You are on page 1of 33

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

PAGES 1^33

2013

doi:10.1093/petrology/egt040

Journal of Petrology Advance Access published July 17, 2013

Magma Dynamics and Petrological Evolution


Leading to the VEI 5 2000 BP Eruption of El
Misti Volcano, Southern Peru

COLLEGE OF EARTH, OCEAN, AND ATMOSPHERIC SCIENCES, OREGON STATE UNIVERSITY, CORVALLIS, OR,

97331-5506, USA
2

DEPARTMENTO DE GEOLOGIA, UNIVERSIDAD NACIONAL DE SAN AGUSTIN, AREQUIPA, PERU

RECEIVED JULY 7, 2011; ACCEPTED JUNE 17, 2013

surface in 5 days at ascent rates of at least 0023 m s1. Further


decompression-driven crystallization is recorded in plagioclase rims
and microlite growth that may have contributed to a rapid increase
in viscosity leading to explosive eruption. This VEI 5 plinian eruption shares characteristics with other explosive events at El Misti
on a time scale of 2000^4000 years, suggesting periodic rechargedriven explosive activity.

Magma dynamics and time scales during the VEI 5, 2000 BP eruption of El Misti volcano, southern Peru (EM2000BP) are investigated to address cyclic explosive activity at this hazardous volcano.
The 14 km3 of pumice falls and flows have abundant mingled
pumice of high-K, calc-alkaline rhyolite and andesite composition.
Phenocryst zoning and compositions reveal mutual exchange of
plagioclase between the two magmas; amphibole in the rhyolite was
derived from the andesite. Amphiboles in the andesite are predominantly unrimmed crystals whereas those in the rhyolite mostly exhibit
reaction rims. Phase equilibria indicate that the andesite formed at
900^9508C and 2^3 kbar pressure and was water-saturated with
51^60 wt % H2O, broadly similar to El Misti magmas overall.
Amphibole, plagioclase, Ti-magnetite, and two pyroxenes were the
crystallizing phases. A separate rhyolite magma existed higher in
the crust at a temperature of 816  308C and 5% H2O in which
only plagioclase and Fe^Ti oxides were stable. The lack of cognate
amphibole in the rhyolite despite H2O saturation requires that it
staged above the stability limit of amphibole (5100 MPa).
Exchange reactions in amphibole (dominantly pargasitic) and
trace element partitioning in plagioclase indicate that both andesite
and rhyolite magmas were broadly constant in temperature and
H2O content. These constraints suggest that the initially separate
rhyolite and deeper andesite magmas interacted by an initial andesite
recharge event that resulted in mingling and crystal exchange. A
period of 50^60 days is required for amphibole introduced into the
rhyolite to develop reaction rims owing to decompression.These rims
are dominated by plagioclase, a consequence of the Al-rich nature of
the amphibole.The lack of reaction rims on amphibole in the andesite
implicates a second, more-forceful and voluminous eruption-triggering recharge event during which andesite rose rapidly from source to

Major composite cones, among the most hazardous volcanoes on the planet, are the integrated product of a prolonged history of effusive cone building activity
punctuated by explosive eruptions and edifice collapses
(Davidson & de Silva, 2000). Although the eruptive style
and attendant hazard is dominated by effusion, the rare
explosive eruptions are often the most voluminous and
hazardous. Understanding the controls on this transition
in activity is central to our efforts to fully address magmatic and volcanic evolution and hazard mitigation. Two
important clues to this effort are that explosive activity is
cyclic or quasi-cyclic (Matthews et al., 1997; Davidson &
de Silva, 2000; Ruprecht & Worner, 2007) and involves
recharge, suggesting that the rhythm of open magmatic
systems is a dominant driver.

*Corresponding author. Telephone: 541 737 8199; Fax: 541 737 2064;
E-mail: ftepley@coas.oregonstate.edu.

 The Author 2013. Published by Oxford University Press. All


rights reserved. For Permissions, please e-mail: journals.permissions@
oup.com

El Misti; explosive eruption; amphibole reaction rims;


trace element partitioning in plagioclase; magmatic time scales; recharge

KEY WORDS:

I N T RO D U C T I O N

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

FRANK J. TEPLEY, III1*, SHANAKA DE SILVA1 AND GUIDO SALAS2

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

MONTH 2013

GEOLOGIC A L S ET T I NG
El Misti (162948S, 714098W; 5822 m above sea level) is a
major volcanic edifice of the Central Volcanic Zone of the
Andes (Bullard,1962; de Silva & Francis,1991a) in southern
Peru lying less than 15 km from the city of Arequipa
(Fig. 1). It is located within the Andean arc, and its history
is one of constructive dome growth, lava flows and explosive volcanism, endangering the growing population
center of Arequipa nearby (de Silva & Francis, 1991a;
Thouret et al., 2001; Harpel et al., 2011).
The geological history of El Misti is one typical of
Andean arc volcanoes. Based on extensive field mapping,
40
Ar/39Ar and 14C dating of rocks and organic material,
Thouret et al. (2001, and references therein), Paquereau
Lebti et al. (2006) and Ruprecht & Worner (2007) have
pieced together a comprehensive volcanic history for El
Misti. The earliest remnant (c. 112 ka) of El Misti is an
eroded stratovolcano (Misti 1) that unconformably overlies
lavas and volcaniclastic deposits of Chachani Volcano
(Paquereau Lebti et al., 2006; Ruprecht & Worner, 2007).
Upon this edifice lie successive edifices, termed Misti 2,
Misti 3, and Misti 4, and lava flows and pyroclastic debris
erupted since 112 ka. Historically, the volcano-building
events of El Misti are associated with alternating growth
and destruction of andesitic and dacitic domes and lava
flows with dome collapses and associated pyroclastic
flows, intermixed with explosive episodes, and avalanche
deposits (Thouret et al., 2001, and references therein;
Ruprecht & Worner, 2007). Thouret et al. (2001) suggested
that, on average, ash falls occur every 500^1500 years,
with pumice fallout-producing eruptions every 2000^4000
years. The 2000 BP eruption is a plinian eruption producing
pumice falls and flows with varying proportions of
banded pumice of rhyolite and andesite compositions
amounting to 14 km3 of material (05 km3 dense rock
equivalent; Thouret et al., 2001; Harpel et al., 2011). Extensive lahars were generated by interaction of pyroclastic
flows with snow on the volcano, attesting to the potential
hazard of explosive eruptions at El Misti (Harpel et al.,
2011).
Over the course of its history, El Misti has produced
relatively homogeneous andesites and dacites with only
a few rhyolites. Thouret et al. (2001) noted that the

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

eruptions at El Misti. Herein we report the results of a detailed petrological study of the 2000 BP eruption at El
Mistithe first of its kind. We establish the magmatic conditions of the andesite and rhyolite reservoirs, and the
physical, chemical, and mineralogical signals of their
interaction, and provide constraints on the timing of the
event that led to the eruption. This work provides valuable
petrological context to a case study of the stratigraphy
and volcanology and hazard assessment of the eruption
(Harpel et al., 2011).

It has long been recognized that magmatic recharge can


trigger explosive eruptions of a perched magma through
thermal, mass, and volatile exchange that results in pressurization of the system, and viscosity changes that result
in rheological and mechanical eruptive thresholds being
exceeded (e.g. Sparks et al., 1977; Blake, 1984). However, detailed studies of magmatic systems and single eruptions
reveal that these first-order results can be achieved in a
myriad of ways: mafic^mafic, mafic^silicic, and silicic^silicic interactions (Sparks et al., 1977; Eichelberger, 1978;
Feeley & Dungan, 1996; Eichelberger et al., 2000; de Silva
et al., 2008). Resident and recharge magma may or may
not achieve thermal and chemical equilibrium (e.g.
Pichavant et al., 2007). Exchange of crystals and redistribution is common (e.g. Davidson & Tepley, 1997; Ruprecht
et al., 2008). The scale of mixing and its controls on thermal
exchange and rates of equilibration, and changes in viscosity are particularly important (Huppert et al., 1982; Sparks
& Marshall, 1986; Snyder & Tait, 1995; Ruprecht &
Bachmann, 2010). All these processes are recorded in the
juvenile materials and revealed through detailed multiscale petrological studies (e.g. Tepley et al., 1999). When
based on a strong stratigraphic and volcanological foundation, such petrological studies form a crucial part of the
overall hazard assessment.
El Misti volcano (herein referred to as El Misti) in
southern Peru is one of the most hazardous volcanoes in
South America (de Silva & Francis, 1991a, 1991b; Thouret
et al., 2001; Harpel et al., 2011). Here, a population of
4800 000 live in Perus second largest city, Arequipa,
within 15 km of El Mistis summit vents. During its 112
kyr eruptive history, at least three major and several smaller explosive eruptions have punctuated the effusive background activity. Reconnaissance of these eruptions has
revealed macroscopic evidence for magma mingling
(Legros, 1998; Thouret et al., 2001) and petrological studies
of plagioclase from various eruptions have revealed that
these eruptions are preceded by multiple magma recharge
events that eventually precipitated the respective eruptions
(e.g. Ruprecht & Worner, 2007). If these observations hold
up to detailed scrutiny, the processes that drive explosive
volcanism at El Misti can be placed in the broader context
of the magmatic evolution of the system. To date no fully
contextual detailed study of an explosive eruption at El
Misti has been conducted.
The most recent explosive eruption at El Misti is the
VEI 5, 2000 BP eruption (Thouret et al., 2001; Harpel
et al., 2011), the products of which are exposed in multiple
drainage canyons on the south and west flanks of the volcano. This was a plinian fall and flow mixed rhyolite^
andesite tephra eruption, hypothesized to have involved a
recharge event based on abundant macroscopic and microscopic evidence from mixed pumices. As such the eruption
serves as a potential model for the other explosive

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

70

(a)

Caribbean Plate

(b)

10

NVZ
0

10

El Misti

30

Nazca Plate

SVZ
40

50

Ch
ile

Ris
e

Antarctic Plate
70

(c)

Fig. 1. (a) Map showing the location of El Misti in South America and its location in the Central Volcanic Zone. (b) Image of El Misti and the
surrounding region (from Harpel et al., 2011). Irregular white regions in bajadas are pyroclastic-flow deposits from the EM2000BP eruption.
Outlined by a white line is the city boundary of Arequipa. (c) Photograph of El Misti taken from downtown Arequipa, illustrating the proximity of a large population center to a potentially explosive volcano.

range of textures and mineralogy (compositions). Of these


four were chosen to represent the end-member textures
and compositions: two were dominantly of the white rhyolite component, and two others were composed primarily
of the black to brown andesite component. Each representative sample contains some mingled white and dark

heterogeneous textures of the banded andesites and rhyolites of the 2000 BP eruption are unique to El Misti compared with other volcanoes in the region, in both texture
and the presence of a distinct mineral suite. For the purposes of our study, 50 samples were collected from throughout the eruption stratigraphy and studied to establish the

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

zca
Na

ile Trench
Peru-Ch

20

CVZ
ge
Rid

JOURNAL OF PETROLOGY

VOLUME 0

component. The results reported in this study are from thin


sections of these four samples.

NUMBER 0

MONTH 2013

R E S U LT S
Lithology and whole-rock textures

A total of seven samples from the two lithological endmembers were analyzed for their major- and trace-element
compositions at the GeoAnalytical Lab at Washington
State University, by X-ray fluorescence spectroscopy
(XRF) and inductively coupled plasma mass spectrometry
(ICP-MS) techniques. Details of the techniques and their
associated analytical errors have been given by Johnson
et al. (1999) and Knaack et al. (1994), respectively.
Petrographic descriptions of the four representative samples provide records of the constituent phases, their abundance, and their textural relationship to the other phases.
Detailed analyses of minerals and glasses were performed
at Oregon State University using a CAMECA SX-100
electron microprobe (EMP) equipped with five wavelength-dispersive spectrometers (WDS) and high-intensity
dispersive crystals for high-sensitivity trace element analysis. Minerals and groundmass glasses were analyzed
using 15 keV accelerating voltage, 30 nA sample current,
and 1 mm beam diameter for mineral phases and 5 mm for
groundmass glasses. Counting times ranged from 10 to
60 s depending on the element and desired detection limit.
In all cases, zero-time intercept functions were applied to
reduce the effects of alkali migration. Data reduction was
performed online using a stoichiometric PAP correction
model (Pouchou & Pichoir, 1984). Back-scattered electron
(BSE) images were obtained using the same instrument
using the CAMECA Peak Site software. Precision measurements for the most significant elements in the glass,
feldspar, amphibole, pyroxene, and Fe^Ti oxides routines
are listed in Tables 5, 8, 6, 3 and 4, respectively.
Because some amphiboles in the selected samples exhibit
reaction rims whereas others do not, several amphiboles
from each lithology were selected for in situ trace element
analysis to determine population identity. Following EMP
analysis, analysis spots were chosen where EMP data
existed and in selected cores, mid-sections and rims of the
amphiboles. The analyses were carried out by laser ablation (LA)-ICP-MS in the Keck Collaboratory for Plasma
Spectrometry, Oregon State University, using a NewWave
DUV 193 nm ArF Excimer laser at 5 hz frequency, 15 ns
pulse duration and 50 mm beam size attached to a VG PQ
ExCell Quadrupole ICP-MS system and following the
techniques outlined by Kent et al. (2004). Concentrations
of single trace elements were calculated employing 43Ca
as an internal standard relative to the USGS glass standard BCR-2G. External errors are dependent on elemental
concentrations in the samples; however, calculated errors
are typically 5% for Sc, Cr, Rb, Y, Zr, Nb, La, Ce, Pr,
Nd, Sm, Eu, Gd, Dy, Er, Yb, and Pb, and 10% for V, Sr,
and Ba (1s).

Whole-rock geochemistry
Of the 50 collected samples of the eruption, seven samples
were chosen for major- and trace-element analysis. These
were chosen to check and supplement existing data from this

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

Both pyroclastic flow and fall deposits contain juvenile


clasts that display abundant evidence for magma mingling.
Two end-member lithologies, a plagioclase^amphibole
rhyolite and a plagioclase^amphibole andesite, are found
intimately mingled at different scales. Both are moderately
porphyritic. No pure end-member clasts were found, and
all the clasts show some mingling. The rhyolite forms a distinct pervasively micro-vesicular pumiceous lithology,
whereas the andesite occurs as a more obviously vesicular
scoriaceous lithology. A wide range of mingling relationships can be seen, from rhyolite-dominated to andesitedominated (Fig. 2). Evidence of mingling is abundant in
hand specimen as millimeter-scale wisps and selvages.
Andesite within dominant rhyolite tends to be in linear
wisps, selvages, and bands, but more complex relationships
are displayed as the lithologies become more andesitedominated. Complex sheath folding relationships can be
seen and thicker (centimeter-scale) bands of rhyolite show
clear evidence of ductile deformation with recumbent
folds (Fig. 2). Complex crenulation develops on rhyolite
selvages included in andesite. In several instances we
found that some of the rhyolitic wisps were rooted in
dense rhyolitic clasts (centimeter scale), which were being
disaggregated at their margins and being incorporated
into the andesite. Some grey selvages may represent a
hybrid lithology. We did not find any systematic stratigraphic variations in the distribution of the lithologies in
either the fall or flow deposits.
Diverse textural features characterize both the rhyolitic
pumice and andesitic scoria in thin section. Some clasts
show uniform distribution of a range of vesicle sizes
throughout the slide, but more commonly, particularly in
the rhyolite, heterogeneous clasts show distinct regions
where small bubbles (diameters 5^25 mm) predominate
and are surrounded by a matrix with intermediate-size to
coarse vesicles (75^100 mm and 175 mm diameters, respectively). Independent of the degree of heterogeneity, a
marked predominance of intermediate-size to coarse vesicles is conspicuous within some slides. The andesitic
scoria is characterized by largely equant to sub-spherical
vesicles with limited evidence for bubble deformation.
However, bubble deformation is ubiquitous in the rhyolite
pumice, which typically exhibits bands of elongated vesicles crossing larger regions with more equant bubbles.
The bands tend to range in width from 50 to 500 mm,
suggesting the presence of localized shear zones on a
range of scales.

