Category Archives: Under the microscope

Mineralogy of carbonates; diagenetic settings

Facebooktwitterlinkedininstagram

Micritised bioclasts cemented by isopachous calcite followed by drusy calcite.

Carbonate diagenesis; How limestones form.

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

Of all the common rock-forming minerals, carbonates are the most reactive chemically. The transformation of loose sediment to hard limestone involves chemical reactions that, depending on the conditions (ionic concentrations, pH, degree of saturation, temperature) promote precipitation or dissolution of minerals, most commonly calcite, Mg-calcite, aragonite and dolomite. These reactions take place at the surface (e.g. sea floor) and at all stages during sediment burial and uplift. Limestone diagenetic pathways are complicated; this is part of the attraction for those who study them (notwithstanding the opportunity to conduct field work in places like Bahamas).

Some general requirements for diagenesis to proceed are:

  • The thermodynamic stability and metastability of precipitating phases are determined by pressure, temperature and chemical composition of the fluids (including partial CO2 pressures, pH, and Mg/Ca ratios).
  • Reactions take place in water: sea water, modified sea water, fresh water and saline brines. These fluids are never static; they flow, delivering new solute (ions in solution) to sites of precipitation, and removing dissolved solids to other sites in the permeable sediment.
  • Carbonate diagenesis at all burial depths is, like any other rock type, governed by subsurface fluid flow and evolving fluid compositions. Fluid flow itself is governed by hydraulic gradients that are generated by topography, sediment compaction and tectonic loads.
  • If the conditions change, any phase that becomes thermodynamically unstable will dissolve. This applies particularly to metastable phases like aragonite and high Mg-calcite, where changes in the fluid environment can render them unstable and prone to dissolution. Fluid composition will change when, for example, shallow sea floor carbonates are exposed to meteoric fresh water as sea level falls, or during burial where original seawater is modified by reactions involving carbonates, siliciclastics and, importantly organic matter.

Geologists like to subdivide things and the carbonate diagenetic realm is no different (it helps to simplify a complex world). Three diagenetic environments are frequently cited, with a fourth located at the transition to metamorphism (each environment will be treated in greater detail in separate articles), and depicted in the diagram below:

  1. Seafloor environments (almost syndepositional), including the first few centimetres or metres of burial.
  2. Shallow burial – meteoric environments
  3. Deep burial environments

Diagram of carbonate diagenetic environments featuring meteoric, vadose, submarine, and burial diagenetic realms

Sea floor diagenesis

Precipitation of aragonite and high-Mg calcite cements on the sea floor or in the first few centimetres to metres beneath it, is common in tropical settings, less so in cooler waters. It is the region influenced by ambient seawater compositions; it is also referred to as the marine phreatic zone.

This, the earliest stage of carbonate diagenesis is promoted by low sedimentation rates, high levels of mineral supersaturation in seawater and rapid exchange of atmospheric CO2 with aerated seawater. Microorganisms like bacteria and photosynthesizing algae also contribute, either as mediators or by direct precipitation. Some microorganisms such as endolithic bacteria and algae have a dual role in that their substrate boring activities tend to destroy primary grains but leave micrite rinds and infills in their wake.

Seafloor cementation can take place from the supratidal-intertidal zone (marine vadose zone)  (e.g. crusts, beachrock), reef and shallow subtidal platform (mainly within the photic zone), to the outer platform and slope (beyond the photic zone). At greater depths, first aragonite (the aragonite compensation depth – ACD – is 2-3km) then calcite (CCD is 4-6km) begin to dissolve because of high CO2 partial pressures.

 

Meteoric environments

Platform and reef deposits exposed by a fall in relative sea level will be subjected to an influx of fresh water. Carbonate deposits subjected to meteoric conditions undergo significant changes to their mineralogy and texture.

In the meteoric environment, groundwater flow through permeable aquifers is driven largely by gravitational potential energy, usually referred to as topography-driven flow. Hydrogeologists have long recognized three distinct zones within meteoric settings; James and Choquette (reference at bottom of the page) included these zones in their early model of carbonate diagenesis:

  • The watertable, below which all porosity is completely saturated; this is the phreatic zone.
  • Above the watertable is the unsaturated or vadose zone where pore spaces are mostly air-filled but are periodically wetted by rising watertables (may be seasonal) and infiltration of surface water.
  • Where aquifers intersect the coast there is a zone of fresh water and seawater mixing. The location of the phreatic mixing zone in relation to the shoreline depends on the hydraulic gradient (or hydraulic head) in the aquifer and how far the aquifer extends offshore. It is not uncommon for fresh water to flow 100s, even 1000s of metres offshore, beneath a platform or shelf.

Aragonite and high-Mg calcite are unstable in fresh water which means that any clasts (skeletal fragments, ooids, foram tests) and cements that contain these minerals will begin to dissolve. In some cases, secondary porosity will form. Also common is the replacement of clasts and cements by low-Mg calcite.

Karst landscapes form during prolonged exposure to meteoric conditions where even the low-Mg calcites dissolve.

 

Deep burial environments

Deep burial usually refers to the interval below the influence of marine phreatic and meteoric fluids (10s to 100s of metres deep) to several 1000m depth (depending on the geothermal gradient). The principle physical process is compaction that rearranges sediment frameworks and drives interstitial fluids to other parts of the sedimentary basin. The influence of temperature on the promotion and rates of chemical reactions becomes increasingly important with depth.

Fluid composition is primarily modified seawater, modification that can eventually produce saline brines. Many different reactions come to play as burial depth and temperature increase. Compaction enhances pressure solution where significant volumes of rock are dissolved and the solute transferred to other parts of the sedimentary basin; what remains are stylolites. Other reactions involve dehydration of clays, clay mineral transformations (e.g. kaolinite-smectite), and perhaps most significantly, the diagenesis of organic matter. All these reactions contribute to changes in pH and alkalinity that effect calcite and dolomite stability.

Carbonate diagenesis is dominated by precipitation of calcite and ferroan calcite that replace aragonite and high-Mg calcite. Recrystallization of calcite spar tends to mask and even obliterate original depositional fabrics. Dolomitization is also common during burial diagenesis and like calcite overprints earlier grain and cement fabrics (i.e. those formed in the sea floor and meteoric environments.

 

Companion diagenetic environments

A schematic of siliciclastic diagenesis in the context of principle diagenetic reactions and pH buffering with increasing burail depths and temperatures. Modified from Surdam et al. 1989

It is important to recognize that the diagenesis of carbonates during burial is not divorced from broadly similar processes taking place in siliciclastic rocks. Precipitation of quartz and clays, and dissolution of feldspar are important determinants for evolving fluid compositions that significantly effect carbonate mineral stability (mostly low Mg-calcite). Organic matter in carbonates and siliciclastics has a profound effect on deep burial diagenetic pathways. Complex organic compounds like kerogens begin to break down at about 60o C. The rate of organic diagenesis increases markedly at 80o C – the lower burial temperature limit of the oil-generation window. Important byproducts of these reactions are organic acids that modify pH and control alkalinity. Note too that pH buffering and alkalinity are also influenced by silicate transformations, particularly those involving smectite, kaolinite and illite. The diagram above (from Surdam and others, 1989) summarizes the progression of diagenetic reactions commonly observed in siliciclastic rocks, in relation to organic maturation, organic acids and pH buffering.

 

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

Mineralogy of carbonates; Pressure solution

 

There is a vast, and for the most part excellent literature on carbonate diagenesis. Here are a few classic and more recent texts that provide much more detail on the subject.

Robin G.C. Bathurst, 1976. Carbonate Sediments and their Diagenesis. Elsevier, Developments in Sedimentology, 12. 658p. An example of the longevity and utility of one of the best on this topic. Now also as an ebook.

Noel James and Phillip Choquette.1984. Diagenesis 5. Limestones; Introduction and subsequent articles on sea floor, meteoric and burial diagenesis. The Canada Geoscience Series on Carbonate Diagenesis is available from the CGS archive.

Noel James and Brian Jones. 2015. The origin of carbonate sedimentary rocks. American Geophysical Union, Wiley works, 464p.An excellent recent update.

Peter Scholle and Dana Ulmer-Scholle, 2003. A colour guide to the petrography of carbonate rocks: grains, textures, porosity, diagenesis. AAPG Memoir 77. Loaded with images.

R.C. Surdam, L.J. Crossey, E.S. Hagen, & H.P. Heasler. 1989. Organic-inorganic interactions and sandstone diagenesis. AAPG, v.73, p. 1-23.

SEPM Strata. Diagenesis and porosity. Part of SEPM’s online stratigraphic web contructed originally by Christopher Kendall. An excellent resource for pretty well anything sedimentological and stratigraphic. Continually updated.

Erik Flugel. 2010. Microfacies of carbonate rocks: Analysis, interpretation and application. Springer. The ebook is cheaper

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of carbonates; basic geochemistry

Facebooktwitterlinkedininstagram

Tabulation of the important equilibria in the carbonic acid - carbonate system

Some basic geochemistry of carbonates, carbonic acid and carbon dioxide

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

Sedimentary carbonate petrology is concerned first and foremost with the precipitation and dissolution of mineral phases; principally calcite, aragonite and dolomite. Both processes involve chemical reactions and the two primary requirements for these reactions are:

  • thermodynamics – there needs to be sufficient energy to drive the reactions, and
  • an excess or deficit of dissolved mass that, for the minerals of interest, includes  Ca2+(aq), Mg2+(aq), and CO32-(aq) (aq = aqueous)

All sedimentary carbonate reactions at the surface or during sediment burial take place in water: fresh water, sea water, or concentrated brines. Furthermore, the chemical composition of these fluids can evolve, for example during burial (sea water to brine), or uplift and exposure (sea water to fresh meteoric water). Such changes are commonly manifested as cement stratigraphy (e.g. high-Mg calcite overlain by low-Mg calcite), or where clasts, matrix and early cements are replaced by new generations of calcite or dolomite.

Evolution of cements and fluid chemistry in a cool-water limestone, NZ.

We are also interested in carbonate reactions because of the present trend towards climate change and the possibility that ocean chemistry may change; ocean acidification tops the list here. To get beyond the hyperbole, to determine whether this is a real possibility or not, we need to understand some basic chemistry.  Some introductory concepts are outlined herein.

To delve deeper into the complexities of the carbonate system, have a read of the contributions cited below. Continue reading

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

The mineralogy of carbonates; classification

Facebooktwitterlinkedininstagram

Oncoidal grainstone, cryptalgal boundstone and rudstone in Proterozoic carbonates

A post in the How to… series on carbonate mineralogy – limestone classification

The classification of carbonate rocks, like that of sandstones, has gone through a few iterations. Two schemes have stood the test of trial and error, in the field and through microscopes; both were compiled in the late 1950s – early 60s, each serves a slightly different purpose, both are still popular. They are the classification schemes of R. Folk (1959, 1962), and R. Dunham (1962). These two schemes form the backbone of modern carbonate classification and are summarized here, but keep in mind that quite a few iterations have been published.

 

Folk’s classification scheme

Folk’s scheme for carbonates is in some respects like the one he devised for terrigenous sandstones. It is based on the proportions of matrix, in this case lime mud, and framework components that are mostly allochems (grains that have been subjected to some degree of transport). At one end if the spectrum there is pure mud, or micrite; at the other end clast-supported frameworks with no mud. The spectrum of textures also reflects the energy of the depositional system. The pore volumes in clast-supported limestones are filled with cements, commonly sparry calcite mosaics, some with micritic cements. The degree of grain sorting in these sparites also increases. Grain size follows a relatively simple size-class like the Wentworth scale:

  • Lutites, 0.004 – 0.062 mm, that are approximately equivalent to mud (silt and clay)
  • Arenites, 0.062 – 1.00 mm (very fine- to coarse-grained sand)
  • Rudites, 1.00 mm to boulder range

Folk’s limestones are classified as either micrites or sparites with qualifiers added for the kinds of allochems (ooids, pelloids, fossils and intraclasts) and grain size. Identification of the allochems usually requires a microscope. Hence, Folk’s scheme is best applied to thin sections.

R.L Folk's classification scheme for carbonates

Folk extended this basic classification to include the percentages of micrite and spar cement (diagram below). The cut-off percentage between pure micrite and a micrite with allochems is 1%, 1-10% skeletal fragments is a fossiliferous micrite, 10-50% a sparse biomocrite, and >50% a packed biomicrite. Other qualifiers like pelloids and ooids use the same designations. Thus, an oosparite may be an unsorted oosparite, or if matrix-supported a sparse oomicrite.

Rigid structures like reefs and bioherms were placed in a category of their own – biolithite. The category dismicrite refers to micrite that has been disturbed by burrowing or erosion where voids have been filled by sparry calcite (i.e. disturbed micrite). However, it may be difficult to distinguish between a dismicrite and a micrite that has undergone partial recrystallization to calcite spar.

Folk's limestone classification scheme based on texture and depositional conditions

Dunham’s classification scheme

This scheme is more applicable to outcrop, hand specimen and drill core. It too is based on textural attributes but only those acquired during deposition (which means cements are excluded). His scheme uses three basic components:

  • The proportion of mud that differentiates between muddy limestones and grainstones (the latter having no mud),
  • The percentage of grains giving us a mudstone, wackestone or packstone, and
  • The presence of binding agents (mostly biological) giving us a boundstone.

Both Folk and Dunham apply a separate catch-all category of biolithite and boundstone respectively, for limestones containing fossils and inorganic structures bound by algal mats and more rigid frameworks like corals, stromatoporids and bryozoans. One could apply a qualifier, such as stromatoporoid boundstone, but this gives little paleoenvironmental information on the kinds of structures involved. While working on some Devonian reefs in the Canadian Arctic, Embry and Klovan realized that additional classification categories were required to fully describe the limestone structures they encountered. Their classification became an expansion of Dunham’s scheme, the basics of which have also stood the test of time, albeit with the odd modification.

Embry and Klovan added three new categories:

  • Bafflestone where organisms trap sediment. By their own admission, Embry and Klovan note that identification of organisms responsible for trapping is equivocal.
  • Boundstone where material is bound by encrusting and binding organisms, such as calcareous algae and cyanobacterial mats, and
  • Framestone that is constructed by framework-building organisms like corals and stromatoporids.

Embry and Klovan's limestone classification, modified from Dunham

So which scheme does one use? The popularity of either scheme depends on which text or journal paper you read – some say Dunham’s scheme, others Folk’s scheme. It is also the case that the two are used interchangeably. If your study focuses on field and outcrop, then Dunham is the logical choice; if the focus is petrographic and thin section then Folk. And if you incorporate outcrop and microscope work, then perhaps use both schemes with cross-referenced rock names.  To some extent it depends on personal preference.

Carbonate students should also look at an evaluation of these two popular schemes by Stephen Lokier and Mariam Junaibi. (2016).  This paper (open access) looks at some of the modifications to both Folk and Dunham schemes.  They find Dunham’s scheme (or modifications thereof) is the most popular. Their own modified scheme is shown below; note that Bafflestone is no longer used here.

A modification by Lockier and Junaibi 2016, of Dunham's limestone classification scheme

*From Sedimentology, v 63, p. 1843-1885, Figure 12 – see link above)

 

Other post in this series:

Mineralogy of carbonates

Mineralogy of carbonates; skeletal grains

Mineralogy of carbonates; non-skeletal grains

Mineralogy of carbonates; lime mud

Mineralogy of carbonates; carbonate factories

Mineralogy of carbonates; basic geochemistry

Mineralogy of carbonates: Stromatolite reefs

 

Important literature contributions

R.J. Dunham, 1962. Classification of carbonate rocks according to depositional texture. In W.E. Ham (editor), Classification of carbonate rocks. American Association of Petroleum Geologists Memoir 1, p. 108-121.

A.F. III. Embry and J.E.Klovan, 1971. A Late Devonian reef tract on northeastern Banks Island, N.W.T. Bulletin of Canadian Petroleum geology, v.19, p. 730-781.

R.L. Folk, 1959. Practical petrographic classification of limestones. Bulletin American Association of Petroleum Geologists, v. 43, p. 1-38.

R.L. Folk, 1962. Spectral subdivision of limestone types. In W.E. Ham (editor), Classification of carbonate rocks. American Association of Petroleum Geologists Memoir 1, p. 62-84.

C.St.J.C. Kendall and P. Flood, 2011. Classification of carbonates. In D. Hopley (editor) Encyclopedia of Modern Coral Reefs; Structure, Form and Process. Springer, p. 193-198.

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

The mineralogy of carbonates; lime mud

Facebooktwitterlinkedininstagram

Mud-dominated stromatolite reef, Paleoproterozoic, Belcher Islands.

A post in the How to… series on carbonate mineralogy – carbonate mud

Micrite – a sensible contraction of microcrystalline calcite

Lime muds are important components of most shallow carbonate-forming environments: platforms, platform margins and slopes, back-reef and fore-reef settings, and lagoons. They also accumulate in shallow alkaline lakes and as oozes on the deep ocean floor. Like their terrigenous cousins, they accumulate as discrete beds and as matrix to coarse-grained carbonate frameworks.

Lime mud, or micrite includes aragonite and calcite (and perhaps dolomite) having crystal dimensions less than 4-5 microns. Thus, any detailed study requires an SEM or similar technology. Their grain size, and hence high unit-volume surface area means they are quite reactive and prone to recrystallization to coarser calcite-dolomite spar mosaics.

Today, the largest volumes of carbonate mud are produced in shallow warm seas at depths usually less than 10 m. All carbonate mud is produced by precipitation, either directly (i.e without biological intervention), or as a by-product of biological activity. Although there is a degree of certainty about this, the relative contribution of each source remains contentious, even after decades of study. Volumetrically less significant but perhaps no less important sources of mud include bioerosion and mechanical breakdown of skeletal material.

Calcareous greeen algae Penicillus and Halimeda, producers of aragonite mud

Several species of green algae produce micron-sized aragonite needles and plates that are shed when the alga dies. Of these, the codiacean algae Halimeda and Penicillus are the most prolific producers. For example, in Florida Bay and reef measurements of Penicillus’ aragonite productivity (as needles 2-3µm long) indicate that this species alone can account for all the mud deposited there (Stockman et al. 1967). Other codiacean species plus breakdown of skeletal material also contribute mud, but at this location Penicillus is it. The aragonite produced in these shallow waters is also transported to tidal flats and farther offshore.

Halimeda growing amongst reef corals and sponges, and a watchful Moray Eel.

Across the ditch, Grand Bahama Bank also hosts swaths of aragonitic mud. Calculations like those penned for Florida, demonstrate that most of this mud (on a volumetric basis) could also have come from codiacean algae. However, this is disputed by Milliman and others who contend that a large proportion of the micron-sized aragonite (up to 75%) precipitated directly from sea water because:

  • The aragonite crystal morphology is more like that produced by direct precipitation, and
  • The strontium content of the mud is too high – Sr/Mg ratio ≈ 4 compared to average values of ≈ 2 for codiacean algae.

One phenomenon that has aroused interest as a potential source of aragonite mud is the spontaneous development of milky white clouds of suspended carbonate, or whitings. They form over shallow platforms and persist only as long as it takes for the suspension to disperse or fall out of suspension. Two mechanisms are frequently invoked as explanations:

  • Where bottom muds are stirred up by feeding fish (this has been observed), and
  • By direct precipitation from seawater. Measurement of whitings over Grand Bahama Bank show that most of the suspension is micron-sized aragonite, but up to 20% high Mg-calcite may be present (Shinn et al 1989).

Some support for the direct precipitation hypothesis is linked to examples of whitings in areas devoid of bottom-dwelling red and green algae (e.g. Trucial Coast). However, this does not preclude the possibility that suspended algae may play a role. J.W. Morse has suggested that en mass precipitation of aragonite in normal seawater can occur if phytoplankton photosynthesis (during algal blooms) reduces dissolved CO2.

Extensive whitings of suspended carbonate particles off the coast of Qatar, plus a greenish algal bloom.

The Bahamian aragonite is also transported to tidal flats and shallow lagoon, and to the platform margin where it accumulates from suspension on the adjacent deep water slope; carbonate mud that accumulates in this way is called hemipelagic mud.

During the Late Paleozoic, the calcareous algae niche was occupied by phylloid algae that were responsible for building mud mounds. Classic examples are found in Pennsylvanian cyclothems of Kansas and Oklahoma. Here they form part of the regressive limestone facies, where bladed and leaf-like phylloids acted as structural frameworks for the buildups. The facies has been well summarized by Phil Heckel (and lots of references therein).

Precambrian carbonate mudrocks (mostly dolomitized) are invariably associated with stromatolites of all shapes and sizes, from simple laminar mats to complex buildup reef-like structures like the example shown at the top of the post . It is generally understood that their construction was promoted by prokaryotic cyanobacteria (Phanerozoic red and green algae are all eukaryotes). Sedimentological evidence indicates that, like their Phanerozoic counterparts, they too inhabited shallow marine shelf, platform and inter-supratidal environments (here’s a link to the Stromatolite Atlas).

Digitate stromatolites that originally were built from carbonate mud precipitated or trapped by cyanophyte algae. Belcher Islands.

However, carbonate mud was deposited on the outer edges of platforms and adjacent slopes. An example from Belcher Islands (Costello Formation) consists of thin bedded dololutites and calcilututes, with the odd calci-turbidite, but no evidence of cryptalgal laminates. These deposits are also thought to have accumulated from mud suspended in the water column (i.e. hemipelagic) that originated in the carbonate mud ‘factory’ on the shallow platform.

Hemipelagic mud deposited on a Proterozoic slope - the mud was probably sourced from an adjacent carbonate platform

Deep sea oozes are composed primarily of coccoliths and planktonic foraminifera that live in uppermost ocean waters (in the photic zone), and when dead gradually sink to the sea floor. Oozes accumulate in oceanic regions that are devoid of terrigenous sediment.

links to other posts in this series

Mineralogy of carbonates

Mineralogy of carbonates; skeletal grains

Mineralogy of carbonates; non-skeletal grains

Mineralogy of carbonates; classification

Mineralogy of carbonates; carbonate factories

Mineralogy of carbonates; basic geochemistry

Mineralogy of carbonates: Stromatolite reefs

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

The mineralogy of carbonates; non-skeletal grains

Facebooktwitterlinkedininstagram

Ooid bars and sandwaves on Great Bahama Bank

A post in the How to… series on carbonate mineralogy – non-skeletal components

Ooids, pisoids, oncoids and pelloids comprise the most common non-skeletal component of limestones. The first three are characterized by enveloping laminations that range from perfectly concentric spheres to strongly asymmetric grains. They are all products of calcite and aragonite precipitation at the sediment-water interface or within the vadose zone of shallow burial. Although non-skeletal, there is mounting evidence that their formation is influenced, even promoted by biological activity, particularly by bacteria, fungii and algae.

 

Ooids

Having the opportunity to study modern ooids must be a real treat; a fortuitous circumstance that puts the sedimentologist fair and square in the balmy Bahamas or Trucial Coast (Arabian Gulf). Oh well, I got to spend field seasons in the Canadian Arctic.

Aragonite ooids from Joulters Cay, Great Bahama Bank

Modern sea-going ooids presently form in tropical settings that promote precipitation of aragonite and high-Mg calcite. Ooids are spherical to subspherical grains, characterized by concentrically layered, micron-sized calcite or aragonite crystals. Bahamian ooids are generally 0.5 mm diameter and less. Their cortex consists of micron-sized aragonite crystals that are organized roughly tangential to grain surfaces. The cortex usually envelops a nucleus of skeletal debris, foraminifera or fecal pellets. Variations on this theme are found in some saline lakes where the cortex is a radial array of acicular microcrystals (calcite and aragonite) around a nucleus (e.g. Great Salt Lake, Utah). Ooids are also documented from Precambrian iron formations and Phanerozoic iron-rich deposits where the cortex consists of iron oxides.

Swaths of ooid sediment blanket the Great Bahama Bank beneath clear waters no deeper than 5-10 m. Along the Trucial Coast they accumulate in shallow high energy tidal deltas associated with barrier islands and lagoons.  Ooids are subjected to waves and tidal currents, that polish, and sort them according to grain size. Bedforms abound, ranging from large sand-waves (subaqueous dunes) to ripples. Ooids may also be mixed with skeletal debris and pelloids (limestones with ooid frameworks are called oolites). The propensity for ooids to form in shallow, high energy depositional settings makes them excellent paleoenvironmental indicators.

Ancient ooids also display concentrically layered cortices but invariably they are composed of radially fibrous or acicular calcite. In thin section and under polarized light the radial fabric gives rise to an extinction cross centered on the ooid nucleus, that rotates with the microscope stage. This begs the question ‘were the ooids originally calcite (high or low Mg) or aragonite?’. Those that were originally calcite may preserve in fine detail the radial pattern and concentric layering. In contrast, calcite that has replaced aragonite tends to lose some (but not all) of this textural detail. The question is complicated further if the calcite in either case shows any degree of recrystallization – this usually results in larger crystals that destroy original textures and fabrics. Dolomitization tends to exacerbate this problem.

Concentrically layered Carboniferous ooids, indented by pressure solution

The origin of ooids has been a topic of discussion for at least the last two centuries: explanations ranging from fish eggs to mechanisms invoking snowball accretion of aragonite, from the purely physico-chemical (with no biological interference) to purely biological. Take a look at the excellent summary of past and current ideas in a review by Harris, Diaz and Eberli (2019). What is known is that, whether aragonitic or calcitic, they are the product of direct precipitation from seawater or saline lake water. Research over the last few decades has also shown that biological mediation by microbes is also a critical factor in the formation of ooids.

The association of organic matter, algal filaments, bacteria and fungii enmeshed with aragonite crystals in ooid cortices has been known since the 1950s.  These early discoveries promoted the idea that the organic matter and microorganisms changed the chemical microenvironment at sites of crystal growth (pH and the activity of Ca2+ and CO3 2-). While it is likely these biochemical processes do operate, it is now understood that microorganisms play a more complicated role in carbonate biomineralization by creating an organic template that fixes Ca2+ and CO3 2- ions and provides sites for crystal nucleation.

 

Pisoliths and pisolites

Vadose pisolite in a Proterozoic supratidal calcihe, Belcher Islands

Pisoliths (or pisoids) bear a superficial resemblance to ooids, but are larger (>2 mm) and concentric layering is more irregular. A classic case of misinterpreted pisoliths lies in the Permian Capitan Reef complex (Texas). Originally interpreted as sedimentary and of algal origin (concentric layering of algal laminates), R. Dunham (1969) concluded that they had formed as concretionary structures within the rock column, specifically the vadose zone of subaerially exposed carbonate deposits. The Capitan Reef structures grew by progressive in-situ calcite and aragonite precipitation (the cemented rock is a pisolite). The vadose zone occurs between the land surface and the watertable (sometimes called the unsaturated zone). Sediment within the vadose zone is wetted when the watertable rises, and by surface water infiltration during precipitation events. Thus, carbonate precipitation is a periodic process associated with alternating sediment wetting and drying. Pisoliths may grow competitively producing fitted contacts. Downard elongation of laminae, another characteristic feature, is probably a response to the gravity-driven seepage.  There may also be periods of carbonate leaching that remove segments of the pisoliths; subsequent regrowth will result in discordant sets of carbonate layers.

Acetate peel of fitted and elongate vadose pisoids, from a Proterozoic caliche

Vadose zone pisoliths are excellent indicators of subaerial exposure and ancient soil-forming conditions, particularly carbonate paleosols or caliches.

 

Oncoids and oncolites

Oncoids are rounded, spherical to oblate, laminar growths of algae around a nucleus (shells, mud intraclasts, broken lumps of algal crust). Their dimensions are measured in centimetres. The algae usually associated with oncoids are the cyanobacteria (of stromatolite fame). They  grow at the sediment surface in supratidal (including sabkhas), intertidal and shallow subtidal environments that are agitated by currents and waves, and perhaps frequented by storm surges. Oncoid laminae grow asymmetrically while the oncoid is at rest, and discordantly when growth has been interrupted by clast jostling and rolling.

Cryptalgal oncoids mixed with laminate rip ups from storm surge across a supratidal flat, Proterozoic, Belcher Islands.

Pelloids

The term pelloid refers to any grain that is an aggregate of micro- to cryptocrystalline carbonate. Most pelloids are spheroidal and sand sized; there is little or no internal structure. If they are known to be fecal then they are called pellets. Otherwise, the term pelloid should be used.

Micrite peloids packed around a skeletal fragment, cemented by sparry calcite

Pellets are very common in lower energy settings like lagoons; walk over any tidal flat strewn with gastropods and there will be countless mounds and ribbons of fecal pellets. Fecal pellets contain high proportions of organic matter that break down to organic acids during burial. Their softness also means that grains are susceptible to compaction. Closely packed pellets will tend to merge into a single mass to the point where individual grains become indistinguishable; the end product may be a micrite that is indistinguishable from non-pelloidal deposits.

Like other micrites, pelloidal limestones are susceptible to recrystallization. Note that the term pelloid is also used for skeletal fragments that have been micritised by endolithic algae and fungi.

links to other posts in this series

Bivalve shell morphology for sedimentologists

Gastropod shell morphology for sedimentologists

Cephalopod morphology for sedimentologists

Mineralogy of carbonates

Mineralogy of carbonates; skeletal grains

Mineralogy of carbonates; lime mud

Mineralogy of carbonates; classification

Mineralogy of carbonates; carbonate factories

Mineralogy of carbonates; basic geochemistry

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of carbonates – skeletal grains

Facebooktwitterlinkedininstagram

modern cool water shell hash

A post in the How to… series on carbonate mineralogy

Like sandstone, limestones are made up of framework components, variable amounts of matrix, and cements. Frameworks, or allochems, consist of skeletal fragments, non-skeletal clasts like ooids, pellets, and lime-mud rip-ups (micrite) that have all been subjected to some degree of transport (sometimes very little). Unlike sandstone, the lithification of allochems begins at the sea floor and continues during burial; the products of lithification involve the transformation of calcium carbonate polymorphs, and precipitation of cements (calcite, aragonite, dolomite), sometimes referred to as orthochems. Skeletal and non-skeletal grains plus cement are all used to describe and classify limestones.

The skeletal components of limestone are largely sourced from invertebrates, a wonderfully diverse group of animals that are domicile on the sea floor, within its sediment, and in the water column above. Invertebrate skeletal material may survive intact, or as fragments resulting from mechanical breakdown, bioerosion, or the end product of a vertebrate’s dinner. The preservation potential of all this material is high.

Devonian crinoid limestone, Maryland

Corals, the foundations of reef complexes, are a dominant component of tropical limestones (reef top, fore-reef, back-reef and lagoon).  During the Paleozoic this niche was also occupied by Stromatoporids, and during much of the Precambrian by stromatolites buildups, or reefs. Echinoids, various molluscs (particularly bivalves and gastropods), brachiopods, benthic foraminifera and sponges are common co-inhabitants. Calcareous algae as encrusting forms and the lime-mud producers Halimeda and Penicillus are important contributors to the carbonate factory.

Structural components of coral reefs; Porites, bryozoa, purple sponges

Cool and temperate-water limestones lack coral reefs although solitary corals do occur. Instead, these limestones form from accumulations of bivalves, gastropods, echinoids,  forams, and associated attached or encrusting benthic critters like barnacles, bryozoans and calcareous algae; Halimeda, Penicillus and other lime-mud producers do not inhabit cool-temperate seas.  Although not reef-like, some species such as oysters commonly construct significant buildups; the Oligocene Te Kuiti Group in New Zealand has some excellent examples.

All these critters, phytoplankton and algae inhabit the photic zone, the uppermost layer in lakes and seas where photosynthesis can take place. The depth to which light penetrates is highly dependent on water turbidity. In clear water 50% of light is absorbed at one metre depth and 90% at 20m. There is little significant light below 200m depth. The photic zone is of critical importance for algae photosynthesis – algae are a primary food source for many invertebrates or join with them in symbiotic relationships.

In tropical-subtropical environs, sea-floor cements include calcite (high- and low-Mg varieties) and aragonite. Sea floor cementation of temperate water carbonates does occur but usually with calcite alone.

Some examples of the important limestone skeletal components are shown below. However, it is important to recognize the limitations when identifying individual fossil grains in hand specimen and with an optical microscope:

  • The fragments you are viewing will probably be at some random, unknown orientation.
  • Depending on orientation, the internal structural organization of crystals may be similar across a range of species.
  • The original skeletal fabrics may be disrupted and even completely obliterated by recrystallization or mineral replacement. Dissolution of skeletal grains also produces moldic porosity.
  • In general, your identification will probably get no further than class level – if you can distinguish bivalve from gastropod, even bivalve from brachiopod, you are doing well.

Before launching into the unknown, it is worthwhile spending time with thin section – grain mounts of known invertebrates, microfossils and calcareous algae. These will avail you of their attributes and help demonstrate some of the problems noted above.

 

Microscopic skeletal structures

Skeletal structure in most invertebrates (shells, corallites, tests) consists of microscopic and submicroscopic calcite and aragonite crystal aggregates organized in layers. The orientation and composition of crystallites is highly variable among phyla, classes and even genera. Molluscs commonly have two or three layers of aragonite or alternating calcite-aragonite layers. Nacreous layers that commonly show beautiful coloration on shell inner surfaces (common in bivalves, gastropods and ammonoids), are always aragonitic. Here the aragonite crystals are stacked plates or prismatic clusters, or both. Other layer types in molluscs might include foliated calcite crystals (where the extinction figure sweeps across the skeletal grain), and various combinations of herringbone-like growths of microscopic needle crystals.

Left: micrite-filled borings in a bivalve fragment. Right: barnacle fragment showing good plicate structure

Brachiopod shells generally have two layers – modern genera are mostly low Mg calcite. Modern echinoderms tests consist of interlocking plates and spines; each plate and spine is a single high Mg calcite crystal. Echinoid plates and spines are highly porous and under the microscope appear like large ‘poikilitic’ calcite crystals. Calcite cements tend to grow as syntaxial overgrowths on these crystals (i.e. the cements are in optical continuity with the host crystal).

Oblique sections through echinoderm spines in modern Hawaiian beachrock. Note the distinctive radial pattern of single crystal growth and pores between each radial 'spoke'.

Oblique sections through echinoderm spines in modern Hawaiian beachrock. Note the distinctive radial pattern of single crystal growth and pores between each radial ‘spoke’.

Modern Scleractinian corals are composed of aragonite fibers or needles organized into radiating clusters, or spherulites. Extinct groups such as Rugose and Tabulate corals consist of low Mg calcite fibers, but these may have recrystallized from high Mg calcite or aragonite. In thin section, the fibrous clusters plus the overall coral structure that includes the radiating septa (solid walls separating the cavities) are useful identifiers.

Thin section micrograph of cool water solitary corals. The original aragonite has been replaced with Low magnesium calcite. The appertures are lined with micrite

Modern bryozoans consist of high Mg calcite, aragonite, or both as microscopic and submicroscopic crystals. In thin section fragments of bryozoans are best recognized by their fenestrate structure – a regular arrangement of chambers, usually in pairs, that housed each animal (bryozoans are colonial structures).

Thin section of cool water bryozoan limestone cemented by coarse calcite spar

Foraminifera are mostly calcitic, although some groups like Globigerina and related families are aragonitic. Like the other phyla, foraminifera wall structures are variable – common types are fibrous (normal to the test wall), granular, and microgranular or micrite-like.

Fibrous calcite structure of the walls of a benthic foraminifera. Cross polars

 

Thin section of planktic foraminfera ooze, showin fibrous and prismatic calcite (replacing aragonite) in chamber walls

Calcareous algae are another equally diverse group that are important limestone contributors. Common encrusting forms like Coralline algae are mostly high Mg calcite where individual crystals are submicroscopic that collectively form intricate arcuate or wavy layers. One common species, Lithothamnion, commonly encrusts pebbles and cobbles and typically is found along high energy rocky coasts.

Coralline algae that have grwn as crudely concentric rhodoliths. Miocene, Waitemata Basin, NZ

Calcareous green algae such as Halimeda and Penicillus produce aragonite needles. Across the shallow sea floor (within the photic zone) carpets of green algae produce large volumes of aragonitic lime mud, or micrite – more on this in the companion Lime mud post.

Planktonic algae such as Coccoliths are composed of low Mg calcite. They too are important contributors to lime mud – but you will need a scanning electron microscope to see them.

SEM micrographs of a siliceous diatom (left), and calcareous coccoliths (right)

links to other posts in this series

Bivalve shell morphology for sedimentologists

Gastropod shell morphology for sedimentologists

Cephalopod morphology for sedimentologists

Mineralogy of carbonates

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

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of Carbonates

Facebooktwitterlinkedininstagram

Tetiaroa coral atoll, French Polynesia, atop an extinct submarine volcanoe, or seamount. The fore reef drops precipitously to the deep ocean floor.

This is the first of a How To…series…  on carbonate rocks – the mineralogy of calcite, aragonite and dolomite.

If the arithmetic is correct, limestones and dolostones make up about 5-10% of sedimentary rocks in Earth’s crust.  They host freshwater aquifers, mineral deposits, hydrocarbons, and some pretty spectacular landscapes (like caves and karst landforms). We treat carbonate rocks separately from sandstones for several reasons:

  • Their framework components are primarily of biological origin, either directly as skeletal material or indirectly where organisms like algae and bacteria mediate carbonate mineral precipitation,
  • All these critters inhabit the photic zone, The depth to which light penetrates is highly dependent on water turbidity. In clear water 50% of light is absorbed at one metre depth and 90% at 20m. There is little significant light below 200m depth. The photic zone is of critical importance for algae photosynthesis – algae are a primary food source for many invertebrates or join with them in symbiotic relationships.
  • Carbonate chemistry is fundamentally different to the siliciclastics, and dominated by three minerals – calcite, aragonite, and dolomite,
  • They tend to be chemically reactive at Earth’s surface and during all stages of burial,
  • Limestones form at all latitudes, but the biomass responsible for their accumulations is most prolific in tropical and subtropical environs.
  • Cementation of limestones begins at the sea floor and continues during burial. Original components like aragonite and high-magnesian calcite are usually replaced by calcite or dolomite during burial. In contrast, cementation of arenites does not begin until after burial.
  • Oceanic limestones (like foraminiferal ooze) tend not to accumulate at depths greater than about 4000-5000m – the calcite-aragonite compensation depth.

Carbonate mineralogy: optical properties

Calcite:

  1. CaCO3
  2. Hexagonal crystal system.
  3. Crystals shapes are varied: commonly as rhomohedra and scalenohedra but can occur as clustered prismatic needles.
  4. Under Polarized light it has high birefringence and appears to twinkle as the microscope stage is rotated. Relief changes from high to low during stage rotation.
  5. Excellent rhombohedral cleavage.
  6. Twin lamellae are common and generally parallel or oblique to the long rhombohedral diagonal.
  7. Uniaxial negative.
  8. Sedimentary calcite can incorporate up to 19 mole% magnesium in its crystal lattice. High magnesian calcite has >5 mol% Mg; low magnesian calcite <5%. The magnesium content of skeletal calcite seems to fall into two groups; <5 mole %, and 11-19 mole %. Ferroan calcite has up to 2-3 mole% Fe2+. Common methods used to distinguish these various forms are noted below.

Calcite twinning in thin section, plain polarised light.

Aragonite:

  1. CaCO3. Can have up to 10,000 ppm strontium (high Sr is common in many recent corals, ooids and calcareous algae.
  2. Aragonite is a polymorph of calcite but belongs to the orthorhombic crystal system.
  3. Mostly commonly as acicular and fibrous cements in limestones, and a common component of invertebrate and protist skeletal material.
  4. High birefringence.
  5. Biaxial negative.

SEM micrograph of radial aragonite clusters in intertidal sands, Auckland, NZ

 

Dolomite:

  1. CaMg(CO3)2. The crystal lattice contains alternating layers of CaCO3 and MgCO3. Highly ordered dolomite contains equal amounts of Ca and Mg. Protodolomites, found in modern hypersaline lagoons and sabkhas, and recently discovered in some calcareous algae, have disordered crystal lattices (on XRD diffractograms) and less Mg.
  2. Hexagonal crystal system, usually as rhomobedra. Commonly seen as a replacement of calcite and aragonite, in some cases the original textures are preserved, and in others they are completely obliterated.
  3. Twinning less common. Twins parallel the short and long rhombohedral diagonals.
  4. Very high birefringence and usually very high relief.
  5. Uniaxial negative

Some popular methods for detecting and distinguishing dolomite and the polymorphs of calcite include:

Staining

Chemical staining of thin sections and rock slabs may seem a bit unsophisticated in an age where the technology for detailed analysis of mineral assemblages is readily available. But for the common carbonate minerals the method can be a very useful first pass at identification. The two stains in common use, Alizarin red-s (ARS) and Potassium ferricyanide (PF), are cheap and easy to use, even in the field where they may give an indication of stratigraphic trends in mineralogy, for example calcite/dolomite ratio, or changes in ferroan calcite. ARS is used to distinguish between calcite (stains pink-red) and dolomite (no stain). The intensity of blue PF stains in ferroan calcite increases with increasing Fe content; ferroan dolomite stains in green hues.

Staining of bivalves reveals replacement of aragonite by ferroan calcite (blue) and non-ferroan calcite (pink)

XRD

Diffraction peaks in an X-ray diffractogram measure the spacings between planes of atoms in a crystal lattice; they are referred to as d-spacings. A characteristic peak in Ca-Mg carbonates is the ‘104 peak’ that increases or decreases according to crystal chemistry. XRD measurement of the 104 peak is one of the more useful methods for identifying the magnesium content of calcites and identifying basic mineralogy in fine-grained deposits.

 

Cathodoluminescence

Pure calcite and dolomite do not luminesce when exposed to a high energy electron beam. However, certain impurities in the crystal lattice may become excited enough to luminesce, emitting light at visible wave lengths. The most common luminescence activator in carbonates is manganese (Mn2+), which emits orange-red to orange-yellow light. Manganese replaces calcium in the calcite lattice during precipitation. As little as 10-20 ppm Mn will activate luminescence. Note that iron (as Fe2+) will quench Mn luminescence if it is present in concentrations greater than a few 10s ppm. Iron as Fe3+, cobalt (Co2+) and nickel (Ni2+) will also quench Mn luminescence.

The amount of Mn2+ incorporated into the lattice will vary during diagenesis in concert with changes in the Mn2+ concentration of the interstitial fluids. This is commonly manifested as zoning in crystals, where the emission colours vary from one zone to another. Analysis of luminescence patterns in carbonate crystals can provide valuable clues about diagenetic history, particularly the evolution of cements, evolving fluid composition, and changes to rock porosity and permeability.

Calcite spar cement stratigraphy revealed by cathodoluminescence. The changes in cement composition (presence of iron and manganese) reflects changes in fluid composition.

Changes in limestone composition through time

We tend to think of limestones as a Phanerozoic phenomenon – those fabulous Devonian, Carboniferous and Permian reefs and platforms. In fact, sedimentary carbonates accumulated across the last 3.5, perhaps even 4 billion years of Earth history. The Precambrian geological record is littered with extensive platform carbonates associated with stromatolite buildups. Most have been dolomitized but there is textural and geochemical evidence of them having been originally calcitic and aragonitic.

Production of skeletal and non-skeletal limestone during the Phanerozoic has not been steady and continuous but has fluctuated on a scale that reflects major changes in plate tectonics. Such grand scale cyclicity was recognised by Fred Mackenzie and John Morse (1992)  who noted a tenuous coincidence between high sea levels and high rates of addition of new crust at mid-ocean spreading ridges, and low sea levels with low accretion rates. A burgeoning database now allows us to be more confident about this grand-scale cyclicity. We now know that the coupling of plate tectonics and sea level also produced changes in sea water composition, specifically the Mg/Ca ratio. High rates of accretion result in a lowering of the Mg/Ca ratio because Mg is removed from seawater and Ca is added during hydrothermal alteration of hot new crust. Geochemical experiments and theoretical analyses also tell us that seawater at low Mg/Ca ratios favours precipitation of calcite cements with low Mg content; many molluscs and modern corals also enjoy these conditions. The opposite, high Mg/Ca ratios, are more likely during low sea floor accretion rates (and low sea levels) and promote precipitation of high magnesian calcite and aragonite cements. High Mg/Ca ratios also favour critters like brachiopods, some extinct corals, modern echinoderms and calcareous algae. Thus, 1st-order low sea level cycles coupled with high Mg/Ca ratios tend to favour aragonite seas and high sea levels with low Mg/Ca ratios calcite seas.

Phanerozoic history of seawater compositions with respect to aragonite versus calcite prone ocean waters

The posts that follow in this carbonate series will deal with: the skeletal and non-skeletal frameworks of limestones, lime mud or micrite, limestone classification, reefs and platforms, cool water limestones, cementation, recrystallization and neomorphism, porosity and permeability – this will do for a start.

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

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Mineralogy of sandstones: Porosity and permeability

Facebooktwitterlinkedininstagram

well sorted sandstone with about 20% porosity. Each grain has a dusting of diagenetic clays

Porosity and permeability – the flow of water and other geofluids

This is part of the How To…series  on the Mineralogy of sandstones

Nearly all geological processes require the presence of water in one form or another. Most sedimentation occurs in water (aeolian deposits are the obvious exception). Sediment burial and compaction involve the expulsion of water. Diagenesis would not take place in the absence of water; hydrocarbons would not migrate to traps and minerals would not be concentrated in ore bodies. Aqueous fluids under pressure reduce cohesion and friction promoting rock deformation.  Metamorphism would be painfully slow, even by geological standards if it were not for the transfer of mass in hot aqueous fluids.

All these processes require not only the presence of water, but its continual movement or flow. Below Earth’s surface, the residence and flow of aqueous fluids requires two fundamental rock-sediment properties:
– voids, commonly in the form of intergranular pores and fractures, and
– connectivity among the voids.
The first of these is referred to as porosity; the second as permeability.

There are two main types of porosity: intergranular porosity that characterizes sands, gravels and mud, and fracture porosity in hard rock. Fracture porosity forms during brittle failure of hard rock or cooling of lava flows. Fracture networks that are connected can provide pathways for fluid flow even when the host rock is impervious (e.g. granite, basalt, indurated sandstone). Highly productive aquifers are not uncommon in fractured bedrock.

Fracture porosity in a columnar jointed lava flow, British Columbia

Intergranular porosity is the void space between detrital grain contacts and is expressed as a percentage of the total sediment-rock volume. It is a dimensionless number (i.e. it has no units of measure). All sediments begin life with some porosity.  Well sorted beach, river and dune sands have initial porosities ranging from 30% – 40%, muds as high as 70%. These values represent the total void space, namely the large pores plus lots of microporosity in tiny nooks and crannies between grains and crystals. Hydrogeologists have found it useful to define effective porosity as that which permits easy movement of fluid. This excludes microporosity where surface tension forces inhibit flow. Effective porosity is always less than total porosity. Follow this link to a simple experiment designed to measure porosity.

As sediment is buried, the grains settle (i.e. they become more closely packed) as they begin to compact.  The reduction in porosity by mechanical compaction continues during sediment burial, in concert with the precipitation of cements (chemical diagenesis).  This is particularly evident during the compaction of mud. The high initial porosity of mud is due to micro-pores between clay particles that have dimensions measured in microns. Compaction compresses the clays and drives off the interstitial water. Compaction (porosity-depth) curves for mud, like the example shown below, typically show a loss of porosity that at shallow depths is almost exponential, becoming approximately linear at depths where shale forms; total porosities in shale are extremely low.

 

Shale porosity - depth (compaction) curve

The conduits for fluid flow (water, oil, gas) from one pore space to another are the narrow connections adjacent to grain contacts. These connections are commonly referred to as pore throats. Pore throats are susceptible to blockage during sediment compaction (lithic sandstones are prone to this) and by cementation, particularly clay cements.

 

Schematic of sandstone burial sequence with compaction and loss of porosity

Diagram of pore-filling cements and occlusion of porosity

Porosity can also be enhanced during burial diagenesis. The primary mechanism for formation of secondary porosity is the dissolution, or partial dissolution of framework grains like feldspar and carbonate bioclasts. Many of these secondary pores are larger than the associated intergranular pore spaces; this is an important diagnostic clue to their identification. Likewise, carbonate and clay cements may be prone to dissolution, resulting in enhanced post-depositional porosity.

 

Secondary porosity caused by the dissolution of feldspar

Burial depths and temperatures where formation of secondary porosity is encountered commonly coincide with chemical reactions involving the break-down of organic matter. By-products of these reactions include carbon dioxide (and carbonic acid) and organic acids like acetic acid. There is a fundamental shift in pH and chemical equilibria, particularly for carbonates, and this promotes dissolution.

Secondary porosity can also form during subaerial exposure of rock and by bioturbation. However, the secondary porosity seen in most ancient sandstones is the product of  burial diagenesis.

Permeability measures the ease with which a fluid flows through sediment or rock. The flow of fluid from one part of a rock to another, or from an aquifer to a bore hole, depends on the connections among pores and fractures. It is possible for a rock or sediment to have high porosity but low permeability if the intergranular or intercrystal connectivity is low – mud and shale are prime examples. In coarse-grained sediments that are devoid of clay, there is a good correlation between porosity and permeability.  This relationship does not apply where there are significant amounts of clay.

Permeability can be expressed in two ways. Henry Darcy’s pivotal experiments with sand-filled tubes (in 1856) established an empirical relationship between hydraulic gradient (that basically is an expression of hydraulic potential energy) and discharge. The proportionality constant in this relationship is called the hydraulic conductivity (K) (a label borrowed from electrical theory), that has units of distance and time (cm/s, feet/s). In mathematical terms, hydraulic conductivity is expressed as a velocity, also known as the Darcy velocity. Hydraulic conductivity is the standard expression of permeability in groundwater studies. Its value depends not only on the connectivity of pores but also on the dynamic viscosity and density of the fluid (viscosity measures the resistance to flow – crude oil is more viscous than water). Thus, for any porous medium the value of K will be different for water and oil, a factor that is important in groundwater remediation.

The hydrocarbon industry deals with fluids of highly variable viscosity (water, oil, gas) and has opted for a standard expression of intrinsic permeability (k) that depends only on the porous medium. The unit is the Darcy that mathematically reduces to units of area (ft2, m2). It is basically a measure of pore size (the oil industry commonly uses the term millidarcy). Frequently used conversions to Darcys are:

1 m2 = 1.013 x 1012 Darcy

1 Darcy = 9.87 x 10-13 m2

Hydraulic conductivity (K) and intrinsic permeability (k) are related by fluid density and dynamic viscosity such that:

k (m2) = K (m/s) x (1.023 x 10-7 m.s) (the time components cancel)

Typical permeability values for unconsolidated sediment and some rock equivalents are shown in the table below.

Table listing typical values of permeability, expressed as hydraulic condictivity and in Darcys

As you can see, the permeability of shale is extremely low. This is the reason why shale beds make good seals to hydrocarbon reservoirs, and aquitards to confined aquifers. Fluid flow in shales and well-cemented sandstones or limestones can be enhanced by hydraulic fracturing. This process (fracing) is front and centre of shale oil production (notwithstanding all the pros and cons of this industrial process). But that is a story for another time.

 

Here are three excellent texts that detail the theoretical aspects of the above:

P.A. Allen and J.R. Allen, Basin Analysis: Principles and Applications. Blackwell 2005

C.W. Fetter.  Applied Hydrogeology, 2001. PrenticeHall

P.A.Domenico and F.W.Schwartz Physical and Chemical Hydrogeology,1998 John Wiley & Sons

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

The provenance of detrital zircon

Facebooktwitterlinkedininstagram

detrital zircon from Pleistocene dune sands, northern NZ. Note relatively sharp crystal faces

This post is part of the How To…series – using zircon geochronology to decipher provenance

Zircon is a common accessory mineral in igneous and metamorphic rocks so it’s not surprising that it is also a common constituent of sedimentary heavy mineral suites. Detrital zircon has assumed a remarkable popularity over the last 2-3 decades as a provenance indicator because:

  • crystals contain measurable amounts of uranium (U), lead (Pb) and thorium (Th) isotopes and can therefore be dated radiometrically,
  • zircon is resistant to chemical and mechanical change – crystals can survive multiple sedimentary cycles (i.e. episodes of erosion from source rocks, deposition, burial and uplift, whereupon the whole process begins anew), and
  • they commonly contain multiple stages of crystal growth that record magmatic, metamorphic and depositional episodes.

Continue reading

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Provenance and plate tectonics

Facebooktwitterlinkedininstagram

This post is part of the How To…series – Provenance, sedimentary basins and plate tectonics

Deciphering the history of sedimentary basins is one of the more exciting tasks geologists can undertake; the provenance of sandstones plays an important part in this adventure.

Sedimentary basins are crustal structures. They are regions of long-term subsidence, responding to tectonic and sediment loads, cooling in the crust and upper mantle, and tectonic dislocation along crustal structures like transform faults. The processes that create sedimentary basins and the sediments that fill them are inextricably linked to plate tectonics.

The idea that the composition of sedimentary rocks was related to large-scale crustal processes was acknowledged by Charles Darwin, Charles Lyell, Henry Sorby and others, but it wasn’t until the 1950s – 60s that sedimentologists began to develop empirical models of sandstone provenance. Robert Folk, Francis Pettijohn, Robert Dott and contemporaries observed links between sediment composition and tectonic domains (or provinces), such as stable cratons and geosynclines – this was the era before plate tectonics.

For geoscientists, the discovery and development of plate tectonic theory changed everything. This is where William R. Dickinson comes into the picture. Dickinson recognized the fundamental link between sedimentary basins and plate tectonics, particularly at plate boundaries. He developed models that relate the modal composition of sandstones to plate tectonic provinces such as collision orogens, magmatic arcs, forearc and foreland basins, and stable cratons. It is important to remember that these models are based on empirical evidence – analysis of 1000s of thin sections that he and many others had recorded from diverse locations.

The models are based on ternary plots like those used by Dott to classify sandstones. Dickinson and his co-workers used different combinations of the quartz, feldspar and lithics end-members to emphasize certain characteristics of the sediment and the source rocks. Both models shown here use the full suite of minerals. Other plots used only the lithic components, or polycrystalline quartz and lithics.

W.R. Dickinson's QFL plots relating provenance to plate tectonics

The Qt-F-L plot combines all varieties of quartz (mono- and polycrystalline quartz, including chert) as a single category and as such emphasizes the maturity of the sediment. Deposits with greater volumes of quartz are generally considered more mature, where mechanical and chemical weathering during sediment transport and deposition have removed less- stable components like feldspar and lithics.

In the Qm-F-L plot, polycrystalline quartz is shifted to the lithic field and in so doing emphasizes the source rocks and production of rock fragments (the quartz component consists only of monocrystalline varieties). Lithics here are key indicators of reworked orogenic provinces along continent-continent and continent-magmatic arc collision provinces. Here, erosion of sedimentary cover and volcanic rocks tends to produce greater proportions of rock fragments.

The second set of diagrams shows typical plate tectonic configurations that correspond to the various QFL fields. The diagrams are highly simplified. In addition to pigeonholing sandstone compositions, the plots provide a useful means of documenting systematic changes as uplift and erosion expose deeper crustal rocks.

QFL plot and plate tectonic cross-section to illustrate provenance from a craton

 

QFL plots and cross-sections of collisional orogen provenance

 

QFL plot and cross-section illustrating provenance from magmatic arcs

For example, unroofing a fold and thrust belt along a collision margin will yield an initial rush of lithics derived from the deformed sedimentary cover. Gradual exposure of a metamorphic core will yield increasing volumes of quartz (mostly polycrystalline) and a new suite of heavy minerals. Likewise, unroofing a magmatic arc complex (the “dissected arc” field in these plots) will provide abundant volcanic lithics followed by more felsic sediment from the deeper intrusive rocks. This is shown schematically in the cartoon below.

Dissection and unroofing of a magmatic arc and typical provenance attributes

The Dickinson plots, like any scientific model, are highly simplified versions of the real world. No two collisional orogens are alike, no two magmatic arcs are exact duplicates. Finding exceptions to any of the models does not indicate their failure – quite the opposite. Their value lies in providing a direction for investigation.

Dickinson’s models have been through several iterations, but their basic structure has survived 40 years of intense scrutiny by geoscientists. They are still useful starting points for unravelling the links among sediment composition, sedimentary basins and plate tectonics.

Check out the companion article – Provenance of sandstones

 

Some useful texts and papers:

Petrology of Sedimentary Rocks, Sam Boggs Jr. 2012

Dickinson, W. R. and C. A. Suczek, 1979, Plate tectonics and sandstone compositions: American Association of Petroleum Geologists Bulletin, v. 63, p. 2164–2182.

W. R. Dickinson, 1988. Provenance and Sediment Dispersal in Relation to Paleotectonics and Paleogeography of Sedimentary Basins. In New Perspectives in Basin Analysis, Editors, Karen L. Kleinspehn & Chris Paola,  Springer-Verlag, pp 3-25.

R.V. Ingersoll, T. F. Lawton, and S.A. Graham, 2018. Tectonics, Sedimentary Basins, and Provenance: A Celebration of the Career of William R. Dickinson. Geological Society of America, Special Paper v.540, 757 pages

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin