Category Archives: Under the microscope

Optical mineralogy: Some terminology

Facebooktwitterlinkedininstagram
Radial clusters of calcite and void-filling botryoidal chalcedony in mineralized wood; Crossed polars.

Radial clusters of calcite and void-filling botryoidal chalcedony in mineralized wood; Crossed polars.

Petrographic (optical, polarizing) microscopes open doors to the architecture of rocks – crystals, grains, cements – a riot of colour and texture. Here is some common terminology that students will encounter on their journey through the lens. The entries have a semblance of order, beginning with the structure and symmetry of crystals, segueing through the important optical properties that form the basis for mineral identification.

My own experience at undergraduate level with the theory behind optical mineralogy was a bit chequered. Looking down the microscope was exhilarating, but marrying the kaleidoscope of shape and colour with the theory wasn’t as simple as my tutor made it out to be.  So, for anyone beginning their foray into this subject, I hope the notes below help. Optical mineralogy is key to understanding the mineralogical and chemical evolution of rocks, their compositional and textural changes, their paragenesis.  So persevere!

 

The terminology

Unit cell: At the atomic scale, the arrangement of atoms that represents the fundamental structure of a mineral in crystal form. The crystals we see consist of a three-dimensional array of stacked unit cells. This means that the overall shape of the crystal mimics its unit cell. The simplest unit cell is a cube; cubes of the same size will stack perfectly. Not all polygonal geometries allow such stacking, for example cells with triangular sides will stack neatly together, but those with 5-sided faces (pentagons) will not. Consideration of the unit cells and their symmetry form the basis for definition of the 6 (or 7) crystal systems.

Unit cells are the foundations of all crystalline materials. The crystals we see are built from 3-dimensional arrays of these cells

Unit cells are the foundations of all crystalline materials. The crystals we see are built from 3-dimensional arrays of these cells

Crystallographic axes: Three or four axes about which a crystal can be rotated through 360o.  The axes intersect at a single point (the centre of symmetry). They are labelled according to their lengths. If axes are the same length, then they are referred to as a1, a2, a3 etc. If they have different lengths, they are labelled a, b, and c. Thus, in the cubic (isometric) crystal system they are labelled a1, a2, a3, and in the tetragonal system a1, a2, c. The hexagonal system is the only one with four axes. Angles between axes are labelled α, β, γ.

Axes of rotation that determine 6-, 4-, 3-, and 2-fold crystal symmetry; face-centered axes for isometric and hexagonal structures, crystal face terminations, and crystal face intersections

Axes of rotation that determine 6-, 4-, 3-, and 2-fold crystal symmetry; face-centered axes for 6- and 4-fold structures, crystal face terminations for 3-fold, and crystal edge intersections for 2-fold symmetry.

Crystal symmetry: Symmetry describes the shape of an object and can be represented both mathematically and visually. In crystallography, the two most useful forms of symmetry are (mainly because they are the easiest to visualize):

  1. Axes of rotation (crystallographic axes) where a particular crystal face will be repeated during rotation through 360o. The number of repetitions for a 360o rotation can be 2, 3, 4, or 6, that are referred to as two-fold, three-fold, four-fold, and six-fold (axial) symmetry respectively.
  2. Planes of symmetry where two parts of a crystal are mirror images. For an analogy, think of this concept in terms of the common bilateral symmetry in many living organisms, such as people, and many classes of mollusc. Note that planes of symmetry are NOT the same as twin planes.
  3. Additional elements of symmetry include: A centre of symmetry, where a crystal face is reflected from one side to another or is repeated by inversion, and an axis of rotary inversion.
Elements of symmetry: Mirror planes (bilateral symmetry), centers of symmetry, and axes of rotary inversion.

Elements of symmetry: Mirror planes (bilateral symmetry), centers of symmetry, and axes of rotary inversion.

Crystal systems: There are 6 crystal systems based on combination of the elements of symmetry; a seventh system – trigonal – is sometimes considered a subclass of the hexagonal system. There are 32 crystal classes based on combinations of the symmetry elements. The defining criteria are axial lengths, the angles between axes, and axial symmetry (the number of repetitions about an axis).

  • Cubic (or Isometric) system: The most symmetric group. All three axes are the same length and are at right angles to each other.

a1 = a2 = a3                  α = β = γ = 90o

2, 3, and 4-fold symmetry depending on the class

Common crystal forms: cubes, octahedra, dodecahedra: e.g., Halite, pyrite, fluorite, garnet

Properties of isometric, tetragonal, and hexagonal crystal systems

  • Tetragonal system: Liken this group to isometric crystals stretched along the c

a1 = a2 ≠ c                    α = β = γ = 90o            Mostly 2- and 4-fold symmetry

Common crystal forms: Prisms, bipyramids with or without prisms: e.g., zircon, chalcopyrite, rutile

 

  • Hexagonal system: This system has 4 axes, 3 of which are perpendicular to c

a1 = a2 = a3 ≠ c             Angles between a1 = a2 = a3 = 120o

6-fold symmetry

Common crystal forms: Prisms, bipyramids: e.g., apatite, beryl,

The Trigonal subsystem has one 3-fold axis or rotation. Three important examples are quartz, calcite and dolomite, commonly formed as bipyramids, rhombohedra, and scalenohedra.

 

  • Orthorhombic system:

a ≠ b ≠ c                      α = β = γ = 90o

2-fold symmetry

Common crystal forms: Prisms, bipyramids: e.g., olivine, cordierite, hypersthene

 

Properties of orthorhombic, monoclinic, and triclinic crystal systems

  • Monoclinic system:

a ≠ b ≠ c                      α = γ = 90o, β ≠ 90o

2-fold symmetry

Common crystal forms: Prisms, pinacoids (flattened prisms): e.g., orthoclase, diopside, sphene, staurolite, most amphiboles.

 

  • Triclinic system: The least symmetric group

a ≠ b ≠ c                      α ≠ β ≠ γ ≠ 90o

No axes of symmetry!

Common crystal forms: Prisms, bipyramids: e.g., microcline, plagioclase, kyanite

 

Plain polarized light: (PPL) The light transmitted through a polarizer, located below the microscope condenser (the condenser focuses this light through an opening in the rotary stage). The polarizer filters out all light frequencies other than those that vibrate in a single plane. The polarizers here are oriented E-W (Note the upper polarizer is oriented N-S).  Minerals in thin section examined under PPL show important identifying characteristics such as crystal shape, cleavage, breakage patterns, relief, and pleochroism. The light path for PPL is shown diagrammatically below.

 

Crossed nicols (polars): The upper polarizer, between the objective lenses and the Bertrand lens, filters out all remaining frequencies present in plain polarized light (PPL). The nicols can be moved in and out of the light path. If you look through the oculars when the polarizer is in the light path (i.e., crossed nicols) then no light will reach the oculars – all will be black. However, if a thin section is placed between the lower and upper polarizers, most minerals will reorient the PPL such that some of this light will pass through the upper polarizer; this light will contain slow (ordinary) and fast (extraordinary) vibration directions that will arrive at the eye pieces at slightly different times. The resulting interference produces the kaleidoscope of colour among all the minerals present. Minerals that permit the passage of light are called anisotropic; those that do not are isotropic.

The path for polarized light in a standard petrographic microscope. The thin section can be viewed in plan polarized light when the upper polarizer is removed from the light path. When the upper polarizer is in place (crossed nicols/polars), the colour of an anisotropic mineral will depend on its birefringence.

The path for polarized light in a standard petrographic microscope. The thin section can be viewed in plan polarized light when the upper polarizer is removed from the light path. When the upper polarizer is in place (crossed nicols/polars), the colour of an anisotropic mineral will depend on its birefringence.

Extinction: As the microscope stage is rotated, under crossed nicols, either the fast or slow vibration direction of light exiting an anisotropic mineral will be blocked by the upper polarizer – at this point no light is transmitted through polarizer and it appears black – the transmitted light has been extinguished. This alignment occurs four times during one 360o rotation (because the fast and slow rays are perpendicular), and therefore each extinction event is 90o apart.

 

 

Extinction angle: The angle at which extinction occurs relative to crystal habit or prominent cleavage, may vary between 0o and 89o. The angle can be easily measured using the grid on the rotating stage. Extinction angles can be used to help identify minerals.

  • Parallel extinction: Extinction parallel to the crystal elongation direction or cleavage, that also parallels the crystallographic c axis. A common example is muscovite
  • Inclined extinction: The most common type, between 0o and 89o.
  • Symmetrical extinction: In minerals that have two prominent cleavage planes (such as calcite) – if the extinction angles measured from each cleavage are the same, then extinction is symmetrical.

 

Birefringence: Plain polarized light that passes through a mineral is resolved into mutually perpendicular fast and slow rays that will each have different indices of refraction as (i.e., their refraction paths and velocities will be different). Birefringence is the maximum difference between these two index values. Under crossed nicols, the difference is manifested in the ‘intensity’ of interference colours. For example, quartz has low birefringence showing typically as first-order greys, whereas high birefringent calcite displays 3rd-order colours. A typical chart showing the range of birefringence values and colours is shown below.

The standard Michele-Levy chart of birefringence and thin section thickness

The standard Michele-Levy chart of birefringence and thin section thickness

Isotropic minerals: Unlike anisotropic minerals, this group cannot reorient plain polarized light which means that no light will pass through the upper polarizer; they will appear black through all rotations of the stage. All isometric minerals (cubic system) are isotropic (e.g., garnet, fluorite, halite, spinels). Note that an anisotropic mineral oriented at right angles to its optic axis will appear isotropic (because there is no resolution of fast and slow rays along the optic axis, i.e., no birefringence).

 

Anisotropic minerals: Minerals in thin section reorient plain polarized light, resolving it into two vibration directions that will pass through the upper polarizer when nicols are crossed. One direction contains a fast light ray (also called the extraordinary ray), the other a slow ray (ordinary ray); the fast and slow rays are perpendicular to each other.  In thin section minerals appear coloured and will go into extinction four times during a 360o rotation of the stage (and thin section). Anisotropic minerals are further divided into uniaxial and biaxial based on the presence of one or two optic axes.

 

Optic axes: Axes along which plain polarized light is NOT split into fast and slow rays. In other words, there is NO birefringence parallel to an optic axis. Minerals having a single optic axis are uniaxial; those having two are biaxial.

 

Interference figures: These appear as curved isogyres or crosses when a mineral is viewed under crossed nicols at high magnification with the Bertrand lens inserted. There are two basic types:

  1. Uniaxial crosses that do not break up or rotate as the stage is rotated.
  2. Biaxial isogyres or crosses that rotate and move with the stage; crosses will break into two curved isogyres.

Determining the +ve or -ve signs for either case is described under the Uniaxial and Biaxial headings.

 

Biaxial minerals: Anisotropic minerals where plain polarized light entering at any angle, other than along two optic axes, is resolved into two planes of polarized light; these two planes each contain the fast and slow rays. The resulting colour depends on the different in the refractive indices of these two light paths – i.e., the birefringence. Note that mineral sections perpendicular to the optic axes will appear isotropic. Some typical interference figures are shown below.

Common biaxial interference figures manifested as curved or crossed isogyres

Common biaxial interference figures manifested as curved or crossed isogyres

Minerals may be positively or negatively biaxial, depending on the orientation of fast and slow rays. Determining the sign (+ve or -ve) is done as follows:

  1. Zoom in on the mineral of choice to the highest possible magnification (under crossed nicols).
  2. Insert the Bertrand lens. You will (hopefully) see interference figures manifested as curved isogyres or isogyre crosses.
  3. Rotate the stage and observe how the isogyres rotate; a cross will break up into separate isogyres.
  4. Orient an isogyre until it is concave to the NE.
  5. Insert the accessory plate (a plate containing a gypsum wafer) and note the change in colour on the immediate inside of the isogyre, from first-order grey to:
  • 1st-order blue, if the mineral is biaxial +ve
  • 1st-order yellow, if it is biaxial -ve.
Determining the sign for biaxial minerals from interference figures, using an accessory plate

Determining the sign for biaxial minerals from interference figures, using an accessory plate

Uniaxial: Anisotropic minerals where plain polarized light entering at any angle other than along a single optic axis, is resolved into two planes of polarized light; these two planes contain the fast and slow rays. The resulting colour depends on the different in the refractive indices of these two light paths – i.e., the birefringence. Note that mineral sections where the optic axis is vertical will appear isotropic under crossed polars. However, when the Bertrand lens is inserted, you will see a dark cross that is also centered.

