Foley1988 The Genesis of Continental Basic Alkaline Magmas

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

The Genesis of Continental Basic Alkaline Magmas-

An Interpretation in Terms of
Redox Melting

by STEPHEN F. FOLEY
Max-Planck-1nstitut fur Chemie, Abt. Kosmochemie, Saarstrasse 23, 6500 Mainz, F.R.
Germany

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


ABSTRACT
Redox melting has recently been proposed as a major process for the generation of partial melts in
the upper mantle by interaction of upwelling CH 4 -rich fluids from a reduced deep asthenosphere with
relatively oxidized (/ O 2 «IW + 2 — 4 log units) overburden. Oxidation of CH 4 produces H 2 O and solid
carbon, and melting is caused by the increase in aHaO, which depresses the mantle solidus temperature.
An assessment is made here of the ability of the redox melting model to account for the great variety
of alkaline basic magmatism observed in continental regions. The hypothetical cases of a stable
continental cratonic region and a developing continental rift are treated for the cases of variable
degrees offluidintroduction and melting. Events caused by reduced fluid introduction into continental
regions are considered to form important contrasts to those outlined in general accounts of oceanic
redox melting processes, mainly due to the discontinuous and incipient nature of melting beneath the
continents, and the absence of a fluid-free melt layer. Important conclusions are: (1) In regions with
heat flow too low to cause partial melting, extensive amphibole and/or mica crystallization may occur.
(2) Initial melts will be CO 2 -rich, produced by either buffered CO 2 + H 2 O-rich fluids or by
disequilibrium redox melting, corresponding to kimberlitic, melilititic, or ultramafic lamprophyric
composition depending on pressure and tectonic setting. (3) More extensive melting will always be in
H 2 O-rich environments, leading to the volumetric predominance of basanitic-nephelinitic and
alkali-basaltic compositions in alkaline volcanic fields. (4) Episodic reactivation of melted regions will
produce more reduced and 'depleted' low A12O3, Na 2 O, and CaO melt types without the initial CO 2 -
rich melting phase; in cratonic regions lamproites, rather than kimberlites, should result. (5) Tensional
faulting in rift environments promotes enrichment at shallow levels by both solidification of melts and
by metasomatic introduction of carbonates produced by oxidation of reduced fluids around fractures.
(6) Partial melting of these enriched areas at a later stage of rift development occurs preferentially due
to the presence of water in: (a) minerals of the enriched zones; and (b) produced by the oxidation of
methane. Resulting melts may be highly alkaline (ultrapotassic) to high melt fractions, but this
alkalinity will be diluted by interaction with surrounding peridotitic rock prior to melt escape, so that
alkalinity should be broadly proportional to the density of enrichment. (7) The direct influence of redox
melting wanes with progressive rifting, and in the oceanic stages of rifting it is subordinate to
decompression melting.
Further assessment of the redox melting processes, particularly as regards mantle enrichment
events, awaits clarification of enrichment types from studies of nodules and alkaline volcanic rocks.

SCOPE
No concensus exists at present on the oxidation state of the mantle, with oxygen fugacity
estimates varying over five orders of magnitude between around FMQ to IW or below,
although the majority of authors favour the higher fOl part of this range (see discussions by
Arculus & Delano, 1981, Ryabchikov et al, 1981; Eggler, 1983,1987; Haggerty & Tompkins,
1983; Arculus, 1985; Woermann & Rosenhauer, 1985; Haggerty, 1986; Taylor & Green,
1986, 1987a; Green et al, 1987a; O'Neill & Wall, 1987).
[Journal of Petrology, Special Lithosphere Issue, pp. 139-161, 1988] © Oxford University Press 1988
140 S T E P H E N F. F O L E Y

A probable role for reduced COH fluids (CH 4 + H 2 O) is recognised in the formation of
diamonds (Deines, 1980; Taylor, 1987a), and, coupled with the refractory nature of mineral
assemblages associated with diamond, this indicates that at least some parts of the lower
lithosphere in cratonic regions are of reduced, geochemically depleted nature (Haggerty &
Tompkins 1983).
A widespread mechanism for magma production in the upper mantle termed redox
melting and controlled by the introduction of reduced gases from the lower mantle by the
remnants of primordial degassing has recently been outlined by Taylor (1985), Taylor &
Green (1986) and Green et al. (1987a,/?). In this hypothesis, melting of upper mantle
peridotites is triggered by oxidation of reduced C-O-H fluids (CH 4 + H2), producing water
which causes the solidus temperature of the peridotite to drop sharply, thus initiating
melting caused by volatile introduction rather than temperature increase. Operation of the

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


redox melting mechanism depends on the existence of a reduced deep asthenosphere
( / O 2 ^ I W + 1-2 log units) overlain by more oxidized overburden, so that oxidation of CH 4
and H 2 to H 2 O and solid carbon will occur at the interface. The general application of the
hypothesis to natural magmas in terms of melt compositions and tectonic settings has been
made for oceanic and continental regions by Green et al. (1987a,fr).Preliminary assessment
of the hypothesis applied to continental ultrapotassic rocks has been made by Foley et al.
(1986), and to lamproites and kimberlites by Foley (1986, 1987a, b). The purpose of this
contribution is to investigate in more detail the possible extent and influence of redox
melting in producing both: (i) the wide range of alkaline rocks seen in stable cratonic areas
and in continental rift regions; and (ii) mantle enrichment events believed to be represented
in xenoliths and suggested to have occurred from studies of alkaline whole-rock and mineral
chemistry.
Because the redox melting model rests on the proposition of a reduced deep astheno-
sphere overlain by more oxidized overburden, it is at first sight the opposite of the model of
Haggerty & Tompkins (1983) and Haggerty (1986), in which a reduced cratonic lower
lithosphere overlies an oxidized asthenosphere. The reason for the difference between the
models lies in the evidence for deep mantle oxidation states which they use: the Haggerty &
Tompkins model relies on relatively oxidized fOl estimates for high temperature xenoliths
and megacrysts in kimberlites from cratonic regions, whereas the redox melting model
couples evidence from cratonic regions and oceanic regions, and is thus more globally based.
The evidence for a reduced deep mantle is recent and controversial, and is reviewed here
since it runs contrary to more popular models for mantle fOr The petrogenetic models
offered in later sections of this paper are intended to illustrate an alternative rather than a
replacement for existing models—the future of the different models awaits more evidence for
mantle oxidation state, particularly from degassing at ocean ridges, and from diamond
inclusion studies.

EVIDENCE FOR THE REDUCED UPPER MANTLE


Studies of both rock and water compositions at mid-ocean ridges (MOR) and mantle
derived nodules in basic volcanics have recently provided evidence for/ 02 conditions in the
mantle more reduced than those corresponding to FMQ or the carbonate-bearing peridotite
assemblages EMOG or EMOD (Eggler & Baker, 1982). Analyses of mid-ocean ridge basalts
(MORB) showed them to have elevated contents of 3 He, which is believed to be primordial
in origin (Lupton & Craig, 1975), and to contain reduced carbon species (CO and CH 4 plus
rarer higher hydrocarbons; Zolotarev et al, 1979a, b; Byers et al., 1983, 1984). Ocean water
above MOR was found to be rich in 3 He, and this enriched water forms extensive plumes
centred on MORs which can be detected thousands of kilometres from the ridges (Lupton &
THE GENESIS OF CONTINENTAL BASIC ALKALINE MAGMAS 141

Craig, 1981; Lupton, 1983). These plumes are also enriched in methane (Welhan & Craig,
1979), and excellent correlations exist between 3 He, CH 4 content and temperature anomal-
ies at hydrothermal sites along the ridges (Kim, 1983). 5l3C measurements in CH 4 of — 15 to
— 20 are further evidence that part of the methane in hydrothermal systems is mantle derived
(Welhan & Craig, 1983; Taylor, 1985). Biogenic and thermogenic methane are isotopically
light (<513C = - 8 0 to - 4 0 ; Claypool & Kvendolen, 1983), whereas Kaminskiy et a/.'s (1976)
estimate for <)13C in mantle methane is approximately — 15. All these lines of evidence are
needed to assign a mantle origin to CH 4 , since there is no doubt that biogenic methane is
present in hydrothermal systems, and in many cases dominates (e.g., Belviso et al, 1987;
Merlivat et al, 1987).
Christie et al. (1986) have analysed separated glasses from MORBs for FeO and total iron
and, comparing them with whole-rock iron analyses obtained by Carmichael & Ghiorso

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


(1986), they found the glasses to be significantly more reduced, representing oxygen
fugacities on average two orders of magnitude below the whole-rocks and with very little
overlap (Fig. 1). The association of elevated primordial 3 He in MORBs and reduced MORB
glasses with correlated 3 He, non-biogenic CH 4 and temperature anomalies in MOR
hydrothermal systems forms the strongest evidence that oxidation states in the mantle are
well below FMQ.
The discrepancy between glass and whole-rock FeO analyses (Fig. 1) also demonstrates
that much oxidation takes place at a very late stage. Calculations modelling oxidation in
COH fluids have emphasized that oxidation of a magmatic fluid between IW and FMQ or
NNO can be achieved by surprisingly small amounts of hydrogen loss or oxidation of
carbon species, and that these processes occur dominantly under near-surface conditions
(Sato, 1978; Mathez, 1984; Foley et al, 1986). Rapid near-surface oxidation implies that
evidence for mantle CH 4 in hydrothermal systems may be disproportionately important due
to its low survivability. In continental regions, where the depth of oxidized overburden is
much greater, 3 He anomalies may be the only surviving indication of reduced fluid
degassing.
Accumulation of data on He isotopes in oceanic regions has shown that MORs are quite
uniform in their helium isotopic ratio (expressed as /?/KA = ( 3 He/ 4 He sample )/( 3 He/ 4 He atm )) at
about 7-9 (summarized in Lupton, 1983). MORB glasses indicating low oxygen fugacities
exist in both the Pacific and Atlantic oceans (Fig. 1; Byers et al, 1986; Christie et al, 1986)
and so are taken to be widespread. Minor systematic differences have been noted by Byers et
al (1984) at the Galapagos spreading centre, but these are small relative to the range of
oxygen fugacities under discussion. A number of Atlantic glasses lack the reduced carbon
species, but this may be attributed to shallow level degassing (Byers et al, 1986).
In contrast to the uniformity of MOR 3 He measurements, 3 He in ocean waters and
volcanic gases associated with ocean islands is in most cases similar to MORs, but in about a
third of the measured localities is much more enriched, ranging up to more than 30 RA (Kurz
et al, 1983; Rison & Craig, 1983; Craig et al, 1984). These occurrences were modelled by
Green et al (1987b) as concentrations of primordial degassing, i.e. 'plumes' of reduced fluids,
which may also occur in some 'hotspots' beneath continental regions. Subduction zone
magmatism is also associated with uniform 3 He signatures of about 5-6 RA (Craig et al,
1978; Lupton, 1983) indicating that primordial fluid introduction is also associated with
subduction zones, since the old oceanic crust would be mostly outgassed and could
contribute only about 0.1 RA (Lupton, 1983). The suggestion that subduction zones are a
focus of degassing from the deep mantle (Green et al, 1987a, b) is consistent with conclusions
from studies of volcanic geochemistry of island arcs (Foden & Varne, 1980; Wheller et al,
1987), which suggest involvement of a mantle-derived component.
In continental regions evidence for a reduced mantle is much harder to find, which is
142 STEPHEN F. FOLEY

GLASS AVG. ROCK AVG.


(n = 44) MW FMQ (n = 86)
10

ATLANTIC

6-

4-

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


2-

-3.0 -2.0 -1.0 0.0 1.0


10 -, GLASS AVG. ROCK AVG.
(n = 31) MW FMQ (n=61)

PACIFIC

6-

4-

2-

ANNO -1.0 0.0 1.0


-3.0 -2.0

FIG. 1. Frequency histograms of the oxidation state of MORB glasses and whole rocks as estimated from
F e 3 + / F e 2 + (after Christie et al, 1986). S denotes intrinsic fOl measurements for the Skaergaard layered intrusion.
ANNO is fOl difference from the nickel-nickel oxide buffer in log units.

consistent with the depth of oxidised overburden being much greater. If the/ O 2 below the
depth of origin of kimberlites (approximately 200 km) is of the order of IW+ 1 —2 log units,
then evidence for the reduced state must survive unoxidized an emplacement through 200
km of overburden, as against 0 2 km in the case of MORs. Xenoliths sampled by the melts
must come from the overlying, predominantly oxidized layer, and this is reflected in/ O 2
estimates from mineral equilibria (Eggler, 1983; O'Neill & Wall, 1987). Oxygen fugacities in
the xenolith samples are further constrained to lie above IW, since below this/ O2 , Ni-rich
metal should precipitate, causing much lower Ni contents in silicate minerals than are
observed (O'Neill & Wall, 1987). Intrinsic fo2 measurements on xenoliths giving values
around IW (e.g., Arculus & Delano, 1981) may be, due to excess carbon giving false fOl
readings (Mattioli et al, 1987).
T H E G E N E S I S O F C O N T I N E N T A L BASIC A L K A L I N E M A G M A S 143

Nevertheless, some evidence for reducing conditions at deeper levels beneath continental
regions can be found in inclusions in diamonds. Although many diamonds are believed to
originate in the lower lithosphere (Haggerty, 1986), others have higher pressure equilibrated
mineral inclusions and may come from much deeper levels. Moissanite (SiC), which is stable
only at fOl substantially lower than IW, and magnesiowustite/ferropericlase have been
described as inclusions in diamonds by Moore et al. (1986), and these have mineral
associations (garnets with substantial clinopyroxene in solid solution) indicative of pressures
in the region of 100 kb (Moore & Gurney, 1986; Moore et al, 1986). These are thus believed
to be samples from the deep mantle which were emplaced above the kimberlite melt
generation depth by convective overturn (Moore & Gurney, 1986). They presumably owe
their survival to the strength of diamond as a pressure container.
For the redox melting model (Taylor, 1985; Taylor & Green, 1986) it is proposed, by

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


analogy with the evidence summarized above for oceanic regions, that deep levels of the
asthenosphere are more reduced than the/ O2 represented by the xenolith population, which
originates at shallower levels. It can be seen that the discrepancy between this model and the
oxidized asthenosphere model of Haggerty & Tompkins (1983) and Haggerty (1986) is not as
large as at first appears, since the asthenosphere/ 02 estimates for the latter model come from
xenoliths and megacrysts in kimberlite. Redox melting should occur at the depth where
reduced fluids first encounter oxidized material (see next section); the overlying, predomi-
nantly oxidized region is not homogeneous in its oxidation state, but contains reduced
regions (Green et al, 1987a, b).

THE REDOX MELTING MECHANISM


The redox melting hypothesis was formulated by Taylor (1985) from: (i) theoretical
calculations of COH fluid speciation at high pressures and temperatures at variable oxygen
fugacities (Holloway, 1981; Taylor, 1987b); and (ii) evidence for a more reduced mantle than
the / O z « FMQ commonly accepted for the upper mantle. Results of the fluid speciation
calculations can be represented on a plot of/O2 against Xc (Fig. 2) which is a measure of the
C/(C + H 2 ) ratio normalised to 100, and is thus analogous to the more familiar mg-number
calculation. The carbon saturation surface (CSS) has a maximum XHlO value at/ O z = IW + 1
— 2 log units which separates a region of reduced fluids (CH 4 + H 2 O + H 2 ) from a region of
oxidized fluids (H 2 O + CO 2 ). Figure 3 depicts the distribution of individual fluid species on

Graphite Saturation Curve in C-O-H

10 20 30 40 50 60 70 80 90 100
Fluid phase 100 C / (C+H 2 )

FIG. 2. Fluid compositions coexisting with the graphite saturation curve in the system COH (Taylor, 1985; curve
shown for 30 kb, 1600 K). Heavy arrows show the approximate positions of major fluid species. IW = iron-wustite
buffer. Reduced fluids at (A) will consist of CH 4 + H 2 O, and on encountering oxidized regions will oxidize towards
(B). After (B), H 2 O and C are produced from the oxidationof CH 4 and fluids will proceed towards the XUlO
maximum at (C) with continued oxidation.
144 STEPHEN F. FOLEY

IW

20 30 40 50 60 70 80 90
% fluid species

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


FIG. 3. Variation in abundance of the fluid species with changing oxygen fugacity at conditions corresponding to
those of Fig. 2. At IW, CH 4 > H 2 O, whereas two log/ O2 units higher at CW, H 2 O accounts for > 90 mol.% of the
fluid. At higher/ O2 corresponding to FMQ or EMOG, fluids will be CO 2 + H 2 O mixtures. Note that CH 4 and CO 2
coexist at equilibrium only where both are very minor consitituents of the fluid phase.

the CSS over the range of/Oz corresponding to Fig. 2, showing that the carbon species
involved are dominated by CH 4 at oxygen fugacities below the H 2 O maximum and by CO 2
at oxygen fugacities above the H 2 O maximum. It is this 'nose' of high XH2O fluids that forms
the theoretical basis for the redox melting model: any fluid such as A in Fig. 2 which oxidizes
due to interaction with its surroundings during ascent will eventually become carbon
saturated at B and its composition will then proceed along the CSS towards C, according to
the reaction

increasing its H 2 O content and precipitating solid carbon as it goes. The increasing OLHJ0 will
depress the solidus of the peridotite and may eventually result in melting at some point
before C is reached.
For most cases the intrinsic oxygen fugacity of the upper mantle peridotite which plays
host to the invading reduced fluids will be close to that of the upper surface of the CSS, as
indicated by the nodule data noted above. The^-composition trace A-B-C in Fig. 2 refers
only to the fluid composition. Once melting starts it is unlikely that in most circumstances
sufficient volatiles will be present to saturate the melt in a fluid phase. Thus, in its simplest
form, the redox melting model can be envisaged as a layered model with/ O2 decreasing with
increasing depth: the/ O2 of the XHlO maximum corresponds to a fluid-free melt region, with
reduced fluids present below and either more oxidized fluids or hydrous and/or carbonated
peridotites present above (Green et al., 1987a; Taylor & Green, 1987/)). The oxidized layer
will in reality be a mixture of oxidized and more reduced regions.
An important aspect of the model is that experimental investigations of volatile solubility
mechanisms can be used to show that a variety of oxidation states can be borne by the melt
phase in the fluid absent region. In the/ O2 region just above the XHlO maximum CO 2 in the
fluid will dissolve as COl~ groups balanced by mono- or divalent cations (Brey & Green,
1975; Eggler & Rosenhauer, 1978), whereas in the/ Oz region below the XHlO maximum CH 4
will be the dominant carbon species and will dissolve as elemental carbon balanced by
structural changes in the silicate network resulting in an increase in Si/O of the network
(Taylor & Green, 19876).
Figure 4 summarises the simplest picture of the redox melting hypothesis which probably
holds better for oceanic areas than continents because of the more marked redox
THE GENESIS OF CONTINENTAL BASIC ALKALINE MAGMAS 145

Continent
Mid Ocean Oceanic
R/RA up to 5
Ridge Island
craton
'';~J£*2&&±'i'", " CrUSt ^ ' ^ 7 ^ ^ ^ ^ ^ ;

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


CH

DIW±1 ®-IW*(1-2) D 3 E3-IW-4 51 Reduced fluid


Increasing f 0 — > • i!Oxidation zone —^partial melting

FIG. 4. Cartoon depicting the essential ingredients of the redox melting model as developed and discussed by
Taylor (1985), Taylor & Green (1986), and Green et al, (1987a, b). Upwelling reduced fluids are detected at mid
ocean ridges as 3 He correlated with non-biogenic CH 4 , but undergo oxidation at much deeper levels in other
geological environments so that CH 4 is not detected at the surface. R/RA = typical helium isotopic signatures for
fluids from the various tectonic environments (see text for data sources). Oxidation of CH 4 produces H 2 O
which triggers melting at the oxidation front. Interaction between CH 4 and oxidized subducted material may
produce water which affects melting relations in subduction environments and beneath some ocean islands
(Green et al. 19876).

inhomogeneity expected under the continents. In its purest form, the redox melting model
requires only input of volatiles and not an increase in heat flow, although the two are in
reality likely to go together as may be indicated by the high 3 He anomalies associated with
'hotspots' such as Hawaii and Iceland (Kurz et al., 1983, 1985).
The supply of volatiles, i.e. localization of degassing, coupled with the heat flux, is critical
in controlling melting. In continental regions both will be more variable than in oceanic
regions, so that melting may not occur in some areas of fluid introduction, but may be
widespread in others, such as rift zones. Dependence of melt character on continuity or
episodism of fluid introduction, and on the abundance of fluids has been suggested by Green
et al. (1987a, b) and Foley (1987a, b), and the coupling of progressive reduction with
geochemical depletion of the mantle has been stressed (Foley, 1^86; Taylor & Green, 1986,
1987a). Taylor & Green (1986) suggested that it is this reduction-depletion process that
leads to the association of diamond with the depleted garnet harzburgite paragenesis.
Compositions of fluid inclusions in diamond are compatible with precipitation of diamond
on both the oxidized and reduced sides of the carbon saturation surface (Taylor, 1987a),
showing that reduction processes may be important in their origin.

