Glauconite

Glauconite

Earth-Science Reviews Elsevier Publishing Company, Amsterdam - Printed in The Netherlands Glauconite S.G. McRae ABSTRACT McRae, S.G., 1972. Glaucon...

3MB Sizes 450 Downloads 216 Views

Earth-Science Reviews Elsevier Publishing Company, Amsterdam - Printed in The Netherlands

Glauconite S.G. McRae

ABSTRACT

McRae, S.G., 1972. Glauconite. Earth-Sci.Rev., 8:397-440.. The term glauconite has been employed in two senses. It has been used most commonly as a morphological term for sand-sized greenish grains found in sedimentary rocks, but also as a name for a specific mineral species, a hydrated iron-rich micaceous clay mineral. The two uses are not synonymous, since not all morphological glauconite consists exclusively of mineral glauconite, nor is the latter restricted in its occurrence to such pellets. Mineral glauconite in sensu lato is a random interstratification of nonexpanding 10 A layers and expanding montmorillonitic layers. The amount of expandable layers may be over 50 % but it is customary to restrict the name mineral glauconite in sensu stricto to varieties with less than 10 % expandable layers. The variation in amount of expandable layers explains many of the observed variations in the properties of glauconite including chemical composition (especially potassium content), thermal characteristics, cation exchange capacity, colour, refractive index and specific gravity. Mineral glauconite is believed to form by the progressive absorption of potassium and iron by a degraded layer silicate lattice of low lattice charge and elimination of other silicate-lattice types under suitable environmental conditions, of which the most critical seems to be the redox potential. The catalytic activity of marine organisms is no longer thought to be essential, although decaying organic matter and empty foraminiferal tests supply the ideal environment for glauconite genesis. The process of glauconitization is arrested by rapid sedimentation, so that there is a relationship between the variety of mineral glauconite formed and the nature of the host rock. Glauconite is found associated particularly with marine transgressions. Morphological glauconite grains are believed to form as casts, faecal pellets or by aecretionary growth, but may have their morphology modified by subsequent re-working. A number of characteristic internal and external morphologies have been recognised. The wide range of environmental conditions suitable for its formation and its common detrital occurrence debars the use of glauconite in palaeo-environmental studies. Its major use in geology is for the absolute age dating of sedimentary rocks by the K - A r method. Glaueonitic deposits have no present day commercial value, but soils formed on glauconitic parent materials are notable for their fertility. Glauconite weathers by loss of potassium to produce a montmorillonitic or vermiculitic product with the release of, or oxidation of, structural iron, so that the grain has the appearance of having weathered,to limonite.

INTRODUCTION

This review attempts to provide a synthesis of the current views on the nature, m i n e r a l o g y , o c c u r r e n c e a n d f o r m a t i o n o f g l a u c o n i t e . It is n e c e s s a r y first to define t h e use o f t h e t e r m " g l a u c o n i t e " as it h a s a c q u i r e d a dual c o n n o t a t i o n ( B u r s t , 1958a, 1 9 5 8 b ) . It h a s b e e n m o s t c o m m o n l y u s e d b y geologists as a m o r p h o l o g i c a l t e r m for r o u n d e d , sandsized, g r e e n i s h , e a r t h y - l o o k i n g grains f o u n d i n s e d i m e n t a r y rocks. It h a s also b e e n emp l o y e d as a n a m e for a specific m i n e r a l species, a h y d r a t e d i r o n - r i c h m i c a c e o u s clay

398

S.G. McRAE

mineral related to the illites. It is not clear if the first uses of the name as "la glauconite" (Brongniart, 1823) or "glaukonit" (Keferstein, 1828) were meant as morphological or nfineralogical terms, but as Burst (1958b) has noted, the suffixes "-it" or "-ite", by subsequent convention, pressure for a milaeralogical interpretation. Current French nomenclature (Millot, 1964) reserves "glauconite" for the mineralogical interpretation and uses "la glauconie" for the morphological. This distinction is not made in English, and considerable confusion has arisen since the two uses of the word are not synonymous. Not all greenish earthy pellets from sedimentary rocks - morphological glauconite - consist exclusively of mineralogical glauconite, nor is the latter restricted in its occurrence to such pellets. Much of the variability in the reported physical and chemical properties of glauconite is due to the mineralogical variability of the pellets, and the validity of such data is questionable unless mineralogical information is supplied. The term "glauconite" used alone is thus best regarded as a morphological term and will be used as such in this review, while the description "mineralogical glauconite" will be employed, where justified, to distinguish the mineral species. The name glauconite has also been used erroneously for a closely related mineral, celadonite, occurring in altered volcanic rocks. Althotigh structurally very similar to glauconite (Hendricks and Ross, 1941; Pirani, 1963b; Wise and Eugster, 1964) it differs chemically from true mineral glauconite (Wise and Eugster, 1964; Foster, 1969). The term celadonite has been used erroneously for magnesium- (Urban, 1957) and iron(Malkova, 1956) rich varieties of glauconite. Skolite, a name suggested for aluminous varieties of glauconite (Smulikowski, 1936) has not been widely adopted (Lazarenko, 1956). The name glauconite, from the Greek TXavKo¢, bluish or pale green, was first used by Brongniart (1823), although the credit is sometimes given to Keferstein (1828). Previously the material had been called greensand, greenearth, grtinerde, terre verte, etc. Many of the early investigations on glauconite were made on the widespread occurrences in the Cretaceous and Tertiary of the N.E. United States (Bailey, 1856; Cook, 1868; Clark, 1894; Bagg, 1898) but a fresh impetus in the study of glauconite was given by the voyages of oceanographic survey ships (Von Gtimbel, 1886; Murray and Renard, 1891; Collet and Lee, 1906a, 1906b; Murray and Phillippi, 1908 quoted by Hadding, 1932; Correns, 1937) which provided information on glauconite in modern marine sediments.

f

Since graduation in 1965 from the University of Aberdeen S.G. McRae has been lecturer in geology and soil science at Wye College, University of London. His research has been concerned mainly with the relationships between geology and soils in Southeast England, and one aspect of this, the effect of indigenous glauconite on soil fertility, was the subject of a doctoral thesis, presented in 1971.

GLAUCONITE

399

Other notable contributions to the knowledge of glauconite and its occurrence, and often including good historical reviews, have been made by Clarke (1916, 1920, 1924), Mansfield (1919a, 1919b, 1920, 1922), Schneider (1927), Hadding (1932), Gruner (1935), Galliher (1935a, 1935b, 1939), Steenhuis (1937-1939), Hendricks and Ross (1941), Smulikowski (1954), Cloud (1955), Warshaw (1957), Burst (1958a, 1958b), Valeton (1958), Hower and co-workers (Toler.and Hower, 1959; Hower, 1961; Pratt (1961, 1962a, 1962b, 1963a, 1963b), Manghnani and Hower, (1964a, 1964b), Porrenga (1963, 1966, 1967), James (1966), Triplehorn (1966a), and Fairbridge (1967). MORPHOLOGICALGLAUCONITE

External morphology Morphological glauconite occurs mainly in the form of sand-sized grains (Fig. 1) usually in the size range of 100-500 /am (Williams et al., 1954; Teodorovitch, 1961; Pratt, 1963a; Mero, 1965). It may also occur in finely divided form as pigmentary glauconite and as a secondary mineral infilling cracks and hollows or replacing pre-existing minerals. The schemes of Cayeux (1897, 1916, 1932) were the first to attempt to classify glauconite on a morphological basis (see Millot, 1964 for details).

Fig. 1. Typical glauconitepellets from the Chloritic (Glauconitic)Marl, Lower Cretaceous, Folkestone, Kent.

400

S.G. McRAE

A more comprehensive and less genetic classification is that of Triplehorn (1966a): (1) Spheroidal or ovoidal pellets. This variety is very common, and many of the grains comprising the New Jersey greensands are of this type. (2) Tabular or discoidal pellets. A rather rare variety. (3) Mammillated pellets, of irregular shape, consisting of small rounded knobs separated by shallow sutures. Konta (1967) considers these to be a peculiar case of lobate (type 4) pellets, and Triplehorn (1967) has agreed that categories 3 (mammillated) and 4 (lobate) could be combined in a general category of lobate pellets. (4) Lobate pellets, very irregular with deep radial cracks which are triangular in cross section.They are said to resemble pop-corn. Such grains are common. (5) Capsule-shaped pellets (Triplehorn, 1966a, 1967; Konta, 1967). These are commonly thought to have formed from faecal pellets (Takahashi and Yagi, 1929; Takahashi~ 1939) but continued attrition of ovoidal pellets (category 1) could also result in grains with capsule or ellipsoidal shape. (6) Composite pellets. These are normally fairly large, up to 3 - 4 mm in diameter, and consist of smaller grains of glauconite and detrital minerals embedded in a glauconite matrix which is usually paler in colour. These grains are relatively uncommon. (7) Vermicular grains. These have also been called caterpillar, zebra, concertina, accordion or booklet grains, and form by the alteration of a flake of mica, probably biotite (Galliher, 1935a, 1935b, 1939; Carozzi, 1951; Seed, 1965), but possibly muscovite (Ichimura, 1940; Wermund, 1961 ; Hodgson, 1962). (8) Fossil casts and internal moulds, notably of foraminiferal tests, but also of sponge spicules, echinoderm spines and other shell debris. Grains with these morphologies may represent replacement structures or may be the end product of the transformation of detrital material infilling empty tests and shells. (9) Pigmentary glauconite, coating the surfaces and sometimes penetrating cracks and cleavages in other minerals. This may be due to direct replacement by glauconite of other minerals or may be a surface coating of finely divided glauconite derived from precipitation or alteration of finely divided detritus, or from disintegration of larger glauconite grains. Further categories may prove to be necessary to accommodate other distinctive morphologies, for example the "quarter-moon" and "cap" pellets of Wermund (1961). As will be discussed in a later section, the morphology of glauconite may be a valuable guide to its mode of origin. Differences in morphology may also be related to the mineralogical nature of the pellets (Ehlmann et al., 1963), although it is frequently found that pellets of very different morphologies are of similar composition (Barackman, 1964; Tapper and Fanning, 1968).

Internal morphology Associated with the external morphologies of glauconite listed above may be characteristic internal structures. The nomenclature schemes of Carozzi (1958, 1960), Wermund

GLAUCONITE

401

(1961) and Triplehorn (1966a) may be combined to give the following classification of internal structures: (1) Random microcrystalline - homogeneous aggregates of overlapping micaceous flakes without any preferred orientation. This is the commonest internal structure. (2) Oriented microcrystalline or possibly monocrystalline - pellets which tend toward unit extinction in polarized light. They seem (Warshaw, 1957) to be lamellar aggregates with a high degree of parallel orientation. (3) Micaceous or vermicular (Wermund, 1961) structures - similar to (2) but with the presence of incipient micaceous cleavage. All vermicular pellets have such an internal structure but not all pellets with micaceous internal structure are vermicular pellets. (4) Coatings on grains - chiefly on grains of heavy minerals and somewhat resembling oolites (see Carozzi, 1960 p.48). (5) Organic replacement structures - a great variety of fibrous, perforate or lamellar structures, sometimes preserving the internal structure of the organisms (Cayeux, 1897, 1935; Smulikowski, 1954; Triplehorn, 1961 ; Ehlmann et al., 1963), sometimes not. (6) Fibroradiated rims - glauconite grains with rims of different colour and structure compared to the bulk of the grain were reported by several early investigators (Murray and Renard, 1891 ; Cayeux, 1897; Bagg, 1898). These rims, termed the corona by Cayeux (1897) are peripheral zones of very small elongate crystals with radial orientation and apparently subdivided by numerous cleavage planes. In these fibroradiated rims (Triplehorn, 1966a) concentric layering is common and they may be continuous or discontinuous around the periphery, with an abrupt junction to the rest of the pellet. The rims are also more birefringent than the rest of the pellet (e.g., Atanasov, 1962). An extensive study of such rims has been made by Zumpe (1962, 1971) on glauconite from the "Chloritic Marl" of the Isle of Wight, showing the presence of two and sometimes three types of microstructure. Both Zumpe and Triplehorn believe these rims to have formed by accumulation or precipitation on a pre-existing grain rather than by post-depo.~itional alteration of the pellets. Thin sections of glauconlte may also show the presence of inclusions, for example pyrite (Bagg, 1909), calcite (Kharkwal, 1966) or iron oxides (Murray and Mackintosh, 1968). Similar material may be present as coatings to the grains (e.g., Hancock, 1961 and the occurrences of "limonite" weathering rims) or fill cracks, especially in the lobate type of pellet (Adams, 1963; Drobnyk, 1965). These cracks and hollows have been attrib,uted to either e.xpansion of the pellets (Pratt, 1963b; Norris, 1964)or to partial desiccation (Warshaw, 1957; Ehlmann et al., 1963; Bailey and Atherton, 1969). A detailed study of internal cracks in glauconites from Northern Ireland (Bailey and Atherton, 1969) has shown linings of fiorous crystallites of apatite or glauconite enclosing a monocrysta Uine glauconite or haematite filling. Colour and op tical properties

As implied by its name, glauconite is usually green in colour, although the shade may vary from dark green, almost black, to a pale greenish-yellow (e.g., Butt, 1932; Needl 1am,

402

S.G. McRAE

1934; James, 1966). More rarely grains which are red, white or grey have been noted (e.g., Wermund, 1961). The shade of green is related to the relative amounts of iron and aluminium present (Burt, 1932; Dallwitz, 1952; Borchert and Braun, 1963; Caill~re and Lamboy, 1970) and particularly to the ratio of ferrous to ferric iron (Weyl, 1951 ; Keller, 1953, 1958) or the presence of tetrahedral ferrous iron (Hoebeke and Dekeyser, 1955; Dekeyser, 1956). Changes in the amount and form of iron and aluminium accompany the formation of mineral glauconite from a grey or pale-brown precursor (see p.000) and as the mineralogy of the material approaches that of true mineral glauconite so the colour becomes greener (Collet and Lee, 1906a; Alexander, 1934; Galliher, 1935a, 1935b; Glass, 1953; Ehlmann et al., 1963; Porrenga, 1963, 1966, 1967). Mature, well-crystallised mineral glauconite is frequently dark green in colour (e.g., Hoskins, 1895; Pratt, 1963a, 1963b; Murray and Mackintosh, 1968; Caill~re and Lamboy, 1970). Oxidation of glauconite by weathering or by heating in an oxidising atmosphere will change the colour (at least of an outer skin) to a rusty brown. Such grains are frequently reported as limonite. Glauconite in thin section is usually a more vivid green than the grain colour and may be pleochroic (Collet and Lee, 1906a; Ross, 1926; Hendricks and Ross, 1941 ; Hutton and Seelye, 1941; Warn and Sidwell, 1953; Keller, • 958; Atanasov, 1962; Lerbekmo, 1963). The aggregate polarisation shown by most grains renders detailed optical work impossible and reported optical data refer ordy to grains with vermicular or oriented microcrystalline structures. This data includes a maximum birefringence of about 0.020, a biaxizl negative interference figure with 2 V up to about 40 ° and an acute bisectrix nearly perpendicular to (001) (Lacroix, 1893; Ross, 1926; Schneider, 1927; Hadding, 1932; Hutton and Seelye, 1941; Grim et al., 1951; Winchell and Winchell, 1951; Milner, 1962; Bentor and Kastner, 1965). The only optical feature of diagnostic value for glauconite is the index of refraction. This may range from 1.56 (Dallwitz, 1952) to 1.64 (Toler and Hower, 1959) and is related to chemical and mineralogical properties of the grains especially to the percentage of Fe203 (Schneider, 1927; Hendricks and Ross, 1941; Hutton and Seelye, 1941; Toler and Hower, 1959) and to the percentage of expandable layers in the mineral (Sabatier, 1949; Toler and Hower, 1959; Pratt, 1962b) as shown in Fig. 2 and 3.

Other properties The specific gravity of glauconite grains varies trom 2.3 to 2.9 depending on mineralogical composition and degree of drying (Sabatier, 1949; Mursky and Thompson, 1958; Bentor and Kastner, 1965; Murray and Mackintosh, 1968), the presence of impurities and the degree of oxidation of the surface (Pentland, 1931; Gildersleeve, 1932; Needham, 1934; Sidwelt, 1936). Even apparently similar grains may have different specific gravities and it is commonly found during an attempted separation of glauconite grains from other minerals using heavy nquids that some glauconite grains will float while apparently identical grains will sink. In addition, glauconite may interact with some heavy liquids,

GLAUCONITE

403

1"64

1"63 ~ ' ~ +

+ +

+

++ 1"62 nz

++ 1-61

1-60

1 59

1-58

+

5I

10 percent

J 15

expandable

L

20

25I

++ t

30

35

4~0

layers

Fig.2. Relation between index of refraction (nz) and percentage of expandable layers in glauconites (r = - 0.94; from Toler and Hower, 1959).

especially those containing thallium salts, e.g., Clerici solution (Gruner, 1935; Hutton and Seelye, 1941; Hutton, 1950). The magnetic susceptibility of glauconite is relatively high and varies with the content of iron (Sabatier, 1949; Bentor and Kastner, 1965), as shown in Fig.4 below. Glauconite ffains may be effectively separated from other mineral grains using an isodynamic separator (Ehlmann et al., 1963) or even a simple magnetic field (Mansfield, 1920). It is important however to separate every glauconite ~rain from a sample by repeated applications of a magnetic field, i f a bias towards separation of grains high in iron is to be avoided (Sabatier, 1949). Glauconite is reported to have a hardness of 2.0 on Moh's scale (Milner, 1962, p.111), but the hardness is likely to be quite variable (Hutton and Seelye, 1941). MINERAL GLAUCONITE

X-ray diffraction studies of glauconite Early workers, using only the results of total chemical analyses, could conclude little more than that glauconite grains consisted of "hydrous potassium aluminium silicate".

404

S.G. McRAE

1"64

1.63

~

+

+ 1.62 nz

1.60

1"59

1-58

i

i 10

i

J

L

i 15 wt.

percent

i

i 20

i

i

i

*

2;

Fe20 3

Fig. 3. Relation between index of refraction (nz) and wt. percent Fe 203 in glauconites (r = -0.82). The solid lines are estimated curves for change in n Z with change in Fe203 for equal amounts of expandable layers. The number on each line is the amount of expandable layers to which that line refers (from Toler and Hower, 1959). With the advent of X-ray diffraction techniques, however, the similarity of glauconitic material to the micas was established (Ross in Schneider, 1927; Gruner, 1935; Tyler, 1936, Maegdefrau and Hofman, 1937; Galliher, 1939; Mehmel, 1939), but it was generally assumed that glauconite grains were homogeneous. A few early investigators, however, recognised the interstratified nature of mineral glauconite (e.g., Gruner in Galliher, 1939; Nagelschmidt, 1944), but it remained for Warshaw (1957) and Burst (1958a, 1958b) to establish clearly the mineralogical variability of grains classed morphologically as glauconite. These workers found by X-ray diffraction of oriented specimens that many glauconite grains consisted of a random interstratifi'cation of nonexpanding 10 A micaceous layers and expanding layers assumed to be montmorillonitic in nature. This material may be called mineral glauconite in sensu lato, with considerable variation in the amount of interlayering with expandable layers. Only a few samples of morphological glauconite grains were found to consist of material with sufficiently little ( 5 - 1 0 %) expandable layers to be classed as mineral glauconite in sensu stricto. In addition, Burst found some grains which did not consist of mineral glauconite but were instead comprised of heterogeneous mixtures of two or more clay minerals o f unrelated structure. The X-ray classffi-

GLAUCONITE

405

~o weight

/

31

+

Fe203

+

27

+

25

_+_

+

-t-

23

FeO

21 19 17 15 2'1

2~3 2L5 2'7 magnetic

2'9

311 3L3 3'5

susceptibility

3'7

39

C.G.S. × IO - 6

Fig.4. Relationship between magnetic susceptibility and total iron content in glauconite (from Bentor and Kastner, 1965). cation scheme of glauconite pellets proposed by Burst (1958a, 1958b) is shown in Table I together with a similar, but slightly more comprehensive, scheme due to Bentor and Kastner (1965), developed from that of Warshaw (1957). Further studies on the type and extent of interlayering in glauconite were made by Hower (1961) who found that interlayering was "almost exclusively between expandable (montmorillonitic) layers and nonexpandable 10 A layers". That the expandable layers are montmoriUonitic rather than vermicutitic was confirmed by the relationship between the percentage of expandable layers and the percentage of potassium in the samples (Manghnani and Hower, 1964a). Using the cation exchange capacity techniques developed by Alexiades and Jackson (1965) to differentiate between montmoriUonite and vermiculite McRae (1971) also showed the absence of vermiculitic layers in mineral glauconite. Hower (1966) has observed that mineral glauconites are dominantly randomly interstratified whereas illites and illite/montmorillonites which appear to form as a result of the deep burial of argillaceous sediments are dominantly ordered. It is convenient to regard mineral glauconite in sensu lato as a more or less continuous sequence of minerals with increasing amounts of expandable layers, and to restrict the term mineral glauconite in sensu stricto to material in which the amount of expandable layers is less than 10 % (Hower, 1961 ; Bentor and Kastner, 1965), although Burst (1958a, 1958b) used a limit of 5 % expandable layers for this category. The limits of expandable

406

S.G. McRAE

TABLE I X-ray classification of glauconite pellets

Burst (1858a, 1958b)

Hower (1961)

Bentor and Kastner (1965)

Well-ordered - non-swelling high potassium lattice showing sharp symmetrical peaks characteristic ofa micaceous 10 A lattice at 10, 5 and 3.3 A. This constitutes the type mineral glauconite

<~ 10 % expandable layers (montmorillonite)

Class 1: mineral glauconite (a) well-ordered 1 M with symmetric and sharp diffraction at 10.1, 4.53, 3.3A. Reflections (112) and (112-) are always present

Disordered - non-swelling, low potassium lattice, micaceous and monomineralic but with subdued peaks, displaying broad bases and asymmetric sides

10-20 % expandable layers (montmorillonite)

(b) disordered 1 Md with asymmetric basal diffractions broadened at the base. Reflections (112) and (112) are absent

Interlayered clay mineral - extremely disordered, expandable, low potassium montmorillonite type lattice

> 20 % expandable layers (montmorillonite)

Class 2: interlayered glauconite, d (001) > 10.15 A

Mixed mineral ~ mixtures of two or more clay minerals as normal constituents - the most frequent combinations being of illite with montmorillonite, and illite with chlorite

Class 3 : mixed mineral (a) two or more clay minerals (b) mixtures of clay with non clay minerals

There is no separate classification for pellets containing minerals which could be classed as impurities

Class 4 : green pellets containing no glauconite

layers f o u n d by H o w e r (1961) to correspond to the categories o f mineral glauconite recognised by o t h e r workers are also s h o w n in Table I. Fully i n d e x e d X-ray diffraction data o f r a n d o m l y orientated crystals o f mineral glauconite have b e e n given b y Sabatier (1949), G r i m et al. (1951) and Warshaw (1957), the last having been taken as the diffraction pattern for t y p e glauconite by the ASTM p o w d e r index file. R a n d o m l y orientated X-ray diffraction patterns have also been used to determine the p o l y m o r p h s present ( S m i t h and Yoder, 1956), which are invariably IM, an ordered one-layer m o n o c l i n i c p o l y m o r p h , if the mineral glauconite has a low percentage o f e x p a n d a b l e layers, and IMd, a disordered one-layer m o n o c l i n i c p o l y m o r p h , if a higher

GLAUCONITE

407

percentage of expandable layers is present (Warshaw, 1957; Burst, 1958a, 1958b; Hower, 1961; Tyler and Bailey, 1961; Bentor and Kastner, 1965; Velde, 1965; Bailey and Atherton, 1969). These polymorphs are typical of low temperature formation of the lattice (Yoder and Eugster, 1955; Velde, 1965). Values of the (060) line in randomly orientated diffraction patterns suggest a generally dioctahedral nature (Hoebeke and Dekeyser, 1955; Zen, 1959; Tyler and Bailey, 1961 ; Bentor and Kastner, 1965), and may aid in the differentiation of mineral glauconite and illite (Zen, 1959; Bentor and Kastner, 1965). The single crystal X-ray studies of vermicular pellets by Seed (1965) show the more regular internal structure of this type of pellet as compared with other types of pellet. The unit cell parameters of glauconite have been given (Frank-Kamensky, 1960) as: a b c /3

= 5.25 kX = 9.09 kX = 20.07 kX =95 °

Thermal characteristics of mineral glauconite Dehydration curves of mineral glauconite show weight losses between ambient and 150 °C and between 400 ° and 60() °C (Ross, 1926; Takahashi, 1939; Nutting, 1943; Sabatier, 1949; Kampioni-Zakrzewska, 1957). These weight losses show up as endothermic peaks on Differential Thermal Analysis (D.T.A.) curves. The typical shape of the D.T.A. curve was first given by Grim and Rowland (1942) and has been generally confirmed by later workers although an additional exothermic peak at approximately 350 °C has been recognised by Hoebeke and Dekeyser (1954a, 1954b). An idealised D.T.A. curve is shown in Fig.5 together with the actual peak temperatures found by various workers. The reactions giving these peaks have been studied, notably by Hoebeke and Dekeyser (1954a, 1954b): Peak A - due to loss of adsorbed water and not regarded as an important diagnostic peak. McRae and Lambert (1968) have noted that the size of this peak is proportional to the percentage of expandable layers. Sabatier (1949) has shown that if unground glauconite grains are used in D.T.A. a low temperature shoulder appears on this peak representing water sorbed in pores within the grain. Peak B - due to oxidation of structural Fe z+ (Hoebeke and Dekeyser, 1954a, 1954b) associated with a loss of - OH ions. This does not occur in vacuo (Bentor and Kastner, 1965). Peak C - due to dehydroxylation, the peak height and position being affected by the heating rate. McRae and Lambert (1968) have observed that the position of the peak is related to the percentage of expandable layers. Peaks D and E - recrystallisation to spinel minerals (Hoebeke and Dekeyser, 1954a, 1954b; Mikheev and Stulov, 1955).

408

S.C. McRAE

"~-

EXOTHERMIC PEAKS

B

E

AT " ~

_

_

~ ~/ ~

~ \ 5OO~ '

ENDOTHERMICPEAKS

~~

' ~

~

~ Temp°C

c

Fig. 5. G e n e r a l i s e d D . T . A . c u r v e o f g l a u c o n i t e . Investigators

Peak temperatures (°C) reported (A)

(B)

(C)

(D)

(E)

Grim and Rowland

(1942)

110

-

550

-

950

Sabatier

(1949)

190

-

575

-

960

K a u f f m a n a n d Dilling

(1950)

150

-

550

-

-

G r i m e t al.

(1951)

145

340

600

865

915

d e B r u y n a n d v a n d e r Marel

(1954)

100

300

570

900-940

-

Hoebeke and Dekeyser

(1954a)

-

350

570

-

925

Zaprozhtseva

(1954)

150-175

-

560-575

865

-

Honda

(1957)

150

450

510

-

940

C l o o s e t al.

(1961)

170-245

355-390

600-620

-

860-930

Tedrow

(1966)

150

-

560

-

925-975

Drobnyk

(1965)

150

-

560

-

-

Parry and Reeves

(1966)

150

-

650

-

-

GLAUCONITE

409

Infra-red studies of mineral glauconite Few infra-red investigations of mineral glauconite have been made, but the work of Cloos et al. (1960), Owens and Minard (1960), Ehlmann et al. (1963), Manghnani and Hower (1964b) and Young et al. (1968) may be cited. The study of Manghnani and HOwer (1964) is of particular interest since it relates the I.R. spectra to the percentage of expandable layers as determined by X-ray diffraction. Changes brought about in mineral glauconite by, for example, acid attack, heat treatments or reducing conditions have also been studied by I.R. methods (e.g., Touillaux et al., 1960; Cloos et al., 1961; Zabelin, 1962).

Electron microscopy and diffraction studies of mineral glauconite Several electron microscope photographs of mineral glauconite have been published (e.g., Kitazaki, 1951, MiJller and Behne, 1954; Gritzaenko, 1956; Pfefferkon et al., 1956; Burst, 1958b; Aleixandre Ferrandis and Gonzalez Pefia, 1960; Cloos et al., 1961 ; Bentor et al., 1963; Bentor and Kastner, 1965) but, in general, have contributed little to the knowledge of glauconite. Well-ordered mineral glauconites are usually lath-shaped while disordered and mixed mineral pellets tend toward equilateral plates. Electron diffraction studies have been made by Kitazaki (1951).

Chemical investigations of glauconite A very large number of total and partial chemical analyses of glauconite grains have been reported in the literature. Collections of these have been compiled by, amongst others, Clarke ('1915, 1916, 1920, 1924), Ross (1926), Hadding (1932), Hendricks and Ross (1941), Smulikowski (1954), Borchert and Braun (1963) and James (1966). The conversion of these analyses to formQlae has reflected the cUrrent state of knowledge of the structure of glauconite, with early workers (e.g., Ross, 1926; Hallimond, 1928) presenting formulae as mixtures of oxides. Later, chemical analyses were converted to structural formulae similar to those for the micas (e.g., Gruner, ]935; Smulikowski, 1936; Hendricks and Ross, 1941; Hutton and Seelye, 1941). A representative formula for glauconite given by Hendricks and Ross (1941): (K,Cax/2 ,Na)o .s 4 (Alo .47 Fe 0.973 +Fe o. x92 +Mgo.4 o )(8i3.65 Alo. 3 s )O 10 (OH)2 may be compared with that given for illite by Grim et al. (1937): • Ko.s s(All.3sFeo.3v 3+ Feo.o4 2+ Mgo .34)(S13.41Alo .s 9)01 o(OH)2

This shows the general similarity of the two minerals, but with glauconite containing more iron and less aluminium and with greater layer charge requiring more interlayer potassium for charge balance. A serious criticism of most published formulae is that they do not take into account the

410

S.G. McRAE

known interlayering within the structure which may cause large differences in structural formulae, as shown by Hower (1962, quoted by Manghnani and Hower, 1964a) in two representative formulae:

(Alo.2 s Fe i .103 +Ro.6 s 2 ÷XSi3.6 s Alo.3 s )O1 o(OH)2 Xl.o 1+ ( ( 5 % expandable layers containing Ko.9o) (All .30Feo.s o 3 +Ro.2 o2+)(Si3.6 s Alo.3 s )O1 o (OH):

Xo.s s l +

(40 % expandable layers containing Ko :a s) where R 2+ is the sum of magnesium and ferrous iron, and X ~+ is the number of moles of univalent interlayer cation which includes the amount of potassium given for each formula. With a few notable exceptions (Bentor and Kastner, 1965; Murray and Mackintosh, 1968; Bailey and Atherton, 1969) formulae are rarely corrected for the presence of impurities such as non-structural iron oxides (Mehra and Jackson, 1960; Bentor and Kastner, 1965), exchangeable cations (Foster, 1951), inclusions of other minerals, e.g., apatite and calcite, or the presence of non-glauconitic clay minerals (Alexiades and Jackson, 1966). In addition, the trial and error allocation of atoms between tetrahedral and octahedral layers to produce charge balance rarely takes into account the cation exchange capacity which is frequently large, and may also cause misallocation of aluminium (Bentor and Kastner, 1965), ferrous iron (Hoebeke and Dekeyser, 1955; Cloos et al., 1960) and perhaps other atoms. In spite of these objections, however, some marked trends in the chemical composition of glauconites can be observed, in particular the close inverse relationship between the potassium content of mineral glauconite and the percentage of expandable layers present in the lattice (Burst, 1958a, 1958b; Hower, 1961; Manghnani and Hower, 1964a; McRae and Lambert, 1968), as shown in Fig.6. It would appear that all glauconite pellets with > 8% K20 and most with > 7% K20 consist of mineral glauconite sensu stricto, with less than 10% expandable layers. Only half the specimens of Hendricks and Ross (1941) were thus likely to have been genuine mineral glauconite sensu stricto. Their generalised formula, given above, is thus incorrect as a formula for mineral glauconite sensu stricto, for which more representative formulae are:

Ko.79 Cao.o s (Alo.3 s Fe1.063 +Feo.2 s 2 +Mgo.41 )(Si3.61 Alo.a 9)01 o(OH)2 or Ko. 76 (Na,Ca)o. 1a (Alo .4 o Feo .s 73 +Feo .492 +Mgo .4 o)(Sia .42 Alo .s a )01 o (OH)2 i.e., those of the Bonne Terre and Franconia mineral glauconites of Cambrian age, respectively, given by Burst (1958a) as type mineral glauconites. Selective chemical dissolution studies made by Alexiades and Jackson (1966) have shown the Franconia glauconite, which should by defintion contain no chlorite, to be rich in chlorite. There is thus still a need to establish a type mineral glauconite sensu stricto.

GLAUCONITE

5

411

3 samples [-

4-

6

j

....." °."'" ~ " ~ ~ ' M C R a e & Lambert (1968)

-I-y"

~.-'"'" 4-

°"

.....'" \

~

7

§

8 ~-

~ - F . . . ' Y

1

0

10

20

30

40

50

percent expandable layers

Fig.6. The relationship between the percentage of expandable layers determind by X-ray methods and potassium content of glauconites. The dashed line is that given by Manghnani and Hower (1964a) and the solid line refers to the points determined by McRae and Lambert (1968). Mineral glauconites have a rather constant content of about 4 % MgO, i.e., 0.35-0.45 Mg ions per unit half cell (Hendricks and Ross, 1941; Millot, 1964; Bentor and Kastner, 1965). Non-marine occurrences of glauconite (see p.415) are of a variety richer in Mg and poorer in A1 than marine varieties (Tyler and Bailey, 1961). There seems a narrow range for the Fe 3 +/Fe 2 + ratio of 4.1 to 6.2 (Hendricks and Ross, 1941) with non-marine varieties having a rather low ratio (Tyler and Bailey, 1961). Most of this iron is believed to occupy octahedral sites, although evidence, such as the inverse relationship between FeO and A1203 contents (Bentor and Kastner, 1965) suggests that some Fe 2+ or Fe 3+ may occupy tetrahedral sites (Sabatier, 1949; Hoebeke and Dekeyser, 1955; Cloos et al., 1960). The number of octahedral cations frequently exceeds two per unit half cell and it seems likely that mineral glauconite is not a true dioctahedral mineral (Yoder, 1959; Bentor and Kastner, 1965). The amount of aluminium substituting for silicon in tetrahedral sites is normally greater than 0.21 ions per unit half cell. The amount of aluminium in octahedral sites is, however, very variable - from 0.12 to 1.27 ions per unit half cell (Hendricks and Ross, 1941). Traces of other elements are also found in glauconite grains 0-lower, 1961 ; Bentor and Kastner, 1965; Yasyrev, 1966; Balashov and Kazakov, 1968), although they may not be truly structural. The content of strontium, barium, vanadium and calcium seems related

412

S.G. McRAE

to the amount of expandable layers present (Hower, 1961; Bentor and Kastner, 1965) but other elements such as boron, rubidium and chromium do not show this trend. Inclusions of other minerals such as futile (Allen, 1937) and apatite (e.g., Hendricks and Ross, 1941) may account for some or all of the titanium, calcium and phosphate frequently reported in glauconite grains. The chemical relationships between mineral glauconite sensu lato, celadonite and other dioctahedral hydrous 2 : 1 layer silicates may be seen in the triangular diagram given initially by Yoder and Eugster (1955) but modified by Bentor and Kastner (1965) to eliminate from the field of glauconite those species with substantial amounts of expandable layers.

I'O/~$'0

7

@

I "2

~

1.4/~

x~" 0 0

/ / I'6 I I

o 0

_ ~J \-"\

/

~"/"

\

"'"\ \ I .

.......

-

TETRAHEDRAL

¢~ \ O's

o tmor

O.CD

.

o 0~

7"06`

.....9?" \ \"'" I

B

o c9

"j

0 ~

t

% - Ll'Jx

~

o ~

0

R -~3

Fig.7. Plot of tetrahedral R 3+ and octahedral R 3+ in atomic proportions of dioctahedral micas and related minerals. Field A (dotted line) is that originally given for glauconite by Yoder and Eugster (1955); field B (solid line) is the revised field suggested by Bentor and Kastner (1965).

Layer change and cation exchange capacity of mineral glauconite The distribution of ions between octahedral and tetrahedral sites suggests that the two layers contribute more or less equally to the overall charge of the mineral glauconite unit, although the octahedral layer tends to have a slightly higher charge than the tetrahedral layer (Foster, 1956, 1960). The overall negative charge is frequently not balanced by the

GLAUCONITE

413

interlayer potassium, giving rise to a net negative charge (the cation exchange capacity). Several of the mineral glauconites studied by Bentor and Kastner (1965), however, had a higher charge in the tetrahedral layer than in the octahedral, thought to be due to ferrous iron and especially magnesium partly filling vacant spaces within the octahedral layer. An increase in interlayer cation content has been observed (Tyler and Bailey, 1961) to be accompanied by a decrease in octahedral cations, i.e., an increase in net octahedral charge. This is believed to be a real trend whereas the apparently similar trend in montmorillonites has been shown (Foster, 1951) to be due to misallocation of Mg between structural and exchange sites. The cation exchange capacity (c.e.c.) of glauconite varies inversely with the amount of potassium and hence directly with the percentage of expandable layers (Manghnani and Hower, 1964a; McRae and Lambert, 1968). The range of c.e.c.'s found by these investigators is from 5 to 39 milli-equivalents per 100 g, a range easily covering the values earlier given by Grim et al. (1951), Carroll (1959) and Owens and Minard (1960). Manghnani and Hower (1964a) found that the solution of the equation of c.e.c, as a function of % expandable layers gave values of 8.2 and 62 mequiv./100 g for zero and 100 % expansible layers respectively. Exclusion of samples with > 30 % expansible layers which seemed to have anomalously low c.e.c, values gave a 100 % expansible layer c.e.c. value of 90 mequiv./100 g, a result consistent with the expandable layers being montmorillonitic in nature. The low results for the highly expandable glauconites, mostly of geologically recent range, was attributed to blocking of exchange sites by organic material or by the presence of fixed cations in the expandable layers. The selectivity of the exchange sites for monovalent cations has been studied somewhat inconclusively by Sabatier (1949), and for trace element cations by Libor (1962a, 1962b) and Libor and Varga (1963). Studies of glauconite as a "scavenger" for the nuclear waste products caesium and strontium by Schnepfe et al. (1964) have shown that the selective uptake of these from solution can be increased by increasing the pH of the medium. The c.e.c, of glauconite may also contribute to the total c.e.c, of sediments and soils in which it is found, in addition to that from the colloidal fraction. Tedrow (1966), for example, has studied a glauconitic soil whose sand fraction, containing glauconite, had a c.e.c, of 19 to 31 mequiv./100 g. THE OCCURRENCE OF GLAUCONITE Glauconite, both morphological and mineral, is restricted in its occurrence and formation to sedimentary rocks. It is most commonly found in marine sediments, but has also been reported from lacustrine and other alluvial occurrences. The glauconite found in sediments may be authigenic, i.e., may have formed more or less in situ or may be allogenic, i.e., detrital. The distinction between these two origins will usually involve consideration of the petrology of the rock (e.g.,Warn and SidweU, 1953; Wermund, 1961),the morphology of the grains (Light, 1952; Carozzi, 1958, 1960; Owens and

414

S.G. McRAE

Minard, 1960; Wermund, 1961; Adams, 1963; Triplehorn, 1966a; Zumpe, 1971), the mineralogy of the grains and associated sediment (Lebauer, 1964; Bell and Goodell, 1967), or the ratio of ferrous to ferric iron and the amount of interlayer ions (Owens and Minard, 1960). The detrital origin of glauconites can also be shown by K - A r dating (Allen et al., 1964; Bodelle et al., 1969) if the age of the host rock is known. Detrital glauconite may have been derived more or less contemporaneously from a nearby zone in the basin of deposition in which authigenic glauconite is forming especially under alkaline pH conditions. It may alternatively have been derived from submarine outcrops of pre-existing glauconitic rocks (e.g., Pratt, 1962a, 1962b; Bell and Goodell, 1967; Murray and Mackintosh, 1968). Derivation involving fluvio-detrital transport is possible (Allen et al., 1964; Bodelle et al., 1969) but unlikely (Wermund, 1964). The physical and chemical limits of glauconite formation have been discussed by Hadding (1932), Cloud (1955) and Lochman (1957). Their work has formed the basis of the limiting conditions set out in the following sections.

Stratigraphic range Glauconite may be found in sedimentary rocks of all ages from Precambrian to the present day. Cloud (1955) believed that glauconite was not present in Precambrian rocks but Precambrian occurrences have since been noted, e.g., Schaub (1955) and many workers using glauconite for K - A r dating (see below). It is particularly common in sediments of Cretaceous and Tertiary age, and areas of these rocks notably in S.E. England (Hallimond, 1922; Oakley, 1943; McRae and Lambert, 1968), N. France (Cayeux, 1897, 1916, 1935), N.E. United States (Cook, 1868; Mansfield, 1919a, 1919b, 1920, 1922; Zodac, 1945), New Zealand (Hutton and Seelye, 1941), Australia (Feldtmann, 1934; Carroll, 1939, 1941) and Russia (Glinka, 1896) are rich in glanconite. Glauconite is also fairly common in Lower Palaeozoic rocks, but relatively rare in rocks of other geological ages. It has been noted (Conway, 1942; Smulikowski, 1954; Hower, 1961) that glauconites from older rocks have more potassium and are less frequently mineralogically heterogeneous than younger glauconites. Lower Palaeozoic glauconites are thus more likely to consist of mineral glauconite sensu stricto and the two glauconites suggested by Burst (1958a, 1958b) as type examples are both of Cambrian age. This variation in potassium content with age has been attributed to a greater salinity in seas of earlier geological epochs (Tarr, 1920; Conway, 1942), but is probably due to lithological differences in the main types of host rock found in each era (Hower, 1961). There is evidence that glanconite may gain potassium in postdiagenic changes (Hower, 196t). It has also been shown (Smulikowski, 1954) that there is a reduction in the ratio of ferric to ferrous iron with increasing geologic age of glauconites. The use of glanconite in isotopic age dating of rocks will be discussed in a later section.

GLAUCON1TE

415

Present areal distribution Glauconite may be found forming today in all the oceans of the world, the limits of 65 ° S to 80 ° N given by Cloud (1955) delimiting simply the Antarctic continent and Arctic ice. It is found forming mainly on continental shelves away from large streams but does not seem restricted solely to the areas near granite- or biotite-bearing rocks as suggested by Galliher (1935a, 1935b, 1939) and Berthois and Dangeard (1939), although the potassium required for its formation would be plentiful in such areas.

Salinity It was believed for long that glauconite was restricted in its occurrence to a marine environment of normal salinity, although Cloud (1955) did admit of a possible theoretical occurrence in potassium-rich saline lakes. Glauconite from lacustrine and other continental deposits has been described by Jung (1954), Dyadchenko and Khatuntseva (1955, 1956), Keller (1958, 1959), Nicolas (1961), Parry and Reeves (1966) and Porrenga (1968). pH conditions which are slightly alkaline (pH 7 - 8 ) seem to be favourable for glauconite genesis (Dapples, 1967; Fairbridge, 1967).

Oxygenation The formation of glauconite has been reported in environments ranging from moderately anaerobic (Lochman, 1949; Cloud, 1955) to strongly oxidising (Van Andel and Postma, 1954); it is most common around Eh 0, and although some workers (e.g.,Teodorovich, 1946, 1947, 1954 - quoted by Larsen and Chilingar, 1967; Borchert, 1960a, i960b) favour a slightly oxidising environment (Eh +50 to +400 mV) most would follow the view that a slightly reducing environment (Eh 0 to - 2 0 0 mV) is more conducive to glauconitization (Lochman, 1949; Krumbein and Garrels, 1952; Cloud,1955; Burst, 1958a, 1958b; Dapples, 1967; Fairbridge, 1967). Sometimes (Emery, in Cloud, 1955) it can be shown that although the macro-environment may be strongly oxidising, the actual glauconitization occurs in a more reducing micro-environment, such as is accorded by the empty tests of forams. The oxygenation requirements are most commonly met where there is slow bacterial decomposition of organic matter (see next section) although such conditions may be met in other ways, for example by escaping gases from coal (Seed, 1965). More complicated mechanisms have been proposed, including the movement of oxygenated water to reducing areas causing local zones suitable for glauconite formation (Tyler and Bailey, 1961; Leclaire, 1964; Aldinger, 1965). Alternatively, a precursor of glauconite or "pro-glauconite" may itself be moved from an oxidising to a reducing milieu (Millot, 1949) or vice versa (Krotov, 1953, quoted by Chilingar, 1956; Zaporozhtseva, 1954; Adams, 1963).

416

S.G. McRAE

Organic matter of bottom sediments It is generally agreed (Cloud, 1955) that formation of glauconite is facilitated by the presence of decaying organic matter providing the required oxygenation conditions described above. Such an environment is a favourable habitat for sediment-ingesting organisms with low oxygen requirements, but there has been considerable discussion as to whether or not the presence of organisms is essential for glauconite genesis. Many early workers, (e.g., Bailey, 1856; Ehrenberg, 1856; Murray and Renard, 1891) believed that an active, living organic agency was required, whereas others (e.g., Von Giimbel, 1886; Prather, 1905), while admitting the observed fact that glauconite pellets are often of casts of forams or coprolitic pellets, believed that these occurrences were only specific cases in a general mode of origin which did not absolutely require the presence of organisms. The currently accepted view is that the presence of organisms is not essential for the formation of mineral glauconite but that empty tests and shells can form suitable microenvironments for the mineralogical transformations, and the formation of grains of glauconite may in many cases involve organic activity, e.g., faecal pellets.

Temperature The view of Cloud (1955) that glauconite formation is more generally favoured by cool water is incorrect (Fairbrige, 1967) for its most widespread occurrence today is in warm shelf seas. The precise temperature limits of glauconite formation have not been established but a minimum temperature of about 15 °C (Porrenga, 1967) and a maximum of 20 °C (Takahashi and Yagi, 1929) would seem reasonable limits. Water temperature together with oxygenation conditions serve to explain the depths at which glauconite formation is observed (see next section). The relative absence of glauconite in preMesozoic oceans (Conway, 1942) may be due to the prevailing relatively warm conditions favouring the formation of chamosite rather than glauconite (Braun, 1962) or may instead be due to the postulated higher pCO2- (Fairbridge, 1967). Favourable conditions for the formation of glauconite are said to occur where cold and warm currents meet (Murray and Phillippi, 1908 - quoted by Cloud, 1955; Goldman, 1921a, 1921b; Hummel, 1922; Gill, 1927; Takahashi and Yagi, 1929; Borchert, 1960a,1960b) probably because of the high organic productivity resulting from such conditions. In general it seems unwise to regard glauconite as a reliable temperature index (Cloud and Barnes, 1948).

Depth Glauconite is unlikely to form in water shallower than about 15 m because of waveinduced turbulence (Murray and Renard, 1891; Cloud, 1955; Fairbridge, 1967). Below this depth its formation is governed by the temperature and oxygenation conditions (e.g., Borchert, 1960a, 1960b, 1964; Porrenga, 1963, 1966, 1967; Giresse, 1965). In tropical

GLAUCONITE

417

areas, the warm surface water conditions seem to prevent glauconitization until cooler waters at depths of 250 m and deeper are encountered (Porrenga, 1967). In non-tropical areas, glauconitization is commonly found at depths as shallow as 30 m (Cloud, 1955; Porrenga, 1967). Glauconite formation is rarely noted in water deeper than that encountered at the shelf margin (about 500 m - Fairbridge, 1967). Glauconite grains found at depths greater than this (e.g., 4,000 m by Murray and Renard, 1891) are likely to have been transported to these localities and not to represent in situ formation at these depths. It would appear that, unless other conditions such as those of temperature and oxygenation can be established independently, glauconite cannot be used as an indicator of water depth at the time of sediment deposition.

Turbulence Evidence of turbulence is commonly found in the associated sediments (Hadding, t932; Lochman, 1949; Cloud, 1955), although continued marked turbulence would seem incompatible with the requirement for a particular narrow range of oxygenation conditions. It is most probable that where glauconite grains are found in association with agitated sediments they are allogenic (Wermund, 1961), i.e., introduced more or less contemporaneously from other parts of the sea-floor nearby or, less likely, derived from pre-existing rock strata. It is then possible to postulate the absence of turbulence at the actual loci of glauconite formation, or the restriction of glauconite formation to nonturbulent times (Lochman, 1949). Odin (1971) has suggested that the formation of glauconite is favoured by alternating flow of the water which allows formation of the granules and transport of ions and of food for organisms whose catalytic activity is believed essential.

Source materials The starting materials for glauconite formation stated by Cloud (1955) are, in general, micaceous minerals, preferably of a degraded nature, or bottom muds of high iron content. In view of the variety of theories for the formation of glauconite to be reviewed later, it would seem advisable to reserve discussion of the source materials for glauconite formation until then.

Sedimentary influx and nature of associated sediment The formation of glauconite is favoured by a low or even a negative sedimentation rate (Murray and Renard, 1891; Goldman, 1921a, 1922; Hadding, 1932; Cloud, 1955; Miiller, 1967). Such conditions, together with other favourable environmental conditions, may be provided during a marine transgression, so that glauconite is commonly found associated with unconformities and other stratigraphic breaks (Goldman, 1921a, 1921b; ReckFroUo, 1963). In modem shelf areas it is almost invariably found associated with relict deposits dating from the inundation of these shelf areas due to rising sea level following

418

S.G. McRAE

the last glaciation (e.g., Allen, 1965; Clarke, 1969). It has been suggested (Rech-Frollo, 1963) that the large amounts of organic matter from inundated terrestrial soils during a marine transgression may help to provide favourable environmental conditions for glauconite genesis. Glauconite forms most readily at the water/sediment interface and rapid sedimentation would arrest tile glauconitization process by burying the developing glauconite grains. Thus glauconite grains from sediments with high influx of detritus, e.g., argillaceous sandstones and marls frequently contain extraneous clay minerals and the mineral glauconite has a higher percentage of expandable layers (Hower, 1961). By contrast, glauconite grains from clean sandstones, limestones and dolomites with low sedimentation rates are almost exclusively monomineralic, consisting only of mixed-layer illite/ montmorillonite with a low percentage of expandable layers. The glauconite associated with geological unconformities is frequently of a highly ordered form (Burst, 1958b) suggesting slow burial of the glauconite. It would be interesting to establish from modern shelf areas whether the mineralogical nature of the glauconite was related to the rate of sedimentation following the transgression, and if possible to establish the time required for the glauconitization process to reach completion. The lithological rock types in which authigenic glauconite is most common are calcareous detrital sediments such as calcareous sandstones and impure granular limestones, while it is rare in pure clay rocks, pure quartz sandstones or chemically precipitated carbonates. It may occur as detrital glauconite in a wide range of sedimentary rocks, particularly if calcareous, since it seems highly stable in alkaline conditions (Fairbridge, 1967). As a detrital mineral it may become associated with other minerals with which its genesis is incompatible e.g., chamosite (Chilingar, 1956). Glauconite is commonly found associated with phosphatic deposits (e.g., Goldman, 1922; Wetzel, 1937; Kazakov and lsakov, 1940; Carozzi, 1958; Litvinenko, 1965; Rooney and Kerr, 1967; Bailey and Atherton, 1969) and more rarely with other nonclastic deposits such as fluorides (Simpson, 1934) and celestine (SrSO4)(Mogarovsky, 1963a, 1965). It is sometimes found in association with altered volcanic ash (Ross, 1926; Atanasov et al., 1964; Hallam and Sellwood, 1968; Seed, 1968) although such occurrences may really be of celadonite (e.g., Rossetti and Sitzia, 1956; Pirani, 1963a, 1963b; Kubovics, 1964; Peyronel Pagliani and Pagnani, 1965). Green grains, identified as glauconite usually only on morphological evidence, may be found in beds of sedimentary iron ore (e.g., Spurr, 1894; Cayeux, 1906, 1924; Clarke, 1924; Hallimond, 1925; Eckel, 1938; Taylor, 1959; James, 1966; Amstutz and Bubenicek, 1967). Some reported occurrences of glauconite (Spurr, 1894; McCallie, 1908; Galliher, 1935b)have been shown to be of chloritic minerals such as greenalite (Clarke, 1903; Leith, 1903; Jolliffe, 1935) or chamosite (Brown, 1914; Caill~re and Giresse, 1966). Conversely some "chloritic material" in iron ore deposits has proved to be glauconite (Spurr, 1915; Ross, 1916). The name chlorite has also been used erroneously for the well-crystallised mineral glauconite of the Bonneterre Formation in Missouri (Tarr, 1918) and of the "Chioritic Marl" at the base of the Chalk in S.E. England which is in reality a glauconitic marl (Jukes-Browne, 1903, pp.12-14; McRae and Lambert, 1968).

GLAUCON1TE

419

THE FORMATIONOF GLAUCONITE Theories explaining the formation of glauconite have to account for both the mineralogy and the morphology of the material formed. Most theories have concentrated on the mineralogical aspects of glauconite genesis. These will be reviewed first, followed by a review of the mechanisms proposed to account for the morphology.

The formation of mineral gtauconite The early theories of mineral glauconite genesis proposed the co-precipitation of'Mg, Fe, A1 and Si gels to which potassium was attracted (Murray and Renard, 1891 ; Collet, 1908). Equally vague were theories proposing the alteration of mud (e.g., Collet, 1908; Takahashi and Yagi, 1929, Takahashi, 1939). Galliher (1935a, 1935b, 1936, 1939), following earlier observations of Hummel (1923) and Alexander (1934) that there was an association between the presence of biotite and the formation of glauconite, showed that biotite could be converted to glauconite in a marine environment. This theory has received widespread support, and while it is almost certainly correct for booklet or accordion grains of glauconite (e.g., Hough, 1940; BaUance, 1964; Seed, 1965) the view of Galliher that biotite is the universal precursor of glauconite is over-stated, and the theory is best regarded as a rather specific case of the modern "layer lattice theory". According to the "layer lattice theory" (Burst, 1958a, 1958b) the formation of mineral glauconite requires simply: (1) a degraded layer silicate lattice; ( 2 ) a plentiful supply of iron and potassium; and (3) a suitable redox potential. Suitable redox potentials can be produced by decaying organic matter and even in a generally oxidising environment can occur in faecal pellets or within foramiferal tests and other debris, hence the frequent glauconitization of these materials. Other environmental factors, e.g., temperature, rate of sedimentation and parent material may influence the type of mineral glauconite formed. The process of glauconitization consists of absorption of potassium and iron by the degraded lattice leading to a reduction in the amount of expandable layers so that given suitable chemical conditions and sufficient time the material would approach the composition and structure of mineral glauconite sensu stricto. The precise mechanisms involved in the diagenetic ~hange from a degraded layer lattice to mineral glauconite have been further studied by Hower (1961). The starting material for glauconitization should have a low lattice charge since high lattice charge material readily adsorbs potassium from sea water and collapses to a non-expandable "lattice (Weaver, 1958), whereas expandable layers seem to be necessary for the initial adsorption of iron prior to its diffusion into the octahedral layer replacing aluminium ions. The model predicts that the increase in net lattice charge during glauconitization should take place by substitutions in the octahedral layers, and that the corresponding rise in the amount of charge-balancing interlayer cations (chiefly potassium) would be related to the net octahedral charge but be unrelated to the tetrahedral charge. This has been demonstrated by Hower (1961), as has the expected relationship between the contents of

420

i

S.G. McRAE

iron and potassium in mineral glauconites. This latter relationship has not, however, been confirmed by Bentor and Kastner (1965) or Foster (1969) who consider the incorporation of iron into the structure and the fixation of potassium to be unrelated processes. The model predicts that burial of the mineral glauconite should arrest the glauconitization process to give a low-potassium disordered mineral glauconite, which would explain the higher potassium (Conway, 1942; Smulikowski, 1954; Hower, 1961) and iron contents (Smulikowski, 1954) of glauconites from older geological epochs where the dominant environments were those with slow sedimentation rates. The higher potassium content o f older glauconites may also be due partly to post-diagenetic gains of potassium (Evernden et al., 1961; Hower, 1961; Hurley, 1961; Hower et al., 1963). The process of glauconitization tends to eliminate from the pellets any layer lattices not of the mineral glauconite type, e.g., kaolinite or chlorite, and other extraneous materials, making mature glauconites virtually monomineralic. It is believed (Burst, 1958a; Hower, 1961) that the mineralogical character of the glauconite pellets reflects only in part the source material, for example the tetrahedral charge, but that the evidence for diagenetic changes controlling the resulting mineralogy is overwhelming. During glauconitization the colour of the glauconite becomes darker green and the degree of hydration and refractive index decrease. These and mineralogical changes which confirm the hypotheses of Burst (1958a, 1958b) and Hower (1961) have been observed in modern glauconitic sediments by a number of workers (Glass, 1953; Pratt, 1961, 1962a, 1962b, 1963a, 1963b; Ehlmann et al., 1963; Porrenga, 1966). An alternative theory proposed by Ehlmann et al., (1963) in which weathered mica layers act as templates on which glauconite layers form, seems a specific case of a more general "epigenetic substitution" theory proposed initially to explain the replacement of calcite by mineral glauconite (Cayeux, 1924, 1932, 1935, 1938). Such replacement has been frequently reported (e.g., Dangeard, 1940; Houbolt, 1957; Wermund, 1961; Mogarovsky, 1963b, 1965; Kharkwal,1966; Lamboy, 1968). In situations where both silicate and carbonate materials are present, substitution apparently takes place preferentially in the carbonates (Seed, 1968). In the Qbsence of carbonates, however, mineral glauconite has been observed replacing such diverse non-micaceous silicate lattices as quartz (Hadding, 1932; Ichimura, 1940; Schreyer,1956; Tresise, 1960), orthoclase (Cayeux, 1897; Collet and Lee, 1906a; Murray and Phillippi, 1908; Ichimura, 1940; Dyadchenko and Khatuntseva, 1955, 1956; Gierwielaniec and Turnau-Morawska, 1961 ) plagioclase (Dyadchenko and Ktiatuntseva, 1955, 1956; Tresise, 1960), amphiboles (Glinka, 1896), pyroxenes (Gooch, 1876, quoted by Hadding, 1932; Glinka, 1896; Takahashi, 1939), and olivine (Gooch, 1876 quoted by Hadding, 1932). Fragments of silicate rocks such as andesite (Hendricks and Ross, 1941), phonolite (Ross et al., 1929), rhyolite (Ojakangas and Keller, 1964) and sandstone (Tresise, 1960) may also be replaced by mineral glauconite. It is possible that some of the above observations may not be due to actual replacements but to alteration of detrital clay as coatings of or infillings in cleavages and other cracks. The work of Ojakangas and Keller (1964) and Dapples (1967), however, provides ample evidence of actual replacement. Mineral glauconite may also

GLAUCONITE

421

replace phosphatic material (Hadding, 1932; Dyadchenko and Khatuntseva, 1955, 1956; Ojakangas and Keller, 1964) and be itself replaced by carbonate in some instances (e.g., Wermund, 1961; Ojakangas and Keller, 1964). Some examples of secondary glauconite have an unusual mineralogy (Tyler and Bailey, 1961; Wise and Eugster, 1964; Young et al., 1968).

The jbrrnation of morphological glauconite The typical morphology of glauconite may be due to: (1) Alteration of internal fillings to organic debris. (2) Conversion of remains in faecal pellets. (3) Transformation of biotite. (4) Agglomeration of clay-sized material. (5) Precipitation on or alteration of mineral surfaces to give pigmentary glauconite. (6) Replacement processes. These processes are not mutually exclusive. Many workers on glauconite (e.g., Bailey, 1856; Ehrenberg, 1856; Clark, 1894; Bagg, 1898, 1909; Collet, 1908;Mansfield, 1919b; Boswell, 1924; Dryden, 1931; Needham, 1934; Plumley and Graves, 1953; Burst, 1958a; Wermund, 1961; Pratt, 1962b; Ehlmann et al., 1963; Rao, 1964; Drobnyk, 1965; Seed, 1965; Kharkwal, 1966; Bell and Goodell, 1967) have noted that the shapes of many of the glauconite pellets are reminiscent of Foraminifera, and some have postulated that the production of glauconite is by the alteration of infillings in empty foram tests. Other workers (e.g., Von Giimbel, 1886; Prather, 1905; Schneider, 1927; Hadding, 1932; Cayeux, 1938) found that pellets with foraminiferal shapes were rare or absent in the glauconites which they studied, and therefore did not support this hypothesis. As discussed above, empty chambers in Foraminifera and other shell debris can provide microenvironments suitable for the m'ineralogical conversion of detrital clay to glauconite and the resulting material may be expected to reflect this origin in its morphology, at least initially. The relative importance of this mode of origin remains, however, an open question. It may be that many glauconite grains are formed in this way and that the general absence of clearly recognisable casts may be due to subsequent attrition. Conversely, it may be that this mode of origin is in general rather unusual and occurs only in isolated cases. Shapes other than Foraminifera, e.g., Radiolaria (Murray and Mackintosh, 1968), pieces of Echinoderma (Bailey, 1856; Berry, 1940; Wermund, 1961; Pratt, 1962a; Seed, 1965), Mollusca (Bailey, 1856; Mansfield, 1919b; Hadding, 1932; Wermund, 1961 ; Pratt, 1962a), Bryozoa (Bailey, 1856), and Porifera (Hadding, 1932; Pratt, 1962a) have been noted and may be internal casts or may be replacement structures. The origin of morphological glaucolaite as altered faecal pellets, giving ellipsoidal or capsule shaped grains was first suggested by Takahashi and Yagi (1929) and Takahashi (1939). From these and similar observations made on modern occurrences of glauconite (Galliher, 1932; Wetzel, 1937; Moore, 1939; Pratt, 1962a, 1962b; Porrenga, 1963, 1966, 1967; Murray and Mackintosh, 1968) it seems reasonable to ascribe faecal pellet origins

422

S.G. McRAE

to capsule-shaped grains from ancient sedimealts (Wermund, 1961; K6ster, 1964; Parry and Reeves, 1966). Passage of biotite flakes through the gut of marine organisms during their conversion to glauconite (Galliher, 1935a, 1935b, 1936, 1939) may produce a bulbous grain of glauconite showing relatively little of its original morphology, although distinctive vermicular grains obviously dependent on the morphology of the parent biotite are often found. It may be expected that alteration of clay galls or shale flakes might also give vermicular grains but it seems more probable that grains with discoidal (Wermund, 1961) or other morphologies (Lebauer, 1964) result. Glauconite grains may be formed by the accretion of material of colloidal dimensions, although workers who have suggested this mode of origin (e.g., Teodorovitch, 1961; Odin, 1971) are rather vague as to how the accretions form, suggesting simply "current action" or similar process. Bailey and Atherton (1969) have provided an attractive hypothesis which also explains the common occurrence of phosphatic material in glauconite pellets. These authors have postulated that there is an initial flocculation of proglauconite (or anteglauconite) and colloidal phosphate as chance encouters in a dispersed suspension and that these primary floccules then serve as centres of accretionary flocculation which sink to the sea floor when of sufficient size (hence the observed minimum size of glauconite grains). On the sea bed, the grains would continue to grow while wave- or current-induced rolling of the accretionary masses during growth would result in a pellet morphology and, by reasons of abrasion, a maxium grain size would be reached. Final maturation of the grain is believed to involve the mineralogical changes as suggested by Burst (1958a, 1958b) and Hower (1961) together with expulsion of water. There is little doubt that the final morphology shown by glauconite grains is dependent not only on their mode of origin but also to the attrition they have undergone. The more delicate vermicular pellets or fossil casts are unlikely to survive much transport and will almost certainly be of local derivation. Spheroidal or ovoidal pellets are frequently of detrital origin. Attrition may also be responsible for pigmentary glauconite (Collet, 1908; Carroll, 1939) although direct precipitation or replacement is more likely (Hadding, 1932). Glauconite morphology may also be modified by compaction of the host sediment (Carozzi, 1960), by accretion to give mammillated or lobate pellets and by inclusion within subsequent composite gtauconite grains.

The experimental synthesis of glauconite No attempts seem to have been made to verify the theories of Burst (1958a, 1958b) and Hower (1961) by experimental synthesis of mineral glauconite. Attempts to verify earlier theories were largely unsuccessful (Calderon et al., 1894; Chaves and del Pulgar, 1896; Collet, 1908; Caspari, 1910; Murray and Hjort, 1912, p.189; Takahashi, 1939) although Birdsall (1951) has claimed to have formed poorly crystalline glauconite by autoclaving co-precipitated gels. No attempts to produce morphological glauconite grains seem to have been made.

GLAUCONITE

423

THE VALUE AND USES OF GLAUCONITE It is proposed in this section to discuss the value of glauconite in geological studies, dealing in particular with its common use in the age dating of sedimentary rocks. The commercial uses to which glauconite may be put will also be reviewed although it must be pointed out that glauconite has little commercial value at the present day. Its greatest use has probably been in agriculture and its value in soil fertility will be discussed.

Glauconite in geological studies The value of glauconite as an indicator of particular palaeo-environments has been frequently stressed, although the rather wide range of conditions possible for its genesis (Cloud, 1955) and its common non-authigenic occurrence would seem to make it less useful than commonly supposed. Examples of the possible use of glauconite in palaeoenvironment studies have been given by Lochman (1957), Burst (1958a) and Wermund (1961), including discussions of the disadvantages of glauconite for such studies. Glauconitic sediments may also be used as correlation tools (Burst, 1958a) and even the mere presence or absence of glauconite may be sufficient for correlation (Triplehorn, 1966a). Some workers (Triplehorn, 1961; Wermund, 1961)have been unable, however, to use observed variations in mineralogy or morphology of glauconite to aid correlations. Triplehorn (1966b) has suggested that the common occurrence of concentrations of glauconite at stillstands or marine transgressions may assist in tracing diastems or unconformities. He has also suggested some possible geochemical applications if detailed differences in chemical composition could be related to differences in the chemical and physical conditions prevailing at the time of origin. At present, however, it seems that the view of Wermund (1961) that glauconite "is neither a satisfactory indicator of broad sedimentary environments nor a creditable parameter for stratigraphic correlation" should prevail. In recent years, a great deal of interest in glauconite has centred on its use for K - A r and R b - S r dating, and hence establishing broad stratigraphic correlations on the basis of similar ages. The literature dealing with this aspect of glauconite is voluminous, and no attempt will be made to review it comprehensively here. Good reviews are given by Curtis and Reynolds (1958), Lipson (1958), Kulp (1959), Rubinshtein et al. (1959), Evernden et al. (1960), Hurley et al. (1960), Evernden et al. (1961), Gulbrandsen et al. (1963), Kazakov (1964), Degens (1965), McDougall et al. (1965), Hurley (1966) and Dalrymple and Lamphere (1970) for K - A r dating; and by Cormier (1956), Cormier et al. (1956), Herzog et al., (1958) and Hurley et al. (1960) for R b - S r dating. It is perhaps worth recording the limitations to which the use of glauconite in K - A t dating are subject: (1) The glauconite grains must be syngenetic with the deposit. Glauconite may occur as detrital grains in relatively younger host rocks (e.g., Allen et al., 1964; Bodelle et al., 1969) or as a secondary mineral in relatively older rocks (e.g., Tyler and Bailey, 1961). It

424

S.G. McRAE

seems advisable (Wermund, 1961) to precede age determinations of glauconite with a microscopic examination of the rock to establish the relative age relationships of glauconite and host. (2) There should be minimal contamination of the glauconite being used for dating with detrital material (Curtis and Reynolds, 1958; Lipson, 1958; Hurley et al., 1960; Kazakov, 1964; Curtis, 1966). (3) Gains or losses of potassium by the glauconite, other than by conversion to argon, should be negligible or should be predictable. Post-diagenetic gains of potassium have been noted by Evernden et al. (1961), Hower (1961), Hurley (1961) and Hower et al. (1963). Losses of potassium by weathering may also occur (Evernden et al., 1961). (4) Losses of argon, either during the history of the glauconite or during sample pretreatment, should be small or should be predictable. Considerable work has been carried out on the diffusion of argon in glauconite with a view to predicting losses (Amirkhanov et al., 1958; Curtis and Reynolds, 1958; Kazakov and Polevaya, 1958; Hurley et al., 1959, 1960; Evernden et al., 1960, 1961; Murina and Sprintsson, 1961; Hower et al., 1963; Sardarov, 1963; Kazakov, 1964; Fechtig and Kalbitzer, 1966; Kazakov and Teplinsky, 1966; Nikolayeva et al., 1969). From these works it seems that glauconite loses Ar more readily than other micas and that the losses depend on the degree of crystallinity of the glauconite, its water content, and on the temperature at which the glauconite is held. Temperatures as low as 150 °C, if sufficiently prolonged, e.g., by deep burial, may cause substantial losses of argon. (5) Gains of argon may also occur (Curtis and Reynolds, 1958; Kazakov and Teplinsky, 1966) but are thought to be rare. The above factors usually combine so that glauconite gives K - A r ages which are anomalously young (Hurley et al., 1960; Hurley, 1961, 1966; Gulbrandsen et al., 1963; Baadsgaard and Dodson, 1964; Dodson et al., 1964; Kazakov, 1964; Dalrymple and Lamphere,1970). Fortunately the lowering in age is somewhat consistent at about 10-20 % (Hurley et al., 1960). Anomalously high glauconite K - A t ages may also occur (Kazakov, 1964). In spite of these limitations, glauco.nite is the only mineral that can be used routinely to date a sedimentary rock directly (Dalrymple and Lamphere, 1970). It may be usefully employed to date sediments of ages greater than 1" 106 years and may extend well into the Precambrian (Polevaya and Kazakov, 1960; Polevaya et al., 1961; Vinogradov and Tugarinov, 1962; Gulbrandsen et al., 1963; Webb et al., 1963;Kazakov et al., 1965; McDougall et a1.,.1965; Tugarinov et al., 1965; Hurley, 1966). The oldest glauconite so far dated, with an age of 1600" 106 years B.P., is from northern Australia (Webb et a1.,1963; McDougall et al., 1965). Rb-Sr dating of glauconite has received less attention. The ions involved are associated with the expandable layers of glauconite (Hower, 1961 ; Hurley, 1961) and are thus more susceptible to gains and losses than structural potassium and argon. For glauconite it would seem that Rb-Sr dating is inherently less satisfactory than K - A r dating.

GLAUCONITE

425

Commercial uses of glauconite

Glauconite has been suggested, usually in times of war (Ashley, 1917; Mansfield, 1920; Anon, 1941 ; Oakley, 1943), as a source of potassium for industrial use and Buck (1918), Charlton ( 1918), Holmes (1919), Shreve ( 1921 ) and Mansfield ( 1922) g'ive details of some extraction processes. Glauconite has also been used as a water-softener (Simpson, 1934; Ingleson and Harrison, 1941; Ingleson and Sullivan, 1941; Rabinovich, 1942; Oakley, 1943; Skow, 1961). Water-softening by glauconite may also occur naturally; waters derived from wells sunk into the Greensand, for example, are much softer than those from wells in the overlying chalk (Milner, 1962, p.589). Recently Schnepfe et al. (1964) have investigated its possible use as a "scavenger" for radioactive isotope wastes. Other suggested uses have been as a colouring agent in glass making (Cook, 1868; Oakley, 1943) and the paint industry (Simpson, 1934; Pocock, 1942; Oakley, 1943), as a sorbent for oils, and to "humanise" or decalcify bovine milk (Simpson, 1934). It may also be employed as a raw material in the manufacture of slag-wool, an insulating material (Carroll-Porcynski, 1958). Glauconite in agriculture

Glauconite or sediments rich in glauconite have been used, or advocated, as fertilisers, particularly in the N.E. United States (Cook, 1868; Bagg, 1898; Skeen, 1925; Skow, 1961) but also in South Carolina (Sloan, 1908), Virginia (Skeen, 1925), Wisconsin (Twenhofel, 1936), Southern England, France and Belgium (Penrose, 1888; Simpson, 1934; Oakley, 1943), Russia (Fedorovsky, 1924), Japan (Swanson, 1949), Australia (Simpson, 1934) and New Zealand (Anonymous, 1941). Glauconite seems to act mainly as a slow-acting source of potassium (Cook, 1868; Lipman and Blair, 1917; True and Geise, 1918; Skeen, 1925; Twenhofel, 1936; Graham and Turley, 1947; McRae, 1971), magnesium (Jurkowska, 1956; McRae, 1971) and occasionally iron (Spencer, 1951). Benefits from additions of greensands or greensand marls may also be due to the presence of adventitious phosphate (Cook, 1868; Topley, 1875; Penrose, 1888; Lipman and Blair, 1917; Joffe, 1923) or calcium carbonate (e.g., McCall and Smith, 1920), or to the improvement of soil texture (Mansfield, 1922). Glauconitic materials are no longer used as fertilisers, but glauconite may contribute to soil fertility where it occurs naturally in the soil. Observations that soils which contain indigenous glauconite are usually very fertile have been made in many parts of the world, e.g., the United States (Morse, 1819 in Cook, 1868; Skeen, 1925; Eckel, 1938), the West Indies (Ahmad et al., 1968), Australia (Simpson, 1934), France (Cayeux, 1905), Belgium (Schreiber, 1907) and England (Topley, 1872, 1875; Oakley, 1943). Although few studies have been made on the actual benefits conferred on soils by indigenous glaueonite, the most likely effect is to improve or modify the potassium status of the soils (Schreiber, 1907; Chaminade, 1936; Joffe and Kolodny, 1937; Levine, 1939; Joffe and Levine, 1947a, 1947b; Levine and Joffe, 1947a, 1947b; Ahmad et al., 1968; McRae,

426

S.G. McRAE

1971). Other beneficial effects of the presence of glauconite in soils may be an enhanced cation exchange capacity (Tedrow, 1966), improved water-holding capacity (Tedrow, 1966) and high available phosphate (Ahmad et al., 1968; McRae, 1971), although this last effect may not be due to the glauconite per se. THE ALTERATIONOF GLAUCONITE

Metamorphism Glauconite survives deep burial but is apparently unstable under conditions of folding (Dapples, 1967). Its reaction to more intense metamorphism is unknown.

Natural weathering It seems that, in general, acid weathering of mineral glauconite, especially in soils, involves loss of potassium with the production of expandable layers of either a vermiculitic or montmorillonitic nature (Cole, 1940-1941, 1943; Nagelschmidt, 1944; Wurman, 1960; Hutcheson and Haney, 1963; Ahmad et al., 1968). Further weathering to a kaolinitic mineral may also occur (Hutcheson and Haney, 1963; Carson and Kunze, 1967). During these transformations, structural iron may oxidise to produce a brown colour (De Coninck and Laruelle, 1960, 1964) often present only as a rim to each grain. Some iron may be liberated to accumulate as hydrated iron oxides usually referred to limonite (e.g., Cayeux, 1905; Gildersleeve, 1932; Needham, 1934; Monroe, 1947; Wood, 1956) and also tending to occur as a rim to each grain. Only rarely is there complete alteration to iron oxide (Wolff, 1967). Occasionally other iron-rich weathering products of glauconite have been suggested, e.g., siderite (Jans and Van Calster, 1962), haematite and magnetite (Prather, 1905) and jarosite (Tyler, 1936; Bentor and Kastner, 1965). The rim of oxidised glauconite or iron oxides may serve to protect the grain against further weathering but usually in situ acid weathering of glauconite disrupts the pellet morphology (Wurman, 1960; Cloos et al., 1961; De Coninck and Laurelle, 1960, 1964; Moreau, 1965). Glauconite" is easily destroyed during fluvial transport (Wermund, 1964) but may survive several stages of re-working if the conditions remain alkaline, e.g., in the calcareous Cretaceous sediments of N.E. United States (Fairbridge, 1967). In weathering stability it may be regarded as similar to biotite (Jackson and Sherman, 1953).

Artificial weathering Alteration of glauconite may also be produced, sometimes inadvertently, by a variety of laboratory procedures. Grinding to a fine size causes release of potassium from glauconite (Chaminade and Drouneau, 1936; Rouse and Bertramson, 1949; McRae, 1971) and also oxidation of ferrous iron (Bien and Goldberg, 1956). Shaking in water causes disaggregation of grains (Mansfield, 1920, 1922; Hutton and Seelye, 1941) and ultrasonic

GLAUCONITE

427

vibrations in water can cause complete disruption of the grains. The latter treatment is therefore a very useful way to prepare glauconite in a fine grained size for mineralogical examination (McRae and Lambert, 1968). Alternate freezing and thawing releases potassium from glauconite (Graham and Turley, 1947), as does heating to high temperatures (Legg and Axley, 1958). Acids severely attack glauconite, especially if concentrated (Hoskins, 1895; Fraps, 1912; Mansfield, 1922; Leaf, 1958; Stahlberg, 1960; Dell'anna, 1964). It has been noted (Hutton and Seelye, 1941; Cloos et al., 1960, 1961; Gastuche, 1960; Touillaux et al., 1960; Zabelin, 1962) that solution of glauconite by acids selectively removes certain components. This would invalidate any chemical analyses carried out on material pretreated With strong acids (e.g., Schneider, 1927). Acid attack also affects the cation exchange characteristics of glauconite (Schnepfe et al., 1964). Selective removal of potassium from glauconite can be effected by ion-exchange resins, especially if H + saturated (Arnold, 1958) and by the use of various salt solutions (Andr6, 1916; Gruner, 1935; Libor,1962a, 1962b, 1964). In these studies and those involving acids (Leaf, 1958) glauconite behaves similarly to other dioctahedral minerals such as muscovite and illite which release potassium much less readily than trioctahedral minerals such as biotite. REFERENCES Adams, J.K., 1963. Petrology and origin of the lower Tertiary formations of New Jersey. J. Sediment. PetroL, 33:587-603. Ahmad, N., Jones, R.L. and Beavers, A.H., 1968. Genesis, mineralogy, and related properties of West Indian soils. I. Montserrat Series derived from glauconitic sandstone, Central Trinidad. J. Soil Sci., 19:1-8. Aldinger, H., 1965. Uber den Einfluss yon Meeresspiegelschwankungen auf Flachwassersedimente in Schw~ibischen Jura. Tschermaks Mineral. Petrogr. Mit., 10:61-68. Aleixandre Ferrandis, V. and Gonzalez Pefia, J.M., 1960. Contribuci6n al estudio de la glauconite. An. Edafol. AgrobioL, 19:615-634. Alexander, A.E., 1934. A petrographic and petrologic study of some continental shelf sediments. J. Sediment. PetroL, 4: 12-22. Alexiades, C.A. and Jackson, M.L., 1965. Quantitative determination of vermiculite in soils. Proc. Soil Sci Soe. Am., 29:522-527. Alexiades, C.A. and Jackson, M.L., 1966. Quantitave clay mineralogical analysis of soils and sediments. Clays Clay Miner., 14:35-52. Allen, J.R.L., 1965. Late Quaternary Niger Delta, and adjacent areas: sedimentary environments and lithofaeies. Bull. Am. Assoc. Pet. Geologists, 49:547-600. Allen, P., Dodson, M.H. and Rex, D.C., 1964. Potassium-argon dates and the origin of Wealden giauconites. Nature, 202:585-586. Allen, V.T., 1937. A study of Missouri glauconite. Am. Mineralogist, 22:1180-1183. Amirkhanov, K.I., Brandt, S.B., Bartnitzky, E.N., Gurvich, V.S. and Gasanov, S.A., 1958. On the preservation of radiogenic argon in glauconites. DokL Akad. Nauk. S.S.S.R., 118:328-330 (in Russian). Amstutz, G.C. and Bubenicek, L., 1967. Diagenesis in sedimentary mineral deposits. In: G. Larsen and G.V. Chilingar (Editors), Diagenesis in Sediments. Elsevier, Amsterdam, pp. 417-475. Andr6, G., 1916. D~placement de la potasse et de l'acide phosphorique contenus dans certaines roches par quelques substances employ6es comme engxais. C.R- Acad. ScL, 162:133-136.

428

S.G. McRAE

Anonymous, 1941. Potash supplies in wartime. N.Z.J. Agric., 62:171 - 172. Arnold, P.W., 1958. Potassium uptake by cation-exchange resins from soils and minerals. Nature, 182:1594-1595. Ashley, G.H., 1917. Notes on the greensand deposits of the eastern United States. Bull. U.S. Geol. Surv., 660: 27-49. Atanasov, G., 1962. Glauconites from the Jurassic in Bulgaria. God. Sofii. Univ., Biol.-Geol. Geogr. Fak., 55:142 157 (in Bulgarian, with English abstract). Atanasov, G., Stefanov, D., Trashliev, S. and Goranov, A., 1964. Bentonitic clays from near Dimitrovgrad. God. Sofii. Univ., Geol. Geogr. Fak., 5 7 : 7 5 - 9 1 (in Bulgarian, with English abstract). Baadsgaard, H. and Dodson, M.H., 1964. Potassium-argon ages of sedimentary and pyroclastic rocks. Q.J. Geol. Soc. Lond., 120S:119 127. Bagg, R.M., 1898. The Cretaceous Foraminifera of New Jersey. Bull, U.S. Geol. Surv., 88:11-89. Bagg, R.M., 1909. Casts of Foraminifera in the Carboniferous of Illinois. Bull. Ill. State Geol. Surv., 14:263-271. Bailey, J.W., 1856. On the origin of greensand and its formation in the oceans of the present epoch. Proc. Boston Soc. Nat. Hist.,.5:364-368. Bailey, R.J. and Atherton, M.P., 1969. The petrology of a glauconite sandy chalk. J. Sediment. Petrol.; 39: 1420-1431. Balashov, Yu.A. and Kazakov, G.A., 1968. Source of the rare earths in Pacific Ocean glauconite. Dokl. Akad. Nauk. S.S.S.R., 179: 181-183. Ballance, P.F., 1964. Streaked-out mud ripples below Miocene turbidites, Puriri formation, New Zealand. J. Sediment. Petrol., 34:91-101. Barackman, M.A., 1964. A Study of the Mineral Glauconite in Apalachicola Bay, Florida. Thesis, Florida State Univ., Tallahassee, Fla., 44 pp. Bell, D.L. and Goodell, H.G., 1967. A comparative study of glauconite and the associated clay fraction in modern marine sediments. Sedimentology, 9: 169-202. Bentor, Y.K. and Kastner, M., 1965. Notes on the mineralogy and origin of glauconite. J. Sediment. Petrol., 35:155 166. Bentor, Y.K., Bodenheimer, W. and Heller, L., 1963. A reconnaissance survey of the relationship between clay mineralogy and geological environment in the Negev (Southern Israel). J. Sediment. Petrol., 33: 8 7 4 - 903. Berry, C.T., 1940. Glauconite pseudomorphs after Ophiuran plates. Science, 91:449. Berthois, L. and Dangeard, L., 1939. La glauconie dans les s~diments n~ritiques de la M6diterran~e. Geol. Meere Binnengewiiss. , 3: 5 3 2 - 541. Bien, G.S. and Goldberg, E.D., 1956. Polarographic determination of ferrous and ferric iron in refractory minerals. Anal. Chem., 28:97-98. Birdsall, P., 1951. Recherches sur les conditions de formation des aluminio-silicates ferreux d'origine secondaire. C.R. Acad. Sci., 233:1371-1372. Bodelle, J., Lay, C. and Parfenoff,. A., 1969. Age des glauconies cr~tac~es du sud-est de la France (2: ValiSe de l'Est~ron Alpes-Maritimes). R6sultats pr~liminaires de la m~thode potassium-argon. C.R. Acad. ScL, 268D:1576-1579. Borchert, H., 1960a. Genesis of marine sedimentary iron ores. Trans. Inst. Min. Metall., 6 9 : 2 6 1 - 279. Borchert, H., 1960b. Geosynklinale Lagerstfitten, was dazu gehiSrt und nicht dazu gehiSrt, sowie deren Beziehungen zu Geotektonik und, Magmatismus. Freiberg. Forsehungsh., C79:8-61. Borchert, H., 1964. Uber Faziestypen yon marinen Eisenerzlagerst~itten. Ber. Geol. Ges. D.D.R., 9:163-193. Borchert, H. and Braun, H., 1963. Zum Chemismus yon drei Glaukonittypen. Chem. Erde, 23: 82-90. Boswell, P.G.H., 1924. The petrography of the sands of the Upper Lias and Lower Inferior Oolite in the West of England. Geol. Mag., 61 : 2 4 6 - 264. Braun, H., 1962. Zur Entstehung der marin-sedimentaeren Eisenerze. Z. Erzbergb. Metalhiitten wes., 15:613-623. Brongniart, A., 1823 M~moire sur les Terrains de S~diments Sup~rieurs Calear~o-Trapp~ens du Vicentin. Levrault, Paris, 85 pp. Brown, T.C., 1914. Origin of oolites and the oolitic texture in rocks. Geol. Soc. Am. Bull., 25:745-780.

GLAUCON1TE

429

Buck, E.C., 1918. Bibliography on the extraction of potash from complex mineral silicates. Metall. Chem. Eng., 18:33-95. Burst, J.F., 1958a. "Glauconite" pellets: their mineral nature and applications to stratigraphic interpretations. Bull. Am. Assoc. Pet. Geologists, 42:310 -327. Burst, J., 1958b. Mineral heterogeneity in "glauconite" pellets. Am. Mineralogist, 43:481 497. Burt, F.A., 1932, Formative processes in concretions formed about fossils as nuclei. J. Sediment. Petrol., 2:38-45. Caill~re, S. and Giresse, P., 1966. Etude min6ralogiquc de diverses "glauconies' actuelles Nouvelle contribution h la gen6se des minerals de fer s6dimentaires. C.R. A c a d Sci, 263D: 1804- 1907. Caill6re, S. and Lamboy, M., 1970. Etude min6ralogique de la glauconite du plateau continental au nord-ouest de l'Espagne. C.R. Acad. Sci., 270D:2057-2060. Calderon, D.S., Chaves, D.F. and Del Pulgar, P., 1894. Contributions al estudio de la glauconite. An. Soc. Esp. Hist. Nat., 23:8. Carozzi, A., 1951, Glauconitisation de la biotite dans le Crdtac6 moyen des chaines subalpines et du Jura. Cab. G6ol. Thoirv, 12:101 103. Carozzi, A.V., 1958. Micro-mechanisms of sedimentation in the epicontinentat environment. J. Sediment. Petrol., 28: 133-150. Carozzi, A.V., 1960. Microscopic Sedimentary Petrology. Wiley, New York, N.Y., 485 pp. Carroll, D., 1939. A note on the mineralogy of the Gingin Chalk, Western Australia. Proc. Geol. Assoc., 5 0 : 2 2 7 - 2 3 4 . Carroll, D., 1941. Heavy residues from some Upper Cretaceous sediments at Gingin, Western Australia. J. Sediment. Petrol., 11:85-91. Carroll, D., 1959. Ion exchange in clays and other minerals. Geol. Soc. Am. Bull., 70:749 780. Carroll-Porcynski, C.Z., 1958. Inorganic Fibres. National Trade Press, London, 353 pp. Carson, C.D. and Kunze, G.W., 1967. Red Soils of East Texas developed in glauconitic sediments. Soil Sci., 104:181-190. Caspari, W.A., 1910. Chemistry of submarine glauconite. Proc. R. Soc. Edinb., 30:364 373. Cayeux, L., 1897. Contribution a l'6tude des terrains s6dimentaires. MOrn. Soc. Geol. Nord (Lille), 4: 163-164. Cayeux, L., 1905. Constitution de la terre arable. Du r61e de l'analyse min6ralogique dans l'analyse des terres. Ann. Soc. Geol. Nord (Lille), 4: 134-162. Cayeux, L., 1906. Gen6se d'un minerai de fer par d6composition de la glauconie. C.R. Acad. Sci, 142:895 897. Cayeux, L., 1916. introduction h l'6tude p6trographique des roches s6dimentaires. MOrn. Carte G6ol. 'Ft. Cayeux, L., 1924. Les Minerais de Fer Oolithique de France. Pt. 2. Minerais de Fer Secondaires. B6ranger, Paris, 1051 pp. Cayeux, L., 1932. Les mani6res d'6tre de la glauconie en milieu calcaire. C.R. Acad. Sei., 195: 1050-1052. Cayeux, L., 1935, Les Roches Sddimentaires de France - Roches Carbonat6es. Masson. Paris, 447 pp. Cayeux, L., 1938. Role des foraminif6res et du carbonate de chaux darts la fixation de certains substances min6rales. C.R. Acad. Sc~, 206: 81-85. Chaminade, R., 1936. Sur le passage ~ l'6tat non exchangeable du potassium dans les sols. C.R. Acad. Sci., 203:682--684. Chaminade, R. and Drouneau, G., 1936. Recherches sur la m6canique chimique des cations exchangeables. Ann. Agron., 6 : 6 7 7 - 6 9 1 . Charlton, H.W., 1918. Recovery of potash from greensand. Am. J. ScL, 45: 142. Chaves, D.F. and Del Pulgar, P., 1896. Sintesis de silicatos de hierro. An. Soc. Esp. Hist. Nat., 24: 157-158. Chilingar, G.V., 1956. Joint occurrence of glauconite and chloxite in sedimentary rocks: a review. Bull. Am. Assoc. Pet. Geologists, 4 0 : 3 9 4 - 3 9 8 . Clark, W.B., 1894. Origin and classification of the greensands of New Jersey. J. Geol., 2:161-177. Clarke, F.W., 1903. Composition of glauconite and greenalite. Monogr. U.S. Geol. S u ~ . , 43:243 247. Clarke, F.W., 1915. Analyses of rocks and minerals from the laboratory of the United States Geological Survey. Bull. U.S. Geol. Surv., 5 9 1 : 3 7 6 pp.

430

S.G. McRAE

Clarke, F.W., 1916. The data of geochemistry (3rd edition). Bull U.S. Geol. Surv., 616:821 pp. Clarke, F.W., 1920. The data of geochemistry (4 th edition). Bull, U.S. Geol. Surv., 695:832 pp. Clarke, F.W., 1924. The data of geochemistry (5th edition). Bull. U.S. Geol. Surv., 770:841 pp. Clarke, R.H., 1969. Quaternary sediments off south-east Devon. Q. J. Geol. Soc. Lond., 125:277-318. Cloos, P., Gastuche, M.C. and Croegaert, M., 1960. Cin6tique de la destruction de la glauconite par l'acidc hydrochlorique, 6tude preliminaire. Int. Geol. Congr. 21st, Copenhagen, 1960, Rep. Sess. Norden, 24:35-50. Cloos, P., Fripiat, J.J. and Vielvoye, L., 1961. Mineralogical and chemical characteristics of a glauconitic soil of the Hageland region (Belgium). Soil Sci., 9 1 : 5 5 - 6 5 . Cloud, P.E. and Barnes, V.E., 1948. Paleoecology of the Early Ordovician sea in Central Texas. In: Committee on a Treatise on Marine Ecology and Paleoecology, Rep. Nat. Resour. Counc. U.S., Washington, D.C., 8: 29-83. Cloud, P.E., 1955. Physical limits of glauconite formation. Bull, Am. Assoc. Pet. Geologists, 39:484 492. Cole, W.F., 1940-1941. X-ray analysis (by the powder method) and microscopic examination of the products of weathering of the Gingin Uppper Greensand. J. Proc. R. Soc. West. Aust. 27:229-243. Cole, W.F., 1943. X-ray analysis of some soil colloids from Gingin, Western Australia. Soil Sci., 56:153-171. Collet, L.W., 1908. LesD~potsMarin~ Doin, Paris, 320 pp. Collet, L.W. and Lee, G.W., 1906a` Recherches sur la glauconie. Proc. R. Soc. Edinb., 26: 238-278. Collet, L.W. and Lee, G.W., 1906b. Sur la composition chimique de la glauconie. C.R. Acad. SoL, 142:999-1001. Conway, E.J., 1942. Mean geochemical data in relation to oceanic evolution. Proc. R. lr. Acad., 48B: 119-159. Cook, G.H., 1868. Geology of New Jersey. Board of Managers, State Geological Survey, Newark, N.J., 900 pp. Cormier, R.F., 1956. Rubidium-strontium ages of glauconite. Geol. Soc. Am. Bull., 67:1812. Cormier, R.F., Herzog, L.F., Pinson, W.H. and Hurley, P.M., 1956. Rubidium-strontium agc determinations on the mineral glauconite. Geol. Soc. Am. Bull,, 67: 1681-1682. Correns, C.W., 1937. Die Sedimente des equatorialen Atlantischen Ozeans. Deutsche Atlantische Expedition "Meteor", 1925-1927, 3: 2 7 8 - 298. Curtis, G.H., 1966. The problem of contamination in obtaining accurate dates of young geologic rocks. In: O.A. Schaeffer and J. Z~ihringer (Editors), Potassium Argon Dating. Springer, Berlin, pp. 151-162. Curtis, G.H. and Reynolds, J.H., 1958. Notes on the potassium argon dating of sedimentary rocks. Geol. Soc. Am. Bull., 69:151-160. Dallwitz, W.B., 1952. A note on glauconitic minerals of low refractive index from Lower Tertiary beds in South Australia and Victoria. In: M.F. Glaessner and E.A. Rudd (Editors), Sir Douglas Mawson Anniversary Volume: Univ. Adelaide, Adelaide, S.A., pp.55-62. Dalrymple, G.B. and Lamphere, M.A., 1970. Potassium-Argon Dating. Freeman, London, 258 pp. Dangeard, L., 1940. Sur diverses mani~res d'etre de la glauconie dans la couche verte du Bajocien normand. C.R. Acad. Sci., 211:264-265. Dapples, E.C., 1967. Diagenesis of sandstones. In: G. Larsen and G.V. Chilingar (Editors), Diagenesis in Sediment~ Elsevier, Amsterdam, pp. 91 - 125. De Bruyn, C.M.A, and Van der Marel, H.W., 1954. Mineralogical analysis of soil clays. Part II: Examples of mineral analysis by X-ray diffraction and differential thermal analysis. Geol. Mijnbouw, 16:407-428. De Coninck, F. and Laruelle, J., 1960. Influence of mineralogical composition on podzol formation. Trang Int. Congr. Soil Sci., 7th, Madison, Wisc., 1960, 4:157-164. De Coninck, F. and Laruelle, J., 1964. Soil genesis in sandy glauconiferous soils of the Belgian Campine. Tran~ Int. Congr. Soil Sc£, 8th, Bucharest, 1964, 5:373-382. Degens, E.T., 1965. Geochemistry of Sediments: a Brief Survey, Prentice Hall, Englewood Cliffs, N. J., 342 pp.

GLAUCONITE

431

Dekeyser, W., 1956. Clay minerals research at Ghent University. Clays Clay Miner., 4:372 380. DeU'anna, L., 1964. La glauconite nei calcari cretacei della Pensinola Salentina. Period. Miner., 33:521 545. Dodson, M.H., Rex, D.C., Casey, R. and Alien, P., 1964. Glauconite dates from the Upper Jurassic and Lower Cretaceous. Q. J. GeoL Soc. Lond., 1205:145-158. Drobnyk, J.W., 1965. Petrology of the Paleocene-Eocene Aquia formation of Virginia, Maryland, and Delaware. J. Sediment. PetroL, 35:626-642. Dryden, A.L., 1931. Glauconite in fossil for~rniniferal shells. Science, 74: 17. Dyadchenko, M.G. and Khatuntseva, A.Y., 1955. On the problem of the genesis of glauconite. DokL Akad. Nauk. S.S.S.R., 101:151-153 (in Russian). Dyadchenko, M.G. and Khatuntseva, A.Y., 1956. Cases of formation of glauconitc in a continental environment. Zap. Vses. Mineral O-va, 8 5 : 4 9 - 5 7 (in Russian). Eckel, E.B., 1938. The brown ores of eastern Texas. Bull. U.S. GeoL Surv., 902:151 pp. Ehlmann, A.J., Hulings, N.C. and Glover, E.D., 1963. Stages of glauconite formation in modern foraminiferal sediments. J. Sediment. PetroL, 33: 87-96. Ehrenberg, C.G., 1856. Uber den Grtinsand und seine Erl~iuterung des organischen Lebens. Abh. Preus~ A kad. Wiss., 51:85-176. Evernden, J.F.. Curtis, G.H., Kistler, R.W. and Obradovich, J,, 1960. Argon diffusion in glauconite, microcline, sanidine, leucite and phlogopite, Am. J. ScL, 258:583-604. Evernden, J.F., Curtis, G.H., Obradovich, J. and Kistler, R., 1961. On the evaluation of glauconite and illite for dating sedimentary rocks by the potassium-argon method. Geochim. Cosmochim, Acta, 23:78-99. Fairbridge, R.W., 1967. Phases of diagenesis and authigenesis. In: G. Larsen and G.V. Chilingar (Editors), Diage~esis in Sediment~ Elsevier, Amsterdam, pp. 19-89. Fechtig, H. and Kalbitzer, S., 1966. The diffusion of argon in potassium-bearing solids. In: O.A. Schaeffer and J. Zahringer (Editors), Potassium-Argon Dating. Springer, Berlin, pp.68 107. Fedorovsky, N.M., 1924. Essay on Applied Mineralogy (Application of Minerals in Different Branches of Industry and Agriculture). Supreme Council of National Economy, Leningrad, 178 pp. (in Russian). Feldtmann, F.R. 1934. The glauconite deposits at Gingin, South-West Division. Prog. Rep. Geol. Surv. West. Aust. (1933), 6 - 7 . Foster, M.D., 1951. The importance of exchangeable magnesium and cation-exchange capacicy in the study of montmorillonitic clays. Am. Mineralogist, 36:717-730. Foster, M.D., 1956. Correlation of dioctahedral potassium micas on the basis of their charge relations. Bull U.S. GeoL Surv., 1036-D:57-67. Foster, M.D., 1960. Layer charge relations in the dioctahedral and trioctahedral micas. Am. Mineralogist, 45:383-398. Foster, M.D., 1969. Studies of celadonite and glauconite. U.S. GeoL Surv., Pro/'. Pap.. 614F i 17 pp. Frank-Kamensky, V.A., 1960. A crystallochemical classification of simple and interstratified clay minerals. Clay Miner. BulL, 4: 161-172. Fraps, G.S., 1912. The active potash of the soil and its relation to pot experiments. Bull Tex. Agric. Exp. Stn., 1945:39 pp. Galliher, E.W., 1932. Organic structures in sediments. J. Sediment. Petrol., 2:46-47. Galliher, E.W., 1935a. Geology of glauconite. Bull Am. Assoc. Pet. Geologists, 19:1569-1601. Galliher, E.W., 1935b. Glauconite genesis. GeoL Soc. Am. Bull., 46: 1351-1365. Galliher, E.W., 1936. Regional petrology of glauconite. Proc. GeoL Soc. Am., 345. Galliher, E.W., 1939. Biotite-glauconite transformation and associated minerals. In: P.D. Trask (Editor), Recent Marine Sediments. Am. Assoc. Pet. Geologists, Tulsa, Okla., pp.513 515. Gastuche, M.C., 1960. Acid dissolution techniques as related to the structure of clay minerals, oxides and gels. Trans. Int. Cong. SoilSci., 7th, Madison, Wisc., 1960, 4:499-501. Gierwielaniec, J. and Turnau-Morawska, M., 1961. Origin of glauconite in the Cretaceous transgressive sediments of Kudowa-Spalona region (Sudeten Mrs.). Arch. Miner., 25:261-275 (in Polish, with English abstract). Gildersleeve, B., 1932. Some stages in the disintegration of glauconite. Am. Mineralogist, 17:98-103.

432

S.G. McRAE

Gill, J.E., 1927. Origin of the Gunflint iron-bearing formation. Econ. Geol., 22:687 728. Giresse, P., 1965. Observations sur la prGsence de "glauconie" actuelle darts les sGdiments ferrugineux peu profonds du bassin gabonais. C.R. Acad. ScL, 260:5597 5600. Glass, H.D., 1953. Clay Mineralogy o f the Coastal Plain Formations o f New Jersey. Thesis, Columbia Univ., Palisades, N.Y., 203 pp. Glinka, K.D., 1896. Der Glauconite, sein Entstehung, sein chemiseher Bestand, und die A r t und Weise seiner Verwiterung. lnstitut Agronomique ae Novo. Alex Russie, St. Petersburg, 128 pp. (in Russian with German summary). Goldman, M.I., 1921a. Lithologic subsurface correlation in the "Bend" series of north-central Texas. U.S. Geol. Surv., Prof. Pap., 129:1 22. Goldman, M.I., 1921b. Association of glauconite with unconformities. Geol. Soc. Am. Bull., 32: 25. Goldman, M.I., 1922. Basal glauconite and phosphate beds. Science, 56:171 -173. Graham, E.R. and Turley, H.C., 1947. Soil development and plant nutrition: Ill The transfer of potassium from the non-available to the available form as reflected by the growth and composition of soy beans. Proc. Soil Sci. Soc. Am., 12:332-335. Grim, R.E. and Rowland, R.A., 1942. Differential thermal analysis of clay minerals and other hydrous minerals. Am. Mineralogist, 27:746 761. Grim, R.E., Bray, R.H. and Bradley, W.F., 1937. The mica in argillaceous sediments. Am. Mineralogist., 22:813-829. Grim, R.E., Bradley, W.F. and Brown, G., 195l. The mica clay ruiner'Ms. In G.W. Brindley (Editor), X-ray Identification and Crystal Structure o f Clay Minerals. Mineralogical Society, London, pp.138- 172. Gritzaenko, G.S., 1956. The application of the electron microscope to the study of fine dispersed minerals.Mineral. Sb., L 'roy 10:81-87 (in Russian). Gnmer, J.W., 1935. The structural relationship of glauconite and mica. Am. Mineralogist, 20:699-714. Gulbrandsen, R.A., Goldich, S.S. and Thomas, tt.H., 1963. Glauconite from the Precambrian Belt Series, Montana. Science, 140:390-391. Hadding, A., 1932. The Pre-Quaternary sedimentary rocks of Sweden. IV. Glauconite and glauconitic rocks. Acta Univ. Lund, 28: 1-175. Hallam, A. and Sellwood, B.W., 1968. Origin of fuller's earth in the Mesozoic of southern England. Nature, 220:1193-1195. Hallimond, A.F., 1922. On giauconite from the Greensand near Lewes, Sussex; the constitution of glauconite. Mineral. Mag., 19:330-333. Hallimond, A.F., 1925. Iron ores: Bedded ores of England and Wales - petrography and chemistry. Mem. Geol. Surv. Spec. Rep. Miner. Resour. G.B. 29:1 139. Hallimond, J., 1928. The formula of glauconite. Am. Mineralogist, 12:589-590. Hancock, J.M., 1961. The Cretaceous system in Northern Ireland. Q. J. Geol. Soc. Lond., 117:11 33. Hendricks, S.B. and Ross, C.S., 1941. Chemical composition and genesis of glauconite and celadonite.Am. Mineralogist, 26:683 708. Herzog, L.F., Pinson, W.H. and Cormier, R.F., 1958. Sediment age determination by Rb/Sr an~ysis of glauconite. Bull. Am. Assoc. Pet. Geologists, 42:717-733. Hodgson, E.A., 1962. Origin of glauconite in some sandstones of the Plantagenet beds, Cheyne Bay, Western Australia. J. Proc. R. Soc. West. Aust., 45:115 116. Hoebeke, F. and Dekeyser, W., 1954a. Evolution thermique de la glauconite. C.R. Aead. Sci., 238: 1143- 1144. Hoebeke, F. and Dekeyser, W., 1954b. L'oxydation de la glauconite .5 basse tempGrature. C.R. Acad. SoL, 238:2171-2173. Hoebeke, F. and Dekeyser, W., 1955. La glauconie. C.R. Rech. Inst. Encour. Rech. Sci. Ind. Agric., 14:103-121. Holmes, A., 1919. Non-German sources of potash. GeoL Mag., 56:251 254, 340-350. Honda, S., 1957. Glauconite from the Oga peninsula, Akita Prefecture. J. Miner. Soc. Jap., 3:124 129. Hoskins, A.P., 1895. On giauconite from Woodburn, Carrickfergus, Co. Antrim. Geol. Mag., 32: 317--321.

GLAUCONITE

433

Houbolt, J.J.H.C., 1957. Surface Sediments of the Persian Gulf near the Qatar Peninsula. Lounz, New York, N.Y., 113 pp. Hough, J.L., 1940. Sediments of Buzzards Bay, Massachusetts. J. Sediment. Petrol., 10:19-32. Hower, J., 1961. Some factors concerning the nature and origin of glauconite. Am. Mineralogist, 46:313-334. Hower, J., 1966. Order of mixed-layering in illite/montmorillonites. Clays Clay Miner., 15:63-74. Hower, J., Hurley, P.M. and Pinson, W.H., 1963. The dependence of K - A r age on the mineralogy of various particle size ranges in a shale. Geochim. Cosmochim. Acta, 27:405-411. Hummel, K., 1922. Entstehung eisenreicher Gesteine durch Halmyrolysis. Geol.Rundsch., 13:50 78. Hummel, K., 1923. Die Oxford-Tuffte der lnsel Buru und Ihre Fauna. Palaeontographica SuppL, IV: 122 123. Hurley, P.M., 1961. Glauconite as a possible means of measuring the age of sediments. Ann. N. Y. Acad. Sci., 91:294-297. Hurley, P.M., 1966. K - A r dating of sediments. In: O.A. Shaeffer and J. Z~ihringer (Editors), Potassium-Argon Dating. Springer, Berlin, pp. 134-151. Hurley, P.M., Pinson, W.H., Cormier, R.F. and Fairbairn, H.W., 1959. Age study of Lower Palcozoic glauconites. J. Geophys. Res., 64:1109. Hurley, P.M., Cormier, R.F., Hower, J., Fairbairn, H.W. and Pinson, W.H., 1960. Reliability of glauconite for age measurement by K - A r and Rb Sr methods. Bull. Am. Assoc. Pet. Geologists, 44: 1793-1808. Hutcheson, T.B. and Haney, D.C., 1963. Solum parent material relationships for three soils profiles developed from sedimentary deposits of Ordovician age. Proc. Soil Sci. Soc. Am., 27:441 445. Hutton, C.O., 1950. Studies of heavy detrital minerals. GeoL Soc. Am. Bull., 61:635-710. Hutton, C.O. and Seelye, F.T., 1941. Composition and properties of some New Zealand glauconites. Am. Mineralogist, 26:595 604. lchimura, T . , 1940. Some glauconitic rocks from Taiwan (Formosa). Mere. Fac. Sci. Agric. Taihoku Univ., 22:25-63 (in Japanese). lngleson, H. and Harrison, A., 1941. Effect of temperature on the exchange capacities of some base-exchange materials used in water softening. J. Soc. Chem. lnd. Lond., 60:87 92. lngleson, H. and Sullivan, W.H., 1941. The preparation of base-exchange materials from some British clays and minerals. J. Soc. Chem. Ind. Lond., 60:11 - 15. Jackson, M.L. and Sherman, G.D., 1953. Chemical weathering of minerals in soils. Adv. Agron., 5:219 318. James, H.L., 1966. Chemistry of the iron-rich sedimentary rocks. U.S. GeoL Surv. Prof Pap., 440W: 61 pp. Jans, H. and Van Calster, P., 1962. Note sur la presence de sid6rose dans le Diestien de la Campine. Bull. Soc. Beige GkoLPa!kontol. t-lydroL, 71:545 -550. Joffe, J.S., 1923. Acid phosphate production by the Lipman process. IlL The use of greensand marl as the inert material in building up sulphur-floats mixtures. Soil ScL, 15:93-97. Joffe, J.S. and Kolodny, L., 1937. The distribution and fixation of potassium in the profile of brown podzolic soils and sandy podzols. Proc. SoilSci. Soc. Am., 2:239 241. Joffe, J.S. and Levine, A.K., 1947a. Fixation of potassium in relation to exchange capacities of soils: II Associative fixation of other cations, particularly ammonium. Soil ScL, 63:151-158. Joffe, J.S. and Levine, A.K., 1947b. Fixation of potassium in relation to exchange capacity of soils: Ili Factors contributing to the fixation process. Soil ScL, 63:241-247. Jolliffe, F., 1935. A study of greenalite. Am. Mineralogist, 20:405-425. Jukes-Browne, A.J., 1903. The Cretaceous Rocks of Britain. Vol. II. The Lower and Middle Chalk of England. Mem. GeoL Surv. U.K., 568 pp. Jung, J., 1954. Les illites du bassin oligocbne'de Salins (Cantal). Bull Soc. Fr. Minkral. Cristallogr., 77:1231 1249. Jurkowska, H., 1956. Some results from investigations on manurial value of glauconite. Zesz. Nauk. Wyzsz. Szk. Roln. Krakow, 1:91-101 (in Polish, with English abstract). Kampioni-Zakrzewska, M., 1957. Sur glauconie des marnes cretac~es des environs de Zurawna. Arch. MineraL, 13:9-19 (in Polish, with French abstract).

434

S.G. McRAE

Kauffman, A.H. and DiUing, E.D., 1950. Differential thermal curves of certain hydrous and anhydrous minerals, with a description of the apparatus used. Econ. Geol., 45:220-244. Kazakov, G.A., 1964. The use of glauconite to determine the absolute age of sedimentary rocks. Khim. Zemnoi Kory, Akad. Nauk. S.S.S.R., Tr. Geokhim. Konf., 2:539-551 (in Russian). Kazakov, A.V. and Isakov, E.N.,and Isakov, E.N., 1940. Glauconite from the Egorov phosphate deposits U.S.S.R. Zap. Vse~ Mineral. O-va, 6 9 : 2 8 - 4 0 (in Russian). Kazakov, G.A. and Polevaya, N.I., 1958. Some preliminary data on elaboration of the post-Precambrian scale of absolute geochronology based on glauconites. Geochemistry, 374-387. Kazakov, G.A. andTeplinsky, G.I., 1966. On the capture of argon by the crystal structure of glauconite. Geochem. lnt., 3: 93. Kazakov, G.A., Knorre, K.G. and Prokot'yeva, L.N., 1965. Absolute age of Precambrian sedimentary rocks in the Oleneksk Uplift, East Siberia. Geochem. Int., 2:985. Keferstein, C., 1828. Deutschland, geognostischgelogisch dargestellt. Weimar, 5:510 511. Keller, W.D., 1953. Illite and montmorillonite in green sedimentary rocks. J. Sediment. Petrol., 23:3-9. Keller, W.D., 1958. Glauconitic mica in the Morrison formation in Colorado. Clap's Clay Miner., 5: 120-128. Keller, W.D., 1959. Glauconitic mica in the Morrison formation in Colorado, U.S.A. Int. Geol. Congr., 2 0th, Mexico, 1956, Sect. 11A: 121-131. Kharkwal, A.D., 1966. Glauconite in the Subathu Beds (Eocene) of the Simla Hills of India. Nature, 211:615-616. Kitazaki, U., 1951. On the crystal structure of glauconite, with special reference to its genesis - notes on the glauconite-bearing strata of the Noto Peninsula Pt. 2. Misc. Rep. Res. Inst. Nat. Resour. Tokyo, 23:38-47. Konta, J., 1967. Remarks to some terms in the paper "Morphology, internal structure, and origin of glauconite pellets". Sedimentology, 8: 169-170. KiSster, H.M., 1964. Glaukonit aus der Regensburger Oberkreideformation. Beitr. Mineral. Petrogr., 11:614-620. Krumbein, W.C., and Garrels, R.M., 1952. Origin and classification of chemical sediments in terms of pH and oxidation-reduction potentials. J. Geol., 60: 1-33. Kubovics, I., 1964. Primary glauconite in igneous rocks. Acta Geol. Hung., 8:19-35. Kulp, J.L., 1959. Absolute age determinations of sedimentary rocks. Proc. World Pet. Congr., 5th, New York, N. Y., 1959. Sect. 1, 37:689-705. Lacroix, A., 1893. Mineralogie de'la France et de ses Colonies, 1. Baudry, Paris, 612 pp. Lamboy, M., 1968. Sur un processus de formation de la glauconie en grains ~ partir des d~bris coquilliers. R61e der organismes perforants. C.R. Acad. ScL, 266D: 1937 1940. Larsen, G. and Chilingar, G.V., 1967. Chapter 1, Introduction. In: G. Larsen and G.V. Chilingar (Editors), Diagenesis in Sediments. Elsevier, Amsterdam, pp. 1-17. Lazarenko, .E.K., 1956. Problems of nomenclature and classification of glauconite. Vopr. Mineral. Osad. Obraz., Lvov. Gos. Univ., 3 - . 4 : 3 4 5 - 3 7 9 (in Russian). Leaf, A.L., 1958. Determination of available potassium in soils of forest plantations. Proc. Soil Sci. Soc. Am., 22:458-459. Lebauer, L.R., 1964. Petrology of the Middle Cambrian Wolsey Shale of southwestern Montana_ J. Sediment. Petrol., 34:503-511. Leclaire, L., 1964. Contribution h l'~tude des conditions physiocochimiques favorable ~ la gen~se de la glauconie darts le d~troit de Sicile. C.R. Acad. ScL, 258:5020-5022. Legg, J.O. and Axley, J.H., 1958. Investigation of a thermal method fox the determination of fixed potassium in soils. Proc. Soil Sci. Soc. Am., 22:287-290. Leith, C.K., 1903. The Mesabi iron-bearing district of Minnesota. Monogr. U.S. Geol. Surv., 43:240. Lerbekmo, J.F., 1963. Petrology of the Belly River formation, Southern Alberta foothills. Sedimentology, 2:54-86. Levine, A.K., 1939. The Relationship of Potassium Fixation to the Exchange Capacity of Soils. Thesis, Rutgers Univ., New Brunswick, N.J., 97 pp. Levine, A.K. and Joffe, J.S., 1947a. Fixation of potassium in relation to exchange capacity of soils: IV. Evidence of fixation through the exchange complex. Soil ScL, 63: 3 2 9 - 335.

GLAUCONITE

435

Levine, A.K. and Joffe, J.S., 1947b. Fixation of potassium in relation to exchange capacity of soils: V. Mechanism of fixation. Soil Sci., 6 3 : 4 0 7 - 4 1 6 . Libor, O., 1962a. Studies on Hungarian glauconites. Magy. K~m. Foly., 6 8 : 5 4 3 - 5 4 5 (in Hungarian with German abstract). Libor, O., 1962b. Investigations on Hungarian glauconites. VI. Cation exchange on glauconites. Magy. Kkm. Foly., 68:545-547. (in Hungarian with German abstract). Libor, O., 1964. A study of glauconite disaggregation. F61dt. KOzl, 9 4 : 3 6 2 - 3 7 0 (in Hungarian with English abstract). Libor, O. and Varga, E., 1963. Fixation and description of biologically active trace elements on glauconite. Agrokem. Tala/tan, 12:621-630 (in Hungarian with English abstract). Light, M.A., 1952. Evidence of authigenic and detrital glauconite. Science, 115:73-75. Lipman, J.G. and Blair, A.W., 1917. Vegetation experiments on the availability of phosphorus and potassium compounds. Rep. N. J. Agric. CoIL Exp. Stn., 353-368. Lipson, J., 1958. Potassium-argon dating of sedimentary rocks. GeoL Soc. Am. Bull,, 69: 137-150. Litvinenko, A.U., 1965. On the knowledge of glauconite. Litol. Polez. Iskop., 2:38 50 (in Russian). Lochman, C., 1949. Paleoecology of the Cambrian in Montana and Wyoming. Committee on a Treatise on Marine Ecology and Paleoecology, Rep. Nat. Resour. Counc. U.S., Washington, D.C., 9:31-71. Lochman, C., 1957. Paleoecology of the Cambrian in Montana and Wyoming. In: H.S. Ludd (Editor), Paleoecology. Mem. Geol. Soc. Am., 67:117-162. Maegdefrau, E. and Hofman; V., 1937. Glimmerartige Mineralien als Tonsubstanzen. Z. Kristallogr. Kristallgeom., 98:31-59. Malkova, K.M., 1956. Celadonite from the Bug region. Mineralog. Sb., L'vov, 10:305-318 (in Russian). Manghnani, M.H. and Hower, J., 1964a. Glauconites: cation exchange capacities and infrared spectra. Part I. The cation exchange capacity of glauconite. Am. Mineralogist, 4 9 : 5 8 6 - 5 9 8 . Manghnani, M.H. and Hower, J., 1964b. Glauconites: cation exchange capacities and infrared spectra. Part II. Infrared absorption characteristics of glauconites. Am. Mineralogist, 49:1631-1642. Mansfield, G.R., 1919a. Preliminary report on potash exploration in New Jersey greensands. Rep. N.J. Dep. Conserv. Dev., 99-104. Mansfield, G.R., 1919b. General features of the New Jersey glauconite beds. Econ. Geol., 14:555-567. Mansfield, G.R., 1920. The physical and chemical character of New Jersey greensand. Econ. Geol., 15:547--566. Mansfield, G.R., 1922. Potash in the greensands of New Jersey. Bull U.S. Geol. Surv., 727:146 pp. McCall, A.G. and Smith, A.M., 1920. Effect of manure sulphur composts upon the availability of the potassium of greensand. J. Agric. Re~, 19: 239-256. McCallie, S.W., 1908. Fossil iron ores of Georgia. Bull Geol. Surv. Ga., 17:199 pp. McDougall, I., Dunn, P.R., Compston, W., Webb, A.W., Richards, J.R. and Bofinger, V.M., 1965. Isotopic age determinations on Precambrian rocks of the Carpentaria region, Northern Territory, Australia. J. Geol. Soc. Aust., 12:67-90. McRae, S.G., 1971. Some Studies on Glauconite in Soils with Special Reference to Selected Soils in South East England. Thesis, London Univ., London, 372 pp. McRae, S.G. and Lambert, J.L.M., 1968. A study of some glauconites from Cretaceous and Tertiary formations in south-east England. Clay Miner., 7:431 440. Mehmcl, M., 1939. Datensammlung zur Mineralbestimmen mit R~Sntgenstrahlen. Teil I. Fortschr. Mineral Kristallogr. Petrogr., 23:91 - 118. Mehra, O.P. and Jackson, M.L., 1960. Iron oxide removal from soils and clays by a dithionite-citrate system buffered with sodium bicarbonate. Clays Clay Miner., 7:317 327. Mero, J.L., 1965. The Mineral Resources of the Sea. Elsevier, Amsterdam, 312 pp. Mikheev, V.I. and Stulov, N.N., 1955. On the products of high-temperature heating of layer silicates. Zap. Vses. Mineral O-va, 84: 3 - 2 9 (in Russian). Millot, G., 1949. Relations entre la Constitution et la Genkse des Roches S~dimentaires Argileuses. GOologie Appliqu~e et Prospection Mini~re, 2. Univ. Nancy, Nancy, 352 pp. Millot, G., 1964. G~ologie desArgiles. Masson, Paris, 499 pp.

436

S.G. McRAI.I

Milner, H.B., 1962. Sedhnentary Petrography (4th revised edition), 2. Allen and Unwin, London, 715 PP. Mogarovsky, V.V., 1963a. Glauconisation of argillaceous rocks as a form of alteration (near-ore) in one of lhe deposits of celestine of the Middle Tadjikie depression. Dokl. Akad. Nauk. S.S.S.R., 151:1178 1181. Mogarovsky, V.V., 1963b. Epigenetic glauconite of red sediments of Goulissai, Tadjikistan S.E. Zap. Vses. Mineral. O-va 92:718-720 (in Russian). Mogarovsky, V.V., 1965. Second discovery of epigenetic glauconite in association with celestine. Litol. Polez. Iskop., 1:118 119 (in Russian). Monroe, W.tI., 1947. Rapid oxidation of glanconite in glauconitic sand. Bull Ant. Assoc. Pet. Geologists, 31 : 1509- 1511. Moore, tt.B., 1939. Faecal pellets in relation to marine deposits. In: P.D. Trask (Editor), Recent Marine Sediments. Am. Assoc. Pet. Geologists, Tulsa, Okla., pp.516-524. Moreau, P., 1965. Sur quelques observations sddimentologiques concernant le Cdnomanien de la rdgion d'Angou16me. C.R. Acad. Sci, 261:3173 3175. Miiller, G., 1967. Diagenesis in argillaceons sediments. In: G. Larsen and G.V. Chilingar (Editors), Diagenesis in Sediments. Elsevier, Amsterdam, pp. 127 177. Mi2111er, W., and Behne, W., 1954. Zur Morphologie des Glaukonits. Naturwissensehaften, 41:575. Murina, G.A. and Sprintsson, V.D., 1961. Retention of radiogenic argon in glauconites. Geochemistry, 489- 494. Murray, J. and ttjort, J., 1912. The Depths o f the Ocean. Macmillan, London, 821 pp. Murray, J.W. and Mackintosh, E.E., 1968. Occurrence of interstratified glauconite-montmorillonoid pellets, Queen Charlotte Sound, British Columbia. Can. J. Earth Sci., 5:243 247. Murray, J. and Phillippi, E., 1908. Die Grundproben der "Deutschen TieJ~'ee-Expedition. " Wissenschaftliche l'irgebnisse der Deutschen Tiefsee Expedition auf dem Dampfer "Valdivia'", 1898 99, Jena, 1908. X:79 206. Murray, J. and Renard, A.F., 1891. Report on Deep-Sea Deposits based on the Specimens collected during the Voyage o f H.M.S. "Challenger" in the Years 1872 t() 1876. H.M.S.O., London, 525 pp. Mursky, G.A. and Thompson, R.M., 1958. A specific gravity index for minerals. Can. Mineralogist, 6:273 287. Nagelschmidt, G., 1944. X-ray diffraction studies on illite and bravaisite. Mhteralog. Mag., 27:59 61. Needham, C.E., 1934. The petrology of the Tombigbee sands of Eastern Mississippi. J. Sed#nent. Petrol., 4:55 59. Nicolas, J., 1961. Sur la presence de "Glauconie" en Bretagne Centrale. Colloq. Int. Cent. Natl. Red*. Sei. (Paris'j, 105:197 206. Nikolayeva, I.V., Klyarovskiy, V.M., Grigor'Yeva, T.N., and Borodayevskaya, Z., 1969. Influence of epigenesis on the structural and chemical properties and absolute age of glauconite. Dokl. Akad. Nauk. S.S.S.R., 85:90-93. Norris, R.M., 1964. Sediments of Chatham Rise. Bull. N.Z. Dep. SeL ind. Res., 159:40 pp. Nutting, P.G., 1943. Some standard thermal dehydration curves of minerals. U.S. Geol. Surv., Pro/~ Pap., 197E:197 217. Oakley, K.P., 1943. Glauconite sand of Bracklesham Beds, London Basin. Wartime Pare. Geol. Surv. Engl. Wales, 33: 1-27. Odin, G.S., 1971. Sur la gen~se de glanconies et leur signification sddimentologique d'aprbs l'dtude detaillde du sondage du Mont Cassil (Nord). C.R. Acad. Sci., 272D:697 699. Ojakangas, R.W. and Keller, W.D., 1964. Glauconitization of rhyolite sand grains. J. Sediment Petrol., 34:84 90. Owens, J.P. and Minard, J.P., 1960. Some characteristics of glauconite from the coastal plain formations of New Jersey. U.S. Geol. Surv., Prof. Pap., 400B: 430 432. Parry, W.T. and Reeves, C.C., 1966. Lacustrine glauconitic mica from pluvial Lake Mound. Lynn and Terry counties, Texas. Am. Mineralogist, 51 : 229- 235. Penrose, R.A.F., 1888. The nature and origin of deposits of phosphate of lime. Bull. U.S. Geol. Surv., 46:143 pp. Pentland, A., 1931. The heavy minerals of the Franconia and Mazomanie sandstones, Wisconsin. J. Sediment. Petrol., 1:23 --36.

437

GLAUCONITE

Peyronel Pagliani, G. and Pagnani, G., 1965. Tuff con glauconite nell alta Val Trompia (Brescia). Rend. Soc. Mineral ltal., 21:247 255. Pfefferkon, G., T h e m a n n , H. and Urban, H., 1956. Electron cn Mikroskopische Untersuchungen zur lnnerstruktur von Glaukonit KiSrner. Naturwissensehaften, 43:535. Pirani, R., 1963a. Studio petrografico dei prodotti delle manifestazione eruttive di Alcama (Monte Bonifato) e di Segesta (Monte Barbaro). Mineral. Petrogr. Acta, Bologna,. 19:1- 18. Pirani, R., 1963b. Sur fillosilicato dei levelli eruttivi di Monte Bonifato di M c a m o e di Monte Barbaro di Segesta e sulla validita di uso della nomenclatura binomia; glauconite celadonite. Mineral. Petrogr. Acta, Bologna, 9:31 78. Plumley, W.J. and Graves, R.W., 1953. Virgilian reefs of the Sacramento Mountains. New Mexico. J.

Geol.,61:l

16.

Pocock, R.W., 1942. Ochres, u m b e r s and other natural earth pigments of England and Wales. lCartime Pare. Geol. Surv. Engl. Wales, 21 : 1 - 19. Polevaya, N.I. and Kazakov, G.A., 1960. Nwe data on the geochronology of the late Precambrian. DokL Akad. Nauk. S.S.S.R., Earth Sci. Sect., 135: 162-- 165. Polevaya, N.I.. Murina, G.A. and Kazakov, G.A., 1961. Utilization of glauconite in absolute dating. Ann. N. Y. A cad. Sci., 9 1 : 2 9 8 - 3 1 0 . Porrenga, D.H., 1963. Chamositc in Recent sediments of the Niger and Orinoco deltas. Geol. Mi/nb., 44:400--403. Porrenga, D.H., 1966. Clay minerals in Recent sediments of the Niger delta. Clays Clay Miner., 14:221 233. Porrenga, D.H., 1967. Glauconite and chamosite as depth indicators in the marine environment. Mar. Geol., 5 : 4 9 5 - 5 0 1 . Porrenga, D.H., 1968. Non-marine glauconitic illite in the Lower Oligocene of Aardeburg, Belgium. Clay Miner., 7 : 4 2 1 - 4 3 0 . Prather, J.K., 1905. Glauconite. J. Geol., 1 3 : 5 0 9 - 5 1 3 . Pratt, W.L., 1961. The origin and distribution of glauconite and related clay-mineral aggregates off southern California. Natl. Coastal Shallow Water Res. Conf., Washington, D . C , I st, 656 - 658. Pratt, W.L., 1962a. Glauconite from the sea floor of central and southern California. Geol. Soc. Am., Spe~:. Pap., 73:58. Pratt, W.L., 1962b. Origin and distribution of glauconites and related clay aggregates on the sea floor off southern California. Bull. Am. Assoc. Pet. Geologists, 46:275. Pratt, W.L., 1963a. The Origin and Distributon of Glauconite from the Sea Floor off Southern CaliJbrnia. Thesis, Univ. S. Calif., Los Angeles, Calif., 285 pp. Pratt, W.L., 1963b. Glauconite from the sea floor off southern California. In: R.L. Miller (Editor), Essays in Marine Geology in Honor o tK.O. Emery. Univ. S. Calif. Press, Los Angeles, Calif., pp. 97-119. Rabinovich, S.D., 1942. Glauconite deposits in the Novo-Lialin district on the eastern slope of the Urals. Dokl. Akad. Nauk. S.S.S.R., 3 7 : 6 3 - 6 5 . (in Russian). Rao, M.S., 1964. Some aspects of continental shelf sediments off the east coast of India. Mar. Geol., 1:59 87. Rech-Frollo, M.M., 1963. Les conditions de formation de la glauconie en relation avec le Flyscb. C.R. Acad. Sci., 2 5 7 : 3 0 1 1 - 3 0 1 3 . Rooney, T.P. and Kerr, P.F., 1967. Mineralogic nature and origin of phosphorite, Beaufort County, North Carolina. Geol. Soc. Am. Bull,, 7 8 : 7 3 1 - 7 4 8 . Ross, C.S., 1916. The "chloritic" material in the ores of southeastern Missouri. Econ. Geol., 11:288-290. Ross, C.S., 1926. The optical properties and chemical composition of glauconite. Proc. U.S. Natl. Mux, 6 9 : 1 - 1 5 . Ross, C.S., Miser, H.D. and Stephenson, L.W., 1929. Water-laid volcanic rocks of early Upper Cretaceous age in southwestern Arkansas, southeastern O k l a h o m a and northeastern Texas. U.S. Geol. Surv., Prof. Pap., 1 5 4 F : 1 7 5 - 2 0 2 . Rossetti, V. and Sitzia, R., 1956. Le terre verdi nell' eruttivo terziario della Sardegna centro-occidentale. Period.~Miner., 25 : 17 1 - 2 0 8 .

438

S.G. McRAt:

Rouse, R.D. and Bertramson, B.R., 1949. Potassium availability in several Indiana soils: its nature and me th otis of evaluation. Proc. Soil Sci, Soe. Am., 14:113-123. Rubinshtein, M.M., Chikvaidze, B.G., Khutsaidze, A.L. and Gel'man, O.Ya., 1959. On the use of glauconite for the determination of the absolute age of sedimentary rocks by the argon method. DokL Akad. Nauk. S.S.S.R., 12:77-83 (in Russian). Sabatier, M., 1949. Recherches sur la glauconie. Bull Soc. Fr. Mineral. Cristallogr., 72:474-542. Sardarov, S.S., 1963. Preservation of radiogenic argon in glauconite. Geochemistry, 937 944. Schaub, H.P., 1955. Physical limits of glauconite formation. Bull. Am. Assoc. Pet. Geologists, 39:1878-1879. Schneider, H., 1927. A study of glauconite. J. Geol. 35:289-310. Schnepfe, M.M., May, I. and Naeser, C.R., 1964. Cesium and strontium sorption studies on glauconite. U.S. GeoL Surv., Prof. Pap., 501B:95-99. Schreiber, C., 1907. Glauconite as a fertiliser. Bull Agric., Brux., 23(9):656 665. Schreyer, W., 1956. Dolomit und Glauconit an der Basis des marinen Burdigals in der niederbayerischen Molasse. GeoL BL Nordost Bayern, 6: 169-171. Seed, D.P., 1965. The formation of vermicular pellets in New Zealand glauconites. Am. Mineralogist, 50:1097 1106. Seed, D.P., 1968. The analysis of the clay content of some glauconitic oceanic sediments. J. Sediment. Petrol., 38: 229- 231. Shreve, R.N., 1921. Action of lime on greensand. J. Ind. Eng. Chem., 13:693. Sidwell, R., 1936. Mineral study of Kiamchi formation of west Texas. J. Sediment. Petrol., 6:31 - 34. Simpson, E.S., 1934. Glauconite; its distribution and uses. Chem. Eng. Min. Rev., 26:391-394. Skeen, J.R., 1925. Greensand as a source of potassium for green plants. Am. J. Bot., 12:607-616. Skow, M.L., 1961. Minor nonmetals. Minerals Yearb., 1387-1389. Sloan, F., 1908. Catalogue of the mineral localities of South Carolina. Bull. S.C. Geol. Surv., 2:505 pp. PP. Smith, J.V. and Yoder, H.S., 1956. Experimental and theoretical studies of the mica polymorphs.

Mineralog. Mag., 31 : 209 - 235. Smulikowski, K., 1936. Skolite, Un nouveau mindral du groupe de glauconie. Arch. Miner., 12:145 178. Smulikowski, K., 1954. The problem of glauconite. Arch. Miner., 18:21-120. Spencer, D.M., 1951. Watercress investigations. Rep. Natl. Veg. Res. Stn, 61 68. Spurt, J.E., 1894. The iron-bearing rocks of the Mesabi Range. Bull. Minn. GeoL Surv., 10:259 pp. Spurt, J.E., 1915. Origin of certain ore-deposits. Econ. GeoL, 10:473-475. Stahlberg, S., 1960. Studies on the release of bases from minerals and soils. Ill. The release of potassium by boiling normal hydrochloric aci.d. ActaAgric. Seand., 10:185 -204. Steenhuis, J.F., 1937-1939. Glauconiet. Natuurh. MaandbL : 26-28. Swanson, C.L.W., 1949. Preparation and use of composts, night ~oil, green manures and unusual fertilizing materials in Japan. Agron. J., 41:275-282. Takahashi, J., 1939. Synopsis of glauconitization. In: P.D. Trask (Editor), Recent Marine Sediments. Am. Assoc. Pet. Geologists, Tulsa, Okla., pp.503-512. Takahashi, J. and Yagi, T., 1929. Peculiar mud-grains and their relation to the origin of glauconite. Econ. Geol., 24:838-854. Tapper, M. and Fanning, D.S., 1968. Glauconite pellets: similar X-ray patterns from individual pellets of lobate and vermiform morphology. Clays Clay Miner., 16: 275-283. Tarr, W.A., 1918. Glauconite in dolomite and limestone of Missouri. GeoL Soc. Am. Bull., 29: 104. Tarr, W.A., 1920. The origin of glauconite. Science, 51:491-492. Taylor, J.C.M., 1959. The clays and heavy minerals of the Shotover lronsand series. Proc. Geol. Assoc., 70:239 253. Tedrow, J.C.F., 1966. Properties of sand and silt fractions in New Jersey soils. Soil SeL, 101:24-30. Teodorovitch, G.I., 1961. Authigenic Minerals in Sedimentary Rocks. Consultants Bureau, New York, N.Y., 120 pp. Toler, L.G. and Hower, J., 1959. Determination of mixed layering in glauconites by index of refraction. Am. Mineralogist, 44:1314-1318.

GLAUCONITE

439

Topley, W., 1872. On the agricultural geology of the Weald. J,R. Agric. Soc., 8:241 267. Toplcy, W., 1875. The geology of the Weald. Mere. Geol. Surv. U.K., 503 pp. Touillaux, R., Fripiat, J.J. and Toussaint, F., 1960. Etude en spectroscopic infrarouge des mindraux argileux. Tranx Int. Congr. SoiI Sci., 7th, Madison, Wise., 1960, 4:460- 467. Tresise, G.R., 1960. Aspects of the lithology of the Wessex Upper Greensand. Proc. Geol. Assoc., 71:316 339. Triplehorn, D.M., 1961. The Petrology of Glauconite. Thesis, Univ. Illinois, Chicago. Ill., 117 pp. Triplehorn, D.M., 1966a. Morphology, Internal structure and origin of glauconite pellets. Sedimentology 6: 247-266. Triplehorn, D.M., 1966b. Glauconite provides good oil search data. Worm Oil, 162:94-97. Triplehorn, D.M., 1967. Morphology, internal structure, and origin of glauconitc pellets: a reply. Sedimentology, 8: 170-171. Tree, R.H. and Geise, F.W., 1918. Experiments on the value of greensand as a source of potassium for plant culture. J. Agrie. Res., 15:483-492. Tugarinov, A.I., Shanin, L.L., Kazakov, G.A. and Arakelyants, M.M., 1965. On the glauconite ages of the Vindhyan System. Geochemistry, 504. Twenhofel, W.H., 1936. The greensands of Wisconsin. Econ. Geol., 31:472--487. Tyler, S.A., 1936. Heavy minerals of the St. Peter sandstone in Wisconsin. J. Sediment. PetroL, 6:55

84.

Tyler, S.A. and Barley, S.W., 1961. Secondary glauconite in the Biwabic iron-formation of Minnesota. Econ. Geol., 56:1033-1044. Urban, H., 1957. Der Glaukonit in den Unterkreideformation yon Westfalen und seine Gencse. Tonind..Ztg. Keram. Rundsch., 81:363-371. Valeton, I., 1958. Der Glaukonit und seine Begleitminerale aus dem Terti~r yon Walsrode. Mitt. Geol. Staatsinst. Hamb., 27: 88-131. Van Andel, T.J. and Postma, H., 1954. Recent Sediments of the Gulf of Paria. Reports of the Orinoco Shelf Expedition. North-Holland, Amsterdam, 245 pp. Velde, B., 1965. Experimental determination of muscovite polymorph stabilities. Am. Mineralogist, 50:436-449. Vinogradov, A.P. and Tugarinov, A.I., 1962. Problems of geochronology of the Precambrian in eastern Asia. Geochim. Cosmochim. Acta, 26:1283-1300. Von Giimbel, C.W., 1886. Uber die Natur und Bildungsweise des Glaukonits. Sber. Bayer. Akad. Wiss., 16:417 449. Warn, G.F. and Sidwell, R., 1953. Petrology of the Spraberry sands of West Texas. J. Sediment. Petrol., 23:67-74. Warshaw, C.M., 1957. The Mineralogy of Glauconite. Thesis, Penn. State Univ., University Park, Pa., 155 pp. Weaver, C.E., 1958. The effects and geologic significance of potassium "fixation" by expandable clay minerals derived from muscovite, biotite, chlorite and volcanic material. Am. Mineralogist, 43:839-861. Webb, A.W., McDougall, I. and Cooper, J.A., 1963. Retention of radiogenic argon in glauconites from Proterozoic sediments, Northern Territory, Australia. Nature, 199:270 271. Wermund, E.G., 1961. Glauconite in early Tertiary sediments of Gulf Coastal Province. Bull. Am. Assoc. Pet. Geologists, 45: 1667-1696. Wermund, E.G., 1964. Geologic significance of fluvio-detrital glauconite. J.Geol., 72:470 476. Wetzel, W., 1937. Die koprogenen Beimengungen mariner Sedimente und ihre diagnostische und lithogenetische Bedeutung. Neues Jahrb. Mineral GeoL PaldontoL Beilbd., 78:109 122. Weyl, W.A., 1951. Light absorption as a result of the interaction of two states of valency of the same element. J. Phys. Colloid Chem., 5 5 : 5 0 7 - 5 1 2 . Williams, H., Turner, F.J. and Gilbert, C.M., 1954. Petrography: an Introduction to the Study of Rocks in Thin Section~ Freeman, San Francisco, Calif., 406 pp. Winchell, A.N. and Winchell, H., 1951. Elements of Optical Mineralogy, Part H. Wiley, New York, N.Y., 4th ed., 551 pp. Wise, W.S. and Eugster, H.P., 1964. Celadonite: synthesis, thermal stability and occurrence. Am. Mineralogist, 49:1031-1083.

440

S.G. McRAE

Wolff, R.G., 1967. X-ray and chemical study of weathering glauconite. Am. Mineralogist, 52:1129 1138. Wood, G.V., 1956. The heavy mineral suites of the Lower Greensand of the Western Weald. Proe. Geol. Assoc., 67: 124-137. Wurman, E., 1960. A mineralogical study of a gray-brown podzolic soil in Wisconsin derived from glauconitic sandstone. Soil Sci., 89: 38-44. Yasyrev, A.P., 1966. Distribution of trace elements in glauconites of the Russian platform. Dokl. Akad. Nauk. S.S.S.R., 168:197-199. Yodel tt.S., 1959. Experimental studies on micas; a synthesis. Clays Clay Miner., 6:42 60. Yodel H.S. and Euster, H.P., 1955. Synthetic and natural muscovites. Geochinr Cosmochim. Aeta, 8:225 280. Young, B.R., Harrison, R.K., Sargeant, G.A. and Stevenson, I.P., 1968. An unusual glauconite associated with hydrocarbon in reef-limestones, near Castleton, Derbyshire. Proe. Yorks. Geol. Soe., 36:417-434. Zabelin, V.A., 1962. Contribution to the question of the structure and genesis of the glauconite from Saratov. Dokl. Akad. Nauk. S.S.S.R., 144: 142-144. Zaporozhtseva, A.S., 1954. About joint occurrence of glauconite and chamosite in rocks. Dokl. Akad. Nauk. S.S.S.R., 97:903 905 (in Russian}. Zen, E., 1959. Mineralogy and petrography of marine bottom sediment samples off the coast of Peru and Chile. J. Sediment. Petrol., 29:513 539. Zodac, P., 1945. Greensands in New Jersey. Rocks Miner., 20:215. Zumpe, H.H., 1962. The Petrology of the Chloritic Marl (Glauconitic Marl) and the ttighest Upper Greensand of the Isle of Wight and Swanage Bay. Thesis, London Univ., London, 195 pp. Zumpe, H.H., 1971. Microstructures in Cenomanian glauconite from the Isle of Wight, England. Mineralog. Mag., 38:215 224. (Accepted for publication Augt~st 7, 1972)