You are on page 1of 46

Short course

INTRODUCTION TO PHYSICAL VOLCANOLOGY AND


VOLCANIC TEXTURES

Prof. Dr. Christoph Breitkreuz


cbreit@geo.tu-freiberg.de

Centre for Volcanic Textures


www.geo.tu-freiberg.de/dynamo/CVT.htm

Institut fr Geologie/Palontologie
TU Bergakademie Freiberg
Freiberg, Oct. 3rd 7th, 2005

Background and intention of the course


Based on experience in the investigation of fossil and (sub-)recent volcanic and
volcanoclastic successions this short course puts emphasis on the recognition of
volcanic textures in outcrop, rock slab and thin section, and on the potentials and
limits of their genetic interpretation. The course neglects other important aspects of
volcanology such as magmatic petrology, geochemistry, isotope geochemistry,
geotectonic setting, and volcanic hazards.
The following text emphasises the processes which lead to the formation of volcanic
textures. The textures itself will be presented and discussed during the short course.

Contents
1. Some physical characteristics of magma and lava

2. Types of eruption

3. Volcanic forms

11

4. Volcanic fragmentation and fragments

18

Volcanic glass and its alteration

20

5. Volcanic transport and deposition: Lava and pyroclastics

22

Lava flow

22

Pyroclastic transport and deposition

27

Magma mingling and mixing

36

6. Emplacement, cooling and alteration textures in SiO2-rich lava


and high-grade ignimbrites
Concept of carapace and core facies
7. References

37
42
43

1. Some physical characteristics of magma and lava


In order to understand processes of magma ascent, eruption and fragmentation,
some physical properties and behaviours of silicate melts are considered here. These
are also relevant for the understanding of nucleation and growth of bubbles and
crystals, and of lava flow.
The viscosity of silicate melts depends
strongly on composition and, to a lesser
degree, on temperature, pressure, and on
the content of bubbles, crystals and lithics
(Fig. 1.1). SiO2 and Al2O3 occur in silicate
melts in the form of tetrahedrons which tend
to connect at their corners with each other
forming chains or networks (framework
builder). Therefore SiO2-rich magmas are
extremely viscous. All other components
(MgO, FeO, CaO, K2O, Na2O etc.) and
especially volatiles like H2O and CO2
interrupt these tetrahedron structures and
therefore lower the viscosity (framework
modifier). Figure 1.2 shows that in rhyolitic
melts, the extraction of a few % of H2O
results in an dramatic increase of viscosity.

Fig. 1.1 Relationship between viscosity


and temperature for some magmas. The
rhyolite was glassy or liquid through the
entire temperature range (From Cas &
Wright 1987, after Murase & McBirney
1973).

Table 1.1 Estimates of eruption temperatures for some common magmas (After Cas & Wright 1987).

4
Temperature

estimates

of

some

common

terrestrial magmas during eruption are given in


Table 1.1. Higher temperature lowers the viscosity
of silicate melts because the stronger vibration of
the components prevents network formation and
existing bonds are less stable.
Silicate melts have also large differences in
density (Fig. 1.3). These are mainly related to the
content of mafic elements like Fe and Mg. Melts
are less dense than their product after cooling (the
magmatic rock). E.g. granitic melt has about 2.2 g
cm-3 whereas granite has about 2.7 g cm-3. Apart
from some very low-viscosity lavas (e.g. alkaline
basaltic)

most

flow

processes

of

magma

Fig. 1.2 The effect of H2O on the viscosity

of (a) granitic and (b) basaltic melts at


varying temperatures (From Cas & Wright
1987, after Murase 1962).

Fig. 1.3 Densities of some molten volcanic


rocks with varying temperature at atmospheric
pressure (From Cas & Wright 1987, after
Murase & McBirney 1973).

5
ascent/emplacement and of lava are laminar (Reynolds Number < 1). Generally,
magma, lava and some pyroclastic flows behave as a Bingham plastic fluid.
Deformational behaviour of magma
or hot glass is complex. Below the
glass transition temperature Tg,
glass reacts always in a brittle
manner upon deformational stress.
Above Tg, the behaviour depends
on the deformation rate (Fig. 1.4).
Thus, at the same temperature,
swift deformation leads to brittle
fragmentation

whereas

slow

deformation yields ductile flow. Fig.


1.7 depicts the cooling history of a
90 m thick rhyolitic lava.
Fig. 1.4 Relation between deformation rate and
deformation style of magma or lava depending
on temperature; Tg = glass transition temperature

Nucleation

and

growth

of

crystals can take place in silicate


melts

during

ascent

and

emplacement or during eruption


and cooling of lava and pyroclastic
deposits. As mentioned before, the
physical presence of vesicles and
crystals increases the viscosity of
the melt. However, these processes
also

change

the

chemical

composition of the remaining melt,


and thus its viscosity. Nucleation of
vesicles and crystals is much easier
in SiO2-poor melts than in SiO2-rich
Fig. 1.5 Relation between super cooling (T) and
crystal nucleation and growth rate in a granitic melt
(Swanson et al. 1989)

melts; among other controls, this is


due to different diffusion rates.
Therefore, in a rhyolitic melt it takes

6
strong supercooling (T = temperature below the liquidus of the system) and time to
generate and grow crystals (Lofgren 1970, 1971a,b; Swanson et al. 1989; Fig. 1.5).
Thus, eruption of rhyolitic and phonolitic melts often results in the formation of glassy

Depth (m)

material like glass shards, pumice and obsidian.

Fig. 1.6 Bubble growth and water oversaturation in an ascending rhyolitic magma. The curves
define oversaturation in % as a function of the depth in the system during magma ascent. The
labels on the curves refer to the ascent or rise rates. The initial conditions are 4 km (a) and 1 km
(b), which correspond to initial water concentrations dissolved in the magma of 3.72 and 1.86
wt.%, respectively. Reproduced from Proussevitch and Sahagian (1996) (From Dingwell 1998).

Vesicles form as a consequence of volatile oversaturation (mainly H2O) in silicate


melts. The nucleation and growth of vesicles is controlled by pressure and
temperature. Again, low diffusion rates and very high viscosities of SiO2-rich melts
result in a time lag between a decompressional event during ascent and the
formation and growth of bubbles. Consequently, H2O oversaturation of the melt is
high and bubbles remain small in rapidly rising melts (Fig. 1.6). Once formed,
vesicles do not rise in viscous melts, however, in silica-poor melts vesicles rise
easily.

Fig. 1.7 Temperature profiles


of the Ben Lomond flow at
different time steps as derived
by numerical modelling with
the emplacement temperature
of 850C; Tg = glass transition
temperature, FVP = finely
vesicular pumice, U.OBS =
upper obsidian, TZ =
transition zone, RHY = stony
rhyolite, L.OBS = lower
obsidian, BRX = breccia (from
Stevenson et al. 2001)

2. Types of eruption
Eruption type strongly controls textures apparent in volcanic or volcanoclastic
products. The explosivity and the degree of fragmentation during eruption depends
on magmatic factors, such as volatile content, eruption rates, viscosity, and
deformation rate, as well as on the presence of external water at the vent
(groundwater, lake, sea, glacier) and the degree of water-melt interaction. Eruptions
in which external water is involved are called phreatomagmatic eruptions. The
classification of eruptions is depicted in Fig. 2.1.
Figure 2.2 schematically shows a highly explosive magmatic eruption. Above a
certain depth in the conduit (< 4 km), vesicles may nucleate in the rising melt due to
decompression (first boiling). During this process, embayed phenocrysts may
fragment if vesicles form inside the embayment (Fig. 2.3; Best and Christiansen
1997). Depending on the volatile content of the melt this process can culminate into a
melt foam with up to 75 vol% of bubbles. At this point, the yield strength of SiO2-rich
viscous melts is surpassed, due to the pressure inside the bubbles and due to high
deformation rates, and it starts to fragment. Above the fragmentation surface, a hot
low-viscosity dispersion of gas and fragments is present in the conduit which strongly
expands and accelerates upwards. The dispersion reaches the mouth of the conduit
with high, sometimes supersonic velocities and forms an eruption column.

Fig. 2.1 (a) D-F plot used to characterise different


types of pyroclastic fall deposit (after Walker 1973b,
and updated in Wright et al. 1980). (b) Cartoon
explaining D-F plot in terms of eruption column height
and explosiveness (after Cas & Wright 1987).

In

the

Fig. 2.2 Schematic diagram of a


explosive volcanic system showing
different regions and rheological
regimes from non-vesiculated magma
to eruption plume. (From Fisher &
Schminke 1984, after Wilson et al.,
1980).

lower

part of the eruption column (gas thrust region, Fig.


2.4) the upward movement of the dispersion
originates from the momentum generated in the
conduit above the fragmentation surface. In the gas
thrust region large volumes of ambient air get
Fig. 2.3 Quartz phenocryst in
a laccolithic rhyolite
fragmented during first
boiling; the resulting
fragments rotated to different
optical orientations and were
annealed; scale bar is 0.5 mm.

incorporated. In contact with the hot gas and


fragments, the air heats up and expands. Therefore
the dispersion gradually slows and expands to form
the

convective

region.