Note too that the N-S axis of the interference cross parallels the fast (extraordinary) ray, and the E-W axis the slow (ordinary) ray.

Minerals may be positively or negatively uniaxial, depending on the orientation of fast and slow rays. Determining the sign (+ve or -ve) is done in the same way as biaxial minerals, except:

  1. The interference figure is a cross that DOES NOT split into curved isogyres when the stage is rotated. Nor does the cross rotate. If the mineral is oriented such that the optic axis is vertical, then the cross will not move across the field of view as the stage is rotated. Mineral sections oblique to the optic axis will show variable positions of the cross, but will remain oriented N-S/E-W.
  2. When the gypsum plate is inserted the colour in the NE quadrant will become blue if uniaxial +ve, and yellow if uniaxial -ve. The most common uniaxial +ve mineral is quartz.
Common uniaxial interference figures.

Common uniaxial interference figures.

Relief: Light is refracted, or ‘bent’ when it passes from one transparent material to another (air to water; air to thin section mineral); refraction corresponds to a change in direction and either an increase or decrease in velocity. The refractive index (RI) of a mineral is the ratio of the velocity of light in air to that in the mineral. Indices differ amongst minerals such that:

  • Isotropic minerals have a single RI
  • Uniaxial minerals have two RI along different axes, and
  • Biaxial minerals have 3 RI.

The different RIs for any mineral are manifested as relief in plain polarized light, where one mineral may appear to stand above or below its neighbours. The orientation of uniaxial and biaxial minerals will also determine their relief. This is a useful criterion for mineral identification. For example, most ferromagnesian minerals have relatively high relief compared to quartz and feldspar; calcite and dolomite relief is even greater. However, quartz and feldspar have very similar relief and the distinction, particularly for untwinned potassium feldspars, must rely on criteria such as cleavage and interference figures.

There is a marked difference in the relief of plagioclase (white) and the darker coloured, high-relief pyroxenes in this gabbro. View is 3 mm wide.

There is a marked difference in the relief of plagioclase (white) and the darker coloured, high-relief pyroxenes in this gabbro. View is 3 mm wide.

Pleochroism: In plain polarized light, some minerals change colour during rotation of the stage. Colour changes may be subtle or intense. Pleochroism is a useful criterion for mineral identification. Common examples in sedimentary rocks are:

  • Amphiboles in various shades of green and yellow – hornblende has fairly intense pleochroism in shades of olive green, glaucophane in shades of blue and blue-green.
  • Pyroxenes in various shades of green and brown.
  • Micas – muscovite is usually colourless but may be pleochroic in pale reds or browns, Biotite is strongly pleochroic in shades of brown, as seen in the video below. In this example the biotite is from a well foliated greenschist (Otago Schist, New Zealand); the view is under plain polarized light. A small, partly rotated garnet has nudged next to the biotite blades.

 

 

Cleavage: A plane of weakness within a crystal that will break with relative ease. It is caused by weak bonds between planes of atoms within a crystal lattice; the pattern of weakness repeats regularly through a crystal. Some minerals have poor or no cleavage (e.g., quartz, olivine); others have excellent cleavage along several lattice planes (e.g., calcite, feldspar). Cleavage is a defining characteristic of a mineral, particularly in thin section.

Two prominent sets of cleavage typical of pyroxene, that intersect almost at right angles in sections oriented perpendicular to the c-axis. Plain polarised light.

Two prominent sets of cleavage typical of pyroxene, that intersect almost at right angles in sections oriented perpendicular to the c-axis. Plain polarized light.

Twinning: A symmetrical intergrowth of two separate crystals of the same mineral, that share the same mineral lattice. In thin section under crossed nicols, each twin segment will go into extinction at different rotations of the microscope stage. There are many kinds of twinning. For example, plagioclase may show albite, carlsbad, or pericline twins individually or as combinations in the same crystal. Important optical properties of twins that help mineral identification include extinction angle (whether straight or inclined), and 2V angles. Note that twin planes are NOT the same as planes of crystal symmetry.

A nice selection of plagioclase crystals in gabbro, showing prominent albite, pericline and carlsbad twinning. Crossed nicols. View is 3 mm wide.

A nice selection of plagioclase crystals in gabbro, showing prominent albite, pericline and carlsbad twinning. Crossed nicols. View is 3 mm wide.

Crystal zoning: Zoning commonly displays as concentrically arranged crystal growths, where the composition changes outwards from the crystal interior. The zones maintain the same crystallographic and optic axes. The changes in composition involve substitution of certain cations. For example, in calcite, Fe2+ and Mn2+ substitute for Ca2+, and in plagioclase sodium may substitute for calcium such that the inner core of a crystal is a calcium anorthite and the outer zone is a sodium albite. Zoning indicates changing fluid or magma crystallization conditions. Zoned crystals may also be twinned.

 

Here are some excellent sites and texts that provide greater detail

The Open University

Raith, M.M., Raase, P., and Reinhardt, Jurgen. (2011) Guide to Thin Section Microscopy. Open access PDF

Johnson, E.A., Liu, J. C., and Peale, M. Introduction to Petrology. Licensed under CC-BY-SA 4.0. Source

Heinrich, E.WM. 1965. Microscopic identification of minerals. McGraw Hill. An old text but still gold.

 

Other posts in this series

Miller Indices in crystallography

Lithic grains in thin section – avoiding ambiguity

Ignimbrites in outcrop and thin section

Sandstones in thin section

Greywackes in thin section

The mineralogy of sandstones: porosity and permeability

The mineralogy of sandstones: Quartz grains

The mineralogy of sandstones: Feldspar grains

The mineralogy of sandstones: lithic fragments

The mineralogy of sandstones: matrix and cement

The provenance of detrital zircon

The provenance of sandstones

Provenance and plate tectonics

 

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Greywackes in thin section

Facebooktwitterlinkedininstagram

Typical greywacke in outcrop and thin section

Greywacke petrography

On the outside, greywacke is generally presented in varying shades of grey commonly with greenish hues, hammer-ringing hard, even splintery indurated rock. Most originate as deep- water successions of turbidites and debris flows, in seemingly endless stratigraphic piles 100s to 1000s of metres thick. Typically, they lack body fossils but may contain diverse trace fossils on bedding.

Greywackes in thin section can be difficult to work with. By definition they contain at least 15% matrix, but the distinction between framework grains and matrix can be difficult.  In part this is because there is an almost continuous grain-size gradation from the coarsest to finest particles; what constitutes matrix becomes a moot point.  The other problem with wacke matrix is that it tends to be chemically reactive and given the depths of burial, may be completely masked by diagenetic products such as clays, sericite, and chlorite. Labile constituents like feldspar, mica and ferromagnesian minerals will be prone to alteration and even complete replacement by calcite, chlorite, and low metamorphic grade minerals like prehnite.

Sericite: is a flaky, white mica and common alteration product of feldspar. In thin section it usually presents as fine ragged crystals (rather than the more uniform muscovite flakes), concentrated along feldspar cleavage planes, or distributed across the entire crystal. It has high birefringence and appears to sparkle against the dull background of altered grains and matrix.

Chlorite: is easily distinguished from sericite by its low birefringence, varying shades of green (in PPL), and crystal habit that is also variable, from fibrous, spherulitic or vermiform (worm-like). May be pleochroic in shades of green and yellow. It is commonly associated with low grade metamorphism and hydrothermal alteration.

Refer to this link for a reminder of optical mineralogy terminology

 

Greywacke

Two samples from the Omarolluk Fm. Paleoproterozoic, Belcher Islands, Hudson Bay. The notes apply to both sets of images, plain polarized light (PPL) on the left, crossed nicols on the right.

Greywacke, Omarolluk Fm. Paleoproterozoic, Belcher Islands, Hudson Bay.

Thin section of greywacke, Omarolluk Fm. Paleoproterozoic, Belcher Islands, Hudson Bay. PPL and crossed nicols.

Thin section of greywacke, Omarolluk Fm (#2). Paleoproterozoic, Belcher Islands, Hudson Bay. PPL and crossed nicols.

Thin section of greywacke, Omarolluk Fm. (#2) Paleoproterozoic, Belcher Islands, Hudson Bay. PPL and crossed nicols.

  1. Grain size ranges from 400μm (0.4 mm) to clay, with a mean hovering between 50-100μm.
  2. Framework grains are angular to surrounded.
  3. Poorly sorted.
  4. Framework is a mix of clast-supported and matrix-supported.
  5. The rock is texturally and mineralogically immature.
  6. Monocrystalline (m) and polycrystalline quartz (p) are generally clear in PPL (about 30-40% in total). Some of the silt and fine sand sized quartz may be derived by breakdown of larger polycrystalline grains.
  7. Monocrystalline quartz grains have variable degrees of strain, ranging from straight to sweeping extinction. This suggests that most of the crystal deformation was inherited from the source rocks, rather than being postdepositional.
  8. Plagioclase (pl – albite and carlsbad twinning) and potassium feldspar (k) (total about 30%) tend to be cloudy in PPL because of alteration to sericite (examples in circled area), chlorite, and clays. Chlorite gives grains and matrix a greenish hue. There are a few small patches of calcite replacement of feldspar (cc)
  9. Lithics (15% in Omarolluk 1; 20-25% in Omarolluk 2) consist of mud-siltstone and cherty mudstone (ls) and volcanic (lv), the latter showing relict feldspar lath textures.
  10. Depositional matrix and diagenetic cement about 15%.
  11. Muscovite flakes (mu <0.5%) are mostly altered to chlorite and bent around stronger grains (arrows). The few small pyroxene grains are highly altered.
  12. Paragenesis: Deposition → compaction (continuous) → feldspar alteration to sericite → clay precipitation → chlorite precipitation.

Greywacke

Greenland Group, Ordovician, NZ. Bar scale 350μm (0.35 mm).

Image credit: JJ Reed P35079, GNS Science. (2004). Petlab.

Thin section of greywacke, Greenland Gp. PPL on left, crossed nicols on right. Bar scale 0.35 mm.

Thin section of greywacke, Greenland Gp. PPL on left, crossed nicols on right. Bar scale 0.35 mm.

This sample gives us a nice illustration of a greywacke that has been subjected to low grade metamorphism, manifested as changes to grain framework and incipient structural fabrics.

  1. Quartz grains, about 40%, are angular and range from medium sand size to silt.
  2. Matrix appears at first glance to be very high – 40-50% but zoom in on the images to see the outlines of lithic grains (a few examples shown by yellow arrows). Many of the grains have amalgamated into diffuse masses, in part combining with the original matrix that probably was of similar composition to the lithics. For many lithic grains, the original size and shape will have changed because of advanced compaction, metamorphic reactions, and the development of an incipient structural fabric – cleavage. In this case, the distinction between matrix and post-depositional textures is fraught.
  3. Opaque minerals < 1%.
  4. The linear fabric, accentuated by the dark, evenly spaced, subparallel bands is incipient cleavage, a penetrative fabric that results from grain dissolution and precipitation through the entire sample (red arrows). Zoom in to see how the cleavage is diverted around quartz grains. Accompanying the cleavage is a subtle alignment of very fine mineral precipitates – probably a mix of chlorite and illite.
  5. Paragenesis: Deposition → compaction (continuous) → clay precipitation → formation of cleavage → clay + chlorite precipitation.

 

Greywacke (feldspathic wacke) – Crossed nicols only.

Rakaia terrane, Late Triassic to Permian (part of the Torlesse composite terrane).

Image credit:, P29217, GNS Science. (2004). Petlab.

Thin section of greywacke, Rakaia Terrane. Crossed nicols only. Bar scale is 0.5 mm

Thin section of greywacke, Rakaia Terrane. Crossed nicols only. Bar scale is 0.5 mm

  1. Coarse-grained, clast sizes ranging from 800μm (0.8 mm) to silt sized.
  2. Angular, poorly sorted, mostly clast-supported.
  3. Quartz 15-20%, a mix of polycrystalline (p) and monocrystalline grains (m).
  4. Feldspar 40-50% is mostly untwinned potassium feldspar (k – good cleavage) – there are a few small albite twinned plagioclase grains (pl). Some alteration to sericite (s).
  5. Lithics 15-20% (l).
  6. Matrix 15%.
  7. A small fracture traverses the upper half of the image (f1). The fracture is filled with quartz (?) where it cuts across a large feldspar grain.
  8. A second structural discordance along the left margin is more diffuse (f2); here there appears to be grain breakage and possibly some fracture-filling mineral precipitation that parallels the trend of the structure. It also appears to truncate the quartz-filled fracture.
  9. Paragenesis: Deposition → compaction (continuous) → feldspar alteration → cement precipitation → deformation fracture 1 → deformation fracture 2.

 

Greywacke

A typically 'dirty looking' greywacke where the distinction between framework and matrix is obscured by clay diagenesis and feldspar alteration. There is a high lithic content (l), plus a poorly sorted array of angular quartz grains (q) and K-feldspars (k), including untwinned and perthite twinned varieties. Some of the feldspars have been altered significantly and look superficially like lithic grains. Overall, the sample is texturally and mineralogically immature.

A typically ‘dirty looking’ greywacke where the distinction between framework and matrix is obscured by clay diagenesis and feldspar alteration. There is a high lithic content (l), plus a poorly sorted array of angular quartz grains (q) and K-feldspars (k), including untwinned and perthite twinned varieties. Some of the feldspars have been altered significantly and look superficially like lithic grains. Overall, the sample is texturally and mineralogically immature. Left: Plain polarized light; Right: Crossed polars.

Volcanic lithic greywacke

A coarse-grained greywacke with large lithic grains that have felted micro-feldspar textures and the occasional plagioclase phenocryst. K-feldspar is also abundant in untwinned (k) and gridiron twinned crystals (g), all showing significant alteration (probably to clays), giving them a dirty brown appearance in PPL. Quartz grains are subordinate, angular, and include a few polycrystalline types (p). There are a few coarse grained opaques (o). The rock overall is typically very poorly sorted and mineralogically immature.

A coarse-grained greywacke with large lithic grains that have felted micro-feldspar textures and the occasional plagioclase phenocryst. K-feldspar is also abundant in untwinned (k) and gridiron twinned crystals (g), all showing significant alteration (probably to clays), giving them a dirty brown appearance in PPL. Quartz grains are subordinate, angular, and include a few polycrystalline types (p). There are a few coarse grained opaques (o). The rock overall is typically very poorly sorted and mineralogically immature. Left: Plain polarized light; Right: Crossed polars.

Other posts relevant to this article

The mineralogy of sandstones: porosity and permeability

The mineralogy of sandstones: Quartz grains

The mineralogy of sandstones: Feldspar grains

The mineralogy of sandstones: lithic fragments

The mineralogy of sandstones: matrix and cement

The provenance of detrital zircon

The provenance of sandstones

Provenance and plate tectonics

Sandstones in thin section

Ignimbrites in outcrop and thin section

Lithic grains in thin section – avoiding ambiguity

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Sandstones in thin section

Facebooktwitterlinkedininstagram
Calcite cemented subarkose, Proterozoic Altyn Fm. southern Alberta

Calcite cemented subarkose, Proterozoic Altyn Fm. southern Alberta. Crossed nicols.

The selection of sandstones in plain polarized light (left) and crossed nicols (right) shown here illustrate common petrographic attributes. Abbreviations are explained in the caption to each set. For explanations of terminology see the relevant articles in this series.

Arenite classification is from Pettijohn, Potter and Siever, (1973).

Sandstone classification

Sandstone classification

The assessment of rock paragenesis (diagenetic history) is based, where possible, on cement stratigraphy, cross-cutting relationships of cements and replacement textures, recrystallization textures, and any structural fabrics that might be present.

Refer to this link for a reminder of optical mineralogy terminology

 

Common textural descriptions in thin section

The two images below are annotated to show the common textural properties of sandstones. The thin section is shown only in plain polarized light; in most cases it is easier to see textural attributes  in PPL than under crossed nicols. A scanning electron micrograph (SEM) of a rock from the same formation reminds us that all physical aspects of a rock are three dimensional.

Sandstone example used to show common textures in thin section. An SEM from the same formation is shown for comparison. Cretaceous Ellerslie Fm. Albert foreland basin.

Sandstone example used to show common textures in thin section. Right: An SEM from the same formation is shown for comparison. Cretaceous Ellerslie Fm. Albert foreland basin. Bar scale 0.3 mm. Left plain polarized light.

  1. Grain size. Medium sand.
  2. Sorting. Well to moderately well-sorted – most grains are a similar size and shape.
  3. Roundness. Well-rounded to subrounded. Care needs to be taken to distinguish original grain boundaries. The SEM image shows several grains with sharp terminations and straight edges that in this example are the result of quartz overgrowths (arrows); i.e., they are diagenetic textures (part of the cementation history) rather than depositional textures.
  4. Framework. In thin section grains are consistently in 2-D contact, and in the SEM we see 3-D contacts. This is an example of a grain-supported framework.
  5. Matrix. Intergranular pores contain small amounts of silt-sized quartz-feldspar and clays. Distinguishing detrital clays (matrix) from diagenetic clays is difficult in thin section. However, an SEM of the same rock (below) shows kaolinite ‘books’ and filamentous illite crystals that overlie the quartz overgrowths – these are diagenetic products.
  6. Porosity. In thin section about 15% (the blue material is resin); 15-20% in the SEM.
  7. Mineralogy. Quartz 80-90% (clear, irregular fractures but no cleavage), lithics (l) 10-15% – dirty brown colours in PPL, feldspar <5% (good cleavage in PPL). Each component can be described in terms of its composition (mono- or polycrystalline quartz, volcanic or sedimentary lithics, plagioclase or potassium feldspars) – see other posts in this series for details on how to do this.
  8. Cements. Quartz overgrowths are the first major cement component. In thin section detrital clays are seen as a thin dark rind on grain boundaries (red arrows). In the SEM, clay dustings and tendrils cover all grains (detailed analysis of similar samples shows illite and kaolinite). There are no calcite or dolomite cements in these examples although they are common in many sandstones.
  9. Rock name. In accord with the classification scheme above: = sublitharenite.
SEM of quartz overgrowth, kaolinite and illite cements in a lithic arenite

SEM of euhedral quartz overgrowths (q), kaolinite ‘books’ (k) and  illite cements (i) in a lithic arenite.

Quartz arenite:

Mukpollo Fm. Paleoproterozoic. Belcher I.

quartz arenite

Quartz arenite. Left plain polarized light. Right Crossed nicols

  1. Coarse-grained, well-sorted, grains well rounded. Clast-supported framework. Original matrix was <10%. The rock is texturally and mineralogically supermature.
  2. Well-rounded monocrystalline quartz grains (m); original grain boundaries (black arrows) overgrown by syntaxial quartz cement (the cements move into extinction with the parent grain – they are optically continuous).
  3. There are two polycrystalline quartz grains in this view (p). They too have syntaxial overgrowths that are in optical continuity with each subcrystal (red arrows).
  4. Polycrystalline grain shapes tend to be less well rounded because they are liable to break along sub-grain boundaries or deform during compaction.
  5. Contacts between cements in the monocrystalline varieties are usually straight-flat crystal faces (c). Contacts between polycrystalline overgrowths are irregular and interpenetrating.
  6. Effective porosity is almost zero in this rock. Fluid flow along grain and cement overgrowth boundaries would have been mostly diffusive.
  7. Provenance: Archean felsic basement rocks. Most of the original feldspar was removed by mechanical abrasion during transport and sediment reworking.
  8. Paragenesis: A relatively simple progression from deposition → compaction (continuous) → quartz cement precipitation during advective fluid flow.

Quartz arenite:

Mukpollo Fm. Paleoproterozoic. Belcher I.

quartz arenite

Quartz arenite. Left plain polarized light. Right Crossed nicols

  1. Well-sorted, grains well rounded. Clast-supported framework. Original matrix was <10%. The rock is texturally and mineralogically supermature.
  2. As above, dominated by monocrystalline quartz (m – 80%), plus 15% polycrystalline quartz (p), and about 5% feldspar (pl). The lithic clast (l) has been deformed by compaction.
  3. Quartz overgrowth cements are syntaxial (arrows).
  4. Plagioclase is partly altered to sericite (a common white mica alteration product) and some chlorite. Some albite twinning is still visible. The alteration products tend to concentrate along cleavage – this makes the distinction between quartz and feldspar a little easier.
  5. There are two patches of dolomite pore-filling cement (d) that overgrow the quartz cements, and on close inspection show incipient replacement of quartz.  Thus, the dolomite is a late diagenetic component.
  6. Provenance: Archean felsic basement rocks. Most of the original feldspar was removed by mechanical abrasion during transport and sediment reworking.
  7. Paragenesis: Deposition → compaction (continuous) → quartz cement precipitation during advective fluid flow → a change to carbonate-dominated fluids and precipitation of late dolomite cement at deep burial.

Quartz arenite (polycrystalline quartz)

Taratu Fm. Otago, NZ. Bar scale 1 mm.

Quartz arenite dominated by polycrystalline grains

Quartz arenite dominated by polycrystalline grains. Left plain polarized light. Right Crossed nicols

  1. This sandstone, although >90% quartz, is made up almost entirely of polycrystalline grains where the subcrystals in each grain are equant.
  2. Grains are well rounded, but several show evidence of breakage along subcrystal boundaries.
  3. Arenite is texturally submature and mineralogically mature.
  4. Grain size is bimodal – medium sand size (about 50%) and very fine sand – silt sized grains. Many sand grains appear to ‘float’ in the fine fraction.
  5. The finer size fraction is much the same size as the whole grain subcrystals.
  6. We can infer that the fine fraction was derived by breakage of the sand-sized polycrystalline grains.
  7. There are no quartz overgrowth cements. Intergranular cements are probably kaolinite (according to other studies) but distinguishing this in thin section is difficult.
  8. Compaction has produced incipient fractures along crystallite boundaries in some polycrystalline quartz grains (arrows).
  9. The polycrystalline quartz was derived from Otago Schists.
  10. Image credit: S.G. McMillan, GNS, Petlab database sample OU62786.

 

Subarkose (plain polarized light only. Bar scale 1mm)

Kapuni  Gp. (Paleocene-Eocene), Taranaki Basin.

Subarkose under plain polarized light

Subarkose under plain polarized light

  1. Poorly sorted, angular to surrounded grains. Fully clast-supported framework.
  2. Texturally submature, mineralogically immature.
  3. Feldspar 70%, lithics 15%, monocrystalline quartz 15% (m), the occasional white mica flake.
  4. Feldspar is a mix of plagioclase (pl) and potassium feldspar (k). Plagioclase can be distinguished by alteration bands that mimic twinning. Alteration tends to be more advanced in the plagioclase grains as a mix of sericite and finely crystalline calcite. The variable degree of alteration among the feldspars suggests that some of it may be inherited from the sediment source.
  5. In this sample there is little evidence of quartz or feldspar overgrowth.
  6. Most grains have a thin, dark rim that is probably incipient clay cement. Fibrous chlorite (cl) occurs as the occasional pore filling cement that overlies the dark clay rims.
  7. In plain polarized light, feldspar can be distinguished from quartz by prevalence of cleavage, and broken, angular grain boundaries that coincide with cleavage planes.
  8. Lithics are predominantly mudstone-siltstone. Compaction deformation of lithics is visible where they have been moulded or fractured around stronger grains (arrows).
  9. Blue resin has filled intergranular porosity (About 10-12%).
  10. Provenance. Primarily felsic plutonic rocks, (granite, granodiorite). Possibly first cycle where removal of mechanically weak feldspar by reworking and abrasion was limited.
  11. Paragenesis. Deposition → compaction (continuous) → +/- feldspar alteration and minor dissolution → precipitation of clay cements → +/- chlorite cement
  12. Image credit: GNS, Petlab database sample P46297.

 

Glauconitic calcareous subarkose

– Glauconite peloids up to 0.5mm across (10%). Some glauconite grains are deformed around stronger quartz and feldspar grains (by compaction). – Quartz grains (70-75%) are mostly monocrystalline (there are a few polycrystalline grains) and show a mix of unstrained and strained extinction. Grains are angular to well rounded, but some of the angularity is due to calcite replacement of quartz (some examples shown by white arrows). – Untwinned K-feldspars (5-10%) show varying degrees of sericite alteration. Some appear superficially like lithic grains. – Lithics (I) (~5-8%). Be careful to distinguish actual lithics from altered feldspar grains. – White micas (1-2% – red arrows) show some alteration to chlorite (bluish interference colours). – The framework clasts and micas have a crude alignment along the top left/bottom right diagonal. – Cement is almost entirely coarse calcite. Contact between the calcite cement and framework grains is irregular, indicating silica replacement.

Left plain polarized light. Right Crossed nicols

1. Glauconite peloids up to 0.5mm across (10%). Some glauconite grains are deformed around stronger quartz and feldspar grains (by compaction).
2. Quartz grains (70-75%) are mostly monocrystalline (there are a few polycrystalline grains) and show a mix of unstrained and strained extinction. Grains are angular to well rounded, but some of the angularity is due to calcite replacement of quartz (some examples shown by white arrows).
3. Untwinned K-feldspars (5-10%) show varying degrees of sericite alteration. Some appear superficially like lithic grains.
4. Lithics (I) (~5-8%). Be careful to distinguish actual lithics from altered feldspar grains.
5. Cement is coarse calcite spar. Contact between the calcite cement and some framework grains is irregular, indicating replacement of silica by Calcium carbonate.

 

Lithic arenite (litharenite)

Monster Fm. Upper Cretaceous, Yukon.

Lithic arenite

Lithic arenite. Left plain polarized light. Right Crossed nicols

  1. Moderately well sorted, angular to subrounded grains; clast-supported framework.
  2. 85-90% lithic grains (ch), <5% monocrystalline quartz (m) and plagioclase (pl), and about 10% matrix.
  3. In PPL the lighter coloured clasts are mostly chert (ch) and muddy chert. The darker brownish grains are carbonaceous (cs).
  4. Ductile (indentations and bending) and brittle deformation (fractures) at many grain boundaries caused by compaction (white arrows).
  5. The distinction between diagenetic clay cement and matrix is difficult. Calcite is rare.
  6. Provenance. Primarily siliciclastic mudstone, cherty mudstone, carbonaceous or coal-bearing sources. The abundance of lithics, that are mechanically weaker than quartz and feldspar, suggests probably first cycle sediment.
  7. Paragenesis. Deposition → compaction (continuous) → precipitation of clay cements.

 

Chert lithic arenite (litharenite)

Cadomin, Lower Cretaceous, Alberta foreland basin.

Litharenite - lithic arenite

Litharenite – lithic arenite. Left plain polarized light. Right Crossed nicols

  1. Moderately well sorted, angular to subrounded, clast-supported framework.
  2. 65-70% chert and muddy chert (ch), 10-12% carbonaceous siltstone (cs), 10% quartz (m,p), 10-12% matrix, 1-2% feldspar (k).
  3. Chert grains show varying degrees of recrystallization (cx); some will be inherited from the source rock.
  4. The monocrystalline quartz grain marked m’ shows a relic grain boundary (arrow), and syntaxial overgrowth that is corroded on its outer edge. This was probably inherited from the source.
  5. The distinction between diagenetic clay cement and matrix is difficult at this scale of magnification.
  6. Some ductile deformation at grain-to-grain boundaries caused by compaction – note the indentation of chert by a quartz grain (white arrow, top left).
  7. Provenance. Recycled sedimentary; may include chert-bearing limestones where the carbonate fraction was winnowed or removed during sediment transport.
  8. Paragenesis. Deposition → compaction, loss of porosity, some ductile deformation (continuous) → +/- chert recrystallization → precipitation of clay(?) cements.

 

Sublitharenite

Proterozoic Altyn Fm, southern Alberta

Recrystallized chert in a sublitharenite

Recrystallized chert fragment in a sublitharenite. Left plain polarized light. Right Crossed nicols

  1. The main features in this image pair are the chert clasts (ch). The pebble fragment (right) shows varying degrees of recrystallization. There is a noticeable band of recrystallized chert that includes some radial crystal clusters (cx); under crossed nicols these show a characteristic extinction cross that rotates with the microscope stage. The crystal band terminates at the clast margin. There are a few other chert and muddy chert clasts (dirty appearance) that do not show any significant recrystallization. Together, these features suggest that the recrystallization was inherited from the source rock, rather than being a diagenetic product in the sandstone.
  2. Most quartz grains are monocrystalline (m); the single potassium feldspar has gridiron twinning (k).
  3. Calcite cement (cc) fills most of the pore space; easily identified because of its high birefringence, high relief, and prominent cleavage. The corrugated contact between the quartz grains and calcite is a result of incipient dissolution and replacement of silica.
  4. Paragenesis. Deposition → compaction, loss of porosity (continuous) → +/- clay precipitation → precipitation of pore-filling calcite cement and incipient dissolution of silica.

Other posts relevant to this article

The mineralogy of sandstones: porosity and permeability

The mineralogy of sandstones: Quartz grains

The mineralogy of sandstones: Feldspar grains

The mineralogy of sandstones: lithic fragments

The mineralogy of sandstones: matrix and cement

The provenance of detrital zircon

The provenance of sandstones

Provenance and plate tectonics

Greywackes in thin section

Ignimbrites in outcrop and thin section

Lithic grains in thin section – avoiding ambiguity

 

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Glossary: Geochemistry and diagenesis

Facebooktwitterlinkedininstagram

Abiotic: Physical and chemical conditions not directly associated with life forms, but interact with biotic conditions to form ecosystems. For example, salinity, pH, temperature, precipitation. The term includes organic compounds present in abiotic conditions such as comets. Cf. prebiotic.

Acid: A substance that releases or donates a proton when dissolved in water. The proton is a hydrogen ion that in solution associates with an H20  molecule to form H30+ , but is usually written as H+ . Acids react with bases (bases contain hydroxyl ions – OH ). Water may act as an acid or a base. Solutions with excess H+ are acidic, such that pH < 7.

Activity (geochemical): Sometimes referred to as effective concentration. The activity of an ion is the ratio of its concentration versus some standard concentration and is therefore dimensionless (unlike concentration). The ratio is calculated using an activity coefficient. It is used in equilibria because it expresses the amount of an anion or cation that is available for reaction; compare concentration that measures the total amount of an ion. In a solution like sea water there are many different cations and anions, all reacting to collisions of various kinds. For example, the CO32- anion may collide with cations other than Ca2+ (Na+, Mg2+, K+ and so on), such that the amount of CO32- available to react with Ca2+ is less than the measured concentration. In other words, the amount of CO32- available in real solutions depends not just on its overall concentration, but also on its environment. For this reason, it is preferable to use activities in thermodynamic calculations, such as equilibrium constants. The activity of solids is usually taken as 1.

Activity coefficient:  The activity coefficient (γ) for a specific ion species is related to the degree of ionic interaction with other species in solution. For dilute solutions γ approaches 1 because there are few ion interactions (γ is dimensionless). Thus, the γ value for HCO3 in fresh river water averages about 0.95, but in sea water is much lower (0.57) because of ionic interactions. Activity (a) is calculated for specific ions from the relationship:

a = γ m where m is concentration.

 

Aeolianite: Dune sands cemented by calcite are an example of shallow meteoric-vadose zone diagenesis. Dune sand mineralogy may be siliciclastic or bioclastic, or a mix of both. Most common in subtropical to tropical coastal dunes.

Aerobic conditions: Reactions that directly utilise available oxygen, the most obvious being respiration in life forms, where oxygen is used in metabolic reactions to generate energy (e.g., from food). In sediments this generally is associated with the metabolic activity of microbes – a distinction is made between these types of reaction and oxidation reactions that do not require intermediary metabolic activity in life forms. Cf. Anaerobic conditions.

Alizarin Red-S: This is a soluble organic acid that reacts with calcium. Distinguish between calcite (stains pink-red) and dolomite (no stain) can be easily done using this stain, on rock slabs or thin sections.

Alkalinity: Alkalinity is a measure of the amount of acid that can be added to an aqueous solution without causing significant changes to the pH; also referred to as the acid neutralizing capacity or buffering capacity. The total alkalinity of seawater is primarily determined by the major anions:

mHCO3 + 2mCO32- + minor constituents like borate, phosphate, and silicate anions.

 

Anaerobic conditions: conditions where metabolic reactions in life forms do not require molecular oxygen. In sediments, such reactions are commonly generated by microbes that reduce oxygen-bearing compounds like sulphate (to sulphide), nitrate (to nitrite or ammonia), and carbon dioxide to methane. Sediment where these conditions persist tend to be green-black, and may have mineral sulphides (e.g., iron, manganese). One example is the sediment beneath wetlands, including marginal marine mangrove wetlands. Cf. aerobic conditions.

Anoxic conditions: Usually applied to aqueous environments (water masses as well as connate water) where there is none, or insufficient dissolved oxygen for respiration; usually measured at less than 0.5 ml/L. Under these conditions, the sources of oxygen via bacterial reduction are from nitrates and sulphates. Once these sources are depleted carbon dioxide becomes an important source during reduction to methane. Deep waters in lakes where there is no turnover of the water mass, can become anoxic. Anoxia are also implicated in some of Earth’s major extinctions, such as the Late Permian – Triassic event. Early Precambrian oceans and lakes were probably anoxic. e.g., März and Brumsack, 2015.

Base: A base is a substance that gains a proton in aqueous solution. This can be written in a generalized way as  H+ + OH = H20.  Water can act as a base or an acid. Solutions with excess OH are basic with pH > 7.

Botryoidal cement: In limestones, this cement form is presented as radial clusters of fibrous or bladed calcite or aragonite that precipitate in more cavernous porosity. Common examples are found in reef frameworks, and fenestrae that form by mineral dissolution, gas bubbles, and crystal expansion (e.g. halite-gypsum crystal growth in sabkhas). Fenestrae are common in some cryptalgal laminates and mud mounds containing Stromatactis.

Brines: Generally used for natural waters more saline than seawater. The main dissolved salt is sodium chloride (NaCl), but calcium and magnesium sulphates are also important constituents, and there are several important trace elements, such as lithium. The primary mechanism for brine concentration in ocean basins and saline lakes is evaporation. The saturation level for NaCl is about 357 ppt (normal seawater is 32 ppt).

