Download Pressures of Crystallization of Icelandic Magmas

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts
no text concepts found
Transcript
JOURNAL OF PETROLOGY
VOLUME 49
NUMBER 3
PAGES 465^492
2008
doi:10.1093/petrology/egm089
Pressures of Crystallization of Icelandic Magmas
DANIEL F. KELLEY* AND MICHAEL BARTON
DIVISION OF EARTH AND PLANETARY DYNAMICS, SCHOOL OF EARTH SCIENCES, THE OHIO STATE UNIVERSITY,
COLUMBUS, OH 43210, USA
RECEIVED FEBRUARY 20, 2007; ACCEPTED DECEMBER 14, 2007
Iceland lies astride the Mid-Atlantic Ridge and was created by seafloor spreading that began about 55 Ma. The crust is anomalously
thick (20^40 km), indicating higher melt productivity in the
underlying mantle compared with normal ridge segments as a result
of the presence of a mantle plume or upwelling centered beneath the
northwestern edge of the Vatnajo« kull ice sheet. Seismic and volcanic
activity is concentrated in 50 km wide neovolcanic or rift zones,
which mark the subaerial Mid-Atlantic Ridge, and in three flank
zones. Geodetic and geophysical studies provide evidence for magma
chambers located over a range of depths (15^21km) in the crust,
with shallow magma chambers beneath some volcanic centers
(Katla, Grimsvo« tn, Eyjafjallajo« kull), and both shallow and deep
chambers beneath others (e.g. Krafla and Askja). We have compiled
analyses of basalt glass with geochemical characteristics indicating
crystallization of ol^plag^cpx from 28 volcanic centers in the
Western, Northern and Eastern rift zones as well as from the
Southern Flank Zone. Pressures of crystallization were calculated
for these glasses, and confirm that Icelandic magmas crystallize
over a wide range of pressures (0001 to 1 GPa), equivalent to
depths of 0^35 km. This range partly reflects crystallization of
melts en route to the surface, probably in dikes and conduits, after
they leave intracrustal chambers. We find no evidence for a shallow
chamber beneath Katla, which probably indicates that the shallow
chamber identified in other studies contains silica-rich magma
rather than basalt. There is reasonably good correlation between the
depths of deep chambers (417 km) and geophysical estimates of
Moho depth, indicating that magma ponds at the crust^mantle
boundary. Shallow chambers (571km) are located in the upper
crust, and probably form at a level of neutral buoyancy. There are
also discrete chambers at intermediate depths (11km beneath the
rift zones), and there is strong evidence for cooling and crystallizing
magma bodies or pockets throughout the middle and lower crust that
might resemble a crystal mush. The results suggest that the middle
and lower crust is relatively hot and porous. It is suggested that
crustal accretion occurs over a range of depths similar to those in
recent models for accretionary processes at mid-ocean ridges.
*Corresponding author. Telephone: (614)-292-2721. Fax: (614)-292-7688.
E-mail: [email protected]
The presence of multiple stacked chambers and hot, porous crust
suggests that magma evolution is complex and involves polybaric
crystallization, magma mixing, and assimilation.
KEY WORDS: Iceland rift zones; cotectic crystallization; pressure;
depth; magma chamber; volcanic glass
I N T RO D U C T I O N
There is considerable interest in the depths of magma
chambers beneath active volcanoes in Iceland (e.g.
Soosalu & Einarsson, 2004; Sturkell et al., 2006) for four
main reasons. First, knowledge of the depths of magma
chambers is important for interpreting precursory activity
(seismic, deformation, gas emissions, etc.) to volcanic eruptions, and therefore for forecasting eruptions (e.g. Marti &
Folch, 2005). Second, knowledge of the depths of chambers
provides constraints on models for magma evolution,
because phase relationships and melt compositions vary as
a function of pressure (e.g. O’Hara, 1968; Thy, 1991; Grove
et al., 1992; Yang et al., 1996). Third, knowledge of the distribution of magma bodies is important to understand thermal
gradients, which affect variations in density and seismic
velocity in the crust (e.g. Kelley et al., 2005). Fourth, knowledge of the locations and sizes of magma chambers is essential to understand mechanisms of crustal accretion and
differentiation (e.g. Pan & Batiza, 2002, 2003).
Various methods have been used to estimate magma
chamber depths in Icelandic crust (Table 1). Most recent
studies utilize geodetic techniques (see Sturkell et al.,
2006), and yield results that in many cases (e.g. Krafla,
Grimsvo«tn, Katla, possibly Torfajo«kull) agree with those
estimated using geophysical methods (see Table 1),
although a magma chamber identified at a depth of 7 km
beneath Hengill^Hro¤mundartindur from geodetic data
ß The Author 2008. Published by Oxford University Press. All
rights reserved. For Permissions, please e-mail: journals.permissions@
oxfordjournals.org
JOURNAL OF PETROLOGY
VOLUME 49
NUMBER 3
MARCH 2008
Table 1: Summary of estimates of magma chamber depth
Center
Method
Depth (km)
Krafla
Seismology
3–57
Einarsson, 1978
Geodesy/gravity
3
Bjornsson et al., 1979
Geodesy
3
Tryggvason, 1980
Magnetotelluric
3, 5–7
Björnsson, 1985
Tryggvason, 1986
Geodesy
3, 5–10, 420
Geodesy
3
Ewart et al., 1991
Seismology
3
Brandsdottir et al., 1997
Geodesy
3
Sigmundsson et al., 1997
Geodesy
3, 45
Arnadottir et al., 1998
Geodesy/gravity
25
Rymer et al., 1998
Geodesy
24, 21
de Zeeuw-van Dalfsen et al., 2004
Geodesy/gravity
28, 21
de Zeeuw-van Dalfsen et al., 2006
424, 21
Depth range
Theistareykir
Petrology
30
Slater et al., 2001
Petrology
10–31
Maclennan et al., 2001
Depth range
Bláfjall
10–31
Petrology
7–10, 14
Schiellerup, 1995
Petrology
2–7, 14–26
Schiellerup, 1995
42, 26
Depth range
Askja
Geodesy
15–35
Tryggvason, 1989
Geodesy
15–35
Rymer & Tryggvasson, 1993
Geodesy
15–35
Sturkell & Sigmundsson, 2000
Geodesy/gravity
3, 16
de Zeeuw-van Dalfsen et al., 2005
Geodesy
3, 20
Pagli et al., 2006
Geodesy
3, 16
Sturkell et al., 2006
415, 20
Depth range
Grimsvötn
Reference
Gravity/magnetics
15–4
Gudmundsson & Milsom, 1997
Geodesy
416
Sturkell et al., 2003
Teleseismic P-wave delays
3–4 km
Alfaro et al., 2007
415, 4
Depth range
Kistufell
Petrology
435
Breddam, 2002
Torfajökull
Magnetotelluric
10–15, 15–25
Eysteinsson & Hermance, 1985
Gudmundsson, 1988
Magnetotelluric, geotherm
43
Seismology
8
Soosalu & Einarsson, 1997
Petrology
47
Gunnarsson et al., 1998
Seismology
8, 415
Soosalu & Einarsson, 2004
Seismology
514
Soosalu et al., 2006a
Geodesy
8
Sturkell et al., 2006
43, 25
Depth range
(continued)
466
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Table 1: Continued
Center
Method
Depth (km)
Hekla
Geodesy
8
Kjartansson & Grönvold, 1983
Magnetotelluric
8
Eysteinsson & Hermance, 1985
Geodesy
9
Sigmundsson et al., 1992
Geodesy
65
Linde et al., 1993
Geodesy
50–60
Tryggvason, 1994a,b
Geodesy
465
Jónsson et al., 2003
Seismic
?14
Soosalu & Einarsson, 2004
Geodesy
11
Sturkell et al., 2006
45, 414
Depth range
Katla
Seismology
3
Gudmundsson et al., 1994
Geodesy
47
Sturkell et al., 2003
Seismology
2
Soosalu et al., 2006b
42, 47
Depth range
Eyjafjallajökull
Geodesy
35
Sturkell et al., 2003
Geodesy
63
Pedersen & Sigmundsson, 2004, 2006
435, 63
Depth range
Vestmannaeyjar
Einarsson & Bjornsson, 1980
Seismology
15–25
Seismology
10–15
Gebrande et al., 1980
Petrology
18–28, 28–35
Furman et al., 1991
Petrology
10–17
Thy, 1991
410, 35
Depth range
Hengill–Hrómundartindur
Reference
Petrology
8–11
Hansteen, 1991
Geodesy
7
Sigmundsson et al., 1997
Geodesy
7
Feigl et al., 2000
Depth range
7–11
Króksfjördur
Petrology
55
Thingmuli
Petrology
35
Jónasson et al., 1992
Frost and Lindsley, 1992
Snaefell
Petrology
13
Hards et al., 2000
Austerhorn
Petrology
2
Furman et al., 1991
Estimated from Schiellerup (1995, fig. 4).
has not been identified from seismic data (Soosalu &
Einarsson, 2004). There are shallow chambers beneath
Katla, Grimsvo«tn and Eyjafjallajo«kull, deep chambers
beneath Vestmannaeyjar, both shallow and deep chambers
beneath Krafla and Askja, and intermediate depth chambers beneath Hekla and Torfajo«kull. These data, along with
evidence for lateral transport of magma along fissures and
mixing of magmas from different centers (Sigurdsson &
Sparks, 1978, 1981; McGarvie, 1984; Mrk, 1984; Fagents
et al., 2001), suggest extremely complex magma dynamics.
Geodetic and geophysical methods are most useful for
locating chambers beneath active volcanoes, whereas
petrological methods allow the depths of magma chambers
beneath both active and inactive volcanic centers to be
determined. Various petrological techniques are used to
determine the pressure, and hence depth, of crystallization
467
JOURNAL OF PETROLOGY
VOLUME 49
(Table 1), but some of the results reported for Icelandic
magmas (Table 1) are only qualitative. Nevertheless, petrological estimates agree well with those obtained using
other methods when direct comparison is possible
(e.g. Hengill, Torfajo«kull, and Vestmannaeyjar).
In this paper we report new estimates of the pressures of
crystallization of Icelandic magmas and use these to determine the depths of magma chambers. Pressures of crystallization are determined from experimentally established
phase equilibrium constraints using the method described
by Yang et al. (1996). The results differ from those obtained
in previous studies in two respects. First, quantitative pressure estimates were obtained using the compositions of
glasses, which unambiguously represent pre-eruptive liquid
compositions. Second, pressures were determined using
glasses from 28 localities so that the results are applicable
to a wide geographical area, and can be interpreted in
terms of crustal thickness. We discuss the implications of
the results for the structure and accretion of the crust, for
geothermal gradients, and for magma evolution.
G E O L O G I C A L B AC KG RO U N D
Iceland lies astride the Mid-Atlantic Ridge (MAR) and is
characterized by crust that is 20^40 km thick (Bjarnason
et al., 1993; Darbyshire et al., 1998, 2000a, b; Menke et al.,
1998; Kaban et al., 2003; Foulger et al., 2003; Leftwich
et al., 2005) compared with the 71 09 km thick normal
MAR crust (White et al., 1992; Bown & White, 1994). The
anomalously thick crust indicates higher melt productivity
in the underlying mantle compared with normal ridge segments (White & McKenzie, 1995). The higher melt productivity, together with geochemical differences between
Icelandic basalts and normal mid-ocean ridge basalt
(N-MORB; e.g. Schilling, 1973), provides evidence for a
mantle plume or upwelling centred beneath the northwestern edge of the Vatnajo«kull ice sheet (Breddam et al.,
2000; Darbyshire et al., 2000a; Leftwich et al., 2005;
Fig. 1). Alternative hypotheses to account for high melt productivity in the sub-Icelandic mantle have been discussed
by Foulger & Anderson (2005) and Foulger et al. (2005).
The full spreading rate for Iceland is 18^20 mm/year
(LaFemina et al., 2005). The subaerial ridge forms 50 km
wide neovolcanic or rift zones characterized by abundant
seismic and volcanic activity (Fig. 1). The Western Volcanic
Zone (WVZ) and the Northern Volcanic Zone (NVZ)
formed at 7 Ma following eastward relocation of the
spreading axis from the Snaefellsnes area (Hardarson
et al., 1997). The WVZ can be traced across the Reykjanes
Peninsula extensional leaky transform to the Reykjanes
Ridge (RR), whereas the NVZ is offset from the
Kolbeinsey Ridge (KR) by the 120 km long right-lateral
Tjo«rnes Transform Fault. Propagation of the NVZ to the
SW at 3 Ma (Steinthorsson et al., 1985) formed the
Eastern Volcanic Zone (EVZ), so that spreading is
NUMBER 3
MARCH 2008
partitioned between the WVZ and EVZ in southern
Iceland. These rift zones are connected by the complex
South Iceland Seismic Zone (SISZ), which accommodates
left-lateral transform motion, and by the Mid-Iceland Belt
(MIB), which may be a leaky transform (Oskarsson et al.,
1985; LaFemina et al., 2005) or a non-transform relay zone
(Sinton et al., 2005).
The neovolcanic zones are mostly covered by basalt
flows younger than 08 Ma. About 30 en echelon volcanic
systems have been recognized, which mostly consist of central shield or composite volcanoes transected by fissure
swarms. Eruptions from both volcanoes and fissure systems
occur in the WVZ and NVZ, but central volcanoes are
lacking in the EVZ. Many central volcanoes have calderas
with typical dimensions of 3 km 4 km, whereas fissure
swarms are 5^20 km wide and 40^150 km long, and are
often characterized by crater rows formed by scoria and
spatter. Other volcanic features include small lava shields,
tuff rings and maars, as well as hyaloclastite ridges, hyaloclastite cones and table mountains (tuyas) produced by
sub-glacial and sub-aqueous eruptions. Olivine tholeiites
and tholeiites dominate the volcanic products, but intermediate and acid volcanic rocks are also erupted, especially at central volcanoes.
Volcanic activity also occurs in three off-rift flank zones
(Fig. 1): the Western Flank Zone (WFZ), the Southern
Flank Zone (SFZ), and the Eastern Flank Zone (EFZ).
Eruptions occur at large shield or composite volcanoes
with calderas, and at small lava shields, tuff rings and
scoria cones. However, the extensive fissure swarms characteristic of the rift zones are absent, and eruptions
produce transitional and alkaline basalts along with intermediate and silicic compositions (Meyer et al., 1985;
Oskarsson et al., 1985). Volcanic products along the NVZ,
EVZ and SFZ change from tholeiitic basalts in northern
and central Iceland to Fe^Ti-rich transitional basalts in
southern Iceland and alkali basalts in Vestmannaeyjar.