Here

cooler,

highly

expanded dispersion exists with a density lower than


that of the ambient atmosphere and which supports

the volcanic fragments by strong turbulence. The bouyant eruption column rises up to

9
the level where its density equals
that of the ambient atmosphere (up
to 65 km). Volcanic fragments fall
out of the column or are drifted by
winds in the respective atmospheric
levels

according

to

their

aerodynamic properties. When the


turbulent dispersion becomes too
dense the entire eruption column or
Fig. 2.4 A schematic volcanic eruption column,

parts

of

it

showing the variation of velocity with height and


relative importance of buoyancy and momentum.
Buoyancy carries column to height HB; lateral
spreading takes place above HB. Momentum drives
some material upwards to a maximum height HT
(Orton 1996, from Self & Walker, 1994).

pyroclastic

collapse
flows.

forming
Important

boundary conditions for column


collapse are depicted in Figure 2.5.
Magmatic
distinguished
explosivity

eruptions
according
into

are
to

their

hawaiian,

strombolian and plinian (Fig. 2.1).


Vulcanian

eruption

are

characterized by small to moderatesized volcanic outbursts which last


seconds to minutes. They form
above conduits in which viscous,
volatile-rich magma rises at a slow
rate and vesiculates.
When ascending magma gets in
contact with ground-, lake-, or
seawater

phreatomagmatic

eruptions may result (surtseyan,


phreatoplinian, Fig. 2.1). Depending
Fig. 2.5 Stability fields for convecting and
collapsing columns in terms of vent radius and
magmatic volatile content. Magma discharge rate
(right side) is largely a function of vent radius (left
side) whereas exit velocity (bottom) is largely a
function of volatile content (top) (Orton 1996, from
Wilson, Sparks & Walker, 1980).

on the depth of interaction with


magma

and

on

the

pressure

conditions, the water is heated up

10
(> 100C), becomes a supercritical fluid, or vapor. The hot fluid infiltrates the magma
along a fine network of cracks that form during movements or seismic shocks. A thin
vapor film (Leidenfrost phenomen) may thermally isolate the hot magma fragments
from the coolant, temporarily.

Fig. 2.6 Interrelations of explosive energy, water-magma ratio, style of volcanic activity and
volcaniclastic fragments in basaltic hydrovolcanic eruptions (largely after Wohletz & Sheridan,
1983). The smallest fragments are produced in Taalian eruptions when most thermal energy is
transferred to mechanical energy. The shape of shards produced (1-5) depends on the viscosity of
the magma and its degree of vesiculation: blocky shards (1) of poorly vesicular magma are most
common; irregular, globular and spherical shards (2-4) indicate fluidal melts; platy and cuspate
shards are part of vesicle walls and develop if vesiculated, generally more viscous magma
interacts with water. Non-explosive quench fragmentation can occur in any environment. For
instance, views of hyaloclastites are from the Mid-Atlantic ridge (Schmincke et al., 1978) and a
Permian example where lava flowed over unconsolidated nearshore marine sediments (From
Orton 1996).

If this vapor film breaks down vigorous explosions happen which send supersonic
shock waves through the magma. In the course of these true explosions (in the
definition of physics) and of other dynamic processes, the magma and eventually

11
material from the conduit wall are being strongly fragmented (high F-values in Fig.
2.1). Phreatomagmatic eruptions produce relatively cool eruption columns which,
compared to magmatic eruptions, do not rise very high into the atmosphere.
Therefore, dispersal of volcanic fragments (D in Fig. 2.1) remains low. Fig. 2.6
emphasises the water/magma ratio in phreatomagmatic eruptions of basaltic magma
and the type of resulting fragments.

3. Volcanic forms
Since volcanic textures are more controlled by eruption style than by the form of
volcanoes, the latter is only briefly considered here. The form of a volcano depends
on the prevailing eruption style(s), (volcano-)tectonic and surface processes (erosion,
land slides, vegetation, glaciation, marine abrasion...). There are monogenetic
volcanoes, which form during a relatively short time displaying one or two eruption

Fig. 3.1 Types of volcanic landforms. Vertical


exaggeration 2 to 1 (polygenetic) and 4 to 1
(monogenetic). Relative sizes are only
approximate (From Orton 1996, after Simkin et
al., 1981).
Styles.

And

polygenetic

there

are

volcanoes,

complex
which

or

have

Fig. 3.2 Schematic cross sections showing

complex structures and which are long-

differences between the three types of crater


formed by phreatomagmatic eruptions and for
maars also by phreatic activity (After Cas &
Wright 1987).

living (up to millions of years; Fig. 3.1).


Scoria cones, tuff rings, tuff cones and

12

Fig. 3.3 Schematic cross section through an extinct maar volcanoe


emphazising the diatreme architecture; From Volker Lorenz, 2004

maars are monogenetic volcanoes which are typical for subaerial, explosive, SiO2poor intraplate volcanism. Scoria cones, the most abundant terrestrial volcanic form,
are built up by the fall out products of strombolian eruptions (Fig. 2.1). Tuff rings, cones and maars are the products of different phreatomagmatic eruptions (Fig. 3.2)
the type depending mainly on the depth of magma-water interaction. In particular
during maar-forming eruptions, the volcanic explosions cut deep into the country rock
leading to the formation of a diatreme (Fig. 3.3). The architecture of a diatreme and
its internal facies is controlled, among others, by the soft vs. hardrock nature of the
host material.

13

SiO2- and volatile-poor magmas tend to produce low- to non-explosive hawaiian


eruptions, when external water is absent or scarce. Here, high magma production
rates lead to the formation of shield volcanoes or plateau basalt fields. SiO2-rich
but volatile-poor magmas (such as rhyolitic, phonolitic) can extrude during the final
eruption stage of a complex volcano, but they also can form individual monogenetic
volcanoes. In both cases, lava domes, mesa lavas, coules or lava flows form,
depending on viscosity, magma production rate and topography. Complex volcanoes
are generally restricted to geotectonic settings where large amounts of intermediate
to SiO2-rich magmas are produced in one place over long time, such as at
convergent plate margins and hot spots on continental lithosphere.
Stratovolcanoes
typically form from the
eruption of intermediate
magmas.

Pyroclastic

deposits from explosive


(phreato-)plinian
tions,
well

erup-

lava(-domes)
as

as

volcanoclastic

sediments built up a cone


with steep slopes. Within
the

stratovolcano

complex

network

conduits,

dykes,

a
of
sills,

Fig. 3.4 General caldera cycle (after Lipman, 1984). Stage 1

stocks and hydrothermal

precaldera volcanism develops clusters of small intermediate


stratovolcanoes, Stage 2 eruption of zoned magma chamber
develops caldera. Ash flow tuffs interfinger with caldera collapse
breccia whereas a thin outflow sheet extends outward from the
caldera, Stage 3 postcaldera deposition of volcanics and
sediment and resurgent doming (From Orton 1996).

circuits develops. As a
consequence,
canoes
experience

stratovol-

tend

to
edifice

failures (sector collapse) resulting in extended debris avalanches. The stratovolcano


proper (cone facies) is surrounded by the volcanic ring plain. With ancient volcanic
successions, the cone facies is typically strongly eroded and alterated, whereas in

14
the volcanic ring plain facies the evolution of the former stratovolcano is preserved
best.

Fig. 3.5 Evolution of Scafell Caldera, English


Lake District (after Branney & Kokelaar, 1994).
The caldera developed atop basaltic to andesitic
lavas (e.g. Lingcove Fm.) that formed a
composite low-profile shield-like volcano.
Schematic section from the Langdale area shows
relative thickness of facies from the various
stages. These are: A emplacement of
Whorneyside ignimbrite and initial subsidence; B
inundation of vent leads to phreatoplinian
eruptions of Whorneyside bedded tuff; C onset of
widespread piecemeal subsidence and eruption of
Long Top Tuffs; D continued subsidence and
deformation of hot ignimbrites; E eruption of
high-grade ignimbrites of Crinkle Crags tuffs; F
development of a caldera lake, with subaqueous
volcaniclastic sediments and tuffs, and intrusion
of rhyolite domes (From Orton 1996).

Very large eruptions from mid- to upper


crustal magma chambers can result in
subsidence of the chamber roof, and a
caldera forms. The round to elliptic

shape of many calderas is related to ring fractures that form prior or during the initial
eruption (Fig. 3.4). Other complexes have a network of faults which are used as
magma conduits and along which vertical block movements take place during
eruption (piecemeal caldera, trapdoor caldera, Fig. 3.5).

15
When basaltic lava flows into a lake or the sea, strong explosions can occur. The
resulting fragments sometimes pile up to litoral cones, which of course do not have
magmatic roots.
Considering magma production rates, subaquatic volcanism is the most important
volcanic activity on earth. It includes volcanism at the mid ocean ridges (MOR), at
seamounts, in lakes and, as a special case, below ice. In comparison to subaerial
volcanoes, the main difference is the abundance of external water. Thus
phreatomagmatic and other hydromagmatic processes are dominant. Consequently,

Fig. 3.6 Schematic illustration of characteristics of volcanoes and the central rift of mid-oceanic
spreading ridges with different spreading rates (From Cas & Wright 1987, after Macdonald 1982).

subaquatic

volcanism

associated

with

abundant

volcanoclastic

mass

which

the

carry

magmatic
fragments

and

is
flows,

phreatohyaloclastic

downslope

the

volcanic edifice and beyond. In


water depth below 500 1000
m, the hydraulic pressure is
generally

high

enough

to

Fig 3.7 Sketch of summit area of a seamount near the


East Pacific Rise (From Orton 1996, after Londsale &

prevent nucleation of vesicles in

Batiza, 1980).

ascending

magma

and

16
outflowing lava (pressure compensation level, PCL), thus explosive volcanism and
associated explosive fragmentation is absent in deep ocean regions (deep water
stage, Fig. 3.10). However, recent studies indicate the existence of deep marine
explosive volcanism (e.g. Gill et al. 1990). Furthermore, the presence of watersaturated soft sediment is a special feature in many subaquatic volcanic sites.
Contact of magma/lava with wet sediments may lead to the formation of peperites
(see next chapter).
Fig. 3.6 depicts volcanic forms at fast and slow spreading mid-ocean ridges (MOR).
At MORs, and other subaquatic volcanic sites, low production rates of basaltic
magma leads to the formation of pillow lavas and pillow breccias (Fig. 3.7). High
rates result in lava sheets. Intermediate to SiO2-rich subaquatic volcanism produces
complex structures of domes, cryptodomes, sills, dykes and abundant fragmentation
(Fig. 3.8). These settings are also of large economic importance (e.g. Kuroko-Type,
volcanic hosted massive sulfide deposits etc., Allen 1992). Fig. 3.9 highlights the very
complex processes and textures which can develop in englacial volcanoes (e.g. in
Iceland).

Fig 3.8 Various forms of subaqueous silicic lavas and domes. (a) Intrusive and
partially extrusive domes. (b) Vent-top submarine dome. (c) Lava lobehyaloclastite complex (From Orton 1996, collated by Cas, 1992).

17

Fig. 3.9 A-D Sketches illustrating the environmentally controlled eruptive


processes of an active englacial volcano in a narrow icebound lake (From
Werner & Schminke, 1999).

Fig. 3.10 Evolution of a volcano on oceanic lithosphere; the


curves depict the change of the height of the edifice with time
(modified after Schmincke and Bogaard 1991)

18

4. Volcanic fragmentation and fragments


Volcanic activity is associated with different forms of fragmentation, which will be
described briefly in this chapter. In addition, all other fragmentation processes which
are known from non-volcanic areas to take place during weathering and erosion, may
also occur in volcanic zones. These so-called epiclastic processes are not
considered specifically in this course.
The following volcanic fragmentation processes can be distinguished (adapted from
R. Cas 1991, IAVCEI Commission on volcanogenic sedimentation, Newsletter 4):

pyroclastic
magmatic
phreatomagmatic*
phreatic*
hydrothermal*

quench clastic or hyaloclastic*

hydrauliclastic*

autoclastic

All processes marked * involve magma-water interaction, and therefore are grouped
under the term hydroclastic fragmentation by many authors (Fig. 4.1). Pyroclastic
fragmentation

comprises

all

explosive

volcanic

processes.

These

include

fragmentation of melt by bubble expansion, phreatomagmatic/phreatic explosions


and by expansion of vapor (Chapter 2). Swift deformation of the erupting material and
particle

collisions

cause

further

fragmentation.

Phenocrysts

with

melt-filled

embayments can experience fragmentation by bubble formation inside the


embayments (Fig. 2.3). Active geothermal fields may be subject to eruptions, e.g.
triggered by earthquakes. The seismic shocks allow the hot or supercritical pore
fluids to expand to vapor. The associated fragmentation of fresh and alterated rocks
in the geothermal field is called hydrothermal.

19

Fig. 4.1 Processes by which subaerial hydroclastic flow and fallout deposits
originate (From Fisher & Schminke 1984).

Quench clastic or hyaloclastic fragmentation is caused by the volume decrease


associated with the rapid cooling of melt in contact with water (e.g. on the surface of
subaquatic lava flows, pillow lavas) or air. In basaltic melts, quenching can cause
rapid crystal growth (skeletal growth) into characteristic forms (e.g. swallowtail
plagioclase). Hydrauloclastic fragmentation results from forceful intrusion of
hydrothermal fluids and usually produce complex in situ vein network breccias. When
magma intrudes wet sediments a combination of quenching, phreatomagmatic and
hydrauliclastic fragmentation may take place. A mixture of magmatic fragments and
deformed sediment (or vice versa) forms which is called peperite (Kokelaar 1982,
Hanson & Wilson 1993).
Autoclastic fragmentation can occur in moving magma, lava or dense pyroclastic
flows. E.g. the cooler top of a flowing lava, which may consist of crystallized or glassy
material or of melt with higher rigidity (see Fig. 1.7), can break due to gravity or
hydraulic force while the inner, more fluid part flows in a ductile manner.
Pyroclastic fragments (or pyroclasts) comprise highly vesicular glass (pumice),
bubble wall fragments (glass shards), less vesicular glass (scoria), crystal(fragment)s, crystals with attached glass, and lithics (Fig. 2.6). Material that formed in

20
the magma is called essential or juvenile. Lithics comprise any other volcanic or
non-volcanic material that did not form in the erupting magma and which was
incorporated into the magma before and during eruption, or picked up by a lava flow,
a pyroclastic surge or flow. A grainsize classification of pyroclasts is given in Table
4.1:

Table 4.1 Granulometric classification of pyroclasts and of unimodal, well-sorted pyroclastic


deposits (From Fisher & Schminke 1984, after Schmid, 1981).
During (phreato-)magmatic eruptions large dense fragments can travel from the vent
on ballistic trajectories through the air (ballistic bombs). In the course of this, molten
fragments may vesiculate, leading to the formation of breadcrust bombs, or they
may get rounded aerodynamically. Accretionary lapilli consist of concentric layers
of fine ash, sometimes with a larger fragment in the centre. They form in moist ash
clouds during phreatomagmatic eruptions and can be found associated with
subaerial pyroclastic fall out, surge and flow deposits (Schumacher and Schmincke
1991, 1995). Similarily, amoured lapilli have a lapillus as a core, covered with layers
of ash.

Volcanic glass and its alteration


The formation and preservation of volcanic glass during eruption and cooling of
silicate melts depends on cooling time and diffusion rates. SiO2-poor melts can

21
crystallize very quickly. Thus, most basalts have a more or less fine-grained
crystalline groundmass. Only upon very quick cooling in direct contact with water or
air, basaltic glass forms (sideromelan). SiO2-poor glass which is clouded by Fe- and
Ti-oxide microliths is called tachylite. During eruption of SiO2-rich melts glassy
pumice, glass shards and glassy lava (obsidian) may form. The different colours of
pumice and obsidian originate from microliths that form during early stages of
eruption (Paulick and Franz 1997). Also some pyroclastic flow deposits can have
glassy zones, called vitrophyre. If not devitrified during cooling (s. Chapter 6),
volcanic glass is generally subject to quick alteration during weathering and
diagenesis/hydrothermal activity. However, there are some rare examples of
Mesozoic and even Paleozoic glass. Exposed to water, glass takes in slowly water
(up to 7 wt% and more), this process is called hydration. During first stages of
subaquatic alteration, sideromelan transforms to silicate gel or fibrous minerals,
called palagonite (Fig. 4.2). Crystallisation of volcanic glass to clay minerals (mainly
smectite, chlorite, illite) is favored under conditions of high (pore-)water circulation.
The presence of stagnant, highly saline and alkaline pore water favors the formation
of zeolites. Other abundant alteration products are quartz, albite, calcite, and ore
minerals.

Fig. 4.2A-D Sequence of three alteration stages (B to D) of tephritic phonolite glass


from the Pliocene Roque Nublo Formation (Gran Canaria, Canary Islands) (From
Fisher & Schminke 1984, after Brey and Schmincke, 1980).

22

5. Volcanic transport and deposition: Lava and pyroclastics


Volcanic transport and deposition/emplacement comprise coherent flow, such as lava
(and some high-grade pyroclastic flows) and clastic processes (pyroclastic fall, surge, and flow).
Lava flow
The different forms of lava flow
and the resulting textures are
strongly dependent on viscosity,
flow rate, flow temperature and
relief. SiO2-poor lava flows are
associated with hawaiian, and
subordinately
eruptions.

with

strombolian

Low-viscosity

flows

develop

smooth,

sometimes

wrinkled

surfaces

(pahoehoe

lava,

ropy

autobrecciated

lava),
tops

rough
form

on

viscous lavas (aa lava; Fig. 5.1).


Fig. 5.1 Longitudinal sections through the two main
types of subaerial basaltic lava flow. (From Cas &
Wright 1987, after Lockwood & Lipman 1980).

In

SiO2-poor

lava

flows,

differences in viscosity stem from

different chemical compositions, temperature or different microlith content (Cashman


et al. 1999). During flow and cooling, volatile oversaturation causes nucleation and
growth of vesicles, also called amygdales. During late stage of crystallisation, small
domains of melt differentiation and segregation may form (Anderson et al. 1984). In
lava fields with high magma production rates (e.g. plateau basalts) individual lava
sheets can be considerably inflated by later inflow, a process which should be
considered during petrologic studies (Fig. 5.2).
The textures which form in SiO2-rich lava and lavadomes (mainly rhyolitic-dacitic,
phonolitic-trachytic) are of particular interest, since in ancient volcanic zones the
distinction between SiO2-rich lavas and high grade ignimbrites is a difficult task.
Scientific drilling through subrecent lavas and their conduit systems revealed that
many SiO2-rich melts vesiculate during ascent in depths of < 4 km (first boiling),

23
letting escape much of the volatiles into porous sediment or volcanic breccia
(Eichelberger et al. 1986). Ascending further, the melt foam collapses again (see also
Manley 1996).

Fig. 5.2 Schematic cross sections of emplacement of a generic inflating pahoehoe sheet
flow. Vertical scale varies from 1-5 m for Hawaiian flows to 5-50 m for the CRB Flows
(CRB = Columbia River Basalt). (a) Flow arrives as a small, slow-moving, lobe of
molten lava held inside a stretchable, chilled viscoelastic skin with brittle crust on top.
Bubbles are initially trapped in both the upper and basal crusts. (b) Continued injection
of lava into the lobe results in inflation (lifting of the upper crust) and new breakouts.
During inflation, bubbles rising from the fluid core become trapped in the viscoelastic
mush at the base of the upper crust, forming horizontal vesicular zones. The growth of
the lower crust, in which pipe vesicles develop, is much slower. Relatively rapid cooling
and motion during inflation results in irregular jointing in the upper crust. (c) After
stagnation, diapirs of vesicular residuum form vertical cylinders and horizontal sheets
within the crystallizing lava core. Slow cooling of the stationary liquid core forms more
regular joints. (d) Emplacement history of flow is preserved in vesicle distribution and
jointing pattern of frozen lava (From Self et al. 1997).

Moving viscous melts develop flow foliation in the conduit and during emplacement.
In contrast to intrusions, which have a closed onion-like flow foliation architecture,
subaerial lava(-domes) have complex flower- or bowl-shaped flow structures (opensurface system; Figs. 5.3, 5.4 and 5.5). Early (high T) flow foliation develops a variety
of textures such as gradation in phenocryst content and size, and alignment of
phenocrysts, vesicles and lithics. Upon further cooling, deformation becomes more
and more resctricted to discrete foliation planes which experience wrinkling, folding,
tearing and brecciation (look again at Figs. 1.4 and 1.7!).

24
Due to the extreme rigidity, extended top and basal breccias form by
autobrecciation (Fink and Griffiths 1998). Already during outflow and later, during
final emplacement and cooling of the lava, remaining volatiles exsolve from the melt
(second boiling) and a three-fold layering can develop discordantly to the flow
banding: Often, the inner part is coarsely vesiculated (coarsely vesicular pumice,
CVP) and is overlain by dense non-vesicular glass (obsidian), the top is finely
vesicular (finely vesicular pumice, FVP; Fig. 5.6). Since the obsidian and FVP
layers

are

denser

than

the

underlying CVP, conditions of


Raleigh-Taylor-Instability

exist,

which can give rise to diapiric


movement

of

CVP

and

fragmentation in all layers (Figs.


5.7, 5.8).

Fig. 5.3 (a) and (b) Cross sections through two of the
rhyolite domes in southern Lipari. Horizontal scale same
as vertical. (c) 3-dimensional view of the growth of a
rhyolite dome (From Cas & Wright 1987, after
Richardson 1978).

Fig. 5.4 Cross section through the length of the Rocche Rosse obsidian coule, with generalised
flow foliation patterns (From Cas & Wright 1987, after Hall 1978).

25
Fig. 5.6 Cross section through four

Fig. 5.5 Schematic diagram showing development of


foliation attitudes in vent area of dome. (1) Viscous dome
emplaced. Shallow surface fractures develop. (2) Fractures
nearest center deepen preferentially. (3) Fractures propagate
inward as lava spreads laterally, causing most of upper
surface to become a fracture surface. (4) Later stage of
growth. Flows have developed. Most of flow still capped by
fracture surface. Compression during flow forms surface
folds. Flow stratigraphy not indicated. (5) Detail of vent area
showing uplift and outward rotation of blocks as lava
continues to rise (From Fink 1983).

rhyolite
flows.
Column
A:
Traditional view of rhyolite flows
(thickness approximate)(modified after
Christiansen and Lipman, 1966;
Williams and McBirney, 1979; and
Ekren et al., 1984). Columns B-E:
Textural zonation in flows discussed in
this article. FVP = finely vesicular
pumice; OBS = obsidian; CVP =
coarsely vesicular pumice; RHY =
lithoidal rhyolite. Except in Middle
Dome, FVP zone is divided into
breccia (above) and coherent rock
(below), and basal breccia into welded
(black matrix; above) and nonwelded
(open matrix; below). Internal breccias
(not shown) occur extensively in RHY
layer in Middle Dome and in several
thin bodies just above RHY layer in
Banco Bonito flow. Diagonal lines
between columns show correlations of
textural zones among various flows
(From Manley & Fink 1987).

26

Fig. 5.7 Rhyolitic obsidian flow profiles. (a)Stratigraphy for a typical 35 m thick rhyolitic
obsidian flow. (b) Density profile, based on measured densities of samples of coarse pumice,
obsidian, and fine pumice. Note density inversion at contact between obsidian and coarse
pumice. (c) Temperature profile, assuming constant internal temperature. Note steep surfacetemperature gradient. (d) Viscosity profile, based on (b), (c), and laboratory viscosity
measurements. Note rapid decrease of viscosity with depth near the upper surface (After
Friedman and others, 1963, from Fink 1983).

Fig. 5.8 Schematic diagram


showing simultaneous rise of
coarse pumice diapirs and
inward propagation of fractures.
Fracture axis corresponds to
former anticlinal axis and lies
parallel to flow direction (From
Fink 1983).
.

27
Pyroclastic transport and deposition
During (phreato-)magmatic eruptions three types of pyroclastic transport and
deposition occur separately or sometimes synchroneously (see Figs. 5.9, 4.1):

Pyroclastic
o - Fall
o - Surge
o - Flow
Fig. 5.9 Processes
by which subaerial
pyroclastic flow and
fallout
deposits
originate
(From
Fisher & Schminke
1984).

Pyroclastic fall out particles are called tephra and include ballistic bombs directly
from the vent and material falling out from the eruption column or from ash clouds
associated with pyroclastic flow and surge. Fall out deposits are relatively well-sorted,
clast-supported, and massive to parallel-bedded. Sometimes proximal tephra is hot
enough during deposition to weld together (fused tuff, or lapilli stone). Since
aerodynamic fractionation of the tephra occurs during transport in the eruption
column and during wind drift, caution has to be taken when different fall layers are
compared petrographically or chemically. Fallout deposits from tephra that settled
through the water column (sea, lake), may display prominent grading (normal with
dense fragments, inverse with vesicular fragments).
Pyroclastic mass flows have a large spectrum of forms. One end member of this
spectrum are surges. Base surges form during phreatomagmatic eruptions from
eruption column collapse or from lateral blasts directly at the mouth of the vent.

Fig. 5.10 Classification 28


of base surge bedform
and
internal
crossstratification variations
related to depositional
rate (relative to transport
rate; vertical axis) and
surge temperature and
moisture
content
(horizontal axis); flow
comes from left (From
Cas & Wright 1987,

Fig. 5.11 Proximal to distal (LFS) and vertical (VFS) facies variation in pyroclastic surge
beds (i.e. single depositional units) on Songaksan tuff ring (after Chough & Sohn, 1990).
The lateral facies sequence (LFS 1) was distilled from common downcurrent facies
transitions in flank deposits whereas vertical facies sequences (VFS) are distilled using
Harpers (1984) method of facies sequence analysis. VFS1 and VFS2 are from proximal
near-vent deposits, probably where short-lived pyroclastic surges were overladen by
suspended sediment fallout (compare with Lowe, 1988). LFS1, and its vertical expression
(VFS3) indicates downcurrent decrease in particle concentration, grain size, and
suspended-load fallout rate with a resultant increase in traction and sorting processes (from
Orton 1996).

29
These are relatively cool (100 400C), dilute, highly turbulent and moist dispersions
which travel with high speed over the landscape, normally less than 15 km. Surge
deposits are ill-sorted, parallel to cross-bedded. Surges can form antidunes (Fig.
5.10, 5.11). Since the degree of fragmentation is high in phreatomagmatic eruptions
(remember Fig. 2.1), many base surge deposits are characterized by a high portion
of fine ash. Ground surges and ash cloud surges are associated with pyroclastic
flows (see below).
Pyroclastic flows are dense, turbulent to laminar, hot mixtures of pyroclastic
fragments, gas and entrained air. The deposits are normally ash-rich, very ill-sorted,
massive to faintly parallel-bedded (Fig. 5.12). They originate from eruption column
collapse during (phreato-)magmatic eruptions, or from explosive or gravitative
collapse of lavadomes and lava fronts (Fig. 5.13). While moving, the pyroclastic flow
entrains ambient air at the front and looses hot gas and fine ash at its top (leading to
pyroclastic fractionation), the latter giving rise to an ash cloud manteling the flow
(Fig. 5.14). At the base of the ash cloud, lateral bursts may develop which form ashrich, parallel to cross-bedded ash-cloud surge deposits. Surges of fine ash and gas
can burst out at the front giving way to the formation of fine parallel- to cross-bedded
deposits (ground surge deposit). Pumice-rich flows are associated with plinian
eruptions and with many caldera-forming eruptions. The deposits are called
ignimbrites.

Fig. 5.12 Idealized products of subaerial pyroclastic flows. P, pumice clasts; L, lithic clasts. Note
concentration of pumice at the top of the flow unit and lithics at the base in all cases (from Sparks,
Self & Walker, 1973; Sheridan, 1979; Fisher & Heiken, 1982; Branney, Kokalaar & Mc Connell,
1992). Note absence of Layer 1 deposits from block and ash flows and lava-like ignimbrites (From
Orton 1996)

30

Fig.

5.13

Mechanisms
generating
pyroclastic
flows. The pyroclastic
flow proper is a high
particle
concentration
underflow. The ash cloud
gives rise to other deposits
(From Cas & Wright
1987).

Fig. 5.14 Schematic


diagram showing the
structure and idealised
deposits
of
one
pyroclastic flow (From
Cas & Wright 1987).

31
Flows that form as the consequence of lava(dome) collapse are called block-andash-flows. Since the classic work of Ross and Smith (1961) on ignimbrites, the
concepts of pyroclastic flows have developed considerably (e.g. Chapin and Elston
1979,

Branney

and

Kokelaar

2002).

Among

others, Ross and


Smith developed
the model of flow
units

and

cooling

units.

Deposits from a
rapid succession
of

pyroclastic

flows behave as
one cooling unit
(see

Fig.

5.15

and chapter 6).


Fig. 5.15 Cross section through part of the Upper Bandelier Tuff, showing
welding and crystallisation zones. The ignimbrite is a compound cooling
unit, and shows an upward increase in the degree of welding in cooling
units I-III. Recognisable flow units are much thinner in units IV and V, and
nearer the source they pass into densely welded tuff, continuing the trend
towards higher temperature of emplacement of successive pumice flows
that is more clearly shown by units I-III. Note, the topography that the
ignimbrite fills in is cut into older ignimbrites and basement, including
Precambrian (cross section is approximately normal to movement direction
of the pumice flows) (From Cas & Wright 1987, after R. L. Smith & Bailey
1966).

Today,

distinction

is

made

between

the flow unit and


the depositional
unit (similar to
the

concept

of

turbidity currents

and turbidites). Branney and Kokelaar (1992) emphasized that there is a spectrum
ranging from highly turbulent expanded flows (EF) to dense high-grade, sometimes
lava-like flows (DF). At the base of the former, particles are sedimented continuously
(aggradation), whereas the latter moves as a viscous mass, the movement of which
eventually freezes to form a highly welded rheomorphic ignimbrite (Fig. 5.12).
Kobberger and Schmincke (1999) disagreed in this point, presenting an
aggradational model for a rheomorphic ignimbrite on Gran Canaria (Figs. 5.16, 5.17).

32

Fig. 5.16 Steps of the depositional and rheomorphic flow history of ignimbrite D by sketching the major
stages of deposition including the successive vertical changes of pyroclast strain, the development of fabrics
crucial to define the related deformation mechanism, and the progression in ascending and descending
lithification fronts. The model is based on the configuration of a gently but evenly inclined basal topography
during deposition and rheomorphic flow (From Kobberger & Schminke 1999).

Fig. 5.17 Model of sedimentation and initiation of rheomorphic flow during the early accumulation
phase of ignimbrite D on a slightly inclined basal topography. Note that pure flattening (compaction) is
the first step of deformation affecting the freshly deposited pyroclastic material in both cases. Shear
flow, in this configuration, is always secondary and depends on a critical load pressure as well as on the
speed of the upward prograding cooling/lithification front (From Kobberger & Schminke 1999).

33

Fig. 5.18 Proximal-distal depositional facies model for the products of ignimbrite-forming
eruptions based on the Bandelier Tuffs (From Cas & Wright 1987, after Wright et al. 1981).

Deposits from expanded flows (and from base surges: Fig. 5.11) display longitudinal
and lateral facies differentiations. Near the vent, large dense blocks which cannot
be transported by the flow form a coarse-grained lag-deposit (co-ignimbritic breccia,
Fig. 5.18). Thickness and, to a certain extend, grain size decrease distally and fine
ash deposits become important. When pyroclastic flows move down a valley, the
resulting deposits are thick and rich in large clasts (pumice, lithics) in the valley
whereas on the slopes the deposits get finer and eventually pinch out. Upon
deposition, the unit starts to deflate and cool, a process which can take month and
years (see next chapter). Additional gas is released from the cooling and eventually
devitrifying glass fragments. If the pyroclastic fragments are hot enough during
deposition primary welding can occur. Deposits from cooler flows can also develop
zones of welding where the rising gas lowers the glass transition temperature
(secondary welding). In both cases pumice and glass shards stick together and
compact (welding-compaction). Stretched or sheared pumice are called fiamme
(Fig. 5.17). Moderate welding-compaction produce an eutaxitic texture, textures
from strong welding-compaction are called parataxitic (Fig. 5.19a).
Above the welded zone, tiny crystals e.g. of cristobalite, tridymite, and feldspar grow
from the rising gases (vapor-phase crystallization, Figs. 5.15, 5.20) and lead to a
rapid solidification of large parts of the deposit (sillar facies). Frequently, the rising
gas concentrates in discrete pipes in which the fine ash is blown away and lapilli are
left behind (degassing pipes, Fig. 5.21). Around degassing pipes and in zones of
enhanced gas flow, vapor-phase crystallisation leads to a strong solidification of the
deposit (fumarole mounds, Fig. 5.20). Due to volume decrease upon cooling, welded
and sillar zones frequently develop columnar jointing.

34

Fig. 5.19 A Illustration of the welding profile in the Garth Tuff. B Idealised profile for vapourphase crystallisation in the Garth Tuff. C Idealised profile for spherulitic and granular crystalline
fabrics in the Garth Tuff. Boundaries between zones are gradational. D Timing of textural
development relative to the cooling history of the Garth Tuff (From Mc Arthur et al. 1998).

35
Fig. 5.21 Occurences of gas segregation
structures in pyroclastic flow deposits. 1, Pipes
and pods generated initially or formed entirely
by intraflow gas sources during emplacement;
1a, formed by continued post-emplacement gas
flow; 2, formed from heated ground water and
incorporating fluviatile pebbles; 3, formed
above burnt vegetation and logs (From Cas &
Wright 1987).

Fig.

5.20 Zones of vapour-phase


crystallisation and devitrification in the
Bishop Tuff. Fumarole mounds project from
top of vapour-phase zone through nonwelded ash (From Cas & Wright 1987, after
Sheridan 1970).

In ancient volcanic successions, the distinction between low grade pyroclastic flow
deposits and volcanosedimentary mass flow deposits (lahars) is difficult to
impossible. The presence of degassing pipes, columnar jointing, welding and
devitrification textures (see below) and in situ cooling joints of large clasts (jig saw
fit) indicate a hot emplacement. Measurements of the anisotropy of the magnetic
susceptibility (AMS) in large clasts can help to prove a hot deposition.
There is much debate on whether pyroclastic flows can enter water or whether
pyroclastic flows can form and flow under water. Figs. 5.22 and 5.23 illustrate the
complex pyroclastic and sedimentary processes involved with explosive volcanism
near or under water (s. also Kokelaar & Kniger 2000, and White 2000).

36

Fig. 5.22 Processes by


which
subaqueous
pyroclastic flows and
fallout originate (From
Fisher
&
Schminke
1984).

Fig. 5.23 Schematic diagram


illustrating processes of a high
mass-discharge
subaqueous
explosive eruption. No relative
scales are implied (From Orton
1996, after Kokelaar & Busby,
1992).

Magma mingling and mixing


Prior to eruption, processes of magma
mingling and mixing may occur. In fact,
these processes often initiate eruptions.
Xenocrysts with reaction halos indicate
magma mixing. Magma mingling is
indicated by streaky lava and pumice,
and by pyroclastic deposits containing
two or more types of essential clasts
(pumice, lava).

37

6. Emplacement, cooling and alteration textures in SiO2-rich lava


and high grade ignimbrites
There is a large variety of textures to be found in SiO2-rich volcanic rocks and in
sediments containing volcanic fragments. Unfortunately, in high grade ignimbrites
and lavas (and subvolcanic intrusions!) many common textures develop which makes
a distinction difficult.
As mentioned above, magma can experience vesiculation during ascent (first
boiling). Second vesiculation (second boiling) can take place in densely welded parts
of ignimbrites and in lavas when the cooling melt or glass becomes volatileoversaturated (Fig. 6.1; see also 5.6 for lava, Fig. 5.19b for ignimbrite).
Processes of texture formation in SiO 2 -rich volcanics
during ascent, eruption, emplacement and cooling*
800C

600C

phenocryst growth
?

40 0C
200C
Autobrecciation (carap ace)

crys tallisa tion of spherulit es, lithophysae etc.

upon decom pression


('firs t boilin g'

upon cooling
'secon d boilin g'

10

200

cooling joint form ation


H ydration of glass
pe rlite formation under q uenchin g

microlith form ation


vesiculation

D iagenesis, hydro therma l


alt eration, tectonics

d evitrifica tion of glass, re crystallisation


o f spheru lites and pheno crysts

brittle-ductile transition of glass T g


(depen ds on composition; ap prox. 65% of T m elt)
*During em placem ent and cooling, a th ermal g radient form s
t hat causes a spacial zonation of the above p rocesse s.

O ther proc esses:


- vapor phase crystallisation
- free fluid proce sses (ju venile a nd m et eoric):
- hy drofra cturing
- shifting o f liquid us
- etc.

mod ified afte r L ofgren 197 1, Eich elbe rg er e t al. 1986, S wans on et al. 1 989, Dav ie s & Mc Phie 1996 , Manley 1996, McA rt hur et al. 1998

Fig. 6.1
Due to low diffusion rates, it take time and strong supercooling (T) to generate
crystals in rhyolitic glass. During early stages of cooling, microliths (tiny crystals of
Fe-, Ti-oxides, mafic or felsic silicates) form in SiO2-rich lava and in welded zones of
ignimbrites, which grow along flow or stretching foliation. In intermediate lavas (basic
dacites, andesites, trachytes) these keep growing to develop a more or less
completely crystallized groundmass (trachytic texture).

38
Further cooling can lead to partial or complete crystallisation of the glass. Quench
crystals (mainly tridymite, christobalite and/or feldspar) grow to spherulites or
axiolites, often using a crystal, a bubble, a flow foliation plane or a crack as a
nucleation surface (Fig. 6.2). Experiments (Lofgren 1971b) and textural observations
indicate that spherulites form in hot glass above Tg. Depending on growth rate and
nucleation density, they develop radiate or microcrystalline textures. Often fine
films of Fe-, Ti-phases form radiating between the crystals or they mark concentric

Fig. 6.2 A Axiolitic structure; spherulitic fibres radiate from a plane. B Fan spherulite with fibres
radiating from a point. C Bowtie (or wheat sheaf) spherulite. Two fan-like arrays are joined at their
apices. D Plumose spherulite showing extensive side branching. Unlike dendrites, branching does not
occur on crystallographic axes. E Spherical spherulite. F Pectinate texture defined by fine axiolites
growing inward from the walls of a juvenile pyroclast, in this case a tricuspate shard. Scale will vary
according to the size of the fragment. G Lithophysal structure with fibres radiating outward from a
central hollow. H Lithophysae with concentric hollows arranged parallel to the crystallisation front
(modified after Lofgren, 1974, from Mc Arthur et al. 1998).

growth fronts of the spherulites. In many cases these radiate or concentric ore
mineral textures survive later recrystallisation and are the only trace of the formerly
existing spherulites. The width of the quench crystals correlates positively with
formation temperature (Lofgren 1971a,b). In densely welded deposits, spherulites
and axiolites grow discordantly to the vitroclastic texture. However, in less- or nonwelded domains pectinate textures can develop with fibrous crystals growing
inwards from the glass shard surfaces (Fig. 5.19c).

39
In volatile-rich glass, crystallisation of spherulites leads to concentration of volatiles in
a halo at the crystallisation front. Many spherulites are associated with curvilinearshaped cavities which form prior or during spherulite formation. These so-called
lithophysae (McArthur et al. 1998, Fig. 6.2) may be filled by crystals during later
cooling or diagenesis. In contrast to vesicles which form due to volatile
oversaturation, lithophysae form due to tension in the glass.
Crystallisation can occur in the form of isolated spherulites, form domains with sharp
devitrification fronts or be complete (Fig. 5.19). It can occur in a single event or multiphase due to complex emplacement, cooling and fluid migration history (remember
that devitrification is controlled by T, which is dependent of temperature and volatile
saturation, Fig. 1.5). The central part of thick ignimbrites and lavas stays hot for a
long time since devitrification as any crystallisation is exothermic. Here the spherulitic
texture can recrystallize to a equigranular mosaic of quartz and feldspar (core
facies, see below; Lofgren 1970, McArthur et al. 1998, Fig. 5.19).

Fig. 6.3 Schematic sketches summarized from thin sections, showing the
relationship of (a) false nonwelded
shards to the original texture in
classical perlite, and (b) false welded
shards to the original texture in banded
perlite. To aid explanation, the portion
of the original perlitic texture marked
by heavy lines is repeated in the
diagrams of false shard textures. The
false shards may be preserved as
siliceous segments between phyllosilicate-altered
perlitic
fractures
(center), or as phyllosilicate altered
sections of the fractures themselves
(right) (from Allen 1988).

40
During cooling, perlite cracks can develop which are often marked by clay minerals.
Davies and McPhie (1996) presume that perlite forms when the hot rock/glass gets in
contact with external water. However, some aspects of perlite formation remain
unclear. In places, these perlite cracks can resemble false vitroclastic (pyroclastic)
textures (Fig. 6.3, Allen 1988, Doyle 2001).

Fig. 6.4 Series of sketches showing formation of fragmental


textures. A Upon final decompression at the Earths surface,
rhyolite lava inflates to a finely vesicular pumice grading
down into less-vesicular lava. Cooling contraction begins to
fracture the pumice. B Compression and extension during
flow emplacement further disrupt the lava flow carapace. C
Continued flow movement causes loose pumice blocks to
grind against each other, producing fine-grained debris [pumice fragments, ash shards
(from pumice), glass chips, and phenocryst fragments] which collects around and
beneath the blocks. D Boxes show locations of close-up views E-G. E Abrasion of
pumice blocks produced glassy debris; some of the finest ash may be carried away by
wind. F Accumulations of breccia deeper in the flow are compressed and deformed. G
Debris-filled fractures which extend below the limit of disrupted blocks can be
deformed and folded as flow advance continues. H Crystallization in the flow interior
can cause devitrification of the cooled upper portion of the flow by re-heating,
producing crystalline material which had brecciated when glassy (from Manley 1996).

False vitroclastic textures from perlite are not the only trap door for students of
ancient SiO2-rich volcanic rocks as many aspects of this chapter revealed. They can
also develop as the consequence of fracturing and alteration of lava (Fig. 6.4, and
Allen 1988). Frequently, a network of tuffbreccias is present in SiO2-rich lavas. They
probably form during phreatic eruptions when the lava covers a pond or wet
sediment. The tuffbreccia often includes obsidian fragments, and, if hot enough, this
clastic fabric may experience ductile deformation. Furthermore, SiO2-rich lavas tend

41
to develop vesicular textures during early stages of cooling (see above, Fig. 5.6).
Compaction of the vesiculated glass during cooling and burial, and autoclastic or
phreatic fragmentation during emplacement and cooling can lead to textures that are
similar to those found in ignimbrites. Table 6.1 summarizes the status of this ongoing
discussion.
Table 6.1: Diagnostic characteristics of felsic lavas and pyroclastic flow deposits
(from Paulick and Breitkreuz 2005; largely based on several previous compilations
provided by Allen (1988), Henry and Wolff (1992), Manley 1996, Manley and Fink
(1987), and McPhie at al. (1993)).
Lava:
core facies

Lava:
carapace facies

Pyroclastic flow
deposit

Facies association

Grading into carapace


facies.

Grading into autoclastic


facies at top and base.
Resedimented synvolcanic
breccia at the flanks.

Typically associated with


co-genetic, lithic and
crystal-rich ground surge
and ash-cloud surge
layers.*

Internal organisation

Coherent, homogeneously
porphyritic, homogenous
groundmass, locally with
flow banding.

Coherent to brecciated
homogeneously
porphyritic, heterogeneous
groundmass; autoclastic
breccia are poorly sorted,
strictly monomictic, nongraded with common jigsaw fit textures.

