You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/31408796

Liquid Immiscibility and the Evolution of Basaltic Magma

Article  in  Journal of Petrology · October 2007


DOI: 10.1093/petrology/egm056 · Source: OAI

CITATIONS READS
114 545

5 authors, including:

Ilya V. Veksler Alexander Borisov


Helmholtz-Zentrum Potsdam - Deutsches GeoForschungsZentrum GFZ Russian Academy of Sciences
125 PUBLICATIONS   2,376 CITATIONS    96 PUBLICATIONS   1,608 CITATIONS   

SEE PROFILE SEE PROFILE

Donald B. Dingwell
Ludwig-Maximilians-University of Munich
948 PUBLICATIONS   19,215 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

high pressure high temperature mineral physics View project

Ash aggregation in a variety of volcanic settings View project

All content following this page was uploaded by Ilya V. Veksler on 03 June 2014.

The user has requested enhancement of the downloaded file.


JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 PAGES 2187^2210 2007 doi:10.1093/petrology/egm056

Liquid Immiscibility and the Evolution of


Basaltic Magma

ILYA V. VEKSLER1*, ALEXANDER M. DORFMAN2,


ALEXANDER A. BORISOV3,4, RICHARD WIRTH1 AND
DONALD B. DINGWELL2,5
1
GEOFORSCHUNGSZENTRUM POTSDAM, DEPARTMENT 4.1, TELEGRAFENBERG, 14473 POTSDAM, GERMANY
2
EARTH AND ENVIRONMENTAL SCIENCES, UNIVERSITY OF MUNICH, THERESIENSTRASSE 41, 80333 MUNICH,
GERMANY
3
INSTITUTE OF GEOLOGY AND MINERALOGY, UNIVERSITY OF COLOGNE, ZU«LPICHERSTRASSE, 49 B, 50674 COLOGNE,
GERMANY
4
INSTITUTE OF GEOLOGY OF ORE DEPOSITS, PETROGRAPHY, MINERALOGY AND GEOCHEMISTRY,
RUSSIAN ACADEMY OF SCIENCES, STAROMONETNY 35, 109017 MOSCOW, RUSSIA
5
GEOLOGICAL AND ENVIRONMENTAL SCIENCES, STANFORD UNIVERSITY, STANFORD, CA, 94305-2115 USA

RECEIVED APRIL 2, 2007; ACCEPTED AUGUST 17, 2007


ADVANCE ACCESS PUBLICATION OCTOBER 6, 2007

This experimental study examines relationships between alternative to the Middle Zone. Thus, magma unmixing might be an important
evolution paths of basaltic liquids (the so-called Bowen and Fenner factor in the development of the Fe-enrichment trend documented in
trends), and silicate liquid immiscibility. Synthetic analogues of the cumulates of the Skaergaard Layered Series.
natural immiscible systems exhibited in volcanic glasses and melt
inclusions were used as starting mixtures. Conventional quench KEY WORDS: liquid immiscibility; Skaergaard; layered intrusions;

experiments in 1atm gas mixing furnaces proved unable to reproduce experimental petrology
unmixing of ferrobasaltic melts, yielding instead either turbid,
opalescent glasses, or crystallization of tridymite and pyroxenes. I N T RO D U C T I O N
In contrast, experiments involving in situ high-temperature centrifu- Very few ideas in the history of igneous petrology have
gation at 1000g (g ¼ 98 m/s2) did yield macroscopic unmixing experienced such dramatic turns in acceptability as the
and phase separation. Centrifugation for 3^4 h was insufficient to concept of igneous petrogenesis by liquid immiscibility.
complete phase segregation, and resulted in sub-micron immiscible A popular idea at the dawn of modern petrology (e.g. Daly,
emulsions in quenched glasses. For a model liquid composition 1914), it fell into disrepute in late 1920s (Greig, 1927; Bowen,
of the Middle Zone of the Skaergaard intrusion at super-liquidus 1928), was revived by new evidence in the 1970s and 1990s
temperatures of 1110^11208C, centrifugation produced a thin, silicic (Roedder & Weiblen, 1970, 1971; De, 1974; McBirney &
layer (645 wt% SiO2 and 74 wt% FeO) at the top of the main Nakamura, 1974; Philpotts,1976, 1982; Dixon & Rutherford,
Fe-rich glass (46 wt% SiO2 and 21 wt% FeO). The divergent 1979; Roedder, 1979), and has remained largely ignored
compositions at the top and bottom were shown in a series of static during the last two to three decades. Nevertheless, a modern
runs to crystallize very similar crystal assemblages of plagioclase, re-evaluation of petrogenetic and geochemical aspects
pyroxene, olivine, and Fe^Ti oxides. We infer from these results that of silicate immiscibility has been attempted recently
unmixing of complex aluminosilicate liquids may be seriously kineti- (e.g. Jakobsen et al., 2005; Bogaerts & Schmidt, 2006;
cally hampered (presumably by a nucleation barrier), and thus con- Schmidt et al., 2006; Veksler et al., 2006). Observations
ventional static experiments may not correctly reproduce it. In the reported in some of the recent papers imply that the pheno-
light of our centrifuge experiments, immiscibility in the Skaergaard menon of melt unmixing may hold a key to a number of
intrusion could have started already at the transition from the Lower major, long-standing problems of igneous petrology.

ß The Author 2007. Published by Oxford University Press. All


rights reserved. For Permissions, please e-mail: journals.permissions@
*Corresponding author. E-mail: veksler@gfz-potsdam.de oxfordjournals.org
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

One early recognized and hotly debated problem of temperatures below 10508C, and very advanced stages of
igneous petrology has been the Bowen^Fenner controversy. crystallization. As a consequence, immiscibility has been
The controversy dates back to the classical works by Bowen relegated to a minor role in igneous petrogenesis. Unless it
(1928) and Fenner (1929), and it deals with the general is proven that immiscibility in natural ferrobasaltic melts
chemical trend of evolution of sub-alkaline basaltic can start earlier (i.e. at about 50^60% crystallization of
magma. Simply put, Fenner and his followers since that a common tholeiitic liquid, and temperatures around
time have advocated a liquid differentiation trend towards 1070^11008C), unmixing will remain insignificant in the
strong enrichment in Fe-oxides at constant or slightly Bowen vs Fenner debate.
decreasing silica content. Bowen led proponents of the In this study, we examine the experimental evidence
alternative claim that the content of Fe-oxides in basaltic for and against earlier unmixing of ferrobasaltic liquids.
liquids never achieves the concentrations inferred from We revisit the thermodynamic and phase equilibria con-
the Fenner trend, and the liquid evolution is characterized straints on the Bowen and Fenner types of liquid evolution.
rather by silica enrichment. The trends, which are also We present new results for static and dynamic (centrifuge)
called tholeiitic and calc-alkaline, both exist in nature experiments, which demonstrate that unmixing of multi-
and result in the formation of distinct volcanic rock asso- component basaltic liquids appears to be seriously experi-
ciations and plutonic complexes (e.g. Wager & Deer, 1939; mentally hampered by a nucleation barrier, formation of
Yoder & Tilley, 1962; Carmichael, 1964). The processes colloidal emulsions, and metastable crystallization. Our
behind these opposing trends have been traditionally results show that conventional static experiments on small
explained in terms of fractional crystallization, with melt droplets or rock chips in metal wire loops may be
particular emphasis on (1) the onset of magnetite crystal- unable to correctly reproduce the onset of liquid immisci-
lization, which depends on the oxidation state of magma bility in evolved, Fe-rich basaltic or dacitic liquids.
(e.g. Osborn, 1979), and (2) cotectic proportions of olivine, The onset of macroscopic immiscibility may be pushed
plagioclase and clinopyroxene (Grove & Baker, 1984). to lower temperatures by metastable crystallization,
especially in charges with high initial volume proportion
Others have stressed the effects of melt chemistry;
of crystals (e.g. in experiments on rock chips). In some
for example, the role of alkalis (Irvine, 1976; Shi, 1993).
cases, a second immiscible liquid phase may not nucleate
Silica and Fe-oxides, the definitive components of the
at all, or not grow beyond the sub-micron, colloidal stage,
Bowen and Fenner trends, are also key players in the
and after hours and days it may remain unrecognizable in
occurrence of silicate liquid immiscibility (Roedder, 1951).
final glasses from fine exsolutions formed during quench-
During the last surge of interest in silicate melt unmixing
ing. High-temperature in situ centrifugation is, however,
20^30 years ago, it was established that certain ferrobasal-
capable of separating immiscible emulsions, and thus
tic liquids were unstable, unmixing to form two contrast-
revealing stable, super-liquidus immiscibility even when
ing liquid compositions, one with a SiO2 content of less
the result is a very fine emulsion. In light of our centrifuge
than 45 wt% and very high total iron [FeO(t) ¼
experiments, we propose that unmixing of natural ferroba-
FeO þ Fe2O3] at 25^30 wt% or more, and the other with
saltic liquids may in fact start at a much earlier stage of
SiO2 contents at or exceeding 55^60 wt%, and FeO(t)
crystallization than was previously thought. Our starting
at about 5^10 wt% (e.g. Philpotts, 1976, 1979, 1982). mixtures were based on compositions of natural immisci-
Observations both in natural rocks and in laboratory ble glasses, and so we begin with a brief overview of those
experiments have confirmed the stable coexistence of cer- natural melt inclusions that appear to record the early
tain silicic and Fe-rich liquids with gabbroic mineral stages of unmixing. In conclusion, we examine the petro-
assemblages (Roedder & Weiblen, 1970; De, 1974; genetic implications for tholeiitic magmas, mostly using
McBirney & Nakamura, 1974; Dixon & Rutherford, 1979; the classical example of the Skaergaard intrusion, East
Philpotts, 1979; Roedder, 1979; Longhi, 1990). Silicate Greenland, and propose that the existence of the divergent
liquid immiscibility in ferrobasaltic melts results in the for- trends primarily stems from the strong non-ideality of
mation of (1) extremely Fe-rich, and (2) silicic conjugate ferrobasaltic and ferrodacitic liquids.
liquids. Thus, it has the potential to place the Bowen vs
Fenner debate on a new basis, possibly linking the diver-
gent liquid evolution trends. In the presence of immiscibil- S I L I C AT E L I Q U I D
ity, the trends of FeO(t) and silica enrichment do not I M M I S C I B I L I T Y I N M AG M A S
appear as mutually exclusive alternatives, but rather as
A N D SELECTION OF
complementary evolution paths of immiscible conjugate
liquids in a heterogeneous magma. In this debate experi- E X P E R I M E N TA L C O M P O S I T I O N S
mental evidence is crucial. It should be pointed out that Volcanic glasses and melt inclusions
reliable experimental demonstrations of immiscibility Droplet exsolution textures in natural volcanic glasses have
in basaltic compositions have been restricted so far to been traditionally regarded as the strongest evidence

2188
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

Table 1: Examples of natural immiscible liquids (quenched glasses) in mesostasis of basaltic lavas, and in melt inclusions
(wt% oxides)

Groundmass glasses Melt inclusions in plagioclase Inclusions in native iron

Lfe1 Lsi1 Lfe2 Lsi2 Lfe2 Lsi2 Lfe3 Lsi3 Lfe4 Lsi4

SiO2 415 733 528 718 494 797 4847 6842 3933 7218
TiO2 58 08 12 07 40 06 433 155 282 060
Al2O3 37 121 74 102 34 106 285 81 752 1338
FeO 310 32 193 40 226 35 2145 726 1978 410
MnO 05 00 02 00 04 00 06 006 160 026
MgO 09 00 83 11 96 06 66 192 628 106
CaO 94 18 90 65 100 12 968 284 1423 290
Na2O 08 31 10 22 00 22 19 233 060 075
K2O 07 33 08 35 04 16 093 284 021 140
P2O5 35 007 00 00 02 00 10 00 753 037
Total 978 9767 1000 1000 1000 1000 9781 9532 9990 9701

References: 1Philpotts (1982); 2Philpotts (1981); 3Krasov & Clocchiatti (1979); 4Ryabov (1989). Lfe, Fe-rich glass; Lsi,
silicic glass.

supporting silicate liquid immiscibility. Classical examples inclusions could be a consequence of arrested crystalliza-
were presented by Philpotts (1982), who carried out tion of clinopyroxene in small inclusion cavities because of
a detailed study of the mesostasis of basalts and andesites the absence of suitable sites for nucleation. Such an expla-
from numerous localities worldwide, and described glassy, nation is, however, difficult to apply to the case studied by
or partly crystallized spherical droplets dispersed in Krasov & Clocchiatti (1979). Those workers attempted to
a second glass of a different composition. The compositions homogenize inclusions in An70 plagioclase phenocrysts
of the droplets and matrix glasses were shown to compare from subduction-related andesite^dacitic lava of the
well with experimental Fe-rich and silica-rich immiscible Karymsky volcano, Kamchatka, Russia, and observed
conjugate melts, consistent with an origin by liquid immis- two conjugate silicate liquids at temperatures of up to
cibility. The average compositions of the coexisting phases 1250^12808C. Importantly, some of the immiscible inclu-
in the mesostasis of tholeiitic lavas according to Philpotts sions had been originally crystallized to fine-grained
(1982) are listed in the first two columns of Table 1. Both daughter crystal assemblages, and immiscibility was
phases are notably depleted in MgO, and imply equilibra- shown to be reversible in heating^cooling experiments
tion at low temperature. Experiments on similar basaltic around 12808C. The compositions of the quenched immis-
compositions (Roedder & Weiblen, 1970; McBirney & cible glass phases reported by Krasov & Clocchiatti (1979)
Nakamura, 1974; Dixon & Rutherford, 1979; Philpotts, are also presented in Table 1. The glasses are characterized
1979; Philpotts & Doyle, 1983; Longhi, 1990), confirmed by elevated MgO contents, and a lower contrast in
the late-stage, low-temperature origin of the groundmass SiO2 and Fe-oxide contents. These features are consistent
phase assemblages forming the immiscible globules. with the equilibration of immiscible liquids at higher
According to those experiments, immiscibility in terres- temperatures.
trial basaltic lavas probably occurs below 10208C. Perhaps the most convincing example of high-
Significantly less evolved, more refractory compositions temperature silicate liquid immiscibility is provided by
have been reported, however, in glassy melt inclusions melt inclusions in native iron, which, although exotic in
hosted by olivine and plagioclase phenocrysts (Roedder & terrestrial environments, do occur in a few basaltic sills
Weiblen, 1970; Krasov & Clocchiatti, 1979; Roedder, 1979; and lava flows. Grains, nuggets, and large blocks of native
Fujii et al., 1980; Philpotts, 1981, 1982; Shearer et al., 2001). iron and minor cohenite (Fe3C) have been reported
Philpotts (1981, 1982) argued that the more primitive glass in basaltic lavas on Disco island, West Greenland (Bird
compositions in plagioclase-hosted inclusions may be due et al., 1981), and in the roof endocontact zone of the
to metastable unmixing. He proposed that early exsolution Khutungunsky sill of the Siberian trap province (Ryabov,
of pyroxenitic, Fe-rich liquid in plagioclase-hosted melt 1988, 1989). The occurrences of native iron are believed to

2189
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

result from reduction of basaltic magma by carbon at Two of our starting mixtures were based on the average
the contact with organic-bearing sediments. Some iron compositions of the immiscible glasses of the iron-hosted
nuggets contain numerous inclusions of rock fragments, melt inclusions. The compositions RY40 and RY20
single silicate crystals, and spherical glass droplets showing (Table 2) were calculated as mixtures between the average
clear, classical immiscibility textures, as can be seen in Fe-rich (Lfe) and silica-rich (Lsi) iron-hosted glasses
Fig. 1. The photograph shows that the Fe-rich and silica- published by Ryabov (1989) (Table 1) in mass proportions
rich glasses are very distinct, and, in those inclusions of 3:2 and 4:1. In other words, the composition RY40 upon
where low-temperature devitrification is minor, the inter- complete melting should theoretically yield 40 wt% of
faces between the glasses are razor-sharp. Representative the conjugate liquid Lsi, and 60 wt% Lfe, whereas the
electron microprobe data on the glasses in the Siberian composition RY20 should give 20 and 80 wt% respectively
samples from Ryabov (1989) are also presented in Table 1. of the same liquids.
Very similar compositions have been reported from Disco Three other starting mixtures were designed in a similar
island (Bird et al., 1981). The elevated MgO contents of the way from compositions of other immiscible glasses listed in
glasses are similar to those of the plagioclase-hosted Table 1. The composition SF-1 is based on the analyses of
melt inclusions studied by Krasov & Clocchiatti (1979; low-temperature immiscible groundmass globules and
see also Table 1), and imply a high-temperature origin. glasses published by Philpotts (1982), and compiled in the
two leftmost columns of Table 1. Accordingly, the composi-
tion has the lowest MgO content among the mixtures used
in this study. The mixture SF-1 has been used in previous
Lfe
centrifuge experiments (Veksler et al., 2006), where it
Lfe produced roughly equal proportions of Lfe and Lsi immis-
Lsi cible phases. The composition KC is based on electron
microprobe analyses of the Fe-rich and silica-rich glasses
of the plagioclase-hosted melt inclusions published by
Krasov & Clocchiatti (1979). The mass proportion of the
Lfe and Lsi (see the analyses in Table 1) is 7:3. The mixture
Fe FJ is based on electron microprobe analyses of immiscible
glassy inclusions in plagioclase phenocrysts from an
olivine-bearing andesite of the Akita-Komagatake volcano,
0.5 mm Japan, which were published by Fujii et al. (1980). The anal-
yses were carried out using defocused beam, and the compo-
Fig. 1. Natural inclusions of immiscible Fe-rich (Lfe) and silica-rich sition FJ represents an average mixture of the Fe-rich and
(Lsi) glasses in native iron, Siberian Traps; reflected light. The Fe-rich silica-rich immiscible liquids in unidentified proportions.
glass forms a rim around the silica-rich glass, and there are also a few
smaller spherical droplets of Lfe inside the Lsi. Glass compositions Notably, the MgO content in the bulk of the inclusions is
reported by Ryabov (1989) are given in Table 1. the highest ever reported among immiscible liquids.

Table 2: Compositions of synthetic starting mixtures normalized to 100 wt%

SF-1 KC FJ FJ20 RY40 RY20 MZ-1 MZ-2 MZF MZS

SiO2 566 561 4994 5545 5313 4522 500 480 4566 6456
TiO2 20 36 201 161 196 296 51 49 541 180
Al2O3 60 456 339 529 999 96 100 96 978 1325
FeO 240 1816 2316 1853 1477 169 172 2051 210 739
MgO 14 535 961 768 424 548 40 384 415 232
CaO 60 786 1027 822 982 1148 100 960 1065 634
Na2O 05 209 081 150 067 138 28 269 244 281
K2O 20 155 032 134 069 058 06 058 048 137
P2O5 00 072 049 039 473 64 03 029 043 015
Total 985 1000 1000 1000 1000 1000 1000 1000 1000 1000


The remaining 15 wt% comprises a mixture of more than 30 trace elements [see Veksler et al. (2006) for details].

2190
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

By their very nature, plagioclase-hosted melt inclusions The disparate results of these reconstructions are
must be saturated in their host plagioclase. The composi- illustrated in Fig. 2. In general, one may conclude that the
tions KC and FJ are, however, surprisingly low in alumina, disagreement among researchers is the strongest when it
and, as we learned from our experiments on these mixtures comes to the late stages of magma evolution, especially
(see below), do not produce plagioclase on the liquidus. after the onset of ilmenite and magnetite crystallization.
To compensate for possible post-entrapment modification The disagreement between the models is, however, not so
as a result of precipitation of plagioclase on inclusion dramatic regarding the starting composition of the
walls [see Veksler (2006) for discussion], we prepared Skaergaard parental magma (see also Nielsen, 2004), and
another mixture FJ20, which has been derived from the the models also agree reasonably well at the early stages of
composition FJ by the addition of 20 wt% of pure magma evolution, up to about 50^60% crystallization.
anorthite. This stage corresponds roughly to the transition from the
Lower to the Middle Zone of the Layered Series marked
Liquids of the Skaergaard intrusion by the disappearance of olivine from the liquidus mineral
Since the study by Wager & Deer (1939), the Skaergaard association and the appearance of liquidus pigeonite
intrusion of layered gabbros has been the prime example [see Wager & Brown (1968) and McBirney (1989) for
of the Fenner, tholeiitic, or Fe-enrichment trend. Thus, a detailed description of the Skaergaard stratigraphy].
Skaergaard has remained for decades at the centre of the The liquid composition at this stage has been well con-
Bowen^Fenner controversy. The intrusion is located in strained by detailed experimental studies (McBirney &
East Greenland, and after many years of comprehensive Naslund, 1990; Toplis & Carroll, 1995; Thy et al., 2006),
research was once termed ‘the most intensely studied body thermodynamic modelling (Ariskin, 2002), and mass-
of rock on Earth’ (McBirney & Naslund, 1990). Numerous balance calculations (Nielsen, 2004). Our starting mix-
attempts have been made to reconstruct liquid lines of des- tures MZ-1 and MZ-2 (Table 2) are based on estimates of
cent in the intrusion by: (1) mass-balance calculations the liquid composition at the top of the Lower Zone.
(Wager & Brown, 1968; Hunter & Sparks, 1987; Nielsen, The composition MZ-1 is very similar to the glasses pro-
2004); (2) experimental simulation of crystallization or duced in experiments by Toplis & Carroll (1995) at tem-
partial melting of Skaergaard rocks (McBirney & peratures close to 11008C. The composition MZ-2 differs
Naslund, 1990; Toplis & Carroll, 1995; Thy et al., 2006); from MZ-1 only by its somewhat higher FeO content,
(3) tracing specific elements using mineral^melt partition which brings it closer to the mass-balance estimation of
coefficients (Tegner, 1997; Jang et al., 2001); (4) geochemical the Middle Zone liquid by Nielsen (2004). The remaining
thermometry of cumulus mineral assemblages (Ariskin, two starting mixtures in Table 2, the Fe-rich composition
1999, 2002); (5) studies of crystallized melt inclusions in MZF and silica-rich composition MZS, are explained
cumulus minerals (Hanghj et al., 1995; Jakobsen et al., below, where we describe the results of our centrifuge
2005, 2007); (6) comparisons with compositions of coeval experiments.
and spatially associated dykes (Brooks & Nielsen,1978,1990).

E X P E R I M E N TA L A N D
Toplis & Carroll 1995 A N A LY T I C A L M E T H O D S
30 Wager & Brown 1968 Preparation of starting mixtures
Hunter & Sparks 1987 All the starting mixtures listed in Table 2 were synthesized
McBirney & Naslund 1990
by fusion in Pt crucibles of carefully weighed and mixed
FeO, wt. %

20
reagent-grade chemicals [SiO2, Al2O3, MgO, TiO2,
CaCO3, Na2CO3, K2CO3 and Ca3(PO4)2]. To avoid
strong oxidation and Fe losses to the Pt crucibles, Fe-free
glasses were first prepared by repeated fusions in an
10 electric furnace once at 9008C, and then twice at 13508C.
The Fe-free glasses were then crushed to a fine-grained
powder (grain size less than 3 mm) and mixed and
ground with reagent-grade FeO in an agate mortar under
0 acetone. Fine powders of the glass^FeO mixtures were used
40 50 60 70 80
SiO2, wt. %
for the experimental charges.

Fig. 2. Liquid lines of descent in the Skaergaard intrusion in terms of Static experiments
total iron (FeO) vs silica (SiO2) variations. Average compositions of
immiscible Fe-rich and Si-rich melt inclusions in apatite reported by Static experiments were conducted in a 1atm vertical tube
Jakobsen et al. (2005) are shown by filled and open stars. furnace in which fO2 was controlled by CO^CO2

2191
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

gas mixtures. A PtRh6^PtRh30 (type B) thermocouple instrument in wavelength-dispersive spectrometry (WDS)


was employed for temperature measurements. The uncer- mode at 15 kVaccelerating voltage and 15 nA beam current
tainties of cited log fO2 values and temperature do not with spot sizes ranging from 1 to 3 mm depending on the
exceed 02 and 28C, respectively. In most cases, silicate properties and compositions of the analyzed materials.
melts were suspended from Re loops, prepared from com- Counting time for all the elements was set to 20 s on peak
mercially available Re ribbon (Rhenium Alloys, Inc.). and 10 s on background. The following synthetic and nat-
Rhenium has been shown to be an excellent loop material, ural standards were used for the calibration: orthoclase
preventing iron loss from a charge even at very reducing (Al and K), rutile (Ti), wollastonite (Si and Ca), albite
conditions (Borisov & Jones, 1999). The samples were and jadeite (Na), apatite (P), hematite (Fe), diopside and
either (1) directly placed in a pre-heated furnace at the periclase (Mg).
required temperature or (2) first over-heated by 50^1508C
and then cooled to run temperature at a rate of 10^508C Transmission electron microscopy (TEM)
per hour. After sufficient dwell exposure at constant TEM was carried out using a FEI TecnaiTMG2 F20
temperature, the samples were quenched in air within a X-Twin electron microscope operated at 200 kV. Focused
few seconds. ion beam milling (FIB) was applied to cut site-specific
TEM-foils from the area of interest. FIB preparation was
Centrifuge experiments
performed under ultrahigh vacuum conditions in an
The experimental equipment and Fe^Pt double containers oil-free system FEI FIB200TEM. TEM-ready foils with
for the centrifuge experiments have been described in
dimensions 20 mm  10 mm  01 mm were cut directly
detail previously (e.g. Dorfman et al., 1996; Veksler et al.,
from microprobe epoxy mounts using a gallium-ion beam.
2006). Briefly, the equipment represents a small cylindrical
The TEM foils were placed on a perforated carbon film on
wire-wound electric furnace mounted on a centrifuge,
a copper grid. Carbon coating to prevent charging of the
which provides acceleration of up to 1000g (g ¼ 98 m/s2)
TEM samples was not applied.
over maximal run durations of 5 h. Temperature is con-
trolled by three Pt^PtRh10 (type S) thermocouples at the
top, bottom and the middle of the container, respectively.
R E S U LT S
Comparisons between the thermocouple readings imply
that ‘top to bottom’ temperature gradients in the samples Static experiments on synthetic analogues
did not exceed 28C. Inner containers made of ‘iron’ of natural immiscible glasses
(more precisely, low-carbon steel with C concentration of The results of the static experiments are summarized in
less than 06 wt%), wrapped and sealed inside platinum Tables 3 and 4. The most important result in the context
foil (Veksler et al., 2006) were used in most of the runs. of this study is that none of the products of the conven-
However, a few runs were carried out in nickel or graphite tional static runs showed unambiguous, macroscopic signs
containers. At the start of centrifuge runs, starting mix- of liquid immiscibility. The composition SF-1 that had
tures were first completely melted and homogenized at repeatedly unmixed at almost identical T and fO2 condi-
high, super-liquidus temperatures and the slowest rotation tions in our previous centrifuge runs (Veksler et al., 2006)
speed (50g) for at least 30 min, but usually for 1h or more. quenched in the loops to beads of turbid, milky glass, occa-
The purpose of the rotation at this stage was to prevent sionally with trace amounts of small (3^10 mm), dendritic
melts from creeping up the container walls by capillary tridymite crystals. Over-heating to 14008C (run L-34) did
forces. Complete melting and homogenization was checked not result in any visible changes in the optical appearance
by a few test runs in which containers were opened and the of the glass. TEM images of foils cut from one of the
glasses were examined after the heating at slow rotation. turbid glasses (Fig. 3b) revealed sub-micron heterogeneity
The main centrifugation stage at 1000g and experimental comprising silica- and Fe-rich amorphous phases.
temperature followed after the homogenization and lasted Compositional TEM scans and profiles showed sharp
for a few hours. Quenching was performed in air by turn- co-variations in Fe, Ca and K at the sub-micron scale
ing off the heating, while the centrifuge rotation continued with areas enriched in Fe and Ca being depleted in K,
at a constant speed. During quenching, temperature inside but a more uniform distribution of Al and Si. The Fe-rich
the sample dropped from 1200 to 11008C in 10 s, from 1100 and Si-rich phases form 3D sponge-like interconnected
to 10008C in 15 s, and further cooling of the sample to textures that are usually interpreted as products of spino-
8008C took about 15^2 min. dal phase separation (Vogel, 1985; Shelby, 2005). Thus, it
appears that although the composition SF-1 did not pro-
Electron microprobe analyses duce macroscopic immiscible liquid (glass) segregations
Run products were mounted in epoxy, ground and in static runs, it falls within the spinodal region of the
polished and studied by electron microprobe at the liquid miscibility gap, and consequently cannot be
GeoForschungsZentrum Potsdam using a Cameca SX50 quenched to a homogeneous glass.

