Phreatomagmatic eruptions through unconsolidated coastal plain sequences, Maungataketake, Auckland...

18
Phreatomagmatic eruptions through unconsolidated coastal plain sequences, Maungataketake, Auckland Volcanic Field (New Zealand) Javier Agustín-Flores a, , Károly Németh a , Shane J. Cronin a , Jan M. Lindsay b , Gábor Kereszturi a , Brittany D. Brand c , Ian E.M. Smith b a Volcanic Risk Solutions, Massey University, Private Bag 11 222, Palmerston North 4442, New Zealand b School of Environment, The University of Auckland, New Zealand c Earth and Space Sciences Department, University of WA, USA abstract article info Article history: Received 23 October 2013 Accepted 26 February 2014 Available online 12 March 2014 Keywords: Monogenetic Phreatomagmatism Maar Tuff ring Shallow diatreme Unconsolidated substrate Maungataketake is a monogenetic basaltic volcano formed at ~8589 ka in the southern part of the Auckland Volcanic Field (AVF), New Zealand. It comprises a basal 1100-m diameter tuff ring, with a central scoria/spatter cone and lava ows. The tuff ring was formed under hydrogeological and geographic conditions very similar to the present. The tuff records numerous density stratied, wet base surges that radiated outward up to 1 km, decelerating rapidly and becoming less turbulent with distance. The pyroclastic units dominantly comprise ne-grained expelled grains from various sedimentary deposits beneath the volcano mixed with a minor compo- nent of juvenile pyroclasts (~35 vol.%). Subtle lateral changes relate to deceleration with distance and vertical transformations are minor, pointing to stable explosion depths and conditions, with gradual transitions between units and no evidence for eruptive pauses. This volcano formed within and on ~ 60 m-thick Plio/Pleistocene, poorly consolidated, highly permeable shelly sands and silts (Kaawa Formation) capped by near-impermeable, water-saturated muds (Tauranga Group). These sediments rest on moderately consolidated Miocene-aged permeable turbiditic sandstones and siltstones (Waitemata Group). Magmawater fuelled thermohydraulic explosions remained in the shallow sedimentary layers, excavating ne-grained sediments without brittle fragmentation required. On the whole, the resulting cool, wet pyroclastic density currents were of low energy. The unconsolidated shallow sediments deformed to accommodate rapidly rising magma, leading to development of complex sill-like bodies and a range of magmawater contact conditions at any time. The weak saturated sedi- ments were also readily liqueed to provide an enduring supply of water and ne sediment to the explosion loci. Changes in magma ux and/or subsequent stabilisation of the conduit area by a lava ring-barrier led to ensuing Strombolian and re-fountaining eruption phases. Future eruptions in littoral environments around Auckland are likely to be of this type, producing base surges that rapidly decrease in energy over short runout distances (~ 1 km). © 2014 Elsevier B.V. All rights reserved. 1. Introduction Fragmentation of ascending magma in phreatomagmatic eruptions is primarily due to highly energetic, explosive fuelcoolant interactions (FCI) (Sheridan and Wohletz, 1983; Wohletz, 1983), also called molten fuelcoolant interactions (MFCI) by Zimanowski et al. (1997). Around 60% of the thermal energy of the involved magma (the fuel) when it in- teracts explosively with water (the coolant) is converted into shock waves that have the potential to disrupt surrounding magma and/or country rock (Raue, 2004). The types of non-juvenile fragments in the phreatomagmatic deposit sequences, as well as the morphology of the diatreme, may be controlled by the degree of excavation into the substrate (White and Ross, 2011). In turn, the extent of excavation and tuff ring/crater structure are inuenced by the composition and strength of this substrate (Lorenz, 2003). Recent studies have also showed that the nal architecture of the crater and underlying diatreme are strongly dependent on the energy release of individual explosive eruptions as well as their relative position in reference to the surface (Valentine et al., 2011; Valentine, 2012; Valentine and White, 2012). The hydroge- ology and the rheology of the country rocks are still expected to play a major role in the shape, size and facies architecture of the craters and diatremes, although the exact nature of this role is not well constrained. The importance of the physical and/or hydrogeological properties of the substrate in phreatomagmatic eruptive processes has been ad- dressed by many authors (White, 1991, 1996; Lorenz, 2003; Sohn and Park, 2005; Auer et al., 2007; Németh et al., 2010, 2012; Valentine, 2012). Some of the features regarded as typical of small basaltic erup- tions through soft/unconsolidated substrates include: the common syn-eruptive downward slumping of crater walls; the formation of broad, shallow champagne glass-shaped craters; and the formation Journal of Volcanology and Geothermal Research 276 (2014) 4663 Corresponding author at: Massey University, Private Bag 11 222, Palmerston North 4442, New Zealand. Tel.: +64 6 3505701x2563. E-mail address: [email protected] (J. Agustín-Flores). http://dx.doi.org/10.1016/j.jvolgeores.2014.02.021 0377-0273/© 2014 Elsevier B.V. All rights reserved. Contents lists available at ScienceDirect Journal of Volcanology and Geothermal Research journal homepage: www.elsevier.com/locate/jvolgeores

Transcript of Phreatomagmatic eruptions through unconsolidated coastal plain sequences, Maungataketake, Auckland...

Journal of Volcanology and Geothermal Research 276 (2014) 46–63

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

j ourna l homepage: www.e lsev ie r .com/ locate / jvo lgeores

Phreatomagmatic eruptions through unconsolidated coastal plainsequences, Maungataketake, Auckland Volcanic Field (New Zealand)

Javier Agustín-Flores a,⁎, Károly Németh a, Shane J. Cronin a, Jan M. Lindsay b, Gábor Kereszturi a,Brittany D. Brand c, Ian E.M. Smith b

a Volcanic Risk Solutions, Massey University, Private Bag 11 222, Palmerston North 4442, New Zealandb School of Environment, The University of Auckland, New Zealandc Earth and Space Sciences Department, University of WA, USA

⁎ Corresponding author at: Massey University, Private4442, New Zealand. Tel.: +64 6 3505701x2563.

E-mail address: [email protected] (J. Agustín-Flo

http://dx.doi.org/10.1016/j.jvolgeores.2014.02.0210377-0273/© 2014 Elsevier B.V. All rights reserved.

a b s t r a c t

a r t i c l e i n f o

Article history:Received 23 October 2013Accepted 26 February 2014Available online 12 March 2014

Keywords:MonogeneticPhreatomagmatismMaarTuff ringShallow diatremeUnconsolidated substrate

Maungataketake is a monogenetic basaltic volcano formed at ~85–89 ka in the southern part of the AucklandVolcanic Field (AVF), New Zealand. It comprises a basal 1100-m diameter tuff ring, with a central scoria/spattercone and lava flows. The tuff ring was formed under hydrogeological and geographic conditions very similar tothe present. The tuff records numerous density stratified, wet base surges that radiated outward up to 1 km,decelerating rapidly and becoming less turbulent with distance. The pyroclastic units dominantly comprisefine-grained expelled grains fromvarious sedimentary deposits beneath the volcanomixedwith aminor compo-nent of juvenile pyroclasts (~35 vol.%). Subtle lateral changes relate to deceleration with distance and verticaltransformations areminor, pointing to stable explosion depths and conditions, with gradual transitions betweenunits and no evidence for eruptive pauses. This volcano formed within and on ~60 m-thick Plio/Pleistocene,poorly consolidated, highly permeable shelly sands and silts (Kaawa Formation) capped by near-impermeable,water-saturated muds (Tauranga Group). These sediments rest on moderately consolidated Miocene-agedpermeable turbiditic sandstones and siltstones (Waitemata Group). Magma–water fuelled thermohydraulicexplosions remained in the shallow sedimentary layers, excavating fine-grained sediments without brittlefragmentation required. On the whole, the resulting cool, wet pyroclastic density currents were of low energy.The unconsolidated shallow sediments deformed to accommodate rapidly rising magma, leading to developmentof complex sill-like bodies and a range of magma–water contact conditions at any time. The weak saturated sedi-ments were also readily liquefied to provide an enduring supply of water and fine sediment to the explosion loci.Changes in magma flux and/or subsequent stabilisation of the conduit area by a lava ring-barrier led to ensuingStrombolian and fire-fountaining eruption phases. Future eruptions in littoral environments around Auckland arelikely to be of this type, producing base surges that rapidly decrease in energy over short runout distances (~1 km).

© 2014 Elsevier B.V. All rights reserved.

1. Introduction

Fragmentation of ascending magma in phreatomagmatic eruptionsis primarily due to highly energetic, explosive fuel–coolant interactions(FCI) (Sheridan andWohletz, 1983; Wohletz, 1983), also called moltenfuel–coolant interactions (MFCI) by Zimanowski et al. (1997). Around60% of the thermal energy of the involvedmagma (the fuel) when it in-teracts explosively with water (the coolant) is converted into shockwaves that have the potential to disrupt surrounding magma and/orcountry rock (Raue, 2004). The types of non-juvenile fragments in thephreatomagmatic deposit sequences, as well as the morphology of thediatreme, may be controlled by the degree of excavation into thesubstrate (White and Ross, 2011). In turn, the extent of excavation and

Bag 11 222, Palmerston North

res).

tuff ring/crater structure are influenced by the composition and strengthof this substrate (Lorenz, 2003). Recent studies have also showed thatthe final architecture of the crater and underlying diatreme are stronglydependent on the energy release of individual explosive eruptions aswell as their relative position in reference to the surface (Valentineet al., 2011; Valentine, 2012; Valentine and White, 2012). The hydroge-ology and the rheology of the country rocks are still expected to play amajor role in the shape, size and facies architecture of the craters anddiatremes, although the exact nature of this role is not well constrained.

The importance of the physical and/or hydrogeological properties ofthe substrate in phreatomagmatic eruptive processes has been ad-dressed by many authors (White, 1991, 1996; Lorenz, 2003; Sohn andPark, 2005; Auer et al., 2007; Németh et al., 2010, 2012; Valentine,2012). Some of the features regarded as typical of small basaltic erup-tions through soft/unconsolidated substrates include: the commonsyn-eruptive downward slumping of crater walls; the formation ofbroad, shallow “champagne glass”-shaped craters; and the formation

47J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

of a shallow diatreme (Lorenz, 2003). However, it is difficult to link thesubsurface hydrogeological/lithological conditions with diatreme anderuptive conditions and processes, largely due to the lack of coupledsubsurface and surficial deposit exposures.

Coastal erosion beside the Manukau Harbour in the southern AVF(Fig. 1) has carved an almost complete cross-section on the SW flank of

Fig. 1. Location of Maungataketake volcano in the Auckland Volcanic Field, along with other nanot have evidence of phreatomagmatic activity (e.g. Rangitoto); names in bold fonts stand fornames in italic bold fonts are those that show phreatomagmatic activity only (e.g. Onepoto). Thabove sea level in the shown area is 260 m at the summit of Rangitoto volcano.

the Maungataketake tuff ring (Fig. 2). This is one of several southAuckland volcanoes that lie near present-day sea level and are underlainby tens to ~60+mof saturated, soft, Pliocene to Recent muds, sands andgravels that rest on a N300 m-thick, poorly consolidated turbiditicsandstone/siltstone sequence (Kermode, 1992). The Auckland area hasnot experienced active faulting since 0.3 Ma (Kenny et al., 2012) and no

med volcanoes. Underlined names in regular fonts represent volcanoes that apparently dovolcanoes that include both magmatic and phreatomagmatic activity (e.g. Pupuke); ande city of Auckland practically stretches out within the entire region. Themaximumaltitude

Fig. 2. Plan view of Maungataketake volcano showing five key sites (M1 to M5) locatedalong a NW–SE-direction cliff. Letter A indicates the highest edge of the tuff ring rim(broken-lined bigger contour) and letter B the extent of quarry area (broken-lined smallercontour). For the purpose of this study, proximal,medial, and distal deposits are representedby sites M1–M2, M3, and M4–M5 respectively.Modified from Brand et al. (2014).

48 J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

significant tectonic uplift since 1Ma (Alloway et al., 2004). These consid-erations and the inferences drawn from paleo-environmental studies(Marra et al., 2006) indicate that the current hydrogeological conditionscorrespond to those existing when Maungataketake volcano erupted.These conditions may have been predominant during several periodsof the Last Interglacial where several high sea level stands are knownbetween 80 and 140 ka (Pillans, 1983). Although the ages onMaungataketake are not conclusive, someof the reported ages fall with-in this interval. For example, Marra et al. (2006), based on opticallystimulated luminescence and biostratigraphy report ages of 140 ±14.2 and 177 ± 23.4 ka. However, the most recent Ar–Ar age yieldedan age of 87.4 ± 2.4 ka (Leonard et al., unpublished data).

Maungataketake presents a case scenario of an eruption through ahighly deformable, fine-grained, water-saturated substrate. Its tuff ringsequence is presented here, along with information about the geologicaland hydrogeological setting, and an eruption scenario is constructed.Concise geochemical information is provided in support the physicaleruption scenario and to help infer the conditions during thephreatomagmatic eruption phase. This study improves the understandingof phreatomagmatic eruptions (and related hazards) in the soft-sedimentdominated coastal plains of the southern AVF, and it is an example forbasaltic volcanism in similar near shore soft-sediment settings.

2. Geological and hydrogeological setting

Maungataketake (previously also known as Ellett'sMountain, beforequarrying operations destroyed it) is a small basaltic volcano located onthewestern edge of the northern Manukau Lowlands at the SW edge ofthe AVF (Fig. 1) in the North Island of New Zealand. The AVF comprisesabout 52 individual eruption centres over a 360 km2 area (Kermode,1992; Allen and Smith, 1994; Spörli and Eastwood, 1997; Haywardet al., 2011). The minimum DRE volume of the eruption products ofthe AVF is 1.7 km3, based on measurements from a Light Detectionand Ranging (LiDAR) Digital SurfaceModel in combinationwith geolog-ical mapping (Kereszturi et al., 2013). Eruptions have been sporadic

since 250 ka, but primarily since ~50 ka (Allen and Smith, 1994;Molloy et al., 2009; Bebbington and Cronin, 2011).

The AVF is almost completely urbanised with the City of Auckland(population ~1.5 million) built on top of it (Fig. 1). The outstandingfeature of the AVF is the predominance of phreatomagmatism, withtuff rings andmaars and scoria cones preceded by phreatomagmatic ac-tivity present throughout the field (approximately 75% of the volcanoesshow evidence of phreatomagmatism). Explosive phreatomagmaticeruptions producing violent base surges are considered to be the mosthazardous events expected to threaten the city's inhabitants (Allenand Smith, 1994; Németh et al., 2012).

At least 20 volcanic centres have been identified in the ManukauLowlands. Based on observation of the preserved pyroclasticsuccessions, 16 of these involved variable degrees of phreatomagmaticexplosive phases. The northern Manukau area is located within afault-bounded graben formed within Early Miocene Waitemata Grouprocks (Kenny et al., 2012). These rocks comprise N300 m-thick inter-bedded turbiditic sandstones and pelagic siltstones with subordinatebreccia and conglomerate units (Ballance, 1974; Hayward, 1979,1993; Raza et al., 1999). The undulating Waitemata paleosurface waspart of an ancient fluvial system that irrigated the Manukau area inthe Pleistocene (Searle and Mayhill, 1981; Kermode, 1992).

Waitemata rocks are weakly indurated with a compressivestrength of ~5 MPa (Spörli and Rowland, 2007). Jointed siltstoneand sandstone beds and conglomerates with a range of thicknessesact as aquifers (Simpson, 1987; Sheridan, 2006; Irwin, 2009; Kennyet al., 2012), often confined by low-permeability mudstones(Scoble and Millar, 1995; Crowcroft and Smaill, 2001; Sheridan,2006). They are low yield aquifers that are heterogeneous withanisotropic hydraulic properties. From a combination of pumptesting and laboratory values, the average hydraulic conductivities(K) in x and y directions are Kx = 10−6 and Kz = 10−10 for theWaitemata rocks (Crowcroft and Smaill, 2001; Delamore and Partners,2003; Sheridan, 2006). Usingfield observations on jointing and beddingpatterns and porosity laboratory studies of Waitemata rocks, Sheridan(2006) concludes that the anisotropic hydraulic conductivity values ofWaitemata aquifers reflect the rock jointing and bedding patterns.

Across theManukau area, theWaitemata Group sediments (especiallywithin the graben) are unconformably overlain in depressions by poorlyconsolidated Pliocene shallow marine and estuarine beds of the KaawaFormation (Viljevac et al., 2002; Edbrooke et al., 2003). This sequenceforms confined aquifers with high permeability (average K = ~10−5,but variable) that are favourable for groundwater storage (Viljevacet al., 2002), although around the Maungataketake area the thickness ofKaawa sediments is not known for certain.

The overlying low-permeability, water-saturated, recent TaurangaGroup sediments confine the Kaawa Formation aquifers in many places(Viljevac et al., 2002). The lowermost sequences of the Tauranga Groupconsist of pumiceous sand, mud, silt, and carbonaceous peat depositedin fluvial, lacustrine, and estuarine environments (Kermode, 1992;Edbrooke et al., 2003). This forms a sedimentary infill of variablethickness, reaching up to 60 m in the area (Kermode, 1992; Edbrookeet al., 2003). Extrapolation from borehole data (PETLAB database, G.N.S.Science, New Zealand, http://pet.gns.cri.nz/pet/index.jsp) suggests thatthe thickness of sediments overlying the Waitemata Group in theMaungataketake area might be 40–60 m. Tauranga Group sedimentsinterfinger with the volcanic eruptive products of the AVF, showingthat the Manukau Lowland volcanoes, including Maungataketake,erupted on a broad,flat coastal plain. Tauranga Group andKaawa Forma-tion are generically called Plio-Pleistocene sediments for this study.

3. General architecture of Maungataketake volcano

The base of the Maungataketake complex comprises a tuff ring withan irregular rim of approximately 1100 by 1300 m, with a long axisoriented NW–SE (Fig. 2). The outermost deposits define a 2000 by

49J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

1700 m area of 2.8 × 106 m2. The total bulk volume of the present tuff is~20.9 × 106m3, which gives a Dense Rock Equivalent eruptive volume of7.2 × 106 m3 (Kereszturi et al., 2013). The present-day tuff ring reachesup to ~25 m in maximum height, and its base at M3 (Fig. 3) roughlycoincides with the maximum high tide levels in the Manukau Harbour.

Following the emplacement of the tuff ring, there was lavafountaining from closely-spaced vents that built a lava spatter cone upto ~73m from its base (Searle, 1959). The cone,which has now been re-moved by quarrying, had two craters oriented NE–SW, with the majorcrater to the E (Searle, 1959). From the cone remnants, Conybeer(1995) documented stratified to crudely stratified, variably agglutinat-ed, grey to red-grey scoria and spatter. Lava flows were produced byfire fountaining and also emerged at the base of the cone on thewesternside, covering tuff to the NW and E (Searle, 1959). The lava flows do notextend beyond the limits of tuff deposits and consist of vesicular basalt.Despite the removal ofmost of the coneby quarrying, strong gravity andmagnetic anomalies occur at Maungataketake (Cassidy and Locke,2010), and are attributed to a large solidified magma body pondedunder the former vent area.

4. Methods and terminology

Tuff ring deposits were examined from proximal to distal locationsalong a low coastal cliff oriented roughly perpendicular to the craterrim (Fig. 2). A systematic lithological description was carried out at fivesites (Fig. 2). The outcrops on the NW section are discontinuous andwere not used in the reconstruction. The uppermost phreatomagmaticsequence is highly weathered and/or eroded at several places along thecliff. The sites are located in the proximal inner wall of the tuff ring(SiteM1), the ring crest (SiteM2), its outer wall (SiteM3), and the distalfan (Sites M4 and M5) (Fig. 2).

Pyroclasts are defined by Fisher and Schmincke (1984) as rock frag-ments ‘produced by many processes connected with volcanic eruptions’,without reference to the causes of the eruption or origin. Pyroclastscan be juvenile (fresh erupted magma) or country-rock fragments

Fig. 3. Schematic correlation of logs and identified units (U), overlying peaty soils (PS), along clVertical exageration is ~30 times. The vertical axis represent the maxium thickness of the expo(that is approximately 3 m above low tide). Site M1 is located in the inner slope of the tuff ring(see Fig. 2). The other locations correspond to the outer tuff ring deposits. A phreatomagmwheathering, and/or transport/deposition mechanisms. Frequency histograms of 12 selected saof grain size fractions are shown in the uppermost graph.Note thehighprevalence of fine ash (~

(non-juvenile), which in this casemay also include older volcanic clasts.Following this, the poorly to moderately consolidated Maungataketakedeposits were classified as tuff and lapilli tuff (after Schmid, 1981)(Table 1). Particleswere grouped following the subdivision for pyroclastsize terms of Sohn and Chough (1989) (see Table 1). b1 cm-thick andN1 cm-thick layers are described as laminae and beds respectively(following Ingram, 1954).

Base surge is defined for the text as a low concentration, turbulent,pyroclastic density current (a diluted pyroclastic density current,DPDC) of phreatomagmatic origin and that consists of two phase (solidsand gas) or three phase (solids, gas, and water) systems. Current(s) orflow(s) may be used as a synonym of base surge throughout the text,and we infer that no other type of pyroclastic density currents wereemplaced at Maungataketake.

Most of the deposits are moderately consolidated, and thus notsuitable for sieving. Some poorly consolidated-tuff samples weredisaggregated by soaking in water and lightly crushing. These weremechanically sieved between 1 and 4.0 ϕ at 0.5 ϕ intervals. Particlessmaller than 4.0 ϕwere analysed using a Horiba Partica LA-950 LaserDiffraction Particle Size Distribution Analyzer. Grain size distributionparameters (Inman, 1952; Folk and Ward, 1957) were calculatedusing the application SFT (Sequential Fragmentation/Transport applica-tion) (developed by Ken Wohletz, available at http://www.ees1.lanl.gov/Wohletz/SFT.htm). These parameters may in this case be affectedby errors resulting from the manual disaggregation of samples, butmost units show comparable distributions to those of other tuff rings(Fig. 3). Particles from the 2 ϕ and 3 ϕ fractions were cleaned withHCl 10% and rinsed with acetone in an ultrasonic bath for 30–60 s.Morphological and compositional characteristics of fragments werecharacterized in the lab by viewing samples (loose grains and thin sec-tions) under light, petrography, and scanning electron microscopy (FEIQuanta 200 environmental scanning electron microscope operatedunder 20 kV, Massey University Microscope Centre), as well as by appli-cation of Energy-dispersive spectrometry (EDS) (Edax 10 mm2 detector.with data processed by Edax Genesis 5.21 software). Counts of 500 grains

iff exposures. It represents an approach of the cross section of the tuff ring from NW to SE.sed sequence. The base of the tuff deposits at site M3 roughly coincide with high tide level(see Fig. 2). Site M2 is likely conformed by deposits of the highest edge of the tuff ring rimatic sequence may or may not be present in one or more exposures due to slumping,mples are displayed indicating the position of samples in the exposures. The percentages70 vol.%) in the analysed samples that represent the pyroclastic successions of the tuff ring.

Table1

Nom

enclatureof

depo

sittyp

esan

dgrainsize.A

lsotheprop

ortion

sof

diffe

rent

sizesan

dcompo

sition

offrag

men

tswithinthede

positsaresh

own,

aswellasthemorph

olog

icalan

dsometexturalch

aracteristicsof

juve

nilean

dno

n-juve

nilefrag

men

ts.

Dep

osittype

Clastsize

~vol.%

within

depo

sittype

~vol.%

intotal

depo

sit

~Propo

rtionof

accide

ntalto

juvenile

Morph

olog

yan

dorigin

ofaccide

ntal

frag

men

tsMorph

olog

yof

juve

nile

frag

men

tsTe

xturean

dve

sicles

ofjuve

nile

frag

men

ts(0.5

to2.5mm)

Not

laye

red,

scattered

frag

men

tsBlock/bo

mb

(−6to

−8ϕ)

70b10

9:1

Subrou

nded

toroun

dedW

aitemata

sand

ston

e(Fig.7

a,b)

andpo

orly-

cons

olidated

,san

d/siltag

greg

ated

ofPlio-Pleistocene

sedimen

ts(Fig.7

a).

Softag

greg

ates

dono

tsh

owplastic

deform

ationor

internal

structures.

Den

se,sub

angu

larto

subrou

nded

,po

orly

vesicu

lar.Few

breadc

rust-

type

bombs.

Mod

alcompo

nents(Fig.4

a,b):

–Ca

-Aug

ite,tabu

larmicrolites

(usu

ally

b78

×8μm

):up

to46

vol.%

–Ca

-Aug

ite,elon

gatedmicroph

enocrysts/

phen

ocrysts:

b20

vol.%

–Eu

hedral

olivineph

enoc

rysts(70–

600μm

):8–

47vo

l.%–

Oxide

s:b10

vol.%

–Side

romelan

ematrix:

15–40

vol.%

Vesicles:

roun

dto

subtly

elon

gated,

exhibitwea

kde

form

ation(Figs.4a

,b,5

e,f,g,

h,i)

–W

allthickne

ss:b

100μm

–Mod

alpe

rcen

tage

:5.2–18

.2vo

l.%–

Vesicle

leng

th:3

9μm

(med

ian)

–Vesicle

eccentricity:1

.7(m

edian)

Coarse

lapilli

(−4to

−6ϕ)

30

Lapilli

tuff

Med

ium

lapilli

(−2to

−4ϕ)

b5

20–40

2:8

Frag

men

tsof

softrock

Waitemata

sand

ston

e,sand

/silt

aggreg

ates

ofPlio-Pleistocene

sedimen

ts.

Suba

ngular

tosu

brou

nded

,poo

rly

vesicu

lar.So

meclasts

areco

ated

byafine

film

ofaccide

ntal

ash.

Some

grains

show

somede

gree

ofpa

lago

nitization

.

Fine

lapilli

(−1to

−2ϕ)

10–20

Coarse

ash

(1to

−1ϕ)

60–80

Tuff

Med

ium

ash

(4to

1ϕ)

Upto

4060

–80

9:1

Dow

nto

3ϕfraction

:Blocky,

opaq

ueindividu

algrains

ofqu

artz

andfeldspar

(N40

vol.%

)(Figs.5a

,b,c).Silic

a-rich

,no

n-ve

sicu

larglassfrag

men

ts(Fig.5

a,b).

Roun

dag

greg

ates

ofsm

allerfrag

men

ts(Fig.5

a).Fragm

ents

ofPlio-Pleistocene

sedimen

tsdo

minate.Notethepresen

ceof

accretiona

ryalpilli

(Fig.5

d)

Dow

nto

3ϕfraction

.Sub

angu

larto

subrou

nded

,poo

rlyve

sicu

larfrag

men

ts(Fig.5

a,c,e,f,g).A

dheringfine

ash

particlesan

dsecond

arymineralsgrow

ths

occu

ron

theou

tersurfaces

andinside

the

cavities

ofve

sicles

(Fig.5

h,i).R

aresu

rficial

cracks

arepresen

t.So

mede

gree

ofpa

lago

nitization

.

Fine

ashb4ϕ

Upto

60

50 J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

were performed in the 2 and 3ϕ fraction sizes for the componentry anal-yses. Powder X-ray diffraction analyses were carried out, showing quartzand feldspar along with few clay minerals in the b2 ϕ fraction. The grainsize and componentry data of the consolidated lapilli tuff were obtainedfrom field observations and thin section analyses.

Using petrographic and scanning electron microscopy on six thinsections (see Section 5.1 for the definition of units and sample details),the textural and vesicle characteristics of 15 juvenile sideromelanegrains were described (Table 1, Fig. 4). A modal analysis performed foreach grain included 800 count points (results are shown in Fig. 4a andbriefly described in Table 1). The samples have overall identical texturaland vesicularity characteristics throughout the entire deposit sequence.

The methods used for whole rock and glass chemistry analyses arepresented in Appendix A. The compositional data are intended forcomparison of this eruption with others in the AVF.

5. Results

The tuff deposits generally show poor sorting, polymodal distribu-tions, and a roughly negative skewness (Table 1 and Fig. 3). Variationsin grain size distribution and particle componentry (Table 1) are usedto characterize stratigraphic divisions within the tuff. The characteristicsof juvenile fragments provide information about magma fragmentation(Wohletz, 1983; White, 1996; Zimanowski et al., 1997) (Table 1; Fig. 5),but the properties of non-juvenile fragments supply clues into theinvolvement of the host rock and related aquifers (e.g., Valentine, 2012;Lefebvre et al., 2013; van Otterloo et al., 2013) (Table 1; Fig. 5). Thesepropertieswere used to define lithofacies (Table 2) and lithostratigraphicunits (Table 3), which in turn provide the basis for the reconstruction ofthe eruptive history of the Maungataketake tuff ring.

5.1. Stratigraphy and sedimentary characteristics of the tuff ring deposits

The exposures studied along the cliff extend from proximal to distalreaches and are thought to be representative of other sectors of the tuffring which have similar dimensions (Fig. 2).

Ten lithofacieswere recognized (following Sohn and Chough, 1989).Bedforms and other depositional features were described that reflectthe main transport mechanisms (cf., Sohn, 1997; Valentine and Fisher,2000; Branney and Kokelaar, 2002. Three main modes of transport areconsidered here (these definitions do not strictly follow the authors'definitions): (1) lower traction carpet (from Sohn, 1997), characterizedby a non-turbulent regime where friction between fragments isubiquitous; (2) bed load (Valentine, 1987) (similar to upper tractioncarpet of Sohn, 1997), where the transport is mainly by saltation androlling in a non-turbulent to turbulent regime; and (3) suspendedload, characterized by turbulent transport of fragments (Valentine,1987). Bedforms were related to the mode of transport in a first-ordermanner, recognising that bedforms in some transport regimes will rep-resent the deposition of grains only at the boundary layer of a progres-sively aggrading current (Branney and Kokelaar, 2002). A general viewis that lower traction carpet, bed load, and suspended load modes oftransports may promote respectively the formation of planar, inverselygraded or massive beds; massive and/or planar to cross-bedded layerswith internal trains of coarser fragments; and cross-bedding and/orcross-lamination. However, from the sedimentary characteristics ofMaungataketake deposits these relationships are not clear-cut and donot always follow this general schema (see Table 2 and Section 6). Asflow distance increases, and at obstacles or slope-breaks, turbulencepatterns decrease or change (Valentine, 1987; Sulpizio and Dellino,2008), but within a single current, at a single point in any specifictime, turbulence increases upwards. Also, grain concentration andgrain size will decrease with distance and upwards within flow (Sohnand Chough, 1989). These will be reflected in bedforms and are consid-ered in the interpretation of bedforms of lithofacies in Table 2, and inturn in interpretation of the lithostratigraphic units (Table 3).

51J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

Lithostratigraphic units (U) (henceforth called units) (Table 3,Figs. 3, 6, 7) are deposit packages of stratawith lower and upper bound-aries that represent a more or less distinct change in grain size andcomponentry. These changes were most easily identified at the proxi-mal sites. A correlation was established with the distal deposits(where some vertical changes are faint) based on distinct stratigraphicmarkers (Fig. 3). The units may be characterized by at least onelithofacies with some showing variations grading from proximal to dis-tal locations (Figs. 3, 6, 7). A total of 6 units were identified. Based onstratigraphic and sedimentary characteristics, the phreatomagmaticphase of the Maungataketake eruption is interpreted to have occurredin four sub-phases (Phase(s), PH). Table 3 shows each PH representedby its units, which are described and interpreted. The sedimentary char-acteristics of the lithofacies contained in each unit are more extensivelydescribed in Table 2. The information in Tables 2 and 3 is the basis forthe following reconstruction of the Maungataketake eruption.

Macroscopically, the whole deposit displays a massive to plane-parallel, slightly undulating bedding. In detail, however, faint lamina-tion is widespread, along with plane-parallel/slightly undulatingbedding with a dominance of cross-lamination. Distal deposits aremostly plane-parallel bedded and laminated. Accretionary lapilli arepresent in varying amounts throughout the deposit, but are moreabundant in the upper U6 (Table 3) and units of the distal deposits.

Fig. 6c shows a tree log within the tuff. Tree logs and moulds arepresent within U1 deposits (there is a standing tree trunk betweensites M3 and M4; not shown). Also, there are fallen trees along thelow tide beach (not shown). According to Hayward and Hayward(1995) and Marra et al. (2006) the standing trees belonged to a liveforest (forest 2, Fig. 8) that was disturbed by the eruption, whereasthe lower fallen units are part of a fossil forest dominated by differentspecies (forest 1, Fig. 8) that was already dead at the time of the erup-tion. Hayward and Hayward demonstrated that the main orientationof the fallen logs was random, indicating that they were not knockeddown by the base surge blast. During this study we concurred withthe view that there were two forests. Therefore, the presence of fallentree logs will not be discussed here and we will assume that only forest2was disturbed by the eruption. Brand et al. (2014) discuss a theoreticalapproach on the dynamic pressure of the initial base surges generated atMaungataketake eruption and its impact on forest 2.

6. Eruption reconstruction

Amodel for this eruption (Fig. 8) integrates all textural, morphologic,lithologic and sedimentary characteristics contained in Tables 1, 2, and 3.

Fig. 4. a) Graph representing the modal analysis (vol.%) (800 point counting exclusively on sposition of samples increases rightward. Some sample numbers are represented by more thanunit): sample 27 (from upper U2 at site M1), sample 26 (from middle U3 at site M1), samplelower U6 at site M5), and sample 41 (from upper U6 at site M5); b) graph representing theTable 1 for more information).

Following Table 3 the reconstruction of the eruption will be describedaccording to four different phreatomagmatic phases. The lithofaciesand units mentioned in the following 4 subsections are described withmore detail in Tables 2 and 3 respectively.

6.1. Phase 1. Vent opening and shallow explosions

TheMaungataketake vent probably openedwhen the total vaporiza-tion energy exceeded the limit of containment (Sheridan and Wohletz,1983). Since the sedimentary record of the lower U1 includes thepresence of lithofacies T1 (Table 2), composed primarily of disaggre-gated material from the unconsolidated Plio-Pleistocene sediments(Tables 1, 2), and the apparent absence of tuff breccia horizon, it is likelythat the vent opening explosions disrupted primarily the ~60 m-thickunconsolidated Plio-Pleistocene sediments (Fig. 8a). However, althoughnot evident in the deposit record, deeper disrupted material and earlytuff deposits could have been recycled into the diatreme due toslumping into the widening crater (cf., Valentine, 2012). A series of tur-bulent to non-turbulent,wet, base surges formed thebasal deposits (U1,Table 3). These first currents entered a live forest (forest 2) (Fig. 8a, b).Someof the treeswere knocked downby the impact of the currents (seeFig. 6c) (Brand et al., 2014), or later fell under the weight of tephra.Subsequent explosions were likely also shallow seated (overallpresence of T1, see above), within unconsolidated sediments, whichalso likely continually slumped inwardly from unstable crater walls(cf., Auer et al., 2007) (Fig. 8c). At M3, the boundary between U1 andU2 is sharp with erosive characteristics (Table 2, Fig. 7c) that suggest achange in the eruption dynamics. However at M4/M5 sites this bound-ary is more gradational, which rules out any suggestion of elapsing timebetween the deposition U1 and U2.

U2 reflects deposition from distinctively density stratified basesurges (lithofacies LT1/T2) (Tables 2, 3)with higher contents of juvenileparticles (Table 3, Fig. 6c, d, e). This could indicate a highermagma effu-sion rate, or a reduction in external water (cf., Houghton et al., 1999).Since the juvenile-rich, lower part of the currents (lithofacies LT1 inU2, Table 3; Figs. 6c, 7a, b) show evidence of lower traction carpet trans-port, the contribution of fallout was probably never dominant (Table 2),orwas erased ormodified by themultiple base surges produced throughthe eruption. Shallow seated explosions may have been dominant at thisstage as indicated by the overall presence of Plio-Pleistocene sedimentsand lack of deeper Waitemata Group fragments. At distal sites (Figs. 6d, eand 7d), U2 shows low-amplitude (0.5 m) undulations that may be theresult of the paleo-surface topography, but this is not clear.

elected juvenile grains displayed in thin sections) of 15 juvenile fragment. Stratigraphicone juvenile grain. Specific sample position is as follow (see Fig. 3 for location of site and28 (from upper U3 at site M1), sample 31 (from middle U5 at site M2), sample 40 (frommaximum diameter of vesicles in the same juvenile fragments described above. (See

Fig. 5. a) Stereo light microscope image of 2ϕ fraction-size fragments from unit 2 (U2) atmedial distance from the vent (between sitesM3 andM4; Fig. 2). Distinct juvenile fragments arewithin the broken-lined circle. Subangular shapes dominate. The encircled particles within the solid line are: a country-rock lithic, and quartz and feldspar crystals. b) Stereo light micro-scope image of 3 ϕ fraction-size fragments from U1 at medial distance (same as above). Note the great majority (N90 vol.%) of accidental crystals (mainly quartz or feldspathic grains ormineral aggregates) (as coined in Sohn et al., 2009). c) Thin section image (plane parallel light) of juvenile particles from U3 at site M1 (Fig. 2). The bigger grain (S) corresponds to asubangular-to-sub-spheric sideromelane glass. Tachylite grains (T) are also present. Euhedral and subhedral olivine phenocrysts (Ol) are common in the fragments. d) Thin-sectionimage (plane parallel light) of tuff from U6 at location M2 showing rim-type accretionary lapilli. Open arrows point to juvenile fragments and filled arrow indicates an accidental lithic(siltstone). Quartz and feldspar grains (bright irregular shapes) are pervasive. e), f), and g) Scanning electronmicroscope (SEM) images of typical juvenile grain morphologies. Fragmentsare from U2 and U3 at medial distance (between sites M3 and M4; Fig. 2). They show subangular-to-subrounded edges and low vesicularity. Observe the similarity of shapes of juvenilefragments between SEM images and the stereo light microscope/thin section images. h) and i) SEM images of vesicles of juvenile particles from U1 at themedial distance. The vesicles aresub-spherical and show surface adhering and secondary mineral growths. Thick vesicle walls and smooth inner surfaces are present in most juvenile fragments.

52 J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

6.2. Phase 2. Excavation into the Waitemata Group rocks

A gradational boundary between U2 and U3 and the absence ofjuvenile-rich lithofacies (Table 3, Figs. 6c, 7a, b) suggest an uninterruptedactivity that in general progressed towards a return (as U1) to the dom-inance of accidental ash-laden base surges that deposited U3 (composedof lithofacies T3 and T6) (Table 3, Figs. 6c, 7a, b). The prevalence ofaccidental ash characterized by Plio-Pleistocene sediments and the initialabsence ofWaitemata rock fragments could indicate that shallow seatedexplosions prevailed. However, in the mid-section of U3 (Fig. 6d, 7a, b),there is possible evidence of deeper excavation into the underlyingWaitemata Group deposits (Fig. 8d), with small blocks and coarse lapilliof this lithology present (Figs. 6c, 7a, b). These are, however, roundedclasts, and thus likely represent Pleistocene fluvial deposits on top ofthe Waitemata paleosurface, consistent with other reconstructions of

the paleogeography of this area (Searle and Mayhill, 1981; Kermode,1992). Thus explosions probably did not penetrate deeper into theWaitemata rocks, but excavated the fragments at the boundary betweenthe Waitemata Group and the Plio-Pleistocene sediments at depths ofapproximately down to 100 m below the eruption paleosurface(Fig. 8d). Whether these deeper explosions were the result of ventshifting is not clear, but the presence of coarse fragments of Waitematarocks decreasing towards distal exposures (M4, Fig. 6d) suggests thatthe vent probably either did not migrate or migration was negligible.Regardless of the locus of these explosions, they were accompanied bygreater slumping of crater wall material (the water-saturated, unconsoli-dated Plio-Pleistocene sediments were readily removed) filling the ventwith liquefied sediments (e.g., Sohn and Park, 2005) (Fig. 8c). This situa-tion prevailed with further shallow seated explosions (Fig. 8e) after theevent marked by the excavation of the Waitemata rocks.

Table 2Nomenclature, characteristics and interpretations of lithofacies assigned for the Maungataketake phreatomagmatic deposits. The definitions (with their references) for the interpretations on transport processes from bedform sedimentarycharacteristics are explained in Section 5.1. The citations supporting the mode of transport processes are omitted in the table if they are expressed in Section 5.1. Photographs related to the depiction of lithofacies are referred in the table andcontained in Figs. 6 (general view) and 7 (closer view).

Lithofacies (thickness (m)/accidentalto juvenile ratio (based on vol.%)

Grain size Internal structure Interpretation

Accidental-fragment dominated,poorly consolidated, matrixsupported tuff (T)

T1 (Figs. 6c, b, e, 7a, b, f)0.2 to 0.5 m/9:1

Fine, grey ash (~up to 55 vol.%) and medium ash(~up to 35 vol.%) of accidental origin. Small block/medium lapillus-size fragments of accidental originare relatively scarce (b5 vol.%). Coarse ash/lapillijuvenile fragments.

Beds with subtle low-angle, discontinuous cross-laminationand occasional dune bedding, containing some accretionarylapilli. It includes continuous, parallel/sub-parallel trains/bands composed of juvenile coarse ash/lapilli.

Soft deformation, and the presence of accretionary lapilli indicateliquid water during transport and deposition (T1, T3, T7).Trains of coarse juvenile fragments suggest that transport involvedcollision and saltation of fragments (bed load mode of transport).(T1, T3, T6, T7).The presence of cross-lamination and cross-bedding point toinvolvement of transport in the turbulent regime. (T1, T2, T3, T7).T2 and, to some extent T4, may represent the deposition of aprogressive aggrading flow that has a more-concentrated basalsegment (“bipartite base surge”; Sohn, 1997; Sohn and Chough,1989; Chough and Sohn, 1990; Vazquez and Ort, 2006).Crudely bedded, diffuse stratification dominated by frictional tocollisional and saltation modes where turbulence was not relevant.(T4, T5).Overall plane parallel bedding suggest the dampening ofturbulence. (T6, T7).Diffuse boundaries point to deposition of rapidly decelerating,progressive aggrading currents (Dellino et al., 1990). (T4).LT2 and T2 in distal outcrops (Fig. 6d, e) shows undulations likelydue to irregular surface at the time of deposition.

T2 (Figs. 6a, b, c, d, e, 7a, b, c, d, e, f)0.1to 0.15 m/9:1

Fine, grey to light brown ash (~up to 55 vol.%)and medium ash (~up to 35 vol.%) of accidentalorigin. No accidental clasts of lapillus/block size.

Beds containing subtle cross-lamination, contain rare accretionarylapilli. Juvenile trains (coarse to medium ash) are subtle, absent,or discontinuous.

T3 (Figs. 6a, c, 7a, b, d, g)0.2 to 0.5 m/9:1

Fine, grey to light brown ash (~up to 55 vol.%) andmedium ash (up to 40 vol.%) of accidental origin.It may contain coarse accidental blocks (up to 40 cm)(b10 vol.%).

Beds with laminations that pinch and swell laterally and areoften truncated by overlying laminations. Cross-laminationand twisted trains composed of juvenile coarse ash arecommon. Pervasive soft deformation. Accretionary lapilli isalso present

T4 (Figs. 6c, 7c)0.5 m/8:2

Fine, light brown ash (~up to 40 vol.%) and mediumash (~up to 30 vol.%) of accidental origin.

Planar beds composed of segments of accidental-rich ashalternating with sections containing diffuse juvenile coarse ash(30–40 vol.%). Diffuse boundaries. Contain some accretionary lapilli.

T5 (Figs. 6b, c, 7c, e)0.3 to 0.5 m/7:3

~70 vol.% of juvenile clasts with grain sizes rangingfrom fine lapilli (4 mm) to coarse ash (1 mm to0.3 mm) embedded in a matrix composed of fine tomedium-sized, light brown ash of accidental origin(up to ~30 vol.%).

Crudely bedded which contain diffusely distributed juvenilefragments. Beds are not affected by impact structures.Accretionary lapilli are not observed.

T6 (Figs. 6d, 7d, h)1.0 m/6:4

Fine-to-medium sized, light brown ash of accidentalorigin (40 vol.%) that alternate with sections thatconsist of medium-to-coarse, juvenile ash.

Planar beds that consists of layers of fine grained ash ofaccidental origin alternating with juvenile-rich layers arrangedin trains or thin beds of irregular thickness (from few mm toup to 3 cm). Subtle soft deformation and abundant accretionarylapilli.

T7 (Figs. 6a, d, h)0.5 to 1.5 m/9:1

Fine-to-medium sized, light brown ash of accidentalorigin (up to 70 vol.%). b30 vol.% juvenile coarse ash(b1 mm).

Planar beds with subtle low-angle, discontinuous cross-lamination and sub-parallel trains of juvenile coarse ash. b5 mm-diameter, spherical/oval accretionary lapilli are ubiquitous.Virtually devoid of lapillus and block size fragments.

Juvenile-fragment dominated,consolidated, clast supportedlapilli tuff (LT)

LT1 (Figs. 6c, d, e, 7a, b, d, f)0.1 to 0.15 m/2:8

N80 vol.% of coarse ash to fine lapilli juvenilefragments. Ash to block size accidental fragmentsusually comprise less than 10 vol.%.

b15 cm beds that are sub-parallel and/or slightly undulating,laterally continuous. Normal or reverse graded, and may showscour surfaces or concave structures due to load of overlyingdenser, juvenile material (bed load structures).

Overall massive deposits or variable grading suggest rapidprogressive sedimentation of a more concentrate base portion(LT1 or LT2) of a base surge that has an upper, diluted, finer-grainedsection (T2) or a single flow related to a collapsing jet (LT3) (RossandWhite, 2006). Fall out may have occurred, but was neverdominant.A frictional mode of transport (lower traction carpet) whereturbulence is practically absent may be dominant. At times, the basalboundary layer was characterized by an erosive, high shear stress.(LT1, LT2, LT3).LT2 lithofacies was affected by intense soft deformation.The palagonitization and armoured/accretionary lapilli shown inLT3 may suggest that the flow contained liquid water.

LT2 (Figs. 6a, c, 7a, b, d, g)0.05 to 0.4 m/2:8

Beds are ungraded, normal or reverse graded. Intense softdeformation is evident and beds can be affected by impacts ofblocks of accidental origin (up to 40 cm).

LT3 (Fig. 6b, e)0.5 m/1:9

Ungraded, normal or reverse graded, and may show load structuresat the base. Diffuse stratification. Fine-grained, light coloured ashirregularly distributed as sub-horizontal, sub-parallel, and laterallycontinuous laminae (b1 cm thick). Scarce accretionary/armoured lapilli.

53J.A

gustín-Floresetal./JournalofV

olcanologyand

Geotherm

alResearch276

(2014)46

–63

Table 3Nomenclature, characteristics, and interpretation of lithostratigraphic units for theMaungataketake phreatomagmatic deposits. For a better understanding of this table, it is recommended to turn to Figs. 2, 3 and 6 in order to get a better understandingof the context regarding the spatial distribution of sites, lithofacies, and units.

Phase Litho-stratigraphicunits

1. Associated Lithofacies2. Average thickness/approximate

accidental to juvenile ratio3. Boundary contact with lower unit4. Dominant bedform facies

Interpretation of units

Proximal (sites M1, M2) Medial (site M3) Distal (sites M4, M5)

PH1 (vent opening/shallow explosions)

U1(Figs. 3, 6c, e)

Not present 1. T1/LT12. 1.5 m/9:13. Sharp and irregular,

moderately erosional4. Generally massive, plane-

parallel/cross laminated.Few impact sags

1. T12. 0.5 m/9:13. As previous4. Generally massive

Vent opening explosions generated fine-grained ash-laden basesurges. The flows were wet, either due to condensation of steamand/or original sediment pore water. Ash particles were cohesive,forming accretionary lapilli. The base surges rapidly deceleratedwith distance (cf., Wohletz and Sheridan, 1979; Zimanowski andWohletz, 2000; Sulpizio and Dellino, 2008).

U2(Figs. 3, 6c,d, e, 7a, b, d)

1. LT2/T22. b0.5 m/5:43. Not visible4. General plane-parallel/

cross laminated

1. LT1/T22. 0.5 m/5:43. Sharp with scour surfaces

and/or load structures4. General parallel-slightly

undulating bedded, plane-parallel/cross laminated tuff

1. LT1/T22. 0.5 m/5:43. Sharp to gradational4. Undulating bedding

The presence of a “bipartite base surge” (see Table 2 fordefinition) with a lower coarser-grained bed (LT1) overlain by afiner-grained layer (T2) suggests that U2 was formed by a seriesof density stratified base surges (e.g. Valentine, 1987). Eachcurrent produced a deposit couplet (cf., Vazquez and Ort, 2006).With distance from the vent, the flows became relatively wetterand topographically controlled.

PH2 (deeperexcavation)

U3(Figs. 3, 6a, c, d)

1. LT2/T32. Variable (usually b0.8 m)3. Gradational4. Laminated (intense soft

deformation). Crosslamination

1. LT2/T32. 0.7 m/7:33. Gradational4. Intense soft deformation of

laminated beds. Visible andcommon impact bedding sags

1. LT1/T62. 0.5–0.7 m/7:33. Gradational4. Parallel-slightly undulating,

laminated tuff

This phase starts with lithofacies T3 deposits (sites M1 and M3,Fig. 6a, c), indicating base surges with an unsteady and pulsatorybehaviour (cf., Sulpizio and Dellino, 2008). This was followed bydeposition of LT2 along with the first appearance of ballisticblocks derived from Waitemata Group lithologies (Figs. 6c and7a, b), which indicates a deepening locus of explosions and/orshifting of vent.

PH3 (shallowseated explosions)

U4(Figs. 3, 6b, c, d, 7c, d)

1. T32. ~0.5 m/8:23. Gradational4. Laminated (intense soft

deformation). Crosslamination

1. T42. 0.5 m/8:23. Gradational4. Plane-parallel/diffusely

bedded. The fine tuff iscross-laminated. Devoid ofimpact structures

1. T6/T12. 0.3 m/6:43. Gradational4. Plane-parallel bedded/

laminated. Devoid ofimpact structures

The eruption proceeded without interruption as suggested fromthe gradational contact from U3 to U4, depositing a rhythmicsequence of base surges (T4). With distance, the currents rapidlydeflated, reduced in turbulence and particle load, which arerepresented in lithofacies T6 (Fig. 7a). The absence of WaitemataGroup block/lapilli-sized accidentals may suggest that the depositedmaterial did not originate from freshly excavatedWaitemata hostrock (Valentine andWhite, 2012).

U5(Figs. 3, 6b, c, 7c, d)

1. LT3/T2/T42. 1.0 m/2:83. Sharp slightly erosional

with load structures4. General crudely bedded.

Plane-parallel/crosslaminations in T2

1. T5/T22. 0.7 m/3:73. Sharp to gradational4. Diffuse bedded/massive.

Devoid of impact structures

1. T62. b0.5 m/4:63. Gradational4. Plane-parallel bedded,

laminated. Devoidof impact structures

LT3 shows that the eruption shifted suddenly to a phase ofproduction of coarser and more common juvenile fragments,either due to a reduction of external water interaction or anincrease in magma ascent rate (cf., Houghton et al., 1999). Anoutstanding feature of U5 is the progressive transformation ofsediment character from proximal to distal sites (Fig. 6b, c, d)which may indicate the deposition of a collapsing jet that wastransformed into a more diluted density current.

PH4 (vent stabilisationand waning)

U6(Figs. 3, 6a, b, d, 7d)

1. T7/T52. Up to 1.5 m/9:13. Sharp4. Parallel-undulating bedded,

cross-laminated. Devoid ofimpact bedding sags

1. T7/T52. b0.5 m/9:13. Gradational4. Plane-parallel bedded, cross-

laminated. Devoid of impactbedding sags. Abundantaccretionary lapilli

1. T7/T52. 1.5 m/9:13. Gradational4. Plane-parallel bedded, cross-

laminated. Devoid of impactbedding sags. Abundantaccretionary lapilli

U6 is characterized by lithofacies T7 and T6 (Fig. 6a, d), which showsa regular pattern of deposition from PDCs. In general, the successionexhibits regular deposition patterns that suggest stabilisation of thevent with explosions remaining in the water-saturated poorlyunconsolidated sediments as Plio-Pleistocene sediments fragmentform the bulk of the deposit.

54J.A

gustín-Floresetal./JournalofV

olcanologyand

Geotherm

alResearch276

(2014)46

–63

Fig. 6. Photographs and corresponding logs for key sitesM1 toM5 (see Fig. 2 for location of sites of sites and Fig. 3 for log correlation). Solid red lines indicate the boundary between units (U)(see description and interpretation of units in Table 3). Uncertain boundaries are marked by dashed red lines. Some accidental blocks are delineated in dotted-fine black lines, as in photo-graphs a) and c). Accretionary lapilli is found in all locations concentrated along thin horizons and otherwise scattered within the tuff. Lithofacies outlined in photographs are explained inTable 2 and shown with more detail in Fig. 7. PS in c) and e) represents the peaty soil underlying the phreatomagmatic deposits. The number followed by “aht” at the base of each log rep-resentsmetres above high tide (the differency betweenhighand low tide is approximately 3m)with high tide levels roughly coincidingwith the lowermost part of the tuff deposits at siteM2.

55J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

6.3. Phase 3. Shallow seated explosions

As indicated by the gradational boundaries between upper U3 andlower U4 (Fig. 7a, b), the shallow explosions continued without inter-ruption (Fig. 8e), promoting the deposition of U4. The evidence of thiscan be seen in the sedimentary characteristics of deposits U4 (Table 3)showing the absence ofWaitemata-pebble fragments and the rhythmicsuccession of deposits from low-energy base surges (lithofacies T4,Table 2; Figs. 6c, 7c). The change from lithofacies T4 to lithofacies T6with distance (Table 3, Fig. 6c, d) suggests decreasing flow energy(changes in the shear stress rate of the flow).

The shallow seated explosionswent on and the conditions promotedthe generation of a juvenile-rich phase that deposited U5. The presenceof massive/crudely bedded, coarser grained lithofacies (LT3 and T5 atM2 and M3 respectively; Fig. 6b, c) indicates a less energetic phasewith shallower explosion locus where the involvement of accidentaldisaggregated sediments was less important. On the other hand, thelateral sedimentary changes in U5 from proximal to distal sites(coarse-grained, crudely bedded to relatively finer-grained, diffuse bed-ded) (Table 3, Figs. 6c, b, d, 7c, d) do not typically represent a typicalbase surge, instead U5 deposits at M2 (Fig. 6b) may have been relatedto a concentrated collapsing jet (Ross andWhite, 2006) that became di-luted with distance (see changes of lithofacies within U5 from Fig. 6b toc, d). Although the sharp boundary between U4 and U5 at site M2(Fig. 6b) may indicate a break, the gradational boundaries observed atM4 (Figs. 6d, 7d) between these same units are the evidence of thelack of pauses during this eruption phase. The same reasoning is appli-cable for the transition from U5 to U6 (Figs. 6b, d, 7d). Therefore thewhole sequence shows continuous activity with no time breaks (or atleast time breaks that were not long enough to exhibit clear evidenceof elapsing time as described at other sites by Sohn and Park, 2005).

6.4. Phase 4. Vent stabilisation and waning of eruption

The uppermost unit (U6) is a fine-grained, plane-parallel bedded,regular sequence devoid of impact sags (lithofacies T7, Table 2). We

infer that the lack of ballistic blocks and the rhythmic sequence oflithofacies T7 indicate a stabilisation of the vent (Fig. 8e) (cf., Némethet al., 2012). The prevalence of fragments from the Plio-Pleistocene sed-iments suggests that the explosions remained shallow seated wherewater was abundant, which is attested by the ubiquitous accretionarylapilli and the presence of vesicular tuff beds. The phreatomagmaticphase was followed by construction of a complex scoria cone andshort lava flows (Fig. 8f). At site M1 there is evidence (not shown) ofan erosional unconformity between the phreatomagmatic depositsand the scoria deposit that may suggest some time gap between thephreatomagmatic and magmatic phases, but the lack of more contactexposures makes it difficult to infer the time that elapsed betweenthem. Geochemically, however, both eruptive phases seem to belongto a single magma batch with rock compositions that show trendsconsistent with differentiation (see Section 7).

7. Maungataketake whole rock and glass chemistry

The whole rock chemical composition shown by Maungataketakesamples (16 from the scoria cone; 5 of the rare coarse lapilli ballisticsfrom the tuff ring) falls toward the low-silica high alkali end of thespectrum of compositions observed from the AVF (McGee et al.,2011). They are low SiO2 (41–45 wt.% loss free) alkali basalts with in-termediateMg-numbers (62–65) (Fig. 9a, b, Table A.1). On variation di-agrams (Fig. 9a, b), Maungataketake basalts show trends that areconsistent with a differentiation process. All major elements exceptAl2O3, but including SiO2, show negative trends with MgO (and Mg#)(Fig. 9a, b). Similarly mantle compatible trace elements (Ni, Cr) showpositive trends with MgO, whereas incompatible trace elements (LILE,HFSE) show a negative trend. Samples from the tuff ring fall towardthe relatively evolved (low MgO, higher SiO2, high alkalis and incom-patible trace elements) end of the compositional spectrum, but thereis not a clear correlation observed between stratigraphic position ashas been observed in other volcanoes of the Auckland Volcanic Field(Crater Hill, Smith et al., 2008; and Motukorea, McGee et al., 2012).

Samples used for microprobe analyses (glass and crystal composi-tions) come from throughout the stratigraphic sequence, but do not

Fig. 7. Photographs showing the lithofacies (for the descriptions and interpretations see Table 2. Lithofacies are also shown in Fig. 6) and boundaries between units inmore detail. a) Photograph offallen block displaying part of the stratigraphic sequence exposed as in photograph b). Both photographs represent siteM3 (Figs. 3, 6c) and are roughly at the same scale (measuring tape is 1m).Red arrows indicate the boundaries between units (see Table 3 for description and interpretation of units). The uppermost part of unit 1 (U1) showsmore concetrated trails of juvenile fragments,sometimes exhibiting impact sags or bedload structures. Note in U2 the upward thinning of beds containing lithofacies LT1; also their grain size diminishes in the same direction. At the contactbetweenU1andU2 it is possible to observe scourmarks and irregular base of beds. Soft sediment deformation is distinct in themiddle sectionofU3. Thewhite-outlined squarewithin a) highlightslithofacies T1. W and PP indicates blocks of Waitemata rock and block aggregates of Plio-Pleistocene sediments respectively. c) Plane parallel bedding is evident, but juvenile-rich beds (darkerlayers) form diffuse bedding as the transition to more accidental rich tuff is gradual. This photograph is part of site M3 (Fig. 6) and the two white-outlined squares highlight lithofacies T4 andT5. d) Photograph of a site located a few metres downflow direction from site M4 (Figs. 3, 6d). Red arrows indicate the boundaries between units. From the middle section of U3 upwards thewhole sequence exhibits plane-parallel beds with distinct, but subtle aggradation. e) Detail of lithofacies LT3, T2, and T5 (Table 2) within U5 at site M2 (Figs. 3, 6c). Note the load structures ateh base of LT3. f) and g) Detail of lithofacies LT1, LT2, T2, and T3 (Table 2) within U2 at site M3 (Figs. 3 and 6). Coin in f) is 2.6 cm and measuring tape in g) is 20 cm. h) Detail of lithofacies T6and T7 (Table 2) within U4, U5, and U6 at M4 (Figs. 3 and 6).

56 J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

Fig. 8. Cartoons that represent a simplifiedmodel of theMaungataketake eruptionhistory.Not to scale; the vertical scale shown in the cartoons is displayed only for approximate thickness referenceof the existing lithologies. The orientation of the section is roughly NNE–SSW intersecting at site M2 (Fig. 2). For a more detailed sequence of the eruption see Section 6 in the text. a) Conditionsprevious to the vent opening and initiation of phreatomagmatic explosions at depth as a consequence of dike emplacement. Note that there are two forests on the surface. b) The vent opens(phase 2) and forest 2 is obliterated and buriedwithin the lower deposits of U1. The fallen and buried trees seen on the right section of the vent are speculative as their presence cannot be verifiedin the field. c) Slumping of crater walls into the vent and promotion of shallow seated explosions (continuation of phase 2 and deposition of U1–U2). d) Excavation into theWaitemata rocks andejection of blocks. Generated base surges are associated with important amounts of water content. e) Filling the vent with slumpedmaterial from the crater walls (dominant Plio-Pleistocene sed-iments).Overall prevalenceof shallowseatedexplosions that characterizephases3and4. Culminationof thephreatomagmatic activity atMaungataketakewith thedepositionofU6. f) Shifting to themagmatic phase of the eruption. g) Current conditions atMaungataketakewhere the scoria cone has been extensively quarried and thewestern section of the ruff ring has been eroded by the sea.

57J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

58 J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

correspond directly to the samples representing whole-rock composition(see Appendix A for details). The glass shard compositions tend to bemore SiO2-rich than the whole-rock samples; CaO shows a positive trendwith MgO (Fig. 9c), but shows a wider range than MgO. TiO2 (Fig. 9d)and FeO describe a positive trendwithMgO, indicating late stage crystalli-zation of oxides. A high rate of nucleation (on shifting the liquidus tohigher temperatures) is triggered by the exsolution of the last H2O fromthe melt in the very shallow conduit (e.g., Schipper et al., 2010). Microlitecomposition is shown in Fig. 9e, and indicates Ca-rich pyroxene.

8. Discussion

8.1. Magma fragmentation and host rock disruption

Sedimentary features such as accretionary lapilli, soft-sedimentdeformation, vesiculated tuff, and plastering of fine ash onto obsta-cles throughout the deposit sequence point to deposition mainlyfrom “wet” base surges. Juvenile pyroclasts show particle coatingsor adhering fine particles (e.g., Sheridan and Wohletz, 1983; Cioni

Fig. 9. a) and b) Major element (wt.%) variation diagrams (whole rock). For comparison, the fieHill (broken line) (Smith et al., 2008) volcanoes is indicated (see Fig. 1 for location of volcanosamples (see Appendix A for details of samples) listed in an ascending stratigraphic order (froof measurements of pyroxene crystal (microlite/microphenocrysts) compositions. The data inc

et al., 1992) and include sideromelane glass (Fisher and Schmincke,1984) (Fig. 8c, i) (Table 1). These all indicate phreatomagmatic frag-mentation. Although there is an absence of non-vesicular, equant,blocky juvenile fragments (2–3 ϕ) regarded (from experimentaldata) to be representative of interactive particles resulting from FCIfragmentation (cf., Sheridan and Wohletz, 1983; Wohletz, 1983;Zimanowski et al., 1997), this rationale may not be satisfactorily ap-plicable in natural cases where vesiculated fragments may be alsothe result of FCI (e.g., Cioni et al., 1992). The majority (N80 vol.%)of juvenile particles throughout the sequence are 0–2 ϕ in size andare equant and subangular, microlite-rich and poorly vesicular(Table 1; Figs. 4a, 5c, e, f, g). These characteristics could indicate gen-erally low-energy FCIs with high water–magma (w/m) mass ratios(cf., Sheridan and Wohletz, 1983; Wohletz and McQueen, 1984).Also, the juveniles may be the result of “secondary fragmentation” asthe result of disruption of surroundingmagma by shockwaves producedin the MFCI processes (Zimanowski et al., 1991; Raue, 2004). Additionally,recycling of juvenile andnon-juvenile clastsmayhaveoccurred (Houghtonand Smith, 1993), contributing to the homogenous morphology.

ld occupied by sample data fromMotukorea (dotted line) (McGee et al., 2012) and Crateres). c) and d) major element (wt.%) variation diagrams of glass compositions of selectedm U3 to U6), and from proximal to distal locations (form site M1 to site M5). e) Plottinglude microprobe measurements on inner, central, and outer crystal spots.

59J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

Raue (2004) reports that N60% of thermal energy of magma in sucheruptions is converted into shock waves that may disrupt surroundingmagma and/or country rock. At Maungataketake, the morphological,stratigraphic, and sedimentary data suggest that the vent opening ex-plosions occurred at shallow depths (around 60–100 m), primarilywithin Pleistocene–Holocene Tauranga Group and Kaawa Formationsediments. Part of themechanical energy produced in the FCI was prob-ably expended in the excavation of this soft host material. Due to theunconsolidated nature of the capping sediments, slumping and refillingof the vent area by the saturated and unconsolidated deposits, appearsto have been common atMaungataketake,with periodic replenishmentof fine sediment and groundwater to the eruption site. The fact that ac-cidental ash (which comprises up to 80 vol.% of the deposit) is princi-pally composed of Plio-Pleistocene sedimentary grains and grain sizedistribution of the tuff is constant laterally and vertically (Fig. 3) sug-gests the primary involvement of the unconsolidated Tauranga Groupand the Kaawa Formation sediments. Other than the rare coarse, round-ed, Waitemata rock fragments in U3 (Table 3, Section 6.2), there is noclear evidence of the presence of this lithology in the tuff. However,downward disruption of the Waitemata Group rock units into adiatreme cannot be excluded.

8.2. Water availability within the host material

The explosive interaction ofmagmawith water is themost commonmechanism by which tuff rings are formed (Chough and Sohn, 1990;White, 1990). In many locations the volcanic edifices related tophreatomagmatic activity are located at coastal margins where shallowsea water or groundwater is present [e.g., Jeju Island, South Korea(Sohn, 1996); Marion and Prince Edward Islands (Verwoerd andChevallier, 1989); Ambae Island, Vanuatu (Németh and Cronin,2009)], whereas scoria cones are formed on higher and drier groundinland. In the AVF, phreatomagmatic activity is characteristic of thelow-lying, coastal areas such as theManukau Lowlands. The thicknessesof the Tauranga Group, Kaawa Formation, and Waitemata Group areonly inferred by extrapolations from borehole data. Kaawa Formationsediments could even be quite thin (b10 m) (Viljevac et al., 2002), butthese are highly controlled by an irregular paleosurface and are farthicker in paleovalleys (Edbrooke et al., 2003). Kaawa Fm. sedimentshave excellent conditions for water storage and good hydraulic proper-ties (average hydraulic conductivity, K = ~10−5). The Tauranga Groupsediments (~30–50 m thick) are almost impermeable and host onlylow-yield bores, although they are water-saturated (Kermode, 1992).The underlying Waitemata sediments have lower hydraulic conduc-tivity, but may still hold water forming a heterogeneous aquifer (forexample, a 200 m depth, 90–100 mm-diameter well can yield30–300 m3/day; Crowcroft and Bowden, 2002). The combined aqui-fers extend to at least 500 m depth (Crowcroft and Bowden, 2002). Asdescribed above, the sea levels and water tables in the area at the timeof the eruptionwere similar to the present ones (currently, groundwateris found ~2–6m below the surface). Although quantitative data of wateravailability is lacking, it appears that water from the unconsolidated sed-iments and the Kaawa aquifers was sufficient to promote the first FCIsduring the vent opening. As shallow-seated explosions were dominantin the construction of the tuff ring, this shallow sediment was obviouslystill able to yield enough water to sustain subsequent FCI explosions. Acomplex interaction between magma and water-laden sediments (orsediment-laden water) (cf., White, 1996) thus took place during mostof the course of the phreatomagmatic eruption.

8.3. Unconsolidated water-saturated sediments and FCI

As stated in the past Section 8.2, it is likely that soft host sedimentwas mainly involved throughout the eruption. In such cases, condi-tions for the explosive mixing of water, unconsolidated sedimentand magma are more complex (White, 1996; Schipper et al., 2011).

The conventional view (from experimental results) is that a largefine-ash fraction is produced at an optimal water/magma massratio of ~0.3 (Sheridan and Wohletz, 1983; Wohletz, 1983, 1986).Higher or lower w/m ratios mostly generate overall coarser juvenilefragments (e.g., Dellino et al., 1990; Brand and Clarke, 2009). This isonly valid for simple FCI models for volcanic eruptions (White,1996).White (1996) presents amodelwhere the explosivity ofw/m in-teractions depends strongly on the physical properties of the coolant(which is a mixture of sediment and water) rather than w/m mass ra-tios alone. Using stratigraphic, sedimentary, and hydrogeological datain the reconstruction of four phreatomagmatic volcanoes, Sohn(1996) arrives at a similar conclusion. Intrusion of magma in water-saturated and unconsolidated sediments often generates explosivewater–magma interaction, despite initial inhibition. Once started,sediment-laden coolants may promote MFCI explosivity by increasingthe availability of nucleation sites (White, 1996).

The opening of the vent of the Maungataketake phreatomagmaticeruption probably occurred 60–100 m below the pre-eruptive surface,where magmamixed with pore water and sediment. Slumping and liq-uefaction of the unconsolidated host materials repeatedly clogged thevent with more water-saturated sediment, which would have alsoserved tomaintain a confining hydrostatic pressure. Shallow explosionsrelated to the emplacement ofmultiple dykes, dyke branches, small sillsor arrested jets ofmagmamay havewidened the area ofmagma contactwith water saturated sediments (as in the model by White, 1996). Alarge-sub-volcanic presence of cooled magma indicated by geophysicalstudies (Cassidy and Locke, 2010) shows that significant sills or lacolithsdeveloped. Deeper explosions may have been promoted by water re-leased from the Kaawa–Waitemata sediments plus also possibly fromtheWaitemata aquifers enclosing the deeper conduit. The boundary be-tween consolidated (Waitemata Group) and unconsolidated sediments(Plio-Pleistocene sediments) may have caused the least principal stresson amagma dyke to deviate towards a vertical position (e.g., Lorenz andHaneke, 2004), generating possible dyke splays, or temporary pondingin sills. The homogeneity in both the textural characteristics (Table 1)and the composition of juvenile (Appendix A) may indicate shallowstalling. This is suspected in other similar settings based on exposeddyke and sill complexes in association with preserved diatremes (e.g.,Martin and Németh, 2007). As a consequence, further complex geome-try of contact betweenmagma and saturated sedimentwould be gener-ated, causing many areas with differing magma–coolant contact anddiffering w/m ratios.

Observations made here point to the fact that w/m ratios need to beconsidered with caution. For example, U3 (Table 3) (with intense softdeformation) has features that could be interpreted as higher levels ofwater involved in the phreatomagmatic explosions, compared to thoseof units 1, 2, 4, and 6 (Table 3), when in reality this just reflects thewater contained at the moment of deposition. This does not necessarilyreflect the w/m ratios in the FCIs involved, but could be due towater di-rectly ejected from the vent area without being vaporized (cf., White,1996) and added to the generated base surges. Water that never tookpart directly in a FCI could have been transported with the excavatedwater-saturated substrate during the course of the phreatomagmaticeruption. The wet sedimentary features of the tuffs indicate that basesurge temperatures during this eruption were low and often locallybelow boiling point. The Maungataketake eruption was very stable,with only subtle transitions, mainly related to minor changes in thedepth of shallow-seated (mostly b100 m) explosions.

8.4. Duration and waning of the phreatomagmatic eruption

The broadly uniformmineralogical assemblages and textural charac-teristics, the absence of evidence for time breaks in the depositionalrecord (Section 6), and the overall small volume of the tuff ring suggesta brief eruption. The ~1000 m-diameter, low rim tuff ring compareswell with historic maar forming-eruptions documented, e.g., the

60 J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

Ukinrek maars between 0.5 and b11 days (Kienle et al., 1980); Taala few days (Moore et al., 1966); and Ambrym AD 1913 eruptionb12-hour growth, with 4 additional days of activity (Németh andCronin, 2011). Based on the eruptive volumes of the AVF tuff rings(Kereszturi et al., 2013), the eruption could have lasted a maximum of16 days, based on an average eruption rate of 5 m3/s. Therefore,Maungataketake volcano was likely formed over only a few days.

The rapid deceleration and loss of turbulence and energy of the basesurges is evidenced by the presence of the general plane-parallel bed-ding in the distal deposits (cf., Valentine and Fisher, 2000) (Table 2).Wet ash particles often aggregate, accelerating their deposition anddamping turbulence (Branney and Kokelaar, 2002). Even if the depositreflects only a wet condition of the flow-boundary zone, the constanteruptions through water-saturated substrate (Tauranga Group andKaawa Formation) could have always provided water to the flows (acontinuous three-phase system).

It is not clear whether the transition to subsequent “dry” activity atMaungataketake was associated with a break in eruptive activity. How-ever, the entire Maungataketake eruption was fed from a single magmabatch. In any case, the last phreatomagmatic unit (U6) (Table 3) indi-cates that the vent had become very stable. It seems hardly likely thatwater supply to the vent area ceased. Thus, themain reason for thewan-ing of the phreatomagmatic eruption was probably due to a rise inmagma eruption rate, and/or the isolation of the rising magma conduitfrom the substrate by a lava moat or similar.

9. Conclusion

Maungataketake volcano in the southern part of the AucklandVolcanic Field was formed by a distinctive type of phreatomagmaticeruption that occurs in saturated, shallow, deformable and fine-grained sediments. This is characteristic of low gradient alluvial basins,deltas and coastal margins, where the water table is high, and fine-grained porous unconsolidated sediments may be hundreds of metresdeep. The Maungataketake phreatomagmatic explosions were domi-nantly shallow (b100 m depth) and associated with an abundance ofwater and liquefiable sediment. As the eruption progressed, magmarose into this deformable material, spreading laterally and developingcomplex contact geometries with the saturated host deposits. Low-energy FCI explosions excavated the wet sediments forming base surgesthat travelled up to ~1 km from the ventwith rapidly outwardly decreas-ing turbulence and velocities. A tuff ring was thus built by the depositionfroma series of relativelywet, highly pulsating, density stratified,moder-ately turbulent base surges. Sedimentary and pyroclast features exhibitsubtle changes either laterally or vertically. The homogeneity of the acci-dental composition (Plio-Pleistocene sediments) and the gradationalvertical changes through the sequence appear to indicate excavationand refilling of the vent area with water-saturated sediments via wallcollapse and inward flow of liquefied unconsolidated sediment.

The contrast in mechanical and hydrological properties of the hostmaterial at the boundary between the consolidated Waitemata Groupand the deformable Pleistocene sediments generated conditions formagma to spread from a simple dyke geometry and form sill structuresand a complex contact betweenmagma and saturated sediments beforeand after the vent opening. Water was available in excess, and shallowseated explosions occurred at locations differing slightly in depth andlocation, forming an irregular tuff ring.

This phase of the Maungataketake eruption was likely brief, lastingonly hours to days, and it was followed by a phase of scoria and lavaflow production only after the magma eruption rates rose significantly,or water was blocked from the conduit. This example demonstrates oneof the lower-hazard variants of explosive phreatomagmatic eruption,contrasting with those occurring with deeper and more focussed FCIexplosions, such as those within theWaitemata Group rocks elsewherein the AVF (e.g., Motukorea volcano, McGee et al., 2012). Thusphreatomagmatic hazard models need to be more closely tailored to

the known features of the host geology in an area, e.g., with those occur-ring in deep, saturated loose sediments, such as low-energy coastal andfluvial settings, having a smaller radius of potential destruction.

Acknowledgements

JAF, SJC and KN are supported by the New Zealand Natural HazardsResearch Platform and JAF and JML by the DEVORA (DeterminingVolcanic Risk in Auckland) project.We also thank the School of Environ-ment and the Institute of Earth Sciences and Engineering at AucklandUniversity for support, aswell as Kate Arentsen for prompt and valuableassistance; Natalia Pardo and Anja Moebis for assistance in laboratorywork/discussions; Tracy Howe for assistance with the borehole data;Bob Stewart for carrying out the XRD analyses; Doug Hopcroft andCarlos Linares for technical assistance; and Jose Rivera, Marc Adamson,and Donald Hsieh for providing accommodation in Auckland. Wethank Rafaello Cioni, Claus Siebe, and two anonymous reviewers fortheir suggestions to improve this manuscript.

Appendix A. Chemical analyses

Whole-rock analyses

Samples fromMaungataketake Volcano analysed formajor and traceelements were collected from the tuff cone (3 samples) and the centralscoria/lava cone (16 samples) to represent the range of materialserupted from the volcano; these data are presented in Table A.1. Thesamples are juvenile bombs from within tuff sections, scoriaceousblocks and lava. Rock fragments were crushed in a tungsten carbidering grinder to b200 μ mesh. Major elements and some trace elementswere analysed by XRF on fused glass discs made using Lithium BorateSpectrachem 12–22 flux, using a Siemens SRS3000 sequential X-rayspectrometer with a Rh tube at the University of Auckland. Minortrace elements were measured on a Laser Ablation Inductively CoupledMass Spectrometer (LA-ICP-MS) at the Australian National Universityusing stacks of XRF discs following the procedure of Eggins et al.(1998). NIST 612 was run every 15 samples and used for calibration,and the silica content obtained by XRF used in data reduction. BCR-2Gwas used as an external standard with every analytical session and 28international standards provided a further check on the method. BCR-2 data (n = 143) is b15% 2SD, and accuracy b10% for all elementsexcept Cu, Y, Zr, Tb and Hf which are b17% and Cr which is b26%. XRFdata is reported for Cr, Ni and Zr.

Sideromelane glass/crystal analyses

10 juvenile grainswere selected (from thin sections) for microprobeanalyses. These grains represent 5 samples taken from different sites(see Fig. 3 in main text for site and unit locations): sample U3a (frommiddle U3 at site M1) (2 grains), sample U3b (from upper U3 at siteM1) (3 grains), sample U5 (frommiddle U5 at siteM2) (2 grains), sam-ple U6a (from lower U6 at site M5) (2 grains), and sample U6b (fromupper U6 at site M5) (1 grain). 6 spot measurements per grain wererun on different glass sites. An average glass composition of each grainis shown in Table A.2. These compositions have been plotted and areshown in Fig. 5c, d. Olivine phenocrysts (not shown) and Pyroxenemicrolites/microcrysts were also analysed and are represented inFig. 5e. 15 pyroxene crystals were analysed from the same 5 samplesmentioned above.

The samples were analysed by a Jeol JXA-8900R electron micro-probe at the Laboratorio Universitario de Petrología (LUPI), Institutode Geofísica, UNAM, México City. Measuring conditions were a beamcurrent of 10 nA, an accelerating potential of 20 kV, and a beamdiameter of 15 μm for glass and 1–5 μm for crystal analyses. Duringanalyses, Na and Kwere analysed using 10 s counting times, whereasa 40 s-counting-time was used for other elements.

Table A.1Whole rock major and trace element analyses for Maungataketake samples. Mg# calculated as mole percent Mg/Mg + Fe x100. SC = scoria cone sample; TR = tuff ring sample.

Type SC TR SC SC TR SC TR SC TR SC SC SC SC SC SC SC TR SC SC SC

Sample AVF-238 AVF-869 AVF-247 AVF-248 AVF-870 AVF-237 AVF-859 AVF-249 AVF-863 AVF-245 AVF-240 AVF-239 AVF-246 AVF-243 AVF-244 AVF-252 AVF-861 AVF-250 AVF-242 AVF-241

SiO2 41.80 42.07 42.23 42.30 42.41 42.50 42.61 42.94 43.18 43.72 43.73 43.75 43.88 44.33 44.37 44.40 44.50 44.51 44.55 44.65TiO2 2.88 2.91 2.85 2.84 2.83 2.84 2.90 2.88 2.73 2.82 2.80 2.80 2.76 2.69 2.69 2.67 2.60 2.66 2.70 2.69Al2O3 11.88 11.86 11.95 11.92 11.89 12.04 12.12 12.51 11.99 12.95 12.86 12.95 13.19 13.20 13.09 13.14 12.05 13.24 13.13 13.26Fe2O3 2.23 2.29 2.21 2.22 2.24 2.24 2.28 2.24 2.26 2.23 2.22 2.22 2.23 2.20 2.20 2.20 2.15 2.18 2.21 2.21FeO 11.17 11.43 11.04 11.11 11.19 11.21 11.39 11.21 11.32 11.14 11.09 11.09 11.16 10.99 11.00 10.99 10.74 10.92 11.03 11.03MnO 0.21 0.20 0.20 0.20 0.20 0.20 0.20 0.21 0.19 0.20 0.20 0.20 0.20 0.20 0.20 0.20 0.18 0.20 0.20 0.20MgO 12.38 11.66 11.75 11.81 11.68 11.37 11.92 11.05 11.33 10.70 10.78 10.25 10.32 10.27 10.31 10.46 10.43 10.21 10.53 10.31CaO 11.40 11.48 11.31 11.34 11.29 11.34 11.52 11.32 10.76 11.11 10.97 11.07 10.84 10.63 10.71 10.76 11.27 10.67 10.68 10.72Na2O 3.76 4.37 4.21 4.13 4.20 4.13 3.18 3.47 4.07 3.34 3.37 4.18 3.84 3.67 3.77 3.45 3.89 3.64 3.24 3.11K2O 1.44 0.92 1.45 1.33 1.29 1.31 1.08 1.38 1.39 1.06 1.25 0.79 0.93 1.19 1.01 1.12 1.45 1.15 1.11 1.18P2O5 0.84 0.82 0.79 0.79 0.79 0.82 0.80 0.79 0.77 0.73 0.72 0.71 0.67 0.64 0.63 0.62 0.73 0.62 0.63 0.65Mg# 66.37 64.52 65.48 65.45 65.05 64.38 65.10 63.71 64.08 63.12 63.39 62.21 62.23 62.47 62.56 62.90 63.37 62.50 62.97 62.49Cs 0.31 0.28 0.33 0.30 0.34 0.30 0.40 0.17 0.38 0.32 0.20 0.17 0.32 0.26 0.20 0.53 0.57 0.20 0.23 0.21Ba 264.9 309.8 310.4 309.9 331.0 314.9 336.1 226.1 309.6 268.8 235.4 266.7 314.9 229.6 271.7 317.6 310.9 223.5 233.6 234.9Rb 20.1 20.9 21.5 20.8 24.4 22.1 18.5 15.8 21.7 16.9 17.4 16.8 22.7 15.1 16.6 24.9 24.4 15.4 16.9 16.1Sr 680.6 771.6 743.0 727.0 764.3 745.6 704.8 613.4 672.0 677.7 621.7 673.6 748.0 615.8 700.6 678.1 671.7 617.8 619.1 621.8Pb 1.67 5.35 2.64 2.56 3.66 2.59 7.74 2.07 4.07 2.25 2.17 1.95 1.48 0.75 1.41 2.99 5.33 2.06 2.13 2.26Th 5.18 6.14 6.03 5.85 6.65 6.25 5.52 4.36 4.79 5.19 4.63 5.26 6.28 4.51 5.50 5.82 4.87 4.37 4.47 4.65U 1.44 1.66 1.61 1.64 1.70 1.70 1.76 1.16 1.53 1.51 1.36 1.46 1.69 1.29 1.42 1.59 1.85 1.24 1.27 1.26Zr 208.2 237.0 220.3 222.1 233.0 225.9 236.0 188.0 232.0 207.9 195.6 211.4 221.4 188.8 212.0 210.9 228.0 187.2 191.6 205.6Nb 63.7 71.4 74.7 72.5 75.3 76.1 72.6 53.2 62.5 64.4 56.3 65.1 75.7 54.8 65.9 66.0 59.7 53.1 55.5 55.9Hf 4.80 5.45 5.04 5.09 5.73 5.34 4.52 4.41 3.94 4.77 4.58 4.83 5.13 4.42 5.03 4.80 4.36 4.38 4.57 4.48Ta 3.97 4.65 4.60 4.55 5.01 4.73 4.28 3.29 3.76 3.99 3.56 4.19 4.71 3.40 4.10 4.20 3.56 3.34 3.47 3.53Sc 23.7 23.9 22.0 22.9 23.5 22.6 20.5 23.8 17.8 23.6 24.2 23.5 22.0 23.6 23.9 20.5 18.0 24.5 23.7 24.0V 286.1 241.0 296.3 292.8 234.0 289.9 238.0 280.5 220.0 298.8 283.4 279.1 275.1 275.0 276.1 272.7 213.0 284.5 268.6 210.5Cr 417.3 332.0 375.8 337.1 344.0 373.8 354.0 281.3 300.0 293.1 293.5 310.3 378.2 283.7 304.9 331.8 277.0 289.6 284.3 264.2Co 62.4 69.6 64.8 69.8 72.9 70.3 70.3 58.2 76.7 63.7 65.0 68.3 66.4 59.4 70.1 77.0 76.5 69.8 67.7 68.9Ni 185.6 300.0 235.1 213.3 308.0 218.4 318.0 193.8 317.0 205.8 194.1 197.8 237.7 196.7 208.4 240.7 306.0 199.5 196.2 228.4Cu 52.7 88.7 53.9 55.0 1489.9 54.6 60.1 53.1 57.4 60.6 50.1 46.9 28.8 66.9 38.7 51.7 57.9 52.1 54.9 52.6Zn 87.2 105.3 93.9 94.9 103.6 96.2 118.0 92.1 108.7 93.6 88.3 88.5 94.4 91.2 98.4 93.7 114.8 89.0 89.9 89.4Ga 39.1 37.4 41.6 41.6 38.6 42.4 30.5 34.7 29.2 39.0 36.5 38.8 41.9 34.7 39.8 41.3 28.6 34.6 35.7 35.8Y 22.5 24.3 22.3 23.4 24.4 22.9 19.9 22.0 17.9 22.8 22.6 22.7 22.7 21.9 23.0 21.4 18.8 22.0 22.0 22.3La 42.7 52.4 49.5 48.4 52.4 50.2 46.0 36.2 41.6 43.6 39.0 44.6 49.5 37.1 44.9 44.3 41.4 36.3 37.9 38.6Ce 82.6 97.5 93.5 92.2 96.8 94.9 96.5 71.9 87.8 83.5 75.8 84.8 93.7 73.2 85.7 85.0 86.6 71.4 74.1 74.6Pr 9.73 11.29 10.91 10.84 11.23 11.16 10.54 8.45 9.41 9.73 8.87 10.01 11.02 8.69 10.02 9.91 9.30 8.52 8.73 8.80Nd 39.76 45.67 44.80 44.26 46.19 45.66 40.02 34.95 37.76 40.04 36.80 40.23 44.41 35.67 41.54 40.36 36.83 35.09 36.28 36.49Sm 7.95 9.01 8.75 8.80 9.12 8.63 7.72 6.97 7.40 7.90 7.56 8.25 8.51 7.17 8.28 8.08 7.68 7.15 7.30 7.33Eu 2.52 2.78 2.75 2.70 2.74 2.79 2.59 2.26 2.41 2.54 2.36 2.62 2.68 2.33 2.64 2.48 2.33 2.37 2.38 2.32Gd 7.33 7.95 7.51 7.34 8.13 7.73 6.84 6.60 5.98 7.07 6.87 7.08 7.66 6.56 7.27 7.01 5.89 6.66 6.56 6.85Tb 0.98 1.05 0.99 1.02 1.04 1.04 0.87 0.93 0.75 1.02 0.96 1.00 1.01 0.93 1.02 0.93 0.82 0.93 0.92 0.94Dy 5.28 5.75 5.57 5.55 5.90 5.37 5.12 4.99 4.07 5.29 5.16 5.19 5.40 5.10 5.43 5.20 4.29 5.02 5.17 5.05Ho 0.94 0.96 0.90 0.92 0.99 0.95 0.77 0.87 0.70 0.94 0.91 0.91 0.91 0.87 0.92 0.87 0.75 0.88 0.89 0.88Er 2.20 2.23 2.13 2.19 2.42 2.15 1.91 2.23 1.70 2.25 2.22 2.22 2.17 2.12 2.32 2.06 1.65 2.21 2.13 2.24Tm 0.27 0.30 0.25 0.28 0.28 0.25 0.22 0.28 0.20 0.27 0.28 0.27 0.26 0.28 0.28 0.25 0.24 0.28 0.27 0.27Yb 1.63 1.62 1.53 1.60 1.75 1.55 1.31 1.71 1.18 1.66 1.66 1.68 1.44 1.69 1.66 1.41 1.40 1.66 1.69 1.64Lu 0.20 0.22 0.19 0.20 0.22 0.20 0.19 0.22 0.15 0.23 0.21 0.23 0.19 0.21 0.22 0.18 0.17 0.22 0.22 0.22

61J.A

gustín-Floresetal./JournalofV

olcanologyand

Geotherm

alResearch276

(2014)46

–63

Table A.2Average glass composition of selected juvenile clasts fromMaungataketake volcano.Mg# calculated asmole percentMg/Mg + Fe × 100. Each sample number (seeAppendix 2 for sampledetails) can be represented by one up to three single grains which are indicated by letters (a, b, c).

Sample 26a 26b 28a 28c 28d 31a 31a 40a 40b 41

SiO2 44.85 46.16 41.92 46.74 46.84 42.88 43.03 45.16 46.01 44.61TiO2 3.32 3.30 3.27 3.04 3.16 3.41 3.43 3.39 3.13 3.44Al2O3 15.75 15.90 15.56 16.44 16.23 15.48 15.48 15.49 16.11 15.58FeO 13.29 13.20 12.84 12.52 12.74 13.19 13.27 13.10 12.94 13.20MgO 4.05 4.10 4.01 3.78 4.06 4.21 4.22 4.20 3.85 4.33CaO 11.31 11.35 11.25 10.77 11.08 11.69 11.67 11.61 10.95 11.73Na2O 5.16 5.04 4.91 5.12 4.95 3.88 3.86 4.04 4.20 4.11K2O 1.93 1.93 1.86 1.99 1.93 1.92 1.87 1.89 1.89 1.89Total 99.67 100.98 95.63 100.40 100.98 96.67 96.82 98.88 99.09 98.89Mg# 35.22 35.63 35.77 36.21 35.01 36.28 36.20 36.38 35.22 36.90

62 J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

References

Allen, S.R., Smith, l.E.M., 1994. Eruption styles and volcanic hazard in the AucklandVolcanic Field, New Zealand. Geosci. Rep. Shizuoka Univ. 20, 5–14.

Alloway, B., Westgate, J., Pillans, B., Pearce, N., Newnham, R., Byrami, M., Aarburg, S., 2004.Stratigraphy, age and correlation of middle Pleistocene silicic tephras in the Aucklandregion, New Zealand: a prolific distal record of Taupo Volcanic Zone volcanism. N. Z.J. Geol. Geophys. 47, 447–479.

Auer, A., Martin, U., Németh, K., 2007. The Fekete-hegy (Balaton Highland Hungary) “soft-substrate” and “hard-substrate” maar volcanoes in an aligned volcanic complex —

implications for vent geometry, subsurface stratigraphy and the paleoenvironmentalsetting. J. Volcanol. Geoth. Res. 159, 225–245.

Ballance, P.F., 1974. An inter-arc flysch basin in northern New Zealand: Waitemata Group(Upper Oligocene to Lower Miocene). J. Geol. 82, 439–471.

Bebbington, M.S., Cronin, S.J., 2011. Spatio-temporal hazard estimation in the AucklandVolcanic Field, New Zealand, with a new event-order model. Bull. Volcanol. 73,55–72.

Brand, B.D., Clarke, A.B., 2009. The architecture, eruptive history, and evolution of theTable Rock Complex, Oregon: from a Surtseyan to an energetic maar eruption.J. Volcanol. Geoth. Res. 180, 203–224.

Brand, B.D., Gravley, D., Clarke, A., Lindsay, J., Boomberg, S.H., Agustin-Flores, J., Németh, K.,2014. A combined field and numerical approach to understanding dilute pyroclasticdensity current dynamics and hazard potential: Auckland Volcanic Field, NewZealand. J. Volcanol. Geoth. Res. 276, 215–232.

Branney, M.J., Kokelaar, P., 2002. Geol. Soc. Lond. Mem. 27 (143 pp.).Cassidy, J., Locke, C.A., 2010. The Auckland volcanic field, New Zealand: geophysical evi-

dence for structural and spatio-temporal relationships. J. Volcanol. Geoth. Res. 195,127–137.

Chough, S.K., Sohn, Y.K., 1990. Depositional mechanics and sequences of base surges,Songaksan tuff ring, Cheju Island, Korea. Sedimentology 37, 1115–1135.

Cioni, R., Sbrana, A., Vecci, R., 1992. Morphologic features of juvenile pyroclasts frommagmatic and phreatomagmatic deposits of Vesuvius. J. Volcanol. Geoth. Res. 51,61–78.

Conybeer, M., 1995. Physical Volcanology of Maungataketake volcano. MSc. Thesis TheUniversity of Auckland, New Zealand (100 pp.).

Crowcroft, G., Bowden, D., 2002. Auckland Water Resource Quantity Statement 2002.Technical Publication Number 171. Auckland Regional Council (120 pp.).

Crowcroft, G., Smaill, A., 2001. Auckland. In: Rosen, M.R., White, P.A. (Eds.), Groundwatersof New Zealand. New Zealand Hydrological Society Inc., Wellington, New Zealand,pp. 303–313.

Pattle Delamore and Partners LTD, 2003. Groundwater modelling of theWaitematas nearThree Kings Quarry. Unpublished internal report prepared for Winstone Aggregates,February 2003, New Zealand.

Dellino, P., Frazzetta, G., La Volpe, L., 1990. Wet surge deposits at La Fossa di Vulcano:depositional and eruptive mechanisms. J. Volcanol. Geoth. Res. 43, 215–233.

Edbrooke, S.W., Mazengarb, C., Stephenson, W., 2003. Geology and geological hazards ofthe Auckland urban area, New Zealand. Quat. Int. 103, 3–21.

Eggins, S.M., Rudnick, R.L., McDonough, W.F., 1998. The composition of peridotites andtheir minerals: a laser-ablation ICP-MS study. Earth Planet Sci. Lett. 154, 53–71.

Fisher, R.V., Schmincke, H.-U., 1984. Pyroclastic RocksSpringer-Verlag, Berlin (472 pp.).Folk, R.L., Ward, W.C., 1957. Brazos river bar: a study in the significance of grain size

parameters. J. Sed. Petrol. 27, 3–26.Hayward, B.W., 1979. Eruptive history of the early to midMioceneWaitekere Volcanic arc

and paleogeography of the Waitemata Basin, northern New Zealand. J. R. Soc. N. Z. 9,297–320.

Hayward, B.W., 1993. The tempestuous 10 million year life of a double arc and intra-arcbasin — New Zealand's Northland Basin in the early Miocene. In: Balance, P.F. (Ed.),Sedimentary Basins of the World. South Pacific Sedimentary Basis, 2. Elsevier,Amsterdam, pp. 113–142.

Hayward, J.J., Hayward, B.W., 1995. Fossil forest preserved in volcanic ash and lava atIhumatao and Takapuna, Auckland. Tane 35, 127–142.

Hayward, B.W., Murdoch, G., Maitland, G., 2011. Volcanoes of Auckland, the EssentialGuideAuckland University Press, New Zealand (234 pp.).

Houghton, B.F., Smith, R.T., 1993. Recycling of magmatic clasts during explosive eruptions:estimating the true juvenile content of phreatomagmatic volcanic deposits. Bull.Volcanol. 55, 414–420.

Houghton, B.F.,Wilson, C.J.N., Smith, I.E.M., 1999. Shallow-seated controls on styles of explo-sive basaltic volcanism: a case study from New Zealand. J. Volcanol. Geoth. Res. 91,97–120.

Ingram, R.L., 1954. Terminology for the thickness of stratification and parting units insedimentary rocks. Geol. Soc. Am. Bull. 65, 130–165.

Inman, D.L., 1952. Measures for describing the size distribution of sediments. J. Sed. Petrol.22, 125–145.

Irwin, M., 2009. Synthesis of existing structural data for the Auckland Volcanic Field. Insti-tute of Earth Science and Engeneering Technical Report 1–2009.01. The University ofAuckland (207 pp.).

Kenny, J.A., Lindsay, J.M., Howe, T.M., 2012. Post-Miocene faults in Auckland: insightsfrom borehole and topographic analysis. N. Z. J. Geol. Geophys. 55, 323–343.

Kereszturi, G., Németh, K., Cronin, J.S., Agustin-Flores, J., Smith, I.E.M., Lindsay, J., 2013. Amodel for calculating eruptive volumes for monogenetic volcanoes — implicationfor the Quaternary Auckland Volcanic Field, New Zealand. J. Volcanol. Geoth. Res.266, 16–33.

Kermode, L.O., 1992. Geology of the Auckland Urban Area. Scale 1:50,000. Institute ofGeological and Nuclear Sciences Geological Map 2, 1 Sheet and 63 p. Institute of Geo-logical and Nuclear Sciences Ltd., Lower Hutt, New Zealand.

Kienle, J., Kyle, P.R., Self, S., Motyka, R.J., Lorenz, V., 1980. Ukinrek Maars, Alaska, I. April1977 eruption sequence, petrology and tectonic setting. J. Volcanol. Geoth. Res. 7,11–37.

Lefebvre, N.S., White, J.D.L., Kjarsgaard, B.A., 2013. Unbedded diatreme deposits revealmaar–diatreme-forming eruptive processes: Standing Rocks West, Hopi Buttes,Navajo, Nation, USA. Bull. Volcanol. 75, 739. http://dx.doi.org/10.1007/s00445-013-0739-9.

Lorenz, V., 2003.Maar–diatreme volcanoes, their formation, and their setting in hard-rockor soft-rock environments. GeoLines 15, 72–83.

Lorenz, V., Haneke, J., 2004. Relationship between diatremes, dykes, sills, laccoliths, intru-sive–extrusive domes, lava flows and tephra deposits with unconsolidated water-saturated sediments in the late Variscan intermontane Saar-Nahe Basin, SWGermany. In: Breitkreuz, C., Petford, N. (Eds.), Physical Geology of SubvolcanicSystems — Laccoliths, Sills, and Dykes. Blackwell Sciences, Oxford, pp. 75–124.

Marra, M.J., Alloway, B.V., Newnham, R.M., 2006. Paleoenvironmental reconstruction of awell-preserved stage 7 forest sequence catastrophically buried by basaltic eruptivedeposits, northern New Zealand. Quat. Sci. Rev. 25, 2143–2161.

Martin, U., Németh, K., 2007. Blocky versus fluidal peperite textures developed in volcanicconduits, vents, and crater lakes of phreatomagmatic volcanoes in Mio/Pliocenevolcanic fields of Western Hungary. J. Volcanol. Geoth. Res. 159, 164–178.

McGee, L.E., Beier, C., Smith, I.E.M., Turner, S.P., 2011. Dynamics of melting beneath asmall-scale basaltic system: a U–Th–Ra study from Rangitoto volcano, Aucklandvolcanic field, New Zealand. Contrib. Mineral. Petrol. 162, 547–563.

McGee, L.E., Millet, M.-A., Smith, I.E.M., Németh, K., Lindsay, J.M., 2012. The inception andprogression of melting in a monogenetic eruption: Motukorea volcano, the AucklandVolcanic Field, New Zealand. Lithos 155, 360–374.

Molloy, C., Shane, P., Augustinus, P., 2009. Eruption recurrence rates in a basaltic volcanicfield based on tephra layers in maar sediments: implications for hazards in theAuckland volcanic field. Geol. Soc. Am. Bull. 121, 1666–1677.

Moore, J.G., Nakamura, K., Alcaraz, A., 1966. The 1966 eruption of Taal Volcano. Science151, 955–960.

Németh, K., Cronin, S.J., 2009. Phreatomagmatic volcanic hazards where rift-systemsmeetthe sea, a study from Ambae Island, Vanuatu. J. Volcanol. Geoth. Res. 180, 246–258.

Németh, K., Cronin, S.J., 2011. Drivers of explosivity and elevated hazard in basaltic fissureeruptions: the 1913 eruption of Ambrym Volcano, Vanuatu (SW-Pacific). J. Volcanol.Geoth. Res. 201, 194–209.

Németh, K., Cronin, S.J., Haller, M.J., Brenna, M., Csillag, G., 2010. Modern analogues forMiocene to Pleistocene alkali basaltic phreatomagmatic fields in the PannonianBasin: “soft substrate” to “combined” aquifer controlled phreatomagmatism in intra-plate volcanic fields. Cent. Eur. J. Geosci. 2, 339–361.

Németh, K., Cronin, S.J., Smith, I.E.M., Agustín-Flores, J., 2012. Amplified hazard of small-volume monogenetic eruptions due to environmental controls, Orakei Basin,Auckland Volcanic Field, New Zealand. Bull. Volcanol. 74, 2121–2137.

Pillans, B., 1983. Upper Quaternary marine terrace chronology and deformation, SouthTaranaki, New Zealand. Geology 11, 292–297.

Raue, H., 2004. A new model for the fracture energy budget of phreatomagmatic explo-sions. J. Volcanol. Geoth. Res. 129, 99–108.

63J. Agustín-Flores et al. / Journal of Volcanology and Geothermal Research 276 (2014) 46–63

Raza, A., Brown, R.W., Ballance, P.F., Kamp, P.J.J., 1999. Thermal history of the earlyWaitemata Basin and adjacent Waipapa Group, North Island, New Zealand. N. Z.J. Geol. Geophys. 42, 469–488.

Ross, P.S., White, J.D.L., 2006. Debris jets in continental phreatomagmatic volcanoes: afield study of their subterranean deposits in the Coombs Hills vent complex,Antarctica. J. Volcanol. Geoth. Res. 149, 62–84.

Schipper, C.I., White, J.D.L., Houghton, B.F., 2010. Syn- and post-fragmentation textures insubmarine pyroclasts from Lo'ihi Seamount, Hawai'i. J. Volcanol. Geoth. Res. 191,93–106.

Schipper, C.I., White, J.D.L., Zimanowski, B., Büttner, R., Sonder, I., Schmid, A., 2011. Experi-mental interaction of magma and “dirty” coolants. Earth Planet. Sci. Lett. 303, 323–336.

Schmid, R., 1981. Descriptive nomenclature and classification of pyroclastic deposits andfragments: recommendations to the IUGS Subcommision on the Systematics ofIgneous Rocks. Geology 9, 41–43.

Scoble, R., Millar, A., 1995. Kumeu–Hobsonville Groundwater Resource AssessmentReport (TP60)Auckland Regional Council (now Auckland Council) Technical Publica-tion, Auckland Council, New Zealand (70 pp.).

Searle, E.J., 1959. The volcanoes of Ihumatao and Mangere, Auckland. N. Z. J. Geol.Geophys. 3, 870–888.

Searle, E.J., Mayhill, R.D., 1981. City of Volcanoes: A Geology of AucklandLongman Paul,Auckland, New Zealand (195 pp.).

Sheridan, G.J., 2006. The Macroporosity and Microporosity Components of theWaitemataGroup Rocks, Auckland. The University of Auckland, New Zealand, New Zealand(M.Sc. thesis 105 pp.).

Sheridan, M.F., Wohletz, K.H., 1983. Hydrovolcanism: basic considerations and review.J. Volcanol. Geoth. Res. 17, 1–29.

Simpson, K.A., 1987. Rock Properties and Slope Processes of the Waitemata Group SoftRock, Along Sections of the East Coast of Auckland. The University of Auckland,New Zealand (M.Sc. thesis 144 pp.).

Smith, I.E.M., Blake, S., Wilson, C.J.N., Houghton, B.F., 2008. Deep-seated fractionationduring the rise of a small-volume basalt magma batch: Crater Hill, Auckland, NewZealand. Contrib. Mineral. Petrol. 155, 511–527.

Sohn, Y.K., 1996. Hydrovolcanic processes forming basaltic tuff rings and cones on ChejuIsland, Korea. Geol. Soc. Am. Bull. 108, 1199–1211.

Sohn, Y.K., 1997. On traction-carpet sedimentation. J. Sed. Res. 67, 502–509.Sohn, Y.K., Chough, S.K., 1989. Depositional processes of the Suwolbong tuff ring, Cheju

Island (Korea). Sedimentology 36, 837–855.Sohn, Y.K., Park, K.H., 2005. Composite tuff ring/cone complexes in Jeju Island, Korea:

possible consequences of substrate collapse and vent migration. J. Volcanol. Geoth.Res. 141, 157–175.

Sohn, Y.K., Son, M., Jeong, J.O., Jeon, Y.M., 2009. Eruption and emplacement of a laterallyextensive, crystal-rich, pumice-free ignimbrite (the Cretaceous Kusandong Tuff,Korea). Sed. Geol. 2009, 190–203.

Spörli, K.B., Eastwood, V.R., 1997. Elliptical boundary of an intraplate volcanic field,Auckland, New Zealand. J. Volcanol. Geoth. Res. 79, 169–179.

Spörli, K.B., Rowland, J.V., 2007. Superposed deformation in turbidites and syn-sedimentary slides of the tectonically active Miocene Waitemata Basin, northernNew Zealand. Basin Res. 19, 199–216.

Sulpizio, R., Dellino, P., 2008. Sedimentology, depositional mechanisms and pulsatingbehaviour of pyroclastic density currents. In: Gottsman, J., Marti, J. (Eds.), CalderaVolcanism: Analysis, Modelling, and Response. Developments in volcanism, vol. 10.Elsevier, Amsterdam, pp. 57–96.

Valentine, G.A., 1987. Stratified flow in pyroclastic surges. Bull. Volcanol. 49, 616–630.Valentine, G.A., 2012. Shallow plumbing systems for small-volume basaltic volcanoes, 2:

evidence from crustal xenoliths at scoria cones and maars. J. Volcanol. Geoth. Res.223–224, 47–63.

Valentine, G.A., Fisher, R.V., 2000. Pyroclastic surges and blasts. In: Sigurdsson, H.,Houghton, B.F., McNutt, S., Rymer, H., Stix, J. (Eds.), Encyclopedia of Volcanoes.Smithsonian Inst Press, Washington D.C., pp. 571–580.

Valentine, G.A., White, J.D.L., 2012. Revised conceptual model for maar–diatreme: subsur-face processes, energetics, and eruptive products. Geology 40 (12), 1111–1114.http://dx.doi.org/10.1130/G33411.1.

Valentine, G.A., Shufelt, N., Hintz, A., 2011. Models of maar volcanoes, Lunar Crater (Nevada,USA). Bull. Volcanol. 73, 753–765.

Van Otterloo, J., Cas, R.A.F., Sheard, M., 2013. Eruption processes and deposit characteris-tics at the monogenetic Mt. Gambier Volcanic Complex, SE Australia: implications foralternatingmagmatic and phreatomagmatic activity. Bull. Volcanol. http://dx.doi.org/10.1007/s00445-013-0737-y.

Vazquez, J.A., Ort, M.H., 2006. Facies variation of eruption units produced by the passageof single pyroclastic surge currents, Hopi Buttes volcanic field, USA. J. Volcanol. Geoth.Res. 154, 222–236.

Verwoerd, W.J., Chevallier, L., 1989. Contrasting types of surtseyan tuff cones on Marionand Prince Edward islands, southwest Indian Ocean. Bull. Volcanol. 49, 399–417.

Viljevac, Z., Murphy, G., Smaill, A., Crowcroft, G., Bowden, D., 2002. South Aucklandgroundwater, Kaawa Aquifer Recharge Study and Management of the Volcanic andKaawa AquifersAuckland Regional Council (now Auckland Council) TechnicalPublication, Auckland Council, New Zealand (Number 133, 193 pp.).

White, J.D.L., 1990. Depositional architecture of a maar-pitted playa: sedimentation in theHopi Buttes volcanic field, northeastern Arizona, U.S.A. Sed. Geol. 67, 55–84.

White, J.D.L., 1991. Maar–diatreme phreatomagmatism at Hopi Buttes, Navajo Nation(Arizona), USA. Bull. Volcanol. 53, 239–258.

White, J.D.L., 1996. Impure coolants and interaction dynamics of phreatomagmaticeruptions. J. Volcanol. Geoth. Res. 74, 155–170.

White, J.D.L., Ross, P.S., 2011. Maar–diatreme volcanoes: a review. J. Volcanol. Geoth. Res.201, 1–29.

Wohletz, K.H., 1983. Mechanisms of hydrovolcanic pyroclast formation: grain-size,scanning electron microscopy, and experimental studies. J. Volcanol. Geoth.Res. 17, 31–63.

Wohletz, K.H., 1986. Explosive magma–water interactions: thermodynamics, explosionmechanisms, and field studies. Bull. Volcanol. 48, 245–264.

Wohletz, K.H., McQueen, R.G., 1984. Experimental studies of hydrovolcanic explosions.Explosive Volcanism: Inception, Evolution and Hazards. National Academy Press,Washington D.C., USA, pp. 158–169.

Wohletz, K.H., Sheridan, M.F., 1979. Amodel of pyroclastic surge. Geol. Soc. Am. Spec. Pap.180, 177–193.

Zimanowski, B., Wohletz, K.H., 2000. Physics of phreatomagmatism—II. Terra Nostra 6,535–544.

Zimanowski, B., Fröhlich, G., Lorenz, V., 1991. Quantitative experiments on phreatomagmaticexplosions. J. Volcanol. Geoth. Res. 48, 341–358.

Zimanowski, B., Büttner, R., Lorenz, V., Häfele, H.-G., 1997. Fragmentation of basaltic meltin the course of explosive volcanism. J. Geophys. Res. 102, 803–814.