A N A LY T I C A L M E T H O D S

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

(a)

Summary of rhyolite and andesite


petrology

(b)

The rhyolite-dominant samples are 50% vesicles, 40%


groundmass, including glass and microlites, and 10%
phenocrysts (4500 mm) and microphenocrysts (100^
500 mm) (Fig. 4). The dominant phenocryst and microphenocryst phases include sub-equal amounts of plagioclase
(6%) and amphibole (2%) with lesser amounts of pyroxene (1%), Fe^Ti oxides (1%), and high-SiO2 rhyolitic
glass (72^78 wt % SiO2). Plagioclase phenocryst and
microphenocryst compositions range from An30 to An85
and display simple to complex normal and oscillatory
zoning. Plagioclase microlite (100 mm) compositions
range from An28 to An63 encompassing two populations
of normally zoned microlites: one in the range An43 to
An63 and another in the range An28 to An44.
The andesite-dominant samples contain 40% vesicles,
50% groundmass of equal proportions of glass and microlites, and 10% phenocrysts and microphenocrysts (Fig. 4).
The groundmass comprises plagioclase, pyroxene, Ti-magnetite, and andesitic to rhyolitic glass (60^72 wt % SiO2).
Plagioclase phenocryst and microphenocryst compositions
range from An30 to An88, showing a similar but slightly
greater compositional range than the rhyolite samples;
they exhibit similar complex textural features. The plagioclase microlites range from An63 to An43 (Table 2).
Amphiboles occur in both lithologies as strongly pleochroic crystals 514 mm in length and are pargasitic
in composition (Mg# 075). They commonly contain
Fe^Ti oxide inclusions. Amphiboles in the andesite are euhedral and slightly zoned, whereas those in the rhyolite
are rimmed by plagioclase, pyroxene and Fe^Ti oxide reaction products of variable thickness (50^600 mm). The reaction rim occurs where the amphibole is in contact with
melt rather than with other crystalline phases.
Pyroxene phenocrysts and microphenocrysts are present
in both lithologies, accounting for 51% of the phenocrysts
in the rocks. Orthopyroxenes in the rhyolite are 51mm in
size, are euhedral to subhedral, and sparsely distributed.
They range in composition from En78 to En82 with an

(c)

Fig. 2. Hand samples illustrating the various textures of the tephra:


(a) sample containing thick globs of rhyolite in andesite matrix; (b)
thin wisps of rhyolite in a gradational rhyolite^andesite matrix; (c)
sample containing examples of both.

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

eruption. All samples are typical medium- to high-K calc-alkaline CentralVolcanic Zone (CVZ) andesites and rhyolites
(Fig. 3; Table 1). Our samples from the EM2000BP eruption
fall in a similar range in aplotof K2O v. SiO2 as other samples
from the 112 kyr history of the volcano and from the CVZ
in general (Legros, 1998; Legrende, 1999; Ruprecht &
Worner, 2007; Mamani et al., 2010; Fig. 3). Similarly, for other
major or trace elements, our samples fall within the data envelope of other CVZ volcanoes.Theyare characterizedby selective enrichment in large ion lithophile and alkaline earth
elements, attesting to the probable involvement of subduction-zone fluids, and lower abundances of rare earth elements
(REE) and high field strength elements, compared with typical mid-ocean ridge basalt, confirming their arc affinity.

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

MONTH 2013

6
CVZ
WR rho
WR and
EMP rhyo glass
EMP and glass

HIGH-K

MEDIUM-K

2
1

LOW-K

0
50

55

60

65

70

75

80

SiO2 wt%
Fig. 3. K2O vs SiO2 diagram for whole-rocks (filled circles and squares distinguished as rhyolite and andesite) and glasses (determined by
EMPA; open circles and squares) from EM2000BP, and their positions relative to a complete sampling of the Central Volcanic Zone (Mamani
et al., 2010). Most rocks are High-K calc-alkaline.

average composition of En80 (Table 3). In the pumiceous


andesite rocks, microphenocrysts of both clinopyroxene
and orthopyroxene are present as 51% of the crystals in
the rocks. Orthopyroxene occurs as euhedral or subhedral
microphenocrysts, and ranges in composition from En79
to En82, with an average composition of En80. The clinopyroxene occurs as microphenocrysts, and like the orthopyroxene, is euhedral to subhedral and 51mm in length.
Its compositions range between Wo42 and Wo47, with an
average composition of Wo45 (Table 3). There are no
Fe^Ti oxide pairs in the andesite, therefore we used coexisting pyroxene pairs to determine magma temperatures.
These coexisting pyroxenes yield temperatures of
940  408C for the andesite based on the thermometer of
Putirka (2008).
In the rhyolite samples, Fe^Ti oxides (both ilmenite and
magnetite) occur as discrete microlites, as inclusions in
amphibole and pyroxene, and as symplectites in the reaction rims of amphibole. Crystals are typically small (2^
20 mm), accounting for 1% of the mode. For temperature
calculations in the rhyolite, groundmass ilmenite and magnetite were used, and yielded temperatures of 816  308C
in the rhyolite based on the oxide thermometer of Ghiorso
& Evans (2008) (Table 4). The Fe^Ti oxides are in equilibrium based on the method of Bacon & Hirschmann (1988).
We determined glass compositions in the four targeted
thin sections by EMP analysis. The rhyolite glass compositions have a compositional range varying between
725 and 764 wt % SiO2 whereas the glass in the andesite ranges from 62 wt % SiO2 to 72 wt % SiO2

(Table 5). The full dataset of whole-rock XRF analyses


and EMP phase chemistries is provided as Supplementary
Data.

P H E N O C RY S T T E X T U R E S A N D
C O M P O S I T I O N A L PAT T E R N S
Amphibole
Amphibole occurs as ubiquitous crystals throughout both
rhyolite and andesite lithologies with a modal abundance
of about 2% in each lithology. The most obvious difference
between the two is that most (490%) of the amphiboles
that reside in the rhyolite have reaction rims of plagioclase,
pyroxene, and Fe^Ti oxides, or occur as ragged clusters,
whereas most (490%) of those in the andesite do not
display a reaction rim. Different rim widths may reflect
differential sectioning of crystals rather than any processrelated phenomenon. Most amphiboles in both lithologies
show some evidence of minor compositional zoning based
on EMP analyses and BSE images.
Amphiboles in both lithologies of the EM2000BP eruption show a relatively small range of major element variation, forming relatively tight trends over 3 wt %
absolute spread in SiO2 (Fig. 5; Table 6). Compositions of
the amphiboles from both lithologies are similar, although
amphibole from the andesite defines the full range for
almost all major oxides.
Cation abundances, used to constrain the amphibole
classification and decipher petrogenetic processes, were
calculated assuming a formula cation sum of 15 excluding

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

K2O wt%

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

SiO2
TiO2

EM

EM

EM

EM

EM

EM

EM

EM

007

008

009

085

094

098

099

0401

5997
0890

6076
0799

6056
0822

6121
0774

6064
0804

6093
0807

6963
0359

5999
0776

Al2O3

1771

1775

1770

1767

1786

1779

1564

1757

FeO*

574

541

552

529

540

531

264

559

MnO

009

009

009

009

009

009

007

010

MgO

295

259

267

246

256

252

105

331

CaO

589

563

571

553

570

563

292

598

Na2O

431

443

440

438

441

434

386

425
218

K2O

215

225

222

231

223

228

370

P2O5

030

030

030

029

030

030

014

025

Total

10000

10000

10000

10000

10000

10000

10000

10000

La

2475

2546

2520

2568

2527

2551

3172

2599

Ce

5001

5103

5072

5104

5090

5121

5738

5128

Pr

614

618

613

617

618

619

613

611

Nd

2417

2422

2385

2382

2411

2413

2118

2344

Sm

454

445

445

439

448

443

353

427

Eu

126

123

126

125

126

125

090

122

Gd

357

339

340

330

338

339

264

330

Tb

048

045

046

045

046

045

037

046

Dy

242

235

241

236

236

235

217

246

Ho

044

043

044

042

044

042

042

046

Er

108

107

109

107

104

108

116

118

Tm

015

015

015

015

015

015

018

016

Yb

091

090

091

086

090

090

122

102

Lu

014

014

014

014

014

014

020

Ba

907

928

917

925

921

927

1092

016
941

Th

232

255

248

267

239

257

763

330

Nb

585

599

606

611

606

613

708

522

1128

1100

1108

1091

1098

1107

1153

1193

Hf

401

406

409

401

403

403

404

391

Ta

036

037

036

037

037

037

061

030

041

043

042

046

041

043

122

044

Pb

1314

1370

1350

1398

1341

1381

2395

1299

Rb

375

403

396

424

394

411

948

442

Cs
Sr
Sc
Zr

086
836
93
151

091
834
81
153

090
835
86
153

097
829
77
154

086
850
81
154

092
840
78
155

248
513
51
145

Plagioclase
Based on EMP analyses and backscattered electron images,
plagioclase phenocrysts, microphenocrysts and microlites
in both rock types define a large compositional range.
Phenocryst sizes range from 01 to 15 mm; we define
the boundary between phenocryst and microphenocryst
at 05 mm and microlites as 01mm. We group the
crystals based on composition into two broad groups: a
Low-An group, which ranges from An60 to An30, and a
High-An group, which ranges from An88 to An65. This
classification is based on An content frequency analyses of
the total plagioclase dataset (Fig. 9), supported by a MgO
wt % frequency histogram. Mirroring the compositional
variations are two broad classes of textural varieties based
on crystal morphology and texture: clear crystals, and
crystals with alternating sieved or dusty and clear portions.

066
840
125
150

*Total Fe given as FeO.

Na and K (15eNK). Amphiboles in both the rhyolites and


andesites are pargasitic [nomenclature of Leake et al.
(1997)]. They show a moderate but significant range in
AlIV from 165 to 2 atoms per formula unit (p.f.u.), although, as in the major oxides case, the amphiboles from
the andesites tend to anchor the high and low end of the

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

trends. Variations of AlVI and (Na K)A with AlIV are


small and form nearly horizontal trends (Fig. 6). Mg#
[Mg/(Mg Fe2)] varies between 07 and 08 with no
distinction between amphiboles in rhyolite or andesite,
and decreases with increasing AlIV, as seen in other studies
(e.g. Rutherford & Devine, 2003). Core and rim data
from both lithologies show no preference for higher or
lower AlIV in variation with AlVI, (Na K)A and Mg#.
Compositional zoning within any single amphibole
phenocryst represented by either core-to-rim transects or
single EMP spots is illustrated in Fig. 7. The chemical variations within single crystals are shown in relation to BSE
images, and representative amphibole samples in both
rhyolite and andesite are illustrated. Representative compositions are given in Table 6.
Selected amphibole crystals from both rhyolite and andesite were chosen for detailed in situ LA-ICP-MS trace
element analysis primarily to determine whether a correlation exists between amphiboles in the different lithologies.
Further, analysis locations were chosen coincident with
electron microprobe locations to utilize the major-element
microprobe data for trace element calibration (see Kent
et al., 2004) and to evaluate the chemical differences between different zones in the phenocrysts. Chondrite-normalized REE patterns of in situ LA-ICP-MS data for
amphiboles from both rhyolites and andesites show a classic convex form; the patterns and normalized concentrations are nearly identical for 495% of the samples,
regardless of the host-rock lithology (Fig. 8; Table 7). Light
REE (LREE) and middle REE (MREE) abundances
(La/SmN vs LaN; not shown) and LREE and heavy REE
(HREE) abundances (La/YbN vs LaN; not shown) also
demonstrate that although variations in normalized concentrations are present, they are minor. Lastly, there are
minor variations between compatible Yand slightly incompatible Sr when the grouped data are considered. In all
cases, there is no distinction between the chemical signatures of amphiboles hosted in the rhyolite or andesite.

Table 1: Representative whole-rock major and trace element


compositions from the 2000 BP eruption of El Misti Volcano

JOURNAL OF PETROLOGY

(a)

VOLUME 0

NUMBER 0

MONTH 2013

(d)

andesite

rhyolite

(e)

(b)

rhyolite

andesite

(f)

(c)

andesite

andesite

Fig. 4. Photomicrographs of amphibole with and without reaction rims and Low-An and High-An Group plagioclase in both rhyolite and andesite. Field of view in all images is 2 mm, and all images are in crossed polars. (a) Boundary (dashed line) between rhyolite and andesite
with reacted and unreacted amphibole, respectively. (b) A rhyolite-hosted amphibole with reaction rim. (c) A clear elongate amphibole residing
in andesitic melt. (d) A complexly zoned plagioclase crystal in rhyolite host from the Low-An group. (e) A plagioclase crystal with complex
dusty core and clear outer rim from the High-An group, in andesitic host. (f) A predominantly clear plagioclase crystal with minor zoning, a
member of the Low-An group, in andesitic host.

Low-An plagioclase crystals are morphologically clear and


simple, and tend to represent the rhyolite, whereas HighAn plagioclase crystals are complexly zoned and textured,
and tend to reside in the andesite. Occasionally these general host^plagioclase relationships are reversed, attesting
to crystal exchange between the two hosts. Figure 10

illustrates some of the various plagioclase types and sizes,


and Table 8 gives the representative compositions.

Low-An plagioclase group (An60^30)


The Low-An plagioclase phenocryst group have maximum
core An contents of An60, and minimum rim An contents

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

rhyolite

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

Table 2: Summary of rhyolite and andesite petrography and compositions


Rock type, proportions

Phases

Characteristics

Rhyolite
small and intermediate to coarse

525 mm, 75100 mm

40% groundmass

glass

7278 wt % SiO2

microlites

An2844; An4363 bimodal

10% phenocrysts,

plagioclase (6%)

An3085 complexly zoned

microphenocrysts

amphibole (2%)

pargasite with reaction rims

FeTi oxides (1%)

8168C  308C

pyroxene (1%)

En7882, av. En80

Andesite
40% vesicles

equant to sub-spherical

50% groundmass

glass

6272 wt % SiO2

microlites

An4363

10% phenocrysts,

plagioclase (8%)

An3088 complexly zoned

microphenocrysts

amphibole (2%)

pargasite euhedral

pyroxene (51%)

En7982, av. En80 Wo4247, av. Wo45

2-pyx temperature

9408C  408C

of An30. This population of phenocrysts is generally normally zoned, although most of those either imaged (BSE)
or analyzed show one zone of increased An content outboard of the core before decreasing to values An40^30
near the rims (Fig. 10a). Backscattered electron images
show that the low-An cores have rounded interior borders
that changed immediately to higher An values, although
these higher values are a few mol % An higher.
Texturally, these phenocrysts are generally simple, euhedral, clear crystals, although there are sparse crystals with
mottled cores.
The trace elements Mg, Ti and Fe were measured simultaneously with major elements during analysis transects.
Generally, their concentrations are low, given the incompatibility of these elements in plagioclase. Transects of
FeO show two patterns: one pattern shows little variation
regardless of changing An content (2B and 11E), whereas
the other pattern is antithetic to An content (e.g. samples
5I and 10G). MgO concentrations are more variable than
FeO, but they display the same patterns relative to An
content.

in their complex textural features, characterized by obvious dusty or sieved portions alternating with clear portions. In most examples, the cores of these crystals are
sieved or dusty and alternate with clear portions outwards
towards the rim; in a few cases, the cores of these crystals
are clear. We see no compositional differences between the
crystals with clear cores versus dusty cores.
Trace element concentrations are generally higher and
their distribution patterns are different from those of the
Low-An group. In contrast to the Low-An group, FeO
variations in all cases are regular and unchanging regardless of the variations in An content (Fig. 10b). However,
MgO appears to be sensitive to changes in An content,
producing an anti-correlation in MgO.