Vitriclastic, poorly sorted,


matrix supported, typically
density graded.
Organisation may vary
substantially within one
emplacement unit* and
may be obscured due to
welding and
rheomorphism.

Xenoliths/lithics

Rarely abundant,
concentrated in lenses.

Rarely abundant,
concentrated in lenses.

Typically common to
abundant.

Phenocysts

Evenly distributed.

Evenly distributed in clasts


and matrix.

Typically more abundant in


matrix than in
pumice/volcanic lithic
clasts due elutriation of
ash during flow.

Broken phenocrysts

Uncommon, optically
continuous.

Broken and dislocated


Common within the matrix
phenocrysts may be locally throughout entire unit.
present in autoclastic
breccia.

Groundmass

Homogeneous
microcrystalline
(equigranular) locally with
remnant ghost textures of
pre-existing spherulites.

Heterogeneous, glassy,
vesicular to pumiceous
and/or devitrified
(spherulites, lithophysae);
flow foliation is common.

Welding textures

Absent.

Welding textures may be


present in autoclastic
breccia, especially in the
basal breccia.

Abundant shards, crystals


and crystal fragments,
however vitriclastic may be
obscured due to welding
crystallisation and
alteration.
Typical: welding zonations,
rheomorphism may
eradicate welding textures
in thick units.

*: The internal organisation of pyroclastic flow deposits varies dramatically depending on erupted volume,
temperature, composition and facies (proximal to distal, central to marginal, valley pond to overbank) and are
variably controlled by the topography.

42
Concept of carapace and core facies
Chapters 5 and 6 revealed that quite similar textures may form in different types of
cooling (sub-)volcanic bodies, such as high-grade ignimbrite, lava and subvolcanic
intrusions. Distinction between these rocks is difficult, especially with scattered small
outcrops or drill cores. Faulting, deformation, erosion and metamorphism complicate
this attempt. At the same time, this distinction is crucial for the reconstruction of the
volcanic and stratigraphic evolution.
Cooling ignimbrite, lava and sill/laccolith develop a core facies with a homogeneous
more or less equigranular crystallized groundmass (Fig. 5.5, 5.15, 5.19). The core
facies is surrounded by the inhomogeneous carapace which is characterized by
domains of coherent glass and breccias. Partial or complete crystallisation with
spherulites and lithophysae, vesiculation and perlitic cracks leads to very
inhomogeneous groundmass textures. Lava and ignimbrites typically have a thick
carapace, the carapace/core ratio is high. In contrast, laccolith and sills have a small
marginal carapace facies (low carapace/core facies). This concepts also helps to
estimate the degree of erosion of a volcanic unit. In conglomerates rich in volcanic
clasts, a careful facies analysis may reveal a lot of information about the volcanic
activity in the source area.

43
7. References
Textbooks and special volumes:
Branney, M., and Kokelaar, P., (2002). Pyroclastic density currents and the sedimentation of
ignimbrites: Geol. Soc. London, Mem., v. 27, 143pp.
Breitkreuz, C. & Petford, N. (Eds.)(2004). Physical Geology of High-Level Magmatic Systems.- Geol.
Soc. Spec. Publ. 234, 253pp.
Cas, R. A. F. and Wright, J. V. (1987). Volcanic successions - Modern and ancient. London, Allen &
Unwin, 528pp.
Chapin, C. E. and Elston, W. E., (Eds.) (1979). Ash-flow tuffs, Geol. Soc. Amer. Spec. Pap., 180,
211pp.
Fink, J. H. (Ed.) (1987). The emplacement of silicic domes and lava flows. Geol. Soc. Amer. Spec.
Pap., 212, 145pp.
Fisher R.V. and Schmincke H.-U. (1984). Pyroclastic rocks. Berlin, Springer Verlag, 472pp.
Francis, P. (1993): Volcanoes - A planetary perspective.- Oxford Univ. Press, Oxford, 443pp.
Gifkins, C., Herrmann, W., and Large, R. (2005). Altered volcanic rocks - a guide to description and
interpretation:, CODES University of Tasmania, Australia, 275 pp.
Heiken, G. and Wohletz, K. (1992). Volcanic ash.- Univ. Calif. Press, Berkeley, 246pp.
McPhie, J., Doyle, M. and Allen, R. (1993). Volcanic Textures - A guide to the interpretation of textures
in volcanic rocks. Hobart, CODES Univ. Tasmania, 196pp.
Orton, G. J. (1996). Volcanic environments. In: Reading H. G. (ed.): Sedimentary environments Processes, facies and stratigraphy. Oxford, Blackwell Sci., 485-567.
Ross, C. S. and Smith, R. L. (1961). Ash-flow tuffs: Their origin, geologic relations and identification.USGS Prof. Pap. 354-F (Also available as reprint: New Mexico Geol Soc. Spec. Publ. 9: 81pp.
1980).
Schmincke, H.-U. (2004): Volcanism.- Springer, Berlin, Heidelberg, 324pp.
Self, S. and Sparks, R.S.J. (eds.)(1981). Tephra studies.- Nato Advanced Study Inst. Ser., C 75,
481pp.
Sigurdsson, H. (chief editor)(2000). Encyclopedia of volcanoes.- Academic Press, 1417pp.

Other references
Allen, R. L. (1988). False pyroclastic textures in altered silicic lavas, with implications for volcanicassociated mineralization.- Economic Geol., 83: 1424-1446.
Allen, R. L. (1992). Reconstruction of the tectonic, volcanic, and sedimentary setting of strongly
deformed Zn-Cu massive sulfide deposits at Benambra, Victoria.- Econ. Geol., 87: 825-854.
Anderson, A. T., Swihart, G.H., Artioli, G. and Geiger, C.A. (1984). Segregation vesicles, gas filterpressing, and igneous differentiation.- J. Geol., 92: 55-72.
McArthur, A. N., Cas, R. A. F. and Orton, G. J. (1998). Distribution and significance of crystalline,
perlitic and vesicular textures in the Ordovician Garth Tuff (Wales).- Bull. Volcanol., 60: 260-285.
Best, M. G. and Christiansen, E. H. (1997). Origin of broken phenocrysts in ash-flow tuffs.- Geol. Soc.
Amer. Bull., 109: 63-73.
Branney, M. J. and Sparks, R. S. J. (1990). Fiamme formed by diagenesis and burial-compaction in
soils and subaqueous sediments.- J. Geol. Soc. London, 147: 919-922.
Branney, M. J., Kokelaar, B. P. and McConnell, B. J. (1992). The Bad Step Tuff: a lava-like
rheomorphic ignimbrite in a calc-alkaline piecemeal caldera, English Lake District.- Bull. Volcanol.,
54: 187-199.
Branney, M. J. and Kokelaar, P. (1992). A reappraisal of ignimbrite emplacement: progressive
aggradation and changes from particulate to non-particulate flow during emplacement of high-grade

44
ignimbrite.- Bull. Volcanol., 54: 504-520.
Buesch, D. C. (1991). Changes in depositional environments resulting from emplacement of a largevolume ignimbrite.- SEPM spec. Publ., 45: 139-153.
Buesch, D. C. (1992). Incorporation and redistribution of locally derived lithic fragments within a
pyroclastic flow.- Geol. Soc. Amer. Bull., 104: 1193-1207.
Cashman, K. V., Thornber, C., and Kauahikaua, J. P., 1999, Cooling and crystallization of lava in open
channels, and the transition of Pahoehoe Lava to Aa: Bull. Volc., v. 61, p. 306-323.
Chapin, C. E. and Lowell, G. R. (1979). Primary and secondary flow structures in ash-flow tuffs of the
Gribbles Run paleovalley, central Colorado.- Geol. Soc. Amer., Spec. Pap., 180: 137-154.
Davies, B. K. and McPhie, J. (1996). Spherulites, quench fractures and relict perlite in a Late Devonian
rhyolite dyke, Queensland, Australia.- J. Volc. Geotherm. Res., 71: 1-11.
Dingwell, D. B. (1998). Magma degassing and fragmentation: recent experimental advances. In: A.
Freundt and M. Rosi (eds.): From magma to tephra - modelling physical processes of explosive
volcanic eruptions. Amsterdam, Elsevier: 1-24.
Doyle, M. G. (2001). Volcanic influences on hydrothermal and diagenetic alteration: evidence from
Highway-reward, Mount Windsor Subprovince, Australia. Econ. Geol., 96: 1133-1148.
Eichelberger, J. C., Carrigan, C. R., Westrich, H. R. and Price, R. H. (1986). Non-explosive silicic
volcanism.- Nature, 323: 598-602.
Fink, J. H. (1980). Surface folding and viscosity of rhyolite flows.- Geology, 8: 250-254.
Fink, J. H. (1983). Structure and emplacement of a rhyolitic flow: Little Glass Mountain, Medicine Lake
Highland, northern California.- Geol. Soc. Amer. Bull., 94: 362-380.
Fink J.H. and Bridges, N. T. (1995). Effects of eruption history and cooling rate on lava dome growth.Bull. Volcanol., 57: 229-239.
Fink, J.H. and Griffiths, R.W., 1998. Morphology, eruption rates, and rheology of lava domes: insights
from laboratory models. J. Geophys. Res., 103(B1): 527-545.
Fisher, R. V., Orsi, G., Ort, M. and Heiken, G. (1993). Mobility of a large-volume pyroclastic flow emplacement of the Campanian ignimbrite, Italy.- J. Volc. Geotherm. Res., 56: 205-220.
Freundt, A. (1998). The formation of high-grade ignimbrites, I: Experiments on high- and lowconcentration transport systems containing sticky particles.- Bull. Volcanol., 59: 414-435.
Green, J. C. (1989). Physical volcanology of mid-Proterozoic plateau lavas: The Keweenawan North
Shore Volcanic Group, Minnesota.- Geol. Soc. Amer. Bull., 101: 486-500.
Gill, J., et al. (1990): Explosive deep water basalt in the Sumisu Backarc Rift.- Science 248, 12141217.
Hanson, R. E. and Wilson, T. J. (1993). Large-scale rhyolite peperites (Jurassic, southern Chile).- J.
Volcan. Geotherm. Res., 54: 247-264.
Heiken, G. and Wohletz, K. (1987). Tephra deposits associated with silicic domes and lava flows.Geol. Soc. Amer. Spec. Pap., 212: 55-76.
Henry, C. D. and Wolff, J. A. (1992). Distinguishing strongly rheomorphic tuffs from extensive silicic
lavas.- Bull. Volcanol., 54: 171-186.
Henry, C. D., Kunk, M.J., Muehlberger, W.R., and McIntosh, W.C. (1997). Igneous evolution of a
complex laccolith-caldera, the Solitario, Trans-Pecos Texas: Implications for calderas and subjacent
plutons.- Geol.Soc. Amer. Bull., 109: 1036-1054.
Hildreth, W. (1983). The compositionally zoned eruption of 1912 in the Valley of Ten Thousand
Smokes, Katmai National Park, Alaska.- J. Volcan. Geotherm. Res., 18: 1-56.
Kobberger, G. and Schmincke, H.-U. (1999). Deposition of rheomorphic ignimbrite D (Mogn
Formation), Gran Canaria, Canary Islands, Spain.- Bull. Volcan., 60: 465-485.
Kokelaar, B. P. (1982). Fluidization of wet sediments during the emplacement and cooling of various
igneous bodies.- J. geol. Soc. London, 139: 21-33.
Kokelaar, P. and Busby, C. (1992). Subaqueous explosive eruption and welding of pyroclastic
deposits.- Science, 257: 196-201.
Kokelaar, P. and Kniger, S. (2000): Marine emplacement of welded ignimbrite: the Ordovician Pitts
Head Tuff, North Wales.- J. Geol. Soc. London, 157, 517-536.
Lofgren, G. (1970). Experimantal devitrification rate of rhyolite glass.- Geol. Soc. Amer. Bull., 81: 553-

45
560.
Lofgren, G. (1971a). Spherulitic textures in glassy and crystalline rocks.- J. Geophys. Res., 76: 56355648.
Lofgren, G. (1971b). Experimentally produced devitrification textures in natural rhyolitic glass.- Geol.
Soc. Amer. Bull., 82: 111-124.
Manley, C. R. and Fink, J. H. (1987). Internal textures of rhyolite flows as revealed by research
drilling.- Geology, 15: 549-552.
Manley, C. R. (1995). How voluminous rhyolite lavas mimic rheomorphic ignimbrites: Eruptive style,
emplacement conditions, and formation of tuff-like textures.- Geology, 23: 349-352.
Manley, C. R. (1996). In situ formation of welded tuff-like textures in the carapace of a voluminous
silicic lava flow, Owyhee County, SW Idaho.- Bull. Volcanol., 57: 672-686.
Marsaglia, K. M. (1993). Basaltic island sand provenance.- Geol. Soc. Amer. Spec. Pap., 284: 41-65.
Marshall, R. R. (1961). Devitrification of natural glass.- Geol. Soc. Amer. Bull., 72: 1493-1520.
McArthur, A. N., Cas, R. A. F. and Orton, G. J. (1998). Distribution and significance of crystalline,
perlitic and vesicular textures in the Ordovician Garth Tuff (Wales).- Bull. Volcanol., 60: 260-285.
Moore, I. and Kokelaar, P. (1998). Tectonically controlled piecemeal caldera collapse: A case study of
Glencoe volcano, Scotland.- Geol. Soc. Amer. Bull., 110: 1448-1466.
Paulick, H., and Franz, G. (1997). The color of pumice: case study on trachytic fall deposit, Meidob
volanic field, Sudan: Bull Volcanol, v. 59, p. 171- 185.
Paulick, H. & Breitkreuz, C. (2005): The Late Paleozoic felsic lava-dominated large igneous province
in North East Germany: volcanic facies analysis based on drill cores.- Int. J. Earth Sci. 94.
Reidel, S. P. and Tolan, T. L. (1992). Eruption and emplacement of flood basalt: An example from the
large-volume Teepee Butte Member, Columbia River Basalt Group.- Geol. Soc. Amer. Bull., 104:
1650-1671.
Rosi, M. (1992). A model for the formation of vesiculated tuff by the coalescence of accretionary
lapilli.- Bull. Volcanol., 54: 429-434.
Schmincke, H.-U. and Bogaard, P. van den (1991). Tephra layers and teprha events.- In Einsele, G,
Ricken, W. and Seilacher, A. (Eds.): Cycles and events in stratigraphy.- Springer Verlag, 392 429.
Schumacher, R. and Schmincke, H.-U. (1991). Internal structure and occurrence of accretionary lapilli
- a case study at Laacher See Volcano.- Bull. Volcanol., 53: 612-634.
Schumacher, R., Schmincke, H.-U. (1995). Models for the origin of accretionary lapilli.- Bull.
Volcanol., 56: 626-639.
Self, S., Thordarson, T. and Keszthelyi, L. (1997). Emplacement of continental flood basalt lava flows.
In: J. J. Mahoney and M. F. Coffin (eds.): Large igneous provinces - continental, oceanic, and
planetary volcanism. Geophys. Monograph. 100: 381-410.
Sheridan, M. F. (1970). Fumarolic mounds and ridges of the Bishop Tuff, California.- Geol. Soc. Amer.
Bull., 81: 851-868.
Sheridan, M. F. and Ragan, D. M. (1976). Compaction of ash-flow tuffs. In : G. V. Chilingarian and K.
H. Wolf (eds.): Compaction of coarse-grained sediments, II. Amsterdam, Elsevier: 677-717.
Smith, R. L. (1960). Ash flows.- Geol. Soc. Amer. Bull., 71: 795-842.
Sparks, S. (1976). Grain size variations in ignimbrites and implications for the transport of pyroclastic
flows.- Sediment., 23: 147-188.
Stevenson, R. J., Dingwell, D. B., Bagdassarov, N. S., and Manley, C. R. (2001). Measurement and
implication of "effective" viscosity for rhyolite flow emplacement: Bull. Volc., v. 63, no. 4, p. 227-237.
Streck, M. J. and Grunder, A. L. (1995). Crystallization and welding variations in a widespread
ignimbrite sheet; the Rattlesnake Tuff, eastern Oregon, USA.- Bull. Volcanol., 57: 151-169.
Swanson, S. E., Naney, M. T., Westrich, H. R. and Eichelberger, J. C. (1989). Crystallization history of
Obsidian Dome, Inyo Domes, California.- Bull. Volcanol., 51: 161-176.
Walker, G.P.L. (1971). Grain-size characterisation of pyroclastic deposits.- J.Geol., 79:696-714.
Werner, R. and Schmincke, H.-U. (1999). Englacial vs lacustrine origin of volcanic table mountains:
evidence from Iceland.- Bull. Volcan., 60: 335-354.

46
Wilson, C. J. N. (1985). The Taupo eruption, New Zealand - II. the Taupo ignimbrite.- Phil. Trans. R.
Soc. Lond., A 314: 229-310.
Wilson, C. J. N. and Walker, G. P. L. (1985). The Taupo eruption, New Zealand - I. General aspects.Phil. Trans. R. Soc. Lond., A 314: 199-228.
White, J.D.L. (2000): Subaqueous eruption-fed density currents and their deposits.- Precambr. Res.,
101, 87-109.

Scientific journals
Journal of Volcanology and Geothermal Research
Bulletin of Volcanology (International Association of Volcanology and Chemistry of the Earths Interior,
IAVCEI)
Useful links
IAVCEI: http://www.iavcei.org/
Societ Volcanologique Europenne: http://www.sveurop.org/
USGS Cascade Volcano Observatory: http://vulcan.wr.usgs.gov/home.html
Institute of Nuclear Science, New Zealand: http://www.gns.cri.nz

You might also like