GEOPHYSICAL ANOMALIES
The seismic low velocity zone (LVZ) and the electrical high conductivity layer (HCL) are
important geophysical anomalies for understanding upper mantle processes. Although the
146 S T E P H E N F. FOLEY

most widely accepted interpretation of these anomalies is a partial melt layer (Anderson,
1962; Green & Liebermann, 1976), alternative explanations for the HCL have been put
forward, namely: (i) a water-rich fluid layer (Tozer, 1981); (ii) a layer including solid carbon
(Duba & Shankland, 1982); and (iii) a layer rich in hydrous minerals (Waff, 1974).
If the redox melting mechanism operates, all four of these features should result, and a
combination of them may explain the magnitude of the anomalies which may otherwise be
difficult to explain by any of the mechanisms alone because of the requirement for
interconnectivity for electrical conductivity. Carbon is produced by oxidation of CH 4 , and is
probably present as a thin film on crystal surfaces (Mathez et al., 1986). The water produced
exists initially as fluid and promotes partial melting. Hydrous phases would be formed both
above and below the melt layer (if present) since the stability of hydrous phases is greatest at
intermediate aHlO (Holloway, 1973).

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


The localization and intensity of magmatic activity at MOR would tend to concentrate
the mechanisms suggested to cause the geophysical anomalies into a relatively closely
defined layer, which may account for the better definition of the anomalies in oceanic
regions. In continental regions, anomalies are more poorly denned and at deeper levels
except in regions of magmatic activity such as rift zones, where different geophysical
anomalies again coincide (Banks & Beamish, 1979; Cook et al., 1979).
In continental areas with lower geothermal gradient, the poor definition of geophysical
anomalies is suggested to be due to either the lack of one or more of the four contributing
factors, or to the smudging out of these contributors over a much larger depth range during
the complex history of events leading to the development of cratonic regions.

THE RANGE OF ALKALINE MELT COMPOSITIONS IN THE MANTLE


Primary melt compositions of alkaline volcanics
Geochemical and experimental work on the genesis of basaltic magmas from upper
mantle peridotitic source compositions has outlined a petrogenetic grid (Green, 1970, 1971)
from which a range of melt compositions may be simplistically represented as a fun'ction of
the degree of partial melting. In addition to the degree of melting, alkaline melt compositions
will be particularly affected by the abundance and composition of volatile components, by
geothermal gradients and so the pressure and temperature of melting, and by variations of
the source mineralogy and composition (i.e. the limitations of the peridotite melting model).
In oceanic regions, the alkaline, low-silica end of the 'basalt spectrum' includes nephelini-
tes and melilite nephelinites which are found on oceanic islands (Clague & Frey, 1982), but
carbonatites, which occur rarely on oceanic islands, are not considered to be primary melts
(e.g., Stillman et al, 1982). Alkaline rocks in continental regions have some similarities to
those in oceanic regions as shown by the Cameroon Line which includes similar alkaline
rocks emplaced through both oceanic and continental crust (Fitton & Dunlop, 1985), as well
as specialities such as in cratonic and rift environments. As the degree of partial melting
decreases, volatiles and mantle enrichment events (MEE) play a greater role, and are
probably essential in the genesis of the most alkaline low-silica rocks such as kimberlites,
lamproites, olivine melilitites, alkaline and ultramafic lamprophyres, and carbonatites.
Anomalously high incompatible element contents may occur even in picritic rocks which
probably represent large degrees of partial melting (Bristow, 1984), emphasising that MEE
may play a role in the genesis of rocks throughout the 'basalt spectrum'.
Many regional studies of alkaline rocks which attempt to define primary or parental
magmas for volcanic fields have led to conclusions that varieties of basanite, olivine
THE G E N E S I S OF C O N T I N E N T A L BASIC ALKALINE MAGMAS 147

nephelinite, or melilitite are the most likely candidates (e.g., Spencer, 1969; Wimmenauer,
1974; Ferguson & Cundari, 1975; Frey et al, 1978; Phelps et al, 1983; Pouclet et al, 1984;
Duda & Schmincke, 1985), but that more than one primary magma similar enough to merit
the same name may be involved (e.g., basanites; Ferguson & Cundari, 1975, or nephelinites;
LeBas, 1978a).
In other areas an ultramafic magma type either associated with carbonatites or
originating in a carbonate-rich environment is invoked (Upton, 1974; Currie, 1976; Brey,
1978; Rock, 1986). Amongst these, olivine melilitites and kimberlites, which may form a
continuum by variation of pressure of melting (Brey, 1978), are the best studied, but primary
ultramafic lamprophyre compositions probably also exist (Malpas et al, 1986; Rock, 1986).
Relationships between such ultramafic lamprophyre compositions and kimberlites, meliliti-
tes, and possibly rift-related ankaratrites (Rock, 1986) remain to be investigated. Primary

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


lamproitic magmas appear to require CO 2 -poor conditions of origin (Jaques et al., 1984).
Recognition of primary magmas is hampered by uncertainties in the original volatile
contents of many rocks. For example, alkaline lamprophyres are possibly equivalent to
volatile-rich basanites (see Rock, 1984). Van Bergen et al. (1983) describe minette inclusions
as potassic basanites which were unable to lose water, and the existence of primary minettic
magmas is suggested from other studies (Roden, 1981; Esperanga & Holloway, 1986).
Recent fluid dynamic modelling suggests that variations in the degree of partial melting
may continue to be important at very low melt fractions, although the compositional
expression of such variation is obscure at present (McKenzie, 1984). Alkalies and volatiles
have the effect of lowering melt viscosities (Bottinga & Weill, 1972; Scarfe, 1986), so that
volatile-rich alkaline melts will be more easily extracted at low melt fractions. Further
research in these fields should assist in determining primary melt compositions.

MEE and the composition of non-volcanic alkaline magmas


Summarizing primary alkaline melt compositions is a more extensive problem than
merely considering rocks seen at the Earth's surface. Magmas originating as low degree
partial melts in continental regions will experience great difficulty in finding a path to the
surface, particularly in cratonic regions. This difficulty expresses itself as the concentration
of alkaline rock occurrences around fundamental crustal fracture systems (Arseneyev, 1963;
Dawson, 1971), which are believed to have facilitated emplacement of the magmas. Melts
which solidify at depth, serving as mantle enrichment events should therefore be considered.
Magmas which solidify at mantle depths will probably form vein systems with high aspect
ratio (Waff, 1980; Bailey, 1984) rather than solidifying in magma chambers.
Probable samples of such magmas are provided by amphibole- or mica-bearing nodules
in alkali basalts (see also Stolz & Davies, this volume), and MARID nodules and megacrysts
in kimberlites (see also Waters & Erlank, this volume); these potentially provide important
information on a number of mantle processes. Xenoliths in alkali basalts carry a number of
features consistent with an origin by crystallisation of melt at high pressures: (i) low mg-
numbers of pyroxenes and the presence of hydrous minerals and apatite indicate fraction-
ation sequences (Wass, 1979; Irving, 1980; Menzies, 1983); (ii) composite xenoliths show
veins of material rich in hydrous minerals restricted to 10-15 cm width (Irving, 1980;
Wilshire et al, 1980); (iii) Alvi/Aliv in pyroxene and Ti/Al in amphibole are compatible with
high pressure crystallization (> 15 kb; Wass, 1979). The abundance of such nodules supports
the interpretation of their origin as conduit linings, whereas the variety attests to the range of
volatile-enriched melts involved. Megacrysts and minerals in MARID nodules in kimberli-
148 S T E P H E N F. F O L E Y

tes also have ranges of mg-numbers and have been suggested to be crystallization products
of 'protokimberlite' melts (Dawson & Smith, 1977; Harte, 1983).
However, complex histories for such xenoliths need to be unravelled, as the melts may be
modified by processes such as flow crystallization (Irving, 1980) and exchange of compo-
nents with the wall rocks. In the present context, though, it is pertinent to consider which
type of magmas they may represent, and thus the likely conditions of origin of these magmas.
Since the nodules may bear evidence from the very smallest melt fractions, they may include
information about magma types which never reach the surface as volcanics. The only major
distinction which can be made at present is between nodules with Al-rich clinopyroxene and
amphibole and those with Al-poor assemblages, including Al-poor clinopyroxene and
richteritic amphiboles. The former are generally attributed to basanitic melts (Irving, 1980;
Wass, 1979; Menzies et al., 1985), whereas the latter may be crystallization products of

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


protokimberlite or lamproitic melts (Dawson & Smith, 1977; Waters, 1987). Further
research may distinguish more types from the minor chemical fluctuations seen in other
nodules (e.g., Wilshire et al, 1980).

REDOX MELTING IN THE SUBCONTINENTAL UPPER MANTLE


Operation of the redox melting mechanism requires pre-existence of areas of different
oxidation states. In the discussion that follows continental crust and subcontinental
lithosphere will be assumed to be in the late stages of development prior to the development
of alkaline volcanism. The oxidation state of the subcontinental upper mantle (SCUM;
Eggler, 1986) is taken to be generally of the order of EMOG or QFM as indicated by
intrinsic/ O2 measurements on mantle mineral assemblages. Pressure estimates ofss 10 kb for
granulite terrains indicate that well developed and heterogeneous crust existed during
Archean times (Ellis, 1983). The development of these regions in the Earth's early history will
not be discussed here, since some precursory development of enriched mantle appears to be
necessary for many alkaline rock types. This is indicated by the restriction of strongly
alkaline rocks to relatively recent periods of Earth history: the oldest ages for carbonatites
are 2-6 Ga (Sage & Watkinson, 1986), kimberlites 1 -6 Ga (Bristow et al., 1986) and lamproites
1-2 Ga (Scott, 1981), but most occurrences of each of these types are considerably younger.
The consequences of reduced fluid introduction will be outlined here for two contrasting
cases; cratonic regions, in which episodic melting events may effect the same mantle region
leading to a variety of melt types produced from a single area, and rift regions, in which the
tensional regime will promote access of fluids and melts to progressively higher levels as the
rift develops. Volcanism in other continental regions can be compared to one or other of
these extreme cases.

Cratonic regions
The likely effects of reduced fluids and consequent melts in cratonic regions can be
demonstrated by considering progressive fluid introduction into a simple layered model for
the upper mantle with an oxidized SCUM overlying a reduced deep asthenosphere (Fig. 5).
The oxidation state of the SCUM is in detail more complex, containing more reduced
regions at lower lithosphere levels (Haggerty, 1986), but the layered model serves as a useful
schematic guide, because interaction with upwelling reduced fluids would occur where
oxidized material is first encountered.
As discussed briefly by Foley (1987c/), the exact behaviour will depend on whether fluid
supply is episodic or continuous, and on the volumes of fluid introduced. The geothermal
THE GENESIS OF CONTINENTAL BASIC ALKALINE MAGMAS 149

STABLE CONTINENTAL REGION

CRUST

OXIDISED
LITHOSPHERE

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


Reduced,
depleted,
REDUCED COj-absent
,ASTHENOSPHERE melting

H 2 0 - rich fluid Parfial melf zone


Mica, amphibole - bearing zone
CH, * H 9 0 fluid Mantle enrichment event
from solidification of melt
• due to H20 from oxidised ChL-

FIG. 5. Redox melting processes in a stable continental region. (A) Oxidation of reduced fluids at very low
geothermal gradients results in extensive hydrous mineral development; (B) Initial melting is either in the presence
of buffered CO 2 + H 2 O fluids, or is disequilibrium melting in the presence of CH 4 and carbonates, but in either case
gives rise to low-silica, CO 2 -rich magmatism such as kimberlite or melilitite: (C) More extensive melting produces
larger volumes of H 2 O-rich magmatism such as basanite or alkali basalt (ii). Melts produced at the leading edge of
the oxidation front remain CO2-influenced and may resemble nephelinites or lamprophyres (i). Magmas failing to
reach the surface will cause basanitic MEE seen as nodules in alkali basalts and basanites. (D) During reactivation
at a later stage, H 2 O may induce melting in a reduced, depleted region of the mantle, and melts may resemble
lamproites if this region has experienced intervening incompatible element enrichment. MEE produced from these
melts will bear a characteristic low Al 2 O 3 signature. The lithosphere may be of more variable oxidation state than
depicted in these schematic cartoons (Figs. 5-7), and the oxidation front may not correspond to the litho-
sphere/asthenosphere boundary, particularly in cratonic regions.

gradient will exert further control, determining whether or not production of water by
oxidation of the fluids will cause melting.
Figure 5A depicts the situation for low geothermal gradients, where CH 4 -rich fluids will
oxidize precipitating carbon and producing H 2 O-rich fluids (see Fig. 2). Hydrous phases will
also precipitate in the zone of oxidation, but a water-rich fluid is likely to persist since only
small amounts of H 2 O are taken up in hydrous phases, and the amount of these is limited by
the K content (Eggler & Holloway, 1977; Wyllie, 1978). The surviving H 2 O-rich fluid will
invade the overlying layer and eventually crystallize in metasomatic reactions producing
hydrous phases. The precipitation of hydrous phases could occur at all levels in cratonic
lithosphere: the presence of fluorine will stabilize amphibole and mica to greater pressures
than those commonly quoted from fluorine-free experiments (Holloway & Ford, 1975; Foley
et ai, 1986).
With the introduction of more fluid and/or a higher geothermal gradient, increasing aHzO
will eventually cause partial melting (Fig. 5B). Since the H 2 O produced at the oxidation
front will migrate rapidly upwards, melting will occur within the overlying oxidized zone.
The nature of melts produced will be a function of the lithosphere composition and the rate
150 S T E P H E N F. F O L E Y

of oxidation of the fluid. Taking oxygen fugacities indicated by mantle mineral assemblages
to be typical of SCUM/ O2 , carbonates are likely to be present in the subsolidus assemblage
(Eggler, 1975; Kushiro et al, 1975). If the H 2 O + CH 4 fluid moves slowly and oxidizes
thoroughly to a H 2 O-rich composition, melting should occur in the presence of H 2 O + CO 2
fluids and will be restricted to low silica compositions which may be kimberlitic at higher
pressures and melilititic at lower pressures (Brey, 1978). It remains to be seen whether
carbonatitic melts are represented amongst these low degree partial melts; these have been
suggested from a number of simple system experimental and theoretical studies (e.g., Wyllie
& Huang, 1976; Wendlandt & Eggler, 1980), but support from experiments on natural
compositions is lacking. The melt composition should be buffered (Eggler & Holloway,
1977; Eggler, 1978; Wyllie, 1978) and the amount of low silica melt produced will depend on
the buffering capacity of the carbonate, which is in turn dependent on which carbonate is

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


present (magnesite has a much greater buffering capacity than dolomite; Olafsson & Eggler,
1983).
In the case of rapid movement of fluid, or where sharp boundaries in the oxidation state of
the SCUM exist, the fluid encountering the carbonate-peridotite will still contain appreci-
able CH 4 . Carbonate should be rapidly consumed by decomposition to graphite or diamond
by reactions such as
MgSiO 3 + MgCO 3 + C H 4 ^ Mg 2 SiO 4 + 2C + 2H 2 O.
Water introduced by the reduced fluid and from the reaction of CH 4 with carbonate may
cause aH2O to be high enough for melting to commence while carbonate persists. In this
disequilibrium process, restricted amounts of low silica melts may be produced before
continued CH 4 introduction causes all the carbonate to react out. This interpretation
contrasts with that of Taylor & Green (1986, 1987a) and Foley (1987a) for equilibrium
conditions, in which carbonate is expected to be exhausted slightly before the onset of
melting (see Taylor & Green, 1986, fig. 2). After exhaustion of the carbonate, more
voluminous melts will be higher silica and H 2 O-rich in character, probably basanitic or
nephelinitic, but should not resemble kimberlite or melilitite which require involvement of
CO 2 in their genesis (Brey & Green, 1975; Brey, 1978; Wyllie, 1979, 1980).
If melts produced at this stage do not reach the surface (Fig. 5B(ii)), they will result in sills
or vein systems containing hydrous phases and possibly carbonates. At pressures lower
t h a n * 18 kb, carbonates are unstable in peridotites and free CO 2 will be produced which
will then be able to metasomatise wall rocks more distant than those affected by exchange
with the melt (Menzies et al., 1985). Solidification of CO 2 -rich, H 2 O-bearing melts may
therefore result in a solid vein assemblage superficially resembling that produced by
crystallization of H 2 O-rich melts.
Continued increase in the amount of water produced at the oxidation front will cause
more extensive melting and may lead to the development of a partial melt-bearing layer with
gradational characteristics (Fig. 5C). The scenario outlined for Fig. 5B will apply to the
leading edge of this zone (Fig. 5C(i)), at which carbonate-bearing peridotites may be
encountered in the country rocks. With more water available, melting can occur at two levels
of this variably oxidized region: under relatively oxidizing conditions as described above,
and also at a deeper, more reduced level corresponding to the intersection of the peridotite
solidus with the underside of the X HzO maximum in the carbon saturation surface (Taylor &
Green, 1986). Melts produced under the latter conditions will tend to be more silica-rich due
to the depolymerizing effect of H 2 O + CH 4 volatile mixtures on melts (Taylor & Green,
1987b), and resemble basanitic or nephelinitic compositions (Fig 5C(ii)). A mantle layer with
laterally heterogeneous oxidation states may produce both types of melt at a single level.
THE GENESIS OF CONTINENTAL BASIC ALKALINE MAGMAS 151

At higher degrees of partial melting the melt pockets will coalesce and a combined melt
layer will exist which is comparable to the fluid-free melt layer described for oceanic regions
by Green et al. (1987a, b). Such a melt layer is capable of supporting a (predominantly
vertical) gradation of oxidation states due to the differing solution mechanisms of CH 4 and
CO 2 in silicate melts: CH 4 dissolves as reduced carbon with complementary reduction of the
silicate network (Taylor & Green, 1987b) whereas CO 2 dissolves as discrete carbonate
groups (Brey, 1976). Therefore the lower oxidation state can be transmitted vertically
through the melt layer and cause progressive reduction at higher levels. Once a melt layer is
established, melting of the overlying oxidized region will be of the 'oxidized' type described
by Taylor & Green (1986) in the presence of H 2 O > C O 2 fluids. Reduced fluids from below
will no longer have access to the carbonated peridotite unless the exceptional case of fluid
saturated melt exists.

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


Partial melting environments producing very low-silica compositions in CO 2 -rich
conditions therefore seem to be restricted to very low degrees of partial melting where
melting of carbonate-bearing rocks is caused by direct access of fluids. The redox melting
model thus predicts that in relatively stable continental areas any voluminous alkaline
volcanism should be of H2O-rich character, which is consistent with alkaline provinces where
melilitites are subordinate to basanites, nephelinites, and alkali basalts.
Melts from water-rich regions of melting which do not reach the surface (Fig. 5B) should
solidify to produce high pressure assemblages approximating basanitic compositions. These
may correspond to amphibole- and clinopyroxene-rich nodules in alkali basalts, studies of
which frequently invoke basanitic melts as parental to the nodules (Wass, 1979; Irving, 1980;
Duda & Schmincke, 1985).
Discussion so far has centred on progressive, continued redox melting. In the case of
episodic melting, CH 4 -rich fluids which are oxidizing to produce the increase in aH2O may
encounter a region which has previously experienced partial melting and thus is relatively
reduced and geochemically depleted (Fig. 5D). Melts produced will bear a depleted major
element signature and are likely to resemble lamproites (Foley, 1986, 1987a), provided that
an intervening MEE has occurred to supply K, Ti, light ion lithophile, and related elements
(Jaques et al., 1984).
The distinction expected between melts produced in oxidized, fertile conditions and in
reduced, depleted conditions may correspond to the Al-rich and Al-poor nodule assem-
blages outlined above. In this case the parental melts to xenolith types would probably be
basanitic and lamproitic respectively. However, both melt and wall-rock chemistry and
mineralogy will be important in determining the nature of MEE sampled by xenoliths. For
example, Al-poor lamproitic melts crystallizing in an Al-rich peridotite would probably not
produce the richterite-bearing assemblages; lamproites are undersaturated in alumina and
extensive wall-rock reaction would occur.
If the country rocks are also geochemically depleted, the low A12O3 environment (with
Cr 2 O 3 derived from the wall rocks) will favour crystallization of characteristically Al-poor
mineral assemblages such as those containing potassic richterite, priderite and/or crichton-
ite series minerals (Jones et al., 1982; Haggerty, 1983), provided the pressure of crystallization
is suitable.
The degree of geochemical depletion and intrinsic oxidation state in the region of melting
may explain the difference in nature between kimberlitic and lamproitic melts (Foley,
1987a, b; Fig. 6). Olivine lamproites may originate at similar depths to kimberlites but from a
more depleted and reduced mantle composition resulting from an earlier period of melting.
Many leucite lamproites may originate from similar source compositions by a similar
melting mechanism to olivine lamproites, but at shallower depths (Foley, 1986; Foley et al.,
152 STEPHEN F. FOLEY

KIMBERLITE/ LAMPROITE PRODUCTION BY REDOX MELTING

KIMBERUTIC
m MELT m ' LEUEITE ~ MADUPITIC
LAMPROITE

OXIDISED
L1THOSPHERE

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


( C H 4 • 0 2 — 2 H20 * C )

REDUCED S CH 4 * H 2
ASTHENOSPHERE S fluid

• IW IVM • IW*(1-2) • -IW-3 ~IV\K -- MELT


INCREASING OXYGEN FUGACITY > . ZONE OF
' PARTIAL MELTING

FIG. 6. Contrasting origins for kimberlites and lamproites according to the redox melting model. Kimberlites are
produced at the first stage of melting where oxidation fronts are relatively sharp and the oxidized zone is fertile; this
corresponds to Fig. 5B. Lamproites are produced by later episodes of melting and so bear a depleted Ca, Na, Al
signature, although incompatible elements have been enriched in the intervening period resulting in a mica-bearing
source for the later melting episode. The reduced nature of this region results in a H2O-rich, CO2-free melting
environment, so that leucite lamproites with >50 wt.% SiO 2 are produced at lower pressures. Olivine lamproites
have lower SiO 2 due to their greater depth of origin, and rare CO 2 -bearing 'lamproites' may originate where
oxidised blocks persist in generally reduced, depleted regions.

1986). Silica-poor madupitic lamproites, exemplified by Leucite Hills madupites and


lamproites from West Greenland (Scott, 1979), may, if primary, be hybrids formed by
melting of mixed oxidized, carbonate-bearing material and reduced, depleted material
(Fig. 6). Oxidized blocks may well persist within a generally depleted region of the mantle,
and be selectively attacked by later episodes of fluid introduction using new pathways.
The MEE (post-dating the depletion event) required for lamproite sources by petrological
and isotopic studies (McCulloch et al, 1983; Jaques et al, 1984; Fraser et al, 1985) may also
be caused by activation of relatively oxidized, more fertile blocks at deeper levels. Otherwise,
a MEE-melt from a reduced and uniformly depleted mantle would have to scavenge
incompatible elements from a very large area. Invocation of a fluid as an enriching agent is
less appealing from a geochemical viewpoint (Pearce, 1983; Foley et al, 1987), and CH4-rich
fluids are not likely to carry high volumes of solute (Taylor, 1985). Ancient subducted blocks
caught beneath continental regions may give rise to oxidized regions which would be
preferentially melted on contact with upwelling reduced fluids (cf. Green et al, 1987/? for
ocean island volcanism); this may explain isotopic and geochemical features of some
continental rocks which have been suggested to be related to a subduction zone component.
THE GENESIS OF C O N T I N E N T A L BASIC ALKALINE MAGMAS 153

Continental rift regions


Continental rift regions contain the largest diversity of volatile-rich alkaline igneous rocks
of any tectonic setting, including intrusive complexes and flood basalts ranging from
nephelinite ( + / — carbonatite) to phonolite and trachyte. In this section the ideal case of a
gradually propagating rift from stable continental to oceanic rift is taken to illustrate the
possible consequences of redox melting.

Progressive continental rifting


The initial stages of rifting will be largely similar to stable continental environments
(Fig. 7A); access of reduced fluids would cause low-silica, CO2-influenced melting as
described for Fig. 5B. However, the earliest magmatism in rifts is not similar to cratonic

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


areas: kimberlites and lamproites are generally not seen (where references to kimberlites in
rift environments are made, these are mostly due to semantic difficulties in the use of the term
'kimberlite'; Mitchell, 1986), an exception being the Labrador Sea rift which cuts through an
Archean craton. The rocks seen in the early stages of rifting, the melilitites, nephelinites and
ultramafic lamprophyres (alnoites and aillikites) may be the equivalent melts produced in
the presence of CO 2 -H 2 O fluids but in regions of thinner lithosphere, i.e. lower pressures.
Most carbonatites probably do not represent mantle-derived melts, since they are associated
with nephelinites with which they may be related by liquid immiscibility (e.g., LeBas, 1977).
However, some carbonate-rich ultramafic lamprophyres (<30 wt. % SiO2) may represent
mantle-derived melts (Malpas et al., 1986; Rock, 1986), and many of these occur in regions of
normal faulting interpreted to represent incipient rifting (e.g., Doig, 1970; Grikurov et al.,
1980).
In rift zones the disequilibrium redox melting process may be relatively more important
than buffered CO2-rich melting, since the abundance of tensional faulting will promote
access of fluids to higher levels. Nephelinite-carbonatite or ultramafic lamprophyre
magmatism generally forms widely dispersed centres of relatively small volume (Doig, 1970;
Rock, 1986) which is consistent with activation of new regions of the mantle. It may be
important here to distinguish between the two types of nephelinite (olivine-rich and olivine-
poor) outlined by LeBas (1978a, 1987). The Group II nephelinites (olivine-poor) are
clinopyroxene-rich and associated with carbonatites and have been suggested by LeBas
(1978b) to originate in a high CO 2 /H 2 O environment. Nephelinites produced at the early rift
stages (Fig. 7A) may be restricted to Group II, whereas later melting in more H 2 O-rich
conditions may lead to Group I nephelinites, which are associated with less alkaline basaltic
lavas.
As rifting progresses the oxidation front encounters steadily higher levels of oxidized
lithosphere. The advancing fluid front continues to promote CO 2 -rich melting, but melts
will be progressively richer in silica as the pressure of melting decreases. At higher degrees of
melting where conditions are richer in H 2 O (cf. above and Fig. 5C) melts may resemble
basanites, Group I nephelinites or alkaline lamprophyres. If rifting continues, fluids will
continue to move towards higher levels, so that the degree of partial melting at any one level
should remain low. Production of depleted types such as lamproites in the manner depicted
in Fig. 6 would require collapse and restarting of the rifting mechanism, which helps to
explain their rarity in rift environments.
At intermediate stages of rifting melting due to redox reactions will be difficult to isolate
from those due to increasing heat flow. That both occur together is implied by the eventual
development of oceanic rifts from continental rifts. The progressive updoming of the
asthenosphere roof is accompanied by thinning and extension of the lithosphere by extensive
154 STEPHEN F. FOLEY

CONTINENTAL RIFT REGION


- Nephelinite (H), Basanire,
/_ Melilitire Nephelinite(H) (I)

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


Peripheral
"nrtched f JJ* Basanite, alkaline Tholeiite
f f Nephelinite(I) salf

FIG. 7. Redox melting processes in the progressive development of an archetypal continental rift. (A) Initial CO 2 -
rich melting results in melilitites and nephelinites rather than kimberhtes due to the lower pressures relative to
cratonic areas. (B) Basanites and nephelinites follow at greater degrees of partial melting. The two nephelinite
groups (LeBas, 1978a) can be assigned to early CO 2 -rich and later H 2 O-rich melting stages. (C) At later stages,
melting at shallower depths occurs in regions of altered mantle composition due to MEE during earlier stages of
rifting. Magmatism produced at this stage is likely to exhibit local characteristics dependent on the nature of these
enrichment events. Tensional faulting associated with extension will assist emplacement of alkaline, low-degree
partial melts. (D) The transition to production of oceanic crust marks the dominance of decompression melting,
producing tholeiites, over volatile-induced melting. Peripheral activity will include alkaline magmas produced in an
analogous manner to (B) and (C). Symbols as in Fig. 5.

tensional faulting. The recurrence of alkaline magmatism in a single region over long time
periods indicates that this process is probably controlled by lithosphere plate dynamics
(Bailey, 1977; Giret & Lameyre, 1985). The deep faults will form low pressure areas which
attract fluid from surrounding areas beneath the lithosphere (Bailey, 1980). Griesshaber et
ah, (1987) describe gases collected in the Rhine graben as having higher 3 He around graben
faults indicating access of mantle derived fluids along the fractures. Fluids which enter
fractures as reduced CH 4 + H 2 O mixtures may become oxidized, with associated carbon
THE GENESIS OF CONTINENTAL BASIC ALKALINE MAGMAS 155

precipitation, to CO 2 + H 2 O mixtures by interaction with convecting crustal fluids resulting


in pervasive carbonate mineralisation around the fractures (Green et ai, 1987a). Once the
fracture system is established, local reduction of the immediate environment should permit
access of reduced fluids to higher levels which were recently enriched in carbonate, with
consequent CO 2 -rich melting.
At the stage represented by Fig. 7C, melting has progressed to higher levels and may
involve mantle material highly altered by enrichment events which may originate either by
solidification of melts from earlier stages of rifting (Fig. 7A-B) or from metasomatic reaction
with fluids around fractures. Melting in these environments is envisaged to occur as
indicated in Fig. 8: melts are likely to have solidified in a system of blind fractures likened to
a stock work by Bailey (1984). Note that the intrinsic oxidation state of the vein system will
be little lower than fOl % EMOG, because the melting took place in H 2 O>CO 2 -bearing

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


conditions. The oxidation state of melts is expected to be controlled by the presence of Fe 3 +
and carbonates, in contrast to reduced melts where the effect of Fe 3 + may be minimal

cl TT r/
FIG. 8. Model of melt separation in an enriched mantle region. (A) Alkaline melts solidify in the upper mantle
producing zoned vein systems rich in amphibole and/or mica, possibly with CO 2 -rich outer zones (Irving, 1980;
Menzies et al., 1985). (B) Reduced fluids mostly, but not exclusively, use the earlier pathways. Melting occurs
preferentially in the enriched vein systems due to their lower solidus temperatures and to water produced from the
oxidation of CH 4 in addition to that prexisting in hydrous minerals. Melt tends to coalesce into horizontal melt
layers, and (C) interacts with peridotitic wall rocks (including minor melting of the peridotite) prior to escape to the
surface. Melts produced in the vein systems will be highly alkaline, and the alkalinity of the volcanic rock will be a
function of the density of MEE in the source region.
156 S T E P H E N F. F O L E Y

compared to that of fluid species (Taylor & Green, 1986; Green et al, 1987a). The reduced
fluids will often use inherited fractures and so their introduction into hydrous mineral
bearing mantle should be disproportionately common (Fig. 8B). Enriched mantle will be
preferentially melted during later periods of fluid introduction due to the lower melting
temperatures of hydrous mineral-bearing assemblages (Bailey, 1984). This effect may be
enhanced by fluids re-using earlier passageways, and by H 2 O produced from oxidation of
reduced fluids.
MEE resulting from earlier partial melt infiltration will have introduced K and related
incompatible elements relative to Na, and later partial melts of the enriched mantle will
therefore inherit this K-rich signature. Experiments on mica-clinopyroxenite, a likely MEE
assemblage, show that partial melts may be ultrapotassic to fairly high degrees of partial
melting (20-30%; Lloyd et al, 1985). The difference between Na-rich volcanism in, for

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


example, the eastern branch of the East African rift valley (Baker, 1987) and K-rich
volcanism in the western branch may lie in the degree of enrichment of the source. Primary
melts to western rift volcanics are considered to be potassic and to be derived from
mineralogically unusual mantle materials (Ferguson & Cundari, 1975; Lloyd & Bailey, 1975;
Pouclet, 1980; Kampunzu et al, 1983; Pouclet et al, 1983).
At higher degrees of partial melting melts are thought to spread laterally into horizontal
layers (Fig. 8C) in accordance with gravitational and compaction constraints on melt
distribution (Waff, 1980; Shankland et al, 1981; Maal0e & Scheie, 1982; Richter &
McKenzie, 1984). The incorporation of melts from unenriched parts of the mantle will
substantially dilute the alkalinity of all melts except those produced in regions of densely
enriched mantle (i.e. high per cent MEE per unit volume). The effect of local variations in
density and nature of MEEs may generally be limited to relatively minor distinctions in melt
types, such as the distinction between K-rich and K-poor basanite within a volcanic field
(Ferguson & Cundari, 1975; Pouclet et al, 1983). The basanitic nature (as opposed to
melilitic or basaltic) will be determined more by the degree of partial melting.
Melts produced at the rift periphery at this stage may resemble low-silica melts produced
during the fluid-influenced preliminary rifting stages. In Greenland, these include micaceous
kimberlites (Emeleus & Andrews, 1975)emplaced contemporaneously with the development
of an oceanic rift in which basaltic melt production was presumably more influenced by high
heat flow.
Figure 7D represents the last stage of rift development—the conversion to production of
oceanic crust in the rift centre. This will be immediately preceded by development of only
marginally alkaline to tholeiitic subaerial volcanics such as those now found at the rift edge
in western Greenland (Clarke, 1975; Larsen, 1977). Heat flow is by this stage very high, and
the transfer to low-alkali basaltic volcanism probably represents the domination of large-
scale decompression melting over that triggered by fluid interaction. Melt products include
picrites which represent high degrees of partial melting (Clarke, 1970). The established
oceanic ridge is of interest to continental redox melting in that it provides evidence of the
reduced state of the asthenosphere where it approaches near surface levels (Fig. 4, above).
Even at the stage of establishment of oceanic crust, the last vestiges of peripheral alkaline
volcanism may persist, and these may continue to be CO 2 + H 2 O-rich and involve enriched
mantle regions. The alkaline lamprophyres of Disko Island (Larsen, 1981) are examples of
melts produced at this stage.

ACKNOWLEDGEMENTS
This paper was written whilst the author was in receipt of a stipendium from the Max-
Planck-Gesellschaft. I am grateful to E. A. Mathez and A. L. Jaques for reviews of the
T H E G E N E S I S O F C O N T I N E N T A L BASIC A L K A L I N E M A G M A S 157

manuscript which led to substantial changes in presentation. Fig. 1 is reproduced with the
kind permission of D. M. Christie and the Elsevier Scientific Publ. Co., Amsterdam.

REFERENCES
Anderson, D. L., 1962. The plastic layer in the Earth's mantle. Sci. Am. No. 7, 12-19.
Arculus, R. J., 1985. Oxidation status of the mantle: past and present. Ann. Rev. Earth, planet. Sci. 13, 75-95.
Delano, J. W., 1981. Intrinsic oxygen fugacity measurements: techniques and results for spinels from upper
mantle peridotites and megacryst assemblages. Geochim. cosmochim. Ada 45, 899-913.
Arseneyev, A. A., 1963. The laws of the distribution of kimberlites in the eastern part of the Siberian platform. Dokl.
Akad. Nauk Earth Sci. Secton 137, 355-7.
Bailey, D. K., 1977. Lithospheric control of continental rift magmatism. J. geol. Soc. Lond. 133, 103-6.
1980. Volcanism, Earth degassing and replenished lithosphere mantle. Phil Trans R. Soc. Lond. A297, 309-22.
1984. Kimberlite: "The mantle sample" formed by ultrametasomatism. In: Kornprobst, J. (ed) Kimberlites I:

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


Kimberlites and Related Rocks. Amsterdam, Elsevier: 323-33.
Baker, B. H., 1987. Outline of the petrology of the Kenya rift alkaline province. In: Fitton, J. G., & Upton, B. G. J.
(eds) Alkaline Igneous Rocks, Geol Soc London Special Publ. 30, 293-311.
Banks R. J., & Beamish, D., 1979. Melting in the crust and upper mantle beneath the Kenya rift: evidence from
geomagnetic deep sounding experiments. J. geol. Soc. Lond. 136, 225-33.
Belviso, S., Jean-Baptiste, P., Nguyen, B. C, Merlivat, L., & Labeyrie, L., 1987. Deep methane maxima and 3 He
anomalies across the Pacific entrance to the Celebes Basin. Geochim. cosmochim. Acta Vol. 51, 2673-80.
Bottinga, Y., & Weill, D. F., 1972. The viscosity of magmatic silicate liquids: a model for calculation. Am. J. Sci. 272,
438-75.
Brey, G. P., 1976. CO 2 solubility and solubility mechanisms in silicate melts at high pressures. Contr. Miner. Petrol.
57,215-21.
1978. Origin of olivine melilitites—chemical and experimental constraints. J. Volcanol. geotherm. Res. 3,
61-88.
Green, D. H., 1975. The role of CO 2 in the genesis of olivine melilitite. Contr. Miner. Petrol. 49, 93-103.
Bristow, J. W., 1984. Picritic rocks of the North Lebombo and south-east Zimbabwe. Spec. Publ. Geol. Soc. S. Africa.
13, 105-23.
Smith, C. B., Allsopp, H. L, Shee, S. R., & Skinner, E. M. W., 1986. Setting, geochronology and geochemical
characteristics of 1600 m. y. kimberlites and related rocks from the Kuruman province, South Africa. Fourth
Int. Kim. Conf. Extended Abstracts: Geol. Soc. Australia Abs. 16, 112—4.
Byers, C. D., Muenow, D. W., & Garcia, M. O., 1983. Volatiles in basalts and andesites from the Galapagos
spreading center, 85° to 86°W. Geochim. cosmochim. Acta 47, 1551-8.
Christie, D. M., Muenow, D. W., & Sinton, J. M., 1984. Volatile contents and ferric ferrous ratios of basalt,
ferro-basalt, andesite and rhyodacite glasses from the Galapagos 95.5°W propagating rift. Ibid. 48, 2239^15.
Garcia, M. O., & Muenow, D. W., 1986. Volatiles in basaltic glasses from the East Pacific Rise at 21°N:
implications for MORB sources and submarine lava flow morphology. Earth planet. Sci. Lett. 79, 9-20.
Carmichael, I. S. E., & Ghiorso, M. S., 1986. Oxidation-reduction relations in basic magma: a case for homogeneous
equilibria. Ibid. 78, 200-10.
Christie, D. M., Carmichael, I. S. E., & Langmuir, C. H., 1986. Oxidation states of mid-ocean ridge basalt glasses.
Ibid. 79, 397-411.
Clague, D. A., & Frey, F. A., 1982. Petrology and trace element geochemistry of the Honolulu volcanics, Oahu:
implications for the oceanic mantle below Hawaii. J. Petrology 23, 447-504.
Claypool, G. E., & Kvendolen, K. A., 1983. Methane and other hydrocarbon gases in marine sediment. Ann. Rev.
Earth planet. Sci. 11, 299-327.
Clarke, D. B., 1970. Tertiary basalts of Baffin Bay: possible primary magma from the mantle. Contr. Miner. Petrol.
25, 203-24.
1975. Tertiary basalts dredged from Baffin Bay. Can. J. Earth Sci. 12, 1396^05.
Cook, F. A., McCullar, D. B., Decker, E. R., & Smithson, S. B., 1979. Crustal structure and evolution of the southern
Rio Grande rift. In: Riecker, R. E. (ed) Rio Grande Rift: Tectonics and Magmatism. Washington, D.C: A.G.U,
195-208.
Craig, H., Lupton, J. E., & Horibe, Y., 1978. A mantle helium component in circum-Pacific volcanic gases: Hakone,
the Marianas, and Mt. Lassen. In: Alexander, E. C , & Ozima, M. (eds) Terrestrial Rare Gases. Tokyo: Japan
Science Soc. Press, 3-16.
Kim, K-R., & Rison, W., 1984. Easter Island hotspot: I. Bathymetry, helium isotopes, and hydrothermal
methane and helium. EOS 65, 1140.
Currie, K. L., 1976. The Alkaline Rocks of Canada. Bull. Geol. Survey Canada 239, 228pp.
Dawson, J. B., 1971. Advances in kimberlite geology. Earth Sci. Rev. 7, 187-214.
Smith, J. V., 1977. The MARID (mica amphibole-rutile-ilmenite-diopside) suite of xenoliths in kimberlite.
Geochim. cosmochim. Acta 41, 309-23.
Deines, P., 1980. The carbon isotopic composition of diamonds: relationship to diamond shape, color, occurrence
and vapor composition. Ibid. 44, 943-61.
Doig, R., 1970. An alkaline province linking Europe and North America. Can. J. Earth Sci. 7, 22-8.
158 S T E P H E N F. F O L E Y

Duba, A. G., & Shankland, T. J., 1982. Free carbon and electrical conductivity in the Earth's mantle. Geophys. Res.
Lett. 9, 1271-4.
Duda, A., & Schmincke, H-U., 1985. Polybaric differentiation of alkali basaltic magmas: evidence from green-core
clinopyroxenes (Eifel, FRG). Contr. Miner. Petrol. 91, 340-53.
Eggler, D. H., 1975. Peridotite-carbonate relations in the system CaO MgO-SiO 2 -CO 2 . Yb. Carnegie Inst., Wash
74, 468-74.
1978. Stability of dolomite in a hydrous mantle, with implications for the mantle solidus. Geology 6, 397—400.
1983. Upper mantle oxidation state: evidence from olivine-orthopyroxene-ilmenite assemblages. Geophys.
Res. Lett. 10, 365-8.
1986. Kimberlites: how do they form? Fourth Int. Kim. Conf. Ext. Abs: Geol. Soc. Australia Abstracts 16, 155-9.
1987. Solubility of major and trace elements in mantle metasomatic fluids: experimental constraints. In:
Menzies, M. A., & Hawkesworth, C. J. (eds) Mantle Metasomatism. Academic Press, 21-41.
Baker, D. R., 1982. Reduced volatiles in the system C-O-H: implications to mantle melting, fluid formation,
and diamond genesis. In: Akimoto, S., & Manghnani, M. (eds) High Pressure Research in Geophysics. Tokyo:
Center for Academic Publications, 237-50.
Holloway, J. R., 1977. Partial melting of peridotite in the presence of H 2 O and CO 2 : principles and review. In