Calcite compensation depth (CCD): As ocean water depths increase, the partial pressure of CO2 increases and the temperature decreases – in both cases CaCO3 becomes increasingly soluble. An important consequence of this convergence is a decrease in CaCO3 saturation to the point where calcite and aragonite begin to dissolve. For calcite, the depths range from about 4.6 to 5.1 km. Aragonite is more soluble and the ACD depths are about 3 km. This means that the sea floor at or below these depth limits will tend to be devoid of calcareous sediment (particularly microfossils like foraminifera and coccoliths).

Calcite divide (geochemistry):  The stage during evaporation of brines where calcite precipitation determines the succession of minerals in waters subsequently depleted in Ca2+ and CO32-. It determines whether the brine subsequently evolves as HCO3 rich or  HCO3 poor.

Caliche: Also called calcrete. Soil horizons in which carbonate precipitation results in a hardened crust. They develop in regions in which evaporation exceed precipitation, where periods of wetting alternate with drying. Thus, carbonate textures commonly show evidence of dissolution and reprecipitation. A common product is vadose pisoids that also show evidence of multiple episodes of dissolution and precipitation. They can develop in alluvial-lacustrine and intertidal-supratidal settings.

Capillary zone:  In hydrogeology, also called the capillary fringe.  It is a relatively narrow interval above the watertable where surface tension forces on aquifer materials cause water to rise and partly fill pore spaces. The capillary fringe is part of the unsaturated, or vadose zone.

Carbonates: The most diverse group of sediments and sedimentary rocks, usually presented as limestones and dolostones. Carbonate precipitation (and dissolution) is based on the chemical equilibria involving CO2, HCO3, CO32-, and H2CO3. Their primary mineralogy includes calcite and aragonite polymorphs (CaCO3), and dolomite (Ca.Mg [CO3]2). Carbonate formation at Earth’s surface is intimately associated with biological production where precipitation is either induced directly by organisms, or indirectly promoted by the activity and metabolism of organisms. Organisms involved in carbonate production range from microbial to large invertebrates.

Carbonic acid: A weak acid that forms naturally from the reaction:

CO2 + H2O = H2CO3

It is the primary cause of slight acidity of rain (pH 5.5 to 5.8). It is an important component in the series of carbonate equilibria, particularly for pH buffering.

Chalcedony: A fibrous form of microcrystalline quartz, or chert. It commonly form radial clusters. Under crossed polars, extinction patterns are sweeping or radial.

Chemical equilibria: Chemical reactions normally written with the reactants on the left and products on the right. The two are separated by either:

  • An equal sign indicating equilibrium, where forward reactions (to the right) equal reverse reactions, or
  • By two opposing arrows that indicate forward and reverse reactions.

Equilibria should be charge and mass balanced. The quantities of reactants and products are written as concentrations or activities.

Chemical equilibrium: At equilibrium there is no net gain or loss of reactants (by convention, the left side of the equation) or products and no net change in energy. Note that this does not mean the system is static – even at equilibrium there are still collisions between ions (all reactions in solution involve collisions), but collisions on the left equal those on the right side of the equation.

Chemical facies: (hydrogeology) This is a useful concept to demonstrate the chemistry of groundwater in relation to aquifer rock-sediment composition, and the evolution of groundwater chemistry as it flows from one rock type to another. For example, flow from sandstone to limestone aquifers will be accompanied by a change in HCO3 and pH, plus the concentrations of cations like calcium and magnesium.

Chemical kinetics: Also called Reaction kinetics. This is the study of reaction rates and reaction pathways, and hence is distinct from thermodynamics that deals with energy transfer during reactions and is independent of rate. Kinetics is a measure of the rate of change (in concentration or activity) of both reactants and products, in reversible and irreversible reactions. It is particularly important in reactions that are slow relative to mass/solute transport. A good example if these conditions is the precipitation of dolomite under surface conditions – the reaction is thermodynamically favoured, but kinetically is very slow. Kinetics is related to thermodynamics in terms of equilibrium constants, the activation energy of reactions (i.e. Gibbs free energy), and temperature. As a general rule, the rate of 1st-order reactions doubles for every 10º increase in temperature.

Chemocline: A boundary within a water column at which there is a fairly abrupt change in chemical gradient. Examples include the boundary between fresh water and seawater, or changes in REDOX conditions, from oxidation to reducing.

Chemotroph: Organisms that obtain their metabolic energy and synthesize biomass (such as carbohydrates) from reduced elements like sulphur, sulphide, and ferrous iron, instead of sunlight.

Chlorite: has low birefringence, in varying shades of green (in PPL), and crystal habit that is also variable, from fibrous, spherulitic or vermiform (worm-like). May be pleochroic in shades of green and yellow. It is commonly associated with low grade metamorphism and hydrothermal alteration. In greywackes and other mud rocks it is a common replacement for clay matrix, micas, and  ferromagnesian minerals.

Clathrate: A general term for gas molecules that become trapped in an ice crystal cage. There are no chemical bonds between the gas and water ice and the gas can be released upon melting. Also called gas hydrates. Vast amounts of methane are trapped this way beneath the sea floor and in permafrost.

Cleavage: A plane of weakness within a crystal that will break with relative ease. It is caused by weak bonds between planes of atoms within a crystal lattice; the pattern of weakness repeats regularly through a crystal. Some minerals have poor or no cleavage (e.g., quartz, olivine); others have excellent cleavage along several lattice planes (e.g., calcite, feldspar). Cleavage is a defining characteristic of a mineral, particularly in thin section.

Compaction:  The process where sediment particles, once deposited, are pushed closer together to form a more tightly knit framework. Compaction begins almost immediately following deposition and continues during sediment burial. The normal compressive stress in this case is applied by the overlying sediment. Because porosity is also reduced, an additional requirement for compaction to take place is the release of interstitial water through aquifers. If fluid cannot escape (for example because of permeability barriers) then the rock body will not compact, and internal fluid pressures will rise – this is called overpressure. Mudrocks can compact to less than a tenth their depositional thickness. More rigid frameworks like sandstones compact far less. See also pressure solution, lithic fragments.

Crystallographic axes: Three or four axes about which a crystal can be rotated through 360o.  The axes intersect at a single point (the centre of symmetry). They are labelled according to their lengths. If axes are the same length, then they are referred to as a1, a2, a3 etc. If they have different lengths, they are labelled a, b, and c. Thus, in the cubic (isometric) crystal system they are labelled a1, a2, a3, and in the tetragonal system a1, a2, c. The hexagonal system is the only one with four axes. Angles between axes are labelled α, β, γ.

Crystal symmetry: Symmetry describes the shape of an object and can be represented both mathematically and visually. In crystallography, the two most useful forms of symmetry are (mainly because they are the easiest to visualize):

  1. Axes of rotation (crystallographic axes) where a particular crystal face will be repeated during rotation through 360o. The number of repetitions for a 360o rotation can be 2, 3, 4, or 6, that are referred to as two-fold, three-fold, four-fold, and six-fold (axial) symmetry respectively.
  2. Planes of symmetry where two parts of a crystal are mirror images. For an analogy, think of this concept in terms of the common bilateral symmetry in many living organisms, such as people, and many classes of mollusc. Note that planes of symmetry are NOT the same as twin planes.
  3. Additional elements of symmetry include: A centre of symmetry, where a crystal face is reflected from one side to another or is repeated by inversion, and an axis of rotary inversion.

Crystal systems: There are 6 crystal systems based on combination of the elements of symmetry; a seventh system – trigonal – is usually considered a subclass of the hexagonal system. There are 32 crystal classes based on combinations of the symmetry elements. The defining criteria are axial lengths, the angles between axes, and axial symmetry (the number of repetitions about an axis).

Cubic (or Isometric) crystal system: The most symmetric group. All three axes are the same length and are at right angles to each other.

a1 = a2 = a3                  α = β = γ = 90o

2, 3, and 4-fold symmetry depending on the class

Common crystal forms: cubes, octahedra, dodecahedra.e.g., Halite, pyrite, fluorite, garnet

Diagenesis:  The sum of physical and chemical processes in sediment, beginning soon after deposition at or immediately below the sediment-water interface, and continuing at depth in concert with increased burial temperatures, lithostatic and hydrostatic pressures, and changing fluid composition.

Dissequilibrium compaction:  Under normal conditions of compaction, fluid that is driven from pore spaces escapes without a significant increase in pore pressure – i.e. hydrostatic conditions prevail.  However, rapid deposition of low permeability deposits can impede fluid flow and under these conditions pore pressures increase; this process is called disequilibrium compaction. In many basins, this occurs at about 3km burial depths. Disequilibrium compaction is enhanced by cementation and tectonic compression.

Dispersion: In geofluids this is the process where dissolved and insoluble compounds move from their source or point of origin; observed in groundwater flow, diagenesis, and metamorphism. In these contexts there are two primary mechanisms – mechanical dispersion, and molecular diffusion.

Drusy cement: Cements consisting of calcite rhomb mosaics that line and fill pores, intraskeletal chambers, and more cavernous porosity. The size of calcite rhombs commonly increases towards the center of void spaces. Intercrystalline boundaries tend to be planar. They are common in meteoric and burial environments where they may overlie earlier fibrous or bladed cements.

Equilibrium constant: For a specific reaction, equilibrium constants are the ratio of product activities (or concentrations) divided by reactant activities; they can be determined experimentally (assuming a reaction is at equilibrium) or using thermodynamic considerations (where activities must be used). The general expression for a reaction involving ionic species in solution is:      aA + bB ↔ cC + dD, where a, b, c, and d are the stoichiometric values for each ion (e.g. 2H+).

K = cC + dD/ aA + bB at equilibrium.

In a real aqueous solution, we can determine whether a reaction will proceed to the left or right: if  cC + dD/ aA + bB is <K the reactants will convert to products (the reaction goes to the right. The opposite occurs if cC + dD/ aA + bB >K.

K is strongly dependent on temperature and pressure.

 

Euxinic conditions: Ocean waters that are depleted in dissolved oxygen (anoxic) and are sulphidic. The sulphide is primarily dissolved H2S. Euxinia can occur in highly stratified water bodies, such as lakes and enclosed seas where there may be an the anoxic layer occurs beneath shallower waters with varying amounts of dissolved oxygen. However, euxinia may also have occurred in larger oceanic water masses in the geological past.

Evaporative pumping:  In arid regions, intense evaporation at the surface creates a hydraulic gradient in shallow subsurface aquifers, inducing lateral groundwater and/or seawater flow to replace lost fluid. Vertical capillary flow through the unsaturated zone (above the watertable) transfers these saline fluids from the aquifer to the surface.

Ferric (iron): Fe3+, or Iron III. is the common oxidized state of iron. It is the primary form of iron in limonite (FeO(OH)·nH2O) and hematite (Fe2O3). Magnetite (Fe2+ Fe3+2 O3 contains both iron II and iron 111. The oxidised state produces the red colouration in red beds and red shales.

Ferrous (iron): Fe2+, or Iron II. This is the common reduced state of iron in aqueous solution and common minerals like siderite (FeCO3), iron sulphate (FeSO42-), iron sulphide (FeS), and pyrite (FeS2). It combines with iron III in magnetite, and substitutes for calcium (Ca2+) in ferroan calcite, and for magnesium in ferroan dolomite. Iron II is largely responsible for the greenish hues of reduced shales.

Geothermal gradient: Temperature generally increases with depth in the crust; the gradient for a particular location is stated as the temperature increase per unit depth. The global average is 3o C/ 100 m although there can be large departures from these values in regions of geothermal and volcanic activity, or regions that have cooled significantly over geological time, such as old oceanic crust.

Goldschmidt classification: The grouping of elements according to their place in the periodic table and their preferred mineral-forming phases. The four main groups are:

  • Lithophile elements – those that bond readily with oxygen; tend to concentrate in the crust: Al, At, B, Ba, Be, Br, Ca, Cl, Cr, Cs, F, I, Hf, K, Li, Mg, Na, Nb, O, P, Rb, Sc, Si, Sr, Ta, Th, Ti, U, V, Y, Zr, W, plus the Lanthanides.
  • Siderophile elements – iron-loving, mostly avoid oxygen, concentrated in the core and mantle: Au, Co, Fe, Ir, Mn, Mo, Ni, Os, Pd, Pt, Re, Rh, Ru.
  • Chalcophile elements – bond with sulphur to form insoluble sulphides – low affinity for oxygen. The elements: Ag, As, Bi, Cd, Cu, Ga, Ge, Hg, In, Pb, Po, S, Sb, Se, Sn, Te, Tl, Zn
  • Atmophile elements – H, C, N, noble gases: mostly form gases.

Greenhouse effect: The heating of an atmosphere when gas molecules absorb certain frequencies of solar infrared energy. On Earth this involves water vapour, carbon dioxide, methane, and to a lesser extent nitrous oxide. Molecular oxygen and nitrogen do not absorb infrared energy. Carbon dioxide and water vapour absorb energy at different frequencies. Note that the amount of water vapour in the atmosphere depends on temperature, unlike carbon dioxide.

Groundwater residence time: The time from recharge (usually at the surface) to discharge. Residence times are briefest in unconfined aquifers, ranging from days to years. In regional groundwater flow systems these times are measured in 105 to 106 years. Groundwater dating utilises trace compounds such as fluorocarbons, isotopes like ³H (tritium from atmospheric atomic device testing), and cosmogenic isotopes such as Carbon-14, and Beryllium-10.

Gypsum divide: The stage during evaporation of brines where gypsum precipitation determines the succession of minerals in waters subsequently depleted in Ca2+ and SO42-.  It determines whether the brines evolve as SO4 rich – Ca poor, or SO4 poor.

Halides: Anions of the Chemical Periodic Table halogen group: Fluoride F‾, chloride Cl‾, bromide Br‾, Iodide I‾, and astatide At‾. Many inorganic halides are water-soluble; most organic halides are not.

Hexagonal crystal system: This system has 4 axes, 3 of which are perpendicular to c axis.

a1 = a2 = a3 ≠ c             Angles between a1 = a2 = a3 = 120o

6-fold symmetry. Common crystal forms: Prisms, bipyramids. e.g., apatite, beryl. The Trigonal subsystem has one 3-fold axis or rotation. Three important examples are quartz, calcite and dolomite, commonly formed as bipyramids, rhombohedra, and scalenohedra.

Holomict: Lakes or seas in which there is mixing of surface and deeper waters. Bottom waters tend to be oxygenated Cf. Meromict.

Hydrolysis: Also called dissociation. The reversible reaction where H20 splits into a hydrogen ion and a hydroxyl ion, as in H20 = H+ + OH. The equilibrium constant is written as:

Kw = (H+).(OH)/( H20). The activity of H20 is usually taken as 1.0, so that Kw = (H+).(OH). At 25ºC K= 10-14.0 . Where the concentration, or activity of (H+) > (OH) is acidic, and (H+) < (OH) is basic. This is the basis for the pH scale, calculated as the -log10  of the activities.

Hypersaline: Having salinity greater than seawater (>35 parts/1000). Modern hypersaline environments are most common between the tropics but are found in such diverse places as the Antarctic dry valleys. Plant and animal life require specialized adaptations to survive these conditions. Prolonged hypersalinity may result in evaporite deposits in lakes and seas.

Karst:  A landscape of gullies, canyons, and steep-sided pinnacles resulting from intense meteoric diagenesis (dissolution) of thick limestones. The relief on karst landforms ranges from 1-2 m to 100s of metres. The corresponding subterranean structures include sinkholes, caverns and underground streams.

Kerogen: Kerogens are complex organic polymers that form during the breakdown of organic matter during the early stages of sediment burial. Three main types are identified depending on the O/C and H/C ratios of the polymer molecules: Type 1 is derived from algal organic matter, Type II from mainly marine micro-organisms, and Type III from plant material. Kerogen itself begins to break down at temperatures around 60o-80oC, as part of the organic diagenetic-maturation process. Identification of the kerogen types preserved in hydrocarbon deposits provides a good indication of the original organic matter.

Lithification: The combination of compaction and cementation that produces hard, hammer-ringing rock from loose, uncompacted sediment. Lithification depends on a complex association of physical and chemical processes. Cementation can occur at very shallow depths in the case of carbonates, or at different stages of burial depending on temperature, and rock – fluid chemistry. Compaction begins soon after deposition and continues at depth.

Lithophile elements: One of the Goldschmidt classification groups of elements that readily bond with oxygen. This means they tend to be concentrated in Earth’s crust and probably the crusts of other rocky planets – common examples are Na, Ca, Mg, Si, Al, K. as well as some of the transition metal elements like Fe, Mn. cf. siderophiles.

Lysocline: The ocean water depth where the dissolution of calcite is first observed in sediment. Its identification requires detailed observation of dissolution textures and is somewhat subjective. It lies above the calcite and aragonite compensation depths; the lysocline should, theoretically, be close to the saturation levels for both minerals.

Magnesium calcite: Also called magnesian calcite. In the calcite crystal lattice, magnesium can occupy the position of calcium, up to about 20 mole percent. Two varieties predominate in carbonate sediments and limestones: Low magnesium calcites (LMC) with <4 mole % Mg), and high magnesium calcites (HMC) with 11-19 mole % Mg). HMC commonly recrystallize to LMC during burial diagenesis.

Mechanical dispersion: In geofluids, this occurs when solute molecules are carried from the source by local eddies around grains or through fractures; this kind of tortuosity takes place at a scale much smaller than the en masse advective flow.  Cf. Molecular diffusion.

Meromict: A stratified lake or enclosed sea where the layers do not mix. Bottom water layers may become anoxic as dissolved oxygen is used up by organisms. In saline waters it applies to salt crystals that precipitate within saturated layers and then sink to the bottom.

Meteoric diagenesis (carbonates): Diagenesis of limestone under fresh-water conditions, both in the vadose (unsaturated) zone, and below the watertable. It is largely controlled by the degree of fresh- water seepage and groundwater flow. Vadose zone diagenesis is dominated by dissolution that, if prolonged, produces caverns, sinkholes (dolines), subterranean streams, and spectacular karst landforms. Dissolved calcium carbonate may reprecipitate as cement and fracture-fill in the saturated zone, and as stalactites-stalagmites in caves.

Molecular diffusion: When a solute gradually mixes with solvent molecules; in geofluids this is primarily water. The process does not involve the physical flow of water, but depends on solvent-solute properties such as polarity and charge, and vibration energies. Cf. Mechanical dispersion.

Monoclinic crystal system:  a ≠ b ≠ c                      α = γ = 90o, β ≠ 90o

2-fold symmetry.Common crystal forms: Prisms, pinacoids (flattened prisms).e.g., orthoclase, diopside, sphene, staurolite, most amphiboles.

Neomorphism: Defined by R. Folk in 1965 as the transformation between one mineral and itself or a polymorph. In other words, neomorphism is a product of recrystallisation where the bulk composition does not change, only the size and/or shape of crystals. It is common in carbonate lithologies and involves recrystallisation of both framework clasts and cements. As such it tends to cross-cut pre-existing textures and fabrics; relict textures may be preserved. Aggrading neomorphism is common in micrites where crystals increase in size in a more-or-less radial fashion.

Nitrogen cycle: The natural transfer of nitrogen and nitrogen compounds from air to soils, vegetation, water and back to the atmosphere. The natural cycle is complicated by human interventions via fertilizers (nitrates) and industrial nitrogen oxides that saturate soils and leach into shallow groundwater and surface waters. Most of the natural nitrogen fixing is done by microbes.

Nitrogen fixing: This is an important process for plant uptake of nitrogen. Plants do not get their fill of nitrogen from the air, but from soil and plant microbes (fungi, bacteria) that convert molecular nitrogen in air (N2) to water soluble compounds, principally nitrates (NO3 ). Plants utilize this soluble form, taking it up via their roots.

Oil generation window: The temperature range 80° – 120°C where hydrocarbon maturation to liquid oil from  sedimentary organic carbon, is most rapid and most productive. At an average geothermal gradient of 30°C/km, the top of the window occurs at depths of about 3 km. Organic matter subjected to temperatures >120°C is prone to gas formation.

Oil migration: Hydrocarbon production in deeply buried sediments, begins in organic-rich sediment, such as oil shale. Once formed (by a series of complex chemical reactions), the oil (and gas) migrate from the shale or mudstone to more porous and permeable rocks such as sandstones and limestones. Migration is driven buoyancy forces and the flow of deep subsurface groundwater. Migration will continue until the oil is trapped (resulting in an oil field). Oil and gas that isn’t trapped will eventually find its way to the surface or sea floor and escape.

Oil seep: Oil, sometimes accompanied by gases like methane, that leak to the surface via fractures or faults, driven of buoyancy forces, or as a part of spring flow.  The hydrocarbons may be sourced from oil-prone porous rock, or from actual subsurface oil pools.

Orthorhombic crystal system:  a ≠ b ≠ c                      α = β = γ = 90o

2-fold symmetry. Common crystal forms: Prisms, bipyramids.e.g., olivine, cordierite, hypersthene

Oxidation: The process where an atom provides electrons to another atom of a different element; and oxidized element has lost electrons. Oxidation always occurs with reduction (REDOX reactions). An oxidized element (atom) is capable of gaining electrons, in which case it becomes reduced; the initial oxidized element is referred to as a reducing agent. Thus Fe2+  is more reduced than Fe3+ ; in the mineral pyrite FeS2  iron is in the 2+ state and sulphur -1 state.

Ozone: When oxygen molecules (O2) in the stratosphere are bombarded with high energy ultraviolet light (UV) the molecule splits into two oxygen atoms. Each of these atoms in turn reacts with O2  to produce ozone, or O3.  Ozone is responsible for absorbing some of the harmful UV radiation that would otherwise reach the surface of the Earth.

Paleothermometer: Geological, paleontological and chemical tools used to determine the temperature conditions and thermal history of ancient environments, and more deep-seated processes associated with sedimentary basins, igneous and metamorphic events. They are components of rocks such as minerals, isotopes, fossils, and fluids that provide us with either a direct measure or proxies of paleotemperatures. Common examples include vitrinite reflectance of coals, fossil colour, radiogenic blocking temperatures, stable isotopes of oxygen and carbon, fission tracks, and fluid inclusions.

Paragenetic sequence: In sedimentary petrology, the sequence of mineral components precipitated (and dissolved) during diagenesis. Sequential changes in mineral composition and/or crystallographic form reflect evolving fluid compositions, fluid flow, burial temperatures, and compaction. It is analogous to cement stratigraphy.

Pendant cement: Stalactite-like cements that accumulate on the low point of grains during gravity drainage of interstitial fluid. They are common in carbonates subjected to vadose zone diagenesis.

pH: Literally the ‘potential of hydrogen’, is a measure of the acidity or alkalinity of an aqueous solution. It is expressed as:

pH = -Log10 (aH+) where aH+ is the activity of H+ in solution.

This means that high concentrations of H+ have low pH values. The pH range is 0 to 14; a neutral solution has pH = 7. An acidic solution has a pH <7.0; an alkaline solution >7.0. Pure water at 25oC has a pH of 7; rain a pH of 5.0 to 5.5 (i.e. slightly acidic because of dissolved CO2), and seawater 7.5 to 8.1. The variations are partly dependent on temperature and its influence on the carbonate equilibria.

pH buffering: Carbonate equilibria do not operate in isolation. If the amount of dissolved CO2(aq) is increased this does not mean that the amount of H+(aq) will increase by the same amount because some of the CO2 forms H2CO3 (aq), some HCO3(aq), and some CO32- (aq), such that the amount of H+ added is small. In other words, the cascade of equilibria acts to buffer the system against large changes in pH.

Phase diagram: The graphical representation of different states for a compound, as solid, liquid, or gas. The phase diagram for water is plotted as pressure against temperature; the triple point where all three phases coexist is at 0.01oC and 608 pascals (0.006 atmospheres). For carbon dioxide the diagram also shows gas, solid and liquid phases, plus a supercritical liquid phase.

Photosynthesis: A process that converts sunlight energy to chemical energy in plants, cyanobacteria, and algae. One of the chemical products is molecular oxygen(O2), that in plants is formed from carbon dioxide reacting with water in plant cells to produce sugars and oxygen. It is generally understood that most of Earth’s free oxygen was produced during the Precambrian by cyanobacterial stromatolites.

Photic zone: The uppermost layer of the oceans and lakes where light penetrates; the base of the zone is at about 1% of incident sunlight. On average it is about 200 m deep. It is the layer where more than 95% of photosynthesis by marine organisms takes place.

Ppb: Parts per billion

Ppm: The abbreviation for parts per million. For water this equates to 1mg/Litre.

Ppt: The abbreviation for parts per thousand. Also written as ‰.

 

Piper diagram: A matrix of three triangular plots that map the chemical compositions of water. It is based on normalized percentages of major cations (Calcium, magnesium, potassium, and sodium), and carbonate-bicarbonate, sulphate, and chloride anions. It is useful for tracking the source of groundwater flows in aquifers derived from different rock types, and the evolution of chemical speciation.

Pressure solution: The dissolution of rock components (framework clasts and cements) as a result of differential compressive stress. Common products of pressure solution are stylolites. Conditions required for dissolution to take place are:

  • Differential compressive stresses develop at intergranular contacts,
  • Interstitial fluids must be undersaturated with respect to the mineral phase under stress,
  • Dissolved components are transported from the grain contacts to regions of lower compressive stress; this requires efficient fluid movement, and
  • The solute reprecipitates some distance from its point of origin.

Reaction kinetics: See Chemical kinetics.

 

Recrystallisation:  In sedimentary rocks this involves the transformation or replacement of a mineral with itself, and usually entails changes in crystal size and shape (but not bulk composition): as in micritic calcite to sparry calcite, or aragonite to its polymorph calcite. The term was originally coined for the process of annealing in metals, which is a dry process. Recrystallisation in sedimentary rocks is always a wet process that involves dissolution of a mineral at grain boundaries, followed by precipitation. It tends to cross-cut original textures, destroying them in the process.

REDOXReactions in which oxidizing and reducing agents combine; thus one atom is oxidized and the other reduced simultaneously. For example, in the sour, toxic gas hydrogen sulphide (H2S), 2 H atoms lose an electron each to the sulphur atom; 2H+ S2-.

Reduction: When an atom gains electrons it becomes reduced. It has electrons to spare and can donate them to the atom of another element that has a deficit of electrons (i.e. it is oxidized). A reduced element that donates electrons is a reducing agent.  Cf. Oxidation, REDOX.

Saline lake brines:  Unlike seawater, terrestrial brines have widely variable compositions, depending on local soil and bedrock compositions, groundwater chemistry, and the degree of evaporitic drawdown. Typical brines contain Na+, Ca2+, Mg2+, Cl, SO42-, HCO3, CO32-, and SiO2, but concentrations are highly variable. pH ranges from highly alkaline to highly acidic. Evaporation pathways produce a succession of different minerals. See also calcite-gypsum divides.

Saturation: Saturation (Ω) is the ratio of the measured ion activity (or concentration) product and the standard solubility product (Ksp) for a mineral. If Ω >1 then the solution is supersaturated with respect to the mineral; if Ω <1 then it is undersaturated and the mineral will dissolve. If Ω = 1 then the mineral is at equilibrium with the solution.

Saturation depth: In ocean chemistry this boundary identifies when seawater becomes unsaturated with respect to calcite (or aragonite). The saturation depth is determined by comparing the measured solubility product of either the activity or concentrations of Ca2+ and CO32- in seawater samples, with the equilibrium solubility product at the same temperature and water pressure.

Secondary porosity: Porosity that is created during burial diagenesis by the dissolution of chemically reactive grains such as carbonates and feldspars. Secondary porosity can enhance the overall porosity of a rock, particularly if primary intergranular pore volumes have been occluded by cements. Secondary pores may be larger than those formed during deposition, where entire grains are dissolved. Partial dissolution along twin or cleavage planes in minerals like feldspar, will result in irregular grain boundaries.

Sequestration: Storage of solid, liquid or gas so that it cannot disperse, or escape. Of recent concern is sequestration of carbon in various forms, particularly CO2 and methane. Natural sequestration occurs on rocks (coal, limestones), soils, and permafrost. Artificial sequestration of supercooled CO2 in certain rock formations (such as depleted oil fields) is considered as one means of controlling CO2 emissions.

Sericite: A flaky white mica and common alteration product of feldspar. In thin section it usually presents as fine ragged crystals (rather than the more uniform muscovite flakes), concentrated along feldspar cleavage planes, or distributed across the entire crystal. It has high birefringence and appears to sparkle against the dull background of altered grains and matrix.

Siderophile elements: Literally iron-loving elements, they include the high density transition metals that bond with iron in solid and molten states. They can also bond with sulphur and carbon. As such they tend to concentrate in Earth’s core and to a lesser extent the mantle. They re rare in the crust. Most, except for Fe and Mn have a low affinity for oxygen. The list includes Ag, As, Bi, Cd, Cu, Ga, Ge, Hg, In, Pb, Po, S, Sb, Se, Sn, Te, Tl, Zn – Sulphur is also a volatile element and at Earth’s surface combines with oxygen to form sulphate anions.

Solubility product: Solubility product expresses whether a mineral will dissolve or precipitate in aqueous solutions, at specified temperatures and pressures. For example, aragonite in seawater, the reaction is CaCO3(solid) ↔ Ca2+(aq) + CO32-(aq). At equilibrium the solubility product is

Ksp = (aCa2+).(a CO32-) / (a CaCO3 solid)

The activity of the solid calcite is 1, such that the constant at equilibrium becomes:

Ksp = (aCa2+).(a CO32-)

(aCa2+).(a CO32-) is also called the activity product. In real solutions, if (aCa2+).(a CO32-) is >Ksp, then aragonite will precipitate; if <Ksp it will dissolve. See also saturation.

Solute: A chemical compound that has dissolved in a solvent. In geofluids, the solvent is primarily water; common solutes are various chlorides, sulphates, hydroxides, nitrates and phosphates. In all these compounds, the solute will consist of cations and an anions surrounded by water molecules.

Solute transport: The movement or flow of dissolved mass in a fluid, usually water. The primary mechanisms of transport are advective flow and diffusion. Transport is usually accompanied by chemical reactions.

Solvent: A liquid (usually) capable of dissolving and maintaining solutions of solid compounds. Water is the most prominent geofluid solvent. Organic solvents are important for industrial processes.

Stalactite:  Tubes, straws. and threads of calcite that hang from the ceiling in the drip zone of caves. Groundwater, initially undersaturated with respect to calcite can, with sufficient transfer of atmospheric CO2, become supersaturated, promoting precipitation. Pillars or columns form when stalactites meet stalagmites, their cave-floor counterpart. They are a type of speleothem, a group of cave precipitation structures that includes cave wall linings (drapery), flowstone, and cave pearls. Stalactites and stalagmites can also form from dripping lava.

Stalagmite: Commonly conical shaped mounds of calcite that grow from cave floors as a result of the steady drip of seepage groundwater. They are the cousin of stalactites.

Strong acid: See weak acid.

 

Structure grumeleuse: A term introduced by Lucien Cayeux in 1935, refers to clotted limestone textures where isolated, diffuse patches of micrite are surrounded by coarser neomorphic spar; the overall texture appears clotted. At times it can be difficult to distinguish between this recrystallisation texture and primary peloidal limestones.

Stylolite:  Saw-tooth like, discordant seams that signify pressure solution of rock components (framework clasts and cements). They are most common in carbonates but can form in siliciclastic rocks. They represent differential compressive stresses at grain-to-grain contacts, the dissolution and mass transfer of carbonate by diffusion and fluid flow. Stylolites commonly parallel bedding (from normal compressive stress) but also form oblique to bedding.

Tetragonal crystal system: Liken this group to isometric crystals stretched along the c axis.

a1 = a2 ≠ c                    α = β = γ = 90o            Mostly 2- and 4-fold symmetry. Common crystal forms: Prisms, bipyramids with or without prisms. e.g., zircon, chalcopyrite, rutile

Thermocline: The ocean layer extending from about 200m to 1000m depth where the temperature decreases rapidly. Below the thermocline the water temperature varies little from about 4o

Triclinic crystal system:  The least symmetric group. a ≠ b ≠ c                      α ≠ β ≠ γ ≠ 90o

No axes of symmetry!  Common crystal forms: Prisms, bipyramids. e.g., microcline, plagioclase, kyanite

Triple point: On a phase diagram, it is the point in pressure-temperature space where solid, liquid and gas phases of a compound coexist.

Tufa: A natural, surface precipitate of calcium carbonate in alkaline lakes, rivers, springs and geothermal hot pools, promoted by degassing of CO2 as the waters exit to the surface. Degassing of CO2 results in an increase in pH, and concomitant increase in the stability of CO32- and HCO3 aqueous species and the degree of calcite saturation. It is also possible that microbial activity also promotes precipitation. Tufas tend to be highly porous; they can encase dead critters and vegetation. Travertines are a denser form of surface calcite precipitation. Extensive deposits are typically terraced.

Twinning: A symmetrical intergrowth of two separate crystals of the same mineral, that share the same mineral lattice. In thin section under crossed nicols, each twin segment will go into extinction at different rotations of the microscope stage. There are many kinds of twinning. For example, plagioclase may show albite, carlsbad, or pericline twins individually or as combinations in the same crystal. Important optical properties of twins that help mineral identification include extinction angle (whether straight or inclined), and 2V angles. Note that twin planes are NOT the same as planes of crystal symmetry.

Unit cell: At the atomic scale, the arrangement of atoms that represents the fundamental structure of a mineral in crystal form. The crystals we see consist of a three-dimensional array of stacked unit cells. This means that the overall shape of the crystal mimics its unit cell. The simplest unit cell is a cube; cubes of the same size will stack perfectly. Not all polygonal geometries allow such stacking, for example cells with triangular sides will stack neatly together, but those with 5-sided faces (pentagons) will not. Consideration of the unit cells and their symmetry forms the basis for definition of the 6 (or 7) crystal systems.

Vadose zone:  The portion of an unconfined aquifer above the watertable where pore spaces are air-filled (and approximately at atmospheric pressure). It is synonymous with unsaturated zone.

Vitrinite reflectance: Vitrinite is a component of coal that forms by thermal alteration of plant tissues.  The intensity of reflection from a polished surface of vitrinite samples increases with coal rank. The reflectance is measured and compared with standard values d to determine coal rank.

Weak acid – strong acid: A general classification that depends on how easily an acid donates a proton (H+ ) to a water molecule to form H3O+ . A weak acid will partially dissociate (i.e. split into its constituent H+ and anion, leaving some of the acid in solution. All the reactions involving carbonate and carbonic acid are weak acid reactions. Strong acids dissociate completely – they donate all their H+ . Common examples include hydrochloric acid (HCl) and sulphuric acid (H2SO4).

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of carbonates; Pressure solution

Facebooktwitterlinkedininstagram

pressure solution at ooid contacts

Pressure solution and stylolite formation in carbonates.

This is part of the How To…series  on carbonate rocks

Pressure solution takes place in lithified rock when stress at intergranular and intercrystalline boundaries exceeds normal hydrostatic pressure. Differential stress can develop at almost any stage of sediment burial and compaction, during tectonic deformation and metamorphism. This post will describe pressure solution in carbonates, but it is worthwhile remembering that it can occur in many sediment types, as well as deformed rock. For example, cleavage, a penetrative fabric commonly found in folded metamorphosed strata, develops via a combination of shortening, pressure solution of certain mineral components and precipitation of new minerals. It is generally understood that pressure solution will occur if three conditions are met:

  • Differential compressive stresses develop at intergranular contacts.
  • Dissolved components are transported from the grain contacts to regions of lower compressive stress; this requires efficient fluid movement, and
  • The solute reprecipitates some distance from its point of origin.

Diagram showing the prgression of pressure solution and burial compaction

In sedimentary rocks, pressure solution is commonly manifested as:

  • Discontinuous grain-to-grain contacts, where one grain crosscuts or penetrates another grain. Intergranular pressure solution was first described by Henry Clifton Sorby in 1863. Pressure solution in ooid grainstones is a good example because we know the original shape of the grains involved. Intergranular pressure solution can develop between quartz grains but is a slower process than in carbonates because of its much lower solubility.

 

Pressure solution in ooid grainstone; single and compound ooids. Note the interpenetration stylolites.

 

  • One of the earliest descriptions of stylolites was presented by K.F. Klöden, a German ‘naturalist’ who in 1928 provided us with the name stylolite, but interpreted them as fossil traces. We now know that stylolites represent pressure solution across an extensive surface, producing contacts with complex sawtooth and interdigitate geometries. Surfaces can extend laterally many metres; the relief on a stylolite surface ranges from millimetres to 10s of centimetres. Stylolite surface orientation is determined by the principal compressive stress which is close to vertical if the stress is borne of sediment compaction – therefore, surfaces commonly parallel bedding but may diverge from bedding if tectonic deformation is involved. The presence of stylolites indicates dissolution of carbonate and a reduction in stratigraphic thickness; losses of 10% thickness are common.

Stylolite seam in dolostone. There are four additional but more crypticseams

Stress, strain, and fluid flow

During compaction grains become closer packed, reducing porosity and expelling connate fluids. Sediment rigidity increases, aided by cementation. Fluid pressures at this stage are close to hydrostatic (the pressure due to the overlying column of water and air). At some point in this process differential stresses in excess of hydrostatic develop at grain contacts, raising the potential for calcite dissolution. Dissolution will be a maximum at grain contacts where pressure and strain energy are maximized. The ensuing solute concentration gradient between the contact and adjacent pore space will promote diffusion along a thin fluid film between the grains. Note, if solute cannot be transported from the contact then dissolution will cease. The ‘thin film’ hypothesis was first suggested by P.K. Weyl in 1959. It remains a popular mechanism for explaining solute transport by aqueous ion diffusion.

 

Diagram depicting pressure solution, stress, strain and diffusion of solute

Diffusion of solute will increase the activity (or concentration) product in the adjacent pore spaces, promoting reprecipitation of calcite (or dolomite); this may take place in the immediately adjacent pore spaces, or solute may be moved farther afield by advective flow (i.e. fluid transport en masse) and precipitate some distance away. Where precipitation occurs will depend on the efficiency of advective flow and the calcite activity product.

Reprecipitation reduces the effective porosity.

 

The influence of the mechanical properties of grains

One prerequisite for pressure solution to take place is that the grains in contact must be rigid. During compression (compaction) and at the relatively low burial temperatures where pressure solution occurs (less than 200oC) rigid grains will act elastically (unless strain rates are high in which case they behave as brittle materials). Elastic strain energy will contribute, along with compressive stress, to calcite dissolution.

If the materials are ductile then pressure solution is inhibited. Common examples in siliciclastics include lithic grains, and in carbonates non-lithified pellets and mud intraclasts.

As the pressure solution seam increases in length-area, there is a concomitant decrease in effective stress and strain rate. Thus, as seams grow, the tendency for dissolution to decrease acts in concert with a reduction in effective porosity resulting from reprecipitation (i.e. the porosity that permits fluid flow), as a kind of self-regulating process.

 

The influence of insoluble materials

Stylolites are usually outlined by dark seams of clay, organic matter, and iron oxides – it’s what makes them so recognisable. Clays are particularly important here because they enhance the transfer of dissolved mass away from grain contacts. Most of the insoluble residues were originally dispersed through the host rock and concentrated along the seam as carbonate minerals dissolved.

The role of clays has been debated at length (lots of topics in geology are debated at length). Do clays act as passive bystanders, left behind after all the action, or do they participate in the pressure solution process? A model presented by E. Aharonov and R. Katsman (2009), based on numerical simulation (and not experimental data) posits that clays act as a kind of catalyst to pressure solution. In this model, both clays and compressive stress act together to the extent that stylolites will not propagate if clays are not present.

 

The consequences of pressure solution

  • Pressure solution is the geochemical response to differential stress.
  • It is an important part of the diagenesis continuum that produces lithified rock.
  • It results in loss of stratigraphic thickness.
  • It promotes precipitation that reduces porosity.

Links to other posts in this series:

Mineralogy of carbonates; skeletal grains

Mineralogy of carbonates; non-skeletal grains

Mineralogy of carbonates; lime mud

Mineralogy of carbonates; classification

Mineralogy of carbonates; carbonate factories

Mineralogy of carbonates; basic geochemistry

Mineralogy of carbonates; cements

Mineralogy of carbonates; sea floor diagenesis

Mineralogy of carbonates; Beachrock

Mineralogy of carbonates; deep sea diagenesis

Mineralogy of carbonates; meteoric hydrogeology

Mineralogy of carbonates; Karst

Mineralogy of carbonates; Burial diagenesis

Mineralogy of carbonates; Neomorphism

 

References etc.

E. Aharonov and R. Katsman, 2009. Interaction between pressure solution and clays in stylolite development: Insights from modelling. American Journal of Science v. 309, p. 607-632.

Robin G.C. Bathurst, 1976. Carbonate Sediments and their Diagenesis. Elsevier, Developments in Sedimentology, 12. 658p. Particularly Chapter 11

R. Tada and R. Siever, 1989. Pressure solution during diagenesis. Annual Review of Earth and Planetary Sciences, v. 17, p. 89-118. 

H.C. Sorby, 1863. On the direct correlation of mechanical and chemical forces. Proceedings of the Royal society of London, v.12, p. 583-600. Henry Clifton Sorby was one of the architects of modern sedimentology

P.K. Weyl. 1959. Pressure solution and the force of crystallization – a phenomenological theory. Journal of Geophysical Research, v 64, p. 2001-2025.  Weyl promoted the thin fluid film idea in this seminal paper.

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of carbonates; Neomorphism

Facebooktwitterlinkedininstagram

Neomorphic calcite in bivalve

This is part of the How To…series  on carbonate rocks

Diagenesis of carbonate sediment begins soon after deposition and continues at all depths until metamorphic processes takes over (a statement that is about as diffuse as the boundary between diagenesis and metamorphism). Burial can also be expressed as a function of changing temperature, pressure, and fluid composition,  and is manifested as physical and chemical changes such as:

  • compaction and pressure solution.
  • replacement of metastable and unstable minerals like aragonite,  high-Mg calcite, and evaporites by calcite,
  • dolomitization,
  • creation of secondary porosity, and
  • neomorphism – recrystallization.

Continue reading

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of carbonates; Burial diagenesis

Facebooktwitterlinkedininstagram
Variables of temperature, pH, and organic solvents in sandstone-limestone burial diagenesis. From Surdam et al. 1989

Modified from Surdam et al, 1989

 

This is part of the How To…series  on carbonate rocks

Some general statements about the diagenesis of limestones during burial.

As sediment is buried, the combined effects of temperature, pressure and changing fluid composition act to lithify, overprint, and even obliterate all previous diagenetic histories. In limestones, the original sediment components and any early cements formed during seafloor or shallow meteoric diagenesis, are cemented, replaced by more stable carbonate phases, or recrystallized.  Burial diagenesis is governed primarily by:

  • Increasing temperature as a function of the local geothermal gradients,
  • Increasing hydrostatic pressures with depth and sediment/water loads,
  • Reactions resulting from organic maturation where pH buffering and changes in pCO2 drastically shift the stability of carbonates,
  • Reactions involving silicates (especially clays) that also change fluid composition and pCO2,
  • Tectonically induced faulting and folding that can alter permeability pathways, and
  • Tectonic uplift that reduces ambient temperatures and pressures and exposes indurated rock to meteoric conditions.

Continue reading

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of carbonates; Beachrock

Facebooktwitterlinkedininstagram

Beachrock on a Rarotonga shore

 

Early lithification of beach sand.  This is part of the How To…series  on carbonate rocks

One of the most recognizable products of seafloor cementation is beachrock; lithified beach sand. Modern beachrock is common on tropical coasts, those that are wave-washed and quieter shores sheltered by reefs and island barriers. It is less likely to be found bordering cooler seas. Beachrock forms in clean carbonate and siliciclastic-volcaniclastic sands. Lithification occurs so rapidly (months, years)  that one can find all sorts of interesting relics entombed – shells, fish skeletons, coconuts, the flotsam and jetsam of wars, boats that have come to grief, or the refuse strewn by centuries of ocean travelers.

Beachrock affords a domicile for algae and invertebrates that would not normally enjoy living on a soft sand beach; thus, its formation may change existing biotas. It also provides a protective carapace to a beach, reducing the impact of waves. Formation of beachrock potentially changes the beach dynamics. Continue reading

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of Carbonates; Sea floor diagenesis

Facebooktwitterlinkedininstagram

Radial clusters of fibrous aragonite from an Abu Dhabi sabkha

Diagenesis on the sea floor – This is part of the How To…series  on carbonate rocks

 

Seafloor carbonate diagenesis encompasses the vadose intertidal-supratidal,  the phreatic shallow subtidal lagoon and platform, deeper water slopes below the photic zone, to bathyal ocean floors. This post deals with the shallow end of the action.

Physical and chemical changes to carbonate sediment begin almost immediately; shells are bored by sponges and fungi, reef corals are grazed by sea urchins (what a lovely Dickensian expression), aragonite and calcite precipitate, skeletal debris dissolves. Biological processes go hand in hand with the abiotic. I like James and Choquette’s description of this important beginning to diagenesis – (paraphrased) loose sand that becomes cemented as a hardground changes the living conditions for a myriad critters and plants. Thus, the formation of beachrock or subtidal hardgrounds displaces a bunch of gastropods that prefer to live on soft sand.

Diagenesis at the seafloor involves fluids that are the same or similar composition to surface waters: unmodified seawater beneath open platforms, brackish fluids in seawater-fresh groundwater mixing zones, and hypersaline fluids in supratidal and sabkha environments. Fluid drive at the surface and shallow subsurface is provided by: Continue reading

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of carbonates; cements

Facebooktwitterlinkedininstagram

Fibrous aragonite bundles acting as a cement in intertidal shell sand

This is part of the of  How To…series…  on carbonate rocks

The diagrams and images of carbonate crystal habits and cements are descriptive and intended to provide essential background to other posts that detail the different diagenetic environments.

Carbonate diagenesis is like a game of two halves: one part involves mineral dissolution, the other precipitation.  The two commonly go hand-in-hand; it all depends on the changing fortunes of thermodynamic stability and interstitial fluid flow as the game progresses.

Cements precipitate in available pore space: intergranular, intragranular (like the whorls of gastropods, the septa of corals,  or the chambers of foraminifera), larger voids like those developed in reef frameworks, and microporosity such as pore throats between grains. Neomorphism and mineral replacement involve dissolution and precipitation that change existing cement fabrics and sediment frameworks and hence are not confined to pore space.

Carbonate cements are as varied as the diagenetic environments in which they form – the sea floor, meteoric, deep burial and everywhere in between. The crystal shapes of CaCO3, it’s polymorphs and chemical variants range from needle and whisker-like, to blocky spar. The transitions from one crystal form to another, their growth in open pores, and replacement by stable carbonate phases is what makes carbonate petrography so fascinating. Continue reading

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin