You are on page 1of 27

Earth-Science Reviews 104 (2011) 213–239

Contents lists available at ScienceDirect

Earth-Science Reviews
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / e a r s c i r ev

Searching for travertines, calcretes and speleothems in deep time: Processes,


appearances, predictions and the impact of plants
A.T. Brasier ⁎
Scottish Universities Environmental Research Centre, Rankine Avenue, East Kilbride, Scotland, UK, G75 0QF

a r t i c l e i n f o a b s t r a c t

Article history: Common models for modern calcite precipitation in and around caves, soils, springs and streams involve CO2
Received 21 April 2010 supplied by thick, high pCO2 biogenic soils which were probably thin or non-existent before vascular plants.
Accepted 27 October 2010 Indeed plant-influenced chemical weathering might have caused accelerated terrestrial carbonate production
Available online 4 November 2010
from the Devonian onwards. However terrestrial carbonates have also been documented from the Archaean,
Proterozoic, Cambrian, Ordovician and Silurian. Mechanisms which could have caused non-marine carbonates
Keywords:
to precipitate without organic-rich soils are described, and some geological events likely to have influenced
calcrete
travertine
non-marine carbonate precipitation up to the origin of vascular plants are highlighted. As organisms have
speleothem evolved, so have the petrographic characteristics of non-marine carbonates; some examples of this are also
terrestrial vegetation given here.
Palaeozoic © 2010 Elsevier B.V. All rights reserved.
Precambrian

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
2. Definitions of travertine, speleothem and calcrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
3. Vascular plants and weathering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
3.1. Significance of plant roots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
3.2. Chemical weathering in the rhizosphere and the role of fungi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
3.3. Did the origin of plant roots cause increased terrestrial carbonate abundance? . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
4. Precipitation of calcretes, SSL carbonates and speleothems before plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
4.1. Calcite dissolution and precipitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
4.2. Calcite precipitation models including carbonic acid-rich, high pCO2 soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.3. Calcite precipitation mechanisms without high pCO2 soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.3.1. Degassing mechanisms: temperature rise. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.3.2. Degassing mechanisms: deep-sourced CO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.3.3. Degassing mechanisms: turbulence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.3.4. Evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.3.5. Common-ion effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
4.3.6. Freeze-out . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
4.3.7. Precipitation from boiling springs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
4.4. Biological influences on non-marine carbonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
4.4.1. Cyanobacteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
4.4.2. Fungi and lichens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
4.4.3. Vascular plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
4.5. Temporal changes in non-marine carbonate precipitation up to the Devonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
5. Recognising Silurian and older non-marine carbonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
5.1. Petrographic changes of non-marine carbonates through time? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
5.2. Further examples of Proterozoic terrestrial carbonates from North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
5.3. Do non-marine carbonates contain a record of pre-Silurian land surface habitation? . . . . . . . . . . . . . . . . . . . . . . . . . 233
5.4. Are Silurian and older non-marine carbonates usually recognised and reported? . . . . . . . . . . . . . . . . . . . . . . . . . . 233

⁎ Tel.: +44 1355 270 147; fax: +44 1355 229 898.
E-mail addresses: alex_brasier@yahoo.co.uk, a.brasier@suerc.gla.ac.uk.

0012-8252/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.earscirev.2010.10.007
214 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

5.4.1. Cave carbonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233


5.4.2. Spring, stream and lacustrine deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
5.4.3. Calcretes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234

1. Introduction ‘travertine’ have drifted from descriptive terms which can be simply
applied to ancient deposits. Recent articles on travertine and tufa have
Terrestrial carbonates are frequently used in palaeoclimatic and tended to focus upon genetic definitions, based for example upon the
palaeoenvironmental reconstructions (e.g. Retallack, 1997; Andrews source of carbon dioxide, or water temperature, at the time of
and Brasier, 2005; Andrews, 2006; Fairchild et al., 2006). Most of what deposition (e.g. Pedley, 1990; Pentecost and Viles, 1994; see also Jones
is currently known about such carbonates, however, has come from and Renaut, 2010) or upon the chemical mechanism involved in
the Devonian to Holocene record of soil (e.g. Esteban and Klappa, precipitation (e.g. Pentecost, 2005). Thus ‘travertine’ has been
1983; Wright and Tucker, 1991; Wright, 2007), spring, stream, lake reserved by some (mostly European) researchers as a term for
(e.g. Pedley, 1990; Ford and Pedley, 1996; Jones and Renaut, 2010) warmer water hydrothermal precipitates whereas ‘tufa’ has been
and cave carbonates (e.g. Fairchild et al., 2006; Frisia and Borsato, reserved for cooler water deposits (e.g. Pedley, 1990; note many
2010), so that pathways leading to terrestrial carbonate lithogenesis North American authors use ‘travertine’ in both cases). For temper-
have largely been viewed in terms relating to the influence of vascular ature-based definitions, water temperatures have been measured
land plants with roots. Much less is known about the character and directly in active deposits, or estimated indirectly from included
distribution of terrestrial carbonates before this time. The purpose of organisms and fossils (e.g. Pedley, 1990; Koban and Schweigert,
this review is to explore some questions relating to non-marine 1993). Definitions of cool water tufa frequently mention a likelihood
carbonate processes both before and after the arrival of terrestrial of, or even require, macrophyte content. Clearly, such definitions
vascular plants in the Silurian (e.g. Edwards, 1996). cannot be applied with ease to terrestrial carbonates deposited before
Terrestrial carbonates are commonly thought to have become the radiation of macroscopic land plants in the Silurian.
established as a result of the influence of vascular plant root systems Another definition of ‘travertine’ has recently suggested restricting
upon silicate mineral weathering. For example, it has been suggested use of the term to carbonates formed by degassing (see, Pentecost,
that weathering by vascular plants (from the Silurian onwards) and 2005), thereby excluding many spring, stream and lacustrine
particularly their roots (since the Devonian) could have released carbonates formed by the common-ion effect (Section 4.3.5) even
significant amounts of calcium ions from silicate rocks, leading towards though such deposits can be petrographically indistinguishable from
a burgeoning of non-marine carbonate precipitation from the Siluro– those formed principally by CO2 degassing. Restricting ‘travertine’ to
Devonian onwards (e.g. Knoll and James, 1987; Berner et al., 2003; deposits formed by degassing would clearly make usage of this term
Alonso-Zarza and Tanner, 2010). A major prediction of this hypothesis is almost impossible to apply to rocks from the distant past, where such
that pedogenic calcite should be rare and lack root-related features in phenomena cannot be directly observed or confidently inferred.
any examples older than Devonian. Thus while definitions based upon specific physical processes can be
Hitherto, predictions about the nature of the Silurian and older record applied to young and active deposits (e.g. Pedley, 1990; Andrews and
of terrestrial carbonates have barely been explored. There are a few Brasier, 2005; Andrews, 2006) they cannot be confidently used where
exceptions to this. Jutras et al. (2009), for example, reported both only the broad habitat of deposition is known (e.g. a spring or a lake) or
‘pedogenic calcite’ and stable K-rich minerals from Middle Ordovician where the physical conditions themselves are moot. To add confusion,
palaeosols in Nova Scotia. They believe this to be due to more alkaline some Proterozoic marine deposits now better called seafloor crusts
groundwaters prior to the presumed emergence of acid soils during the (Grotzinger and Knoll, 1995) have previously been referred to as ‘tufas’
Silurian. Investigating the possible role of vascular plants in enhancing (e.g. Grotzinger, 1986; Sami and James, 1994; see Riding, 2008), while
silicate weathering and raising levels of terrestrial carbonate precipitation similarly ancient (?hydrothermal) spring deposits have (perhaps
is clearly important for reconstructing global carbon budgets through correctly) been called travertine (Melezhik et al., 2004; Rainbird et al.,
deep time. 2006). The problem with application of these terms is further illustrated
When considering the evolution of such ancient terrestrial by Fig. 1, which shows there is often a degree of ambiguity in the naming
carbonates, a succession of questions emerge. For example, to what of a specimen: a theoretical continuum exists between tufa/travertine,
extent do vascular plant roots affect calcium ion availability and calcrete and speleothem based on varying degrees of biological
terrestrial water alkalinity? In which environments could non-marine involvement and habitats of deposition. Such deposits might be relatively
carbonates exist without the (direct or indirect) influence of vascular easy to name where observed forming today, but precise, restrictive
plants? How could cave calcite precipitation proceed without the high definitions can be a hindrance to understanding once depositional
pCO2 soils which seem integral to Quaternary speleothem genesis? conditions are lost in time.
How have the physical appearances of non-marine carbonates Studies of terrestrial carbonates in deep time therefore require
changed through time? And could non-marine carbonates contain definitions that are applicable to deposits across a wide time span, so
an important record of pre-Silurian land surface habitation? that temporal trends can be plotted from Proterozoic to Recent.
Alternative phrasing to tufa and travertine, such as ‘spring carbonate’,
2. Definitions of travertine, speleothem and calcrete ‘stream carbonate’ and ‘lacustrine carbonate’, can be applied where
depositional environment can be deduced from the stratigraphy. This
The terms ‘tufa’ and ‘travertine’ have been used since Roman times less restrictive terminology is relevant to discussions of deposition in
and have therefore acquired many different meanings. Tufa originally different terrestrial settings in this review (e.g. Section 4). Spring,
meant a porous building stone, including volcaniclastic rocks now stream and lacustrine (‘SSL’) carbonate is thus used here for all
called ‘tuff’. ‘Travertine’ was an alternative Latin name for a calcareous authigenic carbonates precipitated in and around springs, streams or
building stone found around the river Tibur. Definitions of ‘tufa’ and lakes, regardless of the temperature of deposition or cause of
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 215

Fig. 1. Circular diagram showing the theoretical continuum that exists between the concepts of tufa/travertine, calcrete and speleothem. Calcrete alpha-fabrics are usually regarded
as abiological and beta fabrics as biological (for examples see Section 5.1 and Table 2). Tufa and travertine are used respectively for ambient temperature and hydrothermal deposits
in spring, stream and lacustrine environments. They are not used in this review because of the difficulty in deducing water temperature in ancient examples. Speleothem is used for
carbonates deposited in caves. Examples and positions of various rock types within this classification are shown outside the circle.

precipitation which might be implied by using ‘tufa’ or ‘travertine’. these that our current theories about SSL carbonates, speleothems and
The term ‘speleothem’ is similarly used here for carbonate precipi- calcretes have mainly been developed. Later sections examine the
tated in a cave because it is relevant to discussions of carbonate theoretical context for, and some documented examples of, Archaean
deposition in different terrestrial settings, although descriptive to Silurian carbonates of terrestrial origin.
language such as ‘laminar sparry calcite cement’ might be more
useful in reference to individual specimens.
The term calcrete (synonymous with caliche when the latter is 3. Vascular plants and weathering
composed of calcium carbonate) meaning ‘a near surface, terrestrial
accumulation of predominantly calcium carbonate, which occurs in The evolution of vascular plants with rhizomes in the Silurian, and
a variety of forms from powdery to nodular to highly indurated’ (Wright with deeper root systems in the Devonian (see for example Algeo and
and Tucker, 1991) is applied to both pedogenic (soil) carbonates and to Scheckler, 1998; Kenrick and Davis, 2004), might have exerted a
non-pedogenic ‘groundwater’ carbonates (Section 4.1; see also Arakel, major influence upon silicate weathering and thus terrestrial
1986; Wright and Tucker, 1991; Nash and Smith, 2003); the latter being groundwater composition from the Silurian Period onward. This
precipitated by subterranean groundwaters. Even in modern examples could have happened because chemical weathering of calcium
pedogenic and groundwater calcretes are hard to distinguish, with silicates (arguably enhanced by plant roots) produces free calcium
many deposits showing features of both types (e.g., Mack et al., 2000). and bicarbonate ions (e.g. Berner, 1992), for example (Eq. (1)):
The term ‘calcrete’ (which does not require reference to pedogenic or
2þ −
groundwater types) is thus relatively easy to apply in deep time because 2CO2 þ 3H2 O þ CaAl2 Si2 O8 →Ca þ 2HCO3 þ Al2 Si2 O5 ðOHÞ4 ð1Þ
the word in itself is not too restrictive, implying no specific set of
depositional circumstances, cause of precipitation or degree of biological Similarly, weathering of magnesium silicates by the same
involvement. The term ‘cement’ is usually applied to mineral pre- phenomena produces further bicarbonate ions, plus free magnesium
cipitates that fill voids or hold sediment together, but in this field of ions (Eq. (2)):
research ‘cement’ can also be used to describe a precipitate which
neither fills voids nor binds (for example ‘tufa cements’ often coat 4CO2 þ 4H2 O þ Mg2 SiO4 →2Mg
2þ −
þ 4HCO3 þ H4 SiO4 ð2Þ
biological substrates), and this broad usage is followed here.
Soils can be regarded as rocks altered by hydrolitic weathering
The net result of this silicate weathering by land plant ecosystems
(Retallack, 2007). This usage allows for discussion of carbonates
can be the precipitation of carbonate minerals (Eqs. (3) and (4)):
formed in and around ‘soils’ before the Devonian evolution of vascular
plants with root systems. For a wider discussion of the evolution of 2þ −
Ca þ 2HCO3 ↔CO2 þ H2 O þ CaCO3 ð3Þ
soils themselves, see reviews in Retallack (1990, 2007, 2008).
The following section begins with questions relating to terrestrial 2þ 2þ −
carbonates of the post-Silurian to modern periods, because it is from Ca þ Mg þ 4HCO3 ↔2CO2 þ 2H2 O þ CaMgðCO3 Þ2 ð4Þ
216 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

If vascular plants caused a significant post-Silurian increase in rates (Meheruna and Akagi, 2006), although the effect of pH alone on
silicate weathering, the geological record should show at least three silicate dissolution may be small (see Drever, 1994).
features compared to preceding times: (1) measurably increased Many studies have shown that plant roots, or their associated
carbonate precipitation (e.g. Eq. (3) and (4)) and (2) increased clay microbes, can cause release of nutrients from mineral grains (e.g. Boyle
levels (e.g. Eqs. (1) and (3)) decreased atmospheric CO2 levels (net and Voigt, 1973; Basu, 1981a,b; Meheruna and Akagi, 2006), although
result of Eqs. (1)–(4); see for example Ebelmen, 1845; Urey, 1952; the direct role of the plant root vs that of symbiotic micro-organisms is
Jutras et al., 2009; Berner, 1992; Berner, 1998). The question not often distinguished (see Kelly et al., 1998). This distinction (which
answered in this section is thus whether it is likely that the arrival can be difficult to make in practice) is important because it has been
of vascular plants (and their ecosystems; e.g. Retallack, 1997) shown that mycorrhizal fungi alone can cause an increase in mineral
increased chemical weathering of silicates (e.g. Basu, 1981a,b; Knoll weathering rates (e.g. Fisher, 1972; Griffiths et al., 1994; Jongmans
and James, 1987), leading to an increase in Ca2+ and HCO−
3 availability et al., 1997) such that high rates of silicate weathering may pre-date
and non-marine carbonate production (e.g. Alonso-Zarza and Tanner, the arrival of vascular plant roots (see for example Humphreys et al.,
2010)? in press). Modern mycorrhizal fungi allow access to mineral sources
that are otherwise physically inaccessible to plant roots (van Breemen
3.1. Significance of plant roots et al., 2000). For example, tubular pores some 3–10 μm wide in
feldspar and hornblende grains (which could be preserved as trace
Microbial and lichen crusts were arguably the only form of fossils) have been attributed to dissolution by organic acids produced
vegetative covering on land before the Ordovician (for example by mycorrhizal fungi (Jongmans et al., 1997; van Breemen et al., 2000).
Campbell, 1979; Golubic and Campbell, 1979; Shear, 1991; Knauth Mycorrhizal fungi are also known to help the release of essential
and Kennedy, 2009; see Section 5.3). A basic feature of land plants, at biolimiting phosphorus ions (e.g. Wallander et al., 1997; Ness and
least from the Middle Devonian onwards, is their formation of root Vlek, 2000; Wallander, 2000) and (importantly here) calcium
systems specialised for anchorage and for uptake of water plus ions (e.g. Blum et al., 2002) from relatively insoluble apatite crystals
biolimiting nutrients. Many of the essential cations (including Ca2+, (Ca5(PO4)3(F,Cl,OH)) as well as potassium ions from micas (e.g. Boyle
Mg2+ and Fe2+) come from rock weathering plus decomposition and and Voigt, 1973; Leyval and Berthelin, 1991; Bonneville et al., 2009).
recycling of organic matter (see for example Knoll and James, 1987; Apatite is a common accessory mineral in many silicate rocks. Blum et
Berner et al., 2003). The ability of plants to grow on rocky, ‘infertile’ al. (2002) estimated that ~35% of calcium in a US forest stream came
slopes itself suggests that plant roots can cause release of nutrients from weathering of apatite by mycorrhizae, with a further ~35% from
from rocks (Boyle and Voigt, 1973). It is thus unsurprising that the weathering of silicate minerals and ~30% from atmospheric sources.
Silurian arrival of vascular plants, followed by the Devonian arrival of Fungal dissolution of apatite (which occurs N103 times faster than
plant roots, has been linked with hypothesised enhanced chemical dissolution of hornblende, plagioclase and K-feldspar; see Blum et al.,
weathering (e.g., Berner, 1992, 1998; Algeo et al., 2001) leading to 2002) may be encouraged by uptake of calcium and phosphorus ions
enhanced soil carbonate formation (e.g. Alonso-Zarza and Tanner, by the fungal hyphae themselves, thereby reducing the concentra-
2010). tions of Ca and P around the dissolving apatite grains (Khasawneh and
One reason for believing chemical weathering of silicate rocks Doll, 1978; Ness and Vlek, 2000). Even so, Le Chatelier's Principle
might have increased since vascular plant diversification in the should only be applied with caution for predictions about increased
Devonian is the relatively high surface area and depth of penetration weathering rates of soil solutions upon silicate rocks (e.g. Drever,
such roots provide (for example Algeo et al., 1995). This is illustrated 1994). This is because soil solutions can often exist in states far from
by a global study of modern root systems by Canadell et al. (1996). equilibrium, meaning that weathering rates are unlikely to increase
They reported living vascular plant roots extending to depths of up to dramatically as a result of localised uptake of cations by microbes or
68 m (in the water-limited central Kalahari desert), while average plants. A precise mechanism for the microbial release of nutrients
maximum tree root depth was 7.0 ± 1.2 m, and global average root from apatite grains has yet to be elucidated. It is believed to involve
depth was 4.6 ± 0.5 m. Current understanding is that Silurian to early the production of oxalic acid by the fungi at their hyphal tips (e.g.
Devonian vascular plant roots were shallow, to depths of only a few Griffiths et al., 1994; Ness and Vlek, 2000; Wallander, 2000).
centimetres, but that depths of around 1 m were achieved by the
Middle to Upper Devonian (see Algeo and Scheckler, 1998; Algeo et 3.3. Did the origin of plant roots cause increased terrestrial carbonate
al., 2001; Kenrick and Davis, 2004). The geological record of soils abundance?
broadly supports this view: early Palaeozoic soils show weathering
concentrated at the surface, whereas mid-Palaeozoic and later soils, The influence of plant roots and (and associated ecosystems; e.g.
often with large root traces, show deeper and greater chemical Retallack, 1997) on local rock weathering, by exposing rock surfaces
weathering with clay contents that extend well below the surface and causing cation and anion release, therefore seems clear. But have
layers (Retallack, 1997). vascular plants brought about a net acceleration of chemical
weathering through geological time (e.g., Basu, 1981a,b; Knoll and
3.2. Chemical weathering in the rhizosphere and the role of fungi James, 1987; Berner, 1992; Algeo et al., 1995, 2001), leading to greater
global availability of Ca2+ and HCO− 3 ions (above and beyond that
The ‘rhizosphere’ (Hiltner, 1904; Darrah, 1993; Gregory, 2006; produced by lichens and fungi), and thus more post-Silurian non-
Hinsinger et al., 2006 and references therein) around plant roots marine carbonates?
exhibits conditions which favour locally enhanced chemical weath- This can be questioned on short timescales since soil also inhibits
ering. Firstly, the pH of the rhizosphere may differ from that in chemical weathering of bedrock by providing a protective clay
surrounding soil by as much as 1–2 units, largely because plant roots mineral barrier (Berner, 1992). But the relative rarity of fossil soils
maintain their charge balance by releasing protons (H+) whenever identified in the geological record probably reflects the ease with
they take up more cations than anions, and vice-versa (see Hinsinger which such clay barriers are eroded, so this barrier effect is less
et al., 2006). Root respiration also produces significant levels of CO2 significant on longer timescales.
(and hence of carbonic acid) which contributes to lowering of In addition to silicate weathering, carbonate bedrock dissolution
rhizosphere pH (Hinsinger et al., 2006) down to values as low as 3 would probably also have accelerated in the relatively deep (and acidic?)
(Arthur and Fahey, 1993). This acidity certainly assists in carbonate post-Middle Devonian rhizosphere. This could have caused some locally
bedrock dissolution and may also help to increase silicate weathering increased groundwater alkalinities and calcium ion activities, resulting in
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 217

re-distribution of carbonate minerals from relatively soluble bedrock to affected these factors, in addition to the major development of
soil and stream precipitates. This is despite global (including marine vascular plants with deep roots, must have exerted some influence on
environment), long term balance between carbonate mineral dissolution the temporal abundance and distribution of non-marine carbonates
and precipitation (see also Berner et al., 2003). (Fig. 2). Abundant calcium sulphates in some pre-Silurian rocks (for
However marine carbonate bedrock abundance and distribution example Melezhik et al., 2005b) provide a further reason for thinking
have not been constant through time: geological events which have that, at least locally, pre-Devonian terrestrial carbonate precipitation

Fig. 2. A timeline from the origin of the Earth to the end of the Devonian, with events of importance to non-marine carbonate processes noted. Note the breaks in scale between the
Hadean (from which no sedimentary rocks are known) and the Archean, and at the Precambrian–Cambrian boundary. Horizontal arrows and vertical lines of the same colour as the
associated text show dates or durations of events, respectively. Isua Gp sediments (e.g. Moorbath et al., 1973) are interpreted as marine, but unconformable surfaces and possible
'soils' (e.g. Johnson et al., 2009) are also reported from ancient times.
218 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Table 1
Non-marine carbonate examples of pre-Devonian age, with some cases (asterisked) where non-marine carbonates might be found on further examination. Gamagara Formation
laterites reported by Gutzmer and Beukes (1998) are not carbonate themselves, but terrestrial sideritic concretions are known from this same formation (Gutzmer, personal
communication, 2010). Petrographic features: Cg = coated grains (sizes and internal structures undifferentiated); Cr = circumgranular cracking; Cu = clay cutans; Cy = crystallaria;
D = desiccation cracks; E = evaporite pseudomorphs; Fg = floating grains; L = laminar crusts; M = massive carbonate bed; N = nodules; O = oncoliths; Oo = ooliths; P = pisoliths;
Pd = peloids; S = stromatolites; Sp = spar cement; Tp = tepee structures.

Reported age Unit Location Description References Selected


petrographic
structures

2.78 to 2.63 Ga Tumbiana Formation Pilbara, lacustrine stromatolites Bolhar and van Kranendonk S
Australia (2007); Awramik and Buchheim
(2009)
~ 2.7 Ga Klippan and Botheville South Africa lacustrine, fluvial and alluvial fan Buck (1980); date from D,Oo,S,Sp
Formations, Ventersdorp with stromatolites Armstrong et al. (1991)
Supergroup
2.6 Ga ‘pre-Chuniespoort Group’ South Africa silcrete and dolocrete crusts Martini (1994)
2.6 Ga basement complex South Africa soil carbonate Watanabe et al. (2000)
2.35 +/− 0.1 Ga Espanola Formation, Bruce Ontario, lacustrine limestone Veizer et al. (1992); Bernstein
“Limestone” Member Canada and Young (1990)
Palaeoproterozoic NW Canada soil and groundwater calcrete Aspler and Donaldson (1986) M,N
2.33 to 2.25 Ga Kuetsjarvi Sedimentary Fennoscandian lacustrine coastal plain, with calcretes Melezhik and Fallick (2001); Cg,E,L,N,P,S,Tp
Formation Russia and hydrothermal travertines Melezhik et al. (2004)
2.2 to 2.0 Ga Gamagara Formation South Africa *paleokarst-hosted laterites Gutzmer and Beukes (1998) L,P
1.85 Ga Missi Group–Amisk Group NW Canada possible pedogenic carbonate in the Holland et al. (1989)
contact ‘Flin Flon paleosol’
1.84–1.79 Ga Baker Lake Group Nunavut, carbonate cements in conglomerates, Mustard and Donaldson
Canada magnesite “chemogenic lake beds” (1990)
1.785 +/− 0.003 Ga Kunwak Formation, Baker Lake Nunavut, hydrothermal travertine Rainbird et al. (2006) E,L,N,Sp
Group Canada
1.3 to 1.865 Ga Murky Formation NW Canada alluvial fan carbonates Hoffman (1976) S
~ 1.27 Ga Dismal Lakes Group Arctic Canada speleothem cements—“marine, ? Glover and Kah (2006) P
phreatic, vadose”
Mesoproterozoic Kanuyak Formation NW Canada lacustrine and fluvial with calcretes Pelechaty et al. (1991); M, Oo,P,Pd,Sp,
Pelechaty and James (1991) Tp
1.199 +/− 0.07 Ga Poll a'Mhuilt Member, Stoer NW Scotland lacustrine limestone Upfold (1984); Turnbull et al.
Group (1996); Stewart (2002)
Proterozoic Staca paleosol (sub-Torridon NW Scotland palaeosol carbonate Retallack et al. (1994); Oo
Group) Williams and Schmidt (1997)
N 1.1 Ga Mescal Limestone Arizona, USA speleothem Skotnicki and Knauth (2007) L,Sp
1.2 Ga Mescal Limestone Arizona, USA calcrete Beeunas and Knauth (1985) L
~ 1.1 Ga Portage Lake Volcanics Michigan, USA calcrete Kalliokoski and Welch (1985) Cg,Cy,Fg,L,N,Sp
1.1 Ga Oronto Group Michigan, USA calcrete in fluvial and alluvial fan Kalliokoski (1986) Cg,Fg,N
sediments
1.087 Ga Oronto Group, Copper Harbor Michigan, USA alluvial fan and stream carbonate Elmore (1983); Davis and O,S
Conglomerate Paces (1990)
700–600 Ma Sarnyere Formation Mali calcrete Bertrand-Sarfati and Cg,Cr,D,N,L,Sp
Moussine-Pouchkine (1983)
Neoproterozoic Contact between Ravensthroat NW Canada *palaeokarst James et al. (2001)
Formation and Hayhook Formation
Neoproterozoic Contact between Sierras Bayas Tandilia *palaeokarst Gomez-Peral et al. (2007)
Group and Cerro Negro System,
Formation Argentina
Neoproterozoic Tsagaan Oloom Formation western *palaeokarst Shields et al. (2002)
Mongolia
Neoproterozoic Jbeliat Group cap carbonate Taudéni basin, *palaeokarst, cavities lined Shields et al. (2007)
NW Africa with barite
Neoproterozoic Krol Group and Infra-Krol Lesser palaeokarst and calcrete Jiang et al. (2003) Fg,N,P
(maximum age 590 Ma) Formation Himalaya,
India
Neoproterozoic to early Contact between Matjies River South Africa *palaeokarst Praekelt et al. (2008)
Cambrian (~ 547 Ma) Formation and Groenfontein
Formation
Neoproterozoic to early Unicoi Formation Appalachian alluvial fan conglomerates (*no Simpson and Eriksson (1989)
Cambrian orogen, USA terrestrial carbonates reported)
Early Cambrian Hawker Group Flinders palaeokarst at unconformity, with James and Gravestock (1990)
Ranges, possible cave pearls and calcretes
Australia
Cambrian Flinders paleosol calcrete Retallack (2008) E,N,Sp
Ranges,
Australia
Cambrian Davis Bore Conglomerate; Officer Basin, calcite cemented conglomerates, White and Youngs (1980); L,S,Tp
Observatory Hill Beds S. Australia playa lake stromatolitic carbonates Southgate et al. (1989)
Lower Cambrian Chapel Island Formation Newfoundland, calcrete Landing and MacGabhann Cy
Canada (2010); Landing et al. (1989)
Lower Cambrian White Point Conglomerate Kangaroo alluvial fan conglomerates (*no Daily et al. (1980)
and Boxing Bay Formation Island, South terrestrial carbonates reported)
Australia
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 219

Table 1 (continued)
Reported age Unit Location Description References Selected
petrographic
structures

Uppermost Lower Cambrian Eastern Anti- calcrete Buggisch and Siegert (1988) Fg
Atlas, Morocco
Middle Cambrian Mount Roosevelt British peritidal to supratidal calcrete, calcite Post and Long (2008) N
Colombia, cemented conglomerates
Canada
Cambrian Guaritas Sequence Brazil alluvial fan calcrete De Ros et al. (1994) L,N, Sp
Uppermost Cambrian La Flecha Formation Argentine calcrete Keller et al. (1989); Buggisch et Cu?,D,E,L,P,S,
Precordillera al. (2003) Tp
Cambro-Ordovician North China *palaeokarst and palaeosols (no Meng et al. (1997)
terrestrial carbonates described)
Cambro-Ordovician Haima Group Oman alluvial fan conglomerates (calcretes Heward (1989)
said to be absent; calcite cement
present)
Lower Ordovician Puddinga Beds Sardinia, Italy *Fluvio-lacustrine and alluvial fan Martini et al. (1991)
conglomerates (no terrestrial
carbonates reported)
Lower Ordovician Beekmantown Group Ontario, vadose cements in palaeokarst (N.B., Dix et al. (1998)
Canada some cements may post-date
Ordovician)
Lower to Middle Maggol Limestone Taebaeksan *palaeokarst (no terrestrial carbonates Ryu et al. (2005)
Ordovician Basin, South reported)
Korea
Middle Ordovician Newmarket Limestone West Virginia, vadose cements in marine carbonates Read and Grover (1977); Grover
USA and Read (1978)
Middle Ordovician Chickamauga Formation Alabama, USA palaeokarst with meteoric cements Tobin and Walker (1994) Sp
Middle Ordovician Dunn Point Formation Nova Scotia calcrete (palaeosol carbonate) Jutras et al. (2009)
Ordovician (Ashgill) Juniata Formation Pennsylvania, calcrete (palaeosol carbonate) Retallack (1985); Retallack Cu,N
USA (2001)
Ordovician Huanghuachong Formation South China *palaeokarst (no terrestrial carbonates Jia-yu and Johnson (1996)
(Caradoc to Ashgill; reported)
capped by early Silurian)
Silurian Dronningveien Siltstone Norway fluvial red beds with concretions Turner (1974); Davies et al. N,P
Member, Ringerike Group (2005)
Silurian Niagara Group Michigan, USA calcretised stromatolites Sears and Lucia (1979) L,P
Silurian (Wenlock) Mill Cove Formation Ireland playa lake muds with incipient calcretes Wright et al. (1992) D,N
Silurian (Ludlow–Pridoli) Bloomsburg Formation Pennsylvania, pedogenic calcrete Mora et al. (1991) Fg,N
USA
Siluro–Devonian Lower Old Red Sandstone Wales and calcretes Allen (1974); Marriott and Cy,N,L
England Wright (1993)

was not limited by low calcium ion activities from a lack of plant root- 2004; Melezhik et al., 2005b). Nevertheless, carbonic acid is
associated weathering. responsible today (and presumably for much of Earth history) for
much carbonate dissolution, and results from mixing of carbon
4. Precipitation of calcretes, SSL carbonates and speleothems dioxide and water (Eq. (5)).
before plants
CO2 þ H2 O→H2 CO3 ð5Þ
The role of high pCO2 soils in precipitation of modern soil, cave and
spring calcites is well studied. Given there are examples of pre- This carbonic acid dissociates twice, first producing bicarbonate
vascular plant terrestrial carbonates preserved in the rock record (HCO− + 2−
3 ) and H ions (Eq. (6)), and then carbonate (CO3 ) and further
(Table 1), one might question how cave (Fig. 3A) and soil (Fig. 4) +
H (Eq. (7)).
calcites formed, and how Silurian and older carbonate depositional
patterns might have differed from those of later times? Here it is þ −
H2 CO3 ↔H þ HCO3 ð6Þ
helpful to examine possible precipitation mechanisms for spe-
leothems, SSL carbonates and calcretes both including and excluding − þ 2−
HCO3 ↔H þ CO3 ð7Þ
high pCO2 soils.
Groundwaters with pH higher than ~9 will have used up all
4.1. Calcite dissolution and precipitation available H+ ions and so predominantly contain carbonate (CO2− 3 )
species, whereas those with pH below ~ 5.5 have a majority of H2CO3
The common model for precipitation of abiotic calcite speleothem molecules over carbonate and bicarbonate ions. Waters with pH of 4
in caves requires an organic-rich high pCO2 soil zone above carbonate or less will have had all of their carbonate converted to carbon dioxide
bedrock (Fig. 3A). To understand why, it is necessary to consider what (Eq. (8); see also Andrews et al., 2004).
governs calcite dissolution and precipitation from carbonic acid-rich
groundwater solutions (see also Andrews et al., 2004). This discussion − þ
HCO3 þ H ↔H2 O þ CO2 ð8Þ
assumes the principle acid in question is carbonic, which may not
always have been the case. For example, local H2SO4 production is Carbonic acid coupled with calcium ions from silicate (see
likely to have increased following oxygenation of the atmosphere and Section 3.1) or carbonate bedrock weathering can thus produce
oxidation of sulphides in the Palaeoproterozoic (e.g. Bekker et al., calcite.
220 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Whether a low ionic strength solution will abiotically precipitate a The solubility product Ksp is an equilibrium constant (e.g. for
mineral is measured by the saturation index (SI — see Eq. (9)). SI calcite, Eq.(10)).
values N1 indicate supersaturation with respect to that mineral and it
may precipitate. Conversely, values of b1 indicate undersaturation 2þ 2−
Ksp ¼ aCa :aCO3 =aCaCO3 ð10Þ
and dissolution is expected.

SI ¼ IAP=Ksp ð9Þ where (for example) aCa2+ and aCO2− 3 are the activities of the
reactants (activity is proportional to concentration), and aCaCO3 is the
where IAP is the ion activity product and Ksp the solubility product of activity of the solid calcium carbonate product (solids are assigned a
the mineral in question (explained further below). value of 1; for further information see Andrews et al., 2004). Because

Fig. 3. Speleothem formation with and without organic-rich soils. Diagrams are not to scale. (A) An organic-rich soil above the cave provides high pCO2 acidic soil solutions which
dissolve carbonate bedrock and precipitate speleothem type carbonate when emerging in a lower pCO2 atmosphere. For degassing and subsequent precipitation to occur the cave
must be open to the (lower pCO2) atmosphere. This model accounts for most speleothem precipitation in warm, humid climates today. (B) A high pCO2 atmosphere (for example in
the Palaeoproterozoic) might allow bedrock dissolution, but degassing and precipitation will not occur unless solutions equilibrate with a lower pCO2 atmosphere. (C) Calcite
precipitation under ice cover by warming of waters and freeze-out. How warming causes precipitation is shown in the plot of Ca concentration vs pCO2 (adapted from Dreybrodt,
1982). In this plot, the curved lines show Ca concentration in equilibrium with pCO2 for given temperatures. For example, waters at 0 °C with a Ca concentration of X1 and pCO2 of Y
are in equilibrium at point A. When these 0 °C calcite-saturated waters enter bedrock and warm to 4 °C, the Ca concentration of the solution remains the same (X1) but pCO2 must
increase to maintain equilibrium at the elevated temperature, moving the composition of the solution along a horizontal line to point B. Degassing of CO2 begins on entering the cave,
returning the composition of the solution to point A, but now at elevated temperature, such that calcite must precipitate to decrease the calcium concentration to X2 and return the
system to equilibrium at point C. For full details see Dreybrodt (1982). When waters freeze Ca ions preferentially remain in the liquid which can also lead to local calcite
supersaturation. (D) Carbonic acid can be produced by mixing of groundwaters with deep-sourced CO2. Such groundwaters degas on emergence in a lower pCO2 atmosphere,
thereby leading to calcite precipitation. (E) Calcite supersaturation caused by the common-ion effect. Solutions saturated with respect to calcite may become supersaturated if extra
calcium is supplied from sources such as dissolution of calcium sulphates or incongruent dissolution of dolomite.
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 221

Fig. 3 (continued).

activities are used rather than concentrations, solubility products can rise above 1 favouring calcite precipitation. According to Le
should be constant for all environments at a given temperature Chatelier's Principle any factor which causes the equilibrium reaction
(though calcite solubility decreases with rising temperature). shown in Eq. (13) to shift towards the right (such as degassing of CO2
The ion activity product (IAP) for calcite can be calculated using or addition of calcium ions) will cause supersaturation of an already
Eq. (11): saturated solution and precipitation of calcite.
2þ 2−
IAP ¼ aCa :aCO3 ð11Þ 2þ −
Ca þ 2HCO3 ↔CaCO3 þ CO2 þ H2 O ð13Þ
The saturation index (SI) for calcite in a given solution at a given
temperature is thus dependent on the activities of calcium and Continued addition of calcium ions does not result in an indefinite
carbonate ions in the solution (Eq. (12)). solution pH rise: the excess Ca2+ ions will neutralise available HCO− 3
ions, and calcite precipitates until equilibrium is restored (Eq. (13)).
2þ 2− This reaction cannot easily be driven too far to the right since carbonic
SI ¼ aCa :aCO3 =Ksp ð12Þ
acid is produced to react with the calcite. Given sufficient supply of
Consequently by adding calcium and carbonate or bicarbonate calcium and carbon dioxide then groundwater pH is buffered by this
ions to a solution, (Eq. (12)) or increasing its temperature, SI values equilibrium reaction.
222 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Fig. 4. Quaternary pedogenic calcrete from central Greece. Atmospheric dust trapped around plants has supplied the calcium. This calcium-rich dust is dissolved at the top of the soil
profile by carbonic and organic acids produced by vascular plants and associated organisms. Calcite has precipitated lower down in the profile because soil waters became
supersaturated here through the evaporation, CO2 degassing and local effects around plant roots. The first (highest) precipitates formed an unconsolidated powder, followed lower
down by nodular growth and finally by coalesced nodules forming a ‘plugged’ horizon. Such zoned profiles are usually found in fine-grained sediments that accumulated through
trapping by vegetation and may usually require vascular plant involvement to form.

4.2. Calcite precipitation models including carbonic acid-rich, high atmospheric pCO2 that arguably pertained through much of the
pCO2 soils Proterozoic to Ordovician (see Section 3.1 and Retallack, 1997),
reaction of atmospheric CO2 with cold water would readily have
Waters percolating through organic-rich high pCO2 soil zones allowed dissolution of limestone or dolomite bedrock until the waters
above caves from the Middle Devonian onwards can be expected to themselves became saturated (Fig. 3B). However these waters would
have dissolved mainly biogenic CO2 and become acidic (Eq. (5); not degas and precipitate carbonate minerals unless they reached a
Fig. 3A), enabling dissolution of the calcium carbonate bedrock. When lower pCO2 environment. This leads us to an important conclusion: a
such drip waters reach a lower pCO2 cave atmosphere they degas CO2 high pCO2 atmosphere in a pre-Silurian world would not be enough by
and become supersaturated with respect to calcite which then itself to cause bedrock dissolution and re-precipitation as speleothem.
precipitates (Fig. 3A; see also Holland et al., 1964; White, 1976; Bar-
Matthews et al., 1991). A similar process causes the distinct vertical 4.3. Calcite precipitation mechanisms without high pCO2 soils
horizonation of pedogenic calcrete-containing soil profiles, where
carbonate is leached from the upper parts of the profile by high pCO2 There are mechanisms capable of causing non-marine calcite
waters, and re-precipitated lower down when supersaturation is precipitation without aide from vascular plants. These processes must
reached (Fig. 4; see Esteban and Klappa (1983) for further examples). have been responsible for many of the Silurian and older deposits
Similarly, some groundwater calcrete precipitation (e.g. Fig. 5 and listed in Table 1.
Goudie, 1996; Nash and Smith, 2003), and calcite precipitation at
springs (see Fig. 6 and Bono et al., 2001; Andrews, 2006) is caused by 4.3.1. Degassing mechanisms: temperature rise
equilibration of relatively CO2-rich waters with the lower pCO2 Water in modern cold and high altitude climates may hold
atmosphere. sufficient CO2 to produce the necessary carbonic acid (see Section 4.1)
Prior to the first land plants in the Ordovician (or perhaps the without need for soil zone CO2 input (Dreybrodt, 1982; Atkinson,
development of deep plant roots in the Middle Devonian), conditions 1983; Frisia and Borsato, 2010), yet the deposits that result are
in caves may well have been different. In conditions of high volumetrically much less impressive than those formed under
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 223

presence of an organic-rich soil above the cave. Tuccimei et al. (2006)


suggest periodic outgassing of deep-sourced CO2 can inhibit spe-
leothem precipitation because waters become too acidic. However
this could only be the case where supply of calcium ions is too low for
the reaction to be buffered (see Eq. (13)). Degassing of deep-sourced
CO2 can also cause carbonate precipitation around springs (e.g.
‘thermogene travertine’ of Pentecost and Viles, 1994; D'Alessandro et
al., 2007) without causing carbonate precipitation through high water
temperatures (e.g. Yoshimura et al., 2004) and might explain some
Palaeoproterozoic hot spring carbonate mounds (Melezhik and
Fallick, 2001).

4.3.3. Degassing mechanisms: turbulence


One of the most important mechanisms for carbonate precipita-
tion around modern rapids and waterfalls (Fig. 6) is degassing of CO2
due to turbulence. Zhang et al. (2001) and Chen et al. (2004)
suggested that calcium carbonate preferentially builds up at waterfall
sites as a result of “sudden hydrological changes”, including enhanced
air–water contact and enhanced flow velocities, thereby resulting in
more rapid CO2 degassing than is found along stretches of streams
with more shallow gradients. Carbon dioxide degassing influenced by
turbulence can occur from springs and spring-fed lakes (Coxon, 1994)
but is limited where pCO2 in the lake and atmosphere are close to
equilibrium (Fig. 6). Environments where degassing is encouraged
through turbulence (such as waterfalls) also tend to be quite dynamic,
with low potential for preservation of any resulting deposit on
geological timescales.

4.3.4. Evaporation
Evaporation causes precipitation by concentrating ions in the
Fig. 5. Quaternary groundwater calcretes in alluvial fan sediments, central Greece.
remaining solution. This is well known as a factor in precipitation of
Calcium was mainly supplied via wind-blown dust and from carbonate bedrock clasts evaporites such as gypsum and “exotic” carbonates (Atkinson, 1983;
in the conglomerates. Calcrete (upper arrow) has precipitated in the less well-drained Harmon et al., 1983) like huntite (Mg3Ca(CO3)4) as well as calcite and
finer grained facies because waters percolate more quickly through coarser clast- aragonite (Thrailkill, 1971) in caves. So long as evaporated water can
supported channel conglomerates (lower arrow). Boundaries of groundwater calcretes
be transported out of the system, mineral precipitation through
are generally sharply defined and they lack the vertical horizonation of pedogenic
forms (see Fig. 3). Similar groundwater calcretes do not require vascular plant evaporation can occur when cold air blowing through the cave is
involvement in their formation, and supersaturation is commonly achieved through heated by the warmer walls (e.g. Harmon et al., 1983). Evidence for
mixing of groundwaters (i.e., the common-ion effect). this process is provided by speleothems that are orientated in the
direction of strong air flow. Even so, the ability of evaporation to
warmer soils providing more biogenic CO2 (see for example Frisia and produce significant amounts of speleothem is questionable (Harmon
Borsato, 2010). Degassing of CO2 following temperature rise can cause et al., 1983).
calcite to precipitate (Dreybrodt, 1982). In this model, waters beneath Calcite supersaturation resulting from evaporation has been
glaciers at temperatures of ~ 0 °C, are able to dissolve calcite. When suggested for precipitation of much carbonate from lakes (e.g. Müller
percolating through cavities in limestone bedrock, these waters warm et al., 1972), often from very shallow and playa lake settings (for
up, and can eventually precipitate calcite owing to its decreasing example Eugster and Hardie, 1975; Risacher and Eugster, 1979; White
solubility with increasing temperature (Fig. 3C). This can happen at and Youngs, 1980; Southgate et al., 1989). Candy et al. (2006) reported
rates of up to 10− 3 cm yr− 1 under modern atmospheric conditions calcrete nodules from the Middle Pleistocene of eastern England,
(Dreybrodt, 1982). For such a process to work, dissolution must demonstrating that the requirement for calcrete formation by
occur in open system conditions, waters must be warmed by several evaporative processes is not just aridity but strong seasonality, with
degrees and cave air pCO2 must be similar to that of the free periods of evaporation exceeding rainfall (perhaps in the summer) and
atmosphere. This mechanism can also work in the absence of vice-versa in the winter. In some vadose zones of alluvial fans, ‘gully-
vegetation, such as when waters at 0 °C enter warm caves during the bed cements’ (Mack et al., 2000) preferentially coat the undersides
winter. Warm waters entering cooler caves may dissolve calcite in of clasts in channel-fill conglomerates, suggesting they form by
the summer, but probably at slower rates than winter deposition evaporation of bicarbonate-rich run-off. Once the interstices in conglo-
proceeds (Dreybrodt, 1982). Warming of waters also causes CO2 merates are cemented, a laminar calcrete sheet forms across the top.
degassing and calcrete precipitation in sub-arctic soils where Mack et al. (2000) suggest gully-bed cements form in the absence of
relatively warm temperatures at shallow depths favour precipitation plants. Evaporation and evapotranspiration (see Section 4.4.3 below)
close to the surface in the summer (Swett, 1974; Dijkmans et al., 1986). may not be so important for some groundwater calcretes precipitated
well below the surface (e.g. Nash and Smith, 2003).
4.3.2. Degassing mechanisms: deep-sourced CO2
Speleothem or SSL calcium carbonates may also precipitate in the 4.3.5. Common-ion effect
absence of vegetation if CO2 is provided by volcanic or other deep The common-ion effect occurs in polymineralic systems with
sources such as CO2 outgassed during metamorphism of sediments calcite and gypsum or calcite and dolomite where solutions are
(Fig. 3D; e.g. Duliński et al., 1995; Pentecost and Viles, 1994; initially saturated with respect to calcite (Fig. 3E). Incongruent
D'Alessandro et al., 2007). However speleothems in modern volcanic dissolution of dolomite or dissolution of gypsum adds extra calcium
settings seem to be rare (Forti, 2005), and seemingly require the ions to the solution, which then becomes supersaturated with respect
224 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Fig. 6. Carbonate deposition in fluvial settings in the absence of vascular plants (based on the perched springline model of Pedley et al., 2003). Rain and snow falling on high ground
percolate down through sediments until they reach an impermeable horizon which then causes groundwaters to flow along the contact until emerging at a spring. In modern
settings, groundwaters have high pCO2 levels as a result of contact with organic-rich soils, and thus much degassing and calcite precipitation occurs at springs. Prior to the Silurian,
soils were thinner and less organic-rich, and atmospheres often had higher pCO2, such that degassing and travertine deposition at springs is likely to have been less significant than
found in Quaternary examples. Trees frequently provide obstacles to flow, thereby increasing turbulence (and hence CO2 degassing) leading to formation of barrages and upstream
ponding of waters to give rise to paludal facies. Prior to vascular plants, such barrage formation is likely to have been less frequent and paludal facies may well have been absent.
Nevertheless cyanobacteria may have produced travertine screens similar to those of Quaternary examples (see Andrews and Brasier, 2005; Brasier et al., in press). With time,
cascade tufas tend to become incised by the down-cutting streams and then abandoned. Travertine precipitation may resume lower down the valley. Such deposits are most likely to
be preserved and recognised as microbial carbonate clasts in conglomerates).

to calcite, causing the latter to precipitate (Eq. (13); see also Wigley, et al. (2004) reported loose, uncemented ‘cryogenic cave calcite’
1973a,b; Lohmann, 1988). Clear examples of this process, which can crystals from the Pleistocene of Central Europe, and Žák et al. (2008)
produce quite substantial deposits, are found where calcite has suggest slower freezing produces larger crystals (up to 20 mm
precipitated in gypsum caves (Calaforra et al., 2008) and where diameter) than fast freezing. Cryptocrystalline carbonate powders
carnotite (K2(UO2)2V2O8.3H2O), which will not form in the presence have been reported from the surface of ice stalagmites and stalactites
of carbonate ions, is found in calcrete. In the latter case, the presence (e.g. Lacelle et al., 2009), but it seems amounts of modern precipitate
of carnotite indicates carbonate ions had been efficiently stripped produced are often quite small, and carbonate flowstones, stalagmites
from groundwaters by precipitating with the excess calcium (Carlisle, and stalactites are not likely to be formed by this process.
1983; see further calcrete examples in Arakel and McConchie, 1982; Freeze-out occurs in modern lakes including those of Antarctica (e.g.
Arakel, 1986; Spötl and Wright, 1992; Nash and McLaren, 2003; Nash Matsubaya et al., 1979) and Canada (e.g. Renaut and Long, 1987), and
and Smith, 2003; Jutras et al., 2007). The common-ion effect has also may have caused vaterite (CaCO3 polymorph) precipitation from an
been suggested as a cause of precipitation of Pleistocene lake tufas in Arctic spring (Grasby, 2003). Canfield et al. (1983) who studied lakes in
Mono Lake, California, USA (Dunn, 1953) and calcite precipitation in Iowa, USA, reported that alkalinity which had risen due to this ‘freeze-
the presence of gypsum from cold springs in ‘dominating Polar out’ effect fell immediately after ice melting, proving the two were
conditions’ of the Canadian Arctic (Omelon et al., 2006; see also linked. Further, Cerling (1984) has suggested this process can cause
Grasby, 2003) as well as hot spring carbonates (Pentecost, 1995). In calcite precipitation in soils. Courty et al. (1994) later reported
an oxic atmosphere, oxidation of sulphides such as pyrite (FeS2) can laminated pendant calcite cements formed by freeze-out below coarse
produce both gypsum and sulphuric acid which may lead to karst clasts in fluvio-glacial sediments, and Vogt and Corte (1996) found
dissolution (e.g. van Everdingen et al., 1985; Hill, 1990) coupled with coatings on grains resulting from melting and re-freezing of permafrost
calcite precipitation via the common-ion effect. with calcium supplied by wind-blown dust. Studies of stable carbon
isotopes have been used to illuminate this process, including the study
4.3.6. Freeze-out of Cerling (1984) who postulated that modern soil carbonates
Freezing of water (Fig. 3C) can also cause calcite (or aragonite; e.g. precipitated from freezing soils show δ13C values strongly influenced
Aharon, 1988) supersaturation and precipitation because calcium by atmospheric compositions because biological respiration rates are
preferentially partitions into the remaining liquid. This happens from low. However this assumes the carbonate precipitates in equilibrium
low ionic strength solutions where ice melts and refreezes below and with surrounding dissolved inorganic carbon and CO2. Clark and Lauriol
within glaciers sliding over limestone bedrock (e.g. Hallet et al., (1992) showed cryogenic calcite can be significantly enriched in 13C as a
1978; Sharp et al., 1990; Fairchild et al., 1993) and from sea ice. In the result of kinetic effects. They reported values of up to +17‰ (PDB). But
latter case calcium carbonate precipitates at just below − 1.9 °C, an important conclusion here is that little or no biological influence is
followed by sodium sulphate (− 8.2 °C) and calcium sulphate at required to precipitate soil carbonates within the freezing zone.
− 10 °C (Richardson, 1976; Anderson and Jones, 1985; Papadimitriou
et al., 2004). Aharon (1988) suggested that weathering of calcium 4.3.7. Precipitation from boiling springs
silicates by carbonic acid (Eq. (1)) was the source of calcium and Boiling water is implicated in deposition of some hydrothermal
bicarbonate ions for aragonite precipitation below Antarctic ice. carbonates since boiling (in addition to the high temperature and thus
Killawee et al. (1998) and Papadimitriou et al. (2004) suggested low calcite solubility) brings about degassing of CO2 (for example
degassing of CO2 bubbles aides calcite precipitation. From caves, Žák Arnórsson, 1989). Even so, hot springs can often deposit calcite at
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 225

Table 2
Petrographic changes in spring, stream or lacustrine (‘SSL’) carbonates, calcretes and speleothems through time. For further speleothem fabrics see Frisia and Borsato (2010). This
table shows many of the fabrics found in non-marine carbonates and indicates likely maximum ages, particularly where biological involvement is required. Recognising pre-
Devonian non-marine carbonates may be hindered by the absence of fabrics considered diagnostic in more recent deposits, such as rhizoliths and rhizolites.

Fabric Spring, Stream and Description Diagram Organism Likely maximum age Reference(s)
Lake (SSL), Calcrete involvement? for example(s)
or Speleothem

Nodules/glaebules Calcrete Spherical to irregular indurated Abiological argillo- Origin of smectite Wieder and
concentrations of carbonate, pedoturbation clays (Archaean) Yaalon (1974)
generally mm to several cm in size

1 cm

Abiogenic laminar Calcrete Laminar calcrete crusts from a few Abiological Origin of calcretes Arakel (1982)
crusts mm to several centimetres thickness (Archaean)

1 cm

Tepees Calcrete Antiformal structures produced by Abiological Origin of calcretes Assereto and
buckling during expansion (Archaean) Kendall (1977)

1 cm
Floating grain fabric Calcrete Silt and sand sized grains floating Abiological Origin of calcretes Esteban and
in a micritic matrix where the matrix (Archaean) Klappa (1983),
appears to be supporting the grains Tandon and
Friend (1989)

1 cm
Circumgranular Calcrete Non-tectonic fractures around grains Abiological Origin of clay Esteban and
cracking produced by shrinking and swelling minerals (Archaean) Klappa (1983)
of clay minerals during desiccation
and hydration

1 mm
Crystallaria Calcrete Calcite-filled cracks, mostly formed Abiological Origin of calcretes Wright and
through desiccation and expansion (Archaean) Tucker (1991)

1 mm
Clay cutans Calcrete Ooid-like coatings of clay around a Abiological Origin of clay Esteban and
grain nucleus minerals (Archaean) Klappa (1983)

1 mm
Mosaics Calcrete Mosaics of calcite crystals of Abiological Origin of calcretes Wright and
different sizes (Archaean) Tucker (1991)

1 mm
Peloids Calcrete SSL Rounded micrite pellets, various Abiological (but see Origin of authigenic Tucker and
origins also faecal peloids) precipitates Wright (1990)
(Archaean)

500 µm

(continued on next page)


226 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Table 2 (continued)
Fabric Spring, Stream and Description Diagram Organism Likely maximum age Reference(s)
Lake (SSL), Calcrete involvement? for example(s)
or Speleothem

Pisoliths Calcrete SSL Concentrically laminated Abiological argillo- Origin of authigenic Esteban and
sphaeroids N 2 mm diameter pedoturbation and precipitates Klappa (1983),
down slope (Archaean) Tandon and
movement Friend (1989)

2 mm
Ooids Calcrete SSL Concentrically laminated Abiological argillo- Origin of authigenic Tucker and
sphaeroids b 2 mm diameter pedoturbation precipitates Wright (1990)
(Archaean)

1 mm
Columnar calcite SSL Elongate crystals wider than Abiological Origin of Kendall and
ten microns speleothems Broughton (1978),
(Archaean) Brasier et al.
(in press)

1 mm

Cave popcorn Speleothem Nodules found on smooth Abiological Origin of Thrailkill (1976)
flowstone, stalagmite or other speleothems
speleothem often containing (Archaean)
near-hemispherical laminations
which thin away from the axis
200 µm

Paper-thin rafts SSL (hot springs) Rafts of crystals with flat upper Abiological Origin of hot Guo and Riding
surfaces produced by rapid springs (Archaean) (1998), Chafetz
precipitation at the water surface et al. (1991)

5 cm

Coated bubbles SSL (hot springs) Millimetre to centimetre-sized Abiological Origin of hot Chafetz et al.
bubbles preserved by rapid springs (Archaean) (1991), Guo and
precipitation of calcite or aragonite Riding (1998)
in hot spring settings, often found
on undersides of rafts

5 cm

Ray/Feather crystals SSL (hot springs) Coarse elongate crystals oriented Abiological (Guo Origin of hot Chafetz and Folk
perpendicular to the depositional and Riding, 1992) springs (Archaean) (1984), Guo and
surface, sometimes forming Riding (1992)
conical radiating patterns

Crystal shrubs SSL (hot springs) Dendritic shrub-like crystals Bacteria (often in hot Origin of bacteria Chafetz and Folk
composed of upward radiating spring environments) (Archaean) (1984), Pentecost
branches (1990)

Sparry calcite fans SSL Speleothem Fan-shaped calcite crystals Abiological Origin of non- Love and Chafetz
(neomorphic marine carbonates (1988), Freytet and
or primary) (Archaean) Verrecchia (1998);
Brasier et al.,
in press)
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 227

Table 2 (continued)
Fabric Spring, Stream and Description Diagram Organism Likely maximum age Reference(s)
Lake (SSL), Calcrete involvement? for example(s)
or Speleothem

Needle Fibre Calcite Calcrete SSL Acicular calcite crystals with Abiological (strong Origin of non- Verrecchia and
lengths at least four times the evaporation, high marine carbonates Verrecchia (1994)
width: monocrystalline supersaturation) (Archaean)
micro-rods and polycrystalline
chains of Verrecchia and
Verrecchia (1994)
0.3 µm
Non-marine SSL Laminar microbial carbonates Cyanobacteria Origin of Andrews and
stromatolites formed in non-marine settings cyanobacteria Brasier (2005),
(Archaean) Irion and Muller
(1968)

3 mm

Micrite coated SSL (Cyano)bacterial sheaths coated Bacteria (including Origin of bacteria Freytet and
filaments in calcite. Organic materials often cyanobacteria) (Archaean) Verrecchia (1998)
decayed, leaving an external mould.

20 µm

Sparry calcite fans SSL Speleothem Fan-shaped calcite crystals Cyanobacteria Origin of Freytet and
cyanobacteria Verrecchia (1998)
(Archaean)

1 mm

Spherulites Calcrete “Calcitic fibro-radial spherulitic Cyanobacteria Origin of bacteria Verrecchia et al.
polycrystals” (Verrecchia et al., and bacteria (Archaean) (1995), Wright
1995) found in laminar crusts et al. (1996)

30 µm

Oncoliths SSL Layered, unattached carbonate Cyanobacteria Origin of Tucker and Wright
coated grain with a cortex of and algae cyanobacteria (1990), Nickel
non-concentric, partially (Archaean) (1983)
overlapping laminae

2 mm
Needle Fibre Calcite Calcrete SSL Acicular calcite crystals with Fungi, lichen Origin of fungi Verrecchia and
lengths at least four times the (Mesoproterozoic?) Verrecchia (1994)
width: coupled rod forms of
Verrecchia and Verrecchia
(1994)

Pisoliths Calcrete SSL Concentrically laminated Fungi, lichen Origin of fungi Beier (1987)
sphaeroidsN 2 mm diameter (Mesoproterozoic?)
composed of Needle Fibre
Calcite

2 mm
(continued on next page)
228 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Table 2 (continued)
Fabric Spring, Stream and Description Diagram Organism Likely maximum age Reference(s)
Lake (SSL), Calcrete involvement? for example(s)
or Speleothem

Lichen stromatolites Calcrete Laminar calcretes formed Lichen Origin of lichens Klappa (1979)
through repeated cycles (Neoproterozoic?)
of lichen action (both biophysical
and biochemical) on bedrock 1 cm
Coated macrophyte SSL External moulds of vascular Vascular plants, Origin of bryophytes Pedley (1990)
stems plant stems bryophytes (Ordovician) and
vascular plants
(Silurian)

1 cm
Faecal peloids Calcrete SSL Rounded micrite pellets Faeces of Origin of terrestrial Esteban and Klappa
resulting from invertebrate invertebrates invertebrates (1983)
defaecation in the soil (Silurian)

500 µm
Insect cocoons Calcrete Calcified insect cocoons Insects Origin of insects Esteban and Klappa
(Silurian) (1983)

1 cm
Rhizoliths Calcrete Organosedimentary Vascular plants Origin of vascular Klappa (1980)
structures produced by roots plant roots (Middle
Devonian)

30 cm
Laminar crust Calcrete Laminar calcrete crusts Calcification of Origin of vascular Wright et al.
rhizolite exhibiting ‘tubular fenestrae’ root mats plant roots (Middle (1988), Alonso-
Devonian) Zarza (1999)

2 mm
Alveolar septal Calcrete ‘Cylindrical to irregular pores, Vascular plant roots Origin of vascular Esteban and
fabric which may or may not be filled and mycorrhizae plant roots and Klappa (1983),
with calcite cement, separated mycorrhizal fungi Wright and
by a network of anastomosing (Middle Devonian?) Tucker (1991)
micrite walls’ (Esteban and
Klappa, 1983).
0.2 mm
Microcodium b Calcrete Calcified plant root cells Vascular plant roots Carboniferous Alonso-Zarza
(Kabanov, et al., et al. (1998)
2008)

0.3 mm
Microcodium Calcrete Calcite crystals resembling Plant roots (Klappa, Cretaceous Klappa (1978),
biological ‘cells in palisades 1978) or ‘mycelial (Kabanov, et al., 2008) Kabanov et al.
around small nucleii’ saprotrophic organism’ (2008)
(Kabanov et al., 2008)

0.5mm

much lower temperatures (e.g. 73 °C maximum reported by Fouke et from Italy), so this process seems responsible for few terrestrial
al., 2000, for Mammoth Hot Springs, Yellowstone, USA; 95 °C deposits today. Hyperthermophile archaebacteria are able to tolerate
maximum reported by Chafetz and Folk, 1984, for travertines of the temperatures up to ~120 °C and may therefore have been involved in
western USA; 39 °C reported by Guo and Riding, 1994, for hot springs precipitation of some hydrothermal carbonates (see, for example
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 229

Fig. 7. Middle Proterozoic redbeds and associated carbonates at Nipigon, Ontario, Canada. (A) Calcite which has replaced gypsum crystals in red mudstones. (B) Calcrete forming a
sheet with some lamination. (C) Stromatolite found in terrestrial (lacustrine) facies. (D) Overview of outcrop at Nipigon, Ontario, showing red mudstones at base, overlain by
silicified stromatolitic limestones with collapse breccias, buff marls and red mudstones with ptygmatic folding and buff-coloured calcitic mudstones with vugs formerly occupied by
gypsum. This section is cut by a dolerite sill (arrowed). Person for scale. (E) Conceptual model for the Nipigon section, with stromatolites around a lacustrine shore, surrounded by
aeolian and fluvial muds containing calcretes and gypsum nodules.

Chafetz and Folk, 1984; Guo et al., 1996; Fouke et al., 2000). Some Pentecost, 1981) but it can lead to tufa precipitation (Eq. 13) in less
authors, on the basis of petrography, suggest that up to 90% of some turbulent settings such as those found in some lakes (e.g. Ohlendorf
hydrothermal travertines are directly or indirectly produced by and Sturm, 2001). Warm lake waters may aide calcite precipitation in
bacteria (e.g., Chafetz and Folk, 1984). However a realistic biogeo- such cases (for example Hodell et al., 1998). Cyanobacteria make
chemical mechanism through which bacteria could cause precipita- another important contribution to SSL carbonate precipitation by
tion has not been detailed. producing extracellular polymeric substances (EPS), known to
promote calcite crystal nucleation in solutions of low ionic strength
4.4. Biological influences on non-marine carbonates (for example Emeis et al., 1987; Riding, 2000; Rogerson et al., 2008;
Pedley et al., 2009). In some cases carbonate build-up at waterfall sites
Cyanobacteria, fungi, lichens and vascular plants all influence non- may not be possible without a cyanobacterial substrate to hold the
marine carbonate build-up, and have likely altered physical appear- precipitate in place (for example Brasier et al., 2010).
ances of these carbonates through time.
4.4.2. Fungi and lichens
4.4.1. Cyanobacteria Fungi are not only responsible for release of calcium from silicates
Photosynthentic uptake of CO2 by cyanobacteria (e.g. Merz, 1992; (Section 3.2), but also have an important role in calcite formation in
Spiro and Pentecost, 1991), algae (e.g. McConnaughey, 1991) and non-marine environments. Calcium oxalate crystals such as whewel-
bryophytes may not be important in fast-flowing streams (e.g. lite (CaC2O4·H2O) and polyhydrated weddellite (CaC2O4·2H2O in its
230 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

simplest form; see Verrecchia et al., 1993) are found in mycorrhizal organism’ not plants, roots or root-associated mycorrhizae was
mats (see Cromack et al., 1979; Verrecchia et al., 1993) and around responsible. Horizontal root mats and their calcification may also
lichen thalli (e.g. Syers et al., 1967; Klappa, 1979; Burford et al., 2003) lead to laminar crusts (Wright et al., 1988), which can be inferred
where they are produced by reaction of biogenic oxalic acid with from the presence of tubular pores formerly occupied by the plant
calcium. Significantly, they promote rhizolith formation in soils rootlets as well as by a ubiquitous alveolar septal structure (see
(Cromack et al., 1979). Dehydration of weddellite to whewellite Table 2). In addition to this process, laminar calcite can also form by
followed by transformation to calcite under the influence of bacteria calcification of root medullas (probably formed during the life of the
(Cromack et al., 1977) was supported as a viable mechanism by root). In these cases, extracellular calcite forms crusts around the
Verrecchia et al. (1993) and Verrecchia and Dumont (1996). cortex, and may be accompanied by post-mortem calcified mucilag-
inous sheets of probable fungal origin (Alonso-Zarza, 1999). It should
4.4.3. Vascular plants not be forgotten that many other laminar calcrete crusts are the result
Vascular plants have affected spring, stream and lacustrine of non-root-associated biological processes, including structures
carbonate facies by providing stable substrates for calcification. Facies produced by cyanobacteria (e.g. Verrecchia et al., 1995) and lichens
models of Pedley (1990), Pentecost and Viles (1994) and Pedley et al. (Klappa, 1979), and other laminar crusts can be explained by
(2003) demonstrate cases of logs initiating barrage build-up by abiological processes such as precipitation from groundwaters
increasing turbulence (and hence CO2 degassing) and providing a ponded on impermeable surfaces.
substrate, as well as calcification of substrates provided by overhang-
ing branches and mosses. Many SSL carbonate facies (such as barrages 4.5. Temporal changes in non-marine carbonate precipitation up to the
and curtains — see Pedley, 1990) are more likely to form in the Devonian
presence of bryophytes and vascular plants because of these effects.
An important effect of the arrival of vascular plants for soil In the absence of thick high pCO2 soils prior to the Devonian, some
carbonates was the entrapment and binding of sediment (e.g. Cotter, precipitation of non-marine carbonates is both possible and expected.
1978). Because many pedogenic calcretes source their calcium from It is true that some of the processes outlined above, such as freeze-out,
wind-blown dust, not chemical weathering of silicate bedrock (for only cause relatively minor and local carbonate accumulations today.
example Machette, 1985; Capo and Chadwick, 1999), the binding of It is also true that all of the described processes can occur in different
sediment would not only produce more alluvial overbank mud host places at the same time. Even so, some notable events might have
materials (e.g. Cotter, 1978; Dunagan and Driese, 1999; Davies and altered the relative importance of these processes through the course
Gibling, 2010), but also allow build-up of wind-blown calcium of Earth history (Fig. 2).
favouring increased calcrete precipitation. For example, it can be argued that hydrothermal surface and near
Enhanced capillary action due to plants drawing up groundwater surface springs may have been more prevalent in Archaean and
(Davies and Gibling, 2010) is unlikely to have significantly increased Palaeoproterozoic times (as suggested by Walter, 1996) than today, in
pedogenic calcrete precipitation because most such calcretes are not part because of the higher heat flow on the hot young earth. If so, then
formed by capillary rise of underlying groundwaters (see Wright and one might predict that more hot springs meant more carbonate
Tucker, 1991). However vascular plants allow evapotranspiration precipitated by boiling processes than is found at present. Archaean
rather than evaporation which might enhance rates of calcrete marine hydrothermal carbonates are well known (e.g. Lindsay et al.,
deposition. For example, evapotranspiration can cause local calcite 2005), but preserved terrestrial deposits seem rarer. Hydrothermal
supersaturation around roots (e.g. Calvet et al., 1975) leading to the travertine is currently not known until ~ 2.2 Ga (Melezhik and Fallick,
formation of calcite sheaths. Hsieh et al. (1998) examined oxygen 2001) and ~ 1.79 Ga (Rainbird et al., 2006).
isotopic compositions of Hawaiian soil waters. They deduced that Aside from the apparent lack of preserved Archaean terrestrial
increasing annual rainfall and decreasing temperature caused water environments, atmospheric pCO2 levels were probably very high until
loss by evaporation to decrease and water loss by transpiration to at least ~2.3 billion years ago (e.g. Bekker et al., 2004). So an apparent
increase. Thus when temperatures cool and humidity increases lack of preserved Archaean terrestrial spring carbonates might also
transpiration still operates and (at least in tropical regions) vascular reflect less precipitation, due to less vigorous equilibration of deep-
plants increase soil water loss. sourced CO2 with the higher pCO2 atmosphere. Further predictions
Rhizoliths (organosedimentary structures produced by roots; are that some CO2 degassing due to turbulence might have led to
Klappa, 1980) are obvious features found in many pedogenic calcretes precipitation of carbonate in Archaean and Palaeoproterozoic braided
that would not be expected without vascular plants (see Table 2). streams (e.g. Buck, 1980), and degassing through temperature rise
Some rhizoliths form when calcium levels in soil solutions are high could have caused some minor precipitation in the shallow subsurface
and exceed the requirements of the plant. In this way calcium (and of cold regions, though there is currently no evidence of the latter.
hence calcite; Eq. (13)) accumulates in the rhizosphere (Hinsinger, Marine carbonate platforms, known from ~ 2.6 Ga onwards
1998; Hinsinger et al., 2006). Plant roots that maintain their charge (Grotzinger, 1989), can present sources of calcium and bicarbonate
balance by taking up protons or releasing bicarbonate ions when they ions for non-marine carbonate precipitation when subaerially exposed
take up more anions than cations (see Sections 3.1 and 3.2) may also and eroded. But the apparently sparse record means any suggestion
create local supersaturation with respect to calcite (e.g. Klappa, 1980). that the origin of the carbonate platform was significant for calcrete
Interestingly, Wright et al. (1995) claimed that some calcretes are production is at best unproven. The oldest reported possible calcretes
“almost completely the result of root activity”, forming through are also found at around ~ 2.6 Ga (Watanabe et al., 2000; see Fig. 2 and
intracellular and extracellular calcification as well as through root Table 1). In this case, Watanabe et al. (2000) suggest calcium was
involvement in peloid formation. Such rocks have been termed derived from nearby granites, and HCO− 3 from the atmosphere.
rhizolites (Klappa, 1980; Wright et al., 1988, 1995). Furthermore, Precipitation was caused by evaporation in a semi-arid setting.
Jaillard (1984) reported that calcified roots may locally form up to a The Huronian glaciation (e.g. Evans et al., 1997) must have
quarter of the soil mass. Intracellular calcification may have produced lowered global sea level and exposed some of these marine platform
Microcodium (see Table 2) in many Moscovian to early Permian and carbonates. This is likely to have produced terrestrial environments
latest Cretaceous to Palaeogene (ages from Kabanov et al., 2008) where palaeokarst surfaces and associated terrestrial carbonates
paleosols (Klappa, 1978; Jaillard et al., 1991; Wright et al., 1995; might be found. Carbonates of possible lacustrine origin in the
Alonso-Zarza et al., 1998; Košir, 2004), although Kabanov et al. (2008) Espanola Formation of Ontario may be an example of this (e.g.
suggest on morphological grounds that a ‘mycelial saprotrophic Bernstein and Young, 1990; Veizer et al., 1992).
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 231

Oxygenation of the Earth's atmosphere (e.g. Bekker et al., 2004) especially where precipitation took place within millimetres or
must surely have influenced non-marine carbonate precipitation? centimetres of the ground surface.
First, this atmospheric revolution seemingly allowed accumulation of Warm, shallow, continent-flooding epeiric seas (Irwin, 1965)
some of the first calcium sulphates in sabkha sediments by ~ 2.1 Ga during the Cambrian and Ordovician, combined with the Cambrian
(Tulomozero Formation, Onega basin, Fennoscandia; Melezhik et al., innovation of marine organisms with carbonate skeletons, produced
2005a,b). With increasing local abundance of gypsum and anhydrite, much marine carbonate bedrock. Subsequent subaerial exposure of
and their consequent dissolution to give groundwaters locally such marine carbonates during the Cambrian likely provided both
enriched in calcium ions and sulphuric acid (the latter causing further calcium and bicarbonate ions for terrestrial carbonates like calcretes
release of calcium ions through carbonate bedrock dissolution; see associated with palaeokarst in the Hawker Group of Australia (James
van Everdingen et al., 1985), the relative importance of the common- and Gravestock, 1990) and calcretes in upward-shallowing peritidal
ion effect as a non-marine carbonate precipitation mechanism is likely cycles of the Cambrian La Flecha Formation, Argentine Precordillera
to have risen. The common-ion effect is able to produce relatively (Keller et al., 1989; Buggisch et al., 2003).
large amounts of carbonate, and taken together with falling The first firm evidence for land plants comes in the form of trilete
atmospheric pCO2 levels (thereby favouring CO2 degassing processes), spores, possibly of terrestrial bryophytes (e.g. Strother et al., 1996;
an overall increase in non-marine carbonate (particularly speleothem Wellman and Gray, 2000). These are first known from the Llanvirn
and groundwater calcrete) abundance would be expected in the rock (Middle Ordovician). As potential substrates for calcification around
record at this time. springs, streams and lakes, and by initiating evapotranspiration, it can
However the timing and scale of production of the first marine be argued that these first plants probably influenced SSL carbonate
sulphates is questioned by Kah et al. (2004), who suggest on the basis of morphology and occurrence, but where are the examples? A lack of
sulphur isotope modelling that atmospheric oxygen and marine roots may have restricted the influence of these first plants on silicate
sulphate levels may initially have remained quite low (the latter being weathering (Algeo et al., 2001). Production of thin organic-rich soils
5–15% of modern levels), only rising gradually to 75% of modern marine likely started to accelerate in the Ordovician, a process which is
sulphate levels by the end of the Proterozoic. If this scenario is correct perhaps reflected in the production of some Ordovician karst-
and the decline in atmospheric pCO2 levels was not initially rapid, then associated cements (e.g. Read and Grover, 1977; Grover and Read,
groundwater calcretes and speleothems produced by the common-ion 1978; Tobin and Walker, 1994) and palaeokarst surfaces (Jia-yu and
effect, as well as through CO2 degassing, might not have been of any Johnson, 1996; Ryu et al., 2005). Glaciation-related eustatic sea level
great volume until the Meso- or Neoproterozoic. It is interesting to note, fall at the end of the Ordovician (e.g. Hambrey, 1985) is believed to
with extremely limited data, that the oldest reported speleothem is not have exposed much Ordovician carbonate bedrock to karstification
found until ~1.2 Ga (Table 1; Glover and Kah, 2006; Skotnicki and (e.g. Jia-yu and Johnson, 1996).
Knauth, 2007), which might support this view. While a few Silurian calcretes are known (e.g. Wright et al., 1992;
One might expect that some non-marine carbonates should have Davies et al., 2005), Silurian SSL carbonates and speleothems have
formed from processes like ‘freeze-out’ at the time of the Cryogenian gone as yet unreported. This might reflect lack of preservation, or lack
‘snowball earth’ glaciations around 710–630 Ma (Kirschvink, 1992), of detection. One can speculate that cyanobacterial influence upon
perhaps in sediments like those with permafrost sand wedges lacustrine and fluvial carbonate precipitation (Fig. 6) is likely to have
reported from Mauritania by Deynoux (1982) and from Norway by decreased during the Silurian in response to increased competition
Edwards (1975). Non-marine carbonates which have been reported from early plants for growing space and increased grazing by
from this time interval include microbially influenced calcretes herbivores (see Shear, 1991).
(Bertrand-Sarfati and Moussine-Pouchkine, 1983). Praekelt et al. A common feature of many (though not all) pre-Devonian calcretes
(2008) suggested South African palaeokarst of late Neoproterozoic to is their occurrence in alluvial fan facies rather than floodplain sediments
early Cambrian age was related to glacio-eustatic sea level fall with (for example Elmore, 1983; Aspler and Donaldson, 1986; Kalliokoski,
consequent subaerial exposure of carbonate bedrock, yet no spe- 1986; De Ros et al., 1994). This may be because alluvial fans were
leothems are described. more widespread than floodplain environments before the stabilising
A problem related to these Cryogenian sediments concerns the effect of plant roots (e.g. Cotter, 1978; Algeo et al., 2001; Davies and
apparent lack of non-marine carbonates immediately following the Gibling, 2010). Terrestrial carbonate production is inferred to have
glaciations, when ‘cap carbonates’(see Shields, 2005, for a review) accelerated when plants with deep roots, their symbionts, and resulting
were widely deposited in marine environments. Some cap carbonate thick organic-rich soils became established in the Middle Devonian
models (e.g. Higgins and Schrag, 2003) invoke extreme alkalinity (e.g. Alonso-Zarza and Tanner, 2010; Davies and Gibling, 2010).
fluxes from terrestrial to marine environments, brought about by
rapid rates of carbonate and silicate weathering, so non-marine 5. Recognising Silurian and older non-marine carbonates
carbonates are surely to be expected. A counter argument – that
generally high atmospheric pCO2 meant acidic terrestrial ground- As shown above and in Table 1, the emergence of vascular plants
waters at a time of only weak CO2 degassing (hence few non-marine (though important) was not a prerequisite for terrestrial carbonate
carbonates) – does not seem to favour this model of cap carbonate precipitation. Many other and earlier events are likely to have
formation (see Shields, 2005). influenced terrestrial groundwater alkalinity and non-marine car-
Later subaerial exposure of the Cryogenian cap carbonates must bonate abundance during the course of Earth history (Fig. 2). If this is
have provided further opportunities for bedrock dissolution and non- true, then such deposits should have a record that stretches far
marine carbonate precipitation: palaeokarst surfaces associated with beyond the Devonian. The following therefore provides a preliminary
exposure of cap carbonates from the Marinoan glaciation in overview of the likely physical appearance of such non-marine
northwestern Canada are described by James et al., 2001, and carbonates in ‘deep time’.
calcretes of similar age and occurrence are described from the Lesser
Himalaya by Jiang et al., 2003. 5.1. Petrographic changes of non-marine carbonates through time?
Biological innovations during the late Neoproterozoic are also
likely to have exerted a major influence upon terrestrial carbonate SSL carbonate, calcrete and speleothem are all influenced by a
deposition. The hypothesised arrival of fungi and lichen at around this variety of biological and environmental factors, each of which has
time, with their microborings and biologically produced calcium varied through time. It follows that petrographic features of non-
oxalates, likely influenced calcrete micromorphology (Table 2), marine carbonates are also likely to have changed through time.
232 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Wright and Tucker (1991) usefully divided calcretes into those with thickness and also reverse-graded. Here, pisoids are cemented into
biological β fabrics and those with abiological α fabrics. β fabrics, clusters of ‘grapestones’, suggesting cementation took place in situ
which include features like alveolar septal fabric, are mostly linked to (Pelechaty and James, 1991). Pisoliths are present in the Kuetsjärvi
the influence of vascular plants and fungi. In contrast, α fabric features Sedimentary Formation of Fennoscandia, of Palaeoproterozoic age
such as circumgranular cracking are often the result of the displacive (Melezhik et al., 2004). It is possible that such ancient pisoliths formed
and replacive nature of calcrete growth, and need not require biology. through an abiotic process of shrinking and swelling of clays (e.g.
Because the timings of major biological innovations and of major Tandon and Friend, 1989) and through downslope transport of grains
environmental changes are relatively well constrained, it is possible to (e.g. Arakel and McConchie, 1982).
ascribe a likely maximum age and (where applicable) a minimum age Some Recent pisoliths are themselves composed of Needle Fibre
to petrographic features found in terrestrial carbonates (see Table 2). Calcite (e.g. Beier, 1987; see Wright and Tucker, 1991), a term which
This table is intended to show how Silurian and older non-marine refers to acicular crystals with lengths at least four times the width
carbonates might be expected to appear. Table 2 can also be compared (see James, 1972; Verrecchia and Verrecchia, 1994; Table 2). Needle
with Table 1 which shows many of the features of currently accepted Fibre Calcite has been explained by both fungal influence and
examples. abiological processes (both types described in Verrecchia and
Data assembled in Tables 1 and 2 reveal that many features of Verrecchia, 1994). In turn, this raises questions about microbial
speleothems, SSL carbonates and calcretes found in post-Silurian rocks participation in the formation of very ancient pisoliths like those
should be recognisable in older rocks, from the late Archaean onward. mentioned above. Interestingly, however, Needle Fibre Calcite is
Calcareous nodules ought to be amongst the oldest kinds of seemingly unknown before the Carboniferous (Wright, 1984). It
carbonate to be found in terrestrial deposits on Earth, though reported remains unclear whether pisoliths which lack Needle Fibre Calcite
examples are quite rare. Plausible examples include those reported require microbial involvement to form.
from Canadian Palaeoproterozoic fluvial and lacustrine sands: these are Carbonate coated macrophyte stems might be theoretically
spherical to elliptical and up to 5 cm diameter, of likely groundwater expected from rocks of Ordovician age onward, following the advent
origin, and (unlike many modern pedogenic nodules) are absent from of bryophyte-like organisms at that time (e.g. Wellman and Gray,
the finer grained alluvial faces (Aspler and Donaldson, 1986). 2000). It seems, however, that they have not been reported from rocks
Mesoproterozoic pedogenic calcrete nodules from Michigan described older than Lower Carboniferous (e.g. Rex and Scott, 1987; Scott and
by Kalliokoski (1986) are much smaller, up to 3 mm diameter and are Galtier, 1988). As the siliceous Rhynie Chert of Lower Devonian age
oval in shape. They are found in association with carbonate coated has shown, mineral-precipitating springs can preserve excellent
siliciclastic clasts, floating grain fabric and tepee structures. From the records of early terrestrial life (e.g. Trewin and Rice, 1992). Hence
same area, Kalliokoski and Welch (1985) described layers of nodules, there are reasons to predict that Silurian, and even Ordovician
the latter being each up to 3 cm diameter, together with mudstones macrophytes were coated by carbonate in calcareous spring deposits
exhibiting crystallaria (Table 2). From a Middle Cambrian supratidal yet to be reported.
environment, Post and Long (2008) illustrated a structureless carbon- Rhizoliths are found abundantly in calcretes from the Upper
ate bed a few tens of centimetres in thickness which probably formed Devonian onwards (for examples see Klappa, 1980; Wright and
through coalscence of calcrete nodules. Ordovician argillaceous Tucker, 1991 and references therein; Balin, 2000). Like the coated
palaeosols have also been found to contain ‘common small yellow’ macrophytes, they are not yet reported from pre-Devonian examples
carbonate nodules (Retallack, 2001). From this admittedly meagre list, (Table 2). It could be that the absence of key macro-morphological
it can be argued that calcrete nodules are neither confined to nor features such as rhizoliths, or indeed fabrics like calcified root mats,
indicative of Devonian and younger alluvial floodplain sediments. alveolar septal fabric and Microcodium, has hitherto hindered the
Similarly, it can be argued that laminar crusts were present in non- recognition of many ancient SSL carbonates and calcretes. Interpreta-
marine deposits from the earliest times. Rare examples from the tions until now have therefore had to remain very tentative. For
Palaeoproterozoic include those from Fennoscandia (e.g. Melezhik et example, James and Gravestock (1990) and Post and Long (2008)
al., 2004). About a billion years younger are the laminated, silicified, have been careful to describe only ‘possible’ Cambrian calcrete
haematite-rich crusts of the ~1.2 Ga Mescal Limestone of Arizona, USA examples. Hopefully this review will help to open up the search for
(Beeunas and Knauth, 1985). Younger still are the layered orbicular early terrestrial carbonates and their fabrics.
crusts reported from the Neoproterozoic of Mali (Bertrand-Sarfati and
Moussine-Pouchkine, 1983). These possible calcretes consist of 5.2. Further examples of Proterozoic terrestrial carbonates from North
alternating dark and clear microspar layers, associated with oncoids America
and sediment-filled cavities.
Early Palaeozoic laminar crusts are also uncommon in the literature. The example illustrated in Fig. 7 displays Proterozoic calcretes
From a middle Cambrian evaporitic playa lake dry mud flat setting, (Fig. 7B) and non-marine stromatolites (Fig. 7C) associated with
Southgate et al. (1989) have reported dolomite crusts represented by redbeds (Fig. 7A; D) of the Mesoproterozoic Sibley Group of the
discontinuous rinds of 0.1 to 2 mm thickness, with sharp tops and Keweenawan Rift system which crop out at Nipigon, Ontario. Facies
gradational bases. A laminated calcrete clast of Cambrian to Ordovician include red coloured alluvial and wind-blown muds containing
age was also illustrated from alluvial fan and braided fluvial sediments gypsum pseudomorphs and laminated calcretes with tepee struc-
by De Ros et al. (1994). The latter cases illustrate, at least, that ancient tures; stromatolitic limestones with evaporite solution collapse
calcrete laminar crusts may be found in ancient terrestrial deposits. breccias; and buff-coloured marls containing ptygmatic folds. These
Further ancient calcrete fabrics, such as pisoids, have been facies point towards a semi-arid system with evaporite (and calcite?)
documented from pre-Devonian rocks. For example, in the Argentine precipitation caused by high evaporation.
Precordillera (Keller et al., 1989), Cambrian pisoidal calcretes are Stromatolites are likely to have developed around the margins of
described as up to 1 metre thick. The pisoids comprise alternations of ephemeral lakes, where it is likely that some of these cyanobacte-
white and brown micritic layers, surrounded by clays. These clay rially-influenced carbonate build-ups formed through of a combina-
layers might be interpreted as pedogenic cutans (Table 2). Mudcracks tion of biological and chemical processes. The availability of calcium
and tepees are also associated with these La Flecha Formation from dissolution of calcium sulphates and wind-blown sources may
calcretes. Older examples of pisoliths have been reported from the have led to some groundwater calcrete precipitation via the common-
Mesoproterozoic Kanuyak Formation of NW Canada (Pelechaty and ion effect, coupled with assistance from evaporation and perhaps
James, 1991; Pelechaty et al., 1991), where beds are similarly of metre some actinobacterial or even fungal encouragement where
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 233

precipitation occurred within a few centimetres of the ground surface, 5.4. Are Silurian and older non-marine carbonates usually recognised
although petrographic evidence of the latter (e.g. Needle Fibre Calcite; and reported?
see Table 2) has not yet been found and might not have been
preserved. It is clear from Table 2 that many of the features associated with
Related Keweenawan non-marine carbonates associated with calcretes and SSL carbonates are produced or influenced by terrestrial
alluvial fan and braided fluvial facies have been reported by Elmore organisms. Comparatively few examples of non-marine carbonate are
(1983) from the Copper Harbor Conglomerate of Michigan, USA. Their reported from pre-Devonian rocks; the literature study of Davies and
reports of gypsum moulds and desiccation cracks are both consistent Gibling (2010) indicates nodular calcretes are more frequently
with a semi-arid setting. The presence of stromatolites plus sparry reported from Siluro–Devonian age alluvial settings than from older
calcite fans as well as ooid grains (Table 2) is similarly consistent with (Cambro-Ordovician) deposits. Whether this apparent sudden in-
their formation in short-lived lakes of this age (Elmore, 1983 and crease in reports of soil carbonate is really due to the arrival of
Table 2). Kalliokoski (1986) also studied calcrete cements from vascular plants (Alonso-Zarza and Tanner, 2010; Davies and Gibling,
Keweenawan conglomerates of Michigan, and likewise concluded 2010) or whether older non-marine carbonates are under-reported
that evaporation was an important factor in their formation. must be considered. So might non-marine carbonates have been
Petrography of these calcretes is consistent with a shallow ground- overlooked?
water origin, and with the age of the rocks, based on features
described in Table 2. It is now interesting to consider whether pre- 5.4.1. Cave carbonates
Silurian non-marine carbonates contain a valuable record of pre- There is a strong literature bias towards studies of Quaternary
Silurian land surface biotas. speleothems because they can be dated by U–Th radiometric methods
and used in palaeoclimatic reconstructions (see Fairchild et al., 2006).
It is predicted that speleothem fabrics ought to be associated with
5.3. Do non-marine carbonates contain a record of pre-Silurian land examples of karst solution (for some examples see Table 1), unless
surface habitation? calcium was routinely transported away by acidic, undersaturated
solutions in the past. Yet while mention of pre-Silurian palaeokarst is
There has been much speculation of pre-Silurian terrestrial land relatively common (for example James and Gravestock, 1990;
cover (e.g. Campbell, 1979; Golubic and Campbell, 1979; Wright, Pelechaty et al., 1991; Horodyski and Knauth, 1994; Gutzmer and
1985; Shear, 1991; Prave, 2002; Knauth and Kennedy, 2009). Biota Beukes, 1998; Kenny and Knauth, 2001; Ryu et al., 2005; Yuan et al.,
from 1 Ga onwards have been suggested to include protists, mosses, 2005; Post and Long, 2008; Keller and Lehnert, 2010) reference to
fungi, liverworts and (from 600 Ma) lichens (Knauth and Kennedy, ‘speleothem’ is rare (Pelechaty et al., 1991; Glover and Kah, 2006;
2009). If these authors are correct, evidence might be found in the Skotnicki and Knauth, 2007). This suggests that a lack of reference to
micro-fabrics of pre-Ordovician calcretes. However Needle Fibre ancient speleothem deposits is not due to a lack of preservation of
Calcite which might indicate the presence of fungi or lichens in fossil exposure surfaces. It may be that preserved examples are seen as
deposits (e.g. Carboniferous examples in Wright, 1984) could be secondary, post-depositional features (not part of the primary
dismissed as abiogenic (e.g. James, 1972) or modern overprint unless depositional story), so referred to as ‘calcareous cement’ or ‘secondary
petrography proves otherwise (see Freytet and Verrecchia, 2002). carbonate’ (e.g. Kenny and Knauth, 2001).
Other structures which might be looked for include microborings,
such as those reported from modern feldspar and hornblende grains 5.4.2. Spring, stream and lacustrine deposits
by Jongmans et al., 1997. Studies of Palaeoproterozoic calcretes have If pre-Devonian groundwaters were predominantly alkaline, with
not hitherto found any unequivocal biological fabrics but Bertrand- the highest pH values in the Middle Ordovician (Jutras et al., 2009),
Sarfati and Moussine-Pouchkine (1983) believed some laminar one might expect more pre-Devonian SSL deposits than currently
fabrics in calcareous soils of the younger (700–600 Ma) Sarnyéré reported (Table 1). But while reports of terrestrial microbial deposits
Formation of Mali to be microbial in origin. are rare, reports of marine stromatolites (e.g. Riding, 2000) are
Lichen-like fossils and possible fungal hyphae of the ~ 551– common, so it is possible to speculate that some terrestrial microbial
635 Ma Doushantuo Formation, China (Yuan et al., 2005) are not carbonates have been mis-identified as marine: the difficulty of
known from carbonates, but from marine phosphorites. However distinguishing lacustrine from marine successions in the absence of
Horodyski and Knauth (1994) reported fossil bacteria preserved in incontrovertible fossil evidence is illustrated by the disagreement
cherts associated with palaeokarst surfaces of the ~ 1.2 Ga Mescal over the Tumbiana Formation (see Awramik and Buchheim, 2009, and
Limestone, Arizona, USA, and Kenny and Knauth (2001) suggest references therein). Confusion with later diagenetic calcite cements is
“secondary carbonate” associated with the karst is depleted in 13C problematic, and early diagenetic processes, such as aggrading
relative to other parts of the section by up to 11‰ (vs PDB) which neomorphism (Love and Chafetz, 1988, but see also Brasier et al., in
they attribute to input of light 12C from photosynthesising commu- press) and micritisation, might hinder the recognition of ancient SSL
nities on the karst surface. Upfold (1984) reported tufted calcitic carbonates. Examples of SSL carbonate are known from the Permian
stromatolites from the Stoer Group, but these have since been re- (Szulc and Cwizewicz, 1989; Stworzewicz et al., 2009), but Devonian
interpreted as thin layers of secondary carbonate interlayed with and Carboniferous reports are still as rare as those from preceding
thin layers of mud containing evaporitic fabrics (Prave, personal times, and this might reflect small amounts of deposition in
communication, 2010). environments unlikely to be preserved.
Proterozoic non-marine stromatolites (Fig. 7C) which might be
considered terrestrial microbial carbonates (see Table 2) include 5.4.3. Calcretes
lacustrine stromatolites of the ~1 Ga Copper Harbor Conglomerate, The lack of Silurian and older calcretes observed by Davies and
Michigan, USA (Elmore, 1983) and ~1.3 to ~ 1.8 Ga stromatolites in Gibling (2010) might be partly due to their definition: “accumulations
alluvial conglomerates of the Murky Formation, Northwest Canada of nodular carbonate in mudstones”. Consequently laminar fabrics are
(Hoffman, 1976) as well as putatively lacustrine stromatolites of the seemingly not included in their assessment. ‘Typical’ calcrete-
Tumbiana Formation of the Pilbara Craton, Australia (Awramik and containing soil profiles formed in the organic-rich modern terrestrial
Buchheim, 2009). Late Ordovician soil of the Juniata Formation, world may be highly variable from place to place, but are easily
Pennsylvania, USA, contains carbonate-sheathed burrows, and Retal- recognised (e.g. Esteban and Klappa, 1983) whereas the significance
lack (1985) speculated the carbonate might be of fungal origin. of cements in alluvial fans (e.g. the Proterozoic Copper Harbor
234 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Conglomerate; Elmore, 1983) might easily be overlooked. Retallack the atmosphere became oxic; freeze-out (today a seemingly minor
(1986) highlights that while weakly developed post-Ordovician soils process) may have been slightly more important during the colder
may be identified by the presence of “fossil roots or trace fossils”, these phases of Earth history. It is also possible that carbonic acid played a
will be lacking in equivalent weakly developed Precambrian examples. In lesser role in the distant past; for example, some surface ground-
some cases, closer examination reveals that calcrete fabrics might easily waters may locally have become quite acidic following oxygenation of
have gone un-noticed, such as “microscopic concretions” or “micro- the atmosphere and oxidation of sulphides in the Palaeoproterozoic.
nodules” which are “similar to pedogenic carbonate” in Proterozoic soils The origin of vascular plants and earlier bryophytes was clearly
of NW Scotland (Retallack and Mindszenty, 1994). significant for terrestrial carbonate precipitation. These plants, together
Many pre-Devonian alluvial fans (e.g. the Cambrian White Point with their symbionts, produced thick organic-rich soils and enhanced
Conglomerate; Daily, 1980) are known, so depositional environments both evapotranspiration and trapping of calcium-rich wind-blown dust.
do not seem to have been lacking. Some of these alluvial fans might be Some other events which may also have been important include the
found to contain calcrete cements on further inspection, but origin of carbonate-precipitating microbes in the Archaean; formation
unfavourable climatic conditions for deposition or preservation (too and exposure of marine carbonate platforms from the late Archaean;
wet or too dry) could also be responsible. This might be true in cases oxygenation of the atmosphere in the Proterozoic, leading to more
where calcrete fabrics have clearly been looked for and not found (e.g. vigorous degassing of CO2 from groundwaters to the lower pCO2
rarity of calcretes in some Canadian Palaeoproterozoic alluvial fans atmosphere, production of H2SO4, CaSO4 and calcite precipitation via the
noted by Aspler and Donaldson, 1986; lack of calcretes in Cambrian common-ion effect; formation and exposure of the Cryogenian ‘cap
peritidal carbonates of the Appalachians, USA, reported by Osleger carbonates’; the origin of fungi, which produce calcium oxalates and
and Read, 1991). release calcium from minerals from the late Proterozoic; and carbonate
There also seems to have been a paucity of studies centred on production by marine organisms since the Cambrian.
calcretes from pre-vascular plant times. For example, numerous Because organisms associated with non-marine carbonates have
articles have been published on terrestrial clastic sediments of the changed through time, the physical appearances of non-marine
Ringerike Group of Norway (such as Turner and Whitaker, 1976; carbonates have also changed through time (Table 2). Known
Turner, 1974; Davies et al., 2005, 2006), but these contain only limited examples of calcretes, SSL carbonates and speleothems formed in
mention of ‘concretions’ and nodules interpreted as calcrete by Davies ‘deep time’ (Table 1) fit the predicted patterns of appearance based
et al. (2005) and Davies and Gibling (2010). Similarly Gutzmer and on likely depositional processes and influential organisms according
Beukes (1998) reported Palaeoproterozoic haematitic laterites from to age and setting; a further example of this is provided by Pro-
South Africa, but associated calcretes (Gutzmer, personal communi- terozoic calcretes and lacustrine stromatolites of the Keweenawan
cation, 2010) have apparently never been worked on. rift system.
Terrestrial carbonates record some evidence of pre-Silurian land
6. Conclusions colonisation in the form of stromatolitic cyanobacterial build-ups in
lacustrine strata, perhaps as long ago as the Archaean. Microbial
A first conclusion from this study is that some definitions currently fabrics have been reported from Neoproterozoic calcretes, and calcite
applied to Quaternary terrestrial carbonates are hard to apply to which might be attributed to fungi from Ordovician soils.
ancient ones. The terms ‘tufa’ and ‘travertine’ imply circumstances There is annecdotal evidence that speleothems, SSL carbonates and
which cannot be easily verified in ‘deep time’ such as locally elevated calcretes from Silurian and older times are under-reported. Possible
water temperatures, and in cases where depositional conditions are reasons for this include an assumption that many or most calcite
unclear, might be avoided in favour of more descriptive terminology. cements are of late diagenetic origin; confusion of terrestrial for
Alternative phrasing such as ‘spring carbonate’, ‘stream carbonate’ marine deposits; or destruction of original textures through early
and ‘lacustrine carbonate’ is less restrictive than ‘tufa’ and ‘travertine’ diagenesis or metamorphism.
and relevant to discussions of deposition in different terrestrial
settings in this review (e.g. Section 4 and Table 2). Acknowledgements
Fungi, like those known to release calcium from apatite, might be
older than vascular plants with roots. The origin of fungi was likely an Helpful comments on an early draft of this manuscript were
important event for chemical weathering of bedrock and non-marine provided by Tony Fallick (SUERC), IGME gave permission to work in
carbonate precipitation. However it is probable that the development Greece. Tony Prave (St Andrews) and Martin Brasier (Oxford). Mike
of vascular plants with large root systems and symbionts in the Simmons and Richard James (Neftex Petroleum) suggested some
Middle Devonian resulted in acceleration of silicate weathering. Low formations which might contain non-marine carbonates for Table 1.
pH levels in the rhizosphere from organic and carbonic acids surely Dineke Brasier assisted with proof reading. The journal editors and
caused increased dissolution of carbonate bedrock. Clay minerals reviewers provided helpful comments. ATB is supported by NERC
produced during silicate weathering probably only provide a short- grant NE/G00398X/1.
lived barrier to chemical weathering.
Thick, organic-rich soils (strongly influenced by vascular plants) References
are today involved in the precipitation of significant volumes of
Aharon, P., 1988. Oxygen, carbon and U-series isotopes of aragonites from Vestfold
speleothem, SSL carbonate and calcrete by encouraging carbonic acid Hills, Antarctica: clues to geochemical processes in subglacial environments.
production leading to carbonate bedrock solution and consequently Geochimica et Cosmochimica Acta 52 (9), 2321–2331.
re-precipitation of calcite in terrestrial environments. In the absence Algeo, T.J., Scheckler, S.E., 1998. Terrestrial-marine teleconnections in the Devonian:
links between the evolution of land plants, weathering processes, and marine
of high pCO2 soils, terrestrial carbonates could still have precipitated anoxic events. Philosophical Transactions of the Royal Society of London. Series B:
by several mechanisms, which may each have varied in their relative Biological Sciences 353, 113–130.
importance through time. Boiling is likely to have been more Algeo, T.J., Berner, R.A., Maynard, J.B., Scheckler, S.E., 1995. Late Devonian oceanic
anoxic events and biotic crises: ‘rooted’ in the evolution of vascular land plants?
important on a young Earth with higher heat flow, yet deposits are
GSA Today 5 (45), 64–66.
seemingly unknown; CO2 degassing is likely to have been less Algeo, T.J., Scheckler, S.E., Maynard, J.B., 2001. Effects of the Middle to Late Devonian spread
vigorous at times of high atmospheric pCO2; evaporation has likely of vascular land plants on weathering regimes, marine biotas and global climate. In:
always been an important factor in surface and shallow environ- Gensel, P.G., Edwards, D. (Eds.), Plants Invade The Land: Evolutionary and
Environmental Perspectives. Columbia University Press, New York, pp. 213–236.
ments; the common-ion effect might have increased in importance Allen, J.R.L., 1974. Sedimentology of the Old Red Sandstone (Siluro–Devonian) in the
through time, particularly with precipitation of calcium sulphates as Clee Hills area, Shropshire, England. Sedimentary Geology 12 (2), 73–167.
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 235

Alonso-Zarza, A.M., 1999. Initial stages of laminar calcrete formation by roots: examples Boyle, J.R., Voigt, G.K., 1973. Biological weathering of silicate minerals. Plant and Soil 38
from the Neogene of central Spain. Sedimentary Geology 126 (1–4), 177–191. (1), 191–201.
Alonso-Zarza, A.M., Tanner, L.H., 2010. Preface. In: Alonso-Zarza, A.M., Tanner, L.H. Brasier, A.T., Andrews, J.E., Marca-Bell, A.D., Dennis, P.F., 2010. Depositional continuity
(Eds.), Carbonates in Continental Settings: Geochemistry, Diagenesis and Applica- of seasonally laminated tufas: implications for δ18O based palaeotemperatures.
tions. Developments in Sedimentology. Elsevier, pp. xi–xii. Global and Planetary Change 71, 160–167.
Alonso-Zarza, A.M., Sanz, M.E., Calvo, J.P., Estévez, P., 1998. Calcified root cells in Brasier, A.T., Andrews, J.E. and Kendall, A.C., in press. Diagenesis or dire genesis? The
Miocene pedogenic carbonates of the Madrid Basin: evidence for the origin of origin of columnar spar in tufa stromatolites of central Greece and the role of
Microcodium b. Sedimentary Geology 116 (1–2), 81–97. Chironomid larvae. Sedimentology. doi:10.1111/j.1365-3091.2010.01208.x.
Anderson, L.G., Jones, E.P., 1985. Measurements of total alkalinity, calcium, and sulfate Buck, S.G., 1980. Stromatolite and ooid deposits within the fluvial and lacustrine
in natural sea ice. Journal of Geophysical Research 90, 9194–9198. sediments of the Precambrian Ventersdorp Supergroup of South Africa. Precam-
Andrews, J.E., 2006. Palaeoclimatic records from stable isotopes in riverine tufas: brian Research 12 (1–4), 311–330.
synthesis and review. Earth Science Reviews 75 (1–4), 85–104. Buggisch, W., Siegert, R., 1988. Paleogeography and facies of the ‘gres terminaux’
Andrews, J.E., Brasier, A.T., 2005. Seasonal records of climatic change in annually (uppermost Lower Cambrian, Anti-Atlas/Morocco). In: Jacobshagen, V. (Ed.), The
laminated tufas: short review and future prospects. Journal of Quaternary Science Atlas System of Morocco. Lecture Notes in Earth Sciences. Springer Berlin,
20 (5), 411–421. Heidelberg, pp. 107–121.
Andrews, J.E., Brimblecombe, P., Jickells, T.D., Liss, P.S., Reid, B., 2004. An Introduction to Buggisch, W., Keller, M., Lehnert, O., 2003. Carbon isotope record of Late Cambrian to
Environmental Chemistry. Blackwell, Oxford. 320 pp. Early Ordovician carbonates of the Argentine Precordillera. Palaeogeography,
Arakel, A.V., 1982. Genesis of calcrete in Quaternary soil profiles, Hutt and Leeman Palaeoclimatology, Palaeoecology 195 (3–4), 357–373.
lagoons, Western Australia. Journal of Sedimentary Petrology 52 (1), 109–125. Burford, E.P., Kierans, M., Gadd, G.M., 2003. Geomycology: fungi in mineral substrata.
Arakel, A.V., 1986. Evolution of calcrete in palaeodrainages of the Lake Napperby area, Central Mycologist 17 (03), 98–107.
Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 54 (1–4), 283–303. Calaforra, J., Forti, P., Fernandez-Cortes, A., 2008. Speleothems in gypsum caves and
Arakel, A.V., McConchie, D., 1982. Classification and genesis of calcrete and gypsite their paleoclimatological significance. Environmental Geology 53 (5), 1099–1105.
lithofacies in paleodrainage systems of inland Australia and their relationship to Calvet, F., Pomar, L., Esteban, M., 1975. Las Rizocreciones del Pleistoceneo de Mallorca.
carnotite mineralization. Journal of Sedimentary Research 52 (4), 1149–1170. Revista del lnstituto de Investigaciones Geologicas, Universidad de Barcelona 30,
Armstrong, R.A., Compston, W., Retief, E.A., Williams, I.S., Welke, H.J., 1991. Zircon ion 35–60.
microprobe studies bearing on the age and evolution of the Witwatersrand triad. Campbell, S.E., 1979. Soil stabilization by a prokaryotic desert crust: implications for
Precambrian Research 53 (3–4), 243–266. Precambrian land biota. Origins of Life 9, 335–348.
Arnórsson, S., 1989. Deposition of calcium carbonate minerals from geothermal Canadell, J., Jackson, R.B., Ehleringer, J.B., Mooney, H.A., Sala, O.E., Schulze, E.D., 1996.
waters — theoretical considerations. Geothermics 18 (1–2), 33–39. Maximum rooting depth of vegetation types at the global scale. Oecologia 108 (4),
Arthur, M.A., Fahey, T.J., 1993. Controls on soil solution chemistry in a subalpine forest 583–595.
in north-central Colorado. Soil Science Society of America Journal 57, 1123–1130. Candy, I., Rose, J., Lee, J., 2006. A seasonally ‘dry’ interglacial climate in eastern England
Aspler, L.B., Donaldson, J.A., 1986. Paleoclimatology of Nonacho Basin (Early during the early Middle Pleistocene: palaeopedological and stable isotopic
Proterozoic), Northwest Territories, Canada. Palaeogeography, Palaeoclimatology, evidence from Pakefield, UK. Boreas 35 (2), 255–265.
Palaeoecology 56 (1–2), 17–34. Canfield Jr., D.E., Bachmann, R.W., Hoyer, M.V., 1983. Freeze-out of salts in hard-water
Assereto, R.L.A.M., Kendall, C.G., 1977. Nature, origin and classification of peritidal tepee lakes. Limnology and Oceanography 28 (5), 970–977.
structures and related breccias. Sedimentology 24 (2), 153–210. Capo, R.C., Chadwick, O.A., 1999. Sources of strontium and calcium in desert soil and
Atkinson, T.C., 1983. Growth mechanisms of speleothems in Castleguard Cave, calcrete. Earth and Planetary Science Letters 170 (1–2), 61–72.
Columbia Icefields, Alberta, Canada. Arctic and Alpine Research 15 (4), 523–536. Carlisle, D., 1983. Concentration of uranium and vanadium in calcretes and gypcretes.
Awramik, S.M., Buchheim, H.P., 2009. A giant, Late Archean lake system: the In: Wilson, R.C.L. (Ed.), Residual Deposits: Surface Related Weathering Processes
Meentheena Member (Tumbiana Formation; Fortescue Group), Western Australia. and Materials: Geological Society of London Special Publication, 11, pp. 185–195.
Precambrian Research 174 (3–4), 215–240. Cerling, T.E., 1984. The stable isotopic composition of modern soil carbonate and its
Balin, D.F., 2000. Calcrete morphology and karst development in the Upper Old Red Sandstone relationship to climate. Earth and Planetary Science Letters 71 (2), 229–240.
at Milton Ness, Scotland. In: Friend, P.F., Williams, B.P.J. (Eds.), New Perspectives on the Chafetz, H.S., Folk, R.L., 1984. Travertines: depositional morphology and the bacterially
Old Red Sandstone: Geological Society of London, Special Publication, 180, pp. 485–501. constructed constituents. Journal of Sedimentary Petrology 54 (1), 289–316.
Bar-Matthews, M., Matthews, A., Ayalon, A., 1991. Environmental controls of Chafetz, H.S., Rush, P.F., Utech, N.M., 1991. Microenvironmental controls on mineralogy
speleothem mineralogy in a karstic dolomitic terrain (Soreq Cave, Israel). Journal and habit of CaCO3 precipitates: an example from an active travertine system.
of Geology 99, 189–207. Sedimentology 38 (1), 107–126.
Basu, A., 1981a. Comment and reply on ‘Weathering before the advent of land plants: Chen, J., Zhang, D.D., Wang, S., Xiao, T., Huang, R., 2004. Factors controlling tufa
evidence from detrital K-feldspars in Cambrian–Ordovician arenites’. Geology 9 deposition in natural waters at waterfall sites. Sedimentary Geology 166 (3–4),
(11), 505–506. 353–366.
Basu, A., 1981b. Weathering before the advent of land plants: evidence from unaltered Clark, I.D., Lauriol, B., 1992. Kinetic enrichment of stable isotopes in cryogenic calcites.
detrital K-feldspars in Cambrian–Ordovician arenites. Geology 9 (3), 132–133. Chemical Geology 102 (1–4), 217–228.
Beeunas, M.A., Knauth, L.P., 1985. Preserved stable isotopic signature of subaerial Cotter, E., 1978. The evolution of fluvial style, with special reference to the central
diagenesis in the 1.2-b.y. Mescal Limestone, central Arizona: implications for the Appalachian Paleozoic. In: Miall, A.D. (Ed.), Fluvial Sedimentology: Canadian
timing and development of a terrestrial plant cover. Geological Society of America Society of Petroleum Geologists Memoir, vol. 5, pp. 361–383.
Bulletin 96, 737–745. Courty, M.A., Marlin, C., Dever, L., Tremblay, P., Vachier, P., 1994. The properties, genesis
Beier, J.A., 1987. Petrographic and geochemical analysis of caliche profiles in a and environmental significance of calcitic pendents from the High Arctic
Bahamian Pleistocene dune. Sedimentology 34, 991–998. (Spitsbergen). Geoderma 61 (1–2), 71–102.
Bekker, A., Holland, H.D., Wang, P.-L., Rumble III, D., Stein, H.J., Hannah, J.L., Coetzee, L.L., Coxon, C.E., 1994. Carbonate deposition in turloughs (seasonal lakes) on the Western
Beukes, N.J., 2004. Dating the rise of atmospheric oxygen. Nature 427, 117–120. Limestone Lowlands of Ireland. Irish Geography 27, 14–27.
Berner, R.A., 1992. Weathering, plants, and the long-term carbon cycle. Geochimica et Cromack Jr., K., Sollins, P., Todd, R.L., Fogel, R., Todd, A.W., Fender, W.M., Crossley, M.E.,
Cosmochimica Acta 56 (8), 3225–3231. Crossley, D.A.J., 1977. The role of oxalic acid and bicarbonate in calcium cycling by
Berner, R.A., 1998. The carbon cycle and carbon dioxide over Phanerozoic time: the role fungi and bacteria: some possible implications for soil animals. Ecological Bulletin,
of land plants. Philosophical Transactions of the Royal Society of London. Series B: Stockholm 25, 246–252.
Biological Sciences 353 (1365), 75–82. Cromack Jr., K., Sollins, P., Graustein, W.C., Speidel, K., Todd, A.W., Spycher, G., Li, C.Y., Todd,
Berner, E.K., Berner, R.A., Moulton, K.L., 2003. Plants and mineral weathering: present R.L., 1979. Calcium oxalate accumulation and soil weathering in mats of the hypogeous
and past. In: Holland, H.D., Turekian, K.K. (Eds.), Treatise on Geochemistry. fungus Hysterangium crassum. Soil Biology and Biochemistry 11 (5), 463–468.
Pergamon, Oxford, pp. 169–188. Daily, B., Moore, P.S., Rust, B.R., 1980. Terrestrial-marine transition in the Cambrian
Bernstein, L., Young, G.M., 1990. Depositional environments of the Early Proterozoic rocks of Kangaroo Island, South Australia. Sedimentology 27 (4), 379–399.
Espanola Formation, Ontario, Canada. Canadian Journal of Earth Sciences 27 (4), D'Alessandro, W., Giammanco, S., Bellomo, S., Parello, F., 2007. Geochemistry and
539–551. mineralogy of travertine deposits of the SW flank of Mt. Etna (Italy): relationships
Bertrand-Sarfati, J., Moussine-Pouchkine, A., 1983. Pedogenetic and diagenetic fabrics with past volcanic and degassing activity. Journal of Volcanology and Geothermal
in the Upper Proterozoic Sarnyéré formation (Gourma, Mali). Precambrian Research 165 (1–2), 64–70.
Research 20 (2–4), 225–242. Darrah, P.R., 1993. The rhizosphere and plant nutrition: a quantitative approach. Plant
Blum, J.D., Klaue, A., Nezat, C.A., Driscoll, C.T., Johnson, C.E., Siccama, T.G., Eagar, C., and Soil 155–156 (1), 1–20.
Fahey, T.J., Likens, G.E., 2002. Mycorrhizal weathering of apatite as an important Davies, N.S., Gibling, M.R., 2010. Cambrian to Devonian evolution of alluvial systems:
calcium source in base-poor forest ecosystems. Nature 417, 729–731. the sedimentological impact of the earliest land plants. Earth Science Reviews 98
Bolhar, R., Van Kranendonk, M.J., 2007. A non-marine depositional setting for the (3–4), 171–200.
northern Fortescue Group, Pilbara Craton, inferred from trace element geochem- Davies, N.S., Turner, P., Sansom, I.J., 2005. Caledonide influences on the Old Red Sandstone
istry of stromatolitic carbonates. Precambrian Research 155 (3–4), 229–250. fluvial systems of the Oslo Region, Norway. Geological Journal 40, 83–101.
Bonneville, S., Smits, M.M., Brown, A., Harrington, J., Leake, J.R., Brydson, R., Benning, L.G., Davies, N.S., Sansom, I.J., Turner, P., 2006. Trace fossils and paleoenvironments of a Late
2009. Plant-driven fungal weathering: Early stages of mineral alteration at the Silurian Marginal-Marine/Alluvial system: the Ringerike Group (Lower Old Red
nanometer scale. Geology 37, 615–618. Sandstone), Oslo Region, Norway. Palaios 21, 46–62.
Bono, P., Dreybrodt, W., Ercole, S., Percopo, C., Vosbeck, K., 2001. Inorganic calcite Davis, D.W., Paces, J.B., 1990. Time resolution of geologic events on the Keweenaw
precipitation in Tartare karstic spring (Lazio, central Italy): field measurements and Peninsula and implications for development of the Midcontinent Rift system. Earth
theoretical prediction on depositional rates. Environmental Geology 41 (3), 305–313. and Planetary Science Letters 97 (1–2), 54–64.
236 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

De Ros, L.F., Morad, S., Paim, P.S.G., 1994. The role of detrital composition and climate on the Grotzinger, J.P., 1986. Cyclicity and paleoenvironmental dynamics, Rocknest platform,
diagenetic evolution of continental molasses: evidence from the Cambro–Ordovician northwest Canada. Geological Society of America Bulletin 97, 1208–1231.
Guaritas Sequence, southern Brazil. Sedimentary Geology 92 (3–4), 197–228. Grotzinger, J.P., 1989. Facies and evolution of Precambrian carbonate depositional
Deynoux, M., 1982. Periglacial polygonal structures and sand wedges in the late systems: emergence of the modern platform archetype. In: Crevello, P.D., Wilson,
Precambrian glacial formations of the Taoudeni basin in Adrar of Mauretania (West J.W., Sarg, J.F., Read, J.F. (Eds.), Controls on Carbonate Platform and Basin
Africa). Palaeogeography, Palaeoclimatology, Palaeoecology 39 (1–2), 55–70. Development. SEPM, Tulsa, Oklahoma, pp. 79–106.
Dijkmans, J.W.A., Koster, E.A., Galloway, J.P., Mook, W.G., 1986. Characteristics and Grotzinger, J.P., Knoll, A.H., 1995. Anomalous carbonate precipitates: is the Precambrian
origin of calcretes in a subarctic environment, Great Kobuk Sand Dunes, the key to the Permian? Palaios 10 (6), 578–596.
Northwestern Alaska, U.S.A. Arctic and Alpine Research 18 (4), 377–387. Grover, G., Read, J.F., 1978. Fenestral and associated vadose diagenetic fabrics of tidal
Dix, G.R., Robinson, G.W., McGregor, D.C., 1998. Paleokarst in the Lower Ordovician flat carbonates, Middle Ordovician New Market Limestone, southwestern Virginia.
Beekmantown Group, Ottawa Embayment: structural control inboard of the Journal of Sedimentary Research 48 (2), 453–473.
Appalachian orogen. Geological Society of America Bulletin 110, 1046–1059. Guo, L., Riding, R., 1992. Aragonite laminae in hot water travertine crusts, Rapolano
Drever, J.I., 1994. The effect of land plants on weathering rates of silicate minerals. Terme, Italy. Sedimentology 39 (6), 1067–1079.
Geochimica et Cosmochimica Acta 58 (10), 2325–2332. Guo, L., Riding, R., 1994. Origin and diagenesis of Quaternary travertine shrub fabrics,
Dreybrodt, W., 1982. A possible mechanism for growth of calcite speleothems without Rapolano Terme, central Italy. Sedimentology 41 (3), 499–520.
participation of biogenic carbon dioxide. Earth and Planetary Science Letters 58 (2), Guo, L., Riding, R., 1998. Hot-spring travertine facies and sequences, Late Pleistocene,
293–299. Rapolano Terme, Italy. Sedimentology 45 (1), 163–180.
Duliński, M., Grabczak, J., Kostecka, A., Weclawik, S., 1995. Stable isotope composition of Guo, L., Andrews, J., Riding, R., Dennis, P., Dresser, Q., 1996. Possible microbial effects of
spelean calcites and gaseous CO2 from Tylicz (Polish Carpathians). Chemical stable carbon isotopes in hot-spring travertines. Journal of Sedimentary Research
Geology 125 (3–4), 271–280. 66 (3), 468–473.
Dunagan, S.P., Driese, S.G., 1999. Control of terrestrial stabilization on Late Devonian Gutzmer, J., Beukes, N.J., 1998. Earliest laterites and possible evidence for terrestrial
palustrine carbonate deposition; Catskill Magnafacies, New York, U.S.A. Journal of vegetation in the Early Proterozoic. Geology 26, 263–266.
Sedimentary Research 69 (3), 772–783. Hallet, B., Lorrain, R., Souchez, R., 1978. The composition of basal ice from a glacier
Dunn, J.R., 1953. The origin of the deposits of tufa in Mono Lake. Journal of Sedimentary sliding over limestones. Geological Society of America Bulletin 89 (2), 314–320.
Research 23 (1), 18–23. Hambrey, M.J., 1985. The late Ordovician–Early Silurian glacial period. Palaeogeogra-
Ebelmen, J.J., 1845. Sur les produits de la décomposition des espèces minérales de la phy, Palaeoclimatology, Palaeoecology 51 (1–4), 273–289.
famille des silicates. Annales des Mines 7, 3–66. Harmon, R.S., Atkinson, T.C., Atkinson, J.L., 1983. The mineralogy of Castleguard Cave,
Edwards, M.B., 1975. Glacial retreat sedimentation in the Smalfjord Formation, Late Columbia Icefields, Alberta, Canada. Arctic and Alpine Research 15 (4), 503–516.
Precambrian, North Norway. Sedimentology 22 (1), 75–94. Heward, A.P., 1989. Early Ordovician alluvial fan deposits of the Marmul oil field, South
Edwards, D., 1996. New insights into early land ecosystems: a glimpse of a Lilliputian Oman. Journal of the Geological Society 146 (3), 557–565.
world. Review of Palaeobotany and Palynology 90 (3–4), 159–174. Higgins, J.A., Schrag, D.P., 2003. Aftermath of a snowball Earth. Geochemistry,
Elmore, R.D., 1983. Precambrian non-marine stromatolites in alluvial fan deposits, the Geophysics, Geosystems 4, 1028. doi:10.1029/2002GC000403.
Copper Harbor Conglomerate, upper Michigan. Sedimentology 30 (6), 829–842. Hill, C.A., 1990. Sulfuric acid speleogenesis of Carlsbad Cavern and its relationship to
Emeis, K.C., Richnow, H.H., Kempe, S., 1987. Travertine formation in Plitvice National hydrocarbons, Delaware Basin, New Mexico and Texas. American Association of
Park, Yugoslavia: chemical versus biological control. Sedimentology 34 (4), Petroleum Geologists Bulletin 74, 1685–1694.
595–609. Hiltner, L., 1904. Über neuere Erfahrungen und Probleme auf dem Gebiete der
Esteban, M., Klappa, C.F., 1983. Subaerial exposure environment. In: Scholle, P.A., Bodenbakteriologie unter besonderer Berücksichtigung der Gründüngung und
Bebout, D.G., Moore, C.H. (Eds.), Carbonate Depositional Environments. American Brache. Arbeiten der Deutsche Landwirtschafts–Geselleschaft 98, 59–78.
Association of Petroleum Geologists, Tulsa, Oklahoma, pp. 1–54. Hinsinger, P., 1998. How do plant roots acquire mineral nutrients? Chemical processes
Eugster, H.P., Hardie, L.A., 1975. Sedimentation in an Ancient Playa-Lake Complex: the involved in the rhizosphere. Advances in Agronomy 64, 225–265.
Wilkins peak member of the Green River Formation of Wyoming. Geological Hinsinger, P., Plassard, C., Jaillard, B., 2006. Rhizosphere: a new frontier for soil
Society of America Bulletin 86, 319–334. biogeochemistry. Journal of Geochemical Exploration 88 (1–3), 210–213.
Evans, D.A., Beukes, N.J., Kirschvink, J.L., 1997. Low-latitude glaciation in the Hodell, D.A., Schelske, C.L., Fahnenstiel, G.L., Robbins, L.L., 1998. Biologically induced
Palaeoproterozoic era. Nature 386 (6622), 262–266. calcite and its isotopic composition in Lake Ontario. Limnology and Oceanography
Fairchild, I.J., Bradby, L., Spiro, B., 1993. Carbonate diagenesis in ice. Geology 21, 43 (2), 187–199.
901–904. Hoffman, P., 1976. Environmental diversity of Middle Precambrian Stromatolites. In:
Fairchild, I.J., Smith, C.L., Baker, A., Fuller, L., Spötl, C., Mattey, D., McDermott, F., E.I.M.F., Walter, M. (Ed.), Stromatolites. Developments in Sedimentology, 20. Elsevier,
2006. Modification and preservation of environmental signals in speleothems. Amsterdam, pp. 599–611.
Earth Science Reviews 75 (1–4), 105–153. Holland, H.D., Holland, H.J., Munoz, J.L., 1964. The coprecipitation of cations with
Fisher, R.F., 1972. Spodosol development and nutrient distribution under Hydnaceae CaCO3-II. The coprecipitation of Sr + 2 with calcite between 90° and 100 °C.
fungal mats. Soil Science Society of America Journal 36 (3), 492–495. Geochimica et Cosmochimica Acta 28 (8), 1287–1301.
Ford, T.D., Pedley, H.M., 1996. A review of tufa and travertine deposits of the world. Holland, H.D., Feakes, C.R., Zbinden, E.A., 1989. The Flin Flon paleosol and the
Earth Science Reviews 41 (3–4), 117–175. composition of the atmosphere 1.8 BYBP. American Journal of Science 289,
Forti, P., 2005. Genetic processes of cave minerals in volcanic environments: an 362–389.
overview. Journal of Cave and Karst Studies 67, 3–13. Horodyski, R.J., Knauth, L.P., 1994. Life on land in the Precambrian. Science 263,
Fouke, B.W., Farmer, J.D., Des Marais, D.J., Pratt, L., Sturchio, N.C., Burns, P.C., Discipulo, 494–498.
M.K., 2000. Depositional facies and aqueous-solid geochemistry of travertine- Hsieh, J.C.C., Chadwick, O.A., Kelly, E.F., Savin, S.M., 1998. Oxygen isotopic composition
depositing hot springs (Angel Terrace, Mammoth Hot Springs, Yellowstone of soil water: quantifying evaporation and transpiration. Geoderma 82 (1–3),
National Park, U.S.A.). Journal of Sedimentary Research 70 (3), 565–585. 269–293.
Freytet, P., Verrecchia, E.P., 1998. Freshwater organisms that build stromatolites: a Humphreys, C.P., Franks, P.J., Rees, M., Bidartondo, M.I., Leake, J.R., Beerling, D.J., in
synopsis of biocrystallization by prokaryotic and eukaryotic algae. Sedimentology press. Mutualistic mycorrhiza-like symbiosis in the most ancient group of land
45 (3), 535–563. plants. Nature Communications 1. doi:10.1038/ncomms1105.
Freytet, P., Verrecchia, E.P., 2002. Lacustrine and palustrine carbonate petrography: an Irion, G., Muller, G., 1968. Mineralogy, petrology and chemical composition of some
overview. Journal of Paleolimnology 27, 221–237. calcareous tufa from the Schwabische Alb, Germany. In: Muller, G., Friedman, G.M.
Frisia, S., Borsato, A., 2010. Karst. In: Alonso-Zarza, A.M., Tanner, L.H. (Eds.), Carbonates (Eds.), Recent Developments in Carbonate Sedimentology in Central Europe.
in Continental Settings: Facies, Environments, and Processes. Developments in Springer Verlag, Berlin, pp. 157–171.
Sedimentology, 61. Elsevier, Amsterdam, pp. 269–318. Irwin, M.L., 1965. General theory of epeiric clear water sedimentation. American
Glover, J.F., Kah, L.C., 2006. Speleothem deposits in a Proterozoic paleokarst, Association of Petroleum Geologists Bulletin 49, 445–459.
Mesoproterozoic dismal lakes group, Arctic Canada, Geological Society of America Jaillard, B., 1984. Mise en évidence de la néogénèse de sables calcaires sous I'action des
Southeastern Section — 55th Annual Meeting Knoxville, Tennessee, USA. Geological racines: incidence sur la granulométrie du sol. Agronomie 4, 91–100.
Society of America Abstracts with Programs 38 (3), 36. Jaillard, B., Guyon, A., Maurin, A.F., 1991. Structure and composition of calcified roots,
Golubic, S., Campbell, S.E., 1979. Analogous microbial forms in recent subaerial habitats and their identification in calcareous soils. Geoderma 50 (3), 197–210.
and in Precambrian cherts: Gloethece coerulea Geitler and Eosynechococcus moorei James, N.P., 1972. Holocene and Pleistocene calcareous crust (caliche) profiles, criteria
Hofmann. Precambrian Research 8, 201–217. for subaerial exposure. Journal of Sedimentary Research 42 (4), 817–836.
Gómez Peral, L.E., Poiré, D.G., Strauss, H., Zimmermann, U., 2007. Chemostratigraphy and James, N.P., Gravestock, D.I., 1990. Lower Cambrian shelf and shelf margin buildups,
diagenetic constraints on Neoproterozoic carbonate successions from the Sierras Flinders Ranges, South Australia. Sedimentology 37 (3), 455–480.
Bayas Group, Tandilia System, Argentina. Chemical Geology 237 (1–2), 109–128. James, N.P., Narbonne, G.M., Kyser, T.K., 2001. Late Neoproterozoic cap carbonates:
Goudie, A.S., 1996. Organic agency in calcrete development. Journal of Arid Mackenzie Mountains, northwestern Canada: precipitation and global glacial
Environments 32 (2), 103–110. meltdown. Canadian Journal of Earth Sciences 38 (8), 1220–1262.
Grasby, S.E., 2003. Naturally precipitating vaterite ([μ-CaCO3) spheres: unusual Jiang, G., Christie-Blick, N., Kaufman, A.J., Banerjee, D.M., Rai, V., 2003. Carbonate
carbonates formed in an extreme environment. Geochimica et Cosmochimica platform growth and cyclicity at a terminal Proterozoic passive margin, Infra Krol
Acta 67 (9), 1659–1666. Formation and Krol Group, Lesser Himalaya, India. Sedimentology 50 (5),
Gregory, P.J., 2006. Roots, rhizosphere and soil: the route to a better understanding of 921–952.
soil science? European Journal of Soil Science 57, 2–12. Jia-yu, R., Johnson, M.E., 1996. A stepped karst unconformity as an Early Silurian rocky
Griffiths, R.P., Baham, J.E., Caldwell, B.A., 1994. Soil solution chemistry of ectomycor- shoreline in Guizhou Province (South China). Palaeogeography, Palaeoclimatology,
rhizal mats in forest soil. Soil Biology and Biochemistry 26 (3), 331–337. Palaeoecology 121 (3–4), 115–129.
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 237

Johnson, I.M., Watanabe, Y., Stewart, B., Ohmoto, H., 2009. Earth's oldest (~3.4 Ga Mack, G.H., Cole, D.R., Trevino, L., 2000. The distribution and discrimination of shallow,
lateritic paleosol in the Pilbara Craton, Western Australia. Geochimica et authigenic carbonate in the Pliocene–Pleistocene Palomas Basin, southern Rio
Cosmochimica Acta 73 (13, Supplement 1), A601. Grande rift. Geological Society of America Bulletin 112, 643–656.
Jones, B., Renaut, R.W., 2010. Calcareous spring deposits in continental settings. In: Marriott, S.B., Wright, V.P., 1993. Palaeosols as indicators of geomorphic stability in two
Alonso-Zarza, A.M., Tanner, L.H. (Eds.), Carbonates in Continental Settings: Facies, Old Red Sandstone alluvial suites, South Wales. Journal of the Geological Society
Environments, and Processes. Developments in Sedimentology, 61. Elsevier, 150, 1109–1120.
Amsterdam, pp. 177–224. Martini, J.E.J., 1994. A late Archaean–Palaeoproterozoic (2.6 Ga) palaeosol on
Jongmans, A.G., van Breemen, N., Lundström, U., van Hees, P.A., Giesler, R., Melkerud, P.A., ultramafics in the Eastern Transvaal, South Africa. Precambrian Research 67 (1–2),
Olsson, M., 1997. Rock-eating fungi. Nature 389, 682–683. 159–180.
Jutras, P., Utting, J., McLeod, J., 2007. Link between long-lasting evaporitic basins and the Martini, I.P., Tongiorgi, M., Oggiano, G., Cocozza, T., 1991. Ordovician alluvial fan to marine
development of thick and massive phreatic calcrete hardpans in the Mississippian shelf transition in SW Sardinia, Western Mediterranean Sea: tectonically (“Sardic
Windsor and Percé groups of eastern Canada. Sedimentary Geology 201 (1–2), 75–92. phase”) influenced clastic sedimentation. Sedimentary Geology 72 (1–2), 97–115.
Jutras, P., Quillan, R.S., LeForte, M.J., 2009. Evidence from Middle Ordovician paleosols Matsubaya, O., Sakai, H., Torii, T., Burton, H., Kerry, K., 1979. Antarctic saline lakes—
for the predominance of alkaline groundwater at the dawn of land plant radiation. stable isotopic ratios, chemical compositions and evolution. Geochimica et
Geology 37, 91–94. Cosmochimica Acta 43 (1), 7–25.
Kabanov, P., Anadón, P., Krumbein, W.E., 2008. Microcodium: an extensive review and a McConnaughey, T., 1991. Calcification in Chara corallina: CO2 hydroxylation generates
proposed non-rhizogenic biologically induced origin for its formation. Sedimentary protons for bicarbonate assimilation. Limnology and Oceanography 36 (4), 619–628.
Geology 205 (3–4), 79–99. Meheruna, A., Akagi, T., 2006. Role of fine roots in the plant-induced weathering of
Kah, L.C., Lyons, T.W., Frank, T.D., 2004. Low marine sulphate and protracted andesite for several plant species. Geochemical Journal 40 (1), 57–67.
oxygenation of the Proterozoic biosphere. Nature 431, 834–838. Melezhik, V.A., Fallick, A.E., 2001. Palaeoproterozoic travertines of volcanic affiliation
Kalliokoski, J., 1986. Calcium carbonate cement (caliche) in Keweenawan sedimentary from a 13 C-rich rift lake environment. Chemical Geology 173 (4), 293–312.
rocks (1.1 Ga), Upper Peninsula of Michigan. Precambrian Research 32 (2–3), Melezhik, V.A., Fallick, A.E., Grillo, S.M., 2004. Subaerial exposure surfaces in a
243–259. Palaeoproterozoic 13 C-rich dolostone sequence from the Pechenga Greenstone
Kalliokoski, J., Welch, E.J., 1985. Keweenawan-age caliche paleosol in the lower part of Belt: palaeoenvironmental and isotopic implications for the 2330–2060 Ma global
the Calumet and Hecia Conglomerate, Centennial Mine, Calumet, Michigan. isotope excursion of 13 C/12 C. Precambrian Research 133 (1–2), 75–103.
Geological Society of America Bulletin 97, 1188–1193. Melezhik, V.A., Fallick, A.E., Hanski, E.J., Kump, L.R., Lepland, A., Prave, A., Strauss, H.,
Keller, M., Lehnert, O., 2010. Ordovician paleokarst and quartz sand: Evidence of 2005a. Emergence of the aerobic biosphere during the Archean–Proterozoic
volcanically triggered extreme climates? Palaeogeography, Palaeoclimatology, transition: Challenges of future research. GSA Today 15 (11), 4–11.
Palaeoecology 296 (3–4), 297–309. Melezhik, V.A., Fallick, A.E., Rychanchik, D.V., Kuznetsov, A.B., 2005b. Palaeoproterozoic
Keller, M., Buggisch, W., Bercowski, F., 1989. Facies and sedimentology of Upper evaporites in Fennoscandia: implications for seawater sulphate, the rise of
Cambrian shallowing-upward cycles in the La Flecha Formation (Argentine atmospheric oxygen and local amplification of the δ13C excursion. Terra Nova 17
Precordillera). Zentralblatt für Geologie und Paläontologie Teil 1 (5/6), 999–1011. (2), 141–148.
Kelly, E.F., Chadwick, O.A., Hilinski, T.E., 1998. The effect of plants on mineral Meng, X., Ge, M., Tucker, M.E., 1997. Sequence stratigraphy, sea-level changes and
weathering. Biogeochemistry 42 (1), 21–53. depositional systems in the Cambro–Ordovician of the North China carbonate
Kendall, A.C., Broughton, P.L., 1978. Origin of fabrics in speleothems composed of platform. Sedimentary Geology 114 (1–4), 189–222.
columnar calcite crystals. Journal of Sedimentary Petrology 48 (2), 519–538. Merz, M.U.E., 1992. The Biology of Carbonate Precipitation by Cyanobacteria. Facies 26,
Kenny, R., Knauth, L.P., 2001. Stable isotope variations in the Neoproterozoic Beck Spring 81–102.
Dolomite and Mesoproterozoic Mescal Limestone paleokarst: Implications for life on Moorbath, S., O'Nions, R.K., Pankhurst, R.J., 1973. Early Archaean Age for the Isua Iron
land in the Precambrian. Geological Society of America Bulletin 113, 650–658. Formation, West Greenland. Nature 245 (5421), 138–139.
Kenrick, P., Davis, P., 2004. Fossil Plants. Natural History Museum, London. 216 pp. Mora, C.I., Driese, S.G., Seager, P.G., 1991. Carbon dioxide in the Paleozoic atmosphere:
Khasawneh, F.E., Doll, E.C., 1978. The use of phosphate rock for direct application to Evidence from carbon-isotope compositions of pedogenic carbonate. Geology 19,
soils. Advances in Agronomy 30, 159–206. 1017–1020.
Killawee, J.A., Fairchild, I.J., Tison, J.L., Janssens, L., Lorrain, R., 1998. Segregation of Müller, G., Irion, G., Förstner, U., 1972. Formation and diagenesis of inorganic Ca–Mg
solutes and gases in experimental freezing of dilute solutions: implications for carbonates in the lacustrine environment. Die Naturwissenschaften 59 (4),
natural glacial systems. Geochimica et Cosmochimica Acta 62 (23–24), 3637–3655. 158–164.
Kirschvink, J.L., 1992. Late Proterozoic low-latitude global glaciation: the snowball Mustard, P.S., Donaldson, J.A., 1990. Paleokarst breccias, calcretes, silcretes and fault
earth. In: Schopf, J.W., Klein, C. (Eds.), The Proterozoic Biosphere. Cambridge talus breccias at the base of upper Proterozoic “Windermere” strata, northern
University Press, Cambridge, pp. 51–52. Canadian Cordillera. Journal of Sedimentary Research 60 (4), 525–539.
Klappa, C.F., 1978. Biolithogenesis of Microcodium: elucidation. Sedimentology 25 (4), Nash, D.J., McLaren, S.J., 2003. Kalahari valley calcretes: their nature, origins, and
489–522. environmental significance. Quaternary International 111 (1), 3–22.
Klappa, C.F., 1979. Lichen stromatolites: criterion for subaerial exposure and a Nash, D.J., Smith, R.F., 2003. Properties and development of channel calcretes in a
mechanism for the formation of laminar calcretes (caliche). Journal of Sedimentary mountain catchment, Tabernas Basin, southeast Spain. Geomorphology 50 (1–3),
Research 49 (2), 387–400. 227–250.
Klappa, C.F., 1980. Rhizoliths in terrestrial carbonates: classification, recognition, Ness, R.L.L., Vlek, P.L.G., 2000. Mechanism of Calcium and Phosphate Release from
genesis and significance. Sedimentology 27 (6), 613–629. Hydroxy-Apatite by Mycorrhizal Hyphae. Soil Science Society of America Journal 64
Knauth, L.P., Kennedy, M.J., 2009. The late Precambrian greening of the Earth. Nature (3), 949–955.
460 (7256), 728–732. Nickel, E., 1983. Environmental significance of freshwater oncoids, Eocene Guarga
Knoll, M.A., James, W.C., 1987. Effect of the advent and diversification of vascular land Formation, southern Pyrenees, Spain. In: Peryt, T.M. (Ed.), Coated Grains. Springer–
plants on mineral weathering through geologic time. Geology 15, 1099–1102. Verlag, Berlin, pp. 308–329.
Koban, C., Schweigert, G., 1993. Microbial Origin of Travertine Fabrics—Two Examples Ohlendorf, C., Sturm, M., 2001. Precipitation and Dissolution of Calcite in a Swiss High
from Southern Germany (Pleistocene Stuttgart Travertines and Miocene Rie- Alpine Lake. Arctic, Antarctic, and Alpine Research 33 (4), 410–417.
döschingen Travertine). Facies 29 (1), 251–263. Omelon, C.R., Pollard, W.H., Andersen, D.T., 2006. A geochemical evaluation of perennial
Košir, A., 2004. Microcodium revisited: root calcification products of terrestrial plants on spring activity and associated mineral precipitates at Expedition Fjord, Axel
carbonate-rich substrates. Journal of Sedimentary Research 74 (6), 845–857. Heiberg Island, Canadian High Arctic. Applied Geochemistry 21 (1), 1–15.
Lacelle, D., Lauriol, B., Clark, I.D., 2009. Formation of seasonal ice bodies and associated Osleger, D., Read, J.F., 1991. Relation of eustasy to stacking patterns of meter-scale
cryogenic carbonates in Caverne de l'Ours, Quebec, Canada: Kinetic isotope effects and carbonate cycles, Late Cambrian, U.S.A. Journal of Sedimentary Petrology 61 (7),
pseudo-biogenic crystal structures. Journal of Cave and Karst Studies 71, 48–62. 1225–1252.
Landing, E., MacGabhann, B.A., 2010. First evidence for Cambrian glaciation provided by Papadimitriou, S., Kennedy, H., Kattner, G., Dieckmann, G.S., Thomas, D.N., 2004.
sections in Avalonian New Brunswick and Ireland: Additional data for Avalon– Experimental evidence for carbonate precipitation and CO2 degassing during sea
Gondwana separation by the earliest Palaeozoic. Palaeogeography, Palaeoclimatol- ice formation. Geochimica et Cosmochimica Acta 68 (8), 1749–1761.
ogy, Palaeoecology 285 (3–4), 174–185. Pedley, H.M., 1990. Classification and environmental models of cool freshwater tufas.
Landing, E., Myrow, P., Alison, P.B., Narbonne, G.M., 1989. The Placentian Series: Sedimentary Geology 68 (1–2), 143–154.
Appearance of the Oldest Skeletalized Faunas in Southeastern Newfoundland. Pedley, M., González Martín, J.A., Ordóñez, S., García Del Cura, M.A., 2003.
Journal of Paleontology 63 (6), 739–769. Sedimentology of Quaternary perched springline and paludal tufas: criteria for
Leyval, C., Berthelin, J., 1991. Weathering of a mica by roots and rhizospheric recognition, with examples from Guadalajara Province, Spain. Sedimentology 50
microorganisms of pine. Soil Science Society of America Journal 55 (4), 1009–1016. (1), 23–44.
Lindsay, J.F., Brasier, M.D., McLoughlin, N., Green, O.R., Fogel, M., Steele, A., Mertzman, Pedley, M., Rogerson, M., Middleton, R., 2009. Freshwater calcite precipitates from
S.A., 2005. The problem of deep carbon—An Archean paradox. Precambrian in vitro mesocosm flume experiments: a case for biomediation of tufas.
Research 143 (1–4), 1–22. Sedimentology 56 (2), 511–527.
Lohmann, K.C., 1988. Geochemical Patterns of Meteoric Diagenetic Systems and Their Pelechaty, S.M., James, N.P., 1991. Dolomitized middle Proterozoic calcretes, Bathurst Inlet,
Application to Studies of Paleokarst. In: James, N.P., Choquette, P.W. (Eds.), Northwest Territories, Canada. Journal of Sedimentary Research 61 (6), 988–1001.
Paleokarst. Springer–Verlag, Berlin, pp. 58–80. Pelechaty, S.M., James, N.P., Kerans, C., Grotzinger, J.P., 1991. A middle Proterozoic
Love, K.M., Chafetz, H.S., 1988. Diagenesis of laminated travertine crusts, Arbuckle palaeokarst unconformity and associated sedimentary rocks, Elu Basin, northwest
Mountains, Oklahoma. Journal of Sedimentary Research 58 (3), 441–445. Canada. Sedimentology 38 (5), 775–797.
Machette, M.N., 1985. Calcic soils of the southwestern United States. Geological Society Pentecost, A., 1981. The tufa deposits of the Malham district, North Yorkshire. Field
of America Special Paper 203, 1–21. Studies 5, 365–387.
238 A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239

Pentecost, A., 1990. The formation of travertine shrubs; Mammoth Hot Springs, Southgate, P.N., Lambert, I.B., Donnelly, T.H., Henry, R., Etminan, H., Weste, G., 1989.
Wyoming. Geological Magazine 127 (2), 159–168. Depositional environments and diagenesis in Lake Parakeelya: a Cambrian alkaline
Pentecost, A., 1995. Geochemistry of carbon dioxide in six travertine-depositing waters playa from the Officer Basin, South Australia. Sedimentology 36 (6), 1091–1112.
of Italy. Journal of Hydrology 167 (1–4), 263–278. Spiro, B., Pentecost, A., 1991. One day in the life of a stream: a diurnal inorganic carbon
Pentecost, A., 2005. Travertine. Springer–Verlag, Berlin. 446 pp. mass balance for a travertine-depositing stream (waterfall beck, Yorkshire).
Pentecost, A., Viles, H., 1994. A review and reassessment of travertine classification. Geomicrobiology Journal 9 (1), 1–11.
Géographie Physique et Quaternaire 48 (3), 305–314. Spötl, C., Wright, V.P., 1992. Groundwater dolocretes from the Upper Triassic of the
Post, R.T., Long, D.G., 2008. The Middle Cambrian Mount Roosevelt Formation (new) of Paris Basin, France: a case study of an arid, continental diagenetic facies.
northeastern British Columbia: evidence for rifting and development of the Kechika Sedimentology 39 (6), 1119–1136.
Graben System. Canadian Journal of Earth Sciences 45, 483–498. Stewart, A.D., 2002. The Later Proterozoic Torridonian Rocks of Scotland: their Sedimentology,
Praekelt, H.E., Germs, G.J.B., Kennedy, J.H., 2008. A distinct unconformity in the Cango Geochemistry and Origin. Geological Society of London Memoir, 24. 125 pp.
Caves Group of the Neoproterozoic to early Paleozoic Saldania Belt in South Africa: Strother, P.K., Al-Hajri, S., Traverse, A., 1996. New evidence for land plants from the
its regional significance. South African Journal of Geology 111 (4), 357–368. lower Middle Ordovician of Saudi Arabia. Geology 24, 55–58.
Prave, A.R., 2002. Life on land in the Proterozoic: Evidence from the Torridonian rocks of Stworzewicz, E., Szulc, J., Pokryszko, B.M., 2009. Late Paleozoic continental gastropods
northwest Scotland. Geology 40, 811–814. from Poland: systematic, evolutionary and paleoecological approach. Journal of
Rainbird, R.H., Davis, W.J., Stern, R.A., Peterson, T.D., Smith, S.R., Parrish, R.R., Hadlari, T., Paleontology 83 (6), 938–945.
2006. Ar–Ar and U–Pb Geochronology of a Late Paleoproterozoic Rift Basin: Support Swett, K., 1974. Calcrete crusts in an arctic permafrost environment. American Journal
for a Genetic Link with Hudsonian Orogenesis, Western Churchill Province, of Science 274, 1059–1063.
Nunavut, Canada. Journal of Geology 114, 1–17. Syers, J.K., Birnie, A.C., Mitchell, B.D., 1967. The calcium oxalate content of some lichens
Read, J.F., Grover, G.A., 1977. Scalloped and planar erosion surfaces, Middle Ordovician growing on limestone. The Lichenologist 3 (03), 409–414.
limestones, Virginia; analogues of Holocene exposed karst or tidal rock platforms. Szulc, J., Cwizewicz, M., 1989. The lower permian freshwater carbonates of the Slawkow
Journal of Sedimentary Research 47 (3), 956–972. graben, Southern Poland: Sedimentary facies context and stable isotope study.
Renaut, R.W., Long, P.R., 1987. Freeze-out precipitation of salts in saline lakes— Palaeogeography, Palaeoclimatology, Palaeoecology 70 (1–3), 107–120.
examples from Western Canada. In: Strathdee, G.L., Klein, M.O., Melis, L.A. (Eds.), Tandon, S.K., Friend, P.F., 1989. Near-surface shrinkage and carbonate replacement
Crystallization and Precipitation. Pergamon, Oxford, pp. 33–42. processes, Arran Cornstone Formation, Scotland. Sedimentology 36 (6), 1113–1126.
Retallack, G.J., 1985. Fossil soils as grounds for interpreting the advent of large plants Thrailkill, J., 1971. Carbonate deposition in Carlsbad Caverns. Journal of Geology 79 (6),
and animals on land. Philosophical Transactions of the Royal Society of London. 683–695.
Series B: Biological Sciences 309, 105–142. Thrailkill, J., 1976. Speleothems. In: Walter, M.R. (Ed.), Stromatolites. Developments in
Retallack, G.J., 1986. Reappraisal of a 2200 Ma-old paleosol near Waterval Onder, South Sedimentology, 20. Elsevier, Amsterdam, pp. 73–86.
Africa. Precambrian Research 32 (2–3), 195–232. Tobin, K.J., Walker, K.R., 1994. Meteoric diagenesis below a submerged platform:
Retallack, G.J., 1990. Soils of the Past. Unwin-Hyman, London. 520 pp. implications for δ13C compositions prior to pre-vascular plant evolution, Middle
Retallack, G.J., 1997. Early forest soils and their role in Devonian global change. Science Ordovician, Alabama, U.S.A. Sedimentary Geology 90 (1–2), 95–111.
276, 583–585. Trewin, N.H., Rice, C.M., 1992. Stratigraphy and sedimentology of the Devonian Rhynie
Retallack, G.J., 2001. Scoyenia Burrows from Ordovician palaeosols of the Juniata chert locality. Scottish Journal of Geology 28, 37–47.
Formation in Pennsylvania. Palaeontology 44 (2), 209–235. Tuccimei, P., Giordano, G., Tedeschi, M., 2006. CO2 release variations during the last
Retallack, G.J., 2007. Soils and Global Change in the Carbon Cycle over Geological Time. 2000 years at the Colli Albani volcano (Roma, Italy) from speleothems studies.
In: Holland, H.D., Turekian, K.K. (Eds.), Treatise on Geochemistry. Pergamon, Earth and Planetary Science Letters 243 (3–4), 449–462.
Oxford, pp. 1–28. Tucker, M.E., Wright, V.P., 1990. Carbonate Sedimentology. Blackwell Science, Oxford.
Retallack, G.J., 2008. Cambrian paleosols and landscapes of South Australia. Australian 496 pp.
Journal of Earth Sciences 55, 1083–1106. Turnbull, M.J.M., Whitehouse, M.J., Moorbath, S., 1996. New isotopic age determina-
Retallack, G.J., Mindszenty, A., 1994. Well preserved late Precambrian paleosols from tions for the Torridonian, NW Scotland. Journal of the Geological Society 153 (6),
Northwest Scotland. Journal of Sedimentary Research 64 (2a), 264–281. 955–964.
Rex, G.M., Scott, A.C., 1987. The sedimentology, palaeoecology and preservation of the Turner, P., 1974. Lithostratigraphy and facies analysis of the Ringerike Group of the Oslo
Lower Carboniferous plant deposits at Pettycur, Fife, Scotland. Geological Magazine Region. Norges Geologiske Undersøkelse 314, 101–131.
124 (01), 43–66. Turner, P., Whitaker, J.H.M., 1976. Petrology and provenance of Late Silurian fluviatile
Richardson, C., 1976. Phase relationships in sea ice as a function of temperature. Journal sandstones from the Ringerike Group of Norway. Sedimentary Geology 16, 45–68.
of Glaciology 17, 507–519. Upfold, R.L., 1984. Tufted microbial (cyanobacterial) mats from the Proterozoic Stoer
Riding, R., 2000. Microbial carbonates: the geological record of calcified bacterial—algal Group, Scotland. Geological Magazine 121 (04), 351–355.
mats and biofilms. Sedimentology 47, 179–214. Urey, H.C., 1952. The Planets: Their Origin and Development. Yale University Press, New
Riding, R., 2008. Abiogenic, microbial and hybrid authigenic crusts: components of Haven, CT. 245 pp.
Precambrian stromatolites. Geologia Croatica 61 (2–3), 73–103. van Breemen, N., Finlay, R., Lundström, U., Jongmans, A.G., Giesler, R., Olsson, M., 2000.
Risacher, F., Eugster, H.P., 1979. Holocene pisoliths and encrustations associated Mycorrhizal weathering: A true case of mineral plant nutrition? Biogeochemistry
with spring-fed surface pools, Pastos Grandes, Bolivia. Sedimentology 26 (2), 49 (1), 53–67.
253–270. van Everdingen, R.O., Shakur, M.A., Krouse, H.R., 1985. Role of corrosion by H2SO4
Rogerson, M., Pedley, H.M., Wadhawan, J.D., Middleton, R., 2008. New insights into fallout in cave development in a travertine deposit—Evidence from sulfur and
biological influence on the geochemistry of freshwater carbonate deposits. oxygen isotopes. Chemical Geology 49 (1–3), 205–211.
Geochimica et Cosmochimica Acta 72 (20), 4976–4987. Veizer, J., Clayton, R.N., Hinton, R.W., 1992. Geochemistry of Precambrian carbonates:
Ryu, I.-C., Oh, C.W., Kim, S.W., 2005. A Middle Ordovician Drowning Unconformity on IV. Early Paleoproterozoic (2.25 ± 0.25 ga) seawater. Geochimica et Cosmochimica
the Northeastern Flank of the Okcheon (Ogcheon) Belt, South Korea. Gondwana Acta 56 (3), 875–885.
Research 8 (4), 511–528. Verrecchia, E.P., Dumont, J.-L., 1996. A biogeochemical model for chalk alteration by
Sami, T.T., James, N.P., 1994. Peritidal carbonate platform growth and cyclicity in an fungi in semiarid environments. Biogeochemistry 35 (3), 447–470.
Early Proterozoic foreland basin, Upper Pethei Group, Northwest Canada. Journal of Verrecchia, E.P., Verrecchia, K.E., 1994. Needle-Fiber Calcite; a critical review and a
Sedimentary Reseach B64 (2), 111–131. proposed classification. Journal of Sedimentary Research 64 (3a), 650–664.
Scott, A.C., Galtier, J., 1988. A new Lower Carboniferous flora from East Lothian, Verrecchia, E.P., Dumont, J.-L., Verrecchia, K.E., 1993. Role of calcium oxalate
Scotland. Proceedings of the Geologists' Association 99 (2), 141–151. biomineralization by fungi in the formation of calcretes; a case study from
Sears, S.O., Lucia, F.J., 1979. Reef-growth model for Silurian pinnacle reefs, northern Nazareth, Israel. Journal of Sedimentary Research 63 (5), 1000–1006.
Michigan reef trend. Geology 7, 299–302. Verrecchia, E.P., Freytet, P., Verrecchia, K.E., Dumont, J.-L., 1995. Spherulites in calcrete
Sharp, M., Tison, J.-L., Fierens, G., 1990. Geochemistry of subglacial calcites: implications for the laminar crusts: Biogenic CaCO3 precipitation as a major contributor to crust
hydrology of the basal water film. Arctic and Alpine Research 22 (2), 141–152. formation. Journal of Sedimentary Research A65 (4), 690–700.
Shear, W.A., 1991. The early development of terrestrial ecosystems. Nature 351, Vogt, T., Corte, A.E., 1996. Secondary precipitates in Pleistocene and present cryogenic
283–289. environments (Mendoza Precordillera, Argentina, Transbaikalia, Siberia, and
Shields, G.A., 2005. Neoproterozoic cap carbonates: a critical appraisal of existing Seymour Island, Antarctica). Sedimentology 43 (1), 53–64.
models and the plumeworld hypothesis. Terra Nova 17, 299–310. Wallander, H., 2000. Uptake of P from apatite by Pinus sylvestris seedlings colonised by
Shields, G.A., Brasier, M.D., Stille, P., Dorjnamjaa, D.-i., 2002. Factors contributing to high different ectomycorrhizal fungi. Plant and Soil 218 (1), 249–256.
δ13C values in Cryogenian limestones of western Mongolia. Earth and Planetary Wallander, H., Wickman, T., Jacks, G., 1997. Apatite as a P source in mycorrhizal and
Science Letters 196 (3–4), 99–111. non-mycorrhizal Pinus sylvestris seedlings. Plant and Soil 196 (1), 123–131.
Shields, G.A., Deynoux, M., Strauss, H., Paquet, H., Nahon, D., 2007. Barite-bearing cap Walter, M.R., 1996. Ancient hydrothermal ecosystems on Earth: a new palaeobiological
dolostones of the Taoudéni Basin, northwest Africa: Sedimentary and isotopic frontier, Evolution of Hydrothermal Ecosystems on Earth (and Mars?). Ciba
evidence for methane seepage after a Neoproterozoic glaciation. Precambrian Foundation Symposium, 202. John Wiley & Sons, Chichester, pp. 112–127.
Research 153 (3–4), 209–235. Watanabe, Y., Martini, J.E.J., Ohmoto, H., 2000. Geochemical evidence for terrestrial
Simpson, E.L., Eriksson, K.A., 1989. Sedimentology of the Unicoi Formation in southern ecosystems 2.6 billion years ago. Nature 408, 574–578.
and central Virginia: Evidence for late Proterozoic to Early Cambrian rift-to-passive Wellman, C.H., Gray, J., 2000. The microfossil record of early land plants. Philosophical
margin transition. Geological Society of America Bulletin 101, 42–54. Transactions of the Royal Society of London. Series B: Biological Sciences 355,
Skotnicki, S.J., Knauth, L.P., 2007. The Middle Proterozoic Mescal Paleokarst, Central 717–732.
Arizona, U.S.A.: karst development, silicification, and cave deposits. Journal of White, W.B., 1976. Cave minerals and speleothems. In: Ford, T.D., Cullingsham, T.D.H.
Sedimentary Research 77 (12), 1046–1062. (Eds.), The Science of Speleology. Academic Press, London, pp. 267–327.
A.T. Brasier / Earth-Science Reviews 104 (2011) 213–239 239

White, A.H., Youngs, B.C., 1980. Cambrian alkali playa–lacustrine sequence in the Wright, V.P., Sloan, R.J., Valero Garcés, B., Garvie, L.A.J., 1992. Groundwater ferricretes
northeastern Officer Basin, South Australia. Journal of Sedimentary Petrology 50, from the Silurian of Ireland and Permian of the Spanish Pyrenees. Sedimentary
1279–1286. Geology 77 (1–2), 37–49.
Wieder, M., Yaalon, D.H., 1974. Effect of matrix composition on carbonate nodule Wright, V.P., Platt, N.H., Marriott, S.B., Beck, V.H., 1995. A classification of rhizogenic
crystallization. Geoderma 11 (2), 95–121. (root-formed) calcretes, with examples from the Upper Jurassic–Lower Cretaceous
Wigley, T.M.L., 1973a. Chemical evolution of the system calcite–gypsum–water. of Spain and Upper Cretaceous of southern France. Sedimentary Geology 100 (1–4),
Canadian Journal of Earth Sciences 10, 306–314. 143–158.
Wigley, T.M.L., 1973b. The incongruent solution of dolomite. Geochimica et Cosmochi- Wright, V.P., Beck, V.H., Sanz-Montero, M.E., Verrecchia, E.P., Freytet, P., Verrecchia, K.E.,
mica Acta 37 (5), 1397–1402. Dumont, J.-L., 1996. Spherulites in calcrete laminar crusts; biogenic CaCO3
Williams, G.E., Schmidt, P.W., 1997. Palaeomagnetic dating of sub-Torridon Group precipitation as a major contributor to crust formation; discussion and reply.
weathering profiles, NW Scotland: verification of Neoproterozoic palaeosols. Journal of Sedimentary Research 66 (5), 1040–1044.
Journal of the Geological Society 154, 987–997. Yoshimura, K., Liu, Z., Cao, J., Yuan, D., Inokura, Y., Noto, M., 2004. Deep source CO2 in
Wright, V.P., 1984. The significance of needle-fibre calcite in a Lower Carboniferous natural waters and its role in extensive tufa deposition in the Huanglong Ravines,
palaeosol. Geological Journal 19, 23–32. Sichuan, China. Chemical Geology 205 (1–2), 141–153.
Wright, V.P., 1985. The Precursor Environment for Vascular Plant Colonization. Yuan, X., Xiao, S., Taylor, T.N., 2005. Lichen-Like Symbiosis 600 Million Years Ago.
Philosophical Transactions of the Royal Society of London. Series B: Biological Science 308 (5724), 1017–1020.
Sciences 309, 143–145. Žák, K., Urban, J., Cílek, V., Hercman, H., 2004. Cryogenic cave calcite from several
Wright, V.P., 2007. Calcrete. In: Nash, D.J., McLaren, S.J. (Eds.), Geochemical Sediments Central European caves: age, carbon and oxygen isotopes and a genetic model.
and Landscapes. Wiley–Blackwell, Oxford, UK, pp. 10–45. Chemical Geology 206 (1–2), 119–136.
Wright, V.P., Tucker, M.E., 1991. Calcretes: an Introduction. In: Wright, V.P., Tucker, M.E. Žák, K., Onac, B.P., Persoiu, A., 2008. Cryogenic carbonates in cave environments: A
(Eds.), Calcretes. Blackwell Scientific Publications, Oxford, UK, pp. 1–22. review. Quaternary International 187 (1), 84–96.
Wright, V.P., Platt, N.H., Wimbledon, W.A., 1988. Biogenic laminar calcretes: evidence of Zhang, D.D., Zhang, Y., Zhu, A., Cheng, X., 2001. Physical mechanisms of river waterfall
calcified root-mat horizons in paleosols. Sedimentology 35 (4), 603–620. tufa (travertine) formation. Journal of Sedimentary Research 71 (1), 205–216.

You might also like