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


Dick, H. J. B. (ed.) Magma Genesis. Oregon Dep Geol Min Ind Bull 96, 5-36.
Rosenhauer, M., 1978. Carbon dioxide in silicate melts: II. solubilities of CO 2 and H 2 O in CaMgSi 2 O 6
(diopside) liquids and vapors at pressures to 40kb. Am. J. Sci. 278, 64-94.
Ellis, D. J., 1983. The Napier and Rayner complexes of Enderby Land, Antarctica contrasting styles of
metamorphisrn and tectonism. In: Oliver, R. L., James, P. R., & Jago, J. B. (eds) Antarctic Earth Science.
Canberra: Australian Academy of Science, 20-4.
Emeleus, C. H., & Andrews, J. R., 1975. Mineralogy and petrology of kimberlite dyke and sheet intrusions and
included peridotite xenoliths from southwest Greenland. Phys. Chem. Earth 9, 179-97.
Esperanca, S., & Holloway, J. R., 1986. On the origin of some mica-lamprophyres: experimental evidence from a
mafic minette. Contr. Miner. Petrol. 95, 207-16.
Ferguson, A. K., & Cundari, A., 1975. Petrological aspects and evolution of the leucite-bearing lavas from
Bufumbira, South West Uganda. Ibid. 50, 25—46.
Fitton, J. G., & Dunlop, H. M., 1985. The Cameroon Line, West Africa, and its bearing on the origin of oceanic and
continental alkali basalt. Earth planet. Sci. Lett. 72, 23-38.
Foden, J. D., & Varne, R., 1980. The petrology and tectonic setting of Quaternary Recent volcanic centres of
Lombok and Sumbawa, Sunda Arc. Chem. Geol. 30, 201-26.
Foley, S. F., 1986, 1987a. The genesis of lamproitic magmas in a reduced fluorine-rich mantle. Fourth Internat Kim.
Conf. Ext. Abs: Geol. Soc. Australia Abs. 16, 173-175 and proceedings volume (in press).
1987b. Experimental studies of the genesis of lamproitic magmas. Terra Cognita 7, 366.
Taylor, W. R., & Green, D. H., 1986. The role of fluorine and oxygen fugacity in the genesis of the ultrapotassic
rocks. Contr. Miner. Petrol. 94, 183-92.
Venturelli, G., Green, D. H., & Toscani, L., 1987. The ultrapotassic rocks: characteristics, classification, and
constraints for petrogenetic models. Earth Sci. Rev. 24, 81-134.
Fraser, K. J, Hawkesworth, C. J, Erlank, A. J., Mitchell, R. H., & Scott-Smith, B. H., 1985. Sr, Nd andPb isotope
and minor element geochemistry of lamproites and kimberlites. Earth planet. Sci. Lett. 76, 57-70.
Frey, F. A., Green, D. H., & Roy, S. D., 1978. Integrated models of basalt petrogenesis: a study of quartz tholeiites to
olivine melilitites from South Eastern Australia utilising geochemical and experimental petrological data. J.
Petrology 19, 463-513.
Giret, A., & Lameyre, J., 1985. Inverted alkaline-tholeiitic sequences related to lithospheric thickness in the
evolution of continental rifts and oceanic islands. J. African Earth Sci. 3, 261-8.
Green, D. H., 1970. The origin of basaltic and nephelinitic magmas-. Trans. Leicester Lit. Phil. Soc. 64, 28-54.
1971. Composition of basaltic magmas as indicators of conditions of origin: application to oceanic volcanism.
Phil Trans R. Soc. Lond. A268, 707-25.
Falloon, T. J., & Taylor, W. R., 1987a. Mantle:derived magmas-roles of variable source peridotite and
variable C - H - O fluid compositions. In: Mysen, B. O. (ed.) Magmatic Processes: Physicochemical Principles.
Geochemical Soc Special Publication 1, 139-54.
Liebermann, R. C , 1976. Phase equilibria and elastic properties of a pyrolite model for the oceanic upper
mantle. Tectonophysics 32, 61-9.
Taylor, W. R., & Foley, S. F., 1987b. The Earth's upper mantle as a source for volatiles. Geol. Soc. Australia
Spec. Publication 12 (in press).
Griesshaber, E., O'Nions, R. K., Oxburgh, E. R., Lodemann, M., & Paces, T., 1987. 3 He distribution in the
Oberpfalz and adjacent areas, West Germany. Terra Cognita 7, 165.
Grikurov, G. E., Orlenko, E. M., & Fedorov, L. V., 1980. Alkaline-ultrabasic rocks of the Beaver Lake region,
eastern Antarctica. Trudi Sovietskoy Antarcticheskoy Expeditsiy 70, 87-99.
Haggerty, S. E., 1983. The mineral chemistry of new titanates from the Jagersfontein kimberlite, South Africa:
implications for metasomatism in the upper mantle. Geochim. cosmochim. Ada 47, 1833-54.
1986. Diamond genesis in a multiply-constrained model. Nature 320, 34-8.
Tompkins, L. A. 1983. Redox state of Earth's upper mantle from kimberlitic ilmenites. Ibid. 303, 295-300.
THE GENESIS OF CONTINENTAL BASIC ALKALINE MAGMAS 159

Harte, B., 1983. Mantle peridotites and processes—the kimberlite sample. In: Hawkesworth, C. J., & Norry, M. J.
(eds) Continental Basalts and Mantle Xenoliths. Nantwich: Shiva, 46-91.
Holloway, J. R., 1973. The system pargasite H 2 O-CO 2 : a model for melting of a hydrous mineral with a mixed-
volatile fluid I. experimental results to 8 kbar. Geochim. cosmochim. Ada 37, 651-66.
—1981. Volatile interactions in magmas. Adv. Phys. Geochem. 1, 273-93.
— —Ford, C. E., 1975. Fluid absent melting of the fluor-hydroxy amphibole pargasite to 35 kilobars. Earth planet.
Sci. Lett. 25, 44-8.
Irving, A. J., 1980. Petrology and geochemistry of composite ultramafic xenoliths in alkalic basalts and implications
for magmatic processes within the mantle. Am. J. Sci. 280A, 389^426.
Jaques, A. L., Lewis, J. D., Smith, C. B., Gregory, G. P., Ferguson, J., Chappell, B. W., & McCulloch, M. T , 1984.
The diamond-bearing ultrapotassic (lamproitic) rocks of the West Kimberley region, Western Australia. In:
Kornprobst, J. (ed) Kimberlites I: Kimberlites and Related Rocks. Amsterdam: Elsevier, 225-54.
Jones, A. P., Smith, J. V., & Dawson, J. B., 1982. Mantle metasomatism in 14 veined peridotites from the Bultfontein
mine. South Africa. J. Geol. 90, 435-53.
Kaminskiy, F. V., Lobkov, V. A., Prasalov, E. M., Beskrovny, N. S., Kudryavtseva, E. I., Anufriev, G. S., & Pavlov,
V. B., 1976. The components of the upper mantle of the Earth in gases of Kamchatka (according to He, Ne, Ar,

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


and C isotopy). Geochem. Int. 13, 35-^48.
Kampunzu, A. B., Vellutini, P-J., Caron, J-P. H., & Bizimana, B. L., 1983. Sur la nature et la signification des laves
basiques du Sud-Ouest de rile Idjwi dans le lac Kivu au Zaire (Rift Africain). C. R. Acad. Sci. Paris Serie II296,
909-912.
Kim, K-R., 1983. Methane and radioactive isotopes in hydrothermal systems. Ph.D. thesis, Scripps Institute of
Oceanography, University of California, San Diego, 206pp.
Kurz, M. D., Jenkins, W. J., Hart, S. R., & Clague, D., 1983. Helium isotopic variations in volcanic rocks from Loihi
Seamount and the Island of Hawaii. Earth planet. Sci. Lett. 66, 388-406.
Meyer, P. S., & Sigurdsson, H., 1985. Helium isotopic systematics within the neovolcanic zone of Iceland. Ibid.
74, 291-305.
Kushiro, I., Satake, H., & Akimoto, S., 1975. Carbonate-silicate reactions at high pressures and possible presence of
dolomite and magnesite in the upper mantle. Ibid. 28, 116-20.
Larsen, J. G., 1977. Transition from low potassium olivine tholeiites to alkali basalts on Ubekendt Ejland, the
Tertiary volcanic province of West Greenland. Meddelelser om Gronland 200, 1-42.
1981. Medium pressure crystallisation of a monchiquitic magma—evidence from meagcrysts of Drever's
block, Ubekendt Ejland, West Greenland. Lithos 14, 241-62.
LeBas, M. J. 1977. Petrogenesis of the carbonatite nephelinite association. In: LeBas, M. J. (ed.)
Carbonatite-Nephelinite Volcanism. London: Wiley, 279-93.
1978a. Nephelinite volcanism at plate interiors. Bull. Volcanol. 41, 459-62.
1978/). Are olivine-poor nephelinites a primary melt product from the mantle? Ibid. 41, 463-5.
1987. Nephelinites and carbonatites. In: Fitton, J. G., & Upton, B. G. J. (eds) Alkaline Igneous Rocks. Geol Soc
London Spec. Publ. 30, 53-83.
Lloyd, F. E., & Bailey, D. K., 1975. Light element metasomatism of the continental mantle: the evidence and the
consequences. Phys. Chem. Earth 9, 389-416.
Arima, M., & Edgar, A. D., 1985. Partial melting of a phlogopite clinopyroxenite nodule from south-west
Uganda: an experimental study bearing on the origin of highly potassic continental rift volcanics. Contr.
Miner. Petrol. 91, 321-9.
Lupton, J. E., 1983. Terrestrial inert gases: isotope tracer studies and clues to primordial components in the mantle.
Ann. Rev. Earth planet. Sci. 11, 371^414.
Craig, H., 1975. Excess 3 He in oceanic basalts: evidence for terrestrial primordial helium. Earth planet. Sci.
Lett. 26, 133-9.
1981. A major 3 He source on the East Pacific Rise. Science 214, 13-8.
Maalyie, S., & Sheie, A. 1982. The permeability controlled accumulation of primary magma. Contr. Miner. Petrol.
81, 350-7.
Malpas, J., Foley, S. F., & King, A. F., 1986. Alkaline mafic and ultramafic lamprophyres from the Aillik Bay area,
Labrador. Can. J. Earth Sci. 23, 1902-18.
Mathez, E. A., 1984. Influence of degassing on oxidation states of basaltic magmas. Nature 310, 371-5.
Blacic, J. D., Beery, J., Hollander, M. & Maggiore, C , 1986. Carbon in olivine: results from nuclear reaction
analysis. J. geophys. Res. 92, 3500-6.
Mattioli, G. S., Wood, B. J., & Carmichael, I. S. E., 1987. Ternary-spinel volumes in the system
MgAl 2 O 4 -Fe 3 O 4 -Fe 8 3 O 4 : implications for the effect of P on intrinsic/ O2 measurements on mantle-xenolith
spinels. Am. Miner. 72, 468-80.
McCulloch, M. T., Jaques, A. L., Nelson, D. R. & Lewis, J. D., 1983. Nd and Sr isotopes in kimberlites and
lamproites from Western Australia: an enriched mantle origin. Nature 302, 400-3.
McKenzie, D. P., 1984. The generation and compaction of partially molten rock. J. Petrology 25, 713-65.
Menzies, M., 1983. Mantle ultramafic xenoliths in alkaline magmas: evidence for mantle heterogeneity modified by
magmatic activity. In: Hawkesworth, C. J., & Norry, M. J. (eds) Continental Basalts and Mantle Xenoliths,
Nantwich: Shiva, 111-38.
160 S T E P H E N F. F O L E Y

Kempton, P., & Dungan, M., 1985. Interaction of continental lithosphere and asthenospheric melts below the
Geronimo volcanic field, Arizona, U.S.A. J. Petrology 26, 663-93.
Merlivat, L., Pineau, F., & Javoy, M., 1987. Hydrothermal vent waters at 13°N on the East Pacific Rise: isotopic
composition and gas concentration. Earth planet. Sci. Lett. 84, 100-8.
Mitchell, R. H., 1986. Kimberlites: Mineralogy, Geochemistry and Petrology. New York: Plenum, 442 pp.
Moore, R. O., & Gurney, J. J., 1986. Mineral inclusions in diamonds from the Monastery kimberlite, South Africa.
Fourth Int. Kim. Conf. Extended Abstracts: Geol. Soc. Australia Abs. 16, 406-8.
Otter, M. L., Rickard, R. S., Harris, J. W., & Gurney, J. J., 1986. The occurrence of moissanite and ferro-
periclase as inclusions in diamond. Ibid. 16, 409-411.
Olafsson, M., & Eggler, D. H., 1983. Phase relations of amphibole, amphibole-carbonate, and
phlogopite-carbonate peridotite: petrologic constraints on the asthenosphere. Earth planet. Sci. Lett. 64,
305-15.
O'Neill, H. St. C , & Wall, V. J., 1987. The olivine-orthopyroxene-spinel oxygen geosensor, the nickel precipitation
curve, and the oxygen fugacity of the Earth's upper mantle. J. Petrology 28, 1169-91.
Pearce, J. A., 1983. Role of the sub-continental lithosphere in magma genesis at active continental margins. In:
Hawkesworth, C. J., & Norry, M. J. (eds) Continental Basalts and Mantle Xenoliths. Nantwich: Shiva, 230-49.

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


Phelps, D. W., Gust, D. A., & Wooden, J. L., 1983. Petrogenesis of the mafic feldspathoidal lavas of the Raton-
Clayton volcanic field, New Mexico. Contr. Miner. Petrol. 84, 182-90.
Pouclet, A., 1980. Contribution a la systematique des laves alcalines, les laves du rift de l'Afrique Centrale
(Zaire-Uganda). Bull. Volcanol. 43, 527^0.
Menot, R-P., & Piboule, M., 1983. Le magmatisme alcalin potassique de l'aire volcanique des Virunga (Rift
occidental de l'Afrique de l'Est). Une approche statistique dans la recherche des filiations magmatiques et des
mecanismes de differenciation. Ibid. 106, 607-22.
1984. Differenciation des laves de l'Afrique Centrale (Rift Ouest). Contribution de l'analyse
statistique multivariee. Neues Jb. Min. Abh. 149, 283-308.
Richter F. M., & McKenzie, D. P., 1984. Dynamic models for melt segregation from a deformable matrix. J. Geol.
92, 729-40.
Rison, W., & Craig, H., 1983. Helium isotopes and mantle volatiles in Loihi Seamount and Hawaiian Island basalts
and xenoliths. Earth planet. Sci. Lett. 66, 407-26.
Roden, M. F., 1981. Origin of coexisting minette and ultramafic breccia, Navajo volcanic field. Contr. Miner. Petrol.
77, 195-206.
Rock, N. M. S., 1984. Nature and origin of calc-alkaline lamprophyres: minettes, vogesites, kersantites and
spessartites. Trans. R. Soc. Edinburgh Earth Sci. Section 74, 193-227.
1986. The nature and origin of ultramafic lamprophyres: alnoites and allied rocks. J. Petrology 27, 155-96.
Ryabchikov, I. D., Green, D. H., Wall, V. J., & Brey, G. P., 1981. The oxidation state of carbon in the reduced
velocity zone. Geochem. Int. 18, 148-58.
Sage, R. P, & Watkinson, D. H., 1986. Alkalic rock-carbonatite complexes of the Precambrian shield of Ontario.
Geol. Soc. Can. Abs. 11, 122.
Sato, M., 1978. Oxygen fugacity of basaltic magmas and the role of gas-forming elements. Geophys. Res. Lett. 5,
447-9.
Scarfe, C. M., 1986. Viscosity and density of silicate melts. In: Scarfe, C. M. (ed) MAC Short Course in Silicate Melts.
Ottawa: Min. Soc. Canada, 36-56.
Scott, B. H., 1979. Petrogenesis of kimberlites and associated potassic lamprophyres from central West Greenland.
In: Boyd, F. R., & Meyer, H. O. A. (eds) Kimberlites, Diatremes and Diamonds: Their Geology, Petrology and
Geochemistry. Am. Geophys. Union, 190-205.
1981. Kimberlite and lamproite dykes from Holsteinsborg, West Greenland. Meddelelserom Gronland Geosci.
4, 1-24.
Shankland, T. J., O'Connell, R. J., & Waff, H. S., 1981. Geophysical constraints on partial melts in the upper mantle.
Rev. Geophys. Space Phys. 19, 394^406.
Spencer, A. B., 1969. Alkali igneous rocks of the Balcones Province, Texas. J. Petrology 10, 272-307.
Stillman, C. J., Furnes, H., LeBas, M. J., Robertson, A. H. F., & Zielonka, J., 1982. The geological history of Maio,
Cape Verde Islands. J. Geol. Soc. Lond. 139, 347-61.
Taylor, W. R., 1985. The role of COH fluids in upper mantle processes: a theoretical, experimental and
spectroscopic study. Ph.D Thesis, University of Tasmania, Hobart.
1987a. A reappraisal of the nature of fluids included in diamonds: a window to deep-seated mantle fluids and
redox conditions. Geol. Soc. Australia Special Publ 12, in press.
19876. A 5-parameter modified Redlich-Kwong equation of state for C-O-H fluids at upper mantle pressure
and temperature. J. Geophys. Res. (submitted).
Green, D. H., 1986, 1987a. Mantle methane and the role of reduced volatiles in 'redox melting' of the mantle.
Fourth Int. Kim. Conf. Extended Abstracts: Geol. Soc. Australia Abs. 16, 211-213, and Proceedings volume (in
press)
19876. The petrogenetic role of methane: effect on liquidus phase relations and the solubility mechanism
of reduced C-H volatiles. In: Mysen, B. O. (ed) Magmatic Processes: Physicochemical constraints. Geochemical
Soc Special Volume 1, 121-38.
T H E G E N E S I S O F C O N T I N E N T A L BASIC A L K A L I N E M A G M A S 161

Tozer, D. C , 1981. The mechanical and electrical properties of the Earth's asthenosphere. Phys. Earth planet. Inter.
25, 280-96.
Upton, B. G. J., 1974. The alkaline province of southwest Greenland. In: Sorensen, H. (ed.) The Alkaline Rocks.
London: Wiley, 221-38.
van Bergen, M. J., Ghezzo, C , & Ricci, C. A., 1983. Minette inclusions in the rhyodacitic lavas of Mt. Amiata
(Central Italy): mineralogical and chemical evidence of mixing between Tuscan and Roman type magmas. J.
Volcanol. geotherm. Res. 19, 1-35.
Waff, H. S., 1974. Theoretical consideration of electrical conductivity in a partially molten mantle and implications
for geothermometry. J. geophys. Res. 79, 4003-10.
— 1980. Effects of the gravitational field on liquid distribution in partial melts within the upper mantle. Ibid. 85,
1815-25.
Wass, S. Y., 1979. Fractional crystallisation in the mantle of late-stage kimberlitic liquids—evidence in xenoliths
from the Kiama area, N.S.W., Australia. In: Boyd, F. R., & Meyer, H. O. A. (eds) The Mantle Sample: Inclusions
in Kimberlites and other Volcanics. Am Geophys Union, 366-73.
Waters, F. G., 1987. A suggested origin of MARID xenoliths in kimberlites by high pressure crystallisation of an
ultrapotassic rock such as lamproite. Contr. Miner. Petrol. 95, 523-33.

Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014


Welhan, J. A., & Craig, H., 1979. Methane and hydrogen in East Pacific Rise hydrothermal fluids. Geophys. Res.
Lett. 6, 829-31.
1983. Methane, hydrogen and helium in hydrothermal fluids at 21 N on the East Pacific Rise. In: Rona,
P. A., Rostrom, K., Laubier, L., & Smith, K. L. (eds) Hydrothermal Processes at Seafloor Spreading Centers.
New York: Plenum Press, 391^09.
Wendlandt, R. F., & Eggler, D. H., 1980. The origins of potassic magmas: 2. Stability of phlogopite in natural spinel
lherzolite and in the system KAlSiO 4 MgO-SiO 2 -H 2 O-CO 2 at high pressures and high temperatures. Am. J.
Sci. 280, 421-58.
Wheller, G. E., Varne, R., Foden, J. D., & Abbott, M. J., 1987. Geochemistry of Quaternary volcanism in the Sunda-
Banda Arc, Indonesia, and three component genesis of island arc basaltic magmas. J. Volcanol. Geotherm. Res.
32, 137-60.
Wilshire, H. G., Pike, J. N., Meyer, C. E., & Schwarzman, E. C , 1980. Amphibole-rich veins in lherzolite xenoliths,
Dish Hill and Deadman Lake, California. Am. J. Sci. 280A, 576-93.
Wimmenauer, W., 1974. The alkaline province of central Europe and France. In: Sorensen, H. (ed.) The Alkaline
Rocks. London: Wiley, 238-71.
Woermann, E., & Rosenhauer, M., 1985. Fluid phases and the redox state of the Earth's mantle: extrapolations
based on experimental, phase-theoretical and petrological data. Fortschr. Miner. 63, 263-349.
Wyllie, P. J., 1978. Mantle fluid compositions buffered in peridotite-CO 2 -H 2 O by carbonates, amphibole, and
phlogopite. J. Geol. 86, 687-713.
1979. Magmas and volatile components. Am. Miner. 64, 469-500.
1980. The origin of kimberlite. J. geophys. Res. 85, 6902-10.
Huang, W. L., 1976. High CO 2 solubilities in mantle magmas. Geology 4, 21 —4.
Zolotarev, B. P., Voitov, G. I., Sarkisyan, I. S., & Cherevichnaya, L. F., 1979a. Distribution of gases and bitumens in
basalts from holes 395 and 396, Leg 45. I nit. Rep. DSDP XLV, 647-51.
Choporov, D. Y, & Voitov, G. I., 1919b. Petrochemistry of basalts and distribution of organic gases: holes 407,
408, 409, 410A, 411, 412 and 413, DSDP leg 49. I nit. Rep. DSDP XLIX, 727-44.
Downloaded from http://petrology.oxfordjournals.org/ at OCLC on December 22, 2014

You might also like