12.07.2015 Views

Nutman et al.2010.pdf - of / [www.ene.ttu.ee]

Nutman et al.2010.pdf - of / [www.ene.ttu.ee]

Nutman et al.2010.pdf - of / [www.ene.ttu.ee]

SHOW MORE
SHOW LESS
  • No tags were found...

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Precambrian Research 183 (2010) 725–737Contents lists available at ScienceDirectPrecambrian Researchjournal homepage: <strong>www</strong>.elsevier.com/locate/precamres≥3700 Ma pre-m<strong>et</strong>amorphic dolomite formed by microbial mediation in the Isuasupracrustal belt (W. Gr<strong>ee</strong>nland): Simple evidence for early life?Allen P. <strong>Nutman</strong> a,b,∗ , Clark R.L. Friend c , Vickie C. Benn<strong>et</strong>t d , David Wright e , Marc D. Norman da School <strong>of</strong> Earth and Environmental Sciences, University <strong>of</strong> Wollongong, Wollongong, NSW 2522, Australiab Beijing SHRIMP Centre, Institute <strong>of</strong> Geology, Chinese Academy <strong>of</strong> Geological Sciences, 26 Baiwanzhuang Road, Beijing 100037, Chinac 45 Stanway Road, Headington, Oxford OX3 8HU, UKd Research School <strong>of</strong> Earth Sciences, Australian National University, Canberra ACT 0200, Australiae Department <strong>of</strong> Geology, University <strong>of</strong> Leicester, University Road, Leicester LE1 7RH, UKarticleinfoabstractArticle history:Received 19 August 2009Received in revised form 22 July 2010Accepted 13 August 2010Keywords:Isua supracrustal beltEarly lifeDolomiteREE + Y chemistryChemical sedimentary rocksPillow lava intersticesChemical (m<strong>et</strong>a)sedimentary rocks in the amphibolite facies ≥3700 Ma Isua supracrustal belt (W. Gr<strong>ee</strong>nland)are mostly strongly deformed, so there is only a small chance <strong>of</strong> the survival <strong>of</strong> features such asstromatolites or micr<strong>of</strong>ossils that would be direct pro<strong>of</strong> <strong>of</strong> a ≥3700 Ma biosphere. Therefore the searchfor evidence <strong>of</strong> ≥3700 Ma life in Isua rocks has focused on chemical signatures, particularly C-isotopes.The new approach presented here is based on whole rock chemistry rather than isotopic signatures. Isuachemical sedimentary rocks have Ca–Mg–Fe bulk compositions that coincide with ferroan dolomite –siderite/Fe-oxide mixtures. Most have low Al 2 O 3 , TiO 2 contents (


726 A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737Fig. 1. (a) Map <strong>of</strong> the Isua supracrustal belt (southern West Gr<strong>ee</strong>nland) showing location <strong>of</strong> discussed samples. (b) D<strong>et</strong>ailed map <strong>of</strong> a small part <strong>of</strong> the eastern part <strong>of</strong> theIsua supracrustal belt, showing the geological relationships b<strong>et</strong>w<strong>ee</strong>n altered intermediate (basaltic-andesite?) pillowed volcanic rocks, chemical sedimentary rocks and anoverlying thrust she<strong>et</strong> <strong>of</strong> boninitic amphibolites. All rocks and the thrust were folded prior to intrusion <strong>of</strong> Ameralik dykes at ca. 3510 Ma.improbable (Appel <strong>et</strong> al., 2003), and the search for early life has concentratedon how to interpr<strong>et</strong> 13 C-depl<strong>et</strong>ed signatures <strong>of</strong> graphitefrom these rocks. Some workers have concluded that these signaturesare evidence that life was already established by 3700 Ma(Schidlowski <strong>et</strong> al., 1979; Mozjsis <strong>et</strong> al., 1996; Rosing, 1999). Unresolvedissues over the interpr<strong>et</strong>ation <strong>of</strong> these signatures includewh<strong>et</strong>her the protoliths <strong>of</strong> the studied samples are really sedimentaryand wh<strong>et</strong>her the graphite formed by inorganic decarbonationreactions instead <strong>of</strong> reduction <strong>of</strong> biogenic precursors (Perry andAhmed, 1977; van Zuilen <strong>et</strong> al., 2002). From one unit <strong>of</strong> Isuaquartzo-feldspathic rocks <strong>of</strong> sedimentary origin, Rosing (1999) documentedfine-grained 13 C-depl<strong>et</strong>ed graphite which he interpr<strong>et</strong>edto be derived from microbial carbon. However, Schoenburg <strong>et</strong> al.(2002) raised the possibility that this graphite could be derivedfrom carbonaceous chondrites, particularly as the same rock unitmight contain a tungsten isotopic anomaly <strong>of</strong> non-terrestrial origin.Thus the C-isotope evidence for Eoarchean life is tantalizing,but not definitive. From another perspective, microbial mediationmay have b<strong>ee</strong>n necessary for deposition <strong>of</strong> banded iron formations(BIF) (Cloud, 1973; Garrels <strong>et</strong> al., 1973) such as those at Isua, perhapsproviding evidence for early life.Based on p<strong>et</strong>rographic, mineralogical and whole rock geochemicalevidence for the existence <strong>of</strong> low-temperature (prem<strong>et</strong>amorphism)dolomite in Isua sedimentary rocks and pillowinterstices, this paper proposes a simple new line <strong>of</strong> evidencefor life: that this earliest dolomite formed by ≥3700 Ma bacterialmediation. This is based on the observation that low-temperaturedolomite formation in modern sedimentary and volcanic rocks hasonly b<strong>ee</strong>n replicated in the laboratory through microbial mediation(Vasconcelos <strong>et</strong> al., 1995; Roberts <strong>et</strong> al., 2004; Wright and Wacey,2005).2. S<strong>et</strong>ting <strong>of</strong> Eoarchaean chemical (m<strong>et</strong>a)sedimentaryrocks in Gr<strong>ee</strong>nland2.1. Itsaq Gneiss Complex regional geologyThe ca. 3000 km 2 Itsaq Gneiss Complex in the Nuuk region <strong>of</strong>W. Gr<strong>ee</strong>nland is dominated by tonalitic rocks, with inclusions <strong>of</strong>amphibolites <strong>of</strong> mostly island arc tholeiite and boninitic chemicalaffinity, showing that it formed from the products <strong>of</strong> magmatic arcsat convergent plate boundaries (<strong>Nutman</strong> <strong>et</strong> al., 2007, 2009; Dilekand Polat, 2008). These rocks were affected by ductile deformationunder amphibolite to granulite facies conditions in the Eoarchaean(3650–3550 Ma) and again under amphibolite facies conditionsin the Neoarchaean (<strong>Nutman</strong> <strong>et</strong> al., 1996). This means that onlya few fortuitous areas escaped most Neoarchaean deformation,allowing rare insights into Eoarchaean geological processes. Thelargest <strong>of</strong> these low Neoarchaean strain areas is in the north <strong>of</strong>the complex around the ≥3690 Ma Isua supracrustal belt (Fig. 1a;e.g. Moorbath <strong>et</strong> al., 1973; Allaart, 1976; <strong>Nutman</strong> <strong>et</strong> al., 1996,2009, 2002; Rosing <strong>et</strong> al., 1996; <strong>Nutman</strong> and Friend, 2009). However,the Isua supracrustal belt contains an Eoarchaean suturezone b<strong>et</strong>w<strong>ee</strong>n a ∼3800 Ma terrane to the south and a ∼3700 Materrane to the north (Fig. 1a; <strong>Nutman</strong> <strong>et</strong> al., 1997, 2009, 2002;Crowley, 2003). Thus its volcanic and sedimentary rocks are nowmostly strongly foliated and/or lineated amphibolite facies tectonites(<strong>Nutman</strong> <strong>et</strong> al., 1996; Myers, 2001), despite having escapedmuch <strong>of</strong> the superimposed Neoarchaean deformation. Thus, onlyrarely are sedimentary layering and volcanic structures preservedin the belt (albeit always still deformed under amphibolite faciesm<strong>et</strong>amorphism; Fig. 2a). It is the Isua rocks that are the focus <strong>of</strong> thispaper.


