23.07.2013 Views

Assembling an Ignimbrite: Mechanical and Thermal Building Blocks ...

Assembling an Ignimbrite: Mechanical and Thermal Building Blocks ...

Assembling an Ignimbrite: Mechanical and Thermal Building Blocks ...

SHOW MORE
SHOW LESS

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

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

<strong>Assembling</strong> <strong>an</strong> <strong>Ignimbrite</strong>: Mech<strong>an</strong>ical <strong>an</strong>d <strong>Thermal</strong><br />

<strong>Building</strong> <strong>Blocks</strong> in the Bishop Tuff, California<br />

Colin J. N. Wilson <strong>an</strong>d Wes Hildreth 1<br />

Institute of Geological <strong>an</strong>d Nuclear Sciences, P.O. Box 30368, Lower Hutt 6315, New Zeal<strong>an</strong>d<br />

(e-mail: c.wilson@gns.cri.nz)<br />

ABSTRACT<br />

The building blocks that control ignimbrite physical characteristics are here defined as mech<strong>an</strong>ical (flow units <strong>an</strong>d<br />

flow packages) <strong>an</strong>d thermal (cooling units). <strong>Ignimbrite</strong> construction from these mech<strong>an</strong>ical <strong>an</strong>d thermal building<br />

blocks is illustrated by a longitudinal profile of the Bishop Tuff in the Owens River Gorge, California. In the Bishop<br />

ignimbrite, flow unit boundaries c<strong>an</strong> be defined only locally, but known or inferred multiple flow units combine to<br />

form eruptive packages showing imbricate, offlapping relationships such that each package is thickest successively<br />

farther from the source. In these packages, the presence or absence of flow unit boundaries does not necessarily reflect<br />

directly the presence or absence, respectively, of time breaks; thus massive ignimbrite need not be the product of<br />

continuous accumulation (progressive aggradation). The intrinsic thermal properties of each package (or parts thereof)<br />

have combined with local thicknesses of accumulated material to control the welding state, measured here by bulk<br />

densities. Four zones of maximum density/welding occur, three of which display imbricate, offlapping relationships<br />

away from the source, in concert with ch<strong>an</strong>ges in package thicknesses. The ignimbrite is thus a single compound<br />

cooling unit in existing terminology, but minima in the density/welding profiles are not at chronostratigraphically<br />

signific<strong>an</strong>t horizons. Thus thermal descriptors, such as simple <strong>an</strong>d compound cooling units, may not have timestratigraphic<br />

signific<strong>an</strong>ce, <strong>an</strong>d use of ignimbrite density/welding profiles to infer emplacement temperatures <strong>an</strong>d<br />

timings is problematic. Development of the Bishop welding zones is, however, explicable by the thicknesses, emplacement<br />

temperatures, <strong>an</strong>d imbricate distribution of the different ignimbrite packages.<br />

Introduction<br />

<strong>Ignimbrite</strong>s (ash-flow tuffs) are <strong>an</strong> abund<strong>an</strong>t <strong>an</strong>d voluminous<br />

product of explosive volc<strong>an</strong>ism, usually<br />

with evolved magma compositions, <strong>an</strong>d generated<br />

by emplacement of concentrated pyroclastic density<br />

currents (pyroclastic flows, sensu stricto). <strong>Ignimbrite</strong>s<br />

may attain volumes of 110 3 km 3 <strong>an</strong>d c<strong>an</strong><br />

show a bewildering variety of morphologies, depositional<br />

facies, <strong>an</strong>d lithological characteristics<br />

(Smith 1960a, 1960b; Ross <strong>an</strong>d Smith 1961; Walker<br />

1983; Wilson 1986; Carey 1991; Druitt 1998;<br />

Freundt et al. 2000). These physical characteristics<br />

must be interpreted to gain insights into the timing<br />

<strong>an</strong>d dynamics of ignimbrite emplacement, interrelationships<br />

with caldera collapse processes, <strong>an</strong>d<br />

the ordering of evacuation from the parental<br />

magma chamber. In this article, we consider two<br />

elements of these physical characteristics that we<br />

M<strong>an</strong>uscript received July 25, 2002; accepted May 8, 2003.<br />

1 U.S. Geological Survey, Volc<strong>an</strong>o Hazards Team, Mailstop<br />

910, 345 Middlefield Road, Menlo Park, California 94025, U.S.A.<br />

[The Journal of Geology, 2003, volume 111, p. 653–670] 2003 by The University of Chicago. All rights reserved. 0022-1376/2003/11106-0003$15.00<br />

653<br />

label the “building blocks” of <strong>an</strong> ignimbrite. The<br />

first is the mech<strong>an</strong>ical structure of the deposit, that<br />

is, how it was constructed by the individual pyroclastic<br />

flows <strong>an</strong>d how this is reflected in the overall<br />

geometry of <strong>an</strong>y larger-scale emplacement packages<br />

<strong>an</strong>d the deposit as a whole. The second is the<br />

thermal structure of the deposit, that is, how the<br />

retained heat of the bulk material varied within the<br />

ignimbrite <strong>an</strong>d subsequently influenced its properties.<br />

In m<strong>an</strong>y (most?) ignimbrites, deciphering the<br />

mech<strong>an</strong>ical structure is hampered by a lack of obvious<br />

internal marker horizons <strong>an</strong>d/or by the effects<br />

of processes that reflect the thermal structure<br />

of the deposit (Smith 1960b; Ross <strong>an</strong>d Smith 1961).<br />

If methods c<strong>an</strong> be found to break down ignimbrites<br />

into their component parts, then qu<strong>an</strong>titative reconstruction<br />

of the parental eruptive pulses would<br />

be possible for a much wider variety of prehistoric<br />

ignimbrites. Here we use part of the 0.76 Ma Bishop<br />

Tuff to show how <strong>an</strong> archetypical partly welded


654 C. J. N. WILSON AND W. HILDRETH<br />

ignimbrite was physically constructed <strong>an</strong>d use the<br />

implications of the field data to address broader<br />

concepts in usage of the terms flow unit <strong>an</strong>d cooling<br />

unit.<br />

Mech<strong>an</strong>ical <strong>Building</strong> <strong>Blocks</strong>: Flow Units. The term<br />

“flow unit” was originally introduced to describe<br />

the body of material generated by the passage of a<br />

single pyroclastic flow (Smith 1960a; Sparks et al.<br />

1973). Flow units were envisaged to be separated<br />

from each other by some kind of boundary, indicative<br />

of a time break separating the individual<br />

flows, <strong>an</strong>d to be typically decimeters to meters<br />

thick. Subsequent work has modified the concept<br />

of flow units in three ways. (1) It was recognized<br />

that different depositional facies or layers could be<br />

associated with the passage of a single pyroclastic<br />

flow (e.g., Sparks et al. 1973; Fisher 1979; Wilson<br />

<strong>an</strong>d Walker 1982; Wilson 1985; Freundt <strong>an</strong>d<br />

Schmincke 1986). (2) It was recognized that ignimbrites<br />

could be divided into simple (one flow unit)<br />

<strong>an</strong>d compound (multiple flow unit) varieties, <strong>an</strong>alogous<br />

to divisions long recognized in lava flows<br />

(Wright 1981; Walker 1983, cf. Walker 1970). (Note,<br />

however, that Smith [1960a] proposed that multiple<br />

flow units could also be generated from emplacement<br />

of a single pyroclastic flow by the superposition<br />

of bodies of material that had taken alternative<br />

travel routes; such flow units also could have developed<br />

within them some degree of depositional<br />

layering.) (3) A more controversial modification has<br />

arisen over the interpretation of ignimbrite sections<br />

where no flow unit boundaries are visible; are they<br />

the product of single flows or of layer-by-layer deposition<br />

(progressive aggradation: Fisher 1966; Br<strong>an</strong>ney<br />

<strong>an</strong>d Kokelaar 1992)? Furthermore, if flow unit<br />

boundaries are taken as evidence for time breaks,<br />

c<strong>an</strong> the absence of such boundaries be used to infer<br />

that deposition of ignimbrite was continuous (Br<strong>an</strong>ney<br />

<strong>an</strong>d Kokelaar 1992), or could deposition have<br />

been punctuated yet leave no clear boundaries?<br />

<strong>Thermal</strong> <strong>Building</strong> <strong>Blocks</strong>: Cooling Units. A characteristic<br />

feature of m<strong>an</strong>y ignimbrites is that they<br />

are emplaced at residual temperatures such that the<br />

juvenile components (pumice <strong>an</strong>d shards) have<br />

welded together (Marshall 1935; Gilbert 1938;<br />

Smith 1960b; Ross <strong>an</strong>d Smith 1961). For a given<br />

juvenile composition, primary controls on the<br />

welding process are thought to be temperature on<br />

deposition <strong>an</strong>d the load stress (Smith 1960b; Riehle<br />

1973), with a possibly signific<strong>an</strong>t control from <strong>an</strong>y<br />

residual volatiles in the glass phase that would reduce<br />

the glass viscosity (e.g., Schmincke 1974). The<br />

welding may r<strong>an</strong>ge from weak cohesion (sintering)<br />

caused by tacking across clast contacts, through the<br />

complete elimination of pore space with accom-<br />

p<strong>an</strong>ying flattening of pumices, <strong>an</strong>d occasionally to<br />

the rheomorphic state where the tuff moves farther<br />

as a viscous mass akin to a lava flow. Because of<br />

the nature of our data, we discuss only nonrheomorphic<br />

welded tuffs because several lines of evidence<br />

(e.g., Chapin <strong>an</strong>d Lowell 1979; Br<strong>an</strong>ney <strong>an</strong>d<br />

Kokelaar 1992; Leat <strong>an</strong>d Schmincke 1993; Freundt<br />

1999) indicate that the onset of signific<strong>an</strong>t welding<br />

during emplacement will lead to further complexities<br />

in the depositional structure of the tuff that<br />

are beyond the scope of this article. However, ignimbrites<br />

similar to the Bishop Tuff, where welding<br />

c<strong>an</strong> be inferred to have largely or entirely occurred<br />

after emplacement, with in situ cooling rates <strong>an</strong>d<br />

lithostatic loads as domin<strong>an</strong>t controls (Br<strong>an</strong>ney <strong>an</strong>d<br />

Kokelaar 1992), are abund<strong>an</strong>t in the geological record<br />

<strong>an</strong>d are amenable to the kind of documentation<br />

employed here.<br />

Welding is often accomp<strong>an</strong>ied <strong>an</strong>d overprinted by<br />

devitrification, recrystallization, <strong>an</strong>d alteration by<br />

a residual vapor phase exsolved from the glass<br />

phase. These processes tend to obscure <strong>an</strong>y mech<strong>an</strong>ical<br />

boundaries <strong>an</strong>d make recognition of flow<br />

units more difficult. However, the overall thermal<br />

structure of <strong>an</strong> ignimbrite, as reflected in variations<br />

of welding state (closely matched, <strong>an</strong>d most conveniently<br />

measured, by ch<strong>an</strong>ges in bulk density)<br />

<strong>an</strong>d <strong>an</strong>y associated crystallization/alteration c<strong>an</strong> be<br />

described quasi-independently of the flow-unit<br />

structure. Smith (1960a, 1960b) thus introduced<br />

the term “cooling unit” for a body of material (one<br />

or more flow units) that was emplaced such that it<br />

had undergone continuous cooling, that is, that the<br />

earliest material was still hot when the latest material<br />

was emplaced. A “simple cooling unit” was<br />

defined by Smith as a body of material that had<br />

accumulated sufficiently rapidly that it cooled as<br />

a single body <strong>an</strong>d had correspondingly simple, idealized<br />

variations of density/welding, with or without<br />

overprinting by devitrification or vapor-phase<br />

crystallization, from base to top. In a simple cooling<br />

unit, the basal <strong>an</strong>d top parts are least welded, the<br />

welding maximum lies within the lower half of the<br />

deposit, <strong>an</strong>d there are no irregularities in welding<br />

intensity with height. A “compound cooling unit”<br />

was defined as <strong>an</strong>y deposit that showed departure<br />

from the hypothetical simple cooling unit profile<br />

qualitatively illustrated by Smith (1960b) <strong>an</strong>d qu<strong>an</strong>titatively<br />

modeled by Riehle (1973). However, despite<br />

such departures, it was taken that the lower<br />

parts of a compound cooling unit had not cooled<br />

entirely before the upper parts were emplaced <strong>an</strong>d<br />

the term “single cooling unit” could be applied.<br />

The term “composite sheet” was introduced by<br />

Smith (1960b) for those deposits in which there was


Journal of Geology ASSEMBLING AN IGNIMBRITE 655<br />

somewhere evidence of signific<strong>an</strong>t time breaks<br />

(e.g., erosion breaks, sediments, lava flows) between<br />

zones of maximum welding intensity, such<br />

that earlier parts of the deposit had effectively<br />

cooled completely before later parts were emplaced,<br />

but that such separate cooling units graded<br />

laterally into single, simple, or compound cooling<br />

units elsewhere. Because the characteristic timings<br />

of flow emplacement are observed or inferred to be<br />

signific<strong>an</strong>tly more rapid th<strong>an</strong> the characteristic<br />

cooling times that would control welding in the<br />

resulting deposits (Riehle 1973), cooling units,<br />

whether simple or compound, are characteristically<br />

developed over greater vertical length scales th<strong>an</strong><br />

flow units, typically meters to hundreds of meters<br />

thick.<br />

Background to Data Sets. In this article, we consider<br />

the structure of <strong>an</strong> archetypical partly welded<br />

ignimbrite, forming part of the Bishop Tuff, in<br />

terms of the fundamental building blocks that created<br />

the deposit as now seen. We use data from the<br />

particularly well-exposed cross sections in the<br />

Owens River Gorge in the eastern part of the Bishop<br />

Tuff (fig. 1). These illustrate how the mech<strong>an</strong>ical<br />

structure of the ignimbrite, coupled with consideration<br />

of its thermal state on emplacement, leads<br />

to complex patterns of rock properties. Detailed<br />

measurements at 10 sections along the gorge walls<br />

(fig. 1) were supplemented at other points by measurements<br />

of lithic sizes, abund<strong>an</strong>ces <strong>an</strong>d lithologies,<br />

<strong>an</strong>d observations of the welding state of the<br />

rocks that were subsequently calibrated to the density<br />

measurements (see below). Although, as mentioned<br />

previously, the thermal state of ignimbrites<br />

on emplacement is expressed by various states of<br />

welding compaction, devitrification, <strong>an</strong>d vaporphase<br />

alteration <strong>an</strong>d recrystallization, we adopt the<br />

view implicit in other studies (Rag<strong>an</strong> <strong>an</strong>d Sherid<strong>an</strong><br />

1972; Riehle 1973) that, in <strong>an</strong>y given deposit, the<br />

degree of welding compaction is the single most<br />

import<strong>an</strong>t qu<strong>an</strong>tifiable parameter that reflects the<br />

emplacement temperature. In turn, we have<br />

adopted bulk density as the simplest way to qu<strong>an</strong>tify<br />

the welding state because of the ease with<br />

which measurements c<strong>an</strong> be made (cf. flattening<br />

ratios of fiamme: Rag<strong>an</strong> <strong>an</strong>d Sherid<strong>an</strong> 1972; Peterson<br />

1979). However, we acknowledge that our ap-<br />

Figure 1. a, Location map showing the eastern part of the Bishop ignimbrite <strong>an</strong>d the positions of the 10 detailed<br />

sections along the Owens Gorge. b, Longitudinal section projected on the true left b<strong>an</strong>k, down the line of the Owens<br />

Gorge, showing the topographic relief of the valley <strong>an</strong>d the gorge walls, main ignimbrite packages (from Wilson <strong>an</strong>d<br />

Hildreth 1997; see also fig. 5), <strong>an</strong>d locations of the density sections. Note that the presence of Ig2Ec on the gorge<br />

rim between sections E <strong>an</strong>d F shows that the thinning over the basement high of older rocks of Ig2Ea <strong>an</strong>d Ig2Eb is<br />

a primary feature <strong>an</strong>d not due to subsequent erosion.


656 C. J. N. WILSON AND W. HILDRETH<br />

Table 1. Summary Stratigraphy of the Bishop Tuff in the Area East of Long Valley Caldera Discussed in This Article<br />

<strong>Ignimbrite</strong> unit Fall unit(s) Notes (ignimbrite)<br />

Ig2Ec F9 185% rhyolite in lithic component<br />

Ig2Eb F9 30%–70%<br />

Ig2Ea Later F8–F9 !20%<br />

Onset of Glass Mountain rhyolite in lithic fraction; incoming of pyroxene-bearing pumice<br />

Ig1Eb F6–earlier F8<br />

Fall horizon (part F6)<br />

Lithic poorer<br />

Ig1Ea Later F2–F6<br />

Fall horizon (F1–earlier F2)<br />

Lithic richer<br />

Note. After Wilson <strong>an</strong>d Hildreth (1997).<br />

proach will become inaccurate for deposits in<br />

which postemplacement diagenetic cementation or<br />

leaching has occurred, <strong>an</strong>d fiamme measurements<br />

may then be the only route to mapping variations<br />

in welding intensity (Roberts <strong>an</strong>d Sidd<strong>an</strong>s 1971; Peterson<br />

1979).<br />

Bishop Tuff Stratigraphic Background<br />

The 0.76 Ma Bishop Tuff consists of two components:<br />

a Plini<strong>an</strong> pumice fall deposit that occurs below,<br />

<strong>an</strong>d interbedded within, a nonwelded to<br />

densely welded ignimbrite (Wilson <strong>an</strong>d Hildreth<br />

1997; table 1). The fall deposit is exposed east of<br />

the source <strong>an</strong>d is divided into nine units (F1–F9)<br />

correlated by bedding <strong>an</strong>d grain-size characteristics<br />

<strong>an</strong>d lithic components (Wilson <strong>an</strong>d Hildreth 1997).<br />

The ignimbrite is well exposed along the walls of<br />

the Owens River valley, the Owens Gorge, <strong>an</strong>d the<br />

Chalk Bluffs escarpment (fig. 1), <strong>an</strong>d material r<strong>an</strong>ging<br />

from nonwelded through densely welded is<br />

widely exposed (Sherid<strong>an</strong> 1970; Rag<strong>an</strong> <strong>an</strong>d Sherid<strong>an</strong><br />

1972; Sherid<strong>an</strong> <strong>an</strong>d Rag<strong>an</strong> 1976).<br />

Our work (Wilson <strong>an</strong>d Hildreth 1997) has developed<br />

ways of subdividing the Bishop ignimbrite<br />

into stratigraphically m<strong>an</strong>ageable units. Clearly defined<br />

flow unit boundaries are scarce in the ignimbrite<br />

in general, <strong>an</strong>d so we introduced the concept<br />

of the eruptive package to encompass ignimbrite<br />

with broadly similar components <strong>an</strong>d lithological<br />

characteristics (but independent of welding zonation),<br />

that was emplaced as multiple pulses or flows<br />

over a definable stratigraphic interval (table 1). In<br />

the Owens Gorge <strong>an</strong>d other areas of the eastern<br />

Bishop Tuff, flow unit boundaries occur in three<br />

main circumst<strong>an</strong>ces: (1) in package Ig1Ea, where<br />

they are frequently demarcated by thin (centimeterto<br />

decimeter-thick) discontinuous-fall or hybridfall<br />

intercalations (fig. 2) or by truncations of degassing<br />

pipes (fig. 3); (2) in distal parts of Ig1Eb <strong>an</strong>d<br />

Ig2, where they are demarcated by the tops of concentration<br />

zones of coarser pumice (fig. 4) or occur<br />

as thin individual flow units intercalated within<br />

fall deposits (Wilson <strong>an</strong>d Hildreth 1997, their fig.<br />

9b); <strong>an</strong>d (3) where continuous decimeter-thick to<br />

meter-thick intercalations of fall material occur,<br />

defining the boundaries between packages Ig1Ea<br />

<strong>an</strong>d Ig1Eb, <strong>an</strong>d Ig1Eb <strong>an</strong>d Ig2Ea (Wilson <strong>an</strong>d Hildreth<br />

1997, their figs. 5, 9, <strong>an</strong>d 10).<br />

In the deposits discussed here, east of the caldera,<br />

earlier pyroxene-free ignimbrite (packages Ig1Ea,<br />

Ig1Eb) was emplaced synchronously with fall units<br />

F2 through all but uppermost F8, <strong>an</strong>d later pyroxenebearing<br />

ignimbrite (Ig2Ea–c) was emplaced synchronously<br />

with uppermost F8 <strong>an</strong>d F9 (table 1; Wilson<br />

<strong>an</strong>d Hildreth 1997). Some minor diachroneity<br />

is present in the Ig1E–Ig2E contact as the onset of<br />

Glass Mountain–derived rhyolite lithics occurs<br />

slightly below the time break at the F8–F9 contact.<br />

Where Ig1E <strong>an</strong>d Ig2E are in direct contact with no<br />

intervening fall material, the boundary between the<br />

two is placed at the onset of the Glass Mountain<br />

rhyolite lithics. The recognition that fall <strong>an</strong>d flow<br />

deposits were coevally emplaced allowed us to infer<br />

overall timings for emplacement of two packages;<br />

ca. 25 h for Ig1Ea <strong>an</strong>d ca. 36 h for Ig1Eb, <strong>an</strong>d hence<br />

to relate average accumulation rates to the presence<br />

or otherwise of flow unit boundaries that would<br />

indicate episodicity in deposition. Variations of<br />

welding in the tuff are superficially straightforward,<br />

with one zone of dense welding being present in<br />

the upper to middle reaches of the Owens Gorge,<br />

two zones in the middle to lower reaches, then a<br />

single zone along the Chalk Bluffs escarpment. Because<br />

of these variations, the Bishop was previously<br />

regarded as containing two cooling units (Sherid<strong>an</strong><br />

1968), which led Hildreth (1979) inappropriately to<br />

term it a composite sheet by the criteria of Smith<br />

(1960b).<br />

Sampling Strategies <strong>an</strong>d Methods<br />

To qu<strong>an</strong>tify the mech<strong>an</strong>ical structure of the ignimbrite,<br />

observations were made of <strong>an</strong>y intercalated


Journal of Geology ASSEMBLING AN IGNIMBRITE 657<br />

Figure 2. a, Photograph of nonwelded to sintered Ig1Ea<br />

at 546607 (grid reference to 100 m in the universal tr<strong>an</strong>sverse<br />

Mercator [UTM] metric grid) to show fall intercalations<br />

(arrowed) in Ig1Ea. Height of cliff is ∼27 m. b,<br />

Close-up view of fall-deposit intercalation at 546607.<br />

fall beds, flow unit boundaries, or stratification that<br />

could be interpreted to represent some kind of hiatus<br />

in deposition. With two exceptions (table 1),<br />

the intercalated fall beds are only centimeters to<br />

decimeters thick, are discontinuous, <strong>an</strong>d c<strong>an</strong>not<br />

consistently be used to regionally subdivide the ignimbrite<br />

packages. In the ignimbrite, measurements<br />

were also made of the maximum lengths of<br />

lithic <strong>an</strong>d pumice clasts or fiamme, <strong>an</strong>d the five<br />

largest values were averaged. The abund<strong>an</strong>ce of<br />

lithics was measured by counting the number of<br />

lithic fragments 15 mm across per square meter,<br />

<strong>an</strong>d their lithologic proportions tallied by counting<br />

rock types (Wilson <strong>an</strong>d Hildreth 1997).<br />

To qu<strong>an</strong>tify the thermal structure of the ignimbrite,<br />

observations of the welding state were made<br />

at m<strong>an</strong>y localities, using the degree of induration,<br />

presence or absence of a eutaxitic texture, <strong>an</strong>d the<br />

relative hardness of the rock. These observations<br />

were then calibrated against bulk density measurements<br />

(table 2) obtained from multiple sets of<br />

samples taken through the 10 detailed sections (A–<br />

J) along the Owens Gorge (fig. 1) <strong>an</strong>d were used to<br />

generate a 2-D profile of welding intensity along<br />

the gorge wall.<br />

Density measurements were made on samples<br />

collected from representative blocks or cored from<br />

typical matrix material, avoiding <strong>an</strong>y obvious local<br />

Figure 3. a, Photograph of flow-unit boundary (arrowed)<br />

in Ig1Ea at 521609, locally demarcated (b)bytruncated<br />

degassing segregation pipes.


658 C. J. N. WILSON AND W. HILDRETH<br />

Figure 4. Photographs of flow units in (a) Ig1Eb in the Owens Gorge at 632466, defined by the tops of swarms of<br />

fiamme marking former pumice concentration zones, <strong>an</strong>d in (b) Ig2Ea at 790475, defined by pumice swarms.<br />

concentrations of lithic or pumice clasts that might<br />

skew the results. Lithic proportions in grain-size<br />

samples of representative ignimbrite from the<br />

Owens Gorge area are all !10 wt%, mostly !5 wt%,<br />

<strong>an</strong>d so the effects of lithics on the sample density<br />

have been neglected. Densities of coherent (sintered<br />

to densely welded) ignimbrite were determined<br />

on representative 300–1800-g blocks, using<br />

water displacement techniques (Archimedes’ principle)<br />

to determine block volumes, with specimens<br />

previously being saturated in water to eliminate<br />

ingress of water during measurement. Ten determinations<br />

were made on each sample, <strong>an</strong>d the resulting<br />

st<strong>an</strong>dard deviations on the density determinations<br />

r<strong>an</strong>ged between 0.001 <strong>an</strong>d 0.010 Mg/m 3 .<br />

Densities of nonwelded (sievable) ignimbrite were<br />

determined from single samples cored in the field<br />

using a carpenter’s brace <strong>an</strong>d bit to excavate a cylindrical<br />

hole of measurable volume (137–296 cm 3 ).<br />

The excavated material was caught in a bag for subsequent<br />

drying <strong>an</strong>d weighing to determine its density.<br />

At one point in section J, block <strong>an</strong>d cored samples<br />

could be obtained from the same level<br />

(everywhere else, the ignimbrite was either too indurated<br />

to excavate with the brace <strong>an</strong>d bit or too<br />

soft to be taken as a coherent block); density values<br />

are 1.262 0.002 (1 SD) Mg/m 3 <strong>an</strong>d 1.240 Mg/m 3<br />

for block <strong>an</strong>d core, respectively. The raw data set<br />

is available from the The Journal of Geology’s Data<br />

Depository free of charge upon request or as <strong>an</strong> Excel<br />

spreadsheet from C. J. N. Wilson.<br />

Mech<strong>an</strong>ical Structure of the Bishop <strong>Ignimbrite</strong><br />

Of the various parameters measured in the ignimbrite,<br />

variations in lithic sizes, numerical abund<strong>an</strong>ces,<br />

<strong>an</strong>d lithologies show up the mech<strong>an</strong>ical<br />

structure best (fig. 5). Coupled with observations of<br />

intercalations of the main fall deposit between<br />

Ig1Ea <strong>an</strong>d Ig1Eb <strong>an</strong>d between Ig1 <strong>an</strong>d Ig 2, these<br />

data show the large-scale cross-sectional structure<br />

of the ignimbrite down the line of the Owens River<br />

(i.e., approximately parallel to the inferred paleoflow<br />

direction) to be a series of offlapping or imbricate<br />

lenses. Internal subdivision of Ig1Eb is discernible<br />

on the basis of lithic counts (fig. 5),<br />

suggesting that the imbricate structure is also paralleled<br />

by growth of the individual packages. Each<br />

main package in turn reaches farther from the<br />

source, with the exception of Ig2Ec, which is poorly<br />

developed in this area (Wilson <strong>an</strong>d Hildreth 1997,<br />

their fig. 12). In a section at 830529 (grid reference<br />

to 100 m in the universal tr<strong>an</strong>sverse Mercator<br />

[UTM] metric grid) on the slopes of the White


Journal of Geology ASSEMBLING AN IGNIMBRITE 659<br />

Table 2. Summary of Field Descriptive Terms Used for Welding Textures in the Bishop <strong>Ignimbrite</strong><br />

Welding term Nature of deposit<br />

Density<br />

(Mg/m3 )<br />

Densely welded Strong eutaxitic texture, dark color if glassy 1ca. 2.00<br />

Moderately welded Clear eutaxitic texture, still relatively soft 1.74–2.00<br />

Poorly welded Some pumice flattening, highly porous <strong>an</strong>d soft 1.49–1.81<br />

Sintered Coherent; requires hammer to fracture; no eutaxitic texture 1.22–1.57<br />

Nonwelded Noncoherent; c<strong>an</strong> be disaggregated between fingers 1.09–1.47<br />

Note. Correlated with density measurements reported in this article.<br />

Mountains east of the Owens Gorge area, a single<br />

centimeter-scale “swash”-like bed of Ig2Eb flow<br />

material occurs within thick F9 fall deposits, at a<br />

stratigraphic level that implies that package Ig2Eb<br />

traveled the farthest of all to the east of the source.<br />

The causes of this offlapping structure in the ignimbrite<br />

c<strong>an</strong>not be related simply to increasing velocities<br />

of the parental flows, since lithic sizes (<strong>an</strong>d<br />

hence inferred velocities <strong>an</strong>d flow-carrying capacities)<br />

in Ig2E are smaller th<strong>an</strong> in Ig1Eb (fig. 5). Instead,<br />

two other factors were probably import<strong>an</strong>t.<br />

First, earlier flows of Ig1Ea <strong>an</strong>d early Ig1Eb would<br />

have ponded in nearer-source low-lying areas,<br />

whereas later flows could travel over a smoother<br />

surface of subdued topographic relief. Second, the<br />

earlier flows are poorer in fine ash-grade material<br />

(fig. 6), regardless of dist<strong>an</strong>ce from the source, <strong>an</strong>d<br />

are thus likely to have deaerated more rapidly <strong>an</strong>d<br />

consequently to have been less mobile (Roche et<br />

al. 2002).<br />

Ch<strong>an</strong>ges in the parameters illustrated in figure 5<br />

serve to highlight two points. First, within Ig1Ea<br />

<strong>an</strong>d Ig1Eb, there do not appear to be major hiatuses<br />

between material separated by flow-unit boundaries<br />

(apart from ch<strong>an</strong>ges in pumice or fiamme abund<strong>an</strong>ce<br />

that define the boundary itself, e.g., fig. 4) or<br />

by fall-deposit intercalations. This implies that the<br />

overall ch<strong>an</strong>ges in lithic sizes <strong>an</strong>d abund<strong>an</strong>ces in<br />

Ig1E were controlled by gradual variations in the<br />

source area, such that flows separated by minutes<br />

to hours share broadly similar characteristics. Second,<br />

at the contact between Ig1E <strong>an</strong>d either Ig2Ea<br />

or Ig2Eb, there are locally abrupt ch<strong>an</strong>ges in measured<br />

parameters (especially the count percentage<br />

of rhyolite in the lithic fraction, Wilson <strong>an</strong>d Hildreth<br />

1997) over dist<strong>an</strong>ces of !0.5 m, such as at<br />

section G (figs. 1, 5), but with no sign of <strong>an</strong>y discontinuity<br />

to reflect <strong>an</strong>y time break or hiatus in<br />

ignimbrite deposition. This is so even though the<br />

presence farther down-valley of F9 fall deposits<br />

along this contact (Wilson <strong>an</strong>d Hildreth 1997, their<br />

fig. 10) demonstrates unequivocally the presence of<br />

a signific<strong>an</strong>t time break. The possibility that ignimbrite<br />

deposition was continuous at section G<br />

but discontinuous down-valley is precluded by the<br />

presence of Ig2Ea down-valley between fall F9 <strong>an</strong>d<br />

Ig2Eb (fig. 5); both F9 <strong>an</strong>d Ig2Ea are absent at section<br />

G. Thus, in the material exposed in the Owens<br />

Gorge, some parts of the Bishop ignimbrite show<br />

that a demonstrable time break may not be accomp<strong>an</strong>ied<br />

by signific<strong>an</strong>t ch<strong>an</strong>ges in the properties of<br />

the ignimbrite (first point above), whereas other<br />

parts show that clearly defined ch<strong>an</strong>ges in ignimbrite<br />

properties may lack <strong>an</strong>y evidence for a signific<strong>an</strong>t<br />

time break, even though such a break c<strong>an</strong><br />

be shown elsewhere to have occurred (second point<br />

above).<br />

<strong>Thermal</strong> Structure of the Bishop <strong>Ignimbrite</strong><br />

Density values measured at each section are plotted<br />

in figures 7–10 as vertical profiles linked to the<br />

horizontal profile along the line of the Owens<br />

Gorge, with independently determined stratigraphic<br />

units <strong>an</strong>d horizons plotted for comparison.<br />

The overall welding state of the ignimbrite along<br />

the walls of the gorge is shown in figure 11. We<br />

draw attention to five features from our data.<br />

1. We recognize four density, <strong>an</strong>d hence welding,<br />

maxima (labeled Zones a through d, from oldest to<br />

youngest). Continuity of these zones between sections<br />

has been ensured by tracing the welding zones<br />

(<strong>an</strong>d stratigraphic marker pl<strong>an</strong>es) along the wellexposed<br />

walls of the Owens River valley. The combination<br />

of stratigraphic controls <strong>an</strong>d good exposure<br />

demonstrates that these four welding zones<br />

are not equally prominent at <strong>an</strong>y one section, <strong>an</strong>d<br />

that the ignimbrite hosting Zones b through d<br />

forms <strong>an</strong> imbricate, overlapping succession, such<br />

that the most prominent development of each<br />

package <strong>an</strong>d its associated welding zone occurs successively<br />

farther from the source (fig. 11). Each<br />

welding zone is developed in a stratigraphically<br />

confined part of the ignimbrite: (a) Ig1Ea; (b) lower<br />

Ig1Eb; (c) upper Ig1Eb; (d) upper Ig2Eb, <strong>an</strong>d thus the<br />

most heat-retentive (i.e., highest emplacement<br />

temperature) material was emplaced to form these<br />

parts of the ignimbrite at certain times. The development<br />

of welding then depends on the local<br />

thickness of each package <strong>an</strong>d the development of


660 C. J. N. WILSON AND W. HILDRETH<br />

Figure 5. Series of longitudinal sections down the line of the Owens Gorge projected on the true left side to show<br />

the data used to define the ignimbrite packages. a, Maximum lithic size (average lengths of the five largest clasts)<br />

in millimeters. b, Numbers of 15-mm-long lithic clasts per square meter. c, The count percent of Glass Mountain<br />

rhyolite in the 15-mm lithic fraction (measurement techniques after Wilson <strong>an</strong>d Hildreth 1997). The dotted line in<br />

b is a contour of 50 lithics/m 2 used to separate lower <strong>an</strong>d upper parts of Ig1Eb <strong>an</strong>d their associated welding zones<br />

(see fig. 11). Note that the approximate paleoflow direction is parallel to the gorge, <strong>an</strong>d so dist<strong>an</strong>ces down the gorge<br />

approximate to radial dist<strong>an</strong>ces from the source.<br />

additional stresses due to subsequent syneruptive<br />

lithostatic load; for example, Zone c attains lower<br />

maximum densities in the upper Owens Gorge (fig.<br />

7, section A) because it was thinner <strong>an</strong>d was buried<br />

under lesser thicknesses of Ig2E material there<br />

compared with the middle <strong>an</strong>d lower reaches of the<br />

gorge.<br />

2. The upper Owens Gorge (sections A–C) is the<br />

only area where all four zones are exposed in proximity,<br />

<strong>an</strong>d Zone b is the best developed. Zone a is<br />

represented by incipient pumice flattening <strong>an</strong>d is<br />

exposed only over a limited area where downcutting<br />

to the east by the Owens River intersects a<br />

westward-thickening wedge of Ig1Ea material.<br />

Zones b <strong>an</strong>d c, both in Ig1Eb, are separately demarcated<br />

in this area by at most a slight reduction<br />

in density (e.g., section A). However, the stratigraphic<br />

control given by variations in numerical<br />

abund<strong>an</strong>ces of lithic fragments (particularly the 50<br />

lithics/m 2 contour; fig. 5) show that the thick<br />

welded Ig1Eb deposits in the middle Owens Gorge<br />

(e.g., sections F <strong>an</strong>d G) correlate only with Zone c<br />

<strong>an</strong>d that material equivalent to Zone b in the upper<br />

parts of the gorge forms only a nonwelded to poorly<br />

welded basal zone to the ignimbrite down-valley<br />

from the prominent high in the basement rocks (fig.<br />

11).<br />

3. In the middle to lower Owens Gorge (sections<br />

G–I), there is a pronounced density minimum between<br />

the maxima of Zones c <strong>an</strong>d d (figs. 7–9), but<br />

the location of this minimum is at a level of no<br />

special stratigraphic signific<strong>an</strong>ce. The basal parts


Journal of Geology ASSEMBLING AN IGNIMBRITE 661<br />

Figure 6. Plot of the contents of !1/16-mm material<br />

versus the me<strong>an</strong> grain size (M z of Folk <strong>an</strong>d Ward 1957)<br />

in sieved samples collected from the four main packages<br />

of Bishop ignimbrite along, or close to, the Owens Gorge.<br />

Samples represent both proximal <strong>an</strong>d distal material in<br />

all packages <strong>an</strong>d also bracket stratigraphically material<br />

that is representative of the densest-welded ignimbrite.<br />

Note the clear distinction in !1/16-mm ash contents between<br />

Ig1E <strong>an</strong>d Ig2E samples.<br />

of Ig2Ea show <strong>an</strong> upward diminishing welding intensity<br />

at sections H (fig. 9) <strong>an</strong>d I, yet at all sites<br />

in the Bishop Tuff where Ig2Ea overlies solely fall<br />

deposits, it is nonwelded. We infer that the emplacement<br />

temperature of Ig2Ea in itself was insufficient<br />

to cause welding, but where it was deposited<br />

on still-hot Ig1Eb, the residual heat from<br />

the underlying deposit was sufficient to cause<br />

“back-welding” of Ig2Ea under the lithostatic stress<br />

of later-deposited material. The density minimum<br />

in the middle Owens Gorge is poorly defined at<br />

sections F <strong>an</strong>d G, largely because Zone d is less<br />

well developed th<strong>an</strong> farther down-gorge as a consequence<br />

of the lesser overlying thickness of Ig2Eb.<br />

4. Wide fluctuations in density are seen in the<br />

nonwelded to sintered material of section J, <strong>an</strong>d the<br />

lowest in situ densities are actually in sintered, not<br />

nonwelded, material (fig. 10). Grain size <strong>an</strong>d component<br />

characteristics of the nonwelded material<br />

ch<strong>an</strong>ge very little, <strong>an</strong>d so the density fluctuations<br />

are not due to variations in either dense (lithic,<br />

crystal) or fine-ash (!1/16 mm) components (fig.<br />

10). We interpret the density values to reflect the<br />

locking in by the sintering processes of bulk densities<br />

at a stage when compaction of the ignimbrite<br />

from its newly emplaced state was incomplete (i.e.,<br />

while the deposit was stationary but still hot <strong>an</strong>d<br />

somewhat exp<strong>an</strong>ded). We infer that nonwelded material<br />

could continue to compact locally by intergrain<br />

movements in response to load stresses (probably<br />

augmented by earthquake-induced ground<br />

shaking in this tectonically active area; Pinter<br />

1995), whereas the structural framework induced<br />

by sintering could resist further compaction. Thus<br />

the mathematically convenient, extrapolated density<br />

value of 1.0 Mg/m 3 adopted by Rag<strong>an</strong> <strong>an</strong>d Sherid<strong>an</strong><br />

