Fulltext d

download Fulltext d

of 12

Transcript of Fulltext d

  • 7/30/2019 Fulltext d

    1/12

    O R I G I N A L P A P E R

    The mechanism of cation and oxygen isotope exchange in alkalifeldspars under hydrothermal conditions

    Dominik R. D. Niedermeier Andrew Putnis Thorsten Geisler Ute Golla-Schindler Christine V. Putnis

    Received: 18 January 2008 / Accepted: 4 June 2008 / Published online: 21 June 2008 Springer-Verlag 2008

    Abstract The mechanism of re-equilibration of albite in

    a hydrothermal fluid has been investigated experimentallyusing natural albite crystals in an aqueous KCl solution

    enriched in 18O at 600C and 2 kbars pressure. The reac-

    tion is pseudomorphic and produces a rim of K-feldspar

    with a sharp interface on a nanoscale which moves into the

    parent albite with increasing reaction time. Transmission

    electron microscopy (TEM) diffraction contrast and X-ray

    powder diffraction (XRD) show that the K-feldspar has a

    very high defect concentration and a disordered Al, Si

    distribution, compared to the parent albite. Raman spec-

    troscopy shows a frequency shift of the Si-O-Si bending

    vibration from*476 cm-1 in K-feldspar formed in normal16O aqueous solution to *457 cm-1 in the K-feldspar

    formed in 18O-enriched solution, reflecting a mass-related

    frequency shift due to a high enrichment of18O in the K-

    feldspar silicate framework. Raman mapping of the spatial

    distribution of the frequency shift, and hence 18O content,

    compared with major element distribution maps, show a

    1:1 correspondence between the reaction rim formed by the

    replacement of albite by K-feldspar, and the oxygen iso-

    tope re-equilibration. The textural and chemical

    characteristics as well as the kinetics of the replacement of

    albite by K-feldspar are consistent with an interface-cou-

    pled dissolution-reprecipitation mechanism.

    Keywords Feldspar replacement Ion-exchange mechanism Interface-coupled dissolution-reprecipitation

    Hydrothermal Fluid-mineral interaction Oxygen isotope exchange

    Introduction

    Understanding the redistribution of elements during

    hydrothermal mineral replacement reactions is fundamen-

    tal for many disciplines in geosciences, e.g. metamorphic

    petrology, economic geology and geochronology. Reac-

    tions induced by hydrothermal fluids are responsible for the

    chemical and mineralogical differentiation of large parts of

    the earths crust. In natural environments, fluid-rock

    interaction results in changes of the bulk rock composition,

    mineral assemblage and the isotopic composition. How-

    ever, the exchange mechanisms of these large scale

    geological processes are not yet fully understood. This

    deficit is largely a result of the complexity of natural fluids

    and the problem of reconstructing the physicochemical

    conditions during re-equilibration reactions in natural

    environments.

    Natural feldspar minerals often show complex re-

    equilibration or alteration textures such as lobate zones of

    different composition and high porosity (e.g. turbid feld-

    spars) that penetrate into primary feldspar (Putnis et al.

    2007). Numerous oxygen isotope analyses have further

    shown that many feldspars were partly re-equilibrated in

    the presence of a fluid phase (e.g. Cole et al. 2004;

    Elsenheimer and Valley 1993, Fiebig and Hoefs 2002). The

    Communicated by J. Hoefs.

    D. R. D. Niedermeier A. Putnis T. Geisler U. Golla-Schindler C. V. PutnisInstitut fur Mineralogie, Westfalische Wilhelms-Universitat,

    Corrensstrasse 24, 48149 Munster, Germany

    Present Address:

    D. R. D. Niedermeier (&)

    Institut fur Petrologie der Ozeankruste, Universitat Bremen,

    Klagenfurterstrasse 2, 28359 Bremen, Germany

    e-mail: [email protected]

    123

    Contrib Mineral Petrol (2009) 157:6576

    DOI 10.1007/s00410-008-0320-2

  • 7/30/2019 Fulltext d

    2/12

    high content of alkali chloride solutions observed in many

    fluid inclusions (Poty et al. 1974; Roedder 1972) has led to

    the suggestion that metamorphic and igneous rocks have

    crystallised in the presence of volatile-rich fluids, enriched

    in Na and K (Orville 1962). Especially in areas of mag-

    matic activity, where the transport of heat is partially

    controlled by hot fluids. The impact of supercritical,

    chloride-rich fluids is obvious. Hence, exchange reactionsbetween albite (NaAlSi3O8) or potassium feldspar (KAl-

    Si3O8) and K- or Na-rich chloride solutions, respectively,

    are important buffering reactions in feldspar-bearing rocks.

    The hydrothermal re-equilibration of feldspar in aqueous

    chloride solutions has thus been a subject of numerous

    experimental studies (e.g. Hovis 1997; Labotka et al. 2004;

    ONeill and Taylor 1967; Orville 1963; Schliestedt and

    Matthews 1987; Weise and Schliestedt 1988). Based on

    hydrothermal experiments made at 1 kbar and 350800C

    ONeill and Taylor (1967) already suspected that the

    exchange of cations and oxygen between alkali feldspars

    and alkali chloride solutions proceeds via a reaction frontthat moves through the parent feldspar crystal. Later,

    Labotka et al. (2004) found a strong spatial correlation

    between the exchange of the cations and oxygen atoms

    during their 18O tracer experiments at 2 kbar, 600C by

    using modern micro-analytical tools. This observation is

    evidence for a complete reorganization of the feldspar

    framework and can be explained by a process that is based

    on the spatial and temporal coupling of congruent albite

    dissolution followed by immediate K-feldspar precipita-

    tion, resulting in an enrichment of Na in the fluid. It has

    been suggested that many re-equilibration processes,

    whose mechanism had been described in terms of volume

    diffusion in the solid state, might be attributable to such a

    coupled dissolution-reprecipitation process (Putnis 2002;

    Putnis and Putnis 2007).

    On the other hand, dry exchange by homogenization of

    molten salt and solid feldspar[800C can be explained by

    solid state diffusion processes in which cations are

    exchanged between the molten chloride and the solid

    feldspar (e.g. Christophe-Michel-Levy 1967; Laves 1951).

    During cation exchange in feldspars, the Al and Si atoms

    do not necessarily need to be rearranged, because the cat-

    ions are incorporated in large sites within the framework.

    At high temperatures such diffusional ion exchange can

    proceed at significant rates so that it can be observed on

    laboratory time scales.

    In a series of experiments, low albite has been treated

    with an aqueous solution of KCl. The KCl solution was

    prepared using H2O enriched with 95%18O as a tracer.

    Confocal micro-Raman spectroscopy was used to investi-

    gate the oxygen isotopic variation within experimentally

    produced K-feldspar reaction rims. The heavier 18O iso-

    tope, substituting into the feldspar structure for16O, lowers

    the vibrational frequencies of the tetrahedral Si(Al)-O

    framework. This principle was used to (1) determine

    whether oxygen isotope exchange accompanies cation

    exchange and (2) to study potential variations of the oxy-

    gen isotope composition across the product K-feldspar

    (see, e.g. Geisler et al. 2005; Putnis et al. 2007). Scanning

    electron microscopy (SEM), electron probe micro-analysis

    (EPMA) and transmission electron microscopy (TEM)were used to analyse the structural and chemical interface

    between pristine Na-feldspar and converted K-feldspar on a

    micro- and nanometer scale.

    Experimental methods

    The starting material for the experiments was a natural

    sample of a low albite from Amelia County, Virginia

    (USA). This pegmatitic albite is nearly 100% Na feldspar

    and fully Al, Si ordered (Salisbury et al. 1987). Two

    different grain sizes of feldspar material were prepared forexperimental runs. First, single crystals of feldspar, with

    targeted grain size of 22.5 mm, were cleaved or broken

    from a large crystal. After cleaving, the crystals were

    cleaned with deionised water and dried in an oven at

    120C. Second, a powder was produced by crushing the

    material in an agate mortar under ethanol. The powder

    was washed in deionised water and wet sieved to remove

    grains finer than 100 lm and coarser than 200 lm. The

    procedure of grinding, washing, and sieving was repeated

    several times until enough material for several experi-

    ments had been produced. The preparation methods of the

    starting material correspond to those used by Labotka

    et al. (2004).

    Hydrothermal experiments were carried out at 600C

    and 2 kbar for run durations between 3 and 14 days. The

    albite was reacted with two molar aqueous KCl solution.

    For the hydrothermal experiments, the sample was filled

    into a gold capsule of 2.8 mm inner diameter and*25 mm

    length. The gold capsules were filled with about 15 mg of

    feldspar powder (grain size 100200 lm) and 30 mg of

    fluid, yielding a fluid to solid ratio of about two. For the

    preparation of TEM samples some experiments were con-

    ducted with single grains of*2 mm in diameter. The

    aqueous solutions for the experiments were prepared with

    dry KCl (grade 99.5%) and deionised water. In some

    experiments 18O enriched H2O (95%18O) was used for the

    preparation of the KCl solution. The capsule was sealed by

    carbon-arc welding and then loaded into a horizontally

    mounted cold-seal vessel. The pressure-vessel was heated

    externally from room temperature to 600 2C within

    *1.5 h while water was used as the external pressure

    medium to keep the pressure at 2 kbars during the

    experiment.

    66 Contrib Mineral Petrol (2009) 157:6576

    123

  • 7/30/2019 Fulltext d

    3/12

    After the experiments, the gold capsules were opened

    and the sample material was washed out with deionised

    water. The reacted powder and the reacted single crystals

    were washed several times with deionised water and dried

    at 120C in an oven. The sample material was then sec-

    tioned for analytical measurements. For electron

    microscopy and Raman measurements, the powder sample

    was embedded into araldite mounts and polished at thesurface. Petrographic thin sections of larger grains were

    prepared for optical microscopy and electron microprobe

    analyses. The specimens were carbon coated for analyses

    by SEM and EPMA. For TEM, thin sections were prepared

    with acetone-soluble adhesive and first were analysed by

    light microscopy and SEM to identify areas of interest for

    TEM specimen preparation. Copper discs with a slit of

    500 lm width and 2 mm in length were glued with epoxy

    resin on these areas, cut out with a Medenbach drill, and

    removed from the thin section using acetone. The specimen

    was ion-beam thinned by a Gatan DuoMill Ion Thinner

    until it was electron transparent.

    Analytical techniques

    Powder X-ray diffraction analyses of the starting material

    and the reacted products were performed with a Philips

    XPert PW 3040 automated diffractometer. The samples

    were finely ground and packed into standard sample

    holders of 2 mm depth to minimize the effect of preferred

    orientation. The operating conditions were 45 kV accele-

    ration voltage at 40 nA tube current. The samples were

    X-rayed using monochromatic Cu-Ka1 radiation

    (c = 1.540598 A). For a good comparison, all samples

    were analyzed under identical conditions with a step size of

    0.02 (2h) within a range between 5 and 70 (2h), where

    the most informative reflections for the feldspar structure

    occur. The 2-theta position was calibrated using silicon as

    an external standard. The counting time was set to 25 s per

    step.

    Samples of starting material and reacted phases were

    analyzed with a JEOL 840 SEM. Element distribution

    images were made with the integrated Oxford INCA EDX-

    system. A JEOL JSM 6300F field emission scanning

    electron microscope (FE-SEM) was used for observations

    of the sample surfaces. The images were captured using the

    analySIS (Soft Imaging System) software. For comparison

    of the sample, grains of the untreated and of the reacted

    feldspar were mounted together onto aluminium stubs with

    Leit-C tabs and subsequently carbon coated.

    Electron probe micro-analyses (EPMA) were performed

    using a JEOL JXA 8600 MX Superprobe at 15 kV

    acceleration voltage and 15 nA beam current. Altogether

    nine elements were measured, using natural mineral

    standards of diopside for Ca and Si, rutile for Ti, micro-

    cline for K, jadeite for a Na, rhodonite for Mn, fayalite for

    Fe, kyanite for Al and forsterite for Mg. Matrix effects

    were corrected using the ZAF correction procedure. A

    JEOL 8900 JXE Superprobe was used for mapping the

    distribution of Na, Al, Si, and K in some reaction products.

    Transmission electron microscope (TEM) investigations

    of starting feldspar and reacted sample were carried outwith a JEOL 3010 TEM that operated at 297 kV. The

    thinned samples were mounted in a double tilt specimen

    holder. The TEM is equipped with a Gatan post column

    energy filter, and an Oxford EDX-system, which enables

    compositional probing of small areas of the sample. The

    crystallographic orientation of the samples was determined

    from diffraction patterns. Bright field images and diffrac-

    tion images were taken on photo negative films.

    A confocal LabRam Raman system was used for a series

    of single Raman measurements and Raman mapping. A

    HeNe (632.187 nm) laser was used for excitation. The

    scattered light was collected in 180 backscatteringgeometry and dispersed by a grating of 1,800 grooves/mm

    after having passed a 100 lm entrance slit. A 1009

    objective with a numerical aperture of 0.9 and a confocal

    hole of 400 lm was used for all measurements, yielding an

    axial (depth) resolution of\3 lm while the lateral reso-

    lution was better than 1.5 lm. The acquisition time for a

    single measurement and for the mapping procedure was 60

    and 24 s, respectively. All measurements were made at

    room temperature.

    Results

    After experimental treatment, the cleaved feldspar grains

    did not change in size, but the surfaces changed to a milky-

    white colour and were fairly friable. The grains that were

    optically transparent before the reaction turned translucent

    and some grains coalesced during the reaction to form

    clusters of numerous single crystals.

    Secondary electron (SE) images of experimentally

    treated albite display high porosity at the grain surface

    (Fig. 1) whereas there was no porosity in the starting

    material. Some areas protrude from the mineral surfaces,

    while at the same time; some pits penetrate deep into the

    host crystal. The surface topography has increased on the

    reacted surface in comparison to the pristine albite surface,

    which only shows smooth cleavage planes (Fig. 1df).

    Substantial change can be observed at the cleavage sur-

    faces, where the change in surface topography is the most

    intensive. Obviously different crystallographic layers have

    been exposed during the reaction. At lower magnification

    these features appear as typical dissolution features of a

    mineral surface. At high magnification the surface of the

    Contrib Mineral Petrol (2009) 157:6576 67

    123

  • 7/30/2019 Fulltext d

    4/12

    product K-feldspar (Fig. 1c) is characterised by fine texture

    apparently made up from small clusters, quite unlike the

    original albite surface at the same magnification. These

    observations suggest new crystal growth is associated with

    the reaction. Feldspar grains that have been experimentally

    treated for longer periods show flat surfaces. In contrast to

    the rough surface in Fig. 1, there is only a minor surface

    roughness visible in Fig. 2. The crystal planes tend to form

    an idiomorphic feldspar crystal habit.

    In backscattered electron (BSE) images of sectioned

    grains a dark grey rim and a brighter core, representing the

    reacted rim and the pristine albite, respectively, can be

    distinguished (Fig. 3). This difference in image contrast

    indicates an abrupt change in the chemistry at the boundary

    between the two phases. This boundary or interface is sharp

    on the micrometer scale and is accompanied by numerous

    pores (Fig. 3). The thickness of the K-rich feldspar rim

    varies conspicuously, e.g. from 10 to 100 lm in a 7-day

    Fig. 1 ac SEM images of the

    surface of a reacted albite grain

    from a 5-day experiment in

    comparison with (df) the

    surface of an untreated grain.

    The images show the cleavage

    surface (001). Note that the

    starting material shows no

    evidence for porosity, whereas

    the reacted material shows a

    rough surface topology

    Fig. 2 SEM photomicrograph of treated feldspar after long-term

    experiment conducted for 13 days. Obvious is a difference between

    areas of high and low growth rates. In the upper right part of the

    picture, plane surface areas have evolved, whereas on the lower left

    side, coarser surfaces are prevailing. Oriented growth of surface areas

    reflects roughly the habit of an idiomorphic single crystal

    68 Contrib Mineral Petrol (2009) 157:6576

    123

  • 7/30/2019 Fulltext d

    5/12

    experiment. EDX elemental mapping demonstrates that theconverted rim primarily contains K but also small amounts

    of Na (Fig. 4). Relics of pristine albite are irregularly

    distributed within the grain (Fig. 4).

    Quantitative analyses of the distribution of Si, Ti, Fe,

    Ca, Mn, Mg, Na, K and Al were made by electron probe

    micro-analysis (EPMA). Due to the spread of the electron

    beam, larger sample areas have been probed and do not

    record the sharp chemical gradient at the interface. Due to

    the loss of Na under the electron beam, higher concentra-

    tions of Si and Al appear in the albite-rich areas when

    normalised to 100%. A line profile across the interface

    between the albite and the product K-feldspar was mea-sured by EPMA (Fig. 5a, b). For each analysis the molar

    percentage of albite, orthoclase and anorthite was calcu-

    lated (Table 1). At the interface the feldspar composition

    abruptly changes from pure albite to pure K-feldspar.

    Interestingly, the pure K-feldspar only occurs directly near

    the interface in the reacted material (Fig. 5b). Further

    behind the interface the product K-feldspar becomes more

    Na-rich and approximates to an average composition

    Or87Ab12An1.

    The minor element Mg concentrations indicate varia-

    tions similar to Ca with higher concentrations in albite-

    bearing areas. The distribution of Ti, Fe and Mn did not

    show any variation between the starting material and theproduct K-feldspar. Electron microprobe element distri-

    bution maps give a broad overview of the distribution of

    Na and K near the interface (Fig. 5c, d). These chemical

    maps confirm the chemical anomaly in the product K-

    feldspar near the interface and show that this anomaly

    extends parallel to the reaction front. The black area in the

    upper parts of Fig. 5c shows the purity of the albite starting

    material, where no K could be detected. The sharp chem-

    ical gradient at the boundary between the pristine albite

    and the K-feldspar is again obvious.

    Raman spectroscopy allows estimation of the 18O con-

    tent and its spatial distribution inside the K-feldsparproduct that formed in a 18O-enriched solution due to a

    mass-related frequency shift of those vibrations that

    involve the motion of oxygen atoms. The most prominent

    bands in the Raman spectrum of K-feldspar occur near 476

    and 513 cm-1 (Fig. 6) and have been assigned to O-T-O

    bending and stretching vibrations of the four-membered

    tetrahedral rings (Matson et al. 1986; McKeown 2005).

    These modes are characteristic for all alkali-feldspars and

    their frequency is almost independent of the NaK ratio

    (Mernagh 1991). However, the degree of crystallinity has

    an influence on the mode frequencies (Matson et al. 1986).

    Nevertheless, since we compare the mode frequencies of

    K-feldspars that formed under identical conditions, it is

    safe to assume that the degree of crystallinity in the K-

    feldspar products is the same. Therefore, the observed red-

    shift of the frequency of these modes in the K-feldspar

    formed in 18O-enriched solution relative to the frequency

    Fig. 3 SEM image of a cross section of the interface between pristine

    albite in the core and reacted K-feldspar in the rim. Note the

    occurrence of numerous pores at the interface

    Fig. 4 a BSE image of a

    reacted albite grain from a 5-day

    experiment. Note the sharp

    interface between the dark Na-

    rich areas in the core and the

    light K-rich areas in the rim.

    The area shown in this imagewas chemically mapped by

    EDX. b Na (white area) and (c)

    K (white area) distribution of a

    reacted albite grain from a 5-day

    experiment

    Contrib Mineral Petrol (2009) 157:6576 69

    123

  • 7/30/2019 Fulltext d

    6/12

    observed for 16O-rich K-feldspar (Fig. 6) can directly be

    linked to the 18O content in the feldspar lattice. It is evident

    from Fig. 6 that the Raman bands of the K-feldspar formed

    in the 18O-enriched fluid shift to significantly lower wavenumbers relative to bands of K-feldspar formed in nor-

    mal 16O-rich solution. The observed shift Dx, of the

    bending mode near 476 cm-1, for instance, is about

    -18 cm-1, reflecting a strong enrichment of18O. The band

    near 478 cm-1 was chosen for Raman mapping. Interest-

    ingly, the Raman map of the frequency shift of this mode in

    the K-feldspar formed in the 18O-enriched solution shows

    even a higher frequency shift close to the interface (Fig. 7),

    which would indicate a higher 18O concentration. An

    important observation is that there is no gradient in the

    frequency shift that could be related to inward diffusion of18O.

    Transmission electron microscopy (TEM) in the bright

    field imaging mode revealed a very different diffraction

    contrast between the pristine albite and the converted K-

    feldspar (Fig. 8a). The parent albite has smooth diffraction

    contours, indicating a relatively defect-free crystal struc-

    ture, whereas the product K-feldspar shows complex

    diffraction contrast and numerous linear features, sug-

    gesting a high defect density. EDX spectra from both sides

    of the interface were taken to clearly identify the two

    feldspar phases. The areas of product K-feldspar display

    numerous structural features that cannot be observed in the

    albite. Observations of the product K-feldspar show no

    macroscopic porosity, but dense dark textures near theboundary to the pristine albite. Some of these structures

    have an elongated appearance, tending into direction of the

    fluid/solid interface, and might represent voids or channels.

    The interface between the two phases is characterised by a

    dark zone that may be caused by high density of disloca-

    tions (Fig. 8a). This dark area generally appears at the

    boundary of the two feldspar phases. Electron diffraction

    patterns taken from both sides of the interface reveal that

    both feldspar phases are structurally similar (Fig. 8b, d)

    and the crystallographic orientation of the K-feldspar rim

    coincides within a degree with that of the albite core. Only

    a slight displacement can be observed in the diffraction

    patterns of the interface area, with splitting between two

    equivalent diffraction spots (Fig. 8c), from each feldspar

    phase. Some parts of the reaction interface show dense

    dislocation arrays suggesting a semicoherent interface,

    while in others there appears to be an open porosity.

    X-ray powder diffraction patterns of the product K-

    feldspar agree very well with powder diffraction patterns of

    a natural sanidine (Weitz 1972). The structural state of the

    untreated and the product K-feldspar was estimated using

    aK

    lowhigh count rates

    Nad

    0

    10

    20

    30

    40

    50

    60

    70

    80

    90

    100

    0 5 10 15 20 25 30 35

    Distance from edge (m)

    Mol%

    b

    Albite

    Orthoclase

    cFig. 5 a BSE image of the area

    within a reacted feldspar grain

    that was chosen for electron

    microprobe traverse along the

    white line (detailed results in

    Table 1) and for electron

    microprobe mapping marked by

    the white rectangle. b Profiles of

    the albite, orthoclase and the

    anorthite component across the

    interface of the product K-

    feldspar and unreacted albite.

    Note the decrease of the albite

    and anorthite component before

    the reaction interface between

    10 and 14 lm. Electron

    microprobe distribution maps of

    c K and d Na near the reaction

    front. Note that the interface

    between the pristine albite and

    the product K-feldspar is sharp

    on the micrometer scale. The

    exceptionally low Na-values

    within the red area are caused

    by the line analyses (b), where

    Na and K were lost under the

    influence of the electron beam

    70 Contrib Mineral Petrol (2009) 157:6576

    123

  • 7/30/2019 Fulltext d

    7/12

    Table1

    Electronmicroprobedata

    ofaline-scanacrosstheinterfacebetween

    pristinealbiteandtheK-feldsparreaction

    product(seeFig.

    5)

    Electronmicroprobedata

    1

    2

    3

    4

    5

    6

    7

    8

    9

    10

    11

    12

    13

    14

    15

    16

    17

    31

    SiO2

    64.6

    8

    64.8

    4

    64

    .31

    64.5

    3

    64.2

    7

    64.0

    2

    64

    .13

    64.5

    0

    64.2

    6

    64.6

    7

    64.7

    7

    64.7

    5

    64.4

    6

    64.2

    8

    68.4

    7

    69.1

    6

    69.1

    9

    TiO2

    0.0

    0

    0.0

    0

    0

    .04

    0.0

    1

    0.0

    0

    0.0

    0

    0

    .00

    0.0

    4

    0.0

    0

    0.0

    0

    0.0

    4

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    1

    Al2O3

    18.3

    6

    18.5

    6

    18

    .27

    18.1

    3

    18.1

    9

    18.3

    6

    18

    .05

    18.3

    4

    18.3

    8

    18.4

    0

    18.0

    2

    18.0

    9

    18.2

    7

    18.4

    4

    19.1

    0

    19.1

    1

    19.4

    3

    FeO

    0.0

    1

    0.0

    3

    0

    .00

    0.0

    1

    0.0

    0

    0.0

    0

    0

    .01

    0.0

    1

    0.0

    0

    0.0

    0

    0.0

    3

    0.0

    3

    0.0

    0

    0.0

    0

    0.0

    0

    0.1

    1

    0.0

    1

    MnO

    0.0

    0

    0.0

    2

    0

    .02

    0.0

    0

    0.0

    0

    0.0

    4

    0

    .00

    0.0

    1

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    1

    MgO

    0.0

    0

    0.0

    0

    0

    .00

    0.0

    0

    0.0

    0

    0.0

    0

    0

    .00

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    0

    0.0

    2

    0.0

    0

    0.0

    0

    0.0

    1

    0.0

    1

    0.0

    1

    CaO

    0.1

    3

    0.1

    9

    0

    .17

    0.1

    4

    0.1

    3

    0.1

    4

    0

    .12

    0.0

    7

    0.0

    8

    0.1

    2

    0.0

    2

    0.0

    0

    0.0

    4

    0.1

    6

    0.2

    4

    0.2

    0

    0.2

    5

    Na2

    O

    1.4

    0

    1.3

    9

    1

    .36

    1.3

    3

    1.2

    6

    1.3

    2

    1

    .12

    1.1

    8

    1.1

    2

    0.7

    9

    0.3

    7

    0.2

    2

    0.1

    4

    0.1

    9

    7.8

    2

    9.9

    4

    10.2

    2

    K2

    O

    14.5

    5

    14.7

    6

    14

    .46

    14.5

    7

    14.6

    9

    14.7

    9

    14

    .79

    14.9

    8

    15

    .00

    15

    .67

    16.2

    9

    16.3

    2

    16.2

    5

    16.0

    3

    2.3

    8

    0.1

    6

    0.2

    0

    Sum

    99.1

    3

    99.8

    0

    98

    .64

    98.7

    2

    98.5

    2

    98.6

    6

    98

    .22

    99.1

    3

    98.8

    5

    99.6

    4

    99.5

    4

    99.4

    2

    99.1

    5

    99.1

    1

    98.0

    3

    98.6

    8

    99.3

    1

    Normalizedtoeightoxygenatoms

    performulaunitfeldspar

    Mg

    0.0

    00

    0.0

    00

    0

    .000

    0.0

    00

    0.0

    00

    0.0

    00

    0

    .000

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    01

    0.0

    00

    0.0

    00

    0.0

    01

    0.0

    01

    0.0

    00

    Ca

    0.0

    06

    0.0

    09

    0

    .008

    0.0

    07

    0.0

    06

    0.0

    07

    0

    .006

    0.0

    04

    0.0

    04

    0.0

    06

    0.0

    01

    0.0

    00

    0.0

    02

    0.0

    08

    0.0

    11

    0.0

    09

    0.0

    11

    Mn

    0.0

    00

    0.0

    01

    0

    .001

    0.0

    00

    0.0

    00

    0.0

    02

    0

    .000

    0.0

    01

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    00

    Fe

    0.0

    01

    0.0

    01

    0

    .000

    0.0

    00

    0.0

    00

    0.0

    00

    0

    .000

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    01

    0.0

    01

    0.0

    00

    0.0

    00

    0.0

    00

    0.0

    04

    0.0

    00

    Na

    0.1

    26

    0.1

    24

    0

    .123

    0.1

    20

    0.1

    14

    0.1

    19

    0

    .102

    0.1

    06

    0.1

    01

    0.0

    71

    0.0

    34

    0.0

    20

    0.0

    13

    0.0

    17

    0.6

    67

    0.8

    41

    0.8

    61

    K

    0.8

    60

    0.8

    69

    0

    .860

    0.8

    66

    0.8

    74

    0.8

    81

    0

    .884

    0.8

    88

    0.8

    91

    0.9

    26

    0.9

    66

    0.9

    68

    0.9

    64

    0.9

    51

    0.1

    34

    0.0

    09

    0.0

    11

    Sum

    A

    0.9

    92

    1.0

    04

    0

    .992

    0.9

    93

    0.9

    95

    1.0

    09

    0

    .992

    0.9

    98

    0.9

    96

    1.0

    03

    1.0

    02

    0.9

    90

    0.9

    79

    0.9

    77

    0.8

    13

    0.8

    63

    0.8

    84

    Si

    2.9

    97

    2.9

    91

    2

    .996

    3.0

    05

    3.0

    00

    2.9

    90

    3

    .004

    2.9

    95

    2.9

    92

    2.9

    96

    3.0

    12

    3.0

    09

    2.9

    99

    2.9

    89

    3.0

    11

    3.0

    17

    3.0

    05

    Al

    1.0

    03

    1.0

    09

    1

    .003

    0.9

    95

    1.0

    00

    1.0

    10

    0

    .996

    1.0

    03

    1.0

    08

    1.0

    04

    0.9

    87

    0.9

    91

    1.0

    01

    1.0

    11

    0.9

    89

    0.9

    83

    0.9

    94

    Sum

    B

    4.0

    00

    4.0

    00

    4

    .000

    4.0

    00

    4.0

    00

    4.0

    00

    4

    .000

    4.0

    00

    4.0

    00

    4.0

    00

    4.0

    00

    4.0

    00

    4.0

    00

    4.0

    00

    4.0

    00

    4.0

    00

    4.0

    00

    Mole%

    feldsparcomponent

    An

    0.6

    0.9

    0

    .8

    0.7

    0.6

    0.7

    0

    .6

    0.4

    0.4

    0.6

    0.1

    0.0

    0.2

    0.8

    1.5

    1.1

    1.4

    Ab

    12.0

    11.7

    11

    .8

    11.5

    10.9

    11.2

    9

    .7

    10.1

    9.6

    6.7

    3.2

    1.9

    1.2

    1.7

    81.3

    97.8

    97.3

    Or

    87.3

    87.3

    87

    .4

    87.8

    88.5

    88.1

    89

    .7

    89.5

    90.0

    92.7

    96.7

    98.1

    98.6

    97.5

    17.3

    1.1

    1.3

    Theaverageanalyticaldatafrompristinealbiteanalyses17

    31aregiveninth

    elastcolumn.

    Theelectronmicroprobeana

    lyseswereconvertedtomolepercentofthefeldsparcomponentsby

    normalizingtoeightoxygenatoms

    performulaunit

    Contrib Mineral Petrol (2009) 157:6576 71

    123

  • 7/30/2019 Fulltext d

    8/12

    the refined lattice parameters and selected diffraction peak

    positions (Kroll and Ribbe 1987). The lattice parameters

    have been obtained by Rietveld refinement using the

    fullprof computer program (Rodriguez-Carvajal 1990).

    The state of Al, Si ordering in alkali feldspars is defined by

    the Al content at the t1o site in the crystal structure. Fol-

    lowing equation was used to determine the occupancy

    value Rt1 for the albite starting material as well as for the

    sanidine product feldspar (Kroll and Ribbe 1987):

    Xt1 t1o t1m

    b 21:5398 53:8405 c

    2:1567 15:8583 c1

    with the lattice parameters for the albite starting material

    given in Table 2, Eq. 1 yields Rt1 = 0.996, i.e. the albite is

    fully ordered. These calculated values are similar to those

    obtained by Smith et al. (1986) for the Amelia low albite

    0.997 Al in T1O, 1.001 Si in T1 m, 1.002 Si in T2O and1.006 Si in T2 m. For the product K-feldspar, we obtain

    Rt1 = 0.569 from the lattice parameters given in Table 2.

    The reaction product is thus highly disordered sanidine

    feldspar.

    Discussion

    The results of this study suggest that a continuous

    exchange by a coupled dissolutionreprecipitation process

    leads to the replacement of albite by K-feldspar. The SEM

    observations of the reacted surface and the electron dif-fraction patterns across the reaction interface demonstrate

    that, although the K-feldspar has recrystallised during the

    reaction, it essentially remains a single crystal with the

    same crystallographic orientation as the parent phase. Such

    an observation is consistent with an interface-coupled

    dissolution-reprecipitation mechanism in which the nucle-

    ation and growth of the product is epitaxial on the parent

    surface (Putnis and Putnis 2007). The interface between

    pristine albite and product K-feldspar is sharp on a nano-

    meter scale, as seen by TEM (Fig. 8a). The TEM bright-

    field image across the reaction interface shown in Fig. 8a

    further reveals a difference in diffraction contrast betweenthe albite parent, showing virtually no extended defects,

    and the K-feldspar product that is characterised by dense

    linear features normal to the reaction interface. These

    features may represent nano-channels that formed along

    dislocations and allowed the fluid reservoir to keep contact

    with the fluid at the reaction front. The linear features must

    originate from the dissolution-reprecipitation reaction,

    because no comparable crystallographic defects could be

    observed in the albite starting material. When albite is

    replaced by K-feldspar, a molar volume increase of 8%,

    caused by the incorporation of the larger K-cation, has to

    be compensated by the crystal lattice at the interface. It is

    the coherent stress that leads to the formation of connected

    dislocations, which provide pathways for chemical

    exchange between the surrounding fluid and the fluid at the

    reaction interface. Lee and Parsons (1997) observed fea-

    tures similar to these nano-channels in natural orthoclase

    that was partially converted to albite. They observed albite

    crystals replacing orthoclase that were delineated by strain

    shadows and proposed a relationship to closely spaced

    dislocations. Also Worden et al. (1990) reported jagged,

    Fig. 6 Representative Raman spectra of two experimentally

    exchanged K-feldspars, one containing common 16O (grey) and one18

    O (black) in the tetrahedral framework. Large-amplitude modes at

    476 and 519 cm-1 shift to lower values near 457 and 497 cm-1,

    respectively

    -13-18 -14-15-16-17

    100 m

    Fig. 7 Raman contour map ofDx from the area marked in the BSE

    image. Dx is the difference between the frequency of the bending

    mode near 476 cm-1 in the K-feldspar formed in an 18O-enriched

    fluid and the frequency of this mode in 16O-rich K-feldspar

    72 Contrib Mineral Petrol (2009) 157:6576

    123

  • 7/30/2019 Fulltext d

    9/12

    incoherent interfaces between albite-twinned albite and

    albite-twinned microcline in naturally altered feldspars.

    After experimental treatment larger open pores are com-

    mon at the interface between the pristine albite and the

    K-feldspar (Fig. 3). These pores must have formed during

    the hydrothermal experiments as they only occur near the

    reaction interface (Figs. 3, 8a) and it is therefore highly

    likely that they are an integral feature of the replacement

    mechanism. It is evident that the reaction rate should

    depend on available fluid pathways through the product

    1m

    Na

    Si

    Al

    O

    O

    Si

    AlK

    Na

    a

    b c d

    Fig. 8 a Bright-field TEM

    image of the interface between

    K-feldspar on the left and albite

    on the right (zone axis [112]).

    The interface has a high contrast

    and appears dark. The product

    K-feldspar phase on the left side

    is characterized by a much

    higher structural complexity.

    Diffraction patterns of (b) the

    albite, (c) the interface area and

    (d) the product K-feldspar (zone

    axis [112])

    Contrib Mineral Petrol (2009) 157:6576 73

    123

  • 7/30/2019 Fulltext d

    10/12

    crystal, i.e. on interconnected porosity. In contrast to the

    porous material seen in Fig. 1, where numerous pores

    appear at the feldspar surface, the surface roughness of the

    grain shown in Fig. 2 is obviously reduced.

    The reaction between albite and a hydrothermal fluid is

    also influenced by the crystallographic properties of the

    host albite. Powder X-ray analyses reveal a change fromtriclinic (C-1) albite structure to monoclinic (C2/m) sym-

    metry in K-feldspar which depends on the Al, Si order. The

    X-ray powder diffraction of the albite starting material

    indicate a very high degree of Al, Si order while the K-

    feldspar product phase is highly disordered. At tempera-

    tures of 600C K-feldspar is Al, Si disordered (Smith and

    Brown 1988) and the difference in Al, Si order between the

    parent and product provides extra thermodynamic driving

    force, in addition to the chemical exchange. Kinetically,

    the reaction between albite and a KCl solution is strongly

    dependent on the crystallographic direction (Holdren and

    Speyer 1985; Weise and Schliestedt 1988), as evidenced by

    the irregular rim thickness (Fig. 3). The reaction rate is

    greater along the (001) crystal surface that shows more

    conspicuous structural changes than other crystal surfaces.

    This observation can be rationalised by the fact that the

    bonding parallel to the (001) lattice plane is weaker than

    along other crystal planes, which is reflected by the pre-

    ferential cleavage of albite along the (001) plane. In

    ordered feldspars the proportion of Al-tetrahedra at the

    (001) surface is higher than at other crystal surfaces

    (Bondham 1967), and the Al-tetrahedra are preferred sites

    of dissolution, because the AlO bonds are energetically

    easier to break than the SiO bonds (Oelkers and Schott

    1995; Xiao and Lasaga 1994).

    Raman spectroscopy gives evidence for the incorpora-

    tion of 18O into the feldspar lattice and thus provides

    further evidence that the K-feldpar precipitated from

    solution. Particularly, the most K-rich area near the reac-

    tion interface shows the highest Raman shift of-18 cm-1

    compared to K-feldspar formed in 16O-rich isotopically

    normalsolution. The area approximate to the grain edge,

    a few micrometer away from the interface, contains more

    Na and has a smaller Raman-shift of-14 cm-1 due to the

    lower 18O content (Fig. 7). The position of the interface

    can be observed as a sharp boundary between albite start-

    ing material and K-feldspar product in the Raman map as

    well as in the electron microprobe map. Observations by

    electron microprobe and Raman spectroscopy show a one-

    to-one correlation between cation and oxygen isotope

    exchange, which is consistent with data of Schliestedt andMatthews (1987) and Labotka et al. (2004).

    The oxygen isotope exchange is not influenced by

    fractionation between sodium and potassium feldspar.

    Apparently both feldspars, Albite and K-feldspar, have

    identical isotopic properties relative to one another

    (Schwarcz 1966; Taylor and Epstein 1962). ONeill and

    Taylor (1967) demonstrated that the isotopic composition

    just tends to approximate equilibrium conditions between

    the feldspar and the fluid. The high concentration of K and18O found close to the interface seems to be similar to 18O-

    enriched alteration zones with sharp chemical gradients to

    relict unreacted areas in polycrystalline, Ta-based pyro-chlore after hydrothermal treatment (Geisler et al. 2005).

    Our results are also consistent with data on natural

    hydrothermally altered plagioclase in volcanic rocks

    which show a direct correspondence between cation and

    oxygen isotope exchange. Cole et al. (2004) have studied

    plagioclase feldspars from Rico Dome, Colorado with

    naturally altered rims of mainly albitic composition. They

    have observed narrow rims enriched in 18O in distal

    feldspars and wide reaction rims or completely exchanged

    feldspars that are depleted in 18O, proximal to the Rico

    hydrothermal system. Cole et al. (2004) have noted that

    simple volume diffusion could not have governed the

    transport during hydrothermal exchange. They compared

    the observed penetration values with hydrothermal based

    volume diffusion coefficients for oxygen in Na- or K-

    feldspars (Farver and Yund 1990). The durations to pro-

    duce d18O profiles, approximating those measured for

    Rico feldspars, are one to two orders of magnitude greater

    than could be achieved by volume diffusion within the

    age of the system.

    Volume diffusion is definitely too slow as an effective

    transport mechanism of K and Na ions and O (Brady and

    Yund 1983; Christoffersen et al. 1983; Cole and Chakra-

    borty 2001; Foland 1974) to explain the high exchange

    rates that could be observed by Raman spectroscopy and

    electron microprobe. The maximum diffusion length for K

    in albite at 600C experiments is less than 0.300 lm within

    67 days under anhydrous conditions (Brady and Yund

    1983; Christoffersen et al. 1983; Giletti and Shanahan

    1997), although under hydrothermal conditions the diffu-

    sivity might be significantly higher (Cole and Chakraborty

    2001, Giletti and Shanahan 1997). However, on small scale

    volume diffusion may operate to equilibrate relict domains

    Table 2 Refined lattice parameters for the albite starting material

    and for the sanidine product phase

    Cell parameter Starting albite Product K-feldspar

    a 8.1430 8.5616

    b 12.7920 13.0204

    c 7.1610 7.1743

    a 94.2425 90.0000b 116.5904 115.9713

    c 87.7035 90.0000

    c* 0.1565 0.1550

    74 Contrib Mineral Petrol (2009) 157:6576

    123

  • 7/30/2019 Fulltext d

    11/12

    of albite feldspar (Fig. 3), or to slowly homogenize areas

    within the reprecipitated K-feldspar.

    The observations of a chemically and structurally sharp

    reaction interface, the correlation between cation and

    oxygen exchange, the generation of defects or nanopores

    associated with the reaction, the recrystallised texture of

    the product K-feldspar and the rate of the reaction are all

    not consistent with volume diffusion as a dominantmechanism, but are entirely consistent with an interface-

    coupled dissolution-reprecipitation mechanism (Putnis

    2002; Putnis and Mezger 2004; Geisler et al. 2005; Putnis

    et al. 2007; Putnis and Putnis 2007).

    The details of the reaction progress by such a mechanism

    is likely to be complex as the product solid continuously

    re-equilibrates with a changing fluid composition.

    The approach towards equilibrium conditions between

    the product feldspar and the composition of the fluid also

    undoubtedly affects the reaction rate. As the replacement

    of albite by K-feldspar proceeds, the composition of the

    fluid changes to higher NaCl contents while the productfeldspar composition continuously re-equilibrates by

    interacting with the fluid. Feldspar-fluid equilibria in the

    system NaAlSi3O8KAlSi3O8H2ONaClKCl (Orville

    1967; Lagache and Weisbrod 1977; Pascal and Roux 1982;

    Rubie and Gunter 1983) define the coexisting compositions

    (Fig. 9). Although the aim of our experiments was to

    determine the mechanism of the replacement, rather than to

    establish equilibrium, Fig. 9 may be used to show the

    possible evolution of the rim compositions in our experi-

    ment. Starting from the initial situation where the fluid

    composition mole% K/(K+ Na) = 100, which would

    coexist in equilibrium with pure K-feldspar, the fluid and

    rim composition would successively re-equilibrate, until

    the experiment was terminated, in our case with a rimcomposition of Or87Ab12An1. At each step of such an

    evolution, re-equilibration is achieved by successive dis-

    solution and reprecipitation reactions. A further caveat

    however, is that such a description assumes that the rim

    composition is in equilibrium with the bulk fluid compo-

    sition, whereas in situ experiments on replacement in salt

    systems have shown that the solid composition which

    reprecipitates is determined by the fluid composition at the

    reaction interface, and that large compositional gradients

    may exist in the fluid while the reaction proceeds (Putnis

    et al. 2005). This suggests that the rate limiting step in such

    replacements may be the mass transfer through the fluidpathways in the solid reaction product.

    Although we have no satisfactory explanation for the

    enrichment of K and 18O in the K-feldspar at the reaction

    interface (Figs. 5, 7), it is a consistent feature of these

    experiments and, in terms of18O, also of other experiments

    (Geisler et al. 2005) and suggests that the fluid at the

    interface must also be enriched in K and 18O during the

    reaction. The lower 18O enrichment away from the reaction

    front may indicate that here the chemical composition of

    the product feldspar re-equilibrated with an increasingly16O richer fluid by secondary processes.

    Acknowledgments We are indebted to Peter Schmid-Beurmann

    who helped with the Rietveld calculations, Arne Janen and Angelika

    Breit for the powder X-ray diffraction measurements and Jasper

    Berndt for the electron microprobe mapping. We are also grateful to

    Herbert Kroll for his help in calculating the state of Al, Si order. We

    thank an anonymous reviewer for a possible explanation for the K and18O enrichment at the reaction interface.

    References

    Bondham J (1967) Structural changes in adularia in hydrolytic

    environments. Medd Dansk Geol For 17:357370Brady JB, Yund RA (1983) Interdiffusion of K and Na in alkali

    feldspars; homogenization experiments. Am Mineral 68:106

    111

    Christoffersen R, Yund RA, Tullis J (1983) Inter-diffusion of K and

    Na in alkali feldspars; diffusion couple experiments. Am Mineral

    68:11261133

    Christophe-Michel-Levy M (1967) Sur le mecanisme de

    lexchange NaK par voie hydrothermale dans lalbite. Bull

    Soc Fr Min Crist 90:411413

    Cole DR, Chakraborty S (2001) Rates and mechanisms of isotopic

    exchange. Rev Mineral Geochem 43:83223. doi:10.2138/

    gsrmg.43.1.83

    600

    0

    20

    40

    60

    80

    100

    0

    Crystal

    Mol % K / (Na + K)

    2M

    Fluid

    Mol%K

    /(Na+K)

    Pressure: 2 kbars

    10080604020

    Fig. 9 Fluidcrystal distribution curve at 600C and 2 kbars pressure

    for a closed system (after Orville 1962). The dashed horizontal line

    represents alkali feldspar immiscibility at 600C. The solid line

    represents equilibrium conditions for solid feldspar (composition on

    the abscissa) coexisting with a fluid (composition on the ordinate).

    The dashed line at 45 represents identical Na/K ratios for fluid and

    crystal. The square symbol lying on the curve marks the observed

    feldspar composition in this study after experiment termination

    Contrib Mineral Petrol (2009) 157:6576 75

    123

    http://dx.doi.org/10.2138/gsrmg.43.1.83http://dx.doi.org/10.2138/gsrmg.43.1.83http://dx.doi.org/10.2138/gsrmg.43.1.83http://dx.doi.org/10.2138/gsrmg.43.1.83
  • 7/30/2019 Fulltext d

    12/12

    Cole DR, Larson PB, Riciputi LR, Mora CI (2004) Oxygen isotope

    zoning profiles in hydrothermally altered feldspars: Estimating

    the duration of water-rock interaction. Geology 32:2932. doi:

    10.1130/G19881.1

    Elsenheimer D, Valley JW (1993) Submillimeter scale zonation of

    delta-O-18 in quartz and feldspar, Isle-of-Skye, Scotland.

    Geochim Cosmochim Acta 57:36693676. doi:10.1016/

    0016-7037(93)90148-P

    Farver JR, Yund RA (1990) The effect of hydrogen, oxygen, and

    water fugacity on oxygen diffusion in alkali feldspar. Geochim

    Cosmochim Acta 54:29532964. doi:10.1016/0016-7037(90)

    90113-Y

    Fiebig J, Hoefs J (2002) Hydrothermal alteration of biotite and

    plagioclase as inferred from intergranular oxygen isotope- and

    cation-distribution patterns. Eur J Mineral 14:4960. doi:

    10.1127/0935-1221/2002/0014-0049

    Foland KA (1974) Alkali diffusion in orthoclase. In Hofmann AW,

    Giletti BJ, Yoder HS, Yund RA (eds) Geochemical Transport

    and Kinetics. Carnegie Inst Washington Publ 634:7798

    Geisler T, Poml P, Stephan T, Janssen A, Putnis A (2005)

    Experimental observation of an interface-controlled pseudomor-

    phic replacement reaction in a natural crystalline pyrochlore. Am

    Mineral 90:16831687. doi:10.2138/am.2005.1970

    Giletti BJ, Shanahan TM (1997) Alkali diffusion in plagioclase

    feldspar. Chem Geol 139:320. doi:10.1016/S0009-2541(97)

    00026-0

    Holdren DR, Speyer PM (1985) pH-dependent changes in the rates

    and stoichiometry of dissolution of an alkali feldspar at room

    temperature. Am J Sci 285:9941026

    Hovis GL (1997) Volumes of KNa mixing for low albite-microcline

    crystalline solutions at elevated temperaturea test of regular

    solution thermodynamic models. Am Mineral 82:158164

    Kroll H, Ribbe PH (1987) Determining (Al, Si) distribution and strain

    in alkali feldspars using lattice parameters and diffraction-peak

    positionsa review. Am Mineral 72:491506

    Labotka TC, Cole DR, Fayek M, Riciputi LR, Stadermann FJ (2004)

    Coupled cation and oxygen-isotope exchange between alkali

    feldspar and aqueous chloride solution. Am Mineral 89:1822

    1825

    Lagache M, Weisbrod A (1977) The system two alkali feldspars-

    KCL-NaCl-H2O at moderate to high temperatures and low

    pressures. Contrib Mineral Petrol 62:77101. doi:10.1007/

    BF00371028

    Laves F (1951) Artificial preparation of microcline. J Geol 59:511

    512

    Lee MR, Parsons I (1997) Dislocation formation and albitization in

    alkali feldspars from the sharp granite. Am Mineral 82:557570

    Matson DW, Sharma SK, Philpotts JA (1986) Raman spectra of some

    tectosilicates and of glasses along the orthoclase-anorthite and

    nepheline-anorthite joins. Am Mineral 71:694704

    McKeown DA (2005) Raman spectroscopy and vibrational analyses

    of albite -From 25C through the melting temperature. Am

    Mineral 90:15061517. doi:10.2138/am.2005.1726

    Mernagh TP (1991) Use of the laser Raman microprobe fordiscrimination amongst feldspar minerals. J Raman Spectrosc

    22:453457. doi:10.1002/jrs.1250220806

    Oelkers EH, Schott J (1995) Experimental study of anorthite

    dissolution and the relative mechanism of feldspar hydrolysis.

    Geochim Cosmochim Acta 59:50395053. doi:10.1016/0016-

    7037(95)00326-6

    ONeill JR, Taylor HP (1967) The oxygen isotope and cation

    exchange chemistry of feldspars. Am Mineral 52:14141437

    Orville PM (1962) Alkali metasomatism and feldspars. Nor Geol

    Tidsskr 42:283316

    Orville PM (1963) Alkali ion exchange between vapor and feldspar

    phases. Am J Sci 261:201237

    Pascal M-L, Roux J (1982) Thermodynamic properties of (Na, K) Cl-

    H2O solutions between 400C and 800C 12 kbar. Geochim

    Cosmochim Acta 46:331337. doi:10.1016/0016-7037(82)

    90224-1

    Poty B, Stalder HA, Weisbrod A (1974) Fluid inclusions studies in

    quartz from fissures of Western and Central Alps. Schweiz min

    petr Mitt 54:717752

    Putnis A (2002) Mineral replacement reactions; from macroscopic

    observations to microscopic mechanisms. Mineral Mag 66:689

    708. doi:10.1180/0026461026650056

    Putnis A, Putnis CV (2007) The mechanism of reequilibration of

    solids in the presence of a fluid phase. J Solid State Chem

    180:17831786. doi:10.1016/j.jssc.2007.03.023

    Putnis CV, Geisler T, Schmid-Beurmann P, Stephan T, Giampaolo C

    (2007) An experimental study of the replacement of leucite by

    analcime. Am Mineral 92:1926. doi:10.2138/am.2007.2249

    Putnis CV, Tsukamoto K, Nishimura Y (2005) Direct observations of

    pseudomorphism: compositional and textural evolution at a

    fluid-solid interface. Am Mineral 90:19091912. doi:10.2138/

    am.2005.1990

    Putnis CV, Mezger K (2004) A mechanism of mineral replacement:

    Isotope tracing in the model system KCl-KBr-H2O. Geochim

    Cosmochim Acta 68:28392848. doi:10.1016/j.gca.2003.12.009

    Rodriguez-Carvajal J (1990) FULLPROF: a program for rietveld

    refinement and pattern matching analysis, abstracts of the

    satellite meeting on powder diffraction of the XV congress of

    the IUCr, p 127, Toulouse, France

    Roedder E (1972) The composition of fluid inclusions. USGS

    Professional Paper 440 Data of geochemistry, 6th edition,

    Chapter JJ, 164p

    Salisbury JW, Walter LW, Vergo N (1987) Mid-Infrared (2.125

    lm) spectra of minerals. 1st edition USGS open file report 87-

    263

    Schliestedt M, Matthews A (1987) Cation and oxygen isotope

    exchange between plagioclase and aqueous chloride solution.

    Neues Jahrb Min 6:241248

    Schwarcz HP (1966) Oxygen isotope fractionation between host and

    exsolved phases in perthite. Geol Soc Am Bull 778:879882.

    doi:10.1130/0016-7606(1966)77[879:OIFBHA]2.0.CO;2

    Smith JV, Artioli G, Kvick A (1986) Low albite NaAlSi3O8 Neutron

    diffraction study of crystal structure at 13 K. Am Mineral

    71:727733

    Smith JV, Brown WL (1988) Feldspar minerals 1: crystal structures,

    physical, chemical and microtextural properties, 828p. Springer,

    Berlin

    Taylor HP, Epstein S (1962) Relation between O18/O16 ratios in

    coexisting minerals of igneous and metamorphic rocks II

    application to petrologic problems. Geol Soc Am Bull 73:675

    694. doi:10.1130/0016-7606(1962)73[675:RBORIC]2.0.CO;2

    Weise P, Schliestedt M (1988) Experimentelle Untersuchungen der

    NaK Austauschreaktionen zwischen Alkalifeldspaeten und

    chloridischen Loesungen. Fortschr Mineral, Beih 66:164Weitz G (1972) Die Struktur des Sanidins bei verschiedenen

    Ordnungsgraden. Z Kristallogr 136:418426

    Worden RH, Walker FDL, Parsons I, Brown WL (1990) Development

    of microporosity, diffusion channels and deuteric coarsening in

    perthitic alkali feldspars. Contrib Mineral Petrol 104:507515.

    doi:10.1007/BF00306660

    Xiao Y, Lasaga AC (1994) Ab initio quantum mechanical studies of

    the kinetics and mechanisms of silicate dissolution H+(H3O+)

    catalysis. Geochim Cosmochim Acta 58:53795400. doi:

    10.1016/0016-7037(94)90237-2

    76 Contrib Mineral Petrol (2009) 157:6576

    123

    http://dx.doi.org/10.1130/G19881.1http://dx.doi.org/10.1016/0016-7037(93)90148-Phttp://dx.doi.org/10.1016/0016-7037(93)90148-Phttp://dx.doi.org/10.1016/0016-7037(90)90113-Yhttp://dx.doi.org/10.1016/0016-7037(90)90113-Yhttp://dx.doi.org/10.1127/0935-1221/2002/0014-0049http://dx.doi.org/10.2138/am.2005.1970http://dx.doi.org/10.1016/S0009-2541(97)00026-0http://dx.doi.org/10.1016/S0009-2541(97)00026-0http://dx.doi.org/10.1007/BF00371028http://dx.doi.org/10.1007/BF00371028http://dx.doi.org/10.2138/am.2005.1726http://dx.doi.org/10.1002/jrs.1250220806http://dx.doi.org/10.1016/0016-7037(95)00326-6http://dx.doi.org/10.1016/0016-7037(95)00326-6http://dx.doi.org/10.1016/0016-7037(82)90224-1http://dx.doi.org/10.1016/0016-7037(82)90224-1http://dx.doi.org/10.1180/0026461026650056http://dx.doi.org/10.1016/j.jssc.2007.03.023http://dx.doi.org/10.2138/am.2007.2249http://dx.doi.org/10.2138/am.2005.1990http://dx.doi.org/10.2138/am.2005.1990http://dx.doi.org/10.1016/j.gca.2003.12.009http://dx.doi.org/10.1130/0016-7606(1966)77[879:OIFBHA]2.0.CO;2http://dx.doi.org/10.1130/0016-7606(1962)73[675:RBORIC]2.0.CO;2http://dx.doi.org/10.1007/BF00306660http://dx.doi.org/10.1016/0016-7037(94)90237-2http://dx.doi.org/10.1016/0016-7037(94)90237-2http://dx.doi.org/10.1007/BF00306660http://dx.doi.org/10.1130/0016-7606(1962)73[675:RBORIC]2.0.CO;2http://dx.doi.org/10.1130/0016-7606(1966)77[879:OIFBHA]2.0.CO;2http://dx.doi.org/10.1016/j.gca.2003.12.009http://dx.doi.org/10.2138/am.2005.1990http://dx.doi.org/10.2138/am.2005.1990http://dx.doi.org/10.2138/am.2007.2249http://dx.doi.org/10.1016/j.jssc.2007.03.023http://dx.doi.org/10.1180/0026461026650056http://dx.doi.org/10.1016/0016-7037(82)90224-1http://dx.doi.org/10.1016/0016-7037(82)90224-1http://dx.doi.org/10.1016/0016-7037(95)00326-6http://dx.doi.org/10.1016/0016-7037(95)00326-6http://dx.doi.org/10.1002/jrs.1250220806http://dx.doi.org/10.2138/am.2005.1726http://dx.doi.org/10.1007/BF00371028http://dx.doi.org/10.1007/BF00371028http://dx.doi.org/10.1016/S0009-2541(97)00026-0http://dx.doi.org/10.1016/S0009-2541(97)00026-0http://dx.doi.org/10.2138/am.2005.1970http://dx.doi.org/10.1127/0935-1221/2002/0014-0049http://dx.doi.org/10.1016/0016-7037(90)90113-Yhttp://dx.doi.org/10.1016/0016-7037(90)90113-Yhttp://dx.doi.org/10.1016/0016-7037(93)90148-Phttp://dx.doi.org/10.1016/0016-7037(93)90148-Phttp://dx.doi.org/10.1130/G19881.1