Mechanisms Governing the Growth, Reactivity and Stability of Iron Sulfides by Francis William Herbert M.Eng, Materials Science, University of Oxford, UK Submitted to the Department of Materials Science and Engineering in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Materials Science and Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY February 2015 c Massachusetts Institute of Technology 2015. All rights reserved. Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Department of Materials Science and Engineering November 20, 2014 Certified by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Bilge Yildiz Associate Professor of Nuclear Science and Engineering Thesis Supervisor Certified by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Krystyn J. Van Vliet Associate Professor of Materials Science and Engineering Thesis Supervisor Accepted by . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Donald Sadoway John F. Elliott Professor of Materials Chemistry Chair, Departmental Committee on Graduate Students 2 Mechanisms Governing the Growth, Reactivity and Stability of Iron Sulfides by Francis William Herbert Submitted to the Department of Materials Science and Engineering on November 20, 2014, in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Materials Science and Engineering Abstract The kinetics of electrochemical processes in ionic materials are fundamentally governed by dynamic events at the atomic scale, including point defect formation and migration, and molecular interactions at the surface. A corrosion system comprising an iron sulfide film (passive layer) formed on iron or steel in contact with an hydrogen sulfide (H2 S)rich fluid can thus, in principle, be modeled by a series of unit reaction steps that control the rate of degradation under given thermodynamic conditions. This overarching thesis goal necessitates a concerted experimental and computational approach to determine the relevant kinetic parameters such as activation barriers Ea and rate constants νo for the homogeneous and heterogeneous reactions of interest. These fundamental values can be obtained experimentally via temperature-dependent measurements on pure, model iron sulfide samples. This thesis therefore consists of three case studies on the stable Fe-S phases pyrrhotite (Fe1-x S) and pyrite (FeS2 ) to identify the elementary corrosion mechanisms and their kinetic parameters. Pyrrhotite is of interest because the off-stoichiometry of this phase leads to relatively rapid bulk processes such as diffusion; pyrite has a comparitively inert bulk but this work showed that it has a chemically labile surface. The first study focuses on two basic, rate-controlling steps in the growth of pyrrhotite: cation diffusion and sulfur exchange at the surface. First, iron self-diffusivity * DFe is determined across the temperature range 170-400 o C through magnetokinetic studies of the diffusion-driven "λ" magnetic transformation, as well as direct tracer diffusion measurements in Fe1-x S crystals using secondary ion mass spectrometry (SIMS). This range encompasses the sponteneous magnetic and structural order-disorder temperature T N = 315 o C in pyrrhotite. The effect of spontaneous magnetization below T N is to increase the Fe vacancy migration energy by a combined 40% increasing Ea for diffusion from 0.83 eV in paramagnetic Fe1-x S to ∼1.20 eV in the fully magnetized state. An extrapolation of the Arrhenius law from the paramagnetic regime would therefore overestimate actual diffusivities by up to 102 times at 150 o C. Second, the surface exchange of sulfur from H2 S into the solid state in Fe1-x S is measured using electrical conductivity relaxation, yielding Ea = 1.1 eV for sulfur incorporation into pyrrhotite. With their similar thermal dependence, there is no clear temperature crossover from cation diffusion- to surface exchange-limiting regimes, or vice versa. Instead, surface exchange is expected to constrain pyrrhotite growth for films under ∼ 100 µm thickness, beyond which diffusion becomes the rate limiting mechanism, independent of external driving factors such as temperature. The second study explores the role of surface electronic states on the electrochemical reactivity of pyrite. Charge transfer between a solid surface and an adsorbate such as H2 S requires the mutual availability of filled/empty electronic states at the same energy level. The semiconducting FeS2 (100) surface is predicted to have intrinsic surface states (SS’s) from Fe and S dangling bonds, as well as extrinsic SS’s related to delocalized defects at the surface, both of which would affect charge transfer characteristics. A novel, broadly-applicable methodology is developed in this thesis to quantify 3 the energy and density of these SS’s, based on experimental scanning tunneling microscopy/spectroscopy (STM/STS) in conjunction with first principles tunneling current modeling. As a result, a decreased surface band gap Eg of 0.4 eV, compared to 0.95 eV in bulk pyrite, is measured. The findings highlight the need to differentiate between bulk and surface electronic structure when assessing heterogeneous reactivity, and have implications for the use of FeS2 in potential technological applications, for example as a photovoltaic adsorber. Finally, the dynamics of point defect formation and clustering on FeS2 (100) under high-temperature, reducing conditions are investigated to understand the stability of the surface under extreme conditions. Synchrotron x-ray photoelectron spectroscopy (XPS) is used to measure a formation energy ∆H f for sulfur vacancies in the topmost atomic layer of 0.1 eV up to approximately 240 o C. Above this temperature, however, point defects are shown to condense into surface pits as measured by scnaning tunneling microscopy (STM). The combined, experimental XPS and STM results are replicated with high precision by a kinetic Monte Carlo (kMC) simulation, developed by Aravind Krishnamoorthy towards his doctoral thesis, of surface degradation on realistic lengthand timescales of 10−10 − 10−7 m and up to several hours, respectively. The findings have implications for the initiation of surface breakdown via pitting in ionic passive films, as well as providing a broader understanding of the non-stoichiometry of the pyrite surface. The common thread is a focus on events at the atomic and electronic scale, with an emphasis on point defects. The results thereby facilitate a bottom-up approach to modeling electrochemical processes such as corrosion in Fe-S phases, in which the unit steps are cast into probabilistic simulation tools. While the three studies here comprise only a partial examination of the atomic-scale events regulating the behavior of Fe-S passive layers, this approach makes inroads towards more accurate component lifetime prediction and the design of robust materials for aggressive environments. Moreover, the fundamental surface and bulk physical chemistry of iron sulfides explored in this work has implications beyond corrosion to other uses of these materials, including potential magnetic devices (Fe1-x S) and earth-abundant photovoltaic and photoelectrochemical adsorbers (FeS2 ). Thesis Supervisor: Bilge Yildiz Title: Associate Professor of Nuclear Science and Engineering Thesis Supervisor: Krystyn J. Van Vliet Title: Associate Professor of Materials Science and Engineering 4 Acknowledgments I am deeply grateful to my co-advisors, Professor Bilge Yildiz and Professor Krystyn van Vliet, for their encouragement, guidance and support. It has been an immense pleasure to witness both Bilge and Krystyn establish themselves with tenure at MIT during my time here and be part of two flourishing laboratories. Meanwhile, Bilge’s passion for solid state chemistry and keen eye for important details, and Krystyn’s diligent and organized approach to high-quality scientific inquiry have greatly inspired me. I also thank Krystyn for teaching me how to spell "properly": the word sulphide seems as alien to me now as sulfide did five years ago. This work would not have been possible without my ever-dependable collaborator and friend Aravind Krishnamoorthy. He is gifted not only with a brilliant scientific intellect, but an immensely humble and generous personality that has made working together on this project a richer experience. His efforts truly allowed our combined computational and experimental approach to become more than the sum of its parts. I am also thankful to my collaborators on this project and others, including: Wen Ma, Yan Chen and Qiyang Lu from the Laboratory for Electrochemical Interfaces at MIT; Peter Albrecht at Brookhaven National Laboratory; Predrag Lasic and Rickard Armiento (Ceder group, MIT); Rupak Chakraborty and Katy Hartman (Buonassissi group, MIT). Thank you to Prof. Randall Feenstra at Carnegie Mellon University for his help deciphering the SEMITIP code for tunneling spectroscopy simulations. I would like to thank all members, past and present, from my two fantastic reasearch groups: the Laboratory for Electrochemical Interfaces (LEI) and the Laboratory for Material Chemomechanics who have taught me so much, from defect chemistry in ionic solids to the mechanics of living cells. In particular, I am very grateful to Roza Mahmoodian for her support and for putting up with my incessant complaining over failed experiments. Also to Bal Mukund Dhar for his infectious enthusiasm and help with CVD, and to Lucy Rands for her help and eagerness as a summer intern. I am indebted to my thesis committee - Prof. Carl Thompson and Prof. Harry Tuller - for their useful comments and constructive criticism. In addition, I greatly thank Prof. Chris Schuh for providing invaluable feedback, despite not sitting on my final committee. BP Plc. had already supported my education for over 20 years when they arrived at MIT to propose this project, so I am delighted they extended their commitment to my graduate studies. In particular, I would like to thank Sai Venkatesweran, Richard Woolam, Steve Shademan and their colleagues for their help and advice. My parents, Richard and Kate, have inspired and guided me my whole life; I would not be here without the opportunities and unwavering support they have provided. And I cannot omit the other four fifths of my band of brothers who are my frame of reference for everything and never stop injecting humour and happiness into my life. Finally, thank you to all those who have made my time at MIT so special outside of the lab. My family away from home; the eclectic and dynamic community at "Martha" (216 Norfolk St): Sam, Katy, Georgie, Jake, Benji, Ines, Chris, Nina, Alex, Federico, Elison, Andre, James, Andre, Stephanie, Rob, Balthazar, Simon, Sebastian, Nico, Melissa, Serjumbi, Aron, all of our other guests, and last but not least Teresa for not losing faith in me after all these years. 5 6 Contents 1 Introduction 13 1.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 1.2 Passivity: a brief introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 1.3 Iron sulfide phases and corrosion products . . . . . . . . . . . . . . . . . . 14 1.3.1 Sour corrosion mechanism: lab and field experience . . . . . . . . 17 1.3.2 Model phases: Pyrrhotite (Fe1-x S) and Pyrite (FeS2 ) . . . . . . . . 19 1.4 Towards a predictive, multiscale corrosion model . . . . . . . . . . . . . . 20 1.4.1 Existing passive film models . . . . . . . . . . . . . . . . . . . . . . . 20 1.4.2 Unit processes controlling Fe-S passive layer behavior . . . . . . . 23 1.4.3 The need for experimentally-derived parameters . . . . . . . . . . 24 1.5 Thesis goals and organization . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2 Growth: cation diffusion and surface exchange as rate-limiting mechanisms in pyrrhotite, Fe1-x S 27 2.1 Background and Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2.2 Pyrrhotite: polytypes and transitions . . . . . . . . . . . . . . . . . . . . . . 29 2.2.1 Structural and magnetic properties . . . . . . . . . . . . . . . . . . . 30 2.2.2 The λ-transition in NC pyrrhotites . . . . . . . . . . . . . . . . . . . 34 2.3 Diffusion-limited λ transition . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2.3.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35 2.3.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2.3.3 Continuous re-ordering of ferrimagnetic superlattice . . . . . . . 40 2.3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45 2.4 Isotope tracer diffusion measurements . . . . . . . . . . . . . . . . . . . . . 45 2.4.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 2.4.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 48 2.4.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 2.5 Sulfur exchange kinetics at the Fe1-x S surface . . . . . . . . . . . . . . . . . 54 2.5.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 2.5.2 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 57 2.5.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 2.6 Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 2.6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 2.6.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 7 3 Reactivity: quantification of electronic band gap and surface states on FeS2 (100) 69 3.1 Background and motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 3.1.1 Electrochemical charge transfer in semiconductor-absorbate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 3.1.2 Surface states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 3.1.3 Scanning tunneling spectroscopy and TIBB . . . . . . . . . . . . . . 74 3.1.4 The FeS2 (100) surface . . . . . . . . . . . . . . . . . . . . . . . . . . 76 3.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3.2.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3.2.2 Computational . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 3.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 3.3.1 Current-separation and current-voltage tunneling spectroscopy . 80 3.3.2 Simulated tunneling spectra based on DFT-calculated DOS . . . . 82 3.4 Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 3.4.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 3.4.2 Implications for other applications of FeS2 , e.g. PV . . . . . . . . . 89 3.4.3 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 4 Stability: dynamics of point defect formation, clustering and pit initiation on the pyrite surface 91 4.1 Background and motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 4.1.1 Chapter goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 4.1.2 Passivity breakdown by pitting . . . . . . . . . . . . . . . . . . . . . 92 4.1.3 FeS2 surface chemistry and non-stoichiometry . . . . . . . . . . . . 94 4.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 4.2.1 Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 4.2.2 Computational . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 4.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 4.3.1 Evolution of pyrite surface structure and chemistry . . . . . . . . . 97 4.3.2 Mechanism of vacancy formation and coalescence . . . . . . . . . 103 4.4 Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 4.4.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 4.4.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 5 Conclusions 109 5.1 Summary of activation barriers . . . . . . . . . . . . . . . . . . . . . . . . . . 109 5.2 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 5.3 Outlook and perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 A Pourbaix diagrams for the Fe-H2 S-H2 O system 113 B Chemical Vapor Deposition of Fe-S 119 B.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 B.2 Methods: CVD setup and apparatus . . . . . . . . . . . . . . . . . . . . . . . 119 B.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 C Diffusivity measurements using thin film samples 8 129 List of Figures 1-1 1-2 1-3 1-4 1-5 1-6 1-7 1-8 Potential E vs. current i (polarization) curve for a generic metal. . . . . . Global sour oil and gas statistics. . . . . . . . . . . . . . . . . . . . . . . . . . Thermodynamic predictions of corrosion products. . . . . . . . . . . . . . Mechanism of iron sulfide formation on steels in H2 S-bearing electrolytes. Iron sulfide stability phase diagram. . . . . . . . . . . . . . . . . . . . . . . . Sulfide corrosion of 4130 carbon steel at 220 o C. . . . . . . . . . . . . . . . Schematic of unit processes in Fe-S passive layers. . . . . . . . . . . . . . . Overview of strategy to construct a non-empirical passive film model. . . 14 15 17 18 20 22 24 25 2-1 2-2 2-3 2-4 2-5 2-6 2-7 2-8 2-9 2-10 2-11 2-12 2-13 2-14 2-15 2-16 2-17 2-18 2-19 2-20 2-21 2-22 2-23 2-24 2-25 2-26 2-27 2-28 2-29 2-30 Collected literature values of Fe self-diffusivity. . . . . . . . . . . . . . . . Structural unit cells of pyrrhotite. . . . . . . . . . . . . . . . . . . . . . . . Pyrrhotite structural and magnetic phase diagrams. . . . . . . . . . . . . Idealized Fe1-x S superstructures. . . . . . . . . . . . . . . . . . . . . . . . . Distributions of vacancies in 4C and NC pyrrhotites. . . . . . . . . . . . . The peak-like λ-transition in NC Fe1-x S. . . . . . . . . . . . . . . . . . . . X-ray diffraction of synthetic pyrrhotites. . . . . . . . . . . . . . . . . . . . Setup of cubic kinetic Monte Carlo (kMC) grid. . . . . . . . . . . . . . . . Temperature-dependent magnetization σ(T ). . . . . . . . . . . . . . . . . Magnetization vs. applied field (σ-H). . . . . . . . . . . . . . . . . . . . . Reversible magnetic transformation at short timescales. . . . . . . . . . Best fits to exponential equation. . . . . . . . . . . . . . . . . . . . . . . . Long-timescale isothermal magnetization. . . . . . . . . . . . . . . . . . . Differential scanning calorimetry (DSC) results. . . . . . . . . . . . . . . Continuous re-ordering towards ferrimagnetic state. . . . . . . . . . . . Cross section of sulfide scale. . . . . . . . . . . . . . . . . . . . . . . . . . . Cu-kα powder XRD pattern. . . . . . . . . . . . . . . . . . . . . . . . . . . Energy-dispersive X-ray spectroscopy (EDS) . . . . . . . . . . . . . . . . . Sources of error considered in statistical analysis of diffusion data. . . Secondary ion mass spectrometry (SIMS) profiles. . . . . . . . . . . . . . Error function solution to diffusion profiles. . . . . . . . . . . . . . . . . . Values for iron self-diffusion coefficient * DFe . . . . . . . . . . . . . . . . . Sputter deposited thin films for ECR experiments. . . . . . . . . . . . . . Electrical conductivity relaxation apparatus setup. . . . . . . . . . . . . . Temperature-pressure equilibrium phase diagram for Fe-S. . . . . . . . . X-ray photoelectron spectroscopy (XPS) from a Fe1-x S thin film sample. Electrical resistance relaxation at 565 o C. . . . . . . . . . . . . . . . . . . Electrical conductivity relaxation results. . . . . . . . . . . . . . . . . . . . Drift, stability and repeatability of ECR experiments. . . . . . . . . . . . Temperature- and film thickness dependence of rate limiting steps. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 30 32 33 33 34 35 36 38 38 40 41 42 43 44 45 46 47 49 49 50 52 55 56 57 59 62 63 65 66 3-1 Charge transfer in electrochemical (corrosion) systems. . . . . . . . . . . 3-2 Band bending effects in STS measurement. . . . . . . . . . . . . . . . . . . 3-3 FeS2 single crystal samples. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 75 78 9 3-4 Distributions of surface states as defined in the SEMITIP program. . . . . 3-5 Scanning tunneling spectroscopy (STM) images of the as-grown FeS2 (100) surface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-6 Current-separation spectroscopy. . . . . . . . . . . . . . . . . . . . . . . . . . 3-7 Current-voltage spectroscopy. . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-8 Pyrite valence band. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-9 Modeling tunneling spectroscopy with surface states . . . . . . . . . . . . 3-10 Density functional theory (DFT)-computed band structures. . . . . . . . . 3-11 Fitting to experimental surface Eg . . . . . . . . . . . . . . . . . . . . . . . . . 3-12 Visualization of FeS2 (100) surface charge q . . . . . . . . . . . . . . . . . . 3-13 Low surface bandgap implications for PV. . . . . . . . . . . . . . . . . . . . 3-14 Preliminary investigations on bulk and 2-dimensional MoS2 . . . . . . . . 4-1 4-2 4-3 4-4 4-5 4-6 4-7 4-8 80 81 82 83 83 85 86 87 88 89 90 Proposed mechanisms of passivity breakdown and pitting. . . . . . . . . . Nanopits formed by vacancy agglomeration. . . . . . . . . . . . . . . . . . XPS sample clamp for FeS2 crystals. . . . . . . . . . . . . . . . . . . . . . . . S 2p photoelectron spectra of FeS2 (100). . . . . . . . . . . . . . . . . . . . Atomic model of the FeS2 (100) surface as viewed side-on. . . . . . . . . . Sulfur monomer vacancy concentration. . . . . . . . . . . . . . . . . . . . . Proportion of the M and S components of the S 2p photoelectron spectra. Scanning tunneling microscopy (STM) images of single crystal FeS2 (100) surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-9 Pits are one half- or one lattice parameter deep. . . . . . . . . . . . . . . . 4-10 Illustration of atomic processes involved in the proposed mechanism of pit formation and growth on pyrite (100). . . . . . . . . . . . . . . . . . . . 4-11 kinetic Monte Carlo simulation results. . . . . . . . . . . . . . . . . . . . . . 93 94 96 98 99 101 101 B-1 B-2 B-3 B-4 B-5 B-6 B-7 Home-made Chemical Vapor Deposition (CVD) system. . . . Description and safety information for Fe and S precursors. Iron sulfide films deposited from Fe(acac)3 and TBDS. . . . Carbon contamination in Fe-S films from Fe(acac)3 . . . . . . Iron sulfide films deposited from Fe(CO)5 and TBMS. . . . . Iron sulfide films deposited from Fe(CO)5 and H2 S. . . . . . Template stripping for ultrasmooth sulfide surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 122 124 125 125 126 127 C-1 C-2 C-3 C-4 Iron self-diffusivity * DFe measurements. . . . . . . . X-ray diffraction of thiin films for ECR experiments. Representative diffusion profiles. . . . . . . . . . . . Oxidation of samples annealed in quartz vials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 131 132 133 10 . . . . . . . . . . . . . . . . . . . . 102 104 105 105 List of Tables 1.1 Stable and metastable iron sulfide phases. . . . . . . . . . . . . . . . . . . . 1.2 Reactions describing the basic unit processes. . . . . . . . . . . . . . . . . . 16 23 2.1 2.2 2.3 2.4 2.5 2.6 2.7 31 37 41 47 51 57 Fe1-x S polytypes: composition and structure. . . . . . . . . . . . . . . . . . Thermodynamic values for pyrrhotite compounds. . . . . . . . . . . . . . . Best fit parameters n and τ. . . . . . . . . . . . . . . . . . . . . . . . . . . . . Isotopic composition of naturally-occurring iron. . . . . . . . . . . . . . . . Iron self-diffusion * DFe measurement results for Fe1-x S crystals. . . . . . . K p (T) values used to calculate sulfur partial pressure. . . . . . . . . . . . Deconvolution of Fe 2p and S 2p x-ray photoelectron spectroscopy (XPS) peaks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.8 Electrical conductivity relaxation results for oxidation experiments. . . . 2.9 Electrical conductivity relaxation results for reduction experiments. . . . 2.10 Key activation energies for pyrrhotite growth. . . . . . . . . . . . . . . . . . 59 61 61 68 3.1 Calculated bulk band gap Eg , and surface Eg . . . . . . . . . . . . . . . . . . 77 3.2 Experimental surface Eg measurements by scanning tunneling spectroscopy (STS). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3.3 Input parameters for tunneling spectroscopy simulations using the SEMITIP program. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79 4.1 XPS core level shift (CLS) for S 2p peak. . . . . . . . . . . . . . . . . . . . . 97 5.1 Summary of experimentally determined activation barriers Ea . . . . . . . 109 A.1 A.2 A.3 A.4 A.5 A.6 Thermodynamic data for species in H2 S-H2 O-Fe system. Input parameters. . . . . . . . . . . . . . . . . . . . . . . . . Fe-H2 O Reactions and reversible potentials. . . . . . . . . Mackinawite-Fe-H2 O system equilibrium reactions. . . . Pyrrhotite-Fe-H2 O system equilibrium reactions. . . . . . Pyrite-Fe-H2 O system equilibrium reactions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 114 115 116 116 117 B.1 CVD of Fe-S phases by other authors. . . . . . . . . . . . . . . . . . . . . . . 122 B.2 Chemical Vapor Deposition conditions for Fe-S phases: literature. . . . . 123 11 12 Chapter 1 Introduction 1.1 Context Vast amounts of energy are consumed in distilling ferrous and non-ferrous metals from their ground states, married to reactive atoms such as oxygen and sulfur in ores. Once fashioned into the components that undergird our energy, transportation and construction infrastructures, the thermodynamic tendency of these metals to regress to their primitive compounds locks us in a Sysyphean struggle against materials degradation and corrosion. But while thermodynamics dictates this ultimatly pessimistic outcome, the slow kinetics of the required reactions ensures that metals and alloys can remain uncompromised for many years, even under the most aggressive chemical, thermal and mechanical conditions. Passivity - the ability of metals to self-protect by forming a thin, inert skin of an ionic compound through a partial reaction at the environment interface - thus constitutes one of our most useful tools in combating degradation. The growth, reactivity and stability of these passive layers all contribute to their overall protectiveness. In this thesis we ask: is it possible to understand and subsequently simulate these characteristics a priori, starting from the atomic scale? We can begin by defining a series of kinetically rate-limiting processes, including mass transport (i.e., diffusion) and surface reactions such as oxidation, reduction and dissolution. By investigating the fundamental unit steps involved, and modeling these steps across multiple length- and timescales, we can aspire not only to predict accurately how metals will behave when placed in harsh environments, but also to design more robust materials that will better self-protect and last longer. 1.2 Passivity: a brief introduction The study of passivity as a scientific discipline perhaps originated with James Keir in 1790, who documented that although iron readily dissolves in dilute nitric acid, it counterintuitively remains inert in concentrated HNO3 . Michael Faraday, as early as 1830, before the invention of the necessary characterization tools, correctly predicted the existence of an ultrathin, electronically conducting surface film that protects the underlying metal. [1] We now have a better understanding of the passive state, as defined by Uhlig [2]: A metal is passive if it substantially resists corrosion in a given environment despite a marked thermodynamic tendency to react. This definition gives us an insight into the essence of passivity; it is a metastable state, one in constant flux. The formation of a passive film does not provide an infinite barrier to metal dissolution in a given aggressive environment, but it does reduce corro13 Schematic polarization curve for passive system Active Passive 1 Transpassive Log i (A cm-2) Pitting (a) Cr, Ni (b) Fe (c) Valve metals Al, Ti, Ta, Zr, Hf (a) (b) 10-6 Eo Ep Epit (c) E1 E (V) Figure 1-1: Potential E vs. current i (polarization) curve for a generic metal, showing the active, passive and transpassive regions, as discussed in the text. The valve metals form extremely stable oxide passive films and display a very large transpassive region. After Marcus et al. [3] sion currents by several orders of magnitude (as much as 106 times for certain alloys such as stainless steels, and metals such as Al, Si, Ti, Ta and Nb, that form very stable passive films). Figure 1-1 shows a schematic of a typical polarization curve for a generic, passive-film forming metal. [3] At a critical passivation potential Ep the corrosion current drops dramatically, coincident with the formation of a passive layer. The thermodynamic conditions for this can be predicted from fundamental electrochemical principles, the basis of the Pourbaix diagram. [4] However, at higher potentials Epit the passive layer can be prone to localized breakdown or pitting. Eventually, localized failure avalanches into a full breakdown of the passive film and the corrosion current shoots up in the transpassive region. The structure of passive films is typically bi- or multi-layered. In the case of an oxide this typically means a thin (<100 nm), adherent, dense inner barrier layer that forms rapidly under the correct anodizing conditions. An outer layer can form by precipitation or solid-state reaction of cations transmitted through the barrier layer with anionic species in the aqueous media, e.g. H2 S, CO3 2- , HS- . Depending on the exact conditions of temperature, pH, concentration of reactive species, pressure etc., the outer layer - which can grow up to 100-1000x the thickness of the inner barrier layer - may contain multiple compounds and different phases. In oilfield brines with dissolved CO2 and H2 S, both FeCO3 and a variety of FeS phases may be present in the outer layer, possibly along with oxides. 1.3 Iron sulfide phases and corrosion products Early oil prospectors had a rudimentary test for the quality of their discoveries: a smudged fingertip to the tongue, hence the terminology still used today: "sweet" or "sour" crude, depending on whether the sulfur content is above or below 0.5 vol The solution- and solid-state chemistry of iron sulfides has been studied by geochemists, microbiologists and thermodynamicists for many decades. A comprehensive review by Rickard and Luther covers much of the accumulated knowledge in this field. [9] The iron sulfide family of phases and their interrelations are complex: up to nine discrete phases have been characterized, as listed in Table 1.1. Several of these are metastable and will convert over time to either stable iron monosulfide (pyrrhotite, Fe1-x S, 0 ≤ x ≤ 0.125 ) or iron disulfide (pyrite, FeS2 ). We limit the discussion here to the three phases which came up most often while reviewing the literature on sour 14 (a) World petroleum reserves with selected crude sulfur content (2013) Urals 1.3-2.5% North Sea Brent 0.4% Thunderhorse 0.9% West Texas 1.9% Saudi 2-3% Nigeria 0.2% Venezuela 1.5-2.3% 110 270 30 Proven oil reserves, bbl (c) Average % S 1.5 OPEC 1.3 1.2 1.1 World Non-OPEC 1 0.9 2015 2025 Year 2035 Crude Oil Sulfur Content High pressure/high temperature Reservoir Temperature oF (b) Souring of world oil production 1.4 Indonesia (Arjuna) 0.1% 600 HPHT-hc 260 oC o Ultra HPHT 400 205 C o 150 C HPHT 200 0 69 MPa 0 Siberian Basin 0.8-1.8% Iran 1-2% 241 MPa US Midcontinent 0.4% 138 MPa Canada- oil sands 3.4% 10 20 40 60 0 Static Reservoir Pressure (kpsi) Figure 1-2: Global sour oil and gas statistics. (a) Proven world crude oil reserves [5], with the average sulfur content highlighted for several representative countries. [6, 7]. (b) The average sulfur content of oil for Organization of the Petroleum Exporting Countries (OPEC) and worldwide. [5] (c) Definitions of high pressure-high temperature well conditions by Schlumberger, Ltd. [8] corrosion: mackinawite, pyrrhotite and pyrite. Mackinwawite, FeS1-x is a 2-dimensional (2D) layered chalcogenide, generally considered to be the initial corrosion product to form under most conditions in H2 S-bearing solutions. [10] Figures 1-3a and b show potential vs. pH (Pourbaix) diagrams for the Fe-H2 S-H2 O system under standard conditions of 1 atm pressure and 25 o C. Full details of the thermodynamic diagrams are provided in Appendix A and Refs. [11–14]. The stable phases pyrrhotite and pyrite are excluded in Figure 1-3a, given that mackinawite manifests itself as the predominant metastable corrosion product for neutral to alkaline solutions. This would correspond to "slightly" sour conditions of T < o C and hydrogen sulfide partial pressures PH2 S < 0.01 MPa [15], or for relatively short exposures up to hundreds of hours. Naturally, the thermodynamic data would predict the stable phases to predominate eventually (Fig. 1-3b), which in reality means after long steady-state exposure or higher temperatures. Evidently the kinetics of iron sulfide formation, dissolution and transformation preclude the prediction of corrosion product phases from thermodynamic principles alone. In particular, the effect of H2 S partial pressure is omitted. Figure 1-3c is a summary of available phase identification from laboratory tests across a range of sour conditions. Mackinawite prevails under slightly sour conditions, but pyrrhotite or a mixture of pyrrhotite and pyrite is more often observed at moderately-highly sour conditions as indicated on the diagram. 15 16 Hexagonal P6̄2c Monoclinic A2/a or Hexagonal P6/mmc Trigonal-hexagonal R3m FeSm , FeS1-x FeSc FeS Fe1-x S (0 ≤ x ≤ 0.17) Fe3+x S4 (0 ≤ x ≤ 0.3) Fe3 S4 FeS2 FeS2 Mackinawite Cubic FeS Troilite Pyrrhotite Smythite Greigite Pyrite Marcasite Orthorhombic Pnnm Cubic Pa3 Cubic Fd3m Cubic F 4̄3m Tetragonal, 2D layer P4/nmm Amorphous/ nanocrystalline FeSm , FeS1-x Amorphous Iron Sulfide Structure Composition Phase M S M M S S M M M Meta- (M)/ Stable (S) Commonly observed in hydrothermal systems; transforms to pyrite. ”Fools gold”: stoichiometric iron disulfide. Contains mixture of Fe2+ and Fe3+ ions.more Sub-phase from pyrrhotite group. Iron-deficient monosulfide. Off-stoichiometry accomodated by iron vacancy superstructures below 315 o C. Fully stoichiometric end member of the Fe1-x S family. Transforms to mackinawite, pyrrhotite or pyrite. Not observed in nature. Commonly observed first corrosion product in sour solutions. Believed by some authors to be simply nanocrystalline form of mackinawite. Comments Table 1.1: Stable and metastable iron sulfide phases. Adapted from Refs. [9, 10]. (a) Pourbaix diagram: metastable Fe-S (b) 1.5 1.5 1 Fe3+ O2 +4 H ++4 0.5 e -=2 H2 O Erev (V) Erev (V) 1 Fe2O3 0.5 0 Fe2+ 2H ++ 0 Fe 2 4 6 pH 8 10 (c) 0 -0.5 2e -= -1 Pyrite Fe2+ Mackinawite -0.5 -1.5 Pourbaix diagram: stable Fe-S H2 12 Pyrrhotite -1 14 -1.5 0 Fe 2 4 6 pH 8 10 12 14 Fe-S scales observed in lab/field tests H2S Partial Pressure (MPa) “Moderately Sour” “Highly Sour” 0 10 -2 10 Mackinawite Pyrrhotite Pyrrh. & Pyrite -4 10 “Slightly Sour” 0 50 100 150 200 Temperature (oC) 250 Figure 1-3: Thermodynamic predictions of corrosion products are only of limited usefulness. (a) pHelectrode potential (Pourbaix) diagram constructed for the Fe-H2 S-H2 O system, excluding the stable phases from reactions. (b) Once the stable products pyrrhotite Fe1-x S and pyrite FeS2 are added, the mackinawite field disappears entirely. These diagrams cannot predict the existence of a given corrosion product a priori. In (c), we show a temperature-H2 S partial pressure PH2 S diagram, overlaid with experimentally-observed phase identifications from several sources. [15–18] Mackinawite is widely observed at lower temperatures and PH2 S ("slightly sour" conditions; at elevated temperatures and partial pressures the product mix shifts to the more stable pyrrhotite and pyrite phases. Exposure time is another factor not considered in these purely thermodynamic diagrams; mackinwaite commonly forms first, but is more likely to transform over longer times up to 100s of hours due to its metastability. 1.3.1 Sour corrosion mechanism: lab and field experience Despite more than seventy years’ worth of investigations into iron sulfide corrosion products, there is still no solid consensus surrounding the mechanism of sour corrosion and the extent of protectiveness conferred by the passive state. The presence of an iron sulfide scale is generally thought to reduce corrosion rates by up to several orders of magnitude, at least upon initial formation on bare steel. However, there is uncertainty regarding the long term stability of otherwise protective films comprising Fe-S. [19] Here we briefly review the observations from laboratory and field experience that allow us to form an empirical picture of the sour corrosion process from a mechanistic perspective. We refererence to Figure 1-4, we broadly describe a multi-stage growth mechanism whereby mackinawite nucleates first, grows to a critical thickness and then 17 (a) Initial formation of bi-layered mackinawite film Porous outer layer 10’s of μm Fe(s)+H2S(aq) = FeSm (s)+2H+(aq)+2e- Inner layer ~ 100 nm Fe2+(aq) +H2S(aq) = FeSm (s)+2H+(aq) Dissolution: 0 + FeS , FeHS Fe-C 30µm (b) Pitting at delamination sites Cl - 20 μm Fe-C (c) Nucleation & growth of stable iron sulfides Figure 1-4: Mechanism of iron sulfide formation on steels in H2 S-bearing electrolytes. (a) A thin, adherent mackinawite film (∼ 100 nm thick) forms via solid state reaction or local precipitation at the bare steel surface. The mackinawite continually dissolves and re-precipitates out of the supersaturated solution in the adjacent boundary layer, building up a much thicker, outer porous layer of mackinawite. [16] (b) because of volumetric stresees, the film may delaminate locally, leading to the formation of pits with bubble-like sulfide deposits. [18] (c) At the repassivated pit sites, a high local concentration of cations in solution may lead to the nucleation of more stable phases: needle-like crystals of hexagonal Fe1-x S or cubic grains of FeS2 . [17] ruptures. Following film rupture, a second stage of FeS film growth occurs during which it is possible that other, more stable Fe-S phases may form. Mackinawite formation: Bare carbon steel - even with a pre-existing oxide pasive film is thought to undergo a solid state reaction (SSR) with H2 S in solution, in other words a direct heterogeneous chemical reaction, to form nanocrystalline mackinawite FeSm (occasionally referred to as "amorphous" mackinawite). The evidence for a SSR revolves around the extremely rapid formation of iron sulfide on the surface over a timescale of several seconds. By contrast, carbonate scale precipitation in CO2 solutions requires minutes to hours for a semi-protective barrier film to form. [16, 19, 20] The formation of a compact mackinawite layer with thickness << 1 mm confers a degree of passivity to the steel, with a corresponding drop in measured corrosion rate on the order of 510 times. The film serves primarily as a diffusion barrier to ionic transport. Above the 18 compact, solid-state reacted mackinawite layer, a thicker - up to hundreds of mm - outer mackinawite layer is typically observed (Figure 1-4b). This outer film is more porous and less protective than the inner one. According to Nesic et al., mechanical instability of the inner film caused by epitaxial stresses as it grows lead to a cyclic process of microcracking, delamination and re-passivation that over time build up a thicker, porous outer layer. [21] Either way, outer film growth rate and eventual thickness depend on the temperature, pH and flow conditions of the solution. Its development leads to a steady further reduction in corrosion rate over periods up to hundreds of hours. [18] The corrosion rate is controlled by rate of mackinawite dissolution, the transfer of ions through the compact inner film and porous outer film as well as mass transport through the liquid boundary layer at the electrolyte-film interface. Re-passivation: At a certain thickness, the mackinawite film can delaminate entirely from the steel surface (Figure 1-4a). The solution in contact with the microcracks, or newly exposed steel surface, becomes supersaturated with Fe(HS)+ or Fe2+ . At this point, there are several developments that can occur, depending on the exact conditions at the damaged film locations. The first possibility is that the steel re-passivates with mackinawite through either SSR, or by precipitation. This is most likely when the solution is less acidic - above pH 6, atypical for oilfield brines - and at low temperatures since ferrous ion and H2 S levels are close to saturation limits for other phases such as troilite and pyrrhotite. Under these conditions the continuous spallation and repassivation of the mackinawite film confers a semi-passive state with general corrosion rates on the order of 1 mm/year. Pitting: Another possibility is that the freshly exposed area of bare alloy becomes susceptible to increased local attack by species in the environment. Particularly where chloride ions are involved in the disruption of the passive state, such a situation has been observed to lead to surface pitting of carbon steels and other alloys. [18, 22] Ex situ examinations of pitted regions have revealed large deposits of iron sulfide directly above the pit (Figure 1-4b); this provides further evidence of increased, localized attack and also suggests that pits may become re-passivated by precipitation from the electrolyte. Pyrrhotite and pyrite nucleation: Finally, certain environments are conducive to the formation of other iron sulfide phases. Mackinawite may undergo direct, solid state transformation to pyrrhotite under reducing conditions, or supersaturated conditions in breaks at the mackinawite layer may lead to the direct nucleation of elcongated troilite needles or hexagonal pyrrhotite plates. [17, 23] Eventually, such a process would lead to the build up of a thick scale (10s-100s of µm’s) of the more stable sour corrosion products pyrrhotite and pyrite. These phases are more than an order of magnitude less soluble than mackinawite. Once they have formed a continuous film, the further formation of mackinawite or other metastable phases is essentially inhibited by the greatly reduced local ferrous ion activity in solution. [15] 1.3.2 Model phases: Pyrrhotite (Fe1-x S) and Pyrite (FeS2 ) Despite the fact that mackinawite is an important phase in the early stages of aqueous iron sulfide formation on steels, this thesis concentrates on pyrrhotite and pyrite. This was motivated in large part by the relative ease with which high-quality and welldefined samples of the stable iron sulfides could be made. Mackinawite is a flaky brown precipitate that can only be formed from solution and oxidises rapidly upon exposure to air, therefore requiring careful handling under inert atmospheres and presenting additional technical challenges. Instead, pyrrhotite and pyrite served as model systems to investigate the physical chemistry of processes involving point defects in iron sulfides. The equilibrium phase diagram and crystal structures of the stable Fe-S phases 19 (a) Fe-S Equilibrium Phase Diagram (c) NiAs-type Fe1-xS (b) Troilite FeS a c b Fe S a c b (d) Pyrite FeS2 a c b Figure 1-5: Iron sulfide stability phase diagram. (a) Equilibrium Fe-S phase diagram. Pyrrhotite refers to a set of iron monosulfide polytypes, described in more detail in Chapter 2. Pyrite is a line compound at the fixed composition Fe:S = 1:2. (b) The crystal structure of the hexagonal pyrrhotite end-member, known as Troilite; (c) other pyrrhotites can be described by a hexagonal, NiAs-like subcell. (d) Cubic pyrite unit cell, described in Chapter 3. is shown in Figure 1-5a. Pyrrhotite Fe1-x S forms a complicated series of polytypes at low temperatures, from stoichiometric troilite (hexagonal FeS, Fig. 1-5b) to a range of vacancy-ordered superstructures based on the NiAs structure (Fig. 1-5c). Pyrite, by contrast, is a line compound that is nominally stoichiometric in the bulk. [24] 1.4 1.4.1 Towards a predictive, multiscale corrosion model Existing passive film models The broader aim of this thesis is to investigate a set of elementary processes at the atomic scale which govern the protectiveness of passive sulfide layers. As we have seen above, real passive layers are complex materials systems: often multi-phase, highly defective and sensitive to changes in electrolyte chemistry as well as global variables such as temperature and stress state. For the purposes of building a passive film model, we must necessarily reduce the system to a series of elementary unit processes that occur on the atmoic scale. Hence, in addition to thermodynamics we need to consider the kinetics 20 of the corrosion process also. The interest in analytical and computational modeling of passive layers dates back at least 70 years. We do not attempt an exhaustive review of all the available models here; instead, let us briefly examine some of the more relevant examples to understand their advantages and shortcomings: High Field Models account for the incipient oxidation of metals. [25–27] In the limit of small passive layer thickness L (on the order of nanometers) the field strength that drops across the film is extremely high, and the oxidation rate is inversely proportional to time t, yielding a logarithmic law: L(t) = ke log(αt + 1) (1.1) where ke and α are temperature- and field- dependent constants and L is film thickness. However, this only holds for very thin passive layers, applicable to the very early stages of oxidation or at low temperatures. Wagner theory describes the growth of passive films in anhydrous environments, i.e. in the absence of scale dissolution. [28] Assuming migration of ions through the film is rate-controlling, a parabolic law is obtained of the form: q (1.2) L(t) = k p t where kp is the parabolic rate constant, a function of the ionic diffusivity D of the mobile species. Dry sulfidation of iron at elevated temperatures > 500 o C is well described by Wagner theory [29–31]; however, it cannot account for scale dissolution in aqueous environments or the case when surface reactions, such as molecular dissociation, are kinetically limiting. The Point Defect Model (PDM) of Macdonald et al. is perhaps the most complete, deterministic analytical framework for describing the growth and protectiveness of generic passive layers. [32,33] Originally inspired by Wagner’s theory of diffusion-limited growth, it further incorporates interfacial reactions and scale dissolution. Interfacial reactions such as cation injection from the metal and redox exchange at the surface typically follow linear kinetics: L(t) = ki t (1.3) where ki is a surface exchange coefficient. The passive layer is treated as a bi-layer of inner, compact and highly defective oxide plus outer scale up to 100x thicker. Besides L(t) and corrosion currents i(t), the PDM can make quantitative predictions about passivity breakdown. Mackinawite formation models that describe iron sulfide scale formation have been proposed by Sun, Nesic et al.. [10, 16] The model assumes a bi-layered mackinawite film comprised of a thin (10 nm), compact inner layer and porous outer scale, 100s of µm thick. The corrosion rate is limited by mass transport (diffusion) through the mass transfer boundary layer at the scale-electrolyte interface, through the liquid in the porous outer scale and finally thorugh the compact inner sulfide. In addition, scale damage by hydrodynamic stresses is considered. However, the model is inherently empirical, requiring several unknown parameters to be fitted using experimental data. Moreover, the evolution of the scale to form stable sulfides at longer times is not considered. Stochastic pitting models have been developed to simulate pit growth kinetics in stainless steels. [34, 35] Based on Monte Carlo simulations of stochastic pitting, they can replicate pitting potentials observed using potentiodynamic experiements. However, the 21 (a) Gaseous Ar-10% H2S (220 oC) (b) Aqueous 10% H2S, 1% NaCl (220 oC) Figure 1-6: Sulfide corrosion of 4130 carbon steel at 220 o C. (a) Amount of sulfur reacted (mass gain of sample) with time in a dry, gaseous environment of 10% H2 S. An initial logarithmic transient due to sufide film formation gives way to a linear regime where the rate of sulfidation is limited by surface exchange of 2− sulfur, described by the overall reaction: H2 S(g) + 2e− ⇔ S FeS + H2(g) . At approx. 225 hours, the linear regime transitions to a "mixed" region in which diffusion of cations begins to limit the overall reaction rate. (b) Corrosion rate as a function of time in aqueous solution containing H2 S. The scale grows under mixed diffusion and interfacial (surface reaction) control. After 60 hours, the rate of scale growth is balanced by scale dissolution to produce a steady state in the corrosion rate. [17] models do not predict pitting initiation through passivity breakdown at the nanoscale, instead the starting point is a just-nucleated pit. To illustrate some of the rate limiting steps identified by the growth models above, Figures 1-6a and b show experimental results for steel sulfidation by H2 S at elevated temperatures of 220 o C, under dry gaseous and aqueous conditions, respectively. In the dry case (Fig. 1-6a), from the amount of reacted sulfur we see a transition from logarithmic sulfide scale growth (High Field Model) at short times, to linear growth (surface reaction, as described by PDM) and finally after a certain thickness to a "mixed" interfacial and diffusion-limiited regime (Wagner theory). By contrast, under aqueous conditions (Fig. 1-6a; the y-axis is now corrosion rate CR), the CR initially falls rapidly as the scale forms. After approximately 60 hours, a steady state CR is reached where the scale growth rate matches that of scale dissolution. 22 Table 1.2: Reactions describing the basic unit processes used to model the kinetics of iron sulfide layer formation, growth and breakdown. After the Point Defect Model. [32] Partial process Metal injection Reaction Parameters 00 − Fe + VFe ⇔ Fe Fe + 2e ν, φ Fe ⇔ Fe Fe + VS•• + 2e− ν, φ (1) (2) Cation diffusion Fe Fe ⇔ Fe Fe ∗ DFe (T, φ) Anion diffusion (1) SS ∗ DS (T, φ) Dissolution 2+ Fe Fe + SS + 2H + ⇔ H2 S + Fe(aq) Surface vacancy formation 2+ Fe Fe ⇔ VFe + Fe(aq) ⇔ (2) SS S2 + 2H + ⇔ VS•• + H2 1.4.2 Kp 00 ∆H Sf Unit processes controlling Fe-S passive layer behavior None of the models described above can successfully predict the transient and steady state passive layer behavior in Figure 1-6 a priori. Moreover, localized breakdown phenomena such as pitting remain largely unaccounted for. The overarching motivation for this project is to work towards a deterministic model. A robust model for such a complicated electrochemical system must be able to predict a priori degradation rates and trends under realistic pipeline corrosion conditions, using parameters that can be measured and compared through experiment. Moreover, the model’s predictive capabilities should serve as a platform for designing more protective materials and chemical inhibitors. Its characteristics should be: 1. Non-empirical: able to calculate passive layer growth rates L(t), transient and steady state corrosion currents i(t) and passivity breakdown (e.g. pitting probability P(t) and stable pit growth rate dx/dt) using a "bottom-up" approach, i.e. by simulating unit steps such as atomic diffusion and charge transfer reactions. The model should have as few empirical (fitted) parameters as possible. 2. Predictive: able to make accurate predictions that can be related to corrosion tests and cover a range of environmental conditions such as temperature, pressure, sulfur concentration and applied potential. 3. Modular: complexity can be added or removed by considering unit processes as separate elements of the model. 4. Bridges length- and timescales: can make predictions from the atomic- and nanometer scale in fractions of a second, to the macroscopic level over timescales of hours and years that can be compared to laboratory experiments and industry tests. As of writing in November 2014, a global Fe-S passive film model is under continued development by Aravind Krishnamoorthy and others in the Yildiz and Van Vliet groups at MIT, and will be presented in more detail in Aravind Krishnamoorthy’s PhD thesis, expected 2015. In Figure 1-7, the processes considered in the model are outlined schematically. In Figure 1-7a, some of the key atomic steps investigated in these two theses thesis are shown; these include ionic diffusion, surface exchange (interfacial charge transfer reactions) and vacancy formation at surfaces under reducing conditions. Figure 1-7b, on the other hand, depicts other important phenomena affecting the stability of passive films. Although beyond the the scope of this thesis and a rudimentary version of the proposed model, they are included for completeness. Finally, the basic chemical reactions that define sulfide passive layer behavior are listed in Table 1.2. 23 (a) Elementary rate-limiting processes in iron sulfide films HSELECTROLYTE Fe-S LAYER H2 x L Dissolution Hads Fe2+ S2- φ = - dV/dx 2e(cathodic) S2- H + VFe (anodic) 0 STEEL 2e- φ Potential drop across film Ionic diffusion Interfacial charge transfer Vacancy formation & pitting (a) Other considerations for generalized passive layer model Solution chemistry & thermodynamics ELECTROLYTE + Chemical inhibitors - Flow/erosion Pit chemistry X- + Fe2+ + ↔ FeX H2S ↔ H + HS Fe-S LAYER g.b STEEL pearlite Phase stability, transformations Voids/pores Interfacial stresses, film delamination Alloy microstructure & chemistry Figure 1-7: Schematic of unit processes in Fe-S passive layers: (a) basic kinetic rate-limiting corrosion processes for a homogeneous iron sulfide passive film on steel. (b) additional considerations for multiphase films, substrates, and different electrolyte chemistries. In addition, mechanical effects such as flow can influence corrosion rate. 1.4.3 The need for experimentally-derived parameters The high-level, proposed approach to formulate a non-empirical iron sulfide passive layer model from fundamental atomic processes is outlined in Figure 1-8. Activation barriers Ea and rate constants νo can be obtained by performing both ex situ and in situ experiments, combined with predictions from first principles by Density Functional Theory (DFT). In situ refers to corrosion tests using sulfidic electrolyte solutions; none were completed in the course of this thesis, but we will come back to potential experiments in Chapter 5. In their simplest incarnation, experiments are designed to measure a given kinetic process as a function of temperature, yielding the desired parameters through fitting to a universal Arrhenius law of the form: P = νo exp[− Ea ] kB T (1.4) where P is the probability (yielding average macroscopic rate) of the given process, and kB is Boltzmann’s constant. A kinetic Monte Carlo simulation calculates the dynamics of local unit steps at the metal-passive layer and passive layer-electrolyte interfaces. 24 Passive layer model: global strategy and thesis contributions Experiments “ex situ” Growth Chapter 2 Reactivity Chapter 3 Density Functional Theory “in situ” Stability Chapter 4 Activation barriers Ea Rate constants νo P = νo exp[- Ea / kBT] Thermodynamics (literature) kinetic Monte Carlo Phase Field model Laboratory tests Field experience Figure 1-8: Overview of strategy to construct a non-empirical passive film model. Contributions of this thesis are shown in red font. Experiments and first-principles Density Functional Theory (DFT) are employed to calculate activation barriers Ea and rate constants νo for the unit processes outlined in Figure 1-7. These are fed into a probabilistic model on the atomic scale, which updates a kinetic Monte Carlo (kMC) simulation. The microscopic fluctuations tend towards certain macroscopic states under an input thermodynamic bias calculated from temperature, chemical potentials, etc. This macroscopic behavior is calculated through a Phase Field model at greater lengthscales than kMC. Finally, the model should allow comparison to lab- and field tests for re-optimization and iteration. The kMC is coupled to a macroscopic Phase Field model which is able to update the structure on greater length- and timescales. Finally, the proposed model will be merely an academic exercise unless it can be usefully compared and contrasted to real results on sour corrosion from laboratory tests and the field. In future studies, these should be fed back into experimental design to improve the collection of useful, non-empirical model inputs. 25 1.5 Thesis goals and organization Having introduced the background and general strategy, the following three chapters dig deeper into some of the phenomena depicted in Figure 1-7a, using the model iron sulfide phases Fe1-x S and FeS2 . No in situ corrosion tests are conducted, nor do we consider the formation of sulfides on steel or iron substrates. Instead, this thesis comprises three individual case studies that test the hypothesis that we can construct a global model by considering local, atomic processes in isolation. The physical chemistry of pure iron sulfide surfaces and bulk processes are investigated and, where appropriate, the implications of the findings beyond passive film behavior are discussed. • Chapter 2 ("Growth") addresses the unit processes of cation diffusion and surface exchange as rate limiting mechanisms in Fe1-x S (pyrrhotite). This phase is interesting due to a magneto-structural transition at 315 o C, below which Fe diffusion had previously not been studied. Is there an effect from magnetic and structural ordering on Fe diffusion in Fe1-x S? Under what conditions does diffusion in pyrrhotite limit the rate of scaling or corrosion with respect to surface reactions, and vice versa? Both phenomena are studied as a function of temperature to extract average activation energies which are cast into a rudimentary pyrrhotite scale growth model. • Chapter 3 ("Reactivity") explores the effect of surface electronic structure on redox charge transfer, by quantifiying the electronic band gap and surface states on FeS2 (100) as a model sulfide phase. For semiconducting passive layers, electron exchange with redox species in the environment occurs by horizontal transfer from (to) occupied (unoccupied) states. How does the surface electronic structure differ from that of the bulk? To probe locally at the surface, the scanning tunneling microscope (STM) is used in spectroscopy mode (STS). However, there is no well-defined protocol for interpreting STS results. A systematic methodology is developed to identify tunneling current contributions from surface states which mediate charge transfer during reactions, extendible to other similar materials. Finally, the implications of these findings beyond passive layer electrochemistry are discussed; for example, regarding earth-abundant, FeS2 photovoltaics for energy production. • Chapter 4 ("Stability") describes the dynamics of point defect formation, clustering and pit initiation on the FeS2 surface. The inherent protectiveness of a passive layer relies on the physicochemical barrier remaining intact against chemical, electrochemical or mechanical stimuli. A key postulate of the PDM is that passivity breakdown originates from vacancy condensation on the cation sublattice of the barrier layer. Can we form a mechanistic picture of this process, informed by experiment? The formation of sulfur defects at elevated temperatures and in UHV on FeS2 is obsered and quantified using x-ray photoelectron spectroscopy (XPS) and in situ STM. The experimental results are used to inform a kinetic Monte Carlo (kMC) simulation for vacancy condensation that predicts the formation of nanocavities at passive film interfaces that can serve as pitting initiation sites. • Chapter 5 "Conclusions" summarises the findings and key contributions from these three case studies above. The insights are contextualized under the scope of the ambitious, global passive film model proposed in Chapter 1. The contribution of such precise physico-chemical studies on atomic processes in model passive layers is revisited and appraised. Finally, the steps required for the future development of a more practical, deterministic passive layer model are discussed, including suggestions for in situ experiments under realistic aqueous conditions. 26 Chapter 2 Growth: cation diffusion and surface exchange as rate-limiting mechanisms in pyrrhotite, Fe1-xS Synopsis Cation diffusion and the surface exchange of sulfur, constituting the predominant kinetically-limiting processes in the growth of pyrrhotite (Fe1-x S) in sour environments, are investigated on model thin film and bulk samples. Diffusion is studied via two methods to understand the influence of the spontaneous magnetic and structural order-disorder transition at the Néel temperature T N of 315 o C in Fe1-x S. The self-diffusivity of iron * DFe in pyrrhotite above T N follows an Arrhenius law with an activation energy Ea = 0.83 eV. First, magnetokinetic measurements of the antiferromagnetic to ferrimagnetic "λ-transition" between 180-210 o C yield Ea = 1.1 eV for cation diffusion. Second, we confirm this higher Ea in magnetic pyrrhotite using 57 Fe tracer diffusion measurements below T N , obtained via secondary ion mass spectrometry (SIMS). These demonstate a downwards deviation from the extrapolated, paramagnetic Arrhenius trend by up to two orders of magnitude at 150 o C. The results are described by a magnetic diffusion anomaly, whereby vacancy formation and migration energies for the cation sublattice are increased by approximately 40% over the paramagnetic state, due to spontaneous magnetization. Finally, we study sulfur exchange kinetics using the electrical conductivity relaxation (ECR) technique on Fe1-x S thin films, in which diffusion is very rapid. The chemical exchange coefficient for sulfur incorporation, kox is found to have an activation energy 1.0 eV. Our experimental activation barrier values for these unit processes can be fed into a multiscale corrosion model to predict growth rates for pyrrhotite scales. The similarity in surface exchange and diffusive barriers suggests that the crossover from the former to latter rate-limiting steps should occur at approximately the same film thickness of 100-1000 µm, independent of temperature. The computational work in this chapter, including the development of kinetic Monte Carlo codes, was discussed and carried out in collaboration with Aravind Krishnamoorthy. 2.1 Background and Motivation Pyrrhotite is a stable iron sulfide phase that forms under higher temperature, more sour conditions (Chapter 1). It was selected for this study because iron diffusion is known to be a relatively rapid process in Fe1-x S, due to its high degree of cation offstoichiometry. In fact, the surface exchange of sulfur can constitute a relatively slower 27 (b) o Temperature ( C) -6 900 700 500 400 200 150 300 -10 X -6 -16 -18 x in Fe1-xS 0.14 0.07 0.03 0.013 0.007 0.004 -8 -9 0.8 -12 -14 Non-stoichiometry X -7 Y ? Fryt† Worrell† Sterten† Hobbins‡ Condit‡ Marusak§ 1.0 1.0 1.2 1000/T (K-1) (c) ? -8 Z 1.5 2.5 2.0 -1 log[*DFe] (cm2s-1) log[*DFe] (cm2s-1) -8 log[*DFe] (cm2s-1) (a) 1000/T (K ) Anisotropy Y -9 II c-axis (x ≈ 0) -10 c-axis -11 -12 1.0 2.0 1.5 -1 1000/T (K ) Figure 2-1: Collected literature values of Fe self-diffusivity. (a) experimental ∗ DFe values obtained using different techniques as listed. The aim of the work in this thesis was to measure a reliable set of data below 300 o C. (b) magnified Fryt data from region “X” on the main graph. Variations in stoichiometry x account for up to one order of magnitude variation in ∗ DFe . (c) magnification of Hobbins data (region “Y”): crystal anisotropy can account for half an order of magnitude variation. Data from: Fryt [31] † , Worrell [37] † , Sterten [38] † , Hobbins [39] ‡ , Condit [40] ‡ , Marusak [41] § . († = sulfurization; ‡ = radiotracer; § = magnetokinetic measurements). kinetic process for pyrrhotite growth under certain conditions. [17,36] Solid-state mass transport (diffusion) produces a parabolic time-dependence of film growth of the form: p X ∼ 4Dt (2.1) where X is film thickness and D is a diffusion coefficient. Surface exchange (the reaction between Fe1-x S and H2 S) on the other hand should produce linear kinetics of the form: X ∼ t/kchem (2.2) where kchem is a surface exchage coefficient. The primary aim of this work was to evaluate and compare the fundamental energy activation barriers for these two processes by considering their Arrhenius-like temperature dependence: Em D ∼ Do exp − kB T Eex kchem ∼ ko exp − kB T (2.3) (2.4) where Do and ko are rate constants, Em is the activation energy barrier for migration, Eex is the barrier to surface exchange reactions and kB is Boltzmann’s constant. The development of a multiscale, non-empirical passive layer model as the ultimate motivation for this project requires experimentally-determined values for these unit process parameters. 28 Diffusivity: effect of order-disorder transition? Fe1-x S undergoes spontaneous magnetic and structural disordering above the critical Néel temperature T N of 315 o C. [42] Previous measurements of Fe self-diffusion *DFe in paramagnetic Fe1-x S above T N have been carried out by several authors using thermogravimetric [31, 37, 38] and radiotracer [39, 40] methods, as compiled in Figure 2-1a. An isothermal spread in *DFe of less than one order of magnitude is observed, stemming from variations in stoichiometry (Fig. 2-1a) or crystalline anisotropy (Fig. 2-1b). Despite this, the results above ∼ 300 o C follow a standard Arrhenius law as a function of temperature, given by D = Do exp[−Q P /kB T ], where Do is a prefactor, QP is the activation energy for diffusion in the paramagnetic lattice, and kB is Boltzmann’s constant. However, already in 1974 Condit et al. predicted that vacancy ordering below T N would lead to an increase in activation barrier for Fe self-diffusion. [40] Indeed, the only available *DFe measurements at temperatures lower than ∼ 300 o C, obtained via a unique magnetokinetic method (Figure 2-1a, data labelled "Z"), may substantiate this claim if it weren’t for a lack of auxiliary data. A secondary goal of this chapter is to clarify the role of ordering at T N on Fe self-diffusion in Fe1-x S. Chapter goals and layout The primary aim of this chapter is to to compare the rates of the kinetically-limiting processes of diffusion (bulk) and sulfur exchange (surface) by studying these processes on controlled samples as a function of temperature. A secondary goal is to resolve the uncertainty over the effect of magneto-structural ordering below 315 o C on cation diffusivity. This chapter is split into five main sections, which are summarized below: • Section 2.2 "Introduction to pyrrhotite polytypes and transitions" describes the closely-coupled structural and magnetic properties of Fe1-x S. • Section 2.3 "Diffusion-limited λ transition" explores cation diffusion via magnetokinetic measurements of a single-phase reordering transformation in nonstoichiometric Fe1-x S, reviewing the experiments that gave rise to the data in region "Z" of Figure 2-1. • Section 2.4 "Isotope tracer diffusion measurements" fills in the missing Fe selfdiffusivity data in the temperature range 170-400 o C through direct diffusivity measurements using secondary ion mass spectrometry (SIMS) on Fe1-x S crystals. • Section 2.5 "Sulfur exchange kinetics at the Fe1-x S surface" describes kinetic measurements of the transfer of sulfur from gaseous H2 S to Fe1-x S thin films, under conditions where diffusion is not rate-limiting. • Finally, in Section 2.6 "Outcomes", the experimentally-determined rates of diffusion and surface reaction are compared to predict from first principles the controlling processes in the sulfidation of iron under a range of scenarios. 2.2 Pyrrhotite: polytypes and transitions In this section, the crystallography and basic physical properties of Fe1-x S are reviewed. This phase characterized by a complicated series of vacancy-ordered structures which form below a common ordering temperature of 315 o C. 29 (a) NiAs unit cell (b) Superstructure AB-plane Fe S C c a B = 2√3a A a =2 Figure 2-2: Structural unit cells of pyrrhotite.(a) NiAs-like unit cell, common for all pyrrhotites. (b) ABplane of superstructure, showing only Fe atoms. Dashed rectangle indicates the unit area A = 2a, B = 23a. 2.2.1 Structural and magnetic properties The term ‘pyrrhotite’ encompasses a set of cation-deficient iron sulfides across the narrow composition range 0 ≤ x ≤ 0.125 in Fe1-x S, where the stoichiometric end-member (0 ≤ x < 0.05) is more specifically referred to as ‘troilite’. A comprehensive review of the known structural and physical properties of pyrrhotites was previously written by Wang and Salveson. [43] All pyrrhotites undergo a spontaneous magnetic ordering transition at a Néel temperature T N = 315 o C which is known to be strongly coupled to the ’β-transition’ or order-disorder transition for Fe vacancies, V Fe . [42] For the remainder of this chapter, we therefore use T N to describe the critical, magneto-structural order-disorder transition temperature in pyrrhotite. Below T N the accommodation of relatively large concentrations of V Fe up to 12.5% produces a series of complex, structurally ordered Fe1-x S superstructures, a principal feature of which is the formation of Kagome nets: tetrahedra sharing apexes in all three dimensions which arise to minimize total vacancy-vacancy interaction energy in magnetically-frustrated systems. [44] Such V Fe ordering in turn bestows a remarkably diverse range of low-temperature physical properties, such as magnetism and electronic conduction, which remain ambiguous despite many years of investigation. The basic unit cell for all Fe1-x S compositions is NiAs-type hexagonal with lattice parameters a and c, and space group P62̄c (Fig. 2-2). However, x-ray diffraction refinement [45–48] and high-resolution electron microscopy [49] studies have identified several low-temperature superstructures based on a supercell of dimensions A = 2a, B = 23a and C = c which can take either hexagonal or monoclinic symmetry. The supercell can be described as a layering of iron AB-planes where vacancy segregation to certain planes creates structures requiring different C-axis repeats to complete unit cell symmetry. The generic superstructure is thus described as NC, where N is an integral or non-integral repeat distance in the C-axis. The polytypes of Fe1-x S described below can be understood with the help of Table 2.1 and the phase diagrams in Fig. 2-3. The fundamental magnetic properties of non-stoichiometric pyrrhotites are known to arise from ferromagnetic (↑↑↑↑) alignment of cations within metal AB-layers and antiferromagnetic (↑↓↑↓) coupling between adjacent layers. [52] The inoccupation of an AB-plane by iron vacancies reduces its overall ferromagnetic moment; net magnetism is hence determined by the periodicity of full and partially-unoccupied layers. These phenomena are discussed together in more detail for each phase below. 2C (Troilite. Fe1-x S: 0 ≤ x ≤ 0.05): Troilite adopts a 2C structure with dimensions A = B = 3a and C = 2c. Magnetic moments on Fe atoms lying in AB-planes are anti-ferromagnetically ordered at room temperature but undergo a spin-flip transition (α-transition, see Fig. 2-3) at the 2C/1C solvus - starting at 140 o C for FeS - to an in-plane ferromagnetic order with antiferromagnetic coupling between adjacent AB-planes, imparting net zero magnetization. 30 Table 2.1: Fe1-x S polytypes: composition and structure. [50, 51]. Type Formula Composition range Symm. Supercell unit cell Comments 1C Fe1-x S Full range Hex. A, 2C Elevated temperature disordered form 2C FeS 0 ≤ x ≤ 0.05 Hex. 3A, 2C Troilite NC Fe1-x S 0.8 ≤ x ≤ 0.11 Hex. 2A, NC 5 ≤ N ≤ 11 4C Fe7 S8 x = 0.125 ± 0.05 Mono. 23A, 2A, 4C “Magnetic” pyrrhotite NA Fe1-x S Unknown Hex. NA, 3C High temp. metastable; 40 ≤ N ≤ 90 MC Fe1-x S Unknown Hex. 2A, MC High temp. metastable; 3 ≤M≤4 Beyond this, magnetic spins fully disorder to paramagnetic at T N = 320 ± 5 o C. 1C (high temperature, disordered form): 1C describes the disordered form of pyrrhotite where vacancies are randomly dispersed and hence no long-range order exists. 1C is the established structure for all compositions at temperatures higher than T N at which magnetic order is lost. 1C is antiferromagnetic below T N , and paramagnetic above T N . 4C (Fe1-x S: x = 0.125 ± 0.05): At the iron-deficient extreme of x = 0.125 ± 0.005, pyrrhotite adopts a monoclinic structure in which the stacking sequence of cation layers alternates between fully occupied and -defective in the sequence (. . . FAFBFCFD. . . ), where F denotes a full layer and A-D are defective layers with different in-plane vacancy arrangements. [43] The quadrupling of the c-axis stacking periodicity leads to the designation of this phase as 4C (Fig. 2-4a). In the 4C superstructure, an uncompensated moment between alternating full and vacant sublattices results in net ferrimagnetism which persists up to T N . The temperature-dependent magnetization up to this point is described by standard Weisstype behavior. Powell et al. demonstrated using neutron diffraction that at T N , coincident with magnetic disordering, a structural disordering towards randomly-distributed vacancies with hexagonal, 1C periodicity also occurs. [42] NA and MC (210 o C < T < 320 o C, unknown composition range): There has been some evidence to suggest that intermediate, metastable pyrrhotites exist above ∼210 o C that can be refined with superstructure cell dimensions Ā = NA with 40 ≤ N ≤ 90 and C̄ = 3C. [45, 48] and subsequently at around 260-300 o C as ‘MC’ (A = 2A and C = MC with 3 ≤ M ≤ 4). [45, 53] Due to the large A-axis repeat units and non-integral C-axis repeat units proposed for NA and MC pyrrhotites, respectively, these phases are at best ill-defined and likely comprise a mixture of metastable, ordered solid solutions that on average resemble the supercell structures described above. NC (Fe1-x S: 0.08 ≤ x ≤ 0.11): In the composition range 0.08 ≤ x ≤ 0.11, Fe1-x S forms a more complex set of hexagonal pyrrhotite superstructures known collectively as ‘NC’, where the repeat distance N of 31 (a) FeS-Fe7S8 phase diagram Liquid 1C + 985 Liquid 1C + S(l) o Temperature ( C) 949 1C + FeS2 TN 325 MC 275 225 TC 1C MC + 4C NA NA + 4C 175 Tα 125 75 6C 11C 5C 2C + NC 50 (b) NC + 4C NC 2C + 1C 49 48 4C + FeS2 4C 33.3 47 at% Fe FeS-Fe7S8 magnetism o 225 175 125 Antiferro- Ferri- ↑↓↑↓↑↓ ↑↓↑↓↑↓↑ c-axis 50 c-axis Tα Antiferro- 75 ↑ TN 325 275 ↑ ↑ ↑ Para- ↑ Temperature ( C) 985 949 →←→← Antiferro- c-axis 49 Ferri- ↑↓↑↓↑↓ ↑↓↑↓↑↓↑ c-axis 48 47 at% Fe c-axis 33.3 Figure 2-3: Pyrrhotite structural and magnetic phase diagrams.(a) showing existence ranges of 1C, NC, 4C, NA and MC pyrrhotite superstructures, as described in the main text. Experimentally determined temperatures for the α-, β− and λ−transition onsets Tα , T N and T C , respectively) are also shown. (b) Approximate magnetic structures superimposed on phase fields; arrows refer to plane-by-plane magnetic moment along the c-axis. After: [45, 53, 54] the NiAs subcell may be either integral or non-integral between 5 and 11. [43] Figure 24b shows an idealized 5C structure with a series of full and vacancy-bearing AB-layers, similar to 4C. However, the periodicity for 5C (...-AFFBFCFFDFA-...) has an antiferromagnetic symmetry due to compensation between sublattices. Although several ideal crystal structure solutions such as this have been proposed for integral N values such as Fe9 S10 (5C), Fe10 S11 (11C) and Fe11 S12 (6C), no exact vacancy distributions have been conclusively determined. [46–48] NC pyrrhotites are better characterized by a distribution of probability of vacancy occupancy (Fig. 2-5). The observation of incommensurate c-axis stacking [56] in some pyrrhotite samples makes it likely that intermediates 32 .. . A (a) Ferrimagnetic 4C 0 (b) Antiferromagnetic 5C 0 F F C 4 F Layer # Layer # 2 B 2 4 6 D 6 8 F 8 A .. . 1 0.75 Occupancy Fe2+ VFe 10 1 0.75 Occupancy .. . A F F B F C F F D F A. .. Magnetic moment (size represents magnitude) Figure 2-4: Idealized Fe1-x S superstructures: (a) 4C with alternating full and partially vacant occupancy of AB-layers. An uncompensated magnetic moment results in net ferrimagnetism. (b) 5C with net magnetic compenstation between full and vacancy-bearing layers; this idealized structure is antiferromagnetic. The labels “F” refer to full Fe layers; A-D are vacancy-bearing layers with different in-plane vacancy arrangements. After Vaughan et al. [55] Fe1-xS vacancy-bearing superstructures 4C 5C Fe2+ Probability of VFe occupancy 11C 6C Figure 2-5: Distributions of vacancies in 4C and NC pyrrhotites.Vacancy distributions are represented by the probability of site occupation for 4C, 5C, 11C and 6C pyrrhotites. [57] (non-integral values of N) are simply mixtures of the well-structured 5C, 11C and 6C polytypes. Irrespective, fully-ordered NC pyrrhotites are consistently antiferromagnetic at room temperature due to a net compensation of magnetic moments between vacant and full cation layers. 33 (b) FeS nanowires 6 c-axis 4 2 300 (c) Natural FeS II c-axis Cooling 3 500 0 300 400 500 T (K) 3 2 1 Heating 400 T (K) σ (memu/g) 6 χ (memu/g) σ (memu/g) (a) Single crystal Fe9S10 8 600 0 100 300 500 o T ( C) Figure 2-6: The peak-like λ-transition in NC Fe1-x S (a) single crystals [64], (b) nanowires [61] and (c) natural, geological samples [51]. 2.2.2 The λ-transition in NC pyrrhotites The unusual magnetic properties of pyrrhotite have long been studied for their fundamental interest. [58] The temperature-dependent magnetization of ordered NC pyrrhotites is characterized by the appearance of a peak during heating, centered around 210 ± 10 o C (Fig. 2-6) that is thought to arise from a structural rearrangement towards a ferrimagnetic superlattice. [59] More recently, Fe1-x S nanowires [60,61] and nanodisks [62] that display the so-called λ magnetic transition have been fabricated by different means and the phenomenon has even been proposed for technological purposes such as phasechange magnetic memory. [63] The kinetics of the λ-transition has been studied before using magnetic techniques, notably by Townsend et al. [64] and Marusak et al. [41]. However, despite confirming the λ-transition to be a diffusion-controlled process, a coherent mechanistic description is still lacking. For example, both authors assumed simplified exponential time-dependence for isothermal magnetization kinetics. 2.3 Diffusion-limited λ transition The λ-transition in NC (specifically, 11C and 6C) pyrrhotites was investigated using temperature-dependent and time-dependent magnetization experiments (σ(T ) and σ(t), respectively). During the first heating ramp from 30-350 o C, σ(T ) for the 11C and 6C polytypes undergoes a peak, similar to that observed for 5C by other authors (Fig. 26), attributed to a rearrangement of the vacancy-bearing sublattice via diffusion. Our work described here constitutes the first systematic investigation into the AF-FI transition in Fe1-x S since 1980. An initial attempt to replicate the earlier experiments of Marusak et al. [41] revealed a more complex time-evolution of the ferrimagnetic superlattice. Instead of a simple exponential fit, we demonstrate the magnetokinetics are better modeled by a phenomenological, stretched exponential function of the form: α(t) = 1 − ex p [−(t/τ)n ] (2.5) where τ descrbes a temperature-dependent relaxation time, and n = 0.45 ± 0.05. Moreover, we describe a kinetic Monte Carlo (kMC) simulation of the λ-transition that reproduces the structural evolution on the experimental timescale from an AF to FI lattice under cation vacancy diffusion alone. The kMC results similarly give a stretched exponential time dependence and help understand the transition as a continuous-ordering transformation. A physical basis for the stretched exponential form of the kinetics is discussed. Finally, we show the temperature dependence of τ in Eq.(2.5) yields an activation energy of 1.1 ± 0.1 eV for the λ-transition, which can be taken as the migration energy for cation diffusion in ordered pyrrhotite. 