M E T H O D S A N D DATA
Method for determining pressure
of crystallization
Various petrological techniques can be used to estimate
the pressure (P) and hence depth (z) of crystallization of
magmas (Table 1). The most appropriate method for use
with a large number of samples is based on comparing
the compositions of erupted melts with those of liquids
lying along P-dependent phase boundaries. Many basalt
magmas crystallize olivine (ol), plagioclase (plag), and
clinopyroxene (cpx), and their compositions can be compared with those of liquids lying along the ol^plag^cpx
cotectic boundary. The effect of pressure on the latter has
been determined experimentally (e.g. O’Hara, 1968;
Grove et al., 1992), and can be seen by recasting melt
468
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Fig. 1. Location of important geological features of Iceland. RR, Reykjanes Ridge; KR, Kolbeinsey Ridge; NVZ, Northern Volcanic Zone;
EVZ, Eastern Volcanic Zone; WVZ, Western Volcanic Zone; WFZ, Western Flank Zone; SFZ, Southern Flank Zone; EFZ, Eastern Flank Zone;
RP, Reykjanes Peninsula; TFZ, Tjornes Fracture Zone, SISZ, South Iceland Seismic Zone, MIB, Mid Iceland Belt; SP, Snaefjellsnes Peninsula; V,
Vatnajo«kull Glacier; L, Langjo«kull Glacier, and sample localities; Th, Theistareykir; Hb, Herdubreid; Bu, Burfell; Bf, Bla¤fjall Ridge; Ha, Halar;
Se, Seljahjalli; As, Askja; Hr, Hrimalda; Gi, Gigoldur; Sp, Sprengisandur; Ki, Kistufell; Kv, Kverkfjoll; Ba, Bardabunga; Gr, Grimsvo«tn; Vd,
Veidivo«tn; Lk, Laki; La, Langjo«kull; Hl, Hlodufell; Ra, Raudafell; E, Efstadalsfjall; Kf, Kalfstindar; Tg, Thingvellir; T, Torfajokull; He, Hengill
(includes Midfell and Maelifell); Ka, Katla; Hk, Hekla; Ge, Geitafell. Shaded areas indicate volcanically active regions. It should be noted that
samples from Sprengisandur were collected to the NW of the Tungnafellsjo«kull volcanic system (Meyer et al., 1985), but are believed to originate
from this center. In addition, pressures have been calculated for two samples from unspecified localities in the NVZ (NE) and the Reykjanes
Peninsula (Rk) (Meyer et al., 1985).
compositions into normative mineral components and projecting phase relations onto pseudoternary planes in the
system CaO^MgO^Al2O3^SiO2. Projection of phase relationships from plag onto the plane ol^cpx^qtz using the
recalculation procedure of Walker et al. (1979) clearly
shows the shift of the ol^plag^cpx cotectic towards ol with
increasing P (Fig. 2). Crystallization pressure can be estimated by comparing the projected compositions of natural
samples with the locations of cotectics on such diagrams,
and this method has been used to estimate crystallization
pressures for Hengill by Trnnes (1990), for Bla¤fjall
Table Mountain by Schiellerup (1995), for Kistufell by
Breddam (2002), and for Theistareykir by Maclennan
et al. (2001).
The shift of the ol^plag^cpx cotectic towards ol and plag
(see O’Hara, 1968; Grove et al., 1992) reflects the different
pressure dependences of cpx^liq, ol^liq and plag^liq
equilibria, with higher pressure favouring earlier
crystallization of cpx. This results in the development of a
trend of decreasing CaO with decreasing MgO in liquids
at an earlier stage of crystallization. Weaver & Langmuir
(1990), Langmuir et al. (1992), Danyushevsky et al. (1996),
Yang et al. (1996), and Herzberg (2004) have proposed
models to quantitatively estimate the crystallization pressure based on such relationships. These models are all calibrated with experimental data, and we have elected to use
that of Yang et al. (1996), who presented equations that
describe the composition of liquids along the ol^plag^cpx
cotectic as a function of P and T. Hence, the composition
of the liquid is used to predict the P (and T) of saturation
with ol, plag and cpx, as illustrated in Fig. 3 using two glass
analyses from Bla¤fjall Table Mountain (Schiellerup, 1995).
A series of liquid compositions that lie on the ol^plag^cpx
cotectic have been calculated from each glass analysis at
increments of 100 MPa. These predicted liquid compositions have been converted to normative mineral
469
JOURNAL OF PETROLOGY
VOLUME 49
Fig. 2. Position of the ol^plag^cpx cotectic at different pressures projected from plagioclase onto the pseudoternary plane Ol^Cpx^Qtz
using the method described by Walker et al. (1979). Pressures of cotectic
given in GPa. Locations of cotectics based on experimental data from
Bender et al. (1978), Walker et al. (1979), Grove & Bryan (1983), Spulber
& Rutherford (1983), Baker & Eggler (1987), Sack et al. (1987), Tormey
et al. (1987), Juster et al. (1989), Kinzler & Grove (1992), Shi (1993),
Thy & Lofgren (1994) and Yang et al. (1996). LOPC, liquid-olivineplagioclase-clinopyroxene cotectic.
components assuming that Fe ¼ FeO and projected from
plag onto the plane ol^cpx^qtz using the procedure of
Tormey et al. (1987) as modified by Grove et al. (1993).
Comparison of observed glass compositions and predicted
liquid compositions indicates crystallization at 100 MPa
for sample 1.8.5 and 670 MPa for sample 1.8.1. These pressures agree with those estimated by Schiellerup (1995) for
crystallization of units BII and BI at this locality.
Rather than use graphical methods, we have calculated
pressures using the procedure described in Appendix A.
An assessment of the accuracy (110 MPa, 1s) and precision (80 MPa, 1s) of the calculated pressures is given in
Appendix B.
Herzberg (2004) noted that pressures obtained with his
method can differ significantly (up to 300 MPa) from
those obtained using the method of Yang et al. (1996). The
methods were calibrated with different sets of experimental data, but even so the reason for this discrepancy is not
clear. We believe that the method used in this paper yields
reliable results for the following reasons: (1) Yang et al.
(1996) obtained results for MORB and Hawaiian samples
that are consistent with those obtained by other methods;
(2) pressures estimated by Maclennan et al. (2001) for
basalts from Krafla and Theistareykir using the Yang et al.
(1996) method agree with those obtained from clinopyroxene geobarometry; (3) pressures calculated for samples
from Midfell in this work agree to better than 60 MPa
NUMBER 3
MARCH 2008
Fig. 3. Comparison of predicted melt compositions saturated with
ol^plag^cpx at pressures between 0001 and 1GPa and observed glass
composition for two samples (1-8-1 and 1-8-5) from Bla¤fjall
Table Mountain (Schiellerup, 1995). *, Predicted melt compositions
at increments of 01GPa for each sample; , analyzed composition of
glass in sample 1.8.1; g, analyzed composition of glass in sample 1.8.5.
Melt compositions have been converted to normative mineral components and projected from plag onto the pseudoternary plane ol^cpx^qtz
using the procedure of Tormey et al. (1987). Comparison of predicted
and observed melt compositions indicates crystallization at
067 GPa for sample 1.8.1 and 01GPa for sample 1.8.5.
with results obtained by Gurenko & Sobolev (2006) using
the method of Danyushevsky et al. (1996), and to better
than 80 MPa with results obtained by these workers
using clinopyroxene geobarometry; (4) the results obtained
in the present study yield estimates of the depths of magma
chambers that are consistent with those obtained using
geodetic, geophysical, or other petrological methods. In
addition, we find that Herzberg’s (2004) method yields
negative and therefore unrealistic pressures for 22% of
the samples used to constrain the depth of magma crystallization in this paper. Nevertheless, it is apparent that more
experimental data are needed to refine petrological methods used for the geobarometry of magmas. In particular,
there is need for additional experiments to more closely
establish the composition of liquids along the ol^plag^cpx
cotectics for a range of basalt compositions over the pressure range 100^1000 MPa. The implications of using the
Herzberg (2004) method to calculate crystallization pressures for Icelandic magmas are briefly discussed below in
the section ‘Interpretation of pressure’.
Samples
Volcanic activity on Iceland produces about
012 km3/year of fresh lava, hyaloclastite, scoria and
spatter. Many recently erupted samples contain glass;
especially those erupted in sub-glacial and sub-aqueous
environments. Glass analyses are preferable to whole-rock
470
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
analyses for calculating the crystallization pressure,
because glasses represent samples of quenched melts.
Therefore, glasses formed from liquids in equilibrium
with ol, plag and cpx should have compositions that lie
exactly on the cotectic at the pressure of crystallization.
Some whole-rock samples represent melts, but others
represent mixtures of crystals and melt. It cannot be
assumed that the latter formed by closed-system crystallization, because the crystals may be of accumulative origin
or may represent xenocrysts (e.g. Trnnes, 1990; Hansteen,
1991; Re¤villon et al., 1999; Hansen & Gro«nvold, 2000). In
such cases, the whole-rock samples do not represent melts
and this explains why many whole-rock compositions are
displaced from glass compositions in pseudoternary projections (Fig. 4). The erroneous assumption that whole-rock
samples represent melts can lead to large errors in pressure
estimates (up to 1000 MPa for the examples shown in Fig. 4).
We have compiled 524 published glass analyses from
localities listed in the Supplementary Data (available for
downloading at http://www.petrology.oxfordjournals.org)
and shown in Fig. 1. Most glasses are from localities in the
NVZ, WVZ, and EVZ, but 44 are from Hekla and Katla
in the SFZ.
We also compiled 201 analyses of glasses in melt
inclusions. The compositions of melt inclusions can be
modified by post-entrapment crystallization and diffusive
re-equilibration with the host mineral (e.g. Danyushevsky
et al., 2002). Therefore, we used only those inclusion glasses
with compositions that plot along and within arrays defined
by groundmass glasses on variation diagrams. Using this
criterion, 64 glass inclusion analyses were selected and used
to supplement data for groundmass glasses.
Compositional data for samples from each locality are
summarized in Table 1 of the Supplementary Data.
Detailed discussion of chemical variations shown by the
glasses is beyond the scope of this paper, and we describe
only the general compositional characteristics along with
evidence that these characteristics are consistent with crystallization of ol, plag and cpx.
Based on their SiO2 to Na2O þ K2O ratios, 584 out
of 588 samples can be classified as basalts (Fig. 5a). The
remaining samples are basaltic andesites (two from
Hengill and two from Askja). In addition, most glasses
(553) are tholeiitic (subalkaline) according to the criterion
proposed by Macdonald (1968) for Hawaiian lavas
(Fig. 5b), and range in composition from olivine tholeiites
(normative Ol up to 18 wt %) to quartz tholeiites (normative Q up to 8 wt %). The remaining 35 glasses plot in the
alkaline field but only two of these (from Hekla) contain
normative nepheline [CIPW norms were calculated with
Fe2O3 and FeO contents fixed assuming log
fO2 ¼ FMQ 1 (McCann & Barton, 2004), where FMQ
is the fayalite^magnetite^quartz buffer] and we therefore
consider all of these samples to represent transitional
Fig. 4. Comparison of whole-rock and glass analyses for samples
from Herdubreid, Geitafell, and Hlodufell in projections from plag
onto the pseudoternary plane ol^cpx^qtz using the procedure of
Tormey et al. (1987). , Glass analyses; *, whole-rock analyses. All
analyses from Moore & Calk (1991). (a) Samples 336, 338, 347 and
350 from Herdubreid. The whole-rock samples have compositions
very similar to those of glasses from the same sample and may represent liquids that crystallized at similar pressure to the liquids represented by the glasses. (b) Samples 85T37 and 85T31 from Geitafell.
The compositions of the whole-rock samples differ from those of
glasses in the same sample and do not represent liquids that crystallized at the same pressure as the liquids represented by the glasses.
(c) Samples 85T3 and 85T12 from Hlodufell. The whole-rock composition of 85T3 is different from that of the glass in the same sample
and does not represent a liquid that crystallized at the same pressure
as the liquid represented by the glass. The whole-rock composition of
85T12 is similar to thatof the glass in the same sample may represent a
liquid that crystallized at about the same pressure as the liquid represented by the glass.
basalts. All glasses show strong enrichment in FeO as
MgO decreases (Fig. 5c).
Chemical variations shown by the glasses can be qualitatively explained by crystallization of ol^plag^cpx (spinel)
(Figs 5 and 6). Quantitative mass-balance models confirm
471
JOURNAL OF PETROLOGY
VOLUME 49
Fig. 5. General chemical characteristics of Icelandic glasses. (a) Total
alkalis vs SiO2 (wt %) with boundaries between magma types from
LeBas et al. (1986). B, basalt; BA, basaltic andesite; TB, trachybasalt;
BTA, basaltic trachyandesite. (b) Total alkalis vs SiO2 showing the
boundary between sub-alkaline (SA) or tholeiitic and alkaline
(A) compositions proposed by Macdonald (1968) for Hawaiian lavas.
(c) MgO vs FeO (wt % total Fe as FeO) illustrating the strong ironenrichment trend developed during differentiation.
NUMBER 3
MARCH 2008
Fig. 6. MgO vs Al2O3, CaO, and CaO/Al2O3 (wt %) illustrating
chemical variations produced by differentiation. Variations of Al2O3,
CaO, and CaO/Al2O3 with MgO allow identification of the mineral
phases that crystallized during magma evolution. The decrease in
Al2O3 with decreasing MgO (Fig. 7a) is consistent with crystallization of ol þ plag spinel, and many Icelandic basalts contain phenocrysts or microphenocrysts of these minerals (e.g. Meyer et al., 1985).
However, the strong decrease in CaO and slight decrease in CaO/
Al2O3 with decreasing MgO (Fig. 7b and c) requires crystallization
of cpx (see also Michael & Cornell, 1998; Herzberg, 2004).
472
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
that removal of these three phases accounts for majoroxide variations at many localities (e.g. Meyer et al., 1985;
Schiellerup, 1995; Maclennan et al., 2001, 2003), but this
does not necessarily mean that all glasses represent liquids
lying along ol^plag^cpx cotectics. The scatter shown on variation diagrams (Figs 5c and 6) indicates that the basalts
do not evolve along a single liquid line of descent (LLD).
This might indicate crystallization along ol^plag^cpx cotectics at different pressures (polybaric crystallization), but
other explanations are also possible, as described in a
later section of this paper (see also Herzberg, 2004). This
requires that results for the various localities are evaluated
in the context of geochemical and petrographic data. Some
samples (24) from the Hengill^Hro¤mundartindur volcanic
complex in the WVZ, as well as from Hrimalda, Gigoldur,
and Sprengisandur in the NVZ, have anomalously high
CaO contents as a result of clinopyroxene assimilation
(CaO/Al2O341; see Fig. 6) as discussed by Trnnes (1990).
These compositions lie outside the range of those used by
Yang et al. (1996) to calibrate the method for pressure calculation. Therefore, we do not report pressures for glasses
with anomalously high CaO/Al2O3.
