You are on page 1of 16

Geological Society, London, Engineering Geology Special Publications 2.

The composition of clay materials


Geological Society, London, Engineering Geology Special Publications 2006; v. 21; p. 13-27 doi:10.1144/GSL.ENG.2006.021.01.02

Email alerting service Permission request Subscribe

click here to receive free email alerts when new articles cite this article click here to seek permission to re-use all or part of this article click here to subscribe to Geological Society, London, Engineering Geology Special Publications or the Lyell Collection

Notes

Downloaded by

Schlumberger on 18 November 2007

2006 Geological Society of London

2.

I he composition of clay materials

Clay materials are composed of solid, liquid and vapour phases. The solid phases are of mineral and organic phases that make up the framework of the clay materials. The mineralogy can be broadly subdivided into the clay and non-clay minerals, including poorly crystalline, so-called 'amorphous' inorganic phases. By definition, minerals are crystalline solids with well-ordered crystal structures but clay minerals and other inorganic phases in clay materials are often poorly crystalline compared to minerals such as quartz and feldspar. Some clay materials may be dominated by one mineral phase, e.g. smectite in bentonites, opal in diatomaceous earths. However, most clay materials are composed of heterogeneous mineral mixtures. Based on the bulk mineral analysis of over 400 samples, Shaw & Weaver (1965) reported the modal mineralogical composition of siliciclastic mudrocks to be: 60% clay minerals 30% quartz and chert 5% feldspar 4% carbonates 1% organic matter 1% iron oxides There is a general increase in the predominance of clay minerals in sedimentary rocks with decreasing grain size 100 -

(Fig. 2.1) (Blatt et al. 1972). However, it needs to be stressed that, whilst clay minerals are usually significant, if not predominant, phases in clay materials, other mineral phases are usually present in varying amounts and can significantly affect the properties and behaviour of the materials. In soils, mineral and organic compositional variations reflect the weathered parent rocks and the physical, chemical and biological factors controlling the soil forming processes (see Chapter 3). The liquid and vapour phases, of which water is usually the most important, occur either as 'bound phases', adsorbed onto the surfaces of the solid particles, or as 'free phases' within pore spaces. In this chapter the nature of the mineralogical, organic and aqueous phases present in clay materials are described plus the nature of swelling and ionic exchange properties in solid phases.

2.1.

Clay minerals

The clay minerals are a group of hydrous aluminosilicates that are characteristically found in the clay ( < 2 gm) fractions of sediments and soils. The majority of clay minerals have sheet silicate structures (see Text Box on p. 14).

80-

~gz .'co
t.kl r ~ r ~---

Sand

Silt

Clay

FIG. 2.1. Detrital mineralogy of sedimentary rocks as a function of ~rain size (after Blatt et al. 1972).

14

THE COMPOSITION OF CLAY MATERIALS

Th~,:she~i Nr i i!i!::if: ii.!ii!,!ii:i :i:iii~if: i :i~iiI i i::~iiiii iii iii if i!~ i iiiiiiili,~il i ~iil,! ~:: ~,:~:~, ~,: ~: :~: ~,i: : iir:;!:ii~ili!ii!!i!i ii!:!?i !if: ~?i ii:: :i?,:i:;i!:%i! :i::~ ~ i~i~::: ! iliii !i::!ii:!!i iiii:!i}:i~:~i::ii: i :i! )i:!i:,:ii!~, :ii::ii/:i: i iii~:!:i:::i::~: !i;: ::i:iii~::i::::i~: i:!, :ii:,:~i: !i::!ilii:ii:::ii,,iiii!!::i

FIG. 2.2. Composite layers in clay mineral structures.

THE COMPOSITIONOF CLAYMATERIALS


2 layer (I:I]type Tetrahedral layer Octrehedral layer 3 layer (2:1)type

Interlayer sites X

KAOLINITE

ILLITES SMECTITES VERMICULITES CHLORITES

FIG. 2.3. General structures of clay minerals.


TABI~E 2.1. Classification of clay minerals

Sheet silicate type 1:1 ~p___~ Kaolin and serpentine 2:1 type Illites Chlorites Smectites (Montmorillonites) Vermiculites Mixed layer clays with smectite/ vermiculite Mixed layer clays without smecfite/vermiculite Palygorskite and sepiolite

Property Non-swelling Non-swelling Non-swelling Swelling Swelling Swelling Non-swelling Non-swelling

There are two types of composite layer structures in the clay minerals: 9 the two layer or 1 : 1 type represented by the kaolin and serpentine groups; 9 the three layer or 2:1 type represented by the illitemica, smectite, vermiculite and chlorite groups (Fig. 2.3 and Table 2.1). 2.1.1. The kaolin and serpentine groups

given to kaolin-rich deposits that occur in different formation environments (see Text Box on p. 16) Kaolinite can range from well crystallized varieties to poorly crystalline forms that can be differentiated by X-ray powder diffraction analysis (see Section 8.2.2). Halloysite is a kaolin group mineral that contains one water layer in the interlayer sites, producing a 10)~ interlayer spacing rather than the 7A spacing of the kaolinite structure. Its general formula is A14Si4010(OH)8"H20. Removal of the water layer causes the halloysite structure to collapse to a 7A interlayer spacing. The collapsed form is referred to as 7A halloysite (or metahalloysite) as opposed to the hydrated 10A halloysite. Halloysites usually have a characteristic tubular morphology though other spheroidal habits have been reported (Fig. 2.5). The tubular morphology arises from the distortion and curving of the 1 : 1 layer structure from which the tubular habit develops. The serpentine group of minerals is the trioctahedral equivalent of the kaolin group. Berthierine (Fe 2+, Mg)6_x (Fe 3+, A1)xSi4_xAlx)O10(OH)8)is the most important member of the serpentine group in studies of sedimentary rocks. It commonly occurs in oolitic ironstones and in pedogenic environments but has often been misnamed chamosite. Chamosite is not a 7A serpentine mineral but a 14A chlorite mineral (see Section 2.1.5). 2.1.2. The illlite-mica group