34 (004) (103) (110) (102) (101) (100) 4C Intensity (arb. units) 4C 5C 5C 11C 11C 6C 6C 2C 30 (c) Composition Measured at%Fe (b) (102) peak (a) X-ray diffraction 52 2C 50 48 46 6C 11C 5C 4C 48 50 at%Fe, Arnold et al. (d) 2C 40 50 2Θ 60 70 42 43 44 45 2Θ Figure 2-7: X-ray diffraction of synthetic pyrrhotites. (a) Comparison of as-synthesized 2C (FeS), 6C (Fe11 S12 ), 11C (Fe10 S11 ), 5C (Fe9 S10 ) and 4C (Fe7 S8 ) samples with labelled pyrrhotite peaks . (b) (102̄) peak, (c) the (102̄) peak position is used to estimate composition with reference to the calibration of Arnold et al. [66]. (d) black pyrrhotites were stored in glass vials and were stable without a change in properties over several months. 2.3.1 Methods Preparation of well-ordered 2C, NC and 4C type pyrrhotites Synthetic pyrrhotite samples of different stoichiometry were prepared by reacting the requisite amounts of iron powder (99.999% purity, 200 mesh, Alfa Aesar, Haverhill, MA) and sulfur granules (99.998% purity, also Alfa Aesar) in quartz tubes, sealed under vacuum to 10-3 mTorr. The target stoichiometries were: Fe7 S8 (4C), Fe9 S10 (5C), Fe10 S11 (11C) and Fe11 S12 (6C). The sealed powders were subjected to an initial heat treatment to allow the elements to fully react [42]: 500 o C for 24 hours, then 800 o C for 48 hours, followed by cooling at 0.5 o C/minute to 250 o C and held for 24 hours before removal from the furnace. The products were removed from the quartz tubes, re-ground with a porcelain pestle and mortar, and re-sealed in fresh quartz tubes under vacuum. A second heat treatment was subsequently applied to ensure each sample relaxed into its lowtemperature, equilibrium ordered structure: 800 o C for 72 hours, followed by cooling at 0.1 o C/min to 250 o C, holding at 250 o C for 24 hours and finally cooled to 125 o C at a rate of 0.1 o C/min, at which point the samples were removed from the furnace and allowed to cool to room temperature. Characterization of synthetic pyrrhotites: X-ray diffraction The structure and phase purity of the as-synthesized powders was determined by x-ray diffraction (XRD) using a PANalytical X’Pert PRO XRPD instrument with Cu-kα radiation (Fig. 2-7). All peaks can be attributed to the hexagonal and monoclinic pyrrhotite structures (ICSD references 53528 and 151766, respectively). The position of the (102) peak was used to confirm the iron content of the samples, using the peak position calibration described outlined by Arnold et al. [65,66]. Only the as-synthesized 2C pyrrhotite did not follow the trend in composition; however this phase was not of primary interest to this work and was not used further. Magnetic measurements Magnetic measurements were obtained using a variabletemperature Vibrating Sample Magnetometer (VSM), with an applied field of 10 kOe. Powders of synthetic pyrrhotite weighing approximately 0.02 g were attached to quartz rods using silver paste adhesive. The sample was purged with N2 during the measurement at a flow rate of 15 standard cubic feet per hour (scfh). For time-dependent mag35 Jump probabilities on kMC grid P1 P2 P1 P1 P1 C B A Figure 2-8: Setup of cubic kinetic Monte Carlo (kMC) grid. One superstructure unit cell is shown; hexagonal symmetry was applied by biasing diffusion probabilities. There are four equivalent jumps, labelled P1 and one non-equivalent jump, P2 > P1 . netization measurements, the sample temperature was first raised to the setpoint, followed by turning on the applied field. The lag between reaching the set temperature and recording the first data point was approximately one minute. The instrument was calibrated using a nickel disk of known magnetization. Differential scanning calorimetry Differential scanning calorimetry (DSC) was performed using a a Q2000 DSC (TA Instruments, New Castle, DE) under a dynamically purged N2 environment. Kinetic Monte Carlo simulations Kinetic Monte Carlo (kMC) simulations were performed on a model block of cation-deficient pyrrhotite to test the mechanistic hypothesis that mass transport of vacancies between AB-planes at elevated temperatures can gradually convert an antiferromagnetic lattice to a ferrimagnetic one. The model aimed to simulate the time-dependent magnetization of a superstructure containing randomlydispersed vacancies as it evolves towards a more 4C-like, layer-by-layer alternating occupancy structure. We defined an order parameter based on the ideal 4C pyrrhotite vacancy distribution shown in Figure 2-4a. to continually assess the magnetism of the structure as it evolved in time through diffusive jumps of V Fe alone. The setup of the kMC model and the execution of unit processes are described step-by-step below. 1. A three-dimensional, cubic Ising model block consisting of 20 x 20 x 20 unit cells was set up. Only iron sites were considered, under the assumption that the sulfur sub-lattice is saturated and therefore does not contribute to mass transport. Individual lattice points can be full (1) or vacant (0) only. The hexagonal symmetry of the NC-type pyrrhotite lattice was imposed by biasing the diffusion paths such that a hop in one of the diagonals was 3 times as unlikely (Fig. 2-8). 2. One in every eleven sites was selected to be “0” at random to simulate an antiferromagnetic, 11C lattice. The start point for the lattice is not ordered; however, there is no unique 11C structure established in the literature. Moreover, we found that imposing a rigid initial structure only added to computation time without affecting the time-evolution results. 3. A vacant site in the structure was selected at random and its nearest neighbor (NN) sites are evaluated as potential jump destinations. The only atomic process modeled by the kMC code was the diffusion of V Fe in the a-, b- and c- directions. Occupied NN’s populated a list of diffusive jump locations; specifically, five non-equivalent diffusion paths with different jump probabilities were considered (Fig. 2-8). The probability P for a jump to any of these sites is calculated in a 36 Table 2.2: Thermodynamic values for pyrrhotite compounds. Enthalpy of formation ∆Hf, 298 K relative to the elements in their standard states at 298 K, absolute third-law entropies S298 K at 298 K, and heat capacity functions Cp . Data from: [67] Compound ∆Hf, 298 K (kJ/mol) S298 K (J/mol.K) Cp (J/(mol.K)) FeS (2C) -100.1 60.3 2437.1 − 9.9T + 0.01T 2 − (41.1 × 106 )T 2 Fe11 S12 (6C) -1148.1 755.2 - Fe10 S11 (11C) -1048.5 693.0 - Fe9 S10 (5C) -950.8 623.5 170.6 − 0.5T + 0.0005T 2 − (3.0 × 106 )T 2 Fe7 S8 (4C) -755.4 486.3 140.5 − 0.7T + (3.1 × 10−7 )T 2 − (3.9 × 106 )T 2 sub-routine, based on: (a) the self-diffusivity or intrinsic activation barrier to migration Em , in the absence of an imposed driving force; (b) an energy bias due to a thermodynamic driving force towards ordering, Etherm , as described below in step (4); (c) a bias due to the magnetic energy in the applied field of 10 kOe, Emag : Emag E ther m Em exp − exp − P = ν. exp − kB T kB T kB T (2.6) 4. Subsequent to each diffusive jump, the occupancy of the seven adjacent supercell AB-planes above and seven below the elected vacant site was assessed (Fig. 28). The closeness of the layer-by-layer occupancy of this volume was compared to the idealized ferrimagnetic 4C-type lattice occupancy (. . . full, vacant, full, vacant. . . ) and was quantified by taking the root mean square (RMS) difference from the ideal 4C structure occupancy. 5. The energy landscape of the simulation was biased such that the structure is thermodynamically driven to evolve towards a more 4C-like structure. A linear bias of the form E ther m = AΘ + B was used, where the parameters A and B provide the difference in Gibbs free energy ∆G between a disordered vacancy structure and the 4C structure at a given temperature of interest, and Θ is the order parameter we assign to the system, with 0 assigned to a randomly-ordered antiferromagnetic lattice and 1 representing the full 4C structure. The free energy G of each phase at a given temperature was approximated via the relation: G = ∆H f ,298K + ZT ZT C p d T − T S298K − 298K Cp T dT (2.7) 298K where ∆H f , 298 K is the formation enthalpy at 298 K, S298 K is the entropy at 298 K and Cp the heat capacity. Values for these thermodynamic parameters, as listed in 2.2, were obtained from Walder and Pelton. [67] 2.3.2 Results and discussion Temperature-dependent magnetization Temperature-dependent magnetization σ(T ) results for the 4C, 11C and 6C Fe1-x S samples are shown in Figure 2-9. The 4C sample followed typical FI, Weiss-type behavior 37 1 (c) 6C (Fe11S12) 2 2 2 M (emu.g-1) 20 (b) 11C (Fe10S11) M (emu.g-1) -1 Magnetization (emu.g ) (a) 4C (Fe7S8) 1.5 3 15 4 3 2 10 1.5 1 0.5 150 200 250 300 350 T (oC) 5 4 3 2 5 0 0 1 100 200 300 o Temperature ( C) 0 1 0.5 150 200 250 300 350 T (oC) 1 100 200 300 o Temperature ( C) 0 100 200 300 Temperature (oC) Figure 2-9: Temperature-dependent magnetization σ(T ). (a) 4C, (b) 11C and (c) 6C pyrrhotite samples. Multiple consecutive forward and reverse sweeps between 30-330 o C, as labelled 1-5 on each of the graphs, were performed until no change from the previous sweep was observed. On (b) and (c) the inset graphs show a magnified region around the peak observed on the first sweep. The heating and cooling rates were both 0.2 o C/min; the applied field was 10 kOe. 11C (Fe10S11) Hysteresis Curves 3 σ (memu/g) 2 150 oC 220 oC 280 oC 325 oC 1 0 -1 -2 -3 -10 -5 0 5 10 -10 -5 0 5 10 -10 -5 0 5 10 -10 -5 0 5 10 H (kOe) H (kOe) H (kOe) H (kOe) Figure 2-10: Magnetization vs. applied field (σ-H). Obtained from 11C pyrrhotite samples held at different temperatures, as indicated. Magnetization reaches a maximum at 220 o C in this series. Inset on each curve is the region around zero applied field; the absence of hysteresis at 325 o C demonstrates paramagnetism. as expected up to T N = 315 o C. This behavior is reversed upon cooling and can be repeated without hysteresis (three heating-cooling cycles are shown in Fig. 2-9a). The as-synthesized 5C sample produced similar results to 4C, indicating that the desired, equilibrium antiferromagnetic (AF) superstructure had not been formed. The 5C powders were not used further of this work. However, the other NC samples, 11C and 6C, displayed markedly different σ(T) behavior (Figs. 2-9(b) and (c)). On the first heating cycle, labelled (1) in the figures, σ started close to 0 emu/g, consistent with AF ordering. The λ-transition peak first appeared at ∼ 180 o C, with a maximum at 210 o C. This is also observed in Figure 210 which shows a series of magnetization vs. applied field (σ-H) curves at different temperatures along the peak. The magnitude of σ for a given H first rose then fell between 150 and 280 o C. At 325 o C the magnetic behavior is paramagnetic, evidenced by a lack of hysteresis. However, during cooling back from 350 o C, the λ peak was not reproduced and σ increased with Weiss behavior back to room temperature. Multiple, repeated heating/cooling cycles (2-5) as indicated on the curves only served to increase overall magnitization further. The maximum room temperature magnetization σRT reached by 11C after several experimental cycles (∼ 13 emu/g) was greater than that of the 6C sample (∼ 10 emu/g); neither reached the maximum of the 4C sample (σRT ∼ 22 emu/g). Thus the final σ for a given structure is limited by the availability of iron vacancy V Fe to maximize the magnetic asymmetry between vacancy-bearing and full layers. 38 Time-dependent magnetization The fact that several heating cycles were required to fully convert the 11C and 6C Fe1-x S to a metastable, FI superstructure (Figs. 2-9b and c) suggests that the kinetics of the λ-transformation are not instantaneous. To understand the transformation kinetics in more detail, we performed isothermal, time-dependent magnetization σ(t) measurements on samples of 11C at various temperatures between 140-220 o C. Short-timescale kinetics (< 10 minutes) and longer-timescale behavior (up to several hours) are discussed separately. Figure 2-11a shows the results of heating one sample consecutively in steps of 10 o C, holding at each temperature for 10 minutes (solid line series). It can be seen that σ in this series was history-dependent: each time the sample was heated, the transformation proceeded from the end-point reached at the previous temperature. Conversely, the dashed curves in Fig. 2-11a were obtained on fresh samples heated directly from 30 o C to the set temperature of 180-220 o C as indicated. The peak shape was also reproduced, but a lack of transformation history means that σ did not reach the same level as the sample subjected to consecutive heating steps. We also observe from Fig. 2-11a that a large proportion of the rise in σ at a given temperature occured very rapidly, within the first 5 minutes and is evidently at least partially reversible, since the curves overlap. The initial reversibility is more clearly seen in Figure 2-11b where the σ(t) of an 11C sample held at 210 o C for 4000 s continuously is compared against that of another sample which was cooled to room temperature three times sequentially at 20 minute intervals between heating to 210 o C. We can infer from this that much of the rapid, initial increase in σ t at a given temperature arose from a small motion of V Fe between adjacent planes, easily reversed upon cooling. However after a few minutes the gradient (dσ/d t) became less steep, i.e. the maximization of σ requires a more coordinated re-shuffling of the vacancy sublattice that remains macroscopically irreversible, or metastable, once cooled back to room temperature. To understand the transformation kinetics in more detail, we performed isothermal, σ(t) measurements on samples of 11C at various temperatures between 140-220 o C for longer times up to 10,000 s. The kinetics of the λ-transition in 5C Fe1-x S, studied via thermomagnetic techniques, has previously been shown to be limited by V Fe diffusion. [64] Assuming exponential growth in σ during the λ-transition, iron self-diffusion coefficients on the order of 10-17 cm2 s-1 have been found by magnetokinetic methods within this transition temperature range [41] that are inconsistently low compared to extrapolated diffusivities obtained from high-temperature sulfidation [30, 31] and radiotracer diffusion studies [O(10-14 cm2 s-1 )]. [40] However, during an initial attempt to replicate these experiments we found that our magnetokinetic data were not described or fit by simple exponential functions. Instead, the stretched exponential in Eq. (2.5) provided a more accurate description (Figure 2-12). Isothermal σ t data were also collected at several temperatures along the λ-transition for longer times up to 10,000 s. Magnetization was converted to ‘phase fraction’ of FI ordering, α F I , according to: αF I = σ t − σi σ f − σi (2.8) where σ t is the measured magnetization at time t, σi is initial magnetization at t = 0 and σ f is final magnetization assuming the transition were allowed to proceed to completion. σ f values for the different temperatures were therefore obtained from curve number (4) in Figure 2-9b, i.e. the maximum FI magnetization at a given temperature T. 39 (a) Short time-dependent magnetization 3.5 210 oC Magnetization (emu/g) Same sample Fresh samples each run 3 o 200 C 220 oC 2.5 190 oC 2 o 180 C o 1.5 170 C o 160 C 150 C 140 oC o 1 (b) 0 10 20 30 40 50 60 70 80 90 Time from start of experiment (min) Reversible initial magnetization 5 o Magnetization (emu/g) 210 C 4 3 2 3 2 Continuous Sequential 1 1 0 1000 2000 Time (s) 3000 4000 Figure 2-11: Reversible magnetic transformation at short timescales. (a) Isothermal, time-dependent magnetization σ(t) of 11C pyrrhotite at different temperatures around the λ transition. The grey, solid curves were obtained on the same sample, consecutively heated to the indicated set temperature, held for 10 minutes and cooled to room temperature (“add-on” magnetization). The dashed lines correspond to a series where each temperature measurement was performed with a fresh sample. (b) σ(t) of two 11C samples: one held continuously at 210 o C for 4000 s (dashed line). The other is sequentially heated to 210 o C three times, with cooling to room temperature between each step. 2.3.3 Continuous re-ordering of ferrimagnetic superlattice Figures 2-13a and 2-13b display the results for the experimental and simulated λtransition magnetokinetics, respectively. For the experimental data, a common value of n = 0.45 ± 0.05 was found to fit all curves reliably. For the kMC results, n = 0.67 ± 0.05; a sensitivity analysis and details of fitting procedures are provided in Table 2-21. Nevertheless, the kMC model, based solely on cation diffusion, accurately replicated the stretched exponential form of the experimental result. The parameters τ and n in Eq. (2.5) are not indicative of any specific atomic mechanism. Generally, τ represents a temperature-dependent relaxation time and n determines the lengthening of τ as the transition progresses (i.e. a deceleration in transition kinetics). Stretched exponential magnetokinetics of this form have been observed for Li2 (Li1-x Fex )N (n = 0.4-0.8) at low temperatures < 20 K, related to finite magnetic moment relaxation. [68] However, this is an unlikely explanation for the kinetics observed in this work at elevated temperatures, where magnetic relaxation should be instantaneous. Alternatively, we can think 40 (b) kinetic Monte Carlo (210 oC) 1 1 0.8 0.8 0.6 0.6 α α (a) Experiment (210 oC) 0.4 0.2 0 0.4 Data Exponential fit Stretched exp. fit 0 5000 10000 15000 0.2 20000 0 0 5000 Time (s) 10000 15000 20000 Time (s) Figure 2-12: Best fits to exponential equation α(t) = 1 − exp [− (t/τ)] and stretched exponential equation α(t) = 1 − exp [− (t/τ)n ] for (a) experimental magnetization data obtained at 210 o C and (b) kinetic Monte Carlo simulated data at the same temperature. Table 2.3: Best fit parameters n and τ in the fitting expression α(t) = αo [1 − exp [− (t/τ)n ]] + (1 − αo ) for the experimental and kinetic Monte Carlo results. Temp. (o C) n t αo Experimental: free parameters 180 0.48 548 ± 9 1.00 190 0.42 134 ± 2 0.99 200 0.40 78 ± 2 0.89 210 0.44 66 ± 3 0.90 Experimental: fixed n 180 0.45 ± 0.05 410 ± 211 1.00 190 0.45 ± 0.05 186 ± 95 0.98 200 0.45 ± 0.05 127 ± 65 0.86 210 0.45 ± 0.05 76 ± 39 0.89 kinetic Monte Carlo: free parameters 180 0.70 737 ± 38 N/A 190 0.73 651 ± 40 N/A 200 0.66 275 ± 19 N/A 210 0.62 150 ± 8 N/A of the λ-transition as a classical nucleation and growth process, described by the semiempirical Johnson-Mehl-Avrami-Kolmogorov (JMAK) relation that takes the same form as Eq. (2.5). For example, the first-order antiferromagnetic-ferromagnetic transition in FeRh has been described by a JMAK equation with n = 0.86. [69] The Avrami exponent n is typically temperature independent and may provide information about the nucleation and growth mechanisms. [70] A value of n = 0.5 to describe the formation of FI phase out of a homogeneous AF lattice would suggest one-dimensional growth with zero nucleation rate, implying small nuclei of FI phase were pre-dispersed in the 41 (a) Experimental ferrimagnetic fraction 1 -3 210 200 190 180 1.1 ± 0.1 eV Data Fit T oC -1 ln( τ ) 0.8 -4 -5 210 oC -6 0.6 1/T x 10-3 (K-1) 200 oC α -7 0.4 190 oC 0.2 180 oC 0 0 (b) 5000 Time (s) 10000 kinetic Monte Carlo results 1 α 0.8 210 oC 200 oC Data Fit 190 oC 0.6 T oC 0.4 180 oC 210 200 190 180 -1 ln( τ ) -5 0.2 -6 -7 0 0 1.1 eV 5000 Time (s) 1/T x 10-3 (K-1) 10000 Figure 2-13: Long-timescale isothermal magnetization. (a) Transformed ferrimagnetic volume fraction α measured over 10,000 s at four temperatures as indicated. We show every 20th point of the raw data as well as a best fit line to the phenomenological, stretched exponential relation α(t) = 1 − exp [− (t/τ)n ], with standard deviation error. Inset: Arrhenius fit of the temperature-dependent fitting parameter τ−1 , with a slope corresponding to a transformation activation energy of 1.1 ± 0.1 eV. (b) kinetic Monte Carlo (kMC) reullts for magnetization transformation at the same temperatures, and fit to a similar stretched exponential fit. Inset: corresponding activation energy of 1.1 eV calculated from computational τ−1 . equilibrium 11C lattice prior to transformation, and that the observed "growth" in α arises from the change in the magnetization order parameter along one axis alone (for example, the c-axis). However, we believe the microscopic transition mechanism can be better described as a diffusion-limited, continuous re-ordering process akin to spinodal decomposition. As such, the FI phase grows out of the AF lattice via an augmentation in small, layer-by-layer vacancy occupancy fluctuations. A second-order transition of this type is consistent with a continuity in enthalpy H but discontinuous heat capacity ∂ H/∂ T during the λ-transition, as measured by differential scanning calorimetry (Figure 2-14). The absence of a pronounced discontinuity in heat Q confirms that the diffusive rearrangement of the λ-transition is second order; that is, the new state of increased symmetry develops continuously from the highly-ordered, lower symmetry 11C or 6C phases. The subsequent, spontaneous disordering at T N is then 1st -order, similarly to the 4C case, but since an already higher degree of disorder exists in the 6C and 11C lattices which have not had time to form into a perfect alternating-layer FI structure, much less diffusion (and therefore latent heat) is required to undergo the 42 Differential Scanning Calorimetry Heat Flow (mW/g) (a) Heat Flow (mW/g) (b) 500 4C Exo. 0 -500 Endo. 200 11C Heat Flow (mW/g) 30.1 J/g 316 oC Heating Cooling 100 2.2 J/g 4.9 J/g 1.4 J/g 4.7 J/g 0 -100 o 210 C Heating -200 (c) 30.5 J/g 314 oC Cooling 200 6C 316 oC Cooling 100 2.2 J/g 5.0 J/g 2.5 J/g 5.0 J/g 0 -100 210 oC Heating -200 50 100 150 200 250 o Temperature ( C) o 316 C 300 350 Figure 2-14: Differential scanning calorimetry (DSC) results from (a) 4C, (b) 11C, (c) 6C synthetic pyrrhotite samples. The heating rate was 10 o C/min; endothermic heat flow is negative on the y-axes. The integrated peak areas for heat uptake or evolution events, in units of J/g and demarcated with dashed lines, are shown on each figure next to the event. full disordering transition. Figures 2-15a-e shows a series of magneto-structural order parameter Θ distributions on kMC lattice points, obtained at various times along the λ-transition at 200 o C. Rather than observing the formation of discrete FI nuclei (red dots in the figure) which grow along one axis and would thus be consistent with the JMAK interpretation, we instead visualize the emergence of regions of intermediate Θ that gradually spread across other lattice points diffusely. This simulated phenomenon is also clearly seen in Figure 2-15f, where we see small fluctuations in layer-by-layer vacancy occupancy augment with time into an alternating-plane, FI structure. The stretched exponential fits in Eq. (2.5) to our σ(t) data take the same form as the Kohlrausch function [71], commonly used to describe non-equilibrium dynamics in disordered condensed matter such as dielectric relaxation [72], relaxation in soft matter [73] and diffusion in complex systems, incluiding glassy materials and H migration in amorphous Si. [74] Although generically semi-empirical, a number of mathematical derivations for the Kohlrausch function have been put forward which provide a more physical basis for the ubiquitously observed stretched exponential behavior. [75] Kohlrausch behavior can arise in the presence a small energy distribution of traps in systems with long-range correlations; this would lead to deviations from "random walk" Brownian motion diffusion in amorphous structures. Alternatively, there may be a time-dependence in populating different energy traps, such that relaxation occurs in stages. [76] Although we do not wish to compare directly the stretched exponential magnetization observed in the λ-transition here to Kohlrausch descriptions of decay in disordered materials, perhaps some parallels may 43 (a) 0 s (b) 7,000 s (c) 12,000 s (d) 45,000 s (e) 100,000 s Order parameter Θ 1 FI c a 0 AF (f) Occupancy 0.95 average occupancy = 0.909 0.9 0.85 0 0s 2 12,000 s 4 6 100,000 s 8 10 12 Layer # along c-axis z-axis 14 16 18 20 Figure 2-15: Continuous re-ordering towards ferrimagnetic state. Local order parameter Θ calculated for individual lattice points in the kinetic Monte Carlo (kMC) simulation at 200 o C, where 0 refers to fully antiferromagnetic (AF) and 1 is ferrimagnetic (FI), corresponding to: (a) 0 s, (b) 7000 s, (c) 12000 s, (d) 45000 s and (e) 100000 s along the λ-transition. (f) layer-by-layer occupancy for 20 planes along the c-axis from the kMC results obtained at different times. As the transition progresses, the difference in occupation between adjacent layers becomes more pronounced. be drawn which give an intuitive physical insight into the phenomenon observed both experimentally and in the kMC results of this work. We can dismiss the possibility of a distribution in traps since Brownian motion was inherently assumed in our kMC model. On the other hand, a time-decay in the rate of magnetization evolution may be more coherently explained by a combination of rapid and subordiante, slower processes. In other words, given sufficient thermal energy, cation vacancies may migrate rapidly to adjacent planes under a large thermodynamic driving force and break the compensating AF lattice symmetry. This leads to small regions with localized FI ordering that nonetheless contribute a large increase in σ on the scale of 10-100s of seconds. At longer times, however, the formation of the optimal FI lattice structure for the available vacancy concentration requires a more labored rearrangement of vacancies into long-range order, decelerating the growth in σ. Only inter-layer V Fe hops contribute to a rise in magnetization. We compared the ratio of inter- to intra-layer hops during the λ-transition in the kMC model and found that it decreased over time. This is consistent with a relaxation in the exponential kinetics for σ; the further the transition progresses, the lower the driving force for vacancy segragation and the smaller the probability of inter-layer jumps. Finally, we turn to the significance of the activation energy of 1.1 ± 0.1 eV measured by fitting the experimental σ(t). A migration barrier for diffusion of 1.2 eV was originally cast into the kMC model. An analysis of the resulting kMC data using the same fitting procedure as for the experiment returned an apparent barrier value of 1.1 eV, confirming that the major rate-limiting step is cation diffusion. We therefore take 1.1 eV to represent a slight underestimate to the activation barrier Em to Fe self-diffusion in magnetically-ordered Fe1-x S superstructures. The overall measured activation energy includes a thermodynamic bias for the transformation on the order of 0.1 eV at 200 o C (calculated from the data in Table 2.2), which serves to lower the diffusion barrier 44 Sulfidation of iron Fe1-xS Marker Iron Figure 2-16: Cross section of sulfide scale formed during the sulfidation of Fe for 20 mins at 800 o C. An inert (Pt) marker indicates the position of the original iron-vapour interface. (100x) [30] slightly. 2.3.4 Conclusions In conclusion, we have investigated the antiferromagnetic to ferrimagnetic λ-transition in NC-type pyrrhotites via magnetokinetic experiments and kinetic Monte Carlo simulations. In contrast to previous reports, the transformation is found to follow a stretched exponential time-dependence. These experimental and computational results together support a description of the λ-transition as a nucleation-free, continuous reordering via diffusion on the cation sublattice. Magnetization initially rises rapidly due to small, localized displacements, but a full optimization of the ferrimagnetic superstructure is a more complex process that emerges only at longer timescales. The migration energy barrier for Fe in magnetic pyrrhotite is shown to be approximately 1.1 eV. The results are helpful in predicting the barrier properties of pyrrhotite scales on ferrous alloys that are exposed to sulfur environments, for example in oil and gas systems. Further, the eludication of the kinetics of the λ-transition encourages continued studies to identify practical applications for this interesting magnetic phenomenon in synthetic Fe1-x S, for example magnetic switching or data recording devices based on earth abundant chemical elements. 2.4 Isotope tracer diffusion measurements Consistent with the metal-deficient nature of Fe1-x S, the diffusion of cations is known to be the primary solid state mass transport mechanism in pyrrhotite scales in both aqueous [36] and dry [29–31] sulfide corrosion conditions. The self-diffusion coefficient for sulfur in Fe1-x S is known to be several orders of magnitude lower than for Fe2+ : *DS = O(10-11 ) cm2 s-1 at 1000 o C compared to O(10-11 ) cm2 s-1 for *DFe . [40]. Hence an inert marker placed at a steel-pyrrhotite interface will remain at the interface as the Fe1-x S scale grows by the outward migration of cations (Fig. 2-16). In the previous study, an activation energy of 1.1 ± 0.1 eV was measured for a diffusion-limited vacancy superstructure rearrangement process in ordered pyrrhotites. This should represent the barrier for Fe migration through a Fe1-x S lattice in which a thermodynamic driving force exists for ordering. In this section, iron self-diffusion *DFe is measured directly in bulk crystals of Fe1-x S by means of iron-57 isotope tracer measurements using secondary ion mass spectrometry (SIMS). 45 (b) (102) peak 30 (110) (103) (101) 40 50 2Θ 60 o (004) 43.65 (101) (100) (100) Intensity (arb. units) (102) (a) Natural Fe1-xS XRD 70 42 43 44 45 2Θ Figure 2-17: Cu-kα powder XRD pattern from a representative sample of research grade natural pyrrhotite. (a) Hexagonal Fe1-x S peaks are labelled on the diagram. Inset: picture of crystals; approximately 5 x 5 x 1 mm3 polished to < 50 nm roughness. (b) the hexagonal Fe1-x S (102) peak’s position of 43.65o indicates an average composition of 48.0 at% Fe. 2.4.1 Methods Sample preparation Natural pyrrhotite crystals were used for diffusion studies.1 Research-grade natural crystals of pyrrhotite were obtained from Ward’s Science (Rochester, NY). Individual specimens were approximately 5x5x5 mm3 and were sourced from North Bend, WA. A representative sample of several individual crystals from the package was ground together into a powder in a porcelain mortar for XRD phase identification (Fig. 2-17a). Hexagonal pyrrhotite was found to be the only phase present. The position of the (102) peak (Fig. 2-17b) was used to estimate an average composition of 48.0 at% Fe, corresponding approximately to Fe11 S12 (6C). [66] The chemical composition of the natural crystals was investigated using energydispersive x-ray spectroscopy (EDS) in a JEOL 6610LV scanning electron microscope (SEM), to check for any large inclusions that may have affected the reliability of the diffusion studies. Figure 2-18 shows a series of chemical maps of a randomly-selected but representative surface, indicating that the majority elements were sulfur and iron. Some minor inclusions of oxide were present in thin seams. An search for potential metallic impurities such as Si and Al came up negative within the detection limit of the instrument (approx. 0.1 wt% [77], implying that the inclusions are likely iron oxides. These were deemed sparse enough to not greatly affect the accuracy of the subsequent diffusion studies. The self-diffusion coefficient of iron in Fe2 O3 is very small below 500 o C (< 10-28 cm2 s-1 . [78] As-received crystals were polished using 1200, 2000 and 4000 grit sandpaper to achieve a flat surfaces followed by an extended polish of 5-10 minutes using 50 nm Al2 O3 suspension. The resulting average surface root mean square (RMS) roughness, as measured by Atomic Force Microscopy (AFM) was < 50 nm; in some 50x50 µm areas RMS roughness was as low as 17 nm. Temperature-dependent magnetization studies as described in Section 2.3.1 confirmed the natural pyrrhotite samples to be antiferromagnetic at room temperature and 1 In fact, diffusion studies were attempted using three different types of pyrrhotite samples: Chemical vapor deposited (CVD) thin films (∼ 200-500 nm thickness, Appendix B), Sputter deposited thin films (∼ 1000 nm thickness, Section 2.5.1) and polished natural single crystals (100’s of µm thick). In this chapter, only the data collected from the natural single crystals is included, since they are considered the most reliable from this study. 46 Chemical composition of natural pyrrhotite samples (a) (c) SEM (e) (g) S Fe (f) (h) 250 µm (b) (d) Al Si O 25 µm Figure 2-18: Energy-dispersive X-ray spectroscopy (EDS) maps of an unpolished surface from a natural crystal used in this study. The scanning electron microscope (SEM) picture is reproduced and overlaid with chemical analysis maps of sulfur, iron, oxygen, silicon and aluminium. Table 2.4: Isotopic composition of naturally-occurring iron. [79] Isotope 54 Nat. abundance (at%) 5.85 Half-life > 3.1 × 10 56 Fe 22 years Fe 57 Fe 59 Fe 91.75 2.12 0.28 Stable Stable Stable to undergo a peak-like λ-transition starting at 160 o C as described for synthetic 11C and 6C pyrrhotite in Section 2, followed by a full magnetic disordering at T N = 316 o C. 57 Fe deposition and diffusion annealing To study iron self-diffusion, the stable isotope 57 Fe was used as a tracer. 57 Fe is a stable isotope with abundance in natural iron of around 2% (Table 2.4). Solid chips of 57 Fe (96.06% enriched, Nakima Ltd., Israel) comprising a total mass of 200 mg were evaporated onto the surface of ∼ 40 polished Fe1-x S crystals simultaneously using a Sharon TE-1 thermal evaporator. The resulting deposit thickness, measured in situ with a quartz crystal microbalance and confirmed ex situ by measuring a deposit shadow profile on a glass substrate, was 130 nm. Once coated in 130 nm of 57 Fe, pyrrhotite crystals were subjected to diffusion annealings at a range of temperatures between 170-400 o C for different lengths of time. Annealing was performed in a horizontal quartz tube furnace under a dynamic H2 S:H2 atmosphere in a molar ratio of 1:3500 (i.e. 0.03% H2 S) which was within the sulfur partial pressure stability window for Fe1-x S for the relevant range of temperatures (see Fig. 2-25 in Section 2.5.1). Maintaining this dynamic, reducing atmosphere prohibited the formation of oxide on the deposit or oxidatoin of the crystals themselves. Crystals were mounted on a custom stage and inserted into the hot part of the furnace for a predetermined period of time before being removed outside the tube furnace for cooling under the H2 S/H2 atmosphere. A thermocouple allowed in situ temperature profile recording for each annealing run (e.g. Fig. 2-19a). Specimens that were left for periods longer than 15 days to anneal were sealed under vacuum in quartz tubes and placed in heated silicone oil baths, rather than in the dynamic atmosphere quartz tube furnace. 2 2 Only the 170 o C sample was annealed in a quartz tube. Thin film samples made by CVD and sputter deposition were also annealed in quartz tubes. The surfaces of these specimens turned bright blue indicating the formation of a surface oxide on the 57 Fe deposit of 10 nm thickness; for the thin film samples this affected the diffusion results, lowering them by approximately three orders of magnitude. For more details, please 47 Secondary Ion Mass Spectrometry (SIMS) Depth profiling of as-annealed specimens was performed in a CAMECA IMS-5f dynamic SIMS at the Materials Research Laboratory of the University of Illinois, UrbanaChampaign. Up to 12 specimens could be mounted in a spring-loaded holder simultaneously to expose a flat face to the primary ion beam. 57 Fe and 56 Fe depth profiles were established by using a 10keV O- primary beam. In addition to these two species, the instrument was calibrated to detect any metallic impurity elements (mainly Si, Al, Ca, Mg) during each measurement. Only areas of the samples where the signal from these impurities was below the detection limit of 10 ppm were used for diffusion analysis. Anions such as sulfur and oxygen could not be simultaneously measured along with cations using the O- primary beam and we do not include an analysis of these species in this work. The 57 Fe concentration [57 Fe ] as a function of sputtering time was determined from the intensities I of the secondary positive ions 56 Fe + and 57 Fe + using 2.4.1 [57 Fe] = I(57 Fe) I(57 Fe) + I(56 Fe) (2.9) Conversion of sputtering time to profile depth was achieved by measuring the depth of the SIMS craters using a Dektak profilometer and assuming a constant sputtering rate (Figs. 2-19b and c). The main factors contributing to error in depth profiling are surface roughness and sample tilt. Tilt can be seen clearly in Figure 2-19b with a distortion of the square SIMS profile. However, the sputter rate was found to be correlated with primary beam current, which could only be controlled accurately to within ± 25 nA but was displayed on the instrument for every run (Fig. 2-19d). An ordinary linear regression of the measured sputter rate on primary beam current was used to estimate a standard deviation for diffusion profiles that are subjected to such systematic errors. Diffusion profiles were fit to the error function solution in Eq. (2.12) using the Curve Fitting application in MATLAB R2012b. 2.4.2 Results and Discussion Fitting of diffusion profiles Figure 2-20 compares the [57 Fe ] diffusion profile of an un-annealed sample with the original 57 Fe deposit versus a typical profile of an annealed sample (275 o C, 5 minutes). The original SIMS spectra as a function of time for the annealed sample are inset in the figure. The un-annealed profile contains a sharp interface between the original 57 Fe deposit and the Fe1-x S crystal, located at 130 nm depth. After annealing, a [57 Fe] tail from diffusion extends approximately 800 nm into the crystal. Nevertheless, even after diffusion annealing at 275 o C, a substantial proportion of the initial, pure 57 Fe remained on the Fe1-x S surface. This essentially acted as a semi-infinite source of isotopic iron in the limit of the relatively short annealing times used in this work. We can therefore assume the concentration at the diffusion couple interface (i.e. at 130 nm depth) is fixed at c = co for the duration of the annealing run. The initial and boundary conditions are hence: c(x, t = 0) = 0 (2.10) c(x = 0, t) = co (2.11) and: refer to Appendix C. For the bulk, natural sample the thickness of the surface 57 Fe deposit was too thick for the formation of a surface tarnish oxide to affect the diffusion result. 48 (a) Typical annealing cycle 400 (b) SIMS crater for diffusion profile Time at 95% of Tset Temperatureo(C) 350 300 250 Heating transient 200 Cooling transient 150 100 (c) 100 µm 0 100 200 300 400 Time (s) 500 (d) Depth profile of crater in (b) 2.5 1 -1 Sputter rate (nm.s ) 0 -1 z (µm) Sputter rate linear regression -2 -3 -4 -5 2 1.5 -6 -7 0 100 200 300 x (µm) 400 1 150 500 200 250 Sputter current (nA) 300 Figure 2-19: Sources of error considered in statistical analysis of diffusion data. (a) representative annealing profile. The time used for each measurement was total time within 5% of the setpoint. (b) optical micrograph of a tilted SIMS crater, indicating the position of the depth profile line scan in (c). (d) One way to quantify the effects of sputtering error is to take the entire sample set and measure sputtering rate (= crater depth/sputter time). This should be linearly correlated with primary ion sputter power; the standard deviation is used to estimate error in the depth of the diffusion profiles. SIMS depth profiles Intensity (arb. units) 1 0.6 57 [ Fe] 0.8 0.4 56 8 10 57 Fe Fe 6 10 4 10 0 500 1000 Time (s) Original deposit 130 nm 0.2 Annealed 275 oC 5 min. 0 0 500 1000 Depth (nm) 1500 Figure 2-20: Secondary ion mass spectrometry (SIMS) profiles. The interdiffusion of a 130 nm-thick 57 Fe deposit on Fe1-x S crystals was measured by SIMS. The black curve shows the SIMS profile from an un-annealed sample, indicating the depth of the original deposit. Inset: raw SIMS data for a sample annealed for 5 minutes at 275 o C; converted in the main plot to 57 Fe concentration vs. depth into the crystal (grey curve). 49 Error function fits 1 o 225 C Measurement n Error f fit 0.6 57 [ Fe] 0.8 13000 s 0.4 1800 s Residuals 0.2 0 0.02 0 -0.02 0.02 0 -0.02 0 1000 2000 Depth (nm) 3000 Figure 2-21: Error function solution to diffusion profiles. The tail of each profile was fit to Eq. (2.12). Example fits for two samples annealed at 225 o C for approx. 1800 and 13000 s are shown. Then it can be shown there exists an error function solution to Fick’s second law in the form [70]: c(x, t) = co 1 − er f x 4Dt) (2.12) The diffusion tails beyond the 130 nm original deposit were normalized to an initial, interfacial concentration co = 1 and fit using Eq. (2.12). Figure 2-21 shows an example of fitting two profiles from samples annealed at the same temperature of 225 o C but for different times of 1.8 x 103 and 1.3 x 104 s. The sample annealed longer has a 57 Fe profile that extends further into the sample, as expected. However, the measured * DFe values from fitting were close: 4.6 and 3.4×10−13 , respectively. Wherever possible, more than one measurement at a single temperature was obtained to check for consistency. The best-fit error function solutions are overlaid on top of experimental data with a range of uncertainty in Figure 2-21. Increased diffusion activation energy from spontaneous magnetization Values of * DFe obtained from error function fitting are listed in Table 2.5 and plotted in Arrhenius form on Figure 2-22, alongside representatitive literature data from Fryt et al. [31] and Condit et al. [40] (also shown in Figure 2-1). We include for comparison only those data corresponding to the stoichiometry range of 0.03 ≤ x ≤ 0.1 in Fe1-x S, which is close to the composition of our samples analyzed via SIMS. Above ∼ 300 o C, our measurements are consistent with the previous results, corresponding to an activation energy QP = 0.83 ± 0.03 eV, which represents the mean and standard deviation slopes obtained by regression fitting. From Figure 2-22, we also observe that below T N our * DFe values are considerably lower than the extrapolated Arrhenius trend with a slope of −Q P /kB . A deviation of this type beginning around 300 o C was previously observed by Condit et al., who postulated that vacancy ordering reduced the number of mobile vacancies by fixing Fe vacancies in equlibrium superlattice positions where they have long residence times. Despite this observation, their experiments were not pursued to low enough temperatures to confirm this hypothesis or to quantify the activation energy in the new regime. In this study, we also considered the possilibity that vacancy ordering may produce the observed, anomalous behavior at T ≤ T N . By comparison, however, 50 Table 2.5: Iron self-diffusion * DFe measurement results for Fe1-x S crystals. Temperature (o C) Annealing time (s) * DFe (cm2 s-1 ) 170 1976400 5.14 x 10-16 6.12 x 10-17 186 54000 9.24 x 10-15 3.25 x 10-15 202 72000 3.02 x 10-14 7.75 x 10-15 202 72000 4.99 x 10-14 7.43 x 10-15 205 5762 1.27 x 10-13 1.78 x 10-14 209 145800 3.33 x 10-14 3.96 x 10-15 225 1794 4.63 x 10-13 7.47 x 10-14 225 12900 3.40 x 10-13 4.37 x 10-14 249 3348 1.08 x 10-12 1.36 x 10-13 249 19830 2.22 x 10-12 2.71 x 10-13 251 573 8.39 x 10-13 4.00 x 10-16 288 116 7.56 x 10-12 6.06 x 10-13 298 449 8.77 x 10-12 1.52 x 10-12 326 596 1.89 x 10-11 2.98 x 10-12 350 86 1.53 x 10-10 6.56 x 10-11 352 410 1.76 x 10-10 3.29 x 10-11 376 94 2.87 x 10-10 4.57 x 10-11 377 276 1.54 x 10-10 3.16 x 10-11 403 98 5.19 x 10-10 8.60 x 10-11 Error in (cm2 s-1 ) * DFe oxides that undergo structural order-disorder transitions display large, discontinuous drops in diffusivity, by up to several orders of magnitude, at the critical ordering temperature. [80] Conversely, the change in Arrhenius behavior of * DFe in Figure 2-22 is continuous and does not seem to be consistent with the expected effect of a first-order structural reorganization, and is more reminiscent of the so-called "magnetic diffusion anomaly" observed at the paramagnetic-ferromagnetic critical point or Curie temperature T C in ferromagnetic materials. Self-diffusion [81–83] and solute diffusion [84, 85] in Fe, interdiffusion in Fe-Ni alloys [86] and, to a lesser extent, diffusion in Co [83] all display a deviation from the Arrhenius law extrapolated from the paramagnetic region at T C , characterized by a sharp spike or disontinuity in the effective activation energy Qeff given by: Q e f f = −kB d(ln D(T )) d(1/T ) (2.13) In other words, the magnetic transition has a second-order effect on D, manifested in an abrupt change in Arrhenius slope. Ruch et al. derived a theoretical model for the magnetic diffusion anomaly, based on a constant diffusivity prefactor Do and a magnetic contribution to the activation energy for diffusion that varies with temperature as S(T )2 , where S is the reduced magnetization relative to magnetization at zero Kelvin: S(T ) = M (T )/M (T = 0K). [87] Beginning with a generic formula for the temperature dependence of the diffusion coefficient: 51 Fe self-diffusion in Fe1-xS below ordering temperature Temperature (oC) -6 900 700 400 300 200 150 Paramagnetic QP = 0.83 eV -8 log[*DFe] (cm2s-1) 500 TN = 316 oC -10 Literature -12 Fryt * Condit ** -14 This work Data Model fit Magnetic QM = QP + αS2 (outlier) -16 1.0 1.5 2.0 2.5 1000/T (K-1) Figure 2-22: Values for iron self-diffusion coefficient * DFe obtained in this work are compared to literature data from Fryt et al. [31] and Condit et al. [40]. The known magneto-structural ordering (Néel) temperature T N of 316 o C is indicated with a dashed line. Above T N the activation energy slope in the paramagnetic state is Q P = 0.83 ± 0.03 eV. However, our results below T N deviate below the extrapolated Arrhenius relationship. The data are fit using an activation energy QM for the spontaneously magnetized state that depends on reduced magnetization S as: Q M = Q P + αS(T )2 . The outlying data point at 170 o C marked with an arrow is excluded from the trendline fit. Literature data correspond to pyrrhotite stoichiometries close to the samples used in tihs work (x in Fe1-x S ∼ 0.04) : * Fryt, x =0.04 [31]; ** Condit, average of data from 0.03 ≤ x ≤ 0.1. [40] D = f gd 2 Γ v C v (2.14) where f is a correlation function, g a geometrical factor depending on crystallography, and d the characteristic jump distance, Γ v is defined as the frequency of diffusive hops and Cv as the average vacancy concentration. Γ v depends on S according to: Γ v = νe f f exp − o Em + C S2 (2.15) kB T o where νe f f is an effective attempt frequency, Em is the migration energy in the paramagnetic state, i.e. the difference in energy between an atom in an activated position for a diffusive hop and that of a adjacent atom in an equilibrium position, and C is a constant. Cv is likewise given by: C v = exp − E of + (zJ/2)S 2 kB T (2.16) where E of is the formation energy of a vacancy in a paramagnetic crystal, z is the coordination number and J the exchange coupling constant for magnetic spins. Defining: α= C + 21 zJ Qp (2.17) then the generic diffusion coefficient in a magnetically ordered system can be written: 52 D(T ) = Do exp − Q p (1 + αS(T )2 ) kB T (2.18) The * DFe data in Figure 2-22 are fit to Eq. (2.18) using the temperature variation of the (001)NiAs magnetic reflection in synthetic Fe7 S8 obtained by Powell et al. [42] The factor α thus quantifies the magnetic ordering effect on the vacancy formation and migration energies, leading to an effective ferromagnetic activation energy for diffusion Q M = Q P + αS(T )2 . From our results we obtained a best fit of α = 0.41 ± 0.06. In other words, the activation energy for diffusion at maximum magnetization (0 K) is approximately 1.18-1.30 eV, as opposed to 0.83 eV in the paramagnetic crystal at high temperatures. We assume here that the vacancy concentration remained fixed at the original bulk value of CV = 0.04 during the annealing experiments. In other words, the effectively semi-infinite sample into which the Fe-57 exchanged remained at fixed stoichiometry. In this case, the increase in overall diffusion activation energy can be attributed solely to the magnetic influence on migration barrier, i.e. the constant C in Eq. 2.17. Nevertheless, it is worth adding a small comment regarding the reliability of our results as pure self-diffusivity measurments. Since pure Fe-57 was used as the tracer exchange material on top of the sulfide specimens, a small extent of chemical diffusion undoubtedly would have affected the measured diffusion profiles. The chemical contribution to diffusion arises from a small concentration gradient developing during annealing, altering the stoichiometry in the topmost, interdiffused volume towards a Fe:S ratio of 1:1. Such a chemical diffusion component would thus be expected to increase the diffusion rate and hence lead to a slight overestimate in self-diffusion values * DFe . We did not quantify the extent of chemical diffusion contribution directly in our analysis, but, judging by the overlap of our data with the literature values in Figure 2-22, it is likely to be small. Moreover, an unintended overestimate in * DFe does not change the key observation of a magnetic diffusion anomaly, which leads to values considerably lower than the extrapolated, paramagnetic trend with temperature. The implications of the magnetic diffusion anomaly in pyrrhotite are twofold. First, the growth conditions for Fe1-x S in solution both as a by-product of H2 S corrosion on iron and steels in energy infrastructure, as well as intentionally via solvothermal or electrodeposition methods, are typically below 300 o C. Therefore any a priori prediction of growth rates from a consideration of cation diffusion must account for the effect of spontaneous magnetization, which reduces * DFe by up to 100x at 150 o C when compared to a linear extrapolation of the Arrhenius slope from above the spontaneous ordering temperature. The accuracy of predictive tools for H2 S corrosion as well as more careful control of Fe1-x S nanocrystal synthesis both stand to benefit from these findings. Second, although Fe1-x S has long been studied for its interesting magnetic properties, to the authors’ knowledge there is no example of a successful technological implementation of this material. Since the magnetic switching phenomena of interest in Fe1-x S involve the local, diffusive rearragement of vacancies, knowing the activation energy for Fe vacancy migration and formation in the magnetized state permits the development of simulation tools to design useful devices from this material, such as temperature- and/or electric field-driven magnetic switching and memory applications. 2.4.3 Conclusions In conclusion, 57 Fe tracer diffusion measurements were performed to determine iron self-diffusivity * DFe in Fe1-x S as a function of temperature in the range 170-400 o C, which extends across the known magnetostructural order-disorder transition temperature T N = 315 o C. Our results for * DFe above T N agree well with measurements from previous studies in paramagnetic and structurally-disordered pyrrhotite, with an activa53 tion energy Q P = 0.83 eV. However, below T N , iron self-diffusivity deviates downwards from the extrapolated paramagnetic Arrhenius trend by approximately a factor of 10 at 200 o C and 100 times at 150 o C. This can be rationalized by considering a magnetic ordering effect on Fe vacancy migration energy, which increases the overall diffusion activation energy by up to 41% or to approximately 1.18-1.30 eV in the fully magnetically ordered state. To our best knowledge, this work constitutes the first description of a magnetic diffusion anomaly in an ionic compound or a ferrimagnetic material. More practicallly, the knowledge of a magnetic contribution to diffusivity allows more accurate crystal growth rate predictions of ferrous sulfide barrier layers encountered in energy systems containing H2 S and other aggressive sulfidizing agents, as well as in the solution-based synthesis of Fe1-x S where temperatures are below 315 o C. 2.5 Sulfur exchange kinetics at the Fe1-x S surface The final section of this chapter addresses the kinetics of sulfur transfer from the gas phase into solid pyrrhotite. Aside from ionic diffusion through the sulfide corrosion product barrier, this reaction is known to be the other major rate-limiting mechanism for sulfide corrosion on steels. [17, 36] For example, with H2 S as the primary sulfurbearing molecule, the net rate of sulfur transfer reaction can be represented by: H2 S(g) ⇐⇒ H2(g) + S(in Fe1−x S) (2.19) − HS(ad) at the sulfide surface is thought where the dissociation of adsorbed H2 S(ad) or to be the rate limiting step. [37] The kinetics of this process at 600 o C have been investigated using a resistance relaxation technique in sulfidized iron foils by Pareek et al. [88], who similarly looked at sulfur transfer in the other non-stoichiometric sulfides Cu2-x S [89] and Ag2+x S [90]. In this section, the results of an electrical conductivity relaxation (ECR) study on Fe1-x S thin films inspired by those studies are described. However, rather than focus on identifying rate-limiting steps and mechanisms, the primary motivation here was to obtain a series of kinetic rates for sulfur transfer as a function of temperature. Consistent with the theme of this chapter, the aim was to establish an experimental value for the activation barrier of this process such that it can be compared to the kinetics of diffusion. 2.5.1 Methods Preparation and characterization of Fe1-x S thin films Pyrrhotite thin films were sputter deposited in an AJA International ATC-1800 sputter deposition system in the Microsystems Technology Laboratory (MTL) at MIT. The target was 99.9% iron monosulfide (FeS) (Semiconductor Wafer, Inc., Taiwan). 1 x 1 cm2 soda lime glass pieces served as the substrates and 350 nm FeS was deposited at a rate of approximately 1 Å/s using an Ar plasma at 250 W power. Figure 2-23a shows the surface and through-thickness morphology of a typical, as-sputtered film. Grain sizes were on the order of 50-100 nm in plan view; however, the tilted view reveals grains are columnar with lengths on the order of several hundreds of nm. To test the thermal stability of the film morphology, the same sample was imaged after annealing for 4 hours under a dynamic, 0.001% H2 S-H2 environment at 450 o C. All samples used in this work were subjected to a four-hour post-anneal treatment under these conditions after sputtering to equilibrate the microstructure prior to ECR experiements. Although grain boundaries became more clearly defined, there was no major coarsening or morphological change of the grain structure after four hours of post-anealing, and it can be assumed that under subsequent ECR conditions no further structural change occurred that could have affected the measurements. Figure 2-23c shows a Cu-kα radiation XRD 54 (a) As sputtered film (b) H2S annealed 450 C (d) Pole figure scans 300 nm (c) Intensity (arb. units) XRD: texture in annealed film 20 (100) (100) (102) (102) (110) 6x10 3 (110) 2x10 30 40 2Θ 50 3 60 Figure 2-23: Sputter deposited thin films for ECR experiments. (a) scanning electron microscope (SEM) of as-sputtered film (inset: 60 o tilted view). (b) the microstructure coarsened slightly after annealing for 4 hours under a dynamic H2 S atmosphere; no additional changes were observed for further annealing up to 12 hours. (c) conventional Cu-kα x-ray diffraction (XRD) scan revealed only two hexagonal Fe1-x S reflexes, as labelled. (d) pole figure XRD scans confirm the high (100) texture of the samples. scan for such a post-annealed film. Only two peaks are visible that can be attributed to (100) and (110) in hexagonal pyrrhotite (ICSD reference 53528). Interestingly, the 102 peak, the predominant reflex from polycrystalline pyrrhotite was not observed in 2Θ XRD scans. Pole figure scans at fixed 2Θ corresponding to the three peaks (100), 102 and (110) are shown in Figure 2-23d, suggesting a high (100) texture. Similar oriented crystallite growth of pyrrhotite was observed by Birkholtz et al. during Fe-S sputtering under reducing conditions. [91] X-ray photoelectron spectroscopy (XPS) was performed on pre- and post-annealed Fe1-x S films using Al-kα radiation at 1487 eV. Samples were cleaned by Ar+ sputtering at 500 eV in the ultra-high vacuum (UHV) chamber. The base pressure during XPS measurement was < 10-9 Torr. XPS spectra were obtained using a pass energy of 20 eV at steps of 0.1 eV. Fitting of the Fe 2p32 spectra was performed according to the procedure outlined by Pratt et al. [92], for pyrrhotite and based on calculations by Gupta and Sen of theoretical core p levels multiplet structures to distinguish between Fe2+ and Fe3+ in Fe2 O3 . [93, 94] The S 2p spectra were fitted by assuming monosulfide (S 2− ), disulfide (S22− ) and polysulfide (Sn2− ). [92] Electrical Conductivity Relaxation (ECR) A custom-built ECR setup was made using a Thermo Scientific TF55030A-1 tube furnace. A schematic of the experimental apparatus is shown in Figure 2-24. A buffered H2 /H2 S gas mixture was flown through a 40 cm-long quartz furnace at predeterimined flow rates up to 500 sccm, controlled by Omega FVL-Series mass flow controllers. The pure gases used were of 5% H2 -balance N2 and either 100 ppm H2 S-balance He or 4% 55 Schematic of ECR apparatus Pt contact wires Al2O3 stage Fe1-xS sample Figure 2-24: Electrical conductivity relaxation apparatus setup. Schematic of experimental setup for electrical conductivity relaxation (ECR) experiements. The inset picture of the sample stage shows platinum wire screw contacts on a 1 x 1 cm2 , sputtered Fe1-x S thin film sample. H2 S-balance N2 , depending on the desired partial pressure. A four-way valve at the gas inlet for the furnace tube allowed the atmosphere in the hot zone of the furnace to be switched rapidly between two different gas mixtures and hence partial pressures of sulfur. The flush time of the reactor (i.e. time to fully replace the atmosphere with a new mixture) was on the order of 30 s. In the following sections, conductivity relaxation results are presented where the experimental timescale is 102 -103 s. We therefore deem the reactor flush time negligible in the analysis of our results. The Fe1-x S samples supported on glass were contacted directly with spring-like platinum wires in a four-probe format (Fig. 2-24, inset). Each Pt contact was maintaned on the sample by compressing it with stainless steel screws. 3 The contacts were connected via Pt wires to four stainless steel gas/electrical feedthroughs which in turn were connected to a Keithley 2400 source meter for automated 4-probe electrical conductivity measurement. The source meter was controlled by a customdesigned LabView program written by Qiyang Lu (MIT) and run from a standard laptop PC. The program initiated and recorded in situ conductivity measurements of the Fe1-x S samples at intervals of 1-20 s. We do not provide conductivity measurements in this work since we did not perform average four-probe (van der Pauw) electrical measurements between all the electrodes. Instead we kept the current-sourcing wires fixed; even though the conductance was thus sensitive to the placing of electrodes, measuring the relative change in conductance was adequate for these kinetic experiments. Control of sulfur partial pressure Atmospheres of fixed sulfur partial pressure were established in the apparatus by mixing H2 /H2 S gases, as described above. The equilibrium relation H2 S(g) ⇐⇒ H2(g) + 12 S2(g) was assumed, for which the equilibrium constant can be written: K p (T ) = pH2 pS212 pH2 S (2.20) The experimental values of K p (T) as determined by Yuan and Kröger [95], Pitzer [96] and Fryt et al. [31] were compared and found to be consistent within a few percent 3 Additional, e-beam deposited contact pads of 200 nm Au or Pt with an adhesion layer of 10 nm Cr on the samples were also tried; these facilitated the establishment of good electrical contact initially, but were found to be unstable at elevated temperatures of > 350 o C and delaminated. The high conductivity of pyrrhotite itself allowed the formation of decent electrical contacts simply by touching the compressed Pt springs to the sample surface. 56 Table 2.6: K p (T) values used to calculate sulfur partial pressure. Author Expression Ref. − 4570 T + 2.35 − 10806 T − 5.87 log K p (T ) = Yuan and Kröger ln K p (T ) = Pitzer ln K p (T ) = − Fryt et al. [95] [96] 90249−50.08×T RT [31] Thermodynamic stability range of pyrrhotite Temperature (oC) 1000 800 600 500 400 300 4 2 log(fS /bar) 0 Liquid Fe1-xS -4 10 % 1% -8 0.1 HS 2 2 % HS 2 1% H 0.0 2S 01 % HS 2 0.0 -12 0.4 fcc Fe 0.6 bcc Fe 0.8 1.0 1.2 1.4 FeS2 HS 1.6 Fe7S8 1.8 2.0 1- 1000/T (K ) Figure 2-25: Temperature-pressure equilibrium phase diagram for Fe-S. Isomolar H2 S lines are superimposed to show the range of gas mixtures that could be used while maintaining Fe1-x S stability. in the temperature range of interest 300-600 o C (Table 2.6). The mean value given by these three literature sources was used to set the sulfur partial pressures used in ECR experiments. The stability range of pyrrhotite limited the range of sulfur partial pressures that could be used in ECR experiments and had to be carefully controlled to not reduce the samples to pure Fe, or conversely oxidize them to FeS2 . The T-pS2 phase diagram in Fig. 2-25 shows the stability range of Fe1-x S along with the predicted pS2 trends for different H2 /H2 S gas mixtures containing between 0.001-10% H2 S. 2.5.2 Results and Discussion In this section, we review the defect chemistry that allows electrical conductivity relaxation measurements to be made in Fe1-x S, by rapidly changing the partial pressure of sulfur in the atmosphere and allowing the sample to equilibrate over time to a new sulfur activity. Experimental, kinetic measurements for oxidation and reduction of Fe1-x S by sulfur as a function of temperature are presented. Although the influence of sulfur chemical potential difference ∆µS remained unclear, the temperature activation of both oxidation and reduction was evident from the results. Finally, some of the incon57 sistencies and inaccuracies in the current ECR experiments on thin films are discussedincluding surface contamination, thermal history and lack of precise control over gas composition. Mechanism of electrical conductivity relaxation in Fe1-x S Besides the work already mentioned on dignetite Cu2-x S [89] and silver sulfide Ag2+x S, the vast majority of ECR studies investigated mixed ionic-electronic conduction in oxides. [97–103] Briefly, the technique can be described as follows: the electrical conductivity of the sample to be studied must have a dependence on the degree of offstoichiometry. Moreover, the intended sample must be appropriately thin such that diffusional processes are very rapid in the temperature range of interest compared to molecular dissociation and incorporation of the active species at the surface. When these conditions are met, a sudden change in partial pressure of the reactive anion-bearing molecule in the atmosphere (H2 S in this case) will cause the sample to equilibrate over time with the gas, and the change in conductivity during this process can be measured continuously in situ to determine the kinetics of the anion transfer process. We can assume linear surface exchange kinetics to give the mass conversion law [98, 99]: ∂ cS (t) A = − .k [cS (t) − cS (∞)] ∂t V (2.21) where A and V are the surface area and volume of the film, respectively; k is the surface exchange coefficient; cS (t) is the sulfur concentration at time t and cS (∞) the sulfur concentration at the new equilibrium. Eq. (2.21) can be integrated to yield: t cS (t) − cS (0) = 1 − exp − cS (∞) − cS (0) τ (2.22) with: τ= l k (2.23) where l is the thickness of the sample. For Fe1-x S we assume the only ionic defects to be iron vacancies V Fe . For different sulfur valence states m = 2-8, the charge transfer process of interest can be written: m− H2 S(g) + me− ⇔ S(ad) + H2(g) (2.24) m− S(ad) ⇔ S(ad) + me− (2.25) Adopting Kröger-Vink notation for defect equilibria, the adsorbed sulfur atom is incorporated into a lattice site by the formation of a charged Fe vacancy. S(ad) ⇔ SS + VFe 0 00 VFe + 2e ⇔ VFe (2.26) (2.27) The conduction mechanism in iron sulfides has been proposed to involve hole hopping in a band of predominantly d character, requiring holes to be generated by the presence of Fe3+ ions. [52] Figure 2-26a shows the Fe 2p3/2 XPS spectrum of a pyrrhotite sample in the H2 S post-annealed condition which reveals a substantial proportion (up to 40%) of ferric ions in the near-surface. The S 2p peak for the same sample is illustrated in Table 2-26b, revealing a majority of monosulfide S 2− . Figure 2-26c and d 58 Intensity (arb. units) (a) Fe 2p photoemission (b) S 2p (c) C 1s Data Fit 725 715 705 B.E. (eV) As sputtered Annealed 295 275 (d) O 1s 2- 2+ S S22-, Sn2- Fe 3+ Fe 285 As sputtered Annealed 716 708 704 712 Binding Energy (eV) 170 166 162 158 Binding Energy (eV) 538 526 534 530 Binding Energy (eV) Figure 2-26: X-ray photoelectron spectroscopy (XPS) from a Fe1-x S thin film sample. (a) Fe 2p spectrum, deconvoluted into Fe2+ and Fe3+ according to [92]. (b) S 2p spectrum indicating an almost entirely S 2− binding environment. (c) and (d) show the reduction in carbon and oxygen contamination, respectively, after annealing the films under an H2 S atmosphere prior to ECR experimentation. Table 2.7: Deconvolution of Fe 2p and S 2p x-ray photoelectron spectroscopy (XPS) peaks. Peak Chemical State Binding (eV) Fe 2p3/2 Fe(II)-S S 2p Energy FWHM (eV) Area (%) 706.5 1.4 10.9 Fe(II)-S* 707.4 1.8 31.6 Fe(II)-S 708.3 1.4 12.7 Fe(II)-S sattelite 713.3 2.7 4.9 Fe(III)-S 709.2 1.5 18.8 Fe(III)-S* 710.3 1.5 11.4 Fe(III)-S 711.3 1.5 5.6 Fe(III)-S 712.3 1.5 4.2 Monosulfide 2p3/2 161.2 1.2 85.2 Monosulfide 2p1/2 162.3 1.2 - Disulfide 2p3/2 162.4 1.6 6.0 Disulfide 2p1/2 163.6 1.6 - Polysulfide 2p3/2 162.9 2.0 8.8 Polysulfide 2p1/2 164.1 2.0 - confirm that the carbon and oxygen content of the films, respectively, is reduced substantially by annealing in an H2 S atmosphere and that these likely do not contribute to the existence of ferric ions in the lattice (e.g. as Fe2 O3 ). The fitting parameters for the Fe 2p3/2 and S 2p1/2 peak deconvolution are shown in Figure 2.7. The photoemission spectra taken together therefore indicate the presence of Fe3+ and substantiate a conduction mechanism that would involve hole hopping on ferric ion sites. A conduction mechanism of this kind is also observed at high temperatures in metallic conducting oxides such as Nd2 NiO4 , with itinerant electrons in the conduction band. [104] Assuming a conduction mechanism of this type, where semi-mobile holes hop from Fe3+ to Fe3+ ion sites, a net charge neutrality condition can be written: 59 00 2 VFe ≈ p ∝ n−1 ∝ ae−1 (2.28) where p and n are the positive and negative charge carrier concentrations, respectively, and ae is the chemical potential of electrons. The concentration of Fe ions cFe in a molar volume V m of Fe1-x S is: c Fe (t) = 1 − cS (t) = 1 − x(t) Vm (2.29) Therefore: p= 1 − c Fe (t) Vm (2.30) Electrical conductivity se in a p-type dominated conductor is given by: se = pµh F (2.31) where µh is hole mobility, which we assume to be constant during isothermal ECR experiments. From Eqs. (2.29), (2.30) and (2.31) therefore, we can write the normalized conductivity g(t) during the measurements in terms of he concentration of sulfur in the film: g(t) = cS (t) − cS (0) se (t) − se (0) = se (∞) − se (0) cS (∞) − cS (0) (2.32) With reference to Eq. (2.22), the normalized conductivity of a sample undergoing a change in sulfur activity to a new equilibrium is expected to follow a single exponential decay of the form: t g(t) = 1 − exp − (2.33) τ Figure 2-27 illustrates the change of resistance of a Fe1-x S sample subjected to three consecutive changes in gas concentration from 5-10% H2 S (balance H2 ), followed by 10-15% H2 S and finally 15-20% H2 S at 565 o C. The resistance fell in an exponentiallike manner each time the atmosphere is enriched in sulfur and the sample incorporates 00 more S atoms, accommodated by an increased VFe and hence mobile hole concentration. In the following section, as-recorded resistance was converted to normalized conductivity g(t) and fitted to an exponential function following Eq. (2.33). Temperature-dependence of sulfur exchange kinetics The full results of the ECR experiments are listed in Tables 2.8 (oxidation) and 2.9 (reduction). For each experiment, the sulfur partial pressure was changed from an initial value pS2 ,i to a final value pS2 , f as indicated. pS2 values were calculated from the equilibrium constants in Table 2.6. The normalized conductivity data for each were fit to a single exponential function (Eq. (2.33)) and the chemical exchange coefficient kox for oxidation relaxation curves and kred for corresponding reduction were derived from the best fit values of τ. Representative conductivity relaxation curves at a range of temperatures are given in Figures 2-28a and b. The decent exponential fits to the data (Figs. 2-28a and c) confirm that a single activation process is limiting the relaxation of conductivity. For a film thickness d = 350 nm and the * DFe results in Figure 2-1 of this chapter, typical diffusion timescales tdiff can be estimated using t di f f ∼ d 2 /(4DFe ). For the lowest temperature used in the ECR experiments of 390 o C (* DFe ≈ 10-10 cm2 s-1 ), tdiff ∼ 3 s. Diffusion should therefore be on the order of 1000x faster than the observed relaxation process, implying that the observed exponential decay should arise from surface exchange processes alone. 60 Table 2.8: Electrical conductivity relaxation results for oxidation experiments: initial (pS2 ,i ) and final (pS2 ,i ) sulfur partial pressure values; surface exchange coefficient kox from exponential fits. These data are plotted for clearer comparison in Figure 2-28. T (o C) log pS2 ,i (atm) log pS2 , f (atm) kox (x 10-9 cm/s) 390 -14.6 -13.2 16.5 390 -14.6 -12.6 23.9 390 -14.6 -12.6 7.34 390 -13.2 -11.8 5.75 400 -18.0 -17.0 14.4 400 -18.0 -17.0 19.6 400 -19.0 -17.0 24.9 430 -17.0 -16.0 9.86 450 -16.0 -14.0 14.6 450 -17.0 -16.0 62.2 480 -16.7 -14.7 55.1 490 -11.4 -10.2 164 510 -11.4 -10.2 194 513 -11.0 -10.1 94.7 520 -10.9 -9.8 120 520 -16.0 -14.0 314 520 -16.0 -14.0 124 520 -16.0 -14.0 300 565 -8.7 -8.1 670 567 -10.3 -9.1 832 Table 2.9: Electrical conductivity relaxation results for reduction experiments: initial (pS2 ,i ) and final (pS2 ,i ) sulfur partial pressure values; surface exchange coefficient kred from exponential fits. These data are plotted for clearer comparison in Figure 2-28. T (o C) log pS2 ,i (atm) log pS2 , f (atm) k r ed (x 10-9 cm/s) 400 -17.0 -18.0 23.2 400 -17.0 -18.0 22.0 400 -17.0 -18.0 28.6 513 -10.1 -11.0 63.3 520 -14.0 -16.0 76.8 520 -14.0 -16.0 118 520 -14.0 -16.0 115 565 -8.6 -9.1 418 565 -9.1 -10.3 350 61 % H2S in mix Electrical resistance relaxation at 565 oC 20 15 10 5 11.5 Resistance (Ω) 5-10% H2S 11 10-15% H2S 15-20% H2S 10.5 10 0 500 1000 1500 2000 Time (s) Figure 2-27: Electrical resistance relaxation at 565 o C upon three successive changes in H2 S-H2 gas mixture in the furnace atmosphere, as indicated. The clearest observation from the tabulated results is an overall trend of increasing k towards higher temperatures, seen more clearly when the data are presented graphically in Figures 2-28a and b. Ignoring for now the dependence on sulfur partial pressure, we use this trend to determine an average activation energy for the surface chemical exchange process Echem of 1.05 ± 0.20 eV for oxidation and 0.79 ± 0.23 eV for reduction. The relaxation time should also, in theory, be affected by a chemical driving force, as has been shown in studies of bulk oxides. [98, 99] This can be defined as the difference in the equilibrium chemical potentials of sulfur ∆µS in the sample in the inital and final states. Assuming the ideal gas law and zero activity constant: pS2 ,i RT ∆µS = ln (2.34) 2 pS2 , f Tables 2.8 and 2.9 also list the initial and final sulfur partial pressures used in each experiment. There is no definitive correlation between relaxation time and partial pressures in our data. The uncertainty in each reading is evident by comparing experiments o performed at constant temperature. For example, in Table 2.8 the 390 C measurements include two results obtained with log pS2 ,i = −14.6 and log pS2 , f = −12.6, yielding two disparate values of kox = 23.9 × 10−9 and 7.34 × 10−9 cm/s. The other two results for oxidation at 390 o C were taken using different values of pS2 ,i and pS2 , f ; however, the resulting values of kox fall between 23.9×10−9 and 7.34×10−9 cm/s. No conclusive verdict can be reached on the influence of ∆µS from these data. The same is true for other readings obtained at constant temperature. In the following section, some of the inaccuracies that may contribute to this uncertainty are discussed in more detail. The ECR technique for pyrrhotite thin films as described here is judged to be reasonable for obtaining the broad temperature dependence of the oxidation and reduction processes. However, in their present form the kinetic measurements are too variable and ambiguous to reveal more intricate informa62 (a) Normalized conductivity: oxidation (c) Exchange coefficient: oxidation Temperature (oC) 600 550 500 450 400 375 -14 Ea, ox = 1.05 ± 0.20 eV 0.8 -15 o ln(k) (cm.s ) 390 C 400 oC o 480 C 520 oC 565 oC 0.6 0.4 -1 Normalized conductivity 1 -17 ΔPS = 101 atm -18 Fit: σ = 1 - exp[-t/τox.] 0.2 -16 2 2 ΔPS = 10 atm 2 -19 0 0 4000 2000 6000 Linear regression 95 % C.I. 1.1 1.2 1.3 Time (s) (b) 1.4 1.5 1.6 -1 1000/T (K ) Normalized conductivity: reduction (d) Exchange coefficient: reduction Temperature (oC) 600 550 500 450 400 375 -14 400 oC o 520 C o 565 C 0.8 Fit: σ = 1 - exp[-t/τred.] 0.6 0.4 -16 -17 -18 ΔPS = 101 atm -19 ΔPS = 102 atm Linear regression 95 % C.I. 0.2 0 0 Ea, red = 0.79 ± 0.23 eV -15 ln(k) (cm.s-1) Normalized conductivity 1 2 2 2000 4000 6000 1.1 Time (s) 1.2 1.3 1.4 1.5 1.6 1000/T (K-1) Figure 2-28: Electrical conductivity relaxation results: (a) representative oxidation relaxation curves (data points) obtained at different temperatures with best fit exponentials (solid lines). (b) representative reduction relaxation curves. Chemical exchange coefficients for (c) oxidative sulfur transfer from H2 S to Fe1-x S and (d) the reverse reductive transfer into the gas phase: individual data and best-fit Arrhenius line with 95 "%" confidence interval. k values corresponding to the individual curves shown in (a) and (b) are demarcated with arrows. tion over secondary effects such as that of the chemical driving force. Measurement consistency and surface degradation The inconsistencies in our ECR results, even among results obtained under identical conditions as discussed above, underscore the high sensitivity of sulfur exchange reaction to the condition of the sample surface. In Figures 2-29a and b, two sets of in situ resistance measurements are shown that were obtained at 520 o C and 400 o C over the course of several hours each. Several oxidation/reduction cycles were performed by switching the sulfur partial pressures between two consistent values as shown and allowing the sample to equilibrate each time. At first glance, the redox cycling looks fairly repeatable with little hysteresis, despite a constant upwards drift on the order of 63 50-60 S/hour. However, in Figure 2-29c we plot some other results obtained at 400 o C on another sample. This time the conductivity is normalized to the equilibrium conductivity at the start of each cycle. revealing a consistent attenuation with successive cycles that is suggestive of a history-dependent surface degradation process. For example, the percentage change in conductance during the first oxidation step "O1" is approximately 1.7%, but is reduced to only 0.6% for the subsequent reduction step "R1" performed after "O1". By the fifth cycle "O5", the conductance relaxation has reduced to only 0.4%. The gradual degradation in the results after repeated cycling on a single sample rules out sample-to-sample differences to be responsible for the variation in measured k values; all samples were fabricated and pre-annealed identically. The reason is more likely to be related to a degradation in the condition of the sample surfaces with time. Scanning electron microscopy investigations of the Fe1-x S samples after different stages of experiment revealed the columnar film morphology and grain size to be stable at high temperature over the course of 12 hours. Moreover, an XPS investigation of a sample subjected to a typical ECR experiment did not expose any atypical surface chemistry, such as increased carbon, oxygen, or other foreign elements. On the contrary, the carbon and oxygen content of the films was decreased by post-annealing in an H2 S atmosphere after deposition (Fig. 2-26) and subsequently remained stable during further ECR experiments. Surface contamination can thus be ruled out as an explanation for the degradation. The as-deposited Fe1-x S films used in this work had a roughness on the order of 10-20 nm; prolonged annelealing did not measurably change the roughness. The instability and lack of repeatability observed in these sulfide ECR experiments are common also in similar work with oxide samples, particularly where thin films and/or carefully engineered surface chemistry or nanostructuring is involved. [101] For example, Wang et al. observed the oxygen exchange surface kinetics on La0.6 Sr0.4 Fe0.8 Co0.2 O3 (LSFCO) to degrade after annealing at 900 o C. Here also, no changes in surface chemistry were detected by XPS that could provide an explanation for the degradation in oxygen transfer properties. Conversely, Chen et al. found in another study using epitaxial lanthanum strontium cobalt oxide (LSCO) that the oxygen exchange activity increased substantially after annealing. They attributed this to a roughening of the surface, which introduced surface steps and edges. [105] The surface exchange process for Fe1-x S involves a sulfur atom taking a vacant anion site on the surface; grain boundaries and surface steps/edges are known to be catalytically active sites for dissociative adsorption. More work is needed to undersand the high sensitivity of ECR experiments to the surface condition, leading to inconsistent results. 2.5.3 Conclusions ECR was used to monitor the sulfur surface exchange kinetics for Fe1-x S thin film samples subjected to changes in sulfur partial pressure in buffered H2 S/H2 gas mixtures. Results can be fit to a single exponential curve, confirming that surface exchange is the sole rate-limiting process. A clear temperature dependence of the exchange coefficient k allowed us to determine the activation barrier Echem to sulfur exchange of 1.05 ± 0.20 eV for oxidation and 0.79 ± 0.23 eV for reduction. However, the technique is not sensitive enough to determine the secondary effect of chemical driving force on exchange kinetics. Moreover, the measurements are history-dependent: multiple redox cycling leads to an attenuation in the relative resistance change upon subsequent oxidation or reduction steps. There is no clear explanation for this based on changes in surface morphology or chemistry, and more work must be done to understand the influence of surface state on kinetics in more detail. 64 Inconsistencies in ECR measurements Conductance (x 100 S) (a) 14.0 520 oC 1 hour 13.8 60 S Red. 13.6 13.4 ~ 4% change Ox. 13.2 13.0 0 1.5 1 0.5 2.5 2 3 3.5 4 Time (10 s) Conductance (x 100 S) (b) o 16.5 400 C < 1% change Red. 16.4 16.3 Ox. 16.2 0 1 0.5 1.5 2.5 2 3 3.5 4 1.02 O1 t Normalized conductivity σ /σ o Time (10 s) 1.01 O4 1 O3 R3 O2 R2 R1 0.99 0.98 0 2000 4000 6000 8000 Time (s) Figure 2-29: Drift, stability and repeatability of ECR experiments. (a) conductance measurement for daylong repeated oxidation and reduction cycles at 520 o C. The global drift is on the order of 60 S/hour. There are also points at which conductance is unstable and jumps suddenly. (b) repeated redox cycling at 400 o C. The timescale for individual relaxations is longer than for measurements at 520 o C. (c) repeated oxidation (O1-O4) and reduction (R1-R3) cycles on a different sample at 400 o C. Relative conductivity is obtained by normalizing results by the equilibrium conductance at the beginning of each cycle; we see a gradual degradation in the magnitude of the overall change after the first oxidation cycle O1. 2.6 2.6.1 Outcomes Conclusions The primary aim of the work described in this chapter was to compare activation barriers Ea and kinetic rates of the unit processes of cation diffusion (bulk) and exchange of sulfur (surface) in Fe1-x S. The experimentally-determined values of Ea are listed in Table 2.10. Since the two processes have a similar temperature dependence, there is 65 Surface exchange slower at all temperatures until film is 1000 μm thick Fe diffusion S exchange Faster Surface limited 0 Equilibration const. (s-1) 10 Surface limited -5 Mixed 10 Diffusion limited -10 10 1 Slower -0.5 10 Film Thickness (μm) -1 -1.5 100 -2 1000 -2.5 1000/T (K- 1) Figure 2-30: Temperature- and film thickness dependence of rate limiting steps. The "equilibration constant" τeq for both processes, as a measure of the kinetic rate, is plotted as a function of both temperature and also film thickness x. As the film gets thicker from 1 to 1000 µm, the overall rate of both processes becomes slower. However, since the diffusion rate decreases parabolically with x, a crossover from surface exchange limited growth to diffusion limited growth would be expected at around x = 100-1000 µm. no clear transition temperature above which one would be expected to dominate over the other in controlling the rate of sulfidation of iron. However, we can take an analytical approach to understanding rate-limiiting regimes by considering the "equilibration constant" τeq for each process. This can be thought of as the characteristic time constant for the given process to occur, assuming an exponential rate R of the form R(t) = R o exp [−t/τeq ]. For diffusion, we can write τeq = 4D/x 2 , whereas for surface exchange, τeq = kex /x. In Figure 2-30 we plot the equlibration constant for different film thicknesses of 1, 10, 100 and 1000 µm. It can be seen that for 1- and 10 µm-thick films, surface exchange as determined experiementally using ECR is a slower kinetic process (smaller τeq ), irrespective of the temperature. However, due to the parabolic dependence of diffusion on film thickness, above 100 µm, the rates become approximately similar, and a crossover would be expected from surface control to bulk diffusion control in the rate of pyrrhotite growth on iron. Moreover, these basic kinetic rate parameters can be fed into the global kMC and phase field model as described in Chapter 1. The secondary aim was to understand the influence of the critical magneto-structural order-disorder transition temperature T N = 315 o C on Fe diffusion in pyrrhotite. This was studied by two different techniques: first through in situ magnetization measurements of the diffusion-driven, λ magnetic transition and second through tracer diffusion studies using SIMS. Both studies converged on approximately the same Ea in magnetic pyrrhotite below T N , which is up to 40% higher than that for non-magnetic pyrrhotite (Table 2.10). Besides clarifying the effect of T N on diffusion, the results are more practically important in estimating rates of degradation where pyrrhotite forms a passive layer. For example, assuming a simple extrapolation of the paramagnetic Arrhenius law 66 down to 150 o C would overestimate real diffusivities by up to two orders of magnitude. 2.6.2 Future work The self-diffusivity of iron was measured here under dry, gaseous conditions in solid (non-porous) samples. However, real corrosion scales in aqueous conditions could behave markedly differently from this idealized case. It would be of interest to confirm * DFe values at temperatures below 300 o C, but under in situ, aqueous conditions. Hightemperature sulfidation and/or corrosion experiments in the range 100-300 o C are required to confirm empirically the rates of both diffusion and surface reaction found here on pure, dense samples. Such studies are also crucial to validating and improving the bottom-up passive film mode that serves as the overarching goal of this work. More work must be done to explain the sensitivity of the ECR technique to surface conditions, and to determine whether fundamentally accurate kinetic rates are measurable via this method. This would involve repeated oxidation/reduction, with careful surface analysis using XPS and AFM/STM imaging to understand any surface changes that may alter the kinetics. 67 Table 2.10: Key activation energies for pyrrhotite growth. Process T range (o C) Technique Ea (eV) Fe diffusion 315-700 Literature 0.83 ± 0.03 Fe diffusion 185-315 SIMS 0.83 + αS(T )2 Fe diffusion 180-210 Magnetokinetics 1.10 ± 0.10 Surf. Exch. (ox.) 350-600 ECR 1.05 ± 0.20 Surf. Exch. (red.) 350-600 ECR 0.79 ± 0.23 1 α = 0.41 ±0.06 and S(T ) is the reduced magnetization (= 0 at 0 K). 68 1 Chapter 3 Reactivity: quantification of electronic band gap and surface states on FeS2(100) Synopsis The scanning tunneling microscope (STM) is used to investigate the surface electronic structure of pyrite, FeS2 , as a model, semiconducting passive layer phase. The STM allows us to probe controllably the energy levels of FeS2 and quantitatively evaluate the surface electronic features which affect its charge transfer characteristics, with respect to a redox species in the environment such as H2 S. The interfacial electronic properties of pyrite are greatly influenced by the presence of electronic states at the crystal free surface. Scanning tunneling spectroscopy (STS) results are interpreted using tunneling current simulations informed by density functional theory (DFT). Intrinsic, dangling bond surface states located at the band edges reduce the fundamental band gap Eg from 0.95 eV in bulk FeS2 to 0.4 ± 0.1 eV at the surface. Extrinsic surface states from sulfur and iron defects contribute to Fermi level pinning but, due to their relatively low density of states, no detectable tunneling current was measured at energies within the intrinsic surface Eg . These findings help elucidate the nature of energy alignment for electron transfer processes at pyrite surfaces, which are relevant to evaluation of electrochemical processes including corrosion. Finally, the broader utility of the methodology developed in this work for reliably interpreting STS results is discussed. This includes determining the fundamental surface energy band gap for less commonly-studied semiconductors for use in earth-abundant photovoltaics and other applications. Portions of this chapter were published in Surface Science. [106] All DFT calculations in this work were performed by Aravind Krishnamoorthy. 3.1 Background and motivation At the end of Chapter 3 of this thesis, the kinetics of charge transfer between gaseous H2 S and pyrrhotite (Fe1-x S) were investigated experimentally by measuring the electrical response of a sample to changes in sulfur chemical potential. This electrochemical process, a necessary step for the incorporation of sulfur from the environment into a growing iron sulfide passive layer, can be summarized by the cathodic half-reactions: H2 S(g) ⇔ S 2− +2H + and 2H + +e− ⇔ H2(g) where the electrons are transferred from the iron sulfide barrier layer to reduce hydrogen sulfide. The electrical conductivity relaxation experiments in Chapter 3 give a practical, macroscopic measure of reaction rates that can be used to predict average rates of sulfur transfer. However, they do not provide any mechanistic insight into the physics of interfacial electron exchange between 69 the solid film and molecular redox species. Chapter goals In this chapter, we aim to investigate surface reaction in more detail by asking the question: how do the electronic properties at the surface of an ionic solid affect the propensity for charge transfer in an electrochemical system? Instead of pyrrhotite, we investigate the surface of pyrite (FeS2 ) as a model, semiconducting passive film material. FeS2 was chosen because of its broader interest to the electrochemical community: other applications including photovoltaics and battery anodes are discussed briefly in Sections 3.1.4 and 3.4.2. Open questions over the surface electronic structure of pyrite from the literature partly motivated this work, and the existing amount of data helped benchmark our experimental and computational results. The key questions addressed in this chaper are summarized as: (1) can the surface band gap of FeS2 (and by extension, other similar semiconducting materials) be quantified a priori using STS? (2) how do both intrinsic and extrinsic surface states on FeS2 (100), as a model passive layer material, affect charge transfer in electrochemical systems? 3.1.1 Electrochemical charge transfer in semiconductor-absorbate systems The theory of charge transfer during surface reactions between solid, semiconducting materials and molecular adsorbates in corrosion systems is covered in the book Corrosion mechanisms in theory & practice by Marcus et al.. [3] The key points are summarized here to contextualize the remaining work in this chapter. Complex, coupled electrochemical systems such as a metal-passive layer-electrolyte structure can be reduced to a connected set of energy levels. Under an applied bias (electrode potential), it is the transfer of electrons across these energy levels which determines the overall reaction current and therefore rate of metal oxidation during corrosion. Let us imagine a basic system comprising the three interconnected components of a metal, a passive layer, and an electrolyte containing a molecular species which can assume a reduced or oxidized state: Metal: (e.g. Fe in steel) has a high concentration of mobile charge carriers, typically O(1023 cm-3 ). Passive layer: typically semiconducting or insulating, we represent the passive layer with a conduction band, valence band and Fermi level dependent on level of doping. The volume fraction of defects is typically very high for in situ formed passive layers, with high concentrations of intrinsic (point) defects and extrinsic defects (substitutional elements). [32] Charge carrier concentrations are commonly in the range 1015 -1019 cm-3 . Redox couple: a molecule in the electrolyte that reacts with the passive layer can be represented electronically by occupied and unoccupied levels, corresponding to the reduced and oxidized components of the redox system, respectively. Each level is depicted as a Gaussian distribution of states, accounting for the uncertainty in rearrangement energy for the solvation shell around the molecule during a reaction. L represents the energy required to reorganize the shell of H2 O molecules surrounding a redox species after charge exchange (see Fig. 3-1). The charge concentration of the redox couple is related to the concentration of ions in solution. For example, a 1 M solution contains 2NA × 10−3 ≈ 1021 ions.cm-3 . 70 We assume that the Fermi levels of each component in the system align at equilibrium. An additional, applied electrochemical potential biases the energy levels depending on the activity of reduced and oxidized species in the electrolyte, aRed and aOx , respectively: E r ed = E rΘed − RT aRed ln zF aOx (3.1) where E r ed is the half-cell reduction potential at a temperature T, E rΘed is the standard reduction potential, R is the universal gas constant, z is the charge transferred per reaction and F is Faraday’s constant. [3] Due to the high availability of charge carriers in the metal and electrolyte, and relative dearth in the passive layer, the potential difference is accommodated across the passive layer. This leads to potential drops at the metal/passive layer and passive layer/electrolyte interfaces; the passive layer’s semiconducting band structure is offset at these interfaces to maintain a constant EF . In other words, charge accumulates or depletes at the passive layer surfaces, resulting in a space-charge layer and band bending (see detailed introduction to band bending below). At the metal-passive film interface, we usually assume very rapid charge transfer. At the passive layer-electrolyte interface, the potential drop can be divided into the potential dropped within the electrical double layer (Helmholtz layer) of ions in the electrolyte ∆ϕH and the potential dropped within the semiconducting passive layer itself ∆ϕSC (Fig. 3-1a). For a constant thickness Helmholtz layer, ∆ϕH is determined by the pH of the electrolyte. ∆ϕSC similarly depends on the charge carrier concentration of the passive layer, which determines the Debye length β. For example, considering an n-type semiconductor with donor concentration N D : v u t ""o kB T β= eo2 ND (3.2) where " and "o are the relative and vacuum permittivities, respectively; eo is the elementary charge and kB is Boltzmann’s constant. The depth of the resulting space-charge layer in the passive layer dSC (Fig. 3-1b) can then be written: dSC v t 2e (E − E ) o FB =β kB T (3.3) where E is the applied potential and EFB the flat-band potential (no bias). We can therefore see that the width of the space-charge region is potential dependent; in other words, the passive layer surface compensates for any changes in applied electrode potential. Band bending For a negative electrode potential η applied to an n-type semiconductor, the EF of the redox couple is lowered relative to that of the passive layer, and free electrons accumulate at the semiconductor surface until the Fermi levels equilibrate. This can be represented as a downwards bending of the conduction and valence bands at the surface (Fig. 3-1a). Conversely, a positive η gives rise to a depletion of electons at the surface and upwards band bending (Fig. 3-1b). Semiconductor surfaces whose bands are free to shift up/down in this manner are said to have an "unpinned’" EF . 3.1.2 Surface states The reactivity of semiconducting materials can be significantly altered by surface states that are either intrinsic to the crystal termination or arise from the presence of crystalline defects at the surface. 71 72 φ ΔφSC Oxide eη iDir 2λ EF, redox Density of states E Valence band Eg Distance EF VB CB iTunn 2λ Density of states E EF, redox dSC Conduction band + + + -- + + + + + + - - - - - - - (d) Tunneling transfer iTunn (+ve η) EF E = e.φ (b) Band bending at electronic equlibrium Figure 3-1: Charge transfer in electrochemical (corrosion) systems. (a) Potential distribution across a metal-passive layer-electrolyte couple; most of the potential drop is accommodated at the interfaces, resulting in the formation of a space charge layer (∆φSC ) in the passive layer. (b) under electronic equilibrium the Fermi level EF is flat, leading to band bending in the passive layer. (c) negative applied overpotential η : downwards band bending allows direct electron transfer from the passive layer conduction band (CB) to the unoccupied levels in the redox system. (d) positive η : enough upwards bending can make tunneling transfer through the space charge layer possible. Distance VB EF CB ΔφH Electrolyte (c) Direct electron transfer iDir (-ve η) Metal (a) Potential distribution across passive layer Intrinsic surface states: The abrupt discontinuation of periodic potential at semiconductor surfaces can impose severe perturbations to the crystal’s electronic structure. Unless a surface reconstructs to remove dangling bonds and autopassivate, the topmost atoms’ crystalline orbitals destabilize in the direction of their free atom orbital character and energy. Binary semiconductors like FeS2 , whose (100) surface does not reconstruct, are predicted from ligand field models [107] and DFT calculations [108–110] to have two associated intrinsic surface states: one for the anion dangling bonds and one for cations. However, the intrinsic surface states have never been experimentally characterized up to this point. Extrinsic surface states: in reality these states are localized at defects such as anion or cation vacancies on the surface steps, kinks, dislocations or impurities [111, 112]. However there is evidence to suggest that these types of defects (as well as step edges or intersecting dislocations) can affect the surface electronic structure over nanometer distances. Horizontal charge transfer Adiabatic charge transfer between a redox couple and a semiconducting passive layer must be horizontal, that is- must occur at the same energy level. Electrons may be exchanged in either direction: anodic transfer refers to electrons transferred from the electrolyte, and cathodic transfer in the opposite direction. If we ignore for the moment the role of surface states, and other potential transfer pathways such as hopping mechanisms and via bulk, intra-band states, charge exchange with the bulk bands of the passive layer may occur by two mechanisms: direct exchange and tunneling. [3] Direct transfer (see Fig. 3-1a) from the semiconductor requires filled states in the CB to be higher in energy, i.e. overlap, than the empty states (oxidized species) of the redox couple. The more negative η is in this scenario, the greater the overlap of electronic states and charge transfer can increase exponentially. This is the fundamental explanation behind the well known linear Tafel slope on a plot of log(cur r ent) vs. η for corrosion systems. Tunneling transfer (see Fig. 3-1b) refers to quantum tunneling of electrons through the space-charge layer. Thus, the larger the band bending, the smaller the tunneling distance dT and the greater the probability of tunneling T: 16E(Vo − E) 2κd T exp − T= Vo2 h (3.4) p where the coefficient κ describes the dependence on barrier height: κ = 2m∗ (Vo − E), where V o is the total energy barrier, E is the energy level of the tunneling electron and m* is the reduced electron mass. Charge transfer at the passive layer surface is therefore proportional to the density of occupied states of the redox system D(Red) and that of the empty states within the passive layer CB, D(Ox). By integrating over all energies above EF where overlapping occupied and empty states may interfere, the tunneling exchange current i+ is therefore: i+ = −F Z T.D(Ox).D(Red).d E (3.5) The size of the band gap and density of electronic states at the surface is therefore crucial to understanding the propensity for electron exchange. Some oxides that have a large band gap, e.g. Ta2 O5 (Eg > 3 eV), such levels are not available and they do not show redox processes even at very positive applied potentials, i.e. would require high band bending to meet the condition of electron tunneling through the space charge layer. 73 3.1.3 Scanning tunneling spectroscopy and TIBB In the work presented in this chapter, the aim was to understand the role of surface states in determining the surface Eg through quantitative analysis of tunneling spectroscopy (STS) measurements. We adopt the approaches developed in modeling STS data from semiconductor surfaces that was advanced from the late 1980s by R.M. Feenstra and others. Early work began with the traditional cubic tetrahedrally bonded [113] and III-V [114] semiconductors, on which band edges and surface-related features could be determined to within an accuracy of ± 0.03 eV. The concurrent development of tunneling spectrum models based on computations of potential distributions and tunneling current has helped identify the role of other physical phenomena in experimental STS spectra, such as tip-induced band bending (TIBB) [115] and surface states [116]. TIBB greatly affects the STS measurement of unpinned semiconductor surfaces, in which changes in the tip-induced electric field lead to an unrestricted accumulation or depletion of charge carriers at the surface which act to screen the tip potential. In this case, the electron chemical potential µe in the sample shifts freely with applied bias, distorting the CB and VB near the surface. However, if surface states are present on the sample, charges from the bulk bands can fall into them and EF becomes pinned at the level to which the surface states are occupied. STS spectra of EF -pinned surfaces typically yield more consistent band onsets and are less affected by localized quantum effects such as inversion or accumulation currents arising from TIBB. The large, localized electric field from the proximate tip extends through the vacuum region and into the surface of the pyrite sample. Consequently, a fraction of the applied potential can be dropped within the sample itself, causing the valence and conduction bands to bend and obscuring the energy scale of the measured STS spectra. A comprehensive description of TIBB can be found in previous reports [115–118], but the main points are outlined here for completeness. The contact potential ∆ϕ is defined as the difference in work function between the metal tip and the semiconductor: ∆ϕ = ϕm − χ − (EC − E F ) (3.6) where ϕm is the metal work function, χ is the electron affinity of the semiconductor, and EC and EF are the conduction band minimum and Fermi level of the sample, respectively (Fig. 3-2a). Even in the absence of applied bias, a non-zero ∆ϕ leads to band bending and the formation of a depletion region in the sample. For example, a positive ∆ϕ produces upwards band bending in the semiconductor as the Fermi level aligns with that of the tip; negative charge correspondingly accumulates at the surface to screen the potential (Fig. 3-2b). The effect is to shift the energy E of any given state at the surface by an amount ϕo : E − E F = eV − ϕo (3.7) When EF at the surface is unpinned in this manner, the resulting experimental i(V) measurement can yield a very wide apparent surface Eg , as the band edges shift with the sweeping voltage and the onset of tunneling is delayed to more positive (for the CB) or negative (for VB) voltages (Fig. 3-2c). Such a situation arises on defect-free ZnO(110) surfaces [119], where the apparent band gap from STS can be larger than the accepted bulk gap of the material (Fig. 3-2e) , or on GaN(1100) where quantitative Eg determination was not possible [121]. Severe band bending can also introduce large tunneling currents from local states when the semiconductor EF is pushed into the VB (inversion) or CB (accumulation) [115, 122]. The presence of intrinsic (dangling bond) surface states on semiconductors typically limits TIBB by pinning EF (Fig. 3-2d) [116, 123]. These states accumulate charge as they become occupied and effectively screen the electrostatic potential from the tip, reducing the distortion of STS spectra arising from TIBB. By analogy, on metallic materials with freely-available conduction band electrons 74 (a) No tunnel contact TIP (b) EF alignment FeS2 Evac Evac χ φtip Δφ EF,tip EC EF EV EF,tip Δφ φo Wtip E0 E0 (c) No surface states (d) Surface states Evac EF,tip Δφ+eV Δφ+eV EC EF EV EC EF EV (c) ZnO(110) Eapp > Eg, bulk (d) InN(100) Eapp < Eg, bulk Figure 3-2: Band bending effects in STS measurement. Schematic energy band diagrams for a n-type pyrite sample where (a) tip and sample are not in tunneling contact; (b) there is an open circuit tunneling junction. With a positive contact potential ∆ϕ, an upwards band bending of magnitude ϕo occurs. In addition to the parameters described in the text: Evac is the vacuum energy level, Eo the ground level, and EV is the valence band maximum of the semiconductor. W tip is the energy difference between the metal’s Fermi level and the bottom of the metal valence band, typically ∼8 eV for PtIr tips. EF,tip is the tip Fermi level. All filled states are shaded in grey. (c) upon applying a positive sample bias V, further upwards tip-induced band bending (TIBB) occurs if EF is unpinned. However, surface states, e.g. shaded in black in (d), can accommodate enough surface charge to pin EF and minimize TIBB (after Feenstra et al. [115]). (e) no surface states exist on ZnO(110); EF remains unpinned and STS measurement gives an overestimate of Eg due to band bending. [119]. (f) Conversely, extrinsic surface states on InN(110) pin EF and reduce the apparent Eg . [120] at the surface, the tip potential drops entirely at the surface and does not extend into the sample. Similarly, extrinsic states (arising from disorder, defects or unintentional contamination), even at low densities of 0.01 monolayers (3x1014 cm-2 ) or less, can hold enough charge to significantly affect the magnitude of TIBB and pin EF , e.g. on InN(110) (Fig. 3-2f) [120]. Below, we rationalize these two competing effects in our experimental STS spectra by simulating the effect of different surface state features, the characteristics of which are known from DFT simulations. 75 3.1.4 The FeS2 (100) surface Applications of FeS2 beyond sulfide corrosion studies Pyrite or FeS2 is a semiconducting mineral for which the electronic structure has been intensively studied in relation to reactivity in geochemical [124–129] and bio-catalytic [130–132] processes, as well as for photovoltaic (PV) and photoelectrochemical properties [24, 133–136]. Heterostructures of FeS2 and perovskite oxides such as LaAlO3 have recently been proposed as promising devices for spintronics applications [137]. In the context of PV, low open circuit voltages (VOC) of < 200 mV (or ∼21% of the widely accepted bulk band gap of 0.95 eV) have been attributed to poor interfacial electronic properties of synthetic FeS2 systems [24]. (100) surface crystallography and electronic properties The crystal structure of FeS2 (space group Pa3) comprises two interpenetrating cation (Fe2+ ) and anion (S2 2- ) face centered cubic (fcc) sublattices, the latter of which is made up of S2 persulfide dimers aligned along the cube diagonal direction <111>. Pyrite is a compound, dband semiconductor with an electronic structure that can be qualitatively understood with the aid of a simple ligand field model [107]. Each Fe2+ ion in the bulk is octahedrally coordinated by S2 2- ions (symmetry group Oh ), creating a strong ligand field that splits the metal d states into non-bonding, triply degenerate Fe 3d t2g states (dxy , dyz and d x 2 − y 2 ) at the top of the valence band (VB). The conduction band (CB) minimum consists of doubly degenerate Fe 3d eg states (dz 2 and d x 2 − y 2 ) hybridized with S ppσ* orbitals. An indirect, bulk band gap Eg of 0.83-1.01 eV has been measured in synthetic FeS2 using various optical [138,139], photoconductivity [140,141] and x-ray absorption/emission spectroscopy studies [142]. At the unreconstructed (100) surface termination of pyrite, the predominant growth and cleavage face, the symmetry of the Fe2+ site is reduced from Oh to square pyramidal C4v , leading to a loss of degeneracy among the Fe 3d t2g and eg states. These further split into two discrete, intrinsic surface states associated with the Fe dangling bond. Recent density functional theory (DFT) calculations are consistent in identifying these two pronounced surface states to be located around the VB maximum (Fe-d x 2 character) and at the CB minimum (Fe-d x 2 − y 2 ). The magnitude of the surface states decays almost entirely to zero beyond approximately three atomic layers into the bulk [109]. As a result it is theoretically estimated that Eg at the FeS2 free surface is reduced by up to 0.3-0.4 eV, as compared to the bulk value (Table 3.1). We define surface Eg as the energy difference between the extrema of the surface bands that extend into the bulk gap, rather than the gap between empty and filled surface bands which may exist as discrete states within Eg ; this distinction was used by Feenstra et al. for investigating states on Ge(111)c(2x8) surfaces [117]. If the surface states are fully degenerate with the bulk bands (i.e., lie within the fundamental bulk Eg ) they are not considered in the quantification of the surface Eg . In addition to the intrinsic surface states on FeS2 (100), computational studies have identified a series of further surface states that appear within the fundamental surface Eg local to interfacial point defects [108, 109, 132]. We refer to such states as "defect" or "extrinsic" states to differentiate them from intrinsic surface states. Significant concentrations of neutral sulfur monomer vacancies V S have been measured by x-ray photoelectron spectroscopy (XPS) on fractured FeS2 (100) [143–147] as well as in situ ion-bombarded [148] and annealed [106] growth faces. Indeed, the formation energy ∆Hf for V S is estimated to be as low as 0.1 eV experimentally [106] and 0.4-0.42 eV computationally [109, 110], suggesting that up to 20% of surface sulfur sites on FeS2 (100) may be vacant at ambient temperatures of 298 K, and therefore V S electronic states are prevalent. Moreover, neutral Fe vacancies V Fe on the surface have been imaged at the atomic scale by STM and shown to comprise a comparably high fraction of the surface [149]. Via DFT, Zhang et al. predicted a maximum surface Eg of 0.72 eV for stoichiometric (Fe:S = 1/2) FeS2 (100), but only 0.56-0.71 eV and 0-0.3 eV for sulfur-deficient and sulfur-rich surfaces, respec76 Table 3.1: Calculated bulk band gap Eg , and surface Eg for pristine and defective FeS2 (100). Defective surface here refers to the presence of a single sulfur vacancy V S in a single 1 x 1 unit surface supercell. Eg (eV) Bulk 0.87 1.02 0.86 0.90 Pristine Surface 0.40 0.56-0.71 0.55 0.60 Defective Surface 0.27 N/A 0-0.2 0.0 Ref. [132] [110] [108] [109] Table 3.2: Experimental surface Eg measurements by scanning tunneling spectroscopy (STS). Sample/Surface Type Natural, fractured in UHV Natural, fractured in air Synthetic, as-grown surface Synthetic, fractured in air Surface Eg (eV) 0.04 0.20 0.95 0.00 Ref. [151] [152] [50] [153] tively. Other authors have theoretically calculated that V S at the surface can reduce the surface Eg by more than this, even making the surface metallic [109]. Such arguments have been used, for example, to explain the low resistivity (O(10-1) Ω.cm) of manufactured pyrite thin films for PV applications [150]. Despite this recognition that FeS2 (100) interfaces are non-stoichiometric, there remains a need to demonstrate experimentally the effect of defects on the electronic structure. The scanning tunneling microscope (STM) operating in ultra high vacuum (UHV) provides a controllable metal-vacuum-semiconductor tunnel junction to probe these electronic states at the surface. A limited number of STS studies on natural [151,152]and synthetic [50, 153] FeS2 single crystals have produced inconsistent results, with apparent band gaps ranging from ∼ 0 eV to the accepted bulk value of 0.95 eV (Table 3.2), and a lack of detailed insight into the nature of the pyrite surface states. 3.2 3.2.1 Methods Experimental FeS2 single crystal synthesis High purity single crystals of FeS2 were synthesized by chemical vapor transport (CVT) in closed quartz ampoules, based on techniques described in Refs. [139, 154]. Raw materials were procured from Alfa Aesar (Haverhill, MA). A 1:2 stoichiometric mixture of 99.999% pure Fe powder and 99.995% S granules totaling 4 g – along with ∼0.3 g of 98% pure anhydrous FeBr3 - was sealed in an evacuated, 20 cm long quartz tube and heated to 700o C for 15 days to form polycrystalline pyrite aggregates. This precursor pyrite was removed, cleaned in acetone and methanol and resealed in a similar quartz tube with 0.3 g of fresh FeBr3 and a small amount of solid sulfur to provide a sulfur-rich environment for single crystal growth. The quartz tube was placed in a temperature gradient from 700 to 550o C, with the polycrystalline pyrite charge placed at the hot end, and left for up to 30 days. The mechanism of pyrite growth by CVT is described in Ref. [24]. The resulting crystals were typically cuboidal in shape with 5-10 mm edge lengths (Fig. 3-3a) and predominant {100} growth faces as determined by electron backscatter diffraction (EBSD) and single crystal x-ray diffraction (XRD) (Fig. 3-3a). As-grown crystals were checked for phase purity using Raman spectroscopy (Fig. 3-3b) and were found to be n-type semiconducting, with a donor concentration N D in the range 1-5 x 1016 cm-3 by Hall measurement at 21o C. In addition, an indirect Eg of 0.9-0.95 eV was detected by optical absorption on FeS2 single 77 (c) Absorption spectrum: optical Eg > 0.9 eV 1000 (200) 800 30 40 50 60 70 80 90 100 2θ (b) Raman spectroscopy Intensity (arb.) 344 300 -1 α (cm ) Intensity (arb.) (a) Single crystal XRD 600 400 380 200 418 350 400 Raman Shift (cm-1) 450 0.7 0.8 0.9 Energy (eV) 1 Figure 3-3: FeS2 single crystal samples. (a) Co-Kα X-ray diffraction (XRD) pattern. (b) Raman spectrum. (c) Absorption coefficient α measured as a function of photon energy, showing an optical (bulk) band gap Eg = 0.9-0.95 eV. Inset: photograph of FeS2 single crystals prepared by chemical vapor transport (CVT). Each square on the background corresponds to 10 x 10 mm2 . crystals polished down into 200 µm-thick plates (Fig. 3-3c). Absorption measurements were performed with a Perkins Ellmer LAMBDA 1050 Uv/Vis spectrophotometer. Scanning tunneling microscopy and spectroscopy Scanning tunneling microscopy (STM) was carried out using an Omicron VT-AFM system (Omicron Nanotechnology, GmbH, Germany) under UHV at pressures in the 10-10 Torr range. We used electrochemically etched Pt-Ir tips that were annealed at 150o C for 2 hours under UHV to remove absorbed H2 O and hydrocarbons prior to taking measurements. Single crystal, {100} growth faces of FeS2 were investigated by STM subsequent to ex situ cleaning by the following procedure: sealed quartz tubes containing freshly-grown crystals were opened in a glove box under a high purity, 95% N2 – 5% H2 environment to control surface oxidation and were ultrasonically cleaned in acetone and methanol to remove residual Br2 , which was proposed to be a source of contamination in previous STM studies of synthetic pyrite [153]. Samples were clamped in a custom made aluminum stage and transferred to vacuum within < 1 min to minimize exposure to laboratory air. STM and STS results from samples prepared in this way were compared with similar data obtained using in situ fractured, synthetic FeS2 single crystals which are known to have stepped, (100)-oriented faces [155, 156]. The STS results from as-grown and in situ fractured surfaces were quantitatively indistinguishable. X-ray photoelectron spectroscopy X-ray photoelectron spectroscopy-valence band (XPS-VB) spectra were obtained at the U12A beam line of the National Synchrotron Light Source (Brookhaven National Laboratory, Upton, NY), using a photon excitation energy of 210 eV. Single crystal growth faces of FeS2 were prepared in a similar fashion as described above and were cooled in situ under UHV to approximately -170o C before performing XPS-VB measurements. 78 Table 3.3: Input parameters for tunneling spectroscopy simulations using the SEMITIP program. (a) bulk Hall measurement on FeS2 single crystals in this work, (b) effective masses optimized from DFT-computed band structure; (c) contact potential estimated from DFT calculation of work function, supported by experimental evidence in Ref. [24]. Property Donor concentration (a) CB effective mass (b) Heavy hole effective mass (b) Contact potential (c) Tip-sample separation Tip radius 3.2.2 Symbol ND mc mhh ∆ϕ s r Value Used 1x1016 0.09-0.15 me 0.6-2.0 me 1.0-1.2 0.8-1.0 50 Unit cm-3 N/A N/A eV nm nm Computational Density functional theory (DFT) Density functional theory (DFT) calculations for this study were carried out by Aravind Krishnamoorthy. The full details on DFT computational methods on the FeS2 system can be found in Refs. [106, 108]. Tunneling current simulation (SEMITIP) Computations of tunneling current for simulating STS data were carried out using the full three-dimensional (MultInt3) version of the open-source program SEMITIP v.6, courtesy of R.M. Feenstra [157]. The program is a Poisson solver that treats the case of a hyperbolic-shaped tip in tunneling contact with a semiconductor sample. A complete description of the physics involved in the calculations is given in Refs. [115,116,158–160]. Table 3.3 summarizes the key input parameters for tunneling current calculations that are related to the electronic properties of the tip and sample, and the geometry of the tunneling simulation. Given the large number of input variables, we found an efficient approach to modeling proceeded along the following routine: first, all known variables are assigned their experimentally or computationally measured values. Second, the tip-sample separation distance s and tip radius R were estimated based on previous literature [115, 122]. Finally, we allocated to the remaining free variables a physically realistic range of values and performed a sensitivity analysis to optimize the fits (see Appendix 5). In practice, it was found that only the major semiconductor properties such as donor concentration N D , conduction band effective mass mc and heavy hole effective mass mhh , along with the contact potential ∆ϕ, had a significant quantitative influence on the model output of tunneling current. In the tunneling spectrum model, we accounted for the existence of charge accumulating surface states on FeS2 by introducing them explicitly into SEMITIP, either as a pair of Gaussian-distributed functions (Figure 3-4a), or as a uniform band across a predefined energy range (Figure 3-4b). For each of these surface state distributions, we fixed the charge neutrality level EN . Here, EN connotes the energy level below which states are neutral when filled and positively charged when empty, or, conversely, above which they are negatively charged when filled and neutral when empty. In the case of the double Gaussian distribution, the additional variables of centroid energy (the displacements of the states in Energy either side of EN ) and the full width half maximum (FWHM) of the peaks were assigned optimized values for fitting (see Appendix 5). It is important to note that surface states in the tunneling model are treated as completely localized at the surface, i.e., their magnitude does not decay exponentially into the bulk. Surface states thus affect only the electrostatic potential part of the calculation and are not included in the computation of tunneling current. 79 Possible surface state distributions in SEMITIP (a) (b) CB CB EC EC FWHM EF EF EN EN Centroid E EV VB VB EV Figure 3-4: Distributions of surface states as defined in the SEMITIP program: (a) uniform distribution, with charge neutrality level EN and (b) double Gaussian distribution, where FWHM is the full width half maximum of the peaks, and the centroid energy defines their separation either side of EN . Filled states are shaded in grey. VB and CB refer to the bulk valence and conduction bands respectively. 3.3 3.3.1 Results and Discussion Current-separation and current-voltage tunneling spectroscopy STS results were obtained experimentally on single crystal FeS2 (100), measured at various tip-sample separation distances s. Due to the well-known exponential dependence of tunneling current itunn on s, the onset of detectable tunneling current either side of 0 V bias (nominally the VB and CB edges), which give rise to an apparent surface Eg in the data, depends on the initial set point tunneling conditions for STS acquisition. Therefore we normalize the data to the constant tip-sample separation so at which a consistent "gap"’ of approximately 0.5 eV is visible. However, we explain why the quantification of Eg directly from STS spectra in this manner can be misleading, since it does not account for the phenomenon of TIBB, as described in Section 3.1.3. Stable STM images were initially taken at relatively low magnification (500 x 500 nm2 ) to locate sizeable flat terraces for consistent STS data acquisition (Figure 3-5a). The tip was subsequently scanned over 20 x 20 nm2 , or smaller, atomically-flat areas (Figure 3-5b) to obtain tunneling spectroscopic information at various set point currents (iset ) and biases (V set ). The tip was then briefly paused over randomly selected points during which the feedback loop was turned off for 1 ms to acquire current-separation i(s) or current-bias i(V) spectra. The magnitude of the measured tunneling current im as a function of bias voltage V is affected by the vertical tip displacement at the instant of STS acquisition. This separation distance s can be related to the setpoint conditions iset and V set through the simple exponential decay relation i(s) = io ex p(−2κs), where io is a constant and κ is the vacuum tunnel coefficient, otherwise known as the decay constant. κ is approximated for one-dimensional tunneling and reasonable V set by [123, 161]: κ= v t 2m e |eVset | 2 B− + kq 2 (3.8) where me is electron mass, B the effective tunneling barrier and kq the parallel wave vector of the tunneling electrons. The decay constant κ for pyrite was determined via i(s) spectroscopy at a range of different setpoint biases. Figure 3-6a shows the i(s) response at V set = -1.4 V (main image) and V set = 0.4, 1.2 and 2.0 V (inset), each averaged over approximately 20 measurements at different points on the FeS2 sample surface. The magnitude of κ over the range -2V ≤ V ≤ 2V varied linearly from approximately 0.3 Å-1 at large bias to 0.5 Å-1 near 0 V, and was symmetric for negative and positive bias 80 Square pyramidal FeS2(100) surface imaged by STM z (nm) (b) +4 (a) z (pm) +25 0 Fe S -4 0 -25 100 nm 0.5 nm -8 -50 Figure 3-5: Scanning tunneling spectroscopy (STM) images of the as-grown FeS2 (100) surface: (a) showing large atomic terraces with step edges oriented along the <100> direction. Tunneling conditions: V set = - 1.5 V, iset = 0.5 nA. Scanning tunneling spectroscopy (STS) was performed on selected 20 x 20 nm2 flat areas, (b). Fe atoms are resolved on the FeS2 (100) surface. The inset figure displays an atomic model for comparison, with one unit cell of Fe atoms outlined by the dashed square. Tunneling conditions: V set = 0.2 V, iset = 4 nA. (Figure 3-6b). The average effective tunneling barrier B, calculated using Eq. (3.8) and assuming k=0, was 1.2 eV. This corresponds to the average work function between the metallic tip and the pyrite sample at the tunnel junction. Figure 3-7a displays a series of individual i(V) spectra taken at four different values of s, where the set point V set = 1.5 V and iset = 200 pA was arbitrarily chosen as the reference separation so . The other values of s were calculated relative to so using the exponential decay relation for tunneling with the experimentally-determined κ from i(s) spectroscopy. At more positive s (larger tip-sample separation), the measured current around 0 V becomes very small and the Eg appears larger, up to approximately 1.7 eV for s = so + 1.8 Å. To correct for the exponential decay in transmission coefficient for tunneling, the i(V) data are normalized in Figure 3-7b to a constant tip-sample separation distance by converting the measured current im to “distance corrected current” is = im e x p [2κ(V ) s], where s = 0 at the reference separation distance so . Separation distance-normalized data are displayed with a logarithmic current scale to enable discrimination among spectra. The four curves overlap consistently, indicating that throughout the tunneling set point range used in this work the tunneling spectra give a true representation of the tunneling response without metallic behavior due to point contact at very small s, or anomalously insulating behavior at large s. Further, we normalize the data to normalized conductance (di/dV ) /(i/V ) (Figure 3-7c) which is known to approximate the DOS in semiconducting or metallic samples [113]. We calculated (d i/d V )by numerical differentiation from the i(V) response. To correct for the well-known divergence of the direct conductance (i/V) at small values of i, i/V was broadened to i/V using a Gaussian distribution described previously [114]. A first approximation of surface Eg from the i(V) response is 0.5 eV, obtained by taking the average voltage separation between the CB and VB current onsets at 1 pA current, which is approximately the instrument resolution or "noise floor" below which current is not reliably measured in the STM used for this study (Figure 3-7b). Nevertheless, the direct quantification of surface Ebg in this manner does not account for the possible occurrence of TIBB, as described in Section 3.1.3. 81 Determine decay constant κ: 0 (a) i =o i exp(-2κs) Experimental i(s) -0.4 ln(i) (nA) i (nA) -0.2 -0.6 -0.8 -1 0 1.2 -2 -4 -6 2 2 2.0V 1.2V 0.4V 6 4 Δs (Å) 4 s (Å) 6 8 (b) 2κ (Å -1 ) 1 0.8 0.6 0.4 -2 1 -1 0 Sample Bias (V) 2 Figure 3-6: Current-separation spectroscopy: (a) Tunneling current i as a function of tip-sample separation s at a set point bias of -1.4 V (solid line). A fitted exponential function i = io ex p (−2κs) with decay constant κ = 0.80 Å-1 is overlaid on the experimental data (open circles). Inset: experimental data for 0.4, 1.2 and 2.0 V biases, plotted on a log scale. (b) κ variation across the bias range used in this work. The dashed lines are to guide the eye, and do not represent a fit to the data. 3.3.2 Simulated tunneling spectra based on DFT-calculated DOS We interpret the underlying electronic structure in our measured STS results on FeS2 by simulating the tunneling spectra using an explicit calculation of the electrostatic potential across the tip-vacuum-pyrite system, followed by a full numerical integration of the resulting tunneling current. Using DFT as a guide for the position and distribution of intrinsic and defect-related surface states, we explored several different configurations of surface electronic structure as the input for the tunneling spectra computations and optimized the fit to the experimental STS data in each case. We first compare the DFT-calculated DOS for pyrite with the valence band spectrum of a synthetic sample, measured using synchrotron x-ray photoelectron spectroscopy (Figure 3-8). A prominent, Fe 3d-related band and the broad, hybridized Fe 3d and S 2p states between 1-7 eV below EV [107,162] are clearly visible in both the experimental and theoretical data, indicating a general correlation which justifies the use of this DFT data in guiding our analysis. To investigate the origin of the apparent 0.5 eV surface Eg in the STS results we considered the calculated DOS in energy region surrounding the bulk band gap (approximately EV - 0.5 eV ≤ E ≤ EV + 1.5 eV) and present here the results for four different simulated electronic structures, which could conceivably give rise to the experimental tunneling spectra. These four models are based on DFT calculations for the bulk crystal (Figure 3-9a), a pristine (stoichiometric) surface (Figure 3-9d,g), and a defective surface containing both charge neutral V Fe and V S (Figure 3-9j). For the purposes of simulating the tunneling current as a function of bias, each of these characteristic DOS distributions was converted to a simplified representation with inputs for the bulk valence and conduction bands and the requisite surface states. The computed tunneling 82 (c) |current| (pA) (b) so-1.2Å 800 400 0 -400 -800 Raw i(V) data 1.7 eV so so+0.9Å so+1.8Å 0.5 eV Distance corrected i(V)/exp[-2κ.s] 102 101 1 pA noise floor 100 Normalized conductance (di/dV)/(i/V) (arb units) (a) current (pA) Precise Eg not quantifiable straight from i(V) curves -2 -1.5 -1 -0.5 0 0.5 1 Sample Bias (V) 1.5 2 Figure 3-7: Current-voltage spectroscopy: (a) Tunneling current-voltage i(V) curves measured on FeS2 (100) with varying tip-sample separations s. so corresponds to a tunneling set point of V set = 1.5 V and iset = 200 pA. (b) the same i-V data normalized to constant tip-sample separation, plotted on a logarithmic i axis to facilitate comparison. (c) normalized conductance (di/dV ) /(i/V )as a measure of local density of states (DOS). An estimate of surface band gap Eg width using the instrument resolution of 1 pA is ∼0.5 eV. Fe-3d states at valence band edge DOS (arb. units) XPS hν = 210 eV DFT calculated Iron Sulfur Total -10 -5 0 E-EV (eV) 5 Figure 3-8: Pyrite valence band: Experimental valence band spectrum obtained using synchrotron x-ray photoelectron spectroscopy (XPS) at photon energy hυ = 210 eV (top) compared against DFT-computed density of states for FeS2 (100) (bottom), also showing partial DOS contributions from iron and sulfur. The dashed line marks the valence band edge. current that resulted from the different simulated DOS representations was compared to the experimental STS results from Figure 3-7. The four theoretical electronic structures that were matched to the experimental data can be described in more detail as: 1. FeS2 bulk-like density of states as calculated by DFT (Figure 3-9a). In the simpli83 fied STS model, the VB and CB extrema are separated by an assumed bulk band gap of 0.95 eV and no surface states exist (Figure 3-9b). 2. Pristine FeS2 (100) surface density of states, including intrinsic surface states arising from Fe dangling bonds (Figure 3-9d). In the theoretical tunneling spectra input (Figure 3-9e), a double Gaussian distribution of surface states is included, straddling the VB and CB edges. The density of surface states is set to 6.8 x 1014 cm-2 .eV-1 , consistent with the density of Fe dangling bonds at the unreconstructed (100) surface. The charge neutrality level EN is fixed exactly halfway between the VB maximum and CB minimum, while the FWHM and centroid energies of the surface states were optimized within a reasonable range to provide the closest match to experiment. In this model, the intrinsic surface states can accumulate or deplete in charge depending on the applied bias, and thus serve to screen the tip potential, but do not produce itunn . Eg is still 0.95 eV at the surface. 3. Pristine FeS2 (100) surface density of states, similar to that described in Model (2), but we now postulate that intrinsic surface states are homogeneously connected to the bulk VB and CB. Therefore, surface states are not explicitly defined in this model, but rather the bulk Eg in the input is decreased from 0.95 to 0.5 eV, to approximate the tunneling contribution of intrinsic surface states (Figure 3-9h). It is important to note that no surface states were explicitly defined in the computations employing this model so no EF pinning would be expected. We include the gray, double Gaussian states in Figure 3-9vh merely to draw the eye to how they effectively reduce the surface Eg . 4. Defective FeS2 (100) surface density of states. The surface Eg is reduced to 0.5 eV by the intrinsic surface states, as in Model (3), but we also include defect states from V Fe and V S with 12.5% surface coverage each, i.e. density of 8.5 x 1013 cm-2 .eV-1 (Figure 3-9j). In this theoretical tunneling spectrum model, the defect states form a broad band across the width of the reduced Eg and only contribute free charge to screen the tip potential, without contributing further tunneling current (Figure 3-9k). The results for the four different tunneling spectra simulations are presented adjacent to their corresponding DOS graphs in Figure 3-9. Together with each simulated curve we also show the same, repeated experimental STS result obtained using set point tunneling conditions of V set = 1.5 V and iset = 200 pA, equivalent to the reference tipsample separation of so in Figure 3-7. As expected, the first model of bulk-like density of states (Figure 3-9c), which excludes any surface states, gives a poor fit to the experimental results. In the absence of available free charge at the surface to pin EF , the semiconductor bands are free to shift with the applied bias, and the CB (VB) edge is dragged to higher (lower) energies with increasing positive (negative) bias. The apparent surface Eg width is therefore > 0.95 eV. The second model (Figure 3-9f) with surface-localized, non-tunneling intrinsic states seems to approximate the experimental result better, but still does not reliably capture the small size of the zero tunnel current region, nominally corresponding to Eg , which is 0.3-0.5 eV larger than the experimental STS results would suggest. A parametric sensitivity study was conducted, and no adjustment of the relative surface state positions or widths produced a fit better than that shown in Figure 3-9f, as quantified by minimizing the root mean squared (RMS) difference between the simulated and experimental results (see Appendix 5). We conclude that the effect of these intrinsic surface states extends beyond a simple screening of the tip potential distribution, as they are defined in the tunneling spectrum model. Consequently, close replication of the experimental spectra can instead be achieved by defining a narrower forbidden energy region Eg in the model input, simulating the case where significant tunneling current originates from the surface states when the bias is swept across biases in the range of approximately 0.4 V and below. In Figure 3-9i 84 85 (j) (g) (d) (a) 0 EV 0.5 1 E - EV (eV) Intrinsic & defect SS CB Screen potential Tunneling 0.5 1 1.5 E - EV (eV) (k) E = 0.5 eV g (h) (e) VB Eg = 0.95 eV 1.5 -0.5 0 Tunneling from intrinsic SS Intrinsic SS screen tip EC Bulk-like states (b) -500 0 -500 -2 0 500 (l) (i) (f) 500 (c) -1 0 1 Sample Bias (V) Experiment STS Model 2 Figure 3-9: Modeling tunneling spectroscopy with surface states: (Left column) density of states (DOS) calculated by density functional theory (DFT) with (middle column) their corresponding, simplified analogs used in tunneling spectroscopy modeling. The shaded regions correspond to surface states (SS’s) that either just screen tip potential (black) or, alternatively, contribute to the tunneling current (grey) in the tunneling spectra model input. (Rightmost column) the output from each tunneling spectrum model (solid line) is displayed alongside the same, repeated experimental scanning tunneling spectroscopy (STS) result (open circles), corresponding to a tip-sample separation of so as described in the main text. (a), (b) and (c): bulk-like electronic structure with band gap Eg = 0.95 eV and no SS’s. (d), (e) and (f) include intrinsic SS’s at the FeS2 (100) surface, bordering the band edges but not contributing to tunneling current (itunn ) within the fundamental, bulk Eg . (g), (h) and (i): the intrinsic SS’s now mediate tunneling and the band gap is reduced to 0.5 eV. (j), (k), and (l) correspond to a FeS2 (100) surface with reduced Eg of 0.5 eV containing 12.5% iron point defects (V Fe ) and sulfur monovacancies (V S ). -0.5 Density of States (arb. units) Progressively add computed surface states into STS model to match experimental result Current (pA) Intrinsic surface states connected; extrinsic form discrete band 1.5 Energy (eV) (a) (b) (c) 1 0.5 Defect State 0 -0.5 L Γ X K L L Γ X K L L Γ X K L k point Figure 3-10: Density functional theory (DFT)-computed band structures for (a) bulk FeS2 , (b) a 4-layer surface slab with (100) termination, where we show in the red region the additional bands arising from intrinsic surface states, and (c) the same surface slab but containing a single sulfur vacancy V S in the topmost layer. The single band coming from the defect is highlighted with a bold, blue line. the intrinsic surface states are included continuously with the VB and CB in this manner, i.e., the Eg for the simulation is reduced now to only 0.5 eV. This input produces simulated tunneling currents that match the experiment much better around the zero current region. However, the best fit to the experimental data was achieved when we further include a broad, distributed band of defect states that simulate a 12.5% concentration of both V Fe and V S , at the surface (Figure 3-9l). In the model these defect states do not mediate electron tunneling but their effect on the simulated spectrum is to reduce the effect of TIBB by screening the tip potential. Physically, we interpret this to mean that the defect surface states that exist within Eg are too dilute and sparsely distributed to contribute significantly to tunneling and thus further reduce Eg . Nevertheless, their presence sufficiently affects the tunneling spectra through EF pinning. To help understand the lack of measured tunneling current from inter-band defect states, in Figure 3-10 we compare the band structure of bulk FeS2 alongside that of a pristine (100) surface and a defective (100) surface containing a 12.5% concentration of V S , as calculated by DFT. The additional, intrinsic surface state bands - highlighted in red in Figures 3-10b,c - form a dense network of states that overlap continuously with the bulk VB and CB. By comparison, the V S defect state is manifested in a single, isolated band 0.2-0.3 eV below EC . Moreover, we see the minimum energy point on this defect band clearly dips down at the L point of the Brillouin zone. During the tunneling process, the perpendicular wave vector for the electron k⊥ is relatively small compared to those from intrinsic surface bands, where the empty-state minimum is very flat across the entire k point range shown. Together, these facts suggest the tunneling from intrinsic surface states would be much stronger than for the defect states, explaining why we observe the intrinsic surface states directly in the experimental tunneling spectra, but the defect states have a more subtle effect. In order to quantify the width of the surface Eg from the experimental data, a sensitivity analysis of Model (4) was performed with the range 0.3 ≤ Eg ≤ 0.6 as the input. The results plotted with logarithmically displayed current in Figure 3-11 indicate a best fit to experimental tunneling current can conceivably be achieved with Eg = 0.4 ± 0.1 eV. The close match between the experimental and the simulated tunneling spectrum results suggest that the presence of dangling bond surface states on the FeS2 (100) free surface leads to a reduction in Eg by ∼0.5 eV over the accepted bulk Eg of 0.95 eV. These 86 Surface Eg = 0.4 ± 0.1 eV 8 10 0.3 eV Expt. Model 6 Tunneling current (pA) 10 0.4 eV 4 0.5 eV 10 0.6 eV 2 10 0 10 -2 -1 0 1 Sample Bias (V) 2 Figure 3-11: Fitting to experimental surface Eg . Comparison of experimentally measured tunneling currentvoltage i(V) spectrum (open circles, V set = 1.5 V and iset = 200 pA) , with defective FeS2 (100) surface Model (4) and four different magnitudes of Eg as listed beside each curve: 0.3, 0.4, 0.5 and 0.6 eV. The absolute values on the tunneling current scale are arbitrary and the curves have been separated by a uniform multiplication of the current data for ease of comparison. We have omitted showing the experimental and modeled data below a current of 1 pA which is the instrument resolution for experimental data acquisition. intrinsic surface states form continuous bands connected to the bulk electronic states and therefore offer available energy levels into which and from which electron tunneling can occur in the presence of the biased probe tip. In contrast, surface states arising from distributed point defects do not contribute to electron tunneling during STS, but instead provide states across a broad range of energies within the fundamental surface Eg that can accrue additional charge from the bulk. This acts to pin the Fermi level and moderate the amount of TIBB during STS. By accounting for both of these surface contributions in tandem, the theoretical model most accurately replicates the experimental STS data. The requirement for such a seemingly high surface defect concentration of 12.5% in the model raises the question of whether this is realistic for a pristine FeS2 (100) surface. The existence of a vacant site at one in every eight surface-bound sulfur sites, with a correspondingly similar density of vacant iron sites, would imply one cation and one anion vacancy per four surface unit cells. As we have already discussed, there is a large body of XPS evidence suggesting that up to 20% of pristine pyrite surfaces can comprise V S , even at ambient temperatures of 298 K, including our work in parallel to this study [106]. Likewise, high concentrations of surface V Fe have been imaged on pyrite by STM [149]. It is probable that the synthetic samples used in the current work contain a significant concentration of V Fe , given the sulfur-rich conditions of single crystal synthesis. Therefore the assumption of 12.5% as an average for both types of ionic point defect seems plausible. In addition, the effect of extrinsic (S-Br)2defects - a known source of impurity in CVT-grown pyrite [163] - has not been considered in our analysis. Although the surface states associated with these defects would be localized at the defects themselves, there is evidence to suggest that vacancies (as well as other defects such as step edges, substitutional dopants and intersecting dislocations) can affect the electronic structure at nanometer distances through the redistribution of 87 Delocalized defect state Figure 3-12: Visualization of FeS2 (100) surface charge q surrounding a single sulfur point defect (V S ) located at the arrow. We show only the difference in q between the defective and defect-free surfaces, i.e. ∆q (VS − S). Positive and negative 0.014 e/Å3 isosurfaces are colored in red and blue, respectively. Unit cell edges are shown as dashed lines; one defect in this area corresponds to a V S concentration of 12.5%. surface charge [164–166]. Figure 3-12 displays the charge density difference between a supercell of four unit cells of FeS2 (100) containing a charge-neutral V S defect and that of the perfect host, with an isosurface of 0.014 e/Å3 . The delocalized positive and negative charge resulting from the V S defect is distributed over almost the entire 2 x 2 unit cell area of the simulation, equivalent to approximately 10.8 x 10.8 Å2 on the surface. This extent of delocalization would suggest that the charge from a density of only 12.5% V S on the surface could affect the experimental tunneling spectroscopy measurement, regardless of the exact location that the probe tip is placed during the data acquisition. 3.4 3.4.1 Outcomes Conclusions An accurate, quantitative assessment of the surface electronic structure of semiconducting pyrite, FeS2 (100), is necessary for understanding the behavior of pyrite in a wide range of applications, including geochemical, bio-catalytic, and corrosion processes, as well as pyrite’s photovoltaic and photoelectrochemical properties. While scanning tunneling spectroscopy is an ideal tool for this purpose, the analysis is complicated by the well-known effect of tip-induced band bending, the presence of intrinsic surface states, and the additional effects of defect states associated with native ionic vacancies and other defects. We performed systematic STS measurements on synthetic FeS2 single crystals at different tip-sample separations and demonstrated that the apparent surface band gap is consistently 0.4 ± 0.1 eV, or ∼0.55 eV smaller than the widely-accepted bulk band gap of 0.95 eV for pyrite. By basing our tunneling current simulations on methodically varied, simplified DFT-calculated electronic structures, we link the origin of this reduction in Eg to the presence of intrinsic surface states from Fe dangling bonds at the free surface termination. The electronic bands arising from the intrinsic surface states overlap continuously with the bulk bands. In addition, the experimental tunneling spectra results can be modeled most accurately if a second distribution of surface states arising from cation and anion vacancies is incorporated into the tunneling current simulations. These defect states do not contribute significantly to overall tunneling current but have an influence on the tunneling spectra by accumulating charge at the pyrite surface, which screens the tip potential during measurement and pins the Fermi 88 Surface states affect photovoltaic performance Into crystal (100) Free surface Figure 3-13: Low surface bandgap implications for PV. A smaller surface band gap may lower overall pyrite solar cell open circuit voltage and therefore performance by supplying a connected set of energy levels into which excited states can thermalize before being usefully extracted. Image from Ref. [167]. level. Our findings confirm the influence of both intrinsic and defect surface states on the electronic structure of pyrite. The presence of a reduced band gap on the surface, as well as the existence of defect states within the band gap, has implications for electronic processes such as charge transfer during electrochemical redox reactions. 3.4.2 Implications for other applications of FeS2 , e.g. PV One of the several motivations for this work, outside the central scope of this thesis, was the applicability of FeS2 to photovoltaic (PV) solar cells. In a paper written by Lasic, Armiento et al., to which the author of this thesis contributed, an attempt was made to explain the low open circuit voltage (OCV) of pyrite devices, which have not exceeded 0.2 V (out of a potential 0.95 V maximum as dictated by the bulk Eg ). [167] One argument was that the surface states responsible for the lowered surface Eg are connected directly to bulk bands in pyrite. Any photoexcited carriers could then thermalize into lower energy levels at interfaces, before being usfully extracted (Figure 3-13). Since this work was published, another group advanced a theory for a conductive inversion layer at the FeS2 (100) surface, essentially comprising positively-charged surface states. [168] The authors suggested this would be consistent with our STS measurements in this work, and provided evidence that surface states can be passivated, offering hope that pyrite devices with OCV > 0.5 V may one day be feasible. 3.4.3 Future work The next step in improving the methodology for quantifying surface band gaps and surface states directly from scanning tunneling spectroscopy is to modify the SEMITIP code for a direct input of DFT data. Such a collaboration was discussed with Prof. Feenstra, the author of SEMITIP, but this was not pursued further during this thesis. However, the next generation of code should allow direct input of first-principles calculated surface density of states, rather than the relatively coarse, parabolic band inputs currently allowed by the program. Moreover, the code as it stands at the time of writing cannot compute tunneling current from surface states. The next version should be able to deal with tunneling from predicted, charged states that also pin the Fermi level. 89 (a) MoS2: facile STM over large areas z (pm) 50 0 20 Å -50 3.2 Å (c) Preliminary STS inconclusive -2.5V, 1nA MoS2 Bulk Crystal MoS2 Single Layer [di/dV] / [i/V] (arb. units) (b) MoS2 basal plane -2 -1 0 1 Sample Bias (V) 2 Figure 3-14: Preliminary investigations on bulk and 2-dimensional MoS2 . (a) stripped MoS2 offers highquality surfaces for scanning tunneling microscopy. (b) higher-magnification image, showing Mo atoms in hexagonal arrangement. (c) an initial study to compare the surface band gap of multilayered ("bulk") vs. single layer MoS2 was inconclusive; however the methodology for STS quantification outlined in this chapter offers a systematic way to further investigate this and other, similar open questions. Experimentally, we hope that the methodology outlined in this chapter will be used and improved by other researchers to test other interesting materials with poorly-characterized surface electronic properties. For example, as part of a collaboration with the Palacios group of the Department of Electrical Engineering at MIT, the author made some preliminary measurements of the MoS2 surface to compare the surface band gap of singlelayered versus multi-layered samples. Figure 3-14 presents some initial findings; the surface is very conducive to imaging by STM. However, the collected STS data was inconclusive without further analysis using a similar technique to that described in this chapter. The surface Eg of both multi- and single-layered MoS2 was found to be close to the known bulk value of 1.2 eV. Many other chalcogenide and other rare semiconducting surfaces should be convenient for such coupled experimental and computational analysis of their surface electronic structure. 90 Chapter 4 Stability: dynamics of point defect formation, clustering and pit initiation on the pyrite surface Synopsis The collective behavior of point defects formed on the free surfaces of ionic crystals under redox conditions can lead to initiation of local breakdown by pitting. Here, we controllably generated sulfur vacancies on single crystal FeS2 (100) through in vacuo annealing, and investigated the resulting evolution of surface chemistry using synchrotron x-ray photoelectron spectroscopy (XPS). By measuring the S 2p photoemission signal intensity arising from sulfur defects as a function of temperature, the enthalpy of formation of sulfur vacancies was found to be 0.1 ± 0.03 eV, significantly lower than the reduction enthalpy of bulk FeS2 . Above 200 o C, the created sulfur vacancies together with preexisting iron vacancies condensed into nm-scale defect clusters, or pits, on the surface, as evidenced by scanning tunneling microscopy (STM). We provide a mechanistic description for the initiation of pits that requires concerted behavior of both the sulfur and iron vacancies, and validate this model with kinetic Monte Carlo (kMC) simulations. The model probes realistic length and time scales, providing good agreement with the experimental results from XPS and STM measurements. Our results mechanistically and quantitatively describe the atomic scale processes occurring at pyrite surfaces under chemically reducing environments, important in many natural and technological settings, ranging from its role as a passivating film in corrosion to its potential use as a photovoltaic absorber in solar energy conversion. Portions of this chapter were published in Electrochimica Acta [169]. The kMC simulations were developed by Aravind Krishnamoorthy. 4.1 Background and motivation As we have seen in Chapters 2 and 3, a stable passive layer can hinder ion or electron transport to/from the metal, and suppress charge transfer at the surface. However, the inherent protectiveness of a passive layer relies on this physicochemical barrier remaining intact. Passive films can break down due to chemical, electrochemical or mechanical stimuli, resulting in more rapid, localized corrosion that accelerates equipment failure. Localized degradation of the barrier layer can expose bare metal to the corrosive environment. In many cases, the metal will quickly re-passivate. However, in some instances a failure to re-passivate will result in the formation of a stable pit, which can rapidly 91 expand due to localized galvanic corrosion (the exposed, small area becomes anodic relative to the surrounding passivated metal). In the extreme case, a pipe wall may be compromised well before the expected service life of the structure. On stressed components, a pit may serve as the nucleation site for a critical crack. Aggressive anions in the environment such as chloride Cl- are known to promote pitting, as will be discussed in Section 4.1.2 below. Certain theoretical approaches have been developed to predict passivity breakdown, such as the point defect model (PDM) of Macdonald et al. which asserts that pitting is a deterministic process that can be modeled by individual, atomistic processes such as point defect formation and condensation. [32] Despite this, local corrosion rates are often non-linear and highly unpredictable. 4.1.1 Chapter goals In this chapter, we aim to understand: (1) the dynamic process of pit nucleation in ionic passive films at the atomic scale, by studying the surface of pyrite as a model system and (2) whether the reported off-stoichiometry in polycrystalline pyrite can be attributed to easily-reducible surfaces. Both these aspects are introduced in more detail in the following Sections 4.1.2 and 4.1.3: 4.1.2 Passivity breakdown by pitting The first and foremost motivation for this study was to uncover the atomic-scale mechanism behind vacancy condensation into nanometer-scale imperfections that could serve as nuclei for the localized breakdown of ionic passive layers, i.e. pitting. There is experimental evidence that vacancy condensation both at the metal-passive layer interface (Figure 4-2a and b) and the passive layer-electrolyte interface (Figure 4-2c) can lead to the localized breakdown of passivity. Various mechanistic theories and models have been proposed to describe this process [170], the most comprehensive of which is the Point Defect Model (PDM) which serves as the key, non-empirical framework guiding the development of a multiscale model in this project. The PDM has been developed to successfully account for the critical breakdown parameters (critical breakdown voltage, V c , and induction time, tind ) for single breakdown sites on the surface, for the distributions in these quantities, and for their pH- and chloride concentration-dependencies. The PDM also provides the first, mechanistically-based explanation of the role of alloying elements in the inhibition of passivity breakdown and hence localized corrosion. Despite a long-standing correlation of global environmental variables to corrosion rates, microscopic mechanisms of passivity and breakdown are not quantitatively understood. Consequently, localized corrosion and the associated macroscopic failure are typically, but incorrectly, identified as being stochastic in nature. However, passivity breakdown is not a matter of chance, but occurs for a well-defined, mechanistic reason, i.e. a deterministic process. A schematic illustrating some of the proposed causes of pitting is shown in Figure 4-1, with a brief description of each provided in the caption. Below we briefly review some of the motivations for the study of surface point defects in this chapter: Aggressive anions and point defects. Chloride is the most commonly observed and well-known pitting agent, although other halides and SO4 2− can also have an accelerating effect on localized corrosion. [3] While the presence of chloride ion is known to lead to localized attack, preferentially at structural and chemical discontinutities in the barrier layer such as grain boundaries and inclusions [171], the key mechanism responsible for the X − induced pitting is still debated. Two broad, descriptive models account for the action of X − on locally degrading the oxide film: the PDM suggests that anions absorb into oxygen vacancy sites at the oxide surface, leading to the creation of cation vacancies in the oxide (Figure 4-1a). [170] This increase in cation point defect 92 Proposed pitting mechanisms (a) (b) (c) Figure 4-1: Proposed mechanisms of passivity breakdown and pitting. (a) Halide ions X − preferentially adsorb at inhomogeneites in the barrier layer such as nanopits and vacant lattice sites, leading to the creation of cation vacancies. The Point Defect Model (PDM) postulates that the resulting, enhanced metal vacancy flux to the alloy interface cause voiding. (b) enhanced dissolution at surface defects can lead to blistering and rupture due to film growth stresses σ, to expose bare metal. (c) If the metal does not repassivate, a stable pit may grow. Hydrolysis of cations enhances local acidity, attracting more negatively-charged X − ions and propagating further pitting. Large pits may serve as nucleation sites for more catastrophic failure modes, e.g. stress corrosion cracking (SCC). concentration in turn leads to enhanced vacancy flux in localized channels in the oxide barrier layer; if the rate of metal vacancy accumulation at the metal interface cannot be matched by absorption into the metal, macroscopic voids may form at the interface. These cavities can subsequently chemically or mechanically destabilize the overlaying film, creating a pit on the nanometer scale. Surface imperfections as initiation sites. Open questions remain over halide-defect association as described above. However what is clear is that surface imperfections serve as initiation sites for localized corrosion. Microstructural or micro-chemical inhomogeneities in the passive film itself are known to play an important role in pitting, and even in the absence of aggressive anions these can lead to differential rates of dissolution which nucleate metastable pits. [172] Modeling pitting: realistic length- and timescales. Both on metals and on ionic passive films, clustering of point defects (vacancies) is thought to be necessary for pit initiation, the rate-controlling step in overall pitting corrosion [110]. This process has been observed on pure metals [173] and alloys. [174] On the other hand, understanding the surface pitting mechanism on ionic passive films has been a challenge, experimentally limited to only few successfully studied systems such as Ni-O, Ni-OH and Cr2 O3 , [175, 176] and without a concerted modeling and experimental demonstration at the atomic scale. Even insimpler systems such as pure metals and semiconductors, where surface pitting is relatively better understood, modeling of surface degradation is generally limited to the use of empirical kinetic parameters. Matching simulated time scales to experimental ones has remained challenging and has lacked experimental val93 Metal-barrier interface (a) Barrier-electrolyte interface (c) (b) Figure 4-2: Nanopits formed by vacancy agglomeration. (a) Scanning tunneling microscopy (STM) images of vacancy condensation to form nanocavities at intermetallic-oxide interfaces, leading to spallation of the passive layer. [179] (b) nanopits formed by localized dissolution (vacancy formation) at NiO surfaces, as recorded by electrochemical STM [175]. idation of model results. [177, 178] 4.1.3 FeS2 surface chemistry and non-stoichiometry The second motivation for this investigation, which is complementary to the first, regards the unresolved non-stoichiometry of pyrite surfaces, which cannot be explained based on defect chemistry of the bulk material. Native point defect concentrations in bulk FeS2 are generally low (O(106 ) cm3 ) at room temperature. [135] On the other hand, anionic vacancies, specifically sulfur vacancies denoted as V S , are expected to be far more prevalent at free surfaces, with calculated formation enthalpies as low as 0.4 eV. [109, 134] This has led to difficulties in obtaining surfaces with low intrinsic defect concentrations in nanocrystalline pyrite precursors and films. [133, 150] Sulfur deficiency is typically put forward as a source of non-ideal electronic and optical properties in synthetic FeS2 . However, there remains a need to experimentally quantify the formation energy of V S on the surface to understand whether these defects are indeed a significant source of off-stoichiometry in pyrite surfaces. As described in Chapter 4, the crystal structure of pyrite is NaCl-type cubic, with Fe2+ at the cation site and S22− dimers at the anion site, aligned along the cube diagonal <111>. The (100) surface of FeS2 is unreconstructed and is the most stable surface, as shown by low energy electron diffraction (LEED) [151,156] and scanning tunneling microscopy (STM). [149,181] The sulfur S 2p x-ray photoemission peak of pyrite, when accessedusing soft x-ray synchrotron radiation, reveals highly detailed information about the different binding environments of sulfur inthe near-surface region. In addition to the dimer S22− signal from the crystal bulk, pyrite’s S 2p photoelectron spectrum distinguishes two additional, surfacelocalized and coordinately reduced sulfur environments at more negative binding energies. [143, 145, 146, 148] Quantification of these surface-localized defect environments enables further understanding of the two open areas summarized above. 94 4.2 4.2.1 Methods Experimental Samples Single crystal pyrite samples were synthesized by chemical vapour transport (CVT) in the presence of Br as a transport agent, described in Chapter 4. Growth faces were identified to be primarily (100) by electron back-scattered diffraction (EBSD; ZeissSupra-55 scanning electron microscope). Phase purity of the single-crystalline synthesized pyrite was confirmed using Raman spectroscopy and X-ray diffraction. We note that the synthesis of FeS2 single crystals was performed under a high partial pressure of sulfur. Although the pyrite phase has a stoichiometric ratio of Fe/S = 0.5 in the bulk in ambient conditions, such sulfur-rich growth conditions are expected to favor the incorporation of iron vacancies in the bulk, consistent with previous calculations of defect formation energies as a function of sulfur chemical potential. [108, 135] After cooling the as-synthesized crystals at ∼ 5 o C/min under sulfur-rich conditions inside quartz tubes, a non-equilibrium high concentration of V Fe remainedquenched within the bulk FeS2 . We propose this quenched-in iron deficiency to be the source of mobile iron vacancies near/on the surface under the reducing environment (i.e., under a low chemical potential of sulfur) during our subsequent experiments. Elevated temperatures assist in the migration of V Fe to the surface from the bulk that has become supersaturated, as the bulk evolves toward thermal equilibrium with a lower vacancy concentration. The role of such iron vacancies on the surface is discussed in more detail in Section 4.3.2. Soft x-ray photoelectron spectroscopy (s-XPS) Synchrotron x-ray photoelectron spectroscopy (XPS) was performed at Brookhaven National Lab (Upton, NY) at the U12A beamline of the National Synchrotron Light Source (NSLS), in order to deduce the temperature dependence and the formation enthalpy of surface defects. The radiation source was tuneable, monochromatized soft x-rays in the energy range 100-600 eV with a resolution ∆E/E of 2 x 102 –103 and spot size of 1 mm2 on the sample surface. The base pressure in the vacuum chamber remained below 1010 Torr for the duration of the experiment. Clean single crystal FeS2 samples with (100)-oriented growth faces larger than 5 x 5 mm2 were prepared by ultrasonication in acetone then methanol in an inert atmosphere glove box, before being mounted in a mechanical sample clamp (Figure 4-3). During transfer into the UHV environment, the samples were limited to <5 min air exposure to minimize surface oxidation and other contamination. We did not observe any secondary peaks from sulfates or other oxidation products. [144] Controllable, in situ heating between 120-330 o C at steps of 30 o C was achieved using a resistively-heated coil placed behind the sample. The heating apparatus and sample stage were cooled collectively by flowing liquid N2 through the manipulator, which was attached to the sample via a large copper block. Each in situ annealing cycle lasted 150 min, which was adequate for the pyrite surface to reach equilibrium with the UHV base pressure; subsequently, the sample was allowed to cool to approximately -170 o C for XPS measurements in order to quench in the surface chemistry and also minimize phonon broadening of the XPS signal. S 2p photoemission spectra were obtained at excitation energies of 210, 350 and 500 eV with an energy resolution of 100 meV and pass energy of 10 eV. Peak fitting of the S 2p XPS spectra was carried out using CasaXPS Version 2.3.16. Shirley background subtraction was applied to all spectra and individual components were fit with 95% Gaussian-5% Lorentzian peak distributions, unless otherwise specified Scanning tunneling microscopy (STM) Scanning tunneling microscopy (STM) images were collected in order to visualize the evolution of surface defects at elevated temperatures. An STM system (Omicron VT95 (a) XPS sample holder (c) Sample clamp dimensions Screw for T.C. 12 mm 4 mm 20 mm (b) FeS₂ sample in vice (d) Side view 3.5 mm 3.2 mm 25 mm Figure 4-3: XPS sample clamp for FeS2 crystals.(a) circular holder for synchrotron x-ray photoelectron spectroscopy (s-XPS) equipment. (b) custom-built aluminium sample clamp, showing FeS2 single crystal with (100) face exposed. (c and d) sample holder design schematics with approximate dimensions. STM; Omicron Nanotechnology, GmbH, Germany) was used under UHV conditions below 109 Torr. Electrochemically etched PtIr tips were annealed in the chamber at 150 o C to clean them prior to taking measurements. FeS2 samples used for STM/STS were synthesized and prepared in a manner similar to that described above for the XPS experiments. STM images were subjected to a global flattening procedure and horizontal noise removal using SPIP 4.8.4 software from Image Metrology (Denmark). 4.2.2 Computational Density functional theory (DFT) DFT calculations of the FeS2 (100) surface for this work were made by Aravind Krishnamoorthy. Details can be found in Refs. [108, 182] kinetic Monte Carlo (kMC) Kinetic Monte Carlo (kMC) simulations were performed on a model of the pyrite (100) surface to understand the processes responsible for the experimentally observed timeevolution of the surface defect structure under non-equilibrium conditions. The kMC was formulated by FWH and Aravind Krishnamoorthy, and coded by Aravind Krishnamoorthy. A full description is available in Ref. [182]. Briefly, we modeled a small set of elementary processes that were able to reproduce the experimentally-observed defect phenomena, including sulfur and iron vacancy formation. and vacancy diffusion diffusion processes. The probability of each process occuring, J, was given by the Arrhenius equation: J = ν exp (−Ea /kB T ), where ν is the attempt frequency and Ea the activation barrier. The two processes described above were subjected to geometrical constraints based on the FeS2 (100) surface. It is worth mentioning briefly that the activation energy for surface V S formation was based on the value of 0.1 eV obtained from the XPS experiments in this work. 96 Table 4.1: XPS core level shift (CLS) for S 2p peak, relative to the bulk pyrite dimer signal B, and full width half maximum (FWHM) of fitted S 2p peaks (B, S, M and HE) used for quantification. CLS (eV) / FWHM of fitted S 2p components Excitation Energy B S M HE 210 eV 0.00 / 0.77 ± 0.03 -0.64 ± 0.03 / 0.65 ± 0.02 -1.23 ± 0.04 / 0.55 ± 0.02 +1.75 ± 0.02 / 1.7 ± 0.30 350 eV 0.00 / 0.63 ± 0.01 -0.65 ± 0.05 / 0.53 ± 0.01 -1.25 ± 0.05 / 0.57 ± 0.02 +1.80 ± 0.10 / 2.2 ± 0.20 500 eV 0.00 / 0.78 ± 0.02 -0.67 ± 0.02 / 0.77 ± 0.02 -1.26 ± 0.02 / 0.60 ± 0.03 +1.80 ± 0.20 / 2.3 ± 0.30 4.3 Results and Discussion We first demonstrate how the surface of FeS2 evolves under the reducing conditions of the ultra-high vacuum environment and increasingly high temperature, through the initial formation of sulfur monovacancies and the migration to the surface of iron vacancies that were supersaturated in the bulk crystal. At a sufficiently high temperature of ∼ 240 o C, vacant cation and anion siteswere observed to coalesce into pits of < 100 nm in lateral dimension and of either exactly one-half or one lattice parameter depth. We employed the kMC model to substantiate our propoesed atomistic mechanism for this phenomenon at realistic time and length scales. 4.3.1 Evolution of pyrite surface structure and chemistry We examined the formation of individual and clustered point defects on the (100) surface of pyrite under successive reduction at increasing temperatures in the UHV environment. The approach to gather element-specific signatures around an x-ray absorption site included analysis of core holes that result from core-level ionization and x-ray absorption. Figure 4-4 shows a series of four S 2p spectra taken on single crystal FeS2 (100) to illustrate the effects of varying the annealing temperature (210 o C and 330 o C, as labeled), as well as the source excitation energy (350 eV in Fig. 4-4a and b; 210 eV in Fig. 4-4c and d. The experimental data were deconvoluted by peak fitting into three doublet components, consistent with the well-established S 2p features of pyrite. [128,145–148] Here, we adopt the nomenclature introduced by Andersson et al. [148] for the features B, S and M, as described below. The spin-orbit splitting 2p3/2 peak was fixed for each component at 1.18 eV above the corresponding 2p1/2 peak, with an intensity ratio of 1:2 for the doublet pairs. Further constraints used in peak fitting are listed explicitly in Table 4.1. Feature B (“Bulk”), distinguished by a 2p3/2 peak centered at 162.8 eV binding energy (BE), is ascribed to the signal from bulk sulfur S22− dimers (see Fig. 4-5b). Feature S (“Surface”) has a core level shift (CLS) of -0.65 ± 0.05 eV BE relative to B and represents surface S22− dimers (Fig.4-5c). Finally, feature M (“Monomer”) is at a CLS of -1.25 ± 0.05 eV BE relative to B and is related to the monomer defect (or monovacancy) species S 2− at the surface (Fig. 4-5d) which is of primary interest for this work. Quantifying the temperature dependence of proportion of the M contribution in the S 2p spectrum enabled us to identify the sulfur vacancy formation enthalpy on the pyrite surface, as described below. An additional singlet peak, which we refer to as the ‘high energy’ (HE) peak was fit to each spectrum at 1.8 ± 0.2 eV above the main B peak, with a 70% Gaussian-30% Lorenztian distribution. Previous reports of the FeS2 (100) S 2p photoemission have either explicitly or implicitly dealt with a similar feature. For example, Nesbitt et al. fit a single peak in their work and attributed it to 97 98 HE S M 165 164 163 162 Binding Energy (eV) B 161 160 HE 167 166 330oC B S 165 164 163 162 Binding Energy (eV) hν = 210 eV (d) 210oC hν = 210 eV (c) M 161 160 Figure 4-4: S 2p photoelectron spectra of FeS2 (100) at excitation energies hν of (a, b) 350 eV and (c, d) 210 eV after annealing in ultra-high vacuum at 210 o C and 330 o C, respectively. The recorded XPS data points are shown as dots and solid lines mark the enveloping curve from peak fitting. Three features, labeled B, S and M are fit to the experimental data. Components S and M grow in intensity from 210-330 o C and are more prominent when probed with the lower 210 eV excitation energy, consistent with the surface localization of these features. 166 330 C o hν = 350 eV (b) 210 C o hν = 350 eV (a) 167 Intensity (arb. units) Monomer defects (M) form under high temperature, reducing environment Intensity (arb. units) Three types of surface sulfur binding environment: B, S, M (a) (100) surf. (b) (c) (d) S M B B 001 Fe 010 S VS Figure 4-5: Atomic model of the FeS2 (100) surface as viewed side-on. (a) with highlighted sulfur atoms corresponding to x-ray photoelectron spectroscopy features denoted as: (b) Bulk ‘B’ sulfur binding environment with 3 Fe-S bonds and 1 S-S bond, (c) surface ‘S’ environment with one fewer Fe-S bond, and (d) monomer ‘M’ with an adjacent sulfur vacancy and hence no S-S bond. polysulfides (Sn2− ) [143], Mattila et al. stipulated the additional signal in this region to arise from the effect of the core hole [146], and Andersson et al. fitted their pyrite S 2p peaks with asymmetric tails towards higher binding energies to account for this contribution. [148] In our work, we did not observe the high-energy tail of the S 2p spectra to change significantly during the course of the annealing experiments from 120-350 o C. As the experiments were conducted under reducing UHV conditions we dismiss the possibility of surface Sn2− contributing to the signal. We therefore postulate that the fitted HE component is generated by the core hole effect in pyrite, and we used systematic fitting parameters to remove the influence of this peak on the quantification of the other components B, S and M. The total intensity (area) of the HE peak was constrained to be 10% of the total S 2p signal and variation of this constraint by ± 5% had no quantitative effect on the relative proportions of the B, S and M features fitted on each spectrum. The surface sensitivity of XPS is related to the inelastic mean free path λ of the emitted photoelectrons. To minimize λ, the incident photon energy is chosen such that the kinetic energy of the excited photoelectron is of the order 40-50 eV [46]. For the S 2p peak at a BE of ∼ 160 eV this corresponds to an x-ray excitation energy hν of 200-210 eV. Hence the spectra in Fig.4-4c and d obtained using hν = 210 eV originate from the top 4.5 ± 1 Å (2-3 sulfur layers) while those in Fig. 4-4a and b at hν = 350 eV are attributable to the top 11 ± 1 Å (4-5 sulfur layers). For the other excitation energy used in this work, hν = 500 eV, the estimated is 14 ± 1 Å (5-6 sulfur layers). At the 500 eV photon energy, the surface sulfur species of interest to this work have relatively weak signal as compared to the bulk species B; therefore we do not include the 500 eV XPS results in further discussion. Features S and M are most prominent in Fig. 4-4c and d which were obtained using 210 eV excitation energy, indicating that these signals arise from the 1-2 atomic layers of pyrite closest to the (100) free surface of the crystal.The proportion of M as a fraction of the total S 2p signal, denoted as [M], increases with increasing annealing temperature up to 330 o C. This trend is consistent with the formation of surface monomer defects as the sample surface is increasingly chemically reduced. In the intermediate temperature region between 120-240 o C, [M] increased consistently with temperature (Fig. 4-6). To estimate the formation enthalpy of sulfur monomer vacancies, [M] in the most surface-sensitive measurement (hν = 210 eV) was quantified as a function of temperature. We assume the following simple defect reaction to form electronically neutral vacancies under UHV annealing: 99 1 × × FeS2 ⇔ Fe× Fe + SSi + VSii + S2,(g) 2 (4.1) where the conventional Kröger-Vink notation [183] is used andthe subscripts Si and Sii refer, in no particular order, to the two adjacent sulfur sites on any given anion dimer. Although the formal oxidation state of the newly formed sulfur monomer site (Sii here) should be S1- , it is understood this configuration relaxes to the more stable S2 by electron transfer from an adjacent cation, leaving the vacant site electronically neutral. [110, 143] We write the Gibbs free energy change of the defect formation reaction in Eq. (4.1) as ∆G f = ∆H f − T ∆S f , where ∆H f and∆S f are the enthalpy and entropy of vacancy formation, respectively. The sulfur vacancy concentration at a given temperature T can then be written: ª § ∆G f |VS | = pS12 . exp − 2 kB T (4.2) where kB is Boltzmann’s constant, pS2 is the equilibrium partial pressure of sulfur gas, and we assume the activities of the solid to be unity. Finally, we note that a given change in [M] can be used as anestimate for change in [V s ], if we apply the reasonable assumption that the only defect species contributing to the increase in [M] is the monomer vacancy, V s . Figure 4-6 shows the results for the increase in relative contribution attributed to M, as ln[M] vs. 1/T, in the annealing temperature range of 120-240 o C. We present only the data obtained with 210 eV XPS excitation energy, which gives the most surface sensitive results. From the slope of the straight-line fit and Eq. (4.2) we infer the formation enthalpy ∆H f to be 0.1 ± 0.03 eV. Our estimated error in ∆H f is the difference between maximum and minimum slope fits obtained from systematic variations in software-generated quantification of S 2p peak fits (including sensitivity to the selection of the HE fitted component, giving rise to the error bars), and does not reflect experimental error associated with the sample, equipment or measurement. The optimal peak fit for each component was first chosen so as to minimize the total root mean squared error between fitted data and experimental data. Error bars were then generated by methodically varying the peak positions of the M and S components relative to the position of B, across the ranges listed in Table 4.1; we believe this gives a reasonable quantitative estimate of fitting error. However, in this quantification we neglect any clustering of defects at lower temperatures that would effectively suppress the value of [M], as discussed in more detail below. This is reasonable given the high resolution achieved in our STM images, which suggest that these surfaces are not likely to have many small clusters at the lower temperatures. As a result, the calculated ∆H f may only slightly underestimate thetrue activation enthalpy for vacancy formation. Upon annealing to higher temperatures (240 o C - 330 o C) we observed a noticeable deviation from the Arrhenius behavior of [M] that is seen in Figure 4-6. Figure 4-7a shows that the percentage of M reached a maximum at around 240 o C and then dropped by ∼ 3-4% of the total S 2p signal, before leveling off to a roughly constant value for temperatures above 270 o C. Furthermore, the total signal arising from the S component increased in this temperature range up to 270 o C, by a roughly equivalent amount (∼10%) to M, then leveled to a consistent value of around 43% of the total S 2p signal between 270 - 330 o C (Fig. 4-7b). Such non-Arrhenius behavior of [M] and [S] imply a more complicated phenomenon than the otherwise logical hypothesis that surface-localized sulfur dimers (S) are converted directly to monomers (M) by the thermally-assisted breaking of the sulfur-sulfur bonds; in such a case we would expect any increase in M signal to be matched by a corresponding drop in S. We will revisit the mechanisms that could explain such a non-monotonic change in sulfur monovacancies via later description of kMC simulations. However, the model underlying such computational simulations is informed by direct observation of the surface reconstruction of 100 Arrhenius growth in [M] at low temperatures T ( oC) 3.6 240 200 160 120 3.4 ln [M] 3.2 3 2.8 f ΔH = 0.1 ± 0.03 eV 2.6 2 2.2 2.4 - 1) x 10 -3 1/T (K 2.6 Figure 4-6: Sulfur monomer vacancy concentration [M] as a percentage of total S 2p photoelectron spectrum signal vs. inverse temperature. The dashed line is the best linear fit to the data, and error bar generation is discussed in the text. ∆H f is calculated from the slope of this line, assuming the Arrhenius relationship in Eq. (4.2). Non-monotonic defect formation at higher T’s 30 (a) %M 25 20 15 210 eV 300 eV 500 eV 10 200 45 250 300 o Annealing T ( C) 350 250 300 o Annealing T ( C) 350 (b) %S 40 35 30 25 200 Figure 4-7: Proportion of the M and S components of the S 2p photoelectron spectra on FeS2 (100) at 200-330 o C, measured using three different excitation energies: 210, 350 and 500 eV. (a) The fraction of the total signal represented by M increased upto 270 o C then dropped and stayed approximately constant, (b) The fraction of total signal represented by S also increased up to 270 o C but then remained unchanged to 330o C . The dashed lines connecting the data points are shown as a guide for the eye. 101 Nucleation and growth of surface vacancy pits by scanning tunneling microscopy 25oC z (nm) +0.25 (a) 220oC z (nm) +0.5 (d) 0.0 0.0 -0.25 -0.5 300oC (g) z (nm) +2.0 EXPERIMENTAL 0.0 30 nm 50 nm +0.25 (b) 50 nm -1.0 -0.5 +0.1 (e) -2.0 +0.4 (h) 0.0 0.0 0.0 -0.1 10 nm 10 nm -0.4 -0.2 10 nm -0.25 0.0 SIMULATED (c) 10 nm -0.8 0.0 (f) 10 nm -0.25 0.0 (i) 10 nm -0.25 -0.25 Figure 4-8: Scanning tunneling microscopy (STM) images of single crystal FeS2 (100) surfaces (a, b) prior to any in situ annealing in ultra high vacuum, (d, e) after three hours of in situ annealing at 220 o C, and (g, h) after three hours at 300 o C. The surface morphology of atomically-flat terraces at room temperature changes underannealing due to the formation of vacancy clusters. In addition, originally straight terrace edges, visible in the bottom left corner of (a), develop into wavy lines, e.g. as indicated by the white arrow in (d). All STM data were collected at room temperature using tunneling conditions in the range ± (1-2) V and 200-1000 nA. The bottom row of images (c, f, i) display the results of kinetic Monte Carlo (kMC) simulations performed at the same temperatures of (c) 25 o C, (f) 220 o C and (i) 300 o C for comparison with the experimental images. The activation barrier values used in these kMC simulations are from Ref. [182]. pyrite, which we address next. Scanning tunneling microscopy (STM) images of single crystal pyrite (100) surfaces both at room temperature and after in situ annealing to 220 o C and to 300 o C for > 150 min (Fig. 4-8) provided further details on the behavior of defects underlying the results in Figure 4-7. For comparison, we include predictions from our temperature-dependent kMC simulations of point defect formation and clustering. These simulated results are discussed in more detail in the following section. The initial surface condition of the FeS2 single crystals consisted of multiple atomically-flat, featureless (100) terraces (Fig. 4-8a and b). After annealing at 220o C in UHV, we observed two notable changes in surface morphology: first, the straight ledges separating atomic terraces became wavy, with small incursions into the terrace (Fig. 4-8d). Second, there appeared multiple small, irregularlyshaped depressions on the surface of flat terraces that arose from the agglomeration of surface anion and cation vacancies (Fig. 4-8e), henceforth referred to as vacancy clusters or ‘pits’. At 220 o C we observed a dispersion of cluster sizes (widths) from smaller than 1 nm to ∼ 10 nm (Fig. 4-8e). Following annealing at 300 o C, these grew to form a more homogeneous spread of clusters with lateral dimensions consistently between 5-10 nm (Fig. 4-8g, h and i). As described in Section 4.2.1,we believe the source of the observed iron vacancies (V Fe ) to be supersaturated V Fe remaining in the bulk after crystal synthesis under high sulfur chemical potential S. During annealing in the low S experimental conditions in UHV, iron vacancies migrate to the surface to equi102 librate a stoichiometric bulk rid from V Fe . This process effectively ‘provides’ cation vacancies to the surface, leading to a situation similar to the dissolution of metal cations into liquid electrolytes in contact with a passive film. [184, 185] Similar surface vacancy clusters have been observed on metals containing supersaturated vacancies from quenching [173,186,187], non-stoichiometric oxides such as CeO2 [188,189] and also natural pyrrhotite Fe1-x S that have been subjected to heating under vacuum. The pits exhibited a curved, non-faceted morphology that can be rationalized by comparison to the curved terrace steps observed by Rosso et al. on conchoidal fracture surfaces of FeS2 . [151] Normal to the free surface of the crystal, the depth of the defect clusters also changed between 220-300o C , as seen in the image height histograms of Figure 4-9. After in situ annealing treatment at 220 o C, there existed a bimodal distribution of pit depths, with the majority of pixels at the nominal surface level (normalized to 0 nm on the scale shown) and a subset located at a depth of approximately 0.25 nm. After in situ annealing at 300 o C, the pit depth distribution broadened and the mean depth of the minor peak in the histogram is shifted to approximately 0.55 nm. The line traces in Figures 4-9a, b and c provide more detail on the depth of pits: at 220 o C the surface pits were consistently 2.7 ± 0.1 Å deep whereas at 300o C the majority of pits have a depth of 5.4 ± 0.1 Å. Given the lattice parameter of pyrite of 5.41 Å [162], these pit depths correspond to one half and one full lattice parameter, respectively. 4.3.2 Mechanism of vacancy formation and coalescence The growth of defect clusters as evidenced by our STM results, along with the XPS results presented in Figures 4-6 and 4-7, lead us to propose a mechanism involving three distinct phenomena occurring in tandem during the in situ reduction of the FeS2 (100) surface, as visualized in Fig. 4-10. 1. Formation of isolated S monovacancies (V S ) and surface-ward migration of V Fe from the bulk, giving rise to increasing signals of sulfur species M and S, respectively; 2. surface diffusion of V S and V Fe , followed by the stochastic clustering of small numbers of vacancies; 3. growth of vacancy clusters, which are more stable geometric features for the reduced surface compared to dispersed point defects, due to a reduced formation energy for sulfur vacancies at step edges. The proposed mechanism was cast into our kMC model. The last row of images in Figure 4-8 shows the successive formation and growth of surface pits after simulated annealing for 4 h. The model reproduces the flat, nearly featureless surface at ambient temperatures (Fig. 4-8c), while a series of pits with average lateral dimension of 1 x 1 nm2 becomes visible at 220o C (Fig. 4-8f) and larger pits on the order of 2 x 2 nm2 at 300o C (Fig. 4-8i). For a complete discussion of the kMC results, the reader is referred to Ref. [182]. Below, we discuss in more detail the three phenomena that account for this pit initiation mechanism: 1. Formation of sulfur and iron vacancies at and near the surface: Elemental sulfur is highly volatile in comparison to iron, with a melting temperature of 115 o C at 1 atm pressure. We therefore assume sulfur sublimes from pyrite and is dynamically removed from the sample surface under vacuum until the pyrite equilibrates with the ambient sulfur chemical potential, dictated by the chamber pressure of 109 -1010 mbar. We also noted the importance of iron vacancies in our model. The corresponding appearance of V Fe near the surface is indicated by the growth of the S feature of the XPS S 2p spectra while the surface is increasingly reduced. 103 Evolution of vacancy pit depth with temperature “M” (a) z (nm) 0.2 200 oC 0 -0.2 -0.4 -0.6 0 (b) z (nm) 0.2 4 16 8 12 x (nm) 001 “S” 010 240 oC 0 2.7 Å step -0.2 -0.4 -0.6 0 5 (c) 0.2 10 15 20 x (nm) 300 oC z (nm) 0 5.4 Å step -0.2 -0.4 -0.6 10 0 20 Fe x (nm) (d) % of pixels o 300 C -0.8 VS VFe 0.25 nm o 220 C -1 S 0.55 nm -0.6 -0.4 -0.2 0 0.2 0.4 Figure 4-9: Pits are one half- or one lattice parameter deep. (a-c) line traces taken across representative surface features observed in scanning tunneling microscopy (STM) scans at 200, 240 and 300 o C. Inset in each graph are STM images of the nanopits, with grey arrows indicating each line trace. The ball-and-stick atomic diagram next to each graph depics a side-on view of the surface at the pit edges. In (a), we observed single monomers or very small nanopits that could not be resolved on the atomic scale. By 240 o C in (b) the measured pit depth is exactly one-half of a pyrite lattice parameter, or 2.7 Å. Finally, in (c) the step height has grown to a full lattice parameter of 5.4 Å. (d) shows height (z-axis) histograms of STM images obtained after annealing at 220o C (and at 300o C bimodal distributions of pit depths are observed at both annealing temperatures; however the average pit gdepth is ∼0.25 nm in the former case and ∼ 0.55 nm in the latter. Figure 4-5 illustrates how six new S binding environments would accompany the introduction of a single V Fe in the second atomic layer from the (100) surface (NB: only 4 S sites are shown in the plane of the graphic). Since Fe loss through evaporation into vacuum is unlikely at the relatively low annealing temperatures compared to the melting point T m of iron (300o C < 0.2 T m ), the increaseof V Fe requires an alternative explanation. It is known that the formation enthalpy of the V Fe in pyrite increases by up to 1.42 eV as the environment changes from sulfur-rich to one deficient in sulfur. [108] This large change implies that the sizable number of V Fe defects originally present in the pyrite crystal during synthesis 104 Atomic processes in pitting mechanism S2 (g) (b) (a) (d) (c) 001 010 Fe2+ VFe Formation S22- VS Diffusion Figure 4-10: Illustration of atomic processes involved in the proposed mechanism of pit formation and growth on pyrite (100). (a) Formation of surface sulfur monovacanciesVSthrough evaporation into vacuum. (b) Diffusion of V S to a pit site. (c) Agglomeration of vacancies on the iron and sulfur sublattices by diffusion, leading to theinitiation and growth of the pit. Presence of an initiated pit (as depicted in this fig-ure) is not a necessary precursor to the process in (c). (d) Iron vacancies, denotedas V Fe , that are already present in the bulk migrate to the surface during annealing.(e) V S formation at under-coordinated sites surrounding pits has lower formation enthalpy as compared to isolated vacancy formation process in (a). kMC result replicates M defect formation 0.4 [M] 0.3 0.2 0.1 0 KMC XPS (210 eV) 200 240 280 320 Temperature (°C) Figure 4-11: kinetic Monte Carlo simulation results. Simulated values of the sulfur monomer vacancy concentration [M] on the pyrite surface as a function of annealing temperature, obtained by kinetic MonteCarlo simulations (kMC) and indicated by the blue zone. The width of predicted [M] indicated by this blue band is given by variation in kMC energy barrier values over the range described in Ref. [182]. Values of [M] experimentally determined from our XPS measurements (Fig. 4-7) are shown for comparison. in sulfur-rich conditions are not stable in the sulfur-deficient conditions encountered during the annealing process. In order to equilibrate the bulk under these experimental conditions, the oversaturated V Fe point defects migrate from bulk towards the free surface at high temperatures to annihilate. In providing iron vacancies to the surface, coincidentally in this work, the situation is analogous to the dissolution of metal cations from the passive film in liquid electrolytes. [184] 2. Diffusion of sulfur and iron vacancies on the surface: the generation of incursions into pre-existing atomic terrace edges (Fig. 4-8d) and the nucleation of 105 small vacancy clusters on top ofterraces (e.g., as seen in Fig.4-8e), requires the diffusion of both V Fe and V S across the surface. When two or more vacancies encounter each other stochastically, a small cluster is formed which is more stable relative to the dispersed individual vacancies. STM imaging by Rosso et al. has recorded surface diffusion of iron vacancies on natural single crystals of FeS2 at room temperature over time scales of minutes, so diffusive processes are expected to occur with low energy barriers. [149] In the kMC model in this work, a higher barrier was taken for diffusion of vacant surface sites away from the pits compared to diffusion of vacancies towards the pits, simulating the trapping of vacant sites by the initiated pits. 3. Growth of pits: Once a stable pit nucleates, the reduced coordination of sulfur atoms at the newly-created step edges of pits reduces the formation energy for vacancies at these sites, and accelerates the growth of pits. Indeed, our DFT calculations showed that ∆H f for individual V S at a step edge of a pit could be up to ∼ 40% lower than that on an atomically-flat surface. This type of dependence of defect formation on the local atomic configuration has also been observed in sulfide inclusions in pitting corrosion. [190] The formation and expansion of surface vacancy clusters in this manner provides an explanation for the surprisingly low ∆H f for sulfur vacancies of around 0.1 eV that we measured using XPS below 240 o C (Fig. 4-6), as compared to recent theoretical predictions in the range of 0.4-1.44 eV. [108,109,134] Upon raising the temperature to greater than 240 o C, the effect of growing the vacancy clusters is to maintain the XPS signal intensity from M at a roughly constant value. This is because the removal of a sulfur atom from the step edge of a vacancy cluster, while growing the cluster, does not result in the creation of an adjacent monomer M site on the surface structure; therefore the M signal intensity does not increase in the S 2p spectrum. 4.4 4.4.1 Outcomes Conclusions We have investigated the evolution of surface chemistry and morphology on synthetic pyrite single crystals as a function of annealing temperature in reducing conditions, in order to visualize and quantify the mechanisms leading to pit initiation on the surface. The formation enthalpy for sulfur vacancies was found to be tobe 0.1 ± 0.03 eV from the exponential temperature dependence of the sulfur monomer vacancy (V S ) binding environment detected in the S 2p photoelectron spectra. However, at higher temperatures above 200 o C, the sulfur vacancy concentration decreased and deviated from Arrhenius behavior, concurrent with the initiation of nanometer-scale surface pits. The depths of these pits were exactly one-half or one FeS2 lattice parameter, as imaged by STM. To explain this behavior, we propose a mechanism involving the simultaneous formation and migration of vacancies at the surface, with facilitated vacancy formation and agglomeration at step edge sites surrounding pits. A simple kinetic Monte Carlo simulation with thermally activated reactions was used to validate the proposed mechanism. We note two important implications of these findings. First, the observed, concerted agglomeration of point defects from both cation and anion sublattices to initiate nanoscale pits has broad consequences for ionic solids in reducing environments. The dynamics of surface point defects observed under controlled reducing environments offer an atomistic level description of the incipient stages of pit formation in passive films, as postulated in models of surface degradation such as the Point Defect Model. Second, the relatively low defect formation energy that we measure for sulfur vacancies confirms the high chemical reducibility of the FeS2 surface, often linked to poor electronic and electrochemical properties in synthetically grown pyrite. 106 4.4.2 Future work Having observed the formation of surface nanopits by reducing FeS2 in UHV at high temperatures, it would be of interest to study in situ pitting nucleation at lower temperatures (< 100 o C) but under electrochemical conditions, for example in model corrosive solutions and/or under an applied electrode bias. The ideal instrument would be an electrochemical STM, to investigate whether the same atomic processes as described in this chapter apply when a high temperature gradient combined with chemical driving force is replaced by an electrochemical one. Separately, it would be interesting to study the reactivity of surface defects and defect clusters towards molecules such as H2 S and H2 O, building on the work by Guevremont et al. [126] and others. [127] Since the publication of the work described in this chapter, Andersson et al. have reported the formation of monomers on the FeS2 (100) surface under energetic ion bombardment, suggesting that the surface chemical activity can be altered greatly by sputtering. [191] However, their assessment of surface reactivity assumes that many monomer defects form individually, dispersed across the surface. As we have shown here, vacancies at elevated temperatures can cluster and reduce the number of dangling bonds available to catalyze heterogeneous reactions. A concerted experimental and computational approach could help understand the reactivity of real pyrite surfaces in aggressive, electrochemical environments. Experimentally, temperature programmed desorption (TPD) would be a good tool to study the adsorption and reactive properties of surface defects on pyrite. Acknowledgements We gratefully acknowledge support provided by BP Plc. through the BP-MIT Center for Materials and Corrosion Research. We thank S. Yip (MIT), R. Woollam, Steven Shademan and Sai P. Venkateswaran (BP) for discussions on pitting mechanisms in H2 S, P. Lazic, R. Armiento and G. Ceder (MIT) for discussions on relevance of these results to PV performance of pyrite, Klas Andersson (formerly of Stockholm University, Sweden) for discussions on pyrite surface chemistry and XPS, and R. Sun (MIT) and M. Kabir (IISER-Pune) for verification of some of the point defect formation energycalculations. We thank D. Mullins and P. Albrecht at Oak Ridge National Laboratory for the use of the U12A beamline (BrookhavenNational Laboratory) for XPS measurements. The U12a beamline is supported by the Division of Chemical Sciences, Geosciences,and Biosciences, Office of Basic Energy Sciences, U.S. Departmentof Energy, under contract DE-AC05-00OR22725 with Oak Ridge National Laboratory, managed and operated by UT-Battelle, LLC. Use of the National Synchrotron Light Source, Brookhaven National Laboratory, was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-98CH10886. We thank the National Science Foundation for providing the computational resources for this project through the Texas Advanced Computing Center under Grant No.TG-DMR120025. 107 108 Chapter 5 Conclusions 5.1 Summary of activation barriers This thesis explored the bulk and surface defect chemistry of the two stable Fe-S phases pyrrhotite and pyrite, to understand the governing mechanisms behind their protectiveness as passive layers formed on steels under sulfidizing conditions. A key goal was to measure activation barriers Ea , the fundamental descriptor of kinetic rates, for the processes depicted in Figure 1-7a. To that end, a summary of the activation barriers of interest is given in Table 5.1. Table 5.1: Summary of experimentally determined activation barriers Ea in eV for key unit processes investigated in the course of this work. Bold entries are contributed from the experiments described in this thesis; those marked "N/A" were either not measured, or could not be found in the literature. Process Temp. (o C) Pyrrhotite Fe1-x S Pyrite FeS2 Chapter 2: Growth Bulk Fe diffusion * DFe 0.83 αS(T )2 170-700 (a) + N/A Bulk S diffusion * DS > 500 7.9 (b) 2.1 (c) Surface S exchange (ox.) 350-600 N/A Surface S exchange (red.) 350-600 1.05 ± 0.20 0.79 ± 0.23 N/A Chapter 3: Reactivity Bulk band gap Eg, bulk 25 0.30-0.80 (d) 0.95 Bulk band gap Eg, surf 25 N/A 0.40 ± 0.10 Chapter 4: Stability 0.10-0.20 (e) Surface Fe diffusion *DFe, surf 25 Fe surface vacancy ∆H v,Fe 100-300 N/A N/A S surface vacancy ∆H v,S 100-300 N/A 0.10 ± 0.03 (a) α = 0.41 ± 0.06 and S(T ) is the reduced magnetization (= 0 at 0 K), (b) Ref. [40], (c) Ref. [192], (d) Refs. [52, 193], (e) Ref. [149] 109 5.2 Contributions With reference to Table 5.1 and to supplement the "Outcomes" from each of Chapters 2, 3 and 4, the key scientific questions and corresponding contributions of this thesis are summarized below: What are the rate limiting mechanisms in the growth of Fe1-x S on iron? Pyrrhotite scale growth occurs by Fe diffusion from the metal to the surface of the passive layer, where Fe combines with S transferred from molecular form such as in H2 S in the environement. Both Fe diffusion and heterogeneous surface exchange of S have a similar activation energy of ∼1 eV. Hence, regardless of temperature (or electrochemical potential), the sulfidation of iron should always commence with linear kinetics, controlled by the surface reaction. However, since the diffusive transport depends on the square of film thickness, eventually diffusion will become the limiting process in scale growth, marked by a transition to parabolic time dependence. From our studies on phase-pure pyrrhotite thin film and bulk samples, this transition should occur when the film has grown to approximately 100-1000 µm thickness. Does the order-disorder transition in Fe1-x S have an effect of iron diffusivity? Iron self diffusivity * DFe in pyrrhotite had not been systematically studied below the known order-disorder transition at T N = 315 o C. Above T N , iron atoms are paramagnetic and cation vacancies are consequently randomly arranged in the lattice. * DFe in the paramagnetic regime follows an Arrhenius trend with an average activation energy Ep = 0.83 eV. However, below T N iron atoms spontaneously magnetize, imposing a driving force for Fe vacancy ordering that results in a series of complex ordered superstructures at low temperature. The effect of spontaneous magnetization is to increase the Fe vacancy migration barrier. Hence the activation enregy Em for self-diffusion below T N has a magnetic dependence of the form: Em = E p + αS(T )2 , where S is the reduced magnetization of pyrrhotite and α is a constant ∼0.4. Therfore, extrapolation of the paramagnetic trend to lower temperatures would overestimate real diffusivitites in this material, by up to 100 times at 150 o C. How can we quantitatively assess surface electronic structure via scanning tunneling spectroscopy (STS)? The probability of charge transfer between a solid and adsorbate is ultimately dictated by the availability of electronic states at the correct energy levels. FeS2 is a good example of a semiconducting material in which unsaturated "dangling" bonds at the surface introduce additional states to those present in the bulk. A seemingly ideal tool to confirm the existence of surface states experimentally is STM; however, the inherent electronic properties of the surface are easily distorted by the very high proximate electric field of the tip, frequently leading to misinterpretation of values such as surface band gap Eg . This thesis contributes a systematic methodology for quantifying the surface electronic strucutre by matching simulated tunneling current predictions to experimental results from the STM. This approach is extendible to other semiconducting materials with similar characteristics, such as other transition metal chalcogenides. What role do surface states play in charge transfer on FeS2 (100)? Using the technique outlined above, the (100) surface of FeS2 is confirmed to contain intrinsic surface states (SS’s) arising from Fe and S dangling bonds. These sit on the edges of the valence and conduction band, respectively, and have the effect of reducing the surface band gap from 0.95 to 0.4 eV. Since the surface states form continuous bands with the bulk states, charge cannot localize at intrinsic SS’s and they do not pin the Fermi level EF . However, extrinsic surface states from surface point defects and adsorbates sit discretely within 110 the surface Eg and do pin EF . The implications for electrochemical reactions such as oxidation during corrosion are that horizontal charge transfer can occur at smaller overpotentials η via surface states than would be predicted if the bulk electronic structure were evaluated alone. In considering a generalized passive film model, this finding highlights the need to differentiate carefully between bulk and surface electronic structure. What are the atomic-scale dynamics of point defects on destabilized ionic passive film surfaces? According to the Point Defect Model of Macdonald et al. (see Chapter 1), pitting is a deterministic process in which the breakdown of passivity is controlled on the microscopic scale by the interactions of point defects at the metal-passive layer and passive layer-electrolyte interfaces. The formation of cavities at these interfaces requires vacancies on both the anion and cation sublattice to coalesce, via vacancy formation and diffusion processes. We show on a model FeS2 (100) surface that under reducing conditions at elevated temperatures, vacancies can condense into pit-like features which effectively comprise single atomic steps into the surface. While not necessarily potential pit initiation sites per se, their characterization allow us to formulate a real-time and predictive model of the dynamics and interactions of point defects at the surface as it destabilizes under reducing conditions. Can surface vacancies account for reported polycrystalline FeS2 off-stoichiometry? The formation of vacancies in bulk pyrite is calculated to be energetically costly; however, the surface of pyrite is found to readily forgo sulfur, and this has been proposed as one reason for the observed off-stoichiometry in polycrystalline FeS2 prepared by high temperature methods such as sulfurization. The distinct binding energies of surface S atoms that are either bound to another S or adjacent to a vacancy, thereby allowing us to quantify the formation of vacancies with increasing temperature. A very low formation enthalpy ∆H f = 0.1 eV for sulfur vacancies implies that the surface is indeed very susceptible to losing sulfur to the atmosphere at even mild temperatures > 100 o C. Such facile vacancy formation suggests that the pyrite surface in situ is constantly in flux vacancies easily form but are also reactive sites for further sulfidation. More generally, these findings are critical to the fabrication of synthetic pyrite for photovoltaic or other applications where homogeneity and/or full stoichiometry are important. 5.3 Outlook and perspectives This thesis began with the hypothesis that a multiscale model for passive layers can be constructed via a "bottom-up" approach, that is, by studying the elementary physical chemistry at the surface and in the bulk of the ionic materials that comprise the passive film, and determining the key unit processes governing film growth, reactivity and stability. None of the experimental contributions outlined above were realized under authentic sour corrosion conditions or even laboratory simulated ones. Instead, the majority of the work was carried out under either ultra high vacuum (UHV), i.e. where sensitive ion- and electron- detection equipment can operate, or high-temperature, gas environments. Therefore, we must ask to what extent these fundamental studies advance the overarching goal of constructing the universal passive film model introduced in Chapter 1. The problem of the vacuum gap, which asks how practical it is to extend discoveries made under highly-controlled, low pressure conditions to the "real-world" environment, is not unique to the investigations described in this thesis. However, it is still greatly important to understand these microscopic events on surfaces free of contaminants in order to gain insight into behavior on the fundamental level. To accomplish a fully non-empirical, predictive tool for electrochemical systems we must start by studying ionic point defects on clean surfaces and in pure materials before adding 111 further complexity in the form of microstructure, liquid interfaces, applied stresses and so on. Nevertheless, the further advancement of these ideas necessitates bridging this gap to a more realistic environment, for example with in situ aqueous studies on the iron sulfide phases of interest. The structure and bonding of water at surfaces is central to all electrochemical reactions in water-based electrolytes. Therefore some of the further studies suggested in the individual experimental Chapters 2-4 involve the introduction of water: is diffusion and therefore film growth more rapid in porous iron sulfide films formed in solution? In what way does an electrical double layer affect the inherent reactivity of the surface? Can this be modeled accurately from first principles by density functional theory? Would a pyrite surface imaged by electrochemical STM in a corrosive solution, and under an applied bias, undergo surface destabilization that can still be described by the basic vacancy dynamics described here? These constitute a small set of the questions that must be answered if we are to achieve a comprehensive passive film model that has real-world applicability. Let us end with a perspective on the two practical drivers of this project: the simulation of microscopic events as a predictive tool, and as a platform to design more robust materials for service in aggressive conditions. As outlined in Chapter 1, the study of passive films on metals has an almost 300-year old history, and it would be a haughty overstatement to suggest this work is anything but a minor contribution to our overall understanding of the field. However, we have outlined a basic methodology by which the study of local phenomena at the basic level of materials - atoms and electrons can be fashioned into a larger-scale, descriptive model that can serve both these end goals. Through novel insights into the rate-limiting mechanisms of surface reactions and diffusion, and continued advances in computing power, we can forsee the ability one day to predict the integrity of the materials in our vast network of energy infrastructure down to the sub-nanometer scale. While humbled by this lofty goal, it is exciting to consider what is possible through our knowledge of these materials on the fundamental level. 112 Appendix A Pourbaix diagrams for the Fe-H2S-H2O system The calculations in this Appendix give rise to the two Pourbaix diagrams shown in Figure 1-3 in Chapter 1. The standard Gibbs free energies (∆G of ) and enthalpies (∆H of ) of the species involved in constructing the thermodynamic stability diagrams are listed in Table A.1, referring to reference conditions of 25 o C and 1 atm pressure. H2 S dissociation The H2 S activity input to the thermodynamic stability diagrams must account for the dissolution of H2 S gas in the local aqueous environment: H2 S(g) KH2 S/H2 S ⇔ H2 S(aq) (A.1) with a corresponding solubility constant: KH2 S(aq) /H2 S(g) = aH2 S(aq) γH2 S.pH (A.2) 2S where: • γH2 S is the fugacity coefficient of H2 S which can vary between 0.4-1.0. [20] • pH2 S is the partial pressure of H2 S in the system. Aqueous H2 S can proceed to dissociate in two partial steps (Equations A1.3 and A1.4 followed by Equations A1.5 and A1.6): Kd,1 H2 S(aq) ⇔ H + + HS − Kd,1 = aH + aHS − aH2 S(aq) Kd,2 HS − ⇔ H + + S 2− Kd,2 = aH + aS 2− aHS − (A.3) (A.4) (A.5) (A.6) In the diagrams here, the value of γH2 S = 1 is used. The H2 S gas-aqueous solubility constant expression determined experimentally by Suleimenov et al. [197], Equation 113 Table A.1: Thermodynamic data for species in H2 S-H2 O-Fe system. Species H + (aq) G ∆G of (kJ/mol) ∆H of (kJ/mol) 0.00 0.00 [194] H2 S(g) -33.3 H2 S(aq) -27.9 H2 O(l) -237.1 H2(aq) 17.7 O2(aq) 16.5 Fe (s) 0 2+ -20.6 [194] -39.7 [194] -241.80 [194] [194] [194] [194] (aq) -91.5 Fe3+ (aq) -17.2 Fe2 O3 (s) -743.5 Fe3 O4 (s) -1017.4 Fe(OH)2 (s) -492.0 Fe(OH)3 (s) -705.5 FeSm (s) Mackinawite -93.3 -92.0 * FeSp (s) Pyrrhotite -114.5 -101.7 FeS2 (s) Pyrite -160.1 Fe FeHS Reference -92.3 [194] [194] [194] [194] [194] [194] [195] [194] -171.5 [194] + -104.3 -152.7 * [196] Table A.2: Input parameters. Temperature (o C) 25 ptotal (atm) 1 pH2 S (atm) 0.01 -3 a Fe2+ (mol.dm ) 1.8x10-4 (10 ppm) a Fe3+ (mol.dm-3 ) 1.0x10-6 pH2 (atm) 1.0 pO2 (atm) 1.0 114 Table A.3: Fe-H2 O Reactions and reversible potentials. No. Reaction Erev or pH + O O2 + 4H + 4e ⇔ 2H2 O E r ev(O) = E roev(O) − 2.3RT 4F log (a H 2H + + 2e− ⇔ H2 E r ev(H) = E roev(H) − 2.3RT 2F log (a − 1 4 H+ ) (pH2 ) 2 H+ ) − 2.3RT 2F log (a Fe2+ ) − 2.3RT 2F log (a Fe3+ ) Fe(OH)2 + 2H + + 2e− ⇔ Fe + 2H2 O E r ev(3) = E roev(3) − 2.3RT 2F log (a 2 H+ ) 4 Fe3 O4 + 2H2 O + 2H + + 2e− ⇔ 3Fe(OH)2 E r ev(4) = E roev(4) − 2.3RT 2F log (a 2 H+ ) 5 6Fe2 O3 + 4H + + 4e− ⇔ 4Fe3 O4 + 2H2 O E r ev(5) = E roev(5) − 2.3RT 4F log (a 4 H+ ) 6 Fe3 O4 + 8H + + 2e− ⇔ 3Fe2+ + 4H 2 O E r ev(6) = E roev(6) − 2.3RT 2F log E r ev(6) = 2.3RT 4F Fe 2+ + 2e ⇔ Fe E r ev(1) = E roev(1) 2 Fe 3+ + 2e ⇔ Fe E r ev(2) = E roev(2) 3 1 7 8 9 − − + 2Fe2 O3 + 12H + 4e ⇔ 4Fe 2Fe Fe 3+ 2+ − 2+ + 3H2 O ⇔ Fe2 O3 + 6H + 6H2 O + + 2H2 O ⇔ Fe(OH)2 + 2H pH = + E roev(6) − 1 1 1 1 1 (a Fe2+ )3 (aH + )8 (a Fe2+ )4 log (a 12 H+ ) − 21 .log(K8 .a2Fe3+ ) pH = − 21 .log(K8 .a Fe2+ ) (A.7). yields the aqueous H2 S activity for a given partial pressure input for the reference temperature T = 298 K. l o g(KH2 S(aq) /H2 S(g) ) = −634.27 + 0.2709T − 0.11132 × 10−3 T 2 − 16719 − 261.9log T T (A.7) Fe-H2 O equilibrium reactions The reactions demarcating the equilibrium between different species in the Fe-H2 OH2 O system are taken from Ref. [14]. Table A.3 lists the expected water-iron reactions along with the reversible oxygen and hydrogen evolution reactions. The column on the far right explicitly shows the reversible electrode potential, or constant pH line, for each reaction. The standard reversible electrode potential E roev(n) for a given reaction n is calculated here by evaluation of the Gibbs free energy of the reaction ∆G r(n) under standard conditions of 25 o C, 1 atm pressure and unit activities, in turn estimated by the sum of the Gibbs free energies of formation of product species minus reactant species: E roev(n) = − ∆G r(n) zF =− i 1 X pr od. ∆G f − ∆G rf eac t. zF m=0 (A.8) Mackinawite, FeSm Mackinawite has been proposed to form on bare iron or steel via the sequential chemisorption of SH- ions and the following anodic discharge reactions. [20] − Fe(s) + H2 S + H2 O ⇔ FeSH ads + H3 O+ − + FeSH ads ⇔ FeSH ads + 2e− The species FeSH+ can be incorporated directly into the growing layer of mackinawite via: 115 Table A.4: Mackinawite-Fe-H2 O system equilibrium reactions. No. Reaction 10 Erev , pH or K + FeSm + 2H + 2e ⇔ Fe + H2 S(aq) − + 2.3RT 2F log (a 2 H+ ) log (a 2 2 H2 S ) (aH + ) log (a 2 3 H2 S ) (aH + ) 11 Fe2 O3 + 2H2 S(aq) + 2H + 2e ⇔ 2FeSm + 3H2 O E r ev(11) = − 2.3RT 2F 12 Fe3 O4 + 3H2 S(aq) + 2H + + 2e− ⇔ 3FeSm + 4H2 O E r ev(12) = E roev(12) − 2.3RT 2F 13 FeSm + 2H + ⇔ Fe2+ + H2 S(aq) pH = − 12 log 14 Fe(OH)2 + H2 S(aq) ⇔ FeSm + 2H2 O K14 = − aH2 S E r ev(10) = E roev(10) − E roev(11) a Fe2+ aH2 S 1 1 K13 1 aH2 S Table A.5: Pyrrhotite-Fe-H2 O system equilibrium reactions. No. Reaction Erev or pH + aH S 20 − FeS p + 2H + 2e ⇔ Fe + H2 S(aq) E r ev(20) = E roev(20) − 2.3RT 2F log (a+2)2 21 Fe2 O3 + 2H2 S(aq) + 2H + + 2e− ⇔ 2FeS p + 3H2 O E r ev(21) = E roev(21) − 2.3RT 2F log (a+ )2 (a1 Fe3 S4 + 2H + + 2e− ⇔ 3FeS p + H2 S(aq) E r ev(22) = E roev(22) − 2.3RT 2F log (a+2)2 22 23 FeS p + 2H + ⇔ Fe2+ + H2 S(aq) pH = − 12 .log a Fe2+ aH2 S H H 2 H2 S ) aH S H K23 Assume mackinawite → pyrrhotite in solid state reaction + − FeSH ads ⇔ FeS1−x(aq) + xSH(aq) + (1 − x)H + However, for the purposes of a simplified stability diagram, we can assume a series of overall equilibrium mackinawite formation reactions such as a direct, solid state reaction (i.e. electrochemical reaction, No. 10 in Table A.4) or solution-phase precipitation (i.e pure chemical reaction, No. 13 in A.4). Other possible formation reactions of mackinawite are also listed in Table A.4. Pyrrhotite, FeSp It is assumed that pyrrhotite forms by direct, solid-state transformation from mackinawite. Since FeSp is the more thermodynamically stable phase, this reaction will occur spontaneously under all conditions in which mackinawite initially forms and greigite/pyrite are unstable. Therfore addition of FeSp to the consideration of a Fe-H2 S-H2 O system Pourbaix diagram will necessarily displace macknawite. In reality, sluggish kinetics of the solid state transformation can "stabilise" mackinawite to long times observed in experiment and in the field, especially at low temperatures and H2 S partial pressures, illustrating the major shortcoming of overreliance on thermodynamic stability diagrams for corrosion product prediction. The other equilibrium reactions involving pyrrhotite are listed in Table A.5. Pyrite, FeS2 The reactions involving pyrite used for construction of the final stability diagram (incorporating all the considered Fe-S phases) are listed in Table A.6. 116 Table A.6: Pyrite-Fe-H2 O system equilibrium reactions. No. Reaction Erev + (aH2 S )2 (a Fe2+ ) 24 FeS2 + 4H 2H2 S(aq) + E r ev(24) = E roev(24) − 2.3RT 2F log 25 FeS2 +4H + +4e− ⇔ Fe+2H2 S(aq) E r ev(25) = E roev(25) − 2.3RT 4F log E r ev(26) = − 2.3RT 2F log E r ev(27) = E roev(27) − 2.3RT 2F log (a+2)2 aH2 S + 2e − ⇔ Fe + 2+ 26 2FeS2 + 3H2 O + 2H + 2e Fe2 O3 + 4H2 S(aq) 27 FeS2 + 2H + + 2e− ⇔ FeSm + H2 S(aq) 28 29 30 − ⇔ E roev(26) + )4 (aH (aH2 S )2 + )4 (aH (aH2 S )4 + )2 (aH aH S H FeS2 + 2H + + 2e− ⇔ FeS p + H2 S(aq) E r ev(28) = E roev(28) − 2.3RT 2F log (a+ )2 FeS2 +4H + +e− ⇔ Fe3+ +2H2 S(aq) E r ev(29) = E roev(29) − 2.3RT F log E r ev(30) = 2.3RT 4F + 3FeS2 + 4H + 4e 2H2 S(aq) − ⇔ Fe3 S4 + 117 E roev(30) − H log (aH2 S )2 (a Fe3+ ) + )4 (aH (aH2 S )2 + )4 (aH 118 Appendix B Chemical Vapor Deposition of Fe-S B.1 Motivation A chemical vapor deposition (CVD) system was set up to fabricate thin films of pyrrhotite (Fe1-x S), with the primary objective of using them for tracer diffusion studies. To this end, we aimed to make high-purity films supported on non-ferrous substrates that would simulate a thin pyrrhotite corrosion scale on the order of several hundreds of nanometers thick. Besides pyrrhotite phase purity, it was desirable to investigate a film growth technique that could allow control over stoichiometry and that produced samples with low surface roughness. A set of samples fabricated via CVD were used in initial diffusion studies for this thesis, the results of which are presented in Appendix 3. The work described here constitutes an initial, empirical investigation into the effects of deposition parameters (substrate temperature, precursors, flow conditions) on film chemistry and morphology. In this appendix, we describe the setup of a homemade CVD system and the fabrication of different iron sulfide films through the use of combinations of various substrates and organic Fe and S precursors, and as a function of substrate temperature. Finally, we introduce the "template stripping" technique that was used to make Fe-S samples with atomically-flat smoothness. Several potential applications of combined CVD/template stripping approach facilitated by this work include: • deposition of ultrathin Fe-S (or other similar sulfides e.g. Ni-S, Co-S, etc.); • atomically-flat, polycrystalline samples for STM studies; • surface patterning: surface plasmonics with chalcogenides. B.2 Methods: CVD setup and apparatus A schematic of the CVD apparatus is shown in Figure B-1a, along with a photograph of the equipment in Figure B-1b. The entire system was constructed in a fume hood to avoid any exposure to toxic precursors. The flow of inert gas (Ar or N2 -5% H2 ) through the system was controlled electronically by Omega FMA-series mass flow controllers (MFCs). The various iron and sulfur precursors used in this work are described in Figure B-2: iron (III) acetylacetonate (Fe(acac)3 ), iron pentacarbonyl (Fe(CO)5 ), di-tert butyl disulfide (TBDS), tert butyl methyl-sulfide (TBMS) and hydrogen sulfide (H2 S). Liquid precursors were kept in glass vials. Stainless steel tubes, connected to the gas lines via 14 -inch Swagelok fittings, passed through a rubber bung in the vial to carry the precursor vapor into the furnace. The inlet gas tube was not submerged beneath the 119 liquid; due to the relatively high vapor pressures it was sufficient to simply have the gas passing through the vial containing the liquid precursors. When the solid iron precursor Fe(acac)3 was used, approximately 20 g of powder was placed in a stainless steel tube, which was wrapped in heating coils powered by an automatic temperature controller to achieve the desired setpoint. The substrate was placed on an 10 o inclined Al2 O3 holder in a quartz tube placed inside a Thermo Scientific Lindberg Blue M Mini-mite tube furnace. Homemade stainess steel compression fittings served as gas inlet/outlet seals to the quartz tube. The temperature variation from the front to back ends of the tube inside the furnace was up to 200 o C. The actual temperatures measured uisng a thermocouple at various positions along the furnace are shown in Figure B-1c, for furnace setpoint temperatures of 300-600 o C. The deposition temperatures reported henceforth refer to this calibration. Finally, the exhaust was scrubbed by bubbling at the surface of a bleach solution before being sent up the stack of the fume hood. Various substrates were used, including polished, 5 x 5 mm2 SiO2 (y-cut) and MgO(100) (both from MTI Corp., Richmond CA), cleaved Muscovite Mica sheets and NaCl crystals (both from SPI supplies, West Chester PA) and cut, soda-lime glass slides (VWR Int., Radnor PA). In a typical deposition run, the substrate was placed in the furnace in the desired position and heated under flowing inert gas. Five minutes prior to the start of the deposition, the vessels containing the Fe and S precursors were flushed with inert gas at the desired flow rate to remove any residual oxygen. The liquid precursor flow during flushing was diverted directly to the exhaust. To begin deposition, a three way valve was switched to redirect the liquid precursor into the furnace. B.3 Results Table B.1 lists a range of conditions used in several prior CVD syntheses of iron sulfides by other authors. Generally, the target composition was FeS2 for applications in photovoltaic adsorbers. In the following, we describe the phase composition, purity and morphology of sample deposited using different precursors and under the range of conditions outlined in Table B.2. First, a combination of Fe(acac)3 and TBDS was used, producing iron disulfide (FeS2 ) films with substantial carbon contamination. To mitigate this, we switched to Fe(CO)5 as an iron source. Finally, to produce monosulfide (Fe1-x S) films, we switched TBDS for TBMS and finally H2 S, both of which contain a single sulfur atom as opposed to a S-S dimer. Fe(acac)3 and TBDS: mostly FeS2 The initial setup we used was similar to that described by Berry et al., using Fe(acac)3 and TBDS sources. [198] Figure B-3 shows the phase identification and SEM micrographs of as-deposited films using this setup, as evidenced by Raman spectroscopy (Fig. B-3a) and x-ray diffraction (Fig. B-3b)a. Pyrite has Raman peaks at 337 and 370 cm-1 while marcasite is characterized by resonances at 321 and 384 cm-1 . Pyrrhotite (Fe1-x S) has no Raman resonance. At a substrate temperature of 280 o C, the predominant phases present were pyrite and marcasite (both FeS2 . Increasing the substrate temperature to 400 o C increased the volume fraction of marcasite relative to pyrite. At lower temperatures, the deposition of the disulfide phases is facilitated by the pre-existence of a sulfur-sulfur bond in the TBDS. To obtain pure pyrite films, the as-deposited samples were post annealed in sealed and evacuated quartz tubes containing a small amount of sulfur powder, removing any trace of metastable marcasite phase. Finally, at higher temperatures beyond 400 o C, the stability limit for FeS2 is reached and increasing amounts of pyrrhotite were deposited. At 600 o C, pyrrhotite was the only iron sulfide phase observed. 120 (a) Schematic of original CVD apparatus MFC’s Fe(acac) 150oC O-ring seals Furnace 300-600oC MFC 1 MFC 2 MFC 3 Vent Ar Mesh to mix Substrate holder Exhaust gases Scrubber inclined 10o d-TBDS 50oC (b) CVD apparatus without heating coils for precursors MFC’s Precursors Furnace Exhaust (c) Temperature variation across furnace Measured temperature (oC) 700 Tset = 600 oC 600 Tset = 500 oC 500 Tset = 400 oC 400 Tset = 300 oC 300 200 100 0 2 4 6 8 10 12 14 Sample position (inches from gas inlet) Figure B-1: Home-made Chemical Vapor Deposition (CVD) system: (a) schematic drawing of setup for solid iron (III) acetylacetonate (Fe(acac)3 ) and liquid di-tert-butyl disulfide (d-TBDS) sources. (b) photograph of setup in fume hood, indicating the position of the components drawn in the schematic. (c) the furnace temperature was not uniform; the graphs shows the variation as a function of position from the gas inlet of the furnace tube, for furnace set temperatures of 300, 400, 500 and 600 o C. Carbon contamination from Fe(acac)3 However, the use of Fe(acac)3 and TBDS as precursors led to the contamination of the samples with carbon. Figure B-4 indicates the amount of carbon measured in the films as a function of substrate temperature, as measured by electron dispersive spectrometry (EDS) in a JEOL Supra 55 SEM. The amount of carbon deposited increases approximately linearly with increasing temperature up to ∼ 30 at% at a substrate temperature of 550 o C. We believe the source of the carbon to be Fe(acac)3 , since it is a large, complex molecule containing fifteen carbon atoms and only six oxygens; hence the removal 121 IRON PRECURSORS SULFUR PRECURSORS (a) Iron(III) acetylacetonate (c) (Di)tert-butyl disulfide [TBDS] [Fe(acac)3] H 3C H 3C CH3 S S H 3C CH3 CH3 Straw-colored liquid. Harmful to aquatic life. (d) Tert-butyl methylsulfide [TBMS] Red, air stable solid. Harmful if swallowed. H 3C H 3C S CH3 H 3C (b) Iron pentacarbonyl [Fe(CO)5] Straw-colored liquid. Flammable, low toxicity. (e) Hydrogen sulfide [H2S] 1 3 S 1 H 4 H 4 Straw-colored liquid. Pungent gas. Very toxic, highly flammable. Highly toxic, highly flammable. 0 Figure B-2: Description and safety information for Fe and S precursors: (a) and (b) iron precursors. Fe(CO)5 is toxic and requires special handling precautions; however, it is easier to control vapor flow in CVD. (c-e) sulfur precursors. TBDS is useful for making FeS2 because it already contains a sulfur-sulfur dimer bond. TBMS dissociates at elevated temperatures to leave an HS- reactive radical. H2 S is the cleanest S source of all (no carbon). Table B.1: CVD of Fe-S phases by other authors. Fe source S source Substrate T (o C) P (Torr) Phase Ref. Fe(acac)3 TBDS Glass, Si 300 + A a) 760 Py. [198] Fe(CO)5 TBDS Various 580 38 Py. Fe(CO)5 TBDS FeS2 475 Fe(CO)5 TBDS Glass 475 Cp2 Fe b) Fe(CO)5 C3 H6 Pyrex 410 H2 S/S Glass 140 c) [199] d) [200] 38 Py. 38 Py., Po. [201] 38/760 Po. [202] 760 Py. [203] Py = pyrite; Po = pyrrhotite; a) A = post anneal in S2 at 500 o C; b) Cp = η-C2 H5 ; c) cold-walled reactor; d) epitaxial pyrite. 122 123 Temp. (o C) 160 0 0 Type Fe(acac)3 Fe(CO)5 Fe(CO)5 Iron source 5-30 30 300 Flow (sccm) H2 S TBMS TBDS Type 25 25 50-60 Temp. (o C) Sulfur source 350 50-180 200 Flow (sccm) N2 - 5% H2 Ar (400) Ar (500) Carrier gas (sccm) 375-550 400-500 300-600 Substrate T (o C) Fe1-x S Fe1-x S, Fe FeS2 , Fe1-x S Phases Table B.2: Chemical Vapor Deposition conditions for Fe-S phases: literature. 20-100 20-100 100-500 Deposition rate (nm/hr) Best quality and purity Fe deposits < 450 o C C contamination > 300 o C Remarks (a) Fe(acac)3 and TBDS precursors: Raman spectroscopy 321 Intensity (arb. units) 384 Pyrrhotite o 600 C 1 μm 337 370 432 Marcasite + Pyrite 400oC Pyrite + Marcasite o 280 C Pyrite o Annealed S2 500 C 300 350 400 450 -1 Raman Shift (cm ) 250 (b) X-ray diffracion of as-deposited and S2-annealed samples Intensity (arb. units) * * † * * † * * † * Pyrite † Marcasite * * * As-grown CVD o 280 C o Annealed in S2 500 C 20 30 40 50 60 70 Cu-kα 2Θ Figure B-3: Iron sulfide films deposited from Fe(acac)3 and TBDS: (a) Raman spectroscopy and corresponding scanning electron microscope (SEM) images for samples deposited at 600, 400, 280 o C and a pyrite film post-annealed in sulfur vapor at 500 o C. (b) Cu-kα x-ray diffraction patterns for as-grown marcasite/pyrite and post-annealed pure pyrite films. of carbon through reaction to carbon monoxide or dioxide is not 100% efficient. Pyrrhotite (Fe1-x S) films with monosulfide precursors With Fe(CO)5 serving as the iron precursor, the sulfur precursor was changed to TBMS (Fig. B-2d) to avoid the pre-existing S-S bond such as in TBDS, which would encourage the formation of the disulfide phases. TBMS is a volatile but low-toxic liquid that could be kept at room temperature during deposition. Fe(CO)5 , on the other hand is toxic and pyrophoric liquid that requires special handling (see Appendix 4 for MSDS). Fe(CO)5 is very volatile and dissociates readily at temperatures as low as 150 o C, and therefore had to be maintained in an ice bucket at 0 o C to avoid excessive metallic iron deposition. Nevertheless, as evidenced by the XRD results in Figure B-5, some pure iron phase was deposited in films at temperatures below 450 o C. Finally, to increase the sulfide partial pressure and ensure full reaction of the precursors, the TBMS was exchanged for N2 - 4% H2 S gas. All films deposited in the range 300-500 o C with Fe(CO)5 / H2 S 124 Carbon contaminaton from Fe(acac)3 Carbon content (at%) 40 30 20 10 0 400 500 600 200 300 Substrate temperature (oC) Figure B-4: Carbon contamination in Fe-S films from Fe(acac)3 : as measured by electron dispersive spectroscopy (EDS), as a function of substrate temperature during deposition. Fe(CO)5 and TBMS precursors: pyrrhotite/iron films Intensity (arb. units) * *† * * * Hex. Fe1-xS Metallic Fe ** * ** * * * * * † * † ** 30 * 40 200 nm 450oC † † 400oC * † † 300oC 50 60 70 Cu-k α 2Θ 80 90 Figure B-5: Iron sulfide films deposited from Fe(CO)5 and TBMS: Cu-kα x-ray diffraction patterns for films deposited at 300, 400 and 450 o C with decreasing metallic Fe content; corresponding scanning electron microscopy (SEM) images of the as-deposited film surfaces. were pure pyrrhotite. Moreover, this combination of precursors also offers the "cleanest" way to deposit FeS: i.e. without any source of carbon or other likely contamination. The downside is the toxicity of the precursors; beyond Fe(CO)5 , hydrogen sulfide is an extremely toxic gas that requires special precautions and handling. In Figure B-6 we show SEM micrographs of the top surface from samples deposited using Fe(CO)5 / H2 S on glass (Fig. B-6a), muscovite mica (Fig. B-6b and c) and NaCl (Fig. B-6d). The choice of substrate had little obvious effect on the morphology of the as-deposited Fe1-x S; in all cases the grains were roughly equiaxed with sizes between several hundreds of nm and one µm. Finally, the exact stoichiometry of the films could not be measured due to their thickness being on the order of 100-1000 nm. The signal from conventional chemical composition techniques, e.g. EDS, was not strong enough to give adequate counting statistics and distinguish the Fe:S ratio within the Fe1-x S composition of 0 ≤ x ≤ 0.125. 125 Fe(CO)5 and H2S precursors: (a) Glass substrate (b) Mica substrate 1 μm (b) Mica substrate 3 μm (d) NaCl substrate 1 μm 200 μm Figure B-6: Iron sulfide films deposited from Fe(CO)5 and H2 S: scanning electron microscopy (SEM) images of films deposited on (a) glass slide, (b and c) mica, (d) NaCl(100) substrate. Template stripping for atomically-smooth surfaces The root-mean squared (RMS) roughness of as-deposited films using Fe(CO)5 / H2 S precursors (Fig. B-7a) was on the order of 50+ nm, as measured by atomic force microscopy (AFM). Given the original motivation of the CVD project to produce thin film samples for tracer diffusion studies, a surface roughness of this order (i.e. greater than the target tracer deposit thickness of 10 nm) was not conducive to obtaining accurate diffusion profiles. Therefore we explored ways to make the films smoother. By vastly reducing the Fe precursor flow such that the overall deposition rate was just a few nm per hour, the films could be fabricated with RMS roughness < 10 nm. However, this approach was impractical for making samples with total thickness of 100 nm or more. Therefore we employed the template stripping technique, which has been used to make "ultrasmooth", patterned metal films for surface plasmonics [204] or flexible electrodes [205] or TiO2 for various applications. [206] The procedure is outlined in Figure B-7c: the as-grown film is coated with an epoxy resin, onto which a glass piece is pressed to remove air bubbles. The epoxy is left to set for the requisite time, before removing the original substrate using the edge of a razor blade. Due to the low surface energy and chemical incompatibility of the substrates (oxide substrates to sulfide films), the CVD film preferentially adheres to the epoxy, exposing the side grown directly onto the polished or atomically-flat, cleaved substrate surface (Fig. B-7b). This technique worked well on a variety of substrates. An alternative that was also attempted was to deposit on NaCl crystals, and wash away the NaCl using distilled water after supporting the FeS film on a glass slide with epoxy. Although also successful, this technique was not as clean as the oxide substrate version. 126 (b) Smooth side (a) As-deposited surface 2 μm (c) Template stripping process Grow film on flat substrate Strip from substrate Coat with high with razor blade temperature epoxy Figure B-7: Template stripping for ultrasmooth sulfide surfaces. Scanning electron microscopy (SEM) images of (a) as-deposited surface (Fe(CO)5 / H2 S) and (b) template stripped (inverted) surface, as grown on polished SiO2 . (c) schematic of the template stripping process. 127 128 Appendix C Diffusivity measurements using thin film samples Before turning to bulk, natural samples to obtain the diffusivity measurements in this work, we carried out preliminary tests using synthetic thin film Fe1-x S samples, fabricated by either chemical vapor deposition (CVD) or sputter deposition. Shown in Figure C-1, iron self-diffusivity values measured from thin film samples were inconsistently low as compared to data from literature or from our bulk samples measured in this work, and hence were not reported in the main text of this paper. Below, we briefly report how the films were fabricated, describe the diffusion measurement results, and discuss the likely sources of the disparate results. Chemical Vapor Deposition of Fe1-x S films CVD films were grown on soda lime glass pieces according to the procedure outlined in Appendix 2, using Fe(CO)5 and H2 S as precursors. As-deposited films were on the order of 300-700 nm thick. In order to obtain flat surfaces for SIMS analysis, we employed the template stripping technique. [204, 206] The as-grown film was coated in a hightemperature epoxy (stable up to 350 o C), and covered with another glass piece, slightly larger than the substrate size. After curing, the glass piece was removed with a razor blade, and the film preferentially adhered to the cured epoxy, stripping cleanly off the original growth substrate and revealing a surface conforming to the substrate’s original topography. Various polished substrates such as MgO(100), SiO2 , Si and soda lime glass were experimented with; soda lime glass was found to give satisfactory results at the lowest cost. The x-ray diffraction (XRD) pattern of a typical CVD template stripped film is shown in Figure Sputter deposition of Fe1-x S films Sputter deposited films were made according to the procedure outlined in Chapter 2. The XRD pattern is shown in Figure Thin film diffusion measurements: results and discussion Both CVD and sputtered film samples were coated in 10 nm of 57 Fe using thermal evaporation. Annealing runs to produce diffusion profiles were performed by holding the samples in heated nitrate salt baths held at the desired temperature. Samples were vacuum-sealed in quartz tubes during immersion in salt baths. SIMS analysis to obtain diffusion profiles of the annealed samples was as described in the main text. To fit the data, we used the thin film diffusion solution to Fick’s second law [70]: 129 Full Fe self-diffusion results from Ch. 2 and for thin films Temperature (oC) -5 10 900 700 500 400 300 200 150 -10 10 Literature Fryt * log[*D Fe - 1 ] (cm 2 s ) 0.94 eV Condit ** Linear reg. -15 10 -20 10 0.5 This work Bulk crystal Model fit CVD Sputtered 1 1.5 2 2.5 -1 1000/T (K ) Figure C-1: Iron self-diffusivity * DFe measurements obtained from thin film, chemical vapor (CVD) and sputter deposited samples (triangles), alongside the literature data and bulk sample measurements discussed in the main text. Our thin film measurements fall up to 3-4 orders of magnitude lower than the bulk sample results. We attribute the discrepancy to oxide formation from residual oxygen during annealing runs, which binds the 57 Fe deposit as iron oxide and hence reduces the extent of interdiffusivity (discussed in the text below). C(x, t) = Co a−x a+x er f p + er f p 2 2 Dt 2 Dt (C.1) where Co is the concentration at the surface, a is the original deposit thickness, and D is diffusivity. In the limit of a very thin film, a Gaussian approximation to Eq. (C.1) can be given: x2 C(x, t) = p exp − 4Dt 4πDt N (C.2) Figure C-3 shows selected results from thin film profile measurements. For CVD samples (Figures C-3a), we used Eq. (C.2) to fit the results. In Figure C-3b we re-plot the data as C vs. x2 , obtaining a straight line from which diffusivity D was calculated. For the sputter deposited samples, we fit the results using Eq. (C.1). The results of our thin film diffusion experiments are presented in Figure C-1, alongside the literature values and the bulk sample results, as discussed in the main text. Although the slope, i.e. activation energy Q, is consistent with the bulk sample result in the ordered regime below T N , the thin film diffusion coefficients are shifted down by 3-4 orders of magnitude. This difference cannot be explained on the basis of difference in stoichiometry alone (which would only account for approx 10x difference: see Figure 1b 130 XRD of thin film samples (100) (110) Intensity (arb. units) (101) (102) (004) (200) CVD Sputtered 30 40 50 Cu-k α 2 Θ 60 70 Figure C-2: X-ray diffraction of thiin films for ECR experiments. (a) Cu-kα x-ray diffraction (XRD) scans for chemical vapor (CVD) and sputter deposited films, with hexagonal pyrrhotite reflections indicated. (b) scanning electron microscope (SEM) of sputtered film top surface, after post-annealing treatment (inset: 60 o tilted view). (c) pole figure XRD scans confirm the high (100) texture of the sputtered samples. in the main text). Still, a very low iron vacancy concentration would be expected after annealing due to the high volume of pure 57 Fe deposit relative to Fe1-x S in the interdiffused region, that would put * DFe values on the lower limit of the stoichiometric effect. We postulate a further reason for the vastly reduced diffusion coeffieicents: oxidation of the 57 Fe deposit from residual oxygen during annealing. The quartz tubes used to seal the samples during diffusion runs in salt baths were evacuated to ∼ 10-3 Torr, and the sample surface after annealing typically turned from metallic silver to a blue tinge. We performed a control test using a sample annealed under a dynamic flowing atmosphere of 1% H2 S-N2 , where the amount of available oxygen was practically nil. The results are presented in Figure C-4, for two samples (original 57 Fe deposit thickness = 30 nm) annealed at 185 o C. Despite being held for a much shorter time of 5 days compared to 25 days, the sample annealed in the H2 S furnace had a completely flat [57 Fe] profile, indicating that all the deposit had fully diffused through the specimen. Upon removing the sample from the furnace after 5 days, the surface was found to still be metallic silvery in color. Conversely, even after 25 days in the quartz vial, the other sample showed a much lesser extent of diffusion. Moreover, the surface coloration evolved into a blue lustre, indicative of iron oxide formation at the surface. In conclusion, we believe two compounding factors lead to an underestimate of * DFe from our measurements on thin film samples annealed in evacuated quartz vials. First, the residual oxygen in the vials reacted with the 57 Fe deposit surface and tied it up as oxide, preventing rapid interdiffusion with the pyrrhotite substrate. Second, the lack of available sulfur in the atmosphere to react with the deposit meant that vacancies in the samples were "flooded" with Fe, making the samples highly iron-rich. Since Fe diffusion in Fe1-x S occurs via a vacancy-exchange mechanism, removing vacancies would have an additional effect on reducing measured diffusion coefficients. After these initial trials with thin film samples, we decided to utilize bulk natural crystals (giving the results outlined in the main text) rather thanto further pursue thin films annealed under dynamic, H2 S-containing environments. The reason was that for thin tracer deposits of 10’s of nm, the required annealing times to achieve measureable diffusion profiles became impractically short (on the order of seconds). For bulk samples, a much thicker 131 (a) (b) Diffusion profiles for CVD films 0.35 -1 CVD FILMS 0.3 CVD FILMