23.07.2013 Views

01 branney et al 2008 bv.pdf

01 branney et al 2008 bv.pdf

01 branney et al 2008 bv.pdf

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

DOI 10.1007/s00445-007-<strong>01</strong>40-7<br />

RESEARCH ARTICLE<br />

‘Snake River (SR)-type’ volcanism at the Yellowstone<br />

hotspot track: distinctive products from unusu<strong>al</strong>,<br />

high-temperature silicic super-eruptions<br />

M. J. Branney & B. Bonnichsen & G. D. M. Andrews &<br />

B. Ellis & T. L. Barry & M. McCurry<br />

Received: 26 April 2005 /Accepted: 8 March 2007 / Published online: 20 June 2007<br />

# Springer-Verlag 2007<br />

Abstract A new category of large-sc<strong>al</strong>e volcanism, here<br />

termed Snake River (SR)-type volcanism, is defined with<br />

reference to a distinctive volcanic facies association<br />

displayed by Miocene rocks in the centr<strong>al</strong> Snake River<br />

Plain area of southern Idaho and northern Nevada, USA.<br />

The facies association contrasts with those typic<strong>al</strong> of silicic<br />

volcanism elsewhere and records unusu<strong>al</strong>, voluminous and<br />

particularly environment<strong>al</strong>ly devastating styles of eruption<br />

This paper constitutes part of a speci<strong>al</strong> issue dedicated to Bill<br />

Bonnichsen on the p<strong>et</strong>rogenesis and volcanology of anorogenic<br />

rhyolites.<br />

Editori<strong>al</strong> responsibility: W Leeman<br />

Electronic supplementary materi<strong>al</strong> The online version of this article<br />

(doi:10.1007/s00445-007-<strong>01</strong>40-7) contains supplementary materi<strong>al</strong>,<br />

which is available to authorized users.<br />

M. J. Branney (*) : G. D. M. Andrews : B. Ellis<br />

Department of Geology, University of Leicester,<br />

University Road,<br />

Leicester LE1 7RH England, UK<br />

e-mail: mjb26@le.ac.uk<br />

B. Bonnichsen<br />

Idaho Geologic<strong>al</strong> Survey, University of Idaho,<br />

Moscow, ID 83844-3<strong>01</strong>4, USA<br />

T. L. Barry<br />

Volcano Dynamics Group, Open University,<br />

Milton Keynes,<br />

MK7 6AA England, UK<br />

M. McCurry<br />

Department of Geosciences, Idaho State University,<br />

Pocatello, ID 83209-8072, USA<br />

G. D. M. Andrews<br />

Volcanology Lab. & MDRU, EOS,<br />

University of British Columbia,<br />

Vancouver, BC, Canada<br />

that remain poorly understood. It includes: (1) largevolume,<br />

lithic-poor rhyolitic ignimbrites with scarce pumice<br />

lapilli; (2) extensive, par<strong>al</strong>lel-laminated, medium to coarsegrained<br />

ashf<strong>al</strong>l deposits with large cuspate shards, cryst<strong>al</strong>s<br />

and a paucity of pumice lapilli; many are fused to black<br />

vitrophyre; (3) unusu<strong>al</strong>ly extensive, large-volume rhyolite<br />

lavas; (4) unusu<strong>al</strong>ly intense welding, rheomorphism, and<br />

widespread development of lava-like facies in the ignimbrites;<br />

(5) extensive, fines-rich ash deposits with abundant<br />

ash aggregates (pell<strong>et</strong>s and accr<strong>et</strong>ionary lapilli); (6) the<br />

ashf<strong>al</strong>l layers and ignimbrites contain abundant clasts of<br />

dense obsidian and vitrophyre; (7) a bimod<strong>al</strong> association<br />

b<strong>et</strong>ween the rhyolitic rocks and numerous, co<strong>al</strong>escing lowprofile<br />

bas<strong>al</strong>t lava shields; and (8) widespread evidence of<br />

emplacement in lacustrine-<strong>al</strong>luvi<strong>al</strong> environments, as<br />

reve<strong>al</strong>ed by interc<strong>al</strong>ated lake sediments, ignimbrite peperites,<br />

rhyolitic and bas<strong>al</strong>tic hy<strong>al</strong>oclastites, bas<strong>al</strong>t pillow-lava<br />

deltas, rhyolitic and bas<strong>al</strong>tic phreatomagmatic tuffs, <strong>al</strong>luvi<strong>al</strong><br />

sands and p<strong>al</strong>aeosols. Many rhyolitic eruptions were high<br />

mass-flux, large volume and explosive (VEI 6–8), and<br />

involved H2O-poor, low-δ 18 O, m<strong>et</strong><strong>al</strong>uminous rhyolite magmas<br />

with unusu<strong>al</strong>ly low viscosities, partly due to high<br />

magmatic temperatures (900–1,050°C). SR-type volcanism<br />

contrasts with silicic volcanism at many other volcanic<br />

fields, where the f<strong>al</strong>l deposits are typic<strong>al</strong>ly Plinian with<br />

pumice lapilli, the ignimbrites are low to medium grade<br />

(non-welded to eutaxitic) with abundant pumice lapilli or<br />

fiamme, and the rhyolite extrusions are sm<strong>al</strong>l volume silicic<br />

domes and coulées. SR-type volcanism seems to have<br />

occurred at numerous times in Earth history, because<br />

elements of the facies association occur within some other<br />

volcanic fields, including Trans-Pecos Texas, Etendeka-<br />

Paraná, Lebombo, the English Lake District, the Proterozoic<br />

Keewanawan volcanics of Minnesota and the Yardea<br />

Dacite of Austr<strong>al</strong>ia.


294 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

Keywords Snake River . Yellowstone . Intraplate .<br />

Hot-spot . Ignimbrite . Welded tuff . Rheomorphic .<br />

Super-eruption<br />

Introduction<br />

In this paper we define a new category of large-sc<strong>al</strong>e silicic<br />

volcanism that is distinct in character to the more wellknown<br />

pumice-rich, Plinian/ignimbrite type of silicic<br />

volcanism that is described extensively in the literature<br />

(e.g. Cas and Wright 1987). The distinctive category of<br />

volcanism has occurred at sever<strong>al</strong> times in Earth history, but<br />

is under-represented in the literature, possibly because the<br />

unusu<strong>al</strong>, and particularly environment<strong>al</strong>ly devastating,<br />

eruption styles are not well understood. To highlight its<br />

character, we describe the best-preserved example known to<br />

us, which serves as a defining ‘type’ example, and where<br />

further study would be richly rewarded. We then contrast<br />

SR-type volcanism with silicic volcanism elsewhere,<br />

discuss the enigmatic eruption styles and, fin<strong>al</strong>ly, list other<br />

volcanic provinces where SR-type volcanism may be<br />

represented.<br />

Large-volume silicic eruptions involving 10s to<br />

1,000s km 3 of magma are the most catastrophic form of<br />

volcanism and can cause abrupt region<strong>al</strong> obliteration and<br />

climatic perturbation (Sparks <strong>et</strong> <strong>al</strong>. 2005). Typic<strong>al</strong>ly, they<br />

Fig. 1 The bimod<strong>al</strong> Columbia<br />

River–Yellowstone volcanic<br />

province of northwest USA<br />

showing the type area of SR-type<br />

volcanism as defined in this paper<br />

(rectangle) and the location of<br />

Miocene–Pleistocene Lake Idaho<br />

(vertic<strong>al</strong> hachure; SE limits y<strong>et</strong> to<br />

be defined). Northern Nevada<br />

rhyolites after Pierce <strong>et</strong> <strong>al</strong>. (2002).<br />

Rhyolitic c<strong>al</strong>deras: Y, Yellowstone<br />

c<strong>al</strong>dera field; N, Newberry; B,<br />

Burns; M, McDermitt. Approximate<br />

outlines of inferred, largely<br />

conce<strong>al</strong>ed eruptive centres: OH,<br />

Owyhee-Humboldt; BJ, Bruneau-<br />

Jarbidge; TF, Twin F<strong>al</strong>ls; P, Picabo;<br />

H, Heise, after Bonnichsen<br />

<strong>et</strong> <strong>al</strong>. (1989) andMorganand<br />

McIntosh (2005); ages given in<br />

Ma. Locations of outflow ignimbrite–ashf<strong>al</strong>l<br />

successions: R, Rogerson<br />

Graben; C, Cassia Hills;<br />

BH, Benn<strong>et</strong>t Hills; T, Trapper<br />

Creek–Goose Creek Basin. Dark<br />

grey: extent of Columbia River–<br />

Oregon plateau bas<strong>al</strong>ts (including<br />

Steens Mountain, M<strong>al</strong>heur Gorge<br />

and Picture Gorge bas<strong>al</strong>ts). MR,<br />

Magic Reservoir centre<br />

involve (1) Plinian explosive eruptions that produce f<strong>al</strong>lout<br />

layers of pumice lapilli with subordinate lithic clasts,<br />

accompanied by (2) emplacement of voluminous pumice<br />

lapilli-bearing ignimbrites, followed by (3) extrusion of<br />

relatively sm<strong>al</strong>l-volume domes and coulées of degassed<br />

silicic lava. However, the silicic eruption products in some<br />

volcanic fields differ significantly from this gener<strong>al</strong> pattern.<br />

The youngest, and best-preserved, example of this is found<br />

in the centr<strong>al</strong> part of the Snake River Plain in southern<br />

centr<strong>al</strong> Idaho and northernmost Nevada, north-west USA.<br />

There, Miocene rhyolitic volcanism was large-sc<strong>al</strong>e and<br />

catastrophic but the assemblage of volcanic deposits is<br />

markedly different to the familiar pumice-rich assemblage<br />

convention<strong>al</strong>ly associated with large Plinian and ignimbrite<br />

eruptions. Unusu<strong>al</strong> characteristics include a paucity of<br />

pumice lapilli, a predominance of particularly intenselywelded<br />

rheomorphic ignimbrites, layers of par<strong>al</strong>lel-stratified<br />

coarse ash in which large cuspate shards are commonly<br />

clearly visible in the field, and the presence of exception<strong>al</strong>ly<br />

large-volume, low-aspect ratio rhyolite lavas. We recognise<br />

that this association of facies represents a distinctive<br />

category of volcanism, and propose the name ‘Snake<br />

River-type’ (SR-type) volcanism, after the c.13 Ma–<br />

c.8 Ma succession in the centr<strong>al</strong> part of the Snake River<br />

Plain (Fig. 1), where it is best represented. Elements of the<br />

same facies association occur in other parts of the world<br />

and suggest that SR-type volcanism has punctuated Earth


Bull Volcanol (<strong>2008</strong>) 70:293–314 295<br />

history; during such events, the eruption of large volumes<br />

of silicic ash at extremely high temperatures and at high<br />

eruptive mass flux rates is likely to have had a profound<br />

environment<strong>al</strong> impact.<br />

Region<strong>al</strong> s<strong>et</strong>ting<br />

Miocene silicic rocks of south centr<strong>al</strong> Idaho and northernmost<br />

Nevada (Fig. 1) form part of the bimod<strong>al</strong> (bas<strong>al</strong>t–<br />

rhyolite) Columbia River–Yellowstone volcanic province of<br />

northwest USA (Fig. 1). The volcanic activity gener<strong>al</strong>ly<br />

youngs north-eastwards towards the most recent activity at<br />

Yellowstone (Fig. 1) and is thought to relate to the passage<br />

of the North American continent over a stationary hot-spot<br />

(e.g. refs in Pierce <strong>et</strong> <strong>al</strong>. 2002). An over<strong>al</strong>l northeastyounging<br />

of silicic eruptive centres in southern Idaho is<br />

thought to reflect diachronous injection of bas<strong>al</strong>tic magma<br />

into the lithosphere, causing the generation of lithospheric<br />

melts (e.g. Leeman 1982a) and/or the parti<strong>al</strong> re-melting of<br />

earlier injected materi<strong>al</strong> (refs in Christiansen and McCurry<br />

2007). The chemistry of the rhyolitic rocks across the<br />

province broadly reflects the nature of the lithosphere from<br />

which the magma derives, i.e. per<strong>al</strong>k<strong>al</strong>ine to m<strong>et</strong><strong>al</strong>uminous<br />

older rhyolites in the west where the lithosphere comprises<br />

accr<strong>et</strong>ed oceanic terrane materi<strong>al</strong>, to younger m<strong>et</strong><strong>al</strong>uminous<br />

rhyolites of the centr<strong>al</strong> and eastern parts of the Snake River<br />

Plain, where lithosphere is part of the North American<br />

craton (e.g. Wright <strong>et</strong> <strong>al</strong>. 2002). This transition in<br />

lithospheric affinity is reflected in the Sr, Nd, Pb and O<br />

isotopes (e.g. Benn<strong>et</strong>t and DePaolo 1987; Wooden and<br />

Mueller 1988; Leeman <strong>et</strong> <strong>al</strong>. 1992). The voluminous centr<strong>al</strong><br />

Snake River Plain rhyolitic rocks that are the subject of this<br />

paper formed during a major region<strong>al</strong> ignimbrite ‘flare-up’<br />

that her<strong>al</strong>ded continent<strong>al</strong> rifting, with formation of the western<br />

Snake River rift (Fig. 1). They are typic<strong>al</strong> of intracontinent<strong>al</strong><br />

A-type granitic melts; m<strong>et</strong><strong>al</strong>uminous and relatively anhydrous,<br />

with elevated HFSE concentrations and high h<strong>al</strong>ogen<br />

concentrations (e.g. Christiansen and McCurry 2007).<br />

The Columbia River–Yellowstone volcanic province<br />

records diverse eruptive styles, ranging from the effusion<br />

of extensive bas<strong>al</strong>tic lava fields, low-profile bas<strong>al</strong>tic shields,<br />

to silicic extrusion and magmatic and phreatomagmatic<br />

explosivity. Features characteristic of SR-type volcanism<br />

occur in sever<strong>al</strong> parts of the province, but are best<br />

developed in the Miocene succession in the type area<br />

(depicted in Fig. 1). This location is the origin<strong>al</strong> type area<br />

for the ‘Idavada Volcanic Group’ of M<strong>al</strong>de and Powers<br />

(1962) and it is where the term ‘super-eruptions’ was first<br />

coined (Bonnichsen 2000 BBC Horizon); super-eruptions<br />

are catastrophic environment<strong>al</strong>ly devastating eruptions<br />

>300 km 3 (Sparks <strong>et</strong> <strong>al</strong>. 2005). The rhyolitic (70–76%<br />

SiO2) eruptions are inferred to have occurred from broad<br />

eruptive centres (e.g. Bruneau-Jarbidge and Twin F<strong>al</strong>ls<br />

centres; Fig. 1), which are largely conce<strong>al</strong>ed beneath<br />

Pliocene–Pleistocene bas<strong>al</strong>t lavas, so much of our understanding<br />

derives from well exposed and deeply dissected<br />

outflow successions in massifs to the north and south of the<br />

Snake River Plain, such as around Jarbidge, the Cassia<br />

Mountains and the Benn<strong>et</strong>t Hills (Fig. 1; e.g. Bonnichsen<br />

and Citron 1982; Honjo <strong>et</strong> <strong>al</strong>. 1992; McCurry <strong>et</strong> <strong>al</strong>. 1996;<br />

Andrews <strong>et</strong> <strong>al</strong>. 2007; Bonnichsen <strong>et</strong> <strong>al</strong>. 2007). The<br />

stratigraphy and p<strong>et</strong>rology of the individu<strong>al</strong> volcanic<br />

successions are described elsewhere (previous refs), and<br />

the present paper presents a synthesis of the physic<strong>al</strong><br />

volcanology of the area. Field locations are given as grid<br />

references (prefixed GR) using 11 T zone UTM coordinates.<br />

Pliocene silicic volcanic rocks, 2–5 m.y. younger than<br />

those described in this paper, at the Magic Reservoir<br />

eruptive centre (MR on Fig. 1) to the north of the hotspot<br />

track in the centr<strong>al</strong> Snake River Plain are not entirely SRtype<br />

in that they exhibit features such as rhyolite domes and<br />

Plinian pumice f<strong>al</strong>lout layers (Honjo and Leeman 1987;<br />

Leeman 1982b; Leeman 2004).<br />

Snake River-type eruptive activity has never been<br />

observed and so must be inferred from the deposit record.<br />

Therefore, we will describe the volcanic facies that make up<br />

the distinctive SR-type volcanism facies-association. Some<br />

of the features are unusu<strong>al</strong> or have unusu<strong>al</strong> fac<strong>et</strong>s, whilst<br />

others are fairly common in silicic volcanism elsewhere.<br />

The distinctiveness of SR-type volcanism lies in the relative<br />

proportions of the facies; for example, in SR-type ignimbrites<br />

lava-like facies predominate, y<strong>et</strong> in many other<br />

volcanic fields lava-like facies are, typic<strong>al</strong>ly subordinate,<br />

if present at <strong>al</strong>l.<br />

SR-type ignimbrites<br />

We estimate that there are ≥40 extensive welded rhyolitic<br />

ignimbrite she<strong>et</strong>s in the centr<strong>al</strong> Snake River Plain area.<br />

Ignimbrite successions dominate the SR-type volcanism<br />

landscape and give rise to widespread trap-topography and<br />

terracing of multiple cliffs <strong>al</strong>ong the w<strong>al</strong>ls of river canyons<br />

and escarpments (Appendix Fig. 1). Individu<strong>al</strong> escarpments<br />

expose as many as ten ignimbrites, and bases of many of these<br />

successions are not exposed (e.g. Bonnichsen and Citron<br />

1982; Honjo <strong>et</strong> <strong>al</strong>. 1992; Perkins <strong>et</strong> <strong>al</strong>. 1995; McCurry <strong>et</strong> <strong>al</strong>.<br />

1996; Cathey and Nash 2004; Andrews <strong>et</strong> <strong>al</strong>. 2007).<br />

Size and morphology<br />

Ignimbrite outflow she<strong>et</strong>s are up to 200 m thick and include<br />

both simple and compound cooling units (e.g. the Grey’s<br />

Landing ignimbrite and Cougar Point Tuff VII, respective-


296 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

ly). Sever<strong>al</strong>, but not <strong>al</strong>l, of the ignimbrites have low aspectratios<br />

(Fig. 2): they trace later<strong>al</strong>ly for tens of kilom<strong>et</strong>res,<br />

thickening loc<strong>al</strong>ly into basins (e.g. Rogerson Graben) and<br />

tapering to less than 3 m thick across p<strong>al</strong>aeotopographic<br />

highs (Fig. 3). This makes estimation of eruption volumes<br />

difficult: in addition, mapping of individu<strong>al</strong> outflow she<strong>et</strong>s<br />

remains poorly advanced, and thicknesses and extents of<br />

proxim<strong>al</strong> ignimbrite within eruptive centres in the axis of<br />

the Snake River Plain are obscured by younger bas<strong>al</strong>t lavas.<br />

Current understanding is that the ignimbrite eruptions were<br />

similar in sc<strong>al</strong>e to those elsewhere in the volcanic province,<br />

with eruption volumes spanning three orders of magnitude,<br />

from tens to thousands of km 3 (VEI 6–8). For example, the<br />

Sand Springs ignimbrite (Andrews <strong>et</strong> <strong>al</strong>. 2007) has an<br />

estimated volume of 10 km 3 , whereas the ignimbrites of<br />

Jackpot, Big Bluff and Cougar Point Tuff XIII may exceed<br />

1,000 km 3 (Bonnichsen and Citron 1982; Andrews <strong>et</strong> <strong>al</strong>.<br />

2007).<br />

Pyroclast populations<br />

Snake River-type ignimbrites are commonly true tuffs (e.g.<br />

the Backwaters Member; Andrews <strong>et</strong> <strong>al</strong>. 2007). This is in<br />

contrast to ignimbrites elsewhere, which are commonly<br />

-3<br />

10<br />

thickness of unit t (km)<br />

0<br />

10<br />

-1<br />

10<br />

-2<br />

10<br />

-3<br />

10<br />

eruption<br />

volume<br />

10 km<br />

10 km<br />

-3 3<br />

10 km<br />

-2 3<br />

-4 3<br />

-1<br />

10<br />

-2<br />

10<br />

10 km<br />

-1 3<br />

Intermediate lavas<br />

n = 239<br />

-1<br />

10<br />

Bas<strong>al</strong>t lavas<br />

n = 479<br />

aspect<br />

ratio<br />

0<br />

t/d = 1:10<br />

Rhyolite lavas<br />

n = 176<br />

0<br />

10<br />

10 km<br />

1 3<br />

2<br />

area covered by unit A (km )<br />

Fig. 2 Dimensions and aspect-ratios of SR-type ignimbrites and SRtype<br />

rhyolite lavas compared with those from other volcanic fields<br />

(data for non-SR-type volcanics adapted from W<strong>al</strong>ker 1983; Cas and<br />

Wright 1987). SR-type ignimbrites (black spots) span the range from<br />

high to low aspect-ratio type. The field of SR-type rhyolite lavas<br />

(dashed line) is quite distinct from that of non-SR-type rhyolite lavas:<br />

SR-type lavas have similar aspect ratios to many bas<strong>al</strong>t lavas (not<br />

extensive flood bas<strong>al</strong>ts, which are not indicated), and they occupy a<br />

10 km<br />

0<br />

10<br />

0 3<br />

1<br />

10<br />

10 km<br />

composed mostly of massive lapilli-tuff (or lapilli-ash).<br />

Pumice lapilli are mostly absent (Fig. 5), as are fiamme<br />

formed from welded pumice lapilli. A few ignimbrites<br />

contain pumice lapilli loc<strong>al</strong>ly. Sieving of rare, non-welded<br />

parts of massive ignimbrites reve<strong>al</strong>s grainsizes of φ>2, and<br />

sorting v<strong>al</strong>ues of σφ


Bull Volcanol (<strong>2008</strong>) 70:293–314 297<br />

Fig. 3 Characteristic features of<br />

SR-type rhyolite lavas (top) and<br />

SR-type ignimbrites (bottom)<br />

contrasted with those typic<strong>al</strong> of<br />

other volcanic fields. SR-type<br />

lavas (note vertic<strong>al</strong> exaggeration)<br />

are much more voluminous,<br />

more extensive, and have<br />

lower aspect ratios than non-SRtype<br />

silicic lavas. SR-type<br />

ignimbrites are extremely highgrade,<br />

commonly lack pumice<br />

lapilli (or fiamme), with intense<br />

welding, rheomorphism and development<br />

of lava-like facies.<br />

They are distinguishable from<br />

lavas by their tapering margins,<br />

extensive zones containing pervasive<br />

vitroclastic textures, and<br />

the absence of extensive bas<strong>al</strong><br />

autobreccia and blunt, lobate<br />

terminations<br />

SR-type rhyolite lava<br />

~ 300 m<br />

silicic dome<br />

100 m<br />

feeder<br />

dyke<br />

large gas cavities<br />

autobreccia<br />

juvenile, others may be derived from underlying welded<br />

ignimbrites or lavas.<br />

Most SR-type ignimbrites are massive, and some show<br />

some vertic<strong>al</strong> zoning with respect to phenocryst abundance<br />

and/or composition (Wright <strong>et</strong> <strong>al</strong>. 2002; Andrews <strong>et</strong> <strong>al</strong>.<br />

2 km<br />

proxim<strong>al</strong><br />

tephra<br />

columnar jointing<br />

bas<strong>al</strong><br />

vitrophyre<br />

30 km<br />

flow interior flow margin<br />

lithoid<strong>al</strong><br />

rhyolite<br />

silicic coulee<br />

upper<br />

vitrophyre<br />

100 m<br />

autobreccia<br />

‘Typic<strong>al</strong>’ non-SR-type rhyolite lava (e.g. It<strong>al</strong>y, Jemez Mtns, La Primavera)<br />

SR-type ignimbrite<br />

~ 100 m<br />

loc<strong>al</strong><br />

peperites<br />

autobreccia<br />

columnar<br />

jointing<br />

medium-sc<strong>al</strong>e<br />

rheomorphic folds<br />

bas<strong>al</strong><br />

vitrophyre<br />

100 km<br />

lithoid<strong>al</strong><br />

rhyolite<br />

‘Typic<strong>al</strong>’ non-SR-type ignimbrite (e.g. It<strong>al</strong>y, Jemez Mtns, Pinatubo)<br />

~ 100 m<br />

eutaxitic lapilli-tuff<br />

10 km<br />

pumice-rich lapilli-tuff<br />

she<strong>et</strong><br />

joints flow<br />

lobe<br />

autobreccia<br />

5 km<br />

upper<br />

vitrophyre<br />

sub-horizont<strong>al</strong><br />

welding and sm<strong>al</strong>l-sc<strong>al</strong>e<br />

rheomorphic fabric<br />

splay-and-fade<br />

stratification<br />

~ 30 m<br />

feather-edge vitrophyre<br />

2007). These characteristics indicate deposition from<br />

sustained density currents whose lower parts were a<br />

granular fluid (‘granular fluid-based pyroclastic density<br />

currents’ of Branney and Kokelaar 2002). Some contain<br />

moulds of trees (e.g. Greys Landing Ignimbrite at Monu-<br />

5 m<br />

5 m


298 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

Fig. 4 Grain size and sorting<br />

characteristics of SR-type<br />

ignimbrites, compared to those<br />

from other volcanic provinces.<br />

Shaded fields from W<strong>al</strong>ker<br />

(1971); percentages cited are<br />

from 300 ignimbrites; sorting<br />

is described in terms of a<br />

standard deviation,<br />

σφ ¼ ðφ84 φ16Þ=2. SR-type ignimbrites are<br />

gener<strong>al</strong>ly b<strong>et</strong>ter sorted than<br />

ignimbrites from other volcanic<br />

provinces<br />

(less well<br />

sorted)<br />

(no grainsize<br />

variation)<br />

ment Canyon, Twin F<strong>al</strong>ls County). A few non-welded<br />

zones <strong>al</strong>so include horizons, traceable for sever<strong>al</strong> kilom<strong>et</strong>res,<br />

of well-developed traction<strong>al</strong> cross-stratification,<br />

indicative of deposition from currents in which clast<br />

Fig. 5 Non-welded facies of SR-type ignimbrites, showing the<br />

absence of pumice lapilli and lithic lapilli, and the presence of<br />

abundant sm<strong>al</strong>l vitric chips supported in a massive, fine ash matrix. a<br />

Typic<strong>al</strong> SR-type non-welded massive ignimbrite with sm<strong>al</strong>l vitric<br />

grains; Sand Springs ignimbrite, US Highway 93. b Jackpot 6<br />

ignimbrite <strong>al</strong>so contains coarser vitric fragments; from US Highway<br />

93 south of Jackpot, Nevada. Rule shows centim<strong>et</strong>res<br />

sorting (84%-16%) / 2 (phi)<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

99%<br />

96%<br />

0<br />

-10 -8 -6 -4 -2 0 2 4 6<br />

(finer grained)<br />

median diam<strong>et</strong>er (phi)<br />

SR-type ignimbrites of the centr<strong>al</strong> SRP<br />

Sand Springs Ignimbrite<br />

Grey's Landing Ignimbrite<br />

Dry Gulch 1 Ignimbrite<br />

Steer Basin ignimbrite<br />

Cougar Point Tuff 15<br />

Cougar Point Tuff 11b<br />

Cougar Point Tuff 11a<br />

Eastern SRP ignimbrites<br />

Tuff of Kilgore<br />

Tuff of Wolverine Creek<br />

concentrations were too low to cause significant grain<br />

interactions or dampen turbulence even within bas<strong>al</strong> parts<br />

(‘fully dilute’ pyroclastic density currents of Branney and<br />

Kokelaar 2002), and they contain accr<strong>et</strong>ionary lapilli with<br />

concentric laminations as well as cored accr<strong>et</strong>ionary lapilli<br />

in which the concentric laminations enclose a centr<strong>al</strong> vitric<br />

fragment (Fig. 8a and b).<br />

Vitroclastic textures reve<strong>al</strong> that most of the juvenile shards<br />

are relatively thick, platy and cuspate, bubble-w<strong>al</strong>l types<br />

(Fig. 6a). In addition there are rare occurrences of subspheric<strong>al</strong>,<br />

globule-shaped shards that appear to be more<br />

mafic than the surrounding cuspate shards (Fig. 6b). These<br />

are non-flattened despite the eutaxitic nature of the surrounding<br />

cuspate vitroclastic matrix. This suggests that, as with<br />

achneliths in bas<strong>al</strong>tic fire-fountaining eruptions, the globuleshaped<br />

shards were liquid at the time of fragmentation, y<strong>et</strong><br />

solidified prior to deposition, possibly as a result of higher<br />

solidification temperatures than those of the surrounding<br />

welded rhyolitic shards. In many SR-type ignimbrites,<br />

however, welding and rheomorphism has pervasively transposed<br />

and obliterated the vitroclastic textures (Fig. 6c).<br />

Unusu<strong>al</strong>ly high welding intensities, and contact-therm<strong>al</strong><br />

effects<br />

Most Miocene ignimbrites in the centr<strong>al</strong> Snake River Plain<br />

range from high-grade (sensu W<strong>al</strong>ker 1983) to extremely<br />

high-grade (sensu Branney and Kokelaar 1992). Welding is<br />

intense—typic<strong>al</strong>ly off the top of the welding intensity sc<strong>al</strong>e<br />

proposed by Quane and Russell (2005)—and it extends<br />

right to the base of many of the ignimbrites. Contacttherm<strong>al</strong><br />

effects extend to as much as 5 m beneath some<br />

ignimbrites: underlying stratified ash layers are fused to<br />

black vitrophyre (Fig. 7a) and substrate p<strong>al</strong>aeosols are<br />

intensely baked to red-brown terracotta and have developed<br />

columnar joints by cooling-contraction (Fig. 7a). Another<br />

unusu<strong>al</strong> characteristic of the welding in SR-type ignimbrites<br />

is that it som<strong>et</strong>imes extends to, or close to, the preserved


Bull Volcanol (<strong>2008</strong>) 70:293–314 299<br />

Fig. 6 Typic<strong>al</strong> microscopic textures of Snake River-type ignimbrites.<br />

a SEM image of non-welded bi- and tri-cuspate bubble-w<strong>al</strong>l type<br />

shards. Base of Cougar Point Tuff Member XV, Murphy Hot Springs.<br />

b Welded cuspate shards form a eutaxitic fabric, with rare, nondeformed<br />

mafic globule-shaped shards (white arrows). Ignimbrite<br />

from the Mount Benn<strong>et</strong>t Hills (Tuff of Fir Grove GR: 663660 478290;<br />

photomicrogram of thin section in PPL; sc<strong>al</strong>e as in A). c Intense<br />

welding and rheomorphism has transposed and attenuated former<br />

vitroclastic textures, producing a flow-lamination similar to that seen<br />

in silicic lavas (the flow lamination grades into vitroclastic tuff near<br />

the base and top of the ignimbrite: not shown). Isoclin<strong>al</strong> fold pair<br />

indicates top-to-left rheomorphic shear. Bas<strong>al</strong> vitrophyre of the Grey’s<br />

Landing ignimbrite (Rogerson Formation); Grey’s Landing, Twin<br />

F<strong>al</strong>ls County, Idaho<br />

Fig. 7 Field characteristics of SR-type ignimbrites. Columnar-jointed<br />

baked p<strong>al</strong>aeosol (behind notebook) and fused par<strong>al</strong>lel-laminated<br />

ashf<strong>al</strong>l, beneath the massive bas<strong>al</strong> vitrophyre (top) of the Greys<br />

Landing ignimbrite. Arrow indicates base of ignimbrite. Sc<strong>al</strong>e is<br />

10 cm. Backwaters, Idaho, GR:685154 4659673. b Isoclin<strong>al</strong> folding in<br />

flow-laminated, lava-like facies of the Greys Landing Ignimbrite<br />

(location and sc<strong>al</strong>e as in a). c Open rheomorphic fold with<br />

subhorizont<strong>al</strong> fold axis in upper part of ignimbrite folds steep isoclin<strong>al</strong><br />

folds (similar to those in b). Steer basin ignimbrite. Cassia Mountains.<br />

Sc<strong>al</strong>e 5 m high<br />

tops of units, with fusing of overlying stratified ash layers<br />

(e.g. layer of fused accr<strong>et</strong>ionary lapilli - bearing stratified<br />

tuff overlying Jackpot 5; Fig. 8b).<br />

Pervasive and intense rheomorphism<br />

The intense welding of SR-type ignimbrites is commonly<br />

associated with intense rheomorphic deformation. Perva-


300 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

Fig. 8 Ash aggregates in SR-type tephras. a Concentric-laminated<br />

accr<strong>et</strong>ionary lapilli from a non-welded, cross-stratified lapilli-tuff,<br />

deposited from a long-runout, fully dilute pyroclastic density current.<br />

Note abundant angular vitric fragments (black), some coated in fine<br />

ash. Jackpot Member 6; US Highway 93, south of Jackpot, Nevada. b<br />

Welded accr<strong>et</strong>ionary lapilli, some with vitric cores (black). The<br />

accr<strong>et</strong>ionary lapilli are framework supported in some layers, and<br />

supported in fine tuff vitrophyre matrix (dark, bottom) in others.<br />

Jackpot 6, Nevada. c Framework-supported ash coated pell<strong>et</strong>s, some<br />

with vitric cores (left) in ashf<strong>al</strong>l deposits. Lower part of Cougar Point<br />

Tuff XV at Murphy Hot Springs, Idaho. Sc<strong>al</strong>e in cm<br />

sive development of elongation lineations, including<br />

str<strong>et</strong>ched vesicles associated with pervasive flow-lamination,<br />

and open to isoclin<strong>al</strong> sm<strong>al</strong>l- to medium-sc<strong>al</strong>e folds<br />

(Fig. 7b, c), including oblique folds and sheath folds,<br />

indicate that the ignimbrites underwent agglutination and<br />

rapid ductile shear whilst still hot (Branney <strong>et</strong> <strong>al</strong>. 2004). In<br />

ignimbrites elsewhere, welding is commonly thought to<br />

post-date ignimbrite emplacement (‘load welding’ Freundt<br />

1999), but in SR-type ignimbrites, both the welding and the<br />

ons<strong>et</strong> of rheomorphic deformation are thought to have<br />

occurred very rapidly, during deposition (Branney and<br />

Kokelaar 1992). Shear directions recorded by fold-axes<br />

and elongation lineations in some SR-type ignimbrites vary<br />

with height through an individu<strong>al</strong> unit, and are thought to<br />

indicate changing shear directions within a diachronous<br />

rheomorphic shear-zone that ascended through the aggrading<br />

agglutinate while deposition occurred from the base of<br />

an over-riding, sustained and shifting pyroclastic density<br />

current (Branney <strong>et</strong> <strong>al</strong>. 2004; Andrews and Branney 2005;<br />

Andrews 2006). Top surfaces of sever<strong>al</strong> of the ignimbrites<br />

have been folded into tight ogive-like medium-sc<strong>al</strong>e folds,<br />

commonly associated with high-angle thrusts: this indicates<br />

late-stage rheomorphic flow, similar to that seen elsewhere<br />

in silicic lavas and strongly per<strong>al</strong>k<strong>al</strong>ine ignimbrites (e.g.<br />

Sumner and Branney 2002). Intense welding and rheomorphism<br />

in some of the ignimbrites persists even where<br />

the ignimbrite thins to less than 5 m.<br />

Lava-like lithofacies<br />

An unusu<strong>al</strong> characterisitc of SR-type ignimbrites is the<br />

widespread development of lava-like facies. We use ‘lav<strong>al</strong>ike’<br />

as a non-gen<strong>et</strong>ic term (after Branney and Kokelaar<br />

1992) for a lithofacies that looks like lava; it may be<br />

massive or flow-banded (e.g. Fig. 6c) but does not exhibit<br />

vitroclastic or eutaxitic texture (e.g. Fig. 6a, b). In most<br />

volcanic fields lava-like facies constitute a rare or restricted<br />

facies of rheomorphic ignimbrites. However, in SR-type<br />

volcanism lava-like facies dominate entire ignimbrite she<strong>et</strong>s<br />

(e.g. House Creek, Grey’s Landing, Big Bluff, Jackpot 1–5<br />

and Castleford Crossing ignimbrites and Cougar Point<br />

Tuffs 11 and 13; Bonnichsen <strong>et</strong> <strong>al</strong>. 2007). Lava-like facies<br />

of SR-type ignimbrites are indistinguishable from true lava<br />

in both hand specimen and in thin section, and their<br />

pyroclastic origin can be inferred only from the field<br />

relations: for example, rheomorphic ignimbrites gener<strong>al</strong>ly<br />

do not have widespread bas<strong>al</strong> autobreccias, whereas bas<strong>al</strong><br />

autobreccias characterise most rhyolite lavas; <strong>al</strong>so, the lav<strong>al</strong>ike<br />

facies in some ignimbrites grade later<strong>al</strong>ly and/or<br />

vertic<strong>al</strong>ly into less intensely welded, unequivoc<strong>al</strong> vitroclastic<br />

tuff (Bonnichsen and Kauffman 1987; Branney <strong>et</strong> <strong>al</strong>.<br />

1992; Henry and Wolff 1992). A characteristic problem of<br />

SR-type volcanism is that in cases where critic<strong>al</strong> field<br />

relations (e.g. bas<strong>al</strong> contacts and dist<strong>al</strong> feather-edges) are<br />

not exposed, it can be <strong>al</strong>most impossible to d<strong>et</strong>ermine<br />

wh<strong>et</strong>her some predominantly lava-like units are rheomorphic<br />

ignimbrites or true lavas. Although lava-like facies<br />

occur within ignimbrites of some strongly per<strong>al</strong>k<strong>al</strong>ine<br />

volcanic fields (Sumner and Branney 2002), they are a


Bull Volcanol (<strong>2008</strong>) 70:293–314 3<strong>01</strong><br />

relatively subordinate facies of ignimbrites within most<br />

m<strong>et</strong><strong>al</strong>uminous rhyolitic provinces, with the exception of<br />

some thick, c<strong>al</strong>dera-filling ignimbrites (e.g. Batchelor<br />

c<strong>al</strong>dera, Lipman 1984; Scafell c<strong>al</strong>dera, Branney <strong>et</strong> <strong>al</strong>. 1992).<br />

Other characteristics of the ignimbrites<br />

SR-type ignimbrites form cooling units with upper and<br />

lower vitrophyres, each as much as 5 m thick, enclosing a<br />

lithoid<strong>al</strong> (microcryst<strong>al</strong>line) centr<strong>al</strong> zone, which may be as<br />

thick as 150 m (Bonnichsen and Citron 1982). Where the<br />

ignimbrites thin over topography, the vitrophyres merge to<br />

form a single vitrophyre. They characteristic<strong>al</strong>ly contain<br />

later<strong>al</strong>ly extensive zones of spherulites and lithophysae,<br />

ranging from a few millim<strong>et</strong>res up to 50 cm in diam<strong>et</strong>er,<br />

close to the contact with the centr<strong>al</strong> lithoid<strong>al</strong> zone. The<br />

spherulites are typic<strong>al</strong>ly spheroid<strong>al</strong>, but are oblate where<br />

their growth was influenced by the presence of a pronounced<br />

welding foliation, and are prolate where they grew<br />

in vitrophyre with a strong rheomorphic elongation lineation.<br />

Primary porosity has disappeared in intensely welded<br />

facies, but upper parts of many of the ignimbrites variously<br />

contain str<strong>et</strong>ched and spheric<strong>al</strong> vesicles that indicate<br />

continued vesiculation during and after the fin<strong>al</strong> stages of<br />

rheomorphic deformation. Some of the ignimbrites have<br />

developed scoriaceous or coarsely pumiceous zones in<br />

upper parts (e.g. in Cougar Point Tuff 11; Bonnichsen and<br />

Citron 1982: Tuff of Dry Gulch at Rock Creek, GR:72<strong>01</strong>59<br />

4698838), in which the w<strong>al</strong>ls of the elongated vesicles have<br />

become so attenuated that the norm<strong>al</strong>ly black vitrophyre<br />

takes on a golden brown colour, similar to that seen in<br />

bas<strong>al</strong>tic r<strong>et</strong>iculite. Clearly, in SR-type ignimbrites density<br />

cannot be used as a proxy for welding intensity. However,<br />

most SR-type rheomorphic ignimbrites are denser and<br />

much less vesicular than is typic<strong>al</strong> of rheomorphic<br />

ignimbrites elsewhere, such as the eutaxitic W<strong>al</strong>l Mountain<br />

Tuff of Colorado (Chapin and Lowell 1979) and the Green<br />

Tuff of Pantelleria (Mahood 1984).<br />

Flow banding, or lamination, is common in the ignimbrite<br />

vitrophyres and lithoid<strong>al</strong> zones. Some of it may have<br />

originated from he<strong>al</strong>ed fractures, former vesicular zones<br />

and other textur<strong>al</strong> h<strong>et</strong>erogeneities that have become<br />

transposed and attenuated during ductile shear, as in flowbanded<br />

lavas, except that in the case of SR-type ignimbrites<br />

it must have developed at the surface by rheomorphism<br />

after deposition and agglutination, rather than within a<br />

magma-filled eruption conduit. Close-spaced (1–2 cm)<br />

straight or curved ‘she<strong>et</strong>ing joints’ (Bonnichsen and Citron<br />

1982) are abundant within the lithoid<strong>al</strong> zones. Their<br />

orientation is usu<strong>al</strong>ly developed sub-par<strong>al</strong>lel to the flow<br />

banding, apart from at tight fold hinges, and they are<br />

probably developed during static devitrification, mim<strong>et</strong>ic<br />

after the anisotropy of the flow banding.<br />

SR-type volcanism is characterised by a gener<strong>al</strong> paucity<br />

of low-grade (non-welded) and moderate-grade (partly<br />

welded) ignimbrites with extensive zones of non-welded<br />

lapilli-ash, sillar, and non-rheomorphic eutaxitic lapilli-tuff.<br />

This is in marked contrast with many other ignimbrite<br />

fields, such as the Jemez Mountains of New Mexico, Taupo<br />

Volcanic Zone in New Ze<strong>al</strong>and, and the Centr<strong>al</strong> Mexican<br />

Volcanic Belt, where those facies dominate. Non-welded<br />

ignimbrites do occur in the centr<strong>al</strong> Snake River Plain, such<br />

as the ash-pell<strong>et</strong>—bearing ignimbrites in the Cassia Hills<br />

(Fig. 1; see later section), but are not widely reported: this<br />

may in part be due to buri<strong>al</strong> by t<strong>al</strong>us and slope-wash from<br />

overlying, cliff-forming welded units, and partly because of<br />

bioturbation and reworking (e.g. unit below the Three<br />

Creek Rhyolite of Bonnichsen and Citron 1982). Loose,<br />

non-welded ignimbrites would be more prone to erosion and<br />

reworking, and may have contributed to the numerous layers<br />

of rhyolitic volcaniclastic sands (see later section). However<br />

the paucity of eutaxitic and moderately welded facies cannot<br />

be so easily explained in this way. Sever<strong>al</strong> other common<br />

features typic<strong>al</strong> of ignimbrites elsewhere, such as coarse-tail<br />

grading, intern<strong>al</strong> layering and diffuse stratification, and<br />

lithic-rich and pumice-rich lenses (Branney and Kokelaar<br />

2002) <strong>al</strong>so have not been reported in the region.<br />

Extensive layers of par<strong>al</strong>lel-stratified rhyolitic ash<br />

Thin-bedded to laminated, grey to white rhyolitic ash is a<br />

common facies of SR-type volcanism (Fig. 9a) and forms<br />

layers ≤5 m thick. The stratification is formed of sharp and<br />

gradation<strong>al</strong> changes in grainsize; millim<strong>et</strong>re-sc<strong>al</strong>e norm<strong>al</strong><br />

and reverse, vertic<strong>al</strong> distribution grading is common, with<br />

no o<strong>bv</strong>ious over<strong>al</strong>l asymm<strong>et</strong>ry or cyclicity. Individu<strong>al</strong> beds,<br />


302 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

Fig. 9 Typic<strong>al</strong> SR-type ashf<strong>al</strong>l<br />

deposits. a Par<strong>al</strong>lel-laminated<br />

and thin-bedded fine to coarse<br />

ash, showing later<strong>al</strong>ly persistent<br />

stratification. Nat-Soo-Pah,<br />

Twin F<strong>al</strong>ls County, Idaho. M<strong>et</strong>re<br />

ruler. b Fused, largely par<strong>al</strong>lelstratified<br />

tuff of dominantly<br />

ashf<strong>al</strong>l origin, with layers of<br />

abundant fused framework-supported<br />

sm<strong>al</strong>l (


Bull Volcanol (<strong>2008</strong>) 70:293–314 303<br />

Fig. 10 Typic<strong>al</strong> vitroclastic textures of SR-type ashf<strong>al</strong>l deposits. a<br />

SEM image of cuspate, bubble-w<strong>al</strong>l shards. Bas<strong>al</strong> ashf<strong>al</strong>l of Cougar<br />

Point Tuff Member XV, Murphy Hot Springs, Owyhee County, Idaho.<br />

b SEM image of bubble-w<strong>al</strong>l shards from an ashf<strong>al</strong>l f<strong>al</strong>l deposit<br />

overlying the Grey’s Landing ignimbrite, at S<strong>al</strong>mon Dam, Twin F<strong>al</strong>ls<br />

County, Idaho<br />

Bonnichsen 1982; Ekren <strong>et</strong> <strong>al</strong>. 1984; Bonnichsen and<br />

Kauffman 1987; Bonnichsen <strong>et</strong> <strong>al</strong>. 1989; Henry and Wolff<br />

1992) and we summarise only their s<strong>al</strong>ient features.<br />

At least eight rhyolite lavas from the Bruneau-Jarbidge<br />

eruptive centre exceed 10 km 3 each: the Dorsey Creek<br />

Rhyolite (Appendix Fig. 3) exceeds 75 km 3 and is over<br />

40 km long, and the Sheep Creek Rhyolite exceeds 200 km 3<br />

(Bonnichsen 1982). Large rhyolite lavas are present <strong>al</strong>so at<br />

the inferred Twin F<strong>al</strong>ls eruptive centre (e.g. Shoshone F<strong>al</strong>ls<br />

Rhyolite, B<strong>al</strong>anced Rock Rhyolite, Bonnichsen <strong>et</strong> <strong>al</strong>. 1989),<br />

and in the Juniper Mountains of SW Idaho (e.g. the Badlands<br />

Rhyolite, Manley 1996; Manley and McIntosh 2002). Sm<strong>al</strong>l<br />

rhyolite domes and coulées are not characteristic of SR-type<br />

volcanism, <strong>al</strong>though the possibility that some lie conce<strong>al</strong>ed<br />

within subsided eruptive centres cannot be excluded.<br />

Aspect ratios of SR-type rhyolite lavas are unlike those of<br />

viscous rhyolite lavas elsewhere (Fig. 2). They are sufficiently<br />

low to coincide with those of many bas<strong>al</strong>t lavas (Fig. 2).<br />

This is reflected in their long distance run-outs from their<br />

inferred source locations. The unusu<strong>al</strong> dimensions led to<br />

early interpr<strong>et</strong>ations that they were ignimbrites (Ekren <strong>et</strong> <strong>al</strong>.<br />

1984), but they are now thought to be lavas on the basis of<br />

sever<strong>al</strong> criteria, including the presence of widespread bas<strong>al</strong><br />

autobreccias and abrupt, thick, stubby lobate terminations<br />

with thick t<strong>al</strong>us aprons (e.g. see Bonnichsen and Kauffman<br />

1987; Henry and Wolff 1992; Manley and McIntosh 2002).<br />

In contrast, bas<strong>al</strong> autobreccias in rheomorphic ignimbrites<br />

are rare and restricted to locations where rheomorphic flow<br />

has carried the hot agglutinate beyond the origin<strong>al</strong> extent of<br />

pyroclastic deposition (e.g. Sumner and Branney 2002).<br />

The lavas are blocky with thick (>5 m), variously vitric<br />

and coarsely pumiceous lower, margin<strong>al</strong> and upper carapace<br />

autobreccias (Bonnichsen 1982). Hot-state shear of<br />

bas<strong>al</strong>, margin<strong>al</strong>, and intern<strong>al</strong> coarsely pumiceous autobreccia<br />

during flowage has caused loc<strong>al</strong>ised fusing and ductile<br />

shear, with the development of vitroclastic textures that can<br />

superfici<strong>al</strong>ly resemble welded pyroclastic textures (Manley<br />

1996) as is common in viscous blocky lavas elsewhere (e.g.<br />

Iddings 1889; Pichler 1981; Sparks <strong>et</strong> <strong>al</strong>. 1993). Thick,<br />

centr<strong>al</strong> zones of the lavas (Appendix Fig. 3) are lithoid<strong>al</strong><br />

(microcryst<strong>al</strong>line), massive or flow-banded, and dominated<br />

by steep columnar jointing, and low-angle close-spaced<br />

she<strong>et</strong>ing joints that in places form intersecting s<strong>et</strong>s,<br />

producing ‘pencil-type’ jointing (Bonnichsen 1982). Spherulites<br />

and lithophysae are common near the base of the<br />

lithoid<strong>al</strong> zone. Some of the lavas, particularly to the west of<br />

the centr<strong>al</strong> Snake River Plain, exhibit agglutinate textures<br />

with spatter rags that indicate a clastogenic origin by some<br />

form of rhyolitic fire fountaining (e.g. Juniper Mountains;<br />

Manley and McIntosh 2002), <strong>al</strong>though the majority of the<br />

rhyolite lavas are not visibly clastogenic.<br />

Lacustrine-<strong>al</strong>luvi<strong>al</strong> facies<br />

The facies association that characterises SR-type volcanism<br />

in south centr<strong>al</strong> Idaho and northern Nevada includes<br />

widespread evidence for the presence of surface water at<br />

the time of the eruptions.<br />

Aqueously deposited volcaniclastic sediments<br />

Numerous thin layers of aqueously deposited volcaniclastic<br />

sediments are interc<strong>al</strong>ated with ignimbrites across the centr<strong>al</strong><br />

and western Snake River Plain (Appendix Fig. 4a, b). Alluvi<strong>al</strong><br />

and lacustrine facies are represented, with graded beds,<br />

par<strong>al</strong>lel-laminated silts, ripple cross-lamination, scours, and<br />

abundant soft-sediment deformation, loading and dewatering<br />

structures. Many of the par<strong>al</strong>lel-stratified ash layers appear to<br />

have been deposited directly into sh<strong>al</strong>low standing water as<br />

there are loc<strong>al</strong> gradations into facies with minor truncations<br />

and scours that indicate reworking by gentle aqueous currents<br />

or wind, and into rippled sands of the same composition (e.g.<br />

Perkins <strong>et</strong> <strong>al</strong>. 1995). Lacustrine sediments occur b<strong>et</strong>ween the


304 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

Cougar Point Tuffs around Jarbidge; within the Rogerson<br />

Formation near Jackpot and in the Rogerson Graben; and<br />

b<strong>et</strong>ween ignimbrites dated at 13.7 to 8.6 Ma in the Shoshone<br />

Basin, Cassia Hills and Goose Creek Basin (Fig. 1; e.g.<br />

Hildebrand and Newman 1985; Perkins <strong>et</strong> <strong>al</strong>. 1995).<br />

Some of the lacustrine interv<strong>al</strong>s within the Miocene<br />

ignimbrite-dominated successions north and south of the<br />

centr<strong>al</strong> Snake River Plain (Appendix Fig. 4a, b) may reflect<br />

somewhat isolated and/or short-lived bodies of water, as<br />

they form thin (100 m) masses of dominantly vitrophyric,<br />

clast-supported breccia, and ch<strong>al</strong>cedony, jasper and op<strong>al</strong><br />

loc<strong>al</strong>ly fills fractures and other cavities in the lavas, with<br />

extensive groundmass silicification to their deep interiors.<br />

Bas<strong>al</strong>tic pillow lava–hy<strong>al</strong>oclastite deltas<br />

Bas<strong>al</strong>t lavas in centr<strong>al</strong> southern Idaho include hydrated<br />

bas<strong>al</strong>ts (water-affected bas<strong>al</strong>ts or ‘WAB’ of Bonnichsen and


Bull Volcanol (<strong>2008</strong>) 70:293–314 305<br />

Fig. 11 Ignimbrite peperites from the centr<strong>al</strong> Snake River Plain,<br />

Idaho. a Peperite at the base of an ignimbrite (‘Picabo Tuff’), just<br />

north of the plain. Bas<strong>al</strong> vitrophyre (dark) is finely brecciated and<br />

loaded into w<strong>et</strong> silt (p<strong>al</strong>e). b Ignimbrite vitrophyre peperite at the base<br />

of the Wooden Shoe Tuff, Rock Creek Canyon, Cassia Mountains.<br />

Note pervasive injection of disaggregated silt (p<strong>al</strong>e) into fractures in<br />

the glassy welded ignimbrite. c D<strong>et</strong>ail of peperite showing clouds of<br />

admixed angular hy<strong>al</strong>oclasts and silt. Location as for b<br />

Godchaux 2002), pillow lavas and hy<strong>al</strong>oclastites, including<br />

prograding pillow-hy<strong>al</strong>oclastite deltas (Appendix Fig. 4d;<br />

Shervais <strong>et</strong> <strong>al</strong>. 2005). Evidence for the interaction of the<br />

bas<strong>al</strong>t lavas with lake water, streams and associated ground<br />

water has been widely documented (Jenks and Bonnichsen<br />

1989, refs in Bonnichsen and Godchaux 2002).<br />

Phreatomagmatic tuffs<br />

Sever<strong>al</strong> bas<strong>al</strong>tic tuff cones and tuff rings dating back to<br />

Miocene times occur within the region of former Lake<br />

Idaho (Godchaux and Bonnichsen 2002). The tuffs contain<br />

abundant ash pell<strong>et</strong>s and accr<strong>et</strong>ionary lapilli, exhibit w<strong>et</strong>,<br />

soft-state deformation and are associated with aqueously<br />

reworked facies. Both Surtseyan and Ta<strong>al</strong>ian (maar-forming;<br />

Kokelaar 1986) eruption styles are represented and<br />

indicate that the erupting magma interacted explosively<br />

with, respectively, lake water and groundwater. There are<br />

few equiv<strong>al</strong>ent examples of rhyolitic phreatomagmatic<br />

centres because the rhyolitic source vents are buried in the<br />

interior of the centr<strong>al</strong> Snake River Plain. However, the<br />

numerous layers of fine vitric ash containing abundant ash<br />

aggregates that occur within the ignimbrite successions may<br />

derive from water-enhanced explosivity. For example,<br />

white, fine ash - rich ashf<strong>al</strong>l layers and associated nonwelded<br />

ignimbrites containing abundant coated ash pell<strong>et</strong>s<br />

occur in eruption units in the Cassia Hills and Trapper<br />

Creek (Fig. 1). The subaeri<strong>al</strong> vent area of the Wilson Creek<br />

ignimbrite is exposed in the western Snake River Plain, and<br />

proxim<strong>al</strong> water-<strong>al</strong>tered phreatomagmatic tuffs contain<br />

abundant angular chips of older rock types indicating<br />

explosive fragmentation of near-surface, water-bearing<br />

rocks by rising rhyolite magma prior to the main ignimbrite<br />

eruption into Lake Idaho (Ekren <strong>et</strong> <strong>al</strong>. 1984; Godchaux and<br />

Bonnichsen 2002; Bonnichsen <strong>et</strong> <strong>al</strong>. 2007).<br />

Associated bas<strong>al</strong>tic volcanism<br />

The rhyolitic rocks are associated with bas<strong>al</strong>ts, in the form<br />

of scattered tuff rings and co<strong>al</strong>escing low-profile shield<br />

volcanoes with gentle concave slopes (1–2° flanks steepening<br />

to 5° near summits) and som<strong>et</strong>imes spatter ramparts<br />

surrounding Hawaiian-type summit collapse craters (e.g.<br />

Shervais <strong>et</strong> <strong>al</strong>. 2005). Some of the bas<strong>al</strong>t lavas are fed by<br />

fissures <strong>al</strong>ong rift-zones. The bas<strong>al</strong>tic volcanism is transition<strong>al</strong><br />

b<strong>et</strong>ween typic<strong>al</strong> (i.e. steeper) shield volcanoes and<br />

true flood bas<strong>al</strong>ts, and has been termed ‘bas<strong>al</strong>tic plains style<br />

volcanism’ to distinguish it from those end-members<br />

(Greeley 1977, 1982). Most of the visible bas<strong>al</strong>t forms a<br />

cap overlying the Miocene ignimbrites and lavas across the<br />

width of the plain. Because of limited incision, it not known<br />

when bas<strong>al</strong>tic volcanism began in the centr<strong>al</strong> Snake River<br />

Plain and wh<strong>et</strong>her any early bas<strong>al</strong>t effusion occurred<br />

associated with emplacement of the inferred mid-crust<strong>al</strong><br />

mafic intrusion (Rogers <strong>et</strong> <strong>al</strong>. 2002).<br />

Discussion<br />

This section considers the eruptive processes, the nature of<br />

the eruptive centres, and other examples worldwide where<br />

SR-type volcanism may have occurred.


306 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

SR-type eruptive processes<br />

Snake River-type eruptions were voluminous and extraordinarily<br />

environment<strong>al</strong>ly devastating, with large-sc<strong>al</strong>e<br />

explosivity (VEI 6–8) during which vast glassy ignimbrites<br />

were fused across the landscape, and accompanied by<br />

widespread atmospheric dispers<strong>al</strong> of vitric ash. The<br />

extensive, predominantly massive and intensely welded<br />

nature of the ignimbrites is best reconciled with hightemperature,<br />

sustained, granular fluid-based pyroclastic<br />

density currents with run-out distances exceeding 100 km<br />

that were fed from high mass-flux pyroclastic fountaining<br />

eruptions (e.g. Bursik and Woods 1996; Branney and<br />

Kokelaar 2002). High eruptive mass flux, possibly via<br />

fissures associated with c<strong>al</strong>dera subsidence, would have<br />

minimised cooling during transport (e.g. Branney <strong>et</strong> <strong>al</strong>.<br />

1992; Freundt 1999). Non-welded bases of a few of the<br />

ignimbrites may be the legacy of entrainment and mixing of<br />

atmospheric air into leading parts of the currents during<br />

early waxing stages of eruptions while the currents initi<strong>al</strong>ly<br />

advanced across the landscape. Such flow-front cooling<br />

seems to have been minor in cases where the ignimbrites<br />

are welded right to their bases. However, it is doubtful that<br />

such ignimbrites were ever initi<strong>al</strong>ly isotherm<strong>al</strong> as is<br />

commonly assumed in welding models (refs in Russell and<br />

Quane 2005): phenocryst geothermom<strong>et</strong>ry indicates that<br />

initi<strong>al</strong> therm<strong>al</strong> gradients may have existed in the ignimbrites<br />

even before they started cooling (Andrews <strong>et</strong> <strong>al</strong>. 2007).<br />

Gradu<strong>al</strong> vertic<strong>al</strong> changes in the trend of rheomorphic<br />

structures (elongation lineations and sheathfold hinges)<br />

within some of the most intensely welded ignimbrites<br />

indicate that pyroclasts in these examples had viscosities<br />

sufficiently low to enable them to agglutinate and shear<br />

even during deposition (Branney <strong>et</strong> <strong>al</strong>. 2004). The abundance<br />

of lava-like facies indicates that co<strong>al</strong>escence of hot<br />

pyroclasts, with obliteration of the clast outlines (Branney<br />

and Kokelaar 1992) was common: such rapid welding<br />

requires unusu<strong>al</strong>ly low viscosities compared to those<br />

required for post-deposition<strong>al</strong> welding (‘load welding’ of<br />

Freundt 1999) which is thought to characterise ignimbrites<br />

elsewhere (e.g. Ross and Smith 1961). The hot deposits<br />

r<strong>et</strong>ained a sufficiently low viscosity to spread gravitation<strong>al</strong>ly<br />

across low (≤5°) topographic slopes. Exsolution of<br />

volatiles continued, with the growth and rheomorphic<br />

attenuation of ellipsoid<strong>al</strong> vesicles and in some cases<br />

development of golden, thick-w<strong>al</strong>led coarse pumiceous<br />

zones in upper parts of ignimbrites; such frothy lava-like<br />

facies are gener<strong>al</strong>ly rare in ignimbrites elsewhere.<br />

Low viscosity of pyroclasts in some rheomorphic<br />

ignimbrites elsewhere is attributable to strongly per<strong>al</strong>k<strong>al</strong>ine<br />

chemistries in which <strong>al</strong>k<strong>al</strong>is act as n<strong>et</strong>work-modifiers<br />

disrupting silicate polymerisation (Mahood 1984). The<br />

globular shards discovered in the (m<strong>et</strong><strong>al</strong>uminous) SR-type<br />

ignimbrites are found <strong>al</strong>so in some strongly per<strong>al</strong>k<strong>al</strong>ine<br />

rheomorphic ignimbrites (e.g. Johnson 1968; Sumner and<br />

Branney 2002) and indicate pyroclast viscosities sufficiently<br />

low for clast shapes to be controlled by surface-tension<br />

prior to chilling during transport (Branney and Kokelaar<br />

1992), just as in dropl<strong>et</strong>-shaped pyroclasts from Hawaiian<br />

fire-fountaining eruptions (Pele’s tears or ‘achneliths’ of<br />

W<strong>al</strong>ker and Croasd<strong>al</strong>e 1972). However, the SR-type<br />

eruptions are not per<strong>al</strong>k<strong>al</strong>ine and the low viscosities must<br />

result from high emplacement temperatures and possibly<br />

r<strong>et</strong>ention of some dissolved volatiles. This is consistent<br />

with high magmatic temperatures of 830–1,050°C estimated<br />

from pyroxenes, feldspars and oxide phases in SRignimbrites,<br />

and estimates that at least some of the rhyolite<br />

magmas had high fluorine contents (e.g. Honjo <strong>et</strong> <strong>al</strong>. 1992;<br />

Cathey and Nash 2004; Christiansen and McCurry 2007).<br />

Elsewhere, unusu<strong>al</strong>ly low rhyolite viscosities indicated by<br />

rocks thought to record rhyolitic fountaining and agglutination<br />

have <strong>al</strong>so been attributed to high fluorine contents<br />

(Duffield 1990). The high mass-flux eruptions may have<br />

minimised both cooling and the exsolution of certain<br />

volatile species during eruption and emplacement.<br />

The absences in SR-type ignimbrites of features like lithic<br />

and pumice dunes, bedding and pumice concentration zones,<br />

may be because agglutination of sticky particles limited<br />

granular segregation. However, absence of such features <strong>al</strong>so<br />

characterises the rare non-welded facies, and so may simply<br />

reflect the paucity of pumice and lithic lapilli. Paucity of<br />

lithic lapilli is a characteristic of extremely high-grade<br />

ignimbrites elsewhere (e.g. Branney <strong>et</strong> <strong>al</strong>. 1992) and may<br />

reflect minim<strong>al</strong> erosion and entrainment from the vent<br />

margins and from the land surface by the pyroclastic current<br />

which, instead, tended to plaster surfaces with hot agglutinate.<br />

The gener<strong>al</strong> absence of pumice lapilli and fiamme is<br />

difficult to explain (fiamme are common in many rheomorphic<br />

ignimbrites elsewhere) and points to some fac<strong>et</strong> of the<br />

fragmentation process, in which cuspate shards were<br />

generated from the vesiculating magma more readily than<br />

were pumiceous blocks and lapilli. Such a fragmentation<br />

mechanism is not understood. The morphology and large<br />

size of the cuspate shards in both the ignimbrites and the<br />

ashf<strong>al</strong>l layers indicates that bubble sizes were larger than is<br />

typic<strong>al</strong> of rhyolitic micro-vesicular pumice elsewhere; this<br />

may be the result of more rapid diffusion rates of volatiles<br />

within the rhyolite magmas due to their unusu<strong>al</strong>ly high<br />

temperature and low viscosity. Fiamme present loc<strong>al</strong>ly in<br />

some of the rheomorphic ignimbrites (e.g. Cougar Point Tuff<br />

XI) appear to be lenticular zones of late-vesiculated welded<br />

tuff matrix rather than former pumice clasts.<br />

The stratified ashf<strong>al</strong>l deposits with their widespread<br />

dist<strong>al</strong> correlatives have great v<strong>al</strong>ue in long-distance,<br />

terrestri<strong>al</strong> stratigraphy and tephrochronology (Perkins and<br />

Nash 2002) and record widespread atmospheric dispers<strong>al</strong> of


Bull Volcanol (<strong>2008</strong>) 70:293–314 307<br />

large volumes of vitric ash, and is associated with major<br />

mamm<strong>al</strong>ian death assemblages, c.1,400 km to the west of<br />

the eruptive sources (Voorhies and Thomasson 1979; Rose<br />

<strong>et</strong> <strong>al</strong>. 2003). The widespread dispers<strong>al</strong> would have been<br />

favoured by the combination of high eruptive temperatures<br />

with high eruptive mass-fluxes. It may have occurred via<br />

vent-derived convective ash plumes, and/or from phoenix<br />

clouds (co-ignimbrite ash plumes) that lofted from the<br />

large, hot pyroclastic density currents. Many of the ash<br />

layers are coarser-grained than is typic<strong>al</strong> of co-ignimbrite<br />

ashes elsewhere (cf. W<strong>al</strong>ker 1981; Smith and Houghton<br />

1995), even dist<strong>al</strong>ly (see Rose <strong>et</strong> <strong>al</strong>. 2003). This could<br />

reflect the unusu<strong>al</strong>ly large shard-sizes produced by SR-type<br />

explosivity coupled with enhanced elutriation and lofting as<br />

a result of the large therm<strong>al</strong> input by the exception<strong>al</strong>ly hot<br />

pyroclastic currents. However, <strong>al</strong>though some fine vitric<br />

ash layers overlie sever<strong>al</strong> of the ignimbrites, a significant<br />

proportion of the stratified ash occurs beneath ignimbrites<br />

of the same eruption-unit; such ashes could only be of coignimbrite<br />

origin if the leading edges of the pyroclastic<br />

density currents advanced rather slowly across a landscape<br />

that was <strong>al</strong>ready being mantled in co-ignimbrite ash from<br />

earlier stages of the eruption. The stratified ash layers<br />

beneath the ignimbrites could, <strong>al</strong>ternatively, derive from<br />

Plinian explosivity. However, they are finer grained than is<br />

typic<strong>al</strong> for Plinian deposits of a similar thickness: Plinian<br />

deposits thicker than a m<strong>et</strong>re or so are typic<strong>al</strong>ly composed<br />

of predominantly lapilli-sized pumice clasts (e.g. at Vesuvius,<br />

V<strong>al</strong>les, Santorini); because they tend to thin out as<br />

they become more fine grained dist<strong>al</strong>ly, a Plinian layer<br />

more than a couple of m<strong>et</strong>res thick of only ash sized<br />

particles would represent an unusu<strong>al</strong>ly large eruption.<br />

Alternatively, the lack of pumice lapilli may be a fac<strong>et</strong> of<br />

the fragmentation mechanism. One possibility is that the<br />

explosive fragmentation of the vesiculating magma was<br />

enhanced by interaction with m<strong>et</strong>eoric water. This would be<br />

consistent with the widespread evidence for lacustrine<br />

conditions: Snake River Plain volcanic rocks are important<br />

aquifers, and one might expect that water occupied<br />

fractures beneath lakewater, for example in areas of<br />

proxim<strong>al</strong> volcanic subsidence. The abundance of ash pell<strong>et</strong>s<br />

and accr<strong>et</strong>ionary lapilli in the rhyolitic ash <strong>al</strong>so are<br />

consistent with the involvement of extern<strong>al</strong> water. Such<br />

ash aggregates form commonly (but not exclusively) during<br />

phreatomagmatic eruptions, or where there is abundant<br />

moisture in ash plumes such as is generated by surface<br />

evaporation where a hot density current flows across<br />

standing water. The par<strong>al</strong>lel bedding and lamination<br />

suggests unsteady, pulsatory f<strong>al</strong>lout, generated either at<br />

the vent or by convective instabilities within an umbrella<br />

cloud (e.g. Carey <strong>et</strong> <strong>al</strong>. 1988; Branney 1991). Pulsatory<br />

explosivity is <strong>al</strong>so a characteristic of some phreatomagmatic<br />

eruptions, and can arise from intermittent access of<br />

water to the erupting magma. Some of the ash-pell<strong>et</strong>bearing<br />

ash deposits are very fine grained and have a<br />

distinctly phreatomagmatic character (e.g. units beneath the<br />

Tuff of Wooden Shoe Butte, Cassia Mountains).<br />

However, the granulom<strong>et</strong>ry of many of the stratified ash<br />

layers in the Snake River Plain is not typic<strong>al</strong> of phreatomagmatic<br />

ash. The ash layers are relatively well-sorted and<br />

medium to coarse-grained, whereas phreatoplinian ashes<br />

(e.g. Hatepe and Rotongiao ashes of New Ze<strong>al</strong>and; W<strong>al</strong>ker<br />

1981) tend to be finer grained and less well-sorted. The<br />

cuspate shape of the shards indicates that the magma was<br />

coarsely vesicular at the time of fragmentation, and so the<br />

explosive fragmentation was probably driven primarily by<br />

volatile exsolution. Given that the absence of pumice<br />

blocks and lapilli characterises the ignimbrites as well as<br />

the f<strong>al</strong>lout deposits, reconciling the involvement of large<br />

volumes of m<strong>et</strong>eoric water with the very high welding<br />

intensity of the ignimbrites presents som<strong>et</strong>hing of a<br />

paradox. However, intensely welded pyroclastic rocks do<br />

occur in association with non-welded, ash-aggregate bearing<br />

deposits at some flooded c<strong>al</strong>dera volcanoes, like Ta<strong>al</strong> in<br />

the Philippines (Torres <strong>et</strong> <strong>al</strong>. 1995) and Scafell in England<br />

(Branney and Kokelaar 1995).<br />

The abundant obsidian and vitrophyre fragments in the<br />

ignimbrites and ashf<strong>al</strong>l deposits may have been entrained at<br />

w<strong>al</strong>ls of eruption conduits cutting dense proxim<strong>al</strong> obsidian<br />

and vitrophyre. The absence of other lithic lithologies (e.g.<br />

bas<strong>al</strong>ts, basement) suggests that the explosivity occurred at<br />

sh<strong>al</strong>low-levels in eruptive centres that at the time were<br />

characterised by thick accumulations of glassy rhyolitic<br />

products, possibly in a lacustrine s<strong>et</strong>ting. Incorporation and<br />

fragmentation of these lithologies may have been facilitated<br />

by groundwater within fractures in the glassy rocks flashing<br />

to steam. Similar angular glass chips are reported from<br />

Plinian (Brown and Branney 2004) and phreatoplinian<br />

(Smith and Houghton 1995) f<strong>al</strong>lout deposits elsewhere, and<br />

in the latter case a sh<strong>al</strong>low obsidian source at the floor of a<br />

c<strong>al</strong>dera lake was invoked.<br />

A feature of the SR-type rhyolite lavas and ignimbrites<br />

of the centr<strong>al</strong> part of the Snake River Plain is that they have<br />

remarkably low δ 18 O v<strong>al</strong>ues (−1.4 to 3.8‰; Boroughs <strong>et</strong> <strong>al</strong>.<br />

2005) compared to non-SR-type rhyolites elsewhere in the<br />

province. The low v<strong>al</strong>ues have been ascribed to large-sc<strong>al</strong>e,<br />

sh<strong>al</strong>low melting of Idaho batholith basement that had been<br />

extensively hydrotherm<strong>al</strong>ly <strong>al</strong>tered during the Eocene<br />

(Boroughs <strong>et</strong> <strong>al</strong>. 2005). Intriguingly, however, the low<br />

δ 18 O v<strong>al</strong>ues coincide with the SR-type volcanism, with its<br />

association with former lakes: further exploration of the<br />

possible role of sub-lacustrine groundwater or hydrotherm<strong>al</strong><br />

waters in fractures associated with rifting and/or c<strong>al</strong>dera<br />

formation might be illuminating.<br />

The SR-type rhyolite lavas were emplaced onto low<br />

slopes. Emplacement predominantly by ductile effusion


308 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

rather than brittle and semi-brittle extrusion is indicated by<br />

the complex refolded folds and the large ratio of nonbrecciated<br />

to brecciated lava. Their large sizes and runnout<br />

distances, and low aspect-ratios (Fig. 2) indicate that the<br />

lava flows were erupted at far higher mass-flux rates, and<br />

with significantly lower viscosities than is typic<strong>al</strong> for<br />

rhyolite spines, domes or coulées. As with other examples<br />

(e.g. Bracks Rhyolite of Trans-Pecos Texas; Henry <strong>et</strong> <strong>al</strong>.<br />

1990) the aspect ratios of the silicic lavas are consistent<br />

with estimated high magmatic temperatures and anhydrous<br />

compositions. However, it is not clear what drove such<br />

large volumes of magma so rapidly to the surface.<br />

C<strong>al</strong>dera subsidence in the Snake River plain<br />

Explosive eruptions larger than a few km 3 gener<strong>al</strong>ly<br />

produce c<strong>al</strong>deras, and it is likely that the larger SR-type<br />

eruptions were c<strong>al</strong>dera-forming. Silicic c<strong>al</strong>deras are best<br />

identified by thick intra-c<strong>al</strong>dera ignimbrite tog<strong>et</strong>her with<br />

c<strong>al</strong>dera-collapse breccias, and overlying c<strong>al</strong>dera lake sediments,<br />

rhyolite domes, and associated sh<strong>al</strong>low intrusions,<br />

hydrotherm<strong>al</strong> <strong>al</strong>teration, and contemporary faulting (e.g.<br />

Lipman 1984; Branney 1995). Where these are not<br />

exposed, the presence of a c<strong>al</strong>dera may be inferred from<br />

an extant topographic rim, tog<strong>et</strong>her with proxim<strong>al</strong>ly-coarsening<br />

lithic breccias in outflow ignimbrites (e.g. Druitt and<br />

Sparks 1982) and thick, coarse pumice f<strong>al</strong>l layers. Most of<br />

these features have not been recorded in the centr<strong>al</strong> Snake<br />

River Plain, so the location and size of any c<strong>al</strong>deras in the<br />

region is tentative. For these reasons, Bonnichsen (1982)<br />

inferred the presence of broad ‘eruptive centres’ (Fig. 1) that<br />

were considered to have been discr<strong>et</strong>e topographic depressions<br />

1,000s km 2 containing lakes and extensive rhyolite<br />

lavas, prior to buri<strong>al</strong> by younger bas<strong>al</strong>t lavas.<br />

Source locations of most of the ignimbrites are imprecisely<br />

known and cannot be inferred from thickness<br />

variations of outflow she<strong>et</strong>s. Spati<strong>al</strong> distributions of<br />

azimuth orientations of rheomorphic lineations, folds and<br />

kinematic indicators can help locate an eruptive source<br />

(Branney <strong>et</strong> <strong>al</strong>. 2004). For example, lineations in the upper<br />

tuff of McMullen Creek, Cassia Hills (Fig. 1) form a fanshaped<br />

pattern that may indicate northerly source (McCurry<br />

<strong>et</strong> <strong>al</strong>. 1996). However, rheomorphic transport directions are<br />

likely to be influenced by loc<strong>al</strong> topographic slopes rather<br />

than an over<strong>al</strong>l transport direction from source (e.g. lower<br />

units of the tuffs of McMullen Creek; McCurry <strong>et</strong> <strong>al</strong>. 1996).<br />

It has been proposed that loc<strong>al</strong> successions of outflow<br />

ignimbrites derive from a common eruptive centre, for<br />

example that <strong>al</strong>l nine Cougar Point Tuffs and associated<br />

lavas derive from the Bruneau-Jarbidge eruptive centre, and<br />

that the ignimbrites of the Cassia Mountains derive from<br />

the inferred Twin F<strong>al</strong>ls eruptive centre (Fig. 1). In this<br />

scenario, a sm<strong>al</strong>l number of distinct and spaced eruptive<br />

centres (Fig. 1) each underwent numerous large eruptions.<br />

It is equ<strong>al</strong>ly possible, however, that a larger number of<br />

subsidence structures formed, each accompanied by emplacement<br />

of just one large ignimbrite. In this scenario, the<br />

large number of ignimbrites would suggest that the centr<strong>al</strong><br />

and eastern Snake River Plain is a complex of numerous<br />

partly overlapping c<strong>al</strong>deras, rather like the Olympic rings,<br />

in which the locus of subsidence migrated over<strong>al</strong>l northeastwards.<br />

This is consistent with our understanding of the<br />

younger c<strong>al</strong>deras around Yellowstone, which are shingled,<br />

and where each c<strong>al</strong>dera subsided during a single ignimbrite<br />

eruption (Morgan and McIntosh 2005). The diam<strong>et</strong>ers of<br />

some of these c<strong>al</strong>deras (Blacktail, Kilgore, Huckleberry<br />

Ridge, and Yellowstone) are thought to be similar to the<br />

width of the Snake River Plain. In d<strong>et</strong>ail, the tempor<strong>al</strong> trend<br />

is likely to have been rather more complex than a simple<br />

north-eastward migration (e.g. a large ignimbrite flare-up<br />

occurred 11.7–10.0 Ma from a broad region of the centr<strong>al</strong><br />

Snake River Plain, and tempor<strong>al</strong> overlap of eruptions in<br />

widely separate areas occurred during the subsequent ∼4<br />

million years; Bonnichsen <strong>et</strong> <strong>al</strong>. 2007).<br />

The low topography of the Snake River Plain relative to<br />

the adjacent massifs is associated with marked downwarping<br />

of ignimbrites and basement towards the axis of<br />

the plain. This structure has been interpr<strong>et</strong>ed variously as<br />

due to SW–NE extension, therm<strong>al</strong> contraction following the<br />

volcanism, or loading by large mid-crust<strong>al</strong> mafic intrusions<br />

(discussion in Rogers <strong>et</strong> <strong>al</strong>. 2002). The structur<strong>al</strong>ly-defined<br />

subsidence exceeds 4.5 km and, in some locations, 8.5 km<br />

(Rogers <strong>et</strong> <strong>al</strong>. 2002). We propose that this includes a<br />

component of c<strong>al</strong>dera subsidence; the centr<strong>al</strong> and eastern<br />

Snake River Plain may thus be regarded as a volcanotectonic<br />

depression formed by successive, partly overlapping<br />

c<strong>al</strong>dera subsidence events in addition to later<br />

therm<strong>al</strong> adjustments and flexur<strong>al</strong> loading by a mid-crust<strong>al</strong><br />

mafic intrusion (Rogers <strong>et</strong> <strong>al</strong>. 2002). The inward dip of the<br />

basement and ignimbrite she<strong>et</strong>s towards the axis of the<br />

plain may include a component of c<strong>al</strong>dera-related downwarping<br />

(‘downsag’ of W<strong>al</strong>ker 1984; Branney 1995)<br />

peripher<strong>al</strong> strata that steepen with proximity to the c<strong>al</strong>dera<br />

margins is a common feature of c<strong>al</strong>dera subsidence<br />

(Branney and Kokelaar 1992; Branney 1995; Roche <strong>et</strong> <strong>al</strong>.<br />

2000; Kokelaar and Moore 2006), e.g. the inward tilting<br />

around the 18 km diam<strong>et</strong>er Namibian Messum Complex<br />

and Goboboseb mountains, interpr<strong>et</strong>ed as due to c<strong>al</strong>dera<br />

subsidence associated with very large-volume silicic eruptions<br />

(Ewart <strong>et</strong> <strong>al</strong>. 1998).<br />

Given the thickness of the outflow ignimbrite successions,<br />

we expect proxim<strong>al</strong> ignimbrite accumulations >2 km<br />

thick under the axis of the centr<strong>al</strong> Snake River Plain. These<br />

thicknesses are not known (base not seen) but in the eastern<br />

Snake River Plain seismic refraction data suggest rhyolitic<br />

rocks are c. 2.5 km thick (Sparlin <strong>et</strong> <strong>al</strong>. 1982) and borehole


Bull Volcanol (<strong>2008</strong>) 70:293–314 309<br />

INEL-1 pen<strong>et</strong>rated 2.39 km of rhyolitic ignimbrite without<br />

reaching its base; this may represent a c<strong>al</strong>dera fill (Doherty<br />

<strong>et</strong> <strong>al</strong>. 1979).<br />

Other possible occurrences of SR-type silicic volcanism<br />

Features of SR-type volcanism <strong>al</strong>so occur elsewhere in the<br />

Proterozoic and Phanerozoic record. Most, though not <strong>al</strong>l,<br />

examples involve large-volume eruptions of intracontinent<strong>al</strong>,<br />

high-temperature and relatively anhydrous, potassic and<br />

high-Fe, m<strong>et</strong><strong>al</strong>uminous rhyolites and dacites (or latites, e.g.<br />

Marsh <strong>et</strong> <strong>al</strong>. 20<strong>01</strong>) in a bimod<strong>al</strong> association with voluminous<br />

bas<strong>al</strong>tic magmatism. In many cases, the physic<strong>al</strong><br />

volcanology of the deposits, particularly of the non-welded<br />

tephras, is poorly understood.<br />

Elsewhere <strong>al</strong>ong the Yellowstone hot-spot track<br />

Styles of volcanism vary <strong>al</strong>ong the Yellowstone hot-spot<br />

track. The variation reflects tempor<strong>al</strong> changes in both<br />

chemistry and magmatic temperatures of the magmas<br />

generated as the hot-spot encountered differences within<br />

the overriding continent<strong>al</strong> lithosphere (Perkins and Nash<br />

2002). Amongst the resultant great diversity of facies<br />

within the province, some successions outside the area<br />

described in this paper (Fig. 1) <strong>al</strong>so exhibit SR-type<br />

characteristics. In the eastern Snake River Plain, sever<strong>al</strong><br />

rhyolitic ignimbrites are high to extremely high-grade,<br />

rheomorphic, loc<strong>al</strong>ly lava-like, and associated with par<strong>al</strong>lel-stratified<br />

ashes and aqueously deposited volcaniclastic<br />

sands silts (e.g. W<strong>al</strong>cott Tuff, Blacktail Creek Tuff, Tuff of<br />

Kilgore, and associated units; Morgan and McIntosh<br />

2005). Other ignimbrites, however, are more transition<strong>al</strong><br />

in character and exhibit some but not <strong>al</strong>l of the<br />

characteristic features of SR-type volcanism (e.g. the of<br />

Wolverine Creek; Morgan and McIntosh 2005). Wellknown<br />

ignimbrites from the Yellowstone area have both<br />

welded and non-welded parts, eutaxitic fabrics, and restricted<br />

rheomorphic, lava-like facies. They contain abundant pumice<br />

clasts and were erupted at lower temperatures than typic<strong>al</strong> SRtype<br />

ignimbrites. Some Quaternary rhyolites in the volcanic<br />

province (Magic Reservoir centre, Leeman 2004; Big Butte,<br />

Cedar Butte and related rhyolite domes further east,<br />

McCurry <strong>et</strong> <strong>al</strong>. 1999) contrast with the Miocene volcanic<br />

rocks described in this paper, and are not of SR-type. The<br />

13.8–12 Ma Owyhee-Humboldt eruptive centre (Fig. 1)<br />

produced widely dispersed ash, intensely rheomorphic<br />

ignimbrites some with lava-like facies, she<strong>et</strong>-like rhyolite<br />

lavas, rhyolite lavas inferred to have been fountain-fed<br />

(clastogenic) and ascribed to eruptions of low-viscosity, hot<br />

and relatively anhydrous rhyolite magma (Ekren <strong>et</strong> <strong>al</strong>. 1982;<br />

Bonnichsen and Kauffman 1987; Manley1995; Manleyand<br />

McIntosh 2002).<br />

Eocene–Oligocene rhyolites of Trans-Pecos Texas<br />

Associations of extensive, low-aspect ratio silicic lavas with<br />

extremely high-grade rheomorphic ignimbrites, including<br />

some lava-like units of equivoc<strong>al</strong> origin occur in Trans-<br />

Pecos Texas (Henry <strong>et</strong> <strong>al</strong>. 1988, 1989; HenryandWolff<br />

1992) and have characteristics of SR-type volcanism. The<br />

38 Ma–32 Ma mafic to silicic volcanic rocks are related to<br />

subduction beneath the North American continent, and<br />

include c<strong>al</strong>dera-related, large volume, predominantly<br />

<strong>al</strong>k<strong>al</strong>ic low-silica rhyolites. Intense welding and rheomorphism<br />

of ignimbrites (e.g. Buckshot ignimbrite,<br />

Gomez Tuff and in the Barrel Springs and Sleeping Lion<br />

formations, Henry <strong>et</strong> <strong>al</strong>. 1989), and the low aspect-ratios<br />

of lavas (Fig. 2), such as the 25–120 m thick Bracks<br />

Rhyolite, which covers 1,000 km 2 and flowed ≤35 km<br />

from source, have been ascribed to eruptive temperatures<br />

≥900°C (Henry <strong>et</strong> <strong>al</strong>. 1990). Other ignimbrites exhibit<br />

fiamme and lithic clasts, and are less typic<strong>al</strong>ly SR-type.<br />

Cr<strong>et</strong>aceous Etendeka-Paraná volcanic province, Namibia,<br />

Angola and South America<br />

Vast silicic she<strong>et</strong>s of the early Cr<strong>et</strong>aceous Etendeka<br />

volcanic provinces of Namibia and Angola, and the<br />

southern Paraná of South America (Marsh <strong>et</strong> <strong>al</strong>. 20<strong>01</strong>;<br />

Garland <strong>et</strong> <strong>al</strong>. 1995; Milner <strong>et</strong> <strong>al</strong>. 1995; Kirstein <strong>et</strong> <strong>al</strong>. 20<strong>01</strong>)<br />

have features of SR-type volcanism. Individu<strong>al</strong> she<strong>et</strong>s in<br />

Namibia are 70–250 m thick, exceed 25,000 km 2 (Milner <strong>et</strong><br />

<strong>al</strong>. 1995; Jerram 2002), and have estimated magmatic<br />

temperatures exceeding 1,000°C. They have been interpr<strong>et</strong>ed<br />

as predominantly lava-like ignimbrites, <strong>al</strong>though few<br />

unequivoc<strong>al</strong> pyroclastic features are preserved and some<br />

have bas<strong>al</strong> breccias (Milner <strong>et</strong> <strong>al</strong>. 1992; Ewart <strong>et</strong> <strong>al</strong>. 1998).<br />

Some units contain non-flattened globules, some larger than<br />

those in the Snake River ignimbrites. Little has been<br />

published on the f<strong>al</strong>lout tephras, and few of the eruptive<br />

centres are well exposed, with the exception of the Messum<br />

Complex (Ewart <strong>et</strong> <strong>al</strong>. 1998).<br />

Middle Proterozoic Keweenawan volcanics, Minnesota<br />

Low-aspect ratio rhyolite lavas and extremely high-grade<br />

rheomorphic ignimbrites of SR-type have been described in<br />

the Middle Proterozoic Keweenawan plateau volcanics of<br />

Minnesota (Green 1989; Green and Fitz 1993). They reach<br />

250 m in thickness and trace for 40 km, with volumes of<br />

100–400 km 3 . Some are lava-like and their pyroclastic<br />

versus effusive origin is equivoc<strong>al</strong>. Less is known about the<br />

non-welded pyroclastic units, <strong>al</strong>though one 20 cm-thick<br />

layer of pumice lapilli is reported beneath a lava. Former<br />

surface water is indicated by thin units of fluvio-deltaic<br />

sediments and bas<strong>al</strong>tic pillow lavas. Temperatures of c.


310 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

900°C have been invoked for the rhyolites, which are<br />

thought to have derived from crust<strong>al</strong> melting during<br />

continent<strong>al</strong> rifting (Vervoort and Green 1997).<br />

Jurassic Karoo volcanic province, southern Africa<br />

Rhyolite she<strong>et</strong>s with some characteristics of SR-type<br />

volcanism occur within the Lebombo monocline of southern<br />

Africa. The rhyolitic eruptions were large, with a<br />

combined volume of 35,000 km 3 (Cleverly <strong>et</strong> <strong>al</strong>. 1984).<br />

Individu<strong>al</strong> rhyolite she<strong>et</strong>s are ≤60 km, ≤200 m thick, and have<br />

been interpr<strong>et</strong>ed as extremely high-grade, largely lava-like<br />

rheomorphic ignimbrites (Cleverly 1979). They have low<br />

δ 18 O v<strong>al</strong>ues (c. 5.6 ‰) ascribed to deep pen<strong>et</strong>ration and<br />

extensive circulation of m<strong>et</strong>eoric water facilitated by brittle<br />

fracturing accompanying continent<strong>al</strong> rifting (Harris and<br />

Erlank 1992).<br />

Middle Proterozoic Yardea Dacite, south Austr<strong>al</strong>ia<br />

The Yardea Dacite (67% SiO2) of south Austr<strong>al</strong>ia is as much<br />

as 250 m thick and covers an area of 12,000 km 2 (Creaser<br />

and White 1991). It is not known wh<strong>et</strong>her or not it is a single<br />

eruptive unit, but it shows mostly lava-like characteristics<br />

and may be either rheomorphic ignimbrite or unusu<strong>al</strong>ly<br />

extensive lava. It is thought to have been erupted from a hot<br />

(>1,000°C) felsic magma that was relatively poor in H2O.<br />

Unlike most SR-type rhyolites, however, it loc<strong>al</strong>ly contains<br />

abundant lithic fragments, and the magmas are slightly more<br />

mafic and cryst<strong>al</strong> rich. Little is published on the volcanology<br />

of associated facies (Creaser and White 1991).<br />

Ordovician Scafell c<strong>al</strong>dera volcano England, UK<br />

Associations of intensely rheomorphic ignimbrites, par<strong>al</strong>lelstratified<br />

ashes and lacustrine volcaniclastic sediments occur<br />

in Ordovician extension<strong>al</strong> continent<strong>al</strong> arc successions of the<br />

English Lake District (Millward 2004; Branney 2006). Some<br />

of the rheomorphic ignimbrites, such as the Bad Step Tuff in<br />

Scafell c<strong>al</strong>dera are of SR-type, with extensive lava-like<br />

facies, rheomorphic folding, she<strong>et</strong>ing joints, zones of<br />

lithophysae zones, and upper autobreccias (Branney <strong>et</strong> <strong>al</strong>.<br />

1992). Interc<strong>al</strong>ated thin layers of stratified tuffs lack pumice<br />

lapilli. Some have ash pell<strong>et</strong>s, trace for long distances, and<br />

resemble ashf<strong>al</strong>l layers of the centr<strong>al</strong> Snake River Plain.<br />

Some are interpr<strong>et</strong>ed as recording very large-volume<br />

phreatoplinian ashf<strong>al</strong>ls that blank<strong>et</strong>ed a subaeri<strong>al</strong> landscape<br />

and, proxim<strong>al</strong>ly, fell into subsiding syn-eruptive lakes<br />

(Branney 1991). The rheomorphic ignimbrite eruptions are<br />

thought to have been characterized by high temperature, high<br />

mass-flux pyroclastic fountaining generating hot density<br />

currents, and the interc<strong>al</strong>ated non-welded stratified units are<br />

thought to record intermittent explosive interaction with<br />

ground and lake water. Differences with SR-type volcanism<br />

include the presence of lava domes and coulees, an<br />

intrac<strong>al</strong>dera s<strong>et</strong>ting, and ignimbrites that contain pumice<br />

and fiamme (Branney and Kokelaar 1995).<br />

Silurian Glencoe c<strong>al</strong>dera volcano, Scotland, UK<br />

The Lower, Middle and Upper Etive Rhyolites of Glencoe<br />

c<strong>al</strong>dera volcano, Scotland, are each 100–150 m thick,<br />

predominantly lava-like, intensely rheomorphic ignimbrites<br />

with contorted flow lamination and upper autobreccias, and<br />

they occur interstratified with fluvi<strong>al</strong> and lacustrine sediments<br />

and thin layers of stratified rhyolitic ash containing<br />

ash aggregates (Kokelaar and Moore 2006). They are<br />

overlain by more convention<strong>al</strong> eutaxitic ignimbrites. They<br />

are high-K rhyolites with 74–77% silica and thought to<br />

represent crust<strong>al</strong> melts erupted in a region<strong>al</strong> volcanic flareup<br />

in a transtension<strong>al</strong> continent<strong>al</strong> s<strong>et</strong>ting.<br />

Most examples of volcanic successions that exhibit SRtype<br />

features record large-volume, high mass-flux eruptions<br />

of hot, relatively anhydrous, high-K, m<strong>et</strong><strong>al</strong>uminous rhyolite,<br />

at continent<strong>al</strong> large igneous provinces (LIPs). With<br />

some exceptions, relatively little is known about their<br />

physic<strong>al</strong> volcanology; for example, the nature and dispers<strong>al</strong><br />

characteristics of non-welded tephras—such studies have<br />

been hindered by the vast sc<strong>al</strong>e of the phenomena. Clearly,<br />

not <strong>al</strong>l silicic volcanism at continent<strong>al</strong> LIPs is of SR-type.<br />

For example, large-volume Oligocene silicic pyroclastic<br />

eruptions in Yemen–Ethiopia bimod<strong>al</strong> LIP produced low- to<br />

moderate-grade welded ignimbrites in which pumice lapilli<br />

and fiamme are abundant (e.g. Peate <strong>et</strong> <strong>al</strong>. 2005).<br />

Conclusion: a summary of the characteristic features<br />

of SR-type volcanism<br />

The characteristic features of SR-type volcanism as<br />

exhibited by the type example, the Miocene rocks of<br />

south-centr<strong>al</strong> Idaho and northernmost Nevada, are listed (1–<br />

18 and Appendix Table 1). (1) A bimod<strong>al</strong> association of<br />

bas<strong>al</strong>ts (‘bas<strong>al</strong>tic plains-style’ volcanism) with H2O-poor,<br />

m<strong>et</strong><strong>al</strong>uminous rhyolites (70–77% SiO2) that have elevated<br />

contents of HFSE and h<strong>al</strong>ogens. (2) Large, region<strong>al</strong>ly<br />

devastating explosive rhyolitic eruptions (VEI 6–8). (3)<br />

Voluminous, high to low aspect-ratio rhyolitic ignimbrites.<br />

(4) Numerous ashf<strong>al</strong>l layers with par<strong>al</strong>lel thin-bedding and<br />

lamination; extensive (to 1,000s of km) atmospheric<br />

dispers<strong>al</strong> of ash. (5) Unusu<strong>al</strong>ly large-volume, extensive<br />

rhyolite lavas with low aspect-ratios. Domes and coulées<br />

are not characteristic. (6) A paucity of pumice lapilli in the<br />

ignimbrites (or, in welded ignimbrites, of fiamme after<br />

pumice lapilli) and an absence of pumice lenses and<br />

pumice-concentration zones. (7) The ignimbrites are gen-


Bull Volcanol (<strong>2008</strong>) 70:293–314 311<br />

er<strong>al</strong>ly lithic-poor and lack proxim<strong>al</strong> lithic breccias. (8)<br />

Sm<strong>al</strong>l angular glassy fragments are abundant both in many<br />

ashf<strong>al</strong>l layers and ignimbrites. (9) The ignimbrites are true<br />

tuffs rather than lapilli-tuffs, and they are unusu<strong>al</strong>ly well<br />

sorted (σ φ


312 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

Branney MJ, Kokelaar BP, McConnell BJ (1992) The Bad Step Tuff:<br />

a lava-like ignimbrite in a c<strong>al</strong>c-<strong>al</strong>k<strong>al</strong>ine pieceme<strong>al</strong> c<strong>al</strong>dera,<br />

English Lake District. Bull Volcanol 54:187–199<br />

Branney MJ, Kokelaar BP (2002) Pyroclastic density currents and the<br />

sedimentation of ignimbrites. Geol Soc Lond, Memoirs 27:1–152<br />

Branney MJ, Barry TL, Godchaux M (2004) Sheathfolds in<br />

rheomorphic ignimbrites. Bull Volcanol 66:485–491<br />

Brown RJ, Branney MJ (2004) Event-stratigraphy of a c<strong>al</strong>deraforming<br />

ignimbrite eruption on Tenerife: the 273 ka Poris<br />

Formation. Bull Volcanol 66:392–416<br />

Bursik MI, Woods AW (1996) The dynamics and thermodynamics of<br />

large ash flows. Bull Volcanol 38:175–193<br />

Carey SN, Sigurdsson H, Sparks RSJ (1988) Experiment<strong>al</strong> studies of<br />

particle-laden plumes. J Geophys Res 93: 15314–15328<br />

Cas RAF, Wright JV (1987) Volcanic successions: modern and<br />

ancient. Alan & Unwin, London, pp 1–528<br />

Cathey HE, Nash BP (2004) The Cougar Point Tuff: implications for<br />

thermochemic<strong>al</strong> zonation and longevity of high-temperature,<br />

large volume silicic magmas of the Miocene Yellowstone<br />

hotspot. J P<strong>et</strong>rol 45:27–58<br />

Chapin CE, Lowell GR (1979) Primary and secondary flow structures<br />

in ash-flow tuffs of the Gribbles Run P<strong>al</strong>aeov<strong>al</strong>ley, centr<strong>al</strong><br />

Colorado. In: Chapin CE, Elston WE (eds) Ash-flow tuffs. Geol<br />

Soc Am Spec Pap 180:137–154<br />

Christiansen EH, McCurry M (2007) Contrasting origins of Cenozoic<br />

silicic volcanic rocks from the western Cordillera of the United<br />

States. Bull Volcanol (this issue)<br />

Cleverly RW (1979) The volcanic geology of the Lebombo monocline<br />

in Swaziland. Tran Geol Soc S Africa 82:227–230<br />

Cleverly RW, B<strong>et</strong>ton PJ, Bristow JW (1984) Geochemistry and<br />

p<strong>et</strong>rogenesis of the Lebombo rhyolites. Geol Soc S Africa Spec<br />

Pub 13:171–195<br />

Creaser RA, White AJ (1991) The Yardea Dacite—large-volume,<br />

high-temperature felsic volcanism from the Middle Proterozoic<br />

of South Austr<strong>al</strong>ia. Geology 19:48–51<br />

Cummings ML, Evans JG, Ferns ML, Lees KR (2000) Stratigraphy<br />

and structur<strong>al</strong> evolution of the middle Miocene synvolcanic<br />

Oregon-Idaho graben. Geol Soc Amer Bull 112:668–682<br />

Doherty DJ, McBroome LA, Kuntz MA (1979) Preliminary geologic<strong>al</strong><br />

interpr<strong>et</strong>ation and lithologic log of the exploratory geotherm<strong>al</strong><br />

test well (INEL-1), Idaho Nation<strong>al</strong> Engineering Laboratory,<br />

eastern Snake River Plain, Idaho. US Geol Surv Open File Rep<br />

79–1248 pp 10<br />

Druitt TH, Sparks RSJ (1982) A proxim<strong>al</strong> ignimbrite breccia facies on<br />

Santorini volcano, Greece. J Volcanol Geotherm Res 13:141–171<br />

Duffield WA (1990) Eruptive fountains of silicic magma and their<br />

possible effects on the tin content of fountain-fed lavas, Taylor<br />

Creek Rhyolite, New Mexico. In: Stein HJ, Hannah JL (eds) Orebearing<br />

granite systems: p<strong>et</strong>rogenesis and miner<strong>al</strong>ising processes.<br />

Geol Soc Am Spec Pap 246:251–261<br />

Ekren EB, McIntyre DH, Benn<strong>et</strong>t EH, Marvin RF (1982) Cenozoic<br />

stratigraphy of western Owyhee County, Idaho. In: Bonnichsen<br />

B, Breckenridge RM (eds) Cenozoic Geology of Idaho. Idaho<br />

Bur Mines Geol Bull 26:215–235<br />

Ekren EB, McIntyre DH, Benn<strong>et</strong>t EH (1984) High-temperature, largevolume,<br />

lava like ash-flow tuffs without c<strong>al</strong>deras in southwestern<br />

Idaho. US Geol Surv Prof Pap 1272<br />

Ewart A, Milner SC, Armstrong RA, Duncan AR (1998) Etendeka<br />

volcanism of the Gobobseb Mountains and Messum Igneous<br />

Complex, Namibia, Parts I and II. J P<strong>et</strong>rol 39:191–253<br />

Freundt A (1999) Formation of high-grade ignimbrites. Part II. A<br />

pyroclastic suspension current model with implications <strong>al</strong>so for<br />

low-grade ignimbrites. Bull Volcanol 60:545–576<br />

Garland F, Hawksworth CJ, Mantovani MSM (1995) Description and<br />

p<strong>et</strong>rogenesis of the Paraná rhyolites, Southern Brazil. J P<strong>et</strong>rol<br />

36:1193–1227<br />

Godchaux MM, Bonnichsen B (2002) Syneruptive magma–water and<br />

posteruptive lava–water interactions in the Western Snake River<br />

Plain, Idaho, during the past 12 million years. In: Bonnichsen B,<br />

White CM, McCurry M (eds) Tectonic and magmatic evolution<br />

of the Snake River Plain Volcanic Province. Idaho Geol Surv<br />

Bull 30:387–434<br />

Greeley R (1977) Bas<strong>al</strong>tic “plains” volcanism. In: Greeley R, King JS<br />

(eds) Volcanism of the eastern Snake River Plain, Idaho: a<br />

comparitive plan<strong>et</strong>ary guidebook. Washington DC, NASA, pp 23–44<br />

Greeley R (1982) The Snake River Plain, Idaho: representative of a<br />

new category of volcanism. J Geophys Res 2705–2712<br />

Green JC (1989) Physic<strong>al</strong> volcanology of mid-Proterozoic plateau<br />

lavas: the Keweenawan North Shore Volcanic Group, Minnesota.<br />

Geol Soc Amer Bull 1<strong>01</strong>:486–500<br />

Green JC, Fitz TJ III (1993) Extensive felsic lavas and rheoignimbrites<br />

in the Keweenawan Midcontinent Rift plateau volcanics,<br />

Minnesota: p<strong>et</strong>rographic and field recognition. J Volcanol Geotherm<br />

Res 54:177–196<br />

Harris C, Erlank AJ (1992) The production of large-volume, low-δ 18 O<br />

rhyolites during the rifting of Africa and Antarctica: the<br />

Lebombo Monocline, southern Africa. Geochem Cosmochim<br />

Acta 56:3561–3570<br />

Hearst J (1999) Deposition<strong>al</strong> environments of the Birch Creek loc<strong>al</strong><br />

fauna (Pliocene: Blancan), Owyhee County, Idaho. In: Akersten<br />

WA, McDon<strong>al</strong>d HG, Meldrum DJ, Flint MET (eds) Papers on the<br />

Vertebrate P<strong>al</strong>eontology of Idaho Honoring John A White, 1:<br />

Idaho Mus Nat Hist Occ Pap 36:56–93<br />

Henry CD, Wolff JA (1992) Distinguishing strongly rheomorphic tuffs<br />

from extensive silicic lavas. Bull Volcanol 54:171–186<br />

Henry CD, Price JG, Rubin JN, Parker DF, Wolff JA, Self S, Franklin<br />

R, Barker DS (1988) Widespread, lava-like silicic volcanic rocks<br />

of Trans-Pecos Texas. Geology 16:509–512<br />

Henry CD, Price JG, Parker DF, Wolff JA (1989) Mid-Tertiary silicic<br />

<strong>al</strong>k<strong>al</strong>ine magmatism of Trans-Pecos Texas: rheomorphic tuffs and<br />

extensive silicic lavas. Excursion 9A: Silicic volcanic rocks in<br />

the Snake River Plain—Yellowstone Plateau province. In:<br />

Chapin CE, Zidek J (eds) Field excursions to volcanic terranes<br />

in the western United States; Volume I, Southern Rocky<br />

Mountains Region. NM Bur Mines Miner<strong>al</strong> Resourc Mem<br />

47:231–274<br />

Henry CD, Price JG, Rubin JN, Laubach SE (1990) Case study of an<br />

extensive silicic lava: the Bracks Rhyolite, Trans-Pecos Texas. J<br />

Volcanol Geotherm Res 43:113–132<br />

Hildebrand RT, Newman KR (1985) Miocene sedimentation in the<br />

Goose Creek basin, south-centr<strong>al</strong> Idaho, north eastern Nevada<br />

and northwestern Utah. In: Flores RM <strong>et</strong> <strong>al</strong>. (eds) Cenozoic<br />

p<strong>al</strong>eogeography of the western US: Rocky Mountains Section,<br />

Soc Econ P<strong>al</strong>eont Miner<strong>al</strong> 55–70<br />

Honjo N, Leeman WP (1987) Origin of hybrid ferrolatite lavas from<br />

Magic Reservoir eruptive centre, Snake River Plain, Idaho.<br />

Contrib Miner<strong>al</strong> P<strong>et</strong>rol 96:163–177<br />

Honjo N, Bonnichsen B, Leeman WP, Stormer Jr JC (1992)<br />

Miner<strong>al</strong>ogy and geothermom<strong>et</strong>ry of high-temperature rhyolites<br />

from the centr<strong>al</strong> and western Snake River Plain. Bull Volcanol<br />

54:220–237<br />

Iddings JP (1889) The rhyolites. In: Geology of the Yellowstone<br />

Nation<strong>al</strong> Park. US Geol Surv Monogr 32(Part II):356–432<br />

Jenks MD, Bonnichsen B (1989) Subaquesous bas<strong>al</strong>t eruptions into<br />

Pliocene Lake Idaho, Snake River Plain, Idaho. In: Chamberlain<br />

VE, Breckenridge RM, Bonnichsen B (eds) Guidebook to the<br />

geology of Northern and Western Idaho and surrounding area.<br />

Idaho Geol Surv Bull 28:17–34<br />

Jerram DA (2002) Volcanology and facies architecture of flood<br />

bas<strong>al</strong>ts. In: Menzies MA, Klemperer SL, Ebinger CJ and Baker<br />

J (eds) Volcanic Rifted Margins. Geol Soc Amer Spec Pap<br />

362:119–132


Bull Volcanol (<strong>2008</strong>) 70:293–314 313<br />

Johnson RW (1968) Volcanic globule rock from Mount Suswa,<br />

Kenya. Geol Soc Amer Bull 79:647–651<br />

Kimmel PG (1982) Stratigraphy, age, and tectonic s<strong>et</strong>ting of the<br />

Miocene–Pliocene lacustrine sediments of the Western Snake<br />

River Plain, Oregon and Idaho. In: Bonnichsen B, Breckenridge<br />

RM (eds) Cenozoic geology of Idaho. Idaho Bur Mines Geol<br />

Bull 26: 559–578<br />

Kirstein LA, Hawksworth CJ, Garland FG (20<strong>01</strong>) Felsic lavas or<br />

rheomorphic ignimbrites: is there a chemic<strong>al</strong> distinction? Contrib<br />

Miner<strong>al</strong> P<strong>et</strong>rol 142:309–322<br />

Kokelaar BP (1982) Fluidization of w<strong>et</strong> sediment during the<br />

emplacement and cooling of various igneous bodies. J Geol<br />

Soc Lond 139:21–33<br />

Kokelaar BP (1986) Magma–water interactions in subaqueous and<br />

emergent bas<strong>al</strong>tic volcanism. Bull Volcanol 48:275–289<br />

Kokelaar BP, Branney MJ (1999) Inside silicic c<strong>al</strong>deras (Snowdon, Scafell<br />

and Glencoe, UK): interactions of c<strong>al</strong>dera development, tectonism<br />

and hydrovolcanism. Field Guide to the IAVCEI Commission on<br />

Explosive Volcanism Field Workshop 7–18 July, 1999, pp 154<br />

Kokelaar BP, Königer A (2000) Marine emplacement of welded<br />

ignimbrite: the Ordovician Pitts Head Tuff, North W<strong>al</strong>es. J Geol<br />

Soc Lond 157:517–536<br />

Kokelaar BP, Moore ID (2006) Classic areas of British geology:<br />

Glencoe c<strong>al</strong>dera volcano, Scotland. Keyworth, Nottingham,<br />

British Geol Surv, pp 127<br />

Leeman, WP (1982a) Rhyolites of the Snake River Plain—Yellowstone<br />

Plateau Province, Idaho and Wyoming: a summary of p<strong>et</strong>rogen<strong>et</strong>ic<br />

models. In: Bonnichsen B, Breckenridge RM (eds) Cenozoic<br />

geology of Idaho. Idaho Bur Mines Geol Bull 26:203–212<br />

Leeman WP (1982b) Geology of the Magic Reservoir area, Snake River<br />

Plain, Idaho. In: Bonnichsen B, Breckenridge RM (eds) Cenozoic<br />

geology of Idaho. Idaho Bur Mines Geol Bull 26:369–376<br />

Leeman WP (2004) Evolution of Snake River Plain (SRP) silicic<br />

magmas—the Magic Reservoir eruptive center. Geol Soc Am.<br />

Abst with Prog 36: 24<br />

Leeman WP, Oldow JS, Hart WK (1992) Lithosphere-sc<strong>al</strong>e thrusting in<br />

the western U.S. Cordillera as constrained by Sr and Nd isotopic<br />

transitions in Neogene volcanic rocks. Geology 20:63–66<br />

Link PK, McDon<strong>al</strong>d HG, Fanning CM, Godfrey AE (2002) D<strong>et</strong>rit<strong>al</strong><br />

zircon evidence for Pleistocene drainage revers<strong>al</strong> at Hagerman<br />

Fossil Beds Nation<strong>al</strong> Monument, centr<strong>al</strong> Snake River Plain,<br />

Idaho. In: Bonnichsen B, White CM, McCurry M (eds) Tectonic<br />

and magmatic evolution of the Snake River Plain Volcanic<br />

Province: Idaho Geol Surv Bull 30:105–119<br />

Lipman PW (1984) Roots of ash-flow c<strong>al</strong>deras in western North<br />

America: windows into the tops of granitic batholiths. J Geophys<br />

Res 89:88<strong>01</strong>–8841<br />

Mahood GA (1984) Pyroclastic rocks and c<strong>al</strong>deras associated with<br />

strongly per<strong>al</strong>k<strong>al</strong>ine rocks. J Geophys Res B89:8540–8552<br />

Manley CR (1995) How voluminous rhyolitic lavas mimic rheomorphic<br />

ignimbrites: eruptive style, emplacement conditions, and<br />

formation of tuff-like textures. Geology 23:349–352<br />

Manley CR (1996) In situ formation of welded tuff-like textures in the<br />

carapace of a voluminous silicic lava flow, Owyhee County, SW<br />

Idaho. Bull Volcanol 57:672–686<br />

Millward D (2004) The Caradoc volcanoes of the English Lake<br />

District. Proc Yorks Geol Soc 55:145–177<br />

M<strong>al</strong>de HE, Powers HA (1962) Upper Cenozoic stratigraphy of western<br />

Snake River Plain, Idaho: Geol Soc Amer Bull 73:1197–1220<br />

Manley CR, McIntosh WC (2002) The Juniper Mountain Volcanic<br />

Center, Owyhee County, Southwestern Idaho: age relations and<br />

physic<strong>al</strong> volcanology. In: Bonnichsen B, White CM, McCurry M<br />

(eds) Tectonic and magmatic evolution of the Snake River Plain<br />

Volcanic Province. Idaho Geol Surv Bull 30:205–227<br />

Marsh JS, Ewart A, Milner SC, Duncan AR, Miller RMcG (20<strong>01</strong>)<br />

The Etendeka Igneous Province: magma types and their<br />

stratigraphic distribution with implications for the evolution<br />

of the Paraná-Etendeka Flood Bas<strong>al</strong>t Province. Bull Volcanol<br />

62:464–486<br />

McCurry M (eds) Tectonic and magmatic evolution of the Snake<br />

River Plain Volcanic Province. Idaho Geol Surv Bull 30:233–312<br />

Morgan LA, McIntosh WC (2005) Timing and development of Heise<br />

volcanic field, Snake River Plain, Idaho, western USA. Geol Soc<br />

Amer Bull 117:288–306<br />

McCurry M, Bonnichsen B, White C, Godchaux MM, Hughes SS<br />

(1997) Bimod<strong>al</strong> bas<strong>al</strong>t-rhyolite magmatism in the centr<strong>al</strong> and<br />

western Snake River Plain, Idaho and Oregon. In: Link PK,<br />

Kow<strong>al</strong>lis BJ (eds) Proterozoic to Recent stratigraphy, tectonics,<br />

and volcanology, Utah, Nevada, southern Idaho, and centr<strong>al</strong><br />

Mexico. Brigham Young Univ Geol Stud 42:381–422<br />

McCurry M, Watkins AM, Parker JL, Wright K, Hughes SS (1996)<br />

Preliminary volcanologic<strong>al</strong> constraints for sources of high-grade,<br />

rheomorphic ignimbrites of the Cassia Mountains, Idaho:<br />

implications for the evolution of the Twin F<strong>al</strong>ls Volcanic Centre.<br />

Northwest Geol 26:81–91<br />

McCurry M, Hack<strong>et</strong>t WR, Hayden K (1999) Cedar Butte and<br />

cogen<strong>et</strong>ic rhyolite domes of the Eastern Snake River Plain. In:<br />

Hughes SS, Thackray GD (eds) Guidebook to the Geology of<br />

Eastern Idaho: Idaho Mus Nat Hist: 169–179<br />

Milner SC, Duncan AR, Ewart A (1992) Quartz latite rheoignimbrite<br />

flows of the Etendeka Formation, north-western Namibia. Bull<br />

Volcanol 54:200–219<br />

Milner SC, Duncan AR, Whittingham AM, Ewart A (1995) Trans-<br />

Atlantic correlation of eruptive sequences and individu<strong>al</strong> silicic<br />

volcanic units within the Paraná-Etendeka igneous province. J<br />

Volcano Geotherm Res 69:137–157<br />

Pichler H (1981) It<strong>al</strong>ienische Vulkan-Gebi<strong>et</strong>e III: Lipari, Vulcano,<br />

Stromboli. Samml Geol Führ 69:233<br />

Peate IU, Baker JA, Al-Kadasi M, Al-Subbary A, Knight KB,<br />

Riisager P, Thirlw<strong>al</strong>l MF, Peate DW, Renne PR, Menzies MA<br />

(2005) Volcanic stratigraphy of large-volume silicic pyroclastic<br />

eruptions during Oligocene Afro-Arabiam flood volcanism in<br />

Yemen. Bull Volcanol 68: 135–156<br />

Perkins ME, Nash BP (2002) Explosive silicic volcanism of the<br />

Yellowstone hotspot: the ash f<strong>al</strong>l tuff record. Geol Soc Amer Bull<br />

114:367–381<br />

Perkins ME, Nash WP, Brown FH, Fleck RJ (1995) F<strong>al</strong>lout tuffs of<br />

Trapper Creek Idaho—a record of Miocene explosive volcanism<br />

in the Snake River Plains volcanic province. Geol Soc Amer Bull<br />

107:1484–1506<br />

Perkins ME, Brown FH, Nash WP, McIntosh W, Williams SK (1998)<br />

Sequence, age, and source of silicic f<strong>al</strong>lout tuffs in middle to late<br />

Miocene basins of the northern Basin and Range province. Geol<br />

Soc Am Bull 110:344–360<br />

Pierce KL, Morgan LA, S<strong>al</strong>tus RW (2002) Yellowstone plume head:<br />

postulated tectonic relations to the Vancouver slab, continent<strong>al</strong><br />

boundaries, and climate. In: Bonnichsen B, White CM, McCurry<br />

M (eds) Tectonic and magmatic evolution of the Snake River<br />

Plain Volcanic Province. Idaho Geol Surv Bull 30:5–53<br />

Quane SL, Russell JK (2005) Ranking welding intensity in pyroclastic<br />

deposits. Bull Volcanol 67:129–143<br />

Ross CS, Smith RL (1961) Ash-flow tuffs, their origin, geologic<strong>al</strong><br />

relations and identification. US Geol Surv Prof Pap 366:1–77<br />

Russell JK, Quane SL (2005) Rheology of welding: inversion of field<br />

constraints. J Volcanol Geotherm Res 142:173–191<br />

Roche O, Druitt OH, Merle O (2000) Experiment<strong>al</strong> study of c<strong>al</strong>dera<br />

formation. J Geophys Res 105:395–416<br />

Rogers DW, Ore HT, Bobo RT, McQuarrie N, Zentner N (2002)<br />

Extension and subsidence of the eastern Snake River Plain,<br />

Idaho. In: Bonnichsen B, White CM, McCurry M (eds) Tectonic<br />

and magmatic evolution of the Snake River Volcanic Province.<br />

Idaho Geol Surv Bull 30:121–155


314 Bull Volcanol (<strong>2008</strong>) 70:293–314<br />

Rose WI, Riley CM, Dartevelle S (2003) Sizes and shapes of 10-Ma<br />

dist<strong>al</strong> f<strong>al</strong>l pyroclasts in the Og<strong>al</strong>l<strong>al</strong>a Group, Nebraska. J Geol<br />

111:115–124<br />

Shervais JW, Kauffman JD, Gillerman VS, Othberg KL, V<strong>et</strong>ter SK,<br />

Hobson VR, Zarn<strong>et</strong>sske M, Cooke F, Matthews SH, Hanan BB<br />

(2005) Bas<strong>al</strong>tic volcanism of the centr<strong>al</strong> and western Snake River<br />

Plain: a guide to field relations b<strong>et</strong>ween Twin F<strong>al</strong>ls and Mountain<br />

Home, Idaho. In: Pederson J, Dehler CM (eds) GSA Field Guide<br />

Interior Western US 6:27–52<br />

Smith KR, Swirydczuk K, Kimmel PG, Wilkinson BH (1982) Fish<br />

biostratigraphy of late Miocene to Pleistocene sediments of the<br />

western Snake River Plain, Idaho. In: Bonnichsen B, Breckenridge<br />

RM (eds) Cenozoic geology of Idaho. Idaho Geol Surv Bull<br />

26:519–541<br />

Smith RT, Houghton BF (1995) Vent migration and changing eruptive<br />

style during the 1800a Taupo eruption: new evidence from the<br />

Hatepe and Rotongaio phreatoplinian ashes. Bull Volcanol<br />

57:432–439<br />

Sparks RSJ, Stasuik MV, Gardeweg M, Swanson DA (1993) Welded<br />

breccias in andesite lavas. J Geol Soc Lond 150:897–902<br />

Sparks RSJ, Self S, working group (2005). Super-eruptions: glob<strong>al</strong><br />

effects and future threats. Rep Geol Soc London Working Group<br />

(2nd ed) Geol Soc, London, UK, p 24<br />

Sparlin MA, Braile LW, Smith RB (1982) Crust<strong>al</strong> structure of the<br />

Eastern Snake River Plain d<strong>et</strong>ermined from ray trace modelling<br />

of seismic refraction data. J Geophys Res 87(B4):2619–2633<br />

Sumner JM, Branney MJ (2002) The emplacement of a remarkable<br />

h<strong>et</strong>erogeneous, chemic<strong>al</strong>ly zoned and loc<strong>al</strong>ly lava-like rheomorphic<br />

ignimbrite: ‘TL’ on Gran Canaria. J Volcanol Geotherm<br />

Res 115:109–138<br />

Swirydczuk K, Larson GP, Smith GR (1982) Volcanic ash beds as<br />

stratigraphic markers in the Glenns Ferry and Ch<strong>al</strong>k Hills<br />

formations from Adrian, Oregon, to Bruneau, Idaho. In:<br />

Bonnichsen B, Breckenridge RM (eds) Cenozoic geology of<br />

Idaho. Idaho Geol Surv Bull 26:543–558<br />

Torres RC, Self S, Punongbayan RS (1995) Attention focuses on Ta<strong>al</strong>:<br />

decade volcano of the Philippines. EOS Trans Am Geophys<br />

Union 76:241–247<br />

Voorhies MR, Thomasson JR (1979) Fossil grass anthoecia within<br />

Miocene Rhinoceras Skel<strong>et</strong>ons: di<strong>et</strong> in an extinct species. Science<br />

206:331–333<br />

Vervoort JD, Green JC (1997) Origin of evolved magmas in the<br />

Midcontinent Rift System, NE Minnesota: Nd isotope evidence<br />

for melting of Archean crust. Can J Earth Sci 34:521–535<br />

W<strong>al</strong>ker GPL (1971) Grain-size characteristics of pyroclastic deposits.<br />

J Geol 79:696–714<br />

W<strong>al</strong>ker GPL (1981) Two phreatoplininan deposits and their waterflushed<br />

origin. J Volcanol Geotherm Res 9:395–407<br />

W<strong>al</strong>ker GPL (1983) Ignimbrite types and ignimbrite problems. J<br />

Volcanol Geotherm Res 17:65–88<br />

W<strong>al</strong>ker GPL (1984) Downsag c<strong>al</strong>deras, ring faults, c<strong>al</strong>dera sizes, and<br />

increment<strong>al</strong> c<strong>al</strong>dera growth. J Geophys Res 89(B10):395–407<br />

W<strong>al</strong>ker GPL, Croasd<strong>al</strong>e R (1972) Characteristic of some bas<strong>al</strong>tic<br />

pyroclasts. Bull Volcanol 35:303–317<br />

Wheeler HE, Cook EF (1954) Structur<strong>al</strong> and stratigraphic significance<br />

of the Snake River capture, Idaho–Oregon. J Geol 62:525–536<br />

Wood SH, Clemens DM (2002) Geologic and tectonic history of the<br />

Western Snake River Plain, Idaho and Oregon. In: Bonnichsen B,<br />

White CM, McCurry M (eds) Tectonic and magmatic evolution<br />

of the Snake River Plain Volcanic Province. Idaho Geol Surv<br />

Bull 30:69–104<br />

Wooden JL, Mueller PA (1988) Pb, Sr, and Nd isotopic compositions<br />

of a suite of Late Archean igneous rocks, eastern Beartooth<br />

Mountains: Implications for crust–mantle evolution. Earth Plan<strong>et</strong><br />

Sci L<strong>et</strong>t 87:59–72<br />

Wright KE, McCurry M, Hughes SS (2002) P<strong>et</strong>rology and geochemistry<br />

of the Miocen<strong>et</strong>uff of McMullen Creek, centr<strong>al</strong> Snake River<br />

Plain. In: Bonnichsen B, McCurry M, White CM (eds) Tectonic<br />

and Magmatic Evolution of the Snake River Plain Province.<br />

Idaho Geol Surv Bull 30:177–194

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!