(1972) <strong>an</strong>d Sherid<strong>an</strong> <strong>an</strong>d Rag<strong>an</strong> (1976) for nonwelded<br />

material to calculate compactional strain<br />

in ignimbrites <strong>an</strong>d infer the position of the original<br />

ignimbrite surface is not supported by our data.<br />

5. An import<strong>an</strong>t aspect of interpreting the welding<br />

<strong>an</strong>d thermal history of <strong>an</strong> ignimbrite is identifying<br />

whether devitrification <strong>an</strong>d vapor-phase<br />

crystallization follow on from the welding compaction,<br />

or interrupt <strong>an</strong>d terminate it. In several<br />

places in the sections measured, there are sharp<br />

ch<strong>an</strong>ges from vitric to nonvitric tuff, <strong>an</strong>d samples<br />

collected in proximity from each type of material<br />

show no major differences in density (fig. 12).<br />

We thus infer that welding densification was effectively<br />

completed before devitrification <strong>an</strong>d/or<br />

vapor-phase crystallization <strong>an</strong>d that the latter two<br />

processes did not act to prematurely halt densification<br />

(cf. item 4, above). Since the densities of the<br />

minerals created by devitrification are greater th<strong>an</strong><br />

those for the vitric phase, a lack of systematically<br />

increased density in devitrified material must<br />

therefore be accomp<strong>an</strong>ied by some increase in the<br />

microscale porosity of the rock.<br />

Three prominent zones of vapor-phase crystallization<br />

are also present in parts of the Bishop ignimbrite<br />

described here: (1) s<strong>an</strong>dwiched between<br />

the most strongly welded portions of Zones b <strong>an</strong>d<br />

c in the upper Owens Gorge, (2) s<strong>an</strong>dwiched between<br />

the most strongly welded portions of Zones<br />

c <strong>an</strong>d d in the middle to lower Owens Gorge (fig.<br />

10; Holt <strong>an</strong>d Taylor 1998), <strong>an</strong>d (3) above Zone d<br />

in the lower Owens Gorge. In similar fashion to<br />

the zones of denser welding that they overlie, each<br />

vapor-phase crystallization zone becomes successively<br />

most prominent in sections along the horizontal<br />

profile of the Owens Gorge. Parallel to variations<br />

in the welding profiles, vapor-phase<br />

crystallization is prominent in Ig2Ea <strong>an</strong>d lower<br />

Ig2Eb only where they overlie welded <strong>an</strong>d devitrified<br />

Ig1Eb (i.e., Zone c; fig. 9); in contrast, where<br />

they overlie vitric welded ignimbrite or fall deposits,<br />

no signific<strong>an</strong>t vapor-phase crystallization is<br />

seen (fig. 10). Thus the regional distribution of areas<br />

of intense fumarolic alteration reported by Sherid<strong>an</strong><br />

(1970) closely reflect the alteration of Ig2E


Figure 7. Density profiles through the Bishop ignimbrite at seven of the 10 sections (A–F, I) shown in figure 1. For<br />

sections G, H, <strong>an</strong>d J, see figures 8–10, respectively. In each profile, filled circles represent data from nonwelded<br />

(sievable) tuff, <strong>an</strong>d filled squares, data from coherent blocks of tuff.


Journal of Geology ASSEMBLING AN IGNIMBRITE 663<br />

Figure 8. View of east wall of the Owens Gorge at grid reference 624490 showing position of section G, selected<br />

data on densities <strong>an</strong>d lithic lithologies, <strong>an</strong>d positions of welding zones <strong>an</strong>d package boundaries. Symbols are as in<br />

figure 7.<br />

above largely buried f<strong>an</strong>s of Ig1Eb (welding Zone c)<br />

that accumulated down the paleodrainages of the<br />

Owens Gorge <strong>an</strong>d Chidago C<strong>an</strong>yon (Wilson <strong>an</strong>d<br />

Hildreth 1997, their fig. 6).<br />

Discussion<br />

Here we review usage of the terms “flow unit” <strong>an</strong>d<br />

“cooling unit” in the light of our Bishop Tuff data.<br />

In doing this, we emphasize a key factor, namely,<br />

that although these terms remain useful for descriptive<br />

purposes, their qu<strong>an</strong>titative application in<br />

interpreting emplacement histories of ignimbrites<br />

is problematic in several import<strong>an</strong>t respects. To use<br />

either term in a nonsimplistic way requires recognition<br />

<strong>an</strong>d acknowledgment of implicit complexities<br />

that have been widely ignored.<br />

Mech<strong>an</strong>ical <strong>Building</strong> <strong>Blocks</strong>: Flow Units. Recognition<br />

of flow units in ignimbrites fundamentally<br />

depends on recognizing flow-unit boundaries,<br />

which in turn indicate (to some extent) a hiatus in<br />

deposition. Note, however, that the inverse does<br />

not hold, that is, not all discontinuities need represent<br />

flow-unit boundaries. Some discontinuities<br />

may arise, for example, through irregular rates of<br />

deposition at the base of a single short-lived flow<br />

(e.g., layering in the Taupo ignimbrite veneer deposit;<br />

Wilson 1985). The nature of flow-unit boundaries<br />

c<strong>an</strong> be highly variable, including the presence<br />

of material of contrasting depositional mech<strong>an</strong>isms<br />

(especially fall material, that would have taken a<br />

finite time to accumulate; fig. 2; Wilson <strong>an</strong>d Hildreth<br />

1997), the presence of <strong>an</strong> inverse-graded layer<br />

2a (Sparks et al. 1973), or truncation of such features<br />

as degassing pipes or pumice concentration<br />

zones in the underlying material (figs. 3, 4). From<br />

this, two corollaries arise.<br />

First, whether or not a flow-unit boundary is seen<br />

depends primarily on the state of the earlierdeposited<br />

material. If the earlier material was unconsolidated,<br />

then shearing at the base of the later<br />

flow may cause mixing of the earlier <strong>an</strong>d newly<br />

depositing material to obscure <strong>an</strong>y contact; unless<br />

there is some contrast in properties (e.g., grain size<br />

or composition), no flow-unit boundary <strong>an</strong>d hence<br />

no evidence for a hiatus may be present. Processes<br />

that would enh<strong>an</strong>ce consolidation of the substrate<br />

include the following, each of which may act on<br />

different time scales: the presence of abund<strong>an</strong>t<br />

coarse clasts that would supply a rigid framework<br />

(e.g., fig. 4a); degassing <strong>an</strong>d compaction of nonwelded<br />

material; welding-induced cohesion; or deposition<br />

of fall material. The second corollary follows<br />

from this, that there is a spectrum of possible<br />

time breaks, reflected by boundaries of some sort,<br />

within the course of emplacement of <strong>an</strong>y ignimbrite,<br />

from seconds up to (conceivably as long as)<br />

months. Thus, depending on local circumst<strong>an</strong>ces,


664 C. J. N. WILSON AND W. HILDRETH<br />

Figure 9. View of east wall of the Owens Gorge at grid reference 633468 showing position of section H, selected<br />

data on densities <strong>an</strong>d lithic lithologies, <strong>an</strong>d positions of welding zones <strong>an</strong>d package boundaries. See Wilson <strong>an</strong>d<br />

Hildreth (1997, their fig. 10) for a closer view of the fall bed contact from this section. Symbols are as in figure 7.<br />

a hiatus between separate flow units (identified as<br />

such here by differing characteristics) may or may<br />

not result in a visible flow-unit boundary.<br />

Fisher (1966) <strong>an</strong>d Br<strong>an</strong>ney <strong>an</strong>d Kokelaar (1992,<br />

1997) suggested that thick ignimbrite sequences<br />

with no apparent flow-unit boundaries may have accumulated<br />

progressively in a continuous fashion<br />

rather th<strong>an</strong> being the products of a single flow or of<br />

piecemeal accumulation from a number of discrete<br />

flows. Our observations <strong>an</strong>d data for the Bishop Tuff<br />

imply, however, that ignimbrite with no visible<br />

flow-unit boundaries c<strong>an</strong> result from emplacement<br />

of multiple flows separated by time breaks. In other<br />

words, a lack of flow-unit boundaries is not diagnostic<br />

of continuous, progressive aggradation. The<br />

clearest example of this is at section G, where Ig2Eb<br />

overlies Ig1Eb with no visible boundary in the field,<br />

the contact being demarcated to !0.5 m by the incoming<br />

upward of Glass Mountain–derived rhyolite<br />

lithic fragments (figs. 5, 8). Elsewhere in the Bishop<br />

ignimbrite (e.g., sections H <strong>an</strong>d I, fig. 10; also Wilson<br />

<strong>an</strong>d Hildreth 1997, their fig. 10), fall material<br />

<strong>an</strong>d Ig2Ea were deposited at this horizon, indicating<br />

a time gap of hours to tens of hours between ignimbrite<br />

packages Ig1Eb <strong>an</strong>d Ig2Eb. We thus conclude<br />

that the presence of a flow-unit boundary<br />

does indicate a hiatus in deposition but that the<br />

reverse is not necessarily true, <strong>an</strong>d that a lack of<br />

flow-unit boundaries, in the absence of other evidence,<br />

c<strong>an</strong>not be taken to indicate that accumu-<br />

lation was continuous (cf. Br<strong>an</strong>ney <strong>an</strong>d Kokelaar<br />

1997).<br />

<strong>Thermal</strong> <strong>Building</strong> <strong>Blocks</strong>: Cooling Units. In the definitions<br />

<strong>an</strong>d illustrative examples of Smith (1960b,<br />

especially pl. 20), it is implicit that the primary<br />

me<strong>an</strong>s whereby a simple cooling unit is recognized<br />

is the ideality of its patterns of zonation in welding,<br />

devitrification, <strong>an</strong>d vapor-phase recrystallization.<br />

The term “compound cooling unit” is then automatically<br />

applied to <strong>an</strong>y deposit that shows “pronounced<br />

deviations” from this pattern (Smith<br />

1960b, p. 157) but c<strong>an</strong> still be considered to have<br />

accumulated rapidly enough for the whole rock<br />

body to cool as one (i.e., to be a single cooling unit).<br />

However, these definitions, as well as the concept<br />

of a composite sheet, are problematic in several<br />

respects:<br />

1. As the perceived zonations for a single cooling<br />

unit c<strong>an</strong> vary so widely (Smith 1960a, 1960b), the<br />

criterion for “pronounced deviations” <strong>an</strong>d hence<br />

the simple versus compound distinction c<strong>an</strong> be<br />

given only in qualitative terms.<br />

2. Even given a simple welding <strong>an</strong>d alteration<br />

profile, turning such a description into interpretation<br />

of the time-temperature relationships of the<br />

deposited material is not always possible. For example,<br />

several workers have inferred (e.g., Christi<strong>an</strong>sen<br />

1979) or assumed (Riehle 1973; Riehle et<br />

al. 1995) that a simple cooling unit is effectively<br />

isothermal on emplacement, consistent with rapid


Journal of Geology ASSEMBLING AN IGNIMBRITE 665<br />

Figure 10. View of Chalk Bluffs at grid reference 733421. Section J was measured from the fall deposit up to the<br />

“cryptic ledge” (a zone of barely sintered material that marks the approximate contact between Ig2Ea <strong>an</strong>d Ig2Eb) on<br />

this face <strong>an</strong>d above this level in a slip scar 100 m farther east. Selected grain size <strong>an</strong>d componentry data (the latter<br />

measured down to 1/16 mm, with !1/16 mm material assumed to be vitric ash) from material taken at the same<br />

point as the density samples are shown to compare with the density variations. Symbols are as in figure 7.<br />

accumulation <strong>an</strong>d simple patterns of welding <strong>an</strong>d<br />

alteration. If, however, a simple cooling unit is built<br />

up progressively (whether continuously or not),<br />

then the emplacement temperatures may have deviated<br />

signific<strong>an</strong>tly from uniform during accumulation.<br />

For example, the basal <strong>an</strong>d topmost materials<br />

may have been cooler <strong>an</strong>d the interior hotter,<br />

or increasing magma temperatures in a zoned package<br />

may influence emplacement temperatures, yet<br />

the welding <strong>an</strong>d alteration zonation could, in principle,<br />

be simple in the sense used by Smith.<br />

3. The distinction between single <strong>an</strong>d multiple<br />

cooling units is valid in principle but may be problematic<br />

to apply in practice. We label the Bishop<br />

ignimbrite in the Owens Gorge a single cooling<br />

unit because of evidence from the density profiles<br />

that the earlier parts of the ignimbrite (Ig1E) were<br />

still hot when later parts (Ig2E) accumulated. For<br />

example, welding Zone c in Ig1Eb was still hot<br />

enough when buried by Ig2Ea <strong>an</strong>d/or Ig2Eb to have<br />

caused welding in the immediately overlying Ig2<br />

material (sections G–I; figs. 7–9) even though that


666 C. J. N. WILSON AND W. HILDRETH<br />

Figure 11. Longitudinal section down the line of the Owens Gorge projected on the true left side showing the<br />

overall welding structure of the ignimbrite. Density data from the 10 profiles (figs. 7–10) were combined with sparse<br />

point measurements <strong>an</strong>d then matched to field observations at numerous sites along the gorge walls using the<br />

calibration between density, rock texture, <strong>an</strong>d welding given in table 2. The dotted line is the 50 lithics/m 2 contour<br />

derived from lithic numerical abund<strong>an</strong>ce data (fig. 5b); note how this contour serves to demarcate welding Zone b<br />

(best developed at sections A <strong>an</strong>d B) <strong>an</strong>d Zone c (best developed down-gorge from section E), with <strong>an</strong> overlap around<br />

sections C <strong>an</strong>d D.<br />

material was insufficiently hot to weld where it<br />

overlay a cold substrate (section J; fig. 10). We thus<br />

disagree with Sherid<strong>an</strong> (1968), who considered that<br />

there were two cooling units in the same area as<br />

sections G–I, implying that the Bishop ignimbrite<br />

is a composite sheet. In general, we suggest from<br />

our data that only in circumst<strong>an</strong>ces where the earlier<br />

unit was effectively “cold,” that is, unable to<br />

influence the welding profile or alteration state of<br />

the later unit, might the density minimum between<br />

two zones of greater welding actually coincide<br />

with the chronostratigraphic boundary between<br />

the two bodies of material. Only under such<br />

circumst<strong>an</strong>ces could the two bodies then be considered<br />

as separate cooling units. In the case of the<br />

Bishop Tuff, lateral ch<strong>an</strong>ges in welding down the<br />

Owens Gorge from areas with only one maximum<br />

in welding intensity (e.g., section E) to those with<br />

two welding maxima do not reflect splitting of<br />

welding zones (from “simple” to “compound”) or<br />

of cooling units (from one to two), but instead result<br />

from differing development of separate welding<br />

maxima as successive packages of host material become<br />

thickest in turn down-gorge.<br />

Thus, in documenting ignimbrites even as young<br />

<strong>an</strong>d clearly exposed as the Bishop, problems arise<br />

with the qu<strong>an</strong>titative interpretation of the cooling<br />

unit concept. Although the original concepts have<br />

demonstrated their value as a descriptive framework<br />

(e.g., Christi<strong>an</strong>sen 1979, <strong>an</strong>d references<br />

therein), interpretative application of the concept<br />

to infer the timings <strong>an</strong>d temperatures of accumulating<br />

ignimbrite material is fraught with problems.<br />

These arise primarily because the welding intensity<br />

(<strong>an</strong>d <strong>an</strong>y consequent devitrification <strong>an</strong>d vaporphase<br />

alteration) depends on several properties of<br />

the deposited material, including variations in (1)<br />

emplacement temperatures (themselves complexly<br />

related to eruption composition, heat losses during<br />

tr<strong>an</strong>sport, <strong>an</strong>d admixture of lithic material), (2) juvenile<br />

compositions, (3) residual volatile contents,<br />

<strong>an</strong>d (4) load stresses. However, even when such factors<br />

as juvenile composition <strong>an</strong>d residual volatiles<br />

are accounted for (or are assumed to be uniform)<br />

in a particular deposit, a given density/welding profile<br />

has a nonunique relationship to its emplacement<br />

temperature profile because the individual<br />

mech<strong>an</strong>ical building blocks (flow units or emplacement<br />

pulses) <strong>an</strong>d their associated temperatures c<strong>an</strong>not<br />

necessarily be uniquely defined, even in <strong>an</strong> ostensibly<br />

simple cooling unit.<br />

Furthermore, although the descriptive distinc-


Journal of Geology ASSEMBLING AN IGNIMBRITE 667<br />

Figure 12. Plot of density data from five pairs of samples<br />

from across vitric to nonvitric boundaries in portions<br />

of the ignimbrite that show minimal background gradients<br />

in density. We use these data to infer that densification<br />

of the ignimbrite by welding processes acting on<br />

the viscous vitric component had largely or completely<br />

ceased by the time crystallization due to devitrification<br />

or vapor-phase alteration occurred.<br />

tion between simple <strong>an</strong>d compound cooling units<br />

is in principle straightforwardly based on the patterns<br />

of welding <strong>an</strong>d crystallization in the rock<br />

mass, interpreting the genetic <strong>an</strong>d chronologic signific<strong>an</strong>ce<br />

of a compound cooling unit profile may<br />

be complex. First, <strong>an</strong>d most straightforwardly, compound<br />

cooling profiles that show only one welding<br />

maximum that occurs toward the top of the ignimbrite<br />

(e.g., section J in the Bishop; fig. 10) are often<br />

interpreted as due to the superposition of increasingly<br />

hotter material with no perceived time breaks<br />

(Lipm<strong>an</strong> et al. 1966; Br<strong>an</strong>ney <strong>an</strong>d Kokelaar 1997).<br />

However, although the welding zonation is compound<br />

(according to Smith 1960b), the emplacement<br />

dynamics of the deposit may not differ in <strong>an</strong>y<br />

respect from those of a simple cooling unit, other<br />

th<strong>an</strong> in the emplaced material becoming hotter<br />

with time. Second, compound cooling units that<br />

have more th<strong>an</strong> one welding maximum are widely<br />

interpreted as reflecting emplacement at intervals<br />

of time sufficient to allow partial cooling of the<br />

earlier material (Christi<strong>an</strong>sen 1979, p. 29; Riehle et<br />

al. 1995), unless evidence for other sources of cool-<br />

ing, particularly the incorporation of cold lithic material,<br />

is present (Lipm<strong>an</strong> et al. 1989, p. 340).<br />

From our work, we infer that there are only two<br />

extended breaks in emplacement of the Bishop ignimbrite,<br />

represented by the F6 fall material separating<br />

Ig1Ea <strong>an</strong>d Ig1Eb (sections A–C), <strong>an</strong>d the F8–<br />

F9 fall material between Ig1Eb <strong>an</strong>d Ig2Ea (sections<br />

H <strong>an</strong>d I). Using fall-accumulation rate estimates<br />

(Wilson <strong>an</strong>d Hildreth 1997), these time breaks are<br />

estimated as 5–10 h for the former <strong>an</strong>d 3 h for the<br />

latter (with <strong>an</strong> additional but short period represented<br />

by the settling out of ash <strong>an</strong>d trivial erosion<br />

[at one site only] at the level of the F8–F9 contact).<br />

In Ig1Ea, other thin fall horizons <strong>an</strong>d flow-unit<br />

boundaries are seen (fig. 2), but the time breaks are<br />

by inference !3 h. None of these time breaks in<br />

themselves is adequate to permit signific<strong>an</strong>t cooling<br />

of the ignimbrite packages (Riehle 1973), <strong>an</strong>d<br />

no discontinuity in density is seen across <strong>an</strong>y of<br />

these depositional breaks. Thus the whole accumulation<br />

of Bishop ignimbrite in this sector represents<br />

a single emplacement unit from the perspective<br />

of its thermal history; that is, it is a single,<br />

compound cooling unit (Smith 1960b).<br />

However, the density minima in sections A–J,<br />

although defining compound cooling, do not coincide<br />

with or represent time breaks at all. If one<br />

accepts that the prime controls on welding intensity<br />

in the interior of the Bishop ignimbrite are emplacement<br />

temperatures <strong>an</strong>d load stresses, <strong>an</strong>d that<br />

load stresses diminish systematically upward, the<br />

closely spaced vertical density variations must then<br />

reflect differing emplacement temperatures with<br />

time. For example, the lower part of Ig2Eb <strong>an</strong>d all<br />

of Ig2Ea were emplaced at signific<strong>an</strong>tly cooler temperatures,<br />

in the latter case insufficient to induce<br />

welding where overlying a fall-deposit substrate.<br />

Our data imply that models presented by Riehle<br />

(1973) <strong>an</strong>d Riehle et al. (1995) for inferring the emplacement<br />

temperature of simple or compound,<br />

single or multiple cooling units are thus limited in<br />

two respects. First, as discussed earlier, a given density<br />

profile in a real deposit (even <strong>an</strong> apparently<br />

simple cooling unit) c<strong>an</strong> result from a number of<br />

possible combinations of increments of material (or<br />

flow units) with nonuniform temperatures. To<br />

match a real-world density profile of a simple cooling<br />

unit with a given inferred emplacement temperature<br />

thus requires <strong>an</strong> assumption of isothermal<br />

emplacement. Second, in a single compound cooling<br />

unit such as the Bishop ignimbrite, we c<strong>an</strong> show<br />

that the position of the density minimum between<br />

two zones of more strongly welded ignimbrite does<br />

not coincide with <strong>an</strong>y time break, such breaks elsewhere<br />

being unequivocally flagged by interbedded


668 C. J. N. WILSON AND W. HILDRETH<br />

fall material (e.g., sections H <strong>an</strong>d I) or inferred from<br />

sharp ch<strong>an</strong>ges in ignimbrite lithology (e.g., sections<br />

F <strong>an</strong>d G). Thus, defining individual bodies of material<br />

for the purposes of thermal modeling by using<br />

the density minimum between them is inherently<br />

flawed for single cooling units <strong>an</strong>d will lead to artificial<br />

results.<br />

Our work suggests also that the use of the term<br />

“composite sheet,” as introduced by Smith, may<br />

also be problematic, for two reasons. First, it is unclear<br />

whether the composite sheet is me<strong>an</strong>t to apply<br />

to material from a single eruption, or could be<br />

from widely separated eruptions, or both. For example,<br />

the Huckleberry Ridge Tuff is described by<br />

Christi<strong>an</strong>sen (1979, 2001) as a composite sheet<br />

composed of three members, each forming separate<br />

cooling units in distal areas yet merging to yield a<br />

single compound cooling unit in proximal areas.<br />

The whole assemblage was envisaged as being<br />

erupted in “hours to days,” that is, over a time scale<br />

insufficient for signific<strong>an</strong>t cooling. On the other<br />

h<strong>an</strong>d, features such as lava flows or sediments were<br />

suggested by Smith (1960b) as possibly present between<br />

separate cooling units in a composite sheet,<br />

with the implication of subst<strong>an</strong>tial time intervals.<br />

Under such circumst<strong>an</strong>ces it is difficult to envisage<br />

the earlier <strong>an</strong>d later cooling units as being part<br />

of what could be seen as a single eruption. Thus<br />

the grounds for delineating single versus multiple<br />

compound cooling units, <strong>an</strong>d hence identification<br />

of a composite sheet in a form that permits genetic<br />

interpretation, are unclear. Second, the lateral<br />

ch<strong>an</strong>ges from simple to compound, or from single<br />

to multiple cooling units are generally inferred or<br />

assumed to represent the same body of material, in<br />

order for comparisons to be made. However, the<br />

imbricate structure of the packages in the Bishop<br />

ignimbrite along the Owens Gorge show that this<br />

inference or assumption is false in this case, <strong>an</strong>d<br />

that the apparent lateral ch<strong>an</strong>ge of one into two (or<br />

simple into compound) cooling units c<strong>an</strong> be <strong>an</strong> artifact<br />

of welding development in separate, imbricate<br />

packages of ignimbrite. Thus, while application<br />

of the term “composite sheet” remains valid<br />

as a descriptive term for <strong>an</strong> ignimbrite that shows<br />

differing development of cooling units in different<br />

areas, we suggest that <strong>an</strong>y genetic interpretation of<br />

that development requires further information<br />

about relationships between the ignimbrite material<br />

in those areas.<br />

Conclusions<br />

The Bishop ignimbrite exposed in a longitudinal<br />

profile down the Owens Gorge shows four large-<br />

scale imbricate depositional lenses (packages Ig1Ea,<br />

Ig1Eb, Ig2Ea, <strong>an</strong>d Ig2Eb of Wilson <strong>an</strong>d Hildreth<br />

1997), each of which is the product of numerous<br />

individual flow units or pulses of material. The<br />

clearest-defined flow-unit boundaries reflect intervals<br />

of ignimbrite nondeposition represented by<br />

thin fall deposits. Such intervals are inferred to<br />

have lasted for at least several hours <strong>an</strong>d as long as<br />

10–15 h. However, in most such cases, <strong>an</strong>d in other<br />

examples in which partings occur between flow<br />

units, the physical characteristics (e.g., grain size,<br />

maximum lithic sizes, lithic lithologies) of the ignimbrite<br />

do not ch<strong>an</strong>ge signific<strong>an</strong>tly across such<br />

boundaries. Conversely, some time breaks of hours<br />

to tens of hours are demonstrably not accomp<strong>an</strong>ied<br />

by <strong>an</strong>y visible flow-unit boundary or parting. Our<br />

data thus show that material with no visible flowunit<br />

boundaries may reflect quite different emplacement<br />

histories. It may be difficult to show<br />

whether such material accumulated as a single flow<br />

unit, or gradationally by progressive aggradation, or<br />

in punctuated fashion from a series of individual<br />

flows without consideration of other information.<br />

In the Bishop Tuff, the presence of time breaks in<br />

deposition of massive ignimbrite c<strong>an</strong> be demonstrated<br />

from sharp ch<strong>an</strong>ges in lithic lithologies (e.g.,<br />

between Ig1Eb <strong>an</strong>d Ig2Eb at section G; fig. 8) between<br />

units that elsewhere c<strong>an</strong> be shown to be separated<br />

by periods of fall <strong>an</strong>d ignimbrite deposition.<br />

Progressive aggradation may sometimes occur<br />

(Br<strong>an</strong>ney <strong>an</strong>d Kokelaar 1992), but our Bishop data<br />

show that discrete flow units are the norm in building<br />

multiflow packages that may or may not retain<br />

evidence for partings between flows.<br />

The Bishop ignimbrite in the Owens Gorge displays<br />

four zones of maximum density (<strong>an</strong>d correlated<br />

welding) that collectively form a single, compound<br />

cooling unit. Each depositional package was<br />

emplaced at temperatures that varied in time during<br />

the eruption. Maximum welding occurred<br />

where the deposited material was initially hottest,<br />

<strong>an</strong>d the zones of minimum welding reflect the presence<br />

of cooler-emplaced material <strong>an</strong>d not time<br />

breaks when cooling occurred. Thus the distinction<br />

drawn by Smith (1960b) between simple <strong>an</strong>d compound<br />

cooling units in ignimbrites c<strong>an</strong>not necessarily<br />

be interpreted solely in terms of the postemplacement<br />

cooling history. The intensities of the<br />

welding maxima vary with local load stresses (i.e.,<br />

thicknesses of the flow packages) <strong>an</strong>d are therefore<br />

greatest in the middle to lower Owens Gorge <strong>an</strong>d<br />

are unrelated to proximity to the source. In the<br />

most distal section, the onset of sintering locked<br />

in a lower bulk density th<strong>an</strong> that which nonwelded<br />

material could attain by compaction in the 0.76


Journal of Geology ASSEMBLING AN IGNIMBRITE 669<br />

million years since the eruption. Models that determine<br />

ignimbrite densities at the top or bottom<br />

of flows by linear extrapolation of welded-tuff densities<br />

in the central part of the flow are thus likely<br />

to be in error. We suggest that density values for<br />

nonwelded tuff at the tops <strong>an</strong>d bottoms of ignimbrite<br />

units should be separately determined.<br />

Existing definitions of the terms “flow unit” <strong>an</strong>d<br />

“cooling unit” remain valid in theory, are valuable<br />

tools for descriptive purposes, <strong>an</strong>d represent the basic<br />

building blocks for constructing ignimbrites.<br />

However, different kinds of data are required, in<br />

addition to identification of flow packages <strong>an</strong>d density<br />

profiles, to constrain the episodicity <strong>an</strong>d original<br />

temperature of flow emplacement.<br />

Br<strong>an</strong>ney, M. J., <strong>an</strong>d Kokelaar, B. P. 1992. A reappraisal of<br />

ignimbrite emplacement: progressive aggradation <strong>an</strong>d<br />

ch<strong>an</strong>ges from particulate to non-particulate flow during<br />

emplacement of high-grade ignimbrite. Bull. Volc<strong>an</strong>ol.<br />

54:504–520.<br />

———. 1997. Gi<strong>an</strong>t bed from a sustained catastrophic<br />

density current flowing over topography: Acatlán ignimbrite,<br />

Mexico. Geology 25:115–118.<br />

Carey, S. N. 1991. Tr<strong>an</strong>sport <strong>an</strong>d deposition of tephra by<br />

pyroclastic flows <strong>an</strong>d surges. In Fisher, R. V., <strong>an</strong>d<br />

Smith, G. A., eds. Sedimentation in volc<strong>an</strong>ic settings.<br />

SEPM Spec. Publ. 45:39–57.<br />

Chapin, C. E., <strong>an</strong>d Lowell, G. R. 1979. Primary <strong>an</strong>d secondary<br />

flow structures in ash-flow tuffs of the Gribbles<br />

Run paleovalley, central Colorado. In Chapin, C.<br />

E., <strong>an</strong>d Elston, W. E., eds. Ash-flow tuffs. Geol. Soc.<br />

Am. Spec. Pap. 180:137–154.<br />

Christi<strong>an</strong>sen, R. L. 1979. Cooling units <strong>an</strong>d composite<br />

sheets in relation to caldera structure. In Chapin, C.<br />

E., <strong>an</strong>d Elston, W. E., eds. Ash-flow tuffs. Geol. Soc.<br />

Am. Spec. Pap. 180:29–42.<br />

———. 2001. The Quaternary <strong>an</strong>d Pliocene Yellowstone<br />

Plateau Volc<strong>an</strong>ic Field of Wyoming, Idaho, <strong>an</strong>d Mont<strong>an</strong>a.<br />

U.S. Geol. Surv. Prof. Pap. 729-G:G1–G145.<br />

Druitt, T. H. 1998. Pyroclastic density currents. In Gilbert,<br />

J. S., <strong>an</strong>d Sparks, R. S. J., eds. The physics of<br />

explosive volc<strong>an</strong>ic eruptions. Geol. Soc. Lond. Spec.<br />

Publ. 145:145–182.<br />

Fisher, R. V. 1966. Mech<strong>an</strong>ism of deposition from pyroclastic<br />

flows. Am. J. Sci. 264:350–363.<br />

———. 1979. Models for pyroclastic surges <strong>an</strong>d pyroclastic<br />

flows. J. Volc<strong>an</strong>ol. Geotherm. Res. 6:305–318.<br />

Folk, R. L., <strong>an</strong>d Ward, W. C. 1957. Brazos River bar: a<br />

study in the signific<strong>an</strong>ce of grain size parameters. J.<br />

Sediment. Petrol. 27:3–26.<br />

Freundt, A. 1999. The formation of high-grade ignimbrites.<br />

II. A pyroclastic suspension current model with<br />

implications also for low-grade ignimbrites. Bull. Volc<strong>an</strong>ol.<br />

60:545–567.<br />

Freundt, A., <strong>an</strong>d Schmincke, H.-U. 1986. Emplacement<br />

REFERENCES CITED<br />

ACKNOWLEDGMENTS<br />

We th<strong>an</strong>k the U.K. Natural Environment Research<br />

Council (GR3/7032); N.Z. Foundation for Research,<br />

Science, <strong>an</strong>d Technology; <strong>an</strong>d Volc<strong>an</strong>o Hazards Program<br />

of the U.S. Geological Survey for past <strong>an</strong>d<br />

continuing fin<strong>an</strong>cial support. We especially th<strong>an</strong>k<br />

B. Christi<strong>an</strong>sen <strong>an</strong>d P. Lipm<strong>an</strong> for detailed <strong>an</strong>d informative<br />

critiques of <strong>an</strong> early version of this work<br />

<strong>an</strong>d for clarification of m<strong>an</strong>y issues relating to<br />

welded ignimbrites. We also acknowledge reviews<br />

<strong>an</strong>d comments from A. Freundt, G. Walker, V. M<strong>an</strong>ville,<br />

M. Rosenberg, F. Anderson, <strong>an</strong>d <strong>an</strong> <strong>an</strong>onymous<br />

reviewer.<br />

of small-volume pyroclastic flows at Laacher See<br />

(East-Eifel, Germ<strong>an</strong>y). Bull. Volc<strong>an</strong>ol. 48:39–59.<br />

Freundt, A.; Wilson, C. J. N.; <strong>an</strong>d Carey, S. N. 2000. <strong>Ignimbrite</strong>s<br />

<strong>an</strong>d block-<strong>an</strong>d-ash flow deposits. In Sigurdsson,<br />

H.; Houghton, B. F.; McNutt, S.; Rymer, H.; <strong>an</strong>d<br />

Stix, J., eds. Encyclopedia of volc<strong>an</strong>oes. S<strong>an</strong> Diego,<br />

Calif., Academic Press, p. 581–599.<br />

Gilbert, C. M. 1938. Welded tuff in eastern California.<br />

Geol. Soc. Am. Bull. 49:1829–1862.<br />

Hildreth, W. 1979. The Bishop Tuff: evidence for the origin<br />

of compositional zonation in silicic magma chambers.<br />

In Chapin, C. E., <strong>an</strong>d Elston, W. E., eds. Ash-flow<br />

tuffs. Geol. Soc. Am. Spec. Pap. 180:43–75.<br />

Holt, E. W., <strong>an</strong>d Taylor, H. P. 1998. 18 O/ 16 O mapping <strong>an</strong>d<br />

hydrogeology of a short-lived (∼10 years) fumarolic<br />

(1500C) meteoric-hydrothermal event in the upper<br />

part of the 0.76 Ma Bishop Tuff outflow sheet, California.<br />

J. Volc<strong>an</strong>ol. Geotherm. Res. 83:115–139.<br />

Leat, P. T., <strong>an</strong>d Schmincke, H.-U. 1993. Large-scale rheomorphic<br />

shear deformation in Miocene peralkaline ignimbrite<br />

E, Gr<strong>an</strong> C<strong>an</strong>aria. Bull. Volc<strong>an</strong>ol. 55:155–165.<br />

Lipm<strong>an</strong>, P. W.; Christi<strong>an</strong>sen, R. L.; <strong>an</strong>d O’Connor, J. T.<br />

1966. A compositionally zoned ash-flow sheet in<br />

southern Nevada. U.S. Geol. Surv. Prof. Pap. 524-F:<br />

F1–F47.<br />

Lipm<strong>an</strong>, P. W.; Sawyer, D. A.; <strong>an</strong>d Hon, K. 1989. Central<br />

S<strong>an</strong> Ju<strong>an</strong> caldera cluster. Field guide 3: South Fork to<br />

Lake City. In Chapin, C. E., <strong>an</strong>d Zidek, J., eds. Field<br />

excursions to volc<strong>an</strong>ic terr<strong>an</strong>es in the western United<br />

States. Vol. I. Southern Rocky Mountain region. N.<br />

M. Bur. Mines Mineral Resour. Mem. 46:330–350.<br />

Marshall, P. M. 1935. Acid rocks of the Taupo-Rotorua<br />

volc<strong>an</strong>ic district. Tr<strong>an</strong>s. R. Soc. N. Z. 64:323–366.<br />

Peterson, D. W. 1979. Signific<strong>an</strong>ce of the flattening of<br />

pumice fragments in ash-flow tuffs. In Chapin, C. E.,<br />

<strong>an</strong>d Elston, W. E., eds. Ash-flow tuffs. Geol. Soc. Am.<br />

Spec. Pap. 180:195–204.<br />

Pinter, N. 1995. Faulting on the Volc<strong>an</strong>ic Tablel<strong>an</strong>d,<br />

Owens Valley, California. J. Geol. 103:73–83.<br />

Rag<strong>an</strong>, D. M., <strong>an</strong>d Sherid<strong>an</strong>, M. F. 1972. Compaction of


670 C. J. N. WILSON AND W. HILDRETH<br />

the Bishop Tuff, California. Geol. Soc. Am. Bull. 83:<br />

95–106.<br />

Riehle, J. R. 1973. Calculated compaction profiles of<br />

rhyolitic ash flow tuffs. Geol. Soc. Am. Bull. 84:2193–<br />

2216.<br />

Riehle, J. R.; Miller, T. F.; <strong>an</strong>d Bailey, R. A. 1995. Cooling,<br />

degassing, <strong>an</strong>d compaction of rhyolitic ash-flow tuffs:<br />

a computational model. Bull. Volc<strong>an</strong>ol. 57:319–336.<br />

Roberts, B., <strong>an</strong>d Sidd<strong>an</strong>s, A. W. B. 1971. Fabric studies in<br />

the Llwyd Mawr ignimbrite, Caernarvonshire, North<br />

Wales. Tectonophysics 12:281–306.<br />

Roche, O.; Gilbertson, M.; Phillips, J.; <strong>an</strong>d Sparks, R. S.<br />

J. 2002. Experiments on deaerating gr<strong>an</strong>ular currents<br />

<strong>an</strong>d implications for pyroclastic flow mobility. Geophys.<br />

Res. Lett. 29(16), 10.1029/2002GL014819.<br />

Ross, C. S., <strong>an</strong>d Smith, R. L. 1961. Ash-flow tuffs: their<br />

origin, geologic relations <strong>an</strong>d identification. U.S. Geol.<br />

Surv. Prof. Pap. 366:1–81.<br />

Schmincke, H.-U. 1974. Volc<strong>an</strong>ological aspects of peralkaline<br />

silicic welded ash-flow tuffs. Bull. Volc<strong>an</strong>ol.<br />

38:594–636.<br />

Sherid<strong>an</strong>, M. F. 1968. Double cooling-unit nature of the<br />

Bishop Tuff in Owens Gorge, California (abstr.). Geol.<br />

Soc. Am. Spec. Pap. 115:351.<br />

———. 1970. Fumarolic mounds <strong>an</strong>d ridges of the Bishop<br />

Tuff, California. Geol. Soc. Am. Bull. 81:851–868.<br />

Sherid<strong>an</strong>, M. F., <strong>an</strong>d Rag<strong>an</strong>, D. M. 1976. Compaction of<br />

ash-flow tuffs. In Chilingari<strong>an</strong>, G. V., <strong>an</strong>d Wolf, K. H.,<br />

eds. Compaction of coarse-grained sediment. Amsterdam,<br />

Elsevier, p. 677–717.<br />

Smith, R. L. 1960a. Ash flows. Geol. Soc. Am. Bull. 71:<br />

795–842.<br />

———. 1960b. Zones <strong>an</strong>d zonal variations in welded ash<br />

flows. U.S. Geol. Surv. Prof. Pap. 354-F:F149–F159.<br />

Sparks, R. S. J.; Self, S.; <strong>an</strong>d Walker, G. P. L. 1973. Products<br />

of ignimbrite eruptions. Geology 1:115–118.<br />

Walker, G. P. L. 1970. Compound <strong>an</strong>d simple lava flows<br />

<strong>an</strong>d flood basalts. Bull. Volc<strong>an</strong>ol. 35:579–590.<br />

———. 1983. <strong>Ignimbrite</strong> types <strong>an</strong>d ignimbrite problems:<br />

J. Volc<strong>an</strong>ol. Geotherm. Res. 17:65–88.<br />

Wilson, C. J. N. 1985. The Taupo eruption, New Zeal<strong>an</strong>d.<br />

II. The Taupo ignimbrite. Philos. Tr<strong>an</strong>s. R. Soc. Lond.<br />

Ser. A 314:229–310.<br />

———. 1986. Pyroclastic flows <strong>an</strong>d ignimbrites. Sci. Prog.<br />

Oxf. 70:171–207.<br />

Wilson, C. J. N., <strong>an</strong>d Hildreth, W. 1997. The Bishop Tuff:<br />

new insights from eruptive stratigraphy. J. Geol. 105:<br />

407–439.<br />

Wilson, C. J. N., <strong>an</strong>d Walker, G. P. L. 1982. <strong>Ignimbrite</strong><br />

depositional facies: the <strong>an</strong>atomy of a pyroclastic flow.<br />

J. Geol. Soc. Lond. 139:581–592.<br />

Wright, J. V. 1981. The Rio Caliente ignimbrite: <strong>an</strong>alysis<br />

of a compound intraplini<strong>an</strong> ignimbrite from a major<br />

late Quaternary Mexic<strong>an</strong> eruption. Bull. Volc<strong>an</strong>ol. 44:<br />

189–212.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!