Вы находитесь на странице: 1из 58

PL ISBN 83-86726-73-3

PL ISSN 0079-3396

POLSKA AKADEMIA NAUK ODDZIA W KRAKOWIE


KOMISJA NAUK MINERALOGICZNYCH

PRACE MINERALOGICZNE (Mineralogical Transactions) vol. 88, 80 pp. (2000)

ZBIGNIEW SAWOWICZ

FRAMBOIDS:
FROM THEIR ORIGIN TO APPLICATION

88 WYDAWNICTWO ODDZIAU POLSKIEJ AKADEMII NAUK KRAKW

2000

__________________________________________________________________________

CONTENTS Abstract........ Introduction....... Mineral composition....... Pyrite formation........ Framboids formation........ Association with organic matter............ Association of pyrite framboids with other minerals...... Environments of framboid formation........ Distribution in a sediment........... Framboids and related forms........... Framboid outline........ Size of framboid and its microcrystals.......... Morphology of microcrystals in framboid......... Ordering in framboid........... Annular framboids........ Sunflower framboids......... From micro- to polyframboids....... From framboids to euhedra......... Framboidal aggregates.......... Framboids and concretions...... Recrystallization and replacement processes.......... Accumulation and redistribution of metals in framboids......... Accumulation of trace metals.......... Redistribution of trace metals......... Transformation......... Oxidation............. Framboids and fossils........... Pyrite framboids and the origin of life........ Role of pyrite framboids......... Energy and catalysis.......... Accumulation and reaction chambers......... Templates............ Applications........... References........... Photographs.......

POLSKA AKADEMIA NAUK

ODDZIA W KRAKOWIE
PRACE MINERALOGICZNE 88 (2000) PL ISSN 0079-3396

KOMISJA NAUK MINERALOGICZNYCH PL ISBN 83-86726-73-3

ZBIGNIEW SAWOWICZ

FRAMBOIDS: FROM THEIR ORIGIN TO APPLICATION


Institute of Geological Sciences, Jagiellonian University, ul. Oleandry 2A, 30-063 Krakw, Poland; E-mail: zbyszek@geos.ing.uj.edu.pl Key words: pyrite, framboids, sulfides, origin of life, fossils, ore genesis, sedimentary rocks.

ABSTRACT
Framboids, spherical aggregates of minute, usually pyrite, grains are the commonest texture of sulfides in sedimentary rocks. They also occur in many other environments, ranging from magmatic rocks to antiquarian books. Framboidal textures are formed by several minerals including copper and zinc sulfides, greigite, magnetite, magnesioferrite or hematite, as the products of both primary deposition and replacement. Special attention is paid to pyrite framboids in this paper, with detailed descriptions of different framboid-related forms: annular framboids, sunflower framboids, micro- and polyframboids. Framboidal forms are often hierarchically structured over three size scales, with complexities ranging from microframboids, to framboids, and to polyframboids. Pyrite in framboids results from inorganic reactions between dissolved iron and sulfide, with a greigite intermediary. The sulfur is usually biogenic in origin. Simple pyrite framboids are formed during aggregation, possibly enhanced by the magnetic properties of the monosulfide precursor. Further processes, including particulation and organically controlled aggregation, result in more complicated forms such as polyframboids. Pyrite framboids can grow to euhedral grains provided the supply of iron and sulfide is not limited. Pyrite framboids often occur in close spatial relationship with organic matter, silica or carbonates, which influence their formation and growth. Replacement of iron sulfide framboid-related forms by other minerals can be indicative of sediment diagenetic conditions, or additional processes such as mineralization in ore deposits. Framboids influence the distribution of many trace elements. Due to their high specific surface areas, framboids can accumulate these trace elements during growth. Recrystallization of framboids can redistribute many trace elements. An overview of the common association between pyrite framboids and fossils is given, stressing both the importance of organic matter in framboid formation, and of the framboids in the processes of mineralization and preservation of fossils. Iron sulfide framboids may have influenced the earliest stages of life formation on the Earth as a source of energy and catalytic action, by accumulating organic compounds, and by acting as reaction chambers and templates which facilitated reproduction and information transfer. Applications from framboid studies are presented, ranging from determination of the redox conditions in water and sediment columns using size and distribution of framboids, to evaluation of deterioration processes in museum specimens.

INTRODUCTION
Spherical aggregates of minute, usually pyrite, grains resembling raspberries, which were first described as framboids by Rust (1935) from French word framboise, have fascinated many scientists for decades (e.g. Schneiderhohn 1923; Fabricius 1961; Kalliokoski 1974; Wilkin, Barnes, 1997a). Globular pyrite was described already in 1875 by M. Daubree (vide Morse et al. 1987) from the Roman pavement beneath a conduit for mineral waters. Framboidal pyrite is found in all kinds of sedimentary rocks (conglomerates, sandstones, shales, carbonate rocks), metamorphic rocks and in some magmatic rocks (e.g. andesites (Love, Amstutz 1969), of all ages, from Late Archean (Hallbauer 1986) to recent sediments (e.g. Morse, Cornwell 1987) and in most of sediment-hosted and volcanogenic ore deposits (e.g. Cu-, Zn-Pb- or Fe-deposits; Kanehira, Bachinski 1967; Chauchan 1974; Ostwald, England 1979; Taylor 1982; England, Ostwald 1993). Sulfide framboids are especially common in anoxic sediments, where bacterial sulfate reduction (BSR) is very active and an amount of available metals, mainly iron, is sufficient. However, framboidal pyrite is present also in shallow sandy aquifers, where the rate of BSR is three orders of magnitude lower than in marine settings (Jakobsen, Postma 1999). Pyrite framboids may also be found in oxidized marine sediments, where bacterial sulfate reduction is active within reduced microniches (Jrgensen 1977). Framboid is a typical texture of pyrite formed via monosulfide precursors, mackinawite and greigite, mainly in sedimentary rocks (Rickard 1970; Sweeney, Kaplan, 1973; Wilkin, Barnes 1997a). Although framboid-like texture is most common for pyrite, there are also other minerals which under specific conditions may reveal a similar texture, e.g. copper sulfides (Sawowicz 1990), spinels (Simon, Grossman 1997), modern spherular dolomite (von der Borch, Jones 1976), garnet (Bard 1980), precious opal (Darrogh et al. 1965) and phosphoric derivatives of allophane (Nakazawa 1999). Different genetic origins proposed for framboids have ranged from a purely inorganic formation, based on laboratory synthesis (e.g. Berner 1969; Graham, Ohmoto 1994) to a direct biogenic origine (e.g. Schneiderhoehn 1923; Love 1957). Extensive information on pyrite framboids and their genesis can be found in papers of Love and Amstutz (1966), Kalliokoski and Cathles (1969), Rickard (1970), Raiswell (1982), Schallreuter (1984), Sawowicz (1993a), Hamor (1994), or Wilkin and Barnes (1997a). On the basis of earlier papers of the author (Sawowicz 1987, 1990, 1992, 1993a; Kosacz, Sawowicz, 1983; Large, Sawowicz & Spratt, 1999), voluminous literature and many new observations, this paper critically evaluates the current knowledge and attempts to present the mystery, beauty and complex role played by the framboids in nature. The first systematics of framboid-related forms and some new ideas on the formation and transformation of framboids are presented. The important role played by gels and colloids in the formation of framboids and in accumulation of trace elements is stressed. The author proposes a hypothesis that the best possible microenvironments to begin the formation of life on Earth were pyrite framboids. A unique album of SEM (EDS supported) microphotographs of framboidal forms expands knowledge on these famous but hardly visible creatures.

MINERAL COMPOSITION
Framboidal texture is typical of pyrite but it also is found among other minerals, like copper and zinc sulfides, greigite, magnetite, magnesioferrite, hematite, goethite, garnet, and dolomite. The origin of such texture in framboids of different mineralogy not necessarily must

be similar. Some of the non-pyrite framboids can be both primary and/or form through the replacement of pyrite or greigite. Copper sulfide framboids occurring in sediment-hosted ore deposits have been described by Alyanak and Vogel (1974; White Pine) and Sawowicz (1990; Polish Kupferschiefer). Sawowicz (1990) showed that copper sulfides (digenite and chalcocite) from the Kupferschiefer copper-bearing shales occur often as spherules and rarer as framboids. The prevaillence of copper sulfide spherules over framboids suggests faster growth process of microcrystals and their homogenization, in comparison with pyrite framboids. Framboid-like texture in case of copper sulfides can result from their formation via process of aggregation of smaller particles. Some of the copper sulfide framboids from the Kupferschiefer, especially those from the Cu-Fe-S zone, may be pseudomorphs after pyrite framboids (Sawowicz 1992; Oszczepalski 1999). However, in the authorss opinion, based on detailed study of the textures and chemistry of the Kupferschiefer Cu-S framboids and their comparison with pyrite frmboids, majority of the copper sulfide framboids in the Cu-S zone are primary precipitates (Sawowicz 1990). Thompson and Helz (1994) found that chalcocite occurs probably as 10100 nm particles in sulfidic zone of the Black Sea and suggested possibility of existence of copper sulfide cluster complexes in aqueous solution. It is interesting to point that also larger grains of copper sulfides form by aggregation of smaller particles, similar to the formation of spherules from framboids (Figs. 6 d, e; in Sawowicz 1990). Some of the spherules and framboids are subhedral to euhedral. Organic envelopes around some framboids are observed. By comparison with pyrite framboids, the present author suggests that major part of framboidal or spherular primary precipitates of copper sulfides in the Kupferschiefer is of early diagenetic origin. Geerite, copper sulfide (Cu1.6S), in the form of spherules (150-300 m diameter, Fig. 7D) has been synthesized in silica gels, by the reaction between dissolved Na2S and Cu(OH)2 (Reid, Sawlowicz 1992; unpublished), and may correspond with the pyrite microconcretions. Marcasite framboids, formed probably at low pH, have been described from several localities (Ixer, Vaughan 1993; Youngson 1995). Zinc sulfide framboids are rare. Luther et al. (1980) described such forms from sediments containing above 1000 ppm Zn. Lebedev (1967) found ZnS framboids in Iokunzh ore deposit. Degens et al. (1972) found zinc sulfide globules suspended in the water column of Lake Kivu. Several ferrimagnetic minerals have framboidal texture resulting probably from their magnetic properties. Greigite was produced experimentally (Sweeney, Kaplan 1973) and is common especially in young sediments, also lacustrine (Ariztegui, Dobson 1996). Magnetite framboids are recorded both from terrestrial (Bethke, Marshak 1990) and extraterrestrial materials (Kerridge 1970; Jedwab 1971; Zolensky et al. 1996). Magnesioferrite was described from the seafloor hydrothermal silica sinter (Taylor 1982). Spherular forms of dolomite, about 0.2-1 m in size and composed of spherules about 100 nm in diameter, were described from ephemeral lakes in South Australia (von der Borch, Jones 1976). Shapes of some of the aggregates have been modified by rhomboidal outlines due to incipient recrystallization of the dolomite. Similar forms of protodolomite (1-5 m in size) were produced experimentally from carbonate gels (Mueller, Fishbeck 1973). Close spatial relationship between pyrite framboids and spheroidal dolomite crystals of probably bacterial origin was described from the Visean karst system (Nielsen et al. 1997). Oxidation of pyrite framboids is probably relatively common phenomenon. Possible products of pyrite oxidation include goethite, lepidocrocite, szomolnokite, hematite and probably also magnetite. Luther et al. (1982) registered oxidized framboids which contained iron and zinc as the chemical constituents. Hematite (Lougheed, Mancuso 1973) and magnetite (Suk et al. 1990) framboids were described as secondary due to pyrite replacement.

Wilkin and Barnes (1997a) suggested that such a replacement, connected with 36% volume reduction, could hardly preserve details of fine microstructure. However, if we suppose that magnetite replace primary greigite, which would be much simpler, the volume reduction would be only 22% with much bigger chances of structure preservation.

PYRITE FORMATION
Framboidal texture is most typical of pyrite, therefore pyrite formation is shortly discussed here. There are two essential aspects of pyrite framboid formation: the chemical processes leading to precipitation of iron sulfides and the mechanism of particle nucleation, growth and aggregation (discussed in the next section). Pyrite is among the best-studied minerals but its formation is still debatable. Pyrite may be formed via at least three pathways, including the reaction of precursor sulfides with polysulfides, the progressive solid-state oxidation of precursor iron sulfides, and the oxidation of iron sulfides by hydrogen sulfide, with different rate of the formation for each process and for a greigite intermediary (see the review in Rickard et al., 1995). In low temperatures pyrite growth is usually preceded by the formation of unstable iron monosulfides (Berner 1970, 1984; Sweeney, Kaplan 1973; Goldhaber, Kaplan 1974; Rickard 1975; Rickard et al. 1995; Schoonen, Barnes 1991a,b), according to the Ostwald rule where the least stable sulfide precipitate first (Morse, Casey 1988). Natural sulfidic waters are typically saturated or slightly undersaturated with respect to the monosulfides and supersaturated with respect to pyrite (Howarth 1979; Lord, Church 1983, Canfield, Raiswell 1991). Pyrite formation begins with the reaction of bisulfide and ferrous ions to form aqueous complexes of Fe(HS)2 and FeHS-, followed by the formation of solid Fe(HS)2 and subsequent conversion to mackinavite (FeS) and greigite (Fe3S4), which can be very fast process (Rickard 1989). Transformation from greigite to pyrite may occur by sulfur addition (Berner 1967, 1984) or iron loss (Furukawa, Barnes 1995). Wang and Morse (1996) suggested that pyritization of metastable iron sulfides follows a dissolution-precipitation pathway. Schoonen and Barnes (1991a) suggested that the energy barrier for pyrite nucleation inhibites pyrite formation from aqueous solutions. Thus, the precursor iron monosulfide can serve as an active surface. In acidic environment when both FeS2 and FeS particles are present, FeS2 can grow directly from the solution on the FeS2 nuclei or by conversion of FeS particles (Schoonen, Barnes 1991b). Rickard (1997) showed that the pyrite forms as overgrowths on several FeS particles and most pyrite subsequently nucleates and grows on earlier-formed pyrite. Recent experiments (Rickard 1997) suggest that in strictly anoxic systems pyrite can form very fast by the H2S oxidation of iron monosulfide in aqueous solutions at low temperatures. Butler and Rickard (1999) were able to produce framboidal pyrite by such a reaction. Morse and Cornwell (1987) proposed that pyrite formation might proceed via iron monosulfide dissolution and subsequent pyrite precipitation. Organic compounds, e.g. aldehydes, can modify the processes of iron sulfide formation (Butler et al. 1999). In sedimentary environment the major source of sulfur incorporated into iron sulfides is H2S or HS- resulting from BSR. Sulfate reducing bacteria do not precipitate iron sulfides directly and their role is probably limited to the production of bisulfide ions (Benning et al. 1999). Iron hydroxides and oxides supply most of the reactive iron used in the Fe sulfide formation (Boesen, Postma 1988; Canfield 1989). Their solubilization is especially fast in a reducing environment. Recently Taillefert et al. (1999) have shown that reactivity of soluble Fe(III) (<50nm diameter) with sulfides is much higher compared to its solid equivalents, especially when complexed by an organic ligand. On the other hand, dissolution of detrital hematite and magnetite is very slow due to their small specific reaction surfaces and

replacement of the rims of these grains by Fe sulfides may last hundreds and thousands of years (Canfield, Berner 1987).

FRAMBOIDS FORMATION
Laboratory synthesis of simple monodispersed colloids, even under monitored conditions, is still poorly understood. In nature we deal with bigger number of variables. Therefore, all considerations presented in this section on the framboid formation in nature may only be regarded as a presentation of possible processes and their interactions. Different genetic origins have been proposed for framboids, ranging from a purely inorganic origin, based on laboratory synthesis (Berner 1969; Farrand 1970; Sunagawa et al. 1971; Sweeney, Kaplan 1973; Kribek 1975; Graham, Ohmoto 1994), through indirect biogenic formation (e.g. Vallentyne 1963; Skripchenko 1968; Kalliokoski, Cathles 1969; Lougheed, Mancuso 1973), to a direct biogenic origin (e.g. Schneiderhoehn 1923; Love 1957; Fabricius 1961; Locquin, Weber 1978). Recently Wilkin and Barnes (1997a) have suggested four consecutive processes leading to the pyrite framboid formation: 1/ nucleation and growth of initial iron monosulfide microcrystals; 2/ reaction of the microcrystals to greigite; 3/ aggregation of uniformly sized greigite microcrystals; 4/ replacement of greigite framboids by pyrite. The authors opinion is that, although these processes can explain some of the simple framboid features, they cannot explain the formation of more complex framboidal forms and the role of organic matter in this process. The experimental works mentioned above support the opinion that inorganic processes are satisfactory to produce framboids. On the other hand, biogenic processes may both produce sulfur and modify the conditions of framboid formation. A wide range of physical and chemical conditions of framboid occurrences suggests that pyrite framboids can probably be formed in more than one specific process, although there are several features that must be common for all these processes. A number of modifications of the processes in nature seem to be unlimited. It is interesting to note that in some very carefully inspected experiments on pyrite formation framboidal morphology could not be reproduced (e.g. Rickard 1975; Schoonen, Barnes 1991b), whereas in less carefull experiments pyrite framboids were produced in a great number (Farrand 1970; Kribek 1975). The present author postulates that gels and colloids are necessary in the framboid formation. Many observed similarities between well-studied formation of uniform particles of silica and other elements and minerals (e.g. gold) in colloidal systems (Iler 1979; Weitz, Oliveira 1984; Matijevi 1996) and framboid formation suggest that formation of colloids in water may be the important step in the iron sulfide framboid formation. According to Kopelman (1989), gels are neither liquid nor solid but exist at the boundary between these states and consist of a distribution of self-similar clusters over a range of sizes, fractal on all length scales. It should be stressed that colloids often form more readily in dilute solutions (suspension as a sol) than in concentrated ones (formation of heavy precipitate). In nature Fe (II) could be oxidized to Fe (III) and then reacting with seawater would produce oxyhydroxide sol (Kalliokoski, Cathless 1969). Colloids composed of Fe (III) oxyhydroxides can be generated by infiltration of oxygen-rich water into anoxic aquifers containing dissolved Fe (II) (Liang et al. 1993). Iron may be present in interstitial waters also as bacteriogenically bound colloidal species like [Fe(H2O)5]2+ or [Fe(H2O)4(OH)2]+ (Ferris et al. 1987). Coacervate droplets of Fe (OH)2 may be stabilized by different substances, like organic matter or colloidal silica. Colloidal systems are especially sensitive to the presence of organic substances that can act as protective colloids. The experiments of Kretzschmar and Sticher (1998) showed that addition of small amounts of humic acid into the suspension of hematite

particles resulted in aggregation and settling of the particles, whereas addition of larger amounts of humic acid increased colloidal stability due to increased negative surface charge. Humic substances are not only able to form rigid spherocolloids at high concentration and low pH (Ghosh, Schnitzer 1980) but their adsorption may stabilize the dispersed mineral particles (Jekel 1986). One of the most important points about coacervate drops, both of biological or inorganic origin, is that they represent diffusion-limited places, where supersaturation levels can be regulated. Narrow distribution of sizes and uniform growth of thousands of crystals in framboids in short time interval can be explained by a regulated balance between rates of nucleation and of crystal growth, beautifully shown in experimental works of La Mer and co-workers (e.g. La Mer, Dinegar 1950). Growth of colloidal particles involves condensation reactions and aggregation effects. As solutions concentrate, macromolecules form by condensation of small polyanions. These macromolecules aggregate and by further condensation form sol particles. As the particle diameter increases, aggregation becomes a dominant growth mechanism. The more freedom of movement of primary particles, the more prefect spherical clusters develop. Nucleation of a supersaturated solution by the first-formed crystal leads to the separation of many crystals, all of the same size (Kondepudi et al. 1990; Matijevi 1996; Wilkin, Barnes 1997a). Framboidal texture results probably from rapid nucleation in environments where pyrite is strongly supersaturated (Butler, Rickard 1999). Supersaturation can be achieved not only by building up the concentration of dissolved component but also by lowering of solubility (LaMer, Dinegar 1950) which can be realized by changing pH and/or temperature. Any saturated solution may be regarded as composed of randomly oriented molecular complexes dispersed in an essentially homogeneous solution (Grisdale 1965). Similar molecular clusters may then be the building blocks concerned in growth of the crystal or due to attractive-repulsive forces in the formation of th framboid. In biological environments, the activation energy for nucleation can be reduced by lowering the interfacial energy and/or increasing the supersaturation (Mann 1988). In Farrands (1970) experiments, the optimal critical concentration was 1-10 times the solubility of pyrite for diffusion-controlled accretion of material particles to aquire a shape determined by a spherical concentration gradient. The mechanism of the formation of the smallest units is not known as they aggregate and recrystallize probably fast. Sometimes observed the central hole and concentric rings in microcrystals suggest that they may have grown in suspension by depositing of iron sulfide onto a precipitated nucleous, perhaps via spiral growth (Schallreuter 1984; Rickard, Luther 1996). It is worth mentioning that spiral-like growth of framboid was observed in replaced biotite (Morrissey 1972). Circular structures on the surface of some individual microcrystals (Fig. 9D) may indicate crystal growth via the development of screw dislocations (Butler, Rickard 1999). It should be stressed that the colloid aggregation is a universal process, independent of chemical details of the particular colloid system, and governed by diffusionlimited (DLCA) or reaction-limited (RLCA) aggregation conditions (Lin et al. 1989). DLCA causes rapid and RLCA slow colloid aggregation, which in the case of pyrite framboids could be primarily dependent on the rate of supersaturation and viscosity (controlled by organic compounds). Experimental works of Sweeney and Kaplan (1973) and Berner (1969) showed that framboidal texture develops probably during the recrystallization of precursor iron monosulfides to pyrite, where the spheroidal form is related to the mackinawite to greigite transformation and the microcrysts formation with the greigite to pyrite transformation. In addition to surfactant forces, conversion of an iron sulfide to greigite (an inverse spinel; Vaughan, Tossell 1981) can be accompanied by the development of ferrimagnetism, which could encourage colloidal FeS to gather on preexisting sulfide surfaces (Taylor 1982). Quantitative approach was presented by Wilkin and Barnes (1997a) who elegantly used the

well known from the colloidal systems DLVO (Derjaguin, Landau, Verwey, Overbeek) theory to describe the aggregation of iron sulfide microcrystals. This theory describes behaviour of particles in colloidal suspension in terms of the interplay of attractive and repulsive forces. Among the most important are van der Waals, electrical double-layer and magnetic interactions. Numerical models proposed by Wilkin and Barnes (1997a) predict that magnetic greigite particles >0.1m will rapidly aggregate in natural waters. These authors also suggested that in the case of the absence of magnetic interactions, aggregation would be favored in high-ionic strengh solutions. Framboid transformation from greigite to pyrite should not destroy their primary texture. On the other hand, Butler and Rickard (1999) have shown that framboid formation may proceed directly through the iron monosulfide oxidation by H2S without the greigite intermediary. In the authors opinion a simple aggregation model cannot explain all aspects of framboid complexity (microframboid-framboid-polyframboid). As Kalliokoski (in Kalliokoski, Cathles 1969) suggested, it is difficult to imagine formation of polyframboids, comprising subunits of the same size, through the process of agglomeration. The presence of polyframboids in confined spaces, like tests, seems to limit application of aggregation theory. Organic substances probably play an important role in these processes. It is interesting to note that in experiments where organic compounds were excluded, framboids mostly formed by aggregation (Sunagawa et al. 1971), whereas in experiments where gel droplets were stabilized by organic substances, framboids formed by particulation (Kizilshtein, Minaeva 1972; Kribek 1975). Different size scale ranges can result from two or more different successive processes. Sawowicz (1993a) suggested that both aggregation and particulation mechanism could be useful in the formation of framboidal aggregates. Aggregation would play important role in the formation of simpler framboids and particulation in the formation of more complex forms. In this context, very interesting are experiments of Reeder et al. (1996) on aggregation of colloids from a stable sol. Laboratory synthesis of hierarchically structured zirconia incorporated two different colloid aggregation techniques: the polymerization-induced colloid aggregation (PICA) to synthesize micrometer-scale particles (1-5 m) and subsequently, the oil emulsion method (OE) to synthetize aggregates tens of micrometers in diameter (40-150 m). Slow removal of the water by extraction into, and subsequent evaporation from the oil phase, lead to a porous aggregate. By comparison with experiments of Reeder et al. (1996) where the second step of the structure building was the oil emulsion (OE) process, the authors interpretation is that organic gel globules, perhaps containing partly formed greigite, greigite-pyrite or pyrite framboids, aggregate into polyframboids (Fig. 1A). During evaporation of water polyframboids organize themselves. Aggregation to polyframboids (relatively rare, not known from experiments, encountered only in natural environments, especially those with high porosity) may be enhanced by the decay of organic matter which behaves in water like an oil emulsion (Rickard 1969). Interesting observations on the growth of framboids come from the framboids composing molds of fossil skeletons or tests (Fig. 12A, B). Close packing infilling of tests can be the evidence for the growth of framboid constituents after their initial nucleation due to a continuous supply of iron from external sources and production of H2S by BSR in situ. The succession of events could be: formation of framboid-size globules (similar to OE process, Reeder et al. 1996) within the test, nucleation and crystallization, succeeded by growth of crystals in individual framboids (Fig. 1). Temperatures of the framboid formation vary significantly. Most of the framboids are found in low temperature environments but some of them are described from higher temperature localities. Th temperature of framboidal pyrite formation in the hydrothermal field of Japan is within the range 150-210oC (Halbach et al. 1993). Pyrite framboids are

common in black smokers (Fig. 7C). Based on the magnetic aggregation model, Wilkin and Barnes (1997a) suggest that mechanism for framboidal pyrite formation via greigite is unlikely to operate in temperatures above 200oC. In some experiments (Sunagawa et al. 1971; Graham, Ohmoto 1994) framboidal pyrite can also form at higher temperatures (up to 350oC), but a different mechanism can be invoked (Wilkin, Barnes 1997a).

Fig. 1. Hypothetical pathways of the formation of polyframboids and polyframboidal aggregates (A in open space; B in confined space) and their transformation towards spherules and massive grains (for PICA and OE explanation see text) (not to scale). Association with organic matter Although interplay between minerals and organic molecules is common in nature and framboids have been synthesized without the presence of organic matter, the authors opinion is that in natural sedimentary environments organic matter is very important for the framboid formation. Pyrite framboids very often reveal a close relationship with organic matter (Fig. 11E, F), containing organic matter both in their interstices (e.g. Love, Amstutz 1966) and/or as membranes coating framboids (e.g. Sweeney, Kaplan 1973). Pyrite framboids are common in coals (Wiese, Fyfe 1986). As the most common occurrences of framboids are in reducing organic-rich environments, the basic role of organic processes is related to the biogenic sulfate reduction. However, influence of organic molecules and polymer matrix is also important in regulation of growth, size of crystals and their morphology, particle aggregation and solubility of metals (Mann 1988; Bianconi et al. 1991; Matijevi 1996). Organic matter in the interstices of framboid may be the remains of organic material in which pyrite crystallized (Love 1965). Rickard (1969) showed experimentally that during anaerobic organic matter decomposition an oil was formed being a mixture of dead bacteria and their waste products with precipitating iron sulfides. Hudson (1982) suggested that pyrite

10

forming in oil would preferentially crystallize on preexisting pyrite due to surfactant forces. Dameron et al. (1989) showed in their experiments that short chelating peptides control the nucleation and growth of CdS crystallites of diameter 20 . At this small size, the CdS structure differs from that of bulk CdS. The experiments also showed that peptides are effective in terminating the growth of crystallites of a specific size, and that accretion and growth are pH-dependent. It is interesting to note that in experiments, where organics was excluded, framboids of multiple-range size and/or with regularly arranged microcrystals have not been recorded. This might suggest that organic matter is necessary for the oriented growth. Conversion or growth of iron sulfides can be slowed down or even inhibited by the organic matter. Schoonen and Barnes (1991b) showed experimentally that the conversion of precursor amorphous FeS to pyrite may be slowed by the addition of organic ligand (citrate). Laufer et al. (1985) on the basis of electron-microscopic studies hypothetized that growth of pyrite crystals can be inhibited by thin coating of material rich in amorphous carbon. The presence of organic material could block reactive surface sites. Sacs, films, membranes or envelopes (Love 1957; Nuhfer & Pavlovic 1979) often surround pyrite framboids. Organic membrane in which framboids are encapsulated may influence the outer shape (spherical or euhedral) of the framboid. The higher pressure the membrane applies during the framboid growth, the more spherical is the shape of the framboid. Association of pyrite framboids with other minerals The association of framboids with other minerals can be both spatial (matrices and coating of framboids; e.g. Love et al. 1984) and genetical. In the authors opinion the latter case is especially important because acompanying minerals may influence the formation and structure of framboidal forms. Formation processes of framboidal pyrite, colloidal silica and other spherular mineral forms show numerous similarities. Von der Borch and Jones (1976) found that in Coorong sediments spherular aggregates of dolomite are accompanied by the similar aggregates of precious opal and suggested similar physical control as an explanation for morphological similarity. They also supoposed that small amount of associated opaline silica may be significant in the spherular dolomite forming process. Association of pyrite framboids with silica is common in nature. The presence of silica may probably both stabilize the former iron-rich gel (Kizilstein, Minaeva 1972) and preserve pyrite framboids from later recrystallization by isolation from external influences (additional pyrite supply, oxidation, etc.) (Ostwald, England 1977). Unique association of exclusively quartz and framboidal aggregates of pyrite can be found in Rochelinval pyrite beds in Belgium (Love, Vanguestaine 1973) (Fig. 9E, F). The characteristique feature of framboidal pyrite aggregation is that they are regularly arranged, probably as the result of close packing (Fig. 10E, F), sometimes forming large crystals built of framboidal microcrystals (Fig. 7E, F). It is interesting to note that pyrite framboids and framboidal clusters found in chalcedony in weathered andesites at Allandale (Australia) also showed indications of crystallization into pyrite polyhedra and pyrite octahedra (Ostwald, England 1977). There, iron sulfide framboids probably started to form and organize themselves within coloidal silica. Very thin and moulded silica interstices between pyrite framboids suggest that the consolidation of silica was later than the formation of pyrite crystals. Also in the case of precious opal, silica spheres with concentric rings are cemented by silica deposited after packing of the spheres (Sanders 1964). The described Belgium occurrences can be compared to the silica sinter. Taylor (1982) described a similar bed as being the product of precipitation of silica during cooling at an exhalative fluid/seawater interface, which contained magnesioferrite spheroids with framboidal texture.

11

In the present authors opinion in Rochelinval bed high contents of sulfides suggests that precipitation took place under low Eh and high sulfur content conditions. Regular arrangement and close packing in framboidal aggregates may result from their sinking to the bottom in silliceous colloid, similarly like it is suggested for precious opals. The precipitation of silica by cooling could preserved this unique framboidal pyrite texture mainly due to isolation from external supply of dissolved iron and/or sulfur. Association of pyrite framboids with clay minerals, especially kaolinite, is common in sedimentary rocks. Usually clay minerals fill interstices in pyrite framboids or form sheaths around framboids. Nuhfer and Pavlovic (1979) and Scheihing et al. (1978) suggested that sulfide formed within a kaolinitic gel. Pollastro (1981) described from foraminiferal tests spherical aggregates where kaolinite did not fill interstices between framboids but formed small aggregates of kaolinite plates intergrown with pyrite framboids of similar size. He suggested contemporaneous precipitation of kaolinite and pyrite from gels in bacteria-rich microenvironments of test. Sometimes fine grains of clay minerals may also be introduced inbetween pyrite microcrystals in framboid during later diagenesis. Pyrite framboids are often present in carbonate minerals, in massive rocks, in lenses or lamina in mudstones, or replacing carbonate organisms remains (Fig. 6C, D). Different framboidal forms may be present even in a very small volume, varying significantly in shape, size of framboid or aggregate, and their building crystals. These variations may result from specific interactions of iron and sulfur from local and external sources, and amount and quality of organic matter present. Pyrite framboids have also been found in barite from Kuroko-type deposits (Vavelidis 1995) and in authigenic gypsum crystals (Siesser, Rogers 1976) or irregular gypsum grains (Fig. 6B).

ENVIRONMENTS OF FRAMBOID FORMATION


Framboids are present in very different environments, from recent sediments (both marine and lacustrine, and also in deep-sea black smokers), to sedimentary and metamorphic, to magmatic rocks and hydrothermal deposits, but their most typical environment is represented by organic-rich marine sediments. Framboidal pyrite usually starts to form in the water column and/or during the earliest stages of diagenesis where bacterial sulfate reduction begins (Skei 1988; Canfield, Raiswell 1991). Iron sulfide framboids can form both in the water column and in the sediment. In normal marine sediments, pyrite is formed only in the sediment, whereas in euxinic environments pyrite can form both in the water column (up to 100% locally in the modern Black Sea; Lyons, 1997) and/or below sediment-water interface (Canfield et al. 1996; Wilkin, Barnes 1997b; Suits, Wilkin 1998). Raiswell and Berner (1985) distinguished syngenetic pyrite formed in the water column, typically anoxic and often euxinic, and diagenetic pyrite that forms inside of the sediment with anoxic conditions, often under oxic water column. These two types often differ in morphology and isotopic composition, size and possibly trace metal admixtures. Framboidal and euhedral pyrite formed during early and late diagenesis can differ in their isotopic sulfur composition, being generally lighter in the case of framboids formed earlier (Raiswell 1982; Shikazono, Utada 1997). On the other hand, often no significant difference between isotopic signal from the both pyrite forms is found. The present authors explanation is relatively simple. If in the last case euhedra represent also early diagenetic forms of framboid transformation (Sawowicz 1993a), one would not expect any isotopically significant difference. Suits and Wilkin (1998) suggested that in some cases where the typical Fe-S-C parameters do not distinguish between oxic and euxinic sediments, the size distribution of pyrite framboids can be helpful. Pyrite framboids formed in the recent anoxic water column are

12

generally smaller and less variable in size than their counterparts from the sediments (Wilkin et al. 1996). Growth time of framboids both from the water column and from the sediments was estimated by Wilkin et al. (1996). Maximum 7 years (25 m) in the sediment and maximum 0.4 years (1.7 m) for framboids formed in the Black Sea water column were found assuming that framboid growth rates are comparable for syngenetic and diagenetic framboids. However, in the authors opinion this presumption may not always be true because concentrations of the components are probably different in both environments, mainly due to different diffusion rate and processes of iron releasing from different minerals. Pyrite framboids are very common in ore deposits (e.g. sedimentary, SEDEX or hydrothermal of low temperatures) where their relationship to other sulfides can be very helpful in understanding their origin and history. There also are many rarer environments of framboid formation. Pichler et al. (1999) found framboidal pyrite and marcasite (their mineralogy dependent on variations in acidity) in the shallow-water hydrothermal system in Papua New Guinea. There, hydrothermal H2S oxidizes upon contact with oxygenated seawater, forms H2SO4, that dissolves primary Fe-rich grains, leading to the neoformation of iron sulfide framboids. Framboids can also be formed during late diagenesis, e.g. by the pyritization of biotite (Menon 1967) or magnetite (Canfield, Berner 1987). Wilson and Zentili (1999) described formation of framboidal pyrite in Lower Cretaceous Lo Prado Formation (Chile) as the result of biodegradation of petroleum derived from underlying organic-rich shales. Pyrite framboids are also present in sulfidized bauxites (Sinkovec et al. 1994), in sulfidic marsh soils (Fitzpatrick et al. 1996), and in pyrite chimneys formed by the expulsion of ground water on a tidal flat (Orpin 1994). Magnetite framboids are relatively common in extraterrestrial materials (Jedwab 1971; Zolensky et al. 1996). Sporadically, later during diagenesis when sulfide is available, a second generation of framboidal pyrite may form (Sassano, Schrijver 1989). Localities where framboids are found not always represent the environment of their formation. Pyrite framboids have even been found suspended in an oxic zone of the seawater column (Middelburg et al. 1988), into which they could be transported from other sulfide-rich environments. Pyrite framboids as relatively durable can be transported with turbidites and found in layers inside oxidized sediments or in beach deposits (Hossain 1975). They were also found in flood sediments of the Odra river immediately after the flood event in 1997 (Damke et al. 1999). A very unusual environment of pyrite formation was described by Garcia-Guinea et al. (1997). The pyrite framboids formed in the anoxic environment inside antique books where the reduction of iron sulfate (mixed with tannine in ink) took place.

DISTRIBUTION IN A SEDIMENT
In sedimentary rocks different morphological forms of pyrite can be observed. Within a few micrometers of sediments, perfectly formed framboids can be intimately mixed with disseminated euhedral crystals, often of the same size. The intermediate textures between these two types are common (Fig. 5C). The present author postulates that distribution of framboids and euhedra, in laminae or in aggregates, depends on homo- or heterogenity of environment of deposition. When intermediate sulfur species, like polysulfides and thiosulfates, are homogeneously distributed in the solution, conversion of amorphous FeS to pyrite through progressive monosulfide phases should proceed at approximately the same rate throughout the system (Schouten, Barnes 1991b). These could explain typically uniform distribution of framboids or microcrystals in homogeneous sediment. In organic-rich laminated rocks the most common are individual framboids of similar size distributed randomly or occurring in a specific type of

13

lamina. There is no agreement which type of lamina, organic or clayey, is preferred by framboids. Distribution is probably dependent on diffusion interaction between H2S production in organic-rich lamina and availability of dissolved iron from clay-rich lamina. Frizzo et al. (1991), studying recent sandy-silty sediments of the Venice Lagoon, observed that distributions of Fe sulfide framboids may be heavily dependent on alternating and interacting redox and turn-over processes which lead to transformations of Fe sulfides to Fe hydroxides and their later dissolution and/or transformation back to sulfides. When framboids or their precursors settle through coloidal silica beds they can form extensive, sometimes closely packed, framboidal beds (Love, Vanguestaine 1973) (Fig. 10E, F). Storm events can rework pyrite framboids into lags (Twitchett, Wignall 1996). Porosity and structure of sediment play an important role in the distribution of framboids and may significantly influence their size. Coarse-grained rocks with bigger pores and hydrothermally altered rocks with voids often contain polyframboids and/or large framboidal clusters. Sawowicz (1987), studying framboidal pyrite from the Radzimowice schists (Sudety Mts.), has found that metamorphic processes may result in the recrystallization of some framboids or in pyrite overgrowths on framboids, but in many cases, a good preservation of framboidal texture is still recorded. When fossils are common in the rock, framboids are often found within confined spaces such as foraminifera tests, diatom frustules, polychaetes tubes and plant cells (Love 1967; Honjo et al. 1965; Thomsen, Vorren 1984; McNeil 1990) (Fig. 12).

FRAMBOIDS AND RELATED FORMS


Processes related to framboid formation may produce different textural and morphological forms both before the final formation of the framboid and as the result of its growth. The first detailed description of framboid-related forms is presented in this chapter (Fig. 2)

Fig. 2. Examples of framboid-related forms (not to scale).

14

Outline of framboids The origin of the spheroidal form of the framboid is not clear and probably results from several processes. On one hand, spherulization might be the result of pseudomorphism. Degens et al. (1972) observed microcrystalline sphalerite replacing resin globules suspended in Lake Kivu. Javor and Mountjoy (1976) proposed that pyrite framboids replace algal endospores. Framboids may grow by replacement of iron-rich organic globules (Kalliokoski, Cathles 1969) or gas bubbles surrounded by an organic sac (Rickard 1970). Raiswell et al. (1993) presented a three-dimensional model of diffusion-with-precipitation, where the organic matter decay generating hydrogen sulfide in iron-dominated porewaters results in the formation of vacuoles, which in turn produce iron sulfide spheres. Vacuoles of organic matter and H2S would be kept spherical by surface tension forces. H2S gas bubles alone could serve as vacuoles (vacuolation process) when H2S is evolved during the formation of FeS from Fe(SH)2 gel (Rickard 1989). Love and Amstutz (1966) and Rickard (1970) supposed that the curved margins of the outermost crystals cutoff at the margin of framboids result from the sphere membrane present at the place before sulfide crystallization. On the other hand, Berner (1969) and Sweeney & Kaplan (1973) showed that sphericity and microcrystals can be produced synthetically, and spheroidal form may be associated with the transformation of mackinawite to greigite (Sweeney, Kaplan 1973). Spherical magnetic aggregates can develop in weak magnetic fields (Wilkin, Barnes 1997a). Using a parallel with precious opal (Darrogh et al. 1966), the more freedom of movement of primary particles the more perfect spherical cluster develops. Self-organization of mineral clusters is typical not only of iron sulfide framboids but also of other mineral clusters that can form in similar process. For example, Nakazawa (1999) described formation of spherical clusters of phosphoric derivatives of allophane when phosphoric acid was added to the water with suspended allophane. Graham and Ohmoto (1994) were able to precipitate pyrite from solution within spherical droplets of liquid elemental sulfur. Spherical shape of framboids could also be a result of expansion of bubles as the surrounding gel contracted, as described by Eggleton (1987) in the case of a structure of hollow packed spheres built of noncrystalline Fe-Si-Al-oxyhydroxides. As framboids reveal either smooth or rough outlines, it is possible that there are several combined or separate paths to obtain spherical shape. The habit of particles building framboids varies from spherical to euhedral with the latter more common in bigger framboids. The shape of the framboid is not always ideally spherical. Very often framboids are elipsoidal, with the longer axis parallel to schistosity or lamination (Fig. 6A, 11F). Framboids that are in direct contact may have outlines differing from the regular spherical shape (Fig. 12E). They all, or one of them, may be flattened at the contact, similar to the contact of gas bubles or organic globules. All these forms suggest that pyrite (or its precursors) inherited their shapes or they were modified in elastic matrixes as the result of the overburden pressure. Size of framboid and its microcrystals One of the most spectacular features of framboids is the uniform size of microcrystals they are composed of. Uniformity of size and morphology of microcrystals within the individual framboid (Fig. 6C, D, 10D, F) can be explained by a process of simultaneous nucleation and growth at the same rate, well known since the experimental works of LaMer and coworkers on sulfur hydrosols (e.g. LaMer, Dinegar 1950; Matijevi 1996). Size criteria can be applied both for framboids and individuals (microframboids, spherules or crystals) that built them. In the authors opinion the most important factors that modify the size of framboid are: 1) type of sediment; 2) period of growth; 3) availability of iron and sulfur; 4) characteristics of solution in which framboids grow; 5) consistency of the parent gel
15

and probably type of organic matter or membrane which can modify surface tension of gel globules. Type of sediment (fine- or coarse-grained, laminated or massive, porous or non-porous, etc) influence significantly the size of framboids or framboidal aggregates. Heterogenity of the environment may results in relatively wide range of sizes. Polyframboids tend to form only in large voids, at present filled sometimes with silica or carbonate minerals. Framboids formed in foraminifer tests tend to be of similar diameter, whereas those formed in the outer sediment may vary greatly in size (Frizzo et al. 1991). Various sizes of framboids and framboidal aggregates may also result from different supersaturation ratio. A greater supersaturation ratio causes nucleation to predominate over crystal growth, giving smaller crystals. As LaMer & Dinegar (1950) pointed, when the solution is more concentrated a nonuniform growth and a polydispersed sol is the final result. The size of aggregates would also depend on the rate of aggregation. Particles aggregate slowly below the critical coagulation concentration and fast above that level (Thieme, Niemeyer 1996) Availability of iron depends on: concentration of dissolved iron in a water column, which is usually th highest in the iron-rich zone below the oxic/anoxic interface (Brewer, Spencer 1974); mineralogy of iron minerals (iron oxyhydroxides are among the most efficient sources of iron in a sediment; Canfield 1989; Raiswell, Canfield 1998); and possibility of external influx of dissolved iron (Berner 1970, 1980). Availability of sulfur depends usually on the efficiency of bacterial sulfate reduction, which is the major source of sulfide in a water column and sedimentary rocks, and on the diffusion rate. In Raiswells et al. (1993) model of diffusion-with-precipitation, where organic spheres are pyritized, the maximum size of framboids may provide a measure of the maximum concentrations of dissolved iron during early diagenesis. The maximum size of framboids formed in environment with typical concentration of dissolved iron in pore waters would be 50 m which is about the usual maximum size of framboids (Raiswell et al. 1993). LaMer and Dinegar (1950) listed three variables on which the final size of sulfur hydrosol particles and the rate of their growth will depend: number of nuclei, total amount of diffusible sulfur and diffusion coefficient. They would probably be also important in the case of framboids. The diameter ratio between framboid and building microcrystals vary significantly between 5:1 and 30:1 but typically is about 10:1. This proportionality of size might be explained by an inverse relationship between the size of primary droplets and their iron concentration, as reported from the experiments of Farrand (1970) and Kizilshtein & Minaeva (1972). The larger concentrations of iron in solution the bigger number of microcrystals in framboid. Wilkin et al. (1996) proposed that the microcrystal size in pyrite framboids is remnant from iron monosulfide precursors because nucleation of pyrite directly from solution is very difficult (Schoonen, Barnes 1991a). However, as the monosulfide stage is necessary only for nucleation, the size of microcrysts may also depends on the period of growth (in the case of unlimited supply of Fe and S). Limited space inside fossils and surface tension of interfacial (organic or gaseous) membranes may limit the size of precursory gel droplets and later framboids. The size of the microcystals could also be modified by different viscosity of the environment of precipitation (content of organic matter, silica, carbonates or clay minerals). Morphology of microcrystals in framboid Uniformity of size and morphology of microcrystals within individual framboid can be explained by a process of simultaneous nucleation and growth at the same rate (Matijevi 1996). In pyrite framboids the habit of the building microcrystals varies from cubic, to octahedral and pyritohedral, to equent anhedral (globular) (Fig. 5A, 8F, 9A-E). The size of the

16

crystal itself often determines its morphology (Grisdale 1965). In the case of framboids, the smallest units are usually spheres and the bigger are idiomorphic. Experiments of Wang and Morse (1996) have shown that morphology of pyrite crystals formed at room temperature is primarily controlled by the degree of supersaturation in the solution from which pyrite is precipitated. With the increasing supersaturation, pyrite morphology changes from cube, to octahedron, to spherule. Often observed pyrite octahedra modified by cubes suggest changes (decrease) in supersaturation during crystal growth (Fig. 9C). It is worth mentioning that morphologies of the uniform colloidal particles of different solids produced in laboratory might vary due to relatively small changes in experimental conditions (Matijevi 1992). At low magnification many microcrystals appear perfectly smooth whereas, for example, at higher magnifications in SEM their building subunits are sometimes clearly discernible and microcrystals resemble microframboids (Fig. 8F, 9A, B, D). Ordering in framboid The arrangement of uniform microcrystals in the framboid varies from random, linear or concentric to polygonal (Lougheed, Mancuso 1973; Chen 1978) (Fig. 8B, 10C-F). The pattern of regular arrangement of uniform spheres is not unique and even better known from precious opal, latex spheres or some insect viruses (e.g. Sanders 1964; Williams, Smith 1957). Regular arrangement can be very early, as it was observed in case of colloidal hydrated iron oxide (Schiller layers) (Iler 1965; Watson et al. 1962). Recent experiments with metallic molybdenum showed that self-arrangement may start even on the atomic level (10-20 atoms on an edge) and stable particles must be larger than some critical size (Edelstein et al. 1991). Further growth proceeds via collision leading to the formation of larger regular structure. Framboids might be considered as hard spheres systems, where the liquid-solid transition arises largely from packing considerations and is sensitive to the variation of the repulsive forces with distance (Oxtoby 1990). Generally it is believed that ordering of microcrysts in framboid is mainly a close-packing effect (Taylor 1982). There is no agreement if the arrangement follows the face-centred cubic (FCC), body-centred cubic (BCC) or hexagonal close packing (HCP) but all these patterns are very favorable from the energetical point of view, having the minimum interfacial area (Iler 1965, 1979). The type of arrangement is also very sensitive to many other factors, e.g. microgravity (Zhu et al. 1997). Among the mechanisms responsible for regular arrangement of crystals in framboids can be densification of aggregates (Klimpel, Hogg 1991, vide Wilkin, Barnes 1997a) or growth inside of closed space or outer membrane (organic?), causing internal-oriented pressure (surface tension). Electrostatic forces may also stabilize the regular pattern. Closely packed ordered framboids show tendency towards forming polyhedral framboids. Crystal-like surfaces are not observed in framboids without closely packed crystals. Interesting is the comparison with well-studied precious opals. Synthetic spheres of opal can be deformed into polyhedra by compression during packing (Darrogh et al. 1966). In natural opal arrays appear to have remained soft for long periods before being hardened. Worth of considering is the concept that ordering may be the result of anastrophic supramolecular organization with its far-from-equilibrium conditions (Nicolis, Prigonine 1977; Kauffman 1993). Disordered framboids can result from partly destroyed ordered framboids or from not closely packed framboids where tendency toward ordering is weak. Although the regular arrangement of crystals in framboids is relatively common in nature, it has not been found in experiments. Thus, the author suggests that the organic matter, common during the formation of framboids in the sedimentary environments, may play a significant role in the process of an ordered aggregation of sulfide particles. Organic matter matrix may strongly influence regular or irregular arrangement of crysts in the framboid.

17

Mann (1988) suggested that an existing organization of the organic matrix (molecular preorganization) which can be both random or periodic, together with molecular complementarity between the inorganic ions and the local binding sites in the matrix, may be essential for the crystallochemical specificity of mineral formation. This kind of aggregation was postulated by Mann (1983, vide Mann et al. 1990b) and produced during experiments with CdS and organic polymer (Bianconi et al. 1991). Experiments have shown that minerals, e.g. CdS, of uniform size, morphology and crystallographic orientation can be formed in ordered arrays in living cells (Bianconi et al. 1991). In the latter experiments, crystallization in a crystalline polymer produced the regular cubic crystal (1 m in size), formed by the ordered aggregation of smaller particles, whereas crystallization in an amorphous polymer gave spherical forms. It is worth noting that both the ordered framboids (Amstutz et al. 1967) and the single euhedra of pyrite (Ramberg, Ekstrom 1964) may show a preferred orientation of a crystallographically orientated pattern of ordering or crystal faces against the plane of bedding or schistosity of the rock. Annular framboids The term annular framboid was proposed by Kosacz and Sawowicz (1983) to describe hollow framboid consisting of a sphere (a discontinuous, rarer continuous, ring in section), composed of minute pyrite grains (Fig. 2, 12E). The interior can be filled with organic matter, sulfides, silica, or carbonates. Annular framboidal forms occurring in nature may have sizes typical both of framboids and microframboids (0.1-1 m and 5-10 m, respectively). The experiments of Kizilshtein and Minaeva (1972) have shown that annular framboids can be the first step in the development of complete framboids. The hollow interiors of annular framboids may result from: 1/ low availability of building material (iron or sulfur) (Papunen 1966; Love 1967); 2/ inhibiting role of organic compounds, silica or phosphates in the gel droplet interior; 3/ differential dissolution properties of a core (Sawowicz 1992, 1993a). Couture et al. (1998) described globular colloids below 1m (similar to spherocolloids of humic substances; Ghosh, Schnitzer 1980) which in the peat were composed exclusively of carbon, whereas in the drain these globules contained iron precipitate on the surface. One may expect that such forms submitted to sulfidation could result in annular iron sulfide framboids. Some framboids with weak magnetic character probably contain residual greigite cores (Roberts, Turner 1993). Dissolution of this metastable core and/or replacement by, for example, copper sulfides (Sawowicz 1992) may also lead to the formation of annular framboids. Annular pyrite framboids could also be a result of replacement of other minerals. Eggleton (1987) described noncrystalline Fe-Si-Al-oxyhydroxides, e.g. ferrihydrite, which showed a structure of hollow packed spheres of 50-1000 in diameter. Sunflower framboids So-called sunflower pyrite framboids consist of framboidal core and radiating elongated outer crystals (Love, Brockley 1973; Ostwald, England 1977; Kosacz, Sawowicz 1983) (Fig. 2, 5E). Both pyrite and marcasite can build outer crystals and grow continuously from the outermost grains of framboid or be separated from the framboid by thin (e.g. organic) membrane. It seems that in some cases, growth of radial crystals enrooted in framboids can lead to the formation of larger pyrite crystal (Fig. 3). Similar bigger forms (up to several centimeters) are known, for example, from the Cap-Blanc Nez chalk cliffs (marcasite nodules built exclusively of pyrite, Tambuyser 1976). Also the cross-sections of the pyrite stalactites from Zn-Pb Upper Silesian deposits show the similar textures (Fig. 7B). Two

18

different pathways leading to the formation of pyrite or pyrite-marcasite sunflowers are proposed by the author, respectively: 1/ pyrite crystals exhibit a piezoelectrical effect, allowing crystal growth along one axis and contraction along the other (Bang 1994). After formation of framboidal core the outermost grains, not limited by space and necessary components supply, grow radiating ouwards; 2/ changes (also local) in pH stimulate the formation of different FeS2 polymorphs. At pH >4 the framboidal pyrite core is formed, whereas below pH 4 formation of the marcasite radiating crystals in the outer rims begins (Fig. 7A). Such local pH changes may be a result of the microbial activity (Massaad 1974).

Fig. 3. Hypothetical pathway of the transformation from the pyrite framboid to the euhedra (not to scale). From micro- to polyframboids Pyrite polyframboids are common and easily recognizable in sedimentary rocks (e.g. Love 1971) whereas microframboids are rare and can be observed only under high electron microscope magnifications (Fig. 9A, B). Surprising similarities between granules building framboids, polyframboids and Mandelbrots The Fractal Geometry of Nature (1982) stimulated Sawowicz (1993a) to propose a sequence, increasing in complexity, from microframboid (av. 0.1-1 m in size) to framboid (av. 5-10 m) to polyframboid (av. 50-100 m) (Fig. 2). Each of the above forms comprises the simpler framboidal forms or their growth transformations. For example, microframboids, which are spheroidal forms consisting of discrete equant nanocrystals, are structurally similar to the normal framboids but are one order of magnitude smaller and in some case may build the framboids, similarly as it is in the case of microcrystals. It is not impossible that microframboids consist of even smaller nanoframboids (?) (Sawowicz 1993a). Hematite nanospheres of possible colloidal origin were described from the Precambrian BIFs (Ahn, Buseck 1990). Spherical aggregates (about 0.05 m diameter) of cassiterite, built of subunits (30-50), have been synthetized by Ocana

19

and Matijevi (1990, vide Matijevi 1992). There is a tendency for larger aggregates to consist of larger consituent structures, the ratio of microcrystal (or framboid) to framboid (or polyframboid) is typically 1:10. The comparison between microframboids, framboids and polyframboids suggests that the trend towards the formation of idiomorphic grains via a framboidal stage is inversely dependent on the size and complexity of the aggregate, i.e. the euhedra are much easier formed from microframboids (Fig. 8F) than from framboids. Size of the constituents building framboids in polyframboids is usually uniform but in some cases may vary significantly, probably as the result of locally different iron sulfide concentration. LaMer and Dinegar (1950) observed that when the solution is more concentrated a nonuniform growth takes place and a polydispersed sol is the final result. Taking into account a complexity of framboids (three size-scale), their formation as the result of one process is difficult to imagine. Polyframboids consist sometimes of both framboids and euhedra of the framboid size what may suggest a separate growth pathway of individual framboids and latter agglomeration, perhaps by different mechanisms (see PICA and OE processes - Reeder et al. 1996; in the section on framboid formation). On the other hand, framboids forming closely-packed polyframboidal aggregates in the confined space of fossils are usually very uniform (Fig. 12B) and were probably formed in one-step process, possibly mediated by organic material (Fig. 1). From framboids to euhedra Close spatial association of pyrite framboids and euhedra is very common in nature and suggests genetic relationship. The most peculiar link between these two morphologies are multi-faceted framboids, sometimes with regular arrangement of the building microcrystals (Fig. 10A-C). In pyritiferous beds from Rochelinval (Belgium) large euhedral pyrite crystals consist sometimes exclusively of framboidal aggregates (Fig. 7E-F). Even in relatively homogeneous environments, such as laboratory flasks, different stages of framboid growth can be observed at the same time (Kizilshtein, Minaeva 1972). Love and Amstutz (1966) suggested possibility of recrystallization from framboidal to single grain pyrite. Sawowicz (1993a) developed this suggestion and proposed a continuous growth of microcrystals in the framboids (sometimes towards euhedra) as long as they are in contact with the initial solution. Also Wilkin et al. (1996) supposed that the growth of the framboid terminates when it is removed from the site of crystallization, and presented more evidences for continuous growth. They showed that secondary growth does not preferentially occur on either small or large framboids and tends to increase with depth in the sediment. On the other hand, framboids suspended in the water column do not show overgrowths of secondary pyrite (Wilkin, Barnes 1997b). The process of crystal growth may be compared to the Ostwald ripening (Morse, Casey 1988) where larger crystals grow at the expense of smaller ones. A decrease in degree of supersaturation due to precipitation and declining iron availability may result in mackinawite undersaturation and pyrite supersaturation leading to direct pyrite precipitation (Raiswell 1982). At low temperatures, the pyrite-forming solutions are nearly always saturated with respect to iron monosulfides, and pyrite forms through replacement of precursor FeS phases. Once pyrite has formed from FeS, it can grow from solution (Schoonen, Barnes 1991a). Continuous growth from framboid to euhedra may be interrupted by unavailability of iron and/or sulfur, both due to limited supply or isolation from the external environment. The present author postulates the simplified pathways of the evolution, from framboids to euhedra (Fig. 3): A/ amalgamation of the microcrystals inside of the framboid and growth of the outermost radiating crystals to the euhedra; B/ growth of the polygonal framboid, with the stable internal crystallographic pattern, to regular faces of the euhedra; C/ growth and

20

amalgamation of the framboid microcrystals (framboid core) and supply of additional pyrite filling interstices and precipitating as euhedral overgrowth. Pyrite in the interstices is usually derived from external sources but by analogy with silica colloids it also is possible that some pyrite may result from dissolution of pyrite microcrystals in framboid. Iler (1973) found that in aggregation or flocullation of colloidal silica particles, silica dissolves from the particle surface and is deposited around the point of contact to minimize the negative radius curvature, thus forming a coalescence or neck between particles. Such a mechanism of keeping microcrystals together in framboid should also be taken into account, especially that recognition of sometimes only nanometers thick intergrowth may be difficult. The growth of framboids to spherules or euhedra may be or may not be continueous. Sometimes etching of euhedra may give additional information on their formation. A supply of additional pyrite may be recorded by the presence of framboidal core (Fig. 2, 5F) or different optical properties of pyrite forming overgrowth or filling interstices between microcrystals. The various optical properties of pyrite may result from different chemical composition (admixture of trace metals, like As) or porosity. During the crystal growth changes of supersaturation may result in changes of its habit (Murowchik, Barnes 1987). The trend towards the formation of idiomorphic grains via framboidal stage is inversely dependent on the size of primary droplet and complexity of the aggregate. It depends also on mineralogy of framboids. In the Zechstein Kupferschiefer growth of copper sulfide framboids typically leads to the spherules formation, whereas pyrite framboids very often grow to the pyrite euhedra (Sawowicz 1990). More information on framboid-euhedra transformation can be found in Sawlowicz (1993a). The author must stress that euhedral pyrite is probably not always formed via framboid growth. The morphology of pyrite is strongly dependent on the saturation level and the rate of formation. For example, rapid pyrite formation at high saturation levels resulted in framboidal pyrite within the Mediterranean sapropels, whereas below the sapropels, at low saturation levels, slow euhedral pyrite formation occurred (Passier et al. 1999). Experiments of Butler and Rickard (1999) have shown that during iron monosulfide oxidation by H2S pyrite framboids were formed at Eh>-250mV whereas single pyrite crystals at Eh<-250mV, probably as a supersaturation effect related to the position within the pyrite stability field in Eh-pH space. Framboidal aggregates The term framboidal cluster or framboidal aggregate is used here to describe irregular aggregates of small crystals, grains or even framboids, generally uniform in size and closely packed (Fig. 6D-F). The size of such aggregates may vary from several microns to several millimeters. Their boundaries can be very irregular, modified by surrounding minerals (Fig. 6E) or mimic the shape of the form (e.g. void or fossil test) they fill. Framboids in clusters often are cemented by framboidal grains (Fig. 6C). Similarly to the transformation from framboid to euhedra, the framboidal cluster may recrystallize to the massive grain (Sawowicz 1990-Fig.7) (Fig. 5F, 6B). The origin of framboidal aggregates seems to be similar to that of framboids (Fig. 1). Framboids and concretions Framboids and concretions reveal many similarities. The main difference between framboids and concretions can be made on the basis of relationship between microcrystals that build them. In framboids the microcystals are generally separated from each other or only touch themselves. In concretions microcrystals are welded and intergrown (Fig. 8C, D). The size of concretions may vary more than that of framboids, from tens of microns to several centimeters. Big concretions are much more common than concretions of the framboid size. It
21

is natural because bigger crystals cannot be separated and kept together due to attractiverepulsive forces. In the authors opinion it is possible that during early stages of concretion formation their individual components (crystals) had framboidal texture (polyframboids) or at least framboidal core. As the trend towards formation of pyrite euhedra from framboids increases with their size, it can be expected that the growth of crystals in bigger concretions is due to supply of secondary pyrite.

RECRYSTALLIZATION AND REPLACEMENT PROCESSES


Recrystallization of pyrite framboids is a very common process. Kohn et al. (1998) recognized the following paragenetic sequence of framboidal pyrite, occurring as infilling and pseudomorphs of the chambers of the tests of foraminifera: agglomeration of isolated framboids, cementation with interstitial pyrite and, finally, recrystallization with some relics of framboids. Fluid-rich diagenesis or low-grade metamorphism often results in thorough recrystallization of framboids and common formation of euhedra. Such pyrite forms are much more refractory and retain many characteristic features. Etching, as well as optical and electron microscopy with chemical mapping, may reveal chemical and physical textures that may be interpreted in terms of the depositional and the post-depositional history of the deposits (Craig et al. 1998). Replacement of pyrite framboids by other sulfides is especially common in sulfide ore deposits (e.g. Love, Zimmerman 1961; Bartholome et al. 1971; Ncube et al. 1978; MartinezFrias et al. 1997) and such textures can be very useful in genetical studies. On the other hand, well-preserved pyrite framboids often occur in copper sulfide grains and lenses, or these minerals fill interstices in framboids (Fig. 11A, B). The possible explanation for such complex structures is that pyrite framboids crystallized first from colloids of complex composition or that later sulfides enveloped framboids. One of the best examples of the replacement phenomena is the association of framboidal pyrite with Cu- and Cu-Fe-sulfides in the Kupferschiefer bed, described in detail by Love (1962), Kosacz and Sawowicz (1983) and Sawowicz (1992). Pyrite framboids in this famous shale are regarded to be the earliest sulfides and play an important role in the discussion on genesis of the Kupferschiefer-type deposits. Common replacement of pyrite by copper sulfides, especially in the form of atoll textures has been used as the argument for the late diagenetic formation of copper mineralization. However, Sawowicz (1992) postulated that the conversion of monosulfide to pyrite occurs from the rim of the framboid inwards. Along with evidence for the low reactivity of pyrite at low temperatures (Rickard, Cowper 1994) Sawlowicz (1992) proposed that atoll textures (framboids with pyrite rims and their cores replaced by copper and copper-iron sulfides, similar to annular framboids) resulted from the replacement of iron monosulfide or greigite cores after the formation of an outer rim of pyrite crystals. This process was limited to the earliest stages of diagenesis. In the authors opinion similar reasoning could also be applied to explain the formation of atoll textures in the cobaltite framboids (Large et al. 1999). In some cases, replacement of iron sulfide framboids by copper sulfides in the Kupferschiefer bed may lead to the formation of chalcopyrite or digenite framboids which reveal some textural and chemical differences in comparison with primary copper sulfide framboids (Sawowicz 1990). The presence of other metals during formation of pyrite and its precursors may notably influence their formation. Pankow and Morgan (1980) found that low concentrations of copper in solution could significantly slow down the dissolution rate of mackinawite, probably by forming CuS coatings on the mackinawite. Bush and Sullivan (1999) observed thin coatings of monosulfide with an acicular morphology around some framboidal clusters in the Australian Holocene sediments. In the Bendigo gold ore field framboidal pyrite or greigite

22

present in unaltered metasediments becomes progresively recrystallized and is replaced by hydrothermal pyrite, chalcopyrite and sphalerite as mineralization is approached (Li et al. 1998). Some magnetite (Suk et al. 1990), goethite and hematite (Mucke et al. 1999) framboids were described as secondaryresulting from pyrite replacement. On the other hand, pyrite framboids may occur as a replacement product on iron oxide and hydroxide grains (Canfield, Berner 1987; Frizzo et al. 1991; Stanton, Goldhaber 1991). Frizzo et al. (1991) observed also small pyrite framboids growing in cracks and discontinuities of Fe/Mg silicate grains, sometimes replacing small exsolutions of Fe oxides. The present author suggests that replacement of spherical aggregations of different minerals, especially iron-bearing, could be in some cases among the first steps (before sulphidization) of the pyrite framboids formation. Examples of such minerals could be goethitic and chamositic oolites which developed directly from colloidal ferric oxide and silicate precursors (James, Van Houten 1979) or spherical halloysite (Ward, Roberts 1990) and allophanes (Nakazawa 1999). If the precursor iron-rich mineral has a framboidal texture, sulfidization can produce iron sulfide framboids. Very irregular pyrite microcrystals in some framboids (Fig. 5A) could result from such process. It is worth noting that similarly to pyrite framboids also clusters of spherical aggregates (0.4-0.6 m in size) of halloysite may have a crudely polyhedral forms (Ward, Roberts 1990). By comparison with the replacement of siliceous radiolaria skeletons by pyrite in a water column (Sawowicz, Bk 1997; Bk, Sawowicz 2000), it is not impossible that pyrite framboidal aggregates could form through the replacement of silica spherules, similar to those found in precious opals.

ACCUMULATION AND REDISTRIBUTION OF METALS IN FRAMBOIDS


Many elements, like Cu, Zn, As, Sb, Ni, Co, are present in high concentrations in sedimentary rocks, especially in anoxic sediments (Vine, Tourtelot 1970). Typically, a sulfide phase composed mainly of pyrite, usually framboidal, is their main carrier, although there are exceptions. In the authors opinion, concentrations of trace metal in the environment, period of the iron sulfides formation and their transformations or later oxidation seem to be the major factors controlling the content of trace metals in pyrite and in anoxic sediments. Incorporation of trace metals into iron sulfides may proceeds both by coprecipitation and adsorption (Cottnam et al. 1999). Accumulation of trace metals Formation of framboids from gel and their extremely large effective surface make scavenging and adsorption of the chalcophile elements during deposition and diagenesis very efficient. The surface of framboid is much larger than comparable euhedra. When the cube of a solid with 1-cm sides is subdivided to form cubes that are 1m edge length and again agglomerated, the overall surface area of the interface, for the same volume of the solid, is four orders of magnitude greater. In a series of experiments (Watson, Ellwood 1994; Watson et al. 1995) nearly all heavy and transition metals tested were preferentially adsorbed onto the bacterial monosulfide. It is thought that in anoxic sediments trace metals, e.g. As and Co, are concentrated with Fe-sulfides by co-precipitation (Jacobs, Emerson 1985; Landing, Lewis 1991; Soma et al. 1994). Graham and Robertson (1995) found that arsenic is especially concentrated in pyrite framboids. Magnetic and electrostatic properties of pyrite precursors and pyrite itself may also be of major importance for trace metal accumulation. Iron

23

monosulfides are ferromagnetic and, on the other hand, pyrite, as the semiconductor, posseses both cathodic and anodic sites. In pyrite substitution of either Ni or Co for Fe results in n-type conductivity (Shuey 1975) and the substitution of As for S results in p-type conductivity controlling gold precipitation (Moeller, Kersten 1994). It seems that precious metal accumulation and precipitation can be influenced by the semiconductivity of the specific texture, e.g. the cobaltite-framboidal pyrite texture (see Fig. 11C). The significance of these properties is well seen in the case of precious metals. Watson and Ellwood (1994) demonstrated that bacterially precipitated FeS can adsorb a variety of platinum group elements and gold. Concentration of metals in pyrite framboids also depends on concentration of metals in the environment of deposition. Heavily polluted sediments usually contribute significantly to the metal content in sulfide fraction. Luther et al. (1980) found up to 4% of Ni in pyrite framboids and high contents of Mn in iron monosulfide framboids from the polluted sediments of Newark Bay (USA). In these sediments, when the content of Zn is below 1000 ppm, it substitutes iron in framboids but when its content is higher it forms zinc sulfide framboids. Sometimes framboids may be the sites where the concentrations of Co and/or As were elevated to allow local saturation with respect to cobaltite only in the framboid, like in the case of pyrite framboids cemented by cobaltite at the boundary between the Kupferschiefer and the underlying Weissliegendes (Large et al. 1999) (Fig. 11B-D). It is not impossible that pyrite framboids enclosed in other metal sulfides crystallized from gels of complex composition (Rosentsvit, Epshteyn 1963). Duration of the scavenging and adsorption processes, together with the concentration of available trace metals in the solution may control the final concentration in sulfide fraction of the sediment. One can expect that pyrite, often as pyrite framboids, formed relatively rapidly in the water column, would contain less trace metals than pyrite formed during early diagenesis. The indirect evidence comes from the Cenomanian/Turonian black shales from Cismon, Italy (Sawowicz 1996 and in prep.) The high enrichment of some elements (enrichment factors: 10-50 for Cd, Ag, Mo and 5-10 for Cu, Zn, V, As) in these shales is similar to the other Cretaceous settings (Brumsack 1986) but correlates well with TOC (total organic carbon) content and not with sulfide content. Iron sulfides from the Cismon section formed probably in the water column, faster than in reducing sediments. Shorter time of adsorption, lower concentration of trace metals in seawater compared to prolonged release of these elements from other (both organic and inorganic) fractions of the sediment, was a possible reason for lower content of siderophile elements in the sulfide fraction. In addition, a content of metals coprecipitated with syngenetic pyrite may be diminished due to expulsion of some metals during recrystallization of precursor iron sulfide. In a water column these metals would be recycled, whereas in organic-rich sediment would be immobilized. It is interesting to note that specific organic compounds, often related to pyrite framboids, may enhance metal accumulation and subsequent coprecipitation with iron sulfides. Williams and Cusack (1996) suggested that glycocalyx would have lined the Lingulid shell interior and adsorbed dissolved metals which are abnormally concentrated in abundant framboids filling the shell interior. Redistribution of trace metals The present author postulates two different processes leading both to enrichment or depletion in trace metals in sulfide framboids: transformation (recrystallization) and oxidation.

24

Transformation The temporary nature of the adsorptive process and the desorptive expulsion of trace metals during the monosulfide to pyrite transition have been proposed by many authors, e.g. Morse and Arakaki (1993). During the transition, which typically takes place in several stages (Wilkin, Barnes 1997a), a desorption of the adsorbed metals is expected as the structure becomes more ordered and surface area is reduced. The reported concentrations of adsorbed metals in bacterially produced FeS (Watson, Ellwood 1994) are considerably greater than the concentrations reported in framboidal pyrite (Raiswell, Plant 1980; Morse et al. 1987; Landing, Lewis 1991). High contents of manganese were probably lost in the transition of Fe3S4 to FeS2 in polluted sediments of Newark, USA (Luther et al. 1980). The authors opinion is that in the case of relatively low concentration of trace metals in the environment of deposition, as in normal sediments, they can be probably present in sulfide fraction throughout all mineralogical and textural stages of pyrite formation. However, when their concentration is high, like in environment of some sediment-hosted ore deposits, they may be expelled during transformation and form specific textures, built of different metal sulfides. Recrystallization of framboids to euhedra and changes in available surface during growth and transformation can lead to the extrusion of trace metals. In coal seams of Nova Scotia (Canada) arsenic is mostly associated with framboidal and not euhedral pyrite (Mukhopadhyay et al. 1998). Korolev (1958) showed experimentally that 70-96 wt.% of molybdenum was coprecipitated from solutions by iron sulfides but the coarsely crystalline pyrites had only minor amounts of Mo due to its loss in aging process. He also suggested that similar mechanism could be important for Ni and Co. In anoxic marine environments Co is associated with Fe-sulfides and, basing on chemical equilibrium considerations, it is thought that the Co is most likely adsorbed by the Fe-sulfides rather than co-precipitated as a discrete Co-S phase (Jacobs, Emerson 1985; Landing, Lewis 1991). The example of possible Co expulsion and precipitation as cobaltite during transformation from iron monosulfide to pyrite comes from the copper deposits of the Kupferschiefer in Poland (Fig. 11B-D). Large et al. (1999) suggested that metal expulsion from Co-As-rich FeS could result in high trace metal concentrations in the vicinity of the framboid and this could influence ore mineral nucleation. High concentrations of Co and/or As desorbed during this transformation to pyrite could attain locally high concentrations in the framboid and result in cobaltite being precipitated within the framboid. Precipitation of cobaltite under these conditions would depend on cobaltite solubility, the rates of Co and As desorption, availability of sulfur and the rate of diffusion of chemical species to and from the microenvironment of a future framboid. Dill and Kemper (1990) suggested that changes of morphology, from the cube to the octahedron in the Fe bisulfide aggregates, also led to a continuous removal of As and Ni from the pyrite. It must be stressed that the difference between specific surface area of primary colloidal iron monosulfides, through the whole size range of iron sulfide framboids, and of euhedral crystal can be enormous. Comparison between surface areas of hypothetical framboid, built of succesive stages of micro- and nanoframboids, and euhedral cubic crystal of the same diameter gives values of several orders of magnitude. Oxidation Framboidal pyrite oxidation is common (Luther et al. 1982). Passier et al. (1999) estimated that in the Mediterranean sapropels the percentage of initially formed iron sulfides that were re-oxidized varies from 34 to 80%. Mssbauer studies of pyrite oxidation showed that pyrite transforms first to szomolnokite, then oxidizes to lepidocrocite and ages finally to goethite (Huggins et al. 1980).

25

In the Cretaceous/Tertiary (K/T) boundary layers, and in other stratigraphic horizons, framboids composed of iron oxides (primarily iron sulfides) are highly enriched in noble metals (up to 50 times) and other trace elements, relative to their non-oxidized counterparts (Brooks et al. 1985; Sawowicz 1993b). The authors opinion is that oxidation of sulfides in the K/T boundary layer is one of the main processes of redistribution, leading both to concentration and dispersion of specific trace elements. It could result both from oxidation from the overlying water column after restoring oxidizing conditions (e.g. Italian and Spanish sections; Montanari 1991) and/or upward transfer of fluids during later diagenesis (e.g. Danish sections; Schmitz 1985). In the first case formation of metastable metal-rich layer is possible (Pruysers et al. 1991), whereas in the other a long-term circulation of diluted solutions may lead to significant metal enrichment at the redox boundary. The detailed authors study of the Spanish K/T boundary sections (Martinez-Ruiz et al. 1995) showed that two types of spherical aggregates of goethite are present. The first type is represented by typical framboidal aggregates (Fig. 8E, F). The aggregate size ranges from microframboids, to framboids, to polyframboids. The presence of all these framboidal forms suggest their protected (by organic matter?) growth and a very large surface available for metal adsorption (Sawowicz 1993a). The matrix of framboids is significantly enriched in titanium and may be related to replacement processes of a precursor material. The second type is represented by concretions composed of closely intergrown octahedral crystals. Framboidal texture is usually associated with pyrite (and its precursor greigite) and such precursors of goethite are generally suggested in the case of K/T boundary layer but local replacement of framboidal magnetite (occurring in meteorites; Keller et al. 1994) or magnesioferrite of volcanic origin (Taylor 1982) is not excluded. Framboids seem to form earlier during diagenesis than microconcretions. Spherulic form of framboids is typically the intrinsic feature of their authigenic growth but may also result from replacement or refilling of glass spherules or organic remains. The occurrence of pyrite on the surface of some goethite spherules suggests a local resulfidization event during late diagenesis, related perhaps to a decomposition of sulfur-rich organic matter.

FRAMBOIDS AND FOSSILS


Pyritized organic remains are common in the sedimentary record. Among many occurrences are: molds of diatoms (Geroch 1978; McNeil 1990); empty spaces and/or replacements of carbonates in echinoderms (Jensen, Thomsen 1987), ammonites (Wiedman, Neugebauer 1979; Hudson 1982), gastropods and bivalves (Fisher 1986) or whale bones (Bang 1994); delineations of tubes and burrows of polyachetes (Thomsen, Vorren 1984); cells in fossil plants and woods (MartinPenela, Barragan 1995; Garcia-Guinea et al. 1998) and replacements of soft-bodied trilobites (Briggs et al. 1991). Pyrite has also been found in recent foraminifers (Seiglie 1973). Pyrite can adopt various forms, from massive to aggregated, euhedra and framboids. Generally pyrite replaces the organic matrix (soft parts) and carbonate skeletons during all stages of sediment burial history, but may probably also replace silica skeletons during their journey through the anoxic water column. Descriptions of mechanisms of fossil pyritization were given by Canfield and Raiswell (1991), Briggs et al. (1996), Raiswell (1997), Bk and Sawowicz (2000) and Grimes et al. (1999). Here only pyritization of fossils in relation to framboidal pyrite will be discussed. The style of pyritization is determined mainly by concentrations of dissolved iron and sulfide. Optimum conditions for replacements of soft-parts are met in near-surface sediment during iron-dominated stage of diagenesis, with relatively slow bacterial sulfate reduction (Canfield, Raiswell 1991; Raiswell 1997). Formation of framboids in the organic remains depends mainly on availability of organic matter, necessary for bacterial sulfate reduction,

26

and on available iron. There are probably two sources of organic matter. One is the post mortem in situ decomposition of organic matter that forms the living organism, including symbiotic bacteria. The other is the organic matter of the surrounding sediment. The reduction of sulfate occurs mainly at the site of the decomposing organic carbon both in anoxic marine sediments (Berner 1980) and in the water column (e.g. Black Sea - Muramoto et al. 1991). It seems that totally reducing sedimentary environment is not necessary. Restricted, minimally aerobic environments are probably optimum for pyritization of fossils (Curtis 1980; Fisher, Hudson 1985). Different styles of fossil pyritization can be distinguished: a) pyritization of soft parts, b) pyritization of skeletons (carbonates, silica, apatite), and c) pyrite infilling of free spaces (pores, voids, whole skeletons) in fossils. The latest style may not always be genetically related to fossilization, if fossils serve only as available spaces in sediment. Allison (1988) recognized the following modes of soft-part preservation by pyrite, where framboidal pyrite is common: 1/ permineralized tissues, 2/ mineral coats, 3/ mineral casts or molds. In pyritization of fossils framboidal pyrite is found practically only in free-spaces inside skeletons or tissues and replacing degradable organic matter. In the authors opinion the most interesting, from the genetic point of view, are closely packed framboids infilling skeletons or, when skeleton had been dissolved, forming casts (Fig. 12A, B). Such forms are very similar to polyframboids. However, in a typical polyframboid, the sizes of the individual framboids, as well as of the crystals building framboids, often vary. In contrast, in fossils both closely packed framboids and their microcrystals are very uniform in size. This authors interpretation is that they are formed in partly isolated microenvironment and probably in decomposing organic matter (protoplasm) what altogether create more homogeneous and stable environment of precipitation. Organic matter in tests is probably responsible for ofter formation of framboids in tests but not massive and/or structureless pyrite. Williams and Cusack (1996) suggested that the formation of kaolinite within Lingulid shells and pyrite framboids in the infilling of these shells was mediated by different compounds of organic matter, present in different parts of the decomposing body. This authors work shows that the degree of recrystallization of pyrite framboids to euhedra can vary from one to the other part of carbonate shell (Fig. 6C), due to the different carbonate texture, admixtures of other elements and organic compounds. Pyrite framboids often occur together with other morphological forms of pyrite in the same specimen. It probably could be explained by different position in the fossil, very localized organic matter, porosity, several stages of pyritization, different assemblages of bacteria, etc., but no detailed studies have been done so far. In some cases a restriction on the development of the different pyrite forms is based on available space with organic matter (Thomsen, Vorren 1984). Also framboids themselves can occur in very different positions and aggregations. On the other hand, sometimes casts of various species in the same sample reveal different pyrite morphologies, e.g. framboidal and structureless. The author is currently working on this subject and only preliminary suggestions can be made. Morphology seems to be dependent on time of pyritization, type of organic matter (more or less available for BSR, modifier of the microenvironment), porosity and openess of skeleton, and many others. Among the examples of the framboids occurrences are pyritized radiolarian skeletons (Sawlowicz & Bak, 1997). They typically occur in two different positions: (1) in channels (pores) (Fig. 12C); (2) inside the skeleton, attached to an internal surface (Fig. 12D). The size of framboids is around 5 m. It is interesting to point that the average size of crystals in the framboids is similar to that of the silica and pyrite grains, building or replacing the radiolarian skeleton. Sometimes the framboids contain silicate minerals in the interstices. Pyrite framboids present in radiolaria were probably formed after the pyritization of skeletons. They occur in free spaces within the skeletons and could form during diagenesis (even late diagenesis if only BSR was active) of the sediment. Bk and Sawowicz (2000) have used

27

such mode of framboids occurrence as evidence for earlier, i.e. in a water column, replacement of silica skeleton by pyrite, before the formation of framboids. Interesting is the relationship between pyrite and bacteria. Pyrite can replace dead symbiotic bacteria living on fossils or is found inside of living magnetotactic bacteria. In the first case, pyrite often forms linings or encrustations on or within skeletons. In the second case, discrete, organized iron sulfide particles occur in spherical multicellular bacteria of 3-8 m in size (Mann et al. 1990 a, b; Farina et al. 1990), sometimes resembling framboids. It is highly improbable that iron sulfide grains could be closely packed in living bacteria but one can imagine infilling of cells by iron sulfides after the bacteria death, retaining its primary shape. It is also possible that after the bacteria death, released pyrite grains can form framboidal aggregate in sediment, especially at the greigite stage with its magnetic properties (Taylor 1983). Magnetotactic bacteria and their magnetofossils (with magnetite and greigite) can significantly contribute to the magnetic properties of water sediments (Lovley et al. 1987). Microorganisms can also directly influence the precipitation of sulfides. Beveridge and Fyfe (1985) have shown that bacterial components, e.g. cell walls, facilitate mineralization by sequestration of metallic ions from solution and provide local sites for nucleation and growth. Pyrite spherules (framboids) may often be misinterpreted in some paleontological studies. This author considers that especially in older papers, when pyrite framboids were not well recognized, their similarity to bacterial colonies and their spherical shape caused that they were taken for microfossils (Love 1957; LaBerge 1967). In papers dealing with ore deposits where pyrite framboids are common, some authors still use the Schneiderhoehns (1923) term vererzten Bakterien (ore-mineralized bacteria). Growth of pyrite, especially transition from framboid (polyframboid) to euhedral crystal, can modify morphology of different species, e.g. spore exines (Neves, Sullivan 1964). Martill and Unwin (1997) have shown that so-called blood corpuscles from blood vessels of an archosaurian limb bone (perhaps also from dinosaur bones) are pyrite framboids. It is worth noting that spherical aggregations may not necessarily be built of pyrite but can be of iron oxides or hydrooxides, forming pseudomorphs after, or rims on, former framboidal pyrite. Goethite framboids studied by the author (Martinez Ruiz et al. 1995), common at the famous clay layer from the K/T boundary, were sometimes described as extraterrestial or organic origin. A detailed description of pyrite occurring in organic remains may be helpful in explaining the diagenetic history of the remains and surrounding sediments. Jensen and Thomsen (1987) demonstrated that framboids present in secondary voids within the trabeculae of echinoderms had to be proceeded by processes of disintegration and of dissolution. Brett and Baird (1986) proposed a pyrite taphofacies model, which describes formation of different textural forms of pyrite in relation to sedimentation rate and oxygen fugacity. Disseminated framboidal pyrite occupies here the most anoxic and the lowest rate of sedimentation field, whereas pyritized fossils, like steinkerns, occupy the field of dysaerobic and moderate accumulation rate. Layers of pyritic nodules record burial episodes of organic matter in environments characterized by moderate background accumulation of organic-poor sediments. The common association of carbonate concretions with framboidal pyrite indicates that both are produced concurrently in the zone of anoxic sulfate reduction (Brett, Baird 1986). Canfield and Raiswell (1991) suggested that as soft-part pyritization is not expected under euxinic conditions, its presence or absence might be used to distinguish environments with poorly oxygenated bottom waters from euxinic basins. The presence of pyrite, especially fine-grain pyrite framboids, in organic remains may cause different processes of a degradation of the latter, both as the results of the growth of framboids and their oxidation. Bang (1994) found skeleton elements destroyed by the oxidation of pyrite framboids.

28

PYRITE FRAMBOIDS AND THE ORIGIN OF LIFE


Analogies between crystallization and fundamental living processes stimulated the opinions that mineral-organic interactions are the key for the synthesis and evolution of complex organic molecules (Degens 1979; Cairns-Smith 1982; Nakazawa et al. 1993; Kaschke et al. 1994). Hartman (1975) put forward the idea that earliest life was based on iron/sulfur chemistry. De Duve (1991) stresses the possible role of thioesters as the sources of energy for metabolism of primitive cells. One of the most interesting recent theories that life originated in hydrothermal systems was stimulated by earlier discoveries of sulfur metabolizing archaebacteria and black smoker hydrothermal vents on the ocean ridge. Wchtershuser (1988a, 1990) proposed that the growth of pyrite provided both energy and was a substrate for the first chemo-autothrophic metabolism. During the formation of pyrite from iron monosulfide and hydrogen sulfide, simple molecules could have been reduced to more complex organic molecules. Developing that idea Russell et al. (1993) suggested that abiogenic production of organosulfide compounds took place within hydrothermal convection system operating in the Mg-Si-rich crust, as the result of Fischer-Tropsch-like reductive or polymerising syntheses. Different deep-sea environments (Fig. 4), usually rich in iron sulfides, may be regarded as suitable for the above reactions: A/ Corliss et al. (1981) suggested that black smokers hot (up to 360oC) springs along the ocean spreading centres could have been ideal places for origin of life (Fig. 4A); B/ Cairn-Smith et al. (1992) proposed that life originated in deep ocen sulfide mounds, away from spreading centres, where medium temperature (200oC) and high pressure alkaline hot springs waters entered the mildly acid Hadean ocean. Such sulfide mounds could be considered as chemical reactors (Corliss 1990) and have lasted much longer than black smokers, probably of the order of a hundred thousand years (Russell 1983). There, iron sulfide membrane formed botryoids, which were reproducing culture and catalytic chambers or protocells (Russell et al. 1993, 1994). Inside such membranes of growing probotryoids also iron sulfide framboids may grow and recrystallize, prolonging the activity of the whole system. Thus, in the authors opinion one could envision two subsequent steps: 1st - formation of membrane, concentration of ferredoxins and other organics (Russell et al. 1993, 1994); 2nd - when inflation of membrane stops and it recrystallizes, production of life within framboids in the botryoids (Fig. 4B); C/ Jrgensen et al. (1990) described the patches of hydrothermal emissions near black smokers on the 2000m deep sea floor. The temperature gradient was up to 124oC in 75 cm of sediment depth. The patches were black from metal sulfide precipitation. Such parts of sediment can be very rich in framboidal pyrite (Fig. 4C). Pyrite framboids are abundant in all of the above mentioned environments. As framboid formation may be related to gas bubbles, degassing of hydrothermal solutions could be the reason for the formation of billions of framboids inside the smoker or mound structures. Ephemeral on geological scale existence of hydrothermal vents may only be the first stage in the life creation. After the vent death, pyrite and its precursors do not loose any of their potential during diagenesis, changing their own and organic-mineral associations. Such deepsea environments would also be ideal for the earliest organisms acquiring their energy by continuously forming and dumping pyrite (Wchtershuser 1988a, b). The theory of life formation via iron sulfides is presently regarded as one of the most fascinating hypotheses and its thermodynamic basis is relatively well established. In the present authors opinion the iron sulfide framboids should play a crucial role in this theory, as their microenvironments potentially supply all necessary ingredients and are the most convinient places to begin the living story of the Earth. This role is discussed in details in the

29

chapter below. Similarities between iron sulfide framboids and many organic forms are stupendous and innumerable. For example, in Foxs experiments hot (140oC) solutions of proteins that cooled slowly to about 70oC spontaneously formed microspheres 1-2 m in diameter, known as proteinoids (Dott, Prothero 1994). Woese and Wchtershuser (1992) wrote about the origin of life scenario energy source, production of reactive compounds, and condensations to form complex organic structures, occupy the same locale. The iron sulfide (pyrite) framboid is an ideal locale, being the complex site where combined action of clay minerals, iron sulfides and organics can be met, creating the most favorable conditions for life creation.

Fig. 4. Different deep-sea environments of the framboid formation related to the origin of life (for explanation see text) (not to scale). The authorss opinion is that when the origin of life in relation to sulfides is discussed, the following features of pyrite framboids can be taken into account: 1/ size of microcrystals in framboid or size of framboid, varying from 0.1 m to 500 m; 2/ hierarchical fractal-like structure of framboids, from micro- to polyframboids; 3/ various shapes of grains or crystals - spherules, cubes, octahedra, dodecahedra and their modifications; 4/ possibility of transformation from the framboid to the crystal of the same size; 5/ often identical size of billions of microcrystals in huge masses of framboids or framboidal aggregations; 6/ regular arrangement of crystals, on small to large scale; 7/ modification of the morphology of minerals filling interstices in framboids, during growth and arrangement of pyrite microcrystals; 8/ the occurrence of annular (hollow or filled with other minerals or organic matter) framboids; 9/ transformation from FeS to FeS2 in framboids; 10/ high contents of life-important metals in iron sulfides and/or their extrusion during the framboid crystallization; 11/ morphological and textural similarities to the simplest organic and inorganic forms occurring in nature; 12/ different physico-chemical interactions between microcrystals and on their surfaces in framboids, e.g. electrostatic forces, Van der Waals
30

attractions, etc.; 13/ active centres in specific positions in the framboid; 14/ framboid formation from gel; 15/ oxidation of sulfide framboids and resulfidization of oxide framboids. Role of pyrite framboids The role of framboids in the formation of life on the Earth is limited to the earliest stages of the life formation on the Earth, which were necessary to begin the creation of life. The key term origin of life used in the further chapters has that meaning. The author considers that iron sulfide framboids can influence the origin of life as: 1) source of energy (electrons) and catalythic action (substrate for chemoautotrophic metabolism); 2) accumulation and reaction chambers; and 3) templates, facilitating reproduction and passing information. Energy and catalysis Wchtershuser (1988a) developed the theory that the exergonic formation of pyrite from hydrogen sulfide and ferrous ions provided the first energy source for life. The metabolic role involves the reduction of carbon dioxide to form organo-carbon exergonic compounds, where reducing power and energy is derived from the redox conversion of iron monosulfide and hydrogen sulfide to form pyrite. FeS-H2S system may be an alternative source of hydrogen, the important agent for the development of the earliest life on the Earth. Drobner et al. (1990) suggested that a functional evolutionary connection might exist between the hydrogenproducing system FeS-H2S and hydrogen-producing iron-sufur centres of hydrogenases and nitrogenases. In the earlier photosynthetic bacteria the electron donor was probably H2S (Hall et al. 1971). Stumm and Sulzberger (1992) proposed that photosynthetic processes occurring on some inorganic Fe-bearing surfaces (semiconductors) and with iron species may be looked at as primitive alternatives or precursors to biological photosynthesis. Experiments with chemotactic interaction of Thiobacillus sp. with sulfur and with pyrite framboids showed that bacterial activity liberates excess sulfur or sulfide carrying species which maintain bacterial life in the surrounding medium (cooperative leaching) (Rojas-Chapana et al. 1998). These authors suggested that such chemotactic behaviour might have developed as an adaptive advantage. Pyrite is hydrophobic, therefore organic molecules should easily be adsorbed on its surface, as it is commonly observed during flotation procedures. Hall et al. (1971) suggested that ferredoxins were the first enzymes where their precursors could have been an iron monosulfide/disulfide redox catalyst. Wchtershuser (1988b, 1990) proposed that the surface of the growing pyrite could act to catalyse the reactions leading to the formation of complex organic phases and surface metabolists because slightly positively charged surface of pyrite is able to attract and bind any negatively charged compounds, like phosphates, carbonates or organic molecules. These molecules could wander around the surface, mingle and form longer chains, bond to the pyrite surface by multiple ionic bonds. Such interacting molecules could autocatalytically create subsequent molecules and finally replicate themselves, similarly to a simple metabolism. Studies of Stumm and Sulzberger (1992) showed that reactions in Fe-Sorganics systems are very complex. Solid and solute Fe (II) complexes with sulfides and other minerals are very efficient reductants from thermodynamic and kinetic point of view. Electron cycling of Fe (II) Fe (III) transformations can occur over the entire Eh range within the stability of water due to complex reactions of these ions with organic and inorganic ligands. The oxidation of Fe (II) to Fe (III) provides a possible means for reducing nitrites to ammonia (Summers, Chang 1993). For catalysis it is important that framboids offer enormous specific surface in comparison to euhedral pyrite crystals. Crystals building typical pyrite framboid have the surface an order

31

of magnitude bigger than euhedral cubic crystal of the same size. If framboid is composed of smaller microframboids which in turn are built of microcrystals (Sawowicz 1993a) (Fig. 1), their total surface may grow up even further. However, it must be stressed that a great number of possible surfactans occurring in nature can easily and unpredictably modify catalytic properties. Occassionally, even a 1% of surface coverage may alter the properties of a particular surface in a drastic manner what means that, on one hand, even hardly imaginable hypotheses regarding the origin of life could be fulfilled, but on the other hand, all theoretically available or restricted reactions may be difficult to apply. Accumulation and reaction chambers Pyrite framboids and their precursory gels (from iron oxyhydroxides to monosulfides) can efficiently accumulate minerals, metals and organic molecules (Stumm 1993). An interesting feature of iron sulfide droplets is that they can have pH and/or Eh different from the surrounding water. The occurrence of organic substances in mineral coacervates may be importan especially as one of the ways of collective concentration of abiogenic organic compounds in ancient hydrosphere of Earth. Variety of metastable abiogenic organic molecules is supplied by the hydrothermal systems (Shock 1992a, b). Organic molecules may be both included into or adsorbed on framboid. They can arrive from outer environments or be catalysed in situ. Coacervate droplets of primary Fe(OH)2 can concentrate dissolved organic compounds, usually aminoacids. Aminoacids when locked inside biologically precipitated minerals escape further biological degradation and are preserved. This could explain the abnormal concentration of aminoacids in ancient sulfides (Nardi et al. 1994). Extremely large effective surface of gels and iron monosulfides (also magnetic and electrostatic properties of the latter) make scavenging and adsorption of the chalcophile elements during deposition and diagenesis very efficient (Shuey 1975; Jacobs, Emerson 1985; Watson et al. 1995). Also organosulfur compounds may be adsorbed onto the fresh iron sulfide surface (Leja 1982). Organic compounds could accumulate within the botryoid vesicle as a consequence of chemiosmotic force and the activity of the iron monosulfide redox catalysts within the membrane (Russell et al. 1993). It is worth mentioning that in experiments the iron sulfur membranes have 20-40 times bigger durability when abiogenic organic compounds are added to the intial solutions (Russell et al. 1994). On the other hand, Russell et al. (1990) suggested that a critical concentration of polar organics adsorbed on iron sulfide surface may inhibit growth of iron sulfides and organics could become ordered to a liquid crystalline structure. Russell et al. (1988) proposed that iron sulfide botryoids may be reaction vessels controlling organic reactions. Experimentally produced sulfide chimneys were suggested to be culture chambers for the first life forms (Russell et al. 1989). Complex process of the formation of pyrite framboids create several possibilities of acting as reaction vessels and the following succession is proposed by the present author: 1/ formation of mixed-phase membrane (gas bubble or gel droplet), separating two different environments and allowing specific transfer of components, 2/ accumulation of organics, iron (also other metals) and/or sulfur inside the membrane, 3/ formation of primary iron monosulfides, 4/ step-wise transformation to pyrite. Degens (1979) suggested that the decisive step in the generation of life is the formation of a primordial cell capsid which was initiated by gas bubbles. Surfaces of gas bubbles are phase boundaries that extract hydrophobic components and generate ionic membranes (Degens et al. 1972; Degens 1979). Framboids probably can be formed using gas bubbles (Rickard 1970), although the latter need not be necessary. The gel droplets, built of organics and iron oxy/hydroxides (Kalliokoski, Cathles 1969), also form closed spherical volumes and could

32

support different chemical reactions inside them and on their surface. The boundary between water and gas or gel is very active and can concentrate almost every material, like organic molecules and metal ions. Organic synthesis could also have occurred beneath the membrane using energy of the FeS-FeS2 transformation (Wchtershuser 1988a,b; Russell et al. 1993). Thin organic layer around a framboid could act similarly to a gelatineous iron-sulfide membrane in botryoids suggested by Russell et al. (1993). In the authors opinion, whatever is the origin of the spherical form, framboids have such a form from their very beginning and are ideal precursors of the living cells. The size of framboid or crystal arising from that varies from one to several tens of microns, with average values between 5 to 10 m, being also very similar to the size of organic cell. Moreover, framboids with their two or three scale size range (Sawowicz 1993a) (Fig. 2) could be the very complex compartmentalized reaction vessels. During all stages of framboid formation different elements and/or organic compounds can be accumulated in, reorganized and expelled out of the reaction chamber. Copper, nickel, cobalt and molybdenum are common in iron sulfides. Phosphorus, so important for RNA and DNA formation, and silica are often present in the environment of framboid formation. Organic compounds can influence significantly formation and transformation of iron sulfides in the framboid (Mann 1988; Bianconi et al. 1991; Schoonen, Barnes 1991b). Recystallization of microspherules (microframboids) building framboid or of framboid itself to the euhedral crystals often leads to a release of microelements, making them available for incorporation into new phases or formation of their own minerals (e.g. cobaltite formation in framboids interstices; Large et al. 1999). During the formation of framboids different mineralogical phases can coexist at the same stage and internal and external framboid texture may change. Annular (hollow) framboids (Fig. 2), consisting of a more or less continuous sphere built of minute pyrite grains, sometimes filled with other minerals or organic matter, are known from nature (Kosacz, Sawowicz 1983; Sawowicz 1993a). The experiments of Kizilshtein and Minaeva (1972) have shown that they can be the first step in the development of complete framboids. If the formation of pyrite framboids begins from the surface of a gel globule, we can expect several stages where different pairs core/sphere are present, e.g. iron oxyhydroxides (with organic compounds?) /iron monosulfides or iron monosulfides/pyrite. Empty interiors may also result from later processes of dissolution. Drobner et al. (1990) experiments showed that reaction of H2S with FeS produces pyrite shells on pyrrhotite where the latter can easily be dissolved. Sawowicz (1992) suggested that iron monosulfide cores in pyrite framboids might be dissolved or replaced by other minerals. In any case, different physico-chemical properties of the sphere and the core may create a very unique reaction chamber. Templates The presence of organic compounds of abiotic origin does not suffice for the origin of life. They must reproduce and pass information that can be facilitated using templates (Orgel 1986). Cairns-Smith (1982) proposed that the key to replication is a template on which organic molecules could line up in close order, allowing them to polymerize until they could do so on their own, and such a template could be made by clay minerals. Later Russell et al. (1990) and Cairns-Smith et al. (1992) suggested that also iron sulfides could play a similar role. Among all pyrite textures framboids seem to be best prepared for such a role. All stages of framboid formation, from iron oxyhydroxide gel, to iron monosulfide, to pyrite can be important. The following special features of pyrite framboids serving as templates are proposed by the author: 1/ morphology and attraction sites on the surface may change during

33

formation and diagenetic transformations; 2/ individual components may be regularly ordered; 3/ similarity of components in one framboid and over a few size-scale ranges; 4/ framboid growth may organize substances in interstices. The size and morphology of the active surface of crystals in framboid change during its growth as the result of temperature, pH and saturation changes, admixtures of metals, silica and organic compounds, and the rate of recrystallization (Murowchik, Barnes 1987). Lowering of temperature in sulfide reactor from the hydrothermal spring (about 200oC) or smokers (below 360oC) to seawater temperature, and pH changes at the interface between alkaline hydrotermal solutions and midly acid seawater, could introduce several modifications of the iron sulfide morphology during their growth (Cairns-Smith et al. 1992). They also suggested that the covalent character of iron sulfide crystals and their defect structures make them potential information carriers. Positively charged surfaces of the heavy metal sulfides (FeS, FeS2, CdS) bind the anionic organic compounds (Wchtershuser 1990; Dameron et al. 1989). Dameron et al. (1989) found that some peptides are effective in terminating the growth of CdS nanocrystallites of a specific size and, by analogy, this could be true for iron sulphide crystallites. The crystal surfaces of pyrite have a positive charge that could attract negatively charged organic molecules, such as phosphate backbones from nucleic acids. RNA-like polymer, negatively charged, would tend to stick to the edges of, for example, cube of iron sulfide (pyrite). It is interesting to mention that Chen and Seeman (1991) produced a construction of a closed polyhedral cube-like object from DNA. Possibilities to arrange organic compounds on pyrite surface are even more extensive as semiconductor pyrite posseses both cathodic and anodic sites. Pyrite semiconductor n- or p-type properties may change with impurities content (Shuey 1975) what means that content of trace metals in pyrite may signicantly influence a possible adsorption of organic molecules to different sites on pyrite crystals surface. Pyrite needs not to be a direct template. Lahav and White (1980; in Lazard et al. 1987) proposed the adsorbed template model, where adsorption reaction occurs in two steps: adsorption of a biomolecule onto the surface of a mineral and, further, the adsorption of another biomolecule onto the first molecule. Lazard et al. (1987) presented experimental results showing that a primitive prebiotic replication system capable of supporting information transfer reactions could be based on such a template, e.g. gypsum. Oxidation of pyrite framboids, either superficial or total, is common. It is possible that products of such oxidation, for example sulfates or [FenSn]m+ clusters formed on the surface of FeS (Mueller, Schladerbeck 1985), could also be used as adsorbed templates. Multiple and stable configurations are the sine qua non for information storage (CairnsSmith 1982). Framboids are self-reproducing and could retain and/or pass any information from one stage of growth to the other, being modified on the way. Microcystals in framboid are uniform in size and may be packed regularly. The pattern of regular arrangement of uniform spheres is not unique and can be found also in some viruses, similarly like the tendency towards polyhedral outline, probably as the result of close-packing. Pyrite framboid may play a similar role as a microprobotryoid described by Russel et al. (1993, 1994). The difference with framboids is that in contrary to the transfer of information or material from one bubble to the other (Degens 1979; Russell et al. 1993, 1994) they could replicate themselves at one time and on several size ranges. In the case of framboids there is no need for budding of first cell, separation etc. as we have milions of cells ready to be used. The large range of size scale of framboids would be an ideal place for such purpose. Genetic takeover postulated by Cairns-Smith (1982) could be from iron sulfide framboids to nucleic acids. On the other hand, nucleic acids can replicate. We can imagine protein polymers formed on iron sulfide cubes in a framboid and nucleic acids, which fill the interstices between crystals and assemble protein polymers. The simpler version of nucleic acids is a

34

single-stranded RNA, the genetic material of viruses. There is a striking similarity between the structure of some viruses and framboids. Ordered crystallites in framboid are also similar to phospholipid membranes described by McConnel et al. (1984). Organic membrane on sulfide surface could change its properties along with the growth and changes of its substrate. Later surface metabolism products may dissolve or oxidize the pyrite surface they live on, resulting in the detachment of membrane from pyrite. The process can be more advanced due to recrystallization and expulsion of any material, also preorganized organic molecules, from the interstices of the microframboids that build framboid, into the interstices of the framboid itself. During growth, arrangement of pyrite microcrystals and modification of the morphology of minerals and organic compounds filling interstices in framboids may take place. Organic molecules partly ordered on previous crystals may be desorbed and expelled into the regular intergranular network, joining each other. Later iron sulfide can be easily oxidized and/or dissolved, leaving structural organic form. Even if the environment of framboids was not capable to transfer information, then noninformational systems may have been essential for the creation of an environment in which informational replication could get started (Orgel 1992).

APPLICATIONS
The importance of well-defined fine particles (monodispersed colloids) of different materials has been recognized in numerous applications, as in ceramics, catalysis, pigments, recording materials and many others (Matijevi 1994). Only some points of the use of framboids to solve different problems, also practical, will be raised here. The formation of pyrite framboids typically in a water column or during earliest diagenesis is often use as an argument for determination a time position of other sulfide paragenesis, although it should be stressed that very rarely and under specific conditions framboids can also form during late diagenesis. Replacement of pyrite (or its precursors) framboids by other sulfides and recognition of different stages of pyrite formation related to framboids is helpful in studies of ore deposits and diagenetic history of sedimentary rocks. The size of framboids can be indicative of environment of their formation, especially of redox levels. There is a statistically significant difference in the diameter of framboids, formed in the water column and in the sediment (Wilkin et al. 1996). Studies of framboids distributions in Holocene Black Sea sediments showed that the mean diameter of framboids from the Units I and II is 5 m, suggesting their formation within an anoxic and sulfidic water column, whereas the mean diameter of framboids from the Unit III is 10 m and suggests their formation with anoxic sediment porewaters (Wilkin et al. 1997). Wignall and Newton (1998) studied size distribution of pyrite framboids in British Jurassic black shales and found that euxinic conditions are characterized by tiny framboids (<5 m) with a narrow size range. Dysaerobic facies contained larger framboids with a broader size distribution. The use of these distributions can be limited in the case of sedimentary reworking, e.g. removal of the smallest size fraction (Wignall, Newton 1998). OBrien (1995) used the reorientation of clay flakes around the framboids during burial as the evidence that shale fabric is a diagenetic feature, produced after framboid formation. Framboids that are by far the most common texture of authigenic iron sulfides might be used as paleosalinity indicators (Baldwin, Holdren 1979; Berner et al. 1979). Greigite, a ferrimagnetic mineral common in framboids from young and recent sediments, and goethite (often pseudomorph after pyrite) framboids have recently gain an attention due to its significance for paleomagnetic and magnetostratigraphic studies (Hilton 1990; Roberts, Turner 1993; Reynolds et al. 1994; Belkaaloul, Aissaoui 1997).

35

Non-recognition of a presence of pyrite framboids and their oxidized counterparts, coupled with their close relationship with organic matter and/or replacement or dissolution, have often led to erroneous results related to the determination of non-existent organic species or extraterrestrial materials (e.g. problem of K/T boundary sediments; Martinez-Ruiz et al. 1995). Weathering of pyrite generally creates serious environmental hazards, from contaminating fresh water supplies, through acid leaching of other minerals, to causing geotechnical problems due to expansion caused by crystallization of hydrated iron sulfate. Much rarer oxidation of pyrite can be useful. There are two main weathering pathways, involving bacterial and strictly inorganic processes: 1) aerobic oxidation by oxygen; and 2) anaerobic oxidation by ferric iron. Oxidation of pyrite produces sulfuric acid and metal ions, mainly iron but also some environmentally more dangerous metals like arsenic, which is often present in pyrite as an admixture. One of the most serious hazards for environment is the acid mine drainage (AMD), a process of weathering of sulfide from mine tailings or dumps and flooding of underground mines, where oxidation of pyrite plays the main role. The surfaces of sulfide minerals have a strong affinity for dissolved metals (Jean, Bancroft 1986) and surface reactions may play an important role in controlling metals concentrations in inactive mine tailings impoundments. Studies of pyrite grains showed that attenuation of metals varies significantly between different dissolved metals (e.g. Cu vs As), and strongly depends both on pH of tailings waters and on the formation of coatings of secondary minerals, both oxides and sulfides, under different redox conditions (Al et al. 1997). Framboidal pyrite, which is finegrained, has the available surface, and often the close spatial relationship with organic matter, much bigger than euhedral pyrite and, catalyzing the oxidation reaction, may weather much faster than any other pyrite morphologies. The rate of weathering depends also on the type of bacterial culture. Boon et al. (1999) demonstrated that Leptospirillum were able to oxidize both framboidal and euhedral pyrite whereas T.ferrooxidans oxidized only framboidal pyrite. Zhang and Evangelou (1998) found that formation of ferric hydroxide-silica coatings on framboidal pyrite inhibited pyrite oxidation under acid conditions. Recognition of this problem is important especially for all calculations regarding times and rate of degradation. Although having generally damaging effect, pyrite oxidation can be used for reclamation of sodium-rich soils, leaching of low-grade ores by bacteria, or sorption of heavy metals from solution. Also in this case, the recognition of much easier oxidized framboidal pyrite among pyrite-rich wastes added to the soils is important. Biomagnetic particles of FeS have been used for separation of heavy metals from the solution (Watson, Elwood 1994; Watson et al. 1995) Rather unusual is the use of framboids to diminish the danger of explosion during unloading of the tankers (Walker et al. 1997). Iron rust on the inner surface of oil tanks reacts with H2S forming the mixture of iron sulfides that are pyrophoric. It is possible, by exposure of the very finely divided greigite and pyrite to an atmosphere of nitrogen gas containing admixture of oxygen, to transform these fine particles to framboidal morphology, which is less reactive. The presence of pyrite in museum specimens, as admixtures or pyritized organic remains, can cause serious problems with their preservation. Especially framboidal pyrite, due to its fine-grain contents and often, close spatial relationship to organic matter, is potentially very dangerous for the deterioration of specimens. The mechanism of such deterioration has been described in detail by Bang (1994). The galvanic reation between FeS2 and organic matter, in the presence of moisture, creates acidic conditions through the sulfuric acid formation and leads to mechanical and chemical breakdown of framboids and associated fossil tissues, especially those built of carbonates or apatite.

36

ACKNOWLEDGEMENTS
Author wishes to thank A.P. Gie (Manchester), W. abiski, M. Rospondek and A. Piestrzyski for critical reading of the manuscript. J.Weiner kindly read the chapter on the origin of life. Discussions with M.J. Russell (Glasgow) and R. Raiswell (Leeds) are appreciated. Help of J. Faber, A. atkiewicz, A. Lewandowska and W. Obcowski during preparation of photographs and drawings and editorial handling by A. Skowroski are acknowledged. M. Bk, L. Dejonghe (Liege), D.H. McNeil (Calgary) and M. Michalik kindly supplied some of the specimens.

REFERENCES
AHN J.H., BUSECK P.R., 1990 - Hematite nanospheres of possible colloidal origin from a Precambrian banded iron formation. Science, 250, 111-113. AL T.A., BLOWES D.W., MARTIN C.J., CABRI L.J., JAMBOR J.L., 1997 - Aqueous geochemistry and analysis of pyrite surfaces in sulfide-rich mine tailings. Geochim. Cosmochim. Acta, 61, 2353-2366. ALLISON P.A., 1988 - Konservat-Lagerstaetten: Cause and classification. Paleobiology, 14, 331-344. ALYANAK N., VOGEL T.A., 1974 - Framboidal chalcocite from White Pine, Michigan. Econ. Geol., 69, 697-703. AMSTUTZ G.C, PARK W.C., SCHOT E.H., LOVE L.G., 1967 - Orientation of framboidal pyrite in shale. Mineral. Deposita, 1, 317-321. ARIZTEGUI D., DOBSON J., 1996 - Magnetic investigations of framboidal greigite formation: A record of anthropogenic environmental changes in eutrophic Lake St. Moritz, Switzerland. Holocene, 6, 235-241. BK M., SAWOWICZ Z., 2000 Pyritized radiiolarians from the Mid-Cretaceous deposits of the Pieniny Klippen Belt a model of pyritization in an anoxic environment. Geol. Carpathica, 51, 91-99. BALDWIN T., HOLDREN JR G.R., 1979 - Authigenic iron sulfides as paleosalinity indicators. J. Sediment. Petrol., 49, 1345-1350. BANG B.S., 1994 - Framboidal pyrite and associated organic matrices. A risky composite for preservation of fossils. In: Kejser, U.B. (ed. ), Surface Treatment: Cleaning, Stabilization and Coatings. IIC Nordic Group, Danish Section, XIII Congress, Copenhagen, 66-82. BARD J.P., 1980 - Microtextures des roches magmatiques et metamorphiques. Masson-Paris, p.192. BARTHOLOME P., KATEKESHA F., LOPEZ RUIZ J., 1971 - Cobalt zoning in microscopic pyrite from Kamoto, Republic of the Congo, (Kinshasa). Mineral. Deposita, 6, 167-176. BK M., SAWOWICZ Z., 2000 - Pyritized radiiolarians from the Mid-Cretaceous deposits of the Pieniny Klippen Belt a model of pyritization in an anoxic environment. Geol. Carpathica, 51, 91-99. BELKAALOUL N.K., AISSAOUI D.M., 1997 - Nature and origin of magnetic minerals within the Middle Jurassic shallow-water carbonate rocks of the Paris Basin, France: implications for magnetostratigraphic dating. Geophys. J. Int., 130, 411-421. BENNING L.G., WILKIN R.T., KORNHAUSER K.O., 1999 - Sulfate-reducing bacteria and mackinawite stability. In: Ninth Ann. V.M. Goldschmidt Conf., p.26. LPI Contr.No.971, Houston. BERNER R.A., 1967 - Thermodynamic stability of sedimentary iron sulfides. Amer. J. Sci., 265, 773-785.

37

BERNER R.A., 1969 - The synthesis of framboidal pyrite. Econ. Geol. 64, 383-384. BERNER R.A., 1970 - Sedimentary pyrite formation. Amer. J. Sci., 268, 1-23. BERNER R.A., 198 - Early Diagenesis: A Theoretical approach. Princeton Univ.Press, 241pp. BERNER R.A., 1984 - Sedimentary pyrite formation: an update. Geochim. Cosmochim. Acta, 48, 605-615. BERNER R.A., BALDWIN T., HOLDREN G.R., 1979 - Authigenic iron sulfides as paleosalinity indicators. J. Sediment. Petrol., 49, 1345-1350. BETHKE C.M., MARSHAK S., 1990 - Brine migrations across North America the plate tectonics of groundwater. Ann. Rev. Earth Planet. Sci., 18, 287-315. BEVERIDGE T.J., FYFE W.S., 1985 - Metal fixation by bacterial cell walls. Can. J. Earth Sci., 22, 1893-1898. BIANCONI P.A., LIN J., STRZELECKI A.R., 1991 - Crystallization of an inorganic phase controlled by a polymer matrix. Nature, 349, 315-317 BOESEN R.A., POSTMA D., 1988 - Pyrite formation in anoxic sediments of the Baltic. Am. J. Sci., 288, 575-603. BOON M., BRASSE H.J., HANSFORD G.S., HEIJNEN J.J., 1999 - Comparison of the oxidation kinetics of different pyrites in the presence of Thiobacullus ferrooxidans or Leptospirillum ferrooxidans. Hydrometallurgy, 53, 57-72. BRETT C.E., BAIRD G.C., 1986 - Comparative taphonomy: a key to paleoenvironmental interpretation based on fossil preservation. Palaios, 1, 207-227. BREWER P. G., SPENCER D. W., 1974 - Distribution of some trace elements in Black Sea and their flux between dessolved and particulate phases. In: Degens E. T. & Ross D. A. (Eds.), The Black Sea Geology, Chemistry and Biology. Amer. Assoc. Petroleum Geol., 2483-2490. BRIGGS D.E.G., BOTTRELL S.H., RAISWELL R., 1991 - Pyritization of soft-bodied fossils: Beechers Trilobite Bed, Upper Ordovician, New York State. Geology, 19, 12211224. BRIGGS D.E.G., RAISWELL R., BOTTRELL S.H., HATFIELD D., BARTELS C., 1996 Controls on the pyritization of exceptionally preserved fossils: an analysis of the Lower Devonian Hunsrueck Slate of Germany. Amer. J. Sci., 296, 633-663. BROOKS R.R., HOEK P.L., PEEVES R.D., WALLACE R.C., JOHNSTON J.H., RYAN D.E., HOLZBECHER J., COLLEN J.D., 1985 - Weathered spheroids in a Cretaceous/Tertiary boundary shales at Woodside Creek, New Zealand. Geology, 13, 735-740. BRUMSACK H. J., 1986 - Trace metal accumulation in black shales from the Cenomanian/ Turonian boundary event. In: Walliser O. (Ed.), Global Bio-Events. Lecture Notes in Earth Sciences, vol. 8,. Springer-Verlag Berlin - Heidelberg, 337-343. BUSH R.T., SULLIVAN L.A., 1999 - Pyrite micromorphology in three Australian Holocene sediments. Austral. J. Soil Res., 37, 637-653. BUTLER I.B., RICKARD D., 1999 - Framboidal pyrite formation by the oxidation of iron(II) monosulfide by hydrogen sulfide. In: Ninth Ann. V.M. Goldschmidt Conf., p.44. LPI Co tr.No.971, Houston. BUTLER I.B., RICKARD D., OLDROYD A., 1999 - Formation of the thiospinel, greigite, through the oxidation of iron(II) monosulfide in the presence of aldehydes. In: Ninth Ann. V.M. Goldschmidt Conf., p.46. LPI Contr.No.971, Houston. CAIRNS-SMITH A.G., 1982 - Genetic Takeover and the Mineral Origins of Life. Cambridge Univ.Press. CAIRNS-SMITH A.G., HALL A.J., RUSSELL M.J., 1992 - Mineral theories of the origin of life and an iron sulfide example. Orig. Life Evol. Biosph., 22, 161-180. CANFIELD D.E., 1989 - Reactive iron in marine sediments. Geochim. Cosmochim. Acta, 51,

38

645-659. CANFIELD D.E., BERNER R.A., 1987 - Dissolution and pyritization of magnetite in anoxic marine sediments. Geochim. Cosmochim. Acta, 51, 645-659. CANFIELD D.E., RAISWELL R., 1991 - Pyrite formation and fossil preservation. In Taphonomy: Releasing the Data Locked in the Fossil Record (P.A.Allison & D.E.G.Briggs eds.), Topics Geobiol., 9, 337-387. CANFIELD D.E., LYONS T.W., RAISWELL R., 1996 - A model for iron deposition to euxinic Black Sea sediments. Amer. J. Sci., 296, 818-834. CHAUCHAN D.S, 1974 - Diagenetic pyrite from the lead-zinc deposits of Zawar, India. Mineral. Deposita, 9, 69-73. CHEN T.T., 1978 - Colloform and framboidal pyrite from the Caribou deposit New Brunswick. Can. Mineral., 16, 9-15 CHEN J., SEEMAN N.C., 1991 - Synthesis from DNA of a molecule with the connectivity of a cube. Nature, 350, 631-633. CORLISS J.B., 1990 - Hot springs and the origin of life. Nature, 347, 624. CORLISS J.B., BAROSS J.A., HOFFMAN S.E., 1981 - An hypothesis concerning the relatioship between submarine hot springs and the origin of life on Earth. Oceanol. Acta. Proc. 26th Intern. Geol. Congress, Paris, 59-69. COTTNAM C.F., BUTLER I.B., RICKARD D., 1999 - Trace metals in sedimentary iron sulfides: preliminary results. In: Ninth Ann. V.M. Goldschmidt Conf., p.62. LPI Contr.No.971, Houston. COUTURE C., MAVROCORDATOS D., ATTEIA O., PERRET D., 1998 - The genesis and transformation of organo-mineral colloids in a drained peatland area. Phys. Chem. Earth, 23, 153-157. CRAIG J.R., VOKES F.M., SOLBERG T.N., 1998 - Pyrite: physical and chemical textures. Mineral. Deposita, 34, 82-101. CURTIS C.D., 1980 - Diagenetic alteration in black shales. J. Geol. Soc. London, 137, 189194. DAMERON C.T., REESE R.N., MEHRA R.K., KORTAN A.R., CARROLL P.J., STEIGERWALD M.L., BRUS L.E., WINGE D.R., 1989 - Biosynthesis of cadmium sulfide quantum semiconductor crystallites. Nature, 338, 596-597. DAMKE H., HENNING K.H., LEHMANN J., KASBOHM J., PUFF, T., 1999 - Phase composition of flood sediments of the German-Polish Odra river immediately after the flood event in 1997. Acta Hydrochim. Hydrobiol., 27, 357-363. DARROGH P.J., GASKIN A.J., TERRELL A.J., SANDERS J.V., 1965 - Origin of precious opal. Nature, 209, 13-16. DE DUVE C., 1991 - Blueprint for a cell: the nature and origin of life. Neil Patterson Publs., Burlington, North Carolina. DEGENS E.T., OKADA H., HONJO S., HATHAWAY J.C., 1972 - Microcrystalline sphalerite in resin globules supended in Lake Kivu, East Africa. Mineral. Deposita, 7, 112. DEGENS E.T., 1979 - Primordial Synthesis of Organic Matter. In: B.Bolin, E.T.Degens, S.Kempe & P.Ketner (Eds.), The Global Carbon Cycle. J.Wiley & Sons, 57-77. DILL, H., KEMPER E., 1990 - Crystallographic and chemical variations during pyritization in the upper Barremian and lower Aptian dark claystones from the Lower Saxonian Basin (NW Germany). Sedimentology, 37, 427-443. DOTT R.H.JR., PROTHERO D.R., 1994 - Evolution of the Earth. McGraw-Hill, Inc. DROBNER E.H., HUBER G., WCHTERSHUSER R.D., STETTER K.O., 1990 - Pyrite formation linked with hydrogen evolution under anaerobic conditions. Nature, 346, 742744.

39

EDELSTEIN A.S., CHOW G.M., ALTMAN E.I., COLTON R.J., HWANG D.M., 1991 Self-arrangement of molybdenum particles into cubes. Science, 251, 1590-1592. EGGLETON R.A., 1987 - Noncrystalline Fe-Si-Al-oxyhydroxides. Clays and Clay Minerals, 35, 29-37. ENGLAND B.M., OSTWALD J., 1993 - Framboid-derived structures in some Tasman fold belt base-metal sulfide deposits, New-South-Wales, Australia. Ore Geol.Rev., 7-5, 381412. FABRICIUS F., 1961 - Die strukturen des Rogenpyrits ( Koessener Schichten, Raet ) als Beitrag zum Problem der Vererzten Bakterien. Geol. Rundsch., 51, 647-657. FARINA M., ESQUIVEL D.M.S, LINS DE BARROS H.G.P., 1990 - Magnetic iron-sulphur crystals from a magnetotactic microorganism. Nature, 343, 256-258. FARRAND M., 1970 - Framboidal sulphides precipitated synthetically. Mineral. Deposita, 5, 237-247. FERRIS F.G., FYFE W.S., BEVERIDGE T.T., 1987 Bacteria as nucleation sites for authigenic minerals in metal-contaminated lake sediments. Chemical Geol., 63, 225-232. FISHER R., 1986 - Pyrite replacement of mollusc shells from the Lower Oxford Clay (Jurassic) of England. Sedimentology, 33, 575-585. FISHER I.S.J., HUDSON J.D., 1985 Pyrite geochemistry and fossil preservation in shales. In: Whittington H.B., Morris S.C. Eds., Extraordinary Fossil Biotas: Their Ecological and Evolutionary Significance. Phil. Trans. Roy. Soc. London B, 311, 167-169. FITZPATRICK R.W., FRITSCH E., SELF P.G., 1996 - Interpretation of soil features produced by ancient and modern processes in degraded landscapes.5. Development of saline sulfidic features in non-tidal seepage areas. Geoderma, 69, 1-29. FRIZZO P., RAMPAZZO, G., MOLINAROLI, E. 1991 - Authigenic iron sulfides in recent sediments of the Venice Lagoon (Northern Italy). Eur. J. Mineral., 3, 603-612. FURUKAWA Y., BARNES H.L., 1995 - Reactions forming pyrite from precipitated amorphous ferrous sulfide. In: Vairavamurthy M.A. & Schoonen M.A.A. (Eds.) Geochemical Transformations of Sedimentary Sulfur. Amer.Chem.Soc., 194-205. GARCIA-GUINEA J, MARTINEZ-FRIAS J., GONZALEZ-MARTIN R., ZAMORA L., 1997 - Framboidal pyrites in antique books. Nature, 388, 631. GARCIA-GUINEA, J., MARTINEZ-FRIAS J., HARFFY M., 1998 - Cell-hosted pyrite framboids in fossil woods. Naturwissenschaften, 85, 78-81. GEROCH S., 1978 - Lower Cretaceous diatoms in the Polish Carpathians. Ann. Soc. Geol. Pologne, 48, 283-295. GHOSH K., SCHNITZER M., 1980 - Macromolecular structures of humic substances. Soil Sci., 129, 266-276. GOBLE R.J., 1980 - Geerite, Cu1.60S, a new copper sulfide from Dekalb Rownship, New York. Canad. Mineral., 18, 519-523. GOLDHABER M.B., KAPLAN I.R., 1974 - The sulfur cycle. In: Goldhaber E.D. (Ed.) The Sea, vol.5, 569-655, Wiley-Interscience. GRAHAM U.M., OHMOTO H., 1994 - Experimental study of formation mechanisms of hydrothermal pyrite. Geochim. Cosmochim. Acta, 58, 2187-2202. GRAHAM U.M., ROBERTSON J.D., 1995 - Micro-PIXE analysis of framboidal pyrite and associated maceral types in oil shale. Fuel, 74, 530-535. GRIMES S.T., RICKARD D., EDWARDS D., OLDROYD A., AXE L., DAVIES K., 1999 Experimental pyritization of plant cells. In: Ninth Ann. V.M. Goldschmidt Conf., p.107. LPI Contr.No.971, Houston. GRISDALE R.O., 1963 - Growth from molecular complexes. In: Gilman J.J. (Ed.) The Art and Science of Growing Crystals. J.Wiley New York. HALBACH P., PRACEJUS B., MAERTEN A., 1993 - Geology and mineralogy of massive

40

sulfide ores from the central Okinawa trough, Japan. Econ. Geol., 88, 2210-2225. HALL D.O., CAMMACK R., RAO K.K., 1971 - Role for Ferredoxins in the origin of Life and Biological Evolution. Nature, 233, 136-138. HALLBAUER D.K., 1986 - The mineralogy and geochemistry of Witwatersrand pyrite, gold, uranium and carbonaceous matter. In: Anhaeusser C.R. & Maske S. (Eds.) Mineral Deposits of Southern Africa. Geol.Soc.South Africa, 731-752. HAMOR T., 1994 - The occurrence and morphology of sedimentary pyrite. Acta Geol. Hungarica, 37, 153-181. HARTMAN H. 1975 - Speculations on the origin and evolution of metabolism. J. Mol. Evol., 4, 359-370. HLTON J., 1990 - Greigite and the magnetic properties of sediments. Limnol. Oceanogr., 35, 509-520. HONJO S., FISCHER A.G., GARRISON R., 1965 - Geopetal pyrite in fine-grained limestones. J. Sedim. Petrol., 35, 480-488. HOSSAIN A., 1975 - The occurrence of polyframboidal pyrite in a beach sand deposit, Coxs Bazar, Bangladesh. Amer. Mineral., 60, 157-158. HOWARTH R.W., 1979 - Pyrite: its rapid formation in a salt marsh and its importance to ecosystem metabolism. Science, 203, 49-51 HUDSON J.F., 1982 - Pyrite in ammonite-bearing shales from the Jurassic of England and Germany. Sedimentology, 29, 639-667. HUGGINS F.E., HUFFMAN G.P., KOSMACH D.A., LOWENHAUPT D.E., 1980 Mossbauer detection of goethite (alfa-FeOOH) in coal and its potential as an indicator of coal oxidation. Int. J. Coal Geol., 1, 75-81. ILER R.K., 1965 - Formation of precious opal. Nature, 207, 472-473. ILER R.K., 1973 - Colloidal silica. Surf. Colloid Sci., 6, 1-100. ILER R.K., 1979 - Chemistry of silica. Wiley Interscience, New York. IXER R.A., VAUGHAN D.J., 1993 - Lead-zinc-fluorite-baryte deposits of the Pennines, North Wales and Mendips. In: Patrick R.A.D. & Polya, D.A. (Eds.) Mineralization in the British Isles. Chapman & Hall. 355-418. JACOBS L., EMERSON S., 1985 - Partitioning and transport of metals across the O2/H2S interface in a permanently anoxic basin: Framvaren Fjord, Norway. Geochim. Cosmochim. Acta, 49, 1433-1444. JACOBSEN R., POSTMA D., 1999 - Redox zoning, rates of sulfate reduction and interactions with Fe-reduction and methanogenesis in a shallow sandy aquifer, Romo, Denmark. Geochim. Cosmochim. Acta, 63, 137-151. JAMES H.R.JR., VAN HOUTEN F.B., 1979 - Miocene goethitic and chamositic oolites, northeastern Columbia. Sedimentology, 26, 125-133. JAVOR B.J., MOUNTJOY E.W., 1976 - Late Proterozoic microbiota of the Miette Group, Southern British Columbia. Geology, 4, 111-119. JEAN G.E., BANCROFT, G.M., 1986 - Heavy metal adsorption by sulfide mineral surfaces. Geochim. Cosmochim. Acta, 50, 1455-1463. JEDWAB J., 1971 - La Magnetite de la Meteorite DOrgueil Vue au Microscope Electronique a Balayage. Icarus, 15, 319-340. JEKEL M.R., 1986 - The stabilization of dispersed mineral particles by adsorption humic substances. Water Res., 20, 1543-1554. JENSEN M., THOMSEN E., 1987 - Ultrastructure, dissolution and pyritization of Late Quarternary and Recent echinoderm. Bull. Geol. Soc. Denmark, 36, 275-287. JRGENSEN B.B., 1977 - Bacterial sulfate reduction within reduced microniches of oxidized marine sediments. Marine Biol., 41, 7-17. JRGENSEN B.B., ZAWACKI L.X., JANNASCH H.W., 1990 - Thermophilic bacterial

41

sulfate reduction in deep-sea sediments at the Guaymas Basin hydrothermal vent site (Gulf of California). Deep-Sea Res., 37, 695-710. KALLIOKOSKI J., 1974 - Pyrite framboid: animal, vegetable, or mineral? Geology, 2, 26-27. KALLIOKOSKI J., CATHLES L., 1969 - Morphology, mode of formation, and diagenetic changes in framboids. Bull. Geol. Soc. Finland, 41, 125-133. KANEHIRA K., BACHINSKI D., 1967 - Framboidal pyrite and concentric features in ores from the Tilt Cove mine. N. E. Newfounland. Canad. Mineral., 9, 124-127. KASCHKE M., RUSSELL M.J., COEL W.J., 1994 - [FeS/FeS2]. A redox system for the origin of life. Origins of Life and Evol. of Biosphere, 24, 43-56. KAUFFMAN S.A., 1993 - The origin of order: self-organization and selection in evolution. Oxford Univ. Press, New York. KELLER L.P., THOMAS K.L., CLAYTON R.N., MAYEDA T.K., DEHART J.M., MCKAY D.S., 1994 Aqueous alteration of the Bali CV3 chondrite: Evidence from mineralogy, mineral chemistry, and oxygen isotopic composition. Geochim. Cosmochim. Acta, 58, 5589-5598. KERRIDGE J.F., 1970 - Some observations on the nature of magnetite in the Orgueil meteorite. Earth Planet. Sci. Lett., 9, 299-306. KIZILSHTEIN L.J., MINAEVA L.G., 1972 - Origin of the framboidal pyrite. Dokl. Akad. Nauk SSSR, 206, 1187-1189 (in Russian) KOHN M.J., RICIPUTI L.R., STAKES D., ORANGE D.L., 1998 - Sulfur isotope variability in biogenic pyrite: Reflections of heterogeneous bacterial colonization? Amer. Mineral., 83, 1454-1468. KONDEPUDI D.K., KAUFMAN R.J., SINGH, N., 1990 - Science, 250, 975-976. KOPELMAN R., 1989 - Diffusion-controlled reaction kinetics. In: Avnir D. (Ed.) The Fractal approach to heterogeneous chemistry. J.Wiley, New York. 295-309. KOROLEV D.F., 1958 - The role of iron sulfides in the accumulation of molybdenumin sedimentary rocks of the reduced zone. Geochem. J., 4, 452-463. KOSACZ R., SAWOWICZ Z., 1983 - Framboidal pyrite from the copper deposit on the Fore-Sudetic monocline, Poland. Rudy Metale, 8, 292-297 (in Polish) KRETZSCHMAR R., STICHER H., 1998 - Colloid Transport in Natural porous Media: Influence of Surface Chemistry and Flow Velocity. Phys. Chem. Earth, 23, 133-139. KIBEK B., 1975 - The origin of framboidal pyrite as a surface effect of sulfur grains. Mineral. Deposita, 10, 389-396. LABERGE G.L., 1967 - Microfossils and Precambrian iron-formations. Geol. Soc. Amer. Bull., 78, 331-342. LAMER V.K., DINEGAR R.H., 1950 - Theory, Production and Mechanism of Formation of Monodispersed Hydrosols. J. Amer. Chem. Soc., 72, 4847-4854. LANDING W.M., LEWIS B.L., 1991 - Thermodynamic modelling of trace metal speciation in the Black Sea. In: Izdar E. & Murray J.W. (Eds.) Black Sea Oceanography, Kluwer. 125-160. LARGE D. J., SAWLOWICZ Z., SPRATT J., 1999 - A cobaltite-framboidal pyrite association from the Kupferschiefer: possible implications for trace element behaviour during the earliest stages of diagenesis. Mineral. Mag., 63, 353-361. LAUFER E.E., SCOTT J.D., PACKWOOD R., 1985 - Inhibition of pyrite growth by amorphous carbon. Canad. Mineral., 23, 57-60. LAZARD D., LAHAV N., ORENBERG J.B., 1987 - The biogeochemical cycle of the adsorbed template. I: Formation of the template. Origins of Life, 17, 135-148. LEBEDEV L.M., 1967 Metacolloids in endogenic deposits. Plenum Press, New York. LEJA J., 1982 - Surface chemistry of froth flotation. Plenum Press, New York. LI X., KWAK T.A.P., BROWN R.W., 1998 - Wallrock alteration in the bendigo gold ore

42

field, Victoria, Australia: Uses in exploration. Ore Geol. Rev., 13, 381-406. LIANG L., MCCARTHY J.F., JOLLEY L.W., MCNABB J.A., MEHLHORN T.L., 1993 Iron dynamics: Transformation of Fe(II)/Fe(III) during injection of natural organic matter in a sandy aquifer. Geochim. Cosmochim. Acta, 57, 1987-1999. LIN M.Y., LINDSAY H.M., WEITZ D.A., BALL R.C., KLEIN R., MEAKIN P., 1989 Universality in colloid aggregation. Nature, 339, 360-362. LOCQUIN M.V., WEBER C., 1978 - Origine et structure des membranes organiques cellulaires des moneres archeo-paleozoiques. CR 103e Congr. Nat. Soc. Savantes Nancy 1978 (Sect Sci), 2, 27-28. LORD C.J. III, CHURCH T.M., 1983 - The geochemistry of salt marshes: Sedimentary ion diffusion, sulfate reduction, and pyritization. Geochim. Cosmochim. Acta, 47, 1381-1391. LOUGHEED M.S., MANCUSO J.J., 1973 - Hematite framboids in the Negaunee Iron Formation, Michigan: evidence for their biogenic origin. Econ. Geol., 68, 202-209. LOVE L.G., 1957 - Microorganisms and the presence of syngenetic pyrite. Q. J. Geol. Soc. London. 113, 429-440. LOVE L.G., 1962 - Biogenic primary sulfide of the Permian Kupferschiefer and Marl Slate. Econ. Geol., 57, 350-366. LOVE L.G., 1965 - Micro-organic material with diagenetic pyrite from the Lover Proterozoic Mount Isa Shale and a Carboniferous shale. Proc.Yorks. Geol. Soc., 35, 187-202. LOVE L.G., 1967- Early diagenetic iron sulfide in Recent sediments of the Wash (England). Sedimentology, 9, 327-352. LOVE L.G., 1971 - Early diagenetic polyframboidal pyrite, primary and redeposited, from the Wenlockian Denbigh Grit Group, Conway, North Wales, UK. J. Sedim. Petrol., 41, 10381044. LOVE L.G., AMSTUTZ G. .C., 1966 - Review of microscopic pyrite from the Devonian Chattanooga shale and Rammelsberg Banderz. Fortschr. Mineral., 43, 273-309. LOVE L.G., AMSTUTZ G.C., 1969 - Framboidal pyrite in two andesites. N. Jb. Mineral. Mh., 3, 97-108. LOVE L.G., BROCKLEY H., 1973 - Peripheral radial texture in framboids of polyframboidal pyrite. Fortschr. Mineral., 50, 264-269. LOVE L.G., VANGUESTAINE M., 1973 - Polyframboidal pyrite of Rochelinval pyrite beds (Belgian Ardennes). Ann. Soc. Geol. Belgique, 9, 347-360. LOVE L.G., ZIMMERMAN D.O., 1961 - Bedded pyrite and microorganisms from the Mount Isa Shale. Econ. Geol., 56, 873-896. LOVE L.G., AL.-KAISY A.T.H., BROCKLEY H., 1984 - Mineral and organic material in matrices and coatings of framboidal pyrite from Pennsylvanian sediments, England. J. Sedim. Petrol., 54, 869-876. LOVLEY D.R., STOLZ J.F., NORD G.I.JR., PHILLIPS E.J.P., 1987 Anaerobic production of magnetite by a dissimilatory iron-reducing microorganism. Nature, 330, 252-253. LUTHER G.W.III, MEYERSON L.A., KRAJEWSKI J.J., HIRES R., 1980 - Metal sulfides in estuarine sediments. J. Sediment. Petrol., 50, 1117-1120. LUTHER G.W. III, GIBLIN A., HOWARTH R.W., RYANS R.A., 1982 - Pyrite and oxidized iron mineral phases formed from pyrite oxidation in salt marsh and estuarine sediments. Geochim. Cosmochim. Acta, 46, 2665-2669. LYONS T.W., 1997 - Sulfur isotopic trends and pathways of iron sulfide formation in upper Holocene sediments of the anoxic Black Sea. Geochim. Cosmochim. Acta, 61, 3367-382. MANDELBROT B.B., 1982 - The Fractal Geometry of Nature. San Francisco. MANN, S. 1988 - Molecular recognition in biomineralization. Nature, 332, 119-124. MANN S., SPARKS N.H.C., FRANKEL R.B., BAZYLINSKI D.A., JANNASCH H.W., 1990a - Biomineralization of ferrimagnetic greigite (Fe3S4) and iron pyrite (FeS2) in a

43

magnetotactic bacterium. Nature, 343, 258-261. MANN S., SPARKS N.H.C., BOARD R.G., 1990b - Magnetotactic Bacteria: Microbiology, Biomineralization, Palaeomagnetism and Biotechnology. Adv. Microbiol. Phys., 31. 125181. MARTILL D.M., UNWIN D.M., 1997 - Small spheres in fossil bones: Blood corpuscles or diagenetic products? Palaeontology, 40, 619-624. MARTINEZ-FRIAS J., NAVARRO-FLORES A., LUNAR-HERNANDEZ R., 1997 - First reference of pyrite framboids in a Hg-Sb mineralization: The Valle del Azogue mineral deposit (SE Spain). N. Jb. Mineral. Mh., 4, 175-184. MARTINEZ RUIZ F., SAWLOWICZ Z., ORTEGA-HUERTAS M., PALOMO I., 1995 Significance of the iron-rich microspherules for the K/T boundary dispute. In: Montanari & Coccioni R. (Eds.) The Role of Impacts in the Evolution of the Atmosphere andBiosphere with Regard to Short- and Long Terms Changes. Ancona 1995, 117-119. MARTINPENELA A.J., BARRAGAN G., 1995 - Silicification of plant remains in Messinian marine sediments in the Vera Basin (Aimeria, Spain). Eclogae Geol. Helvetiae, 88, 577593. MASSAAD M., 1974 - Framboidal pyrite in concretions. Mineral. Deposita, 9, 87-89. MATIJEVI E., 1992 - Monodispersed inorganic colloids: Achievements and problems. Pure & Appl. Chem., 64, 1703-1707. MATIJEVI E., 1994 - Uniform inorganic colloid dispersions. Achievements and challenges. Langmuir, 10, 8-16. MATIJEVI E., 1996 - Internally and externally composite monodispersed colloid particles. In: Pelizzetti E.(Ed.) Fine Particles Science and Technology, Kluwer Acad. Publ., 1-12. MCCONNELL H.M., TAMM L.K., WEIS R.M., 1984 - Periodic structures in lipid monolayer phase transitions. Proc. Natl. Acad. Sci. USA, 81, 3249-3253. MCNEIL D.H., 1990 - Stratigraphy and paleoecology of the Eocene Stellarima Assemblage Zone (pyrite diatom steinkerns) in the Beaufort- Mackenzie Basin, Artict Canada. Bull. Can. Petrol. Geol., 38, 17-27. MENON K.K., 1967 - Origin of diagenetic pyrite in the Quilon Limestone, Kerala, India. Nature, 213, 1219-122. MIDDELBURG J.J., DE LANGE G.J., VAN DER SLOOT H.A., VAN EMBURG P.R., SOPHIAH S., 1988 - Particulate manganese and iron framboids in Kau Bay, Halmahera (Eastern Indonesia). Marine Chem., 23, 353-364. MOELLER P., KERSTEN G., 1994 - Electrochemical accumulation of visible gold on pyrite and arsenopyrite surfaces. Mineral. Deposita, 29, 404-413. MONTANARI A., 1991 Authigenesis of impact spheroids in the K/T boundary clay from Italy: new constraints for high-resolution stratigraphy of terminal Cretaceous events. J. Sedim. Petrol., 61, 315-339. MORRISSEY C.J., 1972 - A quasi-framboidal form of syn-sedimentary pyrite. Trans. Instn. Min. Metall., 81, B22-B56. MORSE J.W., CORNWELL J.C., 1987 - Analysis and distribution of iron sulfide minerals in recent anoxic marine sediments. Marine Chem., 22, 55-69. MORSE J.W., CASEY W.H., 1988 - Ostwald processes and mineral paragenesis in sediments. Amer. J. Sci., 288, 537-560. MORSE J. W., ARAKAKI T., 1993 - Adsorption of divalent metals with mackinawite (FeS). Geochim. Cosmochim. Acta, 57, 3635-3640. MORSE J.W., MILLERO F.J., CORNWELL J.C., RICKARD D., 1987 - The chemistry of the hydrogen sulfide and iron sulfide systems in natural watres. Earth-Science Rev., 24, 1-42. MUCKE A., BADEJOKO T.A., AKANDE S.O., 1999 - Petrographic-microchemical studies

44

and origin of the Agbaja Phanerozic Ironstone Formation, Nupe Basin, Nigeria: a product of a ferruginized ooidal kaolin precursor not identical to the Minette-type. Mineral. Deposita, 34, 284-296. MUELLER G., FISCHBECK R., 1973 - Possible natural mechanism for protodolomite formation. Nature, 242, 139-141. MUELLER A., SCHLADERBECK N., 1985 - Systematik der Bildung von Elektronentransfer-Clusterzentren {Fen/Sn)m+ mit Relevanz zur Evolution von Ferredoxinen. Chimia, 39, 23-24. MUKHOPADHYAY P.K., GOODARZI F., CRANDLEMIRE A.L., GILLIS K.S., MACNEIL D.J., SMITH, W.D., 1998 - Comparison of coal composition and elemental distribution in selected seams of the Sydney and Stellarton Basins, Nova Scotia, Eastern Canada. Int. J.Coal Geol., 37, 113-41. MURAMOTO J.A., HONJO S., FRY B., HAY B.J., HOWARTH R.W., CISNE J.L., 1991 Sulfur, iron and organic carbon fluxes in the Black Sea: sulfur isotopic evidence for origin of sulfur fluxes. Deep-Sea Res., 38, S1151-S1187. MUROWCHICK J.B., BARNES H.L., 1986 - Marcasite precipitation from hydrothermal solutions. Geochim. Cosmochim. Acta, 50, 2615-22629. MUROWCHICK J.B., BARNES H.L., 1987 - Effects of temperature and degree of supersaturation on pyrite morphology. Amer. Mineral., 72, 1241-1250. NAKAZAWA H., 1999 - Self-organization in phosphoric derivatives of allophane. EuroClay99, Krakow, (Abstr.) NAKAZAWA H., YAMADA H., HASHIZUME H., 1993 - Origin of life in the earths crust, a hypothesis: Probable chemical evolution synchronized with the plate tectonics of the early earth. Viva Origino, 21, 213-222. NARDI S., BINDA P.L., BACCELLE L.S., CONCHERI G., 1994 - Amino acids of Proterozoic and Ordovician sulfide-coated grains from western Canada: Record of biologically-mediated pyrite precipitation. Chem. Geol., 111, 1-15. NCUBE A., ZWEIFEL, H., AMSTUTZ G.C., 1978 - On the occurrences of bravoite, framboidal pyrite, marcasite, pyrrhotite and possible plant remains in the Laiswall Lead and Zinc deposits, Sweden. N. Jb. Mineral. Abh., 132, 264-283. NEVES R., SULLIVAN H.J., 1964 - Modification of fossil spore exines associated with the presence of pyrite crystals. Micropaleontol., 10, 443-452. NICOLIS G., PRIGONINE, I., 1977 - Self-organisation in nonequilibrium systems. J.Wiley, London. NIELSEN P., SWENNE R., DICKSON J.A.D., FALLICK A.E., KEPPENS E., 1997 Spheroidal dolomites in a Visean karst system Bacterial induced origin? Sedimentology, 44, 177-195. NUHFER E.B., PAVLOVIC A.S., 1979 - Association of kaolinite with pyritic framboids. J. Sediment. Petrol., 49, 321-323. OBRIEN N.R., 1995 - Origin of shale fabric-clues from framboids. Northeastern Geol. Environ. Sci., 17, 146-150. ORGEL L.E., 1986 - Did template directed nucleation precede molecular replication? Orig. Life Evol. Biosph., 17, 27-34. ORGEL L.E., 1992 - Molecular replication. Nature, 358, 203-209. ORPIN A.R., 1994 - Pyrite chimneys formed by the expulsion of cold brackish ground water on a tidal flat, Awarua Bay, Southland, New Zealand. Sediment. Geol., 92, 67-77. OSTWALD J., ENGLAND B.M., 1977 - Notes on framboidal pyrite from Allandale New South Wales, Australia. Mineral. Deposita, 12, 111-116. OSTWALD J., ENGLAND B.M., 1979 - The relationship between euhedral and framboidal pyrite in base-metal sulfide ores. Mineral. Mag., 43, 297-300.

45

OSZCZEPALSKI S., 1999 Origin of the Kupferschiefer polymetallic mineralization in Poland. Mineral. Deposita, 34, 599-613. OXTOBY D.W., 1990 - New perspectives on freezing and melting. Nature, 347, 725-730. PANKOW J.F., MORGAN J.J., 1980 - Dissolution of tetragonal ferrous sulfide (mackinawite) in anoxic aqueous systems, 2. Implications for cycling of iron, sulfur and trace metals. Environ. Sci. Technol., 14, 183-186. PAPUNEN H., 1966 - Framboidal texture of the pyritic layer found in a peat bog in SE Finland. Bull. Comm. Geol. Finlande, 222, 117-125 PASSIER H.F., MIDDELBURG J.J., DE LANGE G.J., NOETTCHER M.E., 1999 - Modes of sapropel formation in the eastern Mediterranean: some constraints based on pyrite properties. Marine Geol., 153, 199-219. PICHLER T., GIGGENBACH W.F., MCINNES B.I.A., BUHL D., DUCK B., 1999 Econ. Geol., 94, 281-287. POLLASTRO R.M., 1981 - Authigenic kaolinite and associated pyrite in chalk of the Cretaceous Niobrara Formation, eastern Colorado. J. Sedim. Petrol., 51, 553-562. PRUYSERS P.A., DE LANGE G.J., MIDDELBURG J.J., 1991 - Geochemistry of eastern Meditteranean sediments: Primary sediment composition and diagenetic alterations. Marine Geol., 100, 137-154. RAISWELL R., 1982 - Pyrite texture, isotopic composition, and the availability of iron. Amer. J. Sci., 282, 1244-1263 RAISWELL R., 1997 - A geochemical framework for the application of stable sulfur isotopes to fossil pyritization. J. Geol. Soc., London, 154, 343-356. RAISWELL R., PLANT J., 1980 - The incorporation of trace elements into pyrite during diagenesis of black shales, Yorkshire, England. Econ. Geol., 75, 684-699. RAISWELL R., BERNER R.A., 1985 - Pyrite formation in euxinic and simi-euxinic sediment. Amer. J. Sci., 285, 710-724. RAISWELL R., CANFIELD D.E., 1998 - Sources of iron for pyrite formation in marine sediments. Amer. J. Sci., 298, 219-245. RAISWELL R., WHALER K., DEAN S., COLEMAN M.L., BRIGGS D.E.G., 1993 - A simple three-dimensional model of diffusion-with-precipitation applied to localised pyrite formation in framboids, fossils and detrital iron minerals. Marine Geol., 113, 89-100. RAMBERG H., EKSTROM T., 1964 - Note on preferred orientation of pyrite cubes in grit layers in slate. N. Jb. Mineral. Mh., 8, 246-251. REEDER D.H., CLAUSEN A.M., ANNEN M.J., CARR P.W., FLICKINGER M.C., MCCORMICK A.V., 1996 - An approach to hierarchically structured porous zirconia aggregates. J. Colloid Interface Sci., 184, 328-330. REYNOLDS R.L., TUTTLE M.L., RICE C.A., FISHMAN N.S., KARACHEWSKI J.A., SHERMAN D.M., 1994 - Magnetization and geochemistry of greigite-bearing Cretaceous strata, North Slope Basin, Alaska. Amer. J. Sci., 294, 485-528. RICKARD D.T., 1970 - The origin of framboids. Lithos, 3, 269-293. RICKARD D.T., 1975 - Kinetics and mechanisms of pyrite formation at low temperatures. Amer. J. Sci., 275, 636-652. RICKARD D.T., 1989 - Experimental concentration-time curves for the iron (II) sulfide percipitation process in aqueous solutions and their interpretation. Chem. Geol., 78, 315324. RICKARD D.T., 1996 - Kinetics of pyrite formation by the H2S oxidation of iron (II) monosulfide in aqueous solutions between 25 and 125oC: The rate equation. Geochim. Cosmochim. Acta, 61, 115-134. RICKARD D.T. 1997 Kinetics of pyrite formation by the H2S oxidation of iron (II) monosulfide in aqueous solutions between 25 and 125oC: The rate equation. Geochim.

46

Cosmochim. Acta, 61, 115-134. RICKARD D.T., COWPER M., 1994 - Kinetics and mechanism of chalcopyrite formation from Fe(II) disulfide in aqueous -solution (<200C). Geochim. Cosmochim. Acta, 58, 3795-3802. ROBERTS A.P., TURNER G.M., 1993 - Diagenetic formation of ferrimagneticiron sulfide minerals in rapidly deposited marine sediments, South Island, New Zealand. Earth Planet. Sci. Lett., 115, 257-273. RICKARD D.T, SCHOONEN N.A.A., LUTHER G.W. III, 1995 - Chemistry of iron sulfides in sedimentary environments. In: Vairavamurthy, M.A. & Schoonen, M.A.A. (Eds.) Geochemical transformations of sedimentary sulfur. Am. Chem. Soc. Symp. Series, 612: 168-193. ROJAS-CHAPANA J.A., BARTELS C.C., POHLMANN L., TRIBUTSCH H., 1998 - Cooperative leaching and chemotaxis of thiobacilli studied with spherical sulfur/sulfide substrates. Proc. Biochem., 33, 239-248. ROSENTSVIT A.O., EPSHTEYN G.Yu., 1963 - Crystallization of pyrite from gels of complex composition. Dokl. Akad. Nauk SSSR, 150, 1134-1136. RUSSELL M.J., 1983 - Major sediment-hosted exhalative zinc+ lead deposits: Formation from hydrothermal convection cells that deepen during crustal extension. Mineral. Assoc., Canada, Short Course Handbook, 9, 251-282. RUSSELL M.J., HALL A.J., CAIRNS-SMITH A.G., BRATERMAN P.S., 1988 - Submarine hot springs and the origin of life. Nature, 336, 117. RUSSELL M.J., HALL A.J., TURNER D., 1989 - In vitro growth of iron sulfide chimneys: possible culture chambers for origin-of-life experiments. Terra Nova, 1, 238-241. RUSSEL M.J., HALL A.J., GIZE A.P., 1990 - Pyrite and the origin of life. Nature, 344, 387. RUSSELL M.J., DANIEL R.M., HALL A.J., 1993 - On the emergence of life via catalytic iron-sulfide membranes. Terra Nova, 5, 343-347. RUSSELL M.J., DANIEL R.M., HALL A.J., SCHERRINGHAM J.A., 1994 - A hydrothermally precipitated catalythic iron sulfide membrane as a first step toward life. J. Mol. Evol., 39, 231-243. RUST G.W., 1935. - Colloidal primary copper ores at Cornwall Mines, southeastern Missouri. Jour. Geol., 43, 398-426. SANDERS J.V., 1964 - Colour of precious opal. Nature, 4964, 1151-1153. SASSANO G.P., SCHRIJVER K., 1989 - Framboidal pyrite: Early-diagenetic, latediagenetic, and hydrothermal occurrences from Acton Vale quarry, Cambro-Ordovicion, Quebec. Amer. J. Sci., 289, 167-179. SAWOWICZ Z., 1987 - Framboidal pyrite from the metamorphic Radzimowice Schists of Stara Gora (Lower Silesia, Poland). Mineral. Polonica, 18, 57-67. SAWLOWICZ Z., 1990 - Primary copper sulfides from the Kupferschiefer, Poland. Mineral. Deposita, 25, 262-271. SAWLOWICZ Z., 1992 - Primary sulfide mineralization in Cu-Fe-S zones of Kupferschiefer, Fore-Sudetic monocline, Poland. Trans. Instn. Min. Metall. (Sec.B:Appl. Earth Sci.), 101, B1-B8. SAWLOWICZ Z., 1993a - Pyrite framboids and their development: a new conceptual mechanism. Geol. Rundsch., 82, 148-156. SAWLOWICZ Z., 1993b - Iridium and other platinum-group elements as geochemical markers in sedimentary environments. Palaeogeogr., Palaeoclimat., Palaeoecol., 104, 253-270. SAWLOWICZ Z., 1996 - Geochemical evidence for euxinic conditions in a water column during deposition of the Cenomanian-Turonian boundary sediments at Cismon (Venetian Alps, Italy). In: Drobne K., Gorican S & Kotnik B. Eds. The Role of Impact Processes in

47

the Geological and Biological Evolution of Planet Earth. Postojna 1996, 74-75. SAWOWICZ Z., BK M., 1997 - Pyritization of Radiolaria in anoxic water column, anoxic deposits of the Cenomanian/Turonian Boundary in the Pieniny Klippen Belt, Poland. Mineral. Slovaca, 29, 273-274. SCHALLREUTHER R., 1984 - Framboidal pyrite in deep-sea sediments. Init. Repts. DSDP, 75, 875-891. SCHEIHING M.H., GLUSKOTER H.J., FINKELMAN R.B., 1978 - Intersticial networks of kaolinite within pyrite framboids in the Megis Coal of Ohio. J. Sedim. Petrol. 48, 723732. SCHMITZ B., 1985 - Metal precipitation ion the Creatacous-Tertiary boundary clay at Stevns Klint, Denmark. Geochim. Cosmochim. Acta, 49, 2361-2370. SCHNEIDERHOEHN H., 1923 - Chalkographische Untersuchung des Mansfelder Kupferschiefers. N. Jb. Mineral. Geol. Palaontol., 47, 1-38. SCHOONEN M.A.A., BARNES H.L., 1991a - Reactions forming pyrite and marcasite from solution. I Nucleation of FeS2 below 100oC. Geochim. Cosmochim. Acta, 55, 1495-1504. SCHOONEN M.A.A., BARNES H.L. 1991b - Reactions forming pyrite and marcasite from solution. II. Via FeS precursor below 100oC. Geochim. Cosmochim. Acta, 55, 1505-1514. SEIGLIE G.A., 1973 - Pyritization in living foraminifers. J. Foram. Research, 3, 1-6. SHIKAZONO N., UTADA M., 1997 - Stable isotope geochemistry and diagenetic mineralization associated with the Tono sandstone-type uranium deposit in Japan. Mineral. Deposita, 32, 596-606. SHOCK E.L., 1992a - Chemical environments of submarine hydrothermal systems. Origins. Life. Evol. Biosphere, 22, 67-107. SHOCK E.L., 1992b - Hydrothermal organic synthesis experiments Origins. Life. Evol. Biosphere, 22, 135-146. SHUEY R.T., 1975 - Semiconducting Ore Minerals. Elsevier, Amsterdam. SIESSER W.G., ROGERS J., 1976 - Authigenic pyrite and gypsum in South West African continental slope sediments. Sedimentology, 23, 567-577. SIMON S.B., GROSSMAN L., 1997 - In situ formation of palisade bodies in calcium, aluminum-rich refractory inclusions. Meteor. Planet. Sci., 32, 61-70. SINKOVEC B., SAKAC K., DURN G., 1994 - Pyritized bauxites from Minjera, Istria, Croatia. Nat. Croat., 3, 41-65. SKEI J.M., 1988 - Formation of framboidal iron sulfide in the water of a permanently anoxic fjord-Framvaren, South Norway. Marine Chem., 23, 345-352. SKRIPCHENKO N.S., 1968 - Biogenic pyrite ore deposits. Dokl. Akad. Nauk SSSR, 181, 77179 (in Russian). SOMA M., TANAKA A., SEYAMA H., SATAKE K., 1994 - Characterisation of arsenic in lake-sediments by X-ray photoelectron-spectroscopy. Geochim. Cosmochim. Acta, 58, 2743-2745. STANTON M.R., GOLDHABER M.B., 1991 - An experimental study of goethite sulfidization-relationship to the diagenesis of iron and sulfur. US Geol. Surv. Bull., 1973E, 1-20. STUMM W., 1993 - Aquatic colloids as chemical reactants: Surface structure and reactivity. Colloids and Surfaces, A 73, 1-18. STUMM W., SULZBERGER B., 1992 - The cycling of iron in natural environments: Considerations based on laboratory studies of heterogeneous redox processes. Geochim. Cosmochim. Acta, 56, 3233-3257. SUITS N.S., WILKIN R.T., 1998 - Pyrite formation in the water column and sediments of a meromictic lake. Geology, 26, 1099-1102. SUK D., PEACOR D.R., VAN DER VOO R., 1990 - Replacement of pyrite framboids by

48

magnetite in limestone and implications for paleo-magnetism. Nature, 345, 611-613. SUMMERS D.P., CHANG S., 1993 - Prebiotic ammonia from reduction of nitrite by iron (II) on the early Earth. Nature, 365, 630-633. SUNAGAWA I., ENDO Y., NAKAI N., 1971 - Hydrothermal synthesis of framboidal pyrite. Soc. Min. Geol. Jp. Spec. Issue, 2, 10-14. SWEENEY R.E., KAPLAN I.R., 1973 - Pyrite framboid formation: laboratory synthesis and marine sediments. Econ. Geol., 68, 618-634. TAILLEFERT M., THEBERGE S.M., HOVER V., LUTHER, G.W.III, 1999 - The influence of sulfides on soluble iron(III) in anoxic sediment porewaters. In: Ninth Ann. V.M. Goldschmidt Conf., p.289. LPI Contr.No.971, Houston. TAMBUYSER P., 1976 - Morphology of the pyrite aggregates from Cap-Blanc-Nez (France). Mineralogical Record, 7, 179-181, TAYLOR G.R., 1982 - A mechanism for framboid formation as illustrated by a volcanic exhalative sediment. Mineral. Deposita, 17, 23-36. TAYLOR G.R., 1983 - A mechanism for framboid formation-the role of bacteria. Mineral. Deposita, 18, 129-130. THIEME J., NIEMEYER J., 1996 - Fractal characterization of hematite aggregates by X-ray microscopy. Geol. Rundsch., 85, 852-856. THOMPSON R.A., HELZ G.R., 1994 - Copper speciation in sulfidic solutions at lo sulfur activity: Further evidence for cluster complexes. Geochim. Cosmochim. Acta, 58, 29712983. THOMSEN E., VORREN T.O., 1984 - Pyritization of tubes and burrows from Late Pleistocene continental shelf sediments of North Norway. Sedimentology, 31, 481-492. TWITCHETT R.J., WIGNALL P.B., 1996 - Trace fossils and the aftermath of the PermoTraissic mass extinction: Evidence from northern Italy. Palaeogeogr.,Palaeoclim., Palaeoecol., 124, 137-151. VALLENTYNE J.R., 1963 - Isolation of pyrite spherules from Recent sediments. Limnol. Oceanogr., 8, 16-30. VAUGHAN D.J., TOSSELL J.A., 1981 - Electronic structure of thiospinel minerals: Results from MO calculations. Amer. Mineral., 66, 1250-1043. VAVELIDIS M., 1995 - Framboidal pyrite from the Kuroko-type barite mineralization of the Katsimouti area, Milos Island, Greece. Chemie der Erde, 55, 281-294. VINE J. D., TOURTELOT E. B., 1970 - Geochemistry of black shale deposits - a summary report. Econ. Geol., 65, 253-272. VON DER BORCH C.C., JONES J.B., 1976 - Spherular modern dolomite from the Coorong area, South Australia. Sedimentology, 23, 587-591. WCHTERSHUSER G., 1988a - Pyrite formation, the first energy source for life: a hypothesis. System. Appl. Microbiol., 10, 207-210. WCHTERSHUSER G., 1988b - Before enzymes and templates: theory of surface metabolism. Microbiol. Rev., 52, 452-484. WCHTERSHUSER G., 1990 - Evolution of the first metabolic cycles. Proc. Natl. Acad. Sci. USA, 87, 200-204. WALKER R., STEELE A.D., MORGAN D.T.B., 1997 - Deactivation of pyrophoric iron sulfides. Industr. Engin. Chem. Res., 36, 3662-3667. WANG Q.W., MORSE J.W., 1996 - Pyrite formation under conditions approximating those in anoxic sediments. 1. Pathway and morphology. Marine Chem., 52, 99-121. WARD C.R., ROBERTS F.I., 1990 - Occurrence of spherical halloysite in bituminous coals of the Sydney basin, Australia. Clays & Clay Min., 38, 501-506. WATSON J.H.L., CARDELL R.R. JR., HELLER W., 1962 - The internal structure of colloidal crystals of -FeOOH and remarks on their asemblies in Schiller layers.

49

J. Phys. Chem., 66, 1757-1763. WATSON J. H. P., ELLWOOD D. C., 1994 - Biomagnetic separation and extraction process for heavy metals from solution. Minerals Engineering, 7, 1017-1028. WATSON J. H. P., ELLWOOD D. C., QIXI DENG, MIKHALOVSKY S., HAYTER C. E., EVANS J., 1995 - Heavy metal adsorption on bacterially produced FeS. Minerals Engineering, 8, 1097-1108. WEITZ D.A., OLIVEIRA M., 1984 - Fractal structures formed by kinetic aggregation of aqueous gold colloids. Phys. Rev. Lett., 52, 1433-1436. WIEDMANN J., NEUGEBAUER J., 1979 - Lower Cretaceous ammonites from the South Atlantic Leg 40 (DSDP): Their stratigraphic value and sedimentologic properties. Init. Repts. DSDP, Suppl. to Vols. 38, 39, 40 and 41, 709-734. WIESE R.G. JR, FYFE W.S., 1986 - Occurrences of iron sulfides in Ohio coals. Int. J. Coal Geol., 6, 251-276. WIGNALL P.B., NEWTON R., 1998 - Pyrite framboid diameter as a measure of oxygen deficiency in ancient mudrocks. Amer. J. Sci., , 537-552. WILLIAMS R.C., SMITH K.M., 1957 - A crystallizable isect virus. Nature, 179, 119-120. WILLIAMS A., CUSACK M., 1996 - Lingulid shell mediation in clay formation. Lethaia, 29, 349-360. WILKIN R. T., BARNES H. L., 1997a - Formation processes of framboidal pyrite. Geochim. Cosmochim. Acta, 61, 323-339. WILKIN R. T., BARNES H. L., 1997b - Pyrite formation in an anoxic estuarine basin. Amer. J. Sci. 297, 620-650. WILKIN R. T., BARNES H. L., BRANTLEY S.L., 1996 - The size distribution of framboidal pyrite in modern sediments: An indicator of redox conditions. Geochim. Cosmochim. Acta, 60, 3897-3912. WILKIN R.T., ARTHUR M.A., DEAN W.E., 1997 - History of water-column anoxia in the Black Sea indicated by pyrite framboid size distributions. Earth Planet. Sci. Lett, 148, 517-525. WILSON N.S.F., ZENTILLI M., 1999 - The role of organic matter in the genesis of the El Soldado volcanic-hosted manto-type Cu deposit, Chile. Econ. Geol., 94, 1115-1135. WOESE C.R., WCHTERSHUSER G., 1992 - Origin of Life. In: B.E.G.Briggs & P.R.Crowther (eds.) Palaeobiology. A synthesis. Blackwell Scient. Publ., 3-9. YOUNGSON J.H., 1995 - Sulfur mobility and sulfur mineral preciptation during early Miocene-Recent uplift and sedimentation in Central Otago, New Zealand. New Zealand J. Geol. Geophys., 38, 407-417. ZHAN Y.L., EVANGELOU V.P., 1998 - Formation of ferric hydroxide-silica coatings on pyrite and its oxidation behaviour. Soil Sci., 163, 53-62. ZHU XIANG, LI MIN, ROGERS R., MEYER W., OTTEWILL H., STS-73 SPACE SHUTTLE CREW, RUSSELL B., CHAIKIN P.M., 1997 - Crystallization of hard-sphere colloids in microgravity. Nature, 387, 883-885. ZOLENSKY M.E., WEISBERG M.K., BUCHANAN P.C., MITTLEFEHLDT D.W., 1996 Mineralogy of carbonaceous chondrite clasts in HED achondrites and the Moon. Meteoritics Planet. Sci., 31, 518-537.

50

Fig. 5. A) Aggregation of pyrite framboids built of irregular grains. Note angular shape of some framboids (arrow) (Cretaceous shales, flysch Polish Carpathians). SEM. B) Different forms of framboidal pyrite; a) large (50-60 m) framboids, b) aggregate built of partly recrystallized framboids, c) framboidal aggregate (heavy mineral fraction from the Miocene shales of the Polish Carpathians, sample supplied by courtesy of M. Michalik). SEM. C) Aggregation of pyrite framboids, some of them partly recrystallized (note angular habits and individual euhedra)(Silurian shales, East European Platform, Poland). SEM. D) Masses of pyrite framboids of different size, partly embedded in digenite. Reflected light. E) Pyrite framboids with coarser outermost crystals, resembling sunflower framboids. Reflected light. F) Pyrite grain built of partly recrystallized framboids. Reflected light. (D, E, F- samples from the Lubin Mine, Zechstein Kupferschiefer, Poland).

51

Fig. 6. A) Pyrite framboids and spherules (recrystallized framboids), often oval in shape (Miocene shales, flysch Polish Carpathians). SEM. B) Completely recrystallized pyrite framboids (p) in gypsum (g) lense. SEM. C) Framboidal aggregations of pyrite crystals in calcite, possible replacement of brachiopod (?) shell. In the central part partly recrystallized framboids are visible. SEM. D) Masses of framboidal pyrite and pyrite framboids of different size, partly recrystallized. SEM. E) Framboidal pyrite aggregates with shapes modified by surrounding mica crystals. SEM. (B, C, D, E samples from the Silurian shales, East European Platform, Poland). F) Fragments of framboidal pyrite aggregates. Note interstitial silica (arrows) with ghosts after pyrite crystals (Cretaceous siliceous shales, flysch Polish Carpathians). SEM.

52

Fig. 7. A) Complex texture of pyrite framboid core with surrounding pyrite radial crystals, partly replaced by digenite (Lubin Mine, Zechstein Kupferschiefer, Poland). SEM. B) Fragment of iron sulfide stalactite. Framboidal pyrite cores are surrounded by radial marcasite crystals (Zn-Pb Trzebionka Mine, Triassic, Poland). SEM. C) Iron sulfide fossil chimney built of porous pyrite and partly recrystallized pyrite framboids. Note iron sulfide framboids (arrow) in the central vent (Zn-Pb Silvermine Mine, Ireland) ( photo by courtesy of M.J. Russell). Reflected light. D) Spherules of synthetized geerite (Cu1.6S). SEM. E) and F) Pyrite crystals built of framboidal pyrite, with interstices filled by silica (Rochelinval pyritic beds, Belgium; sample supplied by courtesy of L. Dejonghe). SEM.

53

Fig. 8. A) Elipsoidal pyrite framboids in silica matrix ( Rochelinval pyritic beds, Belgium; sample supplied by courtesy of L. Dejonghe). SEM. B) Pyrite framboid with the framboidal tail. Note the domains of regular arrangement of the microcrystals in both structures. The tail could have been formed in the pressure shadow of the pyrite framboid (heavy mineral fraction from the Miocene shales of the Polish Carpathians, sample supplied by courtesy of M. Michalik). SEM. C) Pyrite microconcretion built of mainly ocahedral crystals (limestones, Crimea Peninsula, Ukrainien). SEM. D) Close-up of the octahedra from the pyrite microconcretion from the Fig. C. Note imperfect growth of octahedra, some of them may be hollow. SEM. E) Goethite pseudomorphs after probably pyrite framboids from the K/T boundary layer from Spain (photo by courtesy of F. Martinez Ruiz). SEM. F) Close-up of goethite octahedra from goethite pseudomorphs from the Fig. E. SEM.

54

Fig. 9. A) Intergrown microcrystals in pyrite framboid. Note rough surface suggesting that they are composed of smaller units. SEM. B) Microcrystals of irregular habit from the pyrite framboid. Note dimples on the surface of some crystals suggesting that they are composed of smaller units and amalgamation of some crystals. C) Irregular habit of pyrite crystals building framboid. Finer grains of pyrite fill interstices. SEM. (A, B, C samples from the Miocene shales, flysch Polish Carpathians). D) Irregular habits of partly intergrown microcrystals from the pyrite framboid. Note that on flat faces of some crystals circular flat-topped islands occurred, as if a droplet of fluid had spread over the surface and then solidified (Pandeya & Tolansky, 1961 and Graf, 1951; vide Grisdale, 1965) (Miocene shales, ForeCarpathian Basin). SEM. E) Framboidal aggregate, some pyrite microcrystals have been removed leaving silica interstitial texture. SEM. F) Interstitial silica framework from framboidal pyrite aggregates (close-up of the fragment from fig. E). Note granular texture, possibly moulds of pyrite microcrystals. SEM. (E, F samples from the Rochelinval pyritic beds, Belgium; sample supplied by courtesy of L. Dejonghe).

55

Fig. 10. A) Aggregate of irregular, partly recrystallized framboids in calcite matrix and angular pyrite framboid (35 m). SEM. B) Crystal-like framboids, composed of very irregular microcrystals of non-stechiometric iron sulfide composition, FeS1.7-1.9. SEM. C) Crystal-like pyrite framboid. Note pentagonal (dodekahedron?) pattern of microcrystals and elongated outermost microcrystals. SEM. D) Pyrite framboid with close-packing arrangement of the microcrystals. Note different arangement of microcrystals in separate layers. SEM. (A, C samples of the heavy mineral fraction from the Miocene shales of the Polish Carpathians, supplied by courtesy of M. Michalik). SEM. E) Close-packed pyrite microcrystals from the pyrite framboidal aggregates. SEM. F) Regular close-packing arrangement of microcrystals building pyrite crystals (B, C, E, F samples from the Rochelinval pyritic beds, Belgium; supplied by courtesy of L. Dejonghe). SEM.

56

Fig. 11. A) Pyrite sherules (recrystallized framboids) in chalcocite lense (Lubin Mine, Kupferschiefer, Poland). SEM. B) Pyrite framboids with cobaltite filling interstices and forming rims on pyrite framboids (partly enclosed by digenite grain). SEM. C) Pyrite framboid with cobaltite filling interstices. SEM. D) Digenite grain enveloping pyrite-cobaltite structures; a) pyrite framboids with cobaltite filling interstices and forming rims on pyrite framboids, b) aggregate of pyrite grains (recrystallized framboids). Note that thin interstices between grains are composed probably of cobaltite and that outer rims of pyrite grains are Corich. SEM. E) Organic matter sheat on framboidal pyrite lense. SEM. (B, C, D, E samples from the Rudna Mine, Kupferschiefer, Poland). F) Elongated pyrite framboid in the organic matter sheat (Cretaceous shales, flysch Polish Carpathians), SEM.

57

Fig. 12. A) Partly cracked Diatom steinkern composed of pyrite framboids. SEM. B) Diatom steinkern composed of pyrite framboids. SEM. (A, B samples from the BeaufortMackenzie Basin, supplied by courtesy of D.H. McNeil). C) Section of radiolarian skeleton. Pyrite framboids and framboidal aggregate (disintegrated framboid?) in the so-called oillamp pores. SEM. D) Pyrite framboids in the interior of the radioarian skeleton (C, D samples from the Cretaceous shales of the Pieniny Klippen Belt, Poland). SEM. E) Annular pyrite framboids in foraminifera chambers. Reflected light. F) Pyrite framboids in foraminifera chambers. Reflected light. (E, F samples from the Polkowice Mine, Zechstein Kupferschiefer, Poland).

58

Вам также может понравиться