Está en la página 1de 31

Petrology of Volcanic Rocks

Part 4: How to Name Volcanic Rocks


J. Nicholls and M.Z. Stout
Theoretical Petrology
How To Name Volcanic Rocks
J. Nicholls and M.Z. Stout

Introduction
Hawaiite, basanite, tholeiite, pantellerite, shoshonite.... names of volcanic
rocks, mysterious, intriguing, seemingly endless, and bewildering. Students,
practioners of Earth science, and even petrologists despair of making sense of
how and what to name volcanic rocks. Earth scientists have confronted this
problem through parts of three centuries. This section of the e-text explains
our thoughts on how to name them. We say less about what to name them
but more about what the names should mean. We dont expect to be more
successful than past petrologists in getting our views and ideas on classifying
and identifying volcanic rocks accepted by the Earth science community. We
do, however, want to present them.

A rock classification worth its salt will advance the science and encapsulate
the state of knowledge. It needs names that carry information about how the
rocks formed that is, the name applied to a particular rock needs to convey
some information about the rocks genesis. Shand (1947), citing John Stuart
Mill, described two kinds of classifications those that simply name the
objects and those that imply a relationship between the objects in the clas-
sification. Most rock classifications are claimed to fit the first category. Shand
(1947) wanted to see the development of the second kind. Ryder (1999) re-
cently echoed Shands cry for a classification with genetic implications, this
time for a classification of Lunar basalts.

The criteria used to identify the rock need to be objective features of its min-
eralogy, chemistry, and texture. Although the names in a classification can
have genetic implications, the identification process uses observable features.
Separate classification from identification. Shand (1947) proposed that the
objective criteria used to identify the rocks in the classification be based on
features related to the physical-chemical processes that formed them. The
most obvious manifestations of the physical-chemistry of magmas are the
phases present in and, equally important, absent from the phase assemblage
in the rock. Because thermodynamics is one of the best predictors of phase
assemblages, it follows that the thermodynamic properties of melts and mag-

120
How To Name Volcanic Rocks

mas can aid in selecting criteria for identifying rocks even if the criteria are
not thermodynamic properties themselves.

The authors of two of the most important classifications in petrology, the


CIPW classification (Cross, et al., 1902) and Shands (1917, 1947) clas-
sification, characterized them as genetic classifications. Other authors have
characterized their classifications as non-genetic or have ignored character-
izing them altogether.

Criteria for Identification


The objective criteria that petrologists traditionally use to identify rocks
carry genetic implications. However, petrologists have pointedly ignored
these implications when selecting the criteria for identifying rocks. In fact, it
has become a point of pride that their classification schemes have no ge-
netic implications. Actually, this is an impossible state to achieve unless the
criteria for identification are truly meaningless with regard to rock forming
processes. For example, suppose six rocks: 1) A rock made of black, dense,
glass and a few light-colored crystals. 2) An off-white, glassy rock full of
holes; the bulk density is so low a hand specimen will float in water. This
rock also contains a few light-colored crystals. 3) A black rock composed of
dark, shiny, aligned, prismatic crystals intergrown with irregular grey-green
crystals that have striations on cleavage surfaces. 4) A white rock with oc-
casional, thin, gray, folded layers. The rock is made of interlocking calcite
crystals. 5) A black, thinly layered rock with organic-looking graphite spots
on the surfaces of the layers. Crystals are too small to see in hand specimen
or with a microscope. 6) A white rock composed of rounded, clear to cloudy
grains held together by small amounts of silica cement. A classification with
no genetic implications could use color as the criteria for identification and
the six rocks could be separated into two groups, black rocks and white
rocks, by anyone, anywhere. However, we expect that people with any train-
ing in the Earth sciences would object to grouping an obsidian, amphibolite,
and shale together on the one hand and collecting a pumice, marble, and
sandstone together on the other. No we expect Earth scientists would call
the obsidian and pumice, igneous; the amphibolite and marble, metamorphic;
and the shale and sandstone, sedimentary. We would rather characterize the
rocks with names that carry genetic implications but please notice that we
arrived at the names by seeing the objective features of the rocks: the phases
and objects present (glass, calcite, prismatic crystals, clear to cloudy grains,
bedding, fossils, cement) and texture (phenocrysts, alignment of crystals,
folds). We use observable features of the rocks to identify them but the fea-
121
How To Name Volcanic Rocks

tures can be interpreted in terms of genesis.

Table 1 contains a list, not exhaustive, of criteria included in various clas-


sifications to identify rocks. Any one classification will use only a subset of
features like those in this list. These criteria are all objective in that they can
be measured, observed, or calculated from observations or measurements.
Each criterion carries genetic information. Some carry more than others.
Some carry almost none color, for example some are loaded with genetic
information silica activity, for example. Consequently, even non-genetic
classifications convey genetic information.

All classifications profess to use grain size to separate plutonic from volcanic
rocks. No one ever seems to point out that grain size is a proxy for geologic
setting. Most often, in practical situations, we know whether a rock body is
volcanic, plutonic, or even hypabyssal from the geologic setting. Why, then,
do we insist on using a proxy, grain size, to tell us whether a rock is volcanic
or plutonic when we already know what kind of rock it is? As a proxy, grain
size works exceptionally well but it is not perfect. One can find fine-grained
plutonic rocks and coarse-grained volcanic rocks, although rarely. Only in
Rocks and Minerals 101 are students told to decide if a rock is volcanic or
plutonic by looking at the grain size.

To emphasize this point: In our experience, when Earth scientists are handed
a fine grained rock, composed of essential plagioclase and augite, from a
plutonic body, a chilled margin, for example, they will call the rock a fine-
grained gabbro or a micrograbbo. They do not call it a basalt. A few will
hedge and use the terms diabase or dolerite but all avoid calling it a basalt. In
other words, Earth scientists will use a modified plutonic name that conveys
the correct implication for the geologic setting rather than a name based on
grain size.

Comparison of Classifications
The Earth science community has embraced the IUGS classification as the
one to use. We compare the IUGS classification, presumably considered non-
genetic, with a classification developed by Shand (1917, 1947), a classifica-
tion he considered genetic.

Shands (1947) Classification


Shand created much of the terminology associated with the ideas and con-
cepts for identifying rocks, such as the term color index (Shand, 1916). Over
122
How To Name Volcanic Rocks

a period of thirty years, he (Shand, 1917; Shand, 1947) also devised a clas-
sification that used many of these features Silica activity, Al2O3 ratios, and
color index. He used the terms dyscrystalline and eucrystalline in his clas-
sification and denoted them by the letters D and X. Today, petrologists have
replaced these terms with volcanic and plutonic, respectively. Those who
advocate a classification with no genetic implications (for example, Kretz,
2004) would likely substitute the terms fine- and coarse-grained for volcanic
and plutonic or dyscrystalline and eucrystalline.

Along with grain size Shand (1947) used silica activity, although he didnt
use the term silica activity itself, to help identify different kinds of rocks.
At the most elementary level, silica activity controls whether quartz will
crystallize, a feldspathoid will crystallize, or whether neither will crystallize.
He used the terms oversaturated, undersaturated, and saturated for rocks with
these characteristics and assigned the letters O, S, and U or V. He thought his
undersaturated rocks could be subdivided into two kinds.

Like several terms used to name igneous rocks, oversaturated, undersaturated,


and saturated, are not used appropriately. The saturation terms should refer
to solutions. Rocks, being mechanical mixtures of phases, are not solu-
tions and technically should not be referred to as saturated, oversaturated,
or undersaturated. Melts are solutions and can be described as saturated or
undersaturated. A melt containing crystals of quartz is silica saturated. When
the melt crystallizes a rock is produced that petrologists have called over-
saturated. A melt that is undersaturated with silica will not contain quartz. If
it is also undersaturated with feldspathoids, there will be no crystals of these
minerals present either. Crystallize the melt to form a rock with neither quartz
nor feldspathoid, then petrologists label it a saturated rock. A melt saturated
with a feldspathoid will be farther from a silica saturated melt, or have a
lower silica activity, than one that carries quartz. A rock formed by crystal-
lizing such a melt has been called undersaturated by petrologists. Ideally,
oversaturated rocks should be called quartz-bearing. Saturated rocks should
be called rocks with neither quartz nor feldspathoids, and undersaturated
rocks should be called feldspathoid-bearing rocks. The question is: What
should be done about the misuse of these terms? The choices seem to be: use
appropriate phrases and words for the same concepts and drop the inappro-
priate terms or keep the old terms, realizing they have become specialized
jargon and dont mean what a non-petrologist would think they mean. We
suggest the latter course but with modifications that make the terms appropri-
ate for the melts from which the rocks formed. Quartz-bearing rocks we call
saturated or silica-saturated to reflect the nature of the melts from which they
123
How To Name Volcanic Rocks

formed. Rocks with neither quartz or other silica polymorph nor feldspathoid
we would call undersaturated. For rocks containing feldspathoid, we propose
the term critically undersaturated, following the precedent set by Yoder and
Tilley (1962) when they introduced the basalt tetrahedron into rock classifica-
tions (Table 2).

Along with grain size and silica activity, Shand (1947) included the ratios
Al2O3 makes with Na2O+K2O+CaO and Na2O+K2O as criteria for identi-
fication in his classification. Shand (1947) gave the classes in the scheme
of Al2O3 ratios (Table 3) the following names and letter designations:
peraluminous (p), metaluminous (m), and peralkaline (k). In his original
description of these characteristics, Shand (1947) included a subaluminous
category. The essential feature of the subaluminous category was equality
between the molecular amounts of Al2O3 and Na2O+K2O. Because an exact
equality seldom happens, we have dropped this category from Table 3.

Shand (1947) introduced ranges of mafic mineral percentages into his clas-
sification. He used the following terms and letters for color index: leucocratic
(0% - 30% mafic minerals, L), mesotype (30% - 60%, M), melanocratic
(60% - 90%, M), and hypermelanic (90% - 100%, H). Our terms for this
variable are shown in Table 4. In Shands (1947) classification, the equiva-
lent of a rock name would be a combination of letters, for example: DOmL.
Translated, the name of such a rock becomes: Felsic, metaluminous, over-
saturated, volcanic rock also known as a rhyolite. Most petrologists would
object to calling a rock a DOmL rather than a rhyolite.

Even before Shands work, the creators of the CIPW norm incorporated the
ideas behind silica saturation, Al2O3 ratios, and color index in their procedure
(Cross, et al., 1902).

The separation of femic and salic mineral formulae in the CIPW norm corre-
sponds with the Shands (1947) idea of color index. Some classifications even
use an index they call the normative color index, which is based on the rela-
tive amounts of salic and femic normative constituents (For example, Irvine
and Baragar, 1971). This index is the percentage of the amounts assigned to
the femic mineral formulae.

The desilication steps of the CIPW procedure correspond to levels of silica


activity. The CIPW procedure has seven desilication steps if there is not
enough SiO2 to make normative Q (quartz) whereas Shands (1947) ideas
lead to the division of silica activity space into three regions saturated,

124
How To Name Volcanic Rocks

undersaturated, and critically undersaturated.

The abundance of Al2O3 determines whether normative corundum or the


normative peralkaline minerals (acmite, Na2SiO3, K2SiO3) or neither are
formed. These criteria correspond to Shands (1947) criteria for identifying
peraluminous, peralkaline, and metaluminous rocks.

The IUGS Classification


The IUGS classification (Le Maitre, 2002) is actually a set of classifications;
each member of the set contains separate criteria for identifying particular
kinds of rocks. Pyroclastic rocks, for example, are identified by grain size
whereas common igneous rocks are identified with two sets of criteria. The
first set is based on the normalized proportions of the felsic minerals: quartz,
plagioclase, alkali feldspar, and feldspathoid; the QAPF diagram. The second
one is a chemical classification that plots silica percentage against the per-
centage of Na2O + K2O; the TAS diagram. The TAS diagram (Figure 1) is an
elaboration of the diagram used by Macdonald and Katsura (1964) to distin-
guish Hawaiian alkali olivine basalts from tholeiite basalts and the diagram
used by Cox, et al. (1979) to characterize volcanic rocks. Irvine and Baragar
(1971) also included a curve in SiO2 versus Na2O + K2O space as a criterion
to separate alkaline from subalkaline rocks in their classification (Figure 1).
The QAPF diagram (Figure 2), recommended by the IUGS committee, is a
modification of the modal classification developed by Johannsen (1931)

The TAS diagram illustrates an unfortunate tendency that permeates several


classifications of volcanic rocks: using a proxy criterion to distinguish rock
types when the primary criterion seems more difficult to apply than the proxy.
Macdonald and Katsura (1964) found that a line on the TAS diagram sepa-
rated Hawaiian tholeiitic rocks from rocks in the alkali olivine basalt suite.
However, the rocks were first identified and distinguished by their different
mineral assemblages. The primary criteria were mineral assemblages, not
the location of their representation on a TAS diagram. Later when Irvine and
Baragar (1971) attempted to separate rocks they called subalkaline (tholeiitic
and calc-alkaline rocks) from alkaline rocks, the different rock types were not
always correctly identified. The IUGS manual (LeMaitre, 2002) identifies al-
kali basalts as having compositions that fall in Field B (Figure 1) on the TAS
diagram and containing normative ne. Other basalts with compositions that
do not contain normative ne but fall in Field B are called subalkaline. The
subfield A on Figure 1 contains data points that belong to alkali basalts 97%
percent of the time. The field labelled Sa contains data points that belong

125
How To Name Volcanic Rocks

to subalkaline basalts 89% of the time. A data point that falls in the space
between the A and Sa fields is three times more likely to be alkaline than
subalkaline. The proxy criteria lead to errors when the primary criteria (data
point in Field B and presence or absence of normative ne) are not used.

The TAS diagram also acts as a proxy for distinguishing nephelinites and
melanephelinites. Nephelinites and melanephelinites fall in both the F and
U1 Fields on Figure 1 (Le Maitre, 2002). The questions arise: What primary
criteria were used to distinguish nephelinites and melanephelinites from
basanites and tepherites in the first place? Why are the primary criteria not
used instead of the TAS diagram? Using a proxy instead of primary criteria
reminds us of a cat chasing its tail; an exercise in frustration.

We are obliged to point out that Macdonald and Katsura (1964) did not
advocate their line on the TAS diagram as the criterion for separating tholeiite
from alkali olivine basalts. Rather, they used mineral assemblage as the crite-
rion. Irvine and Baragar (1971) offered three different criteria for separating
subalkaline from alkaline rocks; the TAS diagram/criterion was the easiest
of the three to use but the least reliable. They suggested the TAS line be used
with caution if the distinction between subalkaline and alkaline rocks was
critical to understanding how rocks formed. The IUGS version carefully de-
fines their primary criterion for distinguishing subalkaline and alkaline rocks
and demonstrates the probability of incorrectly assigning a rock to the proper
category if a line on a TAS diagram (Figure 1) is used to distinguish them.
Identifiers of rocks should use primary criteria rather than the proxies.

Although the classification of pyroclastic rocks and the identification of rocks


with the TAS criteria are considered non-genetic (Le Maitre, 2002), one has
to make a decision based on genetic ideas before the classifications can be
used. For example, before one can apply the criteria for identifying a particu-
lar pyroclastic rock, one must first determine whether it is a pyroclastic rock.
The IUGS manual states:

It [The pyroclastic classification] should be used only if the


rock is considered to have had a pyroclastic origin.

In the case of the TAS classification, the manual states:

It must also be stressed that the TAS classification is purely


descriptive and that no genetic significance is implied.

Yet the manual describes fields labeled On, Sn, and Un, where n is an integer

126
How To Name Volcanic Rocks

from 1 to 3, on the TAS diagram (Figure 1) as corresponding to characters


of rocks that are oversaturated, saturated, and undersaturated. With few
exceptions, undersaturated rocks, those that contain a feldspathoid when
holocrystalline, are not related to saturated and oversaturated rocks. Conse-
quently, these fields on the TAS diagram do have some genetic significance.

The QAPF classification uses two primary criteria to identify rocks. The first
criterion is based on an abbreviated concept of silica activity. The second is
based on the abundance of alkali feldspar in the felsic mineral assemblage.

The version of silica activity included in the QAPF classification recognizes


rocks that contain quartz and rocks that contain feldspathoids but confines
rocks that contain neither to a single line on the diagram that separates rocks
with quartz from those that contain feldspathoid. This leads to problems.
Fields 9 and 10 on the QAPF diagram (Figure 2) enclose the criteria for
identifying basalts and andesites. However, alkali feldspar content and pres-
ence of quartz or feldspathoid are inadequate criteria for distinguishing the
varieties of andesitic and basaltic rocks on Earth or the Moon. Many of these
rocks contain neither alkali feldspar nor quartz nor feldspathoid. Consequent-
ly, the modes of such rocks plot as a point on the QAPF diagram. The IUGS
Subcommittee introduced color index and wt% silica as additional criteria to
separate rocks in Fields 9 and 10 into four categories of basalt and andesite:
basalt, andesite, leuco-basalt and mela-andesite. The manual states:

Although this [selection of criteria] may seem rather unsatis-


factory, it is unlikely that many of these rocks will be classified
using the QAPF diagram, as the modes of most basalts and
andesites are difficult to determine accurately so that the TAS
classification will have to be used.

The modes of basalts and andesites are no more difficult to determine than
the modes of other volcanic rocks. For all practical purposes, the IUGS sub-
committee seems to be suggesting that mineralogical criteria for identifying
volcanic rocks are impractical and that chemical criteria be used instead.

Another group of rocks that cause problems are the mafic to ultramafic vol-
canic rocks with a felsic assemblage consisting of feldspathoids. Often these
rocks are so undersaturated that feldspar crystallization is completely sup-
pressed. The rocks then plot at the F-apex of the QAPF diagram. However,
many of these rocks have a high color index and so contain minimal amounts
of feldspathoids. As a result, in the QAPF classification, names are based on a

127
How To Name Volcanic Rocks

felsic assemblage that makes up less than 10-20% of the rock. Some ex-
amples from Volcano Mountain, Yukon Territory, Canada are listed in Table 5
(Trupia, 1992).

The second criterion in the QAPF classification, the concentration of alkali


feldspar in the felsic mineral assemblage, has genetic implications for a lim-
ited set of rocks those that contain essential alkali feldspar. Alkali feldspar
is important in many felsic rocks, especially those analogous to compositions
in Petrogenys Residua System (NaAlSiO4-SiO2-KAlSiO4 Bowen, 1937).
However, the vast majority of basaltic, volcanic rocks that cover the ocean
basins and large parts of the continents contain no alkali feldspar.

The IUGS classifications appear simple to use and straight-forward. This is


deceptive. The criteria are diverse and the system is complicated because the
genetic implications of the criteria used to identify rocks seem not to have
been considered early in their selection. To account for genetic differences
between rock types, additional criteria are added to the QAPF and TAS clas-
sifications to solve problems in identifying rocks that the primary criteria can-
not handle. These secondary criteria include color index, SiO2 concentration,
plagioclase composition, and amounts of normative constituents.

The Criteria We Recommend


The criteria for identifying volcanic rocks that we propose in this section of
the e-text involves several independent factors: the relative contents of the
mafic and felsic minerals (not phases) in the rock; level of silica activity; the
ratios Al2O3 makes with CaO, Na2O, and K2O; the presence or absence of Ca-
poor pyroxene (orthopyroxene, pigeonite); and glass content. The factors are
independent in the sense that any one factor doesnt require the determina-
tion of other factors to be evaluated. If, in time, these factors prove inad-
equate, new factors can be added to the list. Such new factors might include
plagioclase or olivine composition or the presence of biotite and amphibole
or the alkali feldspar content. None of the criteria we suggest are new to clas-
sifying and naming volcanic rocks; petrologists have written and talked about
them for years. Our contribution, if you can call it that, lies in emphasizing
the importance of the genetic implications attached to each of the criteria.

Mafic Index
One of the most conspicuous differences between volcanic rock types is
the amounts of minerals containing essential Fe and Mg compared to the

128
How To Name Volcanic Rocks

amounts of minerals that lack these elements above the trace or minor ele-
ment level. The percentage of Fe-Mg phases in the rock has traditionally been
called the color index of the rock. For example, the essential mineralogical
difference between trachybasalts and basalts is their respective color indices.
Trachybasalts and basalts can contain the same minerals, just in different
proportions and compositions.

Rocks containing glass create problems in assigning a color index if volumes


of phases, rather than volumes of minerals containing essential Fe-Mg are
assigned to the index. We propose to call the percentage of mafic minerals, in
contrast to the percentage of mafic phases in a rock, the mafic index. Volca-
nic glasses nearly always contain essential Fe-Mg. Glass, however, is not a
mineral because it lacks a repeated atomic structure. We advocate a sepa-
rate number for the glass content (see below) and would add only mineral
volumes to the newly defined mafic index. This means that some obsidians,
for example, could have a high mafic index if the crystal phases are domi-
nated by augite and fayalitic olivine, say. In contrast, some obsidians have
a phenocryst assemblage dominated by felsic minerals and, consequently
would have a low mafic index. Examples of the contrasting mineral assem-
blages in obsidians can be found in the modal analyses compiled by Carmi-
chael (1967). Mafic indices could help delineate petrologic problems associ-
ated with obsidians. For example, why should some obsidians carry felsic
phenocryst assemblages while other obsidians, with similar compositions,
carry mafic phenocryst assemblages? The mafic index of holocrystalline
rocks would be identical to the color index of the rock. Only in rocks contain-
ing glass would the two indices differ. Table 4 shows the terminology associ-
ated with mafic index.

Williams, et al. (1954) state:

...that all rocks transitional members of series and that clear-


cut boundaries are not to be expected.

It is a source of wonder that volcanic rocks contain minerals both with Fe


and Mg and minerals without Fe and Mg. There are some ultramafic rocks
that contain only minerals with Fe and Mg, but most are not volcanic. The
rare ultramafic volcanic and hypabyssal rocks usually have small amounts of
felsic minerals (Figure 3). The dichotomy between FeMg-bearing minerals
and those that dont contain essential FeMg has been part of igneous petrol-

129
How To Name Volcanic Rocks

ogy at least since the latter part of the 19th century (For example, Cross, et
al., 1902). This dichotomy is an essential feature of Bowens (1928) reaction
series.

Although petrologists can concentrate independently on the two crystalliza-


tions streams, felsic and mafic, that characterize most volcanic rocks, the two
streams do not form in isolation. The crystallization histories and inferred
physical conditions of crystallization, for example pressure and temperature,
are derived from both the felsic and mafic mineral assembleages. Criteria for
identifying rocks should include both assemblages.

The genetic implications of the mafic index of a volcanic rock are, apparently,
so obvious they are never mentioned. Volcanic rocks come from magmas that
formed by partial melting and assimilation of pre-existing rocks or by differ-
entiation of more primitive magmas. Felsic rocks are thought to result from
partial melting of intermediate to mafic source rocks or by differentiation of
intermediate to femic magmas. Mafic rocks, on the other hand, are thought to
originate by partial melting of ultramafic rocks in the mantle or differentia-
tion of magmas that contain more Fe and Mg than the mafic rock.

We find it passing strange that a feature so loaded with genetic implications


as mafic index is not used in rock classifications. Hatch, et al. (1972) even go
so far to state:

Colour [Mafic] Index is of little practical use in rock clas-


sification, except to delimit the group of rocks deficient (to the
vanishing point) in felsic components but correspondingly rich
in mafics, and appropriately named Ultramafites.

Even though the mafic index of a rock seems a logical variable for a criterion
for naming volcanic rocks, it is not without its problems. In thin section, the
eye more readily sees the colored minerals, the minerals with high relief,
and the opaque minerals than the clear, low relief, and transparent minerals.
Minerals with high relief, color, and high birefringence are almost always
FeMg-bearing. Consequently, mafic indices are often biased to higher values
than actually exist. Even point counting will give a biased mode and mafic
index unless particular care is taken. An unbiased estimate of a mode from
which a mafic index is derived, requires that the areas underlain by the phases
in the rock be estimated on a random surface (Chayes, 1956). Thin sections
have a finite thickness (0.03 mm is standard) and one mineral can lie beneath
another if the boundaries between grains dip beneath the surface of the sec-

130
How To Name Volcanic Rocks

tion. If a mineral with high birefringence, color, or high relief lies beneath a
clear felsic mineral, a petrologist doing a point count will often count the spot
on the surface as being occupied by an FeMg mineral whereas the spot on
the surface of the thin section is occupied by a felsic mineral (Holmes, 1921).
This phenomenon is called the Holmes effect (Chayes, 1956) and it can add a
significant systematic error to the modal percentages. The effect is not unique
to petrological studies; it also happens in the study of biological sections
(Mandarim-de-Lacerda, 2003). The bias can be eliminated by point counting
rock surfaces in reflected light where the intersection of grain boundaries and
the reflected surface can be detected (Figure 4).

In spite of these problems with accurately determining mafic index, we think


the relative amounts of felsic and mafic minerals carry too much genetic
information to be left out of a good classification scheme.

Silica Activity
Silica activity, a thermodynamic property of a melt, magma, or rock, can be
calculated from the chemical compositions of the melt or minerals that make
the rock. Silica activity varies with pressure and temperature as do physical
properties such as density, heat capacity, or viscosity. Just as a temperature or
pressure must be fixed when calculating a physical property so must they be
fixed when calculating a silica activity. The important point: silica activity is
as much an objective feature of a melt, magma, or rock as a physical prop-
erty.

Carmichael, et al. (1970) first noted that several mineral assemblages, like
nepheline and albite, buffered silica activity and that some of these buffers
corresponded with the desilication steps of the CIPW norm. Thermodynamics
thus became a tool for implementing and evaluating criteria in a scheme of
classification for identifying rocks. Ghiorso and Carmichael (1987) outlined
procedures that could give rise to a set of strictly thermodynamic criteria for
identifying rocks. The required thermodynamic data for the melts and solid
solutions were not available then and are still not available not to implement
their procedures.

Comparing a thermodynamic property like silica activity to a physical


property like density does not really tell us what the thermodynamic property
is. One way to think of activity is to think of it as a concentration; the more
concentrated the component is, the higher its activity. If you evaporate a
solution with a high concentration of salt, for example, the solution will soon

131
How To Name Volcanic Rocks

saturate and crystals of salt will


Calculation of Buffer Curves
form. In the same way, melts
with a high activity of silica are The thermodynamic equation for the ac-
closer to saturation with quartz tivity of silica defined by the nepheline-
than melts with a low activity albite buffer is given by:
of silica. To put it another way,
a melt with a high silica activity Gr1 2 RT ln aSiO2 = 0 or:
can dissolve less quartz, none
2 RT )

ln aSiO2 = Gr1
at all if saturated, than a melt
with a low activity of silica. The
activity of a component is not
where Gr 1 = G f , Ab 2 G f , SiO G f , Ne
truly a concentration; the activ- is the difference in standard state Gibbs
2

ity depends also on the nature energies of of formation for the phases,
of the bonding in the solution. R is the universal gas constant (8.3143
Unfortunately, concentration J/deg-mol), and T is the temperature in
and activity cannot be put into Kelvin. The standard Gibbs energies of
an equal one-to-one correlation formation are tabulated values found
even though concentration is the in tables of thermodynamic properties.
largest contributor to activity. Ultimately, they are derived from experi-
mental measurements (see, for example,
Silica activity can be esti- Anderson and Crerar, 1993, Anderson,
mated in several ways: 1) 1996).
Does the rock contain quartz, a
feldspathoid, or neither? 2) Does the rock contain quartz, olivine and ortho-
pyroxene, plagioclase and nepheline, sanidine and leucite, or perovskite and
titanite? 3) What is the silica activity, as calculated with a thermodynamic
model (For example, Ghiorso and Sack, 1995) in a melt with the composition
of the rock? These three methods of estimating silica activity, presence or
absence of quartz or feldspathoid, presence of a particular mineral pair, and
the thermodynamic activity of silica, provide progrssively more exact values
for the property.

Examples of the relationship between mineral assemblage and silica activity


can be shown in the compositional space: forsterite (Fo), nepheline (Ne), and
silica (SiO2) (Figure 5, upper diagram). The space can be divided into three
subspaces representing three different mineral assemblages: 1) albite (Ab),
enstatite (En), quartz (Q); 2) forsterite, enstatite, albite; and 3) forsterite,
nepheline, albite.

Each mineral assemblage defines a different level of silica activity (lower dia-
gram, Figure 5). In the assemblage forsterite, nepheline, albite the exchange:

132
How To Name Volcanic Rocks

Ne + 2 SiO2 = Ab

defines the activity of silica. As long as both albite and nepheline are present
in the melt, the activity of silica in the melt is confined to the lower curve on
Figure 5. If SiO2 is added to the melt in an attempt to increase the activity of
silica, nepheline will combine with the silica to form albite and the activity of
silica will remain the same. Likewise, if SiO2 is removed from the melt in an
attempt to lower the activity of silica, albite will dissolve and more nepheline
will crystallize to keep the activity of silica the same. The activity of silica
is buffered as long as both mineral phases, nepheline and albite are present.
If enough SiO2 is added or subtracted until one of the mineral phases disap-
pears, then, and only then, can the activity of silica depart from values on
the curve. The nepheline-albite buffer curve divides silica activity space into
three parts: the curve itself where albite and nepheline co-exist with melt, the
region above the curve where albite co-exists with melt, and the region below
the curve where nepheline co-exists with melt.

The exchange:

SiO2 = Q
defines the activity of silica in the assemblage En,Ab,Q. The only difference
in properties connected with this curve compared to the nepheline-albite
curve is the presence of only one mineral phase rather than two. Adding SiO2
to a melt that contains quartz will only cause the precipitation of more quartz.
The activity of silica is restricted to the upper curve on Figure 5 or to values
below the curve. If SiO2 is removed from the melt until all the quartz disap-
pears, the the activity of silica can fall beneath the upper curve on Figure 5.

The nepheline-albite and quartz buffer curves define a region in activity of


silica-temperature space where melts co-exist with mineral assemblages that
lack both quartz and feldspathoid (region shaded pink on Figure 5, lower
diagram). Melts co-existing with quartz plot on the upper curve and melts
co-existing with nepheline fall on and below the lower curve. This three-
fold division of silica activity space corresponds the Shands (1917, 1947)
division of silica activity space into saturated, undersaturated and critically
understurated regions.

The CIPW procedure (Cross, et al., 1902) splits silica activity space into
several regions all are characterized by a buffer curve: Presence of norma-
tive quartz (Q), normative orthopyroxene (hy) plus olivine (ol), normative
albite (ab) plus nepheline (ne), normative orthoclase (or) plus leucite (lc),

133
How To Name Volcanic Rocks

or normative titanite (tn) plus Calculation of Activity of


perovskite (pf) are some of the
buffer assemblages possible in Silica in Melts
CIPW norms that can be extend- Ghiorso and Sack (1995) calibrated a
ed to melts and rocks. At each thermodynamic model for silicate melts.
level of silica activity below The activity of silica can be calculated
the saturation curve for quartz, from this model. The equation for the
two minerals are present. An activity of silica in the model is:
example is shown on Figure 5, n
lower diagram where the middle RT ln aSiO2 = RT ln xSiO2 + wi , SiO2 xSiO2
curve represents the exchange: i=2
n 1 n

Fo + SiO2 = En w xx
i =1 j = i +1
ij i j

According to the normative


The wij are numerical coefficients that
procedure, a melt on this buffer
come from calibrating the model, the xi
curve will co-exist with a norma-
are the mole fractions of the components
tive mineral assemblage contain-
that make the melt, SiO2 is designated
ing the normative equivalent of
component 1. The number of components
orthopyroxene, hy, and olivine,
is n.
ol. The CIPW procedure places
the composition in the triangle
Fo + En + Ab (Figure 5, upper diagram) If a composition doesnt contain ol,
then the CIPW procedure will place the composition in the En + Ab + Q field
with normative hy and Q. Compositions with insufficient SiO2 for creating
normative hy will contain normative ol plus ne, the equivalent of a composi-
tion in the field Fo + Ne + Ab (Figure 5, upper diagram).

Neither Shands (1917, 1947) classification nor the CIPW procedure (Cross,
et al., 1902) can precisely depict the silica activity in magmas without a buf-
fer assemblage. To do that, we need to use the activity of silica in the melt or
in a melt that has the same composition as the rock.

At a given temperature and pressure, the activities of components in a melt


or any other solution depend only on the composition of the melt or solution.
Often, this dependence is mathematically complex. The more components
in the melt, the more complex the mathematical dependence. In addition, the
bonding in the melt affects the complexity of the mathematical dependence.
Fortunately, thermodynamic models of silicate melts are relatively simple and
relatively easy to evaluate. Although the calculations may be tedious, the idea
is conceptually simple: if the pressure, temperature, and composition of the

134
How To Name Volcanic Rocks

melt are known, the activity of silica can be calculated.

Thermodynamic models of silicate melts give us the means to calculate


the activity of silica, at a given temperature, whether or not an appropriate
mineral assemblage is present. The activity of silica exists independently of
the mineral assemblage. The thermodynamic specification of the activity also
satisfies Shands (1917, 1947) call for features of physical-chemistry to be
criteria for identifying volcanic rocks.

Although they likely did not know it at the time, the authors of the CIPW
norm (Cross, et al., 1902) were doing thermodynamics when they created the
desilication steps in the normative calculation. Likewise, when petrologists
search out whether or not quartz or feldspathoid are present in a volcanic
rock, they are searching for objective features that can be understood by ap-
plying concepts from thermodynamics.

Al2O3 Ratios in Volcanic Rocks


Al2O3 is usually the second most concentrated oxide after SiO2 in igneous
rocks and magmas. The influence Al2O3 has on the nature of magmas and
igneous processes depends on the molecular ratio of Al2O3 to other oxides,
namely Na2O, K2O, and CaO. The last three oxides and Al2O3 occur in feld-
spars; essential minerals in the most abundant igneous rocks. Igneous rocks
are placed in groups defined on the ratio of Al2O3 to the other three oxides as
shown in the Table 3. The ratio of Al2O3 to the other oxides reflects informa-
tion about the nature of the source rocks and, possibly, information about
how the ratio in a particular rock came to be.

The names applied to the rock types characterized by the different ratios carry
genetic information; identification of the rock type depends on the observable
chemical and mineralogical features in the rocks. Each of the rock types have
mineralogical features that follow from the Al2O3 ratios with K2O, Na2O, and
CaO in the rock. These are listed in (Table 6). The ratio also determines char-
acteristic features in the CIPW norm. These features can be used to determine
what the ratio is in the rock because the CIPW norm is determined by the
chemistry of the rock (Table 7).

Al2O3 Ratios and Source Rocks for Magmas


The most abundant igneous rocks on Earth are the metaluminous rocks. Their
ultimate origin lies in partial melting of the lithosphere and asthenosphere.
Consequently, the consensus among petrologists is that the source rocks for

135
How To Name Volcanic Rocks

the primary metaluminous magmas are also metaluminous.

The peraluminous rocks also derive their character from their source rocks.
Weathering cycles at the Earths surface separate alkali oxides (Na2O and
K2O) from Al2O3. At its extreme stage, weathering puts the alkali oxides
in the ocean, along with nearly everything else, and leaves bauxite behind.
Igneous processes, as we know them, cannot generate peraluminous mag-
mas from metaluminous ones and vice-versa (but see below). Consequently,
peraluminous granitic rocks have become known as S-type rocks, the S
standing for a sedimentary source, and metaluminous rocks are known as I-
type rocks. These names reflect what we know about the nature of the source
rocks that partly melted to form at least some rhyolitic magmas.

The origins of the peralkaline and hyperperalkaline rocks remain more of a


puzzle than the others. There is a mechanism in igneous petrology capable
of generating a peralkaline condition called the plagioclase effect (Bowen,
1945). In essence, the plagioclase effect arises because crystallization of a
melt consisting of albite (NaAlSi3O8) and a calcium-bearing substance (for
example diopside, CaMgSi2O6) will usually generate a plagioclase solid solu-
tion with an anorthite component (CaAl2Si2O8). Subsequent crystallization of
such a plagioclase will deplete the melt in Al2O3 and increase the concentra-
tion of Na2O and K2O, leading to the peralkaline condition. Carmichael, et al.
(1974) have demonstrated that the compositions of some peralkaline rocks,
their included glasses, and phenocrysts do satisfy the requirements for the
plagioclase effect in nature. Whether the process has been the cause for the
origin of all peralkaline rocks remains open. The origins of the hyperperal-
kaline rocks, in particular are difficult to explain with the plagioclase effect
alone.

Most volcanic rocks on Earth fall into the metaluminous category. Table 8
shows the criteria for identifying these rocks and placing them into an ap-
propriate pigeon-hole. The scheme and its basis in the system Fo-Ne-SiO2 are
presented in an animation at:
http://www.gtwist.ca/Petrology/classification.html.

Presence or Absence of Ca-poor Pyroxene


Within the metaluminous rocks, rocks identified as basalts occur at all three
of Shands (1917, 1947) levels of silica activity (Table 8), suggesting another
criterion is needed to distinguish the different basalt types. Rocks with higher
and lower mafic indices than basalts have been given different names in

136
How To Name Volcanic Rocks

the different categories of silica activity, indicating significant differences


between the categories. Consequently, it is reasonable to think that significant
differences exist in the basalt types.

In the early 1920s petrologists recognized at least two basaltic magma types,
tholeiites or tholeiitic basalts (Le Maitre, 2002) and alkali olivine basalts or
alkali basalts (Le Maitre, 2002). In the Mull Volcanic Complex, two main
basaltic magma types were recognized, the non-porphyritic central type and
the plateau magma type (Bailey, et al., 1924). Kennedy (1933) realized that
basaltic rocks with features similar to the magma types of Mull occurred
world wide. He changed the names to tholeiitic basalt and alkali basalt. Til-
ley (1950) later called alkali basalts, alkali olivine basalts. The relationship
between the two types has been a matter of conjecture and controversy but
recognition of the two types is not in dispute. The two types have tradition-
ally been distinguished by whether or not a Ca-poor pyroxene (pigeonite or
orthopyroxene) is present. Tholeiite basalts carry Ca-poor pyroxenes whereas
alkali olivine basalts do not. An intergrowth of augite (Ca-rich pyroxene) and
orthopyroxene (Ca-poor pyroxene) in a tholeiite basalt is shown on Figure 6.

Tholeiite basalts are characterized by the presence of at least two pyroxenes.


One pyroxene is always an augite, the others will be pigeonite and/or ortho-
pyroxene. Olivine, when present, almost always occurs as a phenocryst and
is absent from the groundmass. An exception is a tholeiite from the Garibaldi
Region, British Columbia (Nicholls, et al., 1982, sample GV-7) that contains
groundmass olivine, pigeonite, and augite. Chemically, tholeiite basalts are
characterized by normative Ca-poor pyroxene with either normative olivine
or normative quartz.

Alkali olivine basalts have been characterized by one pyroxene, an augite,


co-existing with olivine. Olivine often occurs as phenocrysts and is always
present in the groundmass. The augite in alkali olivine basalts is often dis-
tinctly titaniferous. Minor amounts of nepheline and/or alkali feldspar may or
may not accompany plagioclase in the groundmass. Chemically, rocks named
alkali olivine basalt usually contain normative nepheline and olivine. Norma-
tive Ca-poor pyroxene is absent. Some basalts are found in western Canada
(Fiesinger, 1975; Fiesinger and Nicholls, 1977) and western United States
(Stout, 1975; Stout and Nicholls, 1977) with traditional alkali olivine basalt
mineralogy but with the normative features of tholeiite basalt.

In the space defined by temperature and activity of silica (Figure 5, up-


per diagram), rocks in the tholeiite suite should lie between the quartz and

137
How To Name Volcanic Rocks

orthopyroxene-olivine buffers but in terms of the CIPW norm, they are con-
fined to the quartz buffer or to the orthopyroxene-olivine buffer: CIPW norms
will contain either normative quartz (Q) and orthopyroxene (hy) or norma-
tive orthopyroxene (hy) and normative olivine (ol). Normative quartz (Q) and
normative olivine (ol) cannot occur together.