Microlites
Microlites in the rhyolite and andesite also show some
compositional heterogeneity. Frequency histograms show
that there are two populations of microlite compositions,
one in the rhyolite, and one in the andesite. Both populations are normally zoned. The microlite population in the
andesite has cores of An54^63 and rims of An43^63, whereas
the microlite population in the rhyolite has cores of
An41^44 and rims of An28^38 (Fig. 9). These populations of
microlites are distinct in composition from the phenocrysts
in their respective host lavas: they are slightly more
evolved. We attribute these compositional characteristics
to lower pressure final equilibration and lower magma
pH2O values.

High-An plagioclase group (An88^65)


The second broad group of plagioclase phenocrysts includes those with compositions between An88 and An65,
with average core An contents of An80. Rim compositions
depend on whether the crystals are hosted in rhyolite or
andesite. This crystal population is also normally zoned,
but most crystals have complex zoning patterns extending
to the rim. This complex compositional zoning is reflected

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

50% vesicles

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

MONTH 2013

Table 3: Representative compositions of clinopyroxene and orthopyroxene phenocrysts


clinopyroxene

orthopyroxene

EM 10 L

EM 10 L

EM 10 N

EM 10 N

EM 11 M

EM 11 M

EM 2 J

EM 2 J

SiO2

5032

5246

4961

5173

5209

Al2O3

320

115

371

155

121

FeO

877

812

1004

980

MgO

1467

1569

1369

CaO

2086

2104

2050

Na2O

045

043

TiO2

088

MnO

038

Total

9953

EM 10 M

5234

5368

5294

5289

114

098

112

114

929

912

1945

1982

1584

1445

1479

2481

1896

2110

2113

100

050

036

046

039

043

105

059

019

037

029

032

044

9968

9940

9914

9924

EM 10 M

EM 11 L

EM 11 L

5301

5361

5322

100

085

098

1998

2004

1831

1846

2429

2445

2440

2524

2494

100

109

109

104

104

003

004

002

001

003

003

014

012

012

030

026

019

021

035

067

070

066

060

058

061

9941

10075

10003

10052

10041

9986

9950

Typical 1SD: SiO2  012; Al2O3  004; FeO*  001; MgO  004; CaO  005; Na2O  004; TiO2  001; MnO  003.

Table 4: Representative compositions of magnetite and ilmenite phenocrysts


magnetite
Sample:

EM 2 O

ilmenite
EM 2 O

EM 2 O

EM 2 O

EM 2 N

EM 2 N

EM 2 O

EM 2 O

EM 2 O

EM 2 N

EM 2 N

EM 2 N

SiO2

005

005

007

005

004

004

000

001

001

003

003

004

TiO2

653

628

620

625

669

622

3843

3795

3708

3721

3691

3725

Al2O3

192

151

153

150

150

147

014

015

016

020

017

025

V2O3

044

048

045

050

044

048

027

032

031

034

030

035

Cr2O3

010

010

011

007

008

005

000

001

002

000

000

000

FeO*

8182

8247

8364

8345

8294

8223

5479

5456

5564

5429

5295

5357

MnO

039

051

052

053

049

058

057

060

052

057

051

056

MgO

184

173

176

167

167

158

237

238

241

262

253

276

CaO

003

001

002

001

004

001

002

001

001

001

003

003

ZnO

018

007

011

006

009

017

010

003

002

004

007

000

Total

9330

9321

9442

9411

9399

9284

9670

9604

9618

9532

9395

9483

*Total Fe given as FeO.


Typical 1SD: (magnetite) SiO2  001; TiO2  001; Al2O3  0005; V2O3  002; Cr2O3  0025; FeO*  03; MnO  002;
MgO  002; CaO  0004; ZnO  001; (ilmenite) SiO2  001; TiO2  04; Al2O3  002; V2O3  003; Cr2O3  001;
FeO*  065; MnO  01; MgO  001; CaO  001; ZnO  001.

during the Pleistocene. The pumice of the 2000 BP eruption


of El Misti is extensively banded and heterogeneous at
macroscopic and microscopic scales, and vesicle textures in
the respective lithologies record differences in rheology and
the results of the interaction (shearing, vesicle trains, etc.).
Petrographic evidence for crystal exchange is supported
by phenocryst compositions that indicate two populations
of plagioclase phenocrysts and microlites based on composition and texture. Low-An plagioclase, morphologically
clear and simple crystals formed in the rhyolite, and

DISCUSSION
The details of magma mingling
The macroscopic and microscopic lithological, petrographic, and petrological observations presented above are
all consistent with extensive mingling of a relatively hot
(940  408C) andesite and a cooler (816  308C) rhyolite
magma prior to the 2000 BP eruption of El Misti. These
magmas are typical of the calc-alkaline high-K suite of
magmas that have erupted in the Central Volcanic Zone

10

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

Sample:

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

Table 5: Representative glass compositions


Sample:

10

SiO2
TiO2

10

7466

7569

7432

7467

6681

7358

6607

033

033

034

033

079

077

056

075

076

Al2O3

1334

1347

1402

1429

1521

1595

1442

1268

1411

FeO*

149

147

148

141

491

418

226

588

376

MnO

005

007

000

008

008

009

003

013

010

MgO

030

029

018

018

196

096

041

420

217

CaO

115

103

073

082

288

304

155

284

255

Na2O

200

236

307

293

293

351

284

272

281

685

11 clear

11 dark

11 dark

6946

464

466

481

471

338

336

426

356

383

004

004

003

002

037

033

026

032

033

Cl

015

013

019

018

011

014

014

013

013

9818

9956

9917

9963

9946

10084

10035

9929

10002

Total

*Total Fe given as FeO.


Typical 1SD: SiO2  007; TiO2  001; Al2O3  003; FeO*  012; MnO  002; MgO  001;
CaO  003; Na2O  004; K2O  003; P2O5  001.

high-An plagioclase, complexly zoned and textured crystals


formed in the andesite, now occur in both andesite and
rhyolite, attesting to crystal exchange. Based on the mineral
composition data, we infer that amphibole grew in the andesite magma at depth. However, amphibole can now be
found in both the rhyolite and andesite, with the amphibole
in the rhyolite exhibiting reaction rims.
Having established these baseline characteristics, below
we explore the deeper issues of the complex trace element
systematics of the plagioclase and reaction rim development on amphibole in the rhyolite, and how these relate
to the timing and development of the system as a whole.

as will increases or decreases in the pH2O of the system


(e.g. Housh & Luhr, 1991; Lange et al., 2009). Closedsystem processes that effect compositional changes, such as
crystal entrainment in convective currents within a
magma chamber (Singer et al., 1995) or density currents
induced from overburdened sidewall or roof crystallization
(Marsh, 1989), may occur without the interaction of different magmas. Open-system processes, such as magma recharge, change not only the composition of the system
and its temperature but also the equilibrium plagioclase
composition.
Discriminating between competing intensive and extensive variables requires evaluation of the minor and trace
element compositions of the plagioclase, which are less susceptible to changing intensive parameters. In closed systems, equilibrium crystallization of plagioclase and the
associated partitioning of trace elements into plagioclase
will be governed by crystal chemical controls on elemental
partitioning (e.g. Blundy & Wood, 1994) and melt compositional controls (Nielsen & Drake, 1979; Nielsen &
Dungan, 1983). An exception to these rules is the non-equilibrium effects of variable diffusion of trace elements to
and from the crystal^melt interface during rapid crystal
growth that may lead to significant departures from equilibrium element partitioning (Albare'de & Bottinga, 1972;
Shimizu, 1983; Singer et al., 1995). Recharge events bring
about changes in temperature, pressure, pH2O and melt
composition, which may change plagioclase compositions
and the trace element composition of the melt, and therefore the equilibrium partitioning of that trace element
into plagioclase. Recharge may also mix two populations
of plagioclase crystals with different compositions.

Plagioclase trace element systematics


Plagioclase compositions in the 2000 BP eruption display a
wide range of variability; we have distinguished two
groups, the High-An Group and the Low-An Group,
based on their predominant compositions and textures.
However, overlap in An content and, in some cases, textural features limits our ability to definitively discriminate
between the two populations of plagioclase phenocrysts.
In this case, we have turned to trace element concentrations in the plagioclase as an efficient discriminator.
Plagioclase compositions are controlled by the melt
composition and its H2O content, and the intensive parameters, temperature and pressure, of the crystallizing
system (Bowen, 1928; Tsuchiyama, 1985; Housh & Luhr,
1991). Changes in these variables can lead to variations in
plagioclase composition (crystal zoning), and possibly in
the rate of crystal growth (i.e. crystal growth kinetics).
Changes in the temperature of the system will change the
equilibrium composition of the plagioclase in that system,

11

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

K2O
P2O5

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

MONTH 2013

1.0
pargasite

magnesiosadanagaite

0.8
0.6
0.4

Amphiboles in
rhyolite melt

0.2
ferropargasite

ferro-edenite

0.0
7.5

7.0

Amphiboles in
andesite melt

6.5

6.0

5.5

5.0

4.5

Si pfu
14.0

13.0
Amphiboles in rhyolite melt

12.5

Amphiboles in andesite melt

13.0

CaO

Al2O3

13.5

12.5

12.0
11.5

12.0

11.0

11.5

10.5

11.0
41.0

42.0

43.0

10.0
41.0

44.0

42.0

SiO2
18.0

14.0

17.0

MgO

13.0

FeO*

44.0

43.0

44.0

SiO2

15.0

12.0
11.0

16.0
15.0
14.0

10.0

13.0

9.0
8.0
41.0

43.0

42.0

43.0

44.0

SiO2

12.0
41.0

42.0

SiO2

Fig. 5. Upper diagram shows EM2000BP amphibole phenocryst compositions from both rhyolite- and andesite-dominated samples plotted in
the classification scheme of Leake et al. (1997) using the Mg# [Mg/(Mg Fe2)] vs Si p.f.u. (per formula unit) diagram. All are pargasitic
amphibole. Lower diagrams show amphibole variations in Al2O3, FeO* (total iron), CaO and MgO vs SiO2.

the plagioclase concentrations, we use starting melt compositions obtained from the rhyolite and andesite whole-rock
compositions (see Table 1), and temperatures calculated
from oxide pairs in the rhyolite and pyroxene pairs in the

Plotted in Fig. 11 are equilibrium partitioning concentration curves of MgO, TiO2 and FeO based on the plagioclase^melt trace element partitioning experiments of
Bindeman et al. (1998) and Tepley et al. (2010). In modeling

12

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

Mg/(Mg+Fe2+)

edenite

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

Table 6: Representative compositions and structural formulae of rhyolite- and andesite-hosted hornblende types
rhyolite-hosted amphiboles
Sample:

EM 2

EM 2

EM 2

EM 2

EM 2

EM 2

EM 5

EM 5

EM 5

EM 5 A

EM 5

EM 5

EM 5

EM 5

I(2)

I(2)

I(2)

I(3)

I(3)

cl1

cl1

cl2A

cl2A

cl2A

cl2B

cl2B

cl2B

4236

SiO2

4228

4230

4247

4243

4238

4250

4275

4284

4297

4303

4232

4272

4250

4253

TiO2

248

253

244

246

254

243

237

243

240

230

241

240

235

255

250

Al2O3

1292

1298

1260

1256

1254

1236

1219

1259

1236

1222

1252

1228

1226

1234

1251

Cr2O3

002

002

002

000

001

000

000

000

000

017

002

000

007

002

006

FeO*

1203

1222

1139

1200

1191

1207

1175

1181

1199

1064

1171

1167

1133

1192

1193

MnO

010

013

010

015

007

008

011

011

015

011

010

011

009

015

015

MgO

1458

1424

1480

1458

1448

1458

1469

1468

1488

1548

1476

1488

1497

1463

1450

CaO

1181

1173

1158

1156

1156

1162

1155

1171

1173

1141

1157

1145

1137

1147

1137

Na2O

228

230

224

232

231

223

225

231

229

225

226

228

230

227

227

K2O

053

063

060

060

056

052

051

055

055

059

054

053

058

053

058

Cl
Total

002

002

002

002

003

002

002

003

003

002

002

002

001

002

002

9909

9918

9834

9874

9843

9846

9825

9911

9940

9826

9825

9836

9789

9849

9831

Si

6103

6122

6170

6157

6166

6175

6221

6184

6184

6230

6153

6203

6199

6181

6172

AlIV

1897

1878

1830

1843

1834

1825

1779

1816

1816

1770

1847

1797

1801

1819

1828

SUM T

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

AlVI

0301

0336

0328

0307

0318

0293

0312

0328

0281

0316

0299

0305

0307

0295

0321

Ti

0265

0272

0263

0265

0274

0262

0255

0260

0256

0247

0260

0258

0254

0275

0270

Fe3

0320

0228

0224

0236

0203

0279

0218

0213

0281

0192

0282

0227

0212

0225

0202

Cr

0002

0002

0002

0000

0002

0000

0000

0000

0000

0019

0002

0000

0008

0002

0007

Mg

3137

3072

3204

3153

3142

3157

3186

3158

3192

3341

3197

3221

3256

3170

3150

Fe2

0974

1090

0979

1039

1062

1011

1029

1042

0994

0885

0960

0989

0964

1033

1050

Mn

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

SUM C

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

Fe2

0158

0161

0181

0180

0185

0177

0183

0172

0169

0212

0182

0201

0207

0191

0201

Mn

0012

0016

0013

0018

0009

0010

0013

0013

0018

0014

0012

0013

0012

0018

0019

Ca

1827

1819

1803

1797

1802

1809

1800

1811

1809

1770

1802

1781

1777

1786

1775

Na

0004

0004

0004

0004

0004

0004

0004

0004

0004

0004

0004

0004

0004

0005

0005

SUM B

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

Na

0635

0641

0628

0648

0647

0624

0632

0642

0634

0627

0634

0638

0645

0635

0636

0097

0117

0112

0110

0104

0096

0096

0102

0100

0109

0099

0097

0108

0099

0108

SUM A

0733

0757

0739

0758

0750

0720

0727

0744

0735

0736

0733

0736

0753

0734

0745

Mg/(Mg Fe2)

0735

0711

0734

0721

0716

0727

0724

0722

0733

0753

0737

0730

0736

0721

0716

andesite-hosted amphiboles
Sample:

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 11

EM 11

EM 11

EM 11

K(2)

K(2)

4263

SiO2

4199

4195

4162

4388

4274

4380

4303

4227

4272

4258

4245

4321

4269

4277

TiO2

258

253

264

230

243

229

239

247

251

254

254

222

235

244

233

Al2O3

1286

1284

1313

1187

1259

1198

1262

1254

1237

1255

1274

1267

1195

1253

1264

Cr2O3

003

001

000

012

011

050

007

004

005

002

002

014

000

017

002

FeO*

1141

1176

1192

1011

1162

1018

1147

1189

1120

1126

1204

1051

1165

1152

1235

(continued)

13

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

EM 2
I

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

MONTH 2013

Table 6: Continued
andesite-hosted amphiboles
Sample:

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 10

EM 11

EM 11

EM 11

EM 11

K(2)

K(2)

012

014

008

007

016

007

009

011

009

011

011

010

011

010

012

MgO

1486

1467

1441

1615

1487

1603

1503

1442

1532

1487

1442

1586

1468

1502

1451

CaO

1121

1153

1134

1141

1159

1161

1175

1175

1160

1166

1175

1123

1155

1145

1139

Na2O

232

239

237

228

231

229

235

221

236

228

228

239

220

230

223

K2O

054

057

058

052

060

061

053

054

060

056

058

059

059

059

054

Cl

002

001

002

002

002

003

002

002

002

002

002

002

002

001

002

9799

9846

9816

9875

9906

9942

9942

9837

9889

9847

9901

9896

9783

9897

9887

Total
Si

6119

6098

6077

6298

6168

6260

6181

6152

6165

6173

6144

6204

6241

6175

6174

AlIV

1881

1902

1923

1702

1832

1740

1819

1848

1835

1827

1856

1796

1759

1825

1826

SUM T

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

8000

AlVI

0328

0299

0337

0307

0310

0278

0318

0303

0269

0318

0318

0348

0300

0308

0332

Ti

0278

0273

0285

0244

0260

0242

0255

0267

0268

0273

0272

0234

0254

0260

0250

Fe3

0227

0270

0229

0156

0223

0167

0225

0275

0245

0208

0236

0178

0214

0215

0258

Cr

0004

0002

0000

0013

0013

0056

0008

0005

0005

0002

0002

0016

0000

0019

0003

Mg

3228

3180

3136

3456

3199

3416

3219

3128

3297

3214

3112

3394

3198

3233

3133

Fe2

0935

0977

1014

0824

0995

0841

0976

1023

0916

0985

1060

0829

1037

0964

1024

Mn

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

0000

SUM C

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

5000

Fe2

0228

0182

0212

0233

0184

0209

0178

0150

0191

0172

0160

0255

0174

0211

0214

Mn

0015

0018

0010

0008

0019

0008

0011

0014

0011

0013

0013

0012

0013

0012

0014

Ca

1751

1796

1773

1754

1793

1778

1808

1833

1794

1811

1822

1728

1809

1772

1767

Na

0005

0004

0005

0005

0004

0004

0004

0003

0004

0004

0004

0005

0004

0005

0004

SUM B

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

2000

Na

0650

0668

0666

0629

0643

0631

0651

0622

0656

0638

0636

0662

0619

0640

0622

0100

0105

0107

0096

0111

0112

0097

0101

0110

0104

0108

0109

0109

0109

0099

SUM A

0750

0773

0773

0724

0754

0743

0748

0722

0766

0742

0744

0770

0728

0748

0721

Mg/(Mg Fe2)

0735

0733

0719

0766

0731

0765

0736

0727

0749

0735

0718

0758

0725

0733

0717

*Total Fe given as FeO.


Typical 1SD: SiO2  011; TiO2  004; Al2O3  005; Cr2O3  002; FeO*  017; MnO  002; MgO  007; CaO  006;
Na2O  007; K2O  004; Cl  0005.

andesite, and then calculate the trace element equilibrium


concentration in the plagioclase. On each diagram, two
equilibrium-partitioning curves are plotted representing
the equilibrium conditions of plagioclase crystals growing
in the rhyolite (Low-An type) and those growing in the andesite (High-An type), labeled as low Tand high T, respectively. Plotted with these equilibrium-partitioning curves are
MgO, TiO2 and FeO compositions measured simultaneously with An content via EMP analysis. In the MgO and
TiO2 diagrams, two swaths of data are prominent, which
plot on or near the equilibrium concentration lines. The
first observation is that trace element concentrations in
plagioclase allow us to discriminate between the two

populations of crystals. The second observation is that, for


the most part, equilibrium crystallization of plagioclase
occurred, and the large variations in An content in both
clusters are consistent with closed-system evolution associated with small variations in H2O and/or temperature of
the host magma. In contrast, FeO shows large variations in
An content with small or no changes in FeO, which suggest
that other factors, such as fO2, contributed to the partitioning of Fe in plagioclase phenocrysts that did not affect Mg
or Ti partitioning. Based on these observations, we conclude
that plagioclase phenocrysts in both the rhyolite and andesite grew independently of each other in relatively consistent
environments before being mingled together and erupted.

14

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

MnO

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

0.6

AlVI (pfu)

0.5

Amphibole in andesite melt

0.4
0.3
0.2

10K

10H

0.1

(a)

0.0

(Na+K)A (pfu)

1.0

Amphibole textures: the significance of reaction rims


The reaction rims on amphibole in the rhyolite lavas are
composed of intergrowths of plagioclase, orthopyroxene,
clinopyroxene and Fe^Ti oxides; the rims occur only
where amphibole edges are in contact with melt, not other
crystals. The rims are of relatively uniform thickness
around selected amphiboles and generally retain the precursor euhedral shape of the amphibole. These observations suggest that the reaction rims grew inward from the
amphibole edge such that the host melt plays an integral
role in the development of the rim (e.g. Rutherford &
Hill, 1993; Browne & Gardner, 2006; Buckley et al., 2006).
These gabbro-type reaction rims on amphibole are often
interpreted as resulting from volatile exsolution as a consequence of H2O loss during magma decompression during
movement to or storage at shallow depth (e.g. Garcia &
Jacobson, 1979; Rutherford & Hill, 1993; Rutherford &
Devine, 2003). The major percentage of amphiboles in the
andesitic host magma have no reaction rims, whereas the
majority of amphiboles in the rhyolite have reaction rims.
We evaluate the re-equilibration process through a detailed
mass-balance analysis of the amphiboles, their reaction
rims, and the surrounding melt, because this information
has bearing on the mechanism, timing and evolution of
the reaction rims.
We determined the distribution and proportion of phases
in the reaction rims using high-resolution X-ray element
mapping on the OSU electron microprobe. X-ray intensity
maps of Al, Fe, Ca and Mg were produced for selected
reacted amphiboles to determine the spatial distribution
and proportion of phases, in which Al is diagnostic for
plagioclase, Fe for Fe^Ti oxides, and Mg and Ca for clinopyroxene and orthopyroxene. The X-ray images were imported and modified in Adobe PhotoshopTM. The layers
for each element were stacked, the reaction rim on the
inside and outside was outlined, and the interior and exterior pixels were cut away leaving only reaction rim pixels.
Within each element layer, pixels of a limiting threshold
were highlighted and counted, and the total pixels from
each layer were summed. To obtain the areal proportion
of a phase in the reaction rim, each layers pixels were
ratioed to the summed pixels of the reaction rim. Table 9
lists the relative proportion of plagioclase, orthopyroxene,
clinopyroxene and Fe^Ti oxides in five amphibole rims.

0.9
0.8
0.7
0.6

(b)

0.5

Ti (pfu)

0.4

0.3

0.2

(c)
0.1

Mg/(Mg+Fe2+

1.0
0.9
0.8
0.7
0.6

(d)

0.5
1.6

1.7

1.8
AlIV

1.9

2.0

2.1

(pfu)

Fig. 6. Amphibole atomic (p.f.u., per formula unit) compositions and


evaluation of substitution mechanisms. (a) AlVI shows no change
with increasing AlIV in the pressure-sensitive Al-Tschermak substitution, reflecting no change in pressure at time of crystallization and
growth. Included in the diagram are lines labeled 10K and 10H indicating the range of AlIV for two amphibole crystals (10K and 10H)
from the andesite. Temperature-dependent exchanges, such as the edenite exchange (b) and the Ti-Tschermak exchange (c), indicate slight
temperature fluctuations. (d) A slight decrease in Mg# [Mg/
(Mg Fe2)] with increasing AlIV is indicative of growth in a fractionating liquid.

15

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

The advantage of these geochemical discriminators is


that regardless of whether a plagioclase crystal is found in
the rhyolite-dominated end-member or the andesite-dominated end-member, the trace element characteristics
coupled with the An content can reveal the original premixed environment of crystallization: the andesite reservoir or the rhyolite reservoir. With a system dominated by
mixed tephra containing mixed crystal populations, this
gives us the ability to elucidate the mixing process.

Amphibole in rhyolite melt

JOURNAL OF PETROLOGY

(a)

VOLUME 0

NUMBER 0

MONTH 2013

Laser spot
Microprobe spot
Microprobe transect

500 um

1000 um

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

0.80

0.80

EM10 andH

EM10 andI

0.75

Mg#

Mg#

0.75

0.70

0.70

0.65

0.65

200

400

600

1.5

800

1.6

1.7

1.8

1.9

2.0

2.1

2.2

AlIV pfu

Distance (m)

1000 um

200 um
0.80

0.80

EM10 andK

EM11 rhyH
0.75

Mg#

Mg#

0.75

0.70

0.70

0.65

0.65

200

400

600

800

200

Distance (m)

400

600

800

Distance (m)

200 um

500 um
0.80

0.80

EM11 rhyI

EM11 rhyK2

0.75

Mg#

Mg#

0.75

0.70

0.70

0.65

0.65
1.5

1.6

1.7

1.8 1.9
AlIV pfu

2.0

2.1

2.2

1.5

1.6

1.7

1.8

1.9

2.0

2.1

2.2

AlIV pfu

Fig. 7. BSE and compositional diagrams for representative amphibole phenocrysts from rhyolite and andesite. Illustrated are BSE images
plotted with EMP traverse or spot point locations and LA-ICP-MS spot locations. In traverses, the Mg# data are plotted versus distance
from rim, and in spot analyses, the Mg# data are plotted versus AlIV p.f.u.
(continued)

16

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

(b)

Laser spot
Microprobe spot
Microprobe transect

500 um

200 um

EM2 rhyI3

EM2 rhyI
0.75

Mg#

Mg#

0.75

0.70

0.70

0.65

0.65
1.5

1.6

1.7

1.8

1.9

AlIV

2.0

pfu

2.1

2.2

1.5

1.6

1.7

1.8

1.9

2.0

2.1

2.2

AlIV pfu

200 um

200 um

0.80

0.80

EM5 rhy2B

EM5 rhy cl1

0.75

Mg#

Mg#

0.75

0.70

0.70

0.65

0.65

200

400

600

1.5

800

1.6

1.7

1.8

1.9

2.0

2.1

AlIV pfu

Distance (m)

500 um

200 um
0.80

0.80

EM5 rhy2A

EM2 rhyI2

0.75

Mg#

Mg#

0.75

0.70

0.70

0.65

0.65

200

400

600

800

200

400

600

Distance (m)

Distance (m)

Fig. 7. (Continued)

17

800

2.2

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

0.80

0.80

JOURNAL OF PETROLOGY

VOLUME 0

Concentration/
Chondrite

100

hbl melt ! cpx opx plag ilm:

10

Ce

Pr

Nd Sm Eu

Gd

Dy

Er

Yb

hbl ! cpx opx plag mag ilm:

100

They applied this equation to amphiboles from Soufrie're


Hills Volcano and found that the mass balance similarly
followed equation (2), but required an open system in
which some components in the amphibole are exchanged
with the melt, and vice versa. The main difference between
equation (1) and equation (2), therefore, is that wholesale
melt interaction is required in the former, whereas selective
component interaction occurs in the latter.
Applying this method to the El Misti amphiboles, it is
noted that the mineral phase mode in the reaction rim on
El Misti amphiboles is dominated by plagioclase followed
by orthopyroxene, clinopyroxene and Fe^Ti oxides
(Table 9). This is in contrast to amphibole reaction rims
from Mount St. Helens (Rutherford & Hill, 1993) and
Soufrie're Hills (Buckley et al., 2006) in which the dominant
phase is clinopyroxene, followed by orthopyroxene, plagioclase and some oxides. Furthermore, using reaction rim
phase compositions, we were not able to reproduce the
target amphibole composition with similar observed mineral modes in the reaction rim or with low residuals. In
the El Misti case, host melt is required to balance the equation, acting as both an element supplier and element reservoir as observed in the case of the Soufrie're Hills
amphiboles (Buckley et al., 2006). The plagioclase-dominated mode in the reaction rims of the EM2000BP rhyolite
amphiboles is probably controlled by the breakdown of
the Al-rich amphibole pargasite; excess Al and Ca
from the decomposing amphibole may contribute to the
preferential growth of plagioclase and then pyroxene
respectively.

EM10 Andesite

10

1
La

Ce

Pr

Nd Sm Eu

Gd

Dy

Er

Yb

100

EM11 Rhyolite and Andesite

10

1
La

Ce

Pr

Nd Sm Eu

Gd

Dy

Er

Yb

Fig. 8. Chondrite-normalized REE element patterns illustrating the


similarity in trace element abundances between reacted (mostly rhyolite) and unreacted (mostly andesite) phenocrysts. Each frame represents an single hand sample with several LA-ICP-MS analysis points
within one or two amphibole phenocrysts from the listed sample,
regardless of the composition of the glass.

Mass-balance calculations were performed using a multiple linear regression least-squares mixing algorithm
coded in MATLAB (Dymond et al., 1973). The code uses
the chemical compositions of the reaction rim phases
(plagioclase, pyroxene, and Fe^Ti oxides) in oxide weight
per cent to calculate a modal best-fit solution to a target
composition (amphibole composition) with the lowest residuals. The algorithm in this code is similar to Petmix
(Wright & Doherty, 1970) used by both Rutherford & Hill
(1993) and Buckley et al. (2006), and reproduces solutions
to mineral proportions in amphibole reaction rims in
these studies accurately.
Rutherford & Hill (1993) noted that there was no combination of reaction rim phases in their calculation equivalent to the amphibole. To balance their equation and

Magma evolution and dynamics


The detailed petrological evidence, in particular the disparate plagioclase populations and the complex amphibole
provenance involving transfer from andesite to rhyolite
and reaction rim growth, provides a framework within
which we now attempt to piece together the magma dynamics that led to the 2000 BP eruption of El Misti Volcano.

The andesitic magma: phase equilibria constraints


The main crystallizing phases in the EM2000BP andesite
are amphibole and plagioclase with lesser amounts of
Fe^Ti oxides and pyroxenes. Notably, amphibole and, in
some cases, plagioclase contain small inclusions of Fe^Ti

18

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

However, Buckley et al. (2006) re-evaluated the Mount St.


Helens data and determined that using amphibole compositions close to the rim instead of an averaged amphibole
composition, the reaction equation can be written
La

Concentration/
Chondrite

MONTH 2013

reduce their residuals, Rutherford & Hill (1993) needed to


include melt compositions in the equation that took the
form

EM2 Rhyolite

Concentration/
Chondrite

NUMBER 0

EM 2

2024

632

198

540

381

147

108

Nd

Sm

Eu

Gd

Dy

Er

Yb

900

EM 10

111

169

358

412

182

442

170

455

286

119

064

EM 11

Nd

Sm

Eu

Gd

Dy

Er

Yb

Sample:

008

084

Yb

022

005

010

073

123

324

497

173

539

084

100

140

351

493

177

529

1911

326

1697

396

H-3

EM 11

008

012

018

030

006

048

093

024

172

005

005

022

024

004

043

027

014

035

008

1 SE

075

130

276

432

194

540

1825

353

1862

365

I-1

EM 10

122

159

421

588

222

770

2782

535

3675

1165

EM 2

065

087

226

377

141

397

1391

229

1282

258

I-1

EM 11

008

023

009

023

009

029

098

012

056

009

1 SE

013

004

009

020

009

007

030

035

016

028

011

1 SE

075

136

269

374

183

432

1771

314

1732

346

I-2

EM 10

098

317
114

012

447

183

538

1826

324

1776

356

I2-1

007

069

017

087

281

078

620

294

1SE

Sample values and standard errors are in mg g1.

381

142

Dy

Er

018

502

178

030

020

1812

Nd

Sm

512

010

296

Pr

Eu

047

Gd

013

380

1673

Ce

352

1940

1 SE

La

016

007

010

034

009

033

077

008

2014

EM 2
I3-3

008

009

033

012

005

015

047

011

034

014

1 SE

015

009

020

042

005

027

034

012

063

010

1SE

EM 2

055

111

291

370

147

448

1512

270

1407

288

I-2

EM 11

107

140

354

467

192

647

2095

388

2204

460

I-3

EM 10

110

121

372

504

176

621

2092

350

2001

428

I2-2

011

005

025

031

009

021

071

005

037

014

1 SE

008

009

033

012

005

015

047

011

034

014

1 SE

015

009

020

042

005

027

034

012

063

010

1SE

EM 2

077

131

362

536

181

591

2055

382

2056

459

I-3

EM 11

099

144

366

603

192

567

2178

375

2098

470

J(2)-1

EM 10

092

134

279

463

158

503

1939

352

2136

506

I2-3

014

010

031

042

007

037

036

017

038

007

1 SE

014

009

032

018

006

051

151

015

041

015

1 SE

007

008

021

026

014

047

043

017

081

017

1SE

EM 2

009

095

145

336

543

153

501

1853

331

1733

387

K(2)-1

EM 11

007

128
095

020

009

011

024

057

018

130

055

1 SE

014

008

016

019

002

034

035

014

070

016

1SE

287

460

183

547

1950

354

1977

480

J(2)-2

EM 10

138

133

341

564

181

589

2324

374

2060

456

I-1

EM 2

015

018

014

043

007

011

047

010

022

009

1 SE

075

079

227

351

137

400

1479

248

1459

287

K-1

EM 10

072

143

381

539

178

620

2016

362

1923

421

I-2

007

009

020

029

007

044

057

005

044

007

1 SE

013

007

023

033

007

029

056

019

069

017

1SE

EM 2

083

117

349

473

169

528

2029

353

1950

409

K-2

EM 10

102

148

343

420

172

525

1821

341

2145

663

I-3

009

009

016

016

009

025

055

007

033

021

1 SE

013

007

023

033

007

029

056

019

069

017

1SE

EM 10

115

168

417

664

230

822

2912

551

3287

676

K-3

EM 10

065

105

270

341

160

435

1602

263

1453

298

H-1

009

009

016

016

009

025

055

007

033

021

1 SE

011

008

016

022

006

039

038

006

037

007

1SE

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

19

H-2

285

1632

Pr

034

1 SE

015

011

016

068

016

143
026

513

025

241

104

1SE

2173

402

2822

495

333

1572

La

Ce

006

EM 2

I3-2

H-5

1 SE

006

010

024

026

005

061

036

011

063

020

1SE

H-4

EM 10

384

Pr

Sample:

429

1955

Ce

I3-1

La

Sample:

Table 7: REE LA-ICP-MS compositions of amphibole


EM 10

032
016
007
018

434
229
098
079

037
015

408
136

023
145

359

115

047

1 SE

010

007

017

046

007

007

044

005

028

022

1SE

1724

2487

828

K-4

EM 10

094

119

315

466

173

471

1845

293

1698

352

H-2

EM 10

084

161

294

527

153

555

1715

297

1534

348

H-1

EM 11

081

112

262

393

154

508

1660

279

1549

339

H-3

005

010

022

018

008

020

030

010

047

013

1 SE

010

007

017

046

007

007

044

005

028

022

1SE

TEPLEY et al.
EL MISTI 2000 BP ERUPTION

40

60

80

100

An Content

8
6
4
2
0
40

50

60

70

80

An Content

Number

Number

Microlites

Total # analyses = 63
Total # crystals = 35

10

9
2
8
7
6
5
4
3
2
1
0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

Relative probability

12

Relative probability

14

30

160
2
140
120
100
80
60
40
20
0
0.00 0.02 0.04 0.06 0.08 0.10 0.12

MgO wt%

16

20

MONTH 2013

MgO wt%

Fig. 9. Cumulative abundance plots of plagioclase phenocryst and microlite An content and corresponding MgO contents. The relative
frequency of occurrence of compositions and 2s errors for MgO are plotted.

pressures of 35 kbar. These experimental studies suggest


that the EM2000BP andesite magma reservoir was located
at 2^35 kbar pressure (7^12 km depth), 9408C
(calculated in this study) and was H2O saturated with
5^6 wt % H2O. We confirm these data using the plagioclase^liquid hygrometer of Lange et al. (2009), which
yields 5^6 wt % (H2O), and the amphibole thermobarometer of Ridolfi et al. (2010), which yields similar temperature
and water saturation values (Fig. 12). Our results for the
EM2000BP andesite are concordant with conditions of
crystallization of 900^9508C and 2^3 kbar pressure
(Legrende, 1999) under conditions of maximum water
solubility for andesite to dacite magmas of 51^60 wt %
(Ruprecht & Worner, 2007) for El Misti overall.

oxides revealing a paragenetic sequence. There are no relative crystallization clues between plagioclase and amphibole (i.e. no inclusions of plagioclase in amphibole or vice
versa) to indicate saturation order.
Experimental data for andesite and dacite phase equilibria reveal the effect of pressure and melt H2O content
(Eggler, 1972; Eggler & Burnham, 1973; Moore &
Carmichael, 1998; Martel et al., 1999), and fO2
(Rutherford & Devine, 1988; Martel et al., 1999) on amphibole stability. These studies show that crystallization of
plagioclase as the main liquidus phase at low pressure is
suppressed by increasing pH2O, and that the amphibole
stability field increases at high pH2O and temperature
and fO2  Ni^NiO 1 at the expense of clinopyroxene,
orthopyroxene and plagioclase (Moore & Carmichael,
1998; Martel et al., 1999). Water-saturated conditions also
change the plagioclase phase equilibria such that plagioclase compositions increase in An content with increasing
pH2O (Housh & Luhr, 1991; Lange et al., 2009). Phase equilibria for a starting material similar to the El Misti andesite indicate that hornblende is the first crystallizing phase,
followed closely by plagioclase and Fe^Ti oxides at pressures between 2 and 25 kbar at H2O-saturated conditions
(5^6 wt % H2O) and temperatures between 950 and
9758C (Moore & Carmichael, 1998). This temperature
range is similar to our two-pyroxene thermometer calculation for the EM2000BP andesite. A similar relationship is
found in the experiments of Martel et al. (1999) using silicic
andesites from Mount Pelee, although at somewhat higher

The andesite: constraints from amphibole compositions


Variations in formula cation abundances in the amphiboles
allow us to describe the pre-mixing thermal and pressure
history of the crystals and the andesite in which they
grew. Experimental studies by Spear (1981) and Blundy &
Holland (1990) found that variations in AlIV in amphibole
are strongly temperature dependent, expressed in the edenite exchange [SiIV A AlIV (Na K)A)], and the
Ti-Tschermak exchange (2SiIV MnVI 2AlIV TiVI)
(Fig. 6c), which is applicable as long as a Ti-rich phase
such as magnetite or ilmenite is present in the mineral assemblage (Spear, 1981). In EM2000BP amphiboles, the edenite exchange accounts for most of the total observable Al
variation (Fig. 6a), whereas Ti (p.f.u.) vs AlIV (p.f.u.)

20

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

20

Number

Number

160
140
120
100
80
60
40
20
0

NUMBER 0

Relative probability

Phenocrysts

VOLUME 0

Relative probability

Total # analyses = 1850


Total # crystals = 47

JOURNAL OF PETROLOGY

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

(a)

An Content

Low An Content Plagioclase

500 um

An Content

Distance (m)

200 um

Distance (m)
Fig. 10. BSE images, An transects, and corresponding MgO (open circles) and FeO (filled squares) concentration profiles for representative
Low-An (a) and High-An group (b) plagioclase phenocrysts from EM2000BP tephra. White lines on plagioclase (BSE images) represent transect locations. Dashed black line represents average limit of detection for MgO. Also plotted are 1s errors for MgO (open circle) and FeO
(filled square).
(continued)

21

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

fluctuations during crystal growth, possibly as a result of


convective rotation in a small magma body or as a result
of repeated small recharge events of similar composition
into a small magma body.
This model is consistent with the zoning characteristics
of two amphibole crystals (Fig. 6) that show large variation
in AlIV, Mg# and other geochemical parameters.
Amphibole 10H and 10K, a non-rimmed and a rimmed
amphibole both in mixed andesite 10, respectively, together
account for the same range in (Na K)A, AlVI, and
Mg# vs AlIV as the complete amphibole dataset. Their
variation in AlIV also covers the full range exhibited by
the complete dataset. This type of zoning, increasing AlIV
and decreasing Mg# followed by decreasing AlIV and
increasing Mg#, has been shown in experimental studies

shows a slightly positive correlation, indicative of the TiTschermak exchange. Further support for temperature
control on the EM2000BP amphibole compositions is seen
in the trend of decreasing Mg# (05 Mg# values) and
increasing AlIV (04 AlIV p.f.u.) (Fig. 6a and d), consistent with the work of Rutherford & Devine (2003). Finally,
our data show no increase in AlVI with any other geochemical indicator obviating a role for the Al-Tschermak
exchange (2SiIV MgVI 2AlIV AlVI) favored by
increasing pressure (Johnson & Rutherford, 1989; Thomas
& Ernst, 1990; Schmidt, 1992).
Thus, on the basis of previously published experimental
results and our observed mineral chemistry variations, we
suggest that the EM2000BP amphiboles crystallized in a
near isobaric environment with modest temperature

JOURNAL OF PETROLOGY

(b)

VOLUME 0

NUMBER 0

MONTH 2013

High An Content Plagioclase

200 um

Distance (m)

500 um

Distance (m)
Fig. 10. (Continued)

and that zoning and changes in amphibole composition


were the result of crystal growth in a melt of fluctuating
temperature. The amphibole adjusted its composition as
the temperature fluctuated, utilizing whatever exchange
mechanism necessary to maintain chemical and thermal
equilibrium with the surrounding melt. Plagioclase data
from the andesite (and rhyolite) support these conclusions
as trace element partitioning between plagioclase and melt
generally falls along equilibrium partitioning curves
(Fig. 11). Although changing pH2O may play a role in the
changing An contents, it did not play a major role in amphibole crystallization and zoning.

to represent crystallization in a hotter, more Al-rich


magma followed by crystallization in cooler, more
Al-poor magma (Scaillet & Evans, 1999; Rutherford &
Devine, 2003).
Similar zoning has been interpreted by Humphreys et al.
(2006) to be the result of changes in pH2O and its associated
effect on plagioclase composition. In effect, increasing the
pH2O of the melt at constant temperature promotes crystallization of higher An plagioclase and Al-poor amphibole.
In the study by Humphreys et al. (2006), this mechanism of
amphibole^plagioclase compositional exchange is mirrored
in the edenite exchange in (Na K)A vs AlIV (their fig. 12c
and d). In the EM2000BP case, (Na K)A shows only
small variation with AlIV (Fig. 6). Additionally, the plagioclase exchange (SiIV NaA AlIV CaA) (not illustrated)
also shows no variation over the full range of AlIV variations. Both of these observations suggest that plagioclase
crystallization did not affect the amphibole compositions

The rhyolite: bulk geochemistry and phase equilibria


constraints
Above we showed that whole-rock major element variations follow similar evolutionary trends to previously
published data from various eruptions in El Mistis past

22

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

Table 8: Representative compositions of Low- and High-An plagioclase phenocrysts and microlites
Low-An plagioclase phenocrysts
Sample:

EM 2 B

EM 2 B

EM 5 I

EM 5 I

EM 5 I

EM 10 G

EM 10 G

EM 10 G

EM 11 E

EM 11 E

EM 11 E

Inter

Rim

Core

Inter

Rim

Core

Inter

Rim

Core

Inter

Rim

SiO2

5747

5562

5700

5511

5376

5891

5655

5498

5969

5428

5290

5620

Al2O3

2661

2792

2720

2848

2887

2570

2749

2827

2582

2859

2960

2753

FeO*

036

033

040

028

029

056

029

031

035

040

044

043

MgO

002

002

001

001

001

005

001

001

001

003

002

002

CaO

871

1011

907

1044

1131

795

936

1041

712

1085

1211

970

Na2O

582

503

526

484

458

521

538

491

621

469

396

467
043

K2O

058

042

051

037

032

057

050

037

077

035

027

TiO2

003

004

002

002

001

005

002

001

002

002

002

002

Total

9959

9950

9948

9954

9914

9901

9960

9926

9998

9921

9932

9902

An content

44

51

47

53

57

44

48

53

37

55

62

52

High-An plagioclase phenocrysts


Sample:

EM 2 D

EM 2 D

EM 2 D

EM 5 H

EM 5 H

EM 5 H

EM 10 E

EM 10 E

EM 10 E

EM 11 C

EM 11 C

EM 11 C

Core

Inter

Rim

Core

Inter

Rim

Core

Inter

Rim

Core

Inter

Rim

SiO2

4764

4904

5471

4665

5077

5364

4599

4816

5002

4786

4829

5082

Al2O3

3314

3235

2816

3403

3094

2972

3394

3225

3128

3276

3278

3139

FeO

055

062

053

054

064

058

059

057

058

057

057

058

MgO

004

006

006

004

009

007

003

005

006

005

005

007

CaO

1655

1534

1071

1707

1384

1186

1734

1547

1412

1585

1569

1401

Na2O

198

251

483

160

325

433

148

241

306

220

240

314

K2O

006

010

031

005

015

020

004

008

011

007

008

011

TiO2

003

003

003

001

003

003

002

002

003

002

002

003

Total

9999

10005

9933

9999

9971

10043

9942

9902

9927

9937

9987

10015

An content

82

54

85

70

86

78

71

80

78

77

59

71

Low- and High-An plagioclase microlites


Sample:

EM 2

EM 2

EM 2

SiO2

5686

5968

6100

Al2O3

2636

2456

2400

FeO

040

025

MgO

001

CaO

842

EM 5

EM 5

EM 5

EM 10

EM 10

EM 10

EM 11

EM 11

EM 11

5464

5485

5427

5411

5450

5875

5286

5301

5345

2869

2767

2810

2821

2779

2492

2932

2944

2909

031

066

075

069

078

079

136

063

065

060

001

000

007

010

008

007

010

014

008

006

008

622

534

1078

1033

1070

1097

1050

800

1209

1198

1175

Na2O

581

662

694

496

507

481

471

493

493

430

428

447

K2O

063

094

132

028

032

030

030

036

101

019

020

026

TiO2

000

002

003

004

005

004

005

006

018

004

003

004

Total

9859

9832

9896

10014

9915

9899

9923

9907

9935

9952

9967

9974

An content

43

32

27

52

54

55

53

44

60

60

58

54

*Total Fe given as FeO.


