You are on page 1of 18

1

Characterisation of silica minerals in a banded agate:


implications for agate genesis and growth mechanisms
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: Agates of volcanic origin contain a range of microcrystalline silica
minerals; cryptocrystalline silica, chalcedony and quartz, arranged in concentric
bands. Although abundant worldwide, little is known about the genesis of the
characteristic banding patterns. Researchers disagree whether the bands result by
precipitation from siliceous hydrothermal influxes or by in situ crystallisation of a
silica gel. The study combines the use of an array of techniques, including
electron backscatter diffraction (EBSD), cathodoluminescene (CL) and Fourier
Transform infrared (FT-IR) spectroscopy, to characterise the silica minerals
present and investigate the relationships between them in the banding
arrangement. Microstructural and compositional (trace element, water content)
observations reveal the bands form by discrete siliceous fluid influxes. Within
these individual bands a gradual change in mineralogy analogous to silica
chalcedony
quartz. The
diagenesis is observed; cryptocrystalline silica
change is reflected in the degree of crystallinity, trace element (e.g. iron)
distribution and water content. The proposed band growth model can account for
the observed changes in composition with respect to the silica minerals present.
Crystallographic orientation data reveals a strong a axes lineation which is present
throughout the sample, suggesting a universal control on agate growth.

The term agate relates to silica rock species of volcanic origin that are commonly
believed to form in vugs created by the vesiculation of the volcanic host (Moxon et al.,
2006). Agates contain a number silica minerals; crypto-crystalline silica, chalcedony,
fine-quartz and quartz. The microcrystalline silica minerals in agate have relatively high
silica purity, containing typically less than 1 wt.% non-volatile impurities (Flrke et al.,
1991; Graetsch, 1994). These minerals contain varying amounts of water (H2O and Si-OH
groups) which can be used to determine the mineral species present (Graetsch et al.,
1985). The segregation of the silica polymorphs, with their differing microstructural and
compositional characteristics (Table 1.), creates the distinctive concentric banding
patterns observed in typical banded agates. Chalcedony is the silica mineral characteristic
of agates. It is widely accepted that wall-lining chalcedony, characterised by its parallel
fibrous microstructure, forms by nucleation onto the wall of the host cavity (Graetsch,
1994). The fibres in chalcedony are elongate along the a axes toward the centre of agates,
perpendicular to the banding pattern. As a consequence chalcedony exhibits a lengthfast optical character; resulting in variable refractive indices but lower refractive index in
the direction of the fibre (Flrke et al., 1982). In the oscillatory zoning present in agates,
chalcedony is often succeeded by fine-quartz or quartz. Fine-quartz is a microcrystalline
variety of quartz with a granular texture with individual grains typically <20 m (Flrke
et al., 1982).

Variety

Microstructure

Crystal size

Total water
(H2Omol + Si-OH)

quartz

crystalline

> 20 m

~ 0 wt.%

fine-quartz

granular

5 20 m

<1 wt.%

chalcedony
(wall-lining)

parabolic fibre bundles


(length-fast)

< 100 m length

1 2 wt.%

crypto-crystalline

poorly defined

1 10 wt.%

Table 1. Nomenclature and characteristics of SiO2 minerals typically present in agates (adapted from Flrke
et al., 1991).

Despite the worldwide occurrence of agates and numerous investigations,


understanding the formation of agates has proved problematic, notably in explaining the
occurrence of the oscillatory bands throughout the body of an agate. The inability for
agate to be produced synthetically in laboratory conditions has resulted in the proposal of
numerous models to explain the banded texture; accounting for the rhythmic segregation
of non-crystalline opal and chalcedony. It is generally believed that agates originate either
from the direct precipitation of silica minerals from hydrothermal fluids or from the
deposition of an amorphous silica gel which subsequently matures by diagenetic
processes to form chalcedony (Moxon & Reed, 2006).
Numerous authors have suggested that agate banding formed as a result of
crystallisation from hydrothermal fluids (Flrke et al., 1982; Heaney & Davis, 1995). In
these models influxes of siliceous fluids with differing levels of silica saturation and trace
element concentrations precipitate the varying silica minerals present; e.g. quartz
precipitating from a relatively low silica concentration, resulting in the formation of low
defect crystals. Wang and Merino (1995) and Merino et al. (1995) postulated that agates
crystallise from lumps of hydrous silica containing trace elements. The authors model
predicted that silica lumps could crystallise in an oscillatory manner as witnessed in
agates. The model proposed that chalcedony fibres form due to morphological instability
at the crystallisation front caused by cations which act to accelerate crystal growth.
Oscillations in trace element concentrations, as predicted by the model, can thus
accommodate for the formation of oscillating chalcedony and quartz bands.
As highlighted by the conflicting models outlined above, there are still a number of
questions regarding the formation of agates which remain unresolved. Focusing on
microstructural and crystallographic relationships with compositional variations, this
investigation aims to provide a greater insight into understanding the controls on the
banding patterns within agate; with possible implications on the genesis and growth
mechanisms of the individual bands. Specific questions relating to the crystallography
and composition of the silica phases within the concentric bands are addressed in this
paper:
1.

How does mineralogy (microstructures, crystal morphology) change throughout


the individual bands? Do the SiO2 minerals form discrete bands or have they
formed in the same event?

2.

Are the changes in mineralogy controlled by compositional variations (trace


element, water content)? How do the two factors relate to each other?

3.

How do the crystallographic orientations of the different silica minerals in the


bands relate to each other? Are they independent or do they share the same
orientation? What implications could this have on band growth mechanisms?

In order to address the issues regarding the crystallographic texture of the silica
polymorphs in agate banding and how they relate to each other, automated electron
backscatter diffraction (EBSD) was employed for the first time on agate. 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
Lloyd, 1987; Lloyd & Freeman, 1991; Prior et al. 1999. By simultaneously employing
techniques investigating compositional variations; backscatter electron (BSE) imaging,
cathodoluminescence (CL) and infrared (IR) spectroscopy, the acquired EBSD data is
used to relate with observations made with these other techniques.

Methodology
Sample
The sample used was taken from a Lake Superior agate formed within a basaltic host
rock dated at 1.1 Ga (Moxon, 2002). From previous investigations (Moxon et al., 2006)
the sample was known to contain the banded components required for this study. The
sample was made into a 100 m doubly polished fluid inclusion wafer to allow FT-IR and
EBSD analysis on the same sample region. To enable electron backscatter diffraction and
cathodoluminescence analysis the sample was SYTON polished using a colloidal silica
solution (Fynn & Powell, 1979; Lloyd, 1987) to remove any mechanical damage to the
surface. The sample was carbon coated using the EMITECH K950X (<10 nm thickness).
Analytical equipment and techniques
Transmitted-light optical analysis was initially performed using a Nikon 79202
binocular microscope to characterise the sample as a whole and identify the silica
polymorphs and their distribution throughout the sample.
Backscatter electron (BSE) imaging and cathodoluminescence (CL)
Backscatter electron (BSE) and SEM-CL was carried out using a Philips XL30 SEM
fitted with a K.E. Developments Ltd cathodoluminescence detector (D308122). Working
conditions were 10 kV (20 kV for BSE) acceleration voltage and ~1 nA beam current and
16 mm working distance. CL images were obtained by accumulating the signal of 32
frames using a slow scanning beam raster. Qualitative spot element analysis was
undertaken using EDAX (Oxford Instruments).
Electron backscatter diffraction (EBSD)
Automated SEM-EBSD was carried out using a CamScan X500 crystal probe field
emission gun SEM. Working conditions were 20 kV accelerating voltage, 30 nA beam
current and ~25 mm working distance. EBSD patterns were auto-indexed using the
CHANNEL 5 software from hkl Technology (Denmark). The software was used to
display maps and pole figure data. Some data processing was used was used to remove
erroneous data from the maps, though care was taken given the poorly-crystalline nature
of areas of the sample. EBSD was employed as it has the capability to resolve
crystallographic orientations of points at a resolution of better than 0.5 m. Using EBSD

to acquire microstructural information about the crystal structure of the bands can, in turn,
lead to interpretations regarding crystallisation mechanisms in the formation of agates.
Fourier Transform Infrared (FT-IR) spectroscopy
The utilisation of infrared spectra to determine silica mineral phases by water content
and speciation has been outlined previously by Langer & Flrke, 1974; Graetsch et al.,
1985; Kronenberg, 1994. Infrared spectra of the sample were obtained using a liquidnitrogen-cooled Nicolet Centaurus FT-IR microscope (Thermo Electron Corporation).
Binocular lens (x 10) gave an optical image of the sample and rectangular apertures 300 x
300 m were used for taking the measurement. Spectra were collected at 4 cm-1 resolution
with 100 scans collected and averaged. Reflectance spectra were generated and converted
to absorption spectra using OMNIC software (Thermo Electron Corporation). Peak
positions were determined by taking positions of local maxima following linear baseline
correction. The position of these peaks was within 3 cm-1 since a wavenumber
resolution of 4 cm-1 was applied for the IR measurement.

Results
Transmitted-light optics
Through non polarised light the sample (Fig. 1A) appears as a colourless medium with
distinct parallel bands with brown colouration. Under cross polarised light the brown
bands comprise of parallel fibres perpendicular to the banding orientation. The fibres
show a length-fast optical character indicative of chalcedony; resulting in variable
refractive indices but lower refractive index in the direction of the fibre (e.g. Flrke et al.,
1991). The chalcedonic parallel fibrous aggregates exhibit rhythmic extinction, resulting
in a distinctive wrinkle-band texture diagnostic of wall-lining chalcedony (Fig. 1B).

Figure 1. Optical images of typical banding patterns. A) Plain-polarised light photograph showing
brown bands. B) Photograph of area in A through crossed polars showing rhythmic banded fibres.
Arrow indicates growth direction of the agate, with the bands gradually widening. Scale bar 500 m.

The fibres display a gradual increase in width and length over the distance of the
individual bands, with coarser fibres distinguishable from finer fibres due to higher
birefringence and longer twist periodicity. The length of the fibres range from sizes
indeterminable using optical methods to ~100 m where the fibres terminate. Coarser
fibres are distinguishable from finer fibres due to longer twist periodicity and higher

birefringence. The fibres are succeeded by small euhedral quartz crystals. Colourless in
transmitted light, the crystals are approximately 25 m in diameter. The thickness of the
bands shows a gradual width increase trend along the sample (Fig. 2). Bands on the outer
edge of the sample relative to the growth inwards are approximately 50 m wide,
increasing to over 400 m at the inner edge of the sample. As shown in figure 2, the
increase in width is not at a constant rate; with fluctuations evident (e.g. bands 16 17).
500

band width (m)

400
300
200
100
0
1

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25

band #

Figure 2. Graph showing general increase in band width toward the centre of the sample. Widths
determined from optical microscopy and BSE imaging. Arrow denotes growth direction.

