16
182 Microstructural Controls on Monosulfide Weathering and Heavy Metal Release (MIMOS) 1. Introduction and summary The iron monosulfides pyrrhotite (Fe 1-x S) and mackinawite (tetragonal FeS) are both environ- mentally important phases linking the iron and sulfide elemental cycle. This report summarizes studies that aim at understanding on the effect of microstructure, mineral chemistry and cry- stallography of these minerals on the reactivity (Pollok et al., 2008) and provides insights to- wards possible applications. Pyrrhotite comes a close second after pyrite among the most abundant iron sulfides in cru- stal rocks and ore deposits including Ni-Cu, Pb- Zn, Au, and platinum-group elements (PGE). The weathering of pyrrhotites contributes to the release of iron and other metals as well as sulfate to surface waters as a part of the acid mine drainage problem. The relevance of this reaction lies in the high reactivity of pyrrhotite (20 to 100 times faster compared to the disul- fide pyrite; Nicholson & Scharer, 1994 ) which limits the neutralization of produced sulfidic acid by silicate weathering reactions. Further- more, many potentially important implications to the fields of geomagnetics, petrology, envi- ronmental mineralogy, and technical mineral processing are to be expected if detailed know- ledge about structures and phase relations is available. Lately, the flotation behavior of pyr- rhotite is attracting more attention as this mine- ral is an important host for Ni and PGEs (Becker et al., 2010; Ekmekçi et al., 2010). In the first section the coupling between cry- stallographic structure and composition of pyr- rhotite has been studied by the use of trans- mission electron microscopy (TEM) which has been used for the first time to directly visualize details of the (non-integral) stacking sequences using superstructure diffraction spots. The observations result in a new and versatile des- cription of pyrrhotite superstructures which provide a basis to understand variation in phy- sical properties and reactivity of pyrrhotites. In the second section the relevance of different superstructures in terms of reactivity is directly substantiated for the biologically-enhanced dis- solution of a pyrrhotite intergrowth. The diffe- rence in dissolution rates betwee the inter- grown phases is quantified by using topogra- phic data produced by confocal microscopy pro- viding a new view on the reactivity and weathe- ring patterns of this mineral. The third part deals with the evolution of surface roughness for such reacted and heterogeneous surfaces. To extend our possibilities to generate meaningful data for reactive surface area and, hence, dissolution rates, a new program is introduced, which allows a non-biased statistical evaluation of sur- face roughness distributions. The concept pres- ented here can be utilized for any kind of topo- graphic data and might be well applicable to other areas of surface science. Pollok K. (1)*, Harries D. (1), Hopf J. (1,2), Etzel K. (1), Chust T. (1), Hochella M.F., Jr. (3), Hellige K. (4), Peiffer S. (4), Langenhorst F. (1) (1) Universität Bayreuth, Bayerisches Geoinstitut, D-95440 Bayreuth, [email protected] (2) Friedrich-Schiller-Universität Jena, Institut für Mikrobiologie, Neugasse 25, D-07743 Jena (3) Center for NanoBioEarth, Department of Geosciences, Virginia Polytechnic Institute and State University, Blacksburg, Virginia, USA (4) Universität Bayreuth, Lehrstuhl für Hydrologie, D-95440 Bayreuth * Coordinator of the project

Microstructural Controls on Monosulfide Weathering and Heavy

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Microstructural Controls on Monosulfide Weathering and Heavy

182

Microstructural Controls on MonosulfideWeathering and Heavy Metal Release (MIMOS)

1. Introduction and summaryThe iron monosulfides pyrrhotite (Fe1-xS) andmackinawite (tetragonal FeS) are both environ-mentally important phases linking the iron andsulfide elemental cycle. This report summarizesstudies that aim at understanding on the effectof microstructure, mineral chemistry and cry-stallography of these minerals on the reactivity(Pollok et al., 2008) and provides insights to -wards possible applications.Pyrrhotite comes a close second after pyriteamong the most abundant iron sulfides in cru-stal rocks and ore deposits including Ni-Cu, Pb-Zn, Au, and platinum-group elements (PGE).The weathering of pyrrhotites contributes tothe release of iron and other metals as well assulfate to surface waters as a part of the acidmine drainage problem. The relevance of thisreaction lies in the high reactivity of pyrrhotite(20 to 100 times faster compared to the disul-fide pyrite; Nicholson & Scharer, 1994 ) whichlimits the neu tralization of produced sulfidicacid by silicate weathering reactions. Further -more, many po ten tially important implicationsto the fields of geomagnetics, petrology, envi-ronmental mineralogy, and technical mineralprocessing are to be expected if detailed know -ledge about structures and phase relations isavailable. Lately, the flotation behavior of pyr-rhotite is attracting more at ten tion as this mine-ral is an important host for Ni and PGEs (Becker

et al., 2010; Ekmekçi et al., 2010).In the first section the coupling between cry-stallographic structure and composition of pyr-rhotite has been studied by the use of trans-mission electron microscopy (TEM) which hasbeen used for the first time to directly visualizedetails of the (non-integral) stacking sequencesusing superstructure diffraction spots. Theobservations result in a new and versatile des-cription of pyrrhotite superstructures whichprovide a basis to understand variation in phy-sical properties and reactivity of pyrrhotites. Inthe second section the relevance of differentsuperstructures in terms of reactivity is directlysubstantiated for the biologically-enhanced dis-solution of a pyrrhotite intergrowth. The diffe-rence in dissolution rates betwee the inter-grown phases is quantified by using topogra-phic data produced by confocal microscopy pro-viding a new view on the reactivity and weathe-ring patterns of this mineral. The third part dealswith the evolution of surface roughness for suchreacted and heterogeneous surfaces. To extendour possibilities to generate meaningful data forreactive surface area and, hence, dissolutionrates, a new program is introduced, whichallows a non-biased statistical evaluation of sur-face roughness distributions. The concept pres-ented here can be utilized for any kind of topo-graphic data and might be well applicable toother areas of surface science.

Pollok K. (1)*, Harries D. (1), Hopf J. (1,2), Etzel K. (1), Chust T. (1), Hochella M.F., Jr. (3), Hellige K. (4),

Peiffer S. (4), Langenhorst F. (1)

(1) Universität Bayreuth, Bayerisches Geoinstitut, D-95440 Bayreuth, [email protected]

(2) Friedrich-Schiller-Universität Jena, Institut für Mikrobiologie, Neugasse 25, D-07743 Jena

(3) Center for NanoBioEarth, Department of Geosciences, Virginia Polytechnic Institute and State University,

Blacksburg, Virginia, USA

(4) Universität Bayreuth, Lehrstuhl für Hydrologie, D-95440 Bayreuth

* Coordinator of the project

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 182

Page 2: Microstructural Controls on Monosulfide Weathering and Heavy

183

Mackinawite is a metastable, nanoparticularphase that forms by precipitation from aqu e -ous solutions at low temperatures. It is a majorcomponent of acid volatile sulfide (AVS) com-pounds, causing black-colored layers in recentanoxic sediments. Mackinawite is therefore animportant precursor of sedimentary pyrite andcan facilitate in-situ remediation of metals inconnection with (biogenic) sulfate reduction inacidic pit lakes and sediments (Bilek, 2006,Church et al. 2007). It is known as an effectiveabsorber due to its large specific surface area(Ohfuji & Rickard, 2006) but detailed reactionmechanisms important to biogeochemistry arestill missing.The last section provides experimental eviden-ce for the formation mechanisms of sedimen-tary sulfides via mackinawite under anoxicconditions. The results highlight the importan-ce of iron hydroxide surfaces as the locationsof electron transfer for mackinawite forma-tion, but also show the spatial decoupling topyrite formation. In conclusion, the study pro-vides fundamental data that enables to modelbasic reactions at geochemical interfaces.

