You are on page 1of 10

Textures

Textures offer valuable information about natural and industrial processes in which the
material under investigation has formed. Although this book focuses on naturally occurring
minerals, it also documents a small number of artificial mineral phases which may exist in
nature, but are quite rare and were not accessible for investigation from natural sources.

The word “texture” derives from Latin and relates to a weaved material. Thus, we shall
investigate the composition and the fabric of ore materials at different scales (albeit here
only under the microscope). For the purpose of this book, “texture” will only reflect the
context of both the examined grain and its matrix. Grain shape and size, deformation,
abundance, distribution, mineral association (not necessarily paragenesis), and replace-
ments are the main descriptors for the characterisation of the intergranular relationships
(Barton, 1991; Craig, 2001).

Although there can be many ambiguities in the interpretation of the spatial relations
between grains seen under the microscope (applies to macroscopic scales as well), textures
disclose some genetic information and are crucial for mineral dressing, because the nature
of the ore, grain properties, and intergrowths directly constrain processing techniques.
However, is important to remember that the field of view under the microscope only con-
stitutes a minute detail from a large ore body and should, therefore, very cautiously be
taken into account for large-scale extrapolations.

In ores and rocks, a succession of geological events, such as magmatic ore formation -
hydrothermal overprinting - metamorphism - exhumation leading to supergene alteration
(weathering) may re-equilibrate the entire ore body and replace initial minerals, completely
obliterating earlier textures, unless geochemically inert and physically stable minerals are
preserved that reveal the true nature of the precursor (see photo below).

Right: Intercumulus chromite grain (medium grey) display-


ing embayments that are typical for the partial resorption of
a chromite in disequilibrium with the surrounding ultrama-
fic melt. Now, the grain is embedded in the sideritic matrix
(brownish grey) of a listwaenite, which otherwise hardly
bears any resemblance to the original dunitic rock (Fanja,
SW of Muscat, Sultanate of Oman).

Even a slowly declining temperature during the cooling of a magma suffices to put an al-
ready established mineral into a state in which its lattice is no longer stable, leading to an
exsolution process (e.g., hematite, rutile, and ulvöspinel separating from ilmenite). Fe-,
Cu-, and Ag-containing sulfides are even more out of equilibrium than the oxides as soon
as they come in contact with the atmosphere. These latter minerals will then start to oxidise
and produce a large number of secondary phases. Such processes generate a vast array of
minerals displaying various textures which, if understood, provide valuable clues on the
origin of rocks, ores, and minerals and their respective formation conditions.

As Barton (1991) states, “Ore deposits belong to the most complex inorganic features of
our planet. If this complexity is reflected in multiple mineralisations overprinting each
other but not necessarily producing a “valid” (meaningful) paragenesis, a determination of
9
the genesis of the ore is severely impeded and has to rely on individual mineralogical or
geological markers (single grains may yield substantial information). In order to develop a
“feeling” for the genetic hints “frozen” in a sample, it is necessary to delineate typical
characteristics reflecting particular processes.

Since a number of authors, such as Ramdohr (1969), Barton (1991), Craig (2001), and
many others have compiled comprehensive accounts of textures, the descriptions below
have been kept concise and are accompanied by typical images. The following section
briefly outlines the most important textures from different geological environments:
• Magmatic,
• Post-magmatic (Hydrothermal),
• Metamorphic,
• Sedimentary & Soils,
• Supergene, and
• Biomineralisation.

Footnote: All samples shown in the section below have been referenced in the description chapter
with respect to their sample sources and show details of presented photos.

Magmatic Environments
Cumulus
During the cooling of a silicate magma, minerals with a high melting point start to crystal-
lise and form euhedral to subhedral grains, because the surrounding melt poses few obsta-
cles to the developing crystal. Thus, oxides such as chromite, ilmenite, and magnetite plus
PGE alloys separate from the siliceous melt. Gravitation forces these dense minerals to
sink and accumulate (cumulus phase) in deeper parts of the melt body (massive ores can
form), a process that may also carry individual components into places in which the newly
formed mineral is no longer in equilibrium with the melt. This, in turn, causes edge disinte-
gration and results in rounded shapes where original crystal faces may or may no longer
show, resorption embayments can also develop.
Rapidly quenched extrusives (mainly basalts, but also
industrial slags and alloys) often exhibit skeletal crys-
tals, where the grains grow along preferred crystallo-
graphic orientations (axes) and do not find the time to
fill interior gaps. In cases of a very fast cooling, not
even crystal faces can develop and the grains have
rounded outlines and dendritic arrangements. Poikilitic
intergrowths between silicates, sulfides, and oxides are
a common phenomenon.