Electron backscatter diffraction (EBSD)


Microstructure and orientation of bands
The band-contrast (BC) image (Fig. 3A) displays a marked grayscale change in the
individual bands. The BC image exhibits a gradual increase in signal brightness across the
width of the bands in the growth direction, indicating a variation in the strength of the
EBSD patterns across the width of the individual bands. This gradual increase in bandcontrast response is coincident with an increase in grain size across the bands (Fig. 3B).
Although the BC image and grain size distribution show a graded variation in signal
response across the individual bands, the three distinct silica minerals can be determined
due to crystal morphologies present:
1) Isolated equiaxed crystals ~2 m in diameter in zero-solution matrix (see determining
the zero-solution zone at end of section). This component exhibits the darkest bandcontrast response. This component is representative of cryptocrystalline silica.
2) Parallel aligned fibrous elongate crystals grading from 25 100 m length with axial
ratios approximately 1:5 though showing a gradual increase in width toward to tip of
the individual crystallites in the direction of growth. The fibres are elongated parallel
to the growth direction of the bands; indicative of chalcedony.
3) Larger polygonal euhedral quartz crystals up to 50 m diameter. Crystals which
do show a degree of elongation on a particular axis do not exhibit the strong parallel
alignments evident in the chalcedony, though a number of crystals do share
elongate axes sub-perpendicular to the band orientation. A number of crystals show
internal changes in crystal orientation of 60 sub parallel to the banding direction;
these are interpreted as dauphin twins. The crystals exhibit curved crystal edges;

particularly on grain boundaries adjacent to similar larger crystals. The quartz exhibits
the highest band-contrast response in the bands.
This distribution pattern of defined silica minerals is observed in most of the bands
present in the sample, though the relative thickness of these components varies
throughout the sample (e.g. Fig. 5). The corresponding crystal-orientation map (Fig. 3C)
uses colour to qualitatively represent crystallographic orientations. The image reveals that
the chalcedony exhibits the same orientation along the length of the individual crystals. In
a number of locations the orientations appear to be continuous into the quartz crystals
present at the edge of the individual bands. The change in colouration between
neighbouring crystals in this component indicates that the crystals have different
crystallographic orientations relative to each other; resulting in the pseudo-parallel
banded pattern present in figure 3C. However, the colour variation is subtle; implying the
orientation contrast between the crystals is minor. Although the quantity and
concentration of data points in the cryptocrystalline SiO2 is not as high as in the
chalcedony and quartz, the data reveals the crystallographic orientations in this
component are similar to those in displayed by the chalcedony.

10
15
20
Grain size ( m)

25

30

Figure 3. Collection of images generated from


EBSD data for one band in the sample. A) bandcontrast map showing increase in brightness with
increasing crystallinity (numbers denote silica
mineral defined in text; 1) cryptocrystalline silica 2)
chalcedony 3) quartz). B) grain size (circular
diameter equivalent) map showing increase across
band. C) crystallographic orientation map showing
similar crystal orientation along each fibre. Scale
bar of 50 m in C relates to A and B also.

Pole figures of EBSD data representing crystallographic orientations projected in


stereonets are shown in figure 4. The EBSD data has been separated to display the
orientations within the individual band components (Figs. 4A C); regions highlighted in
accompanying BC maps. All figures (4A C) show there is a distinct clustering of
orientation data in the a axes parallel to band growth direction with dispersal of data in
the c axis plot. This suggests that all the band components share a similar crystallographic
orientation with an intense the a axes lineation; indicating a strong crystallographic
orientation throughout the entire band.
a axes

11424 data points

B)

chalcedony

A)

cryptocrystalline silica

c axis

C)

quartz

51740 data points

39216 data points

Figure 4. Pole figure representation of EBSD data. Accompanying selectively highlighted


band-contrast map represents area of data selection for each silica mineral phase. Data
from each component represented on contoured stereographic projections showing
crystallography; plotted for <0001> (c axis) and <11-20> (a axes). a axes data is plotted
upper hemisphere (left) and lower hemisphere (right). A) Orientation data from band
cryptocrystalline silica showing a degree of data dispersal; though a axes lineation still
evident. B - C) Orientation data from chalcedony and quartz; showing tight data cluster in
a axes and dispersed c axis girdle. Scale bar 50 m in A applies to all BC images.

In the cryptocrystalline silica (Fig. 4A) there is a broader dispersal of orientation data in
the a axes. However, clustering is still evident in the band growth direction. The
chalcedony and quartz (Figs. 4B & C) exhibit tighter data clustering in the a axes. This
creates intense data spots in the orientation of the band growth; parallel to the elongate
crystal axes of the chalcedony. This a axes lineation is evident in bands throughout the
sample. The pole figure for the c axis in the cryptocrystalline silica and chalcedony band
components (Figs. 4A & B) shows that the data falls on a girdle perpendicular to the
pronounced a axes orientation. This suggests that the rotation axis creating the dispersal
of c axis orientations lies in the a axes direction. In the quartz (Fig. 4C) the c axis data is
not as dispersed, indicating a difference in crystallographic orientation with respect to the
c axis between the chalcedony and quartz.
Crystallographic Misorientations
Misorientation profiles were used to investigate the apparent rotation of the c axis and
differences in crystallographic orientation across individual crystals. The pole figures
show the c axis perpendicular to the crystal growth direction; misorientation profiles were
taken in this orientation. Profiles perpendicular to the elongate axis of the chalcedony
reveal a gradual change in crystallographic orientation across the individual fibres (Fig.
5A). Similar profiles across the equant quartz crystals (Fig. 5B) do not display these
gradual changes in orientation across the individual crystals. The fluctuations in the
profile result from orientation changes between neighbouring crystals, as represented in
the crystal-orientation map (Fig. 3B), and not changes within the individual crystals.

Figure 5. Misorientation profiles displaying


changing crystallographic orientations along
transects. Profiles relative to first point.
(position shown in band-contrast map
above). A) profile perpendicular to elongate
axis of chalcedony fibres showing gradual
curved changes B) contrasting profile across
equant quartz crystals. Scale bar on map 100
m.

Determining the zero-solution zone


The zero-solution zones present in the sample, denoted as cryptocrystalline silica at the
start of each individual band, result from the initial EBSD maps being unable to yield

crystallographic orientation data from these areas. The software was unable to index these
regions because the crystallographic data recorded were so weak. Qualitative secondary
X-ray analysis (EDAX) identified these banded components as silica (SiO2). There are a
number of possible explanations for this pattern: 1) the silica is quartz with abnormally
high dislocation density, possibly due to deformation. As undeformed crystals are present
in adjacent bands, this possibility is unlikely. 2) The diffraction patterns suffer decay over
time due to crystal lattice damage caused by the electron beam. Although there was
appreciable damage to the sample when the beam was focused on a region for longer time
periods, such depreciation in diffraction pattern quality was also evident in quartz. This
suggests that the time dependent damage is similar to quartz and only shows appreciable
damage after time periods significantly longer than the data acquisition time. 3) The final
possibility is that the SiO2 present is a cryptocrystalline phase and does not have a longrange crystal structure capable of diffracting electrons.
Examining diffraction pattern quality across the bands revealed the diffraction patterns
becoming more pronounced in areas of no zero-solution. This, and the band contrast
results, suggests that the degree of crystallinity within the individual bands is gradually
increasing, from cryptocrystalline SiO2 into fully crystalline quartz.
Backscatter electron (BSE) and cathodoluminescence (CL)
The BSE and CL images (Fig. 6) show there is compositional variability within the
individual bands. The BSE image (Fig. 6A) shows non-uniform signal response in each
band. The image shows an initial lighter zone which grades to darker shades though the
distance of the individual band. This suggests variability in each band. Chemical element
analysis shows the bands to comprise of SiO2 to the limit of the detector. The extents of
the individual bands are defined by sharp contrast changes between the crystalline quartz
(darker) of one band and the cryptocrystalline SiO2 (lighter) of the subsequent band
defined using optical microscopy. A number of these sharp contrast boundaries are coated
by bright 2 4 m thick Fe-rich films which show sharp grading over ~10 m into the
lighter zone. The variability of the bands is also highlighted using CL (Fig. 6B).

Fe

Figure 6. A) Backscattered electron (BSE) image which highlights the bright Fe-rich films which coat the
outer surface of the crystalline quartz bands. B) Cathodoluminescence (CL) image, displaying the
luminescence contrast between quartz bands and non-/cryptocrystalline regions. Scale bar 200 m.

Although CL responses for silica polymorphs are generally weak compared to other
minerals such as feldspars and carbonate (Marshall, 1988), there is a marked contrast in
the degree of luminescence throughout the individual bands. The poorly crystalline band

10

component shows a dull CL response which shows a gradual increase in luminescence


into the chalcedony band component. The crystalline quartz component adjacent to the
chalcedonic component exhibits relatively bright luminescence, appearing noncontinuous with the graded luminescence profile witnessed in the rest of the band. In
summary, there are CL and BSE responses that correspond to band boundaries and
variations within the bands.
Infrared spectroscopy
Infrared spectroscopy is dependent on the response of short range molecular scale
energetic vibrations such as O H stretching and bending. This generates characteristic
spectra for particular mineral phases depending on the molecules present, for example
different water species (Langer & Flrke, 1974). Infrared spectra with broad bands
centred around 3430 cm-1 are generally considered to be related to hydrogen molecular
water (H2Omol) such as liquid water. A sharp peak present at ~3585 cm-1 has been
attributed to Si-OH species (cf. Langer & Flrke, 1974; Kronenberg, 1994; Yamagishi et
al., 1997).
Infrared absorption spectra (3800 2800 cm-1) for the three band components (defined
using EBSD) in the individual bands are presented in figure 7. Cryptocrystalline silica
displays a broad peak at ~3435cm-1 with a prominent peak at 3585 cm-1; indicating the
presence of both H2Omol and Si-OH group water. Chalcedony exhibits the same peak
locations as present in cryptocrystalline silica but the Si-OH peak at 3585 cm-1 is not as
defined. The spectra also display relatively less absorbance; indicating less total water
content than the cryptocrystalline phase. The absorbance spectrum for quartz shows no
signal trace in the 3800 2800 cm-1 range. This indicates no molecular water or silanol
group water is present in this region.

cryptocrystalline silica

0.70

chalcedony
0.60

quartz
0.50

0.40

0.30

0.20

0.10
0.00
3600

3400

3200

Wavenumbers (cm-1)

3000

Figure 7. Absorption spectra (2800 3800 cm-1) showing variation between the three silica minerals
present within each individual band. Peaks at ~3430 and 3585 cm-1 relate to H2Omol and Si-OH groups
respectively.

11

Absorbance spectra (3800 2800 cm-1) taken at regular intervals across the sample
show a gradual decrease in absorbance (Fig. 8A & B). The band with a peak at ~3430 cm1
attributed to H2Omol shows a gradual decrease into the centre of the agate. The 3585 cm-1
silanol peak is also present in all the spectra, though it is not as high as that of the
molecular water (3430 cm-1). The silanol peak becomes more defined toward the centre of
the agate; coincident with H2Omol peak becoming gradually broader. Figure 8C shows that
the decrease in both peak heights across the sample is broadly linear; indicating the loss
of total water (H2Omol and Si-OH) is roughly constant into the centre of the agate.

Figure 8. Diagram showing change in infrared


response across sample. A) backscatter electron
image showing relative position of spectra
acquisition areas (1 4) B) absorbance spectra
(2800 3800 cm-1) for regions 1 4 C) graph
showing linear decrease in absorbance value of
peaks at 3430 and 3585 cm-1; denoting H2Omol
and Si-OH groups respectively.

Discussion
Banding patterns; formation from layers or lump?
Electron backscatter diffraction images show the individual bands displaying sharp
undulating contacts between the quartz component and cryptocrystalline silica component
in the adjacent band. This suggests that the banding pattern in agate result from numerous
events and not from the crystallisation of a single hydrous silica lump (e.g. Wang and
Merino, 1995). There are two arguments against the latter hypothesis which arise from
the results: 1) Assuming oscillating crystallisation from a single lump, the transition from
one band to the next would be gradual, exhibiting symmetrical patterns

12

(crystallographically, chemically, etc). The results do not support this hypothesis. 2)


While the bands do show a gradual increase in width into the sample, in accordance with
oscillatory zoning method purported, the width increases show variations not consistent
with mathematical simulations (Bryxina & Sheplev, 1999). This suggests formation of the
bands by separate precipitation events. Flrke et al. (1982) suggested that the rhythmic
banding resulted from magmatic pulsing of hydrothermal fluids in the cooling volcanic
host. Assuming this mechanism of precipitation, the variations in width appear more
plausible. However, the volume of fluid present would have to be reasonably consistent to
accommodate the decreasing surface area effect which generates the overall band width
increase observed. Although it cannot be proved indefinitely, analogous hydrothermal
systems (e.g. geysers) do display regular rhythmic pulsing of fluids, suggesting such a
process is conceivable.
Trace element variation within individual bands
The sample shows marked compositional variations within individual bands, as
observed using CL, BSE and IR. These compositional variations are seemingly coincident
with variations observed in the crystallographic results from EBSD. Comparing CL and
EBSD results, a number of observations are made. Firstly, the poorly crystalline silica
band components with low band contrast have the dullest luminescence. Secondly, the
gradational increase in band contrast observed in the transition from poorly crystalline to
chalcedony is reflected in the CL response, with luminescence also increasing gradually.
Third, the band components comprised of quartz crystals (those with high-band-contrast
response) have the brightest luminescence. If it is assumed that CL signals in quartz are
largely due to trace element distribution (Marshall, 1988), then there is a discernable
compositional difference between the crystalline bands and the non-crystalline and
chalcedonic regions. The concordance between darker signal response in CL and areas of
low band contrast suggest that a trace element capable of quenching cathodoluminescence
response (e.g. iron) is relatively enriched in the poorly crystalline silica mineral bands
compared to the crystalline components. Based on this premise it is possible that the band
components with uniform bright luminescence, i.e., the quartz, contain the least iron.
Conversely, the poorly crystalline silica and chalcedony regions showing gradation of
luminescence and band-contrast (brighter through into chalcedonic banding) are likely to
correspond to a gradual decrease in iron content through the phase change. Backscatter
electron imaging reveals that the quartz shows the darkest response, suggesting it is
relatively less dense than the chalcedony and cryptocrystalline silica. This is surprising as
quartz is typically denser than chalcedony, which, in turn, is denser than cryptocrystalline
silica (Graetsch, 1994). This also suggests the incorporation of denser trace elements into
the chalcedony and cryptocrystalline silica. These interpretations are bolstered by the
presence of the detectable levels of iron in the poorly crystalline regions using qualitative
secondary X-ray analysis (EDAX).
Water content and speciation
Variations in the infrared spectra indicating water content (H2Omol and Si-OH groups)
of the agate band components show concordance with the band-contrast response. The
low-band-contrast poorly crystalline silica displays the highest water content; the
intermediate band-contrast signal indicative of chalcedony shows a relative decrease in
water content; and quartz (high-band-contrast) is coincident with no water present. This
implies that the amount of water (H2Omol present as fluid inclusions and Si-OH groups in
structural disparities) present in SiO2 minerals is dependent on the degree of crystallinity
of the polymorph. It is possible that the decrease in water may simply be due to reducing

13

surface area as consequence of increasing crystal size. Previous investigations focusing


on water content and speciation in silica minerals (Florke et al., 1982; Graetsch et al.,
1985) have suggested that cryptocrystalline silica minerals easily incorporate water due to
the highly disordered crystal structure present in such minerals. The authors noted that
with increasing structural order and crystal size, water content decreased as it could no
longer be accommodated in the evolving crystal structures. The IR responses from the
sample in this study are correlative with these previous observations on water content in
silica polymorphs, indicating a gradual decrease in water content with increasing
crystallinity in the individual bands. The results also show that the ratio between the
H2Omol and Si-OH peaks is constant for the crypto-crystalline silica and chalcedony (Fig.
7); suggesting increasing crystallinity is not preferential toward fluid inclusion sites
(H2Omol) or structural disparities (Si-OH).
The results show an overall decrease in bulk water content into the centre of the agate
sample. This possibly indicates a progressive decrease in the water content of the
hydrothermal fluids which formed the bands; potentially related to fluid flow processes
dictated by hydrothermal evolution of the volcanic host.
Diagenetic analogues and silica solution implications
Previous authors noted that a problem in validating agate banding as forming from a
viscous silica fluid or gel was the lack of a cryptocrystalline/opaline phase concordant
with observations of siliceous sinters (Heaney, 1993; Inagaki et al., 2003). However, the
EBSD data indicates that a poorly crystalline phase in the bands is present. The
progressive pattern observed in the bands; cryptocrystalline silica
chalcedony
quartz, shares characteristics of silica diagenetic processes related to degrees of
supersaturation with respect to silica (cf. Williams et al., 1985; Williams & Crerar, 1985).
A number of studies have investigated the relative rates of nucleation and growth of silica
polymorphs at different silica concentrations in sedimentary realms (Williams & Crerar,
1985). At higher supersaturation with respect to silica, amorphous and cryptocrystalline
silica grow at orders of magnitude more quickly than quartz. Based on this observation, a
simple way to interpret the varying degrees of crystallinity in the band components would
be due to precipitation from solutions with varying degrees of saturation with respect to
silica. It is conceivable that solutions with initial high levels of silica supersaturation
capable of precipitating cryptocrystalline silica could be derived from hydrothermal
influxes through a basaltic host medium (e.g. Gtze et al., 2001)
A number a previous studies (Flrke et al., 1982; Heaney & Davis, 1995) adopted the
notion whereby varying degrees of silica saturation in the solution dictate the type of
silica polymorph precipitated. Flrke et al. (1982) concluded that the different crystallite
sizes and water content in the individual band components represented precipitation of the
microcrystalline silica species from discrete fluid influxes with varying levels of silica
supersaturation. However, the CL and band-contrast results from this study indicate that
although the SiO2 minerals observed are microstructurally and compositionally distinct,
there is an apparent gradation between the polymorphs. This does not correlate with
Flrke et als (1982) conclusion in which the authors stated that the banding pattern
(cryptocrystalline silica
chalcedony
quartz) forms due to discrete compositionally
distinct hydrothermal fluid influxes which serendipitously exhibit progressively less silica
saturation. From the results showing grading of CL, band-contrast and water content it is
possible that the differing band components form by precipitation from a single fluid
influx in which the solution gradually evolves with respect to silica supersaturation.
Heaney and Davis (1995) speculated that a solution initially supersaturated would be
partly polymerised and rapidly precipitate cryptocrystalline silica and chalcedony. The

14

authors remarked that rapid growth of chalcedony fibres at the crystal front would result
in a gradual depletion of silica in the solution. Assuming this hypothesis the decreasing
silica concentration solution would facilitate the progressively slower growth of
increasingly larger crystals with fewer crystal defects. The results, showing increasing
crystal size and decreasing water content through the individual bands appear to support
this crystal growth process. On this basis, eventually the solution would have a
sufficiently low silica concentration that would allow the slow growth of defect-free
quartz crystals. The lack of IR response for water (H2Omol and Si-OH groups) indicates no
presence of fluid inclusions or structurally held water in the quartz. This possibly suggests
defect-free crystal growth from a monomeric solution. This is supported by the highband-contrast signal indicating high crystallinity.
Effect of trace element concentration on crystal morphology
The EBSD results show three distinct crystal morphologies present in the individual
bands; cryptocrystalline silica, chalcedony and equant quartz crystals. The components
also exhibit differing responses in FT-IR, BSE and CL; suggesting a relationship between
the microtexture and composition of the band components. The misorientation profiles
taken parallel to the c axis of the individual chalcedony fibres (perpendicular to the
growth direction) reveal a graded orientation profile in the individual crystallites (Fig.
5A). This gradual change in crystal orientation is not evident in misorientation profiles of
quartz crystals (Fig. 5B). This suggests that the observed misorientations in chalcedony
are not deformation related, as a deformation overprint in the adjacent quartz crystal band
component would be expected also. It is thus plausible that the observed rotated
crystallographic orientations of the c axis (Fig. 4B) within the individual crystallites are
related to the growth mechanism of chalcedony. There have been numerous investigations
relating to the microstructural characteristics of microcrystalline silica minerals, notably
chalcedony (e.g. Flrke et al., 1983; Graetsch et al., 1987). Miehe et al. (1984) stated that
the banded rhythmic extinction pattern (referred to as Runzelbnderung) and changing
birefringence along fibres observed in chalcedony under crossed-polarised light (Fig. 1A)
resulted from the periodic twisting of the c axes perpendicular the fibre elongation during
growth. Prior to this investigation, EBSD on chalcedony has not been conducted;
therefore it is possible that the graded misorientation profile witnessed using EBSD is a
crystal orientation manifestation of this twisting mechanism.
Some authors have proposed that the twisting characteristic in chalcedony arises due to
the incorporation of trace elements into the crystal lattice. Merino et al. (1995) purported
that Si could be readily substituted with trace ions such as Al and Fe in the growing
fibres. As a result of the relatively larger atomic size of the trace ions, the authors
concluded that the fibres have to grow twisted to accommodate these elements, with
higher trace element concentrations leading to the formation of more tightly twisted
chalcedonic fibres. CL analysis of the sample has shown that there is a trace iron gradient
present in the bands. Interestingly the CL gradient is coincident with optical and EBSD
observations which show a gradual increase in chalcedony fibre length, width and twist
periodicity. It is thus possible that the increasing chalcedony fibre dimensions are a
consequence of a lesser degree of twisting of the fibres due to decreasing iron
concentration throughout the individual bands. By contrast, other authors have proposed
that the twisted morphology of chalcedony fibres has a structural foundation. Heaney
(1993) proposed that the twisting is a result of screw dislocations in the quartz lattice in a
polymerised silica solution. EBSD reveals that within the crystal structure of the
chalcedony there are local misorientations which can be interpreted as dislocation defects
within the crystallites. Prior to this investigation it was not clear whether the source of the

15

twisting in chalcedony fibres is due to trace element incorporation, dislocations or both of


these factors. From the results it is feasible to hypothesise that the incorporation of larger
trace elements could, in turn, disrupt the lattice sufficiently to cause the dislocations
advocated by Heaney (1993). The transition from chalcedony to quartz; reflected in sharp
changes in CL signal and crystallographic misorientation, indicates that trace element and
dislocation densities are inherently linked in determining crystal morphology.
Possible band formation mechanisms
As referred to previously, much attention in the study of agates had focused on the
formation of the individual bands. From the above interpretations, a simple
supersaturation-depletion model (Fig. 9) can account for the changes in composition and
crystal morphology within a band. The model is founded on the premise that the bands
form by precipitation from a siliceous fluid; an assumption justified here and in previous
studies (Flrke et al., 1982; Heaney & Davis, 1995).
The influx of a supersaturated silica fluid into the cavity results in precipitation onto the
wall or earlier influx deposit (Fig. 9A). Large trace elements (i.e. iron) rapidly come out
of solution and form the concentrated horizon observed on earlier bands. The high degree
of super saturation with respect to silica leads to rapid formation of cryptocrystalline
silica, trapping water as fluid inclusions and in structural defects (Williams & Crerar,
1985). Silica saturation decreases at the crystal growth front as rapid deposition outpaces
diffusion to the crystallisation front (Heaney & Davis, 1995). Polymerisation in the
solution decreases, allowing the formation of crystals. The presence of trace elements in
the solution creates elongate chalcedony fibres (Fig. 9B). As trace element concentration
decreases twisting of the chalcedony fibres decreases, facilitating the growth of larger
chalcedony fibres. This is coincident with decreasing silica concentration in the fluid.
Gradually levels of silica saturation and iron in solution become sufficiently low to allow
growth of defect free quartz crystals (Fig. 9C). Variation in relative proportions of
cryptocrystalline silica, chalcedony and quartz in bands throughout the sample suggest
influxes of compositionally diverse fluids with respect to silica saturation and trace
element concentrations.

Figure 9. Schematic diagram showing proposed growth mechanism of individual bands in agate: A) rapid
precipitation of cryptocrystalline silica B) growth of chalcedony C) slow growth of defect free quartz
crystals

16

Controls on crystallographic orientation


The EBSD data reveals that all the components of the individual bands share a similar
crystallographic orientation. The pole figures corresponding to the EBSD data show a
strong lineation in the a axis orientation which is apparent in all the band components
regardless of the crystal morphology and degree of crystallinity present. This is
surprising, especially in the cryptocrystalline silica because although the data reveals the
crystallites formed isolated from each other in the amorphous matrix, they share a
common crystallographic orientation. The orientation data becomes more clustered in the
chalcedony and quartz bands, possibly reflecting the relative increase in crystallinity.
Relating the pole figures to the banding arrangement reveals that the general band
growth direction is parallel to the a axes orientation. It is therefore possible that the
lineation dictates the growth direction of the bands. The cryptocrystalline silica and
chalcedony exhibit elongate crystallites parallel to the intense a axes lineation, suggesting
growth in this orientation. Such growth mechanisms for chalcedony have been
hypothesised in previous investigations (e.g. Heaney, 1993). The textures determined in
the results of this study appear to validate these existing growth mechanism proposals for
chalcedony. In addition, the similar crystallographic orientations in the cryptocrystalline
silica suggest a pre-cursor mineral phase which evolves to form chalcedony. Interestingly,
although the prismatic quartz crystals displaying dauphin twin arrangements appear to
grow parallel to the c axis, they too show an intense a axes orientation. This was observed
in numerous bands across the sample; suggesting a crystallographic orientation control
throughout the entire agate. Although it is conceivable a crystallographic control could be
applied by chalcedony onto quartz within individual bands, to suggest that such a process
could be maintained over a number of bands is unrealistic given the structural variability
of the silica polymorphs. Therefore, the crystallographic orientation observed could
possibly relate to a growth mechanism universal to agates. The existence of similar
textures in agates worldwide suggests this may the case. Therefore further EBSD
investigations on agate samples may give a greater insight into this interesting
phenomenon.

Conclusions
The crystallographic, textural and compositional relationships of silica minerals present
within a banded agate have been examined using EBSD, BSE, CL and FT-IR to reveal
the following:
1.

The boundaries between individual bands within the agate are sharply contrasting,
reflected in crystallography (degree of crystallinity, morphology) and composition
(trace elements, water); indicating formation from discrete siliceous fluid influxes.

2.

Crystallographic information reveals that the degree of crystallinity and


morphology of the silica minerals within the individual bands changes in a
systematic fashion analogous to a diagenetic cycle; initial cryptocrystalline silica
progressively grades to fibrous chalcedonic crystals which, in turn, evolve to form
larger equiaxial quartz crystals.

3.

CL and BSE revealed gradual changes in luminescence and signal response across
the individual bands; these probably reflect decreases in trace element (i.e. iron)
concentration during band growth. These changes are coincident with changing
silica mineral polymorphs present; indicating a trace element control on crystal

17

formation and morphology within the host solution. This is significant in the
creation and evolution of chalcedony fibres.
4.

Pole figures from EBSD revealed there is a strong crystallographic control


throughout all the bands in the agate irrespective of crystallinity or microtexture.
This suggests crystal orientation in agate is dictated by a process independent of
the silica mineral polymorph present.

Acknowledgements
Thanks to Richard Worden, Dave Prior and Angela Halfpenny for helpful scientific
discussions and suggestions throughout the project. Additional thanks to: Terry Moxon
for sample donation; Nick Seaton for EBSD expertise; and Carmel Pinnington for SEM
assistance.

References
BRYXINA, N. A. & SHEPLEV, V. S. 1999. Auto-oscillation in agate crystallization. Math.
Geol. 31, 3, 297 - 309.
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
Microcrystalline Quartz of Volcanic Origin: Agates. Contrib. Mineral. Petrol. 80, 324
- 333.
FYNN, G. W. & POWELL, W. J. A. 1979. The cutting and polishing of electro-optic materials.
London. Adams Hilger, pp. 216.
GTZE, J., TICHOMIROWA, M., FUCHS, H., PILOT, J. & SHARP, Z. D. 2001. Geochemistry of
Agates: A Trace Element and Stable Isotope Study. Chemical Geology. 175, 523 542.
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.
HEANEY, P. J. & DAVIS, A. M. 1995. Observation and Origin of Self-Organised Textures
In Agates. Science. 269, 1562 - 1565.
HEANEY, P. J. 1993. A proposed mechanism for the growth of chalcedony. Contrib. Mineral.
Petrol. 115, 66 - 74.
INAGAKI, F., MOTOMURA, Y. & OGATA, S. 2003. Microbial silica deposition in geothermal hot
waters. Appl. Microbial. Biotechnol. 60, 605 - 611.

18

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.
MARSHALL, J. D. 1988. Cathodoluminescence of geological materials. Unwin Hyman. pp.
146.
MERINO, E., WANG, Y. & DELOULE, E. 1995. Genesis of Agates in Flood Basalts: Twisting of
Chalcedony Fibers and Trace Element Geochemistry. American Journal of Science.
295, 1156 - 1176.
MIEHE, G., GRAETSCH, H. & FLRKE, O. W. 1984. Crystal Structure and Growth Fabric of
Length-Fast Chalcedony. Phys. Chem. Minerals. 10, 197 - 199.
MOXON, T. & REED, S. J. B. 2006. Agate and chalcedony from igneous and sedimentary hosts
aged from 13 to 3480 Ma: a cathodoluminescence study. Mineralogical Magazine.
70, 5, 485 - 498.
MOXON, T. 2002. Agate: a study of ageing. European Journal of Mineralogy. 14, 1109 1118.
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.
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.
WANG, Y. & MERINO, E. 1995. Origin of Fibrosity and Banding in Agates from Flood
Basalts. American Journal of Science. 295, 49 - 77.
WILLIAMS, L. A. & CRERAR, D. A. 1985. Silica diagenesis, II. General mechanisms. Journal
of Sedimentary Petrology. 55, 3, 312 - 321.
WILLIAMS, L. A., PARKS, G. A. & CRERAR, D. A. 1985. Silica diagenesis, I. Solubility
controls. Journal of Sedimentary Petrology. 55, 3, 301 - 311.
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