R E S U LT S
Equilibration pressures
Pressures were calculated for all glasses, but some results
are unrealistic and others are considered unreliable. The
average uncertainty in P calculated for nearly all samples
is 5120 MPa, and is similar to that estimated from experimental data (Appendix B). In contrast, uncertainties calculated for 24 of the samples are substantially greater (up to
180 MPa). We consider these pressures to be unreliable,
and exclude most of them from further consideration. This
has no effect on the conclusions, because similar but more
reliable estimates of P are obtained for other samples from
the same localities. We made an exception for results
obtained for three samples from Kverkfjoll (Hoskuldsson
et al., 2006). Pressures calculated for these samples (790^860
160 MPa) are much higher than those obtained for other
samples from this locality (490 MPa), and may provide
insight into the evolution of magmas at this volcanic center.
Pressures calculated for 55 samples lie between 0 and
^100 MPa and can be interpreted as indicating crystallization at 01MPa if uncertainties are taken into account,
and indeed positive pressures close to 01MPa are obtained
for other samples from these localities. Nevertheless, we
exclude all samples that yield negative pressures from
further consideration.
The results demonstrate that Icelandic magmas crystallize over a wide range of P, from 01MPa to 1000 MPa
(Table 2), indicating that magmas evolve over a range of
depths in the crust and, perhaps, upper mantle. Few pressures exceed 600 MPa, and the majority of magmas appear
to have crystallized at P 300^400 MPa (Fig. 7a). It is
important to note that results obtained using the method
of Herzberg (2004) also indicate that Icelandic magmas
crystallize over a wide range of P, from 01MPa to
750 MPa, although the majority of magmas are predicted
to crystallize at 200 MPa rather than the 300^400 MPa
found here.
There is no correlation between calculated P and MgO
(Fig. 7b), and this is unusual in light of the positive correlation between P and MgO demonstrated for MORB glasses
by Michael & Cornell (1998). Again, it is important to
note that there is also no correlation between MgO and
values of P calculated using the method of Herzberg
(2004). This indicates that the lack of correlation between
P and MgO is not an artifact of our method of calculation.
However, for glasses from most localities there is a positive
correlation between P and MgO, whereas for glasses from
the Hengill complex there is a negative correlation
between P and MgO. The unusual negative correlation
for the Hengill complex samples is attributed to the effects
of crustal assimilation (see discussion below). The apparent
lack of correlation between P and MgO in Fig. 7b is therefore the consequence of plotting results for a relatively
large number of samples (120) from the Hengill complex
together with those for samples from other localities on
the same diagram.
Magma temperatures were calculated using the method
of Yang et al. (1996) (Table 2). Yang et al. (1996) noted that
their geothermometer reproduces temperatures to better
than 208C for most samples. The range of temperatures
obtained for Iceland glasses is 1232^11348C, and is similar
to that shown by MORB glasses. A similar range in T
(1254^10998C) is obtained using the olivine^melt geothermometer of Sugawara (2000). As expected, there are positive
correlations betweenTand P (Fig. 7c) and, with the exception of Hengill (see discussion below), betweenTand MgO.
Examples of results from four localities
Representative results obtained for samples from four
localities are shown in Fig. 8. Results obtained for some
localities suggest magma evolution at one pressure, for
example, 140 80 MPa at Gigoldur and 750 40 MPa at
Kalfstindar (Fig. 8a). Results for other localities such as
Bla¤fjall Table Mountain (Fig. 8b) indicate magma evolution at two distinct pressures (590 70 MPa and
100 50 MPa), whereas results for most localities indicate
magma evolution over a relatively wide range of pressure
rather than at one or two distinct pressures. This is illustrated by results from Laki (380 100 MPa) and Hengill
(260 170 MPa) shown in Figs 8c and 9d. The results for
Hengill are unusual in that many calculated pressures are
close to 01MPa (negative pressures were calculated for
36% of glasses from this locality) and, as described
above, there is a negative correlation between P and MgO.
473
JOURNAL OF PETROLOGY
VOLUME 49
NUMBER 3
MARCH 2008
Table 2: Summary of calculated pressures and temperatures
Zone and locality
Pressure (kbar)
n
Average
Temperature (8C)
Maximum
Minimum
Average
Maximum
Minimum
NVZ
Theistareykir
17
3201
5551
1314
12099
12182
12000
Herdubreid
22
3957
6574
1734
11914
12129
11735
3
3098
3240
2949
11747
11871
11501
19
3292
6950
172
11938
12152
11728
Halar
5
5111
5834
4206
12049
12129
11970
Seljahjalli
4
7756
9494
7019
12187
12321
12123
Askja
8
2728
5424
1248
11602
11807
11344
Hrimalda
2
1760
2400
1120
11982
12022
11942
Gigoldur
7
1704
2278
814
11965
12027
11903
Sprengisandur
7
3572
5927
538
11802
12046
11647
Kistufell
10
4014
5182
1432
12086
12145
11966
Kverkfjoll
6
5077
8594
895
11778
12007
11494
Northeastern
1
3313
–
–
11997
Bardabunga
17
2707
8417
649
11753
12098
11640
Grimsvötn
11
3081
5123
977
11751
11922
11492
Bardabunga–Grimsvötn
28
2842
8417
649
11751
12098
11492
Veidivötn
24
2099
6225
354
11715
11948
11500
Laki
67
3793
7655
1446
11714
11925
11615
Katla
17
4267
6563
1993
11838
12124
11598
Hekla
17
6262
10325
4251
12086
12275
11668
Hekla–Katla
34
5265
10325
1993
11962
12275
11598
Burfell
Bláfjall Ridge
–
–
EVZ
SFZ
WVZ
Langjökull
3
2534
6515
347
11934
12168
11787
Hlodufell
19
3566
5234
1682
11835
11985
11702
Raudafell
12
5167
7216
3129
11997
12106
11839
Efstadalsfjall
30
4769
8134
1981
11981
12222
11805
Kalfstindar
10
7450
7711
6509
12171
12189
12109
Thingvellir
10
3746
6274
2206
11940
12057
11819
Midfell
29
998
3694
04
11869
12014
11768
Maelifell
1
140
–
–
11892
Hengill
90
3148
7539
175
11824
12176
11422
120
2603
7539
04
11836
12176
11422
12
3999
5360
2814
11851
11977
11784
1
2017
–
462
3496
Hengill Complex
–
–
RP
Geitafell
Reykjanes
All data
10325
–
11846
04
11860
–
12321
Samples from Meyer et al. (1985).
Average, maximum, and minimum values based on results obtained for all samples from each locality.
474
–
11344
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Fig. 7. Summary of results obtained for all glasses excluding those
considered unrealistic or unreliable (see text for discussion).
(a) Histogram indicating range of calculated pressures. (b) Plot of
calculated pressure in GPa vs MgO. (c) Plot of T (8C) calculated
using the method of Yang et al. (1996) vs P (GPa).
DISCUSSION
Interpretation of pressure
The calculated pressures reflect the pressure of crystallization only for glasses that represent liquids lying along
ol^plag^cpx cotectics. Glass analyses from some localities
show compositional variations (e.g. increasing CaO and
CaO/Al2O3 with decreasing MgO) that are consistent
with crystallization of ol^plag rather than of ol^plag^cpx.
Glasses from other localities have compositions consistent
with ol^plag^cpx crystallization but occur in samples that
lack phenocrysts or microphenocrysts of cpx. In other
words, petrographic data provide no evidence for ol^plag^
cpx crystallization for these glasses. This is the ‘pyroxene
paradox’ recognized for many MORB and other lava
suites (e.g. Dungan & Rhodes, 1978; Fisk et al., 1982; Grove
et al., 1992; Elthon et al., 1995). There are several possible
explanations for this paradox, including: cotectic crystallization of ol^plag^cpx followed by crystallization of ol and
plag (with dissolution of cpx) during ascent, magma
mixing and crystallization accompanied by assimilation
of gabbroic crust (Meyer et al., 1985; Trnnes, 1990;
Hansteen, 1991; Schiellerup, 1995; Hansen & Gronvo«ld,
2000; Breddam, 2002; Maclennan et al., 2003; Gurenko &
Sobolev, 2006). In principle, the results for each locality
can be filtered to identify those melts in equilibrium with
ol plag spinel. Most of these provide no constraints on
crystallization pressure, but some can be used to place
limits on the pressure(s) of ol^plag^cpx cotectic crystallization as illustrated in Fig. 9. (see also Michael & Cornell,
1998; Herzberg, 2004).
Melts lying along path a^b in Fig. 9a crystallize ol þ plag
and have compositions that do not carry a signature of cpx
crystallization. Nominal pressures calculated for these
melts do not accurately reflect the pressure of magma evolution, but the lowest calculated pressure (for melts with
compositions close to b) represents an upper limit for
cotectic crystallization.
Melts plotting along the path c^d^e in Fig. 9b isobarically crystallize ol^plag^cpx along the cotectic and
then crystallize ol þ plag during ascent. Melts saturated in
ol þ plag with compositions between d and e carry a signature of cpx crystallization, but calculated nominal pressures
do not accurately reflect the pressure of magma evolution.
In this case, the highest pressure (for melts with compositions close to d) represents a lower limit for cotectic
crystallization.
Mixing of ol plag saturated melts with ol^plag^cpx saturated cotectic melts (Fig. 9c, path f^g^h) will produce
hybrids (e.g. path f^h) with the compositional signature
of cpx crystallization (Langmuir, 1989). The lowest pressure
calculated for the hybrid melts represents an upper limit
for cotectic crystallization.
There is petrographic evidence that some Icelandic
magmas, notably those from the Hengill complex, evolve
by crystallization of ol þ plag combined with assimilation
of cpx rather than by cotectic crystallization of ol^plag^cpx
(Trnnes, 1990; Hansteen, 1991; Gurenko & Sobolev, 2006).
If assimilation occurs after melts leave the ol^plag^cpx
cotectic and ascend towards the surface (Fig. 9d, path
475
JOURNAL OF PETROLOGY
VOLUME 49
NUMBER 3
MARCH 2008
Fig. 8. Histograms illustrating raw (unfiltered) results obtained for samples from various localities. (a) Results for Gigoldur (diagonal shading)
indicate crystallization at low pressure, whereas those from Kalfstindar (grey shading) indicate crystallization at high pressure. (b) Results for
Bla¤fjall Table Mountain indicate crystallization at two distinct pressures. (c) Glasses from Laki indicate crystallization over a wide range of
pressure. (d) Results for the Hengill complex (Hengill, Maelifell, Midfell) indicate crystallization over a wide range of pressure, although
many samples appear to have crystallized at P502 GPa.
j^k), as appears to be the case at Hengill, the highest calculated pressure represents a lower limit for cotectic crystallization. It should be noted, however, that it is extremely
difficult to discriminate between magmas that evolve via
ol^plag^cpx cotectic crystallization and those that evolve
via ol þ plag crystallization accompanied by assimilation
of cpx if petrographic or other (e.g. isotopic) evidence for
assimilation is lacking. Indeed, it is possible that the high
CaO contents and high CaO/Al2O3 ratios of some glasses
reflect assimilation of cpx. Consequently, the estimated
pressure of cotectic crystallization may be too low if all
samples from one locality (e.g. Hrimalda) have these characteristics (see Fig. 9d). However, the glasses from most
localities show a range of CaO and CaO/Al2O3, and it is
likely that the highest pressures calculated for these glasses
provide a reasonably accurate estimate of the pressure of
cotectic crystallization.
Plots of P vs MgO for samples from four localities are
shown in Fig. 10 to illustrate interpretation of the results
in light of the relationships described above. For each
locality, plots of CaO and CaO/Al2O3 vs MgO, together
with available petrographic data, were used to discriminate between ol þ plag þ cpx saturated melts and ol plag spinel saturated melts. For both Herdubreid and Hlodufell
(Fig. 10a and b), there is a positive correlation between
P and MgO for glasses in equilibrium with ol plag spinel
whereas there is no correlation between P and MgO for
glasses in equilibrium with ol þ plag þ cpx.
Calculated pressures for ol^plag^cpx saturated melts
from Herdubreid (Fig. 10a) define two arrays parallel to
the MgO axis that are interpreted to indicate cotectic
crystallization at 480 30 MPa and 310 20 MPa.
The low pressure calculated for one sample could be an
aberrant result, possibly caused by analytical error, or
could indicate crystallization along an even lower pressure
cotectic (170 MPa). This interpretation is preferred
because results obtained for other localities (Hlodufell,
Grimsvo«tn, Efstadalsfjall, Askja, Hrimalda, Kverkfjoll,
Theistareykir, Thingvellir) also provide evidence for lowpressure (90^200 MPa) cotectic crystallization.
476
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Fig. 9. Possible interpretations of calculated pressures illustrated using phase relationships projected from plag onto the pseudoternary plane
ol^cpx^qtz with the procedure of Tormey et al. (1987). Continuous lines mark the positions of the ol^plag^cpx cotectic (P in GPa). POP, pressure
of crystallization of ol^plag assemblages; POPC, pressure of crystallization of ol^plag^cpx assemblages; PMax, maximum pressure of crystallization;
PMin, minimum pressure of crystallization; PUL, upper limit for pressure of crystallization; PLL, lower limit for pressure of crystallization.
Arrows indicate changes in melt composition. (a) Melts evolve from a to b by polybaric crystallization of ol^plag. (b) Melts evolve from c to d
by isobaric crystallization of ol^plag^cpx and from d to e by polybaric crystallization of ol^plag. (c) Melts evolve from f to g by polybaric crystallization of ol^plag and from g to h by isobaric crystallization of ol^plag^cpx. The dashed line shows mixing between primitive melt f and evolved
melt h. (d) Melts evolve from i to j by isobaric crystallization of ol^plag^cpx and from j to k by polybaric crystallization of ol^plag accompanied by
assimilation of cpx.
Pressures calculated for ol^plag^cpx saturated melts from
Hlodufell (Fig. 10b) define a single broad array parallel to
the MgO axis, and could be interpreted to indicate cotectic crystallization between maximum (PMax ¼ 380 MPa)
and minimum (PMin ¼ 290 MPa) pressures. However, the
range in pressure is about equal to the expected precision
of the results, and therefore, the results are interpreted
as indicating cotectic crystallization at 340 40 MPa.
As with Herdubreid, the low pressure (170 MPa) calculated for one glass is interpreted to indicate crystallization
along a lower pressure cotectic.
Calculated pressures for ol^plag^cpx saturated melts
from Grimsvo«tn (Fig. 10c) show more scatter than those
from Hlodufell, and the range is greater than estimated precision.The results are interpreted to indicate cotectic crystallization between maximum (PMax ¼ 510 MPa) and
minimum (PMin 100 MPa) pressures, but there is clear evidence for cotectic crystallization at intermediate pressures
(see below). A more realistic estimate of the pressure of lowpressure cotectic crystallization is obtained by averaging
the results for samples that crystallized at similar pressures.
Four samples yield pressures between 100 and 200 MPa,
whereas other samples yield pressures 4300 MPa (Fig. 11c).
The average pressure for the four samples is 150 60 MPa
and is the preferred value for low-pressure cotectic crystallization. Preferred values for low-pressure cotectic crystallization have been calculated for Theistreykir, Bardabunga,
Grimsvo«tn,Veidivo«tn, Katla, Hekla, and Thingvellir. Likewise, results for samples that crystallized at similar high
pressures can be averaged and used to calculate preferred
values for high-pressure cotectic crystallization at some
localities (e.g. Hengill).
Finally, glasses from Kalfstindar represent melts saturated with ol þ plag, and the lowest calculated pressure
provides an upper limit (PUL ¼ 650 MPa) for cotectic crystallization (Fig. 10d). It should be noted that pressures calculated for some ol^plag saturated melts from Herdubreid
and Hlodufell are also higher than those calculated
for ol^plag^cpx saturated melts from these localities
(Fig. 11a and b).
477
JOURNAL OF PETROLOGY
VOLUME 49
NUMBER 3
MARCH 2008
Fig. 10. Interpretation of results for selected localities illustrated by plots of pressure P vs MgO. *, Melts in equilibrium with ol plag spinel
assemblages; , melts in equilibrium with ol^plag^cpx assemblages. PC, pressure of cotectic crystallization; PMax, maximum pressure of crystallization; PMin, minimum pressure of crystallization; PUL, upper limit for pressure of cotectic crystallization. (See text for further discussion.)
We have noted that results from the Hengill complex are
anomalous compared with those obtained from other volcanic centers. Plots of CaO and CaO/Al2O3 vs MgO are
consistent with crystallization of ol þ plag þ cpx. Nominal
pressures calculated for 120 glasses range from 04 to
754 MPa (Table 2), and the simplest interpretation is that
cotectic crystallization occurred over a range of pressures
between these values. However, detailed studies byTrnnes
(1990), Hansteen (1991), and Gurenko & Sobolev (2006)
indicate that this interpretation is oversimplified.
Trnnes (1990) has divided Hengill glasses into four
groups based primarily on MgO and CaO content. We
have followed his subdivision for Groups I and II, but
have combined glass analyses from his two other groups
into a single group (Group III). Nominal pressures for
MgO-rich samples from Group I range from 01 to
311MPa (Fig. 11). These samples contain Al- and Cr-rich
endiopside as resorbed phenocrysts and as components
of partly disaggregated gabbroic nodules, suggesting that
the MgO-rich Group I magmas have assimilated crustal
material. The highest calculated pressure for these
samples (311MPa) represents the lower limit for cotectic
crystallization (e.g. Fig. 9d). On the other hand, nominal
pressures for MgO-poor Group III samples range from
340 to 613 MPa (Fig. 11). Some of these samples contain
augite microphenocrysts, suggesting that the MgO-poor
Group III magmas have crystallized ol þ plag þ cpx along
cotectics between a maximum (613 MPa) and minimum
(340 MPa) pressure. The lower limit of cotectic crystallization for Group I glasses and the minimum pressure of
cotectic crystallization for Group III glasses are virtually
identical (Fig. 11) and provide a tight constraint on lowpressure cotectic crystallization (325 MPa). Pressures calculated for Group II glasses encompass values obtained
for Group I and Group III glasses (Fig. 11). Crystallization
of MgO-rich Group I melts at lower pressure than
the MgO-poor Group III melts produces the unusual
negative correlations between P and MgO and P and T
observed for Hengill samples. The strong evidence that
Group I magmas have interacted with gabbroic crust
suggests that the negative correlation between P and
MgO for Hengill samples results from assimilation combined with crystallization as magmas ascend beneath this
volcanic complex.
478
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Pressures of ol^plag^cpx cotectic crystallization for all
localities are listed in Table 3. A striking feature of the
results (Fig. 12) is the large number of samples (85%) that
have crystallized along either low-P (40^220 MPa) or relatively high-P (430^1030 MPa) cotectics. This near-bimodal
distribution of pressures contrasts strongly with the distribution of nominal pressures obtained using unfiltered
results (Fig. 7a), and illustrates the importance of evaluating geochemical and petrographic data for samples from
the various localities to identify melts that lie along
ol^plag^cpx cotectics. It is also clear that some melts crystallized along ol^plag^cpx cotectics at pressures between 220
and 430 MPa (Table 3).
Effect of H2O
Fig. 11. Interpretation of results for glasses from the Hengill complex.
(a) Histogram of results obtained for MgO-rich (Gp I) and MgOpoor (Gp III) samples. The Gp I samples evolve via polybaric crystallization of ol^plag accompanied by assimilation of cpx, which allows
the lower limit of cotectic crystallization to be determined (PLL). The
Gp III samples evolve via polybaric crystallization of ol^plag^cpx,
which allows the minimum pressure of cotectic crystallization to be
determined (PMin). (b) Plot of P vs MgO (wt %) *, Gp I samples;
, Gp III samples. PMax, maximum pressure of cotectic crystallization; PMin, minimum pressure of cotectic crystallization; PLL, lower
limit for pressure of cotectic crystallization.
In summary, melts that lie along ol^plag^cpx cotectics
can be identified allowing crystallization pressures
to be established. Similarly, melts in equilibrium with
ol plag can be identified and used to place upper limits
(PUL) and/or lower limits (PLL) on crystallization pressures. Finally, melts that provide no constraints on crystallization pressure can be identified and filtered out of
the results.
The method used to calculate pressure is based on experiments carried out under nominally anhydrous conditions,
and it is important to assess the possible effect of water on
the results. Experimental studies show that addition of
H2O leads to expansion of the stability field of olivine
and contraction of the stability field of plagioclase
(Kushiro, 1969; Nicholls & Ringwood, 1973; Baker &
Eggler, 1987; Sisson & Grove, 1993), so that the location of
the ol^plag^cpx cotectic will shift with increasing water
content at constant pressure. However, relatively high
water contents are required to produce large shifts in the
position of the cotectic, and the water contents of
Icelandic magmas are probably too low to have much
effect. Measured water contents range from 01 to 1 wt %
(Jamtveit et al., 2001; Nichols et al., 2002), and agree with
water contents calculated for 88 Icelandic glasses using
the method of Danyushevsky et al. (1996) and
Danyushevsky (2001) (average 018 wt %, range 0^094
wt %). Herzberg (2004) found no correlation between calculated pressures and H2O for MORB with 01^08 wt %
H2O. Additional evidence that calculated pressures have
not been significantly affected by water is provided by the
close agreement (40 40 MPa) between pressures calculated for single samples using two projections (see
Appendix A), one from plag onto the plane ol^cpx^qtz, the
other from ol onto the plane plag^cpx^qtz. Addition of water
causes the ol^plag^cpx cotectic to shift towards cpx (away
from ol) in the plag projection and away from cpx (towards
plag) in the ol projection. In other words, water affects the
stability fields of ol, plag and cpx differently and has a different effect on the positions of the ol^plag^cpx cotectic in
these two projections. Accordingly, pressures calculated
for hydrous melts will be associated with large uncertainties reflecting large differences between pressures calculated from the two projections (see Appendix A, Table A1).
We noted in a preceding section that the uncertainty in
calculated pressure for most samples is similar to that estimated from anhydrous experimental data, and it is concluded that the results reported in Tables 2 and 3 closely
reflect the pressures of crystallization. Indeed, pressures
479
JOURNAL OF PETROLOGY
VOLUME 49
NUMBER 3
MARCH 2008
Table 3: Pressures and depths of cotectic crystallization
Zone and
P
locality
(MPa)
Range
Comment
Range
z
(km)
Range
P
Comment
(MPa)
Range
z
(km)
Range
P
Comment
(MPa)
Range
z
(km)
NVZ
Theistareykir
555
–
Max.
Theistareykir
555
–
Max.
Herdubreid
484
30
195
–
–
–
–
–
–
–
–
131
–
149
27
173
–
170
11
3080
19
109
066
–
–
3100
15
109
051
–
–
186
–
–
–
–
–
175
–
Min.
–
–
Preferred
53
1
Min.
61
–
–
–
Min.
62
–
–
Burfell
–
–
Bláfjall Ridge
529
–
Halar
511
76
180
23
–
–
–
–
–
–
–
Seljahjalli
717
25
253
089
–
–
–
–
–
–
–
Askja
536
9
189
032
–
–
–
–
185
56
65
199
Hrimalda
–
–
–
–
–
–
–
–
176
90
62
23
Gigoldur
–
–
–
–
–
–
–
–
170
90
60
26
Sprengisandur
593
–
Max.
209
–
–
–
–
–
54
–
Min.
19
–
Kverkfjoll
859
–
Max.
289
12
–
–
–
–
90
–
Min.
32
–
Kverkfjoll
822
033
–
–
–
90
–
Min.
Kistufell
518
–
Northeastern
–
–
Bardabunga
842
–
Max.
Bardabunga
842
–
Max.
Grimsvötn
512
–
Max.
Grimsvo¨tn
512
–
Max.
Max.
Preferred
Max.
182
–
3250
–
–
–
3310
–
296
–
Min.
–
114
–
143
–
50
–
117
–
–
–
–
–
EVZ
Veidivötn
622
–
Max.
Veidivo¨tn
622
–
Max.
Laki
766
–
Max.
180
219
269
–
–
–
–
–
–
–
65
–
–
–
–
–
177
62
–
–
–
–
97
–
–
–
–
–
154
58
–
–
–
–
35
–
–
–
–
–
141
64
–
–
–
–
145
–
164
16
1993
–
–
–
177
17
Min.
Preferred
Min.
Preferred
Min.
–
62
–
54
–
–
22
–
20
–
Preferred
50
23
Min.
51
–
–
70
–
–
–
–
–
–
–
–
–
SFZ
Katla
656
Max.
231
4663
45
Hekla
1032
Max.
363
–
4250
–
Hekla
956
Preferred
340
26
5017
48
Max.
74
Preferred
Min.
Preferred
WVZ
Langjökull
652
–
229
–
–
Hlodufell
–
–
–
–
3350
37
Raudafell
–
–
–
–
3130
–
Efstadalsfjall
5170
34
182
126
3220
36
Kalfstindar
6510
–
Upper limit
229
–
–
Thingvellir
6270
–
Max.
221
–
2210
–
Thingvellir
6270
–
Max.
–
–
2570
33
–
–
3110
–
Max.
–
–
–
–
196
088
–
–
–
–
74
–
Min.
26
118
13
1680
–
Min.
59
–
110
–
–
–
–
–
113
13
1980
–
70
–
–
–
–
–
–
–
Min.
–
–
–
–
–
–
Preferred
91
12
–
–
109
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
–
Upper limit
Hengill Gp I
–
–
Hengill Gp II
5920
–
Hegill Gp II
5560
025
Hengill Gp III
6130
–
–
–
3400
–
Min.
120
–
–
–
Hengill Gp III
5030
086
Preferred
177
301
3400
–
Min.
–
–
–
–
5070
26
Preferred
178
091
3460
32
122
114
–
–
–
–
–
–
–
–
–
–
–
2020
–
71
–
Preferred
Max.
Lower limit
Min.
RP
Geitafell
Reykjanes
–
See text for discussion of preferred values of pressures (italics).
480
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Depths of magma chambers and
magma plumbing systems
Fig. 12. Histograms showing averaged results for the pressure of
cotectic crystallization at various volcanic centers. (a) Cotectic crystallization at relatively high pressure. (b) Cotectic crystallization at
relatively low pressure. Results indicating cotectic crystallization at
intermediate pressures (P4022 GPa, 5043 GPa) are omitted for
clarity.
associated with large uncertainties have been filtered out of
the results. It is unlikely that these samples crystallized at
higher water contents than most Icelandic magmas,
because water contents calculated using the method of
Danyushevsky et al. (1996) and Danyshevsky (2001)
(average 028 wt %, range 0^085 wt %) are similar to
those obtained for other Icelandic glasses. Use of this
method also indicates that samples that yield nominally
anhydrous pressures between 0 and ^100 MPa did not
crystallize from magmas containing unusually high water
contents, supporting the interpretation that these samples
crystallized at 01MPa taking uncertainties into account.
The pressures reported in Table 3 are those at which melts
are last saturated with ol, plag, and cpx. Ascending magmas
must pause and crystallize for a sufficiently long period of
time to reach multiple saturation, and this most probably
occurs in magma chambers. The melt is then rapidly
erupted from the chamber and quenched to form glass.
The depths of magma chambers may therefore be calculated from pressures of cotectic crystallization using an
appropriate value for crustal density. We have used a value
of 2900 kg/m3 to calculate the depths listed in Table 3.
Preferred values for the pressure of cotectic crystallization
were used in the calculations wherever possible.
Comparison with results obtained using geodetic and/or
geophysical methods (see Table 1) reveals agreement for
the depths of chambers beneath Askja (65 and 189 km
vs 13^35 and 16^20 km) and Grimsvo«tn (54 km vs
15^4 km), and reasonable agreement for the depth of a
chamber beneath Hengill (115 km vs 7 km). Our estimate
of a minimum depth of 175 km for a chamber beneath
Hekla is greater than the recent estimate of 11km based
on strain data (see Sturkell et al., 2006), but is consistent
with recent analysis of seismic data (Soosalu & Einarsson,
2004), which provide no evidence for a significant magma
body at depths 514 km. However, we find no evidence for
the shallow chamber (2^5 km) identified beneath Katla in
geophysical and geodetic studies. This probably indicates
that this chamber does not contain basalt, because
Soosalu et al. (2006b) suggested that a cryptodome containing relatively viscous (silica-rich) magma occurs at shallow (2 km) depths beneath Katla.
Comparison with results obtained in other petrological
studies (see Table 1) reveals excellent agreement in the
case of Bla¤fjall Table Mountain (62 and 186 km vs 2^7
and 14^26 km) and the Hengill volcanic complex
(11^12 km vs 8^12 km). However, the depths obtained for
crystallization at Kistufell (518 km) are much lower than
those obtained by Breddam (2002) (430 km), and the
reason for this discrepancy is unclear.
Our results clearly suggest that magma chambers
are located at different depths in the Icelandic crust
(Figs 13^15), and this is corroborated by geodetic and geophysical studies. There is evidence for only one chamber
located in the shallow (Hrimalda and Gigoldur in the
NVZ), middle (Burfell in the NVZ, Raudafell in the
WVZ), or deep (Kalfstindar in the WVZ) crust beneath
some centers, but there is clear evidence for two or more
stacked chambers beneath most centers (Figs 13 and 15).
Shallow and deep chambers occur beneath Askja, Bla¤fjall
Table Mountain (including Halar and Seljahalli; see
Schiellerup, 1995) and Langjo«kull (Fig. 15a), whereas shallow, intermediate and deep chambers occur beneath
Herdubreid and Efstadalsfjall (Fig. 15b). Dikes presumably
481
JOURNAL OF PETROLOGY
VOLUME 49
NUMBER 3
MARCH 2008
Fig. 14. Depths of intermediate depth magma chambers along the
rift zones. Error bars show uncertainty in estimated depth. Full versions for abbreviations for volcanic centers are given in the
Supplementary Data and Fig. 1 caption. Volcanic centers are arranged
from south to north along the y-axis. The shaded grey bands show the
range of depth of shallow and deep chambers along the NVZ and
WVZ.
Fig. 13. Depths of shallow and deep magma chambers along the rift
zones. Error bars show uncertainty in estimated depth. Full versions
for abbreviations for volcanic centers are given in the Supplementary
Data and Fig. 1 caption. Volcanic centers are arranged from south to
north along the y-axis. The shaded grey bands show the range of
depth of shallow and deep chambers along the NVZ and WVZ.
The greater depths of chambers beneath Kverkfjoll, Hekla,
Grimsvo«tn-Laki, and Bardabunga^Veidivo«tn, and the lack of evidence for shallow chambers beneath Katla and Hekla, should be
noted.
connect chambers located at different depths, and provide
conduits for magmas to reach the surface.
Results for Theistareykir, Sprengisandur, Kverkfjoll,
Bardabunga^Veidivo«tn, Grimsvo«tn^Laki, Langjo«kull,
Thingvellir, and Hengill are consistent with the presence
of both shallow and deep chambers, and magmas also
appear to have crystallized at intermediate depths
although there is no evidence that this occurred in a discrete chamber (see Fig. 10c). As discussed above, it is possible that these magmas evolved by high-pressure cotectic
crystallization of ol þ plag þ cpx followed by crystallization
of ol þ plag (with partial or complete resorption of cpx)
during ascent. However, petrographic data indicate that
at least some glasses are from magmas that evolved via
ol þ plag þ cpx crystallization, which suggests the presence
of cooling and crystallizing magma bodies or pockets
throughout the middle and lower crust. Multiple, stacked,
discrete chambers might exist beneath some volcanoes as
shown in Fig. 15c. However, it is also possible that crystallization occurs in extensive crystal mush zones as suggested
by Hansen & Gro«nvold (2000). In addition, crystallization
probably occurs at various depths in feeder dikes and conduits during lulls in eruptive activity.
Recent work suggests that MORBs partially crystallize
over a range of pressure from 1 to 1000 MPa (Michael &
Cornell, 1998; Le Roex et al., 2002; Herzberg, 2004). The
average crustal thickness along mid-ocean ridges is
71 09 km (White et al., 1992; Bown & White, 1994),
so that crystallization of MORB must begin in the upper
mantle. Some Icelandic magmas may also crystallize in
the mantle. The glasses have chemical characteristics of
482
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Fig. 15. Schematic representation of plumbing systems beneath
Icelandic volcanoes. There is evidence for two or more chambers
beneath most volcanic centers (see also Gudmundsson, 2000). (a)
Shallow (a) and deep chambers (b) linked by a system of conduits
and dikes (c). A plumbing system similar to this may exist beneath
Askja, Bla¤fjall Table Mountain and, possibly, Langjo«kull. (b) Shallow
(a), intermediate (b) and deep chambers (c) linked by a system of
conduits (d) and dikes (e). A plumbing system similar to this may
exist beneath Herdubreid and Efstadalsfjall. (c) Shallow (a) and
deep chambers (b) with a plexus of small chambers or magma
bodies in the lower crust (c) linked by a system of conduits and dikes
(d). This type of plumbing system probably exists beneath
Theistareykir, Sprengisandur, Kverkfjoll, Bardabunga^Veidivo«tn,
Grimsvo«tn^Laki, Langjo«kull, Thingvellir, and Hengill. The plumbing
systems beneath Katla and Hekla in the SFZ may also be similar to
the middle^lower crustal section.
evolved magmas and must have formed from primary
magmas by crystallization in the deep crust or mantle.
Conclusive evidence for crystallization in the mantle is
lacking, but the high pressures (up to 800 MPa) obtained
for ol plag saturated melts from Herdubreid, Efstadalsfjall, and Kalfstindar (Table 2) are consistent with crystallization at upper mantle depths.
Thickness and structure of Icelandic crust
The factors that determine the depth at which magma
chambers form are complex and are poorly understood,
but include the buoyancy force (reflecting the density contrast between melt and surrounding rocks), any excess
pressure over the buoyancy force, and local variations in
stress conditions in the crust (Gudmundsson, 2000). It can
be assumed to a first approximation that variations in the
buoyancy force, excess pressure and stress conditions are
negligible for magmas entering the base of the crust along
the rift zones. Therefore, the main control of the location of
deep chambers is likely to be a lithological boundary associated with a density discontinuity. These chambers occur
over a remarkably small range of depths for volcanic centers in the NVZ (195 26 km, excluding Kverkvjoll) and
WVZ (209 23 km, including samples from Geitafell)
(Fig. 13b). The small range of depths for the rift zones
(201 25 km) suggests that the deep chambers occur at
or near the crust^mantle boundary (Kelley et al., 2004;
Kelley & Barton, 2005). There is excellent agreement
between magma chamber depth beneath Theistareykir
and crustal thickness inferred from seismic data
(Brandsdo¤ttir et al., 1997; Staples et al., 1997) beneath
nearby Krafla (195 vs 19 km). Likewise, there is very
good agreement between magma chamber depths and
crustal thickness inferred from seismic data (Bjarnason
et al., 1993; Weir et al., 2001) for Geitafell (178 09 vs
16 km), Hengill (189 18 vs 175^24 km), Hekla (34 26
vs 30^35 km), and Katla (231 16 vs 20^25 km).
Seismic data do not provide reliable estimates of the
depth of the crust^mantle transition beneath other centers,
and crustal thickesses must be estimated from gravity data
(Darbyshire et al., 1998, 2000a; Allen et al., 2002; Kaban
et al., 2002; Fedorova et al., 2005; Leftwich et al., 2005) and
from the relationship between Moho depth and height
above sea level given by Gudmundsson (2003). These methods do not allow small-scale variations in crustal thickness
to be resolved, and there is considerable uncertainty in the
estimated crustal thickness beneath most volcanoes. In
addition, the crust^mantle boundary may represent a
transition zone 5 3 km thick (Foulger et al., 2003).
Despite this, there is reasonable agreement between estimated magma chamber depth and the base of the crust
beneath Kverkfjoll (29 vs 30^35 km) and beneath
Bardabunga^Veidivo«tn and Grimsvo«tn^Laki (27^29 km
vs 30^40 km) given that absolute uncertainties in calculated magma chamber depths are 39 km. These volcanic centers are located on or near the Vatnajo«kull ice cap
(Fig. 1) in a region of thicker crust above the thermal and/
or compositional anomaly in the mantle (Darbyshire et al.,
2000a; Kaban et al., 2002; Fedorova et al., 2005; Leftwich
et al., 2005). Also, depths estimated for magma chambers
and the base of the crust are similar for the Bla¤fjall complex (25 km), Thingvellir (22 km vs 20^25 km), and
Kalfstindar (23 km vs 20^25 km).
The agreement between magma chamber depths and
Moho depths provides strong evidence that magmas pond
at the base of the crust beneath most volcanic centers in
Iceland. However, there is poor correlation between
483
JOURNAL OF PETROLOGY
VOLUME 49
Moho depth and magma chamber depth for some centers,
and it is possible that the true, maximum depth of chambers beneath these centers is greater than that reported in
Table 3. As noted above, some ol^plag-saturated melts from
Herdubreid and Efstadalsfjall appear to have crystallized
at higher pressure than the maximum value calculated for
cotectic melts. Also, we are not confident that the maximum depth of chambers has been established for Langjo«kull, Sprengisandur and Askja, because of the relatively
small number of samples available for study.
The depth of shallow chambers is also relatively constant
for volcanic centers along the rift zones (Fig. 13a). The
average depths are 52 16 km for the NVZ, 61 21km
for the WVZ (including samples from the Reykjanes
Peninsula), and 54 06 km for the EVZ. The average
depth for all three rift zones is 55 16 km, indicating
that the chambers are located in upper crust that has an
average thickness of 5^7 km (Darbyshire et al., 2000b;
Foulger et al., 2003). [Various workers use different definitions of upper, middle and lower crust for Iceland (see
Foulger et al., 2003). For this paper, we arbitrarily define
upper crust as that 57 km, middle crust as that from 7 to
15 km, and lower crust as that 415 km.] The upper crust is
a heterogeneous mixture of flows, hyaloclastites, and intrusive rocks that have been metamorphosed under greenschist- to amphibolite-facies conditions. The observed rapid
increase in seismic velocity with depth probably reflects
a decrease in the proportion of hyaloclastites and decrease
in porosity and permeability (Foulger et al., 2003, and references therein). It seems likely that shallow chambers form
at a level of neutral buoyancy at or near the base of the
upper crust, although the exact depth may be controlled
by local variations in lithology and/or stress conditions
(Gudmundsson, 2000).
Additional petrological studies to refine estimates of
magma chamber depth are desirable to provide tighter
constraints on the thickness of the whole crust and of the
upper crust, as well as to define regional variations in
these thicknesses. The occurrence of discrete chambers
located at intermediate depths (Fig. 14) also has important
implications for models of crustal structure. Those in the
NVZ and WVZ occur at a relatively constant depth
(114 04 km). There is no convincing evidence for a seismic discontinuity at this depth, and it seems likely that
these chambers are located at a phase transition (amphibolite^granulite facies?), and/or at a rheological boundary.
Fedorova et al. (2005) suggested that the base of the elastic
lithosphere beneath Vatnajo«kull is 15 km deep, and is
underlain by relatively ‘soft’ lower crust. The ‘intermediate
depth’ chambers beneath Katla and Hekla are significantly
deeper (16^18 km) than the intermediate depth chambers
in the active rift zones, which may indicate a different
lithological structure or stress regime in the crust of
the SFZ.
NUMBER 3
MARCH 2008
The evidence for cooling and crystallizing magma
bodies or pockets throughout the middle and lower crust
is consistent with results obtained by Tryggvason (1986)
for Krafla and by Maclennan et al. (2001) for Theistareykir
(Table 1). The presence of multiple magma chambers at
different depths beneath many volcanic centers implies
high geothermal gradients in the crust (see also Meyer
et al., 1985; Bjarnason et al., 1993; Maclennan et al., 2001;
Kelley et al., 2005; Leftwich et al., 2005). Petrological data
therefore strongly suggest that the middle and lower crust
is relatively hot and porous, in contrast to the relatively
cool, rigid crust proposed in some geophysical studies
(e.g. Bjarnason et al., 1993; Menke & Sparks, 1995). However, receiver function data (Darbyshire et al., 2000b), combined surface- and body-wave constraints (Allen et al.,
2002), combined results from explosion seismology and
receiver functions (Foulger et al., 2003), and combined
results from seismic, topography, and gravity data
(Fedorova et al., 2005) all provide evidence for low-velocity
zones (LVZs) in the crust. These may reflect variations in
temperature or lithology, and/or the presence of melt.
Most workers favor an explanation involving both high
temperatures and the presence of small amounts of melt.
Darbyshire et al. (2000b) suggested that a prominent LVZ
at 10^15 km reflects the presence of partially molten sills in
the lower crust beneath Krafla (see Fig. 15c), whereas
Fedorova et al. (2005) proposed that crust deeper than
15 km beneath Vatnajo«kull contains melt. Allen et al.
(2002) identified LVZs in both the upper (0^15 km) and
lower (415 km) crust. The former are thought to reflect
regions of high temperature and possible melt, whereas
the latter (beneath Vatnajo«kull) are thought to reflect the
thermal halo associated with the fluxing of magma from
the mantle to the upper crust. Although detailed correlation of LVZs with the occurrence of magma chambers is
not possible, these results suggest that seismic and petrological data can be reconciled to develop an internally consistent model for Icelandic crust.
Crustal accretion models
There is no doubt that some crustal accretion occurs at
shallow depths in Iceland by eruption of lavas and hyaloclastites that are buried beneath the products of later eruptions, and by crystallization of gabbros in shallow
chambers. Shallow accretion is one of the models proposed
for the formation of oceanic crust. Melt is supplied directly
from the mantle to a shallow axial melt lens (Sinton &
Detrick, 1992). Some of this melt is erupted whereas
the rest crystallizes to form gabbros that subside to form
oceanic crust. However, it is highly unlikely that the thick
(20^40 km) crust in Iceland forms solely by crystallization
in shallow chambers to form gabbroic crust that subsides in
response to plate extension. It is probable that accretion
also results from crystallization in chambers located at
Moho depths (underplating), and by crystallization in
484
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Fig. 16. Petrological model for Icelandic crust. Shallow chambers
are located near the base of the upper crust, and magma is fed to the
surface via dike systems. Crustal accretion occurs by eruption of lavas
and hyaloclastites, and by crystallization of gabbro in the chamber.
The middle and lower crust consists of smaller chambers, pockets of
magma, and conduits within a mush zone extending to the Moho.
A chamber at the Moho feeds magma to the shallower chambers and
directly to the surface. The mush zone may extend to shallower depths
than shown (i.e. to the base of the upper crust). Crustal accretion
occurs by crystallization of magma within the mush zone and at the
base of the crust to form gabbros. Magma chambers or pockets of
magma may also occur in the underlying mantle. (See text for
details.)
dikes, melt pockets, or discrete chambers at various depths
in the crust (intracrustal accretion). A model for accretion
of Icelandic crust is shown in Fig. 16, and is consistent with
recent seismic reflection and petrological data for midocean ridges (e.g. Pan & Batiza, 2002, 2003; Maclennan
et al., 2004; Singh et al., 2006). These data indicate that
lower crust in axial regions is at least locally porous on an
intergranular scale, and suggest that accretion occurs by
crystallization of small pockets of melt over a range of
depths between the shallow melt lens and the Moho.
Petrological implications
The results of this study are consistent with complex
models of magma evolution that involve polybaric
crystallization, magma mixing, and assimilation.
Polybaric crystallization is expected in plumbing systems
with two or more chambers located at different depths,
and supporting evidence is provided by detailed mineral
chemical studies of lavas that reveal the presence of different generations of minerals that formed at different pressure (e.g. Hansteen, 1991; Hansen & Gro«nvold, 2000;
Maclennan et al., 2001). High-pressure crystallization
yields residual melts that are compositionally different
from those produced by low-pressure crystallization (e.g.
Kinzler & Grove, 1992), so that polybaric crystallization
will produce liquids lying along different liquid lines of
descent (see Figs 6c and 7). It is also likely that these
magmas will mix prior to and during sequential eruptive
episodes, because magmas that pond and partially crystallize in chambers, dikes, and conduits after one eruptive episode will be flushed out and mix with fresh batches of
magma rising from deep chambers or from the mantle
during subsequent eruptive episodes. Evidence for mixing
has been described by Sigurdsson & Sparks (1981),
McGarvie (1984), Mrk (1984), and Fagents et al. (2001).