Right (top): Rounded euhedral chromite in a silicate matrix


(top) and

Right (bottom): Massive chromite (medium grey) containing


only minor amounts of silicates in fractures (darker grey).

10
Cumulus (continued)

Right: Skeletal magnetite in the siliceous matrix of a basalt


(light grey crystals).

Right: Dendritic magnetite in a historic Etruscan slag


(lighter grey), again in a silicate matrix.

Intercumulus (Intergranular)
During fractionated crystallisation, intergranular or
intercumulus textures develop wherever a mineral soli-
difies/crystallizes later than the surrounding matrix,
thus filling the left-over interstitial space. Examples
include various Cr, Fe, and Ti oxides as well as sul-
fides, the latter crystallizing from melts immiscible
with the dominantly silicic magma.

Right (top): Intergranular chromite (lighter grey) in ul-


tramafic silicate matrix (dark grey).

Right (bottom): Interstitial pyrite (cream) and con-


nected droplet in a magnetite matrix (grey).

Exsolution
Generally, exsolution textures imply slow or intermediate cooling rates, during which the
respective solid phase re-equilibrates mineralogically and texturally (Barton, 1970). Misci-
bility gaps in the phase stability of a number of minerals show up during cooling: A
homogenous single-phase high-temperature component breaks down into two or more dis-
tinct minerals. The result is an oriented intergrowth of those components that no longer fit
into the same crystal lattice at the lower temperature; examples are, for instance, surplus Fe
or Ti in ilmenite or extra Fe in Cu-sulfides, such as chalcopyrite.

11
Exsolution (continued)
This process presents a record of the thermal history of
the ore (Ramdohr, 1969). These exsolutions form dis-
crete mineral grains along crystallographically pre-
ferred directions of the host from which they differenti-
ated. Thus, oriented exsolution lamellae of rutile,
ulvöspinel or ilmenite may occur in titanomagnetite. In
other cases, more complex shapes develop.

Right: Finely lamellated and spindle-like exsolution of il-


menite (brownish grey) in hematite crystal (lighter grey).

Right: Cross-like sphalerite exsolution (grey) in chalcopy-


rite matrix (yellowish).

Right: Patchy, almost myrmekitic rutile exsolution (slightly


lighter grey) in ilmenite matrix (even grey surrounding of
rutile).

Post-magmatic Environments
Hydrothermal
Examining hydrothermal textures, it is necessary to dis-
tinguish between vastly different types, although all are
directly deriving from hydrothermal precipitation pro-
cesses involving open space deposition.

Right: Large open pore space of a black smoker wall with


sphalerite (medium grey) and fine specs of galena (lightest
grey), open pores (darkest grey).

The main textural types are:


• Submarine environments at or close to the seafloor where quenching of the hydrothermal
fluids produces sponge-like textures, often with skeletal mineral grains,
• Fault-/fracture- or dissolution-related precipitates; temperature fluctuations in the fluids
are often less pronounced than in the previous case; mineral precipitates are either very
well crystallised or appear as colloform layers sub-parallel to the wall-rock exhibiting
radial textures,

12
Hydrothermal (continued)
• Replacements of a) sediments, where the hydrother-
mal ore minerals mirror the sedimentary textures or
b) pre-existing minerals within hydrothermal ore
deposits where, for instance, a brecciation has created
pathways for successive infiltrating fluids or where a
strongly differing fluid composition (often tempera-
ture-related) places existing minerals in disequilib-
rium with the surrounding environment.

Common to all the described types is the fact that tem-


perature variations in the fluid result in mineralogical
variations that show distinct compositions of the pre-
cipitates. This could result in simple fluctuations, such
as the Fe-content in sphalerite, or the formation of quite
complex mineral suites as seen, for instance, in hydro-
thermal black smokers from inside (hot) to outside
(cold):
• Sphalerite-galena-pyrite-isocubanite-fahlore-enargite,
• Stibnite-realgar-orpiment-marcasite-cinnabar,
• Amorphous silica-barite.

In general, such a succession can develop over large


distances (tens or even hundreds of meters) or occur on
an extremely fine scale (2-3 cm, as observed in black
smokers); telescoping is the key term for this behav-
iour. Nevertheless, these types also share similar char-
acteristics that can grade into each other (e.g., well
crystallised and skeletal forms), making interpretation-
more difficult.

Right (top): Massive layered sphalerite (Schalenblende,


various grey and yellowish to brown internal reflections)
overgrowing jordanite (slightly anisotropic); dike filling.

Right (second from top): Late proustite (red internal reflec-


tions) filling quartz druse (milky); dike filling.

Right (third from top): Digital growth of orpiment (pinkish


medium grey); from black smoker.

Right (second from bottom): Syngenetic skeletal intergrowth


of sphalerite (medium grey) and chalcopyrite (yellow);
black smoker.

Right (bottom): Fibrous radially crystallised kermesite (pink


internal reflections); black smoker.

13
Post-magmatic Environments
Hydrothermal Replacements
A relatively common hydrothermal replacement has been named “chalcopyrite disease”
after the fine-grained impregnation of chalcopyrite in sphalerite (Barton and Bethke,
1987). This texture was originally thought to be the result of exsolution and/or co-precipi-
tation. More recent research, however, revealed that it is related to solid-state diffusion
reactions (DIS = diffusion induced segregation) by which sphalerite is replaced by various
Cu-Fe-S-phases (e.g., chalcopyrite bornite, cubanite, and others) in a bi-directional redox
reaction that first converts Fe2+ into Fe3+ (by increasing S-fugacity) and then flips it back
to Fe2+ (Cu reacts with Fe3+), Zn2+ is liberated in the process (Blesgen et al., 2004).
The replacement occurs at elevated temperatures (still
within the hydrothermal frame) but without fluids
being involved; rather, chalcopyrite or bornite grains
adjacent to the sphalerite directly act as Cu source.

Comparable with the reactions described above, there


are cases of single or multiple reaction rims forming
along a grain boundary. Again, we must consider geo-
chemical gradients leading to such zonal replacements.
At the moment, however, it is not clear whether during
any of these re-equilibrations water is essential (i.e., if
the mechanism is purely solid-state driven or not). Ne-
vertheless, wherever the two proto-minerals are associ-
ated in the hydrothermal system, it can be assumed that
re-equilibration starts soon after the establishment of
these phases, thus leading to the described rims.

Replacement reactions (not only in the hydrothermal


environments), often focus on cracks and cleavages,
along which the respective mineral is then replaced.

Right (top): “Dusty” chalcopyrite (yellow) in sphalerite


(grey); black smoker.

Right (second from top): Larger patches of chalcopyrite DIS


(yellow) in sphalerite (grey), pyrite matrix (light yellow).

Right (second from bottom): Bornite (violet grey) and chal-


copyrite (yellow) act as Cu source for DIS in sphalerite
(grey).

Right (bottom): Native silver (white) reacts with galena


(medium grey, large grain) to form an argentite/acanthite
reaction rim (darker grey); the formation of such rims may
extend into the supergene conditions.

14
Hydrothermal Replacements (continued)

Right: Chalcocite (bluish grey) irregularly replaces chalco-


pyrite (yellow) along fractures.

Metamorphic Environments
Recrystallisation
Hydrothermal minerals, especially when they develop
during fluid quenching on the oceanfloor, may precipi-
tate as more or less amorphous colloform materials.
Metamorphic/metasomatic overprinting then presents
the kinetic energy for re-equilibration and re-crystalli-
sation. Existing minerals re-adjust by increasing their
grain-size and straight boundaries form that meet at
specific angles (often 120º).

Right (top): Recrystallised euhedral quartz (dark grey) in


pyrite (creme) and chalcopyrite (yellow) from a fossil Ku-
roko-type deposit, as seen on the left; colloform amorphous
silica (light internal reflections) next to sphalerite (dark,
brown internal reflections) from a modern Kuroko-type ma-
rine hydrothermal deposit, as seen on the right.

Right (bottom): Four vonsenite grains (dark pinkish to


greenish greys) show the typical metamorphic angles and
straight boundaries.

Fracturing & Deformation


Various kinds of metamorphic processes, such as oro-
genic regional metamorphism or auto-metasomatic ser-
pentinisation, involve small- to large-scale displace-
ments and the related stress. In the most simple case,
this leads to a brecciation of rigid ore minerals; minera-
lising fluids can subsequently pass through the cracks
and deposit later mineral phases or lead to an alteration
of the primary mineral. Where malleable or flexible
minerals are involved, foliation will develop.

Right: Brecciated pyrite (creme) with vein-filling of chalco-


pyrite (yellow).

15
Fracturing & Deformation (continued)
Further stress, even on pliant minerals, ultimately
results in crenulations and kink-banding (see also pres-
sure twins).

Right: Galena (light grey) fills gaps produced by cleavage


of foliated muscovite (dark).

Right: Kink-banding of layered molybdite (olive to brown


pleochroism) and molybdenite (grey bireflectance).

Sedimentary Environments & Soils


Transport
In sedimentary environments, ore textures principally reflect the following:
• Transport that shaped rock/mineral fragments (sharp-edged brecciated or rounded con-
glomeratic),
• Physical movement during the chemical precipitation which formed oolitic/pisolitic
grains (shallow marine) as well as nodular spherulitic particles during “soil creep”,
• Stagnant but more or less unrestricted growth space, where chemical precipitates display
continuous banding, botryoidal or vermiform aggregates (e.g., sinters), and
• Cementation/Alteration binding and replacing sedimentary components, concretionary
accumulations are encompassed here as well.

Right: Conglomerate of uraninite (grey rounded grains) in a


massive matrix of pyrobitumen (almost the same colour as
uraninite), rounded pyrite (creme) with a thin overgrowth of
secondary pyrite, and irregular late stage pore-fillings of
gold (yellow).

Right: Finely laminated iron oolith with a relatively coarse


core of oolite fragment. Such a core, which can be com-
posed of any sedimentary material, such as sand grains or
fossil relics, is characteristic for oolitic/pisolitic materials.
Layering is always very fine, more than 100 layers have
been counted in a manganiferous pisolith of ~2 cm diameter.

16
Transport (continued)

Right: Soil spherulites are vastly different in their micro-


scopic appearance when compared with the previous tex-
ture. They generally bind rounded to sharp-edged particles
in a globular shape (often not entirely regular). A fine lami-
nation is missing, although occasionally a crude irregular
cm-thick layering can be observed (see dashed yellow line).

Sedimentary Replacements
Original sedimentary textures, such as layering, may be
preserved during replacement processes, even if a mild
metamorphic/metasomatic overprinting occurs.

A special case of replacement in sediments is the fos-


silisation of organic/organogenic remains by ore miner-
als. Organic carbon compounds (acting like activated
charcoal) attract and concentrate many base metals in a
reducing environment. An elevated sulfur fugacity then
converts these elements to sulfide minerals, replacing
internal and external shapes of the organism.

Right (top): Relict textures of clay-rich sediments infiltrated


by sulfide carrying solutions show sheet silicates (dark
grey), sphalerite (medium grey), chalcopyrite (yellow) and
pyrite (creme).

Right (center): Plant cells are replaced by pyrite (creme).

Right (bottom): Polychaeta relict (diagonal cut through tube


worm wall) from the former surface of a black smoker is re-
placed by pyrite (creme) and overgrown by chalcopyrite
during the growth of the smoker.

Supergene Replacement
Supergene weathering (oxidation, hydration) products and textures can be as versatile as
there are ore and secondary minerals. Thus, only a few are presented here. In sulfidic sys-
tems, the replacements strongly depend on the prevailing redox conditions that will either
support or hinder the formation of particular minerals. The associated textures vary from
patchy, irregular crack fillings, regular linear features relating to internal structures or
cleavages to full replacements (mostly as pseudomorphs, less common as paramorphs).

17
Supergene Replacement (continued)

Right: A characteristic replacement of chalcopyrite often


encompasses multiple replacement products coinciding with
a loss of Fe and an increasing oxidation potential: bornite-
chalcocite/covellite-tenorite and/or malachite plus azurite;
native copper and cuprite are less common in this suite.

Right (center): A fine irregular network of cryptomelane re-


places a siliceous rock; even quartz grains of a sandstone
have been seen to be corroded in this way.

Right (bottom): Hematite replaces siderite along its cleav-


age planes in a regular network, leaving some unaltered
carbonate in interstitial spaces.

Biomineralisation
Biomineralisation has been identified in various envi-
ronments, but in the context of ore minerals they are
especially intriguing in relation to heavy metal minera-
lisation. They are often associated with filamentous
bacteria, cyanobacteria or fungi (coatings, pseudo-
morphs). Nano-minerals, such as magnetite from mag-
netotactic bacteria, are not discussed here because the
crystallites are too fine-grained to be directly observed.

A case in which biomineralisation has been rejected is


framboidal pyrite. It is now thought to form from mag-
netic precursors, such as greigite (iron-thiospinel;
Fe3S4) which, during mild pore water movement, accu-
mulates into globular structures of various sizes.

Right (top): Cinnabar (red internal reflections) replaces fun-


gal filaments; from a black smoker rim.

Right (bottom): Framboidal pyrite (creme).

18

You might also like