2192
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

Table 3: Conditions of static runs and phase composition of the run products

Run no. Mixture Over-heating, T (8C) log fO2 Duration Run products
T (8C) (h)

L-1 SF-1 1200 1145 –128 32 L þ trace tr


L-4 SF-1 1200 1123 –117 63 L þ trace tr
L-5 FJ 1200 1123 –117 63 L þ pig þ aug
L-6 SF-1 1200 1123 –106 615 L þ trace tr
L-7 FJ 1200 1123 –106 615 L þ pig þ aug
L-8 SF-1 1240 1203 –106 235 L
L-9 FJ 1240 1203 –106 235 L þ pig
L-10 SF-1 none 1156 –112 48 L þ trace tr
L-11 FJ none 1156 –112 48 L þ pig þ aug
L-13 KC none 1200 –106 155 L
L-14 FJ20 none 1200 –106 155 L þ pig
L-15 RY40 none 1200 –106 155 L þ tr
L-16 KC none 1249 –100 9 L
L-17 FJ20 none 1249 –100 9 L
L-18 FJ none 1249 –100 9 L
L-19 RY40 none 1249 –100 9 L þ tr
L-20 KC 1200 1151 –113 14 L þ pig
L-21 RY40 1200 1151 –113 14 L þ tr
L-22 RY40 none 1280 –96 6 L þ tr
L-23 FJ20 none 1280 –96 6 L
L-24 RY40 none 1200 –78 16 L þ tr
L-25 FJ20 none 1200 –78 16 L þ pig
L-26 KC none 1200 –78 16 L
L-27 RY40 none 1169 –111 17 L þ tr
L-29 KC none 1169 –111 17 L þ pig þ tr
L-30 RY40 none 1319 –92 5 L þ trace tr
L-31 RY40 1200 1098 –121 355 L þ tr
L-33 KC 1200 1098 –121 355 L þ pig þ aug þ tr
L-34 SF-1 none 1400 –83 15 L
L-35 SF-1 none 1153 –113 18 L þ trace tr
L-36 SF-1 none 1153 –126 265 L þ trace tr
L-62 MZF none 1081 –112 27 L þ aug þ plag þ ilm
L-63 MZS none 1081 –112 27 L þ aug þ plag þ tr þ mt
L-64 MZF 1080 1022 –122 96 L þ aug þ ol þ plag þ ilm þ mt
L-65 MZS 1080 1022 –122 96 L þ aug þ ol þ plag þ tr þ mt
L-66 MZF none 1062 –116 26 L þ aug þ plag þ ilm þ mt
L-67 MZS none 1062 –116 26 L þ aug þ plag þ tr þ mt
L-69 MZF none 1043 –119 64 L þ aug þ ol þ plag þ ilm þ mt
L-70 MZS none 1043 –119 64 L þ aug þ ol þ plag þ mt

Static experiments on mixing between MZF and MZS


L-51 MZF–MZS 1350 1113 –115 27 Lfe þ Lsi
L-52 MZF–MZS 1350 1113 –115 27 Lfe þ Lsi
L-56 MZF–MZS 1350 1114 –115 75 Lfe þ Lsi
L-57 MZF–MZS 1350 1114 –115 75 L
L-58 MZF–MZS 1350 1090 –119 64 Lfe þ Lsi þ aug
S-141 MZF–MZS 1270 1110 5IW 72 Lfe þ Lsi þ ol þ aug

IW, iron–wüstite buffer. Crystal phases: aug, augite; ilm, ilmenite; mt, magnetite; ol, olivine; pig, pigeonite; plag,
plagioclase; tr, tridymite.

2193
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

Table 4: Electron microprobe analyses of glasses and crystal phases in products of static loop experiments

Run Mixture T (8C) Phase SiO2 TiO2 Al2O3 MgO FeO CaO Na2O K2O P2O5 Total

L-9 FJ 1203 glass 5092 231 366 764 2306 1085 086 029 049 10008
pig 5308 023 015 2357 1893 356 003 000 001 9956
L-11 FJ 1156 glass 4917 255 425 596 2511 981 232 033 052 10000
pig 5164 033 019 1825 2138 802 010 001 003 9995
aug 5299 002 000 2045 058 2441 001 001 002 9850
L-5 FJ 1123 glass 4811 323 543 423 2435 975 137 042 063 9752
pig 5108 0329 021 1954 1865 825 004 003 00 9814
aug 5072 053 040 1464 1727 1491 010 0 00 9857
L-25 FJ20 1200 glass 5564 179 534 688 1838 874 153 119 035 9982
pig 5511 015 020 2695 1520 315 004 000 001 10082
L-29 KC 1169 glass 5736 357 439 541 1821 817 179 137 036 10065
pig 5524 116 121 1799 1975 530 045 036 010 10154
L-33 KC 1098 glass 5822 435 555 266 1921 662 178 177 042 10060
pig 5362 162 163 1359 2184 726 048 049 012 10064
aug 5173 096 024 1413 1928 1436 011 001 002 10085
L-30 RY40 1319 glass 5305 207 1002 433 1525 991 049 039 462 10013
tr 9865 027 038 002 049 013 009 005 001 10007
L-22 RY40 1280 glass 5200 207 1038 431 1497 1001 060 050 488 9974
tr 9738 001 002 000 026 002 001 001 000 9771
L-19 RY40 1250 glass 4910 215 1060 450 1558 1044 071 059 549 9915
tr 9845 030 060 000 044 011 012 011 002 10015
L-24 RY40 1200 glass 4863 220 1067 455 1548 1053 072 065 581 9923
tr 9818 031 068 003 047 014 011 011 002 10004
L-27 RY40 1169 glass 4845 225 1086 470 1611 1065 068 065 587 10021
tr 9821 014 070 018 086 050 008 005 018 10088
L-21 RY40 1150 glass 4795 233 1108 454 1572 1069 073 070 621 9993
tr 9892 003 000 000 030 001 003 000 000 9929
L-31 RY40 1098 glass 4680 226 1118 466 1687 1081 075 070 624 10027
tr 9804 030 082 001 045 017 019 018 004 10019
L-62 MZF 1081 glass 4458 594 977 360 2166 1006 249 051 046 9908
cpx 4924 243 308 1127 1384 1943 043 006 006 9983
plag 5334 048 2754 015 133 1154 487 021 003 9949
L-66 MZF 1062 glass 4702 453 954 256 2256 897 262 078 082 9940
cpx 4854 277 484 735 1864 1544 115 032 031 9936
plag 5494 070 2577 035 260 1027 551 031 009 10053
ilm 007 5040 012 265 4383 043 000 001 000 9752
mt 013 2816 194 219 6537 023 001 001 001 9804
L-69 MZF 1043 glass 5022 368 970 203 2151 821 259 108 116 10018
ol 3245 061 005 1293 5352 062 001 000 008 10026
cpx 5000 172 277 886 1822 1779 068 017 018 10039
plag 5721 051 2495 022 248 885 613 050 008 10093
mt 014 2648 232 152 6820 015 001 001 001 9885
L-64 MZF 1022 glass 5282 267 1029 141 1919 744 245 140 173 9939
ol 3180 046 004 1066 5613 062 002 001 013 9987
cpx 4865 185 249 918 2078 1696 032 001 003 10028
plag-1 5359 044 2798 014 174 1169 473 032 009 10072

(continued)

2194
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

Table 4: Continued

Run Mixture T (8C) Phase SiO2 TiO2 Al2O3 MgO FeO CaO Na2O K2O P2O5 Total

plag-2 4316 107 3519 027 090 1911 064 002 013 10047
mt 028 2740 176 123 6765 024 002 001 001 9859
L-63, MZS 1081 glass 6751 219 1247 180 623 516 220 175 029 9960
cpx 5082 071 253 1442 1576 1555 021 003 004 10006
plag 4937 057 3116 061 132 1349 282 019 006 9959
L-67 MZS 1062 glass 6830 185 1187 150 683 471 230 197 024 9956
cpx 4962 054 145 1289 1939 1550 015 002 002 9956
plag 5449 010 2728 033 101 1183 430 019 002 9955
tr 9644 023 103 022 037 037 008 012 005 9891
L-70, MZS 1043 glass 7244 169 1194 080 453 294 250 260 030 9814
ol 3288 086 027 1382 5010 054 001 002 025 9874
cpx 5063 059 196 1094 2263 1311 023 017 005 10030
plag 6271 088 2041 046 255 738 358 124 015 9936
mt 057 657 464 169 7963 014 000 002 001 9326
L-65 MZS 1022 glass 7057 116 1309 069 544 367 300 264 043 10068
ol 3218 012 011 1235 5417 061 000 002 017 9973
cpx 5104 066 189 989 2310 1174 022 018 003 9874
plag 6309 070 1993 051 330 697 372 153 014 9989
mt 033 725 425 091 8181 008 000 000 000 9464
tr 9883 017 101 001 022 007 027 004 004 10066
L-51 MZF/S 1113 Fe-glass 5380 382 1155 365 1436 873 247 079 030 9945
Si-glass 6279 216 1274 272 892 640 237 137 018 9965
L-52 MZF/S 1113 Fe-glass 5431 382 1151 364 1396 863 244 079 029 9938
Si-glass 6319 199 1317 281 841 648 244 139 016 10004
L-56 MZF/S 1114 Fe-glass 5451 367 1172 329 1368 828 241 086 029 9871
Si-glass 6492 197 1250 237 836 577 215 154 015 9975
L-57 MZF/S 1114 glass 5358 395 1151 341 1458 863 251 082 027 9926
L-58 MZF/S 1090 Fe-glass 5431 385 1190 333 1388 846 249 082 030 9934
Si-glass 6518 189 1321 202 776 557 217 159 016 9955
cpx 5230 111 158 1671 1291 1498 013 002 003 9977
S-141 MZF/S 1110 Fe-glass 4537 508 1019 392 2025 990 257 048 050 9826
Si-glass 6161 220 1238 210 1039 586 229 144 021 9848
ol 3478 029 015 2314 4045 070 006 001 010 9968
cpx 5178 159 153 1503 1005 1885 029 002 002 9914

Oxide concentrations are in wt%.

The composition FJ quenched at 12498C to clear, retained very high FeO(t) contents of about 24^25 wt%,
dark brown, optically homogeneous glass. At 12038C, but no signs of liquid unmixing have been observed in the
it produced a similar glass with a few elongated, euhedral run products. The liquid clearly remained under-saturated
crystals of low-Ca pyroxene corresponding to pigeonite in plagioclase down to at least 11238C. The addition of
(sample L-9; see analyses in Table 4). As the temperature 20 wt% of plagioclase components (mixture FJ20) also
was brought further down to 1156 and 11238C, the size and failed to stabilize plagioclase. Pigeonite remained the only
the bulk volume proportion of the pyroxene crystals liquidus phase in the samples L-14 and L-25 produced from
increased. A strong compositional zoning developed in the mixture FJ20 at 12008C, and the mixture completely
the crystals, with low-Ca, pigeonitic compositions in the melted at and quenched to clear, optically homogeneous
cores, and progressively Ca-rich, augitic compositions at brownish glass from 1249 and 12808C (samples L-17
the rims (samples L-11 and L-5). Glass compositions and L-23).

2195
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

(a) glasses occur as inclusions in plagioclase, but it is not


entirely unexpected in view of the low Al2O3 contents in
the bulk compositions as well as quench partial liquids.
As noted above, the original natural plagioclase-hosted
melt inclusions may have been modified by post-
entrapment diffusion and crystallization of host plagioclase
on inclusion walls (see also Veksler, 2006).
Eight static experiments on the mixture RY40 at
temperatures 1098^13198C and log fO2 ranging from ^
78 to 121 all produced tridymite crystals suspended in
clear, dark brown glass, and no signs of liquid immiscibil-
ity. Unlike the sample L-33 (see above), where tridymite
formed tiny, micron-sized needles, and small (2^10 mm)
1 mm elongated euhedral crystals, in the products of mixture
RY40 it mostly formed strange-looking oval blebs or crys-
tal aggregates measuring up to 04 mm in diameter. The
(b) bulk mass fraction of tridymite in the samples varied from
2% at 13198C (sample L-30) to 13% at 10988C (sample
L-31), and tridymite showed no visible resorption at run
durations of up to 35 h. Crystallization of tridymite as a
single liquidus phase over such a broad temperature inter-
val from a melt with SiO2 content of only 533 wt% is
remarkable, and is discussed in more detail below.
The absence of liquid immiscibility is no less remarkable,
as the failure of melt droplets to unmix starkly contrasts
with spectacular immiscibility textures of the natural
iron-hosted melt inclusions, on which the mixture RY40
was based (Fig. 1).

Centrifuge experiments
200 nm
Optical and electron microscope images of centrifuge
run products are presented in Figs 4^8, and experimental
Fig. 3. Products of a static experiment at atmospheric pressure, conditions and results are summarized in Tables 5 and 6.
11238C and log fO2 ¼ ^106 on starting composition SF-1 (sample As mentioned above, the mixture SF-1 has previously
L-6). The charge quenched to a bead of turbid glass. (a) General been studied and described in detail by Veksler et al.
view of the polished epoxy mount in reflected light. The location of
the fragment studied by TEM is indicated by an arrow. (b) TEM (2006). In this study, we present the results of centrifuge
image showing sub-micron heterogeneity in the glass fragment. experiments on new starting compositions KC, FJ, FJ20,
The high-absorption, Fe-rich amorphous phase is darker. The lightest RY20, RY40, MZ, and MZ-2.
areas are those where the low-absorption silica-rich phase goes
through the whole thickness of the foil.
Our first pilot centrifuge run on starting mixture KC
lasted for a total duration of 90 min (sample C-101 in
Table 5), and was probably not long enough to equilibrate
the charge. Olivine, which was never observed in static
Static experiments on starting mixture KC produced runs, formed at the bottom of the container, whereas tridy-
homogeneous, very dark brown glass at 1200 and 12498C, mite crystallized at the top in a turbid, brownish glass.
and liquidus assemblages comprising low-Ca pyroxene at Glass at the bottom, just above the olivine crystals,
1169 and 10988C (samples L-29 and L-33). The pyroxene was darker and more transparent than at the top. Vertical
crystals are similar to those produced in the experiments electron microprobe (EMP) profiles of the glass over the
on the mixtures FJ and FJ20, and they also showed a height of the sample revealed strong, semi-linear composi-
strong compositional zoning from pigeonitic cores to sub- tional gradients with top-to-bottom variations in SiO2
calcic and high-Ca rims (Table 4). A few smaller tridymite content from 726 to 455 wt%. However, when the total
crystals were found in sample L-33 quenched at 10988C. duration of the centrifuge runs was increased to 180 min,
Plagioclase did not crystallize, and no signs of liquid and initial heating to 11808C was performed at a slower
unmixing were observed in the samples. The absence of rotation speed, the mixture KC quenched to a uniformly
plagioclase from the products of mixtures FJ, FJ20 and turbid, non-transparent brownish glass, which showed no
KC is not consistent with the fact that the original natural detectable compositional gradients in vertical electron

2196
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

(a) (b) to clear glass, and small amounts of low-Ca pyroxene.


Thus, for mixtures FJ and FJ20 the results of centrifuged
cpx
runs are generally in good agreement with static experi-
ments at similar conditions (Tables 3 and 4). No signs of
liquid immiscibility have been detected by either method.
At different run durations and 12008C, the centrifuged
products of the mixture RY40 retained, in agreement
cpx with isothermal static runs, variable amounts of tridymite,
but showed also unequivocal signs of liquid immiscibility
and phase separation. Longer run durations resulted in
a better phase separation. However, immiscibility led to
formation of Lfe^Lsi emulsions, and at the exposures of
2 and 3 h liquid segregation was still incomplete. Vertical
(c) (d)
cross-sections of the quenched samples featured three
Lfe+Lsi+tr Lfe+Lsi+tr distinct zones (Figs 5^7): (1) a bottom zone of dark brown
Fe-rich glass; (2) a middle creaming zone of tridymite crys-
tals suspended in semi-transparent, milky glass with
plumes and clouds representing quenched emulsion of
silica-rich and Fe-rich liquids (Fig. 7); (3) a thin (2^3 mm)
layer of yellowish, honey-coloured, transparent silica-rich
glass at the very top of the sample (Figs 5 and 6).
Lfe Lfe According to the TEM image of a glass foil cut from the
creaming zone of the sample C-110 (Fig. 7b), the Lsi^Lfe
emulsions are characterized by drop-in-matrix morphol-
ogy, which implies immiscibility in the nucleation
2 mm
regime (Shelby, 2005), and a mean droplet size of about
Fig. 4. Examples of centrifuged run products. (a) Sample C-117, 100^150 nm. Some droplets appear to merge by floccula-
mixture KC at 11508C. Pyroxene crystals (cpx) in a turbid glass. tion and coalescence. In addition to the silicate unmixing,
(b) Sample C-106, mixture FJ and 12008C. Large clinopyroxene crys- a small amount of Fe^P liquid metal alloy (FeP) accumu-
tals in a dark brown glass. (c) Sample C-114, mixture RY20 in iron
container at 11508C. Emulsion of two liquids Lfe and Lsi and sus- lated at the bottom of the container in most of the runs.
pended tridymite crystals (tr) separated at the top of the clear The formation of the alloy liquid probably resulted from
Fe-rich glass Lfe. (d) Sample C-116, mixture RY20 in nickel container P2O5 reduction by C residing in the inner low-carbon
at 11708C. Poorly separated emulsion of two liquids and suspended
tridymite crystals. steel container. The formation of the Fe^P alloy had a neg-
ligible effect on the FeO content of the silicate liquids, but
resulted in a significant drop of P2O5 concentrations from
microprobe profiles (see the analyses at the top and bottom 47 wt% in the starting charge to an average of less than
of sample C-105 in Table 6). At 11508C (sample C-117), the 4 wt% in some of the centrifuged glasses. The increase of
mixture produced abundant crystals of low-Ca pyroxene, run duration, and the decrease in the proportion of silica-
and tridymite grains fractionated to the top of the sample rich components (mixture RY20) visibly improved phase
in a brownish^greenish, turbid glass (Fig. 4c). Vertical elec- separation. Tridymite crystals are virtually absent from
tron microprobe profiles revealed significant variations in some of the products of the RY20 mixture (e.g. sample
glass composition, with higher FeO contents at the bottom C-112 in Fig. 5), and a narrow but clear and distinct layer
and more silicic compositions at the top (Table 6). Thus, we of silicic glass formed at the top. The compositions of the
conclude that the composition KC showed some signs of Fe-rich and silica-rich experimental glasses at 12008C
unmixing (probably metastable) in the centrifuge runs, are, however, not so contrasting as those in the natural
but the results were obscured by crystallization and were iron-hosted melt inclusions (Table 1), upon which the
not reproducible at run durations shorter than 2 h. starting mixtures RY40 and RY20 were originally based.
Centrifuge experiments on mixture FJ at 12008C pro- To increase fO2 to the level of the Ni^NiO buffer, and com-
duced clear, brown glasses with large, elongated crystals pletely block the formation of the Fe^P alloy, we carried
of low-Ca pyroxene (samples C-106 and 108). Mixture our a few runs in nickel containers. Two experiments on
FJ20 quenched at 12008C to turbid, greenish glass with no mixture RY20 (samples C-116 and C-118) produced poorly
detectable compositional gradients in the vertical cross- separated Lfe^Lsi emulsions, which quenched to turbid,
section (see analyses in Table 6). In a static run (sample non-transparent glasses, and small amounts of tridymite
L-25), the mixture quenched at the same temperature crystals that floated to the top.

2197
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

(a) (b) Lsi (c)


SiO2 FeO CaO Al2O3 TiO2 MgO P2O5 K2O
6

5
Lfe+Lsi+tr
Height (mm)

1
Lfe
0
0 10 20 30 40 50 60 70 0 1 2 3 4 5
Concentration (wt.%) C−104 Concentration (wt.%)

(d) SiO2 FeO CaO P2O5 (e) (f) MgO Al2O3


TiO2 K2O
6 Lsi

5
Height (mm)

1
Lfe
0
0 10 20 30 40 50 60 70
0 2 4 6 8 10 12
Concentration (wt.%)
C−112 Concentration (wt.%)
Fig. 5. Phase separation and vertical compositional gradients in products of centrifuge experiments on mixtures RY40 and RY20 analogous to
the bulk compositions of the iron-hosted melt inclusions (Fig. 1). (a)^(c) represent compositional plots and a photograph of sample C-104; (d)^(f)
represent analogous plots and a photograph of sample C-112. Run conditions and electron microprobe analyses of the phases are given inTables 5
and 6. Abbreviations for phases are as in Fig. 4.

The model Skaergaard mixture MZ-1at 1110^1120 8C


was, as expected, slightly super-liquidus, and quenched in
iron containers to very dark brown glass, which looked
optically homogeneous. Electron microprobe profiles
tr revealed, however, steep vertical compositional gradients
Lsi for all the major oxides in a narrow layer at the very top
of the samples (Fig. 8) The profiles of the components are
consistent with element partitioning between Fe-rich and
silica-rich immiscible liquids (e.g. Schmidt et al., 2006;
Veksler et al., 2006, and references therein), and imply strat-
ification by liquid immiscibility. As in the other ferrobasal-
Lfe+Lsi tic compositions, liquid unmixing apparently started at a
sub-micron, colloidal scale, and emulsion coalescence and
creaming after 2 or 3 h of centrifugation remained incom-
plete. TEM line scans of foils cut at the top of the sample
C-111 immediately below the thin silica-rich layer revealed
Fig. 6. Back-scattered electron image of a small area at the upper right
corner of sample C-104 (see Fig. 5) showing tridymite crystals (tr) and fine-grained chemical heterogeneity of the glass at a sub-
the diffuse transition between the Fe-rich (Lfe) and silicic (Lsi) glasses. micron scale. Peaks in the Fe scan, with a characteristic

2198
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

(a) reacted with the melt, and changed the liquid composition.
Glass in the nickel container was contaminated at 11208C
by 08^12 wt% NiO, and the charge precipitated small
amounts of Ni-rich olivine and magnetite (see Tables 5
and 6). Vertical compositional gradients in the quenched
glass were not so pronounced as in the iron containers,
but still significant (Table 6). In the graphite container,
FeO reduction resulted in crystallization of metallic iron
on the container walls, and an overall drop in the FeO
concentration in the melt to below 15 wt%. A slight
increase in the silica content in the glass at the very top
(Table 6) may indicate incipient liquid phase separation,
but the gradient is even less pronounced than in the Ni
container, and probably insignificant.
Centrifuge experiments on the second, more Fe-rich
Skaergaard model mixture MZ-2 did not show significant
compositional gradients or any other signs of liquid
unmixing. In an iron container at 11108C, melt quenched
to dark brown homogeneous glass (sample C-120). The
homogeneity of the glass at a sub-micron scale was con-
firmed by TEM study, which did not reveal any composi-
2 mm tional heterogeneity above the background noise. In the
nickel container (sample C-121), the liquid dissolved 08^1
wt% NiO and precipitated Ni-rich magnetite at the
(b) bottom and on the container walls. The glass showed no
500 nm systematic compositional gradients, and appeared to be
homogeneous to sub-micron scale.
In Fig. 9 the compositions of the glasses from the
centrifuge experiments (Table 6) are plotted on the
olivine^feldspar^silica projection, and compared with
the position of the liquid miscibility gap in the
Fe2SiO4  KAlSi3O8  SiO2 system (Roedder, 1951, 1979).
For the projection, molar fractions of oxides (XMO) calcu-
lated from the analyses inTable 6 were combined into norma-
tive components: Fo ¼ XMgO/2; Fa ¼ (XFeO XTiO2)/2;
Ab ¼ 2XNa2O; Or ¼ 2XK2O; An ¼ XAl2O3  2(XNa2O þ
XK2O); Qz ¼ (XSiO2 þ XTiO2)/2 þ 3XP2O5 
(XMgO þXFeO)/2 5XNa2O  5XK2O XAl2O3  XCaO.
Fig. 7. Centrifuged products of the starting mixture RY20 at 12008C Although the projection excludes important normative
(sample C-110). (a) General view of the polished epoxy mount
in reflected light. The location of the fragment studied by TEM is components such as wollastonite, apatite and ilmenite, the
indicated by an arrow. (b) TEM image of the glass fragment compositions of the immiscible Lsi and Lfe glasses from the
showing sub-micron globular heterogeneity. High-absorption, Fe-rich centrifuge experiments plot well inside the immiscibility
amorphous phase is darker.
gap of the simplified synthetic system. The compositions of
homogeneous glasses tend to lie to the left of the two-liquid
field.
width of 20^30 nm and a magnitude at least two times
greater than background noise, corresponded to depres- Static mixing experiments
sions in the Si scan, implying the nucleation of Fe-rich The conflicting results of static and centrifuge experiments
and silica-rich amorphous phases. The total proportion of on mixtures RY40, RY20 and SF-1, as well as the unex-
silica-rich liquid Lsi in centrifuged products of the mixture pected unmixing of the Skaergaard model composition
MZ-1 was small, and the consolidated Lsi layer, which MZ-1, call for a critical examination of the experi-
formed on the top, is a mere 02 mm thick. Longer dura- mental techniques, and a careful check of equilibrium.
tions apparently enhance phase separation. Experiments Equilibrium is normally checked by reversal experiments,
were also performed in nickel and graphite containers and to perform those, we prepared mixtures and synthe-
(samples C-113 and C-122). Unfortunately, both containers sized glasses of the compositions MZF and MZS (Table 2),

2199
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

(a) SiO2 CaO


(b) (c) TiO MgO P2O5 K2O
FeO Al2O3 2
5

4
Height (mm)

0
0 10 20 30 40 50 60 70 0 1 2 3 4 5 6
Concentration (wt.%) C−115 Concentration (wt.%)

(d) SiO2 FeO CaO TiO2 Al2O3 (e) (f) MgO P2O5 K2O Na2O

4
Height (mm)

0
0 10 20 30 40 50 0 1 2 3 4
Concentration (wt.%) Concentration (wt.%)
C−120
Fig. 8. Vertical compositional gradients in centrifuged products of mixtures MZ-1 and MZ-2. (a)^(c) represent compositional plots and a
photograph of sample C-115; (d)^(f) represent analogous plots and a photograph of sample C-120.

which correspond to the end-member glass compositions again rapidly heated to 13508C, held at that temperature for
produced at the bottom and at the top in centrifuge runs a few minutes, and quenched. Finally, the loop with two
on mixture MZ-1 and believed to represent conjugate glasses was placed into gas-mixing furnace pre-stabilized
immiscible liquids Lfe and Lsi. Because the duration of at 1090 or 11148C, and fO2 conditions indicated in Table 3,
our centrifuge experiments was limited to a few hours, we annealed for 1 or 3 days, and quenched. Run durations
carried out a series of control reversal experiments in a longer than 3 days were deemed not practical, because
static setup. For that purpose, pairs of MZF and MZS the losses of more volatile components, such as alkalis or
melt droplets were loaded in Re loops, and Fe^Pt double P2O5, were expected to become significant, and affect bulk
containers identical to those that were used in centrifuge chemistry of the two-liquid charges.
experiments. The idea was to see whether the liquids, In containers (run S-141 in Tables 3 and 4), the mixture
which had been produced in the centrifuge, would mix MZF was first fused at the bottom of the inner iron con-
back when kept in contact in a loop for a sufficient time. tainer to a layer of bubble-free, homogeneous Fe-rich glass
In the loop set-up, approximately half of the loop was in the centrifuge furnace at 11108C. The upper part of the
filled first with the silica-rich mixture MZS. The powdered container was then filled with the silica-rich mixture MZS;
mixture was melted for a few minutes at 13508C and reduc- the outer Pt container was welded shut; the container
ing conditions in a 1atm gas mixing furnace, and assembly was placed in a vertical position into a static fur-
quenched to glass. Then the rest of the loop was filled nace, heated for 5 h at 12708C to fuse the MZS mixture,
with the Fe-rich mixture MZF, and the charge was once and then kept for 72 h at 11108C.

2200
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

Table 5: Run conditions and phase assemblages in products of centrifuge experiments

No. Mixture T1 T2 Exposure (min) Inner Run products

(8C) (8C) Slow Slow Fast Total container


rotation rotation rotation
at T1 at T2 at T2

C-101 KC T1 ¼ T2 1180 30 60 90 iron Emulsion glass þ ol þ tr


C-102 RY40 T1 ¼ T2 1200 30 60 90 iron Lfe þ Lsi þ tr
C-103 RY40 T1 ¼ T2 1200 60 none 60 iron Lfe þ Lsi þ tr
C-104 RY40 T1 ¼ T2 1200 60 120 180 iron Lfe þ Lsi þ FeP þ tr
C-105 KC T 1 ¼ T2 1180 60 120 180 iron Emulsion glass
C-106 FJ T1 ¼ T2 1200 60 none 60 iron L þ pig
C-107 FJ20 T1 ¼ T2 1200 60 120 180 iron Emulsion glass
C-108 FJ T1 ¼ T2 1200 60 120 180 iron L þ pig
C-109 RY40 T1 ¼ T2 1200 120 240 360 iron Lfe þ Lsi þ FeP þ tr
C-110 RY20 T1 ¼ T2 1200 60 120 180 iron Lfe þ Lsi þ FeP þ tr
C-111 MZ-1 1170 1120 30 60 120 210 iron Lfe þ trace Lsi
C-112 RY20 T1 ¼ T2 1170 120 120 240 iron Lfe þ Lsi þ trace tr
C-113 MZ-1 1170 1120 30 30 60 120 nickel Lfe þ trace Lsi þ ol þ mt
C-114 RY20 T1 ¼ T2 1150 120 120 240 iron Lfe–Lsi emulsion þ tr
C-115 MZ-1 1150 1100 30 30 180 240 iron Lfe þ trace Lsi
C-116 RY20 T1 ¼ T2 1170 90 90 180 nickel Emulsion glass þ tr
C-117 KC T1 ¼ T2 1150 30 180 210 iron Emulsion glass þ pig þ tr
C-118 RY20 T1 ¼ T2 1200 60 180 240 nickel Lfe–Lsi emulsion þ tr
C-120 MZ-2 1150 1110 30 60 120 210 iron L
C-121 MZ-2 1150 1110 30 60 120 210 nickel L þ mt
C-122 MZ-1 1150 1120 15 30 135 180 graphite L þ iron
C-123 MZ-1 1150 1120 15 45 360 420 iron Lfe þ trace Lsi

Slow rotation corresponds to the acceleration of 50g (g ¼ 98 m/s2), except in C-101 and C-102 where it was 160g.
Fast rotation corresponds to the acceleration of 1000g.

Despite the simple experimental design, the results effective mixing by convection. The Lsi glasses were also
appeared not as simple and straightforward as anticipated. homogeneous, and the transition zone between the glasses
In four loop runs L-51, 52, 56 and 58 the Lfe and Lsi liquid in most cases was narrow and sharp (Fig. 10b). In a single
droplets did not mix back (Fig. 10); however, they did mix run at a lower temperature of 10908C, groups of thin,
completely in the run L-57 (Table 3). Furthermore, elec- needle-like, Ca-rich pyroxene crystals nucleated in the
tron microprobe analyses of quenched glasses (Table 4) silica-rich glass, and the portion of glass near the crystals
revealed a significant partial mixing in all the loop runs. was further depleted in CaO, MgO, and FeO, and
Notably, it mostly affected the Fe-rich liquid. Regardless of enriched in SiO2 by about 2 wt%. We interpret the crystals
the run durations, and initial mass proportions of the two as the first liquidus phase forming in the Lfe^Lsi mixtures.
starting droplets, the Fe-rich melts dissolved roughly equal It appears that the outcome of the mixing experiments
amounts of Lsi, whereas the composition of the silicic in loops depended on the mass proportion of the Fe-rich
liquid (provided it did not dissolve completely as in the and silica-rich (MZF and MZS) liquid droplets. The
run L-57) changed only slightly. In all the experiments, masses of the MZFand MZS mixtures were carefully mea-
including the run L-57, mixing stopped on the Fe-rich side sured in three runs, L-56, L-57 and L-58, and the MZS/
at the same final melt composition characterized by a SiO2 MZF mass ratios in those runs were 135, 077 and 12,
content of 54  05 wt%. Electron microprobe profiles respectively. Thus, it is not surprising that the Lsi glass
across the samples demonstrated that the final Lfe glasses completely dissolved in the run L-57 in which the
were chemically homogeneous, probably because of initial mass proportion of the silica-rich MZS mixture

2201
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

Table 6: Electron microprobe analyses of the products of centrifuge experiments

No. T (8C) Mixture Phase SiO2 TiO2 Al2O3 MgO FeO CaO Na2O K2O P2O5 NiO Total

C-106 1200 FJ glass 5020 212 331 954 2127 1063 086 030 042 001 9866
pigeonite 5324 054 041 2197 1891 498 002 000 003 002 10012
C-108 1200 FJ glass 4882 213 344 772 2250 1057 111 035 044 002 9909
pigeonite 5182 035 019 2222 1857 505 006 001 002 003 9830
C-107 1200 FJ20 bottom 5468 172 530 781 1756 886 166 133 032 001 9926
top 5513 169 527 754 1718 860 158 130 034 000 9861
C-101 1180 KC bottom 4548 358 339 625 2883 924 153 082 037 000 9948
top 7248 117 1022 187 497 337 211 341 001 001 9960
olivine 3489 016 002 2301 4242 065 001 001 008 001 10124
C-105 1180 KC bottom 5729 374 454 563 1669 832 197 147 036 000 10001
top 5674 381 445 550 1631 824 200 146 035 000 9886
C-117 1150 KC bottom 5237 414 440 568 2067 905 185 124 048 000 9988
top 6009 370 506 443 1521 721 195 183 038 000 9985
Pigeonite 5214 109 038 1866 1885 940 018 006 002 000 10079
C-104 1200 RY40 Lfe 4254 223 1027 478 2260 1064 053 048 436 002 9845
Lsi 6233 225 1180 319 725 804 111 131 089 001 9817
C-109 1200 RY40 Lfe 4504 215 1001 453 2157 1026 059 053 330 001 9800
Lsi 6446 224 1205 342 551 817 114 132 015 001 9846
C-110 1200 RY20 Lfe 4215 276 876 529 2197 1148 078 043 488 001 9851
Lsi 6320 293 1079 380 654 848 155 116 044 001 9891
C-112 1170 RY20 Lfe 4063 297 882 561 2243 1169 071 035 538 001 9861
Lsi 6297 270 1027 334 711 787 158 126 067 000 9861
C-114 1150 RY20 bottom 4177 295 914 588 2257 1142 067 036 507 000 9982
top 4818 254 906 490 1653 1073 108 069 646 000 10016
C-116 1170 RY20 bottom 4006 296 917 566 1835 1196 074 04 850 110 9890
top 4650 253 914 531 1490 1108 095 064 751 090 9946
C-118 1200 RY20 bottom 4046 295 914 573 1778 1189 075 044 851 090 9856
top 4727 261 879 514 1459 1093 095 061 762 085 9935
C-111 1120 MZ-1 Lfe 4767 505 988 375 1784 998 262 056 038 001 9776
Lsi 6246 451 1201 223 775 640 180 134 016 000 9866
C-115 1100 MZ-1 Lfe 4815 516 1010 404 1814 1004 258 056 044 000 9921
Lsi 6492 377 1224 223 728 584 250 149 013 000 10039
C-113 1120 MZ-1 bottom 4806 514 1006 375 1689 1016 257 058 037 088 9845
top 5454 410 1070 310 1323 868 258 093 029 099 9914
olivine 3431 019 006 1704 1562 008 003 001 004 3272 10010
magnetite 032 1957 325 307 5361 000 002 002 001 1480 9466
C-122 1120 MZ-1 Bottom 5082 541 1064 413 1483 1041 269 061 042 000 9995
top 5501 540 1134 369 1184 942 270 074 029 000 10043
C-120 1110 MZ-2 Bottom 4639 485 977 378 2224 942 260 058 039 000 10000
top 4764 481 996 373 2112 925 279 064 038 000 10031

Oxide concentrations are in wt%.



Glass compositions at the bottom and at the top of the container.

was the lowest. The presence of a massive iron container 3 days at 11108C. However, microprobe profiles show that
also seems to play a role in the mixing process, as demon- the transition between optically distinct Fe-rich and Si-rich
strated by the static experiment S-141. In that experiment, glasses is not razor-sharp. The width of the diffusion zone
the MZF and MZS liquids layers did not mix back after between the glasses is about 02 mm. The diffusion zone

2202
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

Qz
(a)

sil

or
opx Lfe

fa
lc
Lsi
Fo+Fa mol. % An+Ab+Or 1 mm

Fig. 9. Glass compositions from centrifuge experiments plotted on


the (Fo þ Fa)^(An þ Ab þ Or)^Qz projection. Homogeneous glasses (b) SiO2 FeO CaO K2O TiO2 P2O5
from samples without phase separation are plotted in black circles. 100
Immiscible glasses are shown by open circles (mixtures RY40 and
RY20), and black triangles (mixture MZ-1). Pairs of the conjugate Lsi
and Lfe compositions are connected by tie-lines. The gray area and Concentration (wt.%)
dashed curves represent the miscibility gap and crystallization fields
in the system Fa^Or^Qz (after Roedder, 1951). (See text for explana- 10
tion of the projection method.)

may have formed during the brief heating to 12708C, and


1
it seems that no further mixing took place during the long
exposure at 11108C. The composition of the bottom Fe-rich
layer remained virtually intact (Table 4). In contrast to the
centrifuge experiments, minor crystallization took place in
0.1
the charge. The bottom Fe-rich liquid precipitated small 0 1 2 3
amounts of olivine crystals (Fo50), and a few pyroxenes Distance (mm)
crystallized in the silicic glass at the container walls, and
Fig. 10. Products of a static mixing run. (a) General view of sample
at the MZF^MZS contact. As discussed in the next sec- L-56 in the Re ribbon loop showing incomplete mixing between
tion, the iron container may function as a buffer of FeO Fe-rich glass (Lfe) and silicic glass (Lsi). (b) Electron microprobe pro-
activity and thus affect the equilibrium between two files across the sample. The interface between the Fe-rich and silicic
immiscible silicate liquids. glasses is manifested by a step in the elemental profiles. (See text for
further discussion.)
Static mixing experiment S-141 is analogous in design to
some experiments performed by Longhi (1990) on Fe-rich
and silica-rich immiscible lunar compositions. Those Lfe reached 54 wt% are not clear. First, the composition
experiments were also performed in iron containers, may represent a sort of kinetic barrier at which the viscos-
and different loading methods were used. In some runs, ity of Lfe becomes too high for convective mixing, forcing
powders of Fe-rich and silica-rich conjugate compositions further mixing to proceed at a much slower rate by diffu-
were loaded in two separate layers, and this is very similar sion only. Alternatively, the composition may represent
to what we did in the run S-141. Longhi (1990) may well a true conjugate liquid. If so, the difference from glass com-
have encountered the same metastability problems as we positions formed in the centrifuge may be attributed
did in our static runs. He reported extensive crystallization to incomplete equilibration in shorter centrifuge runs,
of silicates at the interfaces between glasses of contrasting especially for slowly diffusing network-forming compo-
composition, and noted the absence of a clear sharp menis- nents (see also Veksler et al., 2006). Finally, the composition
cus in some of his experimental products. Crystallization may correspond to the borders of the spinodal region of the
of pyroxene in the mixing zone between the MZF and miscibility gap, where further mixing becomes energeti-
MZS glasses in our run S-141 may be of the same nature. cally unfavourable. In our view, the first, kinetic explana-
The reasons why mixing repeatedly stopped (or dramat- tion is less likely because the run durations were
ically slowed down) in loops when SiO2 concentration in apparently long enough for effective diffusional transport

2203
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

in even more viscous Lsi, and no compositional gradients, reliability of the conventional static loop technique when
which would arise from incomplete diffusion, were it is applied to dry ferrobasaltic, andesitic and dacitic
observed in the quenched Lfe glasses. With some reserva- compositions at temperatures below 11008C. Here, none of
tions, we conclude that immiscibility in the MZ-1 liquid is the unambiguous cases of immiscibility revealed in natural
likely to be stable, although chemical equilibration in the glasses was reproducible in static experiments.
centrifuge was probably incomplete, and true conjugate Centrifuge experiments differed from the static runs not
liquids may somewhat differ in composition from only in their higher acceleration, but also in a significantly
the glasses that were analysed at the top and the bottom greater sample mass, and the presence of massive contain-
of centrifuged samples. ers. We used inner containers made of several materials.
None of them proved to be perfectly inert. However, iron
Crystallization experiments on the Fe-rich appears to be the best choice. The main disadvantage
and silica-rich conjugate liquids of iron is that it sets fO2 values somewhat below the
Thermodynamics requires that equilibrium immiscible iron^wu«stite buffer [see Veksler et al. (2006) for further
liquids should crystallize identical mineral assemblages. discussion], and such conditions are too reducing for the
To test this requirement of our putative immiscible liquids, vast majority of terrestrial magmatic environments
we performed a series of static crystallization experiments (with the rare exception of the magmas described above
on pure compositions MZF and MZS. Run conditions and that precipitate native iron). Equilibration with an iron
the results are summarized in Tables 3 and 4. Plagioclase container may slightly change the FeO(t) of the silicate
and augite are the main phases crystallizing in both mix- charge. However, on the positive side, the dissolution
tures at 1081 and 10628C, and compositions of the solid reaction between silicate melt and metal iron regulates
solutions in the products of the MZF and MZS runs are not only fO2, but also the activity of FeO in the melt.
similar (Table 4). The MZF mixture crystallized at Thus, the FeO component, which is crucial for liquid
10818C, plagioclase An56 and pyroxene Wo42En34Fs24 immiscibility, becomes effectively constrained.
(sample L-62), whereas the mixture MZS at the same tem- Graphite containers at low pressures are even more
perature crystallized zoned plagioclase An75^65 (average reducing than iron and proved to be unsuitable. On the
composition An72) and augite Wo32En42Fs26 (sample other hand, our experiments in nickel containers and in
L-63). The crystalline products also include minor amounts the presence of the more oxidizing Ni^NiO buffer were
of Fe^Ti oxides, and olivine, which appeared in products of hampered, as described above, by Ni contamination and
both mixtures between 1062 and 10438C. All the products crystallization of Ni-bearing oxides and silicates. It is
of the MZS mixture also contain variable amounts of not clear, however, why liquid immiscibility and phase
a pure silica phase, probably tridymite. However, this separation in mixtures RY20 and MZ-1 appear to
tridymite, as discussed below, may be partly or fully meta- be slower in nickel than in iron, as evidenced by less
stable. With falling temperature, plagioclase generally pronounced vertical compositional gradients in the centri-
becomes, as expected, more albitic. However, in the fuged products. Additions of less than 1 wt% of NiO to the
low-temperature MZF products at 10228C (sample L-64) liquid phase(s) are not likely to cause large effects on
plagioclase crystals showed broad compositional variations unmixing, and a greater proportion of the Fe2O3 compo-
clustering around two modes (plag-1 and plag-2 inTable 4). nent in more oxidized melts is expected, if anything, to
It should be noted that equilibrium is hardly achievable broaden the miscibility gap and enhance the unmixing
at temperatures below 10508C, especially in the silicic, (Naslund, 1983). However, there may be other, less obvious
viscous liquids of the mixture MZS. Despite these compli- and indirect effects, of a kinetic nature. If liquid phase
cations at lower temperatures, the main conclusion from separation develops via diffusion-controlled nucleation
the crystallization experiments is that the crystal assem- and growth (and TEM images of quenched glasses,
blages formed from the MZF and MZS mixtures such as that in Fig. 7b, support this view), the process
above 10508C are similar, and so the liquids may indeed of nucleation and droplet growth may, for example, signif-
represent immiscible conjugate phases. icantly slow down in liquids with a higher ferric^ferrous
ratio, as amphoteric, network-forming Fe3þ is likely
to diffuse much more slowly than the network-
DISCUSSION modifying Fe2þ.
Kinetic and thermodynamic limitations
Our results imply that reaching thermodynamic equilib- Metastability and liquid immiscibility
rium in experiments on multicomponent ferrobasaltic The classical, best-studied example of low-temperature
compositions at temperatures below 11008C is a very chal- silicate immiscibility in the SiO2^Al2O3^FeO^K2O
lenging task. The liquids are strongly non-ideal, and system is characterized by a flat solvus emerging by
show a clear tendency towards liquid immiscibility. We maximally 1708C above the liquidus surfaces of tridymite
believe that our results pose serious doubts regarding the and fayalite (Roedder, 1951; 1979; Schmidt et al., 2006).

2204
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

The stable miscibility gap along the fayalite^ferrobustamite common in our centrifuged run products. Although some
cotectic in the Fe2SiO4  CaFeSi2O6  KAlSi3O8  SiO2 turbid glasses in our static and centrifuge experiments may
system was reported to be no more than108C above the liqui- have formed by metastable unmixing during quench
dus (Hoover & Irvine, 1978). Numerous other aluminosili- (e.g. samples C-107 and C-108), those centrifuged glasses
cate systems present examples of metastable immiscibility that developed pronounced vertical compositional gradi-
characterized by S-shaped liquidi and liquid^liquid immis- ents probably formed from stable immiscible emulsions.
cibility domes submerged just below the liquidus surfaces The textures of quenched samples, which can be viewed
(e.g. Irvine, 1976). Such topological relationships imply as snapshots of droplet distribution after different centrifu-
that the Gibbs free energy of two immiscible liquids and gation times, demonstrate that sub-micron droplets were
alternative phase assemblages of a single homogeneous slowly moving during centrifugation over macroscopic
liquid and crystal phase(s) are very close for a broad range distances, and in some samples (e.g. products of the
of petrogenetically relevant basaltic and dacitic composi- mixtures RY20 and RY40) formed dense clouds, and
tions.This situation should be favourable for metastability. eventually merged into compact layers of homogeneous
liquids (Figs 4c, 5 and 6). Such features cannot form in
Tridymite as a metastable phase one or two minutes of quench time, and imply that those
Unusual crystallization of tridymite in mixtures SF-1, fine emulsions of silicate liquids formed by super-liquidus
RY20 and RY40 has precedents in other previously liquid immiscibility, and remained stable for hours. TEM
published experimental studies. For example, Irvine images of quenched glasses allow us to see single droplets
(1976) reported an unexpected appearance of primary at high magnifications, measure their size, and reveal the
liquidus tridymite inside the plagioclase crystalli- morphology of emulsions. According to the TEM images,
zation field on two pseudo-ternary joins of the Mg2SiO4  some turbid glasses produced in static runs may result
Fe2SiO4  CaAl2Si2O8  KAlSi3O8  SiO2 system. Tridy- from spinodal decomposition (Fig. 3), but droplets-
mite crystallization was also reported in the mesostasis in-matrix dispersed morphologies, which are typical for
of natural tholeiitic basalt at the brink of stable liquid centrifuged glasses (Fig. 7b), imply formation and growth
immiscibility (Philpotts & Doyle, 1983). Notably, in the in a nucleation regime (Shelby, 2005).
latter case tridymite crystallized at temperatures below
1030^10608C from a homogeneous liquid with only 57^59
The Bowen and Fenner trends:
wt% SiO2. We propose that precipitation of tridymite may
physicochemical background
in some cases represent a metastable response of silicate
liquids to strong interactions between FeO and SiO2 Oxygen fugacity, melt non-ideality and magnetite stability
components at ionic level (Hudon & Baker, 2002), which Textbook explanations of the Bowen and Fenner trends
in a normal, equilibrium state would have led to liquid (e.g. Morse, 1994) are usually built upon the equation
immiscibility. One could speculate that for complex, 3Fe2 SiO4 þO2 ¼ 2Fe3 O4 þ3SiO2 : ð1Þ
multicomponent liquids it may be sometimes easier, from
the kinetic point of view, to decrease the instability of melt when the reaction takes place between pure crystalline
structure by precipitation of pure crystalline silica than by phases, it defines the fayalite^magnetite^quartz (FMQ)
nucleation of a complex immiscible liquid. Our centrifuge oxygen buffer, and establishes a functional dependence
experiments on mixtures RY20 and RY40 demonstrate between T and fO2. Although natural basaltic systems do
that better phase separation and a greater volume of the not crystallize assemblages of pure fayalite, magnetite
continuous Lsi layer at longer run durations also correlate and silica minerals but rather solid solutions of olivine
with decreasing proportion of tridymite crystals. In and Fe^Ti oxides, with silica remaining in a residual melt,
our view, occurrences of sporadic tridymite and quartz the equation is still useful as a simple, shorthand formula-
precipitation from melts that are high in Fe-oxides and tion of two key principles: (1) magnetite stability in mag-
low in SiO2 deserve a thorough re-examination from the matic systems is a function of fO2; (2) crystallization
point of view of potential liquid immiscibility. of magnetite should lead to a strong increase in the SiO2
content of the liquid.
Immiscible emulsions The link between magnetite crystallization and the
Formation of colloidal emulsions is another form of meta- Bowen (silica enrichment) trend as a result of fundamental
stability, which was often observed in the products of our mass-balance principles is undeniable. However, two
centrifuge and static runs. Sub-micron heterogeneities are things should be kept in mind when equation (1) is applied
widespread in silicate glasses, and they have been usually to complex natural basaltic systems. First, fO2 is surely not
thought to form during quench or to result from metastable the only factor controlling the magnetite stability in multi-
liquid immiscibility (Vogel, 1985; Shelby, 2005). Plumes and component magmas. In absence of liquidus tridymite or
clouds of colloidal droplets, as well as blurred and diffuse quartz, silica activity (aSiO2) is not fixed at unity, and
interfaces between immiscible liquid layers, are very becomes a variable, which must be raised to the third

2205
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

power and incorporated in the equilibrium constant for Table 7: Liquid compositions and crystal assemblages
reaction (1). A detailed analysis of the relationship between
silica activity and oxygen fugacity in layered gabbroic produced by melting of natural Skaergaard gabbros (sample
intrusions has been carried out by Morse (1980, 1990). UB-4145) and crystallization of synthetic glasses
Because of the strong non-ideality of silicate liquids, aSiO2
is a complex function of melt composition. The ferrous^ Sample UB-41451 L-692 Fe-763 L-702
ferric ratio in basaltic liquids certainly depends on aSiO2
and the activities of other components, especially alkalis
SiO2 463 5022 5751 7244
(e.g. Kilinc et al., 1983). In view of the important role of
TiO2 34 368 262 169
melt chemistry, it is hardly surprising that alkalis have
Al2O3 82 970 1113 1194
been shown to have a strong effect on magnetite stability,
MgO 19 203 134 080
and the general trend of melt evolution in petrologically
FeO 260 2151 1631 453
relevant aluminosilicate systems (Irvine, 1976; Shi, 1993).
CaO 100 821 599 294
The second consideration is also related to the strong
non-ideality of ferrobasaltic liquids. Some experimental Na2O 24 259 325 250

models show that magnetite crystallization does not neces- K2O 05 108 164 260

sarily prevent the formation of Fenner-type (Fe-rich) P2O5 13 116 00 030
liquids. Furthermore, in those cases where non-ideality Total 1000 10018 9979 9814
takes the extreme form of stable immiscibility, Fe-rich T (8C) 1056 1043 1050 1043
liquids coexist with silicic, Bowen-type liquids, and various log fO2 –111 –119 –124 –119
assemblages of Fe^Mg silicates, Fe^Ti oxides, and silica Compositions of coexisting solid solutions
minerals. Such relationships have been most vividly Olivine, Fo-no. 346 301 275 330
demonstrated in simplified synthetic systems where liqui- Plagioclase, An-no. 434 453 430 480
dus phases comprise pure fayalite and magnetite, and Pyroxene
aSiO2 is fixed at unity by liquidus tridymite (e.g. Naslund, Wo 417 401 379 255
1983). Naslund (1983) showed that fayalite^tridymite and En 321 277 275 324
magnetite^tridymite cotectics in the system Na2O(K2O)^ Fs 262 322 346 421
FeO^Fe2O3^Al2O3^SiO2 intersected a region of stable Other phases mt, ilm mt, ilm ilm mt, tr
liquid immiscibility over a broad range of fO2 from 10^12
to 02. As a result, a series of Bowen- and Fenner-type alu- References: 1McBirney & Naslund (1990); 2this work;
3
minosilicate liquids were shown to coexist regardless of the Toplis & Carroll (1995). Liquid compositions are in wt%
direction of the magnetite^fayalite reaction (1). This work oxides; crystal compositions in mol% normative
also showed that the miscibility gap greatly expanded, components.
and the compositions of coexisting liquids diverged at oxi-
dizing conditions and in equilibrium with magnetite.
Apparently, this means that the Fe2O3 component causes between FeO and SiO2 components in silicate melts
greater non-ideality in aluminosilicate melts than FeO. [see Hudon & Baker (2002) for a theoretical discussion]
The main conclusion, however, is that the effect of magne- should be universal and reveal itself also in liquid composi-
tite crystallization upon the direction of melt evolution tions lying outside miscibility gaps. In the case that non-
may be secondary and subordinate to strong chemical ideality does not result in stable unmixing or metastable
interactions between components and species in the liquid phase separation, it reveals itself in broad compositional
itself. Strong intrinsic non-ideality of ferrobasaltic liquids variations of activity coefficients. Activities in the
may be a more important factor in magma evolution than liquid phase can be assessed from crystal^liquid equilibria.
the composition of liquidus mineral assemblages or fO2 A few examples illustrating this approach are presented in
conditions imposed on the system. Table 7. Here we compare our samples L-69 and L-70,
the products of the MZF and MZS starting mixtures,
Concentration^activity relationships in complex, with two samples published earlier by McBirney &
multicomponent liquids Naslund (1990) and Toplis & Carroll (1995). The data
Unlike the immiscible synthetic systems referred to above, show that very different liquid compositions with SiO2
experiments on natural basaltic compositions have seldom concentrations from 46 to 72 wt% can be equilibrated at
produced stable liquid unmixing. Above, we have already similar T and fO2 with almost identical crystal phase
expressed our doubts regarding the reliability of conven- assemblages comprising olivine (Fo27^35), plagioclase
tional static techniques in the vicinity of or inside stable (An43^48), clinopyroxene, and Fe^Ti oxides. Notably, the
miscibility gaps. Here we would like to point out that liquids appear to show a strong negative correlation
non-ideality arising from anomalously strong interactions between the SiO2 and FeO components, and positive

2206
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

correlations between FeO and CaO, SiO2 and Al2O3, and crystallization experiments on synthetic glasses and
SiO2 and alkalis. Also notable is that the spectrum of rock powders, which modeled the starting composition of
liquids from silicic to extremely Fe-rich forms in a narrow the Skaergaard magma, produced liquids with FeO(t)
range of redox conditions, and the presence or absence concentrations of no more than 18^22 wt% (e.g. Toplis &
of magnetite does not seem to be a crucial factor. Carroll, 1995; Thy et al., 2006), near the compositional
Importantly, Fenner-type liquids at the Fe-rich side of the path inflections where Fe^Ti oxides start to crystallize,
spectrum are characterized not only by low silica content, after which experimental liquids evolve towards silica
but also by low alkalis and high Ca/Al values. These enrichment. This apparent discrepancy demands an
chemical characteristics appear to arise from activity^ explanation.
composition relationships in complex aluminosilicate Our static experiments on compositions MZF and MZS
liquids where FeO, on the one hand, increases the activity (Tables 3, 4 and 7) confirm that very different liquid com-
coefficient of SiO2, whereas alkalis decrease it (e.g. Shi, positions from ferrobasaltic to rhyolitic can be equilibrated
1993). A very similar pattern of the activity^composition under identical conditions with crystal assemblages typical
relationships is revealed by element partitioning between for cumulates of the Middle Zone (MZ) and Upper Zone
the Lsi and Lfe immiscible liquids [e.g. Tables 1 and 6; (UZ) of Skaergaard. It is also true that very similar
see also Veksler et al. (2006) and references therein]. In mineral assemblages can be found in felsic volcanic rocks
the process of the formation of immiscible liquids, alkalis (Hunter & Sparks, 1987, 1990; Sparks, 1988) characterized
partition to the Bowen-type silicic liquids, whereas FeO by rhyolitic groundmass, bulk-rock SiO2 concentrations
and alkaline earths partition to the Fenner-type liquids, well above 60 wt% and low FeO(t). Because the composi-
which are also often characterized by very low alumina tions MZF and MZS were produced by immiscibility in
contents. centrifuge experiments we propose that the diverging
It was also demonstrated that variations of silica content liquid evolution paths in Fig. 2 skirt a region of stable
in melts along the pseudo-binary join AnDi^SiO2 (where liquid immiscibility. Alternatively, immiscibility may be
AnDi is the eutectic composition in the anorthite^diopside
metastable, and the paths may fan out across a flat and
system) have complex, non-linear effects on Fe solubility
broad liquidus plateau. Flat liquidus surfaces are typical
(Borisov, 2007). The maximum of Fe solubility (that is,
for regions above metastable immiscibility domes
a minimum activity coefficient of FeO) was observed in
(e.g. Irvine, 1976). In any case, strong non-ideality rooted
melts with about 55^57 wt% SiO2, regardless of tempera-
in the anomalous incompatibility between Fe-oxide and
ture and oxygen fugacity. Melts on both sides of the
SiO2 components in silicate melt structures appears to be
maximum show significantly lower Fe solubility and, cor-
the principal reason for the divergent liquid evolution
respondingly, higher FeO activity coefficients. It is clear
paths.
that Lsi melts should lie on the silica-rich side of the max-
Silicate liquid immiscibility has been mentioned but pre-
imum, where increasing silica contents result in decreasing
viously never as a central issue in debates about the Bowen
FeO concentrations.
and Fenner trends (e.g. McBirney, 1975). Immiscibility cer-
Implications for the Skaergaard intrusion tainly operated in the Upper Zone of the Skaergaard
and tholeiitic magmas in general Layered Series, close to the end of magma crystallization.
The most recent major dispute about the Fenner and It has been reproduced in melting experiments on the UZ
Bowen trends in the Skaergaard intrusion took place in cumulates (McBirney & Nakamura, 1974; McBirney &
the late 1980s. It was started by Hunter & Sparks (1987), Naslund, 1990), confirmed by the discovery of silicic and
who argued against the strong Fe-oxide enrichment in the Fe-rich melt inclusions in cumulus apatite from UZ layered
Skaergaard liquid on the basis of mass-balance considera- gabbros (Jakobsen et al., 2005; see also Fig. 2) and indicated
tions. A series of papers followed (Brooks & Nielsen, 1990; also by the common occurrence of large-scale, outcrop-
Hunter & Sparks, 1990; McBirney & Naslund, 1990; sized segregations of melanogranophyre. Our results
Morse, 1990), in which the controversy was examined imply that the process may have started earlier, and
from different angles, and on the basis of different types of demand that the role of silicate liquid immiscibility
evidence. In the context of the present discussion, we should be reconsidered.
would like to emphasize the thermodynamic and experi- We propose that the Fe-enrichment trend in Skaergaard
mental aspects of the debate, and point out that liquid evo- was driven at first by fractional crystallization of troctolitic
lution paths at the high- and low-SiO2 ends of Fig. 2 are and gabbroic cumulates that formed the lower parts of the
directly based on experimental studies. Ferrobasaltic Layered Series (the LZa and LZb of the conventional
glasses with FeO(t) well above 26 wt% were produced Skaergaerd stratigraphy, e.g. McBirney, 1989). This is
in a series of partial melting experiments on natural a classical tholeiitic trend, primarily caused by a high
Skaergaard gabbros (McBirney & Nakamura, 1974; cotectic proportion of plagioclase in the cumulus
McBirney & Naslund, 1990). On the other hand, numerous assemblages, along the lines explained, for example,

2207
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

by Grove & Baker (1984). Such Fe-enrichment continued be finally resolved at the location where it has been debated
until about 60 wt% of crystallization (Nielsen, 2004), most, the Skaergaard intrusion in East Greenland.
which corresponds to crystallization of the upper part
of the Lower Zone (LZc), when the magma became
saturated in Fe^Ti oxides. Any further significant AC K N O W L E D G E M E N T S
Fe-enrichment by fractional crystallization would have This study was inspired by fruitful collaboration and dis-
been impossible, and the liquid would have started to cussions of Skaergaard with Alexey Ariskin, Charles Kent
evolve towards silica enrichment. However, in view of our Brooks, Marian Holness, Jakob Jakobsen, ChristianTegner,
centrifuge experiments on the composition MZ-1, liquid Troels Nielsen, and Gemma Stripp. We are grateful to
immiscibility may have started in the Skaergaard magma Dieter Rhede and Oona Appelt for the help with microp-
at about the same time. The initial proportion of the robe analyses. Reviews by Peter Thy, Tony Morse and two
immiscible silica-rich liquid Lsi in the Fe-rich Skaergaard anonymous reviewers were thoughtful, friendly, and most
magma was probably small, at around 15 wt%, but the helpful. I.V.V. has been supported by grants from the
unmixing and gravitational separation of the conjugate German Science Foundation (DFG).
liquids provided a mechanism for continuous Fe enrich-
ment in the denser liquid at the bottom of the magma
chamber despite the crystallization of Fe^Ti oxides.
R E F E R E NC E S
In our currently preferred scenario, fractional crystalliza-
Ariskin, A. A. (1999). Phase equilibria modelling in igneous petrology:
tion coupled with liquid immiscibility continued in
use of COMAGMAT model for simulating fractionation of ferro-
Skaergraard until the end of crystallization in the basaltic magmas and the genesis of high-alumina basalts. Journal
Sandwich Horizon. We believe that the combined effects of Volcanology and Geothermal Research 90, 115^162.
of crystallization and unmixing may explain many enig- Ariskin, A. A. (2002). Geochemical thermometry of layered series
matic features of the Skaergaard rocks and other layered rocks of the Skaergaard intrusion. Petrology 10, 495^518.
gabbroic intrusions. Studies of plagioclase-hosted melt Bird, J. M., Goodrich, C. A. & Weathers, M. S. (1981). Petrogenesis of
inclusions (Jakobsen et al., 2007) and pyroxene^plagioclase Uifaq iron, Disko island, Greenland. Journal of Geophysical Research
86, 11787^11805.
symplectites (Stripp et al., 2007) have already produced Bogaerts, M. & Schmidt, M. W. (2006). Experiments on silicate melt
circumstantial evidence in favour of early liquid immisci- immiscibility in the system Fe2SiO4^KAlSi3O8^SiO2^CaO^
bility in the Skaergaard intrusion. However, a detailed dis- MgO^TiO2^P2O5 and implications for natural magmas.
cussion of the numerous geological implications of this Contributions to Mineralogy and Petrology 152, 257^274.
hypothesis is beyond the scope of this paper and will be Borisov, A. A. (2007). Experimental study of the effect of SiO2 on
presented elsewhere. cobalt and iron solubilities in silicate melts. Petrology 15, (in press).
Borisov, A. & Jones, J. H. (1999). An evaluation of Re, as an alterna-
Immiscible emulsions of silicate melts, which were rou-
tive to Pt, for 1-bar loop technique: an experimental study at
tinely formed in our experiments, are of special interest 14008C. American Mineralogist 84, 1528^1534.
for future research. Kinetic and thermodynamic factors Bowen, N. L. (1928). The Evolution of the Igneous Rocks. Princeton, NJ:
regulating the formation and stability of such emulsions Princeton University Press.
in complex aluminosilicate melts are at the moment Brooks, C. K. & Nielsen, T. F. D. (1978). Early stages in the differentia-
poorly understood. It is not clear whether such emulsions tion of the Skaergaard magma as revealed by a closely related suite
of dike rocks. Lithos 11, 1^14.
can be sufficiently stable at the timescale of natural mag-
Brooks, C. K. & Nielsen, T. F. D. (1990). The differentiation of the
matic processes to seriously affect magma dynamics. As Skaergaard intrusion. A discussion of Hunter and Sparks (Contrib
discussed above, our results hint at significant effects of Mineral Petrol 95: 451^461). Contributions to Mineralogy and Petrology
melt composition upon the kinetics of unmixing (see also 104, 244^247.
Veksler et al., 2007). Viscosity, interfacial energy and diffu- Carmichael, I. S. E. (1964). The petrology of Thingmuli, a Tertiary
sion are likely to be the key factors, and diffusion-driven volcano in eastern Iceland. Journal of Petrology 5, 435^460.
Daly, R. A. (1914). Igneous Rocks and their Origin. New York: McGraw^
coarsening of silicate emulsions is expected to be slow.
Hill.
In general, it is envisaged that the proposed immiscibility De, A. (1974). Silicate liquid immiscibility in the Deccan Traps and its
between Bowen- and Fenner-type liquids is most likely, and petrogenetic significance. Geological Society of America Bulletin 85,
should be most effective in large, slowly cooling basaltic 471^474.
magma chambers, where the proportion of the Fe-rich, low- Dixon, S. & Rutherford, M. J. (1979). Plagiogranites as late-stage
viscosity liquid is large; where there is enough time for emul- immiscible liquids in ophiolite and mid-ocean ridge suites: an
sion coarsening as a result of coalescence and/or Ostwald experimental study. Earth and Planetary Science Letters 45, 45^60.
Dorfman, A. M., Hess, K.-U. & Dingwell, D. B. (1996). Centrifuge-
ripening; and where other conditions are met for large-scale
assisted falling-sphere viscosimetry. European Journal of Mineralogy
gravitational phase separation. The Skaergaard intrusion 8, 507^514.
may represent an example of such a magma chamber. Fenner, C. N. (1929). The crystallization of basalts. American Journal of
Perhaps the long-standing Bowen^Fenner controversy will Science 18, 225^253.

2208
VEKSLER et al. EVOLUTION OF BASALTIC MAGMA

Fujii, T., Kushiro, I., Nakamura, Y. & Koyaguchi, T. (1980). A note on McBirney, A. R. & Naslund, H. R. (1990). The differentiation of the
silicate liquid immiscibility in Japanese volcanic rocks. Journal of the Skaergaard intrusion. A discussion of Hunter and Sparks (Contrib
Geological Society of Japan 86, 409^412. Mineral Petrol 95: 451^461). Contributions to Mineralogy and Petrology
Greig, J. W. (1927). Immiscibility in silicate melts. American Journal of 104, 235^240.
Science 13, 133^154. Morse, S. A. (1980). Kiglapait mineralogy II: Fe^Ti oxide minerals
Grove, T. L. & Baker, M. B. (1984). Phase equilibrium controls on the and the activities of oxygen and silica. Journal of Petrology 21,
tholeiitic versus calc-alkaline differentiation trends. Journal of 685^719.
Geophysical Research 89(B5), 3253^3274. Morse, S. A. (1990). The differentiation of the Skaergaard intrusion. A
Hanghj, K., Rosing, M. & Brooks, C. K. (1995). Evolution of the discussion of Hunter and Sparks (Contrib Mineral Petrol 95:
Skrgaard magma: evidence from crystallized melt inclusions. 451^461). Contributions to Mineralogy and Petrology 104, 240^244.
Contributions to Mineralogy and Petrology 120, 265^269. Morse, S. A. (1994). Basalts and Phase Diagrams. Melbourne, FL:
Hoover, J. D. & Irvine, T. N. (1978). Liquidus relations and Mg^Fe Krieger.
partitioning on part of the system Mg2SiO4^Fe2SiO4^ Naslund, H. R. (1983). The effect of oxygen fugacity on liquid immis-
CaMgSi2O6^CaFeSi2O6 ^KAlSi3O8^SiO2. Carnegie Institution of cibility in iron-bearing silicate melts. AmericanJournal of Science 283,
Washington Yearbook 77, 774^784. 1034^1059.
Hudon, P. & Baker, D. R. (2002). The nature of phase separation in Nielsen, T. F. D. (2004). The shape and volume of the Skaergaard
binary oxide melts and glasses. I. Silicate systems. Journal of Non- intrusion, Greenland: implications for mass balance and bulk com-
Crystalline Solids 303, 299^345. position. Journal of Petrology 45, 507^530.
Hunter, R. H. & Sparks, R. S. J. (1987). The differentiation of the Osborn, E. F. (1979). The reaction principle. In: Yoder, H. S., Jr (ed.)
Skaergaard Intrusion. Contributions to Mineralogy and Petrology 95, The Evolution of Igneous Rocks. Fiftieth Anniversary Perspectives.
451^461. Princeton, NJ: Princeton University Press, pp. 133^169.
Hunter, R. H. & Sparks, R. S. J. (1990). The differentiation of the Philpotts, A. R. (1976). Silicate liquid immiscibility: its probable extent
Skaergaard Intrusion. A reply. Contributions to Mineralogy and and petrogenetic significance. American Journal of Science 276,
Petrology 104, 248^254. 1147^1177.
Irvine, T. N. (1976). Metastable liquid immiscibility and MgO^FeO^ Philpotts, A. R. (1979). Silicate liquid immiscibility in tholeiitic
SiO2 fractionation patterns in the system Mg2SiO4^Fe2SiO4^ basalts. Journal of Petrology 20, 99^118.
CaAl2Si2O8^KAlSi3O8^SiO2. Carnegie Institution of Washington Philpotts, A. R. (1981). Liquid immiscibility in silicate melt inclusions
Yearbook 75, 597^611. in plagioclase phenocrysts. Bulletin de Mine¤ ralogie 104, 317^324.
Jakobsen, J. K., Veksler, I. V., Tegner, C. & Brooks, C. K. Philpotts, A. R. (1982). Compositions of immiscible liquids in volcanic
(2005). Immiscible iron- and silica-rich melts in basalt petro- rocks. Contributions to Mineralogy and Petrology 80, 201^218.
Philpotts, A. R. & Doyle, C. D. (1983). Effects of magma oxidation
genesis documented in the Skaergaard intrusion. Geology 33,
state on the extent of silicate liquid immiscibility in a tholeiitic
885^888.
basalt. AmericanJournal of Science 283, 967^985.
Jakobsen, J. K., Veksler, I. V., Tegner, C. & Brooks, C. K. (2007).
Roedder, E. (1951). Low-temperature liquid immiscibility in the system
Crystallization of the Skaergaard Intrusion from an emulsion of
K2O^FeO^Al2O3^SiO2. American Mineralogist 36, 282^286.
immiscible iron- and silica-rich liquids: Constraints from the melt
Roedder, E. (1979). Silicate liquid immiscibility in magmas. In:
inclusions in plagioclase. In: Harte, B. & Carpenter, M. (eds)
Yoder, H. S., Jr (ed.) The Evolution of Igneous Rocks. Fiftieth
Frontiers in Mineral Sciences 2007, Programme and Abstracts. Cambridge:
Anniversary Perspectives. Princeton, NJ: Princeton University Press,
The Mineralogical Society of Great Britain and Ireland,
pp. 15^58.
pp. 230^231.
Roedder, E. & Weiblen, P. W. (1970). Lunar petrology of silicate melt
Jang, Y. D., Naslund, H. R. & McBirney, A. R. (2001). The differentia-
inclusions, Apollo 11 rocks. In:. Proceedings of the Apollo 11 Lunar
tion trend of the Skaergaard intrusion and the timing of magnetite
Science Conference. Geochimica et Cosmochimica Acta, Supplement 1, 1,
crystallization: Iron enrichment revisited. Earth and Planetary Science
507^528.
Letters 189, 189^196. Roedder, E. & Weiblen, P. W. (1971). Petrology of silicate melt inclu-
Kilinc, A., Carmichael, I. S. E., Rivers, M. L. & Sack, R. O.
sions, Apollo 11 and Apollo 12 and terrestrial equivalents. In:
(1983). The ferrous^ferric ratio in natural silicate liquids Proceedings of the 2nd Lunar Science Conference. Geochimica et
equilibrated in air. Contributions to Mineralogy and Petrology 83, Cosmochimica Acta, Supplement 2, 1, 507^528.
136^140. Ryabov, V. V. (1988). Chemical composition of immiscible liquids in
Krasov, N. F. & Clocchiatti, R. (1979). Immiscibility in silicate melts natural glasses from traps. Soviet Geology and Geophysics 29(11), 62^70.
and its possible petrogenetic importance, as shown by study of melt Ryabov, V. V. (1989). Liquation in Natural Glasses: the Example of Traps.
inclusions. Transactions (Doklady) of the USSR Academy of Sciences 248, Novosibirsk: Nauka (in Russian).
92^95. Schmidt, M. W., Connolly, J. A. D., Gu«nther, D. & Bogaerts, M.
Longhi, J. (1990). Silicate liquid immiscibility in isothermal crystalli- (2006). Element partitioningçthe role of melt structure and com-
zation experiments. In: Sharpton,V.L. & Ryder, G. (eds) Proceedings position. Science 312, 1646^1650.
of the 20th Lunar and Planetary Science Conference. New York: Shearer, C. K., Papike, J. J. & Spilde, M. N. (2001). Trace element
Pergamon, 13^24. partitioning between immiscible lunar melts: An example from
McBirney, A. R. (1975). Differentiation of the Skaergaard intrusion. naturally occurring lunar melt inclusions. American Mineralogist 86,
Nature 253, 691^694. 238^241.
McBirney, A. R. (1989). The Skaergaard layered series: I. Structure Shelby, J. E. (2005). edn. Introduction to Glass Science and Technology, 2nd
and average compositions. Journal of Petrology 30, 363^397. edn. Cambridge: The Royal Society of Chemistry.
McBirney, A. R. & Nakamura, Y. (1974). Immiscibility in late-stage Shi, P. (1993). Low-pressure phase relationships in the system Na2O^
magmas of the Skaergaard intrusion. Carnegie Institution of CaO^FeO^MgO^Al2O3^SiO2 at 11008C, with implications for the
Washington Yearbook 73, 348^352. differentiation of basaltic magmas. Journal of Petrology 34, 743^762.

2209
JOURNAL OF PETROLOGY VOLUME 48 NUMBER 11 NOVEMBER 2007

Sparks, R. S. J. (1988). Petrology and geochemistry of the Loch Ba Veksler, I. V., Dorfman, A. M., Danyushevsky, L. M., Jakobsen, J. K.
ring-dyke, Mull (N. W. Scotland): an example of extreme differen- & Dingwell, D. B. (2006). Immiscible silicate liquid partition
tiation of tholeiitic magma. Contributions to Mineralogy and Petrology coefficients: implications for crystal^melt element partitioning and
100, 446^461. basalt petrogenesis. Contributions to Mineralogy and Petrology 152,
Stripp, G., Holness, M., Nielsen, T., Veksler, I. & Tegner, C. (2007). 685^702.
Reactive symplectite formation as tracers of late-stage liquids: the Veksler, I. V., Dorfman, A. M., Wirth, R. & Dingwell, D. B. (2007).
Skaergaard Intrusion. In: Harte, B. & Carpenter, M. (eds) Frontiers Unmixing of ferrobasaltic liquids: a study of nucleation, settling
in Mineral Sciences 2007, Programme and Abstracts. Cambridge: The and coalescence of emulsion droplets in centrifuged glasses. In:
Mineralogical Society of Great Britain and Ireland, p. 227. Harte, B. & Carpenter, M. (eds) Frontiers in Mineral Sciences 2007,
Tegner, C. (1997). Iron in plagioclase as a monitor of the differentia- Programme and Abstracts. Cambridge: The Mineralogical Society of
tion of the Skaergaard intrusion. Contributions to Mineralogy and Great Britain and Ireland, p. 213.
Petrology 128, 45^51. Vogel, W. (1985). Chemistry of Glass. Columbus, OH: American Ceramic
Toplis, M. J. & Carroll, M. R. (1995). An experimental study of Society.
the influence of oxygen fugacity on Fe^Ti oxide stability, phase Wager, L. R. & Brown, G. M. (1968). Layered Igneous Rocks. Edinburgh:
relations, and mineral^melt equilibria in ferro-basaltic systems. Oliver & Boyd.
Journal of Petrology 36, 1137^1171. Wager, L. R. & Deer, W. A. (1939). Geological investigation in
Thy, P., Lesher, C. E., Nielsen, T. F. D. & Brooks, C. K. (2006). East Greenland, Part III, The petrology of the Skaergaard
Experimental constraints on the Skaergaard liquid line of descent. intrusion, Kangerdluqssuaq, East Greenland. Meddelser om
Lithos 92, 154^180. Grnland 105(4).
Veksler, I. V. (2006). Crystallized melt inclusions in gabbroic rocks. Yoder, H. S., Jr & Tilley, C. E. (1962). Origin of basaltic magmas:
In: Webster, J. D. (ed.) Melt Inclusions in Plutonic Rocks. Mineralogical an experimental study of natural and synthetic rock systems.
Association of Canada Short Courses 36, 100^122. Journal of Petrology 3, 342^532.

2210
View publication stats

You might also like