Вы находитесь на странице: 1из 14

Eur. J. Mineral.

2005, 17, 7-20

Clay minerals and sedimentary basin history


RICHARD J. MERRIMAN*

British Geological Survey, Keyworth, Nottingham NG12 5GG, U.K.

Abstract: Clay minerals in the mud and soil that coat the Earth’s surface are part of a clay cycle that breaks down and creates rock
in the crust. Clays generated by surface weathering and shallow diagenetic processes are transformed into mature clay mineral
assemblages in the mudrocks found in sedimentary basins. During metamorphism, the release of alkali elements and boron from
clay minerals generates magmas that are subsequently weathered and recycled, representing the magma-to-mud pathway of the
clay cycle. Volcanogenic clay represents an important but hitherto underestimated proportion of recycled clay. Within sedimentary
basins, immature clays are transformed to mature and supermature clay assemblages by a series of reactions that generally obey
the Ostwald Step Rule. Bedding-parallel microfabric generated by these reactions produce significant changes in the physical
properties of deeply buried mudrocks. Clay minerals react to form equilibrium assemblages in 1 x 104 years in some hydrothermal
systems, but immature clays may survive for up to 2 x 109 years in mid-continental rift basins. Clay mineral assemblages and the
b cell dimension of K-white mica can be used to infer the geotectonic settings of sedimentary basins.

Key-words: clay minerals, sedimentary basins, clay cycle, clay maturity.

1. Introduction mature chemically, mineralogically and texturally. Basins


generated at active margins, i.e. where plate boundaries are
Clay minerals are the key constituents of the mud and in collision, commonly experience rapid sediment loading
soil that coat the Earth’s surface, forming a thin buffer zone and the contents are likely to mature by rapid burial,
between the atmosphere, hydrosphere and biosphere and followed by deformation and metamorphism. Where basins
the underlying crust of crystalline rock. This coating of have formed within stable cratons (e.g. intracontinental sag
clay, formed by chemical reactions between water and basins) or on some passive margins of long-lived rifts (e.g.
rock, is the most important product of the process of weath- the Atlantic margins), they experience slow subsidence
ering since it provides the essential constituents for the soil rates and may preserve non-metamorphosed sedimentary
and deep-sea mud supporting biogenic activity that first sequences that span a considerable part of the Phanerozoic
appeared on this planet around 3800 Ma. Without such time scale, with some surviving from Proterozoic times.
reactions rock dust would cover the Earth’s surface as it This review is written from a personal perspective that has
does on the Moon and other planets in the solar system. But evolved from some 40 years of applying clay mineral
while clay minerals may be vital to Earth’s life support studies to geological research. It explores how clay
system, they are also an essential group of minerals in a minerals behave in different basinal settings in response to
clay cycle that regenerates the crust and creates new crust the variable rates and types of maturation processes, and
from the underlying mantle (Merriman, 2002a). how clay maturation can be used to characterize the
Sedimentary basins are an important stage in the clay thermal and tectonic history of sedimentary basins.
cycle, where the predominantly juvenile clays formed by
surface weathering are collected into large deposits and
mature before they progress to the next stage of the cycle. 2. The clay cycle
In the broad sense, sedimentary basins are those areas of
the Earth’s crust where sediments accumulate to some Weathering can be regarded as the starting point of the
considerable thickness (> 0.5 km?) over geological time clay cycle since it produces new clay minerals from rock-
periods. Although they are highly variable in size, they tend forming minerals and also liberates existing clay minerals
to be at least 100 km long and tens of km wide (Einsele, (Fig. 1). Almost all of the common rock-forming minerals
2000). The tectonic setting of sedimentary basins strongly are capable of forming clay minerals during weathering,
influences deposition and the rate at which sediments with the notable exception of quartz. Minerals that initially

*E-mail: rjme@bgs.ac.uk

0935-1221/05/0017-0007 $ 6.30
DOI: 10.1127/0935-1221/2005/0017-0007 © 2005 E. Schweizerbart’sche Verlagsbuchhandlung. D-70176 Stuttgart
8 R. J. Merriman

Fig. 1. The clay cycle showing


different pathways for recycling clay
minerals in the Earth’s crust. After
weathering and pedogenesis,
neoformed and inherited clays are
deposited in a sedimentary basin.
Burial and tectonism transforms
clays to low-grade metamorphic
mudrocks composed mainly of
white mica and chlorite. Basin
inversion and uplift returns low-
grade mudrocks to surface weath-
ering, completing the clay cycle.
Deep burial and high-grade meta-
morphism leads to breakdown of
micas and magma generation.
Surface volcanism and weathering
begins a new clay cycle.

crystallized at the highest temperatures, for example those chlorite, marking the onset of greenschist facies metamor-
in a basalt, are least stable when they cool to surface condi- phism and temperatures above 300ºC.
tions and most susceptible to clay mineral formation. Early Further temperature increases and fabric-forming
models of the clay cycle (e.g. Eslinger & Pevear, 1988) deformation of mudrocks produces coarsely crystalline
tended to ignore the contribution from volcanic activity. metamorphic rocks, such as muscovite-chlorite schists and
Glass is the most voluminous product of volcanic erup- biotite gneisses. Where these rocks are assembled into a
tions, on the land and beneath the oceans, and readily alters mountain belt, crustal thickening causes further tempera-
to clay minerals. A conservative estimate of the volume of ture and pressure increases in the most deeply buried meta-
volcanic ash erupted annually is around 0.15 km3, half the morphic rocks, which begin to melt. Rocks containing
minimum annual global production of magma in volcanic micas and other hydrous minerals are usually the first to
arc settings; this is equivalent to at least a 2 mm-thick layer melt and generate migmatites, a mixture of melted and
of ash over the entire Earth’s surface every 1000 years non-melted rock. Magmas that escape from this zone of
(Sparks, 1999). If most of the annual production of ash is partial melting move upwards to form granitic intrusions in
converted to clay minerals, then this suggests that at least the crust, where they weather to begin another cycle of clay
2 % of the present day global sediment yield from rivers, mineral production. Where water circulates through a
estimated at 7.5 km3/year (Einsele, 2000) is probably system of shrinkage cracks and joints in consolidated
volcanogenic clay. Contributions from other sources, granitic intrusions, clays may also form by hydrothermal
particularly submarine volcanism, are more difficult to esti- alteration (Fig. 1).
mate, but are likely to double the above figure of 2 %. The mud-to-magma pathway of the cycle can be by-
Clay minerals produced by subaerial weathering are passed by the uplift of sedimentary basins or metamorphic
mostly transported into sedimentary basins where they are terrains, returning mudrocks and metapelites to surface
collected and matured as they move to the next phase of the weathering (Fig. 1). The rate at which the different path-
cycle. When clay minerals are buried in sedimentary basins ways complete the clay cycle vary considerably. In some
they undergo a series of transformations in response to the accretionary complexes clays can be recycled in less than 1
geothermal conditions within the basin, and the geotec- Ma, as in the case of the Pliocene-Pleistocene Coast Range
tonic evolution of the basin. In some passive margin basins, of eastern Taiwan (Dorsey et al., 1988). In contrast, the
and also some intercontinental sag and rift basins, burial pathway that results in orogenic magma genesis may take
without pervasive tectonism results in diagenetic transfor- well over 100 Ma to complete the cycle.
mations in response to temperatures that barely reach Clay minerals generated by submarine weathering are
200ºC. At active plate margins, diagenetic burial is usually also recycled. Sea floor basalts, making up nearly 60 % of
succeeded by low-grade metamorphism, due to higher the Earth’s crust, readily react with seawater at low temper-
geothermal gradients or tectonism, or a combination of atures to produce clay minerals which tend to be Fe- and
both. As a result, clay minerals are further transformed to Mg-rich, such as saponite and celadonite (Alt, 1999).
equilibrium assemblages dominated by white mica and Similar Fe, Mg-rich clays also form around submarine
Clay minerals and sedimentary basin history 9

Fig. 2. The origin of clay minerals in


different geotectonic settings. (a) Basin
formed in active margin setting.
Extensional back-arc basins contain
inherited clays from volcanic arc and
hinterland terranes, and also neoformed
volcanogenic and hydrothermal clays
formed in situ. Fore-arc and foreland
basins mainly contain inherited clays
from eroded mountain belts. Inherited
clays in accretionary prisms are trans-
formed to low-grade metamorphic clays
during underplating; transformed clays
also predominate in fold-and-thrust
belts. (b) Basins formed at passive
margins or within plate settings may
retain a higher proportion of neoformed
clays. Inherited clays are predominant in
typical passive margin basins, but
shallow shelf basins on passive margins
may contain a high proportion of
neoformed Fe-rich clays. Neoformed
clays also form around mid-oceanic
ridge systems and in basalt-floored
oceanic basins. Continental rift basins
have significant amounts of neoformed
clays. Inherited clays predominate in
continental sag basins formed on shields
with varied topographic relief.

hydrothermal mounds. These clays and other minerals pass before such activity becomes widespread and more of
generated by basalt-seawater reactions are closely associ- the vast accumulation of muddy sediment around the
ated with marine organic debris, distal clays discharged Atlantic is recycled into magma.
from continental river systems and wind-blown dust.
Together they give rise to the pelagic mud that covers wide
areas of typical oceanic basins. Recycling of oceanic crust 3. Modes of origin
takes place at active plate margins where approximately
half of the subducted sediment is accreted or underplated to Up to 60 % of strata in sedimentary basins are mudrock
generate new crust (Einsele, 2000). The remainder, lithologies such as mudstone, shale and slate (Potter et al.,
together with argilized basaltic rock, is conveyed into the 1980), with clay as the dominant mineral component. Clay
mantle on a slab of oceanic crust. The breakdown of clay minerals can form by a variety of processes that reflect
minerals which are carried on and within the subducted their sensitivity to physical and chemical changes at the
slab has been linked with the generation of magmas, partic- Earth’s surface and in the shallow crust. Within sedimen-
ularly by the release of the fluids containing alkali tary basins clay minerals have three modes of origin
elements and boron that promote partial melting of the (Millot, 1970; Eberl, 1984). Neoformed clay minerals are
mantle wedge beneath volcanic arcs (e.g. Mezler & formed in sedimentary basins by direct precipitation from
Wunder, 2000; Patino et al., 2000). The rate at which solution or via the crystallization of amorphous materials,
basalt-derived clays and pelagic mud complete the clay such as glassy volcanic ash. Inherited clay minerals are
cycle depends on the geotectonic setting of the basin. transported into sedimentary basins with little modifica-
Oceanic basins with active margins, such as the eastern tion, having been formed elsewhere. Transformed clay
Pacific, promote the fastest cycles, in some cases taking minerals develop as sedimentary basin mature, when both
less than 150 Ma to recycle basalt from mid-oceanic fissure inherited and neoformed clays react in response to diage-
volcanoes, perhaps with a thin drape of pelagic mud, to the netic and low-grade metamorphic conditions. Hence a
eruption of new lava above a subduction zone. In contrast, transformed origin indicates that the clay mineral devel-
clay recycling is much slower where certain types of oped from a pre-existing mineral, either by dissolution-
passive margin form the borders of major oceanic basins. precipitation or solid-state replacement, including resulting
Much of the Atlantic Ocean is fringed by passive margin overgrowths. Some well characterized transformations
basins containing mud-dominated sediments that have proceed by a sequence of reactions via a series of
accumulated for up to 200 Ma. Active recycling of these metastable intermediate products, as in the case of the
sediments is restricted to subduction zones in the southern smectite-to-illite transformation (e.g. Altaner & Ylagan,
Atlantic and the Caribbean. Tens of millions of years may 1997; Stixrude & Peacor, 2002). Prograde transformations
10 R. J. Merriman

Fig. 3. Schematic diagram showing the series of clay


mineral reactions that characterize transformations in
mudrock lithologies found in sedimentary basins.
Reaction progress is from top to bottom, indicated by
heavy arrows. Diagonal arrows indicate products
contributed from one reaction series to another. The
products of two or more reaction series constitute a
clay mineral assemblage in a typical mudrock.

are generally characterized by progressive increases in contain a high proportion of neoformed Fe-rich clays such
crystal thickness (“crystallinity”), and decreases in crystal as glaucony or verdine ‘facies’ minerals. Neoformed clays
defect densities, lattice strain and compositional variability also predominate around mid-oceanic ridge systems and in
(Peacor, 1992). These changes, together with the develop- basalt-floored oceanic basins. Continental rift basins or
ment of new microfabrics in clay minerals, convert soft some perimarine basins tend to contain significant amounts
muds to lithified mudstone and shale, and subsequently of neoformed clays, commonly formed in saline or hyper-
slate (Merriman & Peacor, 1999). Retrogressive transfor- saline conditions (e.g. Jones & Galan, 1988). These condi-
mations essentially reduce the crystal thickness of clay tions may also influence the type of clays that occur in
minerals by introducing stacking defects, and also increase continental ‘sag’ basins, but inherited clays normally
defect densities, lattice strain and compositional variability. predominate on shields with varied topographic relief.
Retrogression is the dominant mechanism of clay mineral These within-plate basins preserve some of the oldest non-
weathering and soil formation (Millot, 1970; Środoń, metamorphic clay assemblages.
1999), and biogenic degradation (e.g. McIlroy et al., 2003).
Regional retrogressive events in metamorphic terrains are
associated with fluid movement (e.g. Nieto et al., 1994; 4. Clay mineral transformations
Zhao et al., 1999; Abad et al., 2003).
The different thermal histories and geotectonic settings Prograde clay mineral transformations in a sedimentary
of basins are reflected in the origins of their clay mineral basin are primarily a response to burial and varying degrees
assemblages (e.g. Velde & Vasseur, 1992). Basins formed of tectonic fabric development at temperatures below
at active plate tectonic margins tend to have a greater diver- 300°C. Some transformations produce predictable patterns
sity of clay origins than passive margin or within-plate of change in mineral assemblages that can be used to char-
basins (Fig. 2). An extensional back-arc basin typically acterize the conditions of diagenesis and low-grade meta-
contains inherited clays derived from volcanic arc and morphism in basins. Such patterns were first recognized by
hinterland terranes, and also neoformed volcanogenic and Hower et al. (1976) from their study of the smectite-to-
hydrothermal clays formed within the basin. Basins that illite transformation in the Tertiary shales of the Gulf
develop by rapid subsidence, such as fore-arc and foreland Coast, U.S.A. During the past two decades, a series of
basins, tend to contain predominantly inherited clays studies stimulated by these observations have shown that,
derived from eroded mountain belts. In accretionary prisms as sedimentary basins mature, this transformation is part of
inherited clays are soon transformed to low-grade meta- a more extended series of reactions in 2:1 dioctahedral clay
morphic clay assemblages, especially during tectonic minerals that ultimately transform smectite to muscovite
underplating (e.g. Sample & Moore, 1987; Merriman & via a sequence of metastable intermediate products (e.g.
Roberts, 2001). Transformed clays also predominate in Moore & Reynolds, 1997; Merriman & Peacor, 1999). The
fold-and-thrust belts, which may develop when back-arc or series of reactions: smectite → mixed-layer illite/smectite
foreland basins are contracted and deformed. Basins (I/S) → illite → muscovite (Fig. 3) is characterized by a
formed at passive margins or within plates may retain a progressive increase in crystal thickness (illite “crys-
higher proportion of neoformed clays. Inherited clays in tallinity”), and decreases in crystal defect densities, lattice
passive margin basins may reflect the weathering condi- strain and compositional variability as illite-muscovite
tions and the maturity of the hinterland. For example, juve- becomes better ordered (Peacor, 1992). In marine sedimen-
nile clays such as kaolinite or smectite may be the tary basins this is the most widely reported and best char-
dominant inherited minerals derived by tropical weathering acterized clay mineral reaction series, despite ongoing
whereas illite and chlorite are more likely to be derived controversy on the precise nature of reactants and products
from temperate weathering of a mature hinterland. In (e.g. Peacor, 1998; Stixrude & Peacor, 2002). Although
contrast, shallow shelf basins on passive margins may rarely very abundant, a sub-series of reactions involving
Clay minerals and sedimentary basin history 11

Na-rich 2:1 dioctahedral minerals has been reported in


some mudrock sequences (Fig. 3). Products first appear in
late diagenetic mudrocks as nanometer-scale intermediate
Na/K-mica coexisting with K-micas, and these evolve to
form 2M1 paragonite and muscovite in higher grade slates
(Merriman & Roberts, 1985; Jiang & Peacor, 1993; Li
et al., 1994; Livi et al., 1997).
A similar sequence of transformations in 2:1 trioctahe-
dral Mg,Fe-rich clay minerals is analogous to the smectite-
I/S-illite-muscovite reaction series. Although trioctahedral
saponite (Mg-dominant in octahedral sites) is much less
plentiful than dioctahedral smectite in sedimentary basins,
it is the most common neoformed clay mineral replacing
mafic volcanic glass, microcrystalline mafic lava debris,
and Mg,Fe-silicates including olivine, pyroxene, amphi-
bole and biotite. During prograde transformation saponite
is replaced by corrensite, ordered 1:1 mixed-layered chlo-
rite/smectite (C/S), or vermiculite/chlorite; corrensite is in Fig. 4. Schematic diagram illustrating the Ostwald Step Rule as
turn replaced by trioctahedral chlorite in the late diagenetic applied to reactants and products in the smectite-I/S-illite-
zone or low anchizone (Fig. 3). This series of reactions in muscovite reaction series. Direct transformation from smectite to
trioctahedral clays has been observed in a wide variety of muscovite requires a large activation energy ∆Ga, and is unlikely in
a normal sedimentary basin setting. Transformation via a series of
environments, including the prograde regional metamor-
intermediate products requires smaller activation energies, and is
phism of mudrocks from diagenesis through to the epizone, kinetically more likely in sedimentary basins (modified from
in hydrothermal systems and in hydrothermally metamor- Putnis, 1992; fig. 11.53).
phosed oceanic crust (Alt, 1999; Merriman & Peacor,
1999, and references therein). However, in the majority of
immature mudrocks, trioctahedral smectite is a scarce
component whereas chlorite is a common end-product of equilibrium (Fig. 4). In the low temperature environment of
this reaction series in mature mudrocks. This suggests that sedimentary basins, clay mineral transformations with low
the Fe and Mg necessary for chlorite formation is total energy requirements are likely to be favoured in terms
contributed from other reactions, most likely from diocta- of reaction progress (Stixrude & Peacor, 2002). Activation
hedral smectite where Fe + Mg is usually higher than in energies are minimized in minerals related by similar
reaction product illite (Hower et al., 1976; Drief & Nieto, chemical compositions and/or crystal structure allowing a
2000). sequence of reactants and products (reaction series) to form
Although products of the kaolin 1:1 group of minerals in 2:1 dioctahedral and trioctahedral clay minerals
are reported as minor constituents of clay assemblages in (Merriman & Peacor, 1999). Because of the metastable
many mudrock sequences, it is not clear whether the trans- nature of clay assemblages, the reactants and products of a
formations shown in Fig. 3: kaolinite → dickite/nacrite series may coexist and be closely intergrown. While it is
(+SiO2) → pyrophyllite form a reaction series. Examples likely to be kinetically more favourable for clay minerals to
of kaolinite transforming to the 2-layer polytype dickite are be produced by reactions within a given series, it does not
well known in sandstone lithologies (e.g. Ehrenberg et al., preclude reactions in one series contributing products to
1993; Ruiz Cruz & Andreo, 1996). However, the kaolinite another series, as indicated by the diagonal arrows in
→ dickite/nacrite transformation has not been widely Fig. 3. The example of dioctahedral smectite contributing
recorded in mudrocks, possibly because of problems with to chlorite-forming reactions has been mentioned above.
the XRD identification of polytypic reflections from these Further examples include that of kaolinite-bearing
minerals in the presence of other clay minerals. Using TEM mudstones and shales where illite or chlorite can form by
techniques, Buatier et al. (1997) showed that kaolinite the breakdown products of kaolinite reacting with available
transforms to dickite within black marls in a thrust-fault Fe and Mg (e.g. Boles & Franks, 1979), and dickite
zone. In aluminous, Fe-poor mudrocks, kaolinite will react replacement by illite in sandstones (e.g. Patrier et al.,
in the presence of quartz to form pyrophyllite, and this has 2003).
been mapped as a reaction isograde in very low-grade Mineral fabrics associated with prograde metamorphic
metamorphic terrains (Frey, 1987). reactions have been widely recognized and recorded by
The formation of intermediate metastable mixed-layer metamorphic petrologists. Although such fabrics are easily
minerals in dioctahedral and trioctahedral 2:1 clay minerals observed in most meta-sedimentary rocks with a petrolog-
suggests that this type of reaction progress obeys the ical microscope, the submicroscopic fabrics found in
Ostwald step rule (Morse & Casey, 1988; Essene & Peacor, diagenetic and very low-grade mudrocks require more
1995). The rule recognizes that each phase transition powerful microscopes. During the past two decades, elec-
and/or crystal-structural modification in the reaction series tron microscopic techniques have revealed two types of
reduces the Gibbs free energy of the system, and with each diagenetic microfabric produced by the bedding-parallel
step the system approaches chemical (thermodynamic) orientation of the a-b crystallographic plane of clay
12 R. J. Merriman

isotropic microfabric of diagenetic mudrocks and also


reduces porosity (Worden et al., 2003). These changes
represent the beginning of burial metamorphism.
A second type of bedding-parallel fabric is found in
chlorite-mica stacks that originated as detrital components
of some mud-turbidites (Fig. 6). The stacks are predomi-
nantly the products of the trioctahedral 2:1 reaction series,
and probably represent a variety of ferromagnesian mineral
grains, including biotite, and also mafic volcanic detritus.
Replacement of ferromagnesian and mafic detritus by
trioctahedral smectite occurred either during initial weath-
ering and transport or in the early stages of diagenesis. The
development of thin white mica laminae as intergrowths
with approximately the same orientation of the a-b crystal-
lographic plane as the host trioctahedral phase appears to
Fig. 5. Transmission electron micrograph showing bedding-parallel
have occurred during late diagenesis or very low-grade
microfabric in illite. Note the absence of compaction or deflection
metamorphism (Dimberline, 1986; Milodowski &
microtextures in straight illite crystals enclosing the rigid quartz silt
grain. Gulf Coast Shale, 4745m depth (TEM by Hailiang Dong).
Zalasiewicz, 1991). Some of the white mica laminae may
represent a by-product of the corrensite to chlorite reaction
(Fig. 3), whereas others formed by migration of K and Na
into the stacks (Li et al., 1994). Despite the varied shape of
minerals. Bedding-parallel microfabric in illite appears to the stacks, the stacking planes maintain an overall bedding-
develop when the smectite-to-illite reaction has progressed parallel orientation in mudrocks that lack well-developed
beyond 60 % illite in I/S (Fig. 5). The characteristics of this tectonic fabrics. Both types of microfabric, in illite and
microfabric include: a) straight, defect-free illite crystal- chlorite-mica stacks, are probably related to the approxi-
lites replacing smectite crystallites with a high density of mately ground-parallel isocryst surfaces and mineral zones
defects (e.g. Dong et al., 1997); b) penetration of prograde found in burial metamorphic sequences (Roberts et al.,
illite crystallites into pore space, such as compaction 1996; Neuhoff et al., 1997).
shadows around rigid, non-clay detrital grains; c) termina-
tion of illite crystallites against rigid detrital grains at high
angles to the a-b crystallographic plane (e.g. Warr & Nieto, 5. Basin maturity
1998; Merriman & Peacor, 1999). The development of
bedding-parallel microfabric replaces the random or Several factors contribute interactively to clay mineral
reactions rates in sedimentary basins, including heat flow,
fluid movement and composition, sedimentary overburden
and tectonic deformation. Although the products of clay
reactions cannot be used directly as geothermometers
(Essene & Peacor, 1995), correlation of reaction progress
with organic thermal maturity indicators (Merriman &
Kemp, 1996; Merriman & Frey, 1999) suggests that three
stages of basin maturity can be recognized in terms of clay
mineral evolution (Fig. 7). Immature basins contain sedi-
mentary sequences that have not passed through the oil
window and are dominated by neoformed and inherited
clays of the shallow diagenetic zone. Clays in this type of
basin may have experienced initially rapid reaction rates
during neoformation, for example palygorskite formation
in saline lakes and perimarine basins, but have not been
buried deeply enough to transform to more evolved clay
mineral assemblages. However, some rapidly subsided,
deep basins can remain immature because of low heat flow
and the reaction-inhibiting effects of overpressure on both
Fig. 6. Back-scattered scanning electron micrograph showing
bedding-parallel microfabric in chlorite-mica stacks. The stacks are
clays and organic materials (e.g. Carr, 1998). Clay mineral
composed of chlorite (bright grey) and thin laminae of white mica assemblages in immature basins can provide useful infor-
(mid to dark grey) sharing the same crystallographic a-b stacking mation on sediment provenance and mechanisms of
planes. Primary bedding lamination is orientated ENE-WSW; a neoformation (e.g. Hillier, in press), and also provide
weak spaced cleavage cuts the lamination and is orientated WNW- evidence of the distribution and sources of volcanogenic
ESE. Despite the variation in size and shape of the stacks, the clays (Jeans et al., 2000). Most commercial clay deposits
stacking planes maintain the same overall bedding-parallel orienta- are found in immature basins, including bentonite, ball
tion. Silurian mudstone, central Wales (see Li et al., 1994). clay, flint clay and brick clay.
Clay minerals and sedimentary basin history 13

Fig. 7. Basin maturity chart correlating clay maturity zones and Kübler index of illite “crystallinity” with organic maturity indices and
hydrocarbon maturation zones. Modified from Merriman & Kemp (1996).

Mature basins contain sedimentary rocks that are reaction progress in clays is restored (e.g. Colten-Bradley,
within the oil and gas windows. In terms of clay minerals 1987). Tectonic fabric formation may be initiated in the
they are characterized by the partial transformation of low anchizone by mechanical kinking and rotation of
neoformed and inherited clays in response to deep diagen- detrital grains, and crenulation of bedding parallel micro-
esis. Reaction rates are largely controlled by basin heat fabric in clays (van der Pluijm et al., 1998). As tempera-
flow. Hence high-heat flow basins, such as back-arc or tures increase through the high anchizone and epizone
marginal basins, mature relatively rapidly under a thin (Fig. 7), chemical processes such as dissolution/precipita-
overburden, whereas low heat-flow basins such as fore-arc tion, and mechanical processes including intracrystalline
basins and some passive margin basins are slow to mature deformation and dislocation creep, contribute interactively
even under a thick overburden. The transformation of to the crystal growth of metamorphic phyllosilicates
smectite to illite is largely completed at this stage of basin (Merriman et al., 1995; Árkai et al., 2002). White micas
evolution, and has the effect of converting soft clays and that develop in tectonic fabrics can be used to date
mudstone into well-lithified, brittle mudstones and shales. cleavage-forming events in supermature basins using
Although a bedding-parallel microfabric may be generated 40Ar-39Ar laser microprobe techniques (Sherlock et al.,
by illite crystal growth in the mudstone or shale, tectonic 2003).
fabrics are usually absent or very weakly developed. Three types of observation and measurement, detailed
Reaction progress can be measured using % illite in I/S, or in Merriman & Peacor (1999), can be made to assess reac-
clay “crystallinity” (see below), and used to characterize tion progress in transformed clay mineral as an indication
hydrocarbon zones (e.g. Pollastro, 1993), and estimate of sedimentary basin maturity:
maximum palaeotemperatures (Środoń & Clauer, 2001). The clay mineral assemblage
Dating of illite using the K-Ar technique can be used to Quantification of mixed-layer minerals
model the burial history of mature basins (Środoń, 2002). Clay mineral “crystallinity” – Kübler and/or Árkai index
Supermature basins generally lack any hydrocarbon Clay mineral assemblages. In a typical mudrock, the
potential and represent the early stages of regional meta- clay mineral assemblage comprises the products of two or
morphism. In basins of this type deformation plays a more reaction series and can be used to indicate the state of
crucial part in completing the transformation of metastable reaction progress in terms of metapelitic grade (Merriman
diagenetic and inherited clays to equilibrium assemblages & Peacor, 1999, fig. 2.1) or the clay mineral maturity zones
of white mica (including muscovite, phengite, paragonite) shown in Fig. 7. Because of the metastable nature of clay
and chlorite. Deformation also returns overpressured assemblages, the reactants and products of a series may
basins to hydrostatic pressures, so that any retardation of coexist and be closely intergrown. As a consequence it is
14 R. J. Merriman

Fig. 8. Schematic plot illustrating the time taken for some clay minerals to progress from mature to supermature (very low-grade meta-
morphic) assemblages in different sedimentary basin settings. The x-axis shows typical geothermal gradients for a range of basin settings.
The y-axis is a log-scale estimate of the geological time interval between basin formation and the transformation to a supermature clay
mineral assemblage.

not possible to limit the stability of a particular clay lite thickness measurements and predicted by the Scherrer
mineral to a single zone. As general rule however, the pres- equation (Warr & Nieto; 1998; Merriman & Peacor, 1999).
ence of neoformed clays such as smectite and kaolinite in Progress in the trioctahedral 2:1 reaction series can be also
assemblages is a good indication of an immature stage of be monitored using the Árkai Index of chlorite “crys-
basin evolution (shallow diagenetic zone), whereas an tallinity” (Árkai, 1991; Guggenheim et al., 2002). This
assemblage of mature illite-muscovite and chlorite indi- measurement also provides a useful link with mineral
cates a supermature basin (Fig. 7). Most mixed-layer facies indicators of grade in meta-igneous rocks (Árkai &
minerals are formed when immature clays are transformed Ghabrial, 1997).
to mature mudrocks, i.e. the shallow-to-deep diagenetic
zone transition (Fig. 7), and become scarce or absent in
supermature basins. 6. Clay mineral maturity rates
Quantification of mixed-layer minerals. Early reaction
progress in 2:1 layer silicates (Fig. 3) generates mixed- As discussed in Section 2, the history of clays in sedi-
layer clays with varying proportions of reactant and mentary basins largely depends on geotectonic setting. Clays
product. Measurement of the proportions and degree of in some basins are uplifted and recycled in less than 1 million
ordering in the smectite-I/S-illite reaction series is the most years whereas those in other basinal sequences survive for
widely used indicator of reaction progress for the immature orders of magnitude longer. Since differences in heat flow
to mature stage of basin evolution (Środoń, 1999, and are the main distinguishing characteristic of the various
references therein). NEWMOD computer modelling of the geotectonic settings where basins develop, to what extent
mixed-layer components has become the standard tech- does thermal history influence clay mineral maturity rates?
nique of quantification (Moore & Reynolds, 1997). Heating rates in sedimentary basins, which are in the
Clay mineral “crystallinity”. This type of XRD range 1-10°C/million years, are derived from the product
measurement has been widely used as an indicator of reac- of the geothermal gradient (°C/km) and sedimentation rates
tion progress in mudrock lithologies found in mature and (metres/million years). Where sedimentation rates are
supermature basins. The Kübler index of illite “crys- unknown, as is often the case, basin thermal history is
tallinity” (KI) can be used to indicate reaction progress commonly described in terms of geothermal gradients,
when the smectite-to-illite reaction has progressed to which are typically in the range 15-60°C/km. In Fig. 8,
> 80 % illite. The KI technique measures changes in the typical geothermal gradients for basins developed in
width of the first basal reflection of dioctahedral illite- different geotectonic settings have been used schematically
muscovite at an XRD spacing of approximately 10 Å. The to illustrate how long clays appear to survive as immature
width of the 10-Å peak is inversely related to the thickness or mature clays before transforming to equilibrium assem-
of illite-muscovite crystallites, as shown by TEM crystal- blages typical of supermature basins. It must be empha-
Clay minerals and sedimentary basin history 15

sized that the information used to construct Fig. 8 was diffi- are eliminated the initially amorphous kerogen solids are
cult to come by, and in some cases depended on experi- progressively aromatised to incipient graphite crystals.
enced guesswork, largely because clay mineralogists in the Clay minerals on the other hand are usually crystalline
past paid scant attention to basin type or thermal history throughout the reactions that typically occur during basin
when sampling clays for analysis. Given these shortcom- diagenesis, and the greatest loss is water that may mobilize
ings, Fig. 8 indicates that the oldest immature clay minerals variable amounts of cations such as Ca, Na, K and Si.
are likely to be found in mid-continental ‘sag’ basins or Differences in the reaction kinetics of clay and organic
failed rifts, and may survive for up to 2000 Ma. In basins materials were demonstrated in a study of the Pannonian
of this type a combination of low or medium geothermal Basin System by Hillier et al. (1995). They found that
gradients and non-pervasive tectonic events, such as correlation between R0 % and progress in the smectite-to-
faulting but not fabric-forming events, generally results in illite reaction varied systematically from one sub-basin to
low maturity rates. For example, in the Middle Proterozoic another. In sub-basins characterized by lower heating rates
Lake Superior Basin, a mid-continental rift, the Nonesuch the smectite-to-illite reaction is advanced in relation to
Formation contains immature 1Md illite-rich I/S and a one- R0 %, whereas smectite-to-illite reactions are retarded rela-
layer chlorite, indicating that matrix clays are still at the tive to R0 % at higher heating rates.
deep diagenetic stage of basin evolution 1100 Ma after Similar differences at more advanced stages of basin
deposition (Li et al., 1995). An older example is provided maturity are shown in Fig. 9, where two trends are
by the Middle Proterozoic Kombolgie Supergroup, 1800 displayed. A plot of Kübler indices and R0 % values derived
Ma, deposited in the mid-cratonic McArther Basin, from a wide range of basin types and covering a broad
Northern Territories, Australia. Diagenetic dickite in sand- spectrum of geothermal gradients, from 15-45°C/km, is
stone pores is preserved but locally replaced by deep diage- labeled a ‘normal’ correlation curve. A similar trend was
netic illite when the sequence was buried to depths of 5 km identified by Underwood et al. (1993). A second curve was
or more (Patrier et al., 2003). derived from coal-bearing Carboniferous strata in the
At the higher geothermal gradients shown in Fig. 8, in South Wales Coalfield (White, 1992), and the Clare
basins such as the Pannonian Basin system, clay mineral Coalfield in southern Ireland (Wagner, 2003). The South
maturity rates are relatively rapid, reaching the mature Wales Coalfield appears to have experienced high heat
basin stage by simple burial in 8-10 Ma (Hillier et al., flow (50-75°C/km) that matured coals to anthracite rank
1995). In very high heat-flow geothermal systems, such as (Alderton & Bevins, 1996), and the Clare Coalfield is
the Salton Sea geothermal field, clays reach the super- suspected to have a similar history. The two curves suggest
mature stage during a heating event that lasted 10,000 years that at high heating rates vitrinite matures more rapidly
(Yau et al., 1988; Velde & Lanson, 1993), i.e. up to three than illite but maturation of vitrinite is slower where the
orders of magnitude quicker than relatively high heat-flow normal geothermal gradients of burial diagenesis and very
sedimentary basins. However, at the other end of the range low-grade metamorphism apply.
of typical geothermal gradients, clays in accretionary A significant difference between the mudrocks repre-
complexes assembled under very low heat flow conditions sented by the two curves in Fig. 9 is the development of
may also mature rapidly, particularly where they have been slaty cleavage fabrics. In the case of the ‘normal’ heat-flow
tectonised. In the Barbados accretionary complex, for data, the mudrocks with KI values less than 0.30∆˚2θ have
example, the smectite-to-illite transformation is initiated at a well-developed penetrative slaty cleavage (e.g. Merriman
temperatures of only 30°C below a major décollement zone & Peacor, 1999, fig. 2.1). In contrast, the mudrocks from
(Buatier et al., 1992). Moreover, when clay-rich rocks are which data for the high heat-flow curve were derived lack
subducted and underplated, lattice strain resulting from any cleavage fabric, even those with R0 > 5 % (Wagner,
accretionary tectonism plays a significant role in 2003). Again, this emphasizes the importance of tectonism
progressing clay reactions through the mature → superma- in progressing clay reactions through the mature → super-
ture transformation (Merriman et al., 1995). mature transformation, and underlines the conclusions of
The general conclusion to be drawn from Fig. 8 is that Essene & Peacor (1995) that clay mineral reactions are not
the rate at which clay mineral reactions progress from accurate geothermometers.
mature to supermature assemblages can range from 1 x 104
to 2 x 109 years (or more), and both heat and tectonism
contribute to the more rapid rates of maturation. 7. Geotectonic setting and clay minerals
Although clay mineral maturity and organic maturation,
particularly vitrinite reflectance (R0 %), can be correlated The range of tectonothermal conditions under which
in simplified diagrams, such as Fig. 7, it is widely recog- sedimentary basins evolve (Fig. 8) suggests that the clay
nized that these two types of shallow crustal materials mineral assemblages developed should be indicative of
mature at different rates. Both Robert (1985) and Kisch different geotectonic settings. But can differences that may
(1987) suggested that clay mineral reactions tended to be developed at the immature or mature stages of basin evolu-
slower to respond to temperature increases than vitrinite tion be preserved in supermature clay mineral assemblages
reflectance. This is perhaps to be expected given the that result from low-grade metamorphism?
different nature of reaction progress in these two materials. Very few studies have made detailed regional compar-
Vitrinite maturation progresses by loss of water, carbon isons between clay minerals generated in different geotec-
dioxide, methane and higher hydrocarbons, and as these tonic settings. Recently, Merriman (2002b) reviewed
16 R. J. Merriman

Fig. 9. Plot of vitrinite reflectance R0 % against the Kübler index of illite “crystallinity”. The curve labeled ‘normal’ heat flow represents
a wide variety of basin types, and is based on unpublished and published data (Dalla Torre et al., 1996). The curve labeled high heat flow
is based on data from Carboniferous strata in the South Wales Coalfield (White, 1992), and the Clare Coalfield in southern Ireland
(Wagner, 2003).

regional differences in clay mineralogy found in British found in convergent basin settings (Fig. 10). Mature and
Lower Palaeozoic slate belts formed during Caledonian supermature stages of clay mineral evolution (Fig. 7) are
terrane amalgamation, completed approximately 400 Ma preserved in both settings. The extensional back-arc basins,
ago. The different clay assemblages and clay microfabrics characterized by high heat flow (> 35ºC/km) and
found in extensional basins were contrasted with those volcanism with localized hydrothermal activity, contain a

Fig. 10. Schematic diagram illustrating differences in clay mineralogy typical of convergent and extensional basin settings in British
Caledonian terranes (Merriman, 2002b).
Clay minerals and sedimentary basin history 17

greater diversity of clay minerals, including K-, K/Na- and Table 1. K-white mica b dimensions and pressure facies series
Na-micas and pyrophyllite, as well as chlorite/mica stacks. (Guidotti & Sassi, 1986).
K-white micas in these assemblages are typically alumi- Facies series bÅ Inferred geothermal gradient
nous with b cell dimensions < 9.01 Å. Convergent basins, low pressure <9.000 >35°C km-1
characterized by low heat-flow (< 25ºC/km) tend to have intermediate 9.000-9.040 25-35°C km-1
simple clay mineral assemblages of K-white mica and high pressure >9.040 <25°C km-1
chlorite, and the K-micas are phengitic with b cell dimen-
sions > 9.02 Å (Stone & Merriman, 2004).
The Na-micas and diversity of clay mineral assem-
blages found in the extensional basins may have been estimates, particularly for the Welsh Basin (e.g. Bevins &
generated by interaction between hydrothermal systems Merriman, 1988; Roberts et al., 1996).
and seawater, producing fluids with high Na/K ratios
(Merriman, 2002b). Paragonitic alteration appears to be a The b dimension data presented by Stone & Merriman
feature of modern mid-oceanic hydrothermal systems (2004) suggests that the phengite content of K-white micas
(Honnorez et al., 1998). The Trans-Atlantic Geotraverse generated in mature and supermature sedimentary basins is
(TAG) found that extensively chloritized basalts subse- a useful indicator of geothermal conditions rather than
quently reacted with alkali-enriched fluids with high Na/K geobarometric conditions.
ratios generated by mixing hydrothermal fluids with
seawater. These reactions, at temperatures of 250-360°C,
produced the quartz + paragonite + pyrite assemblages that 8. Summary
characterize the main TAG stockwork (Alt & Teagle, 1998).
Paragonite and mixed Na-K-white micas have also been Clay minerals in sedimentary basins are part of a cycle
recorded from extensional settings in Crete (Christidis et of tectonothermal events that degrade and regenerate rock
al., 2003), and in the Eastern Taurides, Turkey, from very in the Earth’s crust (Fig. 1). Neoformed clay minerals,
low-grade Palaeozoic mudrocks (Bozkaya et al., 2002). generated by surface weathering and shallow diagenetic
Bedding-parallel microfabric in illite, and bedding- processes, are transformed into mature clay mineral assem-
parallel stacking planes in chlorite-mica stacks are an addi- blages as sedimentary basins are filled, deformed and
tional characteristic of mudrocks found in the Caledonian metamorphosed. Because of their crystal-chemical proper-
extensional basins, but not found in the convergent basins ties, neoformed clay minerals are hosts for a suite of alkali
(Merriman, 2002b). These microfabrics appear to develop elements and boron that are released during metamor-
in response to high heat flow causing clay reactions during phism. At active plate margins the release of these mobile
‘static’ (non-tectonic) burial metamorphism in the exten- elements promotes magma generation, and when the prod-
sional basins. The absence of bedding-parallel clay micro- ucts of these magmas are weathered they are recycled
fabrics in Caledonian convergent basins, particularly the largely as clay minerals, representing the magma-to-mud
accretionary complex of the Scottish Southern Uplands, pathway of the clay cycle.
may be due to the timing of tectonic events in relation to The transformation of neoformed and inherited clays
trench deposition in such settings. Deposition, deformation within sedimentary basins can be characterized by a series
and burial metamorphism by thrust stacking are almost of reactions in 2:1 dioctahedral and trioctahedral minerals
synchronous events in the accretionary Southern Uplands (Fig. 3). These reactions generally obey the Ostwald Step
terrane (Stone & Merriman, 2004), and probably prevented Rule, i.e. they represent an approach to chemical equilib-
or disrupted the growth of ‘static’ microfabrics. In their rium through a series of crystal-chemical changes that
place, tectonic microfabrics have developed and are sequentially reduce the free energy of the system (Fig. 4).
commonly parallel or subparallel with bedding because of Bedding-parallel microfabrics in clays generated by these
the effects of bedding plane slip during imbrication of the reactions provide submicroscopic evidence of the early
accreted tectonostratigraphy (Merriman et al., 1995). stages of burial metamorphism in mudrocks (Fig. 5, 6).
In addition to clay mineral assemblages and clay micro- Several techniques can be used to indicate reaction
fabrics, variation in the b unit cell dimension of illite-phen- progress in clays as they mature in sedimentary basins. By
gite-muscovite (K-white mica) have provided reliable correlating clay reactions with organic thermal maturity
indicators of geotectonic setting in British Caledonian indicators three stages of basin maturity can be recognized
terranes. Using data from a range of British Lower (Fig. 7). Immature basins are characterized by neoformed
Palaeozoic mudrocks, Stone & Merriman (2004) showed and inherited clays of the shallow diagenetic zone that have
that the largest b dimensions are found in convergent basin not been buried deeply enough to transform to more
setting, representing the intermediate to high pressure evolved clay mineral assemblages. Most commercial clay
facies series with low geothermal gradients (Table 1). deposits are found in immature basins. Mature basins are
Independent assessments of the heat flow characteristics in within the oil and gas windows, and are characterized by
these basins suggest geothermal gradients of 15-25°C/km. the partial transformation of neoformed and inherited clays
Smaller b dimensions were found in extensional basins, in response to deep diagenesis. The transformation of
corresponding to the low-pressure facies series with the smectite to illite is largely completed at this stage of basin
highest geothermal gradients. Once again the gradients evolution. Supermature basins have reached the early
inferred from Table 1 are in agreement with independent stages of regional metamorphism where metastable diage-
18 R. J. Merriman

netic and inherited clays have been largely transformed to Bevins, R.E. & Merriman, R.J. (1988): Co-existing prehnite-acti-
equilibrium assemblages of white mica and chlorite. nolite and prehnite-pumpellyite facies assemblages in the Tal y
The rate at which clay minerals react to form equilib- Fan Metabasite Intrusion, North Wales: Implications for
rium assemblages ranges from 1 x 104 years in some Caledonian metamorphism. J. metamorphic Geol., 6, 17-39.
hydrothermal systems, to more than 2 x 109 years in some Boles, J.R. & Franks, S.G. (1979): Clay diagenesis in Wilcox sand-
mid-continental rift basins. The most rapid rates are found stones of southwest Texas: implications of smectite diagenesis
along active plate margins, where both heat flow and on sandstone cementation. J. Sediment. Petrol., 49, 55-70.
tectonism contribute to clay maturation (Fig. 8). In Bozkaya, O., Yalcin, H., Goncuoglu, M.C. (2002): Mineralogic and
contrast, immature or mature clay minerals are able to organic responses to stratigraphic irregularities: an example
from the Lower Palaeozoic very low-grade metamorphic units
survive for 2000 Ma or more in mid-continental rift basins
of the Eastern Taurus Autochthon, Turkey. Schweiz. Mineral.
through a combination of low or medium geothermal gradi-
Petrogr. Mitt., 82, 291-302.
ents and non-pervasive tectonic events. Buatier, M.D., Peacor, D.R., O’Neil, J.R. (1992): Smectite-illite
Clay mineral assemblages and the b cell dimension of transition in Barbados accretionary wedge sediments: TEM and
K-white mica can be used to infer the geotectonic settings AEM evidence for dissolution/crystallization at low tempera-
of sedimentary basins (Fig. 10). Mudrocks that evolved in ture. Clays Clay Miner., 40, 65-80.
high heat-flow extensional basins may contain a diverse Buatier, M.D., Travé, A., Labaume, P., Potdevin, J.L. (1997):
assemblage of clay minerals, including K-, K/Na- and Na- Dickite related to fluid-sediment interaction and deformation in
micas and pyrophyllite, as well as chlorite/mica stacks. K- Pyrenean thrust-fault zones. Eur. J. Mineral., 9, 875-888.
white micas in these assemblages are typically aluminous Carr, A.D. (1998): A vitrinite reflectance kinetic model incorpo-
with b cell dimensions < 9.01 Å. In contrast, mudrocks that rating overpressure retardation. Mar. Pet. Geol., 16, 355-377.
evolved in some convergent basins appear to have simple Christidis, G.E., Livi, K.J.T., Manutsoglu, E., Árkai, P. (2003): K-,
clay mineral assemblages of K-white mica and chlorite, and Na- and mixed Na-K-white micas in the Ravdoucha (Tyros)
the K-micas are phengitic with b cell dimensions > 9.02 Å. beds and Quartzite-Phyllite Formation, Crete: an indication for
disequilibrium conditions of very low-temperature metamor-
Acknowledgments: Earlier drafts of the article have bene- phism. Euroclay 2003, Abstracts, p. 66.
fited from reviews by Simon Kemp, Alain Meunier and Jan Colten-Bradley, V.A. (1987): Role of pressure in smectite dehydra-
Środoń. Publication is by permission of the Executive tion-effects on geopressure and smectite-to-illite transforma-
Director, British Geological Survey (N.E.R.C.). tion. Am. Assoc. Petrol. Geol. Bull., 71, 1414-1427.
Dalla Torre, M., De Capitani, C., Frey, M., Underwood, M.B.,
Mullis, J., Cox, R. (1996): Very low-temperature metamor-
References phism of shales from the Diablo Range, Franciscan Complex,
California: New constraints on the exhumation path. Geol. Soc.
Abad, I., Nieto, F., Peacor, D.R., Velilla, N. (2003): Prograde and Am. Bull., 108, 578-601.
retrograde evolution in metapelitc rocks of Sierra Espuna Dimberline, A.J. (1986): Electron microscope and microprobe anal-
(Spain). Clay Miner., 38, 1-23. ysis of chlorite-mica stacks in the Wenlock turbidites, mid
Alderton, D.H.M. & Bevins, R.E. (1996): P-T conditions in the Wales Basin. Geol. Mag., 123, 299-306.
South Wales Coalfield: evidence from coexisting hydrocarbon Dong, H., Peacor, D.R., Freed, R.L. (1997): Phase relations among
and aqueous fluid inclusions. J. Geol. Soc. Lond., 153, 265-275. smectite, R1 illite-smectite, and illite. Am. Mineral, 82,
Alt, J. (1999): Very low-grade hydrothermal metamorphism of 379-391.
basic igneous rocks. in “Low-grade metamorphism” M. Frey & Dorsey, R.J., Buchovecky, E.J., Lundberg, N. (1988): Clay miner-
D. Robinson, eds. Blackwell, Oxford, 169-201. alogy of Pliocene-Pleistocene mudstones, eastern Taiwan:
Alt, J.C. & Teagle, D.A.H. (1998): Probing the TAG hydrothermal Combined effects of burial diagenesis and provenance
mound and stockwork: oxygen-isotope profiles from deep unroofing. Geology, 16, 944-947.
ocean drilling. in “Proceedings of the Ocean Drilling Program, Drief, A. & Nieto, F. (2000): Chemical composition of smectites
Scientific Results 158”, P.M. Herzig, S.A. Humphris, D.J. formed in clastic sediments. Implications for the smectite-illite
Miller, R.A. Zierenberg, eds, 285-295. transformation. Clay Miner., 35, 665-678.
Altaner, S.P. & Ylagan, R.F. (1997): Comparison of structural Eberl, D.D. (1984): Clay mineral formation and transformation in
models of mixed-layer illite-smectite and reactions mechanisms rocks and soils. Phil. Trans. Roy. Soc. Lond., A, 311, 241-257.
of smectite illitization. Clays Clay Miner., 45, 517-533. Ehrenberg, S.N., Aagaard, P., Wilson, M.J., Fraser, A.R., Duthie,
Árkai, P. (1991): Chlorite crystallinity: an empirical approach and D.M.L. (1993): Depth-dependent transformation of kaolinite to
correlation with illite crystallinity, coal rank and mineral facies dickite in sandstones of the Norwegian continental shelf. Clay
as exemplified by Palaeozoic and Mesozoic rocks of northeast Miner., 28, 325-352.
Hungary. J. metamorphic Geol., 9, 723-734. Einsele, G. (2000): Sedimentary Basins: Evolution, Facies, and
Árkai, P. & Ghabrial, D.S. (1997): Chlorite crystallinity as an indi- Sediment Budget. Springer, Berlin, 792 p.
cator of metamorphic grade of low-temperature meta-igneous Eslinger, E. & Pevear, D. (1988): Clay Mineralogy for Petroleum
rocks: a case study from the Bükk Mountains, northeast Geologists and Engineers. Society of Economic Paleontologists
Hungary. Clay Miner., 32, 205-222. and Mineralogists Short Course Notes No. 22.
Árkai, P., Mählmann, R.F., Suchý, V., Balogh, K., Sýkorová, I., Frey, Essene, E. J. & Peacor, D. R. (1995): Clay mineral thermometry - A
M. (2002): Possible effects of tectonic shear strain on phyllosil- critical perspective. Clays Clay Miner., 43, 540-553.
icates: a case study from the Kandersteg area, Helvetic domain, Frey, M. (1987): The reaction-isograd kaolinite + quartz = pyro-
Central Alps, Switzerland. Schweiz. Mineral. Petrogr. Mitt., 82, phyllite + H2O, Helvetic Alps, Switzerland. Schweiz. Mineral.
273-290. Petrogr. Mitt., 67, 1-11.
Clay minerals and sedimentary basin history 19

Guggenheim, S., Bain, D.C., Bergaya, F., Brigatti, M.F., Drits, A., Merriman, R.J. & Kemp, S.J. (1996): Clay minerals and sedimen-
Eberl, D.D., Formoso, M.L.L., Galan, E., Merriman, R.J., tary basin maturity. Mineral. Soc. Bull., 111, 7-8.
Peacor, D.R., Stanjek, H., Watanabe, T. (2002): Report of the Merriman, R.J. & Peacor, D.R. (1999): Very low-grade metapelites:
AIPEA nomenclature committee for 2001: Order, disorder and mineralogy, microfabrics and measuring reaction progress. in
crystallinity in phyllosilicates and the use of the “Crystallinity “Low-Grade Metamorphism”, M. Frey & D. Robinson, eds.
Index”. Clay Miner., 37, 389-393. Blackwell Sciences, Oxford, 10-60.
Guidotti, C.V. & Sassi, F.P. (1986): Classification and correlation of Merriman, R.J. & Roberts, B. (1985): A survey of white mica crys-
metamorphic facies series by means of muscovite bo data from tallinity and polytypes in pelitic rocks of Snowdonia and Llyn,
low-grade metapelites. Neues Jahrb. Mineral.-Abh., 153, N. Wales. Mineral. Mag., 49, 305-319.
363-380. —,— (2001): Low-grade metamorphism in the Scottish Southern
Hillier, S. (in press): Formation and alteration of clay materials. in Uplands terrane: deciphering the patterns of accretionary
“Clay materials used in construction”, G. Reeves & J. Cripps, burial, shearing and cryptic aureoles. Trans. R. Soc. Edinburgh:
eds. Special Publication of the Geological Society of London, Earth Sciences, 91, 521-537.
Engineering Group. Merriman, R.J., Roberts, B., Peacor, D.R., Hirons, S.R. (1995):
Hillier, S., Mátyás, J., Matter, A., Vasseur, G. (1995): Illite/smectite Strain-related differences in the crystal growth of white mica
diagenesis and its variable correlation with vitrinite reflectance and chlorite: a TEM and XRD study of the development of
in the Pannonian Basin. Clays Clay Miner., 43, 174-183. metapelite microfabrics in the Southern Uplands thrust terrane,
Hower, J., Eslinger, E.V., Hower, M.E., Perry, E.A. (1976): Scotland. J. metamorphic Geol., 13, 559-576.
Mechanism of burial metamorphism of argillaceous sediments: Mezler, S. & Wunder, B. (2000): Island-arc basalt alkali ratios:
Mineralogical and chemical evidence. Geol. Soc. Am. Bull., 87, Constraints from phengite-fluid partitioning experiments.
725-737. Geology, 28, 583-586.
Honnorez, J.J., Alt, J.C., Humphris, S.E. (1998): Vivisection and Millot, G. (1970): Geology of Clays. Springer-Verlag, New York;
autopsy of active and fossil hydrothermal alterations of basalt Chapman & Hall, London, 429 p.
beneath and within the TAG hydrothermal mound. in Milodowski, A.E. & Zalasiewicz, J.A. (1991): The origin, deposi-
“Proceedings of the Ocean Drilling Program, Scientific Results tional and prograde evolution of chlorite-mica stacks in
158”, P.M. Herzig, S.A. Humphris, D.J. Miller, R.A. Llandovery sediments of the central Wales Basin. Geol. Mag.,
Zierenberg, eds, 231-254. 128, 263-278.
Jeans, C.V., Wray, D.S., Merriman, R.J., Fisher, M.J. (2000): Moore, D.M. & Reynolds, R.C. Jr. (1997): X-ray Diffraction and
Volcanogenic clays in Jurassic and Cretaceous strata of England Identification of Clay Minerals. Oxford University Press,
and the North Sea Basin. Clay Miner., 35, 25-55. 332 p.
Jiang, W.T. & Peacor, D.R. (1993): Formation and modification of Morse, J.S. & Casey, W.H. (1988): Ostwald processes and mineral
metastable intermediate sodium potassium mica, paragonite paragenesis in sediments. Am. J. Sci., 288, 537-560.
and muscovite in hydrothermally altered metabasites from Neuhoff, P.S., Watt, W.S., Bird, D.K., Pedersen, A.K. (1997):
northern Wales. Am. Mineral., 78, 782-793. Timing and structural relations of regional zeolite zones in
Jones, B.F. & Galan, E. (1988): Sepiolite and Palygorskite. in basalts of the East Greenland continental margin. Geology, 25,
“Hydrous Phyllosilicates”, S.W. Bailey, ed. Miner. Soc. Am., 803-806.
Reviews in Mineralogy, 19, 631-674. Nieto, F., Velilla, N., Peacor, D. R., Huertas, M. O. (1994): Regional
Kisch, H.J. (1987): Correlation between indicators of very-low- retrograde alteration of sub-greenschist facies chlorite to smec-
grade metamorphism. in “Low Temperature Metamorphism”, tite. Contrib. Mineral. Petrol., 115, 243-252.
M. Frey, ed. Blackie & Son, Glasgow, 227-300. Patino, L.C., Carr, M.J., Feigenson, M.D. (2000): Local and
Li, G., Peacor, D.R., Merriman, R.J., Roberts, B. (1994): The diage- regional variations in Central American arc lavas controlled by
netic to low grade metamorphic evolution of matrix white variations in subducted sediment input. Contrib. Mineral.
micas in the system muscovite-paragonite in a mudrock from Petrol., 138, 265-283.
Central Wales, U.K. Clays Clay Miner., 42, 369-381. Patrier, P., Beaufort, D., Laverret, E., Bruneton, P. (2003): High-
Li, G., Mauk, J.L., Peacor, D.R. (1995): Preservation of clay grade diagenetic 2M1 illite from the Middle Proterozoic
minerals in the Precambrian (1.1 Ga) Nonesuch Formation in Kombolgie Formation (Northern Territory, Australia). Clays
the vicinity of the White Pine Copper Mine, Michigan. Clays Clay Miner., 51, 102-116.
Clay Miner., 43, 361-376. Peacor, D.R. (1992): Diagenesis and low-grade metamorphism of
Livi, K.J.T., Veblen, D.R., Ferry, J.M., Frey, M. (1997): Evolution of shales and slates. in “Minerals and Reactions at the Atomic
2:1 layered silicates in low-grade metamorphosed Liassic shales Scale: Transmission Electron Microscopy”, P.R. Buseck, ed.
of Central Switzerland. J. metamorphic Geol., 15, 323-344. Miner. Soc. Am., Reviews in Mineralogy, 27, 335-380.
McIlroy, D., Worden, R.H., Needham, S.J. (2003): Faeces, clay — (1998): Implications of TEM data for the concept of funda-
minerals and reservoir potential. J. Geol. Soc. Lond., 160, mental particles. Can. Mineral., 36, 1397-1408.
489-493. Pollastro, R.M. (1993): Considerations of the illite-smectite
Merriman, R.J. (2002a): The magma-to-mud cycle. Geology Today, geothermometer in hydrocarbon-bearing rocks of Miocene to
18, 67-71. Mississippian age. Clays Clay Miner., 41, 119-133.
— (2002b): Contrasting clay mineral assemblages in British Lower Potter, P.E., Maynard, J.B., Pryor, W.A. (1980): Sedimentology of
Palaeozoic slate belts: the influence of geotectonic setting. Clay Shale. Springer-Verlag, Berlin, 270 p.
Miner., 37, 207-219. Putnis, A. (1992): Introduction to Mineral Science. Cambridge
Merriman, R.J. & Frey, M. (1999): Patterns of very low-grade meta- University Press, 357 p.
morphism in metapelitic rocks. in “Low-Grade Robert, P. (1985): Histoire géochimique et diagenèse organique.
Metamorphism”, M. Frey & D. Robinson, eds. Blackwell Elf-Aquitaine, France. English translation 1988. Dordrecht,
Sciences, Oxford, 61-107. Holland: D. Reidel. 311 p.
20 R. J. Merriman

Roberts, B., Merriman, R.J., Hirons, S.R., Fletcher, C.J.N., Wilson, interaction”. M.B. Underwood, ed. Geol. Soc. Am. Special
D. (1996): Synchronous very low grade metamorphism, Paper, 273, 45-61.
contraction and inversion in the central part of the Welsh Lower van de Pluijm, B.A., Ho, N.C., Peacor, D.R., Merriman, R.J. (1998):
Palaeozoic Basin. J. Geol. Soc. Lond., 153, 277-286. Contradictions of slate formation resolved. Nature, 392, 348.
Ruiz Cruz, M.D. & Andreo, B. (1996): Genesis and transformation Velde, B. & Lanson, B. (1993): Comparison of I/S transformation
of dickite in Permo-Triassic sediments (Betic Cordilleras, and maturity of organic matter at elevated temperatures. Clays
Spain). Clay Miner., 31, 133-152. Clay Miner., 41, 178-183.
Sample, J.C. & Moore, J.C. (1987): Structural style and kinematics Velde, B. & Vasseur, G. (1992): Estimation of the diagenetic smec-
of an underplated slate belt, Kodiak and adjacent islands, tite to illite transformation in time-temperature space. Am.
Alaska. Geol. Soc. Am. Bull., 99, 7-20 Mineral., 77, 967-976.
Sherlock, S.C., Kelley, S.P., Zalasiewicz, J., Schofield, D., Evans, J., Wagner, K. (2003): The effect of deformation on illite crystallite
Merriman, R.J. (2003): Precise dating of low-temperature sizes. in “A Clay Odessey: Proceedings of the 12th International
deformation: strain-fringe analysis by 40Ar-39Ar laser micro- Clay Conference”. E.A. Dominguez, G.R. Mas, F. Cravero, eds.
probe. Geology, 31, 219-222. 179-186.
Sparks, S.J. (1999): Production and distribution of volcanic ash. Warr, L.N. & Nieto, F. (1998): Crystallite thickness and defect
Programme and Abstracts; From Magmas to Mud (and density of phyllosilicates in low-temperature metamorphic
Back). Mineralogical Society Millennium Conference, pelites: A TEM and XRD study of clay-mineral crystallinity-
Reading. p. 9. index standards. Can. Mineral., 36, 1453-1474.
Środoń, J. (1999): Nature of mixed-layer clays and mechanisms of White, S. (1992): The tectonothermal evolution of the South Wales
their formation and alteration. Annu. Rev. Earth Planet. Sci., 27, Coalfield. Ph. D. thesis, University of Wales, Cardiff.
19-53. Worden, R.W., Charpentier, D., Fisher, Q.J., Aplin, A.C. (2003):
— (2002): Quantitative mineralogy of sedimentary rocks with Mineralogical and microstructural variations during mudstone
emphasis on clays and with applications to K-Ar dating. diagenesis, North Sea Upper Cretaceous and Gulf of Mexico
Mineral. Mag., 66, 677-687. Miocene mudstones: chemical compaction during deep burial
Środoń, J. & Clauer, N. (2001): Diagenetic history of Lower mudstone diagenesis. Programme and Abstracts, Wrestling with
Paleozoic sediments in Pomerania (northern Poland) traced Mud. William Smith Meeting 2003, Geological Society of
across the Teisseyre-Tornquist tectonic zone using mixed-layer London, 39.
illite/smectite. Clay Miner., 36, 15-27. Yau, Y.C., Peacor, D.R., Bearne, R.E., Essene, E.J. (1988):
Strixrude, L. & Peacor, D.R. (2002): First-principles study of illite- Microstructures, formation mechanisms, and depth-zoning of
smectite and implications for clay mineral systems. Nature, phyllosilicates in geothermally altered shales, Salton Sea,
420, 165-168. California. Clays Clay Miner., 36, 1-10.
Stone, P. & Merriman, R.J. (2004): Basin thermal history favours an Zhao, G., Peacor, D.R., McDowell, S.D. (1999): ‘Retrograde diage-
accretionary origin for the Southern Uplands terrane, Scottish nesis’ of clay minerals of the Precambrian Freda sandstone,
Caledonides. J. Geol. Soc. Lond., 161, 829-836. Wisconsin. Clays Clay Miner., 47, 119-130.
Underwood, M.B., Laughland, M.M., Kang, S.M. (1993): A
comparison among organic and inorganic indicators of diagen-
esis and low-temperature metamorphism, Tertiary Shimanto Received 2 April 2004
Belt, Shikoku, Japan. in: “Thermal evolution of the Tertiary Modified version received 21 June 2004
Shimanto Belt, Southwest Japan: An example of ridge-trench Accepted 13 September 2004

Вам также может понравиться