A, but the majority have formed stubby lenses and appear to be discontinuous in the K-feldspar host. Although the lamellae are coarse there are few dislocation etch pits, although some pairs are distributed over the micrograph. Other parts of this sample have more etch pits and are more semicoherent. D) A complex, very coarse ternary mesoperthite, bulk composition Ab55.4An17.0Cn0.6Or26.9, which has two, almost perpendicular lamellar orientations. Micropores are almost completely absent.

et al. 1980; Waldron et al. 1993; Evangelakakis et al. 1993; Raase 1998), and they occurred in all four Antarctic terranes studied by Cayzer (2002). Antiperthites also occur in granulite facies rocks (Fig. 21). They are unusually coarse. Although in principle they should be common, antiperthites are not often reported in the petrological literature, possibly because they are usually suboptical in scale (cryptoantiperthites; see Brown and Parsons 1988). The textures shown in Figure 21 consist of blebs (rods in three dimensions) of a range of coarseness and a variety of shapes, and they arise by a combination of exsolution followed by replacement reactions. Aligned rods with a distinctive rhombic cross section (Fig. 21B) appear to have nucleated on Albite twin planes in the plagioclase phase (Cayzer 2002). The strongly ternary bulk composition (caption to Fig. 21) would be highly suggestive of a granulite-facies origin. We have not seen antiperthites of this type in the Fulmar Formation. Reactivity and Preservation.Considering that high-grade granulite facies rocks are not abundant, and often occur as relatively small enclaves in regions of lower-grade metamorphism, granulite-facies alkali feldspars

are remarkably abundant in the Fulmar. This is likely to be because of the general lack of dislocations on regular lamellae, and lack of deuteric recrystallization in feldspars from these very dry rocks. Figure 18D illustrates a small but well-preserved detrital fragment of granulite-facies mesoperthite in contact with a much more corroded alkali feldspar, perhaps originally a granitic patch perthite.
DISCUSSION: PROVENANCE AND RESERVOIR QUALITY

The micrographs above (Figs. 821) illustrate the ability of exsolution microtextures in alkali feldspars to provide provenance information in siliciclastic rocks. They cover the most common textural types that occur worldwide, many of which have been encountered in the Fulmar Formation. The alkali feldspar inventory of the Fulmar Formation contains a preponderance of types which on a North Sea catchment-wide scale would be comparatively rare. This is particularly true of the volcanic sanidines and the mesoperthites and two-stage perthites from granulite-facies rocks. Feldspars from subsolvus granites, which should on the basis of areal extent at outcrop be common, do not predominate

938

I. PARSONS ET AL.

JSR

FIG. 18.BSE images of detrital fragments of mesoperthitic alkali feldspars from the Fulmar Formation, most probably originating in granulite-facies rocks. Ab-rich phase darker gray, Or-rich phase lighter gray. A) Coarse mesoperthite, with Ab . Or. Note slightly sinuous lamellae, and wedge shaped ends to lamellae of both phases. B) Complex, irregular mesoperthite in which lamellae at the left have coalesced. This type of complexity is common in granulite facies rocks. C) Coarse mesoperthite with bulk Or . Ab. The round albitic area at the right shows strong CL (image not shown), suggesting that it is part of the original grain and not diagenetic replacement. D) The feldspar at the right, a very coarse mesoperthite, has resisted dissolution, while the grain at the left, probably a patch perthite, is strongly resorbed.

and are often strongly corroded. The relative abundance of detrital alkali feldspar types is not a simple crustal average, but reflects the relative ease of destruction of the original feldspars during weathering, transport, and diagenesis. The feldspar types that degrade most rapidly do so not because of their chemistry but because of the defect inventory of the grains, particularly the presence of edge dislocations on semicoherent albite lamellae or subgrain boundaries and micropores in deuterically recrystallized feldspars. Both of these microtextural features are absent or rare in the feldspar types which are most frequently preserved. Dissolution is strongly texture-sensitive, and provenance therefore exerts a strong control on the rate of development of secondary porosity and hence on reservoir quality. Velbel (1989) stressed the importance of crystal defects in reactions with solutions close to equilibrium during weathering, and some of the factors controlling the reactions of feldspars with diagenetic pore fluids have been discussed recently by Parsons and Lee (2000). During steady burial of an arkosic sandstone containing a static fluid, over millions of years, we can expect that pore fluids will relatively rapidly reach a metastable equilibrium with all the solids. All equilibria in siliciclastic rocks will strictly be metastable, because the sediment is an adventitious assemblage of solids far from stable equilibrium, and with no prospect of significantly approaching a stable assemblage at diagenetic temperatures, even over hundreds of millions of years. In a sedimentary rock composed of quartz with plagioclase and alkali feldspar, the solution

will become saturated relatively rapidly in SiO2 and feldspar components, the fluid reaching a local metastable equilibrium through the pore network. Parts of feldspar grains which are particularly reactive because of their microtextures will dissolve first, and lead to rapid approach to saturation. Grains which lack rate-enhancing microtexture, even if of the same chemical composition as the reactant grains, will be preserved. Depending on fluid composition, and particularly at alkaline pH, remaining feldspar will be stable with respect to the saturated fluid. If, as must usually be the case, the fluid is not static, the metastable equilibria will involve larger rock volumes, depending on porosity, permeability, and rates of fluid movement. As temperature and pressure increase with burial the fluid will change composition to maintain metastable equilibrium, either by dissolving or precipitating quartz and/or the appropriate feldspar. If fluid movements are slow, reactions will take place in fluids which are very close to equilibrium with the solid phases, conditions under which the free energy contribution of defects is particularly crucial (Velbel 1989). We would expect to see the strong dependence on microtexture that we observe in the Fulmar Formation under these conditions. In basinal sedimentary rocks subject to more rapid fluid fluxes we might expect to see more progressive, and ultimately complete, feldspar dissolution, provided that there is mass transfer of feldspar components out of the formation.

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS

939

FIG. 19.Perthites senso strictu from granulite facies rocks, Parts AC from Brattstrand Bluffs and Part D from the Napier Complex, East Antarctica (Cayzer 2002). A) SE image of an etched (001) cleavage surface with very regular, straight, cryptoperthitic and microperthitic film lamellae, bulk composition Ab22.8An2.0 Cn0.5Or74.8. A few etch pits mark dislocation outcrops on thicker lamellae, but otherwise the microtexture is coherent. B) BSE image of a two-stage microperthite, bulk composition Ab23.0An2.1Cn1.2Or73.8. Coarse rounded blebs of darker gray plagioclase occur irregularly in a matrix which contains lenticular film lamellae at a range of scales. The film lamellae seem to have nucleated on the blebs. Two stage textures are very distinctive of granulite facies rocks. C) SE image of an etched (001) cleavage surface of a similar two-stage perthite, bulk composition Ab23.6An2.6Cn0.7Or73.5, showing film lamellae at various scales, larger ones with etch pits, and oval blebs of plagioclase. The faint bands on the surface of the blebs are Albite twins. D) BSE image of a two-stage texture in which the blebs form trails in a matrix of more lenticular lamellae. This rough alignment of blebs is a common feature of granulite-facies alkali feldspars. The bulk composition is Ab30.2An4.6Cn0.8Or64.4; the high An is unequivocally indicative of a granulite source.

Changing temperature and/or influxes of fluid of different composition could lead to selective dissolution of the Ab- or Or-rich feldspar phases. Extrapolating the experimental work of Orville (1963) from the temperature of geothermal systems to those of diagenesis suggests that albite would be preferentially dissolved over K-feldspar if a cool fluid replaced a hotter one (as we see in Fig. 14). In a fluid column in a temperature gradient albite would precipitate at the hotter end as Kfeldspar dissolved, while K-feldspar precipitated at the cooler end and albite dissolved. Increasing albitization with burial is well established in sedimentary basins (e.g., Milliken 1989). The Fulmar rocks contain examples of diagenetic albitization (Fig. 15), and both stable (Fig. 14A, B) and unstable (Fig. 13A, B, D) K-feldspar overgrowths. We interpret these as indicating changes in fluid composition and changing temperature, the source reactant solids being local reactant parts of detrital grains. Basin-scale mass transfer of SiO2 and Al2O3 is not implied by these reactions, which will occur during local dissolutionreprecipitation reactions in which Na and K are exchanged between the fluid and solids. The persistence of relatively large amounts of detrital feldspar in the Fulmar Formation since the Late Jurassic, now at high temperature and pressure, and the clear role of intracrystal microtexture in defining

reactive and relatively inert grains, implies the presence of fluids in metastable equilibrium, or very close to metastable equilibrium, with the solid assemblage over long periods.
FUTURE DEVELOPMENT

In the context of the evolution of the petroleum-bearing formations of the North Sea it would be instructive to explore the feldspar inventories of Triassic and perhaps Devonian arkosic sandstones which might have been reworked as source rocks for the Fulmar. Our SEM and TEM observations on the Fulmar Formation also show several phases of albitization, authigenic K-feldspar growth, and subsequent dissolution which must represent specific events in reservoir evolution. A systematic basin-wide study of the Fulmar using our techniques would be required to evaluate the temperature and fluid compositional changes that they imply, and their relative timing. On a world-wide scale it would be of great interest to apply these methods to other basins, to establish the provenance of the formations, and to see if the natural sorting by defect inventory which seems to have applied in the Fulmar is a widespread phenomenon.

940

I. PARSONS ET AL.

JSR

FIG. 20. A, B) BSE images of two-stage intergrowths from the Fulmar Formation, indicating a granulite facies provenance. The Ab-rich blebs in Part A (dark gray) have a shared asymmetry, suggesting that they and associated lenticular lamellae have been deformed. Blebs in Part B are symmetrical. C) BSE image of what appears to be a two-stage intergrowth. However, micropores occur in Ab-rich areas, suggesting that they have formed by replacement. D) CL image corresponding with Part C showing that the blebs are completely non-luminescent. This, coupled with the micropores, suggests that they formed by diagenetic replacement, although there is some ambiguity (compare Fig. 6).

ACKNOWLEDGMENTS

The work on Fulmar was supported in part by Minor Service Contract 2200001839 from Shell U.K. Expro. Background images on

feldspars and the role of microtextures in feldspar dissolution came from a series of studies supported by the Natural Environment Research Council, most recently GR3/10290 and currently NER/A/S/2001/01099. N.C. acknowledges an N.E.R.C. research studentship GT04/98/82/ES.

FIG. 21.BSE images of coarse antiperthites from the Rauer Group granulites, East Antarctica (Cayzer 2002). Bulk composition Ab57.8An25.3Cn0.1Or16.8. A) Irregular blebs of Or-rich feldspar (light gray) in plagioclase (Ab67An30Or3, darker gray) host. At the right of the image, near the margin of the grain, rows of fine blebs are aligned along Albite twin boundaries. B) Higher-magnification image of area of fine blebs. The blebs have a roughly rhombic cross section.

JSR

ALKALI FELDSPAR MICROTEXTURES AS PROVENANCE INDICATORS


REFERENCES

941

ALDAHAN, A.A., MORAD, S., AND COLLINI, B., 1987, Clouded-untwinned albite in the Siljan granite, central Sweden: Neues Jahrbuch fu r Mineralogie, Monatshefte, Jahrgang 1987, p. 327335. ARRIBAS, J., CRITELLI, S., LE PERA, E., AND TORTOSA, A., 2000, Composition of modern stream sand derived from a mixture of sedimentary and modern source rocks (Henares River, Central Spain): Sedimentary Geology, v. 133, p. 2748. BASKIN, Y., 1956, A study of authigenic feldspars: Journal of Geology, v. 64, p. 132155. BERNER, R.A., AND HOLDREN, G.R., JR, 1977, Mechanism of feldspar weathering: some observational evidence: Geology, v. 5, p. 369372. BERNER, R.A., AND HOLDREN, G.R., JR, 1979, Mechanism of feldspar weatheringII. Observations of feldspars from soils: Geochimica et Cosmochimica Acta, v. 43, p. 11731186. BLUM, A.E., AND LASAGA, A.C., 1987, Monte Carlo simulations of surface reaction rate laws, in Stumm, W., ed., Aquatic Surface Chemistry: New York, J. Wiley and Sons, p. 255292. BROWN, W.L., 1993, Fractional crystallization and zoning in igneous feldspars: ideal water-buffered liquid fractionation lines and feldspar zoning paths: Contributions to Mineralogy and Petrology, v. 113, p. 115125. BROWN, W.L., AND PARSONS, I., 1984, Exsolution mechanisms and kinetics in an ordered cryptoperthite series: Contributions to Mineralogy and Petrology, v. 86, p. 318. BROWN, W.L., AND PARSONS, I., 1988, Zoned ternary feldspars in the Klokken intrusion: exsolution textures and mechanisms: Contributions to Mineralogy and Petrology, v. 98, p. 444454. BROWN, W.L., AND PARSONS, I., 1989, Alkali feldspars: ordering rates, phase transformations and behaviour diagrams for igneous rocks: Mineralogical Magazine, v. 53, p. 2542. BROWN, W.L., AND PARSONS, I., 1993, Stored elastic strain energy: the driving force for low temperature reactivity and alteration of alkali feldspar, in Boland, J., and Fitz Gerald, J.D., eds., Defects and Processes in the Solid State: Geoscience Applications, The McLaren Volume: Amsterdam, Elsevier, p. 267290. BROWN, W.L., AND PARSONS, I., 1994, Feldspars in igneous rocks: in Parsons, I., ed., Feldspars and Their Reactions, NATO, ASI Series, v. C 421 Kluwer Academic Publishers, p. 449499. BROWN, W.L., BECKER, S.M., AND PARSONS, I., 1983, Cryptoperthites and cooling rate in a layered syenite pluton: a chemical and TEM study: Contributions to Mineralogy and Petrology, v. 82, p. 1325. BROWN, W.L., LEE, M.R., WALDRON, K.A., AND PARSONS, I., 1997, Strain-driven disordering of low microcline to low sanidine during partial phase separation in microperthites: Contributions to Mineralogy and Petrology, v. 127, p. 305313. CARSTENS, H., 1967, Exsolution in ternary feldspars I. On the formation of antiperthites: Contributions to Mineralogy and Petrology, v. 14, p. 2735. CAYZER, N., 2002, Feldspar microtextures and the cooling histories of high-grade terrains [Ph.D. thesis]: University of Edinburgh, U.K., 300 p. EGGLETON, R.A., AND BUSECK, P.R., 1980, The orthoclasemicrocline inversion: a high resolution transmission electron microscope study and strain analysis: Contributions to Mineralogy and Petrology, v. 74, p. 123133. ESKOLA, P., 1952, On the granulites of Lapland: American Journal of Science, Bowen Volume, p. 133171. EVANGELAKAKIS, C., KROLL, H., VOLL, G., WENK, H., MEISHENG, H., AND KOPCKE, J., 1993, Low temperature coherent exsolution in alkali feldspars from high-grade metamorphic rocks of Sri Lanka: Contributions to Mineralogy and Petrology, v. 114, p. 519532. FERRY, J.M., 1985, Hydrothermal alteration of Tertiary igneous rocks from the Isle of Skye, northwest Scotland II. Granites: Contributions to Mineralogy and Petrology, v. 91, p. 283304. FITZ GERALD, J.D., AND MCLAREN, A.C., 1982, The microstructures of microcline from some granitic rocks and pegmatites: Contributions to Mineralogy and Petrology, v. 80, p. 219229. FITZ GERALD, J.D., AND PARSONS, I., 2003, A TEM and SEM study of the heating behaviour of alkali feldspar from a subsolvus granite: Geophysical Research Abstracts, v. 5, abstract 04875, EGS-AGU-EUG Joint Assembly, Nice, France. FUHRMAN, M.L., AND LINDSLEY, D.H., 1988, Ternary feldspar modelling and thermometry: American Mineralogist, v. 73, p. 201215. GLOVER, J.E., AND HOSEMAN, P., 1970, Optical data on some authigenic feldspars from Western Australia: Mineralogical Magazine, v. 37, p. 588593. GUTHRIE, G.D., AND VEBLEN, D.R., 1991, Turbid alkali feldspars from the Isle of Skye, Scotland: Contributions to Mineralogy and Petrology, v. 108, p. 398404. RREZ, F., AND ARANDA-GO MEZ, J.J., HAYOB, J.L., ESSENE, E.J., RUIZ, J., ORTEGA-GUTIE 1989, Young high-temperature granulites from the base of the crust in central Mexico: Nature, v. 342, p. 265268. HOKADA, T., 2001, Feldspar thermometry in ultra-high temperature metamorphic rocks: evidence of crustal metamorphism attaining , 1100uC in the Archaean Napier Complex, East Antarctica: American Mineralogist, v. 86, p. 932938. HUBBARD, F.H., 1965, Antiperthite and mantled feldspar textures in charnockite (enderbite) from S. W. Nigeria: American Mineralogist, v. 50, p. 20402051. KAY, S.M., 1977, The origin of antiperthites in anorthosites: American Mineralogist, v. 62, p. 905912. LEE, M.R., AND PARSONS, I., 1995, Microtextural controls of weathering of perthitic alkali feldspars: Geochimica et Cosmochimica Acta, v. 59, p. 44654488.

LEE, M.R., AND PARSONS, I., 1997a, Compositional and microtextural zoning in alkali feldspars from the Shap granite and its geochemical implications: Geological Society of London, Journal, v. 154, p. 183188. LEE, M.R., AND PARSONS, I., 1997b, Dislocation formation and albitization in alkali feldspars from the Shap granite: American Mineralogist, v. 82, p. 557570. LEE, M.R., AND PARSONS, I., 1998, Microtextural controls of diagenetic alteration of detrital alkali feldspars: a case study of the Shap conglomerate (Lower Carboniferous), Northwest England: Journal of Sedimentary Research, v. 68, p. 198211. LEE, M.R., AND PARSONS, I., 2003, Microtextures of authigenic Or-rich feldspar in the Upper Jurassic Humber Group, UK North Sea: Sedimentology, v. 50, p. 112. LEE, M.R., WALDRON, K.A., AND PARSONS, I., 1995, Exsolution and alteration microtextures in alkali feldspar phenocrysts from the Shap granite: Mineralogical Magazine, v. 59, p. 6378. LEE, M.R., WALDRON, K.A., PARSONS, I., AND BROWN, W.L., 1997, Feldsparfluid interactions in braid microperthites: pleated rims and vein microperthites: Contributions to Mineralogy and Petrology, v. 127, p. 305313. LEE, M.R., HODSON, M.E., AND PARSONS, I., 1998, The role of intragranular microtextures and microstructures in chemical and mechanical weathering: direct comparisons of experimentally and naturally weathered alkali feldspars: Geochimica et Cosmochimica Acta, v. 62, p. 27712788. LEE, M.R., THOMPSON, P., POEML, P., AND PARSONS, I., 2003, Peristeritic plagioclase in North Sea hydrocarbon reservoir rocks: implications for diagenesis, provenance and stratigraphic correlation: American Mineralogist, v. 88, p. 866875. MACHADO, I.F., 1970, Orientation of perthite lamellae in granulite facies rocks: Academia Brasileira de Cie ncias, Anais, v. 42, p. 4567. MACKENZIE, W.S., AND SMITH, J.V., 1956, The alkali feldspars. III. An optical and X-ray study of high temperature feldspars: American Mineralogist, v. 41, p. 405427. MARKS, M., AND MARKL, G., 2001, Fractionation and assimilation processes in the alkaline augite syenite unit of the Ilimaussaq Intrusion, South Greenland, as deduced from phase equilibria: Journal of Petrology, v. 42, p. 19471969. MASON, R.A., 1982, Trace element distributions between the perthite phases of alkali feldspars from pegmatites: Mineralogical Magazine, v. 45, p. 101106. MILLIKEN, K.L., 1989, Petrography and composition of authigenic feldspars, Oligocene Frio Formation, South Texas: Journal of Sedimentary Petrology, v. 59, p. 361374. MILLIKEN, K.L., MCBRIDE, E.F., AND LAND, L.S., 1989, Numerical assessment of dissolution versus replacement in the subsurface destruction of detrital feldspars, Oligocene Frio Formation, South Texas: Journal of Sedimentary Petrology, v. 59, p. 740757. ` A, J.A., 1989, Diagenetic K-feldspar MORAD, S., MARFIL, R., AND DE LA PEO pseudomorphs in the Triassic Buntsandstein sandstones of the Iberian range, Spain: Sedimentology, v. 36, p. 635650. MORAD, S., 1986, Albitization of K-feldspar grains in Proterozoic arkoses and greywackes from Southern Sweden: Neues Jahrbuch fu r Mineralogie, Monatshefte, Jahrgang 1986, p. 145186. MUIR, I.D., AND SMITH, J.V., 1956, Crystallisation of feldspars in larvikites: Zeitschrift fu r Kristallographie, v. 107, p. 182195. OGUNYOMI, O., MARTIN, R.F., AND HESSE, R., 1981, Albite of secondary origin in Charny Sandstones, Quebec: a reevaluation: Journal of Sedimentary Petrology, v. 51, p. 597606. ORVILLE, P.M., 1963, Alkali ion exchange between vapor and feldspar phases: American Journal of Science, v. 261, p. 201237. PARSONS, I., AND BROWN, W.L., 1984, Feldspars and the thermal history of igneous rocks, in Brown, W.L., ed., Feldspars and Feldspathoids; Structure, Properties and Occurrences, NATO, ASI Series, v. C137. Dordrecht, The Netherlands, Reidel Publishing Co., p. 317371. PARSONS, I., AND BROWN, W.L., 1988, Sidewall crystallization in the Klokken intrusion: zoned ternary feldspars and coexisting minerals: Contributions to Mineralogy and Petrology, v. 98, p. 431443. PARSONS, I., AND BROWN, W.L., 1991, Mechanisms and kinetics of exsolution structural control of diffusion and phase behavior in alkali feldspars, in Ganguly, J., ed., Diffusion and Flow in Minerals and Fluids: Berlin, Springer-Verlag, Advances in Physical Geochemistry, v. 9, p. 306346. PARSONS, I., AND LEE, M.R., 2000, Alkali feldspars as microtextural markers of fluid flow, in Stober, I., and Bucher, K., eds., Hydrogeology of Crystalline Rocks: Dordrecht, The Netherlands, Kluwer Academic Publishers, p. 2750. PARSONS, I., BROWN, W.L., AND SMITH, J.V., 1999, 40Ar/39Ar thermochronology using alkali feldspars: real thermal history or mathematical mirage of microtexture?: Contributions to Mineralogy and Petrology, v. 136, p. 92110. PRYER, L.L., AND ROBIN, P.-Y.F., 1995, Retrograde metamorphic reactions in deforming granites and the origin of flame perthite: Journal of Metamorphic Geology, v. 13, p. 114. PRYER, L.L., AND ROBIN, P.-Y.F., 1996, Differential stress control on the growth and orientation of flame perthite: a palaeostress-direction indicator: Journal of Structural Geology, v. 18, p. 11511166. PRYER, L.L., ROBIN, P.-Y.F., AND LLOYD, G.E., 1995, An SEM electron-channeling study of flame perthite from the Killarney granite, Southwestern Grenville Front, Ontario: Canadian Mineralogist, v. 33, p. 333347. PUTNIS, A., 2002, Mineral replacement reactions: from macroscopic observations to microscopic mechanisms: Mineralogical Magazine, v. 66, p. 689708.

942

I. PARSONS ET AL.

JSR

M, O., 1992, Petrological RAMSAYER, K., ALDAHAN, A.A., COLLINI, B., AND LANDSTRO modifications in granitic rocks from the Siljan impact structure: evidence from cathodoluminescence: Tectonophysics, v. 216, p. 195204. RASSE, P., 1998, Feldspar thermometry: a valuable tool for deciphering the thermal history of granulite-facies rocks, as illustrated with metapelites from Sri Lanka: Canadian Mineralogist, v. 36, p. 6786. RAASE, P., 2000, Orientation of exsolution lamellae and rods, and optimal phaseboundaries in antiperthite from pelitic granulites, Sri Lanka: Canadian Mineralogist, v. 38, p. 695705. SAIGAL, G.C., MORAD, S., BJRLYKKE, K., EGEBERG, P.K., AND AAGAARD, P., 1988, Diagenetic albitization of detrital K-feldspar in Jurassic, Lower Cretaceous, and Tertiary clastic reservoir rocks from offshore Norway, I. Textures and origin: Journal of Sedimentary Petrology, v. 58, p. 10031013. SMITH, J.V., 1974. Feldspar Minerals, first edition, Volume 2: Berlin, Springer Verlag, 690 p. SMITH, J.V., AND BROWN, W.L., 1988. Feldspar Minerals, second edition, Volume 1: Berlin, Springer Verlag, 828 p. SMITH, P., AND PARSONS, I., 1974, The alkali-feldspar solvus at 1 kilobar water vapour pressure: Mineralogical Magazine, v. 39, p. 747767. SNOW, E., AND YUND, R.A., 1988, Origin of cryptoperthites in the Bishop Tuff and their bearing on its thermal history: Journal of Geophysical Research, v. 93, p. 89758984. STEWART, D.J., 1986, Diagenesis of the shallow marine Fulmar Formation of the Central North Sea: Clay Minerals, v. 21, p. 537564. STOCKBRIDGE, C.P., AND GRAY, D.I., 1991, The Fulmar Field, Blocks 30/16 & 30/11b, UK North Sea, in Abbots, I.L., ed., United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume: Geological Society of London, Memoir 14, p. 309316. TUTTLE, O.F., AND BOWEN, N.L., 1958, Origin of granite in the light of experimental studies in the system NaAlSi3O8KAlSi3O8SiO2H2O: Geological Society of America, v. Memoir 74, 153 p. VELBEL, M.A., 1989, Effect of chemical affinity on feldspar hydrolysis rates in two natural weathering systems: Chemical Geology, v. 78, p. 245253. WALDRON, K.A., AND PARSONS, I., 1992, Feldspar microtextures and the multi-stage thermal history of syenites from the Coldwell Complex, Ontario: Contributions to Mineralogy and Petrology, v. 111, p. 222234. WALDRON, K.A., PARSONS, I., AND BROWN, W.L., 1993, Solutionredeposition and the orthoclasemicrocline transformation: evidence from granulites and relevance to 18O exchange: Mineralogical Magazine, v. 57, p. 687695.

WALDRON, K.A., LEE, M.R., AND PARSONS, I., 1994, The microstructures of perthitic alkali feldspars revealed by hydrofluoric acid etching: Contributions to Mineralogy and Petrology, v. 116, p. 360364. WALKER, F.D.L., LEE, M.R., AND PARSONS, I., 1995, Micropores and micropermeable texture in alkali feldspars: geochemical and geophysical implications: Mineralogical Magazine, v. 59, p. 507536. WALKER, T.E., 1984, Diagenetic albitization of potassium feldspar in arkosic sandstones: 1984, SEPM Presidential Address: Journal of Sedimentary Petrology, v. 54, p. 316. WILKINSON, M., AND HASZELDINE, R.S., 1996, Aluminium loss during sandstone diagenesis: Geological Society of London, Journal, v. 153, p. 657660. WILKINSON, M., DARBY, D., HASZELDINE, R.S., AND COUPLES, G.D., 1997, Secondary porosity generation during deep burial associated with overpressure leak-off. Fulmar Formation, United Kingdom Central Graben: American Association of Petroleum Geologists, Bulletin, v. 81, p. 803813. WORDEN, R.H., AND RUSHTON, J.C., 1992, Diagenetic K-feldspar textures: a TEM study and model for diagenetic feldspar growth: Journal of Sedimentary Petrology, v. 62, p. 779789. WORDEN, R.H., WALKER, F.D.L., PARSONS, I., AND BROWN, W.L., 1990, Development of microporosity, diffusion channels and deuteric coarsening in perthitic alkali feldspars: Contributions to Mineralogy and Petrology, v. 104, p. 507515. YUND, R.A., MCLAREN, A.C., AND HOBBS, B.E., 1974, Coarsening kinetics of the exsolution microstructure in alkali feldspar: Contributions to Mineralogy and Petrology, v. 48, p. 4555. YUND, R.A., AND ACKERMAND, D., 1979, Development of perthitic microstructures in the Storm King Granite, N.Y.: Contributions to Mineralogy and Petrology, v. 70, p. 273280. YUND, R.A., AND CHAPPLE, W.M., 1980, Thermal histories of two lava flows estimated from cryptoperthite lamellar spacings: American Mineralogist, v. 65, p. 438443. YUND, R.A., AND DAVIDSON, P., 1978, Kinetics of lamellar coarsening in cryptoperthites: American Mineralogist, v. 63, p. 470477. YUND, R.A., ACKERMAND, D., AND SEIFERT, F., 1980, Microstructures in the alkali feldspars from the granulite complex of Finnish Lapland: Neues Jahrbuch fu r Mineralogie-Monatshefte, v. 3, p. 109117. Received 28 April 2004; accepted 21 January 2005.

Das könnte Ihnen auch gefallen