Moreover, crystal aggregates formed by partial crystallization of magma in the plumbing system in an early eruptive
phase can be disrupted and incorporated into later batches
of magma rising from deep chambers or from the mantle.
This yields magmas with complex crystal cargoes derived
from different magma batches, such as those described by
Hansen & Gro«nvold (2000). Interaction between ascending melts and the crystalline products of earlier eruptive
episodes can lead to complex variations in melt composition (Kvassnes et al., 2003; Danyushevsky et al., 2003), and
interaction between ascending melts and clinopyroxene
appears to explain the unusual chemical characteristics
(e.g. high MgO, CaO, CaO/Al2O3) of some magmas
erupted in the Hengill complex and elsewhere along the
rift zones (e.g. Trnnes, 1990; Hansen & Gro«nvold, 2000;
Gurenko & Sobolev, 2006).
Oxygen isotope studies provide strong evidence for
assimilation of hydrothermally altered crust by Icelandic
magmas (e.g. Condomines et al., 1983; Hemond et al., 1988,
1993). Assimilation may be extensive if the middle and
lower crust is hot, as suggested above, because less energy
is required for a fixed mass of magma to assimilate hot
crustal material. Moreover, frequent replenishment with
new batches of magma during mixing events leads to thermal buffering in magma chambers, which also facilitates
assimilation (Cribb & Barton, 1996). Certainly, the high
temperature of basaltic magma emplaced into Icelandic
crust [average 11868C using the method of Yang et al.
(1996) or 11778C using the method of Sugawara (2000)],
together with the relatively small temperature range of
magmas erupted at single volcanic centers (average
348C), suggests that abundant heat is available to drive
assimilation processes. High geothermal gradients coupled
with high magma temperatures will also facilitate melting
of crustal lithologies to form silicic magmas, which are
relatively abundant on Iceland (Gunnarsson et al., 1998).
Our results suggest that silicic magmas can be generated
over a relatively wide depth range in the middle and
lower crust, and this prediction can be tested by additional
studies of appropriate compositions.
CONC LUSIONS
A method to calculate pressures of crystallization of melts
lying along ol^plag^cpx cotectics based on the procedure
485
JOURNAL OF PETROLOGY
VOLUME 49
described by Yang et al. (1996) yields results that are accurate to 110 MPa (1s) and are precise to 80 MPa (1s).
Pressures calculated for Icelandic glasses from 28 volcanic
centers in the Western, Northern and Eastern rift zones as
well as from the Southern Flank Zone indicate crystallization range from 1 to 1000 MPa, equivalent to depths of
0^35 km. Magma chamber depths estimated from these
results agree well with those estimated using other methods for Askja, Bla¤fjall Table Mountain, Grimsvo«tn,
Hengill, and Hekla.
Deep chambers (417 km) occur beneath most volcanic
centers and appear to be located at the Moho, indicating
that magma ponds at the crust^mantle boundary. Shallow
chambers (571km) also occur beneath most volcanic centers. These are located in the upper crust, and probably
form at a level of neutral buoyancy. There is good evidence
for cooling and crystallizing bodies and pockets of magma
throughout the middle and lower crust, which probably
resembles a crystal mush. This strongly suggests that the
middle and lower crust is relatively hot and porous, and
that crustal accretion occurs over a range of depths as
inferred in recent models for crustal accretion at midocean ridges. The presence of multiple, stacked chambers
and hot, porous crust suggests that magma evolution is
complex and involves polybaric crystallization, magma
mixing, and assimilation.
AC K N O W L E D G E M E N T S
We thank Wendy Panero, Ralph von Frese, Hal Noltimier
and Tim Leftwich for their interest and for helpful discussions. Peter Meyer and Haraldur Sigurdsson kindly provided information about sample localities. Thoughtful
comments by David Elliot, Wendy Panero, Ingi Bjarnason,
Claude Herzberg, John Maclennan, and an anonymous
reviewer significantly improved the manuscript. We are
also grateful to Colin Devey and Marjorie Wilson for
their editorial efforts. Financial support from the Ohio
State University and from the Friends of Orton Hall is
gratefully acknowledged.
S U P P L E M E N TA RY DATA
Supplementary data for this paper are available at Journal
of Petrology online.
R EF ER ENC ES
Alfaro, R., Brandsdo¤ttir, B., Rowlands, D. P., White, R. &
Gudmundsson, M. T. (2007). Structure of the Gr|¤ msvo«tn central
volcano under the Vatnajo«kull icecap, Iceland.. Geophysical Journal
International 168, 863^876.
Allen, R. M., Nolet, G., Morgan, W. J., Vogfjord, K., Nettles, M.,
Ekstrom, G., Bergsson, B. H., Erlendsson, P., Foulger, G. R.,
Jokobsdottir, S., Julian, B. R., Pritchard, M., Ragnarsson, S. &
Stefansson, R. (2002). Plume-driven plumbing and crustal formation in Iceland. Journal of Geophysical ResearchçSolid Earth 107, No.
B8, 2163, doi:10.1029/2001JB000584.
NUMBER 3
MARCH 2008
Arnadottir, T., Delaney, P. T. & Sigmundsson, F. (1998). Sources of
crustal deformation associated with the Krafla, Iceland, eruption
of September 1984. Geophysical Research Letters 25, 1043^1046.
Baker, D. R. & Eggler, D. H. (1987). Compositions of anhydrous and
hydrous melts coexisting with plagioclase, augite, and olivine or
low-Ca pyroxene from 1atm to 8 kbar; application to the Aleutian
volcanic center of Atka. American Mineralogist 72, 12^28.
Bender, J. F., Bence, A. E. & Hodges, F. N. (1978). Petrogenesis of
basalts from the Project FAMOUS area; experimental study from
0 to 15 kbars. Earth and Planetary Science Letters 41, 277^302.
Bjarnason, I. T., Caress, D., Flovenz, O. G. & Menke, W. (1993).
Tomographic image of the Mid-Atlantic plate boundary in southwestern Iceland. Journal of Geophysical Research, B, Solid Earth and
Planets 98, 6607^6622.
Bjo«rnsson, A. (1985). Dynamics of crustal rifting in NE Iceland.
Journal of Geophysical Research 90, 10151^10162.
Bjornsson, A., Johnsen, G., Sigurdsson, S., Thorbergsson, G. &
Tryggvason, E. (1979). Rifting of the plate boundary in north
Iceland 1975^1978. Journal of Geophysical Research 84, 3029^3038.
Bown, J. W. & White, R. S. (1994). Variation with spreading rate of
oceanic crustal thickness and geochemistry. Earth and Planetary
Science Letters. Amsterdam: Elsevier 121, 435^449.
Brandsdottir, B., Einarsson, P., Menke, W., Staples, R. K. & White, R.
S. (1997). Faroe^Iceland Ridge experiment; 2, Crustal structure of
the Krafla central volcano. Journal of Geophysical Research, B, Solid
Earth and Planets 102, 7867^7886.
Breddam, K. (2002). Kistufell; primitive melt from the Iceland mantle
plume. Journal of Petrology 43, 345^373.
Breddam, K., Kurz, M. D. & Storey, M. (2000). Mapping out the conduit of the Iceland mantle plume with helium isotopes. Earth and
Planetary Science Letters. Amsterdam: Elsevier, 176, pp. 45^55.
Condomines, M., Groenvold, K., Hooker, P. J., Muehlenbachs, K.,
O’Nions, R. K., Oskarsson, N. & Oxburgh, E. R. (1983). Helium,
oxygen, strontium and neodymium isotopic relationships in
Icelandic volcanics. Earth and Planetary Science Letters 66, 125^136.
Cribb, J. W. & Barton, M. (1996). Geochemical effects of decoupled
fractional crystallization and crustal assimilation. Lithos 37,
293^307.
Danyushevsky, L. V. (2001). The effect of small amounts of H2O on
crystallisation of mid-ocean ridge and backarc basin magmas.
Journal of Volcanology and Geothermal Research 110, 265^280.
Danyushevsky, L. V., Dmitriev, L. V. & Sobolev, A. V. (1996).
Estimation of the pressure of crystallization and H2O content of
MORB and BABB glasses. Mineralogy and Petrology 57, 185^204.
Danyushevsky, L. V., McNeill, A. W. & Sobolev, A. V. (2002).
Experimental and petrological studies of melt inclusions in phenocrysts from mantle-derived magmas; an overview of techniques,
advantages and complications. Chemical Geology 183, 5^24.
Danyushevsky, L. V., Eggins, S. M., Falloon, T. J. & Perfit, M. R.
(2003). Crustal origin for coupled ‘ultra-depleted’ and ‘plagioclase’
signatures in MORB olivine-hosted melt inclusions; evidence from
the Siqueiros transform fault, East Pacific Rise. Contributions to
Mineralogy and Petrology 144, 619^637.
Darbyshire, F. A., Bjarnason, I. T., Flovenz, O. G. & White, R. S.
(1998). Crustal structure above the Iceland mantle plume imaged
by the ICEMELT refraction profile. Geophysical Journal International
135, 1131^1149.
Darbyshire, F. A., Gudmundsson, G. B., Jakobsdottir, S. S.,
Priestley, K. F., Stefansson, R. & White, R. S. (2000a). Central
structure of central and northern Iceland from analysis of teleseismic receiver functions. Geophysical Journal International 143, 163^184.
Darbyshire, F. A., Priestley, K. F. & White, R. S. (2000b). Structure of
the crust and uppermost mantle of Iceland from a combined
486
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
seismic and gravity study. Earth and Planetary Science Letters 181,
409^428.
de Zeeuw-van Dalfsen, E., Pagli, C., Pedersen, R. & Sigmundsson, F.
(2004). Satellite radar interferometry 1993^1999 suggests
deep accumulation of magma near the crust^mantle boundary
at the Krafla volcanic system, Iceland. Geophysical Research Letters
31, 5.
de Zeeuw-van Dalfsen, E., Rymer, H., Sigmundsson, F. & Sturkell, E.
(2005). Net gravity decrease at Askja Volcano, Iceland;
constraints on processes responsible for continuous caldera deflation, 1988^2003. Journal of Volcanology and Geothermal Research 139,
227^239.
de Zeeuw-van Dalfsen, E., Rymer, H., Williams-Jones, G., Sturkell, E.
& Sigmundsson, F. (2006). Integration of micro-gravity and geodetic data to constrain shallow system mass changes at Krafla
Volcano, N. Iceland. Bulletin of Volcanology 68, 420^431.
Dungan, M. A. & Rhodes, J. M. (1978). Residual glasses and
melt inclusions in basalts from DSDP Legs 45 and 46; evidence
for magma mixing. Contributions to Mineralogy and Petrology 67,
417^431.
Einarsson, P. (1978). S-wave shadows in the Krafla Caldera in NE
Iceland, evidence for a magma chamber in the crust. Bulletin
Volcanologique 41, 187^195.
Einarsson, P. & Bjo«rnsson, S. (1980). Earthquakes in Iceland. Jo« kull 29,
37^43.
Elthon, D., Meen, J. K. & Ross, D. K. (1995). Composition variations
of basaltic glasses from the Mid-Cayman Rise spreading center.
Journal of Geophysical Research, B, Solid Earth and Planets 100,
12497^12512.
Ewart, J. A., Bjornsson, A. & Voight, B. (1991). Elastic deformation
models of Krafla Volcano, Iceland, for the decade 1975 through
1985. Bulletin of Volcanology 53, 436^459.
Eysteinsson, H. & Hermance, J. F. (1985). Magnetotelluric measurements across the Eastern Neovolcanic Zone in South Iceland.
Journal of Geophysical Research, B 90, 10093^10103.
Fagents, S. A., McGarvie, D. W. & Greely, R. (2001). Dynamics of
the Ljotipollur phreatomagmatic eruption, S. central Iceland.
EOS Transactions, American Geophysical Union, 82 (Fall Meeting
Supplement, Abstract) V52A^1044.
Falloon, T. J., Danyushevsky, L. V. & Green, D. H. (2001). Peridotite
melting at 1GPa; reversal experiments on partial melt compositions
produced by peridotite^basalt sandwich experiments. Journal of
Petrology 42, 2363^2390.
Fedorova, T., Jacoby, W. R. & Wallner, H. (2005). Crust^mantle transition and Moho model for Iceland and surroundings from seismic,
topography, and gravity data. Tectonophysics 396, 119^140.
Feigl, K. L., Gasperi, J., Rigo, A. & Sigmundsson, F. (2000). Crustal
deformation near Hengill Volcano, Iceland 1993^1998; coupling
between magmatic activity and faulting inferred from elastic modeling of satellite radar interferograms. Journal of Geophysical Research,
B, Solid Earth and Planets 105, 25655^25670.
Fisk, M. R., Bence, A. E. & Schilling, J. G. (1982). Major element
chemistry of Galapagos rift zone magmas and their phenocrysts.
Earth and Planetary Science Letters 61, 171^189.
Ford, C. E., Craven, J. A., Fisk, M. R. & Russell, D. G. (1983).
Olivine^liquid equilibria; temperature, pressure and composition
dependence of the crystal/liquid cation partition coefficients for
Mg, Fe2þ, Ca and Mn. Journal of Petrology 24, 256^265.
Foulger, G. R. & Anderson, D. L. (2005). A cool model for the Iceland
hotspot. Journal of Volcanology and Geothermal Research 141, 1^22.
Foulger, G. R., Du, Z. & Julian, B. R. (2003). Icelandic-type crust.
Geophysical Journal International 155, 567^590.
Frost, B. R. & Lindsley, D. H. (1992). Equilibria amount Fe^Ti oxides,
pyroxenes, olivine, and quartz: Part II. Application. American
Mineralogist 77, 1004^1020.
Furman, T., Frey, F. A. & Park, K. (1991). Chemical constraints on the
petrogenesis of mildly alkaline lavas fromVestmannaeyjar, Iceland;
the Eldfell (1973) and Surtsey (1963^1967) eruptions. Contributions to
Mineralogy and Petrology 109, 19^37.
Gebrande, H., Einarsson, P. & Miller, H. (1980). Seismic structure
of Iceland along RRISP-Profile I. Zeitschrift fu«r Geophysik 47,
239^249.
Grove, T. L. & Bryan, W. B. (1983). Fractionation of pyroxene-phyric
MORB at low pressure; an experimental study. Contributions to
Mineralogy and Petrology 84, 293^309.
Grove, T. L., Kinzler, R. J. & Bryan, W. B. (1993). Fractionation of
Mid-Ocean Ridge Basalt (MORB). In: Phipps-Morgan, J. et al.
(eds) Mantle Flow and Melt Migration at Mid-Ocean Ridges, American
Geophysical Monograph 71, 281^311.
Grove, T. L., Gerlach, D. C., Sando, T. W. & Baker, M. B. (1993).
Origin of calc-alkaline series lavas at medicine lake volcano by
fractionation, assimilation and mixingçcorrections and clarifications (vol 82, pg 407, 1983). Contributions to Mineralogy and Petrology
114, 422^424.
Gudmundsson, A. (1988). Effect of tensile stress concentration
around magma chambers on intrusion and extrusion frequencies.
Journal of Volcanology and Geothermal Research 35, 179^194.
Gudmundsson, A. (2000). Dynamics of volcanic systems in Iceland:
Example of tectonism and volcanism at juxtaposed hot spot
and mid-ocean ridge systems. Annual Review of Earth and Planetary
Sciences 28, 107^140.
Gudmundsson, O. (2003). The dense root of the Iceland crust. Earth
and Planetary Science Letters 206, 427^440.
Gudmundsson, M. T. & Milsom, J. (1997). Gravity and magnetic
studies of the subglacial Grimsvo«tn Volcano, Iceland; implications
for crustal and thermal structure. Journal of Geophysical Research, B,
Solid Earth and Planets 102, 7691^7704.
Gudmundsson, O., Brandsdottir, B., Menke, W. & Sigvaldason, G. E.
(1994). The crustal magma chamber of the Katla volcano in South
Iceland revealed by 2-D seismic undershooting. Geophysical Journal
International 119, 277^296.
Gunnarsson, B., Marsh, B. D. & Taylor, H. P., Jr (1998). Generation of
Icelandic rhyolites; silicic lavas from the Torfajo«kull central volcano.
Journal of Volcanology and Geothermal Research 83, 1^45.
Gurenko, A. A. & Sobolev, A. V. (2006). Crust^primitive magma
interaction beneath neovolcanic rift zone of Iceland recorded in
gabbro xenoliths from Midfell, SW Iceland. Contributions to
Mineralogy and Petrology 151, 495^520.
Hansen, H. & Gro«nvold, K. (2000). Plagioclase ultraphyric basalts in
Iceland; the mush of the rift. Journal of Volcanology and Geothermal
Research 98, 1^32.
Hansteen, T. H. (1991). Multi-stage evolution of the picritic Maelifell
rocks, SW Iceland; constraints from mineralogy and inclusions of
glass and fluid in olivine. Contributions to Mineralogy and Petrology
109, 225^239.
Hardarson, B. S., Ellam, R. M., Fitton, J. G. & Pringle, M. S. (1997).
Rift relocation; a geochemical and geochronological investigation
of a paleo-rift in Northwest Iceland. Earth and Planetary Science
Letters 153, 181^196.
Hards, V. L., Greenwood, P. B., Kempton, P. D. & Thompson, R. N.
(2000). The magmatic evolution of the Snaefell volcanic centre; an
example of volcanism during incipient rifting in Iceland. Journal of
Volcanology and Geothermal Research 99, 97^121.
Hemond, C., Alle'gre, C. J., Condomines, M., Fourcade, S., Javoy, M.
& Oskarsson, N. (1988). Thorium, strontium and oxygen isotopic
487
JOURNAL OF PETROLOGY
VOLUME 49
geochemistry in Recent tholeiites from Iceland; crustal influence
on mantle-derived magmas. Earth and Planetary Science Letters 87,
273^285.
Hemond, C., Arndt, N. T., Hofmann, A. W., Lichtenstein, U.,
Oskarsson, N. & Steinthorsson, S. (1993). The heterogeneous
Iceland plume; Nd^Sr^O isotopes and trace element constraints.
Journal of Geophysical Research, B, Solid Earth and Planets 98,
15833^15850.
Herzberg, C. (2004). Partial crystallization of mid-ocean ridge basalts
in the crust and mantle. Journal of Petrology 45, 2389^2405.
Hoskuldsson, A., Sparks, R. S. J. & Carroll, M. R. (2006). Constraints
on the dynamics of subglacial basalt eruptions from geological
and geochemical observations at Kverkfjoll, NE Iceland. Bulletin of
Volcanology 68, 689^701.
Jamtveit, B., Brooker, R., Brooks, K., Larson, L. M. & Pederson, T.
(2001). The water content of olivines from the North Atlantic
Volcanic Province. Earth and Planetary Science Letters 186, 401^415.
Jo¤nasson, K., Holm, P. M. & Pedersen, A. K. (1992). Petrogenesis of
silicic rocks from the Kro¤ksfjo«rdur central volcano, NW Iceland.
Journal of Petrology 33, 1345^1369.
Jo¤nsson, S., Bjornsson, G., Pedersen, R. & Segall, P. (2003).
Post-earthquake ground movements correlated to pore-pressure
transients. Nature 424, 179^183.
Juster, T. C., Grove, T. L. & Perfit, M. R. (1989). Experimental constraints on the generation of FeTi basalts, andesites, and rhyodacites
at the Galapagos spreading center, 858W and 958W. Journal of
Geophysical Research, B, Solid Earth and Planets 94, 9251^9274.
Kaban, M. K., Flovenz, O. G. & Palmason, G. (2002). Nature of the
crust^mantle transition zone and the thermal state of the upper
mantle beneath Iceland from gravity modelling. Geophysical Journal
International 149, 281^299.
Kaban, M. K., Artemieva, I. M., Mooney, W. D. & Schwintzer, P.
(2003). Density of the continental roots; compositional and thermal
contributions. Earth and Planetary Science Letters 209, 53^69.
Kelley, D. F. & Barton, M. (2005).Thickness, composition, and physical properties of crust in Iceland’s neovolcanic zone. EOS
Transactions, American Geophysical Union 86(52), Fall Meeting
Supplement, V13E-0591.
Kelley, D. F., Kapostasy, D. D. & Barton, M. (2004). Depths of
magma chambers in the Icelandic crust. EOS Transactions,
American Geophysical Union 85(17), Joint Assembly Supplement,
V54A-02.
Kelley, D. F., Leftwich, T. E. & Barton, M. (2005). Petrologic constraints on Iceland’s lower crust. EOS Transactions, American
Geophysical Union 86(18), Joint Assembly Supplement, V13A-02.
Kinzler, R. J. & Grove, T. L. (1992). Primary magmas of
mid-ocean ridge basalts; 1, Experiments and methods.
Journal of Geophysical Research, B, Solid Earth and Planets 97,
6885^6906.
Kjartansson, E. & Gro«nvold, K. (1983). Location of a magma
reservoir beneath Hekla Volcano, Iceland. Nature 301, 139^141.
Kushiro, I. (1969). The system forsterite^diopside^silica with and
without water at high pressures. American Journal of Science 267-A,
269^294.
Kvassnes, H. J., Dick, J. B. & Grove, T. L. (2003). The Atlantis Bank
gabbro-suite was not a ‘normal’ magma-chamber that produced
basalts. Geophysical Research Absracts 5, 12203.
LaFemina, P. C., Arnadottir, T., Dixon, T. H., Einarsson, P.,
Malservisi, R., Sigmundsson, F. & Sturkell, E. (2005). Geodetic
GPS measurements in south Iceland; strain accumulation and
partitioning in a propagating ridge system. Journal of Geophysical
Research 110, 21.
NUMBER 3
MARCH 2008
Langmuir, C. H. (1989). Geochemical consequences of in situ crystallization. Nature 340, 199^205.
Langmuir, C. H., Klein, E. M. & Plank, T. (1992). Petrological systematics of Mid-Ocean Ridge Basalts: Constraints on melt generation beneath ocean ridges. In: Phipps-Morgan, J. et al. (eds) Mantle
Flow and Melt Migration at Mid-Ocean Ridges, American Geophysical
Monograph 71, 183^280.
Le Bas, M. J., Le Maitre, R. W., Streckeisen, A. & Zanettin, B. (1986).
A chemical classification of volcanic rocks based on the total alkalisilica diagram. Journal of Petrology 27, 745^750.
Le Roux, P. J., Le Roex, A. P. & Schilling, J. G. (2002). Crystallization
processes beneath the southern Mid-Atalantic Ridge (40^508S),
evidence for high-pressure initiation of crystallization. Contributions
to Mineralogy and Petrology 142, 582^602.
Leftwich, T. E., von Frese, R. R. B., Potts, L. V., Kim, G. R.,
Roman, D. R., Taylor, P. T. & Barton, M. (2005). Crustal modeling
of the North Atlantic from spectrally correlated free-air gravity
and terrain gravity. Journal of Geodynamics 40, 23^50.
Linde, A. T., Agustsson, K., Sacks, I. S. & Stefansson, R. (1993).
Mechanism of the 1991 eruption of Hekla from continuous borehole
strain monitoring. Nature 365, 737^740.
Macdonald, G. A. (1968). Composition and origin of Hawaiian lavas.
In: Coats, R. R., Hay, R. L. & Anderson, C. A. (eds) Studies in
Volcanology: A Memoir in Honor of Howel Williams, Geological Society of
America. Memoir 116, 477^522.
Maclennan, J., Gronvold, K. & McKenzie, D. (2001). Plume-driven
upwelling under central Iceland. Earth and Planetary Science Letters
194, 67^82.
Maclennan, J., Eiler, J. M., Gronvold, K., Kitchen, N., McKenzie, D.
& Shimizu, N. (2003). Melt mixing and crystallization under
Theistareykir, northeast Iceland. Geochemistry, Geophysics, Geosystems
4, doi:10.1029/2003GC000558.
Maclennan, J., Hulme, T. & Singh, S. C. (2004). Thermal models
of oceanic crustal accretion: Linking geophysical, geological and
petrological observations. Geochemistry, Geophysics, Geosystems 5, doi:
10.1029/2003GC000605.
Marti, J. & Folch, A. (2005). Anticipating volcanic eruptions. In:
Ernst, G. G. J. (ed.) Volcanoes and the Environment. Cambridge:
Cambridge University Press, pp. 90^120.
McCann, V. & Barton, M. (2004).Oxygen fugacities of lavas from
Iceland and implications for mantle redox states. EOS Transactions,
American Geophysical Union 85(17), Joint Assembly Supplement,
U53A-06.
McGarvie, D. W. (1984). Torfajoekull; a volcano dominated by magma
mixing. Geology 12, 685^688.
Menke, W., Brandsdottir, B., Sparks, D. & West, M. (1998). Compressional and shear velocity structure of the lithosphere in
Northern Ireland. Bulletin of the Seismological Society of America 88,
1561^1571.
Menke, W. & Sparks, D. (1995). Crustal accretion model
for Iceland predicts ‘cold’ crust. Geophysical Research Letters 22,
1673^1676.
Meyer, P. S., Schilling, J. & Sigurdsson, H. (1985). Petrological
and geochemical variations along Iceland’s neovolcanic zones.
Journal of Geophysical Research, B 90, 10043^10072.
Michael, P. J. & Cornell, W. C. (1998). Influence of spreading rate and
magma supply on crystallization and assimilation beneath midocean ridges; evidence from chlorine and major element chemistry
of mid-ocean ridge basalts. Journal of Geophysical Research, B, Solid
Earth and Planets 103, 18325^18356.
Moore, J. G. & Calk, L. C. (1991). Degassing and differentiation in
subglacial volcanoes, Iceland. Journal of Volcanology and Geothermal
Research 46, 157^180.
488
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Mrk, M. B. E. (1984). Magma mixing in the post-glacial Veidivoetn
fissure eruption, Southeast Iceland; a microprobe study of mineral
and glass variations. Lithos 17, 55^75.
Nicholls, I. A. & Ringwood, A. E. (1973). Effect of water on olivine
stability in tholeiites and the production of silica saturated magmas
in the island arc environment. Journal of Geology 81, 285^300.
Nichols, A. R. L., Carroll, M. R. & Hoskuldsson, A. (2002). Is the
Iceland hot spot also wet? Evidence from the water contents of
undegassed submarine and subglacial pillow basalts. Earth and
Planetary Science Letters 202, 77^87.
O’Hara, M. J. (1968). Are ocean floor basalts primary magma? Nature
220, 683^686.
Oskarsson, N., Sigvaldason, G. E. & Steinthorsson, S. (1985).
Iceland geochemical anomaly; origin, volcanotectonics, chemical
fractionation, and isotope evolution of the crust. Journal of
Geophysical Research, B 90, 10011^10025.
Pagli, C., Arnadottir, T., Einarsson, P., Sigmundsson, F. & Sturkell, E.
(2006). Deflation of the Askja volcanic system; constraints on the
deformation source from combined inversion of satellite radar
interferograms and GPS measurements. Journal of Volcanology and
Geothermal Research 152, 97^108.
Pan, Y. & Batiza, R. (2002). Mid-ocean ridge magma chamber processes; constraints from olivine zonation in lavas from the East
Pacific Rise at 9830’N and 10830’N. Journal of Geophysical Research,
B, Solid Earth and Planets 107, 13.
Pan, Y. & Batiza, R. (2003). Magmatic processes under mid-ocean
ridges; a detailed mineralogic study of lavas from East Pacific
Rise 9830’N, 10830’N, and 11820’N. Geochemistry, Geophysics,
Geosystems 4, doi:10.1029/2002GC000309.
Pedersen, R. & Sigmundsson, F. (2004). InSAR based sill model links
spatially offset area of deformation and seismicity for the 1994
unrest episode at Ejyafjallajo«kull Volcano, Iceland. Geophysical
Research Letters 31, 5.
Pedersen, R. & Sigmundsson, F. (2006). Temporal developement
of the 1999 intrusive episode in the Eyjafjallajo«kull volcano,
Iceland, derived from InSAR images. Bulletin of Volcanology 68,
377^393.
Re¤villon, S., Arndt, N. T., Hallot, E., Kerr, A. C. & Tarney, J. (1999).
Petrogenesis of picrites from the Caribbean Plateau and the North
Atlantic magmatic province. Lithos 49, 1^21.
Rymer, H. & Tryggvason, E. (1993). Gravity and elevation changes at
Askja, Iceland. Bulletin of Volcanology 55, 361^371.
Rymer, H., Cassidy, J., Locke, C. A. & Sigmundsson, F. (1998). Posteruptive gravity changes from 1990 to 1996 at Krafla Volcano,
Iceland. Journal of Volcanology and Geothermal Research 87, 141^149.
Sack, R. O., Carmichael, I. S. E. & Walker, D. (1987). Experimental
petrology of alkalic lavas; constraints on cotectics of multiple
saturation in natural basic liquids. Contributions to Mineralogy and
Petrology 96, 1^23.
Schiellerup, H. (1995). Generation and equilibration of olivine tholeiites in the northern rift zone of Iceland; a petrogenetic study of
the Bla¤fjall table mountain. Journal of Volcanology and Geothermal
Research 65, 161^179.
Schilling, J. G. (1973). Iceland mantle plume. Nature 246, 141^143.
Shi, P. (1993). Low-pressure phase relationships in the system
Na2O^CaO^FeO^MgO^Al2O3^SiO2 at 11008C, with implications
for the differentiation of basaltic magmas. Journal of Petrology 34,
743^762.
Sigmundsson, F., Bilham, R. & Einarsson, P. (1992). Magma chamber
deflation recorded by the Global Positioning System; the Hekla
1991 eruption. Geophysical Research Letters 19, 1483^1486.
Sigmundsson, F., Einarsson, P., Foulger, G. R., Hodgkinson, K. M.,
Rognvaldsson, S. T. & Thorbergsson, G. (1997). The 1994^1995
seismicity and deformation at the Hengill triple junction, Iceland;
triggering of earthquakes by minor magma injection in a zone of
horizontal shear stress. Journal of Geophysical Research, B, Solid Earth
and Planets 102, 15151^15161.
Sigurdsson, H. & Sparks, R. S. J. (1978). Rifting episode in North
Iceland in 1874^1875 and the eruptions of Askja and Sveinagja.
Bulletin Volcanologique 41, 149^167.
Sigurdsson, H. & Sparks, R. S. J. (1981). Petrology of rhyolitic and
mixed magma ejecta from the 1875 eruption of Askja, Iceland.
Journal of Petrology 22, 41^84.
Singh, S. C., Harding, A. J., Kent, G. M., Sinha, M. C., Combier, V.,
Bazin, S., Tong, C. H., Pye, J. W., Barton, P. J., Hobbs, R. W.,
White, R. S. & Orcutt, J. A. (2006). Seismic reflection images of
the Moho underlying melt sills at the East Pacific Rise. Nature
442, 287^290.
Sinton, J. M. & Detrick, R. S. (1992). Mid-ocean ridge magma chambers. Journal of Geophysical Research, B, Solid Earth and Planets 97,
197^216.
Sinton, J., Gronvold, K. & Saemundsson, K. (2005). Postglacial eruptive history of the Western volcanic zone, Iceland. Geochemistry,
Geophysics, Geosystems 6, doi:10.1029/2005GC001021.
Sisson, T. W. & Grove, T. L. (1993). Temperatures and H2O contents of
low-MgO high-alumina basalts. Contributions to Mineralogy and
Petrology 113, 167^184.
Slater, L., Gronvold, K., McKenzie, D. & Shimizu, N. (2001). Melt
generation and movement beneath Theistareykir, NE Iceland.
Journal of Petrology 42, 321^354.
Soosalu, H. & Einarsson, P. (1997). Seismicity around the Hekla and
Torfajo«kull volcanoes, Iceland, during a volcanically quiet period,
1991^1995. Bulletin of Volcanology 59, 36^48.
Soosalu, H. & Einarsson, P. (2004). Seismic constraints on magma
chambers at Hekla and Torfajo«kull volcanoes, Iceland. Bulletin of
Volcanology 66, 276^286.
Soosalu, H., Jonsdottir, K. & Einarsson, P. (2006a). Seismicity crisis
at the Katla volcano, Icelandçsigns of a cryptodome? Journal of
Volcanology and Geothermal Research 153, 177^186.
Soosalu, H., Lippitsch, R. & Einarsson, P. (2006b). Low-frequency
earthquakes at the Torfajo«kull volcano, south Iceland. Journal of
Volcanology and Geothermal Research 153, 187^199.
Spulber, S. D. & Rutherford, M. J. (1983). The origin of rhyolite and
plagiogranite in oceanic crust; an experimental study. Journal of
Petrology 24, 1^25.
Staples, R. K., Brandsdottir, B., Maguire, P. K. H., McBride, J. H.,
Menke, W. & White, R. S. (1997). Faroe^Iceland Ridge experiment;
1, Crustal structure of northeastern Iceland. Journal of Geophysical
Research, B, Solid Earth and Planets 102, 7849^7866.
Steinthorsson, S., Oskarsson, N. & Sigvaldason, G. E. (1985). Origin of
alkali basalts in Iceland; a plate tectonic model. Journal of
Geophysical Research, B 90, 10027^10042.
Sturkell, E. & Sigmundsson, F. (2000). Continuous deflation of the
Askja Caldera, Iceland, during the 1983^1998 noneruptive period.
Journal of Geophysical Research, B, Solid Earth and Planets 105,
25671^25684.
Sturkell, E., Einarsson, P. & Sigmundsson, F. (2003). Recent
unrest and magma movements at Eyjafjallajo«kull and Katla
Volcanoes, Iceland. Journal of Geophysical Research, B, Solid Earth
and Planets 108, 13.
Sturkell, E., Einarsson, P., Freysteinn, S., Holldor, G., Olafsson, O.,
Pedersen, R., de Zeeuw-van Dalfsen, E., Linde, A. T., Sacks, I. S.
& Stefansson, R. (2006). Volcano geodesy and magma dynamics in
Iceland. Journal of Volcanology and Geothermal Research 150, 14^34.
Sugawara, T. (2000). Empirical relationships between temperature,
pressure, and MgO content in olivine and pyroxene saturated
489
JOURNAL OF PETROLOGY
VOLUME 49
liquid. Journal of Geophysical Research, B, Solid Earth and Planets 105,
8457^8472.
Thy, P. (1991). High and low pressure phase equilibria of a mildly
alkalic lava from the 1965 Surtsey eruption; experimental results.
Lithos 26, 223^243.
Thy, P. & Lofgren, G. E. (1994). Experimental constraints on the
low-pressure evolution of transitional and mildly alkalic basalts;
the effect of Fe^Ti oxide minerals and the origin of basaltic andesites. Contributions to Mineralogy and Petrology 116, 340^351.
Tormey, D. R., Bryan, W. B. & Grove, T. L. (1987). Experimental petrology of normal MORB near the Kane fracture zone; 228^258N,
Mid-Atlantic Ridge. Contributions to Mineralogy and Petrology 96,
121^139.
Trnnes, R. (1990). Basaltic melt evolution of the Hengill Volcanic
System, SW Iceland, and evidence for clinopyroxene assimilation
in primitive tholeiitic magmas. Journal of Geophysical Research 95,
15893^15910.
Tryggvason, E. (1980). Observed ground deformation during the Krafla eruption of March 16, 1980. Nordic Volcanological Institute, University of
Iceland, Series: Nordic Volcanological Institute 0258^7718, 80^05.
Tryggvason, E. (1986). Multiple magma reservoirs in a rift zone
volcano; ground deformation and magma transport during the
September 1984 eruption of Krafla, Iceland. Journal of Volcanology
and Geothermal Research 28, 1^44.
Tryggvason, E. (1989). Ground deformation in Askja, Iceland; its
source and possible relation to flow of the mantle plume. Journal of
Volcanology and Geothermal Research 39, 61^71.
Tryggvason, E. (1994a). Surface deformation at the Krafla Volcano,
North Iceland, 1982^1992. Bulletin of Volcanology 56, 98^107.
Tryggvason, E. (1994b). Observed ground deformation at Hekla,
Iceland prior to and during the eruptions of 1970, 1980^1981 and
1991. Journal of Volcanology and Geothermal Research 61, 281^291.
Walker, D., DeLong, S. E. & Shibata, T. (1979). Abyssal tholeiites from
the Oceanographer fracture zone; II, Phase equilibria and mixing.
Contributions to Mineralogy and Petrology 70, 111^125.
Weaver, J. S. & Langmuir, C. H. (1990). Calculation of phase equilibrium in mineral^melt systems. Computers & Geosciences 16, 1^19.
Weir, N. R. W., Brandsdottir, B., Darbyshire, F., Einarsson, P.,
Gunnarsson, G., Hjorleifsdottir, V., Jonsdottir, K., Mochizuki, M.,
Nakanishi, A., Shimamura, H., Shiobara, H., Smallwood, J.,
Staples, R. & White, R. S. (2001). Crustal structure of the northern
Reykjanes Ridge and Reykjanes Peninsula, Southwest Iceland.
Journal of Geophysical Research, B, Solid Earth and Planets 106,
6347^6368.
White, R. S. & McKenzie, D. (1995). Mantle plumes and flood basalts.
Journal of Geophysical Research, B, Solid Earth and Planets 100,
17543^17585.
White, R. S., McKenzie, D. & O’Nions, R. K. (1992). Oceanic crustal
thickness from seismic measurements and rare earth element inversions. Journal of Geophysical Research, B, Solid Earth and Planets 97,
19683^19715.
Yang, H. J., Grove, T. L. & Kinzler, R. J. (1996). Experiments and
models of anhydrous, basaltic olivine^plagioclase^augite saturated
melts from 0001 to 10 kbar. Contributions to Mineralogy and Petrology
124, 1^18.
A P P E N D I X A : C A L C U L AT I O N
OF PR ESSU R E
Yang et al. (1996, table 3) presented three equations that
allow direct calculation of pressure. These equations are
NUMBER 3
MARCH 2008
used to calculate a series of liquid compositions (LP)
lying along the ol^plag^cpx cotectic for the sample of interest at increments of 100 MPa (see Table A1). The liquid
compositions are converted to normative mineral components using the procedure described by Grove et al. (1993)
assuming that Fe ¼ FeO, and projected from plag onto the
plane ol^cpx^qtz and from ol onto the plane plag^cpx^qtz.
The pressure dependence of each normative mineral
component in the predicted liquids (LP) is found by
regression, and the pressure of crystallization is found
from the regression equations using the projected normative mineral components for the original sample (LS).
Thus for the projection from plag, values of P are calculated from predicted and observed ol, cpx and qtz, whereas
for the projection from ol, values of P are calculated from
predicted and observed plag, cpx and qtz. We have used
these two projections because most basalt melts are saturated with plag and ol, and obtain six values of P for each
sample. The average value is taken as the pressure of crystallization, and all values are used to calculate the uncertainty (1s) associated with the calculated pressure.
An Excel spreadsheet to perform these calculations is
available as Supplementary Data at http://www.petrol
ogy.oxfordjournals.org/.
The approach described above uses all three of the
equations given by Yang et al. (1996) to calculate pressure.
In contrast, Michael & Cornell (1998) used one of the
equations of Yang et al. [1996, equation (2)] to calculate
the pressures of crystallization of MORB. Results
obtained using the two methods are compared in
Appendix B.
A P P E N D I X B : AC C U R AC Y
A N D PRECISION
In principle, accuracy can be determined by comparing
pressures calculated for glasses produced in experiments
with the reported run pressure. A problem with this
approach is that experimental glass compositions show considerable variability, much greater than that expected from
analytical uncertainties, which reflects problems inherent in
experimental studies (e.g. Ford et al., 1983; Yang et al., 1996;
Falloon et al., 2001). These include uncertainties in measurement of pressure and temperature, unknown amounts of
volatiles in nominally anhydrous experiments, loss of Fe2þ
and alkalis from the charge during the experiment, modification of melt compositions by quench crystallization, and
run durations too short for the attainment of equilibrium. It
appears that some experimental glasses reported to be in
equilibrium with ol, plag, and cpx do not have compositions
that represent liquids lying on the ol^plag^cpx cotectic.
Accordingly, large differences are expected between calculated and experimental pressures, leading to poor estimates
of accuracy.
490
KELLEY AND BARTON
CRYSTALLIZATION PRESSURE OF ICELANDIC MAGMAS
Table A1: Example of method used to calculate pressure
Sample 1.7.41
Content (wt %)
SiO2
Al2O3
TiO2
Fe2O3
FeO
MnO
MgO
CaO
Na2O
K2O
P2O5
Total
476
159
123
0
1008
017
918
129
18
01
011
99070
P (kbar)
Plag projection
Ol
Ol projection
Cpx
Qtz
Plag
Cpx
Qtz
Sample2
4278
5189
533
6866
2843
292
Predicted3
2893
7145
038
6107
3914
021
1
3080
6883
037
6226
3754
020
2
3271
6615
114
6347
3591
062
3
3467
6341
192
6469
3427
104
4
3667
6061
272
6593
3261
146
5
3871
5775
354
6718
3092
190
6
4080
5482
438
6846
2921
233
7
4294
5182
524
6974
2748
278
8
4513
4875
612
7105
2572
323
9
4738
4561
701
7237
2394
368
10
4967
4239
793
7371
2214
414
11
5203
3910
888
7507
2031
461
12
5444
3572
984
7645
1846
509
047
034
117
078
058
227
1336
2425
070
4749
2286
062
0001
Slope4
Intercept4
R2
09985
09985
09985
09995
09995
09995
P calc.5
677
682
695
606
636
723
Results: average P ¼ 670 kbar (1s ¼ 042 kbar, range ¼ 117 kbar); the average, 1s, and range were calculated from six
pressures obtained by regression. iPlag–Ol ¼ 030 (where iPlag–Ol is the difference between average P calculated for
Plag projection and average P calculated for Ol projection).
1
Glass analysis from Schiellerup (1995).
2
Projected mineral components for sample 1.7.4 calculated using procedure of Tormey et al. (1987).
3
Projected mineral components calculated at the pressures listed in Column 1 using procedure of Tormey et al. (1987).
4
Slope and intercept from regression of projected mineral components (Columns 3–8) vs P (Column 1).
5
Pressures calculated from regressions and projected mineral components for sample 1.7.4.
The experimental glass compositions used to construct
Fig. 2 are taken from the sources listed in the figure caption. Glass compositions from experiments on natural samples were combined with glass analyses from experiments
by Shi (1993) in the system CaO^MgO^Al2O3^SiO2^
FeO^Na2O and used to map the likely positions of the
cotectic at different pressures (samples with anomalous
compositions, that plot off the main data array at a given
pressure, were discarded). The actual positions of the
cotectic were located using a subset of these glasses that
have compositions that form tightly defined arrays in projections such as those shown in Fig. 2, and therefore appear
to anchor the positions of the cotectic at different pressures.
The glass compositions in both datasets (excluding alkaline compositions and glasses from Shi’s experiments)
were used to test the accuracy of pressures calculated for
sub-alkaline and transitional compositions. For the
extended dataset (91 glasses), calculated pressures agree
with experimental pressures to 120 MPa (1s), which is
about the same as uncertainties for pressures calculated
by similar methods (Herzberg, 2004). For the smaller dataset (59 glasses), calculated pressures agree with experimental pressures to 90 MPa (1s). These results suggest that
reported pressures are accurate to 110 MPa.
Pressures were also calculated using the method
described by Michael & Cornell (1998). For the extended
dataset (91 glasses), pressures calculated using equation (2)
of Yang et al. (1996) agree with experimental pressures
to 160 MPa (1s), whereas for the smaller dataset
(59 glasses), calculated pressures agree with experimental
491
JOURNAL OF PETROLOGY
VOLUME 49
pressures to 120 MPa (1s). These results indicate that
pressures calculated using the method used here are more
accurate than those estimated using the method described
by Michael & Cornell (1988).
Yang et al. (1996) estimated the uncertainty in pressure
estimates (precision) from the standard deviation of the
NUMBER 3
MARCH 2008
mean of replicate electron microprobe analyses, and
obtained a value of about 50 MPa (2s). The uncertainties
in calculated pressures for the experimental datasets determined as described in Appendix A are somewhat larger
than this value, 60^80 MPa (1s).
492
Related documents