A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737 727Isua supracrustal belt carbonates has recently b<strong>ee</strong>n confirmed bytheir seawater-like REE + Y (SN) (rare earth element and yttriumpost-Archaean-average-shale-normalized (SN)) trace element patterns(Bolhar <strong>et</strong> al., 2004; Friend <strong>et</strong> al., 2007). Preservation in somecarbonate rocks <strong>of</strong> seawater-like REE + Y (SN) patterns strongly suggeststhat their chemistry has little or no modification throughhigh temperature m<strong>et</strong>asomatic processes, as can be observed insome marbles (e.g. Hecht <strong>et</strong> al., 1999). This does not preclud<strong>et</strong>he possibility that other carbonates entirely <strong>of</strong> high temperature(600–500 ◦ C) 500 ◦ C) m<strong>et</strong>asomatic origin can also be present (suchas within ultramafic rocks as proposed by Rose <strong>et</strong> al., 1996; s<strong>ee</strong>Fig. 2b).2.3. Isua supracrustal belt carbonate-bearing pillow lavaintersticesFig. 2. (a) Low strain lithon (right) in g<strong>ene</strong>rally strongly deformed BIF (left). Sedimentarylayering is still preserved in the lithon, although the rock has b<strong>ee</strong>n deformedand recrystallised under amphibolite facies conditions. Tip <strong>of</strong> pen for scale in topright <strong>of</strong> image. (b) Early boundinaged carbonate vein (v) cutting Isua m<strong>et</strong>asomatisedultramafic schists (um) dominated by orthoamphibole. This demonstrates carbonateveining, as s<strong>ee</strong>n in orogenic belts <strong>of</strong> all ages. Pen for scale in the upper middlepart <strong>of</strong> the image.2.2. Isua supracrustal belt chemical sedimentary rocks – BIFs andcarbonatesIsua magn<strong>et</strong>ite-banded and quartz-rich rocks have b<strong>ee</strong>n widelyaccepted as having chemical sedimentary protoliths <strong>of</strong> BIF andchert respectively (Moorbath <strong>et</strong> al., 1973; Allaart, 1976; Dymekand Klien, 1988; Bolhar <strong>et</strong> al., 2004; Friend <strong>et</strong> al., 2007). However,the origin <strong>of</strong> the carbonate-bearing rocks that are widespreadin some Isua lithological units has b<strong>ee</strong>n debated. Some workersregard at least some <strong>of</strong> these rocks as sedimentary in origin (Allaart,1976; <strong>Nutman</strong> <strong>et</strong> al., 1984; Dymek and Klien, 1988) whereas othershave proposed that the carbonates are entirely m<strong>et</strong>asomatic(Rose <strong>et</strong> al., 1996; Rosing <strong>et</strong> al., 1996). The former interpr<strong>et</strong>ationacknowledges the presence <strong>of</strong> secondary carbonate formation ormobilisation during post-depositional m<strong>et</strong>asomatism. Clear evidence<strong>of</strong> this is carbonate veins that cut Isua supracrustal beltrocks (Fig. 2b). Such carbonate mobility is common in highlydeformed and m<strong>et</strong>amorphosed rocks <strong>of</strong> all ages. However, an originallysedimentary or diagen<strong>et</strong>ic source for at least some <strong>of</strong> theA focus <strong>of</strong> this paper is a locality <strong>of</strong> low strain with preservedpillow structures in m<strong>et</strong>avolcanic rocks first described bySolvang (1999) tog<strong>et</strong>her with adjacent more deformed carbonaterichchemical m<strong>et</strong>asedimentary rocks (Figs. 1 and 3b). The pillowedvolcanic rocks are basic-intermediate in composition (e.g. analysisG05/22, Table 1). The shapes <strong>of</strong> the best-preserved pillows (Fig. 3a)show that the volcanic rocks face towards the more deformedchert + calc-silicate + carbonate and BIF rocks that wrap aroundtheir northwestern side (Fig. 3b). The pillows become increasinglybrecciated and traversed by carbonate-bearing veins as the contactis approached, but there is no definitive indication that the contactb<strong>et</strong>w<strong>ee</strong>n them is tectonic. Therefore we interpr<strong>et</strong> these chemicalsedimentary rocks to have b<strong>ee</strong>n deposited on top <strong>of</strong> the volcanicrocks, and subsequently h<strong>et</strong>erog<strong>ene</strong>ously deformed tog<strong>et</strong>her(Friend <strong>et</strong> al., 2007). Evidence supporting this is that the m<strong>et</strong>asedimentaryrocks contain rare d<strong>et</strong>rital/volcanic zircons with ages <strong>of</strong>only 3690 Ma (<strong>Nutman</strong> <strong>et</strong> al., 2009; repository Table 1), whereaspillow lava sample G05/22 (GPS 65 ◦ 10.765 ′ N, 49 ◦ 48.211 ′ W –WGS84 datum) from the underlying basic to intermediate m<strong>et</strong>avolcanicrocks yielded a single, prismatic, igneous zircon dated at3709 ± 9Ma(repository Table 1). Therefore the volcanic rocks mustbe marginally older than the overlying m<strong>et</strong>asedimentary rocks(Fig. 4). These volcanic rocks and capping chemical sedimentaryrocks are preserved in an isoclinal fold nose, and are separatedfrom structurally overlying amphibolites <strong>of</strong> boninitic chemicalaffinity (Polat <strong>et</strong> al., 2002) by an early isoclinally folded mylonite(Fig. 1b). The interior <strong>of</strong> the pillow interstices are siliceous and havecarbonate-rich rinds (regr<strong>et</strong>tably not sampled). The interior <strong>of</strong> thepillows commonly show carbonate alteration, and intact vesiclesare filled by quartz or calcite.3. Chemistry <strong>of</strong> Isua supracrustal belt chemical(m<strong>et</strong>a)sedimentary rocks and a pillow interstice3.1. Filtering <strong>of</strong> dataChemical sedimentary rocks can become polluted by terriginousadditions, namely windblown dust, subaqueous d<strong>et</strong>rital input orvolcanic ash (Bolhar <strong>et</strong> al., 2004 and references therein). The REEabundance <strong>of</strong> these pollutants is much higher than in chemical sediments(e.g. Bolhar <strong>et</strong> al., 2004). Therefore only small terriginousadditions will mask the low abundance REE signature related t<strong>of</strong>ormation <strong>of</strong> chemical sedimentary rocks. In the whole rock chemistrythis pollution is readily d<strong>et</strong>ected by increases in Al 2 O 3 andTiO 2 . Hence here we have applied an arbitrary filter <strong>of</strong>


Table 1Whole rock geochemistry.Sample no JHA170728 IF-G# G05/17 G07/22 G07/08 G93/28* G04/54* G04/55* G04/85* G91/75* G05/22 G05/23 G91/25* G91/26R*Sampl<strong>et</strong>ype andprotolithdolomiterock m<strong>et</strong>adolostonechemicalsedimentBIFchemicalsedimentdolomitic-BIFchemicalsedimentdolomitic-BIFchemicalsedimentdolomitic-BIFchemicalsedimentdol. calc-sil.chemicalsedimentdol. calc-sil.chemicalsedimentdol. calc-sil.chemicalsedimentdolomitic-BIFchemicalsedimentchertalteredpillow lavaandesite?pillowintersticechemicalsedimentdolomitic-BIFchemicalsedimentdolomitic-BIFUnit and Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Isua amph-f Akilia gran-f Akilia gran-fgradeLatitude 65 ◦ 05.4 ′ N 65 ◦ 09.869 ′ N 65 ◦ 08.059 ′ N 65 ◦ 08.432 ′ N 65 ◦ 05.62 ′ N 65 ◦ 10.767 ′ N 65 ◦ 10.767 ′ N 65 ◦ 10.277 ′ N 65 ◦ 12.14 ′ N 65 ◦ 10.765 ′ N 65 ◦ 10.765 ′ N 63 ◦ 55.73 ′ N 63 ◦ 55.73 ′ NLongitude 50 ◦ 06.5 ′ W 49 ◦ 48.936 ′ W 50 ◦ 11.464 ′ W 50 ◦ 09.302 ′ W 50 ◦ 09.73 ′ W 49 ◦ 48.227 ′ W 49 ◦ 48.227 ′ W 49 ◦ 48.968 ′ W 49 ◦ 47.40 ′ W 49 ◦ 48.211 ′ W 49 ◦ 48.211 ′ W 51 ◦ 41.08 ′ W 51 ◦ 41.08 ′ WSiO 2 (wt%) 3.19 41.20 66.26 71.28 89.73 71.86 54.04 86.60 53.78 97.07 63.04 85.37 82.97 66.80TiO 2 0.02 0.01 0.01 0.01


A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737 729Fig. 4. 238 U/ 206 Pb versus 207 Pb/ 206 Pb concordia diagram for rare small zircons fromchemical m<strong>et</strong>asedimentary rocks G04/54, G04/85 and IEh50ch/a (<strong>Nutman</strong> <strong>et</strong> al.,2009), plus thr<strong>ee</strong> age d<strong>et</strong>erminations on a single prismatic zircon recovered fromvolcanic pillow G05/22. Errors are depicted at the 2 level. S<strong>ee</strong> online repositoryTable 1 for U–Pb analytical data <strong>of</strong> sample G05/22.3.2. REE + Y signaturesFig. 3. (a) Weakly deformed pillow lavas (p) with quartz + carbonate + tremoliteactinoliteinterstices (int) at GPS 65 ◦ 10.765 ′ N, 49 ◦ 48.211 ′ W. Note stronglyweathered carbonate-filled margins and cracks in the pillows. Pen for scale onleft hand side. (b) Layered quartz + carbonate + tremolite-actinolite rocks. SampleG04/54 and G05/55 came from where the person is standing. These rocks displayseawater-like REE + Y (SN) trace element signatures (Friend <strong>et</strong> al., 2007), and thereforeare derived from chemical sedimentary rocks. However, due to high strainin them, the alternation b<strong>et</strong>w<strong>ee</strong>n quartz and carbonate-rich layers should not beregarded as well-preserved sedimentary layering. (c) Transmitted light photomicrograph<strong>of</strong> pillow interstice sample G05/23. Qtz, quartz; tr, tremolite-actinolite;cc, calcite; cp, chalcopyrite. Field <strong>of</strong> view ∼2.5 mm across. Phases were identified byelectron microprobe <strong>ene</strong>rgy dispersive analyses using the JEOL JXA-8800R instrumentat the Chinese Academy <strong>of</strong> Geological Sciences. Cc + tr would have formed byprogressive reaction b<strong>et</strong>w<strong>ee</strong>n qtz + dol in the pre-m<strong>et</strong>amorphic rock. In this casesilica was present in excess, leaving a dolomite-fr<strong>ee</strong> assemblage.Dolomite-rich rock JHA170728 (Fig. 5a, Table rock JHA170728(Fig. 5a, Table 1) shows an identical PAAS – normalised (REE + Y (SN) )pattern (and in this case elemental abundances) to IF-G, the geochemicalstandard BIF from Isua (Bolhar <strong>et</strong> al., 2004). This clearlyshows that the common origin <strong>of</strong> such Isua carbonate rocks andsilica – Fe rocks <strong>of</strong> accepted sedimentary origin and regarded asa seawater proxy. Regardless <strong>of</strong> silica content and proportions <strong>of</strong>whole rock Fe to Ca–Mg components, all the Isua samples with lowAl 2 O 3 and TiO 2 show REE + Y (SN) seawater-like patterns with clearpositive Y, Eu, La anomalies and g<strong>ene</strong>rally discernable Ce anomalies(Fig. 5b and data in Bolhar <strong>et</strong> al., 2004; Friend <strong>et</strong> al., 2007; s<strong>ee</strong> thelatter for the analytical m<strong>et</strong>hod <strong>of</strong> the new data presented here).This REE + Y (SN) pattern is produced by precipitation directlyfrom seawater, and also from low-temperature groundwaters (e.g.Bolhar <strong>et</strong> al., 2004; Johannesson <strong>et</strong> al., 2006 and references therein).On the other hand, hydrothermal fluids and carbonate–silica rocksdeposited from such fluids have different REE + Y (SN) signatures(e.g. Bolhar <strong>et</strong> al., 2005 and references therein). This is demonstratedhere by Precambrian hydrothermal siderite such as sample177886 from the Pilbara (van Kranendonk <strong>et</strong> al., 2003; analysis4inFig. 5c), plus m<strong>et</strong>asomatic dolomites from the Göpfersgrüntalc deposit (Hecht <strong>et</strong> al., 1999) and modern oceanic hydrothermalbrines (Douville <strong>et</strong> al., 1999; Fig. 6). In the Göpfersgründeposit Hecht <strong>et</strong> al. (1999) analysed progressive g<strong>ene</strong>rations <strong>of</strong>dolomites, from regional dolomitic marble to early hydrothermalvein dolomite and later hydrothermal vein dolomite (analyses 1–3respectively in Fig. 5c). The regional dolomitic marble still displaysa weak Y anomaly and a downward-bowed PAAS-normalisedsignature for the light rare earth elements. In the successiveg<strong>ene</strong>rations <strong>of</strong> hydrothermal dolomite, the positive Y anomalyand downward-bowed PAAS-normalised signature for the lightrare earth elements disappear (Fig. 5c). The Pilbara hydrothermalsiderite <strong>of</strong> van Kranendonk <strong>et</strong> al. (2003) also displays neither a positiveY anomaly nor a seawater-like PAAS-normalised signature forthe light rare earth elements (analysis 4 in Fig. 5c). Thus an impor-


730 A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737Fig. 6. REE=Y (SN) plot for Isua pillow lava G05/22, pillow interstice G05/23 and Isuacarbonate-rich sedimentary rocks G04/54 (Friend <strong>et</strong> al., 2007), Pilbara Warrawoonadolomite-rich pillow interstice is from Yamamoto <strong>et</strong> al. (2004), Mid-Atlantic Ridgehydrothermal fluid is from Bau and Dulski (1999). Also shown for reference arem<strong>et</strong>asomatic dolomites <strong>of</strong> Hecht <strong>et</strong> al. (1999).Fig. 5. PAAS (Post-Archean-Australian-Shale; McClennan, 1989) normalized REE + Ypatterns (REE = Y (SN) ). (a) IF-G Isua BIF (Bolhar <strong>et</strong> al., 2004) and m<strong>et</strong>adolomiteJHA170728 (this paper), modern Mg-calcite microbialite from Heron Island,Great Barrier R<strong>ee</strong>f (Webb and Kamber, 2000) and a Warrawoona ∼3.45 Ga lowtemperaturedolomitized stromatolite (van Kranendonk <strong>et</strong> al., 2003). (b) Isuadolomitic calc-silicate rocks and BIF with a dolomitic component. (c) M<strong>et</strong>asomaticdolomites and host dolomitic marble (Hecht <strong>et</strong> al., 1999) and Pilbara ∼3.45 Gahydrothermal siderite deposited from hydrothermal waters (van Kranendonk <strong>et</strong> al.,2003).tant feature is that hydrothermal fluids and carbonates precipitatedfrom them show no Y anomaly (or it is extremely muted – Fig. 5c).Conversely this is a consistent feature <strong>of</strong> seawater, some groundwatersand the sedimentary proxies deposited from them. Analyticallythe Y anomaly is particularly useful, given the higher abundance<strong>of</strong> Y and the neighboring reference Dy and Ho, compared to theLREE that in seawater proxies occur at an order <strong>of</strong> magnitude lowerabundances.Thus for the Isua samples we consider here, their REE + Y (SN)signature indicates pre-m<strong>et</strong>amorphic low-temperature depositionrather than from hydrothermal fluids. Deposition could haveoccurred in a range <strong>of</strong> surficial (directly from seawater) to nearsurfaces<strong>et</strong>tings (from groundwater during diag<strong>ene</strong>sis). Thereforewe contend from the REE + Y data that the Isua supracrustal beltcontains a diverse suite <strong>of</strong> chemical sediments whose chemistryis controlled by deposition from seawater or during near-surfacediagen<strong>et</strong>ic processes. Their composition ranged from almost puredolomite (JHA170728), to clean chert (G91/75, Friend <strong>et</strong> al., 2007)to magn<strong>et</strong>ite-rich BIF (IF-G, Bolhar <strong>et</strong> al., 2004 and G04/85, Friend<strong>et</strong> al., 2007), and are united in their origin by showing identicalREE+Y (SN) patterns (Fig. 5a and b).3.3. Major element variation <strong>of</strong> Isua chemical sedimentary rocksThe low Al 2 O 3 and TiO 2 samples including data from the literatureand this study (Table 1), show a range in SiO 2 content from 3%to 97% (e.g. the dolomite-rich sample JHA170728 presented here,and m<strong>et</strong>achert G91/75 in Friend <strong>et</strong> al., 2007), and when cast intomolecular proportions, mostly form a linear array from Fe to 0.5Ca,0.5Mg in a Fe–Mg–Ca ternary plot (Fig. 7). Even within the samelithological unit (such as the locality discussed in most d<strong>et</strong>ail inthe east <strong>of</strong> the belt where weathered pillow basalts are capped bychemical sedimentary rocks – Fig. 1b), there is a spread in bulkcompositions from close to ferroan dolomite (sample G04/54 – GPS


A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737 731Fig. 7. Molar Ca, Mg, Fe proportions <strong>of</strong> Isua chemical sedimentary rocks and pillowinterstice G05/23 with low Al 2O 3 and TiO 2. Analyses <strong>of</strong> samples are in Table 1, Dymekand Klien (1988) and Bolhar <strong>et</strong> al. (2004). Pilbara interpillow dolomite-rich sampleis from Yamamoto <strong>et</strong> al. (2004).65 ◦ 10.760 ′ N, 49 ◦ 48.153 ′ W and sample IS625-13B <strong>of</strong> Dymek andKlien, 1988; Table 1), to other more iron rich compositions (sampleG04/55), to BIF (e.g. sample G04/85 – 65 ◦ 10.310 ′ N, 49 ◦ 49.224 ′ W;Fig. 1b; Table 1).This range <strong>of</strong> Ca–Mg–Fe compositions is found in other units<strong>of</strong> definite and proposed chemical sedimentary rocks throughoutthe Isua supracrustal belt (Figs. 1 and 7a, Table 1). The ends <strong>of</strong> thisCa,Mg–Fe array are represented by the sample JHA170728 <strong>of</strong> massivedolomite rock interlayered with m<strong>et</strong>abasaltic amphibolitesand the geochemical standard BIF sample IF-G (Fig. 7). We note thatalthough JHA170728 contains dolomite, this will have b<strong>ee</strong>n compl<strong>et</strong>elyrecrystallised during superimposed m<strong>et</strong>amorphism. For thepurpose <strong>of</strong> description, we have divided this compositional arrayinto four groups <strong>of</strong> rocks; dolomites (JHA170728); dolomitic calcsilicaterocks (e.g. G04/54 and -55), BIF with a dolomitic component(e.g. G07/22, G93/28) and BIF (e.g. IF-G and G04/85; Fig. 7). TheREE+Y (SN) pattern <strong>of</strong> all <strong>of</strong> these groups are seawater-like (Fig. 5,Bolhar <strong>et</strong> al., 2004; Friend <strong>et</strong> al., 2007), that indicates this Ca,Mg–Fearray is a low-temperature, pre-m<strong>et</strong>amorphic, non-m<strong>et</strong>asomaticfeature.We interpr<strong>et</strong> these compositions to have originally containedboth iron rich phases (iron hydroxides and/or siderite) + dolomite,related either to a diagen<strong>et</strong>ic phenomenon or to high strain telescoping<strong>of</strong> chemical sedimentary rock layers <strong>of</strong> different Fe/Mgratio into the scale <strong>of</strong> a single hand specimen. Similar compositionalranges are found in younger, b<strong>et</strong>ter-preserved (lowerm<strong>et</strong>amorphic grade and undeformed) Precambrian chemical sedimentarysequences such as the Transvaal Supergroup <strong>of</strong> southernAfrica, where there is both sedimentary interlayering <strong>of</strong> BIF, chertsand carbonate rocks, and diagen<strong>et</strong>ic dolomite growth (Beukes,1987).3.4. Chert + dolomite protolith <strong>of</strong> an Isua pillow intersticeSample G05/23 (GPS 65 ◦ 10.765 ′ N, 49 ◦ 48.211 ′ W) represents theinterior <strong>of</strong> a pillow interstice. It is devoid <strong>of</strong> a foliation, andin decreasing modal abundance consists <strong>of</strong> quartz + tremoliteactinolite+ calcite with accessory chalcopyrite in textural equilibrium(Fig. 3c). It was taken from m<strong>et</strong>abasaltic rocks underlyingthe chemical sedimentary unit represented by samples G04/54 andG04/55. It has low Al 2 O 3 (0.28 wt%) and TiO 2 (


732 A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737perature m<strong>et</strong>amorphism, m<strong>et</strong>asomatism and ductile deformation.It demonstrates that dolomite precipitation was occurring in differentlow-temperature s<strong>et</strong>tings, whose modern analogs involve theaction <strong>of</strong> anaerobic bacteria. There is also strong independent evidencefor life in the Warrawoona sequences including preservedmicr<strong>of</strong>ossils and undisturbed stable isotope signatures (reviewedby Schopf, 2006).4.2. Significance <strong>of</strong> REE + Y (SN) trace element patterns in IsuadolomitesThe seawater-like REE + Y pattern recurs in chemical sedimentaryrocks throughout the geological record from Isua in theEoarchaean (this paper and Bolhar <strong>et</strong> al., 2004; Friend <strong>et</strong> al., 2007)up to the Holoc<strong>ene</strong>, where it is present in low-temperature s<strong>et</strong>tingssuch as microbialite carbonate cements (Webb and Kamber,2000) or groundwaters (Johannesson <strong>et</strong> al., 2006). Therefore, wecontend that the Isua supracrustal belt contains a diverse suite <strong>of</strong>chemical sediments whose trace element chemistry and the presence<strong>of</strong> dolomite in the protoliths was controlled by seawater ornear-surface diagen<strong>et</strong>ic processes.4.3. Low-temperature dolomite formation mechanisms4.3.1. BackgroundDolomite, broadly defined here to as including ferroan vari<strong>et</strong>ieswith some substitution <strong>of</strong> Mg 2+ by Fe 2+ , is a commonlow-temperature sedimentary and diagen<strong>et</strong>ic mineral in carbonatesediments from sabkha to d<strong>ee</strong>p ocean environments (s<strong>ee</strong> Wrightand Wacey, 2005 and references therein) and occurs in intersticeswithin basalts (Burns <strong>et</strong> al., 2000; Yamamoto <strong>et</strong> al., 2004). Despiteover two centuries <strong>of</strong> intense research, numerous low-temperaturephysico-chemical experiments, including one running for >30 yearsat 1000 times oversaturation, have failed to precipitate dolomite(Machel and Mountjoy, 1986; Hardie, 1987; Wright, 1997; Purser<strong>et</strong> al., 1994; Land, 1998; Wright, 2000; Wright and Wacey, 2004).This is all the more surprising given that thermodynamic considerationsdictate that in the marine environment, not onlyshould dolomite precipitate spontaneously, but any solid calciumcarbonate should also be immediately dolomitized (Lippmann,1973). However, many observed chemical trends in saline solutionsdo not follow thermodynamic predictions (e.g. Kelleher andRedfern, 2002). Although seawater is supersaturated with respectto dolomite, no spontaneous precipitation <strong>of</strong> the mineral has everb<strong>ee</strong>n observed, and it is extremely doubtful that dolomite couldhave spontaneously precipitated from normal seawater in the past(e.g. Wright, 2000; Wright and Altermann, 2000). Precipitation <strong>of</strong>dolomite from a fluid or by replacement <strong>of</strong> a CaCO 3 precursor hasonly b<strong>ee</strong>n produced experimentally at or near hydrothermal conditions(e.g. Gaines, 1980; Lumsden <strong>et</strong> al., 1995; Tribble <strong>et</strong> al., 1995;Land, 1998), y<strong>et</strong> vast amounts <strong>of</strong> sedimentary (low-temperature)dolomite are present in the geological record, especially in thePrecambrian.4.3.2. Solution chemistry <strong>of</strong> dolomiteClearly, the precipitation <strong>of</strong> dolomite from supersaturated seawatersolution under earth-surface conditions is inhibited in someway, and research has shown that this can be attributed to molecularkin<strong>et</strong>ics, which affect the behaviour <strong>of</strong> the component ions <strong>of</strong>dolomite in solution. These kin<strong>et</strong>ic barriers have b<strong>ee</strong>n identified as:(1) The high hydration <strong>ene</strong>rgy <strong>of</strong> the Mg 2+ ion (Lippmann, 1973;Dasent, 1982; Slaughter and Hill, 1991). Cations in aqueoussolutions form hydration shells, which enhance solubility. Themagnesium cation is strongly attached to the oxygen atoms <strong>of</strong>six water molecules that act as electric dipoles whose removalrequires <strong>ene</strong>rgy.(2) The extremely low concentration and activity <strong>of</strong> CO 3 2− (e.g.Garrels and Thompson, 1962; Lippmann, 1973; Slaughter andHill, 1991).(3) The presence <strong>of</strong> even very low concentrations <strong>of</strong> sulphate,which causes ion complexing including the formation <strong>of</strong>strongly bonded neutral MgSO 4 0 and CaSO 4 0 ion pairs(Usdowski, 1967; Baker and Kastner, 1981; Kastner, 1984;Morrow and Rick<strong>et</strong>ts, 1988; Slaughter and Hill, 1991; Wacey,2003; Wright and Wacey, 2004, 2005; Wright and Oren, 2005).Both (1) and (3) significantly increase the solubility <strong>of</strong> the Mg 2+ion, which significantly in experiments is always the last to precipitateout from evaporated brines (e.g. Borchert and Muir, 1964,personal observation). In saline solutions, ion pairs form due toshort-range interactions <strong>of</strong> adjacent ions, attracted by coulombicforces. This complexing reduces the ions’ activities below theirmolarities, making precipitation <strong>of</strong> carbonate minerals unlikely.For example more than 90% <strong>of</strong> total CO 3 2− in seawater is complexedwith hydrated m<strong>et</strong>al cations, chiefly magnesium (Garrelsand Christ, 1965).4.3.3. Sulphate and sulphate reducing bacteriaThe requirement to overcome these inhibitors for inorganicprecipitation led some researchers to investigate s<strong>et</strong>tings wheremodern sedimentary dolomite is forming, specifically to understandhow the kin<strong>et</strong>ic barriers might be overcome in the naturalenvironment. A common link was subsequently observed b<strong>et</strong>w<strong>ee</strong>nmicrobial mediation, organic degradation, raised carbonate alkalinity,removal or absence <strong>of</strong> sulphate and dolomite formation in bothmodern and ancient sediments (e.g. Gieskes <strong>et</strong> al., 1982; Compton,1988; Vasconcelos <strong>et</strong> al., 1995; Vasconcelos and McKenzie, 1997;Wright, 1997, 1999, 2000; Warthmann <strong>et</strong> al., 2000; van Lith <strong>et</strong> al.,2003; Roberts <strong>et</strong> al., 2004; Wright and Wacey, 2004, 2005; Wrightand Oren, 2005; Wacey <strong>et</strong> al., 2007; Kenward <strong>et</strong> al., 2009; Oliveri <strong>et</strong>al., 2010).Sulphate is an abundant component <strong>of</strong> seawater, and formsneutral ion pairs with m<strong>et</strong>al cations, enhancing their solubility;moreover the proportion <strong>of</strong> ion pairs increases with ionic concentration.The work <strong>of</strong> e.g. Walter (1986) has shown that sulphatecan inhibit calcite precipitation, and this is entirely consistent withdolomite inhibition – the same kin<strong>et</strong>ics are involved, with sulphateforming neutral ion pairs with Ca 2+ , but the hydration shell surroundingthe Ca 2+ ion is less tightly bound than that surroundingthe Mg 2+ ion (1926 kJ per mole at infinite dilution).Sulphate reducing bacteria (SRB) can remove all kin<strong>et</strong>icinhibitors to dolomite formation through dissociation <strong>of</strong> MgSO 40ion pairs as a part <strong>of</strong> their m<strong>et</strong>abolic activity, the formation <strong>of</strong>carbonate ions through buffering <strong>of</strong> bicarbonate during organicdiag<strong>ene</strong>sis (Slaughter and Hill, 1991; Wright, 1999), and the disruption<strong>of</strong> hydration shells at electrically charged surfaces <strong>of</strong> degradedorganic material (Kafkafi <strong>et</strong> al., 1967; Slaughter and Hill, 1991;Balarew <strong>et</strong> al., 2001). Charged cell walls attract both Mg 2+ andCa 2+ cations, with subsequent adsorption <strong>of</strong> carbonate ions andthe formation <strong>of</strong> fine-grained carbonates, and with extracellularpolysaccharides serve as nucleation centres for dolomite precipitation(e.g. van Lith <strong>et</strong> al., 2003).Hardie (1987) used the presence <strong>of</strong> sulphate in certain lakewaters to argue against the effectiv<strong>ene</strong>ss <strong>of</strong> sulphate as an inhibitorto dolomite formation. However, the presence <strong>of</strong> sulphate is necessaryfor SRB to m<strong>et</strong>abolise, and dolomite is formed not in aerobiclake waters but in anoxic conditions associated with the zone <strong>of</strong>microbial sulphate reduction. The experiments <strong>of</strong> Usdowski (1967),Baker and Kastner (1981) and Morrow and Rick<strong>et</strong>ts (1988), providecompelling evidence for sulphate inhibition <strong>of</strong> dolomite precipi-


A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737 733tation. Sulphate is <strong>of</strong> course present in seawater, but in modernd<strong>ee</strong>p-sea organic-rich sediments, where continual diffusion <strong>of</strong> seawaterSO 4 2− f<strong>ee</strong>ds and controls populations <strong>of</strong> SRB, dolomite formsin sulphate-fr<strong>ee</strong> anoxic interstitial waters immediately b<strong>ene</strong>ath thezone <strong>of</strong> bacterial sulphate reduction (Gieskes <strong>et</strong> al., 1982; Compton,1988). Where sulphate reduction was absent in d<strong>ee</strong>p-sea samples,no dolomite was reported (Gieskes <strong>et</strong> al., 1982). These observationsprovide the necessary evidence that the removal <strong>of</strong> sulphateby SRB from a solution in which it was originally present can leadto dolomite formation.Observations and data reported in Wright (1999, 2000), Wacey(2003), Wacey <strong>et</strong> al. (2007) and in experiments by Wright andWacey (2004, 2005) also demonstrate that primary dolomite formswhere bacterial sulphate reduction has removed sulphate in thepresence <strong>of</strong> organic matter. For example, high counts <strong>of</strong> SRBrecorded from Coorong distal ephemeral lake sediments (up to <strong>of</strong>3.74 × 10 6 /ml in dolomitic lakes) and enrichment <strong>of</strong> 34 S in residuallakewaters, indicate flourishing microbial populations and bacterialfractionation <strong>of</strong> sulphate during sulphate reduction. Sulphateconcentrations were extremely high (>20,000 mg/l) in the lakewatersamples, but declined dramatically and progressively withdepth through the sulphate reduction zone in dolomitic sedimentpore-water samples. By the end <strong>of</strong> the evaporative cycle, sulphatewas entirely removed. These observations, data and experimentsindicate that it is bacterial sulphate reduction and the consequentbiochemical interactions that drive dolomite formation.As long ago as 1928, Georgii Nadson, a Sovi<strong>et</strong> biologist, recognizedan association b<strong>et</strong>w<strong>ee</strong>n microbial mediation and carbonate(including dolomite) formation, associated with bacterial sulphatereduction. Other early studies in which bacteria were implicated indolomite formation are those by Neher (1959), Neher and Rohrer(1958) and Mansfield (1980). Oppenheimer and Master (1964)reported dolomite formation in association with ‘algal mats’ whileDavies and Ferguson (1975) established a ‘causal relationship’b<strong>et</strong>w<strong>ee</strong>n dolomite and decaying organic matter in experiments.M<strong>et</strong>hanogens and dissimilatory iron reducing bacteria have alsob<strong>ee</strong>n strongly implicated in dolomite formation (e.g. Roberts <strong>et</strong> al.,2004).4.3.4. Abiological mechanismsThe organogenic approach dating back to Nadson was supplantedby a series <strong>of</strong> theor<strong>et</strong>ical physico-chemical models thatbecame popular but remained unsupported by experimental evidence(for reviews, s<strong>ee</strong> Hardie, 1987; Tucker and Wright, 1990;Wright, 2000). When interpr<strong>et</strong>ing dolomite formation in the context<strong>of</strong> these ‘conventional’ models, consideration <strong>of</strong> fundamentalchemical constraints has <strong>of</strong>ten b<strong>ee</strong>n avoided or neglected. Most<strong>of</strong> these models involve dolomitization <strong>of</strong> limestone by circulatingbrines or fluids operating in specific hydrologic systems (e.g.Adams and Rhodes, 1960; Badiozamani, 1973; Folk and Land, 1975;Machel and Mountjoy, 1987).One model commonly referred to as the ‘magnesium pump’ wasinvoked to explain massive dolomitization <strong>of</strong> carbonate platformsby the circulation or tidal pumping <strong>of</strong> normal seawater through carbonateplatforms (e.g. Carballo <strong>et</strong> al., 1987; Land, 1991). However,seawater is unable to act as a dolomitizing solution because <strong>of</strong> thehigh enthalpy <strong>of</strong> hydration <strong>of</strong> the magnesium ion, the low activity<strong>of</strong> the carbonate ion and the presence <strong>of</strong> sulphate – all effectiveinhibitors to dolomite formation. Lippmann (1973) argued that fordolomitization <strong>of</strong> calcite to even begin in normal seawater, all Ca 2+must first be removed by precipitation, and a significant increase incarbonate alkalinity would be required for it to proc<strong>ee</strong>d – changesin chemical composition that are impossible to maintain in normalseawater.A similar case can be made against the mixing zone or ‘Dorag’model (Badiozamani, 1973), in which mixed marine and m<strong>et</strong>eoricwaters produce a solution both undersaturated in calcite and supersaturatedin dolomite. Subsequent high Mg:Ca ratios are unlikelyto lead to dolomitization because there is no mechanism to breachthe high hydration barrier <strong>of</strong> the magnesium ion, nor increase carbonateion activities. This model also predicts that dolomitizationwould be concomitant with the dissolution <strong>of</strong> calcite, but as Hardie(1987) reports, this does not occur in modern mixing zones. Smith<strong>et</strong> al. (2003) provides an example <strong>of</strong> dolomite formation by SRB ina modern mixing zone.Overriding kin<strong>et</strong>ic problems are also inherent in models invokingan evaporative mechanism, including s<strong>ee</strong>page reflux (Adamsand Rhodes, 1960) and evaporative pumping (Hsu and Siegenthaler,1969). Lippmann (1973) showed that evaporation would not favourdolomitization despite consequent high Mg:Ca ratios, because <strong>of</strong>the accompanying decrease in the activity <strong>of</strong> the CO 3 2− fr<strong>ee</strong> ion. Thealready low activity <strong>of</strong> the CO 3 2− ion would be further suppresseddue to its complexing to form neutral ion pairs, giving MgCO 3 0 ,while Mg 2+ would also complex with sulphate, as MgSO 4 0 .Experimental evaporation <strong>of</strong> brines in beakers has not produceddolomite in the sequence <strong>of</strong> precipitated minerals (e.g. Borchertand Muir, 1964, personal observation) clearly indicating that simpl<strong>ee</strong>vaporation does not produce dolomite. Proposals that dolomite isan evaporative mineral (e.g. McKenzie <strong>et</strong> al., 1980), are thus shownto be untenable. Where dolomite occurs b<strong>ene</strong>ath the Abu Dhabisabkha, it does not cross-cut facies, as might be expected froms<strong>ee</strong>page reflux or evaporative pumping, but is located within horizontalbeds associated with cyanobacterial mats (McKenzie <strong>et</strong> al.,1980; Wright, 2000) – suggesting that microbial mediation may beinvolved. More recent work by Bontognali <strong>et</strong> al. (2010) attributesthe Abu Dhabi dolomite formation to a microbially mediated origin.4.3.5. Documentation <strong>of</strong> anaerobic microbial mediationIn contrast, anaerobic microbial mediation has b<strong>ee</strong>n observedto cause the precipitation <strong>of</strong> dolomite in modern sedimentsand groundwater-basalt systems, with several different m<strong>et</strong>abolicpathways and waters <strong>of</strong> different composition (Moore <strong>et</strong> al., 2004;Roberts <strong>et</strong> al., 2004; Douglas, 2005; Wright and Oren, 2005 and referencestherein; Wright and Wacey, 2005). These cases <strong>of</strong> dolomiteformation have b<strong>ee</strong>n replicated under controlled conditions inthe laboratory (e.g. Vasconcelos <strong>et</strong> al., 1995; Roberts <strong>et</strong> al., 2004;Wright and Wacey, 2004, 2005), making a compelling case thatmicrobial mediation is essential for low-temperature precipitation<strong>of</strong> dolomite in saline solutions (Douglas, 2005; Wright and Oren,2005; Wacey <strong>et</strong> al., 2007).The idea that dolomite n<strong>ee</strong>ds thousands or even millions <strong>of</strong>years to form is manifestly false, since recent experiments haveconclusively demonstrated that dolomite can be formed in daysthrough microbial mediation (Wright and Wacey, 2004, 2005). Onthe antiquity <strong>of</strong> microbial sulphate reduction, Shen <strong>et</strong> al. (2001)and Shen and Buick (2004) have used sulphur isotope data to argu<strong>et</strong>hat sulphate reducing microbes had evolved by the early Archaean.The large spread <strong>of</strong> 34 S values <strong>of</strong> microscopic pyrites aligned alonggrowth faces <strong>of</strong> former gypsum crystals in the ∼3.47-Ga North Polebarite deposit <strong>of</strong> northwestern Australia provide the oldest evidence<strong>of</strong> microbial sulphate reduction and the earliest indication<strong>of</strong> a specific microbial m<strong>et</strong>abolism. Microbial ecosystems dominatedthe depositional environments <strong>of</strong> Archaean, Proterozoic andsome Phanerozoic carbonate platforms, in which sulphate reduction,organic diag<strong>ene</strong>sis and dolomite formation were widespread(e.g. Wright and Altermann, 2000; Shen <strong>et</strong> al., 2005; Altermann <strong>et</strong>al., 2006; Gandin and Wright, 2007). Such ecosystems thus actednot only as the ‘engine house’ <strong>of</strong> early carbonate production, butalso provide a ‘process analogue’ for dolomite formation.In conclusion, although seawater is supersaturated with respectto dolomite, the mineral is prevented from precipitating undernormal earth-surface conditions by kin<strong>et</strong>ic inhibitors. Numer-


A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737 735interpr<strong>et</strong>ation <strong>of</strong> the process <strong>of</strong> low-temperature dolomite growthin some Isua rocks is correct, then by 3700 Ma, life was alreadywell established in a range <strong>of</strong> ecological niches, including in asub-surface volcanic pillow interstice environment. Such an environmentwould afford life protection from any surface-sterilizingmassive impacts at that early stage <strong>of</strong> Earth’s history (summarizedby Koeberl, 2006), and from whence surficial environments couldbe readily colonized again.5. Conclusions(1) Dolomite was a pre-m<strong>et</strong>amorphic mineral in some Isua chemicalsedimentary rocks and pillow interstices.(2) The Isua dolomite (e.g. sample JHA170728) has a REE + Y (SN)seawater-like signature identical to international BIF standardIF-G from Isua, showing it formed in a pre-m<strong>et</strong>amorphic lowtemperatur<strong>ee</strong>nvironment, during deposition or diag<strong>ene</strong>sis.(3) Recent low-temperature dolomite precipitation in sedimentarysystems and intra-basalt ground waters apparently requiresmicrobial mediation, and this has b<strong>ee</strong>n replicated under controlledlaboratory conditions. On the other hand, years <strong>of</strong>experimentation has failed to precipitate low-temperaturedolomite by abiotic means. Given our results and theseobservations, we propose that geochemical evidence for prem<strong>et</strong>amorphicIsua low-temperature dolomite is simple, directevidence for microbial life by 3700 Ma.(4) If some Akilia association BIF with a dolomitic component andseawater-like REE + Y (SN) signatures really are ≥3850 millionyears old, then the dolomite component in these rocks is evidence<strong>of</strong> life at the start <strong>of</strong> Earth’s known sedimentary record.AcknowledgementsThis research was funded by Australian Research Council grantDP0342794, 2006–2009 support from the Chinese Academy <strong>of</strong>Geological Sciences and currently an operating grant from theUniversity <strong>of</strong> Wollongong. The Geological Survey <strong>of</strong> Denmark andGr<strong>ee</strong>nland is thanked for permission to publish this paper.Appendix A. Supplementary dataSupplementary data associated with this article can be found, inthe online version, at doi:10.1016/j.precamres.2010.08.006.ReferencesAdams, J.E., Rhodes, M.L., 1960. Dolomitization by s<strong>ee</strong>page refluxion. AAPG Bull<strong>et</strong>in44, 1912–1921.Allaart, J.H., 1976. The pre-3760 m.y. old supracrustal rocks <strong>of</strong> the Isua area, centralWest Gr<strong>ee</strong>nland, and the associated occurrence <strong>of</strong> quartz-banded ironstone. In:Windley, B.F. (Ed.), The Early History <strong>of</strong> the Earth. Wiley, London, pp. 177–189.Altermann, W., Kazmierczak, J., Oren, A., Wright, D.T., 2006. Cyanobacterial calcificationand its rock-building potential during 3.5 billion years <strong>of</strong> Earth history.Geobiology 4, 147–166.Appel, P.W.U., Moorbath, S., Myers, J.S., 2003. Isuasphaera isua (Pflug) revisited. PrecambrianResearch 126, 309–312.Badiozamani, K., 1973. The Dorag dolomitization model – application to the MiddleOrdovician <strong>of</strong> Wisconsin. Journal <strong>of</strong> Sedimentary P<strong>et</strong>rology 43, 965–984.Baker, P., Kastner, M., 1981. Constraints on the formation <strong>of</strong> sedimentary dolomite.Science 213, 214–216.Balarew, C., Tepavitcharova, S., Rabadjieva, D., Voigt, D., 2001. Solubility and crystallizationin the system MgCl 2–MgSO 4–H 2O at 50 and 75 ◦ C. Journal <strong>of</strong> SolutionChemistry 30, 815–823.Bau, M., Dulski, P., 1999. Comparing yttrium and rare earths in hydrothermal fluidsfrom the Mid-Atlantic Ridge: implications for Y and REE behaviour during nearventmixing and for the Y/Ho ratio <strong>of</strong> Proterozoic seawater. Chemical Geology155, 77–90.Beukes, N.J., 1987. Facies relations, depositional environments and diag<strong>ene</strong>sis in amajor early Proterozoic stromatolitic carbonate platform to basinal sequence,Campbellrand Subgroup, Transvaal Supergroup, Southern Africa. SedimentaryGeology 54, 1–46.Beukes, N.J., 2004. Early options in photosynthesis. Nature 431, 522–523.Bolhar, R., Kamber, B.S., Moorbath, S., Fedo, C.M., Whitehouse, M.J., 2004. Characterisation<strong>of</strong> early Archaean chemical sediments by trace element signatures. Earthand Plan<strong>et</strong>ary Science L<strong>et</strong>ters 222, 43–60.Bolhar, R., van Kranendonk, M.J., Kamber, B.S., 2005. A trace element study <strong>of</strong>siderite-jasper banded iron formation in the 3.45 Ga Warrawoona Group. PilbaraCraton – formation from hydrothermal fluids and shallow seawater. PrecambrianResearch 137, 93–114.Bontognali, T.R.R., Vasconcelos, C., Warthmann, R.J., Bernasconi, S.M., Dupraz, C.,Strohmenger, C.J., McKenzie, J.A., 2010. Dolomite formation within microbialmats in the coastal sabkha <strong>of</strong> Abu Dhabi (United Arab Emirates). Sedimentology57, 824–844.Borchert, H., Muir, R.O., 1964. Salt Deposits. Van Nostrand Reinhold, London.Burns, S.J., Baker, P.A., Showers, W.J., 1988. The factors controlling the formationand chemistry <strong>of</strong> dolomite in organic-rich sediments: Mioc<strong>ene</strong> DrakesBay Formation, California. In: Shulka, V., Baker, P.A. (Eds.), Sedimentology andGeochemistry <strong>of</strong> Dolostones, 43. Soci<strong>et</strong>y <strong>of</strong> Economic Paleontologists and Mineralogists,Special Publications, pp. 1–52.Burns, S.J., McKenzie, J.A., Vasconcelos, C., 2000. Dolomite formation and biogeochemicalcycles in the Phanerozoic. Sedimentology 47, 49–61.Carballo, J.D., Land, L.S., Miser, D.E., 1987. Holoc<strong>ene</strong> dolomitization <strong>of</strong> supratidal sedimentsby active tidal pumping, Sugarloaf Key, Florida. J. Sedimentary P<strong>et</strong>rology57, 153–165.Cloud, P.E., 1973. Paleoecological significance <strong>of</strong> banded iron formation. EconomicGeology 68, 1135–1143.Compton, J.S., 1998. Sediment composition and precipitation <strong>of</strong> dolomite and pyritein the Neog<strong>ene</strong> Monterey and Sisquoc formations, Santa Maria Basin area, California.In Shukla, V., Baker, P.A. (Eds.), Sedimentology and Geochemistry <strong>of</strong>Dolostones. The Soci<strong>et</strong>y <strong>of</strong> Economic Paleontologists and Mineralogists SpecialPublication 43, 53–64.Crowley, J.L., 2003. U-Pb geochronology <strong>of</strong> 3810-3630 Ma granitoid rocks south <strong>of</strong>the Isua gr<strong>ee</strong>nstone belt, southern West Gr<strong>ee</strong>nland. Precambrian Research 126,235–257.Dasent, W.E., 1982. Inorganic Energ<strong>et</strong>ics, 2nd ed. Cambridge University Press, NewYork.Dauphas, N., van Zuilen, M., Wadhwa, M., Davis, A.M., Marty, B., Janney, P.E., 2004.Clues from Fe isotope variations on the origin <strong>of</strong> early Archean BIFs from Gr<strong>ee</strong>nland.Science 306, 2077–2080.Davies, P.J., Ferguson, J., 1975. Dolomite and organic material. Nature 255, 472–473.Dilek, Y., Polat, A., 2008. Suprasubduction zone ophiolites and Archean tectonics.Geology 36, 431–432.Douglas, S., 2005. Mineralogical footprints <strong>of</strong> microbial life. American Journal <strong>of</strong>Science 305, 503–525.Douville, E., Bienvenu, P., Charlou, J.L., Donval, J.P., Fouqu<strong>et</strong>, Y., Appriou, P., Gamo,T., 1999. Yttrium and rare earth elements in fluids from various d<strong>ee</strong>p-seahydrothermal systems. Geochimica <strong>et</strong> Cosmochimica Acta 63, 627–643.Dymek, R.F., Klien, C., 1988. Chemistry, p<strong>et</strong>rology and origin <strong>of</strong> banded ironformationlithologies from the 3800 Ma Isua supracrustal belt, West Gr<strong>ee</strong>nland.Precambrian Research 39, 247–302.Folk, R.L., Land, L.S., 1975. Mg/Ca ratio and salinity: two controls over crystallization<strong>of</strong> dolomite. AAPG Bull<strong>et</strong>in 59, 60–68.Friend, C.R.L., Benn<strong>et</strong>t, V.C., <strong>Nutman</strong>, A.P., Norman, M., 2007. Seawater-like trac<strong>ee</strong>lement signatures (REE + Y) <strong>of</strong> Eoarchaean chemical sedimentary rocks fromsouthern West Gr<strong>ee</strong>nland, and their corruption during high-grade m<strong>et</strong>amorphism.Contributions to Mineralogy and P<strong>et</strong>rology, doi:10.1007/S00410-007-0239-z.Friend, C.R.L., <strong>Nutman</strong>, A.P., 2005. Complex 3670-3500 Ma orogenic episodessuperimposed on juvenile crust accr<strong>et</strong>ed b<strong>et</strong>w<strong>ee</strong>n 3850 and 3690 Ma, ItsaqGneiss Complex, southern West Gr<strong>ee</strong>nland. Journal <strong>of</strong> Geology 113, 375–398.Gaines, A.M., 1980. Dolomitization kin<strong>et</strong>ics: recent experimental studies. In Zenger,D.H., Dunham, J.B., Ethington, R.L. (Eds.), Concepts and Models <strong>of</strong> Dolomitization.Soci<strong>et</strong>y <strong>of</strong> Economic Paleontologists and Mineralogists Special Publication 28,81–86.Gandin, A., Wright, D.T., 2007. Evidence <strong>of</strong> vanished evaporites in Neoarchaean carbonates<strong>of</strong> South Africa. In Schreiber, B.C., Lugli, S., Babel, M. (Eds.), EvaporitesThrough Space and Time. Geological Soci<strong>et</strong>y, London, Special Publication 285,341–364. doi:10.1144/SP285.19.Gandin, A., Wright, D.T., Melezhik, V., 2005. Vanished evaporites and carbonateformation in the Neoarchaean Kogelb<strong>ee</strong>n and Gamohaan Formations <strong>of</strong>South Africa. Presidential Review 8, Journal <strong>of</strong> African Earth Sciences 41, 1–23,doi:10.1016/j.jafrearsci.2005.01.003.Garrels, R.M., Christ, C.L., 1965. Solutions, minerals, and equilibria. Harper & Row,New York, 450 p.Garrels, R.M., Thompson, M.E., 1962. A chemical model for seawater at 25 ◦ C and oneatmosphere total pressure. American Journal <strong>of</strong> Science 260, 57–66.Garrels, R.M., Perry, E.A., MacKenzie, F.T., 1973. G<strong>ene</strong>sis <strong>of</strong> Precambrian iron formationsand the development <strong>of</strong> atmospheric oxygen. Economic Geology 68,1173–1179.Gieskes, J.M., Elderfield, H., Lawrence, J.R., Johnson, J., Meyers, B., Campbell, A., 1982.Geochemistry <strong>of</strong> interstitial waters and sediments, Leg 64, Gulf <strong>of</strong> California.In: Curray, J.R., Moore, D.G., <strong>et</strong> al. (Eds.), Initial Reports <strong>of</strong> the D<strong>ee</strong>p Sea DrillingProject, vol. 64. U.S. Government Printing Office, Washington, DC, pp. 675–694.Hardie, L.A., 1987. Perspectives: dolomitization: a critical view <strong>of</strong> some currentviews. Journal <strong>of</strong> Sedimentary P<strong>et</strong>rology 57, 166–183.Hecht, L., Freiberger, R., Gilg, H.A., Grundmann, G., Kostitsyn, Y.A., 1999. Rare earthelement and isotope (C, O, Sr) characteristics <strong>of</strong> hydrothermal carbonates:


736 A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737gen<strong>et</strong>ic implications for the dolomite-hosted talc mineralization at Göpfersgrün(Fichtelgebirge, Germany). Chemical Geology 155, 115–130.Hsu, K.J., Siegenthaler, C., 1969. Preliminary experiments and hydrodynamic movementinduced by evaporation and their bearing on the dolomite problem.Sedimentology 12, 11–25.Johannesson, K.H., Hawkins D.L.Jr., Cortés, A., 2006. Do Archean chemical sedimentsrecord ancient seawater rare earth element patterns? Geochimica <strong>et</strong>Cosmochimica Acta 70, 871–890.Kafkafi, U., Posner, A.M., Quirk, J.P., 1967. De-sorption <strong>of</strong> phosphate from kaolinite.Soil Science Soci<strong>et</strong>y <strong>of</strong> America Proc<strong>ee</strong>dings 31, 348–353.Kasting, J.F., 1993. Earth’s early atmosphere. Science 259, 920–926.Kastner, M., 1984. Control <strong>of</strong> dolomite formation. Nature 311, 410–411.Kelleher, I.J., Redfern, S.A.T., 2002. Hydrous calcium magnesium carbonate, a possibleprecursor to the formation <strong>of</strong> sedimentary dolomite. Molecular Simulation 28,557–572.Kenward, P.A., Goldstein, R.H., GonzÁlez, L.A., Roberts, J.A., 2009. Precipitation <strong>of</strong>low-temperature dolomite from an anaerobic microbial consortium: the role <strong>of</strong>m<strong>et</strong>hanogenic Archaea. Geobiology 7, 556–565.Koeberl, C., 2006. Impact processes on the early Earth. Elements 2, 126–211.Konhauser, K.O., Hamade, T., Morris, R.C., Ferris, F.G., Southam, G., Raiswell, R.,Canfield, D., 2002. Could bacteria have formed the Precambrian banded ironformations? Geology 30, 1079–1082.Land, L.S., 1991. Dolomitization models – seawater mixing zones. (Abs.). In:Dolomieu Conference on Carbonate Platforms and Dolomitization, 144, Karo-Drucle, Eppan, Italy.Land, L.S., 1998. Failure to precipitate dolomite at 25 ◦ C from dilute solutiondespite 1000-fold oversaturation after 32 years. Aqueous Geochemistry 4, 361–368.Lippmann, E., 1973. Sedimentary Carbonate Minerals. Springer, Berlin.Liss, A., Widdel, F., Schnell, S., Heising, S., Ehrenreich, A., Assmus, B., Schink, B.,1993. Ferrous iron oxidation by anoxygenic phototrophic bacteria. Nature 362,834–836.Lumsden, D.N., Morrison, J.W., Lloyd, R., 1995. Role <strong>of</strong> iron and Mg/Ca ratio indolomite synthesis at 192 ◦ C. Journal <strong>of</strong> Geology 103, 51–61.Machel, H.G., Mountjoy, E.W., 1986. Chemistry and environments <strong>of</strong> dolomitization– a reappraisal. Earth Sciences Reviews 23, 175–222.Machel, H.G., Mountjoy, E.W., 1987. G<strong>ene</strong>ral constraints on extensive pervasivedolomitization - and their application to the Devonian carbonates <strong>of</strong> westernCanada. Bull<strong>et</strong>in <strong>of</strong> Canadian P<strong>et</strong>roleum Geology 35, 143–158.Manning, C.E., Mojzsis, S.J., Harrison, T.M., 2006. Geology, age and origin <strong>of</strong>supracrustal rocks at Akilia, West Gr<strong>ee</strong>nland. American Journal <strong>of</strong> Science 306,303–366.Mansfield, C.F, 1980. A urolith <strong>of</strong> biogenic dolomite-another clue in the dolomitemystery. Geochimica <strong>et</strong> Cosmochimica Acta 44, 829–839.McClennan, S.M., 1989. Rare earth elements in sedimentary rocks: influence <strong>of</strong>provenance and sedimentary processes. In: Lipin, B.R., McKay, G.A. (Eds.), Geochemistryand Mineralogy <strong>of</strong> Rare Earth Elements. Mineralogical Soci<strong>et</strong>y <strong>of</strong>America, Washington, DC, USA, pp. 169–200.McGregor, V.R., Mason, B., 1977. P<strong>et</strong>rog<strong>ene</strong>sis and geochemistry <strong>of</strong> m<strong>et</strong>abasaltic andm<strong>et</strong>asedimentary enclaves in the Amîtsoq gneisses, West Gr<strong>ee</strong>nland. AmericanMineralogist 62, 887–904.McKenzie, J.A., Hsu, K.J., Schneider, J.F., 1980. Movement <strong>of</strong> subsurface waters underthe sabkha, Abu Dhabi, UAE, and its relation to evaporative dolomite g<strong>ene</strong>sis.In Zenger, D.H., Dunham, J.B., Ethington, R.E. (Eds.), Concepts and Models <strong>of</strong>Dolomitization. Soci<strong>et</strong>y <strong>of</strong> Economic Paleontologists and Mineralogists SpecialPublication 28, 11–30.Moorbath, S., O’Nions, R.K., Pankhurst, R.J., 1973. Early Archaean age for the Isua ironformation, West Gr<strong>ee</strong>nland. Nature 245, 138–139.Moore, T.S., Murray, R.W., Kurtz, A.C., Schrag, D.P., 2004. Anaerobic m<strong>et</strong>hane oxidationand the formation <strong>of</strong> dolomite. Earth and Plan<strong>et</strong>ary Science L<strong>et</strong>ters 229,141–154.Morrow, D.W., Rick<strong>et</strong>ts, B.D., 1988. Experimental investigation <strong>of</strong> sulfate inhibition<strong>of</strong> dolomite and its mineral analogues. In: Shukla, V., Baker, P.A. (Eds.), Proc<strong>ee</strong>dings<strong>of</strong> Sedimentology and Geochemistry <strong>of</strong> Dolostones. Raleigh, North Carolina,26-28 September 1986. Soci<strong>et</strong>y <strong>of</strong> Economic Paleontologists and Mineralogists,Special Publications, 43, pp. 25–38.Mozjsis, S.J., Arrhenius, G., McK<strong>ee</strong>gan, K.D., Harrison, T.M., Friend, C.R.L., 1996.Evidence for life on Earth before 3800 million years ago. Nature 270,43–45.Myers, J.S., 2001. Protoliths <strong>of</strong> the 3.8–3.7 Ga Isua gr<strong>ee</strong>nstone belt, West Gr<strong>ee</strong>nland.Precambrian Research 105, 129–141.Nadson, G.A., 1928. Beitrag zur Kenntnis der Bakteriog<strong>ene</strong>n Kalkablagerungen.Archiv fuer Hydrobiologie 19, 154–164.Neher, J., 1959. Bakterien in tieferliegenden Gesteinlagen. Eclogae Geologia Helv<strong>et</strong>ica52, 619–625.Neher, J., Rohrer, E., 1958. Dolomitbildung unter Mitwirkung von Bakterien. EclogaeGeologica Helv<strong>et</strong>ica 51, 213–215.Nothdurft, L.D., Webb, G.E., Kamber, B.S., 2004. Rare earth element geochemistry <strong>of</strong>Late Devonian r<strong>ee</strong>fal carbonates, Canning basin. Western Australia: confirmation<strong>of</strong> a seawater REE proxy in ancient limestones. Geochimica <strong>et</strong> CosmochimicaActa 68, 263–283.<strong>Nutman</strong>, A.P., 1980. A field and laboratory study <strong>of</strong> the early Archaean rocks <strong>of</strong>the northwestern Buksefjorden region, southern West Gr<strong>ee</strong>nland. Ph.D. Thesis,University <strong>of</strong> Ex<strong>et</strong>er, UK.<strong>Nutman</strong>, A.P., 2006. Antiquity <strong>of</strong> the oceans and continents. Elements 2, 223–227.<strong>Nutman</strong>, A.P., Friend, C.R.L., 2006. P<strong>et</strong>rography and geochemistry <strong>of</strong> apatites inbanded iron formation, Akilia, W. Gr<strong>ee</strong>nland: consequences for oldest life evidence.Precambrian Research 147, 100–106.<strong>Nutman</strong>, A.P., Friend, C.R.L., 2009. New 1:20,000 scale geological maps, synthesis andhistory <strong>of</strong> investigation <strong>of</strong> the Isua supracrustal belt and adjacent orthogneisses,southern West Gr<strong>ee</strong>nland: a glimpse <strong>of</strong> Eoarchaean crust formation and orogeny.Precambrian Research 172, 189–211.<strong>Nutman</strong>, A.P., Allaart, J.H., Bridgwater, D., Dimroth, E., Rosing, M.T., 1984. Stratigraphicand geochemical evidence for the depositional environment <strong>of</strong> th<strong>ee</strong>arly Archaean Isua supracrustal belt, southern West Gr<strong>ee</strong>nland. PrecambrianResearch 25, 365–396.<strong>Nutman</strong>, A.P., Friend, C.R.L., Benn<strong>et</strong>t, V.C., 2002. Evidence for 3650-3600 Ma assembly<strong>of</strong> the northern end <strong>of</strong> the Itsaq Gneiss Complex, Gr<strong>ee</strong>nland: Implication for earlyArchean tectonics. Tectonics 21, article 5.<strong>Nutman</strong>, A.P., Friend, C.R.L., Horie, H., Hidaka, H., 2007. Construction <strong>of</strong> pre-3600 Macrust at convergent plate boundaries, exemplified by the Itsaq Gneiss Complex<strong>of</strong> southern West Gr<strong>ee</strong>nland. In: van Kranendonk, M.J., Smithies, R.H., Benn<strong>et</strong>t,V.C. (Eds.), Earth’s Oldest Rocks. Elsevier, pp. 187–218.<strong>Nutman</strong>, A.P., McGregor, V.R., Friend, C.R.L., Benn<strong>et</strong>t, V.C., Kinny, P.D., 1996. The ItsaqGneiss Complex <strong>of</strong> southern West Gr<strong>ee</strong>nland; the world’s most extensive record<strong>of</strong> early crustal evolution (3900–3600 Ma). Precambrian Research 78, 1–39.<strong>Nutman</strong>, A.P., Benn<strong>et</strong>t, V.C., Friend, C.R.L., Rosing, M.T., 1997. ∼3710 and ≥3790 Mavolcanic sequences in the Isua (Gr<strong>ee</strong>nland) supracrustal belt; structural and Ndisotope implications. Chemical Geology 141, 271–287.<strong>Nutman</strong>, A.P., Friend, C.R.L., Benn<strong>et</strong>t, V.C., McGregor, V.R., 2000. The early ArchaeanItsaq Gneiss Complex <strong>of</strong> southern West Gr<strong>ee</strong>nland: the importance <strong>of</strong> fieldobservations in interpr<strong>et</strong>ing dates and isotopic data constraining early terrestrialevolution. Geochimica <strong>et</strong> Cosmochimica Acta 64, 3035–3060.<strong>Nutman</strong>, A.P., Friend, C.R.L., Barker, S.S., McGregor, V.R., 2004. Inventory and assessment<strong>of</strong> Palaeoarchaean gneiss terrains and d<strong>et</strong>rital zircons in southern WestGr<strong>ee</strong>nland. Precambrian Research 135, 281–314.<strong>Nutman</strong>, A.P., Friend, C.R.L., Paxton, S., 2009. D<strong>et</strong>rital zircon sedimentary provenanceages for the Eoarchaean Isua supracrustal belt southern West Gr<strong>ee</strong>nland: juxtaposition<strong>of</strong> a ca. 3700 Ma juvenile arc assemblage against an older complex with3920–3800 Ma components. Precambrian Research 172, 212–233.Oliveri, E., Neri, N., Bellanca, A., Riding, R., 2010. Carbonate stromatolites from aMessinian hypersaline s<strong>et</strong>ting in the Caltaniss<strong>et</strong>ta Basin, Sicily: p<strong>et</strong>rographicevidence <strong>of</strong> microbial activity and related stable isotope and rare earth elementsignatures. Sedimentology 57, 142–161.Oppenheimer, C.H., Master, I.M., 1964. Transition <strong>of</strong> silicate and carbonate crystalstructure by photosynthesis and m<strong>et</strong>abolism (abstract). Geological Soci<strong>et</strong>y <strong>of</strong>America, Special paper 76. Abstracts for 1963, 125.Perry, E.C., Ahmed, S.N., 1977. Carbon isotope composition <strong>of</strong> graphite and carbonateminerals from 3.8-AE m<strong>et</strong>amorphosed sediments, Isukasia, Gr<strong>ee</strong>nland. Earthand Plan<strong>et</strong>ary Science L<strong>et</strong>ters 36, 280–284.Pflug, H.D., Jaeschke-Boyer, H., 1979. Combined structural and chemical analysis <strong>of</strong>3,800-Myr-old micr<strong>of</strong>ossils. Nature 280, 483–486.Polat, A., H<strong>of</strong>mann, A.W., Rosing, M.T., 2002. Boninite-like volcanic rocks in the3.7–3.8 Ga Isua gr<strong>ee</strong>nstone belt, West Gr<strong>ee</strong>nland: geochemical evidence forintra-oceanic subduction zone processes in the early Earth. Chemical Geology184, 231–254.Purser, B.H., Tucker, M.E., Zenger, D.H., 1994. Problems, progress and future researchconcerning dolomites and dolomitization. In Purser, B., Tucker, M., Zenger, D.(Eds.), Dolomites: A Volume in Honour <strong>of</strong> Dolomieu. Special Publication <strong>of</strong> theInternational Association <strong>of</strong> Sedimentologists 21, 3–20.Roberts, J.A., Benn<strong>et</strong>t, P.C., González, L.A., Macpherson, G.L., Milliken, K.L., 2004.Microbial precipitation <strong>of</strong> dolomite in m<strong>et</strong>hanogenic groundwater. Geology 32,277–280.Rollinson, H., 2003. M<strong>et</strong>amorphic history suggested by garn<strong>et</strong>-growth chronologiesin the Isua Gr<strong>ee</strong>nstone Belt, West Gr<strong>ee</strong>nland. Precambrian Research 126,181–196.Rose, N.M., Rosing, M., Bridgwater, D., 1996. The origin <strong>of</strong> m<strong>et</strong>acarbonate rocks inthe Archean Isua supracrustal belt, West Gr<strong>ee</strong>nland. American Journal <strong>of</strong> Science296, 1004–1044.Rosing, M.T., 1999. 13 C-depl<strong>et</strong>ed carbon microparticles in >3700-Ma sea-floor sedimentaryrocks from West Gr<strong>ee</strong>nland. Science 283, 674–676.Rosing, M.T., Rose, N.M., Bridgwater, D., Thomsen, H.S., 1996. Earliest part <strong>of</strong> theEarth’s stratigraphic record: a reappraisal <strong>of</strong> the >3.7 Gaa Isua (Gr<strong>ee</strong>nland)supracrustal sequence. Geology 24, 43–46.Rouxel, O., Ono, S., Alt, J., Rumble, D., Ludden, J., 2008. Sulfur isotope evidence formicrobial sulphate reduction in altered oceanic basalts at ODP Site 801. Earthand Plan<strong>et</strong>ary Science L<strong>et</strong>ters 268, 110–123.Schoenburg, R., Kamber, B.S., Collerson, K.D., Moorbath, S., 2002. Tungsten isotop<strong>ee</strong>vidence from ∼3.8 Gyr m<strong>et</strong>amorphosed sediments for early m<strong>et</strong>eorite bombardment<strong>of</strong> Earth. Nature 418, 403–405.Schidlowski, M., Appel, P.W.U., Eichmann, R., Junge, C.E., 1979. Carbon isotope geochemistry<strong>of</strong> the 3.7 × 109-yr old Isua sediments, West Gr<strong>ee</strong>nland: implicationsfor the Archaean carbon and oxygen cycles. Geochimica <strong>et</strong> Cosmochimica Acta43, 189–199.Schopf, W.J., 2006. When did life emerge? Elements 2, 229–233,doi:10.2113/gselements.2.4.229.Shen, Y., Buick, R., 2004. The antiquity <strong>of</strong> microbial sulphate reduction. Earth-ScienceReviews 64, 243–272.Shen, Y., Buick, R., Canfield, D.E., 2001. Isotopic evidence for microbial sulphatereduction in the early Archaean era. Nature 410, 77–81.


A.P. <strong>Nutman</strong> <strong>et</strong> al. / Precambrian Research 183 (2010) 725–737 737Shen, J., Teng, J., Pedoja, K., 2005. Middle and Late Devonian microbial carbonates,r<strong>ee</strong>f and mounds in Guilin, South China and their sequence stratigraphic, paleoenvironmentaland paleoclimatic significance. Science in China, Series D: EarthScience 48, 1900–1912.Slaughter, M., Hill, R.J., 1991. The influence <strong>of</strong> organic matter in organogenic dolomitization.Journal <strong>of</strong> Sedimentary P<strong>et</strong>rology 61, 296–303.Smith, S.L., Whitaker, F.F., Parkes, R.J., 2003. Dolomitization by saline groundwatersin the Yucatan Peninsula. In: The Geom<strong>et</strong>ry and P<strong>et</strong>rog<strong>ene</strong>sis <strong>of</strong> DolomiteHydrocarbon Reservoirs – Final Programme and Abstracts.Solvang, M., 1999. An investigation <strong>of</strong> m<strong>et</strong>avolcanic rocks from the eastern part <strong>of</strong>the Isua gr<strong>ee</strong>nstone belt, West Gr<strong>ee</strong>nland. Geological Survey <strong>of</strong> Denmark andGr<strong>ee</strong>nland (GEUS) Internal Report, Copenhagen, Denmark, 62 pp.Tribble, J.S., Arvidson, R.S., Lane IV, M., MacKenzie, F.T., 1995. Crystal chemistry,and thermodynamic and kin<strong>et</strong>ic properties <strong>of</strong> calcite, dolomite, apatite, andbiogenic silica: applications to p<strong>et</strong>rologic problems. Sedimentary Geology 95,11–37.Tucker, M.E., Wright, V.P., 1990. Carbonate Sedimentology. Blackwell Scientific Publications,482 p.Usdowski, H.E., 1967. The formation <strong>of</strong> dolomite in sediments. In: Muller, G., Friedman,G.M. (Eds.), Recent Developments in Carbonate Sedimentology in CentralEurope. Springer, Heidelberg, pp. 21–32.van Kranendonk, M.J., 2007. A review <strong>of</strong> the evidence for putative Paleoarchean lifein the Pilbara Craton, Western Australia. In: Earth’s Oldest Rocks, Developmentsin Precambian Geology 15. Elsevier, pp. 855–878.van Kranendonk, M.J., Webb, G.E., Kamber, B.S., 2003. Geological and trace elementevidence for a marine sedimentary environment <strong>of</strong> deposition and biogenicity <strong>of</strong>3.45 Ga stromatolitic carbonates in the Pilbara Craton, and support for a reducingArchaean ocean. Geobiology 1, 91–108.van Kranendonk, M.J., Smithies, R.H., Hickman, A.H., Champion, D.C., 2007. Paleoarcheandevelopment <strong>of</strong> a continental nucleus: the east Pilbara terrane <strong>of</strong> thePilbara Craton, Western Australia. In: Earth’s Oldest Rocks, Developments inPrecambian Geology 15. Elsevier, pp. 307–337.van Lith, Y., Warthmann, R., Vasconcelos, C., McKenzie, J.A., 2003. Sulphate-reducingbacteria induce lowtemperature dolomite and high Mg-calcite formation. Geobiology1, 71–79.van Zuilen, M.A., Lepland, A., Arrhenius, G., 2002. Reassessing the evidence for earliesttraces <strong>of</strong> life. Nature 418, 627–630.Vasconcelos, C., McKenzie, J.A., 1997. Microbial mediation <strong>of</strong> modern dolomite precipitationand diag<strong>ene</strong>sis under anoxic conditions (Lagoa Vermelha, Rio deJaneiro, Brazil). Journal <strong>of</strong> Sedimentary Research 67, 378–390.Vasconcelos, C., McKenzie, J.A., Bernasconi, S., Grujic, D., Tien, A.J., 1995. Microbialmediation as a possible mechanism for natural dolomite at low temperatures.Nature 377, 220–222.Von Der Borch, C.C., 1976. Stratigraphy and formation <strong>of</strong> Holoc<strong>ene</strong> dolomitic carbonatedeposits <strong>of</strong> the Coorong area, South Australia. Journal <strong>of</strong> SedimentaryP<strong>et</strong>rology 46, 952–966.Wacey, D., 2003. Microbial mediation <strong>of</strong> dolomite formation: geochemical investigationsin the Coorong region <strong>of</strong> South Australia. DPhil thesis, University <strong>of</strong>Oxford.Wacey, D., Wright, D.T., Boyce, A.J., 2007. A stable isotope study <strong>of</strong> microbialdolomite formation in the Coorong Region, South Australia. Chemical Geology244, 155–174.Walter, L.M., 1986. Relative efficiency <strong>of</strong> carbonate dissolution and precipitationduring diag<strong>ene</strong>sis: a progress report on the role <strong>of</strong> solution chemistry.In Gautier, D.L. (Ed.), Roles <strong>of</strong> Organic Matter in Sediment Diag<strong>ene</strong>sis. TheSoci<strong>et</strong>y <strong>of</strong> Economic Paleontologists and Mineralogists Special Publication 38,1–11.Warthmann, R., van Lith, Y., Vasconcelos, C., McKenzie, J.A., Karp<strong>of</strong>f, A.M., 2000. Bacteriallyinduced dolomite precipitation in anoxic culture experiments. Geology28, 1091–1094.Webb, G.E., Kamber, B.S., 2000. Rare earth elements in Holoc<strong>ene</strong> r<strong>ee</strong>fal microbialites:a shallow seawater proxy. Geochimica <strong>et</strong> Cosmochimica Acta 64, 1557–1565.Whitehouse, M.J., Kamber, B.S., Moorbath, S., 1999. Age significance <strong>of</strong> U–Th–Pbzircon data from early Archaean rocks <strong>of</strong> west Gr<strong>ee</strong>nland – a reassessmentbased on combined ion-microprobe and imaging studies. Chemical Geology 160,201–224.Whitehouse, M.J., Myers, J.S., Fedo, C.M., 2009. Interpr<strong>et</strong>ations <strong>of</strong> >3.8 Ga lifein SW Gr<strong>ee</strong>nland. Journal <strong>of</strong> the Geological Soci<strong>et</strong>y <strong>of</strong> London 166, 335–348.Wright, D.T., 1997. An organogenic origin for widespread dolomite in the CambrianEilean Dubh Formation, north western Scotland. Journal <strong>of</strong> SedimentaryResearch 67, 54–64.Wright, D.T, 1999. The role <strong>of</strong> sulphate-reducing bacteria and cyanobacteria indolomite formation in distal ephemeral lakes <strong>of</strong> the Coorong region, South Australia.Sedimentary Geology 126, 147–157.Wright, D.T., 2000. Benthic microbial communities and dolomite formation inmarine and lacustrine environments – a new dolomite model. In Glenn, C.R.,Lucas, J., Prevot-Lucas, L. (Eds.), Marine Authig<strong>ene</strong>sis: From Global to Microbial.SEPM Special Publication 66, 7–14.Wright, D.T., Altermann, W., 2000. Micr<strong>of</strong>acies development in late Archaeanstromatolites and ooids <strong>of</strong> the Ghaap Group, Republic <strong>of</strong> South Africa. InInsalaco, E., Skelton, P.W., Palmer, T.J. (Eds.), Carbonate Platform Systems:Components and Interactions. Geological Soci<strong>et</strong>y Special Publication 178,51–70.Wright, D.T., Oren, A., 2005. Non-photosynth<strong>et</strong>ic bacteria and the formation <strong>of</strong> carbonatesand evaporites through time. Geomicrobiology Journal 22, 27–53.Wright, D.T., Wacey, D., 2004. Sedimentary dolomite – a reality check. In: Braithwaite,C.J.R., Rizzi, G., Darke, G. (Eds.), The Geom<strong>et</strong>ry and P<strong>et</strong>rog<strong>ene</strong>sis <strong>of</strong>Dolomite Hydrocarbon Reservoirs. Geological Soci<strong>et</strong>y Special Publication, 235,pp. 65–74.Wright, D.T., Wacey, D., 2005. Precipitation <strong>of</strong> dolomite using sulphate-reducing bacteriafrom the Coorong Region, South Australia: significance and implications.Sedimentology 52, 987–1008.Yamamoto, K., Itoh, N., Takuya, M., Tanaka, T., Adachi, M., 2004. Geochemistry <strong>of</strong>Precambrian carbonate intercalated in pillows and its host basalt: implicationsfor the REE composition <strong>of</strong> circa 3.4 Ga seawater. Precambrian Research 135,331–344.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!