You are on page 1of 20

1

Characterisation and the diagenetic transformation of nonand micro-crystalline silica minerals


DAVID R. LEE
Department of Earth and Ocean Sciences, University of Liverpool, 4 Brownlow Street,
Liverpool L69 3GP, UK (e-mail: d.r.lee@liverpool.ac.uk)
Abstract: Non- and micro-crystalline quartz is found abundantly in the
sedimentary realm as non-crystalline opal (opal-A), microcrystalline opal (opalCT/-C) and microcrystalline quartz (chalcedony, quartzine) that form the principle
constituents of cherts and agates. The various silica minerals can be characterised
by their microstructural features; texture, crystallite size, orientation, determined
using microscopic methods (optical, SEM, TEM). The total water content and
concentration ratios of molecular water to silanole water are characteristic for the
different SiO2 phases. During diagenesis the silica minerals follow an opal-A
opal-CT/-C
microcrystalline quartz transformation. The diagenetic
transformation occurs predominantly via a dissolution-reprecipitation mechanism
that is governed by time and temperature. X-ray diffraction and EBSD techniques
can be applied to determine the increasing structural order during diagenesis, with
XRD patterns showing more intense and narrow peaks with increasing
crystallinity. Infrared spectroscopy has been used to demonstrate the behaviour of
the water species present in SiO2 minerals during heating, with molecular and
silanol group waters removed at different temperatures due to differing bonding
relationships with the microstructure. These analytical techniques have the
potential be used simultaneously to characterise each diagenetic phase transition.

Introduction
Authigenic non- and microcrystalline silica minerals are abundant in a number of
forms. Due to the various growth fabrics and microstructures present, characterising
individual SiO2 polymorphs has proved complicated. However, several varieties and subvarieties have been distinguished; amorphous quartz (opal-A), cryptocrystalline quartz
(opal-CT/-C) and microcrystalline quartz (chalcedony, quartzine, etc) (Hesse, 1989;
Flrke et al., 1991; Stamatakis et al., 1991; Graetsch, 1994; Knauth, 1994). Non- and
micro-crystalline silica minerals are pure SiO2 H2O which contain typically less than 1
wt.% non-volatile impurities (Flrke et al., 1991; Graetsch, 1994). Numerous
investigations have been conducted in a bid to determine the microstructures and varying
water contents which characterise the various minerals and how these properties change
during diagenesis (Jones & Segnit, 1971; Flrke et al., 1982; Graetsch et al, 1987; Flrke
et al., 1991; Elzea et al., 1994; Graetsch, 1994; Graetsch et al., 1994).
Non- and micro-crystalline silica minerals are evident in a host of siliceous rocks;
agate and chert (bedded and nodular), with the occurrence of opal and high water content
between and microcrystalline quartz in cherts suggesting that they are diagenetic
precipitates derived from the dissolution of amorphous silica (Siever, 1962; Stamatakis et
al., 1991). At low temperatures amorphous silica (opal-A) is generally precipitated
inorganically or biogenically from natural aqueous solutions. Biogenic opal-A is formed
by the dissolution of organisms; diatoms, radiolaria and siliceous sponges. The high
surface areas of the organisms can give rise to abundantly high solubilities and rates of

solution that can lead to supersaturation and the precipitation of siliceous oozes (Siever,
1962; Williams et al., 1985; Elzea et al., 1994). Inorganic precipitation is evident in
silcrete cements, volcanic systems and precious opal deposits (Williams & Crerar, 1985;
Hesse, 1989; Stamatakis et al., 1991; Hendry & Trewin, 1995; Alexandre et al., 2004;
Haddad et al., 2006).
During diagenesis siliceous deposits undergo mineralogical changes from noncrystalline silica (opal-A) to microcrystalline opal (opal-CT/-C) to microcrystalline quartz
(fig. 1) ( Williams et al., 1985; Flrke et al., 1991; Cady et al., 1996; Lynne & Campbell,
2004) This transformation sequence is the most widely known mechanism for the
formation of chert rocks, with the initial opaline gel converting to crystalline quartz
(Iijima & Tada, 1981; Knauth, 1994; Hattori et al., 1996). In volcanic silica deposits (e.g.
agates) formed at higher temperatures opal-CT/-C seem to form directly without the
earlier opal-A dissolution phase present in the formation of chert (Graetsch et al., 1994).
In general it has been reported that the transformation results in increasing crystallinity
and decreasing solubility of the silica polymorphs (Willey, 1980; Williams et al., 1985).
However, the nature of the different diagenetic schemes is complicated and thus cannot
be readily transferred from one setting to another (Knauth, 1994). This review aims to
characterise non- and microcrystalline silica minerals by consolidating existing
observations and nomenclature as well as summarising current perceptions on the
mechanisms involved in the diagenetic transformation with the application of different
analytical techniques.

Figure 1. Schematic diagram of


diagenetic zones in the sedimentary
realm displaying the transformation of
opal-A to quartz with intermediate
phases (from Williams et al., 1985).

Characterisation of silica minerals


Silica polymorphs exhibit various physical properties which can be used to
characterise the numerous phases present in silicic rocks such as cherts and agates. The
structural variability of SiO2 and varying amounts of water present in the structure in the
past has resulted in ambiguity when defining particular silica phases (Flrke et al., 1991).
Table 1 shows the currently accepted nomenclature proposed by Flrke et al. (1991)
which is adopted in this review.

Microstructure
Due to their structural characteristics micro- and non-crystalline silica minerals can be
divided into three groups: microcrystalline quartz, microcrystalline opals and noncrystalline opal (Graetsch, 1994). After Flrke et al. (1991), the term microcrystalline
relates to a polycrystalline microstructure of individual crystallites which can only be

resolved microscopically (i.e. utilising transmission electron microscopy (TEM),


scanning electron microscopy (SEM), etc).
Non- and microcrystalline opal
Opals are hydrous silica polymorphs that typically contain varying amounts of
amorphous silica, tridymite and cristobalite (Heaney, 1993). Jones and Segnit (1971)
stated that although chemically indistinguishable, naturally forming opaline forms could
be categorised into three distinct groups (opal-C, opal-CT and opal-A) from X-ray
diffraction patterns related to crystal structure. This grouping has been divided into two
categories: (1) microcrystalline opal (opal-C, opal-CT) and (2) non-crystalline opal (opalA) (Graetsch, 1994). Microcrystalline opals, also referred to as cryptocrystalline quartz
(Alexandre et al., 2004) and paracrystalline quartz (Herdianita et al., 2000) are common
diagenetic products of opal-A (Knauth, 1994), exhibiting disordered stacking faults of cristobalite and tridymite (Graetsch et al., 1994). They are grouped as opal-CT or opal-C
depending on the degree of crystallinity and disorder in the crystal structure (Elzea et al.,
1994).
Opal-A
Opal-A is the most common form of highly disordered opaline silica, typically
deposited biogenically (Knauth, 1994). Langer and Flrke (1974) categorised opal-A into
opal-AN (networked opal such as hyalites) and opal-AG (gel-like). Due to the high
temperature genesis of hyalitic opal, this material is largely ignored in this paper. The
latter opal-AG type relate to those precipitated biogenically from aqueous solutions at low
temperatures (<100c) (Stamatakis et al, 1991). Precious and potch opals are categorised
as opal-AG (Flrke et al, 1991). Precious opal is comprised of close packed homometric
spheres of amorphous silica, typically 1 8 m in diameter (fig. 2a); with additional silica
cement partly filling the interstices (Darragh et al., 1966; Herdianita et al., 2000). These
are visible using scanning electron microscopy. Graetsch (1994) stated that the structure
of the spheres is usually highly disordered with abundant stacking faults. Potch opals
comprise of irregularly packed heterometric spheres with the interstices commonly filled
with silica cement (Graetsch, 1994).
Opal-CT
Opal-CT has a greater degree of disorder than opal-C and has been interpreted as
microcrystalline -cristobalite and tridymite in a matrix of amorphous silica (Elzea et al.,
1994). Whereas in opal-C cristobalite stacking appears to dominates, in opal-CT the
amount of cristobalite and tridymite stacking is approximately equal (Flrke et al., 1991),
with Graetsch (1994) stating that opal-CT typically contains 30 50% tridymite type
stacking. This interpretation is bolstered by Rice et al. (1995) who stated that opal-CT is
most likely to be a cristobalite structure incorporating randomly intergrown tridymite,
creating the disordered and poorly crystalline nature of the polymorph. Wilson et al.
(1974) observed that in some crystals of opal-CT the predominant component is tridymite
as opposed to cristobalite. The apparent ambiguity in determining the nature of opal-CT is
shown by De Jong et al. (1987) who stated that NMR spectra determined from opal-CT is
closer to amorphous silica than to cristobalite and tridymite crystal structuring. Massy
opal comprises of small thin platelets forming lepispheric aggregates between 1 and 10
m in diameter (Williams et al., 1985) (fig. 2b), which show random orientation of the
crystals with a high degree of stacking disorder. Herdianita et al. (2000) stated that the
lepispheres appear to pseudo-morph the original opal-A spheres.

opal-A

opal-CT

opal-C

chalcedony

platy, lepidospheric
(isotropic)
close packing of
homometric spheres
irregular packing of
heterometric spheres

Precious opal

Potch opal

fibrous (length-slow)

massy opal

lussatite

platy (length-fast)

1 8 m spheres

1 10 m spheroids

10 100 nm

10 100 nm

~ 100 200 nm

parabolic fibre bundles


(length-slow)

quartzine

lussatine

~ 100 200 nm

>1 m
(typically 50 350
nm)

parabolic fibre bundles


(length-fast)
radiating spherulites
(length-fast)

20 - 5 m

20 50 m

>50 m

Crystal size

granular

crystalline

Microstructure

horizontally banded

wall-lining

mesoquartz

macroquartz

Sub-variety

10 12 wt.%

3 10 wt.%

1 3 wt.%

0.5 2.5 wt.%

Total water
(H2Otot)

Table 1. Nomenclature and characteristics of non- and micro-crystalline silica minerals (adapted from Flrke et al., 1991).

non-crystalline
opal

microcrystalline
opal

microcrystalline
quartz

megaquartz

crystalline quartz

fine-quartz

Variety

Crystal Structure

0.1 0.7

0.02

~0

H2OSiOH
H2Omol

Darragh et al., 1966;


Langer & Flrke,
1974; Graetsch,
1994; Herdianita et
al., 2000

Langer & Flrke,


1974; Williams et al.,
1985; Graetsch,
1994; Knauth, 1994;
Cady et al., 1996,
1998; Alexandre et
al., 2004

Langer & Flrke,


1974; Heaney, 1993;
Graetsch, 1994;
Knauth, 1994; Cady
et al., 1996, 1998;
Moxon et al., 2006

Hesse, 1989;
Graetsch, 1994;
Knauth, 1994

Hesse, 1989; Hendry


& Trewin, 1995

Additional
References

The high degree of disorder in the structure results in an almost isotropic optical
character, with the crystallites showing a mottled appearance in TEM micrographs
(Graetsch, 1994; Cady et al., 1996). The other variety is fibrous opal-CT, termed
lussatite, forms as bundles of length-slow fibres up to 250nm in length which exhibit a
parallel texture (Cady et al., 1996). As a result lussatite displays higher birefringence than
massy opal (Flrke et al., 1991; Graetsch, 1994).
Opal-C
Opal-C is regarded to be the most structured and crystalline of the opal varieties
outlined previously, with limited evidence of disordered tridymite stacking compared to
opal-CT (Jones & Segnit, 1971). Graetsch (1994) stated that the structure of opal-C can
contain between 20 30% tridymite type stacking whereas other studies have shown no
tridymite present in opal-C (Rice et al., 1995). Although Jones and Segnit (1971) stated
that opal varieties are structurally distinct, these variations in determining what exactly
defines opal-C (as well as other opal varieties) highlight the complexities present in the
current nomenclature. Opal-C is composed of platy crystallites, termed lussatine, which
aggregate into regions of parallel texture which show a smooth transition into regions of
random orientation. In crossed polars it shows a patchwork pattern of alternating high and
very low birefringence and displaying a length-fast optical character (Flrke et al., 1991).
Microcrystalline quartz
Quartzine
In older literature referred to as length-slow chalcedony, quartzine forms between
wall-lining chalcedony (Xu et al., 1998) and displays distinctly different properties to
length-fast chalcedony in the fibrous form as proposed by Heaney (1993) and adopted in
this text. Quartzine comprises of fibres, 100 200 nm in length (Cady et al., 1998), which
are elongated parallel to the crystallographic c-axis (Graetsch et al, 1987; Hesse, 1989).
This creates a high refractive index in this direction and is termed length-slow (Flrke et
al., 1982; Cady et al., 1998 Wahl et al., 2002).
Chalcedony
Heaney (1993) reported that the term chalcedony has been applied loosely to a
number of microcrystalline silica polymorphs and that technically the term relates to
microcrystalline fibrous silica <1 m in length. Typically fibre lengths are between 50 and
350 nm (Miehe et al., 1984; Heaney, 1993; Cady et al., 1998). Fibres in chalcedony are
elongated in the direction of the a-axis [110], perpendicular to the crystallographic c-axis
of quartz crystallites (Graetsch et al., 1987). With quartz exhibiting positive optical
properties this results in a lower refractive index in the direction of the fibre. This
negative optical characteristic of the fabric is called length-fast (Flrke et al., 1982;
Graetsch, 1994; Wahl et al., 2002). The crystallographic system in chalcedony appears to
be monoclinic or triclinic, not trigonal as seen in quartz (Heaney, 1993). Length-fast
chalcedony forms either as wall-lining or horizontally banded layers (Graetsch, 1994),
both exhibiting distinct growth fabrics. Wall-lining chalcedony forms parallel-fibrous
aggregates in which the c-axis is twisted around the fibre axis causing distinctive
rhythmic banding of extinction (Runzelbnderung) present between cross polars (Flrke
et al., 1983) (fig. 2c). Horizontal banded chalcedony comprise of close packed radiating
spherulitic fibres which show no evidence of c-axis twisting unlike wall-lining
chalcedony varieties (Miehe et al., 1984). Heaney et al. (1994) proposed from electron
diffraction patterns in chalcedony that the fibres created a randomly orientated fabric.
However, it has been noted by Flrke et al. (1982) that individual quartz crystallites

within a fibrous chalcedonic structure show a better mutual orientation than in a finer
granular structure present in fine-quartz. Using TEM Cady et al. (1998) showed that the
smaller randomly orientated crystallites are aligned in a common direction. It thus
appears that characterisation of chalcedony has a strong dependence on the scale of
observation.
Fine-quartz
Fine-quartz is a microcrystalline variety of quartz with a granular texture which
displays a random mutual orientation of strained grains with sizes typically <20 m.
(Flrke et al., 1982; Hesse, 1989). Due to this fabric, in crossed polars fine-quartz
displays undulatory extinction (Flrke et al., 1991). The grain fabric is typically large
enough to be determined using optical microscope methods (Alexandre et al., 2004).

Figure 2. Images of microtextures present in silica polymorphs. A) SEM micrograph of opal-A exhibiting
typical 48 m diameter spheres with smaller 12 m diameter spheres, scale bar 10 m (Herdianita et al.,
2000). B) SEM micrograph showing opal-CT lepisphere comprised of thin bladed crystals (<10 m
diameter), scale bar 10 m (Rodgers et al., 2004). C) Optical image with crossed polars of wall-lining
chalcedony showing rhythmic extinction banding (Runzelbnderung), scale bar 1 mm (Graetsch et al.,
1987). D) SEM micrograph showing well-developed coarse quartz crystals (~0.2 mm length), scale bar 0.1
mm (Herdianita et al., 2000).

Crystalline quartz
Due to the numerous forms in which it can manifest, crystalline quartz has been
broadly defined as well defined crystals which do not contain the crystallographic

complexities associated with disordered microcrystalline quartz and opals (Knauth,


1994). Crystalline quartz forms prismatic crystals (fig. 2d.) which typically elongate
parallel to the c-axis, displaying a positive refractive index that is determinable using
optical microscopy (Hesse, 1989; Heaney et al., 1994). Hendry & Trewin (1995)
proposed that equant quartz fabric with crystals 10 50 m are termed mesoquartz, whilst
those >50 m termed macroquartz.

Water
Using infrared spectroscopy Langer and Flrke (1974) determined two water types
present in silica minerals; molecular water (H2Omol) and silanol groups (H2OSiOH). Both
molecular and silanol group are divided into type A; isolated non-hydrogen bonded
molecules and hydroxyl groups trapped in the structure as fluid inclusions, and type B;
strongly hydrogen-bonded accumulations of water molecules or hydroxyls either within
the structure or on external and internal surfaces (Flrke et al., 1982; Flrke et al., 1991;
Graetsch, 1994; Herdianita et al., 2000). The total water content and relative ratio of
different water species within the total water content are characteristic for micro- and
non-crystalline silica minerals (fig. 3) (Flrke et al., 1991).

Figure 3. Concentrations of
molecular water (H2Omol)
and silanol group water
(H2OSiOH) in micro- and
non-crystalline
silica
minerals. (adapted from
Flrke et al., 1991)

Non-crystalline opal
Most of the water present in non-crystalline opals is molecular water, with opal-A
containing 10 - 12 wt. % (H2O)mol (Flrke et al., 1991; Knauth, 1994). This relatively
high amount compared to more crystalline silica polymorphs is attributed to the porous
nature of the opal-A. Flrke et al. (1982) stated that H2Omol present in opal is type A, with
liquid water forming fluid inclusions in the microstructure (Graetsch, 1994). Noncrystalline opal has a higher (H2O)SiOH content (0.5 1.0 wt. %) than microcrystalline
opal; with a wt.% H2OSiOH / wt.% H2Omol concentration ratio between 0.1 and 0.7 (Langer
& Flrke, 1974; Alexandre et al., 2004). This relative abundance of hydroxyls has been
attributed to the highly disordered nature of the structure in non-crystalline opal (Flrke et
al., 1991).

Microcrystalline opal
Total water content in microcrystalline opals is dependent on the degree of disorder in
the structure, ranging from 1 3 wt. % in opal-C (Graetsch et al., 1985; Flrke et al,
1991) and 3 10 wt. % in opal-CT (Graetsch et al., 1987; Flrke et al., 1991; Cady et al.,
1996). In microcrystalline opals most of the water present is type B molecular water
present in pores, with 0.1 0.3 wt. % silanol group water present. Due to the relatively
high content of type B molecular water, the majority of the hydroxyls present are type B
silanoles (Flrke et al., 1991). Graetsch et al. (1987) estimated that one third of the type B
silanoles present are surficial, with the remaining two-thirds on internal surfaces. The
ratio of wt.% H2OSiOH / wt.% H2Omol of opal-CT is approximately 0.02 (Langer & Flrke,
1974; Alexandre et al., 2004). As the structure in opal-CT becomes more ordered and
compacted, the water content in microcrystalline opals decreases (Flrke et al., 1991).
Microcrystalline quartz
Microcrystalline silica can be characterised by a higher total water (wt.% (H2O)SiOH and
mol) than crystalline quartz (Flrke et al., 1991). The water content in chalcedony and
quartzine ranges from 0.5 to 2.5 wt. %, with the water content in these minerals decreases
significantly with increasingly crystallite size as structural defects become less common
(Miehe et al., 1984; Moxon et al., 2006). Flrke et al. (1982) stated that the content of
silanol group water in chalcedony ranges from 0.2 to 0.9 wt.% and that molecular water
ranges from 0.5 to 1.4 wt.%. In quartzine silanol group water content ranges from 0.4 to
0.6 wt.% with molecular water content similar to chalcedony (Flrke et al., 1991). Langer
and Flrke (1974) calculated that for microcrystalline quartz (e.g. chalcedony) the relative
concentration ratio of wt.% H2OSiOH / wt.% H2Omol is 0. Almost all the molecular water in
microcrystalline quartz is hydrogen bonded (type B), as are SiOH groups which are
hydrogen bonded on internal surfaces and imperfect crystalline regions, with
concentrations at low and high angle boundaries (Frondel, 1982; Graetsch et al., 1987).
The observed oscillatory nature of the fibres in chalcedony has been attributed to the
zones of low and high concentrations of hydroxyls at structural defects in the
microstructure (Heaney, 1993).
As silica polymorphs become more ordered the total water content (H2Otot) decreases
and the structures become more compact, resulting in increasing density (fig. 4).
Although non- and microcrystalline opals exhibit a broad range of total water contents (1
10 wt.%), they all exhibit densities between 2 and 2.2 gcm-3 (Graetsch, 1994). In more
ordered microcrystalline quartz H2Otot is relatively lower, resulting in greater density
(Graetsch, 1994; Moxon et al., 2006).

Diagenesis of silica polymorphs


As referred to earlier, the diagenetic sequence for silica minerals proceeds; opal-A
opal-CT/-C
microcrystalline quartz (fig. 1), with the polymorphs showing
increasing degrees of structural order and crystallinity. The phase transitions between the
silica polymorphs are complex, with a number of polymorphs present at the same time,
resulting in complicated mineralogical and textural patterns at the microscale (Herdianita
et al., 2000). Ideally only one polymorph of silica can be present at one time; however
kinetics and rates of transformation among the polymorphs are slow enough for
metastable polymorphs to exist (Williams et al, 1985).

Figure 4. Densities of non-crystalline opal, microcrystalline opals and microcrystalline quartz


vs. the total water content (adapted from Graetsch, 1994).

Stamatakis et al. (1991) noted factors which affect the diagenetic transformation of
non-crystalline opal-A through to microcrystalline quartz are: burial depth, time, heat
flow and host rock lithology. In the natural environment within proximity to igneous sills,
opal-A has been observed having transformed directly to quartz, with metastable phases
(opal-CT/-C) formed in sediment away from the heat source (Stamatakis et al., 1991).
The bypassing of phase transformations as outlined above implied that heat flow can have
a significant effect on the diagenesis of silica minerals. Many authors agree that
temperature and time are the dominant factors controlling the rate of silica diagenesis,
with burial history often intrinsically linked with these controls (Jones & Segnit, 1971;
Rice et al., 1995; Cady et al., 1996). The multiple stage diagenetic transformation of opalA, via opal-CT/-C, to quartz has been recognised as a series of complex dissolutionprecipitation events (Stein & Kirkpatrick, 1976; Williams et al., 1985; Williams & Crerar,
1985; Hendry and Trewin. 1995).
Opal-A opal-CT
Jones and Segnit (1971) proposed that the transformation of opal-A to opal-CT is
proceeded by the formation and nucleation of discrete cristobalite and tridymite
sequences which eventually coalescence to form larger crystallites in the non-crystalline
matrix. Williams et al. (1985) stated that the reaction proceeds when temperatures reach
approximately 50C, whereas other studies showed that opal-CT forms at ~2C (Knauth,
1994). However, it is accepted that the transformation to quartz can only occur once
structural reorganisation has reached an advanced state, with increased temperatures
acting as a catalyst for the transition (Robertson, 1977). The proposed reorganisation of
the structure proceeds with an increase in cristobalite ordering that is coincident with a
loss of tridymite from the crystal structure (Williams & Crerar, 1985).
Opal-CT microcrystalline quartz
There is much debate regarding the mechanism responsible for the apparent structural
phase transformation that occurs during the diagenetic transformation of opal-CT to
microcrystalline quartz (Shoval et al., 1997). Williams et al (1985) stated that retention of

10

fabric, observed by the presence of pore filling opal-CT lepidospheres in quartz cherts,
was evidence of solid-state transformation of opal-CT to quartz. Such fabrics have also
been attributed to slow dissolution rates and rapid precipitation in a grain by grain
replacement process (Heaney, 1993; Knauth, 1994). For either mechanism to occur
structural Si-O bonds have to be broken. A key concept in the argument against the solidstate mechanism is that the transformation must break Si-O bonds which would require
increased activation energies in the solid state. It is because of this factor that
reconstructive transitions to quartz are slow, allowing cristobalite to exist in a metastable
state in temperatures far below its thermal stability (Williams et al, 1895; Swainson et al,
1993). Murata et al. (1977) observed 18O/16O ratio changes across diagenetic boundaries
in chert. The authors observed abrupt decreases in 18O across opal-A/opal-CT and opalCT/quartz boundaries and proposed that the isotopic fractionation was due to a solid-state
mechanism. Stein & Kirkpatrick (1976) observed that with water and OH acting as
catalysts in the transition, the initial nucleation and growth of the crystallites was more
suited to a dissolution-reprecipitation mechanism.
However, Cady et al. (1996) observed with increasing diagenesis the boundaries
between opal-CT and quartz become more ordered. TEM micrographs revealed that
lussatite fibres present at the opal-CT quartz boundary areas displayed an increase in
ordering, whereas interboundary regions showed an apparent increase in disorder. The
authors concluded that these observations showed evidence of solid-state structural
reworking. It is therefore currently believed that the earliest quartz may form by a solidstate mechanism before the reaction proceeds predominantly as a dissolutionreprecipitation transformation mechanism once the concentration of silica in the
interstitial pore fluid reaches the solubility of quartz (Cady et al. 1996).
Microcrystalline quartz macrocrystalline quartz
It is widely accepted that the diagenetic transformation of microcrystalline quartz to
quartz is by dissolution-precipitation. The first quartz to precipitate is typically of very
fine grain size with poor crystallinity, though once precipitated the crystal size begins to
increase, resulting in a decrease in surface defects, in the growing crystallites and creating
a more compact structure (Moxon et al., 2006). Hurst (1981) showed that dissolution and
replacement textures formed during diagenesis of quartz are not random but seem to be
governed by the crystallographic properties of the detrital quartz grains. This assertion is
supported by Flrke et al (1982) who showed that individual quartz crystallites formed
within the existing fibrous microcrystalline structure show a better mutual orientation
than those precipitated homogenously.

Analytical techniques
Numerous methods have been applied to observe changes in silica polymorphs during
diagenesis; x-ray diffraction, infrared spectroscopy, electron back-scattered diffraction.
These techniques have been employed to varying extents to examine changes in structural
order, water speciation and crystallinity of the silica phases and during diagenesis (Jones
& Segnit, 1971; Langer & Flrke, 1974; Flrke et al., 1991; Graetsch, 1994).
X-ray diffraction
X-ray diffraction (XRD) has been used extensively to characterise silica polymorphs
by classifying phases due to the degree of crystallographic ordering present in the
structure (Jones & Segnit, 1971). Structural disorder and heterogeneous distribution of
water in the structure causes a reduction in the intensity of XRD patterns, resulting in
increased diffusivity. In microcrystalline quartz and opals this creates distinctive patterns

11

for different silica phases with different structural identities (Flrke et al., 1991). Opal-A,
opal-CT and opal-C are notably structurally distinct and are routinely characterised due to
their XRD patterns (fig. 5). Opal-A produces a prominent single diffuse band centred at
approximately 4 (Jones & Segnit, 1971). Although varying degrees of stacking disorder
and tridymite can be present in opal-CT, it is characterised by two broad reflections; a
broad strong (101) peak centred at 4.09 with a weaker peak at 4.32 , and a subsidiary
reflection at about 2.50 . (Jones & Segnit, 1971; Iijima & Tada,1981; Elzea et al., 1994;
Lynne & Campbell, 2004). It has been observed that during diagenesis opal-CT becomes
more ordered, resulting in the formation of opal-C. In XRD patterns it is characterised by
lower d-spacings than opal-CT, with a more intense single strong (101) peak within 4.04
4.06 , a moderate peak at 2.49 and weak peaks present at ~2.85 and ~3.14
(Mizutani, 1977; Iijima & Tada, 1981; Graetsch et al., 1987). In opal-CT and -C the
intensity bands at 4.1 and 2.5 are representative of cristobalite and tridymite stacking
sequences (Flrke et al., 1991).
The variation in X-ray trace patterns reflects changes in structural order/disorder from
opal-A, which lacks long range order, to mixed opal-CT stacking sequences. This is
reflected in XRD patterns with the broadly shaped trace for non-crystalline opal-A
gradually becoming modified to the sharper peaked XRD pattern of opal-CT, with a
distinct tridymite shoulders upon the formation of opal-C (Florke et al., 1991; Graetsch,
1994; Lynne & Campbell, 2004).
Figure 5. XRD patterns
for micro- and noncrystalline opals. A) OpalA broad XRDP centred at
~4.09 . B) Sharp peaked
opal-A > -CT XRDP
centred at 4.09 . C) opalCT > -A XRDP centred at
4.09 with tridymite
shoulder at 4.23 with
development of subsidiary
peak at ~2.50 . D)
Sharper peaked opal-CT
XRDP centred at 4.09
(adapted from Lynne &
Campbell, 2004).

The full width at half-maximum intensity (FWHM) of the ~4 diffraction band has
been applied as a guide to the degree of lattice order/disorder within each silica phase,
with narrower FWHM values indicative of greater degree of ordering (Elzea et al., 1994;
Graetsch et al., 1994). Typical FWHM values for opal-A are 1.31 0.02, for opal-CT,
0.27 0.06 and 0.05 0.01 for opal-C. During diagenetic transformation, mixtures of
opal-A and opal-CT are evident and can be qualitatively determined due to degree of
broadness and sharpness of peaks. Typical FWHM values for opal-A>opal-CT are 0.5
and for opal-CT>opal-A are 0.26 0.33 (Graetsch et al., 1987; Graetsch et al., 1994;
Lynne and Campbell, 2004). Lynne and Campbell (2004) observed that incremental steps
on X-ray diffraction patterns were coincident with marked morphological transitions on
the microscale. The shift in peak position from 4.0 0.03 to 4.09 has been interpreted
as the initial formation of opal-CT, with the latter peak position reported for opal-CT
(Flrke et al., 1991; Graetsch, 1994; Elzea et al, 1994). From SEM micrographs, the
authors found that the band shift was almost coincident with the formation of platy massy

12

opal-CT, with the mineralogical shift occurring before the morphological signature, in the
form of bladed lepispheres, was evident. With increasing diagenesis and ageing, resulting
in improved crystallinity, the degree of order in the structure of silica polymorphs is
expressed in XRD patterns exhibiting increasing peak intensities and decreasing FWHM
values as the peaks become tighter (fig. 6) (Williams et al., 1985; Elzea et al., 1994;
Moxon et al., 2006).

Figure 6. XRD patterns showing


sharpening
and
narrowing
of
reflections with increasing crystallinity
with age of agates. Sample age; A) 38
Ma B) 133 Ma C) 285 Ma D) 412
Ma E) 1.1 Ga F) 1.84 Ga and G)
macrocrystalline quartz. (adapted from
Moxon et al., 2006)

Infrared spectroscopy
Because X-ray diffraction is only applicable to materials with a long-range order, it is
unable to determine non-crystalline materials (e.g. opal-A). In contrast, Fourier transform
infrared spectroscopy (FT-IR) depends on the response of short range molecular scale
energetic vibrations such as O H stretching and bending (Langer & Flrke, 1974;
Kronenberg, 1994). As a result FT-IR can be applied equally well to crystalline and noncrystalline silica polymorphs, making it ideal for observing diagenetic transformations
from opal-A through to quartz (Rice et al., 1995). During phase changes experienced by
various silica polymorphs through diagenetic processes the behaviour of water present in
the structure can change considerably (Yamagishi et al., 1997). The vibrational motion of
hydroxyl ions and water molecules present in silica exhibit characteristic peaked
absorption bands using IR (fig 7a).
Sharp absorption peaks below 1400 cm-1 are recognised as relating to stretching and
bending of Si-O bonds (Langer & Flrke, 1974). Broad absorption bands between 4000
2500 cm-1 are due to O H stretching vibrations, with the broad asymmetric band at
around 3400 cm-1 due to hydrogen bonded molecular water H2Omol (Graetsch et al. 1985;
Kronenberg, 1994). A band present between 3750 cm-1 and 3500 cm-1 has been attributed
to SiOH groups, with a notable sharp peak observed at 3585 cm-1 (Graetsch et al., 1985).
The sharp peak at 3585 cm-1 is assigned to SiOH located at structural defects and
typically displays a shoulder at 3595 cm-1 (Frondel, 1982). In opal-C an additional is
present at 3650 cm-1 (Graetsch et al., 1985, 1987). Near infrared (NIR) spectra for
microcrystalline quartz (fig. 7b) exhibits a broad absorption band around 4500 cm-1 with
an asymmetric peak around 5200 cm-1. This peak has been attributed to O H bending
and stretching related to the presence of molecular water. The broader band around 4500
cm-1 shows finer peaks at 4520, 4450 and 4350 cm-1 which have been attributed to O H
stretching and Si-OH bending (Langer & Flrke, 1974; Flrke et al., 1982). In opal-C, the
higher energy peak in the ~4500 cm-1 band is absent (Graetsch et al., 1985).

13

Figure 7. a) mid-infrared (IR) and b) near infrared (NIR) absorption spectra for agate showing vibration
modes of water species present (from Yamagishi et al., 1997).

Moxon et al. (2006) observed that IR absorption bands for agates displayed a
reduction in intensity accompanied by broadening of peaks with increasing age due to
increasing crystallinity and reduction in water content during diagenesis. Heating
experiments on microcrystalline quartz have shown that upon heating to relatively low
temperatures (150C), a significant decrease in the 3400 cm-1 band assigned to physically
absorbed molecular water is observed (Langer & Flrke, 1974; Graetsch et al., 1985;
Kronenberg, 1994). This change has been attributed to the rapid loss of loosely bound
molecular water (0.1 0.3 wt.%) from grain boundaries and pore spaces (Graetsch et al.,
1987; Yamagishi et al., 1997). Graetsch et al. (1987) noted that rapidly deposited
horizontally banded chalcedony has a lower tendency to trap molecular water compared
to wall-lining chalcedony. At temperatures above 350C Si-OH absorption bands at
~3740 and 2340 cm-1 appear (fig. 8).

Figure 8. O H stretching vibration bands of H2OSiOH and H2Omol in chalcedony when heated to various
temperatures. Note the decrease in broad 3400 cm-1 band attributed to absorbed water and the subsequent
increase in the 3740 and 2340 cm-1 peaks related to silanol group water (adapted from Graetsch et al.,
1985).

14

Graetsch et al. (1985) attributed this appearance to low level surface hydration following
the previous dehydration event, whereas Yamagishi et al. (1997) stated that the
appearance is due to dehydration of tightly bonded molecular water in the structure,
allowing the release of the non-hydrogen bonded silanol groups that were previously
interacting with the molecular water. Graetsch et al., (1985, 1987) observed once
temperature had reached 600C, between 0.4 and 1.3 wt.% H2Omol held in fluid inclusions
is lost. At higher temperatures the absorption peak at 3585 cm-1 shifts to 3620 cm-1 and
continuously decreased due to further expulsion of SiOH groups from the microstructure.
Focusing on Si-O bond absorption bands at ~470 and ~500 cm-1, Rice et al. (1995)
observed that during simulated diagenesis the 470cm-1 peak (opal-A) gradually shifts to
500 cm-1 (opal-CT) (fig. 9a). The relatively intensity ratio of the two peaks ( = Iopal-A /
Iopal-CT) through time exhibited three distinct phases (fig. 9b). This opal-A to opal-CT
phase transition has been attributed to a dissolution-precipitation process followed by
increasing structural ordering in opal-CT, as cited by Williams et al (1985) and Cady et
al. (1996).
Figure 9a. FT-IR spectra (350
600 cm-1) showing change in
peak position and intensity
through time of 470 and 500 cm-1
absorption bands. From Rice et
al., 1995.

Figure 9b. FT-IR results


interpreted as three stages
present in the opal-A to opal-CT
transformation;
i)
Opal-A
dissolution.
ii)
opal-CT
precipitation.
Abrupt
slope
change indicating change in
reaction mechanism (relatively
fast process). iii) opal-CT
ordering. Abrupt slope change
indicating change of reaction
(relatively slow process). From
Rice et al., 1995.

Electron Backscattered Diffraction


To determine the crystallography of silica polymorphs electron backscattered
diffraction (EBSD) can be employed to reveal microstructural information about the
crystal structures. The technique revolves around the orientation dependent scattering of
electrons at lattice planes within crystalline materials, resulting in distinctive diffraction
patterns depending on the material present (Neumann, 2000). The principles of EBSD
have been presented previously by Venables & Harland, 1973; Lloyd, 1987; Lloyd &
Freeman, 1991; Prior et al. 1999. Recently EBSD techniques have been applied to
problems related to diagenesis of quartz cements (Haddad et al, 2006). The authors

15

utilised electron backscatter patterns (EBSPs) to assess the crystallinity of individual


points in the cement overgrowths (fig. 10). The pronounced zone axis and strong image
present in figure 10a indicated that the material is strongly crystalline and probably
quartz. Figure 10b showed a moderately strong image near identical to the quartz EBSP.
This was interpreted as microcrystalline quartz. Figure 10c exhibited a very weak EBSP
though a pronounced zone axis was still present. This pattern was interpreted as cryptocrystalline quartz such as opal-CT.
The technique allows the determination of the degree of crystallinity in a mineral,
enabling the possible interpretation of the silica phase present. At present EBSD has only
been applied minimally to observe crystallographic changes during the diagenesis of
micro- and non-crystalline silica. As a result definitive characterisations of silica
polymorphs from EBSD are not currently available.

Zone axis
Diffraction
patterns

Figure 10. Individual electron


backscatter patterns (EBSPs) from
individual grains. Image above
depicts diffraction bands from their
lattice planes and the zone axis at
their intersection. A) strong EBSP for
quartz B) EBSP of microcrystalline
quartz C) EBSP of poorly crystalline
silica (e.g. opal-CT).
(adapted from Haddad et al, 2006)

Conclusions
Investigations have shown that non- and microcrystalline silica minerals have
microstructural characteristics which allow differentiation between the numerous
polymorphs. Mineralogical studies indicate three distinct phases; non-crystalline opal,
microcrystalline opal and microcrystalline quartz.

16

Non-crystalline opal-A is highly disordered with relatively high total water content
(>10 wt.%). Due to the variability in structure of non-crystalline opals, definitive
characterisation has proved difficult. This problem is accentuated by the ambiguous
nature of the transformation between opal-A to opal-CT during diagenesis.

Microcrystalline opals contain opal-CT and opal-C. Opal-CT is characterised by


randomly orientated lepispheric aggregates with increasing abundance of parallel
fibrous (lussatite) crystallites as the phase becomes more structurally ordered prior to
the formation of opal-C. This change is accompanied by a reduction of stacking faults
present in the crystal lattice, allowing differentiation between the two varieties.

Microcrystalline quartz; quartzine, chalcedony (wall-lining and horizontally banded)


and fine-quartz, display more ordered structures with pronounced crystallite textures
than the opal minerals. Although rightfully acknowledged as a distinct phase, it is
evident from investigations that microcrystalline quartz encompasses a wide range
of crystalline silica minerals with diverse microstructural characteristics.

Although some questions still remain regarding the mechanisms involved in the
diagenesis of non- and microcrystalline silica minerals to quartz, it appears that the
transformation occurs predominantly as a dissolution-precipitation process whereby;
opal-CT/-C
microcrystalline quartz. The diagenetic sequence is
opal-A
characterised by increasing crystallinity, crystal size, and structural order. Total water
content decreases as the microstructures becomes more ordered and compacted. This is
reflected in the mineral densities which increase as they become more crystalline
X-ray diffraction patterns show that during diagenesis the reflections at ~4 become
more intense and sharper. Opal-A exhibits broad diffuse peaks which become gradually
narrower and more intense with increasing structural order through the microcrystalline
opal phases (opal-CT/-C). In truly crystalline quartz minerals the relatively high degree
of crystallinity and lack of structural defects is shown as sharp, intense isolated peaks
with XRD. The application of IR spectroscopy has shown that the content and speciation
of water present in silica mineral phases changes with heating, with physically absorbed
molecular water (H2Omol) lost at relatively low temperatures (~150C). Typically silanol
group waters (H2OSiOH) are lost from the crystal lattice as structural defects reduce with
heating and diagenesis.
The two techniques facilitate the determination of independent aspects of silica
minerals, structural order and water speciation respectively, during diagenesis. However,
limited information is available regarding how these material properties change with
respect to each other during diagenesis. At present limited studies on silica diagenesis
using EBSD have been undertaken. In conjunction with previously employed techniques
such as IR spectroscopy and XRD, the technique could be employed to potentially
characterise crystallographic transformations with respect to water content or structural
order changes.

References
ALEXANDRE, A., MEUNIER, J.-D., LLORENS, E., HILL, S. M. & SAVIN, S. M. 2004.
Methodological improvements for investigating silcrete formation: petrography,
FT-IR and oxygen isotope ratio of silcrete quartz cement, Lake Eyre Basin
(Australia). Chemical Geology. 211, 261 - 274.
CADY, S. L., WENK, H.-R. & DOWNING, K. H. 1996. HRTEM of microcrystalline opal in
chert and porcelanite from the Monterey Formation, California. American
Mineralogist. 81, 1380 - 1395.

17

CADY, S. L., WENK, H.-R. & SINTUBIN, M. 1998. Microfibrous quartz varieties:
characterisation by quantitative X-ray texture analysis and transmission electron
microscopy. Contrib. Mineral. Petrol. 130, 320 - 335.
DARRAGH, P. J., GASKIN, J., TERRELL, B. C. & SANDERS, J. V. 1966. Origin of precious
opal. Nature. 209, 13 - 16.
DE JONG, B. H. W. S., VAN HOEK, J., VEENMAN, W. S. & MANSON, D. V. 1987. X-ray
diffraction and 29Si magic-angle-spinning NMR of opals; incoherent long- and
short-range order in opal-CT. American Mineralogist. 72, 1195 - 1203.
ELZEA, J. M., ODOM, I. E. & MILES, W. J. 1994. Distinguishing well ordered opal-CT and
opal-C from high temperature cristobalite by x-ray diffraction. Analytica Chimica
Acta. 286, 107 - 116.
FLRKE, O. W., GRAETSCH, H. & MIEHE, G. 1983. Crystalstructure and Microstructure of
Chalcedony. Fortschritte der Mineralogie. 61, 1, 62 - 63.
FLRKE, O. W., GRAETSCH, H., MARTIN, B., RLLER, K. & WIRTH, R. 1991.
Nomenclature of microcrystalline and non-crystalline silica minerals, based on
structure and microstructure. Neues Jahrbuch F r Mineralogie-Abhandlungen.
163, 1, 19 - 42.
FLRKE, O. W., KHLER-HERBERTZ, B., LANGER, K. & TNGES, I. 1982. Water in
Microcrystalline Quartz of Volcanic Origin: Agates. Contrib. Mineral. Petrol. 80,
324 - 333.
FRONDEL, C. 1982. Structural hydroxyl in chalcedony (Type B quartz). American
Mineralogist. 67, 1248 - 1257.
GRAETSCH, H. 1994. Structural characteristics of opaline and micro-crystalline silica
minerals. In: Heaney, P. J., Prewitt, C. T. & Gibbs, G. V. eds. Silica. Physical
Behaviour, Geochemistry and Materials Applications, pp. 209 - 232. Reviews in
Mineralogy, 29.
GRAETSCH, H., FLRKE, O. W. & MIEHE, G. 1985. The Nature of Water in Chalcedony
and Opal-C from Brazilian Agate Geodes. Phys. Chem. Minerals. 12, 300 - 306.
GRAETSCH, H., FLRKE, O. W. & MIEHE, G. 1987. Structural Defects in Microcrystalline
Silica. Phys. Chem. Minerals. 14, 249 - 257.
GRAETSCH, H., GIES, H. & TOPALOVI , I. 1994. NMR, XRD and IR study on
microcrystalline opals. Phys. Chem. Minerals. 21, 166 - 175.
HADDAD, S. C., WORDEN, R. H., PRIOR, D. J. & SMALLEY, P. C. 2006. Quartz cement in
the Fontainebleau Sandstone, Paris Basin, France: crystallography and the
implications for mechanisms of cement growth. Journal of Sedimentary Research.
76, 244 - 256.
HATTORI, I., UMEDA, M., NAKAGAWA, T. & YAMAMOTO, H. 1996. From chalcedonic
chert to quartz chert: diagenesis of chert hosted in a Miocene volcanicsedimentary succession, central Japan. Journal of Sedimentary Research. 66, 1,
163 - 174.
HEANEY, P. J. 1993. A proposed mechanism for the growth of chalcedony. Contrib.
Mineral. Petrol. 115, 66 - 74.
HEANEY, P. J., VEBLEN, D. R. & POST, J. E. 1994. Structural disparities between
chalcedony and macrocrystalline quartz. American Mineralogist. 79, 452 - 460.

18

HENDRY, J. P. & TREWIN. N. H. 1995. Authigenic quartz microfabrics in Cretaceous


turbidites: evidence for silica transformation processes in sandstones. Journal of
Sedimentary Research. A65, 2, 380 - 392.
HERDIANITA, N. R., BROWNE, P. R. L., RODGERS, K. A. & CAMPBELL, K. A. 2000.
Mineralogical and textural changes accompanying ageing of silica sinter.
Mineralium Deposita. 35, 48 - 62.
HESSE. R. 1989. Silica Diagenesis: Origin of Inorganic and Replacement Cherts. Earth
Science Reviews. 26, 253 - 284.
HURST, A. 1981. A scale of dissolution for quartz and is implications for diagenetic
processes in sandstones. Sedimentology. 28, 451 - 459.
IIJIMA, A. & TADA, R. 1981. Silica diagenesis of Neogene diatomaceous and
volcaniclastic sediments in northern Japan. Sedimentology. 28, 185 - 200.
JONES, J. B. & SEGNIT, E. R. 1971. The nature of opal 1. Nomenclature and constituent
phases. Journal of the Geological Society of Australia. 18, 1, 57 - 68.
KNAUTH, P. 1994. Petrogenesis of chert. In: Heaney, P. J., Prewitt, C. T. & Gibbs, G. V.
eds. Silica. Physical Behaviour, Geochemistry and Materials Applications, pp.
233 - 258. Reviews in Mineralogy, 29.
KRONENBERG, A. K. 1994. Hydrogen speciation and chemical weakening of quartz. In:
Heaney, P. J., Prewitt, C. T. & Gibbs, G. V. eds. Silica. Physical Behaviour,
Geochemistry and Materials Applications, pp. 123 - 176. Reviews in Mineralogy,
29.
LANGER, K. & FLRKE, O. W. 1974. Near infrared absorption spectra (4000-9000cm-1) of
opals and the role of water in these SiO2 nH2O minerals. Fortschritte der
Mineralogie. 52, 1, 17 - 51.
LLOYD, G. E. & FREEMAN, B. 1991. SEM electron channeling analysis of dynamic
recrystallisation in a quartz grain. Journal of Structural Geology. 13, 8, 945 - 953.
LLOYD, G. E. 1987. Atomic number and crystallographic contrast images with the SEM: a
review of backscattered electron techniques. Mineralogical Magazine. 51, 3 - 19.
LYNNE, B. Y. & CAMPBELL, K. A. 2004. Morphologic and mineralogic transitions from
opal-A to opal-CT in low-temperature siliceous sinter diagenesis, Taupo volcanic
zone, New Zealand. Journal of Sedimentary Research. 74, 4, 561 - 579.
MIEHE, G., GRAETSCH, H. & FLRKE, O. W. 1984. Crystal Structure and Growth Fabric
of Length-Fast Chalcedony. Phys. Chem. Minerals. 10, 197 - 199.
MIZUTANI, S. 1977. Progressive Ordering of Cristobalitic Silica in the Early Stage of
Diagenesis. Contrib. Mineral. Petrol. 61, 129 - 140.
MOXON, T., NELSON, D. R. & ZHANG, M. 2006. Agate recrystallation: evidence from
samples found in Archaean and Proterozoic host rocks, Western Australia.
Australian Journal of Earth Sciences. 53, 235 - 248.
MURATA, K. J., FRIEDMAN, I. & GLEASON, J. D. 1977. Oxygen isotope relations between
diagenetic silica minerals in Monterey Shale, Temblor Range, California.
American Journal of Science. 277, 259-272.

19

NEUMANN, B. 2000. Texture development of recrystallised quartz polycrystals unravelled


by orientation and misorientation characteristics. Journal of Structural Geology.
22, 1695 - 1711.
PRIOR, D. J., BOYLE, A. P., BRENKER, F., CHEADLE, M. C., DAY, A., LOPEZ, G.,
PERUZZO, L., POTTS, G. J., REDDY, S., SPIESS, R., TIMMS, N. E., TRIMBY, P.,
WHEELER, J. & ZETTERSTRM, L. 1999. The application of electron backscatter
diffraction and the orientation contrast imaging in the SEM to textural problems in
rocks. American Mineralogist. 84, 1741 - 1759.
RICE, S. B., FREUND, H., HUANG, W.-L., CLOUSE, J. A. & ISAACS, C. M. 1995.
Application of Fourier transform infrared spectroscopy to silica diagenesis: The
opal-A to opal-CT transformation. Journal of Sedimentary Research. A65, 4, 639
- 647.
ROBERTSON, A. H. F. 1977. The origin and diagenesis of cherts from Cyprus.
Sedimentology. 24, 11 - 30.
RODGERS, K. A., BROWNE, P. R. L., BUDDLE, T. F., COOK, K. L., GREATREX, R. A.,
HAMPTON, W. A., HERDIANITA, N. R., HOLLAND, G. R., LYNNE, B. Y., MARTIN,
R., NEWTON, Z., PASTARS, D., SANNAZARRO, K. L. & TEECE, C. I. A. 2004. Silica
phases in sinters and residues from geothermal fields of New Zealand. Earth
Science Reviews. 66, 1 - 61.
SHOVAL, S., CHAMPAGNON, B. & PANCZER, G. 1997. The quartz-cristobalite
transformation in heated chert rock composed of micro and crypto-quartz by
micro-Raman and FT-IR spectroscopy methods. Journal of Thermal Analysis. 50,
203 - 213.
SIEVER, R. 1962. Silica solubility, 0 - 200C., and the diagenesis of siliceous sediments.
The Journal of Geology. 70, 2, 127 - 150.
STAMATAKIS, M. G., KANARIS-SOITIRIOU, R. & SPEARS, A. 1991. Authigenic silica
polymorphs and the geochemistry of Pliocene siliceous swamp sediments of the
Aridea volcanic province, Greece. Canadian Mineralogist. 29, 587 - 598.
STEIN, C. L. & KIRKPATRICK, R. J. 1976. Experimental porcelanite recrystallisation
kinetics; a nucleation and growth model. Journal of Sedimentary Research. 46, 2,
430 - 435.
SWAINSON, I. P., DOVE, M. T. & PALMER, D. C. 2003. Infrared and Raman spectroscopy
studies of the - phase transition in cristobalite. Phys. Chem. Minerals. 30, 353 365.
VENABLES, J. A. & HARLAND, C. J. 1973. Electron back-scattering patterns: A new
technique for obtaining crystallographic information in the scanning electron
microscope. Philosophical Magazine. 27, 1193 - 1200.
WAHL, C., MIEHE, G. & FUESS, H. 2002. TEM characterisation and interpretation of
fabric and structural degree of order in microcrystalline SiO2 phases. Contrib.
Mineral. Petrol. 143, 360 - 365.
WILLEY, J. D. 1980. Effects of ageing on silica solubility: a laboratory study. Geochimica
et Cosmochimica Acta. 44, 573 - 578.
WILLIAMS, L. A. & CRERAR, D. A. 1985. Silica diagenesis, II. General mechanisms.
Journal of Sedimentary Petrology. 55, 3, 312 - 321.

20

WILLIAMS, L. A., PARKS, G. A. & CRERAR, D. A. 1985. Silica diagenesis, I. Solubility


controls. Journal of Sedimentary Petrology. 55, 3, 301 - 311.
WILSON, M. J., RUSSELL, J. D. & TAIT, J. M. 1974. A New Interpretation of the Structure
of Disordered -Cristobalite. Contrib. Mineral. Petrol. 47, 1 - 6.
XU, H., BUSECK, P. R. & LUO, G. 1998. HRTEM investigation of microstructures in
length-slow chalcedony. American Mineralogist. 83, 542 - 545.
YAMAGISHI, H., NAKASHIMA, S. & ITO, Y. 1997. High temperature infrared spectra of
hydrous microcrystalline quartz. Phys. Chem. Minerals. 24, 66 - 74.

You might also like