Typical 1SD: SiO2  014; Al2O3  015; FeO*  003; MgO  001; CaO  01; Na2O  005; K2O  0015; TiO2  001.

23

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

EM 2 B
Core

JOURNAL OF PETROLOGY

VOLUME 0

80

Phenocrysts
Microlites
Hi T eq
Lo T eq
2

70
60
50
40
30
20
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

MgO wt%
100

detection limit

(b)

90

An Content

80

Phenocrysts
Microlites
Hi T eq
Lo T eq
2

70
60
50
40
30
20
0.00

0.02

0.04

0.06

0.08

0.10

TiO2 wt%
Fig. 11. Trace-element [MgO (a), TiO2 (b), FeO*(c)] variations vs An content in plagioclase from EM200BP tephra. Plotted are phenocryst
(open symbols) and microlite (shaded symbols) values. Also plotted are equilibrium partitioning curves for andesite (Hi T eq) and rhyolite
(Lo Teq) and partitioning curve uncertainties based on the partitioning behavior of these elements into plagioclase as reported by Bindeman
et al. (1998) and Tepley et al. (2010). Values for each element used to determine the equilibrium partitioning curves are whole-rock values of the
andesite (EM0401) and the rhyolite (EM099), as an estimate of the melt composition, and T is determined through Fe^Ti oxide (rhyolite;
8168C) and two-pyroxene (andesite; 9408C) geothermometry. Limit of detection for each trace element is depicted as a gray dashed line. Also
plotted are 2s errors.
(continued)

24

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

detection limit

100
90

An Content

MONTH 2013

and thus a case can be made that the rhyolite endmember is related to the andesite magma by crystal
fractionation of plagioclase, amphibole, pyroxene and
magnetite. However, we have no definitive evidence that

and within the Central Volcanic Zone as a whole (Fig. 3).


Although the rhyolite is related to the system as a whole,
its petrogenesis is not wholly understood. For El Misti as
a whole, all compositions fall on a liquid line of descent

(a)

NUMBER 0

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

(c) 100

Phenocrysts
Microlites
Hi T eq
Lo T eq

90

70
60
detection limit

50
40
30
20
0.0

0.2

0.4

0.6

0.8

1.0

FeO wt%
Fig. 11. (Continued)

the EM2000BP rhyolite was derived from the EM2000BP


andesite. Whatever its exact relation to the EM2000BP andesite, the rhyolite magma must have separated from an
El Misti andesite magma sometime in the past and staged
at shallower levels in the crust as required by its separate
plagioclase phenocryst population, the lack of cognate
amphibole, its lower temperatures of equilibrium, and its
evolved residual liquid composition. This magma sat in
the upper crust fractionating plagioclase and Fe^Ti
oxides, stagnated and partially solidified. It seems that an
injection of andesitic magma reactivated the rhyolite on a
local level, mingled with it intimately, and a later recharge
induced it to erupt explosively.
The main crystallizing phases in the rhyolite are plagioclase, Fe^Ti oxides and pyroxene, which are similar to
those crystallizing in the rhyolitic component of a zoned
eruption deposit from the 1912 eruption at Novarupta,
Alaska (Hildreth, 1983). Phase equilibria experimental results from the Novarupta rhyolite (Coombs & Gardner,
2001) demonstrate that at temperatures and pressures similar to the El Misti system (T 8168C, P5100 MPa), a
similar mineral assemblage was produced. The experiments also show that amphibole is not on the liquidus
5100 MPa in pressure despite being water saturated.
Plagioclase^melt equilibria (plagioclase rims^adjacent
glass compositions) indicate that the rhyolite was water
saturated with 5 wt % H2O (Lange et al., 2009). This
probably represents the conditions of the rhyolite when it
first formed and not just before eruption, as otherwise it
would have been water-saturated with the potential to
erupt through crystallization-driven overpressure. The

EM2000BP rhyolite must then represent a degassed remnant of some prior episode in El Mistis past, or it had passively degassed. In either case, based on phase equilibria
experiments and plagioclase^melt equilibria, the lack of
cognate amphibole requires that the magma was stored
above the stability limit of amphibole (100 MPa or
53 km at c. 816  308C). This is consistent with the correspondence of the EM2000BP rhyolitic glass to the 05^1
kbar granite ternary eutectic in Petrogenys Residua
System (Tuttle & Bowen, 1958).

Time scales of magma dynamics during EM2000BP:


chemical equilibration and amphibole rim development
Three observations can be used to provide time scales for
the magmatic interactions preceding and during the
EM2000BP eruption of El Misti volcano: (1) the lack of
pervasive chemical equilibration despite minimal diffusive
length scales; (2) the absence of reaction rims on amphibole in the andesite; (3) the presence of reaction rims on
the andesite-originated amphibole mixed into the rhyolite.
The lack of chemical equilibration of andesite and rhyolite melt despite intimate mixing of the two magmas supports very short time scales of interaction prior to
eruption. Macroscopic and microscopic textural evidence
shows that mingling on the crystal scale and significant
folding and stretching of the magmas occurred. The equilibration process is dependent on the diffusivities of major
elements, which are of the order 10^12 m2 s1 (Liang et al.,
1996). Using t  x2/D, for length scales of 1mm to 1cm,
which would be necessary for pervasive equilibration,
chemical equilibration would be reached in the order of

25

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

An Content

80

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

MONTH 2013

1000

Table 9: Amphibole reaction rim phase proportions, and


comparison with other studies
plagioclase

pyroxene*

oxidesy

EM 2 I

68

25

EM 2 I3

64

26

64

30

EM 5 K

61

36

EM 10 K

55

40

Soufrie`re Hills (B)

182

774

45

Soufrie`re Hills (B)

181

783

36

MSH (B)z

237

675

87

MSH (R&H)

43

53

maximum thermal
stability limit

600

400

200

0
700

800

900

1000

1100

T (C)
1150

T (C)

1050

*Clinopyroxene and orthopyroxene are combined as


pyroxene.
yIlmenite and magnetite are combined as oxides.
zRecalculation of Rutherford & Hill (1993) amphibole reaction rim mineral modes of Buckley et al. (2006) using the
method described by those researchers.
Original calculation of reaction rim mineral modes of
Rutherford & Hill (1993).
B, Buckley et al. (2006); R&H, Rutherford & Hill (1993);
MSH, Mount St. Helens amphibole.

950
850

maximum thermal
stability limit

750

EM rhyolite
EM andesite

650
550
2

10

H2Omelt (wt.%)
Fig. 12. P (MPa) vsT (8C), and T (8C) vs H2Omelt (wt %) based on
the amphibole reduction of Ridolfi et al. (2010), showing the coherence
of data from amphiboles residing in both the rhyolite and andesite.
This figure also illustrates T and P of formation for the amphibole,
and the water content in the andesitic melt.

10 days to 1^2 years over those length scales. Meanwhile,


thermal equilibration (10^6 m2 s1) and cooling would
have been achieved orders of magnitude fasterin a
matter of seconds to minutes. Thus, because very limited
hybridization seems to have taken place, the process must
have been arrested rather rapidly, in the order of days.
Although not meant to be definitive, the point here is that
even though diffusive length scales were minimal, chemical equilibration is too slow to compete with freezing of
the system. The interaction between the rhyolite and
andesite could not have been prolonged and had to have
happened just prior to eruption.
The lack of reaction rims on amphibole in the
EM2000BP andesite allows us to assign an upper time
limit on the mixing process during this eruption. Multistep decompression experiments for starting materials
from the 1989 eruption of Redoubt volcano, Alaska, USA,
by Browne & Gardner (2006) suggest that as few as 4
days and as many as 7 days elapsed before reaction rims
developed on amphiboles that were moved outside their
stability range, depending on the rate of decompression. A
similar time frame was demonstrated for the 1980 eruption
of Mount St. Helens volcano, Washington, USA, in the
decompression experiments of Rutherford & Hill (1993).
Therefore, for the EM2000BP andesite, we suggest a

conservative time frame of 5 days for magma migration


from the storage reservoir to the surface.
Reaction rim development on andesite-originated
amphiboles in the EM2000BP rhyolite is controlled by diffusion of constituent material from the melt to the crystal
surface and diffusion of unused elements away from the
crystal surface into the melt (Liang, 2000; Coombs &
Gardner, 2004). Whereas the rate of the reaction will
depend on the temperature and pressure of the system
and the amount of dissolved water in the melt, the rates of
crystal decomposition and crystal rim growth should be
governed by the diffusive exchange of amphibole rim material and the surrounding melt. In a study by Browne &
Gardner (2006), amphibole decompression rim growth
was determined for various systems at a range of temperatures. Those researchers found that reaction rim growth
rate is probably related to melt viscosity and associated
temperature, and that it would take 50^60 days as a minimum for the reaction rims on amphiboles to develop.
Given that the EM2000BP rhyolite has a similar temperature to the 1989 Mount Redoubt dacite (8168C
and 8408C, respectively), we infer a similar time scale

26

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

EM 2 I2

P (MPa)

Sample

EM rhyolite
EM andesite

800

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

of 50^60 days for the development of rims on the amphiboles in the EM2000BP rhyolite.

Microlite crystallization

Magma dynamics and physical model of the EM2000BP


eruption
Based on the above, a petrological model for the El Misti
2000 BP eruption can now be proposed (Fig. 13). Two separate, compositionally distinct magma reservoirs existed beneath El Misti, of andesite and rhyolite composition, each
with their own sets of phenocryst and microphenocryst
populations. The andesite reservoir was located at
7^12 km depth in the crust (200^350 MPa) and the
magma within it was at 9408C and was water saturated
(5^6 wt % H2O). Although sparse, the crystallizing
phases were amphibole, plagioclase, Fe^Ti oxides and pyroxenes. Conditions of crystallization in the reservoir were
relatively constant with minor perturbations in temperature perhaps reflecting small-scale convection currents or
small-volume recharge of similar composition magma.
The rhyolite reservoir is less well constrained but it appears that the only crystallizing phases were plagioclase
and Fe^Ti oxides. The rhyolite reservoir was located at
3 km depth in the crust (5100 MPa) and the magma
within was at 816  308C and either degassed or partially
degassed. The lack of stable hydrous phases and the temperature of the system also suggest low-pressure
(5100 MPa), shallow crustal residence (3 km depth).
We envision the development of the eruption as a twostage process (Fig. 13a). The first stage initiated as a dike
of andesitic magma intruding into the rhyolite magma.
The andesite was water-saturated and would have been
vesiculating as it rose and, on encountering the rhyolite, it
may have vigorously mixed with the latter. Exchange of
minerals occurred between the two magmas during this
intrusion event, but was limited by the large temperature
and viscosity contrasts. Given that the rhyolite resides in

27

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

Plagioclase microlites dominate the groundmass mineral


assemblage and constitute a significant proportion of the
total crystal fraction in both lithologies. Microlite composition histograms show two populations of microlites
(Fig. 9), the compositions of which correlate with the host
magma composition and are lower in An than the phenocryst compositions in the respective host magmas. This
suggests that the microlites grew at lower pH2O during
magma ascent (e.g. Geschwind & Rutherford, 1995).
Moreover, the microlites are normally zoned and texturally display elongate, tabular and swallowtail morphologies that indicate rapid crystallization. This suggests that
the plagioclase microlites crystallized during ascent in
their respective host magmas rather than in a shallow
holding chamber such as in the 1980^1986 eruptions of
Mount St. Helens volcano (e.g. Geschwind & Rutherford,
1995).

the upper crust above the stability limit of amphibole, it is


at this stage that amphiboles and plagioclase from the andesite are mixed into the rhyolite along a thin interactive
zone, and the amphibole, now out of its stability field,
develops reaction rims over a period of 50^60 days.
Eruption was precluded by the relatively small volume of
the recharging magma with respect to the host rhyolite
magma.
Around 50^60 days after the initial recharge event,
spurred on either via continued recharge or by simple
buoyant rise, another larger pulse of vesiculating andesite
magma forced its way through the earlier stalled
recharged zone in the perched rhyolite and initiated the
eruption. In this second stage, further limited crystal exchange may have occurred at the margins of the andesitic
dike or eruption conduit, with amphiboles grown in the
andesite being added to the rhyolite and plagioclase from
each lithology being mutually exchanged.
The time for the EM2000BP second pulse of andesite to
reach the rhyolite at 3 km depth (above the amphibole
stability limit) is indeterminate, but the absence of reaction
rims on the amphiboles in the andesite recharge magma
requires that it travelled from its storage reservoir to the
surface within 5 days. This yields an average ascent rate
of at least 0023 m s1.
The presence of abundant microlites in the EM2000BP
rhyolitic and andesitic groundmasses is consistent with
such ascent rates from low pressures (m s1 and
5100 MPa; e.g. Klug & Cashman, 1994; Metrich &
Rutherford, 1998; Cashman & Blundy, 2000; Martel &
Schmidt, 2003). The formation of plagioclase microlites
was most probably driven by magma undercooling owing
to exsolution of volatiles associated with decompression
(e.g. Muncill & Lasaga, 1988; Hammer, 2008). The loss of
dissolved volatiles has the effect of increasing the relative
liquidus temperature of the magma, thereby decreasing
the An content of any crystallizing plagioclase. The lower
An contents of the rims on both High-An and Low-An
plagioclase populations, when compared with their core
An contents, attest to this process. This process similarly
affects andesite and rhyolite microlite compositions, which
in El Mistis case, show an overall reduction in An content
in comparison with their respective plagioclase phenocryst
populations (Figs 9 and 10).
Given the relatively small volume of the EM2000BP
tephra deposits, we prefer a model in which the magmatic
interactions take place along and within a dike of watersaturated andesite. Mixing between the hot (9408C) andesitic magma and a cooler rhyolitic (8168C) magma will
initially be limited, given the temperature and viscosity
contrasts between them (Huppert et al., 1982; Campbell &
Turner, 1985; Sparks & Marshall, 1986; Turner & Campbell, 1986; Snyder & Tait, 1995); however, shearing along
the edges and progressive physical mixing will allow

JOURNAL OF PETROLOGY

(a)

VOLUME 0

NUMBER 0

MONTH 2013

Stage 2

Stage 1

Eruption

rhyolite crystal mush reservoir

Duration of stagnation:
50-60 days

hbl stability limit (in rhy)

lnitial dike emplacement


and stagnation

hbl stability limit (in and)

Duration of event:
<5 days

Second more forceful


dike emplacement
of recharge magma

andesite reservoir

(b)
eruption

Stage 2

Stage 1

~5 days

e1d
>200 MPa

amphibole stability
(in andesite)

max

max

Stage 2 dike emplacement

ceme

amphibole stability
(in rhyolite)

ike e

mpla

100 MPa

Stag

Pressure

nt

50-60 days

min
max

min
max

Vm/Vf

Vm/Vf
min
max

min
conduit
cross section

min
max

min
conduit
cross section

Time
Fig. 13. Petrogenetic model illustrating andesite reservoir location and perched rhyolite magma lens. (a) Schematic model and simplified development of the 2000 BP eruption of El Misti. Diagram illustrates the initial conditions of each reservoir in relationship to the amphibole stability
limit. Stage 1 is initiated with dike emplacement into and stagnation in the existing rhyolitic mush. Limited magma mixing occurs during this
stage, resulting in mixed crystal populations and development of reaction rims on amphibole. Stage 2 occurs when a stronger recharge pulse reactivates the emplaced dike, causing more magma mixing, mixed crystal populations and eruption. (b) Right panel illustrates the detail associated with mixed magma and crystal exchange. Included are box models schematically illustrating the interaction between andesite and
rhyolite at the initial contact deep in the system (left), and interactions in the conduit during eruption (right). The box models offer acrossdike and conduit qualitative assessments associated with variations in temperature (T), volume of mafic to felsic magma (Vm/Vf) and viscosity
(m). These gradients are steep at the initial contact deep in the system and become more gentle higher in the system with continued shearing
and diffusion of material and heat. The left panel illustrates the pressure^time relationship showing the time scales of eruption based on amphibole stability. Elongate dike represents the initial intrusion of andesite into the rhyolite magma reservoir (Stage 1). This magma resides
above the amphibole stability limit for 50^60 days, before being recharged by another pulse of andesite magma, which initiates evacuation
and eruption in the time frame of 5 days (Stage 2). Diagram is not to scale.

28

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

More in-conduit
magma mixing
and exchange
of crystals

Limited magma mixing


and exchange of crystals

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

Blundy & Cashman, 2001). In the case of the 2000 BP eruption of El Misti, all these processes probably conspired to
cause the explosive eruption, but the fundamental trigger
for the eruption was andesite recharge.

thermal equilibration to commence rapidly (of the order of


seconds to minutes for length scales of millimeters to centimeters respectively), reducing the viscosity contrast
between the two magmas (e.g. Ruprecht et al., 2012;
Fig. 13b). With time (and decreasing depth of the rising
magma in the dike), this will promote magma mingling.
Shear along the interface between the magmas might aid
mixing of liquid and crystals. As seen in the ejecta, interaction between the two magma compositions takes many
forms, from thick toothpaste-like globs of rhyolite in a
matrix of andesite, to thin wisps of alternating andesite
and rhyolite (e.g. Fig. 2), to intermixing melt fractions and
crystal population transfers. We see this as reflecting an
evolving gradient in mingling in space and time from the
edges to the center of the dike, with more intimate mingling the further upwards it travels (Fig. 13b).
Texturally, vesicle shapes and glass distortion provide
evidence for the viscosity differences between the rhyolite
and the andesite. Bubble coalescence, merging of smaller
bubbles into larger bubbles, can be seen in most of the
rhyolite thin sections [donut-like features of Klug et al.
(2002)]. As in Mazama pumices (Klug et al., 2002), interaction between equal-sized bubbles often results in very
thin, planar melt films (1 mm), inferred to be caused by
approximately equal pressures acting on the film from
inside each bubble. These textures suggest significant
shear within the rhyolitic melt that may have been buffered
in the andesite by its lower viscosity. The combined effects
of groundmass crystallization and loss of volatiles from
the melt lead to increased magma and melt viscosity, but
combine with vesiculation to increase the potential for explosive eruption, overcoming the viscous death described
by Annen et al. (2006).
Magma recharge and associated mixing with a preexisting magma is often cited as a triggering mechanism
for volcanic eruptions (Sparks et al., 1977; Eichelberger,
1978; Huppert et al., 1982; Pallister et al., 1992; Suzuki &
Nakada, 2007; de Silva et al., 2008; Kent et al., 2010). This
may be due to a simple hydraulic pressure increase induced
by addition of mass to a magma reservoir (e.g. Blake, 1981,
1984), by exsolution of volatiles from a resident felsic
magma induced through superheating owing to recharge
by hot mafic magma (Sparks et al., 1977), or by cooling of
the more mafic recharge magma forcing saturation and
vapor phase exsolution (e.g. Huppert et al., 1982; Tait et al.,
1989; Pallister et al., 1992; Folch & Marti, 1998). In all
these cases, over-pressurization of small magma chambers
beyond the tensile strength of the wall-rocks is thought to
be the trigger for explosive eruptions (see Gregg et al.,
2012). Alternatively, volatile exsolution and syn-eruptive
crystallization driven by depressurization during adiabatic
rise of magma to the surface can drive explosive eruptions
(Geschwind & Rutherford, 1995; Hammer et al., 1999;
Nakada & Motomura, 1999; Cashman & Blundy, 2000;

The magmatic architecture at El Misti

CONC LUSIONS
The architecture, dynamics, and time scales of andesite^
rhyolite interaction during the 2000 BP, VEI 5 eruption of
El Misti in southern Peru have been revealed through
detailed petrological study. Bulk-rock chemistry, mineral
textures and compositions reveal macroscopic and microscopic evidence for magma mingling and crystal exchange
that record how an initial dike tapping a deep (7^12 km),
hot, water-saturated andesite magma reservoir intruded

29

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

Lastly, we consider the size and nature of the El Misti


magmatic system over the lifetime of the volcano.
Ruprecht & Worner (2007) concluded that a single, large,
often-recharged magma reservoir existed below El Misti
rather than a plexus of smaller, interconnected magma reservoirs and dikes. Given that the eruption history of El
Misti is one dominated by effusive edifice-building andesite
and dacite domes and flows, this interpretation is plausible.
However, the 2000 BP eruption of El Misti was an explosive
eruption precipitated by a recharge event(s) of andesite
into rhyolite. Rhyolite at El Misti is rare and is found only
in the explosive eruptions that punctuate its effusive-dominated eruption history on a time scale of 2000^4000 years
(e.g. Thouret et al., 2001). The similarity of the EM2000BP
juvenile material in composition and magmatic conditions
and in physical appearance to those from the other (less
studied) explosive events during the history of El Misti
allows a model for the explosive events to be presented.
We modify the model of Ruprecht & Worner (2007) to include periods when a small rhyolitic reservoir develops at
shallower levels in the crust. Recharge by andesite results
in an explosive eruptiona fast and transient event in the
history of El Misti. We suggest that these events do not tap
a large single reservoir but perhaps represent the interaction between a deeper long-lived andesitic reservoir and
a small transient shallow rhyolitic magma that may form
cyclically on a 2000^4000 years time scale. This time
frame may represent the period required to produce the
rhyolite from the andesite and segregate it to a high level
in the plumbing system shortly before eruption.
Interaction does not have to be chamber wide, but more
probably occurs along a dike that penetrates, interacts locally and erupts at the surface. If, as in the case of the
2000 BP eruption, the explosive eruption coincides with
periods when El Misti has significant snow cover, the inevitable ash-fall hazard would be magnified by the triggering
of extensive lahars by small pyroclastic flows (Harpel
et al., 2011).

JOURNAL OF PETROLOGY

VOLUME 0

into a cooler, dryer, shallow (3 km) rhyolitic magma


higher in the crust and stalled. During the initial intrusion,
limited exchange of crystals from the two magmas
occurred. Amphibole crystals grown in the andesite
magma were transported into a cooler, shallower, and
chemically different environment where over a period of
at least 50^60 days they decompressed in both the rhyolitic
and andesitic melt to form plagioclase-dominated reaction
rims. A subsequent recharge via an andesitic dike remobilized the small magma storage system and resulted in extensive magma mingling and crystal exchange at a variety
of scales with mingling diminishing away from the andesite dike^rhyolite magma interface. Explosive eruption of
pervasively to minimally banded pumice reveals that although decompression crystallization of plagioclase microlites occurred, there was no wholesale equilibration of
melt and no reaction rims developed on amphiboles in the
andesite from the second recharge event. These observations require that during this latter stage, transport of andesite magma above the amphibole stability zone,
interaction with the rhyolite, and eruption all happened
within a period of 5 days at an average ascent rate of
002 m s1.
The 2000 BP VEI 5 plinian eruption shares characteristics with other explosive events that punctuate the background effusive activity at El Misti with a period of
2000^4000 years. It may therefore serve as a model for explosive events at this hazardous volcano. Our model for
the El Misti system includes the interaction of a deeper
and larger body of andesitic magma with a small rhyolitic
reservoir resulting in cyclic explosive eruptions. The periodicity may represent the time scales required for rhyolite
development, rapid andesite recharge and eruption.

NUMBER 0

MONTH 2013

a visiting scientist supported by Oregon State University,


Department of Geosciences.

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

Albare'de, F. & Bottinga, Y. (1972). Kinetic disequilibrium in trace


element partitioning between phenocrysts and host lava.
Geochimica et Cosmochimica Acta 36, 141^156.
Annen, C., Blundy, J. D. & Sparks, R. S. J. (2006). The genesis of intermediate and silicic magmas in deep crustal hot zones. Journal of
Petrology 47, 937^955.
Bacon, C. R. & Hirschmann, M. M. (1988). Mg/Mn partitioning as a
test of equilibrium between coexisting Fe^Ti oxides. American
Mineralogist 73, 57^61.
Bindeman, I. N., Davis, A. M. & Drake, M. J. (1998). Ion microprobe
study of plagioclase^basalt partition experiments at natural concentration levels of trace elements. Geochimica et Cosmochimica Acta
62, 1175^1193.
Blake, S. (1981). Volcanism and the dynamics of open magma chambers. Nature 289, 783^785.
Blake, S. (1984). Volatile oversaturation during the evolution of silicic
magma chambers as an eruption trigger. Journal of Geophysical
Research 89, 8237^8244.
Blundy, J. & Cashman, K. (2001). Ascent-driven crystallization of
dacite magmas at Mount St. Helens, 1980^1986. Contributions to
Mineralogy and Petrology 140, 631^650.
Blundy, J. & Wood, B. (1994). Prediction of crystal^melt partition coefficients from elastic moduli. Nature 372, 452^454.
Blundy, J. D. & Holland, T. J. (1990). Calcic amphibole equilibria and
a new amphibole^plagioclase geothermometer. Contributions to
Mineralogy and Petrology 104, 208^224.
Bowen, N. L. (1928). The Evolution of the Igneous Rocks. Princeton, NJ:
Princeton University Press.
Browne, B. L. & Gardner, J. E. (2006). The influence of magma ascent
path on the texture, mineralogy, and formation of hornblende reaction rims. Earth and Planetary Science Letters 246, 161^176.
Buckley, V. J. E., Sparks, R. S. J. & Wood, B. J. (2006). Hornblende dehydration reactions during magma ascent at Soufrie're Hills
Volcano, Montserrat. Contributions to Mineralogy and Petrology 151,
121^140.
Bullard, F. M. (1962). Volcanoes of southern Peru. Bulletin of Volcanology
24, 443^453.
Campbell, I. H. & Turner, J. S. (1985). Turbulent mixing between
fluids with different viscosities. Nature 313, 39^42.
Cashman, K. & Blundy, J. (2000). Degassing and crystallization of ascending andesite and dacite. Philosophical Transactions of the Royal
Society of London, Series A 358, 1487^1513.
Coombs, M. L. & Gardner, J. E. (2001). Shallow-storage conditions for
the rhyolite of the 1912 eruption at Novarupta, Alaska. Geology 29,
775^778.
Coombs, M. L. & Gardner, J. E. (2004). Reaction rim growth on olivine in silicic melts: Implications for magma mixing. American
Mineralogist 89, 748^759.
Davidson, J. P. & de Silva, S. L. (2000). Composite volcanoes. In:
Sigurdsson., H. (ed.) Encyclopedia of Volcanoes. London: Academic
Press, pp. 663^681.

AC K N O W L E D G E M E N T S
J. Permenter, C. Harpel, W. Scott, B. Anders, Y. Lavallee
and J. Burns, as well as Ms. C. Harpel-Avendano and
other students from UNSA, were helpful during various
fieldwork sessions at El Misti when these samples were collected. We thank S. Marcott for help with MATLAB
code, H. Diettrich for help with EMPA work, and C.
Bouvet de la Maisonneuve for an internal review of the
paper. A. Allan, P. Ruprecht, M. Rutherford, and M.
Streck provided very thorough and constructive reviews
that are appreciated. These, and G. Worners editorial
handling of the paper, have helped clarify and strengthen
our ideas.

F U N DI NG
This work has been variously supported by the National
Science Foundation (EAR 0087181 to S.d.S.) and the
Volcano Disaster Assistance Program (VDAP) of the US
Geological Survey. This work was initiated when G.S. was

30

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

R E F E R E NC E S

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

Housh, T. B. & Luhr, J. F. (1991). Plagioclase^melt equilibria in


hydrous systems. American Mineralogist 76, 477^492.
Humphreys, M. C. S., Blundy, J. D. & Sparks, R. S. J. (2006). Magma
evolution and open-system processes at Shiveluch volcano: insights
from phenocryst zoning. Journal of Petrology 47, 2303^2334.
Huppert, H. E., Sparks, R. S. J. & Turner, J. S. (1982). Effects of volatiles on mixing in calc-alkaline magma systems. Nature 297,
554^557.
Johnson, D. M., Hooper, P. R. & Conrey, R. M. (1999). XRF analysis
of rocks and minerals for major and trace elements on a single low
dilution Li-tetraborate fused bead. Advances in X-ray Analysis 41,
843^867.
Johnson, M. C. & Rutherford, M. J. (1989). Experimental calibration
of the aluminum-in-hornblende geobarometer with application to
Long Valley caldera (California) volcanic rocks. Geology 17, 837^841.
Kent, A. J. R., Darr, C., Koleszar, A. M., Salisbury, M. J. &
Cooper, K. M. (2010). Preferential eruption of andesitic magmas
through recharge filtering. Nature Geoscience 3, 631^636.
Kent, A. J. R., Stolper, E. M., Francis, D., Woodhead, J., Frei, R. &
Eiler, J. (2004). Mantle heterogeneity during the formation of
the North Atlantic Igneous Province: Constraints from trace
element and Sr^Nd^Os^O isotope systematics of Baffin Island
picrites. Geochemistry, Geophysics, Geosystems 5, Q11004, doi:10.1029/
2004GC000743.
Klug, C. & Cashman, K.V. (1994).Vesiculation of May 18,1980 Mount
St. Helens magma. Geology 22, 468^472.
Klug, C., Cashman, K. V. & Bacon, C. R. (2002). Structure and physical characterization of pumice from the climactic eruption of Mt.
Mazama (Crater Lake), Oregon. Bulletin of Volcanology 64, 486^501.
Knaack, C., Cornelius, S. & Hooper, P. R. (1994). Trace element analysis of rocks and minerals by ICP-MS. Open File Report.
Pulman, WA: Department of Geology, Washington State
University, 18 p.
Lange, R. A., Frey, H. M. & Hector, J. (2009). A thermodynamic
model for the plagioclase^liquid hygrometer/thermometer.
American Mineralogist 94, 494^506.
Leake, B. E., Woolley, A. R., Arps, C. E. S., Birch, W. D., Gilbert, M.
C., Grice, J. D., Hawthorne, F. C., Kato, A., Kisch, H. J.,
Krivovichev, V. G., Linhout, K., Laird, J., Mandarino, J.,
Maresch, W. V., Nickel, E. H., Rock, N. M. S., Schumacher, J. C.,
Smith, D. C., Stephenson, N. C. N., Ungaretti, L., Whittaker, E. J.
W. & Youzhi, G. (1997). Nomenclature of amphiboles: Report of
the subcommittee on Amphiboles of International Mineralogical
Association Commission on New Minerals and Mineral Names.
Mineralogical Magazine 61, 295^321.
Legrende, C. (1999). Petrogene'se du volcan Misti moderne (Sud
Perou): essai de caracterisation de levolution petro-geochimique et
chronologique, DEA thesis, Universite Blaise-Pascal, ClermontFerrand, 82 p.
Legros, F. (1998). Tephrostraphie du volcan Misti (Perou) et
Modelisation des coulees pyroclastiques, PhD thesis, Universite
Blaise-Pascal, Clermont-Ferrand, 139 p.
Liang, Y. (2000). Dissolution in molten silicates: effects of solid solution. Geochimica et Cosmochimica Acta 64, 1617^1627.
Liang, Y., Richter, F. M., Davis, A. M. & Watson, E. B. (1996).
Diffusion in silicate melts: I. Self diffusion in CaO^Al2O3^SiO2 at
15008C and 1 GPa. Geochimica et Cosmochimica Acta 60, 4353^4367.
Mamani, M., Worner, G. & Sempere, T. (2010). Geochemical variations in igneous rocks of the Central Andean orocline (138S to
188S): tracing crustal thickening and magma generation through
time and space. Geological Society of America Bulletin 122, 162^182.
Marsh, B. D. (1989). On convective style and vigor in sheet-like
magma chambers. Journal of Petrology 30, 479^530.

31

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

Davidson, J. P. & Tepley, F. J., III (1997). Recharge in volcanic systems:


Evidence from isotope profiles of phenocrysts. Science 275,826^829.
de Silva, S. L. & Francis, P. (1991a). Volcanoes of the Central Andes. New
York: Springer.
de Silva, S. L. & Francis, P. W. (1991b). Potentially active volcanoes of
PeruObservations using Landsat Thematic Mapper and Space
Shuttle imagery. Bulletin of Volcanology 52, 286^301.
de Silva, S. L., Salas, G. & Schubring, S. (2008). Triggering explosive
eruptionsthe case for silicic magma recharge at Huaynaputina,
southern Peru. Geology 36, 387^390.
Dymond, J., Corliss, J. B., Heath, G. R., Field, C. W., Dasch, E. J. &
Veeh, H. H. (1973). Origin of metalliferous sediments from the
Pacific Ocean. Geological Society of America Bulletin 84, 3355^3372.
Eggler, D. H. (1972). Water-saturated and undersaturated melting relations in a Paricutin andesite and an estimate of water content in the
natural magma. Contributions to Mineralogy and Petrology 34,261^271.
Eggler, D. H. & Burnham, C. W. (1973). Crystallization and fractionation trends in the system andesite^H2O^CO2^O2 at pressures to
10 kb. Geological Society of America Bulletin 84, 2517^2532.
Eichelberger, J. C. (1978). Andesitic volcanism and crustal evolution.
Nature 275, 21^27.
Eichelberger, J. C., Chertkoff, D. G., Dreher, S. & Nye, C. J. (2000).
Magmas in collision: rethinking chemical zonation in silicic
magmas. Geology 28, 603^606.
Feeley, T. C. & Dungan, M. A. (1996). Compositional and dynamic
controls on mafic^silicic magma interactions at continental arc volcanoes: evidence from Cordon El Guadal, Tatara^San Pedro complex, Chile. Journal of Petrology 37, 1547^1577.
Folch, A. & Marti, J. (1998). The generation of overpressure in felsic
magma chambers by replenishment. Earth and Planetary Science
Letters 163, 301^314.
Garcia, M. O. & Jacobson, S. S. (1979). Crystal clots, amphibole fractionation and the evolution of calc-alkaline magmas. Contributions
to Mineralogy and Petrology 69, 319^327.
Geschwind, C. & Rutherford, M. J. (1995). Crystallization of microlites during magma ascent: the fluid mechanics of recent eruptions
at Mount St. Helens. Bulletin of Volcanology 57, 356^370.
Ghiorso, M. S. & Evans, B. W. (2008). Thermodynamics of rhombohedral oxide solid solutions and a revision of the Fe^Ti two-oxide
geothermometer and oxygen-barometer. American Journal of Science
308, 957^1039.
Gregg, P. M., de Silva, S. L., Grosfils, E. B. & Parmigiani, J. P. (2012).
Catastrophic caldera-forming eruptions: Thermomechanics and
implications for eruption triggering and maximum caldera dimensions on Earth. Journal of Volcanology and Geothermal Research 241^
242, 1^12.
Hammer, J. E. (2008). Experimental studies of the kinetics and energetics of magma crystallization. In: Putirka, K. & Tepley, F. J., III
(eds) Minerals, Inclusions and Volcanic Processes. Mineralogical Society of
America and Geochemical Society, Reviews in Mineralogy and Geochemistry
69, 9^59.
Hammer, J. E., Cashman, K. V., Hoblitt, R. P. & Newman, S. (1999).
Degassing and microlite crystallization during pre-climactic
events of the 1991 eruption of Mt. Pinatubo, Philippines. Bulletin of
Volcanology 60, 355^380.
Harpel, C., de Silva, S. L. & Salas, G. (2011). The 2 ka eruption of
Misti Volcano, southern Peruthe most recent plinian eruption of
Arequipas iconic volcano. Geological Society of America, Special Papers
484, 1^72.
Hildreth, W. (1983). The compositionally zoned eruption of 1912 in the
Valley of Ten Thousand Smokes, Katmai National Park, Alaska.
Journal of Volcanology and Geothermal Research 18, 1^56.

JOURNAL OF PETROLOGY

VOLUME 0

NUMBER 0

MONTH 2013

32

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

gathering during magma mixing. Geochemistry, Geophysics, Geosystems


9, Q07017, doi:10.1029/2008GC002022.
Ruprecht, P., Bergantz, G. W., Cooper, K. M. & Hildreth, W. (2012).
The crustal magma storage system of Volcan Quizapu, Chile, and
the effects of magma mixing on magma diversity. Journal of
Petrology 53, 801^840.
Rutherford, M. J. & Devine, D. D., III (1988). The May 18,1980, eruption
of Mount St. Helens, III. Stability and chemistry of amphibole in the
magma chamber. Journal of Geophysical Research 98, 19667^19685.
Rutherford, M. J. & Devine, D. D., III (2003). Magmatic conditions
and magma ascent rate as indicated by hornblende phase equilibria and reaction in the 1995^2002 Soufrie're Hills magma. Journal
of Petrology 44, 1433^1454.
Rutherford, M. J. & Hill, P. M. (1993). Magma ascent rates from
amphibole breakdown: an experimental study applied to the 1980^
1986 Mount St. Helens eruptions. Journal of Geophysical Research
98(B11), 19667^19685.
Scaillet, B. & Evans, B. W. (1999). The 15 June 1991 eruption of Mount
Pinatubo. I. Phase equilibria and pre-eruption P^T^fO2^fH2O
conditions of the dacite magma. Journal of Petrology 40, 381^411.
Schmidt, M. W. (1992). Amphibole composition in tonalite as a function
of pressure: an experimental calibration of the Al-in-hornblende barometer. Contributions to Mineralogy and Petrology 110, 304^310.
Shimizu, N. (1983). Interface kinetics and trace element distribution
between phenocrysts and magma. In: Augustithis, S. S. (ed.) The
Significance of Trace Elements in Solving Petrogenetic Problems and
Controversies. Athens: Theophrastus, pp. 175^195.
Singer, B. S., Dungan, M. A. & Layne, G. D. (1995). Textures and Sr,
Ba, Mg, Fe, K, and Ti compositional profiles in volcanic plagioclase: Clues to the dynamics of calc-alkaline magma chambers.
American Mineralogist 80, 776^798.
Snyder, D. & Tait, S. (1995). Replenishment of magma chambers: comparison of fluid-mechanic experiments with field relations.
Contributions to Mineralogy and Petrology 122, 230^240.
Sparks, R. S. J. & Marshall, L. A. (1986). Thermal and mechanical
constraints on mixing between mafic and silicic magmas. Journal
of Volcanology and Geothermal Research 29, 99^124.
Sparks, R. S. J., Sigurdsson, H. & Wilson, L. (1977). Magma mixing: a
mechanism for triggering acid explosive eruptions. Nature 267,
315^318.
Spear, F. (1981). An experimental study of hornblende stability and
compositional variability in amphibolite. American Journal of Science
281, 697^734.
Suzuki, Y. & Nakada, S. (2007). Remobilization of highly crystalline
felsic magma by injection of mafic magma: Constraints from the
middle sixth century eruption of Haruna Volcano, Honshu, Japan.
Journal of Petrology 48, 1543^1567.
Tait, S., Jaupart, C. & Vergniolle, S. (1989). Pressure, gas content and
eruption periodicity of a shallow, crystallizing magma chamber.
Earth and Planetary Science Letters 92, 107^123.
Tepley, F. J., III, Davidson, J. P. & Clynne, M. A. (1999). Magmatic
interactions as recorded in plagioclase phenocrysts of Chaos
Crags, Lassen Volcanic Center, California. Journal of Petrology 40,
787^806.
Tepley, F. J., III, Lundstrom, C. C., McDonough, W. F. &
Thompson, A. (2010). Trace element partitioning between high-An
plagioclase and basaltic to basaltic andesite melt at 1 atmosphere
pressure. Lithos 118, 82^94.
Thomas, W. M. & Ernst, W. G. (1990). The aluminum content of hornblende in calc-alkaline granitic rocks: A mineralogic barometer
calibrated experimentally to 12 kbars. In: Spencer, R. J. &
Chou, I.-M. (eds) Fluid^Mineral Interactions. A Tribute to H. P. Euster.
Geochemical Society Special Publications 2, 59^63.

Martel, C. & Schmidt, B. C. (2003). Decompression experiments as an


insight into ascent rates of silicic magmas. Contributions to
Mineralogy and Petrology 144, 397^415.
Martel, C., Pichavant, M., Holtz, F., Scailet, B., Bourdier, J.-L. &
Traineau, H. (1999). Effects of fO2 and H2O on andesite phase relations between 2 and 4 kbar. Journal of Geophysical Research 104,
29453^29470.
Matthews, S. J., Gardeweg, M. C. & Sparks, R. S. J. (1997). The 1984
to 1996 cyclic activity of Lascar Volcano, northern Chile: cycles of
dome growth, dome subsidence, degassing and explosive eruptions.
Bulletin of Volcanology 59, 72^82.
Metrich, N. & Rutherford, M. (1998). Low pressure crystallization
paths of H2O-saturated basaltic^hawaiitic melts from Mt. Etna:
implications for open-system degassing of basaltic volcanoes.
Geochimica et Cosmochimica Acta 62, 1195^1205.
Moore, G. & Carmichael, I. S. E. (1998). The hydrous phase equilibria
(to 3 kbar) of an andesite and basaltic andesite from western
Mexico: constraints on water content and conditions of phenocryst
growth. Contributions to Mineralogy and Petrology 130, 304^319.
Muncill, G. E. & Lasaga, A. C. (1988). Crystal-growth kinetics of
plagioclase in igneous systems: Isothermal H2O-saturated experiments and extension of a growth model to explain silicate melts.
American Mineralogist 73, 982^992.
Nakada, S. & Motomura, Y. (1999). Petrology of the 1991^1995 eruption at Unzen: effusion pulsation and groundmass crystallization.
Journal of Volcanology and Geothermal Research 89, 173^196.
Nielsen, R. L. & Drake, M. J. (1979). Pyroxene^melt equilibria.
Geochimica et Cosmochimica Acta 43, 1259^1273.
Nielsen, R. L. & Dungan, M. A. (1983). Low pressure mineral^melt
equilibria in natural anhydrous mafic systems. Contributions to
Mineralogy and Petrology 84, 310^326.
Pallister, J. S., Hoblitt, R. P. & Reyes, A. G. (1992). A basalt trigger for
the 1991 eruption of Pinatubo volcano? Nature 356, 426^428.
Paquereau Lebti, P., Thouret, J.-C., Worner, G. & Fornari, M. (2006).
Neogene and Quaternary ignimbrites in the area of Arequipa,
Southern Peru: Stratigraphical and petrological correlations.
Journal of Volcanology and Geothermal Research 154, 251^275.
Pichavant, M., Costa, F., Burgisser, A., Scaillet, B., Martel, C. &
Poussineau, S. (2007). Equilibration scales in silicic to intermediate
magmasimplications for experimental studies. Journal of Petrology
48, 1955^1972.
Pouchou, J. L. & Pichoir, F. (1984).PAP (f^r^z) correction procedure for improved quantitative microanalysis. In: Armstrong, J. T.
(ed.) Microbeam Analysis. San Francisco, CA: San Francisco Press,
pp. 104^106.
Putirka, K. (2008). Thermometers and barometers for volcanic systems. In: Putirka, K. & Tepley, F. J., III (eds) Minerals, Inclusions
and Volcanic Processes. Mineralogical Society of America and Geochemical
Society, Reviews in Mineralogy and Geochemistry 69, 61^120.
Ridolfi, F., Renzulli, A. & Puerini, M. (2010). Stability and chemical
equilibrium of amphibole in calc-alkaline magmas: an overview,
new thermobarometric formulations and application to subduction-related volcanoes. Contributions to Mineralogy and Petrology 160,
45^66.
Ruprecht, P. & Bachmann, O. (2010). Pre-eruptive reheating during
magma mixing at Quizapu volcano and the implications for the explosiveness of silicic arc volcanoes. Geology 38, 919^922.
Ruprecht, P. & Worner, G. (2007). Variable regimes in magma systems
documented in plagioclase zoning patterns: El Misti stratovolcano
and Andahua monogenetic cones. Journal of Volcanology and
Geothermal Research 165, 142^162.
Ruprecht, P., Bergantz, G. W. & Dufek, J. (2008). Modeling of gasdriven magmatic overturn: Tracking of phenocryst dispersal and

TEPLEY et al.

EL MISTI 2000 BP ERUPTION

Turner, J. S. & Campbell, I. H. (1986). Convection and mixing in


magma chambers. Earth-Science Reviews 23, 255^352.
Tuttle, O. F. & Bowen, N. L. (1958). Origin of granite in the light of experimental studies in the system NaAlSi3O8^KAlSi3O8^SiO2^H2O. Geological
Society of America, Memoirs 74, 153 p.
Wright, T. L. & Doherty, P. C. (1970). A linear programming and least
squares computer method for solving petrologic mixing problems.
Geological Society of America Bulletin 81, 1995^2008.

Thouret, J.-C., Finizola, A., Fornari, M., Legeley-Padovani, A.,


Suni, J. & Frechen, M. (2001). Geology of El Misti volcano near
the city of Arequipa, Peru. Geological Society of America Bulletin
113(12), 1593^1610.
Tsuchiyama, A. (1985). Dissolution kinetics of plagioclase in the melt
of the system diopside^albite^anorthite, and the origin of dusty
plagioclase in andesites. Contributions to Mineralogy and Petrology 89,
1^16.

Downloaded from http://petrology.oxfordjournals.org/ at Niedersaechsische Staats- und Universitaetsbibliothek Goettingen on July 31, 2013

33

You might also like