2. Towards a new structural model forNC-pyrrhotites: Structural and chemicalcharacterization of selected samples byelectron microscopyThe physical properties of a mineral and thereactivity of a mineral surface are determinedby its chemical composition and internal ato-mic structure. In this respect, non-stoichiome-tric pyrrhotite is a striking phase because itsstructure and the physical (e.g. magnetic) pro-perties change dramatically within a small com -positional range. A detailed knowledge aboutthe composition-structure relationship as wellas the structure-property relationship is nee-ded to gain a fundamental understanding ofthis environmentally and technically impor-tant mineral. Pyrrhotite has a NiAs-derivative structure thatcontains sequences of hexagonal close-packedS atom layers stacked in alternation with layersof Fe atoms in octahedral coordination. Non-stoichiometric compositions arise from the pre-

sence of vacancies within Fe layers and spanthe range in-between ~47 to 50 at% Fe (Fe7S8

to FeS), whereas most structural complexityoccurs in the narrow compositional range bet-ween 47 to 48 at% Fe. At temperatures below300 °C various superstructures are observed inthis range due to vacancy ordering among andwithin Fe layers, resulting in a rather complica-ted subsolidus phase diagram (e.g., Nakazawa& Morimoto, 1970). Based on a NiAs-type structure of FeS, the wellunderstood monoclinic 4C structure is derivedby removing 1/4 of Fe atoms from every se condFe layer, leading to the composition Fe7S8. Fourvacancy layers are mutually discernible in thestacked arrangement and the resulting sequen-ce of partially vacant (V) and filled (F) Fe lay-ers is VAFVBFVCFVDF. The spacing between Vlayers of the same type is strictly four timesthe c-axis repeat of the NiAs substructure (oreight Fe layers). Structural descriptions of integral NC-pyrrhoti-te structures can be roughly divided into twoapproaches using either discrete (0 or 1) orpartial Fe site occupancies. The former relatesto an idealized extension of the 4C stackingscheme by introducing additional F layers,which increases both the periodicities of thesuperstructures as well as the Fe/S ratios, whilethe latter proposes in addition to extended lay -er stacking a considerable atomic scale dis or -der of Fe vacancies. Yamamoto & Nakazawa (1982) developed asuperstructure description for an Nc = 5.54pyrrhotite based on a continuous modulationof (partial) Fe site occupancies. In this model ofan aperiodic crystal structure the probability offinding a Fe atom at a given lattice site is go -verned by a continuous, sinusoid-like modula-tion function embedded in a four-dimensionalsu perspace formalism. Aperiodicity resultsfrom the occupation modulation being incom-mensurately related to the substructure peri-odicity. Based on this work Izaola et al. (2007)introduced a generalized superspace approachby replacing the continuous occupancy modu-lation with a step-like function governing dis-crete site occupancies. By virtue of a »close-ness« condition, the model strictly obeys the

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 183

Page 3: Microstructural Controls on Monosulfide Weathering and Heavy

184

sequence of ABCD layering as well as the rulethat two V layers must not follow consecutive-ly without an interjacent F layer, resulting inthe fixed relation 1/Nc = 2x between super-structure Nc value and composition expressedas x in Fe1-xS.A primary motivation for this study was thatthe reported electron diffraction pattern ofnon-integral NC-pyrrhotites strongly resemblethose seen in certain translation interfacemodulated binary alloys such as Au3+xZn1-x

(Amelinckx & Van Dyck, 1992). In these alloysystems integral superstructures arise in stoi-chiometric compounds as Au3Zn from orderingatoms on certain sites of the fcc lattice. Non-stoichiometric compositions and further super-structures are related to variably ordered, non-conservative (»composition-changing«) anti-pha se boundaries (APBs) of these simple su -perstructures. Such translation interface mo -du lation (TIM) is characterized by replacement(or splitting) of the simple superstructure dif-fractions spots (e.g., those of 4C-pyrrhotite)with arrays of satellite reflections – as seen inNC-pyrrhotites.

2.1. Samples and experimental methodsIn this study, two natural pyrrhotite samplesfrom geologically different locations were cha-racterized in detail by electron microscopy. Sam -ples were obtained from the Sta. Eulalia mi -ning district in Chihuahua, Mexico (EUL) andfrom the Nyseter mining area near Grua, Nor -way (NYS). Ar-ion thinning for transmission elec -

tron microscopy (TEM) was carried out usingliquid nitrogen cooling or a low angle thinningsystem. In order to check for preparation arte-facts, powdered and ultramicrotomed sampleswere prepared as well, but no systematic diffe-rences in SAED pattern could be detected. Nc

values were obtained from selected area elec-tron diffraction (SAED) patterns with typicaluncertainties of ±0.01 (2s). TEM superstructu-re dark-field imaging (SDF) was accomplishedby placing an objective aperture of 2.3 nm-1

diameter on strong satellite reflections.

2.2. Results - Electron microprobe analysis(EMPA)Back scatter electron (BSE) imaging of EUL andNYS shows abundant exsolution lamellae inboth samples. In sections the lamellae formtwo sets of lens shaped or sigmoidal bodiesinclined towards the [001] direction of hostpyrrhotite. Lamellae typically measure <200µm in length and up to 30 µm in width and aredarker in BSE images (Fig. 1a). Quantitativeanalyses show the lamellae to have composi-tions near Fe0.875S (or Fe7S8) indicative of 4C-pyrrhotite, while the host pyrrhotites’ Fe con-tents are slightly below Fe0.900S (or Fe9S10) asindicated by x values of 0.106±0.001 (EUL)and 0.102±0.001 (NYS). Contrary to the EULsample NYS shows a small Ni content of 0.05to 0.11 wt%. Exsolution of 4C-pyrrhotite fromNC-pyrrhotite appears to be quite abundant innature and is well documented by petrogra-phic methods.

Figure 1: (a) BSE image of 4C/NC intergrowths in sample EUL. (b) TEM-SDF image of the 4C/NC interface in EUL. APBsterminate in eight-fold node structures. (c) TEM-SDF image of the 4C/NC interface in sample NYS. Nearly identical nodestructures occur at the boundary but additionally changes in APB configurations (arrow) in the NC phase can be seen

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 184

Page 4: Microstructural Controls on Monosulfide Weathering and Heavy

185

2.3. Results - TEM-SAEDThe indexing of superstructure satellite reflec-tions is done by applying a (3+1)-dimensional(hklm) scheme analogous to the one adoptedby Yamamoto & Nakazawa (1982) and com-monly used for the description of one-dimen-sionally modulated structures (Fig. 2). A dif-fraction vector g is thus:

g = h a* + k b* + l c* + m q

m is the satellite order of a superstructure ref -lection and q is the modulation vector relatingeach satellite to a (hkl) substructure reflection.q is generally parallel or subparallel to c*. TheNc value of a superstructure is obtained as thelength ratio c*/q with c* referring to the NiAs-type subcell (c* ≈ 1.72 nm-1). For the EUL andNYS samples coincident Nc values of 4.81-4.87and 4.78-4.96, respectively, have been obtai-ned. In general, the Nc values compare well tothe x values obtained from EMPA assuming therelation 1/Nc = 2x. Based on the hypothesis of an APB modula-tion, the splitting of reflections in SAED pat-terns obtained from zone axes perpendicularto c allows to calculate average spacings anddisplacement vectors of APBs (Amelinckx &Van Dyck, 1992). The geometrical evaluationof pattern obtained from EUL and NYS suggestspacings of 1.58 to 1.74 nm and a displace-ment vector R = 1/8[001]4H (the 4H indexingrefers to a hexagonal cell in which the NiAs aand c dimensions are doubled and quadrupled,respectively).

2.4. Results - TEM-SDFIn order to obtain TEM-SDF images from elec-tron beams diffracted by the superstructure, itwas found to be most effective to place theobjective aperture on the (0004) and (008-4)satellite reflections accessed from near the[210]4H zone axis. The sample foil was tiltedsuch that in the ideal case only diffracted beamsin the (00lm) reciprocal lattice row were excited.In the NC portions of both the EUL and NYSsamples, SDF images reveal dense arrays ofdark stripes being aligned parallel to the (001)planes and solely visible in dark-field images

using strong superstructure reflections. In EULthe stripes appear more ordered and show lesswiggles and irregularities compared to NYS,which displays some weak diffuse diffraction inthe area from which the images were obtai-ned. Further, in NYS the configuration of stri-pes appears to change along certain domainboundaries where the thin stripes, as seen inEUL, change into thicker stripes. The latterdominate the NYS sample and are likely dou-blets of two stripes being too close to be resol-ved under the SDF imaging conditions. Theaverage spacing of stripes in EUL is 1.63±0.05nm and the observed (half-)width of a singlestripe is approx. 0.80 nm. Assuming the thickstripes in NYS to be non-resolvable doublets,the average spacing between individual stripesis 1.94±0.10 nm. The average distance for EUL compares verywell with the expected APB spacing as obtai-ned from the SAED pattern geometries and wetherefore interpret the dark stripes to be APBs.Based on contrast in dark-field images we canexclude them to be stacking faults or twinboundaries, as the former would be seen withsubstructure reflections as well and the latterwould show intensity differences of domainson either site of the boundary. The larger ave-rage spacing in NYS, despite very similar Nc

values, is likely due to the higher disorder withlarger and more irregular spacings betweenstripes which bias the mean distance to highervalues. The average of 1.94 nm corresponds toan Nc value of approx. 4.7, which is compati-ble with diffuse satellite reflections seen insome SAED patterns. Most striking features in SDF images of EULand NYS are the phase boundaries between4C- and NC-pyrrhotite where the stripes formcomplicated but remarkably similar node struc-tures in which the stripes terminate. When thephase interface is oriented mostly parallel tothe stripes, no nodes occur directly at the inter-face, but occasionally appear nearby withinthe NC phase, often associated with crystallo-graphic edge dislocations having the Burgersvector b = [001] of the NiAs cell. In all cases ofnode structures we observed eight stripes toterminate in a single node.

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 185

Page 5: Microstructural Controls on Monosulfide Weathering and Heavy

186

The behavior of stripes in the observed nodestructures and adjacent 4C-pyrrhotite stronglyendorses our interpretation of them beingnon-conservative APBs of the 4C structure.Their orientation is fully compatible with theinferred displacement vector R = 1/8[001]4H

which creates double layers of completely filledFe positions as opposed to single layers beingpresent in the 4C structure. The eight-fold no -des facilitate the reconstruction of the 4C pha-ses from the APB modulated NC phase anddirectly support the proposed displacementvector 1/8[001]4H as its eight-fold multiple willbe the c dimension of the hexagonal 4C unitcell (see discussion section). Apparently thehighly similar node configurations represent aself-organization towards an energeticallyoptimal phase interface.

2.5. DiscussionOur TEM observations provide compelling evi-dence that the structural diversity of non-inte-gral NC-pyrrhotites can be explained in termsof a TIM structure model. The translation inter-faces are APBs based on the Fe sublattice of 4C-pyrrhotite. This model bears resemblance to pre -viously suggested »out-of-step« and anti-phasedefect models (Pierce & Buseck, 1974) but ismore precise on the actual nature of the invol-ved translation interfaces. On the other hand,the superspace approach of Izaola et al. (2007)is fully compatible with our TIM model. In es -sence, this superspace model produces discre-tely modulated superstructures by insertingadditional F layers into the basic 4C layer se -

quence, producing either periodic or aperiodic(but uniform) stacking sequences parallel to(001). Because the so produced double F layerscarry an Fe excess, the resulting phases willhave higher Fe/S ratios than 4C-pyrrhotite andfollow strictly the relation 1/Nc = 2x as spacingsbetween double layers are evenly distributed. For a structure with Nc = 4.8, which is a simplerepresentation for the host NC-pyrrhotite inour EUL and NYS samples, it is possible to deri-ve two well ordered stacking schemes:(DSSDS)4 and (DDSSS)4, where D representsdouble F layers and S single F layers, intercala-ted with the four vacancy bearing layers. The(DSSDS)4 sequence can be derived via thesuperspace model of Izaola et al. (2007) and isuniform, while the (DDSSS)4 sequence cannotbe obtained from the model and has lessequally spaced D layers, leading to slightly lessorder while maintaining the same averagevacancy-vacancy distance. The observed D layer or APB arrangements inEUL are alike the well ordered (DSSDS)4 se -quence, while in NYS domains relating to bothstacking schemes coexist as indicated by thechanges in stripe configuration. Here the do -mains containing the thicker, non-resolvablestripes may correspond to the (DDSSS)4 stak-king in which the closest spacing between Dlayers or APBs is approx. 0.86 nm and thusmore or less equal to the half-width of a singlestripe as observed in EUL. The average ~1.72nm D layer spacing of the idealized (DSSDS)4sequence compares well with the average stri-pe and APB spacings in EUL as obtained from

Figure 2: Diffraction patterns of non-integral NC-pyrrhotite. (a) Schematic representation of zone axis [210]for Nc = 4.85. Grey open circles/indices represent 4C reflections, crosses mark the (generally absent) 4Creflections used for superstructure indexing. (b) Experimental SAED pattern of NYS with Nc = 4.84±0.01

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 186

Page 6: Microstructural Controls on Monosulfide Weathering and Heavy

187

SDF images and SAED patterns, respectively,and consistently reproduces the observedslightly non-equal spacings in the bulk NCphase. It appears reasonable to assume that incase of NYS the energetic difference betweenboth arrangements of D layers in the (DSSDS)4and (DDSSS)4 stacking variants are very smalland either the structure is observed on a tran-sition to a more ordered state as in EUL or theelevated Ni content exerts some control onvacancy arrangements. So far the discussion focused on an idealizedlayer stacking involving parallel layer arrange-ments. SDF images clearly shows that this idealsituation is virtually never fulfilled in the stu-died samples as stripes and therefore APBs areslightly wavy, corrugated and often non-paral-lel. Changes in stacking schemes at given Ncvalues as well as disorder introduced by bentAPBs will not produce strong effects in SAEDpatterns. For the (DSSDS)4 and (DDSSS)4 stak-king schemes the diffraction pattern geome-tries will be the same with only moderate dif-ferences in spot intensities, likely being hiddenby dynamical diffraction effects. Increasingmesoscale disorder of APBs will first lead tovanishing higher order satellite reflections andgradually lead to streaky diffraction along thec* direction when a quasi-periodicity is no lon-ger maintained. While a D layer in the ideallayer model is always an APB of the Fe sublat-tice the converse is not strictly true as isolatedAPBs are capable of moving more or less free-

ly through the crystal and can have even in theconfinement of a modulated structure a muchmore dynamical behavior as represented by astatic layer model. Therefore, describing thelayer stacking as a TIM structure of denselyarranged APBs offers much better prospects ofunderstanding the behavior and properties ofnon-integral NC-pyrrhotites. The need of a more dynamical model clearlybecomes apparent when trying to understandthe node structures observed at NC/4C phaseinterfaces. The termination of eight stripesequivalent to eight APBs in a single node isnecessary to reconstruct the 4C structure froman NC structure with well ordered V layer stak-king. Because the stacking of V layers on theleft side strictly obeys the ABCD sequence, thetermination creates a sequence fault where twoVF layers follow on each other. Conse quent ly,two APBs have to terminate either at a perfectdislocation with b = [001]NiAs = 1/4[001]4H, whichremoves a pair of VF layers, or 4 pairs of APBshave to terminate close-by such that the eight1/8[001]4H displacement vectors add up to the4C lattice repeat. Both variants and combina-tions thereof are seen in our samples and wesuggest that the observed node configura-tions facilitate an energetically most favorablestate when there are no or few dislocationsinvolved. Based on the TIM model we suggest that thereis no significant difference between integraland non-integral pyrrhotites and both types

Figure 3: a) BSE image of the TYS sample. Bright lamellae contain 2 wt% more iron than the matrix. b) SEimage of the surface of the abiotically reacted pyrrhotite cubes after 40 days which shows almost no disso-lution. c) SE image of the pyrrhotite surface reacted with A. ferrooxidans after 40 days. Deep trenches for-med at the position of the stoichiometric FeS lamellae

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 187

Page 7: Microstructural Controls on Monosulfide Weathering and Heavy

188

can be understood as part of a continuum ofstructures in-between Fe7S8 and (at least)Fe11S12, in which variable Fe contents are go -ver ned by changes in densities and orderingstates of APBs. Based on the presented dataand preliminary data on other natural pyrrhoti-te samples we currently do not see evidencefor preferred stability of integral NC-pyrrhoti-tes in exsolution association with 4C-pyrrhoti-te, but the true phase relations may be obscu-red by slow diffusion kinetics below 220 °C,where NC-pyrrhotites first appear (Nakazawa& Morimoto, 1970). However, the very sharpcompositional gradients at the 4C/NC interfa-ces in our EUL and NYS samples point to wellequilibrated assemblages. Therefore, the asso-ciation of Nc ≈ 4.85 pyrrhotite with exsolved4C-pyrrhotite in two of our samples, as well asthe coexistence of Nc = 4.88 pyrrhotite and 4C- pyrrhotite observed by Morimoto et al. (1975),points to a favored stability of NC-pyrrhotiteswithin the range of Nc = 4.8 to 4.9 – at leastwith regards to typical temperatures of theuppermost crust, where the samples residedfor geological timescales. The insight into complexity in the bulk structu-re and the applied TIM model can lead to abetter understanding of physical propertiesand the reactivity of surfaces. Due to the rathertwo dimensional, layer-like structure of pyrrho-tites one clear expectation is a pronouncedanisotropy of surface properties. Preliminaryexperimental results indicate that reactionspeeds on basal (001) faces are considerablydifferent than on prismatic (hk0) surfaces.Owing to pyrrhotite’s excessively variable mag-netic and physicochemical properties andwidespread occurrence in rocks and ore depo-sits, many potentially important implications tothe fields of geomagnetics, petrology, environ-mental mineralogy, and technical mineral pro-cessing (e.g. pyrrhotite flotation) are to be ex pec -ted when the presented detailed characteriza-tion of structures and phase relations is applied.

3. Structural controls on the alteration ofpyrrhotite by Acidithiobacillus ferrooxi-dans under acidic conditonsBased on the detailed characterization proce-dures described above, a well-directed studyon the reactivity of pyrrhotite intergrowth inexperiments with acidophilic microorganismswas conducted. Acidophilic microorganismsplay a dominant role in acid mine drainage(AMD) and for the metal recovery from low-grade ores by bioleaching applications. A num-ber of studies focused on the reaction paths ofchemical and biological metal sulfide oxidation(Rawlings, 2002; Schippers, 2004). However,experimental studies on the biologically-en hanc -ed pyrrhotite oxidation are still scarce (seeBelzile et al. 2004 for review) or focus on bio-leaching applications of low grade ores (Ke &Li, 2006; Jiang et al., 2007), but clearly indica-te the catalytic effect of microorgansmswithout reporting a reliable characterization ofthe starting material and dissolution rates.Schippers et al. (2007) provides a potential pyr-rhotite oxidation rate measured in a mine tai-ling using microcalorimetry, which suggeststhat natural rates can exceed laboratory ratesby up to 4 orders of magnitude. Beside theenhancing effect of microbial consortia, pyr-rhotite dissolution may be also influenced bythe reactivity and intergrowth texture of diffe-rent superstructures. Janzen et al. (2000) stu-died a number of pyrrhotite samples from diffe-rent localities and determined the relative pro -portions of so-called monoclinic (mainly 4C) andhexagonal (NC) pyrrhotite. No systematic trendin dissolution rates were found, however, infor-mation about the intergrowth texture is missingfor these samples.Acidithiobacillus ferrooxidans, an acidophilicproteobacteria that catalyses the oxidation offerrous iron, elemental sulfur and reduced sul-fur compounds is by far the best characterizedbioleaching microorganism and has been usedin the presented study. Cells of A. ferrooxidansproduce an exopolysaccharide (EPS) layer toattach to the sulfide mineral surface (Gehrke etal., 1998) and this attachment is viewed as aprocess by which the bacterial membranecomponents interact directly with metal and

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 188

Page 8: Microstructural Controls on Monosulfide Weathering and Heavy

189

sulfur of the mineral by an enzymatic type ofmechanism (Sand et al. 1995).In this study X-ray photoelectron spectroscopy(XPS), scanning electron microscopy (SEM),and confocal microscopy have been used totrace the chemical and morphological changesat the surface of a well characterized pyrrhoti-te sample. Surface topography has employedto quantify a differential dissolution rate bet-ween intergrown pyrrhotites.

3.1. Samples and experimental methodsA pyrrhotite from Tysfjord region, Norway(TYS) was used in this study. The TYS sampleshows a dense population of 1-3 µm wideexsolution lamellae with a common orientationparallel to (001). They appear brighter in BSEimages compared to the matrix indicating ahigher Fe/S ratio (Fig. 3a). Quantitative analy-ses place them close to stoichiometric FeS (tro-ilite). The matrix pyrrhotite of the TYS samplehas a composition close to Fe0.900S (or Fe9S10)with x values of 0.096±0.001 and thus, thelamellae contain about 2 wt% more iron thenthe matrix.Polished cubes (3x3x3 mm) of TYS pyrrhotitewere dissolved in presence of Acidithiobacillusferrooxidans at pH 2 over 40 days with an a bi -otic control in parallel. A. ferrooxidans ATCC19859, obtained from the American Type Cul -ture Collection (ATCC, Manassas, VA), wasrou tinely cultivated as a pre-culture in ATCC

medium 2039 at 28°C for five days on a rota-ry shaker (300 rpm). For the inoculation of theexperiments 10 µl pre-culture (approx. 2×103

cells/ml) was directly used. Cultures for theexperiments were grown in 50 ml mineral saltsolution with 0.4 g each of (NH4)2SO4, K2HPO4,and MgSO4×7H2O per liter. Cubes were remo-ved from the batch experiments after 1, 4, 7,14, 17, 21, 28 and 40 days, respectively. XPS was performed on a PHI Quantera SXM-03 Scanning XPS Microprobe using an Al K X-ray source with a highly focused beam (<9microns). High resolution scans of the Fe 2p, S2p, and O 1s peaks were used to determinethe bond and valence state information. XPSreference binding energies for pyrrhotite provi-ded by Legrand et al. (2005) were used toassign and fit the spectra. The topography of selected surfaces was mea-sured by the confocal 3D microscopy systemµsurf custom (Nanofocus AG, Oberhausen). Itconsist of a monochromatic LED light source (= 505 nm), a helically-shaped arrangement ofpinholes (Nipkow-Disk), a stage with piezomodule for z movement and a megapixel CCDcamera. The objective used (numerical apertu-re 0.9) provides 160 µm x 160 µm field of viewwith a resolution of 10 nm in z and about 500nm in x/y.

3.2. ResultsA. ferrooxidans develops a visible biofilm on

Figure 4: Overlays of the Fe 2p3/2, O 1s and S 2p narrow region scans of pyrrhotite surfaces abiotic control (left) andbiotic (right) experiments. Labels denote the duration of the experiments

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 189

Page 9: Microstructural Controls on Monosulfide Weathering and Heavy

190

the pyrrhotite cubes after approximately 5 daysof incubation. After 9 days the whitish biofilmcovered the entire pyrrhotite cubes and grewslowly until the experiment was finally stoppedat 40 days. After removing the biofilm by awashing procedure the color has changed todark blue/black and the reflectivity from shinyto matt. In contrast, the control pyrrhotitecubes (abiotic) were still shiny golden in allexperiments with only minor changes in reflec-tivity. XPS analyses (Fig. 4) reveal a significant-ly more oxidized surface in the biotic experi-ment with a clear shift from Fe(II)-S to Fe(III)-Obonding. Compared to the control experiment,the surface is more hydroxylated and showseven traces of sulfate. SEM images (Fig. 3) ofthe pyrrhotite surface exposed to A. ferrooxi-dans reveal long and deep trenches which startto develop already after 4 days of experiment.The position of the trenches is identical to thetroilite lamellae indicating that the reactivity ofthe lamellae is significantly higher than the sur-rounding NC-pyrrhotite. Trenches reach depthsof about 2 µm and a width of 3 µm as deter-mined by confocal microscopy measurementsafter 28 days (Fig. 5). Height profiles suggestthat the lamellae are inclined to the dissolvingsurfaces which lead to the nonsymmetricalshape of trenches. Although information ofthe original surface height is missing, the 3Dtopographic data can provide a differential dis-solution rate of the 2C-pyrrhotite compared tothe NC-pyrrhotite by using the matrix surfaceas a reference height. Five topographic imagestaken from surfaces after 21 and 28 days resul-

ted in a differential dissolution rate of8.8±1.2×10-9 and 1.2±0.3×10-8 mol m-2s-1,respectively.

3.3. DiscussionThe dissolution experiments clearly show thecatalyzing effect of microorganisms at condi-tions that are unfavorable for chemical dissolu-tion. The XPS data indicates a fast and extensi-ve oxidation of the biologically altered surfacewith dominating of Fe(III)-O bonds as well assome sulfate which points to the formation ofa iron hydroxide layer (possibly FeOOH). How -ever, this does not result in a passivation of thesurface as indicated by a continuous deepe-ning of the 2C lamellae which possible can beexplained by the presence of EPS and a rege-neration of ferric iron by the bacteria. For thefirst time, the clear difference in reactivity bet-ween two pyrrhotite superstructure (2C andNC) has been shown by the presented experi-ments. A direct and an indirect mechanism forbiologically-enhanced dissolution have beenproposed in the literature. For the directmechanism, bacteria are attached to the surfa-ce surrounded by EPS which provides comple-xed ferric iron (Rodriguez-Leiva & Tributsch,1988). The indirect mechanism interprets dis-solution as an inorganic process where non-attached bacteria only maintain a certain ferriciron concentration which accelerates the disso-lution (Rodriguez et al., 2003). Whether the dif -ference in dissolution rate is facilitated by apreferred attachment of bacteria due to thedifferent iron content or are caused by diffe-

Figure 5: Surface of the TYS sample after biologically-enhanced dissolution measured for 28 days measuredby confocal microscopy. The height of the ma trix NC-pyrrhotite (~0.5 µm) was used as reference height todetermine the differential dissolution rate from thedissolved volume of the 2C-lamellae. The squares illu-strate the principle of roughness parameter calcula-tion as function of sampling size performed withROUGHNECK.The upper profile clearly shows the non-symmetricalshape of the trenches which is explained by the incli-ned intersection of the lamellae with the surface

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 190

Page 10: Microstructural Controls on Monosulfide Weathering and Heavy

191

rences in reactivity of troilite and NC-pyrrhoti-te alone is not yet been conclusively evaluatedand will be subject of further activity. In anycase, the lamellar structure of the trenches cannot be interpreted as a primary microbial bio-signatures because it is originally caused by theexsolution process during equilibration of thepyrrhotite sample and acts like a templateduring dissolution. In any case, the differential dissolution rate isalready higher than all reported abiotic dissolu-tion rates and highlights the significance of thestructural variability and intergrowth texture.When we combine the differential dissolutionrate with the only available biotic dissolutionrate for pyrrhotite under acidic conditions(8×10-9 mol m-2s-1) reported in Belzile et al.(2004), the resulting net dissolution rate is a -bout one order of magnitude higher than re -ported rates in the absence of additional ferrousFe. Further accelerating factors towards theobserved rates include the combined effect ofbacterial consortia (combinations of iron- andsulfur-oxidizer) as well as the increase of reactivesurface area by the deepening of the trenches.

4. From surface morphology to rates: Anautomated routine to evaluate convergedroughness parameters of heterogeneoussurfacesAn increase of surface roughness during biolo-gically-enhanced dissolution of pyrrhotitewhich, in turn, affects the dissolution rate wascaused by both the microbial activity and thesample heterogeneity. Two components contri-bute to the surface roughness evolution of theTYS sample, namely, the alteration of the NCmatrix and the deepening of trenches at the 2Clamellae. The high differential dissolution ratemay be at least partly explained by increasingsurface area. For the presented surfaces (Fig. 5)we are therefore interested in the evolution ofsurface roughness for the matrix, the lamellaeand the combined surface of the biologicallydissolved pyrrhotite surface, respectively. In geosciences, rates are usually determinedfrom dissolution studies (either flow throughor batch experiments) using a size fraction of

ground and sieved starting material by measu-ring the change in fluid chemistry as a functionof time. The surface area of the starting mate-rial is commonly determined by the BETmethod and used as a constant to normalizerates. Recently, topographic methods havebeen introduced to directly measure sulfidedissolution rates and surface roughness (Astaet al., 2008) and offer new insides into thereaction mechanisms and surface reactivity.Furthermore, it allows a systematic determina-tion of surface roughness parameters. As statistical quantities roughness parametersare strongly dependent on the field of view ofthe used technique or the sampling size. Theconcept of »converged roughness parameter«(Fischer & Lüttge, 2007) was shown to be veryuseful to characterize certain surface buildingblocks by employing recurrent (squared) bisec-tions of the measuring field with an edgelength of a. A surface parameter is defined ascon verged when a flat slope is found in theconvergence graph (roughness parameter ver-sus edge length a). Converged parametersintrinsically depend on the choice of »repre-sentative« areas for their analysis. Further more,artificial measuring points may lead to an ove-restimation of roughness parameters. Here, wepresent the newly developed program ROUGH -NECK, which calculates a number of surfaceroughness parameters from topographic data asfunction of sampling size allowing to analyzeand visualize roughness parameter distributions.

4.1. The program ROUGHNECKThe program ROUGHNECK utilizes the com-mon Surface Data Format (SDF) (Blunt & Stout,2001) as input and output file format and is astand-alone application that can be used onWindows and Linux systems. Our input filesproduced by confocal microscopy consist of984x984 points representing a 160x160 µmfield of view. For this study a quality criteria ofa minimum of 99% of measured points wasused for the topographic data. The residualnon-measured points were linearly interpola-ted. The following 3D roughness parametersare implemented so far: the roughness avera-ge Sa, the root-mean-square roughness Sq, the

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 191

Page 11: Microstructural Controls on Monosulfide Weathering and Heavy

192

peak-to-peak height Sz, the ten-point-heightS10z, the root-mean-square-gradient Sdq, thesurface-skewness Ssk, the surface-kurtosis Sku,and the surface area ratio F. The program usesfreely selectable edge lengths for the desiredsampling area, which is moved pixel by pixelover the entire field of view (cf. Fig. 5).

4.2. Evaluation of ROUGHNECK outputsThe surface area ratio, F=A3D/A2D, which des-cribes the surface area of a three dimensionalsurface normalized by the two dimensional(projected) surface, is used as an example para-meter to illustrate the statistical procedure. Aperfectly flat surface will result in F=1, while arough surface results in F>1. This parameter isa so-called hybrid surface parameter because it

does not mainly depend on the amplitude butis more sensitive to the high frequency modula-tions of the surface as it is measured by theslope between adjacent surface points. ROUGHNECK has been used to calculate theconverged F parameters of the biologically dis-solved pyrrhotite surface after 28 days (Fig. 5).The distribution of roughness parameters (herefrequency N vs. F) can be plotted as histogramby applying discrete bins (Fig. 6a). The maximaof the distributions show the roughness valuefor a significant portion of the measured surfa-ce. For heterogeneous surfaces, a number ofmaxima may indicate various components withdifferent surface roughness. In general, themaximum F value increases with increasingsampling area. At small scale (a=1 to 10 µm),

Figure 6: F value distributionsfor various squared samplingareas with a given edge length.With increasing sampling sizethe maximum of the curves isshifted to higher F values. How -ever, the transition from low va -lues around 1.05-1.1 to valuesaround 1.2 is clearly visible andused in the convergence graph(Fig. 7a). For some curves (e.g. a = 4 µm) up to 3 maxima canbe located indicating the hetero-geneity of surface components

Figure 7: a) Evolution of F as function of the edge length of the squared sampling area calculated for fresh and reactedpyrrhotite surfaces. F at larger sampling size (>10 µm) is controlled by the deepening of 2C lamellae. b) Spatial distribu-tion of F values for a squared sampling size with edge length a=4 µm. High F values (~1.6) can be found at the posi-tions of the dissolved 2C lamellae. Since these areas represent only a small portion of the entire field of view, they arenot detectable in the histogram. Very high F values originate from overlying particles (cf. Fig. 5), but they do not alterthe converged roughness parameter determined from the histogram

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 192

Page 12: Microstructural Controls on Monosulfide Weathering and Heavy

193

the first maxima represent only the roughnessof the NC pyrrhotite matrix which dissolves slo-wer than the 2C lamellae. At larger scale (a=15to100 µm) the F value includes contributionsfrom both because the spacing between thelamellae is smaller than the edge length of thesampling area. The maxima of each distribution can be trans-lated into a convergence graph plotting the Fvalue of the maxima versus the samplinglength a. Comparing predominant roughnessvalues as a function of time allows to trace theincrease in surface roughness of the NC matrixalone (a=1 to 10 µm) as well as the evolutionof roughness for the entire surface (a=15to100 µm). The output data for every samplingarea can also be visualized providing a spatialroughness parameter distribution (Fig. 6b).This is particularly useful if certain surface com-ponents, like the locations of the former 2Clamellae, represent only a small portion of theentire surface and are hence not well repres-ented in the histogram. From the roughnessdistribution image (a=4 µm) a roughness para-meter for the 2C lamellae of 1.6 can be deter-mined. In turn, this also validates that theroughness parameter distribution (Fig. 6a) isnot sensitive for artefacts like overlying parti-cles or inaccurately measured surface points.

5. The reaction pathway of secondaryiron sulfide formation under anoxic con-dition: An experimental study on thereaction of lepidocrocite with dissolvedsulfideAMD waters are typically characterized by highamounts of ferric iron and sulphate. The for-mation of secondary iron sulfides in the sub-surface is an important pathway to reduce thecontaminant concentration and to raise the pHof overlying waters (Bilek, 2006). In this res -pect, the interaction between dissolved sulfideand ferric oxides can be regarded as a key reac-tion leading ultimately to pyrite formation (Can -field et al., 1992, Poulton et al., 2004), butknow ledge on the pathways and on control-ling factors is still limited. Here, results frombatch experiments on the reaction betweenlepidocrocite and dissolved sulfide underanoxic conditions are presented with a specialemphasis on the characterization of nanocry-stalline products by TEM forming at differenttime steps.

5.1. Materials, experimental setup andmethodsKinetic batch experiments were conducted bymixing lepidocrocite (25 mmol/L) and differentconcentrations of dissolved sulfide at roomtemperature and a constant pH of 7 within ananoxic glove box. After 1-2 hours, 24 hours,

Figure 8: Evolution of elemental sulfur (left) and acid extractable ferrous iron (right) during two weeks of reaction forselected experimental runs. The initial sulfide concentration in run 10 and 14 was twice as the concentration in run 15.Mackinawite formation is rapid within the first hours of the experiments, whereas pyrite formation is relatively slow,starts after 2 days and continued to the end of the run

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 193

Page 13: Microstructural Controls on Monosulfide Weathering and Heavy

194

1 week and 2 weeks, respectively, dissolvedFe(II) and S(-II), Fe(II) extractable with 0.5 NHCl, S°, and the total iron content were deter-mined by wet chemistry extraction. At diffe-rent time steps aliquots of the reacting suspen-sion were analyzed by TEM, XRD and Möss -bauer spectroscopy. In order to limit oxidationin air during TEM sample preparation the sus -pension was first sampled in gas-tight vials. Adrop of solution was then taken with a syringeand put onto a Lacey carbon-coated coppergrid. The grid was immediately transferred tothe TEM holder and inserted into the highvacuum of the TEM. The short exposure of thesample to air can be limited to 1-2 minutes atmaximum with this procedure.

5.2. ResultsThe reaction can be divided into three phaseson the basis of wet chemistry (Fig. 8) and TEMobservations (Figs. 9 and 10). In the first minu-tes of the reaction (0-15min) dissolved sulfideis rapidly consumed. It is partly oxidized to S°species in solution and partly consumed by ini -tial mackinawite formation. In a second pha se(15-120 min) a continuous but slower forma-tion of S° is observed while acid extractableFe(II) remained constant. TEM measurementsre vealed the occurrence of a mackinawite rimcovering the lepidocrocite crystals that wasseparate from the lepidocrocite surface by aninterfacial magnetite layer (Fig. 9). The magne-

tite layer can be seen as an intermediate stagelinking two reactions, the formation of macki-nawite which reduces ferric iron at the lepido-crocite surface on one hand and the transportof electrons in the deeper regions of the lepi-docrocite bulk crystal which facilitates reducti-ve dissolution and mackinawite rim growthand the other hand. The third phase is charac -terized by a decrease of acid extractable Fe(II)and S° and results in pyrite formation togetherwith some magnetite (2-14 days). TEM measu-rements indicate that mackinawite completelydissolves from the surface of lepidocrocitewhile the precipitation of pyrite occurs separa-tely from the lepidocrocite surface. The pyriteformation is coupled to mackinawite dissolu-tion and probably accompanied by the forma-tion of FeS clusters that serve as precursors.The conversion of mackinawite to pyrite wastraced by TEM at 72 hours, 1 and 2 weeks. Themackinawite rims get gradually thinner andmore corrugated with time. After 48 hoursamorphous regions composed of iron and sul-fur in variable concentration can be found bet-ween lepidocrocite grains. Pyrite grains reachdiameters between 200 and 500 nm after 2weeks and commonly show striking cubic oroc tahedral outlines (Fig. 10). However, thegrains are composed of smaller cubic buildingblocks which point to a formation by orientedaggregation. In addition to pyrite, some mag-netite grains similar in size and shape could be

Figure 9: Bright field (a) and high resolution (b,c) TEM images of lepidocrocite crystals with sulfur-rich rims after 2 hoursof reaction. The spotted contrast on lepidocrocite grains in (a) is due to nanocrystalline mackinawite. In (b) characteri-stic (001) and (111) lattice fringes of mackinawite (0.52 and 0.23nm, respectively) are visible in the outer rim. A conti-nuous intermediate layer (arrow) shows fringes matching d220 of magnetite/maghemite (arrow). This layer can also benoticed in (a). A nanocrystal of mackinawite in [010] zone axis orientation is shown in (c) together with its calculatedFFT and a simulated diffraction pattern as inset

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 194

Page 14: Microstructural Controls on Monosulfide Weathering and Heavy

195

identified by STEM-EDX and electron diffrac-tion as well.

5.3. DiscussionThe temporal evolution of the chemical speciesconcentrations and solid phases indicate thatthe reaction progress is highly dynamic. TEMdemonstrates that several phases form duringthe reaction towards pyrite. On the basis of theobserved concentrations a number of balancedchemical equations can be formulated invol-ving sorbed iron species as well as polysulfites.However, it is obvious that the system is kineti-cally controlled towards chemical equilibration(pyrite formation) due to the slow reductivedissolution process and the redox disproportio-nation. This study has provided novel insightsinto the redox processes occurring at the sur-face and in the vicinity of the lepidocrocite cry-stals and reveals the occurrence of magnetiteand mackinawite in the initial phase of thereaction covering the lepidocrocite surface. Itfurther demonstrates that dissolution of mak-kinawite occurs prior to formation of pyritepresumably being followed by further reactionwith polysulfides in the absence of dissolvedsulfide. It has also demonstrated that the kine-tics of precursor formation from reductive dis-solution of ferric (hydr)oxides by hydrogen sul-fide and of subsequent pyrite formation arekinetically decoupled. This observation impliesthat in iron-rich aqueous systems of periodicalsulfide formation (e.g. movement of the capil-lary fringe in ground waters) ferric oxides may

undergo »charging« with reduced substancesthat may lead to pyrite (and magnetite) forma-tion even in the absence of dissolved sulfide.This reactions sequence has also implicationsfor the retention of toxic metal by adsorptionon nanocrystalline mackinawite which may getmobilized by further reaction to pyrite.

6. ConclusionThe most important scientific insights obtainedwithin the MIMOS project are:(i) A new structural model (TIM model) has

be en successfully applied to non-sto i chi o -me tric NC-pyrrhotite. It well explains im -portant structural details that are observedin electron diffraction patterns and dark-field images and can lead to a better under-standing of surface and bulk physical pro-perties of pyrrhotite relevant to geomagne-tics, petrology, environmental mineralogy,and technical mineral processing.

(ii) Differences in reactivity caused by small stru -ctural and chemical variations have be endirectly proven for biologically-enhanced dis -solution of pyrrhotite. The intergrowth tex-ture has therefore a direct influence on thedissolution rate by a continuous increase ofreactive surface area. This result is relevantfor the assessment of the weathering ratesof pyrrhotite tailings and bioleaching appli-cations.

(iii) The evolution of surface area was quanti-fied by combining topographic data and a

Figure 10: High resolution images, electron diffraction pattern and EDX spectra of pyrite. Note the aggre-gative nature of the grain consisting of cubic building blocks. Slight misorientations are also reflected inthe diffraction pattern

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 195

Page 15: Microstructural Controls on Monosulfide Weathering and Heavy

196

statistical evaluation of surface roughnessusing the newly developed programROUGHNECK for heterogeneous surfaces.This approach provides new prospects formineral dissolution studies as well as formore general topics dealing with surfacealteration like corrosion.

(iv) Analytical TEM was successfully applied totra cing the reaction pathway of iron sulfi-de for mation. The new results are relevantfor our understanding of processes andelement distributions at geochemicalboundaries which can help to deciphersuccessful remediation strategies for pol-luted waters and soils.

ReferencesAmelinckx, S. & Van Dyck, D. (1992) Electrondiffraction effects due to modulated structu-res. Electron diffraction techniques, IUCr/Ox -ford Science Publications, 2, 309-373.

Asta, M.P., Cama, J., Soler, J.M., Arvidson, R.S.& Lüttge, A. (2008) Interferometric study ofpyrite surface reactivity in acidic conditions.Am. Mineral., 93, 508-519.

Becker, M., de Villiers, J. & Bradshaw, D. (2010)The flotation of magnetic and non-magneticpyrrhotite from selected nickel ore deposits.Miner. Eng., in press, doi:10.1016/j.mineng.2010.07.002

Belzile, N., Chen Y., Cai, M. & Li, Y. (2004) Are view on pyrrhotite oxidation. Journal of Geo -chemical Exploration, 84, 65-76.

Blunt, L. & Stout, K.J. (2001) EC Project FinalReport »Development of a Basis for 3D SurfaceRoughness Standards« SMT4-CT98-2209.

Bilek, F. (2006) Column tests to enhance suphi-de precipitation with liquid organic electron do -nators to remediate AMD-influenced ground -water. Environ. Geol., 49, 674-683.

Canfield D. E., Raiswell R. & Bottrell S. (1992)The reactivity of sedimentary iron minerals to -

ward sulphide. Am. J. Sci., 292, 659-683.

Church, C.D., Wilkin, R.T., Alpers, C.A., Rye, R.O. & McCleskey, R.B. (2007) Microbial sulfatereduction and metal attenuation in pH 4 acidmine water. Geochem. Transac., doi:10.1186/1467-4866-8-10.

Ekmekçi, Z., Becker, M., Bagci Tekes, E. & Brag -shaw, D. (2010) An impedance study of theadsorption of CuSO4 and SIBX on pyrrhotitesamples of different provenances. Miner. Eng.,in press, doi:10.1016/j.mineng.2010.02.007

Fischer, C. & Lüttge, A. (2007) Converged sur-face roughness parameters – A new tool toquantify rock surface morphology and reactivi-ty alteration. Am. J. Sci. 2007, 307, 955-973.Gehrke, T., Telegdi, J., Thierry, D. & Sand, W.(1998) Importance of extracellular polymericsubstances from Thiobacillus ferrooxidans forbioleaching. Appl. Environ. Microbiol., 64,2743 – 2747.

Izaola, Z., Gonzalez, S., Elcoro, L., Perez-Mato,J.M., Madariaga, G. & Garcia, A. (2007) Re vi -sion of pyrrhotite structures within a commonsuperspace model. Acta Cryst. B, 63, 693-702.

Janzen, M.P., Nicholson, R.V. & Scharer, J.M.(2000) Pyrrhotite reaction kinetics: reaction ra -tes for oxidation by oxygen, ferric iron, and fornonoxidative dissolution. Geochim. Cos mo chim.Acta 64, 1511 –1522.

Jiang, L., Zhou, H.Y. & Peng, X.T. (2007) Bio-oxidation of pyrite, chalcopyrite and pyrrhotiteby Acidithiobacillus ferrooxidans. Chin. Sci.Bull., 52, 2702-2714.

Ke, J. & Li, H. (2006) Bacterial leaching of nik-kel-bearing pyrrhotite. Hydrometallurgy, 82,172-175.

Legrand, D.L., Bancroft, G.M. & Nesbitt, H.W.(1995) Oxidation/alteration of pentlandite andpyrrhotite surfaces at pH 9.3: Part 1. Assign -ment of XPS spectra and chemical trends. Am.Min., 90, 1042-1054.

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 196

Page 16: Microstructural Controls on Monosulfide Weathering and Heavy

197

Morimoto, N., Gyobu, A., Tsukuma, K. &Koto, K. (1975) Superstructure and nonstoi-chiometry of intermediate pyrrhotite. Am.Min., 60, 240-248.

Nakazawa, H. & Morimoto, N. (1970) Pyrrho -tite phase relations below 320°C. Proc. JapanAcad., 46, 678-683.

Nicholson, R.V. & Scharer, J.M. (1994): La bor -atory studies of pyrrhotite oxidation kinetics.In: Environmental geochemistry of sulfide oxi-dation. Alpers C.N., Blowes D.W. (eds.), .ACSSymp. Series 550, 14-30.

Ohfuji, H. & Rickard, D. (2006) High resolutiontransmission electron microscopic study of syn-thetic nanocrystalline mackinawite. Earth andPlanetary Science Letters, 241, 227-233.

Pierce, L. & Buseck, P.R. (1974) Electron ima-ging of pyrrhotite superstructures. Science,186, 1209-1212.

Pollok, K., Langenhorst, F., Hopf, J., Kothe, E.,Geisler, T., Putnis, C.V. & Putnis, A. (2008) Micro -structural Controls on Monosulfide Weatheringand Heavy Metal Release (MIMOS), GEOTECH-NOLOGIEN Science Report, 12, 79-88.

Poulton S.W., Krom, D.M. & Raiswell R. (2004)A revised scheme for the reactivity of iron (oxy-hydr)oxide minerals towards dissolved sulfide.Geochim. Cosmochim. Acta, 68, 3703-3715.

Rawlings, D.E. (2002) Heavy metal mining u -sing microbes. Annual Reviews in Micro bio -logy, 56, 65–91.

Rodriguez, Y., Ballester, A., Blazquez, M. A. &Munoz, J. A. (2003) Study of bacterial attach-ment during the bioleaching of pyrite, chalco-pyrite and sphalerite. Geomicrobiol. J. 20,131–141.

Rodriguez-Leiva, M. & Tributsch, H. (1988) Mor -phology of bacterial leaching patterns by Thi o -bacillus ferrooxidans on synthetic pyrite. Arch.Microbiol. 149, 401–405.

Sand W., Gehrke, T., Hallman, R. & Schippers,A. (1995) Sulfur chemistry, biofilm, and the(in) direct attack mechanism – a critical evalua-tion of bacterial leaching. Appl. Microbiol.Biotechnol., 43, 961-966.

Schippers, A. (2004) Biogeochemistry of metalsulfide oxidation in mining environments, sedi-ments and soils. In:Amend, J.P., Edwards, K.J.,Lyons, T.W. (Eds.), Sulfur Biogeochemistry—Past and Present. Special Paper, vol. 379. Geo -logical Society of America, Boulder, Colorado,pp. 49-62.

Schippers, A., Kock, D., Schwartz, M., Bött cher,M.E., Vogel, H. & Hagger, M. (2007) Geomic -robiological and geochemical investigation of apyrrhotite-containing mine waste tailings damnear Selebi-Phikwe in Botswana. J. Geochem.Expl., 92, 151-158.

Yamamoto, A. & Nakazawa, H. (1982)Modulated structure of the NC-type (N = 5.5)pyrrhotite, Fe1-xS. Acta Cryst. A, 38(1), 79-86.

SR16_Mineraloberfl_V8.qxp:Abstract_InfosysII.qxd 08.10.10 13:25 Seite 197