+ All Categories
Home > Documents > Electron microscopy and pyrolysis of kerogens from the Kimmeridge Clay Formation, UK: Source...

Electron microscopy and pyrolysis of kerogens from the Kimmeridge Clay Formation, UK: Source...

Date post: 29-Nov-2023
Category:
Upload: univ-orleans
View: 0 times
Download: 0 times
Share this document with a friend
17
Geochimica et Cosmochimica Acta, Vol. 59, No. 18, pp. 3731-3747, 1995 Copyright © 1995 Elsevier Science Ltd Printed in the USA. All rights reserved 0016-7037/95 $9.50 + .00 0016-7037(95)00273-1 Electron microscopy and pyrolysis of kerogens from the Kimmeridge Clay Formation, UK: Source organisms, preservation processes, and origin of microcycles M. BOUSSAFIR,I F. GELIN,2 E. LALLIER-VERGEs, I S. DERENNE,2 P. BERTRAND, I. * and C. LARGEAU 2 IUniversite d'Orleans, URA CNRS 724, Dept. des Sciences de la Terre, F-45067 Orleans cedex, France 1Laboratoire de Chimie Bioorganique et Organique Physique, URA CNRS 1381, Ecole Nationale Superieure de Chimie de Paris, II rue Pierre et Marie Curie, F-7523I Paris cedex 05, France (Received November 29, 1994; accepted in revised form June 9, 1995) Abstract-Recent studies revealed short-term cyclic variations (microcycles) in total organic carbon (TOC) and the hydrogen index (HI) in the Kimmeridge Clay Formation, an organic-rich deposit considered to be a lateral equivalent of the main source rocks of the North Sea. In addition, three different types of organic matter that all appear to be amorphous when observed by light microscopy (AOM) were recog- nized. Together, these AOM types account for over 80% of total kerogen and their relative abundances show large variations along each microcycle. In the present work, transmission electron microscopy (TEM) observations were carried out on samples (whole kerogens, kerogen subfractions only comprising a single type of AOM, selected rock fragments) corresponding to typical points within a microcycle and obtained via high resolution sampling. The nature and the relative abundances of the products generated by Curie- point Py-GC-MS and off-line pyrolyses of isolated kerogens were also determined for two selected samples corresponding to the beginning and the top of the microcycle. Combination of such ultrastructural observations, including some semiquantitative studies, and the analysis of pyrolysis products allowed ( 1) determination of the ultrastructural features of the three AOM types thus providing what we believe to be the first example of correlations between light microscopy (palynofacies, in situ maceral analysis) and TEM observations on "amorphous" fossil materials; (2) identification of the source organisms and elu- cidation of the mode of formation of the different AOM types in the Kimmeridge Clay; (3) explanation of the variations in their relative abundances taking place along a microcycle and establishment of tight correlations with TOC and HI changes; and (4 ) explanation of the origin of the microcyclic variations in kerogen quantity (TOC) and quality (HI) occurring in the Kimmeridge Clay Formation. Interrelationships between primary productivity, sulphate reduction intensity, and lipid' 'vulcanization" likely played a major role in the control of such variations. 1. INTRODUCTION The aim of this work was to gain information on kerogen genesis and source organisms in the Kimmeridge Clay and on the origin of the short-term cyclic variations, in both kerogen quantity and quality, occurring in this formation. Kimmeridge Clay is a marine deposit, with alternating organic-rich shales and marls, considered as a lateral equivalent of the main source rocks of the North Sea. This Upper Jurassic formation, up to five hundred meters thick, outcrops in Britain as a belt stretching from Dorset to Yorkshire (Williams and Douglas, 1980). It is generally considered that the Kimmeridge Clay was deposited below wave base under a shallow shelf regime (Tyson et al., 1979) during a period of maximum eustatic rise and transgression (Gallois, 1976). Depositional conditions were mainly controlled by the topography of the basin and sedimentation and subsidence rates. The onshore basin topog- raphy led to the sedimentation of organic matter-rich mud- stones and to anoxic conditions in restricted areas only (Tri- bovillard et al., 1994). Previous studies on the bulk chemical features of Kimmeridge Clay kerogens, chiefly based on el- emental analyses and pyrolyses (Williams and Douglas, 1980, 1983; Farrimond et al., 1984; Pfendt, 1984; Eglinton et al., 1986, 1988a,b), showed a type II material with a low degree * Present address: Universite Bordeaux I, URA CNRS 197 Geo- logie & Oceanographie, F-33405 Talence cedex, Fr.ance 3731 of maturity, consistent with a maximal burial depth around 1,500 m (Williams, 1986). In addition, a recent bitumen ex- amination pointed to an important contribution of algae along with a minor contribution of land-derived organic material (Ramanampisoa and Disnar, 1994). Recently, high resolution measurements of total organic carbon (TOC) on Kimmeridge Clay cores from three bore- holes in the Cleveland basin (Yorkshire, UK) revealed a con- spicuous feature (Herbin et al., 1991, 1993). Pronounced, short-term, cyclic variations in TOC were observed with each microcycle corresponding to about 30,000 years. Moreover, it was noted (Ramanampisoa et aI., 1992) that the hydrogen index (HI) also exhibits substantial cyclic variation along with TOC. As a result of such parallel, wave-shaped, changes in TOC and HI, the oil potential of the Kimmeridge Clay formation shows a sharp maximum at the top of each cycle. The first petrographic observations carried out on Kim- meridge Clay kerogens, by light and UV fluorescence mi- croscopy (Tyson et al., 1979; Williams and Douglas, 1980; Smith, 1984), indicated a marked predominance of a seem- ingly structureless matter with bituminite as the main maceral according to Williams and Douglas (1980). A low contri- bution of recognizable elements, mainly of a terrestrial origin (humic material), was also noted (Bertrand et aI., 1990; Huc et al., 1992). Examinations by transmitted and reflected light microscopy were recently carried out on kerogens obtained via high resolution sampling of a microcycle, from one of the
Transcript

Geochimica et Cosmochimica Acta, Vol. 59, No. 18, pp. 3731-3747, 1995Copyright © 1995 Elsevier Science LtdPrinted in the USA. All rights reserved

0016-7037/95 $9.50 + .00

0016-7037(95)00273-1

Electron microscopy and pyrolysis of kerogens from the Kimmeridge Clay Formation,UK: Source organisms, preservation processes, and origin of microcycles

M. BOUSSAFIR,I F. GELIN,2 E. LALLIER-VERGEs, I S. DERENNE,2 P. BERTRAND, I. * and C. LARGEAU2

IUniversite d'Orleans, URA CNRS 724, Dept. des Sciences de la Terre, F-45067 Orleans cedex, France1Laboratoire de Chimie Bioorganique et Organique Physique, URA CNRS 1381, Ecole Nationale Superieure de Chimie de Paris,

II rue Pierre et Marie Curie, F-7523I Paris cedex 05, France

(Received November 29, 1994; accepted in revised form June 9, 1995)

Abstract-Recent studies revealed short-term cyclic variations (microcycles) in total organic carbon(TOC) and the hydrogen index (HI) in the Kimmeridge Clay Formation, an organic-rich deposit consideredto be a lateral equivalent of the main source rocks of the North Sea. In addition, three different types oforganic matter that all appear to be amorphous when observed by light microscopy (AOM) were recog­nized. Together, these AOM types account for over 80% of total kerogen and their relative abundancesshow large variations along each microcycle. In the present work, transmission electron microscopy (TEM)observations were carried out on samples (whole kerogens, kerogen subfractions only comprising a singletype of AOM, selected rock fragments) corresponding to typical points within a microcycle and obtainedvia high resolution sampling. The nature and the relative abundances of the products generated by Curie­point Py-GC-MS and off-line pyrolyses of isolated kerogens were also determined for two selectedsamples corresponding to the beginning and the top of the microcycle. Combination of such ultrastructuralobservations, including some semiquantitative studies, and the analysis of pyrolysis products allowed ( 1)determination of the ultrastructural features of the three AOM types thus providing what we believe to bethe first example of correlations between light microscopy (palynofacies, in situ maceral analysis) andTEM observations on "amorphous" fossil materials; (2) identification of the source organisms and elu­cidation of the mode of formation of the different AOM types in the Kimmeridge Clay; (3) explanationof the variations in their relative abundances taking place along a microcycle and establishment of tightcorrelations with TOC and HI changes; and (4 ) explanation of the origin of the microcyclic variations inkerogen quantity (TOC) and quality (HI) occurring in the Kimmeridge Clay Formation. Interrelationshipsbetween primary productivity, sulphate reduction intensity, and lipid' 'vulcanization" likely played a majorrole in the control of such variations.

1. INTRODUCTION

The aim of this work was to gain information on kerogengenesis and source organisms in the Kimmeridge Clay and onthe origin of the short-term cyclic variations, in both kerogenquantity and quality, occurring in this formation. KimmeridgeClay is a marine deposit, with alternating organic-rich shalesand marls, considered as a lateral equivalent of the mainsource rocks of the North Sea. This Upper Jurassic formation,up to five hundred meters thick, outcrops in Britain as a beltstretching from Dorset to Yorkshire (Williams and Douglas,1980). It is generally considered that the Kimmeridge Claywas deposited below wave base under a shallow shelf regime(Tyson et al., 1979) during a period of maximum eustatic riseand transgression (Gallois, 1976). Depositional conditionswere mainly controlled by the topography of the basin andsedimentation and subsidence rates. The onshore basin topog­raphy led to the sedimentation of organic matter-rich mud­stones and to anoxic conditions in restricted areas only (Tri­bovillard et al., 1994). Previous studies on the bulk chemicalfeatures of Kimmeridge Clay kerogens, chiefly based on el­emental analyses and pyrolyses (Williams and Douglas, 1980,1983; Farrimond et al., 1984; Pfendt, 1984; Eglinton et al.,1986, 1988a,b), showed a type II material with a low degree

* Present address: Universite Bordeaux I, URA CNRS 197 Geo­logie & Oceanographie, F-33405 Talence cedex, Fr.ance

3731

of maturity, consistent with a maximal burial depth around1,500 m (Williams, 1986). In addition, a recent bitumen ex­amination pointed to an important contribution of algae alongwith a minor contribution of land-derived organic material(Ramanampisoa and Disnar, 1994).

Recently, high resolution measurements of total organiccarbon (TOC) on Kimmeridge Clay cores from three bore­holes in the Cleveland basin (Yorkshire, UK) revealed a con­spicuous feature (Herbin et al., 1991, 1993). Pronounced,short-term, cyclic variations in TOC were observed with eachmicrocycle corresponding to about 30,000 years. Moreover,it was noted (Ramanampisoa et aI., 1992) that the hydrogenindex (HI) also exhibits substantial cyclic variation alongwith TOC. As a result of such parallel, wave-shaped, changesin TOC and HI, the oil potential of the Kimmeridge Clayformation shows a sharp maximum at the top of each cycle.

The first petrographic observations carried out on Kim­meridge Clay kerogens, by light and UV fluorescence mi­croscopy (Tyson et al., 1979; Williams and Douglas, 1980;Smith, 1984), indicated a marked predominance of a seem­ingly structureless matter with bituminite as the main maceralaccording to Williams and Douglas (1980). A low contri­bution of recognizable elements, mainly of a terrestrial origin(humic material), was also noted (Bertrand et aI., 1990; Hucet al., 1992). Examinations by transmitted and reflected lightmicroscopy were recently carried out on kerogens obtainedvia high resolution sampling of a microcycle, from one of the

3732 M. Boussafir et al.

three boreholes previously studied by Herbin et al. (1991,

1993): Cleveland basin, Marton 87 well (Ramanampisoa etaI., 1992; Pradier and Bertrand, 1992). The palynofaciesstudy (transmitted light microscopy) revealed that the domi­nant "amorphous" organic material (AOM, over 80% oftotal

organic matter) is in fact comprised of three distinct types ofparticles that can be characterized by differences in textureand color: orange homogeneous flakes with sharp edges (or­ange AOM, Fig. la); brown heterogeneous flocks with fuzzy

outlines (brown AOM, Fig. Ib); opaque aggregates (black

AOM, Fig. Ic). It was observed that important differences in

the relative abundances of the three above types take placealong the cycle. Moreover, Ramanampisoa et al. ( 1992) noteda strong parallel between the variations in TOC and HI values,

on the one hand, and the abundance of orange AOM on the

other hand. The beginning and the end of the cycle are thuscharacterized by a predominance of the black and brown

AOMs associated with land-derived debris, whereas the peak(or top) of the cycle, where maximum TOC and HI occur,

shows a major contribution of orange AOM.

The purposes of the present study were therefore to deriveinformation on (I) the ultrastructural and chemical features,

the source organisms, and the mode of formation of thesedifferent types of so-called "amorphous organic matter"(AOM), (2) the origin of the substantial variations in AOM

relative abundances occurring along the microcycle, (3) the

relationships between such variations and TOC and HIchanges, and (4) the cause ( s) of the cyclic variations ob­served in the Kimmeridge Clay. The conclusions obtainedfrom the study may also aid in understanding the factors thatcontrol organic matter quantity (TOC) and quality (HI) in

marine sediments in general. To this end, transmission elec­tron microscopy observations (TEM) and pyrolytic studies

were carried out on isolated kerogens, corresponding to typ­ical points from the above Kimmeridge Clay microcycle. Bulk

kerogens and kerogen subfractions, obtained by micromanip­

ulation and only comprising a single type of AOM particles,were both examined by TEM. Additionally, in situ TEM ob­servations were also directly performed on untreated rockfragments.

2. EXPERIMENTAL

2.1. Samples

The studied microcycle (90 cm thick, corresponding to about30,000 years, depth in the core 128.2 to 129.1 m) was sampled in theEudoxus zone of the Marton 87 borehole, Pickering Vale, ClevelandBasin, Yorkshire, UK (Herbin et al., 1991). The corresponding geo­logic section is discussed in Herbin et al. ( 1991, 1993), Tribovillardet al. ( 1994), and Desprairies et al. ( 1995). TOC values range fromca. 2% (beginning and end of the microcycle) to 9.5% at the top ofthe microcycle and HI from ca. 250 to ca. 780 mg of hydrocarbons Ig of organic carbon (Ramanampisoa et aI., 1992). The microfaciesobserved by light microscopy (reflected light and UV excitation onpolished sections) is laminated throughout the cycle and no biotur­bation features are observed. The distribution of the organic matterappears to be relatively heterogeneous. In samples with relatively lowTOC values, the organic matter is comprised predominantly of smallirregular or angular particles, from 5 to 20 j.Lm in size, identified asinertinite. For samples with TOC values> 4%, the microfacies arecharacterized by an increase in size and abundance of more or lesscontinuous elongated organic particles (up to 500 j.Lm) which arelargely predominant at the peak of the cycle; when studied by re-

flected light they appear as bituminite maceral (Pradier and Bertrand,1992). For the highest TOC values, the organic matter tends to formthick laminations and to occur within the mineral matrix in greateramounts.

The eight samples examined were selected on the basis of previouslight microscopy and Rock-Eval high resolution studies (centimetricscale) carried out by Ramanampisoa et al. (1992). These samplescorrespond to the beginning of the cycle (128.23 m), the sectionencompassing the increasing TOC values (128.62 and 128.67 m),the top of the cycle (128.74 and 128.75 m), the section coveringdecreasing TOC values (128.79 m), and the end of the cycle (129.09and 129.1 m). In agreement with the previously stated occurrence oflow maturity type II material in the Kimmeridge Clay, elementalanalysis indicated atomic H/C ratios of 1.1 to 1.2 for the selectedkerogens.

Kerogens used for petrographic and pyrolytic studies were isolatedfrom these different rock samples via the classical HFIHCI treatment.Direct TEM observations were then carried out on whole kerogensand on handpicked kerogen subfractions comprising only a singletype of AOM. The subfractions were obtained by picking out AOMparticles under a stereomicroscope, with a microsyringe coupled toa micromanipulator, from aqueous suspensions of whole isolated ker­ogens (Boussafir et al., 1994a). Fragments of untreated rocks domeinated by a single type of maceral (e.g., bituminite), or by the organo­mineral matrix, were also selected by light microscopy, from polishedsections, for in-situ TEM observations.

All the above materials (whole kerogens, kerogen subfractions,and selected rock fragments) were fixed in osmium tetroxide, em­bedded in resin, cut with an ultramicrotome, and stained as describedin Boussafir et al. (1994b) prior to TEM examination using a STEMJeollOO CX.

2.2. Pyrolyses

"Off-line" pyrolyses and identification of pyrolysis products wereperformed according to Largeau et al. (1986). Briefly, kerogens arefirstly heated at 300°C for 20 min so as to eliminate adsorbed compoundsby thermovaporization and, after extraction with CHCI,lMeOH (21 I ),the insoluble residue is pyrolysed at 4OQ°C for 2 h under a helium flow.The released products are trapped in CHC1, at -5°C, separated by col­umn chromatography on activity 2 alumina into three fractions of in­creasing polarity by eluting with hexane, toluene, and methanol, respec­tively. The hexane-eluted, low polarity products are further separated intothree subfractions by lLC on SiOrAgNO, (10%). The various pyro­lysate fractions thus obtained were analysed by GC-MS (DBI capillarycolumn, 60 m X 0.25 mm i.d., oven heated from 120 to 260°C at 4°CImin). Curie point pyrolysis-gas chromatography-mass spectrometry(Py-GC-MS) was performed as described in Derenne et al. (l992a)using a Curie point pyrolyser (FOM3-LX) and a ferromagnetic wire witha Curie temperature of 610°C.

3. RESULTS AND DISCUSSION

3.1. Electron Microscopy

A first series of TEM observations on whole isolated ker­

ogens, corresponding to typical points of the microcycle, re­vealed the presence of various ultrastructures in the eight sam­

ples examined: a massive, gel-like, amorphous material; very

thin lamellae; small ligneous debris; a granular material; anda diffuse amorphous material (Fig. 2). The above ultrastruc­tures were shown to occur all along the microcycle but sharp

differences in their relative abundances were noted. A semi­quantitative study was therefore carried out by examining, for

each kerogen sample, 60 to 75 ultrathin sections. The gel-likematerial was thus shown to exhibit large changes in relativeabundances that parallel TOC and HI variations (and hencevariations in orange AOM relative contribution). This struc­tureless material is only a minor constituent at the end andthe beginning of the cycle (about 5 to 10% of total AOM)

Provenance and preservation of kerogen 3733

FIG. 1. The three different types of AOM observed in transmitted light microscopy (palynofacies) from KimmeridgeClay isolated kerogens: (la) orange AOM; (lb) brown AOM; (Ie) black AOM; (ld) Elongated organic particles ofbituminite "b" associated with framboidal pyrite "fp" as seen by reflected light (UV excitation) in the microtextureof high-TOC samples.

but markedly predominates at the top (up to 75%). Reversechanges in relative abundances were noted, along the micro­cycle, for other types of ultrastructures (lamellae, ligneousdebris) and the latter are only abundant at the end and thebeginning of the cycle. The above semiquantitative featuressuggested that tight correlations may occur between the threetypes of AOM, previously detected by transmitted light mi­croscopy (palynofacies), and the different ultrastructuresidentified at higher magnification in the present TEM study.The occurrence of such correlations was fully confirmed byTEM examination of kerogen subfractions, composed of asingle type of AOM, obtained by picking out particles underthe light microscope.

In fact, when taken together, the above TEM observationsindicate the following.

1) The orange AOM corresponds to the massive gel-like ma­terial. This material is characterized by a completely ho­mogeneous texture without any apparent biological struc­tures and it appears truly amorphous, even at a very high

magnification (up to xSO,OOO). That this material is a dis­crete entity is suggested by well-defined, distinct edges(Fig. 2a). The orange AOM can therefore be defined as a"nanoscopically amorphous organic matter." Such a ma­terial is often associated with pyrite crystals and framboids(Fig. 2b). In addition, Electron Energy Loss Spectrometry(EELS) pin-point analyses (Boussafir et aI., 1995)showed that the orange AOM contains substantial amountsof organic sulphur (atomic SIC ratios of 0.9 to l.l).

2) The brown AOM corresponds to very thin lamellae witha regular thickness ranging from 50 to 400 nm (Fig. 2c).Similar structures, with thickness up to 60 nm, were pre­viously observed, by TEM, in a number of kerogens iso­lated from source rocks and oil shales (Raynaud et aI.,1988; Largeau et aI., 1990a,b; Lugardon et aI., 1991; Der­enne et aI., 1991a, 1993) and termed ultralaminae (Lar­geau et aI., 1990a). No pyrite and no organic sulphur weredetected in the brown AOM.

3) The black AOM corresponds to a mixture dominated byminute ligneous debris (Fig. 2d) and also comprises a dif-

3734 M. Boussafir et al.

FIG. 2. Main types of ultrastructures observed by TEM in Kimmeridge Clay isolated kerogens: (2a, 2b) massive, gel­like, nanoscopically amorphous organic matter associated with pyrite (P), corresponding to the orange AOM, theelectron-lucent areas are voids originating from the elimination of mineral constituents during kerogen isolation; (2c)ultralaminae corresponding to the brown AOM; (2d) ligneous debris (major constitutents of the black AOM); (2e, 2f)diffuse nanoscopically amorphous matter contributing to the black AOM; this continuous organic network appears asa grey material whereas "ghosts" of minerals (quartz: Q; pyrite: P; clays: Cl; coccolithes: C) are seen as white areas.In 2f pyrite has been eliminated via nitric acid treatment.

Provenance and preservation of kerogen 3735

fuse nanoscopically amorphous material; low amounts ofgranular organic matter are also detected. The amorphousdiffuse material was closely associated with mineral con­stituents in the untreated rock and built an organic networkwithin the mineral groundmass. This network is well pre­served following mineral elimination by chemical leaching(Fig. 2e, f ). Such a preservation reveals the occurrence ofa continuous organic network exhibiting a high degree ofcohesion.

The above TEM observations thus revealed major ultra­structural differences between the three types of AOM iden­tified by transmitted light microscopy. The orange AOM istruly amorphous whereas the brown and black forms of AOMcomprise well-defined structures that could not be detected bylight microscopy. The combination of TEM examination onwhole kerogens and on kerogen subfractions obtained by mi­cromanipulation thus provided what we believe to be the firstexample of correlations between "amorphous" organic ma­terials identified in palynofacies studies and ultrastructuresobserved by TEM. Moreover, close relationships can be notedbetween the relatives abundances of these different ultrastruc­tures and TOC and HI values (Fig. 3).

Finally, in-situ TEM observations were carried out on rockfragments selected by light microscopy from polished sec­tions. The purpose of these observations was (I) to attempt

a correlation between the main ultrastructures identified aboveby TEM in isolated kerogens (and hence the AOM types ob­served in palynofacies studies) and the macerals identified bylight microscopy from polished sections of untreated rocksamples and (2) to examine the association of these maceralswith the mineral constituents. Selected bituminite particles,when observed by TEM, show a massive nanoscopicallyamorphous gel-like organic matter (Fig. 4a) embedding clayparticles and pyrite framboids (Fig. 4b). In addition, previousanalyses performed by X-ray mapping (Lallier-Verges et aI.,1993a) revealed a specific enrichment in organic sulphurwithin bituminite particles from the Kimmeridge Clay. Bitu­minite and the orange AOM therefore exhibit the same lackof ultrastructure, the same close association with sulphur, andthe same variability in relative abundances along the micro­cycle. The bituminite maceral thus appears to correspond tothe orange AOM observed in palynofacies studies. In sharpcontrast, the ultralaminar structures (and hence the brownAOM) do not correspond to any recognizable macerals. Theyare thought to be dispersed in the organo-mineral matrix andthus difficult to be identified by light microscopy from pol­ished sections. The ligneous debris observed in the blackAOM present the same morphology at high magnification andthe same ultrastructural features as inertinite macerals. Be­cause of their small size (about I JLm), these debris were notdetected by light microscopy.

•750

•0 00

00

0 0

00000 lP •0 0

0 0 00 0 0 0 •0 00 0 • 0 0

000

o ,0 0mdP

00 0

04 :0

00

0

~Cb0

550

650

150

~ 4500

0fJJ:til 350Es:

250

1098765432

50 +---.......................-r-........-.-.,.....,.-,-....................,.....,--,--.-........-.-...................................-,-....-.-...-.--.-,........................-.....,.....,...-.--.-.,.....,.-,-...,.....,.........-.-~1

Toe %

FIG. 3. Main ultrastructures identified by TEM in Kimmeridge Clay isolated kerogens and relationships with TOCand HI variations along the microcycle (after Lallier-Verges et aI., 1993a). Filled circles represent the samples studiedby TEM. Ultralaminae (b, c) and minute ligneous debris (a) dominate in samples with relatively low TOC values whilethe massive amorphous matter (d) is predominant in kerogens with higher TOC.

3736 M. Boussafir et al.

FIG. 4. In-situ TEM observations of bituminite-rich rock fragments revealing the nanoscopically amorphous natureof bituminite (B) and its close association with clays (Cl). Such a close association between clays and bituminite hasbeen described in other source-rock samples (Bishop and Philp, 1994).

Further studies were carried out by pyrolysis so as to ex­amine the chemical features, the source organisms, and themode of formation of the different constituents identifiedabove in Kimmeridge Clay AOMs.

3.2. Pyrolyses

On the basis of the above petrographic results, two samplestermed!! (128.23 m) and I (128.75 m) were selected forpyrolytic studies; they correspond to the beginning and thetop of the microcycle, respectively. Sample!! is thus char­acterized by lower values for TOC and HI ( 1.8% and 255 mgHC/g Corg ) when compared to I (9.5% and 582 mgHC/g Corg )'

3.2.1. Curie-point Py-GC-MS of isolated kerogens

GC-MS analysis of the 610°C flash pyrolysates revealedcomplex mixtures of products dominated by doublets corre­sponding to n-alkanes and n-alk-I-enes from C7 to C,o (Fig.5). These two dominant series show maxima around CIS, nosignificant difference in their distributions, and relative inten­sities was noted between kerogens!! and I. These observa­tions indicate an abundant contribution of long polymethy­lenic chains in both samples. The GC-MS traces also showthe presence of a large number of compounds eluting betweenthe alkane/alkene doublets. These compounds correspond, asindicated for some of them on Fig. 5, to series of alkylatedthiophenes, alkylated phenols, alkylated benzenes, and to ad­ditional series of linear/branched, (un)saturated hydrocar­bons. The major difference between!! and I flash pyrolysatesis to be found in the higher relative abundance of thiopheniccompounds for I. This difference can be assessed from thethiophene index: ThI = [2,3-dimethylthiophene]/[non-I-ene+ I,2-dimethylbenzene]. This ratio is known to provide aconvenient way for a rapid estimation of organic sulphur con­tent in kerogens (Eglinton et aI., 1990). A significantly higherratio is indeed noted (0.22 instead of 0.12) from I pyrolysatewhen compared to !!, thus indicating a larger contribution of

sulphur-containing moieties in the former sample. Such a dif­ference is also supported by bulk elemental analyses and theatomic Sorg/C ratio is about twice as high for T

3.2.2. Off-line pyrolyses of isolated kerogens

Due to the very high complexity of the effluents producedupon thermal degradation of kerogens!! and I, "off-line"pyrolyses were also carried out so as to obtain further infor­mation on the generated products. Prior to GC-MS analysesthe crude pyrolysate was separated by column chromatogra­phy into three fractions (low polarity products, medium po­larity products, and polar products) accounting for about 35,25, and 40% of the total pyrolysates, respectively. Thereafterthe low polarity, hexane-eluted fraction was further separatedby TLC on Si02-AgNO, and the polar methanol-eluted frac­tion was separated by extraction into acid and non-acid com­pounds.

3.2.3. Low polarity products

The different series of compounds identified in the low po­larity fractions are reported in Table I along with their distri­butions and relative abundances. In addition to the alreadymentioned n -alkanes and n-alk-I-enes series, a number ofother homologous series were identified. They comprise var­ious types of branched alkanes, normal unsaturated hydrocar­bons, and compounds with long normal alkyl chains associ­ated with a cyclopentyl, a cyclohexyl, a benzene, a thiophene,a benzothiophene, or a naphthalene ring (in addition the ben­zene and thiophene rings can be substituted by one or twomethyl groups). Despite the occurrence of these differenttypes of cyclic units, the identification of the low polaritypyrolysis products confirms the important contribution oflong(CH2 )n chains to both!! and I kerogens. Indeed it can benoted that (I) each fraction is dominated by a series corre­sponding to normal compounds and that other series of acyclicunbranched products also occur in substantial amounts, (2)the cyclic structures are always associated with long normal

Provenance and preservation of kerogen 3737

•>.- •'" 14 j • 19c I

Q,I • • • i •- • •.5 • B• ••Q,I • .. •.::... •

CIS •'ai

0"'" • Pr • • •9

\OJ •• 24~ l OH •

I·. er • •

~• ••

I.

WM~~~ }1 I

,I ~IiJ ~1

JJi~ il ' 'II

~,J~ iretention time

• • ~~ 14 J...';j • • 19I: •~ OJ .Pr •.5 •• •• •Q,I •.::... • TCIS ••'ai"'" ~. ••

"- . •

0 :. \ 24• •

\;

I•

•• ·Uw•1· I'I'I~

IiIII,

il

'~ I ~', ~,'~

1\ f

retention timeFIG. 5. Total ion current (TIC) traces of the flash pyrolysates of kerogens Band T. Filled circles and filled squaresindicate the homologous series of n-alk-l-enes and n-alkanes, respectively. Numbers indicate their chain length. Thestructure of some major compounds is indicated and Pr designates prist-l-ene.

3738 M. Boussafir et al.

Table 1. Nature and relative abundances of the homologous series of low

polarity compounds identified in the 400°C pyrolysates of kerogens .Il and I. a.

.Il I

constituents r.a. constituents r.a.

n-alkanes C14-C31 (C17) 1 Cl2-C30 (CI5) 1

isoprenoid alkanes C15-C20 (CI8) 0.16 C15-C20 (CI8) 0.23

dirnethylalkanes I b C17-C31 (C25) 0.08 n.d.

dirnethylalkanes II b C15-C31 (C2I) 0.05 n.d.

3-methylalkanes C14-C30 (C20) 0.10 Cl3-C22 (CI5) 0.04

n-alkylcyclopentanes C14-C26 (C20) 0.27 n.d.

n-alkylcyclohexanes C14-C25 (CI9) 0.05 Cl2-C23 (CI4) 0.13

n-alkenes (E) C14-C26 (C17) 1 Cl3-C25 (CI5) 1

n-alkadienesc C14-C24 (C17) 0.52 Cl2-C22 (CI5) 0.62

n-alkylbenzenes C14-C22 (CI6) 0.16 Cl2-CI8 (CI4) 0.35

n-alkyltoluenes C14-C23 (CI6) 0.18 Cl2-C21 (CI5) 0.28

n-alkyldimethylbenzenes C14-C20 (CI6) 0.16 Cl2-CI8 (CI4) 0.33

2-n-alkylthiophenes Cl2-C24 (CI4) 0.15 CIO-C24 (Cl2) 0.65

2-n-alkyl,5-methylthiophenes Cl2-C24 (CI4) 0.33 CIO-C20 (Cl2) 0.86

n-alkyl, dimethylthiophenes Cl2-C24 (CI5) 0.06 CIO-C20 (Cl2) 0.28

n-alk-I-enes C14-C28 (CI8) Cl2-C28 (CI5)

n-alkenes (Z) C14-C25 (CI8) 0.09 n.d.

n-alkadienesc C14-C22 (CI8) 0.08 C12-C18 (CI4) 0.11

n-alkylbenzothiophenes C8-Cl2 trace C8-Cl2 (Cl I) 0.40

n-alkylnaphthalenes CIO-CI4 trace CIO-CI4 (Cl2) 0.15

'Identifications were carried out following further separation of the hexane-eluted com­pounds into three fractions by TLC. La.: relative abundances of the homologous seriescalculated with respect to the predominant series of each fraction (ratios of maxima); thebracketed values correspond to the maximum of the series. "Based on their mass spectra,these compounds correspond either to 3,7-dimethylalkanes or to 3,w-7-dimethylalkanes(series I) and either to 3,5-dimethylalkanes or to 3,w-5-dimethylalkanes (series II). 'Doublebond positions could not be determined. The alkadienes series of the second and thirdfractions exhibit different retention times and do not correspond to a,w-alkadienes. n.d.:not detected.

alkyl chains, and (3) the abundant series of alkylmethylthio­phenes corresponds to the isomer (2-alkyl,5-methyl) with a"normal carbon skeleton", i.e., to compounds derived fromsulphur incorporation in unbranched precursors (Sinninghe­Damste et aI., 1989; Sinninghe-Damste and de Leeuw, 1990).This major contribution of long polymethylenic chains wasconfirmed, as shown thereafter, by analysis of the toluene­

and methanol-eluted fraction of!! and! pyrolysates; it is alsoconsistent with previous studies on Kimmeridge Clay kero­gens from the Dorset area using solid state DC NMR spec­trometry and RU04 oxidation (Boucher et aI., 1990).

As discussed below, the abundance and the nature of theorganic sulphur compounds (OSC) generated from samples!! and! provide important information. As shown in Table1, although the same series with similar distributions are ob­served in both cases, much larger relative amounts of OSC

are always obtained from T In agreement with flash pyroly­ses, the latter kerogen therefore comprises a markedly highercontribution of sulphur-containing moieties. This differencecan be also illustrated by comparison of the two mass chro­matograms of m/z III of the total pyrolysates of !! and !(Fig. 6), highlighting the series of 2-n-alkyl, 5-methylthio­phenes, and n-alk-l-enes.

It is well documented that a rapid incorporation of sulphurinto various functionalized lipids takes place during early dia­genesis in the presence of reduced sulphur species (reviewedin Sinninghe-Damste and de Leeuw, 1990). Such an incor­poration occurs in lipids comprising carbon-carbon doublebonds but also, as recently demonstrated, in ketones and al­dehydes (Schouten et aI., 1994; Krein and Aizenshtat, 1994).Due to this "natural vulcanization," the above lipids can be­come tightly associated via covalent bonds within insoluble

Provenance and preservation of kerogen

o

3739

o 12•

o

o

o

o

o

19o

o

o

o

19•

o

oo

o

• •

retention time

o

B

• 15

19o

o

oo T

• • 0

• 0

150

19 0• •• 0

retention time

FIG. 6. Mass chromatograms of mlz III revealing the homologous series of 2-n-alkyl, 5-methylthiophenes (filledcircles) and n-alk-I-enes (empty circles) in the total "off-line" pyrolysates of kerogens T and B. Numbers indicate thelength of the alkyl chains. - -

3740 M. Boussafir et al.

Table 2. Nature and relative abundances of the homologous series of medium

polarity compounds identified in the 400°C pyrolysates of kerogens B. and I. a.

B. I

constituents r.a. constituents r.a.

n-alkan-2-ones C14-C31 (C17) ClO-C23 (Cl3) 1

branched alkyl, methylketonesb n.d. Cll-C21 (Cl3) 0.3

n-alky1, ethylketones n.d. ClO-C25 (Cl3) 0.2

n-alkylnitriles Cll-C21 (C14) 0.8 trace

alkylated indoles C9-C12 0.9 C9-Cll 0.1

alkylated quinolines ClO-C12 0.15 ClO-Cl3 0.3

a Toluene-eluted compounds. r.a.: relative abundance of the homologous series calculated withrespect to the predominant series (ratios of maxima). n.d.: not detected. The bracketed valuescorrespond to the maximum of each seriesb The methyl branch is not on carbon 3, but its precise location was not established.

macromolecular structures. As a result such lipids are likelyto be protected against microbial mineralization and thus toefficiently contribute to kerogen formation (Sinninghe-Dam­ste et al., 1989). The main outcome of sulphur incorporationmay therefore be to strongly reduce the degradation thatwould take place if these lipids remained in a free state. Re­garding morphological features, the "vulcanization" processshould lead to amorphous kerogen fractions. As already em­phasized, the orange AOM that occurs in Kimmeridge Claykerogens appears truly amorphous when examined by TEM;it is often associated with pyrite and pin-point analyses indi­cated a substantial content of organic sulphur. In addition, itsrelative abundance sharply increases with TOC and is at amaximum at the top of the cycle, what is to say for sample I.Moreover, as just discussed, the level of organic sulphur-con­taining moieties is higher in kerogen I when compared to .!!.Finally, it is noted that the different series of OSC generatedboth from.!! and I are commonly found in the pyrolysates ofkerogens which formed via sulphur incorporation into lipids.Taken together, the morphological and chemical features de­scribed above therefore strongly point to the formation of theorange AOM occurring in Kimmeridge Clay kerogens vialipid "vulcanization." The nature of the OSC generated uponpyrolysis indicate that such lipids were dominated by productscomprising long polymethyleneic chains as commonly noted,for example, in the total lipid fraction of most microalgae.This formation pathway for the orange AOM is also consistentwith previous observations on Tmax obtained by Rock-Evalpyrolyses (Ramanampisoa and Disnar, 1994) which indicateda slight but regular increase in Tmax when TOC decreases inthe microcycle (422°C for sample I and 427°C for .!!). Dif­ferences in maturity could not be responsible for Tmax varia­tions over this short-term cycle since no hydrothermal or vol­canic alterations occurred and the above differences were con­sidered as reflecting changes in the degree of organic sulphurcontent. In fact, it is well documented that sulphur-sulphurand sulphur-carbon bonds are much weaker than carbon-car­bon bonds (Orr, 1986), thus facilitating thermal degradation.

Indeed, it was previously reported that samples with the high­est Thl also show the lowest Tmax values (Eglinton et ai.,1989). A higher content of orange AOM thus likely accountsfor the lower Tmax observed in kerogen I when comparedto B.

3.2.4. Medium polarity products

The compound series occurring in the toluene-eluted frac­tion of .!! and I pyrolysates are reported in Table 2. In bothcases the fraction is dominated by a series of n-alkan-2-ones.This type of ketone was observed in the pyrolysates of all thealgaenans so far examined and, also, of all the kerogensknown to be derived from the selective preservation of theseresistant biomacromolecules (e.g., Largeau et al., 1986; Aa­viano et ai., 1994). The thermal cleavage of the ether bridgeslinking some of the hydrocarbon chains building up the mac­romolecular skeleton of algaenans accounts for the formationof the ketones (Gelin et ai., 1993). Substantial amounts ofaromatic, N-containing products, corresponding to series ofsubstituted indoles and quinolines, are also formed from .!!and I. However, these compounds have not been so far re­lated to a specific type of source organism and biomolecules;thus, the significance of their abundant presence in both pyrol­ysates has yet to be established.

Normal alkylnitriles are also generated in substantialamounts from kerogen.!! where they account for ca. 6% ofthe total pyrolysate; in sharp contrast, these compounds arenot clearly detected in the case of sample I. It is now welldocumented that such nitriles are specific pyrolysis productsof the algaenans building up the thin resistant outer walls ofnumerous species of green microalgae (Derenne et ai., 1991a,1992b). Unlike the n-alkan-2-ones mentioned above, thesen-alkylnitriles are not produced from all types of algaenanand a complete absence of such compounds was previouslynoted in the case of species (Botryococcus braunii and Te­traedron minimum) exhibiting thick resistant outer walls.These nitriles were also systematically detected in the pyroly-

Provenance and preservation of kerogen 3741

sates of ultralaminae-containing kerogens. Indeed as previ­ously shown, the latter structures, which can only be identifiedby electron microscopy, originate from the selective preser­vation of thin resistant outer walls of microalgae (Derenne etaI., 1991a, 1992c). It was also observed that important dif­ferences in n-alkylnitrile distribution occur depending onwhether the examined samples, i.e., algaenan-composed thinouter walls from extant green microalgae and ultralaminae­containing kerogens, are of a marine or lacustrine origin (Der­enne et al., 1992d). With regard to Kimmeridge Clay kero­gens, light microscopy and TEM observations revealed a ma­jor contribution of brown AOM in kerogen B and the bulk ofthis brown material appears to be composed~f lamellar struc­tures reminiscent of ultralaminae. The nature and the originof these lamellar structures was ascertained, from the pyrol­ysis experiments, by (I) the relatively abundant presence ofn-alkylnitriles in sample.!! pyrolysate and (2) the identicaldistribution of these nitriles when compared to those gener­ated from ultralaminae-containing marine kerogens and fromthin algaenan-composed outer walls of extant marine microal­gae. These morphological and chemical features observedfrom kerogen .!! therefore clearly indicate that the brownAOM of the Kimmeridge Clay was formed via the selectivepreservation of thin resistant microalgal walls. The difficultyof clearly establishing the presence of n-alkylnitriles in T py­rolysate reflects the relatively low level ofbrown AOM i~ thiskerogen.

Insoluble and nonhydrolysable macromolecular constitu­ents have been recently shown to also occur in some speciesof bacteria (Le Berre et al., 1991; Flaviano et al., 1994). Asobserved for algaenans, the so-called bacterans are located incell walls. However the skeleton of the cell wall of these bac­teria is chiefly composed of hydrolysable macromoleculeslike peptidoglycans. Accordingly bacterans, when isolated af­ter dr~stic ~as~ and acid hydrolyses, appear as amorphousmatenals. Smlliarly, the selective preservation of bacteransresults in the formation of amorphous kerogen fractions. In­deed, close chemical relationships have been established be­~ween some bacterans and kerogens dominated by nanoscop­Ically amorphous organic matter, as shown by TEM obser­va~ions (Flavi~o et al., 1994). As discussed above, sampleI IS charactenzed by a low contribution of selectively pre­served algal material whereas such constituents played a ma­jor role in kerogen.!! formation. Pyrolysis results also pointto a substantial contribution ofbacteran selective preservationin th.e genesis of the latter kerogen. Thus, the low polarityfraction of .!! pyrolysate is characterized by the presence ofsignificant amounts of dimethylalkanes (series I) and 3-meth­ylalkanes (Table I). These two series were previously de­tected, with similar relative abundances and distributions(odd-carbon-numbered dimethylalkanes and even-carbon­numbered 3-methylalkanes), in pyrolysates of bacterans andderived kerogens (Flaviano et al., 1994). Sample B constit­uents originating from the selective preservation ofbacteransshould appear as an amorphous material. Accordingly, suchconstituents are likely to be associated with the amorphousmatrix occurring in the black and/or in the brown organicmatter and embedding ligneous debris or minerals and ultra­laminae, respectively. In agreement with the low contributionof black and brown matter in kerogen I, bacteran-related

products are not detected, or in much lower amounts, in thepyrolysate of this sample (Table 1).

3.2.5. Polar products

Fatty acids. The nature and the relative abundances of thefatty acids generated upon pyrolysis of kerogens B and T areillustrated in Fig. 7. In both cases, the acid fracti~n is d~mi­nated by palmitic acid and also comprises substantial levelsof stearic and n -C22 saturated acid. The main series in thesefractions thus corresponds to CWC24 normal saturated acidsthat are characterized by a very strong predominance of theeven-carbon-numbered compounds, with Carbon PreferenceIndexes (CPI)t of 0.12 and 0.10 from samples Band T, re­spectively. Substantial amounts of normal mon~unsat;atedCI6 and CI8 acid are also observed, along with low levels ofdiunsaturated C 18 acids including the w( 9) cis, w( 12) cis iso­mer (linoleic acid). A very low contribution of branched, isoand anteiso, saturated Cn compounds is also noted. Compar­ison of.!! and I (Fig. 7) reveals almost identical compositionsregarding both their nature and their relative abundances, forthe acids released upon pyrolysis.

Fatty acids from extant microalgae and higher plants aregenerally characterized by a marked predominance of evennormal fatty acids, comprising substantial amounts of unsat­urated compounds along with saturated acids. However, it iswell documented that fatty acids, especially the unsaturatedones, are highly sensitive to microbial degradation. Accord­ingly, both fatty acids esterified into lipid structures and oc­curring in a free form are markedly altered during early dia­genesis as reflected by a sharp decrease in the predominanceof the even compounds and a nearly complete disappearanceof the unsaturated acids. Nevertheless, it has been recentlyshown that fatty acids can escape to diagenetic degradationswhen they are included, as esters, in insoluble and non-hy­drolysable macromolecules like algaenans (Largeau et aI.,1986; Kawamura et al., 1986; Fukushima and Ishiwatari,19.88; Derenne et al., 1991b). Such "tightly bound" fattyaCids can only be released by a thermal stress and the corre­s~ondin~ este~ moieties remain nearly unaffected followingdiagenesIs owmg to the very efficient protection provided by~e macromolecular network. As a result, the fatty acid frac­tions generated by pyrolysis of kerogens derived from al­gaenan selective preservation are characterized by a very lowlevel of alteration.

~atty .acids may also be protected from diagenetic degra­dation via a second way, based on sulphur incorporation. As~ready stre~sed, if "vulcanization" reactions take place dur­mg early diagenesis, various lipids can be involved in the~orm.ation ?f insoluble macromolecular structures. Such lip­Ids, mcludmg their fatty acid moieties, are then efficientlyprotected and, owing to this acquired resistance, they will onlyundergo negligible alterations after the "vulcanization" step.

The presence in kerogen pyrolysates of fatty acids whichdistribution reveals a low level of alteration should thereforereflect a contribution of the selective preservation of al­gaenans and/or the occurrence of a fast "vulcanization" dur-

t Calculated according to Bray and Evans (1961).

3742 M. Boussafir et al.

16•

16•

18•

8

24••

retention time

T

••

retention time

FIo. 7. TIC traces of the acid fractions from "off-line" pyrolysates of samples B and T. Filled squares and circlesindicate saturated and unsaturated fatty acids, respectively; filled triangles indicate branched saturated acids.

Provenance and preservation of kerogen 3743

ing early diagenesis. In fact, these two mechanisms of fattyacid protection appear to be important in !! and I kerogens,respectively. The very low level of alteration observed in theacid fractions of their pyrolysates is thus consistent with ( 1)the formation of the brown AOM, that predominates in !!, viaalgaenan selective preservation and (2) the formation of theorange AOM, accounting for the bulk of I, via lipid "vul­canization' , .

Non-acid compounds. GC-MS analysis ofthe non-acid po­lar fractions only indicated the presence of phenolic com­pounds. Their distribution is illustrated in Fig. 8 in the caseof kerogen !!. A similar distribution is also observed fromsample I although the relative abundance of these phenols issubstantially lower (1.5-2 times), as shown by comparisonof the GC traces of the crude pyrolysates. A number of phe­nols, including series of alkyl-substituted compounds, aregenerated upon pyrolysis of both!! and I However, the sub­fractions are dominated by mono- and dimethyl products. Thelatter phenols are well known as typical pyrolysis products ofdiagenetically altered lignins (Nip et aI., 1987). As alreadydiscussed, electron microscopy observations on picked outparticles showed the prevalence of minute ligneous debris inthe black AOM. In addition, the latter material is relatively

-

more abundant in sample !! than in I Such features, addedto the above results on the nature and abundance of phenolsin pyrolysates, indicate that a part of the black AOM of Kim­meridge Clay kerogens is of a terrestrial origin and is derivedfrom lignins.

The information obtained from the electron microscopy andpyrolytic studies, on the ultrastructure, the source organisms,and the mechanism of formation of the different types ofAOM in the Kimmeridge Clay are summarized in Table 3.Based on these results, and on the variations in AOM relativeabundances, a general scheme accounting for the occurrenceof TOC and HI microcycles in the Kimmeridge Clay For­mation can be established.

3.3. Origin of TOC and HI Cycles

Previous studies indicated that cyclic variations in the Kim­meridge Clay can result neither from changes in redox con­ditions (because bottomwater remained continuously anoxicduring deposition, as shown by the lack of bioturbation andthe analysis of inorganic trace elements), nor from changesin the extent of organic matter dilution by detrital and/or bio­genic minerals (Tribovillard et aI., 1992, 1994; Bertrand and

-+3

•1+2

- CIO• 45 C16~

- -• • • • • •

retention time

FIG. 8. Summed mass chromatogram of m/z 107 + 108 revealing the distribution of n-alkylphenols in the total' 'off­line" pyrolysate of sample B. Filled squares indicate the homologous series of ortho-n-alkylphenols and filled circlesindicate the homologous senes of the coeluting meta- and para-n-alkylphenols. Numbers correspond to the followingdimethylphenols: I = 2,4-dimethylphenol; 2 = 2,5-dimethylphenol; 3 = 3,5-dimethylphenol; 4 = 2,3-dimethylphenol;5 = 3,4-dimethylphenol. A similar mass chromatogram was obtained for kerogen I.

3744 M. Boussafir et al.

Table 3. Relationships between the ultrastructural features, the typical

pyrolysis products and the origin of the different types of AOM occuring in

Kimmeridge Clay kerogens.

Ultrastructure and ADM type

Massive nanoscopicallyamorphous orange ADM

Ultralaminae of brown ADM

Minute ligneous debris of blackADM

Nanoscopically amorphous matrixof brown and/or black ADM

Typical pyrolysisproducts *

Organic sulphurcompounds

Alkylnitriles

Alkylphenols

Branched alkanes

Process of formation

Lipid "vulcanization"

Algaenan selective preservation

Altered lignin preservation

Contribution of bacteranselective preservation

* Fatty acids cannot be related with a given ADM type; they can be associated with anyfraction described in this Table except the lignin-derived debris in the black ADM.

Lallier-Verges, 1993). Accordingly, the typical cycles occur­ring in the Kimmeridge Clay probably reflect changes in theprimary productivity of phytoplanktonic species without min­eral tests (Bertrand and Lallier-Verges, 1993; Bertrand et al.,1994). In addition, examination of kaolinite distribution, alsocarried out on Kimmeridge Clay samples from the Marton 87borehole, pointed to a climatic origin for the above changesin primary productivity (Desprairies et al., 1995).

It is well documented that increasing phytoplankton pro­ductivity is associated with faster sinking rates (Wefer,1989). The resultant higher export to deep water is due to theformation of aggregates and flakes (Jackson, 1990). As a re­sult, a larger proportion of the degradable constituents of theprimary microalgal biomass can escape mineralization in theupper part of the water column and thus reach the oxic-anoxicinterface. This increasing supply of metabolizable organicmatter to anoxic bottomwater should promote a prolificgrowth of sulphate-reducing bacteria (an ubiquitous group ofmicroorganisms in anoxic environments containing both sul­phate and organic matter sources, Widdel, 1988; Trudinger,1992; Elsgaard et al., 1994), hence an intense production ofhydrogen sulphide. The latter will rapidly react with availableiron to generate iron sulphides; however, a part may react withvarious lipids according to the already mentioned "vulcani­zation" process. Lipidic compounds that, otherwise, wouldbe heavily degraded may escape mineralization via this pro­cess. Owing to such an acquired resistance, the above lipidswill contribute to kerogen formation and a large increase inTOC in the corresponding sediments is thus achieved. (Thisis also supported by a recent study by Lallier-Verges et aI.,( 1994) concerned with sulphur content in the KimmeridgeClay (total S, organic S, pyrite) and suggesting an importantrole for sulphate reduction intensity in controlling organicmatter accumulation). All these interrelationships betweenprimary productivity, sinking rates, and sulphate-reduction in­tensity, and the resulting control on TOC values, as illustratedin Fig. 9, are fully consistent with the present morphologicaland chemical results. Thus, an increasing "vulcanization"should be reflected by the formation of high TOC sediments

with an abundant content of orange AOM, as observed at thetop of the cycle. Moreover, the morphology of the orangeAOM particles and their close association with sulphate-re­duction markers (iron sulphides) and clays indicate that sucha material was probably generated very early as flocks ormats.

As shown in Fig. 9, decreasing primary productivity willresult in lower sinking rates. Extensive mineralization of thedegradable constituents of microalgae will therefore takeplace in the oxic part of the water column. Accordingly, onlya weak sulphate reduction intensity will develop in anoxicbottomwater. Under these conditions the amount of buriedorganic matter will be relatively low and chiefly comprisedof biomacromolecules exhibiting a high intrinsic resistance todiagenetic degradations such as the refractory constituents ofalgal cell-walls and (to a lesser extent) lignins. Indeed, thekerogen samples corresponding to relatively low TOes, i.e.,at the beginning and the end of the cycle, exhibit a low contentof orange AOM and are dominated by the brown and blackAOMs.

The interrelationships illustrated in Fig. 9 can also accountfor the cyclic changes in HI occurring, along with TOC vari­ations, in the Kimmeridge Clay Formation. Thus, the high HIvalues observed at the top of the TOC cycle result from themarked predominance of the orange AOM. The latter, beingderived from sulphur incorporation into long chain lipids,should be characterized by a high oil generation potential.Resistant biomacromolecules from algal cell walls are knownto be highly aliphatic. As a result, the relatively abundantpresence of brown, ultralarninae-composed AOM in kerogensamples with lower TOC, corresponding to the beginning andthe end of the cycle, will favour high values for HI. In sharpcontrast the black AOM, mainly derived from lignins, shallbe characterized by extremely low HI. The substantial con­tribution of this material in the samples with lower TOC ac­count for their relatively low HI. Moreover, in such sampleswith TOC values around 2%, a mineral matrix effect (Espi­talie et aI., 1980; Katz, 1983) may also contribute to HI low­ering.

Provenance and preservation of kerogen 3745

/Phytoplankton /sinking ratesproductivity - -

Top of cycle

I exportation ofmetabolizable OMto anoxic zone

"vulcanization"

/TOC

orange "AOM"dominates

/IiPid preservation(acquired resistance)

/sulphate - reduction

buried OM mainly #to resistant biomacro­molecules (intrinsicresistance)

\Phytoplanktonproductivity -

End or beginningof cycle

brown and black"AOM" dominate

\Sinking rates _\

exportation ofmetabo~izable OMto anoxIc zone

~_ \ sulphate-reduction

~IG: 9. Postulated rel~tionships between primary productivity, sulphate reduction intensity, TOC and HI microcyclicvanatlons, and the dommant type(s) of AOM in Kimmeridge Clay kerogens.

4. CONCLUSIONS

The combination of high resolution sampling of a Kim­meridge Clay microcycle, of transmission electron micro­scopy observations on total isolated kerogens, on handpickedkerogen particles and on untreated rock fragments, along withflash and off-line pyrolytic studies allowed:

I) Precise information to be derived on the morphologicalfeatures of the three types of so-called "amorphous" or­ganic matter previously defined by light microscopy in theKimmeridge Clay. The orange AOM is truly amorphouswhereas the brown AOM is chiefly composed of ultralam­inae and the black AOM of minute ligneous debris em­bedded within a diffuse nanoscopically amorphous matrix.It was also shown that the maceral bituminite, observed insitu from polished sections, corresponds to the orangeADM identified in palynofacies studies.

2) Elucidation of the sources and modes of formation of thethree above types of organic matter. The orange AOMoriginates from sulphur incorporation into lipids, whereasthe brown AOM is derived from the selective preservationof algaenan-composed thin outer walls of microalgae. Theblack AOM is likely to be highly heterogeneous and it wasshown to comprise contributions from altered lignins andselectively preserved bacterans.

3) Explanation of ( I ) the large changes in the relative abun­dances of the three types of AOM occurring, in Kimmer­idge Clay kerogens, along a ca. 30,000 years microcycleand (2) the origin of the parallel, cyclic, wave-shaped vari­ations in kerogen quantity (TOC) and quality (HI) typicalof the Kimmeridge Clay Formation. Such variationsstemmed from modifications in primary productivity rate.However, the production of microalgal biomass did notexert a direct and simple control on TOC and HI. In fact,

the extent of lipid "vulcanization" was probably a majorparameter that strongly amplified the effects of primaryproductivity variations originating from global environ­mental changes. The values of TOC and HI and the relativeabundances of the different types of AOM in the Kim­meridge Clay appear therefore to be controlled by a bal­ance between a number of interrelated processes con­cerned both with the nature of the deposited biomass anddeposition conditions: (I) the extent of terrestrial materialcontribution (minute ligneous debris of the black mate­rial), (2) the level of resistant biomacromolecules in thesource microorganisms (algaenans and bacterans forming,via selective preservation, the bulk of the brown material(ultralaminae ), and a part of the amorphous matrix of theblack and/or brown ADM's, respectively), and (3) theintensity of lipid "vulcanization" leading to the orangematerial. This intensity being controlled by several factorsincluding the amount of metabolizable organic matterreaching the oxic-anoxic interface, the occurrence of suit­able conditions, in anoxic bottomwater, allowing for a pro­lific growth of sulphate-reducers and a somewhat limitedsupply of iron so that the produced HzS is not entirelytrapped as pyrite. Finally, the first of the above parametersis determined, in tum, by the depth of the oxic part of thewater column and by primary productivity-controlledsinking rates. The OM accumulated in the KimmeridgeClay thus chiefly originates from the contributions of bothmacromolecular compounds with a high intrinsic resis­tance to diagenetic degradations, like algaenans, bacteransand lignins, and lipidic components that acquired a resis­tant nature via "vulcanization." Kerogen quality (HI) re­flects the balance between these different contributions,high HI values being promoted by large levels of "vul­canized" lipids, and/or selectively preserved algaenans.

3746 M. Boussafir et al.

Acknowledgments-Financial support was provided by the ResearchGroup GdR 942 (CNRS, IFP, Total, Elf Aquitaine, Universitesd' Orleans et de Paris-Sud). A special thank is devoted to all thescientists of this Research Group. We are also indebted to D. Jalabertfrom the "Service central de microscopie electronique de I'Univer­site d'Orleans" for its technical assistance. Dr. J. W. de Leeuw(NIOZ, Netherlands) is gratefully acknowledged for permitting ac­cess to the Py-GC-MS instrument. We are grateful to the reviewers,Drs. T. I. Eglinton, B. Horsfield and S. W. Imbus and the AssociateEditor for constructive comments.

Editorial handling: 1. T. Senftle

REFERENCES

Bertrand P. and Lallier-Verges E. (1993) Past sedimentary organicmatter accumulation and degradation controlled by productivity.Nature 364, 786-788.

Bertrand P. et al. (1990) Examples of spatial relationships betweenorganic matter and mineral groundmass in the microstructure ofthe organic-rich Dorset Formation rocks (Great Britain). Org.Geochem. 16,661-675.

Bertrand P., Lallier-Verges E., and Boussafir M. (1994) Enhance­ment of both accumulation and anoxic degradation of organic car­bon controlled by cyclic productivity: a model. Org. Geochem. 22,511-520.

Bishop A. N. and Philp R. P. (1994) The potential for amorphouskerogen formation via adsorption of organic material at mineralsurfaces. 207 1h American chemical Society 1994, National Meet­ing, March 13-18, San Diego, CA, USA.

Boucher R. 1., Standen G., Patience R. L., and Eglinton G. (1990)Molecular characterisation of kerogen from the Kimmeridge clayformation by mild selective chemical degradation and solid state13C-NMR. Org. Geochem. 16,951-958.

Boussafir M., Lallier-Verges E., Bertrand P., and Badaut-Trauth D.( I994a) Etude ultrastructurale de matieres organiques micro-pre­levees dans les roches de la "Kimmeridge Clay Formation"(Yorkshire, UK). Bull. Centres Rech. Explor. Prod. Elf-Aquitaine,pp.275-277.

Boussafir M., Lallier-Verges E., Bertrand P., and Badaut-Trauth D.( I994b ) Structure ultrafine de la matiere organique des rochesmeres du kimmeridgien du Yorkshire (UK). Bull. de la Soc. Geol.de Fr. 165,355-363.

Boussafir M., Lallier-Verges E., Bertrand P., and Badaut-Trauth D.( 1995) SEM and TEM studies on isolated organic matter and rockmicrofacies from a short-term organic cycle of the KimmeridgeClay Formation (Yorkshire, G.B.). Lecture Notes in Eanh ScienceSeries 57, pp. 15-30. Springer-Verlag.

Bray E. E. and Evans E. D. (1961) Distribution of n-paraffins as aclue to recognition of source beds. Geochim. Cosmochim. Acta 22,2-15.

Derenne S., Largeau C., Casadevall E., Berkaloff c., and RousseauB. (1991a) Chemical evidence of kerogen formation in sourcerocks and oil shales via selective preservation of thin resistant outerwalls of microalgae: Origin of ultralaminae. Geochim. Cosmochim.Acta 55, 1041-1050.

Derenne S., Largeau c., and Casadevall E. (1991 b) Occurrence oftightly bound isoprenoid acids in an algal, resistant biomacromole­cule: possible geochemical implications. Org. Geochem. 17,597­602.

Derenne S. et al. (1992a) Similar morphological and chemical vari­ations of Gloeocapsomorpha prisca in Ordovician sediments andcultured Botryococcus braunii as a response to changes in salinity.Org. Geochem. 19,299-313.

Derenne S., Largeau c., Berkaloff C, Rousseau B., Wilhelm c., andHatcher P. G. ( I992b) Non-hydrolysable macromolecular constit­uents from outer walls of Chlorella fusca and Nanochlorum eu­caryotum. Phytochemistry 31, 1923-1929.

Derenne S., Largeau C., and Hatcher P. G. (1992c) Structure of Chlo­rella fusca algaenan: Relationships with ultralaminae in lacustrinekerogens; species- and environment-dependent variations in thecomposition of fossil ultralaminae. Org. Geochem. 18, 417 -422.

Derenne S., Le Berre F., Largeau C., Hatcher P., Connan J., andRaynaud 1. F. (1992d) Formation of ultralaminae in marine ker­ogens via selective preservation of thin resistant outer walls ofmicroalgae. Org. Geochem. 19, 345-350.

Derenne S., Largeau c., and Taulelle F. (1993) Elucidation, viasolid-state 15N NMR, of the origin of the alkylnitriles generatedupon pyrolysis of ultralaminae-containing kerogens. In OrganicGeochemistry (ed. K. 0ygard), pp. 766-770. Falch Hurtigtrykk.

Desprairies A., Bachaoui M., Ramdani A., and Tribovillard N. P.( 1995) Clay diagenesis in organic-rich cycles from the Kimmer­idge Clay Formation of Yorkshire (G. B.): implication for palaeo­climatic interpretations. In Lecture Notes in Eanh Sciences 57, pp.63-91. Springer-Verlag.

Eglinton T. I., Rowland S. J., Curtis C. D., and Douglas A. G. (1986)Kerogen-mineral reactions at raised temperatures in the presenceof water. Org. Geochem. 10, 1041-1052.

Eglinton T. I., Douglas A. G., and Rowland S. 1. (1988a) Release ofaliphatic, aromatic and sulphur compounds from Kimmeridge ker­ogen by hydrous pyrolysis: A quantitative study. Org. Geochem.13,655-663.

Eglinton T. I., Philp R. P., and Rowland S. J. (1988b) Flash pyrolysisof artificially maturated kerogens from the Kimmeridge Clay(U.K.). Org. Geochem. 12,33-41.

Eglinton T. I., Sinninghe-Damste J. S., Kohnen M. E. L., de LeeuwJ. W., Larter S. R., and Patience R. L. (1989) Analysis of maturity­related changes in the organic sulfur composition of kerogens byflash pyrolysis-gas chromatography. In Geochemistry of Sulfur inFossil Fuels (ed. W. L. Orr and C. M. White); ACS SymposiumSeries 429, pp. 529-565. Amer. Chern. Soc.

Eglinton T. I., Sinninghe-Damste 1. S., Kohnen M. E. L., and deLeeuw 1. W. (1990) Rapid estimation of the organic sulphur con­tent of kerogens, coals and asphaltenes by pyrolysis-gas chroma­tography. Fuel 69, 1394-1404.

Elsgaard L., Isaksen M. F., J0rgensen B. B., Alayse A-M., and Jan­nasch H. W. (1994) Microbial sulfate reduction in deep-sea sedi­ments at the Guaymas Basin hydrothermal vent area: Influence oftemperature and substrates. Geochim. Cosmochim. Acta 58, 3335­3343.

Espitalie J., Madec M., and Tissot B. (1980) Role of mineral matrixin kerogen pyrolysis: influence on petroleum generation and mi­gration. AAPG Bull. 64, 59-66.

Farrimond P. et al. (1984) Organic geochemical study of the UpperKimmeridge Clay of the Dorset type area. Mar. Petrol. Geol. 1,340-354.

Flaviano c., Le Berre F., Derenne S., Largeau c., and Connan J.( 1994) First indications of the formation of kerogen amorphousfractions by Selective Preservation. Role of non-hydrolysable mac­romolecular constituents of Eubacterial cell walls. Org. Geochem.22,759-771.

Fukushima K. and Ishiwatari R. ( 1988) Geochemical significance oflipids and lipid-derived substructures interlaced in kerogen. Org.Geochem. 12,509-518.

Gallois R. W. (1976) Coccolith blooms in the Kimmeridge Clay andorigin of the North Sea oil. Nature 259, 473-475.

Gelin F. et al. (1993) Mechanisms of flash pyrolysis of ether lipidsisolated from the green microalga Botryococcus braunii race A. J.Anal. Appl. Pyrolysis 27, 155-168.

Herbin J. P., Geyssant 1. R., Millier C., Melieres F., Ie groupe YOR­KIM, and Penn I. E. (1991) Heterogeneite quantitative et quali­tative de la matiere organique dans les argiles du Kimmeridgiendu val de Pickering (Yorkshire, UK). Cadre sedimentologique etstratigraphique. Rev. Inst. Fr. Petrol. 46, 1-39.

Herbin J. P., Millier c., Geyssant J. R., Melieres F., and Penn I. E.( 1993) Variation of the Distribution of organic matter within atransgressive system tract: Kimmeridge Clay (Jurassic), England.In AAPG 1993 "Petroleum source rocks in a sequence strati­graphic framework (ed. B. Katz and L. Pratt), pp. 67-100.

Huc A. Y., Lallier-Verges E., Bertrand P., Carpentier B., and Hol­lander D. J. (1992) Organic matter response to change of depo­sitional environment in Kimmeridgian shales, Dorset, U.K. In Or­ganic Matter: Productivity, Accumulation and Preservation ofOr­ganic Matter in Recent and Ancient Sediments (ed. J. Whelan andJ. Farrington), pp. 469-486. Columbia Univ. Press.

Provenance and preservation of kerogen 3747

Jackson G. A. ( 1990) A model of the formation of marine algal floesby physical coagulation processes. Deep-Sea Res. 37, 1197-1211.

Katz B. 1. (1983) Limitations of "Rock-Eval" pyrolysis for typingorganic matter. Org. Geochem. 4,195-199.

Kawamura K., Tannenbaum E., Huizinga B. 1., and Kaplan I. R( 1986) Long-chain carboxylic acids in pyrolysates of Green Riverkerogen. In Advances in Organic Geochemistry 1985 (ed. D. Ley­thaeuser and J. RullkOUer); Org. Geochem. 10, 1059-1065. Per­gamon.

Krein E. B. and Aizenshtat Z. (1994) The formation of isoprenoidsulfur compounds during diagenesis: simulated sulfur incorpora­tion and thermal transformation. Org. Geochem. 21, 1015-1025.

Lallier-Verges E., Boussafir M., Bertrand P., and Badaut-Trauth D.(1993a) Selective preservation of various organic matter types asassessed by STEM studies on a cyclic productivity-controlled sed­imentary series (Kimmeridge Clay Formation). In Organic Geo­chemistry: Poster sessions from the 16th International Meeting onOrganic Geochemistry, Stavanger, 1993 (ed. K. 0ygard), pp.384-386. Falch Hurtigtrykk.

Lallier-Verges E., Bertrand P., Biickel D., Hue A. Y., and TremblayP. (1993b) Control of the preservation of organic matter by pro­ductivity and sulphate reduction in Kimmeridgian shales fromDorset (UK). Mar. Petrol. Geol. 10,600-605.

Lallier-Verges E. et al. (1994) Productivity-induced sulfur enrich­ment of organic-rich sediments. 207th American Chemical Society1994, National Meeting-March 13-18, San Diego, CA, USA.

Largeau c., Derenne S., Casadevall E., Kadouri A., and Sellier N.( 1986) Pyrolysis of immature Torbanite and of the resistant bio­polymer (PRBA) isolated from extant alga Botryococcus braunii.Mechanism offormation and structure of Torbanite. In Advancesin Organic Geochemistry 1985 (ed. D. Leythaeuser and J. Rullk­otter); Org. Geochem. 10, 1023-1032. Pergamon.

Largeau C. et al. (l990a) Occurrence and origin of ultralaminarstructures in "amorphous" kerogens from various source-rocksand oil-shales. In Advances in Organic Geochemistry 1989 (ed. B.Durand and F. Behar), pp. 889-896. Pergamon.

Largeau C. et al. (I990b ) Characterization of various kerogens byScanning Electron Microscopy (SEM) and Transmission ElectronMicroscopy (TEM) - Morphological relationships with resistantouter walls in extant microorganisms. Meded. Rijks Geol. Dienst.45,91-101.

Le Berre F., Derenne S., Largeau C., Connan J., and Berkaloff C.(1991) Occurrence of non-hydrolysable, macromolecular, wallconstituents in bacteria. Geochemical implications. In OrganicGeochemistry Advances and Applications in Energy and the Nat­ural Environment (ed. D. A. C. Manning), pp. 428-431. Man­chester Univ. Press.

Lugardon B., Raynaud 1. F., and Husson P. (1991) Donnees ultra­structurales sur la matiere organique amorphe des kerogenes. Pa­lynoscience 1, 69-88.

Nip M., de Leeuw J. W., and Schenck P. A. (1987) Structural char­acterization of coals, coal macerals and their precursors by pyrol­ysis-gas chromatography and pyrolysis-gas chromatography-massspectrometry. Coal Sci. Technol. 11,89-92.

Orr W. L. (1986) Kerogen I asphaltene I sulfur relationships in sulfur­rich Monterey oils. Org. Geochem. 10,499-516.

Pfendt P. A. (1984) Comparison of the general chemical nature ofvarious kerogens based on their reactivities towards bromine. Org.Geochem. 6, 383-390.

Pradier B. and Bertrand P. ( 1992) Etude ahaute resolution d'un cyclede carbone organique des argiles du Kimmeridgien du Yorkshire(GB): relation entre composition petrographique du contenu or­ganique observe in-situ, teneur en carbone organique et qualitepetroligene. C.R. Acad. Sc. Paris 315, 187-192.

Ramanampisoa L. and Disnar J. R ( 1994) Primary control of paleo­production on organic matter preservation and accumulation in theKimmeridge rocks of Yorshire (UK). Org. Geochem. 21, 1153­1167.

Ramanampisoa L., Bertrand P., Disnar J. R, Lallier-Verges E., Pra­dier B., and Tribovillard N. P. (1992) Etude a haute resolutiond'un cycle de carbone organique des argiles du Kimmeridgien duYorkshire (GB ): resultats preliminaries de geochimie et de petro­graphie organique. C. R. Acad. Sc. Paris 314, 1493-1498.

Raynaud J. F., Lugardon B., and Lacrampe-Couloume G. (1988)Observation de membranes fossiles dans la matiere organique"amorphe" de roches-meres de petrole. c. R. Acad. Sci. Paris 307,1703-1709.

Schouten S., van Driel G. B., Sinninghe-Damste 1. S., and de LeeuwJ. W. (1994) Natural sulphurization of ketones and aldehydes: Akey reaction in the formation of organic sulphur compounds. Geo­chim. Cosmochim. Acta 57, 5111-5116.

Sinninghe-Damste 1. S. and de Leeuw J. W. (1990) Analysis, struc­ture and geochemical significance of organically-bound sulphur inthe geosphere: State of the art and future research. In Advances inOrganic Geochemistry 1989 (ed. B. Durand and F. Behar); Org.Geochem. 16, 1077-1101. Pergamon.

Sinninghe-Damste 1. S., Eglinton T. I., de Leeuw J. W., and SchenckP. A. ( 1989) Organic sulphur in macromolecular sedimentary or­ganic matter: I. Structure and origin of sulphur-containing moietiesin kerogens, asphaltenes and coals as revealed by flash pyrolysis.Geochim. Cosmochim. Acta 53, 873-889.

Smith P. M. R. (1984) The use of fluorescence microscopy in thecharacterisation of amorphous organic matter. Org. Geochem. 6,839-845.

Tribovillard N. P., Desprairies A., Bertrand P., Lallier-Verges E.,Disnard J-R, and Pradier B. (1992) Etude ahaute resolution d'uncycle du carbone organique de roches kimmeridgiennes du York­shire (Grande-Bretagne): mineralogie et geochimie (resultats pre­liminaires). C. R. Acad. Sci. Paris 314 (II), 923-930.

Tribovillard N. P. et al. (1994) Geochemical study of organic matterrich cycles from the Kimmeridge Clay Formation of Yorkshire(UK): productivity versus anoxia. Palaeogeogr. Palaeoclim. Pa­laeoecol. 108, 165-181.

Trudinger P. A. ( 1992) Bacteria sulfate reduction: Current status andpossible origin. In Early Organic Evolution: Implications for Min­eral and Energy Resources (ed. M. Schidlowski et al.), pp. 367­377. Springer-Verlag.

Tyson R V., Wilson R C. L., and Downie C. (1979) A stratifiedwater column environmental model for the type Kimmeridge Clay.Nature 277,377-380.

Wefer G. (1989) Particles flux in the Oceans: effects of episodicproduction. In Report of the Dahlem Workshop on Productivity ofthe Oceans: Present and Past (ed. W. H. Berger, V. S. Smetacek,and G. Wefer), pp. 139-154. Wiley.

Widdel F. (1988) Microbiology and ecology of sulfate- and sulfur­reducing bacteria. In Biology of Anaerobic Organisms (ed. J. B.Zehnder), pp. 469-585. Wiley.

Williams P. F. V. (1986) Petroleum Geochemistry of the Kimmer­idge Clay of onshore southern and eastern England. Mar. Petrol.Geol. 3,258-281.

Williams P. F. V. and Douglas A. G. (1980) A preliminary organicgeochemical investigation of the Kimmeridgian oil shales. In Ad­vances in Organic Geochemistry 1979 (ed. A. G. Douglas and1. R. Maxwell), pp. 531-545. Pergamon.

Williams P. F. V. and Douglas A. G. (1983) The effect of lithologicvariation on organic geochemistry in the Kimmeridge Clay, Brit­ain. In Advances in Organic Geochemistry 1981 (ed. M. Bjof0Yet al.), pp. 568-575. Pergamon.


Recommended