By the same token, alkali olivine basalts should contain a feldspathoid


and have normative nepheline (ne) in their norms. Ca-poor pyroxene and
nepheline do not co-exist. Neither do normative orthopyroxene (hy) and
normative nepheline (ne). Idealy, alkali olivine basalts, then, should have
silica activities on or below the albite-nepheline buffer. The normative
minerals that characterize alkali olivine basalts are olivine (ol) and nepheline
(ne). Yet, between the two buffer curves is a region where melts that neither
contain Ca-poor pyroxene nor feldspathoid should exist. Dividing basalts into
just two kinds makes it difficult to reconcile normative and modal mineral
assemblages

Just as we argued that a two-fold division of rock types in the QAPF system
was inadequate, we also argue that the two-fold division of basaltic rocks into
rocks of the tholeiite suite and the alkali olivine basalt suite is inadequate.
Basalts and their associated rocks should be divided into three groups. The
first group includes rocks with Ca-poor pyroxenes in the mafic and interme-
diate varieties with quartz and Fe-olivine in the felsic varieties. These rocks
traditionally belong to the tholeiite suite. The second group contains rocks
without Ca-poor pyroxene and without feldspathoids and quartz. Basaltic
rocks with these characteristics are common Recent volcanic products in
western North America between the Yukon and the Rio Grande rivers and
maybe beyond. Rocks without, quartz, Ca-poor pyroxene, nor feldspathoids
but with small to large mafic indices erupted 2000 years ago at Crater of the
Moon (Stout, et al., 1994). These rocks we would call members of the olivine
basalt suite. The third group of rocks are characterized by the presence of
feldspathoids, typically nepheline and lesser amounts of leucite. These are
rocks of the traditional alkali olivine basalt suite (Table 9).

In Mg-rich melts a fundamental relationship occurs that separates


feldspathoid-bearing rocks from rocks containing a silica polymorph, such
as quartz. Figure 7, upper diagram, shows the three-phase triangles for solid
assemblages in the system forsterite, nepheline, and silica. Feldspathoid-bear-
ing assemblages (Fo, Ab, Ne) are separated from assemblages containing
a silica polymorph (En, Ab, SiO2) by assemblages with neither quartz nor
feldspathoid (Fo, Ab, En). The separation occurs because the tie-line between

138
How To Name Volcanic Rocks

Fo and Ab is more stable than the tie-line between En and Ne (lower diagram,
Figure 7). Because nepheline and enstatite are unstable compared to forsterite
and albite, the Fo-Ab tie-line represents a thermal barrier at temperatures
where liquids exist in the system. Liquids with compositions that fall in the
triangle Fo-Ne-Ab are analogous to alkali olivine basalt magmas whereas
liquids that fall in Fo-Ab-En and En-Ab-SiO2 are analogous to tholeiite
basalt magmas. Ca-poor pyroxenes are absent from alkali basalt assemblages
whereas they are required in tholeiite basalt assemblages. In this simple
system, the two kinds of magmas cannot be related to each other by crystal-
lization processes.

In complex systems with more components, several petrologists have pos-


tulated that the two magma types can be related by fractional crystallization
(For example, Bowen, 1928; Naumann and Geist, 1999). Generally, one can
find reasonable mass balance paths from one magma type to another (Bowen,
1928). Paths from one magma type to the other that satisfy energy balance
constraints are harder to find but they seemingly do exist (Naumann and
Geist, 1999).

Still, one wonders whether the examples of relating tholeiite and alkali ol-
ivine basalt melts by crystal fractionation could be more a matter of impre-
cision in identification rather than phase relations. In the examples, do the
melts labeled tholeiite crystallize Ca-poor pyroxene at low pressures? Do the
melts labeled alkali olivine basalt crystallize a feldspathoid at some stage in
their magmatic history? Perhaps both melts actually have characteristics of
the olivine basalt suite rather than the tholeiite or alkali olivine basalt suites.

The presence or absence of Ca-poor pyroxenes is important enough to be


included in the set of criteria for distinguishing various volcanic rocks. Rocks
that contain Ca-poor pyroxenes carry implications for their genesis signifi-
cantly different from implications associated with similar rocks that do not
contain Ca-poor pyroxene. The presence or absence of Ca-poor pyroxenes
enables us to distinguish three magma types: tholeiite basalt, olivine basalt,
and alkali olivine basalt (Table 9). Commonly associated with the three basalt
types are rock types both more mafic and more felsic than the basalts. These
rocks carry distinguishing chemical and mineralogical features; names ap-
plied to these rock types are listed in Table 9.

Glass in Volcanic Rocks


In 1928 Bowen made the statement about glassy rocks that:

139
How To Name Volcanic Rocks

They are the only rocks of which we can say with complete
confidence that they correspond in composition with a liquid.

We know today that this statement is not always correct; some glasses have
had their compositions altered by interaction with ground waters (Stewart,
1979). Even Bowen (1928) added a caveat to his statement in a footnote that
recognized some constituents, particularly volatile components, could be lost
during cooling. In spite of these possible alterations, many glasses, especially
those associated with modern volcanic rocks do fit Bowens statement they
do have compositions that correspond closely to volcanic liquids.

Because of this correspondence, Carmichael (1962, 1963, 1964, 1965) was


able to deduce crystallization paths and liquid lines of descent for several
magma types. For example, by studying rocks containing large amounts of
glass, Carmichael (1963) was able to distinguish residual liquids that were
the products of fractionation of tholeiite basalt from residual liquids that
formed by processes involving sialic material.

Obsidians are not the only rocks dominated by glass. The 1968 eruptions of
Kilauea, Hawaii, for example, produced rocks that contained 70%-90% glass
(Jackson, et al., 1975). A photomicrograph is shown on Figure 8. The glasses
in the rocks have compositions consistent with formation from more primi-
tive melts by processes dominated by crystal fractionation of olivine (Nich-
olls and Stout, 1988). In addition, at the microscopic scale, glass inclusions
trapped in olivine phenocrysts in the Mg-rich, glassy (>80%) rocks suggest
the primitive melts contained 18% or more MgO (Nicholls, 2000; Nicholls
and Stout, 1988; Wright, 1971).

Trupia (1992) analyzed residual glasses in several nephelinites from Yukon


Territory, Canada and found significant chemical variations across a single
thin section (Some examples of the glass compositions she analyzed are
given in Table 10). Summaries of the chemical variations in the residual
glasses are shown on Figure 9. A large part of the variation can be explained
by differential crystallization of olivine, augite, nepheline, leucite, and apatite
in different combinations (Figure 9, upper diagram). One interpretation of
the data postulates each spatially isolated volume of glass acted as a separate
chemical system, plating out on the surrounding minerals in different propor-
tions. The residual glasses have different characteristics, resulting in different
levels of silica activity (Figure 9, lower diagram).

The residual glasses in the nephelinites did not experience the effects of

140
How To Name Volcanic Rocks

plagioclase crystallization. The mineral that contains abundant CaO is


augite, which contains minor Al2O3. Consequently, the residual glasses
have high concentrations of Al2O3 and very low concentrations of CaO
(Table 10). Many of the residual glasses have normative corundum and are
peraluminous. In some instances, there is insufficient Ca to make norma-
tive apatite and the norms carry excess P2O5. This is the plagioclase effect
in reverse; plagioclase crystallization can lead to the peralkaline condition
where suppression of plagioclase crystallization can lead to the peraluminous
condition. Whether either process operates in nature to an extent large enough
to form rock bodies with these contrasting properties is another question.

Authors of classifications have given short shrift to volcanic glasses and


volcanic rocks that contain glass. Typically, a classification that uses chemi-
cal criteria to identify rocks is extended to rocks containing glass so that
rocks with and without glass have the same name. We ask, is this the best that
can be done? A rhyolite, presumably holocrystalline, is not the same kind of
rock as an obsidian, a rock dominated by glass; the two shouldnt be identi-
fied with the same name. In this instance, at least, we agree with the IUGS
subcommittee on the systematics of igneous rocks:

Rocks should be named according to what they are, and not ac-
cording to what they might have been (Le Maitre, 1989, 2002).

Obsidians might have become rhyolites if they had not congealed to glass.
Consequently, rocks that contain glass should be distinguished from those
that dont.

We could find only six names specifically designated for rocks containing
large amounts of glass: obsidian, pitchstone, perlite, palagonite, limburgite,
and vitrophyre. The IUGS glossary (Le Maitre, 2002) defines obsidian as a
common term for volcanic glass... Most Earth scientists, we expect, would
also think obsidians have compositions similar to rhyolites. Pitchstones are
different from obsidians because they contain more H2O (Obsidians < 1%
H2O, pitchstone > 1% H2O; Le Maitre, 2002). Perlite is a volcanic glass with
numerous concentric cracks (Le Maitre, 2002). Most perlites also contain
lots of H2O and may have had their compositions altered by interaction with
ground water (Stewart, 1979). Palagonite, according to the IUGS Glossary
(Le Maitre, 2002), is a yellow or brown completely devitrified glass. The
yellow-brown material formed by Recent subglacial eruptions of intermediate
and mafic rocks in northern British Columbia and Yukon territory, Canada, is
isotropic glass; it is not devitrified. The IUGS glossary calls limburgite:

141
How To Name Volcanic Rocks

A basic volcanic rock containing phenocrysts of pyroxene, ol-


ivine and opaques in a glassy groundmass containing the same
minerals. No feldspars are present. May be used as a synonym
for a hyalo-nepheline basanite (Le Maitre, 2002).

Finally, a vitrophyre is called a porphyry with a glassy groundmass in the


IUGS Glossary (Le Maitre, 2002). The name has been used for the dense
glass layer in many welded tuffs. The IUGS classification recommends
identifying rocks with criteria based on the amount of glass, with limits at
20%, 50%, and 80% by using the terms glass-bearing, glass-rich, and glassy,
respectively. Rocks with more than 80% glass should get a special name like
obsidian. However, as we have just noted, there are not a lot of names to use.

If we take these definitions literally, some have very specific criteria for
identifying rocks (limburgite) and some could apply to many rocks contain-
ing glass, with both the rocks and glasses having compositions all over the
chemical map from rhyolite to basalt (obsidian, vitrophyre). Other names
(pitchstone and perlite) suggest the glasses have had their compositions
altered. People disagree whether palagonite is even a glass.

Obviously, glasses can only be characterized chemically and in a rock


containing glass, we would like to know how much glass is present. Conse-
quently, we suggest there are two features of glass content that are important
when identifying rocks: How much glass is present and what are its chemical
characteristics. In identifying a rock with glass an Earth scientist should tell
the world how much glass is in the rock, in other words give a glass index.
They should also provide information about the composition of the glass. If
an analysis is available, the analysis should be given.

Even at the most elementary level, rocks identified by their chemistry will
have different names than rocks identified with mineralogical criteria (see
Figure 10). One expects SiO2-rich rocks to be felsic and SiO2-poor rocks to
be mafic. In fact, in an ideal universe it would be nice if rocks plotted in a
diagonal band stretching from the upper left of Figure 10 to the lower right.
They dont and our expectations are dashed. The same rocks frequently have
different names in chemical and mineralogical classifications that employ the
same set of names but assign them with different criteria. Chemical rhyolites
are often not mineralogical rhyolites. The long and persistent tradition of
using the same set of names in mineralogical and chemical classifications
will not stop even though it should. But perhaps, maybe, petrologists can
start providing indications of where the names come from, for example

142
How To Name Volcanic Rocks

rhyolite(chem) or rhyolite(TAS). For rocks whose names come from the TAS
classification, the IUGS recommends the prefix hyalo-, for example hyalo-ba-
salt. The term hyalo-picrite would be a better name for the rock shown in the
photomicrograph of Figure 8 than just plain picrite.

Problems and Propects for Rock Identification


Myths of Rock Classification
Schemes of rock classification and identification have been dogged by several
myths. It is past time they were abandoned.

Myth 1: Rock names should not carry genetic implications. No to be


useful, rock names must carry genetic implications and eliminating genetic
implications from the criteria used to identify rocks is a near impossibility.
The criteria used to arrive at the rock name must be measurable, observable
or derived from measurement or observation. Identification needs to be dis-
tinguished from classification.

Myth 2: There are too many rock names. No the number of names is not a
problem; the problem is too many names have few implications about the re-
lationship between the rock and other rocks. This myth implies that we have
seen all the rock types in the universe let alone here on Earth,

Myth 3: A good classification will last forever. No in spite of the de-em-


phasis on mineralogy and petrology in modern university curricula, petrol-
ogy is a living science. Because new knowledge and concepts continue to
develop, classifications and the criteria used to identify rocks will change.

Myth 4: A good classification will be short, simple, and easy to use. No, not
necessarily a classification should be logical and the genetic reasons for the
criteria used to identify rocks should be clearly explained. Seemingly arbi-
trary criteria for identifying rocks lead to complicated classifications that are
difficult to understand and be taken seriously by non-petrologists.

How to Name Volcanic Rocks


The concepts, ideas, and information for constructing a classification with
genetic implications have existed for more than a century. Our understanding
of the information provided by physical-chemistry has improved over time
but the basic concepts needed to identify rocks in a framework that implies
process and physical condition are unchanged. Why havent we, as petrolo-

143
How To Name Volcanic Rocks

gists, taken advantage of these ideas and devised a classification that incorpo-
rates them?

One reason for considering the important implications connected to rock


names is to provide a window of communication between volcanic petrolo-
gists and the rest of the Earth science community. A second reason is to
provide a research tool that helps delineate problems with our understanding
of volcanic processes.

To properly place a volcanic rock in a meaningful classification, we suggest


that five factors need to be considered in the identification process: Mafic
index, Silica activity, Al2O3 ratio, Presence or absence of Ca-poor pyroxene,
and Glass index with a chemical characterization of the glass. The big three:
mafic index, silica activity, and Al2O3 ratios, are as important to understand-
ing volcanic processes today as they were a century ago. They should be
included in any set of criteria for identifying rocks. By themselves, however,
they are insufficient. We suggest two other criteria but even with these, the
number of criteria may be insufficient. But with these criteria as a starting
point, other criteria can be added to the set to create a practical, logical, and
genetic classification.

References Cited
Anderson, G., 1996, Thermodynamics of natural systems: New York, New
York, John Wiley & Sons, Inc., 382 p.

Anderson, G. M., and Crerar, D. A., 1993, Thermodynamics in Geochemistry,


The Equilibrium Model: New York, New York, Oxford University Press,
588 p.

Bailey, E. B., Clough, C. T., Wright, W. B., Richey, J. E., and Wilson, G. V.,
1924, Tertiary and Post-Tertiary geology of Mull, Loch Aline, and Oban:
Memoir Geological Survey of Scotland.

Bellieni, G., Justin Visentin, E., Le Maitre, R. W., Piccirillo, E. M., and
Zanettin, B., 1983, Proposal for a division if the basaltic (B) field of the
TAS diagram: Subcommission on the Systematics of Igneous Rocks Cir-
cular, v. 38, no. 102.

Bowen, N. L., 1928, The Evolution of the Igneous Rocks: Princeton, NJ,
Princeton University Press.

144
How To Name Volcanic Rocks

Bowen, N. L., 1937, Recent high-temperature research on silicates and its


significance in igneous geology: American Journal of Science, v. 33, p.
1-21.

Bowen, N. L., 1945, Phase equilibria bearing on the origin and differentiation
of alkaline rocks: American Journal of Science, v. 243A, p. 75-89.

Carmichael, I. S. E., 1962, Pantelleritic liquids and their phenocrysts: Miner-


alogical Magazine, v. 33, p. 80-113.

Carmichael, I. S. E., 1963, The crystallization of feldspar in volcanic acid


liquids: Quarterly Journal of the Geological Society of London, v. 119, p.
95-131.

Carmichael, I. S. E., 1964, Natural liquids and the phonolitic minimum: Geo-
logical Journal, v. 4, p. 55-60.

Carmichael, I. S. E., 1965, Trachytes and their feldspar phenocrysts: Mineral-


ogical Magazine, v. 34, p. 107-125.

Carmichael, I. S. E., 1967, The iron-titanium oxides of salic volcanic rocks


and their associated ferromagnesian silicates: Contributions to Mineral-
ogy and Petrology, v. 14, p. 36-64.

Carmichael, I. S. E., 1967, The mineralogy and petrology of the volcanic


rocks from the Leucite Hills, Wyoming: Contributions to Mineralogy and
Petrology, v. 15, p. 24-66.

Carmichael, I. S. E., Nicholls, J., and Smith, A. L., 1970, Silica activity in
igneous rocks: American Mineralogist, v. 55, p. 246-263.

Carmichael, I. S. E., Turner, F. J., and Verhoogen, J., 1974, Igneous Petrol-
ogy: New York, USA, McGraw-Hill Book Company, 739 p.

Chayes, F., 1956, Petrographic Modal Analysis, An Elementary Statistical


Appraisal: New York, USA, John Wiley and Sons, Inc., 113 p.

Cox, K. G., Bell, J. D., and Pankhurst, R. J., 1979, The Interpretation of Igne-
ous Rocks: London, England, George Allen & Unwin, Ltd., 450 p.

Cross, W., 1897, Igneous rocks of the Leucite Hills and Pilot Butte, Wyo-
ming: American Journal of Science, v. 4, p. 115-141.

145
How To Name Volcanic Rocks

Cross, W., Iddings, J. P., Pirrson, L. V., and Washington, H. S., 1902, A quan-
titative chemico-mineralogical classification and nomenclature of igneous
rocks: Journal of Geology, v. 10, p. 555-690.

Fiesinger, D. W., 1975, Petrology of the Quaternary volcanic centers in the


Quesnel High-lands and the Garibaldi Provincial Park areas, British Co-
lumbia [Ph D thesis]: University of Calgary, 145 p.

Fiesinger, W., D., Nicholls, and J., 1977, Petrography and petrology of
Quaternary volcanic rocks, Quesnel Lake region, East-central British
Columbia, Volcanic Regimes in Canada: L. C. Coleman, and J. M. Hall.
Geological Association of Canada Special Paper, R. A. Baragar, p. 25-38.

Ghiorso, M. S., and Carmichael, I. S. E., 1987, Modeling magmatic systems:


Petrologic applications, in Carmichael, I. S. E., and Ghiorso, M. S., eds.,
Thermodynamic Modeling of Geological Materials: Minerals, Fluids,
and Melts: Reviews in Mineralogy, Mineralogical Society of America, p.
467-499.

Ghiorso, M. S., and Sack, R. O., 1995, Chemical mass transfer in magmatic
processes IV. A revised and internally consistent thermodynamic model
for the interpolation and extrapolation of liquid-solid equilibria in mag-
matic systems at elevated temperatures and pressures: Contributions to
Mineralogy and Petrology, v. 119, p. 197-212.

Hatch, F. H., Wells, A. K., and Wells, M. K., 1972, Petrology of the Igneous
Rocks: London, England, Thomas Murby & Co, 551 p.

Holmes, A., 1921, Petrographic Methods and Calculations: London, England,


Thomas Murby and Co., 515 p.

Irvine, T. N., and Baragar, W. R. A., 1971, A guide to the chemical classifica-
tion of the common volcanic rocks: Canadian Journal of Earth Sciences,
v. 8, p. 523-548.

Jackson, D. B., Swanson, D. A., Koyanagi, R. Y., and Wright, T. L., 1975,
The August and October 1968 East Rift eruptions of Kilauea Volcano, Ha-
waii: United States Geological Survey Professional Paper, v. 890, p. 1-33.

Johannsen, A., 1931, A Descriptive Petrography of the Igneous Rocks: Chi-


cago, Illinois, The University of Chicago Press, v. 1, 318 p.

146
How To Name Volcanic Rocks

Kennedy, W. Q., 1933, Trends of differentiation in basaltic magmas: Ameri-


can Journal of Science, v. 25, p. 239-256.

Kretz, R., 2004, Mineral and rock names: A provocative review: The Cana-
dian Mineralogist (Newsletter Insert), v. 42, p. 18, 22.

Le Maitre, R. W., 2002, Igneous rocks: a classification and glossary of terms:


recommendations of the International Union of Geological Sciences, Sub-
commission on the Systematics of Igneous Rocks, Cambridge University
Press, 236 p.

Le Maitre, R. W., 1989, A Classification of Igneous Rocks and Glossary of


Terms: Oxford, England, Blackwell Scientific Publications Ltd, 193 p.

Macdonald, G. A., and Katsura, T., 1964, Chemical composition of Hawaiian


lavas: Journal of Petrology, v. 5, p. 82-133.

Mandarim-de-Lacerda, C. A., 2003, Stereological tools in biomedical re-


search: Anais da Academia Brasileira de Cincias, v. 75, p. 469-486.

Naumann, T. R., and Geist, D., 1999, Generation of alkalic basalt by crystal
fractionation of tholeiitic magma: Geology, v. 27, p. 423-426.

Nicholls, J., 2000, Thermodynamics of a magmatic gas phase 50 years


later: Comments on a paper by John Verhoogen (1949): Canadian Miner-
alogist, v. 38, p. 1313-1328.

Nicholls, J., and Stout, M. Z., 1988, Picritic melts in Kilauea - Evidence from
the 1967-1968 Halemaumau and Hiiaka eruptions: Journal of Petrology,
v. 29, p. 1031-1057.

Nicholls, J., and Stout, M. Z., 1997, Epitactic overgrowths and intergrowths
of clinopyroxene on orthopyroxene: Implications for paths of crystalliza-
tion, 1881 lava flow, Mauna Loa Volcano, Hawaii: Canadian Mineralo-
gist, v. 35, p. 909-922.

Nicholls, J., Stout, M. Z., and Fiesinger, D. W., 1982, Petrologic variations
in Quaternary volcanic rocks, British Columbia, and the nature of the
underlying upper mantle: Contributions to Mineralogy and Petrology, v.
79, p. 201-218.

147
How To Name Volcanic Rocks

Ryder, G., 1999, Naming Lunar Mare basalts: Quo vadimus redux: NASA
Lunar and Planetary Institute Workshop on New Views of the Moon 2:
Understanding the Moon Through the Integration of Diverse Databases,
p. 55-56.

Shand, S. J., 1916, A recording micrometer for geometrical rock analysis:


Journal of Geology, v. 24, p. 394-404.

Shand, S. J., 1917, A system of petrography: Geological Magazine, v. 6, p.


463-469.

Shand, S. J., 1947, Eruptive Rocks: New York, New York, Hafner Publishing
Company, 488 p.

Stewart, D. B., 1979, The formation of siliceous potassic glassy rocks, in


Yoder, H. S., Jr., ed., The Evolution of the Igneous Rocks, Fiftieth Anni-
versary Perspectives: Princeton, New Jersey, Princeton University Press,
p. 339-350.

Stout, M. Z., Nicholls, J., and Kuntz, M. A., 1994, Petrological and mineral-
ogical variations in 2,500-2,000 yr B.P: lava flows, v. Craters of the Moon
Lava Field, Idaho. Journal of Petrology, 35, p. 1681-1715.

Stout, M. Z., 1975, Mineralogy and petrology of Quaternary lavas, Snake


River Plain, Idaho and the cation distribution in natural titanomagnetites
[MSc thesis]: University of Calgary, 150 p.

Stout, Z., M., and Nicholls, J., 1977, Mineralogy and petrology of Quaternary
lavas from the Snake River Plain, Idaho: Canadian Journal of Earth Sci-
ences, v. 14, p. 2140-2156.

Tilley, C., 1950, Some aspects of magmatic evolution: Quarterly Journal of


the Geological Society of London, v. 106, p. 37-61.

Trupia, S., 1992, Petrology of nephelinites and associated ultramafic nod-


ules of Volcano Mountain, Yukon Territory [M Sc thesis]: University of
Calgary, 123 p.

Wade, A., and Prider, R. T., 1940, The leucite-bearing rocks of the west
Kimberley area, Western Australia: Quarterly Journal of the Geological
Society of London, v. 96, p. 39-98.

148
How To Name Volcanic Rocks

Williams, H., Turner, J., F., Gilbert, and M., C., 1954, Petrography, An
Introduction to the Study of Rocks in Thin Section: San Francisco, W.H.
Freeman and Company.

Wright, T. L., 1971, Chemistry of Kilauea and Mauna Loa in space and time:
United States Geological Survey Professional Paper, v. 735, p. 1-40

Yoder, H. S., Jr., and Tilley, C., 1962, Origin of basaltic magmas: An experi-
mental study of natural and synthetic rock systems: Journal of Petrology,
v. 3, p. 342-532.

149

También podría gustarte