The kaolin group of minerals is the most common of the clay minerals with the two layer 1 : 1 type of structure. The term kandite has also been used to describe kaolin group minerals but is not the preferred name. Kaolins have dioctahedral structures with, ideally, no net negative charge on the composite layers and consequently no compensating interlayer cations (Fig. 2.4) or water layers. There are three principal polymorphs (or polytypes), kaolinite, dickite and nacrite of general formula A14Si4Oa0(OH)8. Specific names have also been

Illite is the term generally used for all clay grade micas and most commonly has a dioctahedral 2:1 or three-layer composite layer structure (Fig. 2.6). In the illite structure substitution of [Si4+]iv by [A13+]iv and [R3+]w by [W+] vI produces a net negative charge of about 0.7-1.0 per O10 (OH)2 formula unit. Potassium is the principal interlayer cation and lenses of water may also be present in the interlayer sites (FIR. 2.6).

16

THE COMPOSITIONOF CLAY MATERIALS

FIG. 2.4. Crystal structure of kaolinite (A14Sc40l0 (OH)8).

Illite Ko.s(R3+,.65R2+0.35) (Si3.55A1045)O,0(OH)2 Muscovite KA12(Si3A1)Olo(OH)2 Sericite is also used to generally signify fine-grained micacaeous material of an indeterminate nature that may include illite but also other clay and sheet silicate minerals. Glauconite (sensu stricto) is a dioctahedral Fe-rich illite of general formula K(R 2+067R3+,.33) (Si3.67Alo.33)Olo(OH)2 with Fe 3+ >> A1, Mg > Fe 2+ and Fe 3+ > F e z + However, 'glauconite' is also used to describe any green (pelletal) clay material irrespective of composition (see Text Box on p. 18). The term 'glaucony' has been proposed as a general term when the specific mineral composition is not known (Millot 1970).

Celadonite is the more Mg-rich equivalent of glauconite of ideal formula K(Mg, Fe 2+, Fe 3+)Si40~o(OH)2. Unlike glauconites that are most often found in modern and ancient sediments, celadonites are most commonly formed in association with the alteration of volcanic, usually basaltic, rocks.
2.1.3. Smectites

Smectites are three-layer or 2:1 clay minerals (Fig. 2.7) that are most commonly dioctahedral but trioctahedral varieties do exist. They have a layer charge of 0.2-0.6 per O10(OH)2 unit of structure, which is offset by hydrated interlayer cations, principally Ca and Na. The hydration of the interlayer cations causes the interlayer crystalline swelling that characterizes the smectites. Water is sorbed into the interlayer sites

THE COMPOSITIONOF CLAYMATERIALS

1/

composed predominantly of montmorillonite and/or to indicate genesis of montmorillonites from alteration of volcanic ash. In the UK Fullers Earth is used in a similar way to bentonite though in North America Fullers Earth can be composed of palygorskite rather than smectite (montmorillonite). Dioctahedral smectites: Montmorillonites R+y(A12_y R2+y)Si4Olo(OH)z'nH20 Beidellite R+yAlz(Si4_yAly)O10(OH)2"nH20 Nontronite R+yFe3+2(Si4_yAly)Olo(OH)z'nH20 Trioctahedral smectites: Saponite R+x_y(Mg3_y(A1,Fe 3+)y)(Si4_x Alx) Olo(OH)2"nH20 Hectorite R+x_y(Mg3_yLiy)Si4Olo'(OH)z'nH20
2.1.4. Vermiculite

FIG. 2.5. Photomicrograph of tubular habit of halloysite (from Kohyama et al., 1978, reproduced by permissionof the authors and publishers). in monomolecular sheets (Fig. 2.8). Na and Camontmorillonites can absorb up to three layers at 90% relative humidity but in some instances at 100% humidity, Na-montmorillonites can absorb more and the 2:1 layers disperse. Bentonite is often used synonymously with montmorillonite but is a rock term rather than a mineral name. As a rock term, it used to signify a commercial clay deposit

Vermiculites are similar in structure to the smectites (Fig. 2.9) but have a larger net negative charge on the composite layer of 0.6-0.8 per O10(OH)2. The principal interlayer cations are hydrated magnesium. Vermiculites exhibit swelling properties similar to the smectites but to a lesser extent due to the higher layer charge. Vermiculites are trioctahedral with the general formula: Mg (x-y~/2(Mg,Fe 2+)3_y(A1, Fe3 +)y (Si4_x Alx)O10(OH)2"nH20
2.1.5. Chlorite

The chlorites have a 2:1 type structure with a second octahedral layer in the interlayer sites having a net positive charge to offset the net negative charge on the 2:1 layers (Fig. 2.10).

FIG. 2.6. Crystal structure of muscovite.

18

THE COMPOSITION OF CLAY MATERIALS

FI6. 2.7. Crystal structure of dioctahedral smecite.

FIG. 2.8. One layer and two layer water structures in smectite (from Velde 1992).

THE COMPOSITION OF CLAY MATERIALS

19

FI6. 2.9. Crystal structure of vermiculite.

FIG. 2.10. Crystal structure of chlorite.

The general formula of the chlorites is


(Mg, Fe 2+)6_x(A1, Fe 3+)xSi4_x Al~Olo(OH)8

2.1.6.

Mixed layer clay minerals

The majority of chlorites are trioctahedral but some dioctahedral and mixed dioctahedral (2:1 layers) trioctahedral (interlayers) forms are known and can be identified by X-ray diffraction analysis (see Chapter 8).

Mixed layer clay minerals describe those mineral phases in which different sheet silicate units occur in the stacking sequence. Generally the 2:1 clay minerals are most commonly involved in forming mixed layer clay minerals, though kaolinite-smectites and serpentine-chlorites have been reported. The most commonly reported mixed layer

20
Interstratified Model

THE COMPOSITION OF CLAY MATERIALS

Interparticle Model 10A "Smectite"

Sm

Sm

Exchangecations, H20,organics

)
Sm Sm

(
K+ K+

The mixed layer phases may also show segregations into AB sequences with separate domains of A or B rather than continuously alternating sequences. Generally mixed layer illite-smectites with more than 45% smectite are randomly interstratified, those with 30-45% smectite can be random or ordered and with less than 30% smectite are ordered (Bethke & Altaner 1986).

10A "Smectite"

2.1.7.

Sepiolite and palygorskite

Exchange cations,H2O,organics

2OA "lllite"

Sm

lq Exchangecations, H20,organics

Although sepiolite and palygorskite are often described as 2:1 sheet silicates, the sheets are not continuous but form alternate ribbons of 2:1 sheet silicate structures (Fig. 2.12). Consequently they have fibrous rather than the typically platy habits of the true sheet silicates. Compositionally they are Mg-rich, Mg being the principal cation in the octahedral sites but in some palygorskites aluminium can be the more common octahedral cation,
i.e. [AP + ] v , > [Mg2 +]vi.

I I I
K+ K+ K+ K+ 3OA "lllite"

The name attapulgite is often used for commercial deposits of palygorskite, but palygorkite is the preferred mineral name.

2.1.8.

Swelling properties of clay minerals

J /

/ Tetrahedral layers L Octahedral layers "N Tetrahedral layers

Fro. 2.11. The nature of illite-smectite as defined by the interstratified and interparticle models (after Nadeau & Bain 1986).

clays are the illite-smectites and to a much lesser extent the chlorite-smectites. The interstratification may be random, with no discernible pattern in the stacking sequence, or ordered, e.g. ABABAB, AABAABAA or AAABAAB. Specific names are also given to certain types of ordered mixed layer clay minerals.

The clay minerals can be classified into swelling and nonswelling varieties (Table 2.1). Water and organic molecules may be sorbed onto the surface of and into the interlayer sites of clay minerals. The sorption of water and organic molecules into the interlayer sites is responsible for the characteristic intraparticle swelling properties of smectites and vermiculites (Table 2.1). A 'dry' smectite absorbs water into the interlayer sites in discrete layers with the water forming hydration sheaths around the interlayer cations. At very high humidities Na-smectites when immersed in water exhibit osmotic swelling and may completely dissociate. Complete dissociation of the clay layers occurs when the repulsive forces of the negatively charged clay surfaces exceed the forces of attraction between the hydrated interlayer cations and the clay particles. Vermiculites, having a higher interlayer charge than smectites, show less swelling and will not completely dissociate in water.

THE COMPOSITION OF CLAY MATERIALS

21

PALYGORSKITE
-I--

solution
-I-

+
m

0
m

+-+

+ +

-t-I--

+
-

L
SEPIOLITE

j
m

0
m

++ + + + + +

+ +
+

+
+

-I-

\\ f
I

27~, ~ Tetrahedrallayer Octahedrallayer Tetrahedral layer

Ion concn

n+

Fro. 2.12. Crystal structures of palygorskite and sepiolite.

Illites may have lenses of water in the interlayer sites, out not complete layers and do not show intraparticle swelling. Chlorites are non-swelling clay minerals but there have been reported instances of'swelling chlorites'. Moore & Reynolds (1997) commented that swelling chlorites are such a poorly described phase they were unable to calculate their diffraction characteristics. It seems probable that 'swelling chlorites' are most likely to be mixed layer chlorite/smectites or chlorite/vermiculites rather than a swelling chlorite s e n s u s t r i c t o . Halloysites in the hydrated state have a water layer between the 7N kaolin layers producing their characteristic 10/~ interlayer spacing. Kaolinite in contrast to halloysite does not swell in the presence of water though increasingly disordered kaolinites contain lenses of water. Sepiolite and palygorskite have open channel sites similar to zeolites (see Section 2.2.7) into which water and organic compounds can be absorbed or desorbed without significantly affecting the unit cell dimensions and thus are not regarded as swelling clays. In addition to intraparticle swelling, all clay minerals can show interparticle swelling which is governed by similar factors that control intraparticle swelling, that is the nature of the clay minerals and the nature of and concentration of cations adsorbed onto the clay surface in the diffuse double layer and hydration of

d i s t a n c e from c l a y s u r f a c e
FIa. 2.13. Distribution of ions in the clay-water layer of thickness d and formation water.

surface cations (Guven 1993; Guven & Pallastro 1993; Low 1992). Interparticle associations of clays control their flocculation and dispersion in natural waters.

2.1.9.

Ion exchange properties of clay minerals

The 2 : 1 clay minerals have a net negative charge on their composite layers due to cation substitutions (e.g. A13+ for Si4+; FEZ+; Mg 2+ for AP +, Fe3+). In minerals such as illites and smectites the net negative charge is offset by interlayer cations and in the chlorites by the interlayer octahedral sheet. Kaolins ideally have neutral composite layer structures, but in reality limited cation substitution may occur producing a very small net negative charge on the composite layer offset by a small number of interlayer cations. When suspended in water some of the interlayer cations on the surface of the clay particles go into solution producing a negatively charged clay surface surrounded by a diffuse' double layer' of hydrated cations (Fig. 2.13) (van Olphen 1977).

22

THE COMPOSITION OF CLAY MA'IERIALS

OH
Si
O... H

/O-...HOH
Si

A(-OH + 2 /\O...H
Si

\o /
A I - O-...NON

/\o
Si

\o.
In acid solution positively charged
Anion exchanger

O-...HOH

OH
Si

In alkaline solution negatively charged


Ca tion exchanger

)o
Si

\
OH

/
Si

O_AI'I-2

//

/OAI034 Si

?
AI - O - A I +2

\o
A [ - OAIO~ 4

\ /o \

/
Si

Si

O - A I +2

\OAIO~ 4 In alkaline solution in the presence of AIO4 s" ions negatively charged
Cation exchanger

In acid solution in the presence of AI3+ions positively charged


Anion exchanger

Flc. 2.14. Edge site anion exchange as a function of pH (from Yariv and Cross, 1979).

On the edges of the clay particles the disruption of the clay structure produces broken bond edge sites which may be negatively or positively charged. The nature of the charge is determined by the presence of certain ions, notably H +, OH-, A13+ and A1045-, whose presence is pH dependent, being negatively charged in alkaline solutions and positively charged in acidic solutions (Fig. 2.14). This pH dependent charge accounts for only a small percentage of the total charge in illites and smectites, but is more significant for kaolins and chlorites. In addition to cation exchange it is also possible to have

anion exchange but usually to a lesser extent than cation exchange. The ions principally adsorbed onto the clay surfaces and, to a lesser extent, on the edges of the clay particles, can be exchanged with other available ions. This gives rise to the phenomenon of cation/anion exchange. The degree to which exchange reactions take place depends on the nature of the clay minerals and the concentrations of the cations/anions involved and of other ionic species. Oxides and organic matter commonly coat clay particles and also have ion exchange properties, which are strongly

THE COMPOSITIONOF CLAYMATERIALS TABLE 2.2. Cation exchange capacities of some clay minerals
(meq/lOOg) at p H = 7 (after Drever 1982)

23

Smectite Illite Kaolin Chlorite

80-150 10-40 1-10 < 10

dependent on pH, and can alter the cation/anion exchange capacities of associated clays in natural systems. Ion exchange is a dynamic process governed by the law of mass action. Generally, the greater the concentration of a particular cation in solution, the more readily it will replace equivalent cations adsorbed onto the clay surfaces and interlayers. Cations of higher valence will usually replace those of lower valence and, if of the same valence, smaller sized cations will replace larger sized cations. However, it is important to note that in the diffuse layer, the cations will be hydrated and it is the size of the hydrated cation that needs to be considered. In swelling clay minerals, the interlayer cation sites will also provide exchange sites, whereas for other clays only the outer surface sites will be involved. This explains the higher cation exchange capacities of the smectites compared to the illites (Table 2.2). Because the edge sites are also involved in ion exchange and as the nature of the charge on the edge sites is pH dependent the cation exchange capacity of individual clay minerals are also pH dependent.

It is a dense, relatively hard material composed of finely crystalline quartz crystals or fibrous chalcedonic quartz, usually derived from the diagenetic alteration of volcanogenic or biogenic opaline silica (see Text Box below). Cherts usually occur as either bedded or nodular forms. The bedded forms are commonly associated with volcanic sequences whereas nodular cherts are mainly hosted in chalks or mudrocks. The nodular forms hosted in chalks and other carbonates are specifically referred to as flints. Jasper is a red coloured chert that contains hematite impurities. 2.2.2. Feldspars

As with quartz, feldspars in clay materials are predominantly detrital in origin, but there is petrological evidence to suggest they may also form diagenetic cements around existing feldspar grains. 2.2.3. Carbonates

2.2. Non-clay mineralogy


2.2.1. Quartz and chert Quartz is commonly present in clay materials, predominantly in the silt grade material (Fig. 2.1). It is generally detrital, but some may be biogenic or authigenic in origin. Silicification may occur during early diagenesis with the dissolution of biogenic silica, a possible source for the later inorganic precipitation of quartz (Millot 1970). Diagenefic modification of clay minerals may also cause the mobilization of SiO2 which may reprecipitate as a quartz cement in mudrocks or adjacent sandstone beds. Chert is a general description for fine-grained siliceous sediments of chemical, biochemical or biogenic origin.

According to Shaw & Weaver (1965) there is an average of about 4% by weight of carbonate minerals in siliciclastic mudrocks but this can be highly variable. Calcite (CaCO3) and dolomite (CaMg(CO3)2 are the predominant carbonate minerals in clay materials, though siderite (FeCO3) may also be locally important. Metastable polymorphs of CaCO3, aragonite (orthorhombic) and, less commonly, vaterite (hexagonal) also occur in Recent clay materials. The carbonate may be present as clasts derived from organic skeletal remains and the weathering of pre-existing carbonate rocks or as diagenetic cements and nodules. The source of carbonate for the formation of diagenetic carbonate cements or nodules may be from the recrystallization of primary (skeletal) carbonate material and/or direct precipitation from alkaline bicarbonate-rich pore waters. These processes are considered to be often related to the maturation of organic matter during diagenesis (Curtis 1983). Dolomite and siderite formation are favoured in sulphate-depleted environments (Baker & Kastner 1981; Berner 1981). The presence of authigenic siderite is indicative of anoxic sulphate-depleted conditions at the time of its formation (Table 2.3).

24

THE COMPOSITIONOF CLAYMATERIALS Their nominal formula is SiA1203(OH)4but both have poorly ordered crystal structures and are often described as amorphous. Generally allophane forms spherules and appears to be less well ordered than imogolite with its typically cylindrical habit. The SiO2:A1203 ratio varies from 1-1.2 for imogolite to 1.3-2.0 for allophane (Moore & Reynolds 1997).
2.2.5.3. Ion exchange in oxides and hydroxides. In the oxides and hydroxides the charge on the surfaces arises largely from broken surface bonds producing unshared surface and edge O and OH rather than the cation substitutions that are largely responsible for imparting charge to the clay mineral structures. As described above, charge arising from broken bonds is pH dependent (Fig. 2.14). The pH at which the charge is zero is referred to as the zero point of charge or ZPC; when conditions are more acidic than the ZPC the oxide/hydroxide will be positively charged and negatively charged for conditions more alkaline than the ZPC. For example, the ZPC for boehmite is 8.2 and 6 to 7 for goethite (Eslinger & Pevear 1988).

TABLE 2.3. Classification of oxic-anoxic environments using authigenic Fe-mineral indicators (Berner 1981) Environment Oxic Anoxic (sulphidic) Anoxic (non-sulphidic) (a) post-oxic (b) methanic Authigenic Fe-minerals Hematite, goethite (no organic matter preserved) Pyrite (organic matter preserved) Glauconite, Fe2+/Fe 3+silicates siderite (no sulphides some organic matter preserved) Siderite after pyrite (organic matter preserved)

2.2.4.

Iron sulphides

Pyrite (FeS2) is the most common of the sulphides found in clay materials occurring as veins, nodular aggregates, scattered crystals and infilling/replacing/pseudomorphing skeletal fragments. Sedimentary authigenic pyrite forms in anoxic sulphidic environments (Table 2.3). Weathering of pyrite produces iron hydroxides/oxides and mobilizes sulphate which may be incorporated into sulphate minerals such as gypsum and jarosite. Marcasite is an orthorhombic dimorph of pyrite that has occasionally been reported in clay materials. There is some evidence to suggest that marcasite forms under more acidic conditions than pyrite. Metastable monosulphide (e.g. mackinawite and griegite) are found in modem muds as precursors to later formation of pyrite. Pyrrhotite (Fel_xS) is a stable monosulphide mineral phase that occcasionally has been reported in clay materials but stoichiometric FeS, the mineral troilite, is only found in meteorites.

2.2.6.

Sulphates

2.2.5.

Oxides and hydroxides

2.2.5.1. Iron oxides and hydroxides. Iron oxides and hydroxides are present in various mineral forms in clay materials, notably hematite (c~-Fe203), maghemite (7Fe203), goethite (~-FeO.OH), lepidocrosite (7-FeO'OH). Limonite is a general term for a mixture of iron oxides and hydroxides of a poorly crystalline nature. The presence of these iron oxides and hydroxides produces the characteristic red/brown/yellow colours of many clay materials depending on which mineral is present. Goethite is yellow-brown, lepidocrosite orange, maghemite brown-black and hematite red. 2.2.5.2. Aluminium oxides and hydroxides. Gibbsite (AI(OH)3), boehmite (y-AIO'OH) and diaspore (~A10"OH) are the most common of the aluminium oxides and hydroxides and are typically significant constituents of laterites and bauxites. Boehmite is isomorphous with lepidocrosite and diaspore with goethite. In addition there are aluminium - silicon oxyhydroxides, imogolite and allophane, that characteristically occur in soils derived from the weathering of volcanic "ocks.

Sulphates may form nodules or veins in clay materials, with gypsum (CaSO4"2H20), followed by anhydrite (CaSO4) being the most common, but celestite (SrSO4), barite (BaSO4) and the more water-soluble sodium and magnesium sulphates may also occur. Their occurrence may suggest hypersaline conditions of formation, though gypsum and various iron sulphates may also form during weathering of shales containing pyrite, the oxidation of the sulphide being the source of the sulphate. Jarosite (KFe3(SO4)dOH)6) is a distinctively yellow mineral that characteristically forms as surface coatings and aggregates from the oxidation of pyrite to iron sulphate, followed by further reaction with illite/mica or K-feldspars: 12FeSO4 + 4KAlaSi3Os(OH)2 + 48H20 + = 4KFe3(SO4)2(OU)6 + 8AI(OH)3 + 12Si(OH)4 + 4H2SO4.
0 2

2.2.6.1. Ettringite group. The ettringite group of minerals, mixed calcium carbonate sulphate silicates, and especially thaumasite (Ca3(SO4)(CO3)(Si(OH)6)'12H20), has attracted attention recently because of the observation of the growth of such minerals at concrete-clay boundaries, which can cause significant damage to concrete structures (Thaumasite Expert Group 1999).

2.2.7.

Zeolites

Zeolites are hydrous framework silicates and those that occur in clay materials are predominantly alkali varieties (Table 2.4), They are commonly associated with aluminous smectites formed by the weathering and alteration of

THE COMPOSITION OF CLAY MATERIALS TABLE2.4. Common zeolites occurring in clay materials Natrolite Analcite Phillipsite Erionite Heulandite Clinoptilolite Mordenite Na2A12Si3010"2H20 NaA1Si2Or.H20 (Ca,Na,K) 3A13SisOI6.6H20 (Na2K2CaMg)4_sA19Si27072"27H20 (Na,K)CaaAIgSi27072.24H20 (K,Na)6 AI6Si30072-20H20 (Na,K,Ca)A12Si10024.7H20

25

2.3.

Organic matter

volcanics. Analcite is also characteristically found as an authigenic mineral in hypersaline muddy sediments. Zeolites can be used as indicators of diagenesis and low grade metamorphism of argillaceous sediments. The mixed alkali (K, Na and some Ca) zeolites are progressively replaced by more sodic, less siliceous varieties which with increasing metamorphism are in turn replaced by albite and eventually K-feldspars. The calcic forms persist to higher temperatures. Zeolites have open framework silicate structures with the capability of acting as 'molecular sieves' by absorbing cations, anions and polar organic molecules into the open channels within their structures.

2.2.8.

Phosphates

The most common phosphate minerals occurring in clay materials are the carbonate apatite group (Cas(PO4, CO3)3(OH,F, C1). Phosphates may occur in a cryptocrystalline form referred to sensu lato as 'collophane'. The iron-rich phosphate mineral, vivianite, which has a very distinctive blue colour, may form in lacustrine environments and soils. Phosphates may be bioclastic (e.g. accumulations of vertebrate skeletal fragments), faecal or diagenetic in origin. Authigenic phosphates can precipitate in mudrocks and sandstones as nodules, ooliths, pisoliths, cements or replacements of calcareous skeletal clasts in environments with low clastic sedimentation and high organic productivity enriched in phosphate. In addition to being present in phosphate minerals, phosphorus is sorbed onto clay and other colloidal particles and this mechanism can be a major factor in controlling phophorus mobility in soil and aqueous environments.

2.2.9.

Halides

Halite (NaC1) and, to a much less extent, other halides such as sylvite (KC1) and camallite (KMgC13"6H20) are characteristically found in clay materials associated with evaporitic sedimentary and soil environments. Evaporation of sea water and chloride-rich brines can create a sequence of evaporitic minerals that are often cyclic due to replenishment of the evaporating brines or interbedded with fine-grained siliciclastic or carbonate sediments.

Organic matter is present in a variety of forms in clay materials including discrete organic particles, absorbed onto clay and other associated colloidal particles (e.g. iron oxides) and micro-organisms. The nature, abundance and distribution of organic matter can influence the physico-chemical behaviour of clay materials. Generally the organic materials can be subdivided into labile and refractory organic fractions. Labile organic matter generally does not survive burial whereas refractory organics are non-metabolizable, generally survive burial and can be preserved in ancient sediments as kerogen. Kerogen makes up 95% or more of the organic matter preserved in sedimentary rocks. Kerogen has been defined as the organic constituents that are insoluble in aqueous alkaline and common organic solvents (Tissot & Welte 1984) which is the equivalent of 'humin' in soil science (Tyson 1995). The labile organic components degrade via leaching and decomposition processes. The leaching process involves the breakdown of cellular material by intracellular enzymes producing soluble compounds. The decomposition process relates to the bacterial colonization and the fermenting activity of the bacteria which is greatest under warm, oxic conditions. In addition to bacteria, other micro-organisms may also be present such as protozoa and small algae, fungi and viruses. In soils there are 107 to 108 bacteria per gram of dry mass, 105 to 106 fungi and 104 protozoa (Campbell 1992). These micro-organisms show a wide range of sizes and morphologies and can also exist in various dormant states such as spores and cysts that are resistant to dessication and temperature changes. Micro-organisms can flourish even under apparently very adverse conditions such as hydrothermal vents on mid-ocean ridges with temperatures above 200~ hypersaline soda lakes with a pH > 10 or in acid mine drainage waters with a pH of <2. Some micro-organisms are aerobic, acquiring their energy from reactions with oxygen, which is consequently consumed, leading to oxygen-deficient conditions. Others are anaerobic, generating their energy by reducing inorganic chemical species (sulphate, nitrate, carbon dioxide) or organic species (fermentation reactions) and/or oxidizing other inorganic species (sulphide, ferrous ions) in the absence of oxygen. The bacteria Thiobacillus denatrificans uses sulphide, carbon dioxide and nitrate as energy sources and produces sulphate, sugars and reduced nitrogen compounds (Campbell 1992). This is one of many thiobacillus bacteria that exist in clay materials which derive energy from oxidizing inorganic sulphur usually under aerobic conditions. One variety, Thiobacillus ferrooxidans also oxidizes ferrous ions to ferric in acidic environments and is a key influence in the natural degradation of pyrite (Hawkins & Pinches 1992). The nature and abundance of micro-organisms in clay materials and the inorganic compounds available for

26

THE COMPOSITION OF CLAY MATERIALS

oxidation and/or reduction will have a major influence in determining their chemical behaviour.

2.3.1.

Organic-clay complexes and interactions

Organic-clay complexes represent part of the humin or kerogen organic fraction. Organic molecules are absorbed onto the charged clay mineral surfaces and interlayer sites by: 9 competing with and replacing the water molecules and forming complexes with interlayer cations; 9 being bonded to the cations by bridging water molecules; 9 bonding of organic cations, molecules with a strong tendency to form hydrogen bonds and polar organic molecules onto the charged clay surfaces. Most organic compounds of interest as environmental pollutants are small non-ionic molecules and their sorption onto clay minerals involves weak interactions via van der Waal forces and hydrogen bonding (Sawhney 1996). Such complexes and organo-clay reactions can have significant influences on chemical mobility and catalysis of chemical reactions. These can affect, for example, the chemical behaviour of pollutants, fertilizers, pesticides in soils and the physical properties of flocculation and dispersion in aqueous environments (Johnston 1996). The catalytic behaviour is related to the large surface areas of clays and other colloids, which is a function of mineralogy, crystal size and habit. Clay surfaces also have acidic properties due to the presence of A1 substituting for Si in the tetrahedral sites. The resultant net negative charge can be offset by the presence of charge balancing H3O corresponding to a proton-donating Bronsted acid site. In addition, AP + ions on the edges of clay particle are capable of being electron pair acceptors and thus can act as Lewis acid sites (Rupert et al. 1987). The degree to which this surface acidity is produced is an essential factor in creating the conditions for clay minerals to act as effective catalysts of various organic reactions.

this is most significant for the clay minerals because their fine grain size and platy habit produce very large surface areas and thus give a relatively greater capacity for water adsorption. The presence of adsorbed water layers covering the clay particles produces the characteristic cohesive plastic behaviour of clay materials. The water adsorbed onto the clay mineral surfaces has properties midway between bulk liquid water and ice and forms a structured water layer and a more diffuse less structured water layer. The thickness of the adsorbed surface water layer varies depending on the clay surface charge, the exchange cations and cation hydration and the salinity of the aqueous solution. With increasing ionic strength there is a lesser tendency for cations to diffuse away from the negatively charged clay particles and the absorbed water layer becomes compressed and less diffuse than at lower ionic strengths. The nature and behaviour of the water absorbed into the interlayer sites of clay minerals is a function of: 9 the polar nature of the water molecules; 9 the size and charge of the interlayer cations and their hydration state; 9 the location and value of the charge on the silicate layers of the clay mineral structures. In the interlayer sites of expandable clay minerals the water forms coordinated shells around interlayer cations. In the smectites the water molecules form single, double or triple water layers, depending on the humidity of the surrounding environment. The vermiculites can exhibit single or double water layers depending on the humidity conditions. There are significant differences in the interlayer water present in halloysites compared with the smectites and vermiculites that help explain why rewetting of dehydrated halloysites does not cause the structure to rehydrate. In the hydrated state there appears to be no preferential ordering of the water molecules and only weak interactions with the halloysite surfaces. There is also evidence that dehydration of the halloysite is accompanied by delamination of the halloysite layers thus inhibiting halloysite rehydration (Newman 1987). The absorption of water into sepiolites and palygorskites involves adsorption onto the external surfaces and absorption into the channel sites within the crystal structure, which is analogous to the absoption of water into the channel sites of zeolite structures. The packing of the water molecules absorbed onto the outer surfaces of the sepiolites and palygorskites is much less dense than than that of a surface water monolayer on other sheet silicates (Newman 1987). The density of the water in the interlayer sites of swelling clay minerals is considered to be greater than that of liquid water and to increase with increasing pressure (Skipper et al. 1993). However, Moore & Reynolds (1997) argued that with increasing pressure the interlayer water should have a similar density to that of liquid pore waters.

2.4.

Water

In clay materials water is present in the inter-particle pore spaces, adsorbed onto or absorbed into clay minerals, other minerals and organic matter. Guven (1993) defined three forms of hydration in clay materials: 9 interlamellar hydration involving adsorption of water onto the internal surfaces of clay mineral particles; 9 continuous osmotic hydration involving unlimited adsorption of water onto the internal and external surfaces of primary mineral and organic particles; 9 capillary condensation of water into micro-pores. It needs to be emphasized that all mineral surfaces are capable of adsorbing water onto their surface. However,

THE COMPOSITION OF CLAY MATERIALS

2/

2.5.

Conclusions

Clay materials are c o m p o s e d o f a varying m i x o f solid, liquid and vapour phases, with the clay minerals usually the most abundant mineral phases. The varying properties o f the individual clay minerals and their often c o m p l e x interactions with fluid phases under different physical and c h e m i c a l conditions can play a m a j o r role in controlling the b e h a v i o u r o f clay materials. H o w e v e r , it is simplistic and m i s l e a d i n g to equate clay materials with clay minerals. The nature and a b u n d a n c e o f other minerals and also organic matter, and their varying interactions with fluid phases, can also have a significant effect on the b e h a v i o u r o f clay materials. Clay materials are used in a variety o f applications as construction materials and a detailed k n o w l e d g e o f their compositions is o f fundamental importance if w e are to m a x i m i z e the efficient use o f such materials for a given application.

References
BAKER,P. A. & KASTNER,M. 1981. Constraints on the formation of sedimentary dolomite. Science, 213, 214-216. BERNER,R. A. 1981. A new geochemical classification of sedimentary environments. Journal of Sedimentary Petrology, 51, 359-366. BETHKE,C. M. & ALTANER,S. P. 1986. Layer-by-layer mechanism of smectite illitization and application to a new rate law. Clays and Clay Minerals, 34, 135-145. BLATT, H., MIDDLETON, G. & MURRAY, R. 1972. Origin of Sedimentary Rocks. Prentice-Hall, New York. BURST, J. F. 1958. Glauconite pellets: their mineral nature and applications to stratigraphic interpretations. AAPG Bulletin, 42 (2), 310-327. CAMPBELL, R. 1992. Microbiology of soils. In: HAWKINS,A. B. (ed.) Proc. IGCC 92 Conference on The implications of ground chemistry and microbiologyfor construction, Bristol, 1-17. CURTIS, C. D. 1983. Geochemistry of porosity enhancement and reduction in clastic sediments. In: BROOKS,J. (ed.) Petroleum Geochemistry and Exploration of Europe. Geological Society, London, 113-126. DREVER,J. I. 1982. The Geochemistry of Natural Waters. PrenticeHall, New York. ESL1NGER, E. & PEVEAR, D. 1988. Clay Minerals for Petroleum Geologists and Engineers. SEPM Short Course, 22. GUVEN, N. 1993. Molecular aspects of clay-water interactions. In: GUVEN, N. & PALLASTRO,R. M. (eds) Clay Water Interface and its Rheological Properties. CMS Workshop Lectures, 4, Clay Minerals Society, 2-79. GUVEN, N. & PALLASTRO,R. M. (eds) 1993. Clay Water Interface and its Rheological Properties. CMS Workshop Lectures, 4. Clay Minerals Society. HAWKINS, A. B. & PINCHES,G. M. 1992. Understanding sulphate generated heave from pyrite degradation. In: HAWKINS,A. B. (ed.) Proc. 1GCC 92 Conference on The Implications of Ground Chemistry and Microbiology for Construction. Bristol.

JOHNSTON,C. T. 1996. Sorption of organic compounds in clay minerals: a surface functional group approach. In: SAWNHEY,B. L. (ed.) Organic Pollutants in the Environment. CMS Workshop Lectures, 8. Clay Minerals Society, 2-44. KOHYAMA,N., FUKUSHIMA,K. ~: FUKAMI,A. 1978. Observations of the hydrated form of tubular halloysite by an electron microscope fitted with an environmental cell. Clays and Clay Minerals, 26, 25-40. Low, P. F. 1993. Interparticle forces in clay suspensions: flocculation, viscous flow and swelling. In: GUVEN,N. PALLASTRO, R. M. (eds) Clay Water Interface and its Rheological Properties. CMS workshop lectures, 4, Clay Minerals Society, 157-190. MCKINNON,I. D. R. 1987. The fundamental nature ofillite-smectite mixed layer clay particles: a comment on the papers by P.H.Nadeau and coworkers. Clays & Clay Minerals, 35, 74-76. MILLOT, G. 1970. Geology of Clay. Chapman-Hall, London. MOORE, D. M. & REYNOLDS, R. C. 1997. X-ray difraction and the identification and analysis of clay minerals. Oxford University Press. NADEAU,P. H. & BA1N,D. C. 1986. Composition of some smectites and diagenetic illitic clays and implications for their origin. Clays and Clay Materials, 34, 455-464. NADEAU, P. H., WILSON,M. J., MCHARDY,W. J. & TAIT, J. 1987. Interparticle diffraction: a new concept for interstratified clays. Clay Minerals, 19, 757-769. NEWMAN,A. C. D. 1987. The interaction of water with clay mineral surfaces. In: NEWMAN, A. C. D. (ed.) Chemistry of Clays and Clay Minerals. Mineralogical Society Monograph, 6, 237-274. RUPERT, J. P., GRANQUIST,W. T. & PINNAVAIA,T. J. 1987. Catalytic properties of clay minerals. In: NEWMAN,A. C. D. (ed.) Chemistry of Clays and Clay Minerals, Mineralogical Society Monograph, 6, 275-318. SAWNHEY, B. L. 1996. Sorption and desorption of organic compounds by soils and clays. In: SAWNHEY,B. L. (ed.) Organic Pollutants in the Environment. CMS Workshop Lectures, 8. Clay Minerals Society, 45-68. SAWNHEY, B. L. & REYNOLDS,R. C. 1985. Interstratified clay as fimdamental particles: a discussion. Clays & Clay Minerals, 33, 559. SHAW, D. B. & WEAVER,C. E. 1965. The mineralogical composition of shale. Journal of Sedimentary Petrology, 35, 213-222. SKIPPER,N. T., REFSON, K. & MCCONNELL,J. D. C. 1993. Monte Carlo simulations of Mg-and Na-smectites. In: MANNING,D. A. C. HALL, P. L. & HUGHES, C. R. (eds) Geochemistry of Clay-pore Fluid Interactions, Chapman & Hall, London, 40-61. THAUMASITE EXPERT GROUP 1999. The Thaumasite Form of Sulphate Attack." Risks, Diagnosis, Remedial Works and Guidance on New Construction. Department of Environment, Transport & Regions, HMSO, London. TISSOT, B. P. & WELTE, D. H. 1984. Petroleum Formation and Occurrence. Springer, Berlin. TUCKER, M. E. 1981. Sedimentary Petrology: an Introduction. Blackwell, Oxford. TYsoN, R. V. 1995. Sedimentary Organic Matter. Chapman & Hall, London. VAN OLPHEN,H. 1977. An Introduction to Clay Colloid Chemistry. Wiley Chichester. VELDE, B. 1992. Introduction to Clay Minerals. Chapman & Hall, London. YARIV, S. & CROSS, H. 1979. Geochemistry of Colloidal Systems for Earth Scientists. Sprin~er, Berlin.

You might also like