Sei sulla pagina 1di 28

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

PAGES 147^174

2011

doi:10.1093/petrology/egq076

Silicate Liquid Immiscibility within the Crystal Mush: Evidence fromTi in Plagioclase from the Skaergaard Intrusion
MADELEINE C. S. HUMPHREYS*
DEPARTMENT OF EARTH SCIENCES, UNIVERSITY OF CAMBRIDGE, DOWNING STREET, CAMBRIDGE CB2 3EQ, UK

RECEIVED JANUARY 8, 2010; ACCEPTED NOVEMBER 2, 2010

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

A key target in the study of layered intrusions is to constrain the liquid line of descent of the magma. However, the evolution of the interstitial liquid is rarely considered, and its liquid line of descent is often assumed to be equivalent to that of the bulk magma. Because of extensive sub-solidus and diffusional changes that occur in slowly cooled rocks, clues to the composition of the interstitial liquid can only be obtained using very slowly diffusing trace elements and components. This study uses the Ti concentrations and anorthite contents of interstitial plagioclase to consider the compositional evolution of the interstitial liquid in the Skaergaard Intrusion. Ti^XAn zoning of interstitial plagioclase does not follow the same cryptic variations that develop in plagioclase primocrysts as a function of stratigraphic height, demonstrating that the bulk and interstitial liquid lines of descent are not equivalent. After Fe^Ti oxides start to crystallize, Ti concentrations decrease in both primocryst and interstitial plagioclase as a result of decreasing melt Ti. However, in the interstitial plagioclase within a single thin section, divergent trends develop adjacent to fine-grained interstitial pockets containing diverse mineral assemblages, which are interpreted to represent the crystallized products of late-stage immiscible liquids. These trends vary systematically as a function of stratigraphic height and spatial location within the intrusion. The distribution and compositions of these plagioclase zoning trends are used to comment on the spatial distribution and differential movement of interstitial immiscible liquids within the intrusion.

I N T RO D U C T I O N
Layered intrusions are of interest for many reasons: for example, they reveal information about differentiation and fractionation processes that may be operating beneath currently active volcanoes, they provide insights into the effects of reactive fluids moving within a porous medium and they are commonly host to economic precious metal deposits. The cumulate rocks that define layered intrusions can accumulate by various mechanisms, including the settling of primocrysts (early formed crystals) onto the crystal pile, in situ growth of crystals in the cooling boundary layers, or deposition from crystal-laden plumes (e.g. Wager & Brown, 1968; McBirney & Noyes, 1979; Irvine et al., 1998; Marsh, 2006). In most instances, the result is a layer of cumulus crystals surrounded by interstitial liquid. Once this crystal mush has formed, the interstitial liquid is slowly eliminated through various processes that may include overgrowth on primocrysts, nucleation and growth of oikocrysts and other interstitial material (including new minerals), consolidation and compaction of the mush owing to the weight of overlying crystals and compositional convection, which can result in adcumulus textures. Although post-depositional processes acting within the mush have been discussed in many previous studies (e.g. Tait et al., 1984; Barnes, 1986; Boudreau & McCallum, 1992; Haskin & Salpas, 1992; Meurer & Boudreau, 1998; T egner et al., 2009), relatively little work has addressed the differentiation of the interstitial liquid itself, compared with that of the bulk magma. Studies specifically considering the development of intercumulus

KEY WORDS: layered igneous rock; immiscibility; plagioclase; zoning; cumulate; Skaergaard

*Corresponding author. Present Address: Department of Earth Sciences, University of Oxford, South Parks Road, Oxford, OX1 3AN. T elephone: 44 (0)1865 272020. Fax: 44 (0)1865 272072. E-mail: madeleine.humphreys@earth.ox.ac.uk

The Author 2010. Published by Oxford University Press. All rights reserved. For Permissions, please e-mail: journals.permissions@ oup.com

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

overgrowths and late mineral phases are also relatively sparse (e.g. Meurer & Claeson, 2002; Meurer & Meurer, 2006; T oplis et al., 2008; Humphreys, 2009). Questions include whether the interstitial liquid follows the same liquid line of descent as the bulk magma, what happens to its oxidation state, and whether it undergoes liquid immiscibility. The macro-scale effects of silicate liquid immiscibility have been discussed in detail in previous studies (e.g. Philpotts, 1976, 1981; Roedder, 1978; Ripley et al., 1998). In particular, there is evidence that anorthosites may crystallize from a relatively silicic immiscible liquid that is rich in normative feldspar (Philpotts, 1981; Loferski & Arculus, 1993). However, the likelihood and possible effects of liquid immiscibility occurring within the crystal mush have been neglected. These are important questions, because if any of the interstitial liquid is returned to the bulk magma, perhaps through compaction or compositional convection, then there is the potential to affect the evolving composition of the bulk magma, or even its mode of fractionation, and to create differences in bulk-rock composition between floor and roof cumulates. Reconstructing the liquid line of descent of the magma that formed a layered intrusion is difficult, as demonstrated by the still-continuing disagreement over the liquid line of descent for Skaergaard (e.g. Wager, 1960; Hunter & Sparks, 1987; Brooks & Nielsen, 1990; McBirney & Naslund, 1990; Thy et al., 2006, 2009; Morse, 2008b). Typically, the liquid line of descent is estimated either through mass balance and extraction of bulk cumulate compositions from an estimated starting or final composition (e.g. Wager, 1960; Nielsen, 2004), or by experimental petrology and modelling (McBirney & Nakamura, 1974; T oplis & Carroll, 1995, 1996; Thy et al., 2006, 2009). This is not a viable approach for the interstitial liquid, not least because it may be difficult to demarcate the primocryst and its interstitial overgrowths, and because much of the chemical variation may have been obliterated by diffusion. Experimental work would also be very difficult in the relatively low-temperature, multi-component systems expected in late-stage liquids. Instead, the study of very slowly diffusing minor or trace components can be used to infer the evolving composition of the liquid, taking into account variations in partition coefficients with temperature and composition. If the bulk and interstitial liquids follow similar differentiation paths, then the cryptic chemical variations that occur in primocrysts at different stratigraphic levels should also be observed in the interstitial overgrowths. This study will attempt to address this problem, using the Skaergaard Intrusion, east Greenland, as a case study. The Skaergaard Intrusion has been very well documented over many years, and in addition to changes in primocryst assemblage, there is very clear cryptic compositional variation of the primocryst grains with stratigraphic height.

This study investigates Ti zoning in primocryst and interstitial plagioclase from the Layered Series and Marginal Border Series of the Skaergaard Intrusion, building on preliminary results from the Lower Zone of the Layered Series (Humphreys, 2009). Ti is an ideal marker for the composition of the liquid, because its diffusion rate in plagioclase is extremely slow, as is CaAl^NaSi interdiffusion (Grove et al., 1984). The work attempts to constrain compositional differences between bulk and interstitial liquid evolution and assess the importance of interstitial liquid immiscibility.

T H E S K A E R G A A R D I N T RU S I O N
The Skaergaard Intrusion is located on the eastern coast of Greenland, and formed at $55 Ma (Brooks & Gleadow, 1977; Hirschmann et al., 1997) during the opening of the North Atlantic (Nielsen, 1975). In plan view, the intrusion is c. 8 km 11km in size (Fig. 1) and has a volume of c. 280 km3 (Nielsen, 2004). Over 3000 m of stratigraphy is now exposed as a result of gentle tilting to the SE, which occurred during Early T ertiary subsidence of the continental margin (Nielsen, 1975). The magma was intruded at the contact between Precambrian metamorphic basement and accumulating Eocene flood basalts (Wager & Brown, 1968). The intrusion comprises three main series of layered igneous rocks: the Layered Series (LS), which accumulated on the floor of the intrusion, the Upper Border Series (UBS), which accumulated under the roof, and the Marginal Border Series (MBS), which formed on the steeply dipping walls of the intrusion (Wager & Brown, 1968). The LS and UBS meet at the Sandwich Horizon. The LS is divided into three zones on the basis of the initial presence (Lower Zone, LZ), absence (Middle Zone, MZ) and subsequent reappearance (Upper Zone, UZ) of cumulus olivine (Wager & Brown, 1968). The LZ and UZ are then subdivided on the basis of changes in cumulate mineralogy to give LZa (cumulus Fo-rich olivine plagioclase), LZb ( augite), LZc (oxides) and UZa (cumulus Fa-rich olivine plagioclase clinopyroxene oxides), UZb ( apatite) and UZc ( ferro-bustamite). In addition, the LS includes a Hidden Zone (HZ), which, as the name suggests, is not exposed at the surface, but comprises the most primitive cumulates (olivine plagioclase) and was sampled by the 1966 Cambridge Drill Core. The MBS was initially divided according to the presence of compositional banding, into the Tranquil Division and the Banded Division (Wager & Brown, 1968). However, this banding and subdivision does not clearly relate to the stratigraphic development of the cumulates; therefore Hoovers (1989) use of stratigraphically equivalent zones denoted HZ*, LZa*, etc. is preferred and will be used here. The UBS is subdivided into three main zones, UBS-a, UBS-b and UBS-g, which are

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

148

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

Fig. 1. Geological map of the Skaergaard Intrusion, after McBirney (1989). The locations of the 1966 Cambridge drill core and the Platinova drill cores 90-22 and 90-10 are shown. Continuous lines (KR02 and SP) indicate the locations of traverses through the Marginal Border Series. The star indicates the location of LZc sample 458279 (from T egner et al., 2009).

equivalent to the LZ, MZ and UZ of the LS, and UBS-t, which refers to the outermost part of the UBS and is essentially equivalent to the HZ (Naslund, 1984; Douglas, 1961). In addition to the changes in primocryst assemblage, the compositions of the primocrysts change with stratigraphic height. For example, plagioclase compositions vary from $An70 in HZ to An25 at the Sandwich Horizon (e.g. Wager & Brown, 1968; Maale, 1976). This cryptic variation is generally accepted to reflect closed-system, fractional crystallization of essentially a single batch of magma. Whereas the LS has been heavily studied since early descriptions (Wager & Brown, 1968), the MBS and UBS have been relatively neglected apart from the studies of Naslund (1984) and Hoover (1989). The comprehensive early petrological study of the MBS (Hoover, 1989) showed that primocryst olivine compositions are up to 9 mol % more Fa-rich, clinopyroxene is 5 mol % richer in the diopside component, and plagioclase is 5^8 mol % more calcic, than primocrysts from equivalent stratigraphic levels in the LS (Hoover, 1989). Minor element contents

of plagioclase were not analysed. These observations were ascribed to re-equilibration of primocrysts with Fe-rich interstitial liquid flowing down the margins of the intrusion (Hoover, 1989). In addition, mineral compositions from the upper parts of the MBS (i.e. adjacent to the Upper Border Series) were also observed to be more Fe-rich, and plagioclase more Na-rich, than elsewhere in the intrusion. This was interpreted as the result of either more effective fractionation near the roof or modified magma composition in the upper parts of the chamber (Hoover, 1989). For Skaergaard, there is still disagreement over changes in oxygen fugacity (e.g. T oplis & Carroll, 1995, 1996; Thy et al., 2009), and whether the residual liquid follows the Fenner or Bowen compositional trends; that is, whether it becomes Fe-rich at low SiO2 or tends towards SiO2-rich compositions at relatively lower FeO contents (e.g. McBirney, 1975; Hunter & Sparks, 1987; Brooks & Nielsen, 1990; McBirney & Naslund, 1990; T oplis & Carroll, 1995, 1996; Thy et al., 2006, 2009). It seems plausible that both

149

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

the Bowen and Fenner trends are represented in the compositions of two immiscible liquids that have been shown to coexist by at least UZb (McBirney & Nakamura, 1974; Hanghj et al., 1995; Jakobsen et al., 2005). The postulated earlier occurrence of immiscibility in the residual liquid (LZc, Veksler et al., 2007) is disputed (e.g. Morse, 2008a; McBirney, 2008; Philpotts, 2008; Veklser et al., 2008). Furthermore, the effect, if any, of liquid immiscibility on the resulting fractionation trend is unknown, as is the stage at which immiscibility might occur within the interstitial liquid.

Textural indicators of evolved interstitial liquid


Recent work has demonstrated that evolved, late-stage interstitial liquid can be present in glass-bearing nodules entrained in lava flows as thick films of glass on grain boundaries. In cumulate rocks, pyroxene commonly forms thick grain boundary films and it has been suggested that these represent an original liquid film (e.g. Holness, 2005; Holness et al., 2007a). At Skaergaard, crystallized traces of evolved liquids can also be seen in trails of tiny oxide and apatite crystals, clustered particularly along the margins of plagioclase chadacrysts within clinopyroxene oikocrysts (Fig. 2a). Holness et al. (2011) demonstrate that, following immiscibility in the bulk magma, the solidification of large pockets of interstitial liquid results in the formation of paired (spatially associated) intergrowths of granophyre and an ilmenite-rich assemblage including apatite, skeletal Fe^Ti oxides and clinopyroxene (see Holness et al., 2011, figs 10^12). Granophyre-rich pockets tend to be associated with the plagioclase-rich regions of a sample whereas ilmenite-rich pockets tend to be associated with regions rich in olivine and clinopyroxene (Holness et al., 2011). The paired intergrowths commonly occupy planar-sided pockets, which shows that the liquid was not reactive, and the variations in their spatial distribution are clearest in the MBS. In the Skaergaard Peninsula, which represents the most complete chemical sequence, they become abundant near the MZ*^UZa* boundary. At the same point, evidence of localized reaction between oxides and pyroxene disappears (Holness et al., 2011). In combination the paired intergrowths represent up to 44% of the sample by area in the Layered Series, and more in the evolved parts of the MBS (Stripp, 2009). Samples from the less evolved, outermost parts of the Skaergaard Peninsula traverse contain small, isolated, fine-grained, planar-sided pockets of late-stage interstitial phases (Fig. 2b and c). These are distinct from the paired intergrowths of Holness et al. (2011) in that they are smaller, commonly very fine-grained and highly altered; however, these fine-grained pockets probably also represent the crystallized products of late-stage liquids. Where the contents of the pockets can be identified, they typically include some of quartz, K-feldspar, apatite, non-skeletal Fe^Ti

oxides, pyroxenes, biotite and zircon, as well as plagioclase growing on the margins of the pockets. Their mineralogy is highly variable (assuming that the 2D distribution is representative of the 3D volume), and the pockets represent between 027 and 106% of the samples by area in the Skaergaard Peninsula (Table 1). At lower structural levels in the MBS, near the base of the intrusion (e.g. northern Kraemer Island), the interstitial pockets are more sparsely developed, representing at most 020 area % of the samples (Table 1). Abundant interstitial low-Ca pyroxene is present, commonly surrounding resorbed olivine or adjacent to Fe^Ti oxides (Fig. 2d), and olivine may be partially replaced by oxy-symplectites of magnetite and orthopyroxene (see Holness et al., 2011). In the LS, the interstitial pockets are typically present at 020^035 area % (Table 1), but in the main replacive grain-boundary microstructures are seen instead in the central parts of the LS (LZc^MZ, Holness et al., 2011; see below and Table 1). At the base of the LS low-Ca pyroxene is also present, and the interstitial pockets are less abundant. The coarse-grained, paired granophyric and ilmenite-rich intergrowths become abundant around the MZ^UZa boundary near the margins of the LS and near the UZa^UZb boundary in the centre of the intrusion (Holness et al., 2011). Between LZc and the point at which the coarse-grained, paired intergrowths dominate, there is abundant evidence of reactive melt occupying grain boundaries, manifest as irregular (stepped) clinopyroxene^plagioclase grain boundaries, fish-hook intergrowths of pyroxene and plagioclase, olivine rims on oxide grains, and mafic symplectites; grains of apatite, amphibole and oxides are commonly found within the microstructures. This spatial distribution, as well as the microstructures involved, is described in detail by Holness et al. (2011), who interpret the microstructures as the result of chemical reaction following physical separation of Fe-rich and Si-rich immiscible liquids within the floor cumulates.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

SA M PLES A N D M ET HODS Samples studied


This study investigates a suite of samples from the LS and MBS of the Skaergaard Intrusion (Table 1). The LS was sampled mainly using drill core material, with samples selected to cover the range of stratigraphy. The lower parts of the LS (HZ^LZb) are covered by the 1966 Cambridge Drill Core (Maale, 1976; Holness et al., 2007b), and the upper parts (UZa^UZb) by drill core 90-22 (collected by Platinova Resources Ltd, 1990; T egner, 1997), with one LZc sample (458279) from a surface reference profile (Nielsen et al., 2000; T egner et al., 2009). Drill core 90-22 samples the centre of the intrusion (Fig. 1). Four samples from MZ^UZa were also examined from drill core 90-10,

150

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

(a)
ap ap pl ap ap ap

(b)
px
ox

pl

0.25 mm

100 m

ox

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

K-fsp pl

pl

(c)
pl ol

(d)

qz pl
opx pl opx

pl
200 m 0.25 mm

ol ol

pl

Fig. 2. T extural characteristics of interstitial material from the Skaergaard cumulates. (a) Cross-polarized light photomicrograph showing strings of small, grain-boundary oxides (arrowed) and apatite crystals (labelled) along the margins of plagioclase within clinopyroxene oikocrysts, marking the location of late interstitial liquids (from HZ*, north Kraemer Island). (b) Back-scattered SEM image of fine-grained interstitial pocket containing mafic minerals including apatite, Fe^Ti oxides and pyroxenes, from sample SK46 (LZa*^LZb* boundary) in the Skaergaard Peninsula. (c) Back-scattered SEM image of fine-grained interstitial pocket containing felsic minerals including quartz and alkali feldspar, as well as unidentifiable, brightly coloured weathered material, from sample SK46 (LZa*^LZb* boundary) in the Skaergaard Peninsula. (d) Cross-polarized photomicrograph showing orthopyroxene rims around olivine crystals, characteristic of the lower parts of the MBS (Kraemer Island) and the Layered Series.

which is located at the western margin of the intrusion (Fig. 1). Plagioclase compositions in the MBS were analysed in selected samples from two surface traverses, one on Kraemer Island (sample numbers beginning KR02), which includes zones HZ*^LZb*, and one on the Skaergaard Peninsula (sample numbers beginning SP), which samples the base of LZa* to UZb* (the contact is unexposed, so it is unclear how much of HZ*^LZa* is missing, but the fine-grained sample SP60 is assumed to

represent the base of LZa*). The figures presented here also include plagioclase data for MZ (drill core 90-22) collected by Stripp (2009) as well as data from the Cambridge drill core from Humphreys (2009).

Analytical methods
Plagioclase analyses were obtained using a Cameca SX-100 electron microprobe at the University of Cambridge. A 2 mm, 15 kV beam was used, with 10 nA current for major elements and 100 nA for minor elements. Typical peak

151

T able 1: Locations and details of samples studied


Series mineralogy sympl? pockets? Pockets sympl? replacive (N) Zone Primocryst Evolved Area % Replacive NonLocation Latitude Longitude (W)

Sample

Strat.

height (m)

90-22-421 Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series Layered Series HZ pl ol Y 011 LZa pl ol Y 020 oxy LZb pl ol cpx Y 030 (T1) oxy LZb pl ol cpx (Y) 0061 LZb pl ol cpx Y 033 MZ pl cpx ox Y 013 T1 MZ pl cpx ox T1 MZ pl cpx ox Y 035 (T1) UZa pl ol cpx ox Y 028 (Y?) LZc pl ol cpx ox Y 0032 T1 UZa pl ol cpx ox (Y) 0006 T1 UZa pl ol cpx ox T1 Y Drill core 90-22 Drill core 90-22 Kraemer Island Drill core 90-10 Drill core 90-10 Drill core 90-10 Drill core 90-10 Cambridge drill core Cambridge drill core Cambridge drill core Cambridge drill core Cambridge drill core UZb pl ol cpx ox ap Y 026 T1 Drill core 90-22

1671

Layered Series

UZb

pl ol cpx ox ap

Drill core 90-22

90-22-461.8

1630

90-22-481.8

1610

90-22-824.2

1268

458279

752

68811265"

31842563"

90-10-212

90-10-445

90-10-456

JOURNAL OF PETROLOGY

90-10-524.1

118590

197

118601

1681

118605

1636

VOLUME 52

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

152
Series mineralogy pockets? pockets Zone Primocryst Evolved Area % MBS MBS MBS MBS MBS MBS MBS MBS MBS HZ* LZa* pl ol pl ol LZb* pl ol cpx Base LZa* pl ol LZa*-LZb* pl ol cpx LZb* pl ol cpx LZc* pl ol cpx ox MZ* pl cpx ox Y Y Y Y Y Y Y (Y) UZb* pl ol cpx ox ap 027 036 067 106 047 014 020 0015 oxy oxy T1 T1 (T1)

118653

43

118678

22

Sample

Dist. from

Replacive sympl?

Nonreplacive sympl?

Location

Latitude (N)

Longitude (W)

margin

NUMBER 1

(m)

SP29

594

Skaergaard Peninsula Skaergaard Peninsula Skaergaard Peninsula Skaergaard Peninsula Skaergaard Peninsula Skaergaard Peninsula Kraemer Island Kraemer Island Kraemer Island

6888495" 6888495" 6888497" 6888502" 6888506" 6888489" 68811362" 68811378" 68811382"

3184532" 31845138" 31845177" 31845286" 31845444" 31845542" 31844224" 31844340" 31844401"

SP20

4718

SP16

4273

SP8

3042

SP46

133

JANUARY 2011

SP60

104

KR02-21

2227

KR02-37

805

KR02-4

102

Stratigraphic heights of 90-22 and location of 458279 taken from the reference profile of Tegner et al. (2009). Y, yes this microstructure is present (parentheses indicate rarely); T1, Type 1 replacive symplectites; oxy, oxy-symplectites (Holness et al., 2011).

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

counting times were 20 s for major elements and 30^40 s for minor elements. Zoned grain margins, triple junctions, the margins of planar-sided interstitial pockets (see above) and traverses of optically strongly zoned grain edges were preferentially studied, to identify the compositional variations of interstitial plagioclase from the late stages of fractionation. These features were identified optically or by SEM observation using a JEOL JSM-820 instrument, with a 20 kV beam and c. 15 mm working distance. Data from Stripp (2009) from MZ are included for completeness; these represent traverses across zoned grains but probably do not sample the most extreme compositions.

Interstitial plagioclase: Marginal Border Series, Skaergaard Peninsula


The MBS as exposed on the Skaergaard Peninsula represents a nearly complete transect through the stratigraphy, comprising the base of LZa* to UZb*. As expected, interstitial plagioclase from any given stratigraphic zone typically evolves to substantially more sodic compositions than the primocrysts from that zone: to $An40 in HZ*, through $An35 in LZc*, to $An15 in UZb* (see Electronic Appendix 1, available for downloading at http://www.petrology.oxfordjournals.org). The composition of the most evolved plagioclase analysed from any sample is variable. Minor element concentrations show complex but systematic variations. In general, many interstitial plagioclase compositions plot at lower TiO2 contents for a given XAn than the primocrysts, and this will be described in detail below. However, one of the most interesting features of the interstitial compositions is that in the lower stratigraphic zones, the interstitial plagioclase trend itself splits into two divergent compositional trends, typically adjacent to the planar-sided interstitial pockets described above. Strong normal zoning can be observed adjacent to pockets containing abundant granophyre or quartz, commonly with a clear concave-upward trend of TiO2^XAn. In contrast, strong reverse zoning with decreasing TiO2 is observed in the vicinity of clearly calcic, Fe-rich pockets (typically containing Fe^Ti oxides, apatite, clinopyroxene biotite; Fig. 4). Single spot analyses also fall on these trends. There is a wide variation of plagioclase compositions even around a single pocket, and few plagioclase compositions represent the extreme ends of these normal and reverse trends. Many of the interstitial pockets are not clearly dominated by either Si-rich or Fe-rich minerals, or are strongly altered. As far as can be observed, however, reverse or calcic compositions are not associated with pockets that contain quartz or granophyre, whereas conversely not all pockets that contain Fe-rich minerals (apatite, oxides, pyroxenes, etc.) show reverse zoning. In sample SP60 (base LZa*), the TiO2 and anorthite contents of the interstitial plagioclase are initially indistinguishable from primocryst compositions from higher stratigraphic zones. With decreasing XAn, TiO2 concentrations rise to a maximum of c. 012 wt % at $An56, and then diverge steeply from the primocryst trend, with interstitial plagioclase dipping to lower TiO2 contents whereas the primocrysts show a much shallower decrease (Fig. 5a). The interstitial plagioclase trend then itself follows two divergent compositional trends after An54 (Fig. 5a) as described above. Interstitial pockets represent $047% of the thin section by area in this sample (Table 1). In sample SP46 (LZa*^LZb* boundary), interstitial pockets account for $106% by area (Table 1). Plagioclase compositions follow the XAn^TiO2 primocryst trend

R E S U LT S Primocryst plagioclase
The cryptic compositional variation of plagioclase primocrysts that occurs through the Layered Series stratigraphy at Skaergaard is summarized here, for later comparison with the interstitial compositions. Anorthite contents of primocrysts decrease steadily upwards through the stratigraphy, with a relatively restricted range of compositions observed in each zone, from $An63^69 in HZ, through $An48^53 at LZc, to $An31^39 at UZb (Fig. 3a). This is in agreement with the results of several previous studies (e.g. Maale, 1976; McBirney, 1989) and is consistent with closed-system fractionation of basaltic magma. TiO2 contents of plagioclase primocrysts increase to a maximum of $013 wt % in the LZb sample, then decrease almost linearly with increasing stratigraphic height (decreasing XAn: the primocryst trend, Fig. 3a), as a result of saturation in Fe^Ti oxides (Humphreys, 2009). Primocryst FeO contents decrease with decreasing XAn until approximately UZa levels, then start to increase again (Fig. 3b). This increase in FeO through the Upper Zone is consistent with previous observations (T egner, 1997) and has been interpreted as reflecting the increasing FeO content of the fractionating bulk magma (T egner, 1997; T egner & Cawthorn, 2010). MgO contents are scattered but approximately constant until $An45 and then decrease with XAn (Fig. 3c). In general, primocrysts from Layered Series rocks near the margins of the intrusion (i.e. drill core 90-10) contain slightly more FeO and MgO than those from the centre of the intrusion (drill core 90-22). In the Marginal Border Series, primocryst compositions are very similar to those in the LS in terms of XAn, but tend to be enriched in MgO and FeO, similar to plagioclase from drill core 90-10. This is consistent with the observations of Hoover (1989), who interpreted the Fe-rich compositions as the result of flow of Fe-rich residual liquids down the marginal walls of the intrusion.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

153

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

0.16 0.14 0.12

(a)

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 20 1.0 30 UZb 40 UZa MZ 50 LZc LZb 60 LZa 70 HZ 80
Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

(b)

0.8

wt% FeO

0.6

0.4

0.2

0.0 20 0.10 30 40 50 60 70 80

(c)
0.08

wt% MgO

0.06

0.04

0.02

0.00 20

30

40

50

60

70

80

mol % Anorthite
Fig. 3. Mole per cent Anorthite vs (a) TiO2, (b) FeO and (c) MgO compositions of primocryst plagioclase from the Layered Series. Data from Holness et al. (2011) and this study. Approximate zone boundaries are indicated.

154

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

(a)
pl ox

0.16 0.14 0.12

(b)

X An

70 60 50 40 30 20 10 0 160

px + ap

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 0 20 40 60 80 100 120 140

Distance from pocket margin (m)

100 m

0.16 0.14 0.12

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

(c)

wt% TiO2

0.10 0.08 0.06 0.04

(d)
pl K-fs qz
wt% TiO2

0.02 0.00 10 0.16 0.14 0.12 0.10

towards Si-rich pockets


20 30 40 50

towards Fe-rich pockets

60

70

80

90 70

X An

(e)

X An

60 50 40

0.08 0.06 0.04 0.02 30 20 10 0 350

400 m

0.00 0 50 100 150 200 250 300

Distance from pocket margin (m)


Fig. 4. Continuous zoning profiles towards Fe-rich (a) and Si-rich (d) interstitial pockets from sample SP46 (LZa*^LZb* boundary, Skaergaard Peninsula). Zoning at the margins of Fe-rich pockets shows decreasing TiO2 (filled symbols, b) but increasing (reverse) XAn (open symbols, b) adjacent to the margin (5 50 mm). In contrast, zoning at the margins of Si-rich pockets shows a decrease in TiO2 (filled symbols, e) and decreasing (normal) XAn (open symbols, e). This results in two diverging Ti^XAn trends (c; shows four separate profiles; normal zoning profiles, circles; reverse zoning profiles, diamonds).

until $An50, when the interstitial compositions diverge to lower TiO2, as for SP60. The two interstitial trends observed in SP60 are particularly well defined for sample SP46. Adjacent to pockets containing abundant

granophyre or quartz, there is a very clear, curving trend towards very low TiO2 at An28, whereas plagioclase adjacent to pockets containing mafic phases defines a sharply reversed trend towards $An75 (Fig. 5b). There is also one

155

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

0.16 0.14 0.12

0.16

HZ*

(a)

0.14 0.12 0.10 0.08 0.06 0.04

LZa*

(b)

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 10 0.16 0.14

SP60 KR02_4
20 30 40 50 60 70 80 90

0.02 0.00 10 0.16

SP46 KR02_37
20 30 40 50 60 70 80 90

LZb*

(c)

0.14 0.12 0.10 0.08 0.06 0.04

LZc*

(d)
Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

wt% TiO2

0.12 0.10 0.08 0.06 0.04 0.02 0.00 10 0.16 0.14 20 30 40 50 60

SP8 KR02_21
70 80 90

0.02 0.00 10 0.16

SP16
20 30 40 50 60 70 80 90

MZ*

(e)

0.14 0.12 0.10 0.08 0.06 0.04 0.02

UZa*

(f)

wt% TiO2

0.12 0.10 0.08 0.06 0.04 0.02 0.00 10 0.16 0.14 20 30 40 50 60 70

SP20
80 90

0.00 10

(no samples analysed)


20 30 40 50 60 70 80 90

X An
UZb*

(g)

wt% TiO2

0.12 0.10 0.08 0.06 0.04 0.02

SP29
0.00 10 20 30 40 50 60 70 80 90

X An
Fig. 5. TiO2^XAn core and interstitial compositional data for traverses through the Marginal Border Series on the Skaergaard Peninsula (grey circles) and northern Kraemer Island (black triangles), with Layered Series primocryst compositions for comparison (grey crosses, as shown in Fig. 3). Black continuous lines indicate representative continuous zoning profiles from each dataset. No data are available for UZa. Grey arrow in UZb* (g) points to analyses from ilmenite-rich intergrowths (up to 023 wt % TiO2). Sample SP60 is from base LZa*, Skaergaard Peninsula, shown in (a) for clarity.

156

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

point at low TiO2 and intermediate anorthite content (An45). In LZb* (SP8), interstitial pockets may contain any of apatite, quartz, clinopyroxene, biotite/hornblende, plagioclase, zircon and opaque minerals, and account for $067% by area of the section (Table 1). Interstitial plagioclase compositions lie on the primocryst trend until $An47. Fewer analyses show very evolved compositions than in the LZa*^LZb* sample (SP46), but several analyses plot on a similar normal zoning trend to that identified in SP46, reaching very low TiO2 at $An28. Highly calcic, low-TiO2 compositions were not observed, although points plotting below the primocryst trend at An49, 005 wt % TiO2 are arguably part of a steeper trend towards calcic compositions (Fig. 5c). In LZc* (SP16), the few pockets observed were highly altered or contained opaque minerals, biotite, plagioclase and apatite, and accounted for $036% by area. Interstitial plagioclase follows the primocryst trend until $An45 ( 3 mol % anorthite), when the interstitial compositions diverge towards lower TiO2. In this sample a reverse interstitial trend is not observed, only steep normal zoning (Fig. 5d). Several compositions have anomalously high TiO2 contents; these analyses were from traverses along thin interstitial plagioclase cusps between clinopyroxene and Fe^Ti oxides, and probably reflect diffusive contamination or secondary fluorescence effects. These compositions show very slight reverse zoning (Fig. 5d). For MZ* (SP20), the interstitial plagioclase compositions lie on the primocryst trend with no divergence to lower TiO2; there is no evidence for either of the two distinct interstitial zoning trends seen in the lower stratigraphic zones (Fig. 5e). Instead, zoning profiles towards interstitial pockets ($027% by area) follow the compositional trend defined by the primocrysts. Interstitial plagioclase from UZb* (SP29) again lies on the compositional trend defined by the primocrysts (Fig. 5g). Plagioclase adjacent to the granophyric nonreplacive microstructures described by Holness et al. (2011) has very low XAn ($An25), whereas plagioclase from within the coexisting ilmenite-rich intergrowths has anomalously high TiO2 (4 011wt % and up to 025 wt % TiO2, An18^38). These data were excluded on the basis that these anomalous compositions were probably caused by secondary fluorescence. The other minor elements (Mg and Fe) typically show less clear compositional differences than Ti. MgO contents of interstitial plagioclase in the Skaergaard Peninsula are low and decrease continuously with decreasing XAn, overlapping with the primocryst compositions. Zoning profiles towards silicic pockets typically show constant or decreasing MgO, and zoning profiles towards mafic pockets have little change in MgO contents. For FeO, zoning towards

the felsic pockets or granophyre typically shows constant or slightly decreasing FeO contents, and the up-turn in FeO seen for cumulus minerals at $An40 is not seen (Fig. 6). Zoning towards mafic pockets tends towards higher FeO (Fig. 6). For MZ and above, Fe compositions are scattered and show no clear trend.

Interstitial plagioclase: Marginal Border Series, Kraemer Island


Kraemer Island has a low structural position within the intrusion, so the MBS comprises only HZ*, LZa* and part of LZb* at this location. In terms of MgO and FeO content, the Kraemer Island plagioclase compositions overlap with those from the Skaergaard Peninsula, but FeO contents are scattered towards high values (Fig. 6; Electronic Appendix 2). However, in TiO2 content, interstitial plagioclase from Kraemer Island samples shows some differences compared with the Skaergaard Peninsula. The part of the Skaergaard Peninsula interstitial trend that matches the primocrysts is typically also displayed by interstitial plagioclase from Kraemer Island. However, nearly all the analyses from Kraemer Island plot on the reverse zoning interstitial trend, similar to that defined adjacent to Fe-rich pockets in the Skaergaard Peninsula (Fig. 4). Many more reversed compositions are found compared with the Skaergaard Peninsula, fleshing out the shape of this trend and (for LZb*) extending it (Fig. 5). The planar-sided pockets found in the Kraemer Island samples are much less common than those in the Skaergaard Peninsula (0015^020% by area; Table 1); typically the identifiable minerals are apatite and pyroxenes as well as plagioclase.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

Interstitial plagioclase: Layered Series, Cambridge drill core and drill core 90-22
The compositions of plagioclase from the central Layered Series are shown in Figs 7 and 8, and extend earlier observations (Humphreys, 2009). The results show a complex mixture of the different XAn^TiO2 trends observed in the Marginal Border Series. In HZ (sample 118678, Fig. 7a), the compositions are almost identical to those of sample SP60 on Skaergaard Peninsula (see Fig. 5; Electronic Appendices 3, HZ-MZ, and 4, UZ). Interstitial pockets represent $011% by area (Table 1). LZa plagioclase compositions (sample 118653) are similar to those of HZ, the primocryst trend being followed until $An52, with a maximum TiO2 content of 013 wt % observed. The interstitial compositions separate to low and high XAn contents at low TiO2 (Fig. 7b), as for HZ, but in contrast to LZa* plagioclase. However, the lowest and highest XAn compositions observed in these divergent trends are equivalent to those seen in LZa* (An35 and An75, respectively). Interstitial pockets represent $020% by area (Table 1). Plagioclase compositions from LZb (samples 118605, 118601, 118590) follow the primocryst trend until An49,

157

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

0.8

HZ*

(a)

0.8

LZa*

(b)

wt% FeO

0.6

0.6

0.4

0.4

0.2 0 10 20 30 40 50 60

0.2

SP60 KR02_4
70 80 90

0 10 20 30 40 50 60

SP46 KR02_37
70 80 90

0.8

LZb*

(c)

0.8

LZc*

(d)
Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

wt% FeO

0.6

0.6

0.4

0.4

0.2 0 10 20 30 40 50 60

0.2

SP8 KR02_21
70 80 90

0 10 20 30 40 50 60 70

SP16
80 90

0.8

MZ*

(e)

0.8

UZa*

(f)

wt% FeO

0.6

0.6

0.4

0.4

0.2 0 10 20 30 40 50 60 70

0.2

SP20
80 90

0 10 20 30

(no samples analysed)


40 50 60 70 80 90

X An
0.8

UZb*

(g)

wt% FeO

0.6

0.4

0.2 0 10 20 30 40 50 60 70

SP29
80 90

X An
Fig. 6. FeO^XAn core and interstitial compositional data for the Skaergaard Peninsula and northern Kraemer Island traverses through the Marginal Border Series, with Layered Series primocryst compositions for comparison, as in Fig. 3. Samples and symbols as for Fig. 5.

158

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

0.16 0.14 0.12

0.16

HZ

(a)

0.14 0.12 0.10 0.08 0.06 0.04 0.02

LZa

(b)

wt% TiO2

0.10 0.08 0.06 0.04 0.02 1966-118678 0.00 10 0.16 0.14 0.12 20 30 40 50 60 70 80 90

0.00 10 0.16

1966-118653 20 30 40 50 60 70 80 90

LZb

(c)

0.14 0.12 0.10 0.08 0.06 0.04

LZc

(d)
Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 10 0.16 0.14 0.12 20 30 40 50 60 1966-118590 1966-118601 1966-118605 70 80 90

0.02 458279 0.00 10 0.16 20 30 40 50 60 70 80 90

MZ

(e)

0.14 0.12 0.10 0.08 0.06 0.04

UZa

(f)

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 10 0.16 0.14 0.12 20 30 40 50 60 70 80 90 90-22-983.97 and 995.09 90-22-966.16 and 1000.08

0.02 0.00 10 20 30 40 50 60

90-22-481.8 90-22-824.2 70 80 90

X An UZb (g)

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 10 20 30 40 50 60 70 80 90 90-22-421 90-22-461.8

X An
Fig. 7. TiO2^XAn core and interstitial compositional data for the central Layered Series from drill cores 1966 and 90-22 as well as surface sample (458279), with Layered Series primocryst compositions for comparison (grey crosses; see Fig. 3). Samples are 118678, HZ (a); 118653, LZa (b); 118605, 118601 and 118590, LZb (c); 458279, LZc (d). MZ compositions (e) include unpublished zoning traverses from samples 90-22-966.16 and 90-22-1000.08 (diamonds, Stripp, 2009) and plagioclase zoning towards granophyre veins in samples 90-22-995.09 and 90-22-983.97 (crosses, Stripp, 2009). Upper zone samples include 90-22-824.2 (f, lower UZa), 90-22-481.8 (f, UZa^UZb boundary), 90-22-461.8 (g, lower UZb) and 90-22-421 (g, mid-UZb). Black continuous lines indicate representative continuous zoning profiles from each dataset.

159

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

0.8

HZ

(a)

0.8

LZa

(b)

wt% FeO

0.6

0.6

0.4

0.4

0.2 1966-118678 0 10 20 30 40 50 60 70 80 90

0.2 1966-118653 0 10 20 30 40 50 60 70 80 90

0.8

LZb

(c)

0.8

LZc

(d)
Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

wt% FeO

0.6

0.6

0.4

0.4

0.2 0 10 20 30 40 50 60

1966-118590 1966-118601 1966-118605 70 80 90

0.2 458279 0 10 20 30 40 50 60 70 80 90

0.8

MZ

(e)

0.8

UZa

(f)

wt% FeO

0.6

0.6

0.4

0.4

0.2 90-22-983.97 and 995.09 90-22-966.16 and 1000.08 0 10 20 30 40 50 60 70 80 90

0.2 90-22-481.8 90-22-824.2 0 10 20 30 40 50 60 70 80 90

X An
0.8

UZb

(g)

wt% FeO

0.6

0.4

0.2 90-22-421 90-22-461.8 0 10 20 30 40 50 60 70 80 90

X An
Fig. 8. FeO^XAn core and interstitial compositional data for the central Layered Series. Samples and symbols as for Fig. 7.

160

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

where the interstitial trend diverges towards low XAn and low TiO2. The interstitial trend is similar to that seen for LZa, but no highly calcic compositions were observed except for one point at An55 (Fig. 7c). The mode of interstitial pockets is variable in the three LZb samples, representing $006 to $033% by area; this could reflect sampling issues or inter-sample variations (see below). In LZc (sample 458279) the interstitial trend diverges from the primocryst trend at $An48. The interstitial trend is similar to that seen for LZb until $An55; most compositions become more calcic with decreasing TiO2, although there is also a hint of the normal zoning trend at around An46 (Fig. 7d). One highly calcic interstitial plagioclase composition (An78) was recorded at a pl^pl^cpx triple junction. Interstitial pockets are very rare in this sample, representing only $003% by area (Table 1). In MZ (samples from 96616 m and 100008 m depth in drill core 90-22; data from Stripp, 2009), the primocryst trend is not followed substantially: late-stage interstitial plagioclase plots below the primocryst trend at $An45. Few low-TiO2 compositions were observed, making it difficult to assess how these data fit into the series. Zoning profiles towards isolated granophyre veins in other MZ samples (99509 m and 98397 m in drill core 90-22; data from Stripp, 2009) lie on the primocryst trend until $An38 and then diverge to lower TiO2 at $An32 (Fig. 7e). A few analyses plot at anomalously high TiO2 contents (above the primocryst trend); these may result from diffusive contamination or secondary fluorescence effects and appear similar to those from LZc* (see Fig. 5d). Interstitial plagioclase compositions in UZa (8242 m and 4818 m in drill core 90-22) show variable trends. The lower sample, from 8242 m in mid-UZa, shows two trends of low-TiO2 interstitial compositions that diverge at $An43, one towards $An30 and one towards $An55 (Fig. 7f). In contrast, interstitial plagioclase from 4818 m (at the UZa^UZb boundary) lies on the primocryst trend up to very evolved compositions ($An20), with no evidence of highly calcic, low-TiO2 compositions. Interstitial pockets are very rare (5 001% by area in sample from 8242 m; Table 1). Plagioclase compositions from UZb (samples from 461 8 m and 212 m in drill core 90-22), lie on a linear trend of decreasing TiO2 and XAn, with compositions very similar to those from the UZa^UZb boundary. The trend is very slightly steeper than the equivalent trend for the Skaergaard Peninsula (Fig. 7g). Again, there is no evidence of highly calcic, low-TiO2 compositions. Two analyses plot at higher TiO2; these are probably similar to high-TiO2 compositions from MZ. Interstitial pockets are present in the sample from 4618 m (026% by area), which has no non-replacive symplectites (Table 1). As with the MBS (see above) and the lower parts of the LS (Humphreys, 2009), MgO and FeO contents of

interstitial plagioclase also vary with XAn. MgO contents are lower than in the MBS, and are highest in HZ^LZ (005 wt % MgO), decreasing continuously to $001wt % at UZb. FeO contents (Fig. 8) tend to be slightly lower than in the MBS, as with the primocryst compositions, and decrease from $045 wt % in HZ to $015 wt % in upper UZb. More evolved compositions tend either towards high FeO contents with little difference in FeO, or towards lower FeO and lower XAn. As with the MBS samples, the most evolved plagioclase compositions (An30 and below) do not show the increase in FeO observed in the primocryst compositions. The two distinct trends are most clear for MZ, UZa and UZb. For the MZ samples, the low-Fe trend is defined by plagioclase adjacent to granophyre veins (Stripp, 2009).

Interstitial plagioclase: Layered Series, drill core 90-10


Four samples in MZ and UZa were examined from drill core 90-10, to determine whether spatial position within the Layered Series affects interstitial plagioclase compositions and to look more closely at intra-zone variations (Figs 9 and 10; Electronic Appendix 5). For equivalent stratigraphic positions, the plagioclase compositions are similar to those of drill core 90-22, with core compositions in the range An40^48. In general, plagioclase from drill core 90-10 is slightly more Fe-rich and more Mg-rich than that from drill core 90-22, similar to plagioclase from the MBS. In terms of Ti^XAn variations, there is limited evidence of highly calcic, low-TiO2 compositions, except for a few analyses from 90-10-524.1 (meso-gabbro, lower MZ) and decreasing TiO2 at approximately constant XAn in 90-10-456 (oxide-rich layer, mid-MZ; Fig. 9). This behaviour is similar to that seen for drill core 90-22 in LZc or MZ. Samples 90-10-445 (meso-gabbro, top MZ) and 90-10-212 (mid-UZa) show clear normal zoning with a steep decrease in TiO2 between An45 and An35 (Fig. 9), similar to the zoning adjacent to veins of granophyre in lower MZ (drill core 90-22, Stripp, 2009; see Fig. 7e). There is no clear progression of composition with stratigraphy as expected in these MZ samples; the steepness of the Ti^XAn trend varies non-systematically with stratigraphic height and may depend on the bulk composition of the sample, with the steepest trend observed in the most oxide-rich sample (90-10-456; Fig. 9). The abundance of interstitial pockets varies between these four samples, from 013 to 028% by area where present, but the pockets are absent in 90-10-456 (Table 1).

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

Summary
In general, interstitial plagioclase from a given stratigraphic zone shows a much greater compositional range than primocrysts from that zone. Minor element concentrations, in particular TiO2, vary systematically with XAn, but interstitial plagioclase compositions do not always lie on the

161

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

0.16 0.14 0.12

0.16

MZ

(a)

0.14 0.12 0.10 0.08 0.06 0.04 0.02

MZ

(b)

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 10 0.16 0.14 0.12

90-10-524.1
20 30 40 50 60 70 80 90

90-10-445
20 30 40

90-10-456
50 60 70 80 90

0.00 10

X An UZa (c)
Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

wt% TiO2

0.10 0.08 0.06 0.04 0.02 0.00 10

90-10-212
20 30 40 50 60 70 80 90

X An
Fig. 9. TiO2^XAn core and interstitial compositional data for drill core 90-10, with primocryst compositions for comparison (grey crosses, as in Fig. 3). Samples are from MZ (a, b) and UZa (c). Black continuous lines indicate representative continuous zoning profiles from each dataset.

same Ti^XAn trend as the primocrysts. In late-stage interstitial plagioclase from lower stratigraphic zones the interstitial trend deviates from that of the primocrysts, and the XAn composition at which the two trends separate depends on stratigraphic position. In the least evolved parts of the Skaergaard Peninsula (Marginal Border Series), reverse zoning can be observed adjacent to interstitial mafic pockets, whereas normal zoning is observed adjacent to interstitial silicic pockets in the same sample, thus defining two distinct interstitial trends (single spot analyses also lie on these trends). In structurally lower parts of the intrusion (lower parts of the Layered Series and northern Kraemer Island), the normal zoning trend is poorly developed or even absentthe late-crystallizing interstitial plagioclase is dominated by calcic compositions. In the uppermost parts of the stratigraphy, interstitial plagioclase compositions define continuously decreasing TiO2 and XAn, and lie on the same trend as the primocrysts. In the MBS, this occurs for MZ* and above (there is no information above LZb* for Kraemer Island). However, in the LS this does not occur until near the UZa^UZb boundary and may depend on lithology. Of the other minor elements, MgO contents consistently decrease with XAn, whereas FeO contents decrease in

evolved plagioclase crystallizing near granophyre, but remain high and scattered in plagioclase crystallizing near mafic pockets. FeO and MgO contents are higher in the MBS than in LS.

C RY S TA L L I Z AT I O N O F T H E I N T E R ST I T I A L L IQU I D
The Skaergaard magma is generally agreed to have undergone almost perfect fractional crystallization to form the cumulates that are observed. Thus the variation of immobile elements (such as Ti) with tracers of differentiation (XAn) can be used to track the magma evolution. For example, the partition coefficient for Ti in plagioclase (DTipl) may change slightly with temperature but this cannot explain the sudden downturn in TiO2 in the primocrysts, which coincides with the arrival of primocryst Fe^Ti oxides in LZc. Instead, this is interpreted to represent the point at which Ti becomes compatible in the bulk magma (Humphreys, 2009). These compositional variations in the primocrysts represent an end-member mode of differentiationnear-perfect fractional crystallization of the bulk magmathat can be used as a reference for crystallization of the interstitial liquid. The departure of

162

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

0.8

MZ

(a)

0.8

MZ
90-10-445

90-10-456

(b)

wt% FeO

0.6

0.6

0.4

0.4

0.2

0.2

90-10-524.1
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90

X An
0.8

UZa

(c)
Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

wt% FeO

0.6

0.4

0.2 0 10 20 30

90-10-212
40 50 60 70 80 90

X An
Fig. 10. FeO^XAn core and interstitial compositional data for drill core 90-10, with primocryst compositions for comparison. Samples and symbols as for Fig. 9.

the interstitial trend from that of the primocrysts therefore indicates that differentiation of the mush liquid does not always occur under conditions of near-perfect fractional crystallizationthe interstitial liquid and the bulk magma can follow different liquid lines of descent. This departure from the primocryst trend typically occurs after an initial period when interstitial and primocryst compositions coincide (e.g. Fig. 5b). The XAn composition at which primocrysts and interstitial plagioclase trends diverge decreases with increasing stratigraphic height, suggesting a porosity or permeability control on crystallization conditions. The gradient of the interstitial trends also varies both within samples and with stratigraphic position. However, the uppermost parts of the stratigraphy contain interstitial plagioclase compositions that do not diverge from the primocryst Ti^XAn trend: this suggests that in the latest stages of crystallization, the mush undergoes fractional crystallization much as the bulk magma does. A key question, therefore, concerns the processes controlling the interstitial compositional variations from the lower stratigraphic zones. Modelling is made difficult by continuous solid solution in plagioclase, olivine and pyroxenes; peritectic reactions (e.g. pyroxene^olivine); the unusual, highly Fe-rich melt compositions reached; and the lack of certainty about the original (primocryst)

compositions of mafic phases owing to sub-solidus diffusion. However, in a simplified model that assumes negligible compaction, compositional convection or diffusive processes, Ti can be treated simply as a trace element in a liquid that is undergoing progressive fractional crystallization, with XAn used as a passive marker of both temperature (T) and F, the fraction of liquid remaining (Humphreys, 2009). Using this approach, different bulk partition coefficients for Ti (DB Ti ) can be used to test the effects of different crystallizing assemblages, from a starting point at moderate XAn and high TiO2 (Fig. 11). In the absence of a formula specific to crystallization of the interstitial liquid, a polynomial function derived from experimental petrology studies (T oplis & Carroll, 1996) T K 1295 265F 290F 2 1753 1

is used to determine T as a function of F, and thus derive XAn and DTipl (following Humphreys, 2009): T  C 3:61XAn 899 ln
pl DTi

2 3

0:00535T  C 9:458:

For simplicity, only three phases are considered: plagioclase, Fe^Ti oxides and clinopyroxene. The DTi values for olivine and plagioclase are very similar, so neglecting

163

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

0.16 0.14 0.12

wt% TiO2

0.10

1.
0.08

3.
0.06 0.04 0.02

2. 4.

0.00

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

10

20

30

40

50

60

70

80

X An
Fig. 11. Numerical modelling results for different bulk Ti partition coefficients, superimposed on interstitial data from sample SP46 (LZa*^ LZb* boundary, Skaergaard Peninsula) for comparison (symbols as in figure 5b). The continuous curve (1) represents the model from Humphreys (2009) using DB Ti $ 32 after An54 (equivalent to 20% Fe^Ti oxides 40% plagioclase 40% clinopyroxene). The long-dashed 7 (equivalent to 10% Fe^Ti oxides 50% plagioclase 40% clinopyroxene). The dotted line (3) represents line (2) represents DB Ti $ 1 DB Ti $ 12 (equivalent to 7% Fe^Ti oxides 70% plagioclase 23% clinopyroxene). The short-dashed line (4) indicates crystallization with DB Ti $ 24 (equivalent to 45% plagioclase 40% clinopyroxene 15% Fe^Ti oxides), with a starting composition of 42 wt % TiO2 and assuming initial F of 028.

olivine may overestimate the modal abundance of plagioclase, but plagioclase clinopyroxene Fe^Ti oxides represent most of the interstitial material in most rocks. Despite its over-simplicity (ignoring crystallization of all accessory minerals, for example), the model is useful because it demonstrates that, to a first order, significant variations in Ti^XAn curvature can be produced simply by different bulk partition coefficients (DB Ti ), which may correspond to differing proportions of the interstitial phases (Fig. 11). The curvature essentially depends on the proportion of Fe^Ti oxides crystallizing relative to plagioclase and olivine, because of their strong differences in DTi. Importantly, the normal zoning trends observed in the MBS can be reproduced well, although no unique results can be obtained. For example, profiles from sample SP46 (LZa*^LZb* boundary, Skaergaard Peninsula) can be matched by a low bulk DTi that could equate to crystallization of 45% plagioclase, 40% clinopyroxene and 15% Fe^ Ti oxides (Fig. 11). Therefore, at least in places, interstitial crystallization can be explained by relatively simple fractional crystallization processes. In principle, using the same approach to model equilibrium crystallization will result in a slower decrease of Ti concentrations compared with fractional crystallization: equilibrium crystallization in the mush cannot therefore explain the differences between interstitial and primocryst plagioclase compositions. However, the interstitial reverse zoning cannot be produced by simple fractional crystallization and requires

a different mechanism. This is of course because XAn is calculated as a passive marker of T and F, and therefore decreases monotonically with progressive crystallization. The liquid Ca/(Ca Na) is not modelled explicitly; the simplified dependence of XAn on temperature is taken on the basis of the available experimental data (T oplis & Carroll, 1995, 1996; Thy et al., 2006). However, processes that might cause changes in liquid composition must clearly be considered (see below). Given the apparently strong effects of modal mineralogy on the Ti^XAn gradient, one might expect the primocryst trend to be curved, as predicted by the modelling of Thy et al. (2009), or to show significant changes in slope at each sub-zone boundary to reflect the changes in mineralogy. However, the primocryst trend is almost linear for zones LZc and above, suggesting that DB Ti changed little during differentiation of the upper parts of the intrusion.

Origin of calcic plagioclase rims and reverse zoning


As demonstrated above, the normal interstitial zoning trends observed are consistent with a relatively simple process of fractional crystallization, but the reverse zoning observed in the lower parts of the intrusion is not and therefore requires a different mechanism. In principle, increased XAn in plagioclase could result from increased temperature, increased pH2O, or an increase in the anorthite component of the liquid. For the interstitial

164

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

liquid of a slowly cooled basaltic crystal mush, an increase in temperature during solidification seems unlikely, and although reversely zoned plagioclase has been produced experimentally by metastable crystallization during rapid quenching (Lofgren, reported by Grove et al., 1973), this is not applicable to intrusive gabbroic suites. If pH2O were the cause of the increasing anorthite contents, significant quantities of H2O would need to be present in the liquid (Sisson & Grove, 1993; Koepke et al., 2005), and the absence of hydrous phases associated with the calcic rims argues against this. The presence of adjacent calcic and sodic rims within the same thin section would require the formation of interstitial pockets with widely differing H2O contents; this also seems unlikely. Furthermore, regions of high pH2O should occur preferentially in the outermost parts of the intrusion, where external water could most easily infiltrate the rocks, or in the most fractionated parts (i.e. UZc), where incompatible components are most strongly enriched, and not in the centre of the LS. The most likely cause of the reverse zoning and calcic rims therefore seems to be a change in the composition of the interstitial liquid, as argued by Humphreys (2009). There are several possible causes of changing liquid composition, including convective exchange within the crystal mush, dissolution and reprecipitation during compaction, infiltration metasomatism, and silicate liquid immiscibility within the mush.

of the crystal mush, as demonstrated by the very small proportion of total interstitial plagioclase represented by calcic compositions. This suggests that convection within the crystal mush is not the reason for the observed grain boundary zoning.

Infiltration metasomatism
Metasomatism by expulsion of interstitial liquids from lower parts of the crystal mush during compaction was suggested by Irvine (1980) to explain discrepancies between observed modal and geochemical zoning patterns in the cyclically layered Muskox intrusion, Canada. Compaction also occurs at Skaergaard (e.g. McBirney, 1995; T egner et al., 2009); however, unlike the Muskox Intrusion, the Skaergaard Intrusion represents essentially a single pulse of magma that solidified through continuous fractional crystallization. Residual liquids deeper in the crystal mush are therefore probably more evolved, not more primitive, than those at shallow depths in the mush or even in the bulk magma (although see Boorman et al., 2004), so infiltration metasomatism at Skaergaard would result in normal zoning, not reverse zoning. Injection of new, primitive magma (e.g. Roelofse et al., 2009) is ruled out because of a lack of supporting field evidence, because the reversals are seen in samples covering a wide stratigraphic range, and because the reversals are seen only in the outermost grain margins.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

Compositional convection
Compositional convection within the mush can occur if the residual liquid becomes less dense than the adjacent crystals or overlying melt (e.g. Sparks et al., 1984; Morse, 1986). Convection would result in the supply of fresh, dense, unevolved melt to the crystallization site, and removal of more evolved, less dense melt, and could result in adcumulus textures if the convective velocity is greater than the growth rate of the crystallization front (e.g. Tait et al., 1984; Kerr & Tait, 1986; Tait & Jaupart, 1992). For floor cumulates, the melt composition at the magma^ mush interface will be essentially identical to that in the main magma reservoir, with more evolved residual melt near the base of the mush (lower Ca content, higher FeO, higher alkalis, Thy et al., 2009; although see Boorman et al., 2004). Compositional convection within the interstitial liquid could therefore result in mineral zonation near the base of the mush. This would correspond to decreasing FeO, increasing Mg-number for mafic phases, and increasing XAn (reverse zoning) in plagioclase. However, once the interstitial liquid is saturated with Fe^Ti oxides, TiO2 concentrations in the interstitial liquid will fall below those of the overlying bulk magma, so this reverse zoning would also be associated with increasing TiO2 in plagioclase, whereas the observations show increasing FeO and decreasing TiO2. Compositional convection would also be strongly hindered by the low permeability near the base

Dissolution^reprecipitation
Local dissolution of plagioclase can result from compaction during post-cumulus growth, owing to dissolution of unfavourably oriented plagioclase and reprecipitation of the material in lower stress orientations, as suggested for the Stillwater Intrusion (Meurer & Boudreau, 1998) and the Skaergaard Intrusion (Boudreau & McBirney, 1997). Maale (1976) showed that the dissolving plagioclase grain is always more An-rich than that precipitating, which should result in enrichment of the anorthite component in the interstitial liquid, and could explain local reverse zoning. However, this mechanism cannot account for reversed plagioclase within the steeply dipping, and relatively rapidly cooled MBS, where compaction should be negligible, nor the transition to normal (i.e. non-reversed) overgrowth rims in higher stratigraphic zones of the MBS. Furthermore, T egner et al. (2009) showed that compaction is unimportant in HZ^LZa, where strong reverse zoning is seen.

Silicate liquid immiscibility


Immiscibility of the interstitial liquid was not previously considered (Humphreys, 2009) but could explain the differing compositional trends observed in interstitial plagioclase. Liquid immiscibility in silicate magmas has been demonstrated unambiguously by petrographic observations (e.g. De, 1974; Philpotts, 1979, 1982), as well as

165

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

experimental studies (e.g. Roedder, 1951; Dixon & Rutherford, 1979; Philpotts & Doyle, 1983). Late-stage immiscibility within the Skaergaard magma specifically has also been demonstrated, both by the presence of two distinct populations of melt inclusions in apatite primocrysts (UZb, Jakobsen et al., 2005) and by experimental observations (McBirney & Nakamura, 1974; Veksler et al., 2007). It has been suggested that in fact the magma intersects the two-liquid miscibility gap much earlier in the fractionation process, around upper LZ to upper MZ, when the magma was $45^60% crystallized (Veksler et al., 2007; Holness et al., 2011), although this is controversial (McBirney, 2008; Morse, 2008a; Philpotts, 2008; Veksler et al., 2008). Irrespective of the stage of fractionation at which immiscibility occurs, the magma at Skaergaard exsolves into an Fe-rich liquid that is also rich in Ca and P, and a Si-rich liquid that is also rich in alkalis (e.g. McBirney & Nakamura, 1974; Jakobsen et al., 2005; Veksler et al., 2007). At thermodynamic equilibrium, the chemical potential of each component must be the same in every phase in the system, including both conjugate liquids. However, this does not equate to equal equilibrium concentrations, hence the markedly different compositions of conjugate immiscible liquids and their different trace element concentrations (e.g. Veksler, 2009). Furthermore, each conjugate liquid may crystallize different modal proportions of the minerals present, and partition coefficients can also differ (Roedder, 1978; Philpotts, 1979; Veksler et al., 2006). For minerals with complete solid solution, the effects of liquid immiscibility on an intermediate composition can be thought of in terms of the end-member components. For example, in plagioclase feldspar the Si-rich, alkali-rich liquid would crystallize dominantly the albite (NaAlSi3O8) component whereas crystallization from the Fe-rich, Ca-rich liquid would be dominated by the anorthite component (CaAl2Si2O8). At equilibrium, when both conjugate liquids are present in cotectic proportions, there is no change in the feldspar composition compared with crystallization from a single liquid, but disequilibrium conditions could result in crystallization of more An-rich plagioclase from the Fe-rich liquid and more An-poor plagioclase from the silicic liquid. In the Skaergaard cumulates, the observation of highly discrepant compositional zonation in different parts of the same samples indicates that disequilibrium crystallization must be occurring, resulting in the crystallization of anomalously calcic plagioclase.

C RY S TA L L I Z AT I O N F RO M IMMISCI BLE LIQUI DS


Because of experimental difficulties, the composition of plagioclase in equilibrium with each of the liquids

separately is not known independently, but can be estimated if the anorthite component (XCaAl2Si2O8) and albite component (XNaAlSi3O8) of the conjugate liquids are known. These components can be calculated (Lange et al., 2009) from known or estimated two-liquid pairs (Table 2). The calculated anorthite components vary from 0064 to 1175 for Fe-rich conjugate liquids and from 0076 to 0247 for Si-rich conjugate liquids, whereas the calculated albite components vary from 0047 to 0162 for Fe-rich and from 0286 to 0613 for Si-rich conjugate liquids (Table 2), based on liquid compositions from re-homogenized melt inclusions (Jakobsen et al., 2005), experimental compositions (Veksler et al., 2007) and melting experiments (McBirney & Nakamura, 1974). The higher factors for Si-rich liquids reflect their higher affinity for both calcic and sodic feldspars, but in every case, the nominal XAn for the liquid, calculated as XCaAl2Si2O8/ (XCaAl2Si2O8 XNaAlSi3O8), is substantially lower for the Si-rich liquids (0120^0464) than the Fe-rich liquids (0391^0636, Table 2). This demonstrates that, if the liquids are out of equilibrium with one another, the Fe-rich liquid should crystallize more calcic plagioclase, whereas the Si-rich liquid should crystallize more evolved (albite-rich) plagioclase. Thus the observed reverse zoning and anomalously high anorthite contents of some interstitial plagioclase are compatible with disequilibrium crystallization from an emulsion with greater than cotectic proportions of the Fe-rich component, or from the Fe-rich liquid itself. It is concluded that strong disequilibrium can occur within the mush, and that immiscible melts that exsolve from the bulk residual liquid do not necessarily remain in equilibrium with each other after exsolution, as suggested by Crawford & Hollister (1977). This concept is supported by the results of several previous studies. Ripley et al. (1998) reported calcic plagioclase (An71^79) in nelsonite bodies from the Duluth Complex, Minnesota, which are thought to have formed from an immiscible Fe-rich liquid. Similarly, Loferski & Arculus (1993) reported highly calcic ( An92) rims around plagioclase-hosted melt inclusions that had crystallized to cpx ilm ap titanite, rutile and baddeleyite, in rocks from the Stillwater Intrusion, Montana. In lunar basalts, two distinct plagioclase compositional trends were observed adjacent to patches of high-K mesostasis (to An64) and low-K mesostasis (to An94^96, Crawford & Hollister, 1977). The Loch Ba ring dyke, Scotland, formed by mixing of rhyolitic and mafic melts; the melts contain plagioclase of An21^32 (rhyolite) and An30^65 (mafic) respectively (Sparks, 1988). Disequilibrium crystallization could occur as a result of physical separation, perhaps by gravitational separation of two liquids of differing densities, or by diffusion rates falling below the threshold required to maintain equilibrium (Roedder, 1978). This would require the liquid droplets to

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

166

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

T able 2: Compositions of conjugate immiscible silicate liquids from Skaergaard


Jakobsen et al. (2005) Fe-rich Si-rich McBirney & Nakamura (1974) Fe-rich Si-rich Philpotts (1982) Fe-rich Si-rich Dixon & Rutherford (1979) Fe-rich Si-rich

SiO2 TiO2 Al2O3 FeOt MnO MgO CaO Na2O K2O P2O5 Total n XCaAl2Si2O8 XNaAlSi3O8 XAn liquid

4067 186 787 3085 051 235 897 158 103 025 9594 54 011 008 058

6558 022 1295 863 013 047 2 433 368 003 9802 17 008 051 014

514 24 67 266

655 12 96 119

415 58 37 310 05

733 08 121 32 00 00 18 31 33 007 9767 16 008 061 012

439 461 70 2345 055 232 1018 185 042 487 9915 7 012 009 057

685 170 111 786 016 075 382 286 12

04 67 22 1 17 991

04 31 2 26 03 966

09 94 08 07 35 978 15

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

096 9891 7 015 047 024

010 016 038

011 037 023

006 005 055

The compositions are based on the compositions of homogenized melt inclusions in cumulus apatite (Jakobsen et al., 2005), melting experiments (McBirney & Nakamura, 1974; Dixon & Rutherford, 1979) and analysis of exsolved glasses (Philpotts, 1982). Dataset from Jakobsen et al. (2005). Anorthite components (XCaAl2Si2O8) and albite component (XNaAlSi3O8) of the liquids are calculated following Lange et al. (2009). XAn liquid XCaAl2Si2O8/ [XCaAl2Si2O8 XNaAlSi3O8].

be separated by more than the characteristic diffusion distance over the relevant crystallization timescale, and could be achieved either by complete physical segregation of the immiscible liquids, perhaps through gravitational separation (because the silicic liquid will have much lower density than the Fe-rich liquid), or by partial but incomplete separation at different temperatures and crystallization rates. A droplet separation of 100 mm, for example, might permit equilibrium crystallization at slow crystallization rates (when there is a longer time available for diffusion) or at higher temperatures (when diffusivities are higher), yet at more rapid crystallization rates or low temperatures the same droplet separation could result in disequilibrium and hence changes in mineral composition. This means that, under certain conditions, an emulsion that contains non-cotectic proportions of two immiscible liquids can undergo disequilibrium crystallization to form crystals of anomalous composition. A quantitative decoupling of the effects of crystallization rate and diffusivity is beyond the scope of this study. However, the compositions of crystal rims will depend on the relative rates of droplet exsolution, liquid^liquid segregation, diffusion and crystallization with falling temperature. The permeability of the crystal mush may also be important in driving disequilibrium, by producing pockets

of varying composition, in particular in the least evolved cumulates. Immiscibility in the interstitial liquid will occur at a very late stage of crystallization in the LZ cumulates, when the residual porosity and permeability are very low. Local differences in modal mineralogy could cause the composition of the residual liquid in each pore to vary widely, and residual liquids remaining in the pores may not be in chemical equilibrium. The wide variation of plagioclase compositions that can occur around a single pocket demonstrates that these final stages of solidification can be highly localized. Differing proportions of the two liquids in these residual pockets, or perhaps the interstitial liquid locally entering the miscibility gap from opposite sides of the solvus, could therefore result in the disequilibrium trends observed. However, it is not entirely clear why the initial interstitial trend is so steep (e.g. Fig. 5a). It is possible that the steep gradient reflects high DTi and may also be related to some permeability threshold, but this cannot be constrained further using the data presented here.

Interstitial liquid immiscibility in the Skaergaard Peninsula


The clearest signature of plagioclase crystallization from single conjugate liquids is in sample SP46 from the LZa*^

167

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

LZb* boundary on the Skaergaard Peninsula. In this sample, interstitial plagioclase compositions show clear normal zoning in the vicinity of planar-sided interstitial pockets containing quartz granophyre, but reverse zoning in the vicinity of a mafic pocket dominated by apatite pyroxenes Fe^Ti oxides (see above and Fig. 4); single spot analyses also fall on these trends. Because of the disparate bulk major-element compositions of the interstitial pockets, and their similarity to the crystallized products of immiscible melt inclusions in apatite (Jakobsen et al., 2005), the pockets are interpreted as the crystallized products of trapped immiscible residual liquids (see above and Holness et al., 2011). The normal and reverse plagioclase zoning trends form by disequilibrium crystallization towards pores dominated by the Si-rich and Fe-rich liquids, respectively. This is supported by the compositional similarity to plagioclase in granophyre and mafic replacive symplectites, which represent the crystallized products of immiscible residual liquids (see Holness et al., 2011, fig. 13). Thus any difference between primocryst and interstitial compositions in a given zone is related to whether disequilibrium crystallization occurs within the mush of that zone, in contrast to the bulk magma. It is worth noting that calcic plagioclase or reverse zoning may be present even in the absence of a nearby interstitial pocketreverse-zoned plagioclase commonly occurs on grain boundaries or at triple junctions, in contrast to the very An-poor compositions, which are almost always associated with quartz-bearing interstitial pockets. This could be related to the differing viscosities and/or interfacial energies of the two liquids, with the inviscid Fe-rich liquid able to form thin films on grain boundaries, and the highly viscous Si-rich liquid tending to sit in more equant pore spaces. This is consistent with the morphology of inferred residual liquid films as determined from glass-bearing nodules and cumulate rocks (Morse & Nolan, 1984; Holness, 2005; Holness et al., 2007a). It is of interest to estimate the residual porosity when the interstitial liquid intersects the miscibility gap, and this can be done approximately using the modelling results of Thy et al. (2009). According to their modelled liquids, the Skaergaard magma reaches UZa (by which time experiments and the presence of paired granophyric and ilmenite-rich intergrowths indicate that immiscibility has occurred in the bulk magma; Veksler et al., 2007; Holness et al., 2011), after 66^68% fractional crystallization. An LZa rock (comprising primocrysts plus LZa liquid and following the same liquid line of descent as the bulk magma) with an initial porosity (fi) of around 50%, would reach immiscibility after 66^68% crystallization of the interstitial liquid, leaving a residual porosity (fr) of fi(1^066) to fi(1^068), or 17^16%. Using the same approach, even if immiscibility did not start until the

equivalent of UZb (Jakobsen et al., 2005), and allowing a further 10% crystallization to account for the HZ (see Thy et al., 2009) the interstitial liquid in an HZ rock could still intersect the miscibility gap with a residual porosity in the region of 11^75% (this approach assumes that porosity is not occluded by other processes such as adcumulus growth). Therefore the observation that even the most primitive rocks undergo silicate liquid immiscibility in residual liquid retained in the last interstitial pores is consistent with previous studies demonstrating immiscibility in the bulk liquid (McBirney & Nakamura, 1974; Jakobsen et al., 2005). However, the permeability when this occurs is probably very low, so the exsolved liquids are likely to form in isolated pockets. In more evolved rocks, interstitial liquid immiscibility will occur at increasingly high fr until eventually, at the point where the bulk magma itself exsolves into two liquids, even the initial porosity will be filled with an emulsion rather than a single liquid.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

S PAT I A L D I S T R I B U T I O N O F IM MISCI BLE LIQU IDS IN MBS


The diverse compositions of interstitial plagioclase on grain boundaries and adjacent to interstitial pockets are consistent with disequilibrium crystallization from immiscible liquids during the late stages of solidification. The spatial distribution of these diverse plagioclase compositions can provide insights into the distribution of the two liquids within the crystal mush and throughout the intrusion. The simplest sample traverse to consider is that through the Marginal Border Series on the Skaergaard Peninsula. The MBS crystallized more or less in situ on the walls of the intrusion, and is therefore free from the effects of compaction and current-driven sedimentation, unlike the floor cumulates (Layered Series). The Skaergaard Peninsula traverse represents the most complete cross-section through the evolving cumulates. The presence of the two distinct interstitial plagioclase compositional trends in HZ* indicates that, at the final stages of crystallization, the residual interstitial liquid reached the miscibility gap and the two liquids were separated, probably into spatially distinct pockets, although probably only the most evolved liquids reached immiscibility, at a very low residual porosity. In LZa*, the reverse Ti^XAn trend appears nearly identical to that in HZ*, but the two interstitial trends diverge earlier, suggesting that the residual liquid formed coarser pockets and more evolved compositions. Samples from LZb* and LZc* show the same pattern, although the most extreme compositions were not observed, as reflected in the paucity of very low TiO2 contents. This may reflect a sampling issue or could reflect a genuine low abundance in the rocks (see discussion

168

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

below). However, despite a lack of definition, two distinct trends are still apparent in LZb*, with one clearly heading towards $An30 and the other discernible in several analyses below the primocryst trend at $An50 (see Fig. 5). In LZc* there are two clear trends, one of which defines normal zoning whereas the other (a zoning profile towards a plagioclase^pyroxene^oxide triple junction) does show slight reverse zoning but is affected by diffusive modification or secondary fluorescence leading to anomalously high Ti contents (see Fig. 5). These patterns also indicate that two liquids exsolved during disequilibrium crystallization of the interstitial liquid. For MZ* and above, interstitial plagioclase compositions fall on the primocryst trend. The slope is shallower than that of the normal zoning trend in stratigraphically lower zones, which suggests a lower DB Ti and a difference in abundance of pyroxene or oxides in the crystallizing assemblage. The appearance of abundant, paired granophyre and ilmenite-rich intergrowths within MZ* indicates that the onset of immiscibility in the bulk magma also occurs at approximately this point (Holness et al., 2011). This is earlier than suggested by some researchers (e.g. McBirney & Nakamura, 1974) but may be consistent with experiments (Veksler et al., 2007; although see McBirney, 2008; Morse, 2008a; Philpotts, 2008; Veklser et al., 2008) and compositional variations of melt inclusions in plagioclase near the LZ^MZ boundary (Jakobsen et al., 2006). The coincidence of the primocryst and interstitial compositional trends at MZ* and above is interpreted as the result of crystallization from equilibrium immiscible liquids, as in the bulk magma, resulting in no change in feldspar composition and no distinction between interstitial and primocryst compositions.

adjacent) pockets. For the liquids to be in equilibrium, they must be separated by less than the characteristic diffusion distance relative to the crystal interface. The granophyric and ilmenite-rich intergrowths do clearly occupy separate pockets but these are closely associated spatially; moreover, the porosity (and permeability) will be high when the pockets form, allowing good communication, and cooling rates will be relatively low. Alternatively, the paired intergrowths may in fact represent pockets of incompletely separated emulsion, with varying proportions of the two liquids. Granophyre-rich pockets may contain one or more of apatite, oxides, clinopyroxene, biotite or amphibole, and similarly the ilmenite-rich intergrowths may be intergrown with granophyre or K-feldspar as well as biotite, pyroxenes, apatite and olivine (Holness et al., 2011). Ilmenite-rich intergrowths form from the Fe-rich liquid with a subsidiary component of Si-rich liquid, whereas granophyric intergrowths are dominated by the Si-rich liquid with a lesser Fe-rich component. Either way, both liquids are still in chemical equilibrium, so no change in plagioclase composition resulted. The Ti^XAn data presented in this study cannot differentiate between these explanations; the reality is probably a combination of both.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

Summary
Primitive cumulates from the Skaergaard Peninsula undergo silicate liquid immiscibility during crystallization of the interstitial liquid. This happens when the residual porosity and permeability are low, resulting in isolated pockets containing differing proportions of the two immiscible liquids. Disequilibrium crystallization can occur as a result of rapid crystallization and physical separation of the two liquids, and leads to the formation of two distinct interstitial crystallization trends. At approximately MZ* and above, slower crystallization and inefficient separation of liquids results in equilibrium conditions and precludes the formation of the discrepant plagioclase compositions.

Disappearance of calcic plagioclase or reverse zoning


It is suggested that the primocryst and interstitial compositions coincide when the interstitial liquid no longer undergoes disequilibrium crystallization, but crystallizes in equilibrium as the bulk magma doesin other words, when the feldspar composition is buffered by the presence of both liquids in equilibrium. In the MBS this occurs within upper MZ*, where the appearance of abundant paired granophyre-rich and ilmenite-rich intergrowths (Holness et al., 2011) also implies that the bulk magma has reached the miscibility gap. These intergrowths represent large pockets of Si-rich and Fe-rich liquid, respectively, that were present within the crystal mush in the higher parts of the stratigraphy (Holness et al., 2011) so the absence of two distinct interstitial zoning trends for plagioclase there is surprising; one might have expected to see even clearer evidence for two distinct trends than in LZ. This suggests that chemical communication between the paired intergrowths was sufficiently good to preclude disequilibrium crystallization, despite residing in different (albeit

DISCUSSION Differences between Layered Series and Marginal Border Series


In comparison with the MBS, interstitial variations in LS plagioclase from the 1966 and 90-22 drill cores appear to evolve more slowly with increasing stratigraphic height (except for HZ compositions, which are indistinguishable from those in HZ* on the Skaergaard Peninsula). Specifically, the two distinct interstitial trends tend to split at lower TiO2 in the LS, and the presence of two trends persists right up to the UZa^UZb boundary, instead of dying out in MZ as in the Skaergaard Peninsula. The presence of two distinct interstitial trends up to UZa^UZb suggests that disequilibrium within the mush persisted until

169

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

later in the LS fractionation sequence than in the Skaergaard Peninsula. This is supported by the observation that the grain-boundary replacive microstructures are present until later in the stratigraphy in the LS compared with MBS (see above; Holness et al., 2011). This difference in stratigraphic position must relate to physical differences in the deposition and solidification of horizontal and vertical mushy layers. A plausible qualitative explanation is buoyancy-driven separation of the immiscible liquids within the mush, resulting in preferential loss of Si-rich, alkali-rich liquid from the LS cumulates; such separation of immiscible liquids with different densities has been highlighted in previous studies (e.g. Roedder, 1978; Philpotts, 1981). The first droplets of Si-rich immiscible liquid that nucleate from the evolving residual liquid will be sparsely distributed and probably much smaller than the average pore throat. The Si-rich droplets will be able to ascend buoyantly, provided they are large enough to overcome the yield strength of the liquid, leading to liquid^liquid fractionation. Buoyant droplet rise rate will depend on particle size and the viscosity of the host melt. With time, the droplets will coarsen and coagulate, and the pore throats will reduce in size as a result of crystallization. Eventually, larger droplets may start to block pore throats, and hence reduce permeability, inhibiting porous flow and preventing further fractionation. However, in theory the droplets can be deformed under pressure and squeeze through pore throats that are smaller than the droplet size (e.g. Chung & Mungall, 2009). This implies that liquid^liquid fractionation can continue until a larger critical droplet size if an external pressure is imposed on the system. Si-rich droplets may therefore be able to continue migrating until a later point in the floor cumulates, which are strongly compacted (T egner et al., 2009), than in the vertical Marginal Border Series, where compaction is negligible. Thus disequilibrium crystallization may persist in the LS mush until much later in the stratigraphy, because separation and fractionation of immiscible interstitial liquids can be enhanced by compaction, which is not important in the MBS. This interpretation requires that compaction is relatively rapid compared with solidification and liquid fractionation. It therefore seems likely that some of the inferred compaction (T egner et al., 2009) occurs by consolidation of the mush; that is, a shortening of the vertical mush thickness by mechanical rearrangement of the grain packing, instead of by purely solid-state deformation of the grains themselves (McKenzie, 1984; T egner et al., 2009). However, the result of differential flow between the two immiscible liquids is unlikely to have had a significant impact on the compositional evolution of the bulk magma, because the proportion of Si-rich liquid escaping from the top of the mush is probably very small. Buoyant rise of Si-rich blebs

within the exsolved bulk magma would be much more important, and could be partly the cause of the silicic, K-rich nature of the Upper Border Series (Veksler et al., 2008).

The lower Marginal Border Series (Kraemer Island): a special case?


Most interstitial plagioclase compositions from the lower parts of the MBS (Kraemer Island) are highly calcic and lie on the reverse zoning trends defined by samples from the upper parts (Skaergaard Peninsula). Plagioclase adjacent to an evolved interstitial pocket shows normal zoning in only one sample (LZb*, KR02-21). This striking contrast with the Skaergaard Peninsula suggests that the structurally low parts of the MBS contain a much smaller proportion of the Si-rich liquid component than the upper parts. Although compaction is probably unimportant in the relatively rapidly cooled MBS, differential movement of Si-rich and Fe-rich immiscible interstitial liquids could still occur along the vertical layering [downward flow of Fe-rich residual liquid was suggested by McBirney (1985)]. Unlike in the LS, sinking Fe-rich liquid would not encounter an increase in porosity as it moved downward, which could allow enhanced liquid movement in the MBS. At mid-height in the intrusion, differential loss of Si-rich liquid upward would be compensated by new liquid arriving upward from below, resulting in approximate mass balance and no overall change in the composition of the interstitial liquid. Similarly, any downward loss of Fe-rich interstitial liquid from mid-height mush could be compensated by downward flow from above. However, near the structural base of the MBS, there is no way to replace Si-rich liquid lost upwards into the overlying mush (the converse would presumably be true at the top of the MBS). The scarcity of normally zoned interstitial plagioclase from the structural base of the MBS is interpreted to reflect this porous flow filtering throughout the sidewall mush.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

Sampling and the representative elementary volume (REV)


Interstitial compositions from drill core 90-10 are interesting in that they provide some insight into the importance of modal composition in determining the interstitial plagioclase composition. In none of the MZ samples from 90-10 are the interstitial compositions coincident with those of the primocrysts. Samples from close to the UZa^ UZb boundary were not analysed, so there is insufficient evidence to say whether equilibrium mush crystallization conditions started at the same stratigraphic position near the margins of the intrusion (drill core 90-10) and in the centre (drill core 90-22). However, it is notable that the three MZ samples from drill core 90-10 show strong differences in interstitial plagioclase composition (see Fig. 9)

170

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

within the same stratigraphic zone. Furthermore, the changes in Ti^XAn gradient are not systematic with stratigraphic height, but apparently vary with modal mineralogy: sample 90-10-456, which is the middle sample but has the steepest gradient, is particularly rich in mafic phases (clinopyroxene and Fe^Ti oxides). It is therefore possible that there may be a link between the modal mineralogy of whole-rock gabbro samples and the composition of their interstitial liquids, which could have interesting implications for understanding the formation of ore deposits. However, the variability of these MZ samples also emphasizes the possible biases of sampling effects. For any porous medium, a representative elementary volume (REV) can be defined as the smallest volume from which measurements can be taken, in the expectation that those measurements will be representative of the porous medium as a whole. In most igneous petrology studies (including this one), however, the typical sample size is that of a standard thin section, irrespective of the likely size of the REV. In the context of the current study, this means that factors such as grain size as well as macro-scale homogeneity may be very important in determining how closely any given gabbro sample may approach the likely REV, particularly when the grain margins are the focus of study. For example, the Marginal Border Series rocks are generally coarser-grained than those from the Layered Series, but tend also to be rather more homogeneous in outcrop than those from the Layered Series. The samples from MZ in drill core 90-10 demonstrate the potential difficulties of studying late-stage processes in strongly layered gabbros, although conversely such strongly layered samples may also be able to provide additional constraints. Another potential problem is whether the analyses obtained from each sample are truly representative of the sample itself, particularly given the very small volumetric proportion of material considered. This may be especially important in rocks with a sizeable component of adcumulus growth. For example, the abundance of evolved interstitial pockets measured in thin sections from the lower parts of the LS varies widely, whereas in MBS they are more abundant (see Table 1). Furthermore, not all pockets could realistically be analysed, so low-abundance compositions could easily be missed, particularly as the mineralogy of the pockets can rarely be identified in full. One is also limited to investigating in detail a few snapshots of the cumulates, in contrast to studies that give complete coverage of the full stratigraphy (e.g. T egner & Cawthorn, 2010). Nevertheless, despite these issues, the minor-element compositional trends identified, and the ways in which those trends vary with height, can provide a new way of investigating post-cumulus solidification processes in slowly cooled cumulate rocks.

CONC LUSIONS
Major- and minor-element compositions of interstitial plagioclase from the Skaergaard Intrusion vary systematically with stratigraphic height and spatial location within the intrusion. In particular, decreasing Ti contents in plagioclase are interpreted to reflect falling Ti concentrations in the interstitial liquid. However, interstitial zoning does not follow the same path as the cryptic variations in primocryst compositions, indicating differences between the interstitial and bulk liquid lines of descent. In a single thin section, reverse zoning in interstitial plagioclase may be observed on grain boundaries and adjacent to planarsided interstitial pockets filled with fine-grained mafic minerals (oxides, apatites and pyroxenes), whereas normal zoning is observed adjacent to planar-sided interstitial pockets containing silicic minerals and/or granophyre. The divergent interstitial zoning trends are therefore interpreted as the result of disequilibrium crystallization from pockets of interstitial liquid that have undergone immiscibility. The interstitial liquid intersects the miscibility gap even in the most primitive cumulates at Skaergaard. Fractional crystallization from an equilibrated emulsion results in identical interstitial and primocryst plagioclase compositions in the MBS. However, compositional differences in LS plagioclase suggest differential movement of the two immiscible liquids within the mush, probably as a result of buoyant rise of Si-rich droplets aided by compaction, and a later persistence of disequilibrium conditions in the LS compared with MBS. The study of immobile components of plagioclase can thus be an important way to improve our understanding of the late-stage crystallization processes of cumulate rocks.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

AC K N O W L E D G E M E N T S
Chiara Petrone is thanked for assistance with electron microprobe analysis; Christian T egner and Troels Nielsen are thanked for the loan of samples. Gemma Stripp is acknowledged for permission to use unpublished data. This work has benefited from numerous discussions with Marian Holness, Christian T egner, Rune Larsen, Ilya Veksler and Troels Nielsen, as well as detailed and useful reviews from T ony Philpotts, Alan Boudreau and Grant Cawthorn, and editorial input from Marjorie Wilson, which significantly improved the paper.

FU N DI NG
M.C.S.H. acknowledges the support of a Junior Research Fellowship from Trinity College, Cambridge, and NERC grant NE/F020325/1.

171

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

S U P P L E M E N TA RY DATA
Supplementary data for this paper are available at Journal of Petrology online.

R EF ER ENC ES
Barnes, S. J. (1986). The effect of trapped liquid crystallization on cumulus mineral compositions in layered intrusions. Contributions to Mineralogy and Petrology 93, 524^531. Boorman, S., Boudreau, A. & Kruger, F. J. (2004). The Lower Zone^Critical Zone transition of the Bushveld Complex: a quantitative textural study. Journal of Petrology 45, 1209^1235. Boudreau, A. E. & McBirney, A. R. (1997). The Skaergaard Layered Series. Part III. Non-dynamic layering. Journal of Petrology 38, 1003^1020. Boudreau, A. E. & McCallum, I. S. (1992). Infiltration metasomatism in layered intrusionsAn example from the Stillwater Complex, Montana. Journal of Volcanology and Geothermal Research 52, 171^183. Brooks, C. K. & Gleadow, A. J. W. (1977). A fission-track age for the Skaergaard intrusion and the age of the East Greenland basalts. Geology 5, 539^540. Brooks, C. K. & Nielsen, T. F. D. (1990). A discussion of Hunter and Sparks (Contrib. Mineral. Petrol. 95: 451^461). Contributions to Mineralogy and Petrology 104, 244^247. Chung, H.-Y. & Mungall, J. E. (2009). Physical constraints on the migration of immiscible fluids through partially molten silicates, with special reference to magmatic sulfide ores. Earth and Planetary Science Letters 286, 14^22. Crawford, M. L. & Hollister, L. S. (1977). Evolution of KREEP: further petrologic evidence. Proceedings of the 8th Lunar Scientific Conference. Geochimica et Cosmochimica Acta Supplement 2403^2417. De, A. (1974). Silicate liquid immiscibility in the Deccan Traps and its petrogenetic significance. Geological Society of America Bulletin 85, 471^474. Dixon, S. & Rutherford, M. J. (1979). Plagiogranites as late-stage immiscible liquids in ophiolite and mid-ocean ridge suites: An experimental study. Earth and Planetary Science Letters 45, 45^60. Douglas, J. A. V. (1961). A further petrological and chemical investigation of the Upper Part of the Skaergaard Intrusion, East Greenland. DPhil thesis, Oxford University. Grove, T. L., Walker, D., Longhi, J., Stolper, E. & Hays, J. F. (1973). Petrology of rock 12002 and origin of picritic basalts at Oceanus Procellarum. Geochimica et Cosmochimica Acta 1, 995^1011. Grove, T. L., Baker, M. B. & Kinzler, R. J. (1984). Coupled CaAl^NaSi diffusion in plagioclase feldspar: Experiments and applications to cooling rate speedometry. Geochimica et Cosmochimica Acta 48, 2113^2121. Hanghj, K., Rosing, M. T. & Brooks, C. K. (1995). Evolution of the Skaergaard magma: evidence from crystallized melt inclusions. Contributions to Mineralogy and Petrology 120, 265^269. Haskin, L. A. & Salpas, P. A. (1992). Genesis of compositional characteristics of Stillwater AN-I and AN-II thick anorthosite units. Geochimica et Cosmochimica Acta 56, 1187^1212. Hirschmann, M. M., Renne, P. R. & McBirney, A. R. (1997). 40 Ar/39Ar dating of the Skaergaard intrusion. Earth and Planetary Science Letters 146, 645^658. Holness, M. B. (2005). Spatial constraints on magma chamber replenishment events from textural observations of cumulates: The Rum Layered Intrusion, Scotland. Journal of Petrology 46, 1585^1601. Holness, M. B., Anderson, A. T., Martin, V. M., Maclennan, J., Passmore, E. & Schwindinger, K. (2007a). T extures in partially

solidified crystalline nodules: a window into the pore structure of slowly cooled mafic intrusions. Journal of Petrology 48, 1243^1264. Holness, M.B., Nielsen, T.F.D. & T egner, C. (2007b). T extural maturity of cumulates: A record of chamber filling, liquidus assemblage, cooling rate and large-scale convection in mafic layered intrusions. Journal of Petrology 48, 141^157. Holness, M. B., Stripp, G., Humphreys, M. C. S., Veksler, I. V. & Nielsen, T. F. D. (2011). Silicate liquid immiscibility within the crystal mush: Late-stage magmatic microstructures in the Skaergaard Intrusion, East Greenland. Journal of Petrology (in press). Hoover, J. D. (1989). Petrology of the Marginal Border Series of the Skaergaard Intrusion. Journal of Petrology 30, 399^439. Humphreys, M. C. S. (2009). Chemical evolution of intercumulus liquid, as recorded in plagioclase overgrowth rims from the Skaergaard Intrusion. Journal of Petrology 50, 127^145. Hunter, R. H. & Sparks, R. S. J. (1987). The differentiation of the Skaergaard Intrusion. Contributions to Mineralogy and Petrology 95, 451^461. Irvine, T. N. (1980). Magmatic infiltration metasomatism, doublediffusive fractional crystallization, and adcumulus growth in the Muskox Intrusion and other layered intrusions. In: Hargraves, R. B. (ed.) Physics of Magmatic Processes. Princeton, NJ: Princeton University Press, pp. 325^383. Irvine, T. N., Andersen, J. C. . & Brooks, C. K. (1998). Included blocks (and blocks within blocks) in the Skaergaard intrusion: Geologic relations and the origins of rhythmic modally graded layers. Geological Society of America Bulletin 110, 1398^1447. Jakobsen, J. K., Veksler, I. V., T egner, C. & Brooks, C. K. (2005). Immiscible iron- and silica-rich melts in basalt petrogenesis documented in the Skaergaard intrusion. Geology 33, 885^888. Jakobsen, J. K., Veksler, I. V., T egner, C., Lesher, C. E., Thy, P. & Brooks, C. K. (2006). Evidence for early liquid immiscibility in the Skaergaard intrusion, East Greenland. Geochimica et Cosmochimica Acta 70(S1), A288. Kerr, R. C. & Tait, S. R. (1986). Crystallization and compositional convection in a porous medium with application to layered igneous intrusions. Journal of Geophysical Research 91, 3591^3608. Koepke, J., Feig, S. T. & Snow, J. (2005). Late-stage magmatic evolution of oceanic gabbros as a result of hydrous partial melting: Evidence from the ODP Leg 153 drilling at the Mid-Atlantic Ridge. Geochemistry, Geophysics, Geosystems 6, 1^27. Lange, R. A., Frey, H. M. & Hector, J. (2009). A thermodynamic model for the plagioclase^liquid hygrometer/thermometer. American Mineralogist 94, 494^506. Loferski, P. J. & Arculus, R. J. (1993). Multiphase inclusions in plagioclase from anorthosites in the Stillwater Complex, Montana: implications for the origin of the anorthosites. Contributions to Mineralogy and Petrology 114, 63^78. Maale, S. (1976). The zoned plagioclase of the Skaergaard Intrusion, East Greenland. Journal of Petrology 17, 398^419. Marsh, B. D. (2006). Dynamics of magmatic systems. Elements 2, 287^292. McBirney, A. R. (1975). Differentiation of Skaergaard Intrusion. Nature 253, 691^694. McBirney, A. R. (1985). Further considerations of double-diffusive stratification and layering in the Skaergaard Intrusion. Journal of Petrology 26, 993^1001. McBirney, A. R. (1989). The Skaergaard Layered Series: 1. Structure and average compositions. Journal of Petrology 30, 363^397. McBirney, A. R. (1995). Mechanisms of differentiation in the Skaergaard Intrusion. Journal of the Geological Society, London 152, 421^435.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

172

HUMPHREYS

INTERSTITIAL LIQUID IMMISCIBILITY

McBirney, A. R. (2008). Comments on: Liquid immiscibility and the evolution of basaltic magma, Journal of Petrology 48, 2187^2210. Journal of Petrology 49, 2169^2170. McBirney, A. R. & Nakamura, Y. (1974). Immiscibility in late-stage magmas of the Skaergaard Intrusion. Carnegie Institution of Washington Yearbook 73, 348^352. McBirney, A. R. & Naslund, H. R. (1990). A discussion of Hunter & Sparks. Contributions to Mineralogy and Petrology 104, 235^240. McBirney, A. R. & Noyes, M. N. (1979). Crystallisation and layering of the Skaergaard intrusion. Journal of Petrology 20, 487^554. McKenzie, D. (1984). The generation and compaction of partially molten rock. Journal of Petrology 25, 713^765. Meurer, W. P. & Boudreau, A. E. (1998). Compaction of igneous cumulates Part II: Compaction and the development of igneous foliations. Journal of Geology 106, 293^304. Meurer, W. P. & Claeson, D. T. (2002). Evolution of crystallizing interstitial liquid in an arc-related cumulate determined by LA ICP-MS mapping of a large amphibole oikocryst. Journal of Petrology 43, 607^629. Meurer, W. P. & Meurer, M. E. S. (2006). Using apatite to dispel the trapped liquid concept and to understand the loss of interstitial liquid by compaction in mafic cumulates: an example from the Stillwater Complex, Montana. Contributions to Mineralogy and Petrology 151, 187^201. Morse, S. A. (2008a). Compositional convection trumps silicate liquid immiscibility in layered intrusions: A discussion of Liquid immiscibility and the evolution of basaltic magma by Veksler et al., Journal of Petrology 48, 2187^2210. Journal of Petrology 49, 2157^2168. Morse, S. A. (2008b). Principles of applied experimental igneous petrology: A comment on Experimental constraints on the Skaergaard liquid line of descent by Thy, Lesher, Nielsen, and Brooks, 2006, Lithos 92: 154^180. Lithos 105, 396^400. Morse, S. A. (1986). Convection in aid of adcumulus growth. Journal of Petrology 27, 1183^1214. Morse, S. A. & Nolan, K. M. (1984). Origin of strongly reversed rims on plagioclase in cumulates. Earth and Planetary Science Letters 68, 485^498. Naslund, H. R. (1984). Petrology of the Upper Border Series of the Skaergaard Intrusion. Journal of Petrology 25, 185^212. Nielsen, T. F. D. (1975). Possible mechanism of continental break-up in the North Atlantic. Nature 253, 182^184. Nielsen, T. F. D. (2004). The shape and volume of the Skaergaard intrusion: implications for mass balance and bulk composition. Journal of Petrology 45, 507^530. Nielsen, T. F. D., T egner, C., Thy, P., Fonseca, A. K. M., Jakobsen, J. K., Kristensen, M., Simpson, J. A., Brooks, C. K., Kent, A. J. R., Peate, D. W. & Lesher, C. E. (2000). Retrieval of PLATINOVA drill cores: a new Skaergaard initiative. T ransactions of the American Geophysical Union 81, Fall Meeting Supplement V21E-15. Philpotts, A. R. (1976). Silicate liquid immiscibility: Its probable extent and petrogenetic significance. American Journal of Science 276, 1147^1177. Philpotts, A. R. (1979). Silicate liquid immiscibility in tholeiitic basalts. Journal of Petrology 20, 99^118. Philpotts, A. R. (1981). A model for the generation of massif-type anorthosites. Canadian Mineralogist 19, 233^253. Philpotts, A. R. (1982). Compositions of immiscible liquids in volcanic rocks. Contributions to Mineralogy and Petrology 80, 201^218. Philpotts, A. R. (2008). Comments on: Liquid immiscibility and the evolution of basaltic magma. Journal of Petrology 49, 2171^2175. Philpotts, A. R. & Doyle, C. D. (1983). Effect of magma oxidation state on the extent of silicate liquid immiscibility in a tholeiitic basalt. AmericanJournal of Science 283, 967^986.

Ripley, E. M., Severson, M. J. & Hauck, S. A. (1998). Evidence for sulfide and Fe^Ti^P-rich liquid immiscibility in the Duluth Complex, Minnesota. Economic Geology 93, 1052^1062. Roedder, E. (1951). Low temperature liquid immiscibility in the system K2O^FeO^Al2O3^SiO2. American Mineralogist 36, 282^286. Roedder, E. (1978). Silicate liquid immiscibility in magmas and in the system K2O^FeO^Al2O3^SiO2: an example of serendipity. Geochimica et Cosmochimica Acta 42, 1597^1617. Roelofse, F., Ashwal, L. D., Pineda-Vargas, C. A. & Przybylowicz, W. J. (2009). Enigmatic textures developed along plagioclase^augite grain boundaries at the base of the Main Zone, Northern Limb, Bushveld Complexevidence for late stage melt infiltration into a nearly solidified crystal mush. South African Journal of Geology 112, 39^46. Sisson, T. W. & Grove, T. L. (1993). T emperatures and H2O contents of low-MgO high-alumina basalts. Contributions to Mineralogy and Petrology 113, 167^184. Sparks, R. S. J. (1988). Petrology and geochemistry of the Loch Ba ring-dyke, Mull (N.W. Scotland): an example of the extreme differentiation of tholeiitic magmas. Contributions to Mineralogy and Petrology 100, 446^461. Sparks, R. S. J., Huppert, H. E. & Turner, J. S. (1984). The fluid dynamics of evolving magma chambers. Philosophical Transactions of the Royal Society of London, Series A 310, 511^534. Stripp, G. R. (2009). The late-stage evolution of the Skaergaard Intrusion, East Greenland. PhD thesis, University of Cambridge. Tait, S. R. & Jaupart, C. (1992). Compositional convection in a reactive crystalline mush and melt differentiation. Journal of Geophysical Research 97, 6735^6756. Tait, S. R., Huppert, H. E. & Sparks, R. S. J. (1984). The role of compositional convection in the formation of adcumulate rocks. Lithos 17, 139^146. T egner, C. (1997). Iron in plagioclase as a monitor of the differentiation of the Skaergaard intrusion. Contributions to Mineralogy and Petrology 128, 45^51. T egner, C. & Cawthorn, R. G. (2010). Iron in plagioclase in the Bushveld Complex and the Skaergaard intrusion: implications for iron contents in evolving basic magmas. Contributions to Mineralogy and Petrology 159, 719^730. T egner, C., Thy, P., Holness, M. B., Jakobsen, J. K. & Lesher, C. E. (2009). Differentiation and compaction in the Skaergaard Intrusion. Journal of Petrology 50, 813^840. Thy, P., Lesher, C. E., Nielsen, T. F. D. & Brooks, C. K. (2006). Experimental constraints on the Skaergaard Liquid line of descent. Lithos 92, 154^180. Thy, P., Lesher, C. E. & T egner, C. (2009). The Skaergaard liquid line of descent revisited. Contributions to Mineralogy and Petrology 157, 735^747. T oplis, M. J. & Carroll, M. R. (1995). An experimental study on the influence of oxygen fugacity on Fe^Ti oxide stability, phase relations, and mineral^melt equilibria in ferro-basaltic systems. Journal of Petrology 36, 1137^1170. T oplis, M. J. & Carroll, M. R. (1996). Differentiation of ferrobasaltic magmas under conditions open and closed to oxygen: Implications for the Skaergaard Intrusion and other natural systems. Journal of Petrology 37, 837^858. T oplis, M. J., Brown, W. L. & Pupier, E. (2008). Plagioclase in the Skaergaard intrusion. Part 1: Core and rim compositions in the layered series. Contributions to Mineralogy and Petrology 155, 329^340. Veksler, I.V. (2009). Extreme iron enrichment and liquid immiscibility in mafic intrusions: Experimental evidence revisited. Lithos 111, 72^82.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

173

JOURNAL OF PETROLOGY

VOLUME 52

NUMBER 1

JANUARY 2011

Veksler, I. V., Dorfman, A. M., Danyushevsky, L. V., Jakobsen, J. K. & Dingwell, D. B. (2006). Immiscible silicate liquid partition coefficients: implications for crystal^melt element partitioning and basalt petrogenesis. Contributions to Mineralogy and Petrology 152, 685^702. Veksler, I. V., Dorfman, A. M., Borisov, A. A., Wirth, R. & Dingwell, D. B. (2007). Liquid immiscibility and the evolution of basaltic magma. Journal of Petrology 48, 2187^2210. Veksler, I. V., Dorfman, A. M., Borisov, A. A., Wirth, R. & Dingwell, D. B. (2008). Liquid immiscibility and evolution of

basaltic magma: Reply to S. A. Morse, A. R. McBirney and A. R. Philpotts. Journal of Petrology 49, 2177^2186. Wager, L. R. (1960). The major element variation of the Layered Series of the Skaergaard Intrusion and a re-estimation of the average composition of the Hidden Layered Series and of the successive residual magmas. Journal of Petrology 1, 364^398. Wager, L. R. & Brown, G. M. (1968). Layered Igneous Rocks. Edinburgh: Oliver & Boyd.

Downloaded from http://petrology.oxfordjournals.org/ by guest on February 1, 2013

174

Potrebbero piacerti anche