02.06.2015 Views

Article - MyGeologyPage - University of California, Davis

Article - MyGeologyPage - University of California, Davis

Article - MyGeologyPage - University of California, Davis

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Article</strong><br />

Volume 12, Number 12<br />

28 December 2011<br />

Q12017, doi:10.1029/2011GC003639<br />

ISSN: 1525‐2027<br />

Environmental and chemical controls on palagonitization<br />

Bruce D. Pauly, Peter Schiffman, and Robert A. Zierenberg<br />

Department <strong>of</strong> Geology, <strong>University</strong> <strong>of</strong> <strong>California</strong>, 1 Shields Avenue, <strong>Davis</strong>,<br />

<strong>California</strong> 95616, USA (bdpauly@ucdavis.edu)<br />

David A. Clague<br />

Monterey Bay Aquarium Research Institute, 7700 Sandholdt Road, Moss Landing,<br />

<strong>California</strong> 95039, USA<br />

[1] Palagonitized sideromelane from submarine volcaniclastic, seafloor volcanic, marine phreatomagmatic,<br />

lacustine phreatomagmatic, and subglacial volcanic settings was investigated using in situ microanalysis<br />

to test if palagonite composition and texture are related to depositional environment. Palagonitization extent<br />

varies linearly and inversely with original sample porosity, suggesting that porosity is a controlling factor <strong>of</strong><br />

palagonitization. Water absorbance <strong>of</strong> reflected infrared light varies linearly with water content derived from<br />

electron microprobe totals. Palagonite water content has a linear, inverse relationship to palagonitization<br />

extent. REEs are immobile during palagonitization, so they can be used to construct isocon diagrams<br />

for estimating major‐element concentration changes. Major‐element and overall mass changes during<br />

palagonitization vary widely (particularly for FeO and TiO 2 ) and indicate that palagonitization cannot be<br />

an isovolumetric process. These parameters depend strongly on original sideromelane composition, thus<br />

requiring composition to be taken into account when performing global oceanic cation flux calculations.<br />

Subalkaline sideromelane dissolves much more rapidly than alkaline sideromelane during palagonitization.<br />

Two styles <strong>of</strong> palagonitization, burial‐diagenesis (relatively long‐duration, low water/rock; passive fluid<br />

circulation) and hydrothermal (relatively short‐duration, high water/rock; hydrothermal fluid circulation),<br />

are recognized. Observed palagonite REE concentration gradients indicate that sideromelane dissolution<br />

must continue in the zone behind the advancing palagonitization front. MgO was found to be highly mobile<br />

during palagonitization. Observed palagonite MgO gradients are not developed during sideromelane dissolution,<br />

but instead record initiation <strong>of</strong> syn‐ and/or post‐palagonitization conversion <strong>of</strong> the gel‐palagonite<br />

layer to a phyllosillicate layer, consistent with evolution <strong>of</strong> sideromelane alteration layers toward equilibrium<br />

with the solution.<br />

Components: 13,000 words, 14 figures, 7 tables.<br />

Keywords: alteration layers; equilibrium; glass dissolution; mass change; palagonitization; water content.<br />

Index Terms: 1039 Geochemistry: Alteration and weathering processes (3617); 8424 Volcanology: Hydrothermal systems<br />

(0450, 1034, 3017, 3616, 4832, 8135); 8427 Volcanology: Subaqueous volcanism.<br />

Received 28 June 2011; Revised 27 October 2011; Accepted 29 October 2011; Published 28 December 2011.<br />

Pauly, B. D., P. Schiffman, R. A. Zierenberg, and D. A. Clague (2011), Environmental and chemical controls on<br />

palagonitization, Geochem. Geophys. Geosyst., 12, Q12017, doi:10.1029/2011GC003639.<br />

Copyright 2011 by the American Geophysical Union 1 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

1. Introduction<br />

[2] Palagonitization entails the hydrolytic alteration<br />

<strong>of</strong> sideromelane whereby basaltic glass is dissolved<br />

and hydrated, producing palagonite and various<br />

authigenic minerals, notably zeolites and smectites<br />

[Fisher and Schmincke, 1984; Stroncik and<br />

Schmincke, 2001; Walton and Schiffman, 2003].<br />

Palagonitization is a globally significant geochemical<br />

process. In spite <strong>of</strong> important implications to our<br />

understanding <strong>of</strong> nuclear waste storage [Crovisier<br />

et al., 2003], volcanic edifice stability [Schiffman<br />

et al., 2006], and Martian crustal evolution<br />

[Bishop et al., 2002], the controls and mechanism <strong>of</strong><br />

the palagonitization reaction remain poorly understood.<br />

Palagonite generally is considered to be the<br />

initial, possibly metastable, replacement product <strong>of</strong><br />

sideromelane during palagonitization, with the final<br />

replacement product being smectite [Staudigel and<br />

Hart, 1983; Thorseth et al., 1991; Stroncik and<br />

Schmincke, 2001]. In this study, we follow Walton<br />

and Schiffman [2003] and consider palagonitization<br />

to be the process <strong>of</strong> forming palagonitized<br />

sideromelane, which herein will be referred to simply<br />

as palagonite. We use the term palagonite to<br />

describe the gel‐like, yellow to orange (as viewed in<br />

plane‐polarized light), isotropic replacement product<br />

<strong>of</strong> sideromelane; this is the practical, comprehensive<br />

sense as defined by Stroncik and Schmincke<br />

[2001].<br />

[3] Published major‐element compositional data on<br />

palagonites from different localities, examples <strong>of</strong><br />

which can be found in the works by Honnorez [1967,<br />

1972], Hay and Iijima [1968a], Staudigel and Hart<br />

[1983], Furnes [1984], Bednarz and Schmincke<br />

[1989], Jercinovic et al. [1990], Crovisier et al.<br />

[1992], Daux et al. [1994], Stroncik and Schmincke<br />

[2001], and Walton and Schiffman [2003], show<br />

substantial variation. The cause(s) <strong>of</strong> these compositional<br />

variations remain poorly understood. Although<br />

there is general agreement among previous workers<br />

that dissolution (congruent or incongruent) and<br />

hydration are responsible for the creation <strong>of</strong> palagonite<br />

[Crovisier et al.,1987;Daux et al., 1994; Le Gal<br />

et al., 1999], the mechanism(s) by which the transformation<br />

occurs is not completely understood.<br />

Stroncik and Schmincke [2001] proposed that palagonitization<br />

consists <strong>of</strong> two different reaction stages,<br />

the first being congruent dissolution <strong>of</strong> glass and<br />

simultaneous precipitation <strong>of</strong> palagonite, the second<br />

being an aging process during which palagonite<br />

reacts with fluid and crystallizes to smectite.<br />

Crovisier et al. [2003] summarizes results also<br />

suggesting that palagonitization is not a steady state<br />

process because glass dissolution rates decrease by<br />

up to 5 orders <strong>of</strong> magnitude as the palagonite layer<br />

thickens.<br />

[4] The purpose <strong>of</strong> this study was to investigate<br />

whether palagonitization and the resulting textural<br />

and compositional properties <strong>of</strong> palagonites are<br />

controlled by characteristics <strong>of</strong> the starting materials<br />

and <strong>of</strong> the palagonitization environment. Selected<br />

palagonitized samples from various geologic settings<br />

and localities were systematically analyzed by<br />

optical petrography, electron microprobe analysis<br />

(EPMA), and laser ablation inductively coupled<br />

mass spectroscopy (LA‐ICPMS). An advantage <strong>of</strong><br />

these techniques is the ability to perform in situ<br />

analyses <strong>of</strong> thin section samples <strong>of</strong> palagonite and<br />

adjacent sideromelane in their petrogenetic context.<br />

The widely used assumption that palagonite water<br />

content equals the difference between 100% and the<br />

total <strong>of</strong> the analyzed oxides in EPMA compositional<br />

analyses was tested using in situ reflected‐light<br />

Fourier transform infrared spectroscopy (RL‐FTIR).<br />

[5] Basaltic glass alteration textures currently are<br />

categorized as either biotic or abiotic; Staudigel et al.<br />

[2008] provided a recent review. Thorseth et al.<br />

[1992] suggested a mechanism for the formation <strong>of</strong><br />

bioalteration features in volcanic glass whereby<br />

microbes dissolve glass by changing the local contact<br />

area pH. Biotic versus abiotic glass alteration can be<br />

differentiated based on petrographically observable<br />

microtextures [Furnes et al., 2002, 2007]. Bioalteration<br />

textures generally fall into two categories: granular<br />

and tubular (tunneling) [Furnes et al., 2001;<br />

Cockell and Herrera, 2008]. Staudigel et al. [2008]<br />

provide detailed documentation <strong>of</strong> these textures.<br />

Many recent studies [Furnes et al.,2001;Walton and<br />

Schiffman, 2003; Staudigel and Furnes, 2004;<br />

Staudigel et al., 2004, 2006; Walton, 2008] document<br />

evidence that microorganisms play a role in palagonitization,<br />

particularly in the enhancement <strong>of</strong> sideromelane<br />

dissolution and involvement in precipitation<br />

<strong>of</strong> authigenic phases. Because this study focuses on<br />

abiotic palagonitization, samples with visual biotically<br />

mediated textures were excluded.<br />

2. Sample Selection<br />

[6] The sampling strategy for this study was designed<br />

to test if observed variations in the petrographic and<br />

chemical data are related to varying protolith and/<br />

or palagonitization environment. Localities were<br />

chosen based on prior knowledge <strong>of</strong> the geologic<br />

history and therefore likely conditions attending<br />

sample palagonitization. All <strong>of</strong> the samples have<br />

2<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

undergone palgonitization, and were classified by<br />

similarity <strong>of</strong> geologic setting into 5 environmental<br />

groups. Samples in the Submarine Volcaniclastic<br />

group consist <strong>of</strong> hyaloclastite that was deposited and<br />

subsequently buried in marine environments. The<br />

Seafloor Volcanic samples are from surficial hyaloclastite<br />

deposits or surficial lava flows located near<br />

the submarine edifice where they formed. Samples in<br />

the Marine Phreatomagmatic and Lacustrine Phreatomagmatic<br />

groups consist <strong>of</strong> hyaloclastite formed at<br />

tuff rings and tuff cones. The Subglacial Volcanic<br />

samples are from hyaloclastite deposits formed during<br />

eruptions <strong>of</strong> volcanoes beneath Icelandic glaciers.<br />

Table 1 lists the location, nomenclature, geologic<br />

setting, type, age, and reference(s) for each <strong>of</strong> the<br />

studied samples by environment group. The samples<br />

in this study are <strong>of</strong> Neogene and Quaternary age, the<br />

youngest being the two Surtsey samples, which were<br />

deposited between 1963 and 1968.<br />

3. Analytical Methods<br />

3.1. Petrography<br />

[7] One thin section <strong>of</strong> each sample was pointcounted<br />

using a mechanical stage with 1 mm gridspacing.<br />

The number <strong>of</strong> points counted per sample<br />

varied as a function <strong>of</strong> the available thin‐section<br />

area and ranged from 240 to 464. Point‐counting<br />

categories were sideromelane, lithic and crystal<br />

clasts, palagonite and smectite, undifferentiated<br />

zeolite, pore space, and “other” as necessary (clastic<br />

matrix, non‐replacive smectite, etc.). Minus‐cement<br />

porosity [Walton and Schiffman, 2003], an estimate<br />

<strong>of</strong> initial (depositional) porosity, was determined for<br />

each sample. For this study, minus‐cement porosity<br />

was defined as % pore space + % zeolite − % lithic<br />

and crystal clasts. This parameter can range from<br />

0 to 100 percent; lower values indicate lower initial<br />

porosity and vice versa. Palagonitization extent,<br />

defined here as % zeolite + % palagonite + %<br />

smectite (calculated as lithic and crystal clast and<br />

pore space free) was also calculated for each sample.<br />

This parameter can range from 0 to 100 percent;<br />

higher values indicate a greater extent <strong>of</strong> palagonitization<br />

and vice versa.<br />

3.2. Reflected Light Infrared Spectroscopy<br />

(RL‐FTIR)<br />

[8] Four thin‐section samples, with widely differing<br />

inferred palagonite water contents (based<br />

on electron microprobe compositional data) were<br />

selected for analysis by FTIR‐reflectance methods<br />

[Grzechnik et al., 1996; Moore et al., 2000] using a<br />

Nicolet Magna 750 FTIR spectrometer, fitted with<br />

a SpectraTech Analytical‐IR microscope, at the<br />

U.S. Geological Survey in Menlo Park, CA. The<br />

reflectance spectrum for each sample consisted <strong>of</strong><br />

250 scans performed over the 6000–650 cm −1<br />

range using a 40 mm aperture. Background reflectance<br />

was measured using these same parameters<br />

and a polished gold target. The spectra have a<br />

resolution <strong>of</strong> 32 cm −1 and were corrected to provide<br />

a flat baseline, followed by a Kramers‐Kronig<br />

transformation between 4000 and 2000 wave numbers.<br />

All data collection and transformations were<br />

undertaken with Omnic® s<strong>of</strong>tware provided with<br />

the Nicolet FTIR spectrometer. The Kramers‐<br />

Kronig absorbance due to total water was determined<br />

by measuring the peak height at 3500 cm −1 ,<br />

but this indicates relative sample water content only,<br />

since RL‐FTIR calibration standards for palagonite<br />

water content are not available.<br />

3.3. Electron Probe Microanalysis (EPMA)<br />

[9] Palagonite and sideromelane major‐element<br />

composition was analyzed using a Cameca SX‐100<br />

electron microprobe at the <strong>University</strong> <strong>of</strong> <strong>California</strong>,<br />

<strong>Davis</strong>. Prior to analysis, a ∼250 Å‐thick conductive<br />

carbon coating was applied to the sample thin sections.<br />

All analyses were performed using 15 KeV<br />

accelerating voltage. Sideromelane was analyzed<br />

using a 10 mm rastered beam and 10 nA current.<br />

Elemental thermal stability (particularly <strong>of</strong> Na) in<br />

palagonites as a function <strong>of</strong> current and raster size<br />

was investigated; optimal beam settings for palagonite<br />

analyses were 5 mm raster size and 2 nA current.<br />

Peaks and backgrounds for all analyses were typically<br />

counted for 10 s, and standard ZAF correction<br />

techniques [Schiffman and Roeske, 2002] were used<br />

to convert net intensities relative to oxide and silicate<br />

calibration standards to concentrations. Two types<br />

<strong>of</strong> analyses were performed. Palagonite <strong>of</strong> all<br />

samples was initially analyzed at several spots<br />

roughly in the centers <strong>of</strong> the rinds; adjacent sideromelane<br />

spot analyses were also performed. Also,<br />

for several <strong>of</strong> the samples with rinds <strong>of</strong> sufficient<br />

thickness, spot analyses were performed at a uniform<br />

spacing along a traverse starting at the outer<br />

edge <strong>of</strong> the palagonitized rind, moving inward toward<br />

and perpendicular to the sideromelane contact, continuing<br />

across the contact, and ending with at least<br />

one analysis in sideromelane. The number <strong>of</strong> palagonite<br />

spot analyses possible for each traverse was a<br />

function <strong>of</strong> the rastered beam size and the particular<br />

sample’s palagonitized rind thickness.<br />

3<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Table 1. Sample Geologic Information and Reference(s)<br />

ENVIRONMENT GROUP<br />

Sample (Nomenclature for This Sudy) [Location] Geologic Setting Sample Setting (Rock Type) Age Reference(s)<br />

SUBMARINE VOLCANICLASTIC<br />

Hawaii Scientific Drilling Project<br />

(HSDP)<br />

[Big Island <strong>of</strong> Hawaii] (hyaloclastite)<br />

submerged volcano flank drillcore Pleistocene Walton and Schiffman [2003];<br />

Garcia et al. [2007]<br />

St. <strong>of</strong> Hawaii Geothermal Hole #1<br />

(SOH‐1)<br />

[Big Island <strong>of</strong> Hawaii] (hyaloclastite)<br />

submerged volcano flank drillcore Pleistocene Novak and Evans [1991];<br />

Trusdell et al. [1999]<br />

Palagonia submerged lava flows surface (uplifted) Miocene–Early Pliocene Honnorez [1967, 1972]<br />

[Acqua Amara, Palagonia, Sicily] (hyaloclastite)<br />

Hilina Bench volcaniclastic basin submarine surface Pleistocene Lipman et al. [2002, 2006];<br />

Coombs et al. [2006];<br />

Kimura et al. [2006]<br />

[Offshore, Big Island <strong>of</strong> Hawaii] (volcaniclastic hyaloclastite)<br />

SEAFLOOR VOLCANIC<br />

Davidson seamount submarine push-core 14.8–9.8 Ma <strong>Davis</strong> et al. [2002];<br />

<strong>Davis</strong> and Clague [2003];<br />

Clague et al. [2009]<br />

[Davidson Seamount, Pacific Ocean] (flow breccia)<br />

North Arch volcanic field submarine surface Pliocene–Pleistocene Clague et al. [1990, 2002];<br />

<strong>Davis</strong> and Clague [2006]<br />

[North Arch volcanic field,<br />

<strong>of</strong>fshore Hawaii]<br />

(flow margin; Mn oxide‐encrusted)<br />

O’ahu cone submarine surface 510 Ka Clague et al. [2006]<br />

[O’ahu, northeast <strong>of</strong>fshore flank] (hyaloclastite)<br />

MARINE PHREATOMAGMATIC<br />

Diamond Head tuff ring subaerial surface 360 Ka Winchell [1947];<br />

Clague et al. [2006]<br />

[O’ahu] (hyaloclastite)<br />

Surtsey (less palagonitized sample) tuff cone subaerial surface 1963 AD Thorarinsson et al. [1964];<br />

Jakobsson [1972, 1978];<br />

[<strong>of</strong>fshore Iceland] (hyaloclastite) Jakobsson and Moore [1986];<br />

Jakobsson et al. [2000]<br />

4<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Table 1. (continued)<br />

ENVIRONMENT GROUP<br />

Sample (Nomenclature for This Sudy) [Location] Geologic Setting Sample Setting (Rock Type) Age Reference(s)<br />

Surtsey (more palagonitized sample) tuff cone subaerial surface 1963 AD Thorarinsson et al. [1964];<br />

Jakobsson [1972, 1978];<br />

[<strong>of</strong>fshore Iceland] (hyaloclastite) Jakobsson and Moore [1986];<br />

Jakobsson et al. [2000]<br />

Heimaklettur tuff cone subaerial surface approx. Nine Ka Mattsson and Höskuldsson [2003]<br />

[Westman Islands, Iceland] (hyaloclastite)<br />

LACUSTRINE PHREATOMAGMATIC<br />

Black Point tuff cone subaerial surface 13.61 Ka Benson et al. [1998]<br />

[Mono Lake, eastern <strong>California</strong>] (hyaloclastite)<br />

Pahvant Butte tuff cone subaerial surface 15.3 Ka Oviatt and Nash [1989];<br />

White [1996, 2001]<br />

[South‐central Utah] (hyaloclastite)<br />

Fort Rock tuff ring subaerial surface 100–50 Ka Heiken et al. [1981];<br />

Jensen [1995]<br />

[Southeastern Oregon] (hyaloclastite)<br />

SUBGLACIAL VOLCANIC<br />

Mosfell (less palagonitized sample) tuya subaerial surface 400–500 Ka H. Franzson (personal<br />

communication, 2010)<br />

[Iceland] (Grimsnes district, Iceland) (hyaloclastite)<br />

Mosfell (more palagonitized sample) tuya subaerial surface 400–500 Ka H. Franzson (personal<br />

communication, 2010)<br />

[Iceland] (Grimsnes district, Iceland) (hyaloclaastite)<br />

Ingólfsfjall tuya subaerial surface approx. One Ma Saemundsson et al. [2010]<br />

[Iceland] (hyaloclastite)<br />

5<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 1. Thin section photomicrograph (PPL) <strong>of</strong> HSDP sample R0777–17.4 (hyaloclastite; submarine volcaniclastic<br />

group). Note relatively thick, smooth, gel‐like palagonite rinds developed on sideromelane grains, abundant zeolites<br />

(phillipsite and chabazite), and evidence <strong>of</strong> original porosity reduction due to burial such as abundant fractured<br />

grains and intergranular contacts.<br />

3.4. Laser Ablation Mass Spectroscopy<br />

(LA‐ICPMS)<br />

[10] Trace element analyses were performed using<br />

Laser Ablation‐Inductively Coupled Mass Spectrometry<br />

(LA‐ICPMS) at the <strong>University</strong> <strong>of</strong><br />

<strong>California</strong>‐<strong>Davis</strong> Interdisciplinary Center for Plasma<br />

Mass Spectrometry. A New Wave Research Nd:<br />

YAG 213 nm laser (model UP‐213), equipped with<br />

an optical microscope, was used for laser ablation.<br />

Helium gas was used to carry ablated material to an<br />

Agilent Technologies inductively coupled mass<br />

spectrometer (model 7500a) equipped with a quadrupole<br />

mass analyzer and an electron multiplier<br />

detector. Ar gas was mixed with the He carrier gas at<br />

the entrance to the plasma chamber. Palagonitized<br />

sideromelane can disintegrate easily with application<br />

<strong>of</strong> the laser. In order to prevent uncontrolled ablation,<br />

optimal laser parameters were determined by trial and<br />

error, attempting to maximize count rates and prevent<br />

sample destruction during analysis. Sample thin<br />

sections were ablated using a 20‐micron beam size,<br />

70% laser energy (approximately 3.75 J/cm 2 ), and a<br />

10 Hz laser pulse repetition rate. For each analysis,<br />

the laser dwell time was 35 s. Mass spectrometer<br />

integration times were 10 ms for V, Sr, Zr, and Rb,<br />

30 ms for Sc, Cr, Co, and Cu, and 50 ms for Zn, Nb,<br />

Ba, REE, and Pb. U.S. Geological Survey Geochemical<br />

Reference Materials BCR‐2 (absolute external<br />

calibration) and BHVO‐2 (internal reference) fused to<br />

glasses were used as calibration standards. Intensities<br />

(counts) were converted to concentrations by the<br />

GLITTER s<strong>of</strong>tware [van Achterbergh et al., 2001].<br />

Palagonite and adjacent sideromelane spot and traverse<br />

(when possible) analyses locations were as<br />

described above for EPMA.<br />

4. Results<br />

4.1. Petrography<br />

[11] Significant palagonite microtextural differences<br />

across the sample set were observed with respect to<br />

the abundance <strong>of</strong> pore space (initial and remaining),<br />

abundance <strong>of</strong> zeolitic cements, and thickness <strong>of</strong><br />

palagonitized rinds. Submarine volcaniclastic samples<br />

such as those from HSDP have relatively abundant<br />

zeolites and evidence <strong>of</strong> significant porosity<br />

reduction (likely due to their deep burial) such as<br />

significant fractures and grain to grain contacts<br />

(Figure 1). These samples also have relatively thick<br />

and smooth palagonitized rinds on sideromelane<br />

grains. In contrast, samples from the seafloor volcanic,<br />

marine phreatomagmatic (example Surtsey), and<br />

lacustrine phreatomagmatic groups typically have<br />

relatively rare zeolites, abundant pore space, and thin<br />

palagonitized rinds (Figure 2). The subglacial volcanic<br />

samples have more variability in terms <strong>of</strong> zeolite<br />

abundance, porosity, and palagonitized rind<br />

thickness. Measured palagonitized rind thickness<br />

varies widely across the sample set from 0.020 to<br />

0.600 mm (Table 2).<br />

6<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 2. Thin section photomicrograph (PPL) <strong>of</strong> Surtsey sample 07010405 (the more palagonitized <strong>of</strong> the 2 Surtsey<br />

samples; hyaloclastite; marine phreatomagmatic group). Note relatively thin palagonitized rinds, rare zeolites, and<br />

abundant pore space.<br />

Table 2.<br />

Rind Thickness, Minus‐Cement Porosity, and Palagonitization Extent Results<br />

Environment Group<br />

Sample Name<br />

Sample<br />

Identification<br />

Maximum Rind<br />

Thickness (mm)<br />

Percent Palagonite<br />

and Assoc.<br />

Smectite<br />

Percent<br />

Zeolite<br />

Minus‐Cement<br />

Porosity a (%)<br />

Palagonitization<br />

Extent b (%)<br />

Submarine Volcaniclastic<br />

HSDP R0777–17.4 0.350 28 21 37 85<br />

SOH‐1 4565 0.200 49 17 19 71<br />

Hilina Bench (all grains) S708–9 0.225 47 26 36 90<br />

Hilina Bench (subalkaline grains) S508–7 0.225<br />

Hilina Bench (alkaline grains) S508–7 0.150<br />

Palagonia PAL‐1 0.425 55 20 26 88<br />

Seafloor Volcanic<br />

Davidson T430‐R8 – c 31 16 43 70<br />

North Arch 26D‐6 0.200 37 3 51 57<br />

O’ahu T298‐R17 0.250 32 5 56 53<br />

Marine Phreatomagmatic<br />

Diamond Head 02010405 0.125 32 19 59 66<br />

Surtsey (less palagonitized) 07010404 – d 9 9 71 29<br />

Surtsey (more palagonitized) 07010405 0.150 26 11 63 58<br />

Heimaklettur 07020402 – d 40 12 30 68<br />

Lacustrine Phreatomagmatic<br />

Black Point 08310202 0.020 12 12 48 40<br />

Pahvant Butte 10230304 0.300 35 28 51 84<br />

Fort Rock 08029904 0.600 42 19 50 76<br />

Subglacial Volcanic<br />

Mosfell (less palagonitized) MOS‐1 0.040 14 9 24 35<br />

Mosfell (more palagonitized) MOS‐2 0.060 36 8 16 50<br />

Ingólfsfjall ING‐1 0.325 52 26 34 82<br />

a Minus‐cement porosity (%) = % pore space + % zeolites − % lithic and crystal clasts.<br />

b Palagonitization extent (%) = % zeolites + % palagonite and smectite (calculated as lithic and crystal clast and pore space free).<br />

c Rind thickness not measurable (glassy flow margin; palagonitized in areas, not rinds).<br />

d Rind thickness not measurable (rinds are very thin and incomplete).<br />

7<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 3. Plot <strong>of</strong> palagonitization extent versus minus‐cement porosity. Black solid symbols: submarine volcaniclastic<br />

group; violet solid symbols: seafloor volcanic group; blue solid symbols: marine phreatomagmatic group;<br />

red open symbols: lacustrine phreatomagmatic group; green symbols: subglacial volcanic group. For the submarine<br />

volcaniclastic, seafloor volcanic, and marine phreatomagmatic environment groups, samples with relatively greater<br />

palagonitization extent also have relatively low minus‐cement porosity and vice versa; samples from the other two<br />

groups are more variable. However, the more‐palagonitized Surtsey and Mosfell samples have higher palagonitization<br />

extent and lower minus‐cement porosity than the corresponding less‐palagonitized samples from the same edifices<br />

(colored arrows connecting symbols), implying that samples with lower original porosity (and expected lower permeability)<br />

palagonitized to a greater extent.<br />

[12] Minus‐cement porosity and palagonitization<br />

extent (Table 2) also show wide variations. Minuscement<br />

porosity ranges from 16 (relatively low) to<br />

71% (relatively high initial pore space). Minuscement<br />

porosity is fairly consistent within each<br />

environmental group (except for the Heimaklettur<br />

sample <strong>of</strong> the marine phreatomagmatic group) and is<br />

relatively low for the submarine volcaniclastic and<br />

subglacial volcanic samples. Palagonitization extent<br />

ranges from 29 (least palagonitized) to 90% (most<br />

thoroughly palagonitized). Palagonitization extent<br />

is consistently higher for the submarine volcaniclastic<br />

samples; it is generally lower and more variable<br />

for samples from the other sample groups.<br />

Palagonitization extent versus minus‐cement<br />

porosity is plotted in Figure 3. The submarine volcaniclastic<br />

samples have relatively high palagonitization<br />

extent and relatively low minus‐cement<br />

porosity. The other groups are more variable, but<br />

in general, there is a trend <strong>of</strong> increasing palagonitization<br />

extent with decreasing minus‐cement porosity.<br />

The less‐palagonitized samples from Surtsey and<br />

Mosfell have greater minus‐cement porosities than the<br />

more‐palagonitized samples from these same edifices.<br />

4.2. Water Content Measurement<br />

[13] The Kramers‐Kronig absorbances measured<br />

using RL‐FTIR at 3500 cm −1 (the wave number <strong>of</strong><br />

the total H 2 O absorbance peak) correspond to the<br />

derived (EPMA) water contents (Table 3). Figure 4<br />

shows that this is a nearly linear relationship, with<br />

the line <strong>of</strong> best fit having the equation<br />

Kramers-Kronig absorbance<br />

¼ 0:0081ðwt:% EPMA-derived H 2 OÞ:<br />

[14] The extinction coefficient <strong>of</strong> 0.0081 is in close<br />

agreement with the calibrated extinction coefficient<br />

obtained for basalt glass (0.0084) by King et al.<br />

[2006] using the same method. These workers calibrated<br />

the extinction coefficient using transmission<br />

IR spectroscopy on samples <strong>of</strong> natural sideromelane<br />

from the Manus Basin and the Galapagos Islands.<br />

8<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Table 3. Measured Kramers‐Kronig Absorbance for the 3500 cm −1 Band (RL‐FTIR) and Inferred Water Content (EPMA,<br />

wt. % Oxide)<br />

Location Sample KK Absorbance (RL‐FTIR) Inferred H 2 O (EPMA)<br />

Surtsey (less palagonitized sample) 07010404 0.260 33.4<br />

Davidson 26D‐6 0.246 27.1<br />

HSDP 7231 0.154 20.4<br />

Palagonia PAL‐1 0.130 14.0<br />

The precise extinction coefficient for palagonite<br />

awaits analysis <strong>of</strong> a set <strong>of</strong> palagonite samples with<br />

known absolute water contents. The linear relationship<br />

observed here, however, supports our<br />

assertion that the inferred water content (i.e., 100%<br />

minus the EPMA analysis total) <strong>of</strong> a palagonite<br />

sample represents its water content, and that therefore<br />

palagonite water content can be derived with the<br />

other major‐elements using EPMA. Water content<br />

results are given and discussed below (majorelement<br />

composition).<br />

4.3. Major‐Element Composition<br />

[15] Sideromelane major‐element composition results<br />

(EPMA) are given in Table 4. Sampled sideromelane<br />

compositions are fairly evenly distributed between<br />

alkaline and subalkaline (Figure 5). Averaged palagonite<br />

compositions for each sampled locality (measured<br />

adjacent to the sideromelane analyses in the<br />

center <strong>of</strong> the palagonitized rind) are given in Table 5.<br />

[16] Sample S508–7 from Hilina Bench is unique in<br />

this study because it contains both alkaline and subalkaline<br />

palagonitized sideromelane grains (Figure 6).<br />

These grains are from the pre‐shield alkalic and shield<br />

tholeiitic stages [Sisson et al., 2002, and references<br />

therein], respectively, <strong>of</strong> Kilauea volcano development.<br />

Palagonitization <strong>of</strong> these samples must have<br />

occurred after final deposition within the (Hilina)<br />

volcaniclastic basin, because the hyaloclastite grains<br />

are for the most part sub‐rounded and have relatively<br />

thick and smooth palagonite rinds, which would not<br />

be expected to survive erosion. Palagonitized rinds<br />

for this and other samples were observed to be so<br />

friable that preparation <strong>of</strong> double‐sided thin sections<br />

was not possible without rind separation from the<br />

fresh glass. Therefore, all <strong>of</strong> the grains within this<br />

particular sample must have palagonitized with<br />

identical conditions and for identical duration. The<br />

major‐element compositions <strong>of</strong> sideromelane and<br />

corresponding palagonite for 2 adjacent pairs <strong>of</strong><br />

Figure 4. Plot <strong>of</strong> measured Kramers‐Kronig absorbance (from RL‐FTIR measurements) versus inferred water content<br />

(from EPMA analyses) for samples from the 4 localities indicated (Table 3). A nearly linear, direct relationship<br />

between absorbance and inferred water content is indicated.<br />

9<strong>of</strong>26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Table 4.<br />

Average Sideromelane Major‐Element Compositions (wt. % Oxide), Determined Using EPMA<br />

Sample Group<br />

Location<br />

(Sample ID)<br />

Number<br />

<strong>of</strong><br />

Concentration (wt. % Oxide):<br />

Sideromelane<br />

Analyses SiO 2 TiO 2 Al 2 O 3 FeO MnO MgO CaO Na 2 O K 2 O<br />

P 2 O 5 SO 2 Total<br />

Submarine Volcaniclastic<br />

HSDP (R0777–17.4) 10 52.66 2.60 13.83 10.77 0.18 6.72 10.63 2.31 0.31 0.24 0.03 100.65<br />

SOH‐1 (3135) 9 50.76 2.86 13.63 11.36 0.16 6.09 10.28 2.54 0.48 0.38 0.07 98.60<br />

Palagonia (PAL‐1) 9 52.75 1.72 14.66 9.45 0.13 6.39 9.42 3.08 0.19 0.28 0.11 98.19<br />

Hilina Bench (S508‐7) (See Table 6)<br />

Submarine Volcanic<br />

Davidson (T430‐R8) 9 44.18 3.40 17.39 9.22 0.14 4.59 8.61 4.37 1.79 0.74 0.11 94.53<br />

North Arch (26D‐6) 3 44.64 2.21 13.79 11.44 0.24 6.73 13.77 3.82 0.91 0.35 0.35 98.25<br />

O’ahu (T298‐R17) 9 40.74 3.39 12.99 11.86 0.20 5.24 14.79 5.25 1.82 1.19 0.30 97.77<br />

Marine Phreatomagmatic<br />

Diamond Head (02010405) 9 42.65 2.70 13.84 11.87 0.18 6.38 13.42 3.94 0.98 0.66 0.12 96.74<br />

Surtsey–less palagonitized 9 46.33 2.37 15.70 12.01 0.22 5.82 10.25 3.84 0.55 0.53 0.12 97.74<br />

sample (07010404)<br />

Surtsey–more palagonitized 9 48.53 2.48 14.56 12.24 0.23 5.70 10.74 3.75 0.54 0.50 0.12 99.39<br />

sample (07010405)<br />

Heimaklettur (07020402) 9 46.32 2.66 15.78 12.18 0.21 5.87 10.02 3.71 0.70 0.39 0.10 97.93<br />

Lacustrine Phreatomagmatic<br />

Black Point (09120301) 6 49.90 1.51 18.78 8.32 0.13 5.63 8.59 3.94 1.08 0.36 0.03 98.27<br />

Pahvant Butte (10230304) 5 51.76 1.75 15.17 12.01 0.18 4.87 8.36 3.40 1.13 0.39 0.09 99.12<br />

Fort Rock (08029904) 4 50.24 1.65 15.60 11.12 0.22 5.90 8.78 3.55 0.45 0.34 0.09 97.95<br />

Subglacial Volcanic<br />

Mosfell–less palagonitized 9 47.05 1.50 15.60 10.82 0.21 8.65 12.68 2.17 0.05 0.22 0.02 98.96<br />

sample (MOS‐1)<br />

Mosfell–more palagonitized 11 47.78 1.67 16.01 11.11 0.18 8.70 12.66 2.21 0.05 0.14 0.02 100.53<br />

sample (MOS‐2)<br />

Ingolfsfjall (ING‐1) 9 48.96 1.72 15.04 11.92 0.20 7.90 11.72 2.45 0.14 0.25 0.13 100.44<br />

Figure 5. Total alkalies versus silica plot <strong>of</strong> sideromelane compositions, averaged at each locality sampled. Black<br />

solid symbols: submarine volcaniclastic group; violet solid symbols: seafloor volcanic group; blue solid symbols:<br />

marine phreatomagmatic group; red open symbols: lacustrine phreatomagmatic group; green symbols: subglacial volcanic<br />

group. Sideromelane composition sampled in this study covers a broad range and is fairly evenly distributed<br />

across alkaline and subalkaline compositions.<br />

10 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Table 5.<br />

Average Palagonite Major‐Element Compositions (wt. % Oxide), Determined Using EPMA<br />

Sample Group<br />

Location<br />

(Sample ID)<br />

Number<br />

<strong>of</strong><br />

Concentration (wt. % Oxide):<br />

Palagonite<br />

Analyses SiO 2 TiO 2 Al 2 O 3 FeO MnO MgO CaO Na 2 O K 2 O<br />

P 2 O 5 SO 2 Total H 2 O a<br />

Submarine Volcaniclastic<br />

HSDP (R0777‐17.4) 9 40.80 5.15 9.88 11.50 0.00 0.97 10.07 0.91 0.16 0.10 0.02 79.62 20.38<br />

SOH‐1 (3135) 9 41.87 4.70 9.31 15.45 0.16 2.73 9.23 0.91 0.35 0.37 0.11 85.19 14.81<br />

Palagonia (PAL‐1) 9 44.54 3.27 10.65 16.16 0.11 1.86 7.13 1.83 0.32 0.04 0.03 85.99 14.01<br />

Hilina Bench (S508‐7) (See Table 6)<br />

Seafloor Volcanic<br />

Davidson (T430‐R8) 8 38.25 3.12 10.47 12.32 0.00 1.09 0.79 4.57 1.94 0.21 0.15 72.91 27.09<br />

North Arch (26D‐6) 4 26.40 4.47 4.93 20.96 0.03 0.83 0.71 2.24 0.83 0.00 0.11 61.57 38.43<br />

O’ahu (T298‐R17) 9 31.41 4.87 16.08 16.84 0.00 2.06 2.40 3.07 1.13 1.70 0.14 79.70 20.30<br />

Marine Phreatomagmatic<br />

Diamond Head (02010405) 8 43.06 4.25 7.83 19.07 0.02 0.78 4.30 0.34 0.43 0.08 0.01 80.21 19.79<br />

Surtsey–less palagonitized 9 27.80 3.99 7.55 16.41 0.19 2.22 6.87 0.72 0.16 0.70 0.00 66.64 33.36<br />

sample (07010404)<br />

Surtsey–more palagonitized 9 30.39 3.81 12.84 18.46 0.16 3.31 6.89 0.04 0.04 0.30 0.00 76.39 23.61<br />

sample (07010405)<br />

Heimaklettur (07020402) 6 37.01 4.28 10.40 17.23 0.07 2.31 4.52 0.45 0.66 0.00 0.01 77.73 22.27<br />

Lacustrine Phreatomagmatic<br />

Black Point (09120301) 9 36.99 2.92 9.49 12.98 0.06 1.12 6.76 1.85 0.78 0.20 0.01 73.20 26.80<br />

Pahvant Butte (10230304) 3 36.81 2.91 8.25 16.35 0.05 0.55 8.06 0.68 0.14 0.06 0.01 73.88 26.12<br />

Fort Rock (08029904) 5 34.07 2.27 10.08 15.46 0.01 1.05 4.66 1.27 0.25 0.11 0.01 69.28 30.72<br />

Subglacial Volcanic<br />

Mosfell–less palagonitized 9 42.58 2.58 11.15 17.35 0.07 0.75 9.83 0.17 0.05 0.08 0.03 84.66 15.34<br />

sample (MOS‐1)<br />

Mosfell–more palagonitized 3 46.81 2.79 13.54 13.12 0.05 2.92 6.59 0.19 0.05 0.05 0.03 86.34 13.66<br />

sample (MOS‐2)<br />

Ingólfsfjall (ING‐1) 9 42.33 3.33 11.00 17.23 0.14 3.08 8.34 0.14 0.16 0.11 0.09 85.99 14.01<br />

a Calculated as 100‐Total.<br />

alkaline and subalkaline palagonitized hyaloclastite<br />

grains (Figure 6) in this sample are given in Table 6.<br />

The alkaline versus subaklaline compositions <strong>of</strong> the<br />

sideromelane grains are clearly distinguishable<br />

(Table 6 and Figure 6 (insets)), as are those <strong>of</strong> the<br />

corresponding palagonites (Table 6). Also, it can be<br />

seen in Figure 6 that the palagonitized margins <strong>of</strong> the<br />

subalkaline grains are relatively thick (Table 2) and<br />

deep‐orange colored, whereas palagonitized margins<br />

<strong>of</strong> the alkaline grains are relatively thin and paleorange<br />

colored.<br />

[17] The derived palagonite water contents <strong>of</strong> all<br />

samples are given in Tables 5 and 6, and are plotted<br />

(by sample location and environmental group) in<br />

Figure 7. The averaged group values for water<br />

content range from approximately 14 (subglacial<br />

volcanic group) to 29 (seafloor volcanic group)<br />

wt. % oxide. The values are generally consistent<br />

within the seafloor volcanic, lacustrine phreatomagmatic,<br />

and subglacial volcanic groups, although<br />

the less‐palagonitized Surtsey sample has significantly<br />

higher water content than the other marine<br />

phreatomagmatic samples. The palagonitized rinds<br />

developed on the 2 subalkaline grains analyzed in<br />

Hilina Bench sample S508–7 (Figure 6 and Table 6)<br />

have much higher water content (25% average) than<br />

for the 2 alkaline grains (17% average). Water<br />

content (Tables 6 and 7) versus palagonitization<br />

extent (Table 2) is plotted in Figure 8. Water content<br />

generally decreases with increasing palagonitization<br />

extent. This trend also holds for the less‐ and morepalagonitized<br />

Surtsey and Mosfell samples, with<br />

both sample sets showing lower water content with<br />

higher palagonitization extent. No correlation<br />

between sample age and sample water content was<br />

observed.<br />

4.4. Trace Element Composition<br />

[18] The raw trace element data are presented in<br />

Table S1 in the auxiliary material. 1 Multielement<br />

spider diagrams were prepared for each sample<br />

1 Auxiliary materials are available in the HTML. doi:10.1029/<br />

2011GC003639.<br />

11 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 6. Thin section photomicrographs (PPL) and TAS plot (inset) for Hilina Bench sample S508–7. Sideromelane<br />

<strong>of</strong> hyaloclastite grains (a) 12 and 13 and (b) 18 and 19 are alkaline and subalkaline, respectively (Table 6). Note<br />

that the palagonitized rinds on the subalkaline grains are colored a deeper shade <strong>of</strong> orange than those <strong>of</strong> the alkaline<br />

grains (Table 2), which are relatively pale (the sideromelanes are also distinguishably colored). The palagonitized<br />

rinds developed on the subalkaline grains are also significantly thicker than those on the alkaline grains.<br />

Table 6.<br />

Compositional Analyses (EPMA) <strong>of</strong> Sideromelane and Palagonite Grains in Hilina Bench Sample S508–7 a<br />

SiO 2 TiO 2 Al 2 O3 FeO MnO<br />

Concentration<br />

(wt. % Oxide)<br />

MgO CaO Na 2 O<br />

K 2 O P 2 O 5 SO 2 Total H 2 O b<br />

Sideromelane<br />

S508‐7 G12‐1 (alkaline) 47.81 4.17 12.96 13.02 0.18 6.08 11.13 2.93 0.64 0.46 0.08 99.45<br />

S508‐7 G12‐2 (alkaline) 47.98 4.21 13.50 13.20 0.18 6.08 11.14 2.92 0.75 0.55 0.01 100.52<br />

S508‐7 G18‐1 (alkaline) 46.22 4.68 13.79 13.44 0.16 5.96 11.13 3.25 0.84 0.70 0.07 100.24<br />

S508‐7 G18‐2 (alkaline) 46.32 4.64 13.98 13.53 0.25 5.92 11.15 3.25 0.88 0.66 0.04 100.61<br />

S508‐7 G18‐3 (alkaline) 46.01 4.56 13.63 13.24 0.20 6.02 11.35 3.31 0.89 0.47 0.02 99.70<br />

average alkaline 46.87 4.45 13.57 13.28 0.19 6.01 11.18 3.13 0.80 0.57 0.04 100.10<br />

S508‐7 G13‐1 (subalk.) 52.29 2.30 13.21 11.44 0.18 6.84 10.96 2.27 0.32 0.41 0.02 100.25<br />

S508‐7 G13‐2 (subalk.) 52.24 2.33 13.51 11.46 0.20 6.72 10.93 2.38 0.31 0.49 0.00 100.55<br />

S508‐7 G13‐3 (subalk.) 52.57 2.29 13.40 11.20 0.20 6.74 10.98 2.26 0.33 0.37 0.00 100.34<br />

S508‐7 G19‐1 (subalk.) 53.70 2.20 13.47 10.98 0.20 7.00 10.82 2.36 0.31 0.36 0.00 101.40<br />

S508‐7 G19‐2 (subalk.) 52.95 2.14 13.73 10.82 0.18 7.13 10.91 2.29 0.29 0.32 0.03 100.78<br />

average subalkaline 52.75 2.25 13.46 11.18 0.19 6.89 10.92 2.31 0.31 0.39 0.01 100.66<br />

Palagonite<br />

S508‐7 G12‐1 (alkaline) 36.75 6.97 6.53 12.27 0.10 3.45 3.59 1.95 1.31 0.15 0.09 73.14 26.86<br />

S508‐7 G12‐2 (alkaline) 38.76 7.60 7.24 12.17 0.08 2.86 3.90 1.63 1.05 0.17 0.08 75.53 24.47<br />

S508‐7 G18‐1 (alkaline) 39.08 8.32 8.10 10.51 0.16 2.78 4.73 1.19 1.05 0.14 0.13 76.19 23.81<br />

S508‐7 G18‐2 (alkaline) 39.56 8.33 7.65 11.30 0.04 2.58 4.88 1.98 0.98 0.00 0.14 77.45 22.55<br />

S508‐7 G18‐3 (alkaline) 36.45 7.48 6.42 12.65 0.18 3.10 3.59 1.57 1.15 0.08 0.00 72.67 27.33<br />

average alkaline 38.12 7.74 7.19 11.78 0.11 2.95 4.14 1.66 1.11 0.11 0.09 74.99 25.0<br />

S508‐7 G13‐1 (subalk.) 39.72 4.57 8.03 18.79 0.11 3.14 2.62 1.96 1.46 0.00 0.32 80.71 19.29<br />

S508‐7 G13‐2 (subalk.) 42.01 4.74 9.60 17.23 0.18 2.67 3.39 1.93 1.45 0.15 0.19 83.53 16.47<br />

S508‐7 G13‐3 (subalk.) 42.09 4.79 8.40 17.66 0.07 3.79 2.76 2.20 1.37 0.11 0.07 83.29 16.71<br />

S508‐7 G19‐1 (subalk.) 43.29 4.51 10.22 15.50 0.14 2.87 3.66 2.27 1.45 0.00 0.02 83.91 16.09<br />

S508‐7 G19‐2 (subalk.) 42.46 4.35 9.57 15.96 0.07 2.62 3.33 2.44 1.56 0.01 0.08 82.45 17.55<br />

average subalkaline 41.91 4.59 9.16 17.03 0.11 3.02 3.15 2.16 1.46 0.05 0.14 82.78 17.22<br />

a Grains 12 and 18 consist <strong>of</strong> alkaline sideromelane (dark blue); grains 13 and 19 consist <strong>of</strong> subalkaline sideromelane (red).<br />

b Calculated as 100‐Total.<br />

12 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 7. Plot <strong>of</strong> derived water content (wt. % oxide; EPMA) versus sample location. Black: submarine volcaniclastic<br />

group; violet: seafloor volcanic group; blue: marine phreatomagmatic group; red: lacustrine phreatomagmatic<br />

group; green: subglacial volcanic group. Values within the seafloor volcanic and subglacial volcanic groups are fairly<br />

consistent; each <strong>of</strong> the other groups contains at least one sample with an inconsistent value. Average group values<br />

(indicated by the black lines) range from approximately 14 (subglacial volcanic group) to 29 wt. % oxide (seafloor<br />

volcanic group).<br />

Table 7. Major‐Element and Overall Mass Change (%) for Subalkaline and Alkaline Samples, Determined Graphically Using<br />

Isocon Diagrams (Figure 10 and Figure S2) a<br />

Location<br />

Mass Change (%) Element (Oxide)<br />

SiO 2 TiO 2 Al 2 O 3 FeO MgO CaO Na 2 O K 2 O<br />

Overall Mass<br />

Change (%)<br />

(Subalkaline Sideromelane)<br />

HSDP* −71.7 −31.1 −70.9 −68.4 −96.1 −63.9 −87.4 −77.2 −63.7<br />

SOH‐1* −54.3 6.3 −64.2 −21.8 −63.7 −41.6 −82.5 −49.5 −42.6<br />

Palagonia* −56.3 2.8 −66.5 2.2 −82.7 −59.1 −69.1 −41.4 −47.0<br />

Ingólfsfjall −58.9 −5.4 −62.8 −41.7 −84.3 −61.8 −97.9 −38.9 −53.3<br />

Hilina Bench (subalk. grain)* −48.7 34.0 −59.8 1.0 −64.0 −83.9 −37.6 166.0 −35.9<br />

Avg. Subalkaline Sideromelane −58.0 1.3 −64.8 −25.7 −78.1 −62.1 −74.9 −8.2 −48.5<br />

(Alkaline Sideromelane)<br />

Davidson* −53.0 −47.6 −70.3 −17.5 −88.2 −95.0 −33.6 8.1 −43.2<br />

Surtsey (more palagonitized sample) −56.3 18.5 −33.8 8.7 −64.8 −49.8 −91.8 −91.8 −27.4<br />

O’ahu* −50.8 −8.1 −1.5 3.7 −83.0 −90.6 −69.4 −65.2 −33.0<br />

Black Point −40.8 109.5 −64.6 44.2 −71.3 −45.6 −69.5 −22.5 −23.4<br />

Pahvant Butte −53.0 7.8 −63.2 −5.5 −91.9 −27.8 −89.2 −92.8 −36.3<br />

Diamond Head −17.9 11.1 −57.7 32.3 −88.6 −79.9 −91.9 −65.3 −19.6<br />

Fort Rock −53.8 2.4 −53.7 3.8 −88.6 −71.0 −44.0 −60.7 −29.0<br />

Heimaklettur −55.9 27.6 −66.9 19.7 −83.9 −76.9 −63.5 −69.1 −38.1<br />

Hilina Bench (alkaline grain)* −26.4 54.7 −48.9 −31.9 −59.4 −63.0 −68.1 8.8 −13.0<br />

Avg. Alkaline Sideromelane −45.3 19.5 −51.2 6.4 −80.0 −66.6 −69.0 −50.0 −29.2<br />

*Oceanic sample average (this study) −51.6 1.6 −54.6 −19.0 −76.7 −71.0 −64.0 −7.2<br />

Staudigel and Hart [1983, Table 2]; MORB −50.0 n/a −55.0 −12.0 −67.0 −88.0 −81.0 900.0<br />

a The averaged values for the oceanic samples are compared to MORB values calculated by Staudigel and Hart [1983] at the bottom <strong>of</strong> the table<br />

(see section 5).<br />

13 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 8. Plot <strong>of</strong> palagonite water content versus palagonitization extent. Black solid symbols: submarine volcaniclastic<br />

group; violet solid symbols: seafloor volcanic group; blue solid symbols: marine phreatomagmatic group; red<br />

open symbols: lacustrine phreatomagmatic group; green symbols: subglacial volcanic group. Samples with lower<br />

water contents generally palagonitized to a greater extent, however this trend is not evident in the seafloor volcanic<br />

and lacustrine phreatomagmatic groups. The less‐ and more‐palagonitized sample pairs from Surtsey and Mosfell<br />

show this trend (indicated by arrows connecting symbols).<br />

(Figures 9 and S1). The sampled palagonite REE<br />

patterns among the different environment groups<br />

and for localities within each environment group<br />

are broadly similar. Sr, K, Rb, and Ba in palagonite<br />

relative to sideromelane are highly variable across<br />

the sample set. Sr and K in palagonite vary from<br />

depleted to enriched (relative to sideromelane),<br />

independent <strong>of</strong> environment sample group. Rb and<br />

Ba in palagonite, for the majority <strong>of</strong> the samples,<br />

are depleted (in some cases greatly) relative to<br />

sideromelane. For most <strong>of</strong> the samples, Nb through<br />

Lu (including REEs) palagonite concentration<br />

Figure 9. Chondrite‐normalized [McDonough and Sun, 1995] multielement spider diagrams for (a) a subalkaline<br />

grain and (b) an alkaline grain in Hilina Bench sample S508–7 (submarine volcaniclastic environment group). Colors<br />

and analyses numbers as in Figure S1. Note the difference in Nb through Lu (immobile elements) concentration<br />

relative to sideromelane for the palagonitized subalkaline versus alkaline grain.<br />

14 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 10. Isocon diagrams [Grant, 1986] used for determination <strong>of</strong> elemental and overall mass change for palagonite<br />

developed on (a) a subalkaline grain (plotted trace element analyses Sid. One and average <strong>of</strong> Pal. One and Pal. 2) and<br />

(b) an alkaline grain (plotted trace element analyses Sid. One and average <strong>of</strong> Pal. One and Pal. 2) in Hilina Bench sample<br />

S508–7 (submarine volcaniclastic environment group). For both samples, plotted major‐element analyses and trace<br />

element analyses were carried out using EPMA and LA‐ICPMS, respectively, at identical spots in sideromelane and<br />

palagonite. Plotted palagonite major‐ and trace element analyses were made as close to the center <strong>of</strong> palagonitized rinds<br />

as possible. Plotted sideromelane major‐ and trace element analyses were made at identical spots in sideromelane, which<br />

in all samples is homogeneous with respect to major‐ and trace element composition. Plotted major‐element sideromelane<br />

and palagonite analyses (green symbols) are given in Table S2; plotted trace element analyses (red and blue<br />

symbols) are given in Table S1. C A : factored concentration in altered sideromelane (palagonite); C O : original sideromelane<br />

factored concentration (see text). Isocon line (red) was determined by linear regression using Zr, Nb, Sc, La,<br />

and Nd. The slope <strong>of</strong> the isocon line is significantly steeper for palagonitized subalkaline (35.9% overall mass loss) than<br />

for alkaline sideromelane (13.0% overall mass loss).<br />

patterns are parallel to and show higher concentrations<br />

than the corresponding sideromelane patterns.<br />

Walton et al. [2005], using LA‐ICPMS, found<br />

similar behavior in palagonite relative to sideromelane<br />

REE patterns for HSDP hyaloclastite<br />

samples, as did Daux et al. [1994], using inductively<br />

coupled plasma atomic emission spectrometry<br />

(ICP‐AES) to analyze chemically separated<br />

sideromelane and palagonite fragments from Icelandic<br />

subglacially erupted hyaloclastite samples.<br />

The immobile‐element sections <strong>of</strong> the spider patterns<br />

<strong>of</strong> the present study indicate that essentially<br />

no systematic chemical fractionation <strong>of</strong> these<br />

elements occurred during palagonitization (a few<br />

samples show the LREEs to be slightly less enriched<br />

than the HREEs). This finding is consistent<br />

with Daux et al. [1994], who found no evidence for<br />

REE fractionation among a suite <strong>of</strong> 11 Icelandic<br />

palagonitized hyaloclastites. For samples with<br />

sufficiently thick palagonitized rinds, so that palagonite<br />

analyses could be determined along a traverse<br />

(Pal. 1, Pal. 2, etc. in Figures 9 and S1),<br />

immobile‐element concentration decreases systematically<br />

from the outer palagonitized rim inward<br />

toward the fresh sideromelane. Immobile‐element<br />

(Nb through Lu) concentration relative to sideromelane<br />

is greater for subalkaline (Figures S1a,<br />

S1b, S1c, and S1m) as compared to alkaline<br />

(Figures S1d, S1f, S1g, S1h, S1i, S1j, S1k, and<br />

S1l) palagonite. This is also the case for subalkaline<br />

and alkaline grains analyzed in Hilina Bench<br />

sample S508–7 (Figure 9). Only the North Arch<br />

palagonite trace element pattern (Figure S1e) does<br />

not follow the trends described above, and shows<br />

a distinct Ce anomaly.<br />

4.5. Major‐Element Concentration<br />

Changes<br />

[19] Isocon diagrams [Grant, 1986] were prepared for<br />

each sample in order to investigate major‐element<br />

concentration changes during palagonitization. This<br />

method allowed a test <strong>of</strong> a constant volume or constant<br />

mass assumption based on immobile element<br />

concentrations. The trace‐ and major‐element concentration<br />

analyses (Tables S1 and S2, respectively)<br />

used for these diagrams were from LA‐ICPMS and<br />

EPMA analyses performed at approximately the<br />

center <strong>of</strong> the palagonitized outer margins and at as<br />

close to identical spots as possible for each sample.<br />

15 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 11. Mass loss (Table 7) versus palagonitization extent (Table 2). Samples plot along a general trend <strong>of</strong><br />

increasing mass loss with increasing extent (arrow). Subalkaline samples for the most part have greater mass loss<br />

values than alkaline samples. The alkaline and subalkaline fields overlap for extent values between 70 and 85%,<br />

however in this region, subalkaline mass loss values are generally greater.<br />

The palagonite compositions were plotted against the<br />

analyses <strong>of</strong> adjacent sideromelane. Major‐element<br />

(wt. % oxide) and trace element (ppm) concentrations<br />

were scaled, using appropriate factors, for<br />

plotting. The actual trace‐ and major‐element analyses<br />

used to construct each sample’s isocon diagram<br />

are given with Figures 10 and S2. Zr, Nb, Sc, La, and<br />

Nd concentrations consistently plotted along isocon<br />

lines for almost all <strong>of</strong> the samples (Figures 10 and<br />

S2). These elements therefore can be considered<br />

immobile. Major‐element oxide as well as Rb, V,<br />

and Cr concentrations did not consistently plot<br />

along the isocon lines, hence these elements are<br />

considered mobile. Concentration changes for mobile<br />

elements, as well as overall mass loss, were determined<br />

graphically using each sample’s isocon diagram<br />

and the lever rule [Grant, 1986].<br />

[20] The mass loss results fall into two categories:<br />

palagonites in samples <strong>of</strong> subalkaline sideromelane<br />

(Figures 10a, S2a, S2b, S2c, and S2m) exhibit relatively<br />

higher (35.9 to 63.7%) mass loss; palagonites<br />

in samples <strong>of</strong> alkaline sideromelane (Figures 10b,<br />

S2d, S2f, S2g, S2h, S2i, S2j, S2k, and S2l) exhibit<br />

relatively lower (13.0 to 43.2%) mass loss (Table 7).<br />

In particular, the palagonites in samples <strong>of</strong> subalkaline<br />

sideromelane lost significant amounts Si,<br />

Ti, Al, and Fe, whereas in palagonites <strong>of</strong> samples <strong>of</strong><br />

alkaline sideromelane, relatively less Si and Al were<br />

lost, and, on average, Ti and Fe were gained. These<br />

trends are also observed for the subalkaline and<br />

alkaline grains in Hilina Bench sample S508–7<br />

(Figure 10 and Table 7). Overall mass loss versus<br />

palagonitization extent is plotted in Figure 11. Mass<br />

loss generally increases linearly with increasing<br />

palagonitization extent. Samples <strong>of</strong> subalkaline<br />

sideromelane that palagonitized to a relatively high<br />

extent also had relatively high mass loss. For a given<br />

palagonitization extent, samples <strong>of</strong> subalkaline palagonitized<br />

sideromelane lost more mass than alkaline<br />

samples. The North Arch Zr, Nb, Sc, La, and Nd<br />

concentrations do not plot along a line (Figure S2e),<br />

so the isocon approach cannot be used to determine<br />

valid palagonitization mass changes for this sample.<br />

The North Arch sample was the only Mn oxideencrusted<br />

sample studied (Table 1). Previous studies<br />

indicate that Mn and Fe oxides comprising marine<br />

ferromanganese crusts scavenge trace metals<br />

[Koschinsky and Hein,2003;Hein et al., 1997]. The<br />

anomalous mobile behavior <strong>of</strong> Zr, Nb, Sc, La, and<br />

Nd in the North Arch sample we interpret to be due<br />

to variable sequestering <strong>of</strong> these elements within<br />

the Mn oxide crust. Furthermore, we interpret the<br />

observed negative Ce anomaly to be the result <strong>of</strong><br />

significant oxidative accumulation <strong>of</strong> Ce within<br />

the Mn oxide crust, as observed in ferromanganese<br />

crusts from the Central Pacific by Bau and<br />

Koschinsky [2009].<br />

16 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 12. Major‐element concentration traverse, SOH‐1 (submarine volcaniclastic environment group; subalkaline<br />

sideromelane). (top) BSE photomicrograph showing the palagonite traverse in red and the sideromelane traverse in<br />

brown. Spot‐spacing is given on the lower image. (bottom) Palagonite major‐element concentration (Table S3),<br />

plotted as a percentage <strong>of</strong> the outermost (Spot 1) palagonite concentration, for each spot along the palagonite traverse.<br />

4.6. Major‐Element Concentration<br />

Gradients<br />

[21] Major‐element spot analyses along traverses<br />

through palagonite were performed on 5 <strong>of</strong> the samples<br />

with relatively thick palgonitized rinds. The raw<br />

data are given in Table S3, and the traverse details and<br />

plotted results are given in Figures 12 and S3a–S3d.<br />

In each sample, the sideromelane is homogeneous<br />

with respect to major‐element composition. For all<br />

5 <strong>of</strong> the palagonites, a distinct MgO concentration<br />

gradient, highest at the outer palagonitized margin<br />

and decreasing inward toward the sideromelane<br />

contact to only about 50% <strong>of</strong> the outer margin value,<br />

was observed. Over the same interval, SiO2, Al2O3,<br />

and CaO concentrations generally increased by 5 to<br />

20%, whereas TiO2 and FeO concentrations generally<br />

remained steady or decreased.<br />

5. Discussion<br />

5.1. Palagonitization Extent<br />

[22] Previous workers have devised various means<br />

for the quantification <strong>of</strong> palagonitization extent,<br />

based for example on palagonitized rind thickness<br />

and major‐element concentration changes during<br />

palagonitization. Interpretation <strong>of</strong> palagonitized rind<br />

thickness is not straightforward. Rind thickness has<br />

been shown to relate to both palagonitization duration<br />

[Moore, 1966] and temperature [Moore et al.,<br />

17 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 13. Palagonitization extents determined using 3 different methods (see legend and text) for sampled localities<br />

in the present study (sample groups as indicated). The calculated extents vary widely for most <strong>of</strong> the samples.<br />

Petrography‐based reaction extents are 15 and 29% higher for the more palagonitized Surtsey and Mosfell (see<br />

PPL photomicrograph insets) samples respectively than for the corresponding less palagonitized samples (green<br />

arrows). The same values determined using the SiO 2 /MgO method are 29 and 25% (violet arrows). However, those<br />

determined using the TiO 2 method are −5 and −15% (blue arrows).<br />

1985; Jakobsson and Moore, 1986]. Crovisier et al.<br />

[1992] found palagonitized rind thickness not to be a<br />

good indicator <strong>of</strong> reaction progress. Studies using<br />

major‐element concentration data as the basis for<br />

reaction extent include Staudigel and Hart [1983],<br />

Crovisier et al. [1992], and Stroncik and Schmincke<br />

[2001], in which assumed passive accumulation <strong>of</strong><br />

TiO 2 , SiO 2 /MgO, or MgO, respectively, was used<br />

as the basis for determining elemental changes<br />

during palagonitization. Since the present study<br />

shows that TiO 2 , SiO 2 , and MgO mass change<br />

behaviors vary significantly among different samples<br />

(Table 7), palagonitization extent values based<br />

on these element behaviors cannot be used to<br />

compare palagonitization extent among samples.<br />

Furthermore, SiO 2 , TiO 2 and particularly MgO<br />

concentrations vary as a function <strong>of</strong> analysis location<br />

within a given palagonite rind due to the<br />

observed concentration gradients for these elements<br />

(Figures 12 and S3).<br />

[23] Since palagonitization extent could potentially<br />

depend on many factors, including initial sideromelane<br />

composition, palagonitization duration,<br />

available heat, water/rock (i.e., porosity), fluid<br />

chemistry, and analysis location, a petrographicbased<br />

approach to determining the extent <strong>of</strong> palagonitization,<br />

representing an average extent for<br />

the sample, rather than a chemically based extent<br />

<strong>of</strong> palagonitization seems appropriate for sample<br />

comparison. Here we suggest an alternative,<br />

petrography‐based approach, wherein palagonitization<br />

extent is calculated based on observed modal<br />

data (see section 3.1 and Table 2). Palagonitization<br />

extents calculated from both modal data (this study)<br />

and using the major‐element concentration data<br />

approach used by previous workers are compared in<br />

Figure 13; wide variations are apparent. At Surtsey<br />

and Mosfell, petrographically distinct (Figure 13,<br />

insets) samples were collected from areas that<br />

were less‐ and more‐palagonitized, based on field<br />

observations. Calculated extents using the SiO 2 /<br />

MgO method correlate with the petrographically<br />

observable difference in degree <strong>of</strong> palagonitization<br />

between the samples; the TiO 2 method does not.<br />

However, as mentioned above, the SiO 2 /MgO<br />

method is shown here to depend on analyzed spot<br />

location, particularly for MgO. The petrographic<br />

palagonitization extent provides a basis for comparison<br />

<strong>of</strong> other measured parameters in samples<br />

from different locations that are known to have<br />

palagonitized under different conditions. For<br />

example, in this study, palagonite water content is<br />

18 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

observed to decrease linearly with increasing palagonitization<br />

extent (Figure 8); this finding is discussed<br />

further below.<br />

5.2. Porosity Control <strong>of</strong> Palagonitization<br />

Extent<br />

[24] The observed general trend <strong>of</strong> increasing<br />

palagonitization extent with decreasing minuscement<br />

porosity (i.e., decreasing original porosity;<br />

Figure 3) across the sample set suggests that porosity<br />

plays a controlling role in the extent <strong>of</strong> palagonitization.<br />

Walton and Schiffman [2003] made a similar<br />

finding for palagonitized HSDP hyaloclastites. Less<br />

fluid flow would be expected within low‐porosity<br />

samples (due to the expected relative lack <strong>of</strong> permeability),<br />

so we can speculate that water/rock ratio<br />

(w/r) decreases with decreasing porosity. Therefore<br />

the trend observed in Figure 3 suggests that sample<br />

w/r is inversely related to palagonitization extent.<br />

Pauly [2011] showed that pH (presumably determined<br />

by sample w/r and thus porosity) controls B<br />

isotope fractionation between palagonite and seawater.<br />

In that study, palagonitized sideromelane<br />

samples having the lowest zeolite abundance (suggesting<br />

low palagonitization extent) also had the<br />

lightest d 11 B values, as would be expected for<br />

samples that palagonitized in relatively low pH,<br />

more porous environments (presumably where little<br />

if any porosity reduction due to burial has occurred).<br />

This finding confirms that porosity is an important<br />

control <strong>of</strong> palagonitization extent.<br />

[25] It is also apparent in Figure 3 that within the<br />

lacustrine phreatomagmatic and subglacial volcanic<br />

sample groups, palagonitization extent varies<br />

widely among samples with similar porosity. Palagonitization<br />

extents <strong>of</strong> 84% and 40% were measured<br />

at Pahvant Butte and Black Point (lacustrine<br />

phreatomagmatic samples) respectively, yet these<br />

samples have nearly identical porosities (Table 2).<br />

Evidently, within these sample groups, other factor(s)<br />

besides porosity, presumably including available<br />

heat and palagonitization duration, exert control on<br />

palagonitization extent.<br />

5.3. Non‐isovolumetric Process:<br />

Implications for Major‐Element Changes<br />

[26] Previous workers [Hay and Iijima, 1968a,<br />

1968b; Honnorez, 1967, 1972; Eggleton and Keller,<br />

1982; Staudigel and Hart, 1983; Furnes, 1984;<br />

Bednarz and Schmincke, 1989; Zhou and Fyfe,<br />

1989; Jercinovic et al., 1990; Thorseth et al.,<br />

1991; Crovisier et al., 1992; Daux et al., 1994;<br />

Stroncik and Schmincke, 2001; Walton et al., 2005]<br />

studied major‐ and/or trace element changes during<br />

palagonitization as a means <strong>of</strong> understanding the<br />

geochemical mechanism <strong>of</strong> palagonitization, using<br />

various methods and assumptions. Stroncik and<br />

Schmincke [2001] and Walton and Schiffman<br />

[2003] summarize these studies. This impressive<br />

number <strong>of</strong> studies, spanning nearly 40 years, gives<br />

an indication <strong>of</strong> the (still) elusive nature <strong>of</strong> elemental<br />

mobility during palagonitization.<br />

[27] In the present study, the availability <strong>of</strong> both<br />

major‐ and trace‐element data, obtained using in situ<br />

microanalysis at essentially the same spots within<br />

sampled palagonites, is exploited to determine<br />

major‐element and overall mass changes during<br />

palagonitization, using the isocon method [Grant,<br />

1986]. The observed large major‐element mass changes<br />

(Table 7) are well in excess <strong>of</strong> changes in solidphase<br />

specific gravity that can be attributed to<br />

hydration alone, and require extensive dissolution<br />

and creation <strong>of</strong> microporosity. This indicates that<br />

palagonitization cannot be an isovolumetric process,<br />

as previously suggested based on petrographic evidence<br />

[Walton and Schiffman, 2003; Stroncik and<br />

Schmincke, 2001; Jercinovic et al., 1990]. Even<br />

though, when observed petrographically, preserved<br />

primary texture and absence <strong>of</strong> secondary pores<br />

within palagonitized areas suggests an isovolumetric<br />

process, evidently significant microporosity develops<br />

during palagonitization, potentially in the leached<br />

layers imaged within gel palagonite by Drief<br />

and Schiffman [2004]. These workers concluded<br />

that the leached layers result from selective dissolution,<br />

and form palagonite. The finding that palagonitization<br />

is not an isovolumetric process calls<br />

into question element changes that have previously<br />

been calculated in weight per unit volume [Daux et al.,<br />

1994; Crovisier et al., 1992; Hay and Iijima, 1968a],<br />

and suggests instead that the isocon approach <strong>of</strong> Grant<br />

[1986], in which mobile element concentration changes<br />

are determined based on immobile element concentration<br />

changes, is necessary to understand mass<br />

transfer during palagonitization.<br />

5.4. Major‐Element Changes: Implications<br />

for Seawater Cation Flux<br />

[28] Major‐element concentration changes determined<br />

from immobile‐element isocon lines in the<br />

present study (Table 7) confirm that palagonitization<br />

<strong>of</strong> sideromelane can be a significant source (or<br />

sink) <strong>of</strong> oceanic cations. The averaged element<br />

changes determined for oceanic samples in the<br />

present study are for the most part in reasonable<br />

agreement with elemental changes calculated for<br />

19 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

palagonitization <strong>of</strong> MORB crust by Staudigel and<br />

Hart [1983] (Table 7). The lone exception is<br />

K 2 O, for which average change calculated for<br />

oceanic samples in the present study is −12%,<br />

whereas Staudigel and Hart [1983] calculated<br />

a value <strong>of</strong> 900%. An important observation is<br />

that both overall and specific major‐element mass<br />

changes depend strongly on original sideromelane<br />

composition. In particular, averaged TiO 2 and FeO<br />

changes range from 1.3 and −25.7% respectively<br />

during palagonitization <strong>of</strong> subalkaline sideromelane<br />

(i.e., neutral to a source <strong>of</strong> seawater cations) to 19.5<br />

and 6.4% respectively during palagonitization <strong>of</strong><br />

alkaline sideromelane (i.e., sink <strong>of</strong> oceanic cations).<br />

This finding points out the need to account for variable<br />

mass changes due to sideromelane composition<br />

when performing global oceanic flux calculations.<br />

5.5. Compositional Control<br />

[29] Palagonitized rinds on the subalkaline and<br />

alkaline grains <strong>of</strong> Hilina Bench sample S508–7 were<br />

observed to have distinguishable major‐element<br />

compositions, even though they were palagonitized<br />

under identical conditions (Table 6). In addition, the<br />

average overall mass loss values across the entire<br />

sample set are very different for palagonitization<br />

<strong>of</strong> alkaline (26%) versus subalkaline (48%) sideromelane<br />

(Table 7). It is clear that palagonite<br />

composition depends on original sideromelane<br />

composition. This finding is potentially inconsistent<br />

with Stroncik and Schmincke [2001], who found<br />

palagonite composition to be mainly controlled by<br />

palagonite aging, beginning with the first appearance<br />

<strong>of</strong> palagonitized sideromelane. However, it<br />

should be noted that for the most part the samples<br />

analyzed by Stroncik and Schmincke [2001] are<br />

significantly older than those <strong>of</strong> the present study;<br />

also all samples analyzed in the present study are gel<br />

palagonite.<br />

5.6. Subalkaline Versus Alkaline<br />

Sideromelane Dissolution Rate<br />

[30] The substantial difference in average overall<br />

mass loss values (Table 7) indicate that subalkaline<br />

sideromelane dissolves much more readily than<br />

alkaline. This observation cannot be attributed to<br />

differences in palagonitization duration or conditions,<br />

because subalkaline versus alkaline sideromelane<br />

grains in Hilina Bench sample S508–7,<br />

which must have palagonitized for identical duration<br />

and conditions, lost 35.9% and 13.0% mass,<br />

respectively (Figure 10 and Table 7). The thicker<br />

versus thinner palagonitized rinds that developed<br />

on the subalkaline versus alkaline grains in this<br />

sample (Figure 6) are therefore evidence that subalkaline<br />

sideromelane must dissolve at a relatively<br />

higher rate than alkaline sideromelane during<br />

palagonitization. This is strong evidence that sideromelane<br />

dissolution rate depends on sideromelane<br />

composition. An implication <strong>of</strong> this finding is that<br />

palagonitized rind thickness gives a true indication<br />

<strong>of</strong> palagonitization duration, as suggested by Moore<br />

[1966], only if the original sideromelane compositions<br />

in the sample set are identical.<br />

5.7. Water Content<br />

[31] The observation that palagonite water content<br />

decreases linearly with palagonitization extent<br />

(Figure 8) holds for samples from all <strong>of</strong> the environmental<br />

groups, including localities known or<br />

inferred to have palagonitized over a relatively short<br />

duration (volcanic and phreatomagmatic environments)<br />

and localities known or inferred to have<br />

palagonitized over a relatively long duration (submarine<br />

volcaniclastic environments). In particular,<br />

water contents <strong>of</strong> the less‐ and more‐palagonitized<br />

samples from both Surtsey and Mosfell (subglacial<br />

volcanic group) are higher and lower, respectively<br />

(Figure 8, blue and green arrows). Presumably<br />

palagonitization duration is similar for each sample<br />

pair from Surtsey and Mosfell. This indicates that<br />

water content depends on palagonitization extent<br />

(or possibly temperature), for a given palagonitization<br />

duration (assuming water content to be stable<br />

post‐palagonitization). Drief and Schiffman [2004,<br />

Figure 2d] showed that palagonite consists <strong>of</strong><br />

amorphous and smectite‐like areas. The higher<br />

water contents observed in relatively low palagonitization<br />

extent samples in the present study might<br />

reflect a higher abundance <strong>of</strong> amorphous (and<br />

presumably more hydrated) material relative to<br />

smectite‐like areas. Similarly, the lower water contents<br />

observed in relatively high palagonitization<br />

extent samples might reflect higher degrees <strong>of</strong> crystallinity,<br />

consistent with inferred recrystallization <strong>of</strong><br />

palagonite to smectite (see section 5.9).<br />

5.8. Two Distinct Palagonitization Styles<br />

[32] The submarine volcaniclastic samples have<br />

remarkably similar low porosity values, high zeolite<br />

abundances, thick, smooth, gel‐like palagonite rind<br />

textures, and high palagonitization extents (Figure 2<br />

and Table 2). The submarine volcanic, phreatomagmatic<br />

(both marine and lacustrine) and subglacial<br />

volcanic samples generally are much more<br />

variable in terms <strong>of</strong> these same parameters (Figure 3<br />

20 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

and Table 2); palagonitization extent within these<br />

groups is particularly variable. These observations<br />

suggest two distinguishable and environmentally<br />

dependent palagonitization styles.<br />

[33] Burial‐diagenetic palagonitization. Palagonitization<br />

in submarine volcaniclastic localities is<br />

known, or is inferred, to have occurred significantly<br />

later than initial eruption, fragmentation, and deposition<br />

<strong>of</strong> hyaloclastites [e.g., Walton and Schiffman,<br />

2003]. In particular, the rounded sideromelane<br />

grains deposited in the volcaniclastic basin at Hilina<br />

Bench indicate that palagonitization must have<br />

occurred later (i.e., after transport) than the initial<br />

eruption and deposition <strong>of</strong> hyaloclastite. These<br />

samples have been buried; burial‐diagenesis is the<br />

likely source <strong>of</strong> thermal energy to drive palagonitization<br />

at these localities. Walton and Schiffman<br />

[2003] attributed the high degree <strong>of</strong> palagonitization<br />

observed near the bottom <strong>of</strong> the core to be the<br />

result <strong>of</strong> burial‐diagenesis. Presumably, as samples<br />

at submarine volcaniclastic localities underwent<br />

progressive burial (initial porosity loss, decrease in<br />

w/r), the temperature would have been relatively<br />

low; for example at HSDP palagonitization occurred<br />

at approximately 15°C [Walton and Schiffman,<br />

2003]. The rate <strong>of</strong> palagonitization would be<br />

expected to have been relatively low, but the duration<br />

<strong>of</strong> palagonitization might have been relatively<br />

long; Walton and Schiffman [2003] estimated that<br />

100 mm thick palagonitized hyaloclastite rinds<br />

developed over a 5000‐year interval in their HSDP<br />

samples. These factors in combination would be<br />

expected to produce the observed relatively high<br />

palagonitiation extent, smooth, complete palagonitized<br />

rinds, relatively low water content, and relatively<br />

high REE enrichment in all <strong>of</strong> these samples.<br />

[34] Hydrothermal palagonitization. On the other<br />

hand, palagonitization <strong>of</strong> the seafloor volcanic,<br />

phreatomagmatic, and subglacial volcanic samples<br />

is known or inferred to have occurred essentially<br />

syn‐eruptively or relatively shortly thereafter. The<br />

relatively high temperatures associated with the<br />

volcanic activity at these localities would be<br />

expected to establish a potentially vigorous hydrothermal<br />

system, flushing heated fluid through the<br />

relatively porous volcanic edifice. Palagonitization<br />

at these environments would be expected to have<br />

occurred at a higher rate (presumably due to the<br />

active hydrothermal system) as compared to palagonitization<br />

in submarine environments. For example<br />

at the surface <strong>of</strong> Surtsey, Jakobsson and Moore<br />

[1986] and Jakobsson et al. [2000] mapped rapid<br />

palagonitization (advancement <strong>of</strong> consolidated<br />

tephra areas visible during annual field visits) that<br />

was dependent on the volcano’s hydrothermal system.<br />

Palagonitization duration at these localities<br />

would be expected to be relatively short compared to<br />

that expected for submarine volcaniclastic systems,<br />

due to dependence on the longevity <strong>of</strong> the volcanic<br />

heat source. Palagonitized samples from these<br />

environments therefore would be expected to be<br />

distinguishable from submarine volcaniclastic samples<br />

in terms <strong>of</strong> palagonitized rind texture (relatively<br />

thin, less well‐developed), zeolite abundance (relatively<br />

low). Palagonitization extents for these samples<br />

would be expected to be relatively low,<br />

resulting in higher water contents as was generally<br />

observed.<br />

[35] Of special note however are the rind textures<br />

(complete and fairly smooth), zeolite abundances<br />

(high; Table 2), and palagonitization extents (high;<br />

Table 2) for the Ingólfsfjall (subglacial volcanic) and<br />

Pahvant Butte (lacustrine phreatomagmatic) samples;<br />

in terms <strong>of</strong> these parameters, these samples resemble<br />

submarine volcaniclastic samples. The observed<br />

greater palagonitization extent observed for these<br />

samples potentially could be due to exceptionally<br />

long palagonitization duration or exceptionally high<br />

palagonitization temperatures. Another possibility is<br />

a second stage <strong>of</strong> palagonitization that could occur in<br />

seafloor volcanic, marine phreatomagmatic, and<br />

subglacial volcanic environments due to subsequent<br />

burial‐diagenesis <strong>of</strong> the volcanic edifice. Jakobsson<br />

and Gudmundsson [2008] found that hydrothermal<br />

activity caused tephra palagonitization and consolidation<br />

<strong>of</strong> the edifices during the first years after the<br />

eruptions <strong>of</strong> Surtsey (1963–1967) and Gjálp (1996),<br />

but that alteration and consolidation <strong>of</strong> the outer<br />

slopes likely occurs over thousands <strong>of</strong> years and is<br />

mainly a diagenetic process. In summary, burialdiagenesis<br />

and hydrothermal palagonitization styles<br />

are distinguishable based on palagonitized rind<br />

petrographical and in situ analytical observations,<br />

however care must be taken interpreting the palagonitization<br />

<strong>of</strong> samples from seafloor volcanic,<br />

phreatomagmatic, and subglacial volcanic environments.<br />

Unambiguous distinction <strong>of</strong> palagonitization<br />

style requires additional information regarding<br />

the potential occurrence <strong>of</strong> post‐eruptive burialdiagenesis<br />

in these environments.<br />

5.9. Formation and Evolution<br />

<strong>of</strong> Alteration Layers<br />

[36] Palagonitization arguably must begin at the<br />

outer margin <strong>of</strong> fresh sideromelane shards, where<br />

initially sideromelane is in contact with fluid and<br />

begins to dissolve, and then proceed inward with<br />

21 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Figure 14. Schematic model depicting evolution and migration <strong>of</strong> sideromelane alteration layers. Arrows indicate<br />

evolution direction <strong>of</strong> the gel‐palagonite and phyllosilicate alteration layers with time. The contact between the phyllosilicate<br />

and gel‐palagonite layers likely is gradational (purple area). The observed SiO 2 ,Al 2 O 3 , and CaO (mobile)<br />

and Zr, Nb, Sc, La, Nd (immobile) concentration gradients are consistent with progressive sideromelane dissolution<br />

with time, however the observed MgO (mobile element) concentration gradient is not. The MgO concentration gradient<br />

is consistent with the syn‐ and/or post palagonitization development <strong>of</strong> a phyllosilicate (e.g., smectite) layer,<br />

forming as a consequence <strong>of</strong> the progressive evolution <strong>of</strong> the gel‐palagonite layer to equilibrium with the fluid.<br />

The inferred degree <strong>of</strong> development <strong>of</strong> smectite is highest at the fluid boundary.<br />

time; Walton and Schiffman [2003] showed this to<br />

be true for samples <strong>of</strong> palagonitized hyaloclastite<br />

from HSDP. The observed palagonite REE concentration<br />

gradients in the present study (Figures 9<br />

and S1) suggest therefore that REE concentration<br />

changes with time. This finding contrasts with<br />

Daux et al. [1994], who found that changes in<br />

palagonite REE concentration are independent <strong>of</strong><br />

reaction progress. Because we have found that the<br />

REEs are immobile during palagonitization, mobileelement<br />

mass loss must be greatest at the outer<br />

palagonitized margin, and decrease inward toward<br />

sideromelane. Furthermore, dissolution must continue<br />

behind the palagonitization front in order to account<br />

for the observed REE concentration gradients.<br />

[37] Because MgO was found to be highly mobile<br />

(Table 7), the observed MgO concentration gradients<br />

(decreasing through palagonite toward sideromelane;<br />

Figures 12 and S3) are inconsistent<br />

(opposite direction <strong>of</strong>) with gradients that would be<br />

expected due to cumulative dissolution with time ‐<br />

the MgO concentration gradients do not record<br />

sideromelane dissolution. Rather, they are consistent<br />

with syn‐ and/or post‐palagonitization conversion<br />

<strong>of</strong> gel‐palagonite to phyllosilicate (e.g.,<br />

smectite), beginning at the outer palagonitized<br />

margin and proceeding inward toward sideromelane<br />

with time. Valle et al. [2010] documented the formation<br />

and inward‐thickening with time <strong>of</strong> a phyllosilicate<br />

layer at the outer margin <strong>of</strong> gel‐palagonite<br />

layers that developed on synthetic borosilicate<br />

(nuclear waste containment) glass, and these workers<br />

suggest that similar behavior during alteration <strong>of</strong><br />

natural sideromelane would be expected. A schematic<br />

model depicting the evolution and migration<br />

<strong>of</strong> alteration layers in this framework is given in<br />

Figure 14. Valle et al. [2010] concluded that the gel<br />

(leached) layer developed on borosilicate glass surfaces<br />

is not at equilibrium with the solution,<br />

implying that this layer evolves with time. Honnorez<br />

[1967, 1972, 1978] and Stroncik and Schmincke<br />

[2001] concluded that the ultimate fate <strong>of</strong> gelpalagonite<br />

is conversion to smectite, and Drief and<br />

Schiffman [2004] documented both amorphous and<br />

crystalline areas within the outer palagonitized<br />

margins <strong>of</strong> natural sideromelane samples. The MgO<br />

concentration gradients within palagonites observed<br />

in the present study are consistent with these previous<br />

studies, and likely represent the continuing<br />

22 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

evolution (conversion) <strong>of</strong> the gel‐palagonite layer<br />

toward a phyllosilicate layer, which would be<br />

expected to be approaching equilibrium with the<br />

fluid.<br />

6. Conclusions<br />

[38] 1. Palagonite water content decreases with<br />

increasing palagonitization extent, likely due to<br />

recrystallization <strong>of</strong> palagonite into smectite.<br />

[39] 2. Palagonitization extent varies inversely with<br />

original sample porosity, confirming that porosity<br />

(i.e., associated permeability) is an important controlling<br />

factor <strong>of</strong> palagonitization; however, other<br />

factors are also important.<br />

[40] 3. Zr, Nb, Sc, and REE concentrations in<br />

palagonite and sideromelane plot along isocon<br />

lines, indicating that these elements are immobile<br />

during palagonitization.<br />

[41] 4. Large major‐element mass changes indicate<br />

that palagonitization entails extensive dissolution<br />

and creation <strong>of</strong> microporosity, and therefore cannot<br />

be considered an isovolumetric process, as has<br />

been assumed previously by many workers.<br />

[42] 5. Major‐element mass changes due to palagonitization<br />

for SiO 2 ,Al 2 O 3 , and in particular TiO 2<br />

and FeO depend on the subalkaline versus alkaline<br />

character <strong>of</strong> the sideromelane.<br />

[43] 6. Original sideromelane composition is an<br />

important control <strong>of</strong> palagonite composition.<br />

Therefore it is necessary to take original sideromelane<br />

composition (and resulting expected<br />

mass loss for specific elements) into account when<br />

calculating estimates <strong>of</strong> global seawater cation flux.<br />

[44] 7. Subalkaline sideromelane appears to dissolve<br />

at a much higher rate than alkaline sideromelane<br />

during palagonitization.<br />

[45] 8. Two distinct palagonitization styles, burialdiagenetic<br />

and hydrothermal, are recognized.<br />

[46] 9. Observed palagonite REE concentration<br />

gradients indicate that sideromelane dissolution<br />

continues in the zone behind the advancing palagonitization<br />

front.<br />

[47] 10. Observed palagonite MgO concentration<br />

gradients appear to reflect the early stages <strong>of</strong> synand/or<br />

post‐palagonitization conversion <strong>of</strong> the gelpalagonite<br />

layer to a phyllosilicate (e.g., smectite)<br />

layer, consistent with the evolution <strong>of</strong> sideromelane<br />

alteration layers toward equilibrium with the surrounding<br />

fluid.<br />

Acknowledgments<br />

[48] Michelle Coombs, Mike Garcia, and Don Thomas graciously<br />

provided some <strong>of</strong> the samples used in this study, as well<br />

as thoughtful discussions. José Honnorez provided the Palagonia<br />

sample. Roy Bailey and James White provided very helpful<br />

field guidance at Black Point and Pahvant Butte, respectively.<br />

Svienn Jakobsson and The Surtsey Committee granted access<br />

to Surtsey, and Hjalti Franzson supported and participated in<br />

the Iceland fieldwork. Tony Walton and Jim Moore provided<br />

very useful discussion support. Jake Lowenstern provided lab<br />

access and guidance for the reflected‐light FTIR measurements.<br />

Alicé <strong>Davis</strong> provided expert assistance with petrographic imaging;<br />

Norm Winter and Greg Baxter provided thin section sample<br />

preparation. Sarah Roeske and Brian Joy were <strong>of</strong> tremendous<br />

help during the extensive electron microprobe sessions, and<br />

Michelle Gras was equally helpful in the laser‐ablation ICPMS<br />

lab. This work was supported by NSF‐EAR grant 0125666. The<br />

authors thank José Honnorez and Tony Walton for very helpful<br />

reviews <strong>of</strong> the initial manuscript.<br />

References<br />

Bau, M., and A. Koschinsky (2009), Oxidative scavenging <strong>of</strong><br />

cerium on hydrous Fe oxide: Evidence from the distribution<br />

<strong>of</strong> rare earth elements and yttrium between Fe oxides and Mn<br />

oxides in hydrogenetic ferromanganese crusts, Geochem. J.,<br />

43, 37–47.<br />

Bednarz, U., and H.‐U. Schmincke (1989), Mass transfer during<br />

sub‐seafloor alteration <strong>of</strong> the upper Troodos crust (Cyprus),<br />

Contrib. Mineral. Petrol., 102, 93–101, doi:10.1007/<br />

BF01160193.<br />

Benson, L. V., S. P. Lund, J. W. Burdett, M. Kashgarian, T. P.<br />

Rose, J. P. Smoot, and M. Schwartz (1998), Correlation<br />

<strong>of</strong> Late‐Pleistocene lake‐level oscillation in Mono Lake,<br />

<strong>California</strong>, with North Atlantic climate events, Quat. Res.,<br />

49, 1–10, doi:10.1006/qres.1997.1940.<br />

Bishop, J. L., P. Schiffman, and R. Southard (2002), Geochemical<br />

and mineralogical analyses <strong>of</strong> palgonitic tuffs and altered<br />

rinds on pillow basalts in Iceland and applications to Mars, in<br />

Volcano‐Ice Interaction on Earth and Mars, editedbyJ.L.<br />

Smellie and M. G. Chapman, Geol. Soc. Spec. Publ., 202,<br />

371–392.<br />

Clague, D. A., R. T. Holcomb, J. M. Sinton, R. S. Detrick, and<br />

M. E. Torresan (1990), Pliocene and Pleistocene alkalic flood<br />

basalts on the seafloor north <strong>of</strong> the Hawaiian islands, Earth<br />

Planet. Sci. Lett., 98, 175–191, doi:10.1016/0012-821X(90)<br />

90058-6.<br />

Clague, D. A., K. Uto, K. Satake, and A. S. <strong>Davis</strong> (2002),<br />

Eruption style and flow <strong>of</strong> emplacement in the submarine<br />

North Arch Volcanic Field, Hawaii, in Hawaii Volcanoes<br />

Deep Underwater Perspectives, Geophys. Monogr. Ser.,<br />

vol. 128, edited by E. Takahashi et al. pp. 65–84, AGU,<br />

Washington D. C.<br />

Clague, D. A., J. B. Paduan, W. C. McIntosh, B. L. Cousens,<br />

A. S. <strong>Davis</strong>, and J. R. Reynolds (2006), A submarine perspective<br />

<strong>of</strong> the Honolulu volcanics, Oahu, J. Volcanol.<br />

Geotherm. Res., 151, 270–307.<br />

Clague, D. A., J. B. Paduan, R. A. Duncan, J. J. Huard, A. S.<br />

<strong>Davis</strong>, P. R. Castillo, P. Lonsdale, and A. DeVogelaere<br />

(2009), Five million years <strong>of</strong> compositionally diverse, episodic<br />

volcanism: Construction <strong>of</strong> Davidson Seamount atop<br />

23 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

an abandoned spreading center, Geochem. Geophys. Geosyst.,<br />

10, Q12009, doi:10.1029/2009GC002665.<br />

Cockell, C., and A. Herrera (2008), Why are some microorganisms<br />

boring?, Trends Microbiol., 16(3), 101–106, doi:10.1016/<br />

j.tim.2007.12.007.<br />

Coombs, M. L., T. W. Sisson, and P. W. Lipman (2006),<br />

Growth history <strong>of</strong> Kilauea inferred from volatile concentrations<br />

in submarine‐collected basalts, J. Volcanol. Geotherm.<br />

Res., 151, 19–49, doi:10.1016/j.jvolgeores.2005.07.037.<br />

Crovisier, J. L., J. Honnorez, and J. P. Eberhart (1987), Dissolution<br />

<strong>of</strong> basaltic glass in seawater; mechanism and rate,<br />

Geochim. Cosmochim. Acta, 51, 2977–2990, doi:10.1016/<br />

0016-7037(87)90371-1.<br />

Crovisier, J. L., J. Honnorez, B. Fritz, and J. C. Petit (1992),<br />

Dissolution <strong>of</strong> subglacial volcanic glasses from Iceland: Laboratory<br />

study and modeling, Appl. Geochem. Suppl., Issue,<br />

7, 55–81.<br />

Crovisier, J. L., T. Advocat, and J. L. Dussossoy (2003), Nature<br />

and role <strong>of</strong> natural alteration gels formed on the surface <strong>of</strong><br />

ancient volcanic glasses (natural analogs <strong>of</strong> waste containment<br />

glasses), J. Nucl. Mater., 321, 91–109, doi:10.1016/<br />

S0022-3115(03)00206-X.<br />

Daux, V., J. L. Crovisier, C. Hemond, and J. C. Petit (1994),<br />

Geochemical evolution <strong>of</strong> basaltic rocks subjected to weathering:<br />

Fate <strong>of</strong> the major elements, rare earth elements, and<br />

thorium, Geochim. Cosmochim. Acta, 58, 4941–4954,<br />

doi:10.1016/0016-7037(94)90223-2.<br />

<strong>Davis</strong>, A. S., and D. A. Clague (2003), Hyaloclastite from miocene<br />

seamounts <strong>of</strong>fshore central <strong>California</strong>: Compositions,<br />

eruption styles, and depositional processes, in Explosive Subaqueous<br />

Volcanism, Geophys. Monogr. Ser., vol. 140, edited<br />

by J. D. L. White, J. L. Smellie, and D. A. Clague, pp. 129–142,<br />

AGU, Washington D. C.<br />

<strong>Davis</strong>, A. S., and D. A. Clague (2006), Volcaniclastic deposits<br />

from the North Arch volcanic field, Hawaii: Explosive fragmentation<br />

<strong>of</strong> alkalic lava at abyssal depths, Bull. Volcanol.,<br />

68, 294–307.<br />

<strong>Davis</strong>, A. S., D. A. Clague, W. A. Bohrson, G. B. Dalrymple,<br />

and H. G. Greene (2002), Seamounts at the continental margin<br />

<strong>of</strong> <strong>California</strong>: A different kind <strong>of</strong> oceanic intraplate volcanism,<br />

Geol. Soc. Am. Bull., 114, 316–333, doi:10.1130/<br />

0016-7606(2002)1142.0.CO;2.<br />

Drief, A., and P. Schiffman (2004), Very low temperature<br />

alteration <strong>of</strong> sideromelane in hyaloclastites and hyalotuffs<br />

from Kilauea and Mauna Loa volcanoes: Implications for<br />

the mechanism <strong>of</strong> palagonite formation, Clays Clay Miner.,<br />

52, 623–635.<br />

Eggleton, R. A., and J. Keller (1982), The palagonitization <strong>of</strong><br />

limburgite glass: A TEM study, Neues. Jahrb. Mineral.<br />

Montash., 7, 321–336.<br />

Fisher, R. V., and H.‐U. Schmincke (1984), Alteration <strong>of</strong> volcanic<br />

glass, in Pyroclastic Rocks, pp. 312–345, Springer,<br />

New York.<br />

Furnes, H. (1984), Chemical changes during progressive subaerial<br />

palagonitization <strong>of</strong> a subglacial olivine tholeiite hyaloclastite;<br />

a microprobe study, Chem. Geol., 43, 271–285,<br />

doi:10.1016/0009-2541(84)90054-8.<br />

Furnes, H., H. Staudigel, I. H. Torseth, K. Muehlenbachs, and<br />

O. Tumyr (2001), Bioalteration <strong>of</strong> basaltic glass in the oceanic<br />

crust, Geochem. Geophys. Geosyst., 2(8), 1049, doi:10.1029/<br />

2000GC000150.<br />

Furnes,H.,I.H.Thorseth,T.Torsvik,K.Muehlenbachs,<br />

H. Staudigel, and O. Tumyr (2002), Identifying bio‐interaction<br />

with basaltic glass in oceanic crust and implications for estimating<br />

the depth <strong>of</strong> oceanic lithosphere: A review, in Volcano‐Ice<br />

Interaction on Earth and Mars, editedbyJ.L.Smellieand<br />

M. G. Chapman, Geol. Soc. Spec. Publ., 202, 407–421.<br />

Furnes, H., N. R. Banerjee, H. Saudigel, K. Muehlenbachs,<br />

N. McLoughlin, M. de Wit, and M. Van Kranendonk<br />

(2007), Comparing petrographic signatures <strong>of</strong> bioalteration<br />

in recent to Mesoarchean pillow lavas: Tracing subsurface life<br />

in oceanic igneous rocks, Precamb. Res., 158, 156–176.<br />

Garcia, M. O., E. H. Haskins, E. M. Stolper, and M. Baker (2007),<br />

Stratigraphy <strong>of</strong> the Hawai’i Scientific Drilling Project core<br />

(HSDP2), Anatomy <strong>of</strong> a Hawaiian shield volcano, Geochem.<br />

Geophys. Geosyst., 8, Q02G20, doi:10.1029/2006GC001379.<br />

Grant, J. A. (1986), The isocon diagram–A simple solution to<br />

Gresens’ equation for metasomatic alteration, Econ. Geol.,<br />

81, 1976–1982, doi:10.2113/gsecongeo.81.8.1976.<br />

Grzechnik, A., H. D. Zimmermann, P. L. King, P. F. McMillan,<br />

and R. L. Hervig (1996), FTIR micro‐reflectance measurements<br />

<strong>of</strong> the CO 3 2− ion content in basanite and leucite glasses,<br />

Contrib. Mineral. Petrol., 125, 311–318, doi:10.1007/<br />

s004100050224.<br />

Hay, R. L., and A. Iijima (1968a), Nature and origin <strong>of</strong> palagonite<br />

tuffs <strong>of</strong> the Honolulu Group, on Oahu, Hawaii, in Studies<br />

in Volcanology: A Memoir in Honor <strong>of</strong> Howell Williams,<br />

Mem. Geol. Soc. <strong>of</strong> Am., 116, 331–376.<br />

Hay, R. L., and A. Iijima (1968b), Petrology <strong>of</strong> palagonite tuffs<br />

<strong>of</strong> Koko Craters, Oahu, Hawaii, Contrib. Mineral. Petrol., 17,<br />

141–154.<br />

Heiken, G. H., R. V. Fisher, and N. V. Peterson (1981), A field<br />

trip to the maar volcanoes <strong>of</strong> the Fort Rock—Christmas Lake<br />

Valley basin, Oregon, in Guides to Some Volcanic Terranes in<br />

Washington, Idaho, Oregon, and Northern <strong>California</strong>, U.S.<br />

Geol. Surv. Circ., 838, 119–140.<br />

Hein, J. R., A. Koschinsky, P. E. Halbach, F. T. Manheim,<br />

M. Bau, J. K. Kang, and N. Lubick (1997), Iron and manganese<br />

oxide mineralization in the Pacific, in Manganese Mineralization:<br />

Geochemistry and Mineralogy <strong>of</strong> Terrestrial and<br />

Marine Deposits, edited by K. Nicholson et al., Geol. Soc.<br />

Spec. Publ., 119, 123–138.<br />

Honnorez, J. (1967), La palagonitisation, l’alteration sousmarine<br />

du verre volcanique basique de palagonia (Sicile),<br />

Ph.D. thesis, Univ. Libre de Bruxelles, Brussels.<br />

Honnorez, J. (1972), La Palagonitisation, l’Alteration Sous‐<br />

Marine du Verre Volcanique Basique de Palagonia (Sicile),<br />

131 pp., Birkhaüser, Stuttgart, Germany.<br />

Honnorez, J. (1978), Generation <strong>of</strong> phillipsites by palagonitization<br />

<strong>of</strong> basaltic glass in seawater and the origin <strong>of</strong> K‐rich<br />

deep‐sea deposits, in Natural Zeolites, Occurrence, Properties,<br />

Use, edited by L. B. Sand and F. A. Mumpton, pp. 245–258,<br />

Elsevier, New York.<br />

Jakobsson, S. P. (1972), On the consolidation and palagonitization<br />

<strong>of</strong> the tephra <strong>of</strong> the Surtsey volcanic island, Iceland,<br />

Surtsey Res. Program Rep., 6, 121–128.<br />

Jakobsson, S. P. (1978), Environmental factors controlling the<br />

palagonitization <strong>of</strong> the Surtsey tephra, Iceland, Bull. Geol.<br />

Soc. Den., 27, 91–105.<br />

Jakobsson, S. P., and M. T. Gudmundsson (2008), Subglacial<br />

and interglacial volcanic formations in Iceland, Joekull, 58,<br />

179–196.<br />

Jakobsson, S. P., and J. G. Moore (1986), Hydrothermal minerals<br />

and alteration rates at Surtsey volcano, Iceland, Geol. Soc.<br />

Am. Bull., 97, 648–659, doi:10.1130/0016-7606(1986)<br />

972.0.CO;2.<br />

Jakobsson, S. P., G. Gudmundsson, and J. G. Moore (2000), Geological<br />

monitoring <strong>of</strong> Surtsey, Icelend, 1967–1998, Surtsey<br />

Res. Program Rep., 11, 99–108.<br />

24 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Jensen, R. A. (1995), Roadside Guide to the Geology <strong>of</strong> Newberry<br />

Volcano, CenOreGeoPub, Bend, Oreg.<br />

Jercinovic,M.J.,K.Keil,M.R.Smith,andR.A.Schmitt<br />

(1990), Alteration <strong>of</strong> basltic glasses from north‐central<br />

British Columbia, Canada, Geochim. Cosmochim. Acta, 54,<br />

2679–2696, doi:10.1016/0016-7037(90)90004-5.<br />

Kimura, J.‐I., T. W. Sisson, N. Nakano, M. L. Coombs, and<br />

P. W. Lipman (2006), Isotope geochemistry <strong>of</strong> early Kilauea<br />

magmas from the submarine Hilina Bench: The nature <strong>of</strong> the<br />

Hilina mantle component, J. Volcanol. Geotherm. Res., 151,<br />

51–72, doi:10.1016/j.jvolgeores.2005.07.024.<br />

King, P. L., J. F. Larsen, D. K. Lui, J. Lockard, K. Dalby, and<br />

T. Griffin (2006), Rapid analysis <strong>of</strong> total H 2 O and molecular<br />

H 2 O contents <strong>of</strong> small geologic glasses using reflectance IR<br />

spectroscopy, Geol. Assoc. Can. Abstr. Program, 31, 77–78.<br />

Koschinsky, A., and J. Hein (2003), Uptake <strong>of</strong> elements from<br />

seawater by ferromanganese crusts: Solid‐phase associations<br />

and seawater speciation, Mar. Geol., 198, 331–351, doi:10.1016/<br />

S0025-3227(03)00122-1.<br />

Le Gal, X., J. L. Crovisier, F. Gauthier‐Lafaye, J. Honnorez,<br />

and B. Grambow (1999), Meteoric alteration <strong>of</strong> Icelandic<br />

volcanic glass: Long term changes in the mechanism, C. R.<br />

Acad. Sci., Ser. IIa Sci. Terre Planetes, 329, 175–181.<br />

Lipman, P. W., T. W. Sisson, U. Tadahide, J. Naka, and J. R.<br />

Smith (2002), Ancestral submarine growth <strong>of</strong> Kïlauea volcano<br />

and instability <strong>of</strong> its south flank, in Hawaiian Volcanoes:<br />

Deep Underwater Perspectives, Geophys. Monogr. Ser.,<br />

vol. 128, edited by E. Takahashi et al., pp. 161–191, AGU,<br />

Washington, D. C.<br />

Lipman, P., T. Sisson, M. Coombs, and J.‐I. Kimura (2006),<br />

Piggyback tectonics: Long‐term growth <strong>of</strong> Kilauea on the<br />

south flank <strong>of</strong> Mauna Loa, J. Volcanol. Geotherm. Res.,<br />

151, 73–108, doi:10.1016/j.jvolgeores.2005.07.032.<br />

Mattsson, H., and A. Höskuldsson (2003), Geology <strong>of</strong> the<br />

Heimaey volcanic centre, south Iceland: Early evolution <strong>of</strong> a<br />

central volcano in a propagating rift?, J. Volcanol. Geotherm.<br />

Res., 127, 55–71, doi:10.1016/S0377-0273(03)00178-1.<br />

McDonough, W. F., and S.‐S. Sun (1995), The composition <strong>of</strong><br />

the Earth, Chem. Geol., 120, 223–253, doi:10.1016/0009-<br />

2541(94)00140-4.<br />

Moore, G., A. Chizmeshya, and P. F. McMillan (2000), Calibration<br />

<strong>of</strong> a reflectance FTIR method for determination <strong>of</strong><br />

dissolved CO 2 concentration in rhyolitic glasses, Geochim.<br />

Cosmochim. Acta, 64(20), 3571–3579, doi:10.1016/S0016-<br />

7037(00)00447-6.<br />

Moore, J. G. (1966), Rate <strong>of</strong> palagonitization <strong>of</strong> submarine<br />

basalt adjacent to Hawaii, U.S. Geol. Surv. Pr<strong>of</strong>. Pap.,<br />

550D, 163–171.<br />

Moore, J. G., D. J. Fornari, and D. Clague (1985), Basalts<br />

from the 1887 submarine eruption <strong>of</strong> Mauna Loa, Hawaii:<br />

New data on the variation <strong>of</strong> palagonitization rate with temperature,<br />

U.S. Geol. Surv. Bull., 1163, 11 pp.<br />

Novak, E. A., and S. R. Evans (1991), Preliminary results from<br />

two scientific observation holes on the Kilauea East Rift<br />

Zone, Trans. Geotherm. Resour. Counc., 15, 187–192.<br />

Oviatt, C. G., and W. P. Nash (1989), Late Pleistocene basaltic<br />

ash and volcanic eruptions in the Bonneville basin, Utah,<br />

Geol. Soc. Am. Bull., 101, 292–303, doi:10.1130/0016-<br />

7606(1989)1012.3.CO;2.<br />

Pauly, B. D. (2011), In situ micro‐analytical investigations <strong>of</strong><br />

palagonitization, Ph.D. dissertation, 1779 pp., Univ. <strong>of</strong><br />

Calif., <strong>Davis</strong>, Calif.<br />

Saemundsson, K., H. Jóhannesson, Á. Hjartarson, S. Gardar<br />

Kristinsson, and M. Á. Sigurgeirsson (2010), Geological Map<br />

<strong>of</strong> Southwest Iceland, 1:100000, Icel. GeoSurvey, Reykjavík.<br />

Schiffman, P., and S. Roeske (2002), Electron microprobe<br />

analysis <strong>of</strong> minerals, in Encyclopedia <strong>of</strong> Physical Sciences<br />

and Technology, vol. 5, pp. 293–306, Academic, San Diego,<br />

Calif.<br />

Schiffman, P., A. W. Walton, N. Thompson, and R. J. Watters<br />

(2006), Hyaloclastite alteration and slope stability <strong>of</strong> Hawaiian<br />

volcanoes: Insights from the Hawaiian Scientific Driling<br />

Project’s 3‐km drill core, J. Volcanol. Geotherm. Res., 151,<br />

217–228, doi:10.1016/j.jvolgeores.2005.07.030.<br />

Sisson, T. W., P. W. Lipman, and J. Naka (2002), Submarine<br />

alkalic through tholeiitic shield‐stage development <strong>of</strong><br />

Kīlauea volcano, Hawai’i, in Hawaiian Volcanoes: Deep<br />

Underwater Perspectives, Geophys. Monogr. Ser., vol.128,<br />

edited by E. Takahashi et al., pp. 193–219, AGU, Washington,<br />

D. C.<br />

Staudigel, H., and H. Furnes (2004), Microbial mediation <strong>of</strong><br />

oceanic crust alteration, in Hydrology <strong>of</strong> the Oceanic Lithosphere,<br />

edited by E. <strong>Davis</strong> and H. Elderfield, pp. 608–626,<br />

Cambridge Univ. Press, Cambridge, U. K.<br />

Staudigel, H., and S. R. Hart (1983), Alteration <strong>of</strong> basaltic<br />

glass: Mechanisms and significance for the oceanic crustseawater<br />

budget, Geochim. Cosmochim. Acta, 47, 337–350,<br />

doi:10.1016/0016-7037(83)90257-0.<br />

Staudigel, H., H. Furnes, K. Kelley, Y. Plank, K. Muehlenbachs,<br />

B. Tebo, and A. Yayanos (2004), The oceanic crust as a bioreactor,<br />

in Deep Subsurface Biosphere at Mid‐Ocean Ridges,<br />

Geophys. Monogr. Ser., vol.144,editedbyW.S.Wilcock<br />

et al., pp. 325–341, AGU, Washington, D. C.<br />

Staudigel, H., H. Furnes, N. R. Banerjee, Y. Dilek, and<br />

K. Muehlenbachs (2006), Microbes and volcanoes: A tale<br />

from the oceans, ophiolites and greenstone belts, GSA Today,<br />

16, 4–10, doi:10.1130/GSAT01610A.1.<br />

Staudigel, H., H. Furnes, N. McLoughlin, N. R. Banerjee,<br />

L. B. Connell, and A. Templeton (2008), 3.5 billion years<br />

<strong>of</strong> glass bioalteration: Volcanic rocks as a basis for microbial<br />

life?, Earth Sci. Rev., 89, 156–176, doi:10.1016/j.earscirev.<br />

2008.04.005.<br />

Stroncik, N. A., and H.‐U. Schmincke (2001), Evolution <strong>of</strong><br />

palagonite: Crystallization, chemical changes, and element<br />

budget, Geochem. Geophys. Geosyst., 2(7), 1017,<br />

doi:10.1029/2000GC000102.<br />

Thorarinsson, S., T. Einarsson, G. E. Sigvaldson, and G. Elisson<br />

(1964), The submarine eruption <strong>of</strong>f the Westman Islands,<br />

1963–64, Bull. Volcanol., 27, 435–445, doi:10.1007/<br />

BF02597544.<br />

Thorseth, I. H., H. Furnes, and O. Tumyr (1991), A textural<br />

and chemical study <strong>of</strong> Icelandic palagonite <strong>of</strong> varied composition<br />

and its bearing on the mechanism <strong>of</strong> the glass‐palagonite<br />

transformation, Geochim. Cosmochim. Acta, 55, 731–749,<br />

doi:10.1016/0016-7037(91)90337-5.<br />

Thorseth, I. H., H. Furnes, and M. Heldal (1992), The importance<br />

<strong>of</strong> microbiological activity in the alteration <strong>of</strong> natural<br />

basaltic glass, Geochim. Cosmochim. Acta, 56, 845–850,<br />

doi:10.1016/0016-7037(92)90104-Q.<br />

Trusdell, F. A., E. Novak, S. R. Evans, and K. Okano (1999),<br />

Core lithology from the State <strong>of</strong> Hawaii Scientific Observation<br />

Hole 1, Kilauea Volcano, Hawai’i, U.S. Geol. Surv.<br />

Open File Rep., 99‐389.<br />

Valle, N., A. Verney‐Carron, J. Sterenich, G. Libourel,<br />

E. Deloule, and P. Jollivet (2010), Elemental and isotopic<br />

( 29 Si and 18 O) tracing <strong>of</strong> glass alteration mechanisms, Geochim.<br />

Cosmochim. Acta, 74, 3412–3431, doi:10.1016/j.gca.2010.<br />

03.028.<br />

van Achterbergh, E., C. Ryan, S. Jackson, and W. Griffin<br />

(2001), Data reduction s<strong>of</strong>tware for LA‐ICP‐MS, in Laser<br />

25 <strong>of</strong> 26


Geochemistry<br />

Geophysics<br />

Geosystems G 3 PAULY ET AL.: GEOCHEMICAL CONTROLS ON PALAGONITIZATION 10.1029/2011GC003639<br />

Ablation ICPMS in the Earth Sciences: Principles and Applications,editedbyPaulSylvester,Short<br />

Course Handb. Mineral.<br />

Assoc. Can., 29, 239–243.<br />

Walton, A. W. (2008), Microtubules in basalt glass from<br />

Hawaii Scientific Drilling Project #2 phase 1 core and Hilina<br />

slope, Hawaii: Evidence <strong>of</strong> the occurrence and behavior <strong>of</strong><br />

endolithic microorganisms, Geobiology, 6, 351–364,<br />

doi:10.1111/j.1472-4669.2008.00149.x.<br />

Walton, A. W., and P. Schiffman (2003), Alteration <strong>of</strong> hyaloclastites<br />

in the HSDP 2 Phase 1 Drill Core 1. Description and<br />

paragenesis, Geochem. Geophys. Geosyst., 4(5), 8709,<br />

doi:10.1029/2002GC000368.<br />

Walton, A. W., P. Schiffman, and G. L. Macpherson (2005),<br />

Alteration <strong>of</strong> hyaloclastites in the HSDP 2 Phase 1 Drill<br />

Core: 2. Mass balance <strong>of</strong> the conversion <strong>of</strong> sideromelane<br />

to palagonite and chabazite, Geochem. Geophys. Geosyst.,<br />

6, Q09G19, doi:10.1029/2004GC000903.<br />

White, J. D. L. (1996), Pre‐emergent construction <strong>of</strong> a lacustrine<br />

basaltic volcano, Pahvant Butte, Utah (USA), Bull. Volcanol.,<br />

58, 249–262, doi:10.1007/s004450050138.<br />

White, J. D. L. (2001), Eruption and reshaping <strong>of</strong> Pahvant<br />

Butte volcano in Pleistocene Lake Bonneville, Spec. Publ.<br />

Int. Assoc. Sedimentol., 30, 61–80.<br />

Winchell, H. (1947), Honolulu series, Oahu, Hawaii, Geol. Soc.<br />

Am. Bull., 58, 1–48, doi:10.1130/0016-7606(1947)58[1:<br />

HSOH]2.0.CO;2.<br />

Zhou, Z., and W. S. Fyfe (1989), Palagonitization <strong>of</strong> basaltic glass<br />

from DSDP Site 335, Leg 37: Textures, chemical‐composition,<br />

and mechanism <strong>of</strong> formation, Am. Mineral., 74, 1045–1053.<br />

26 <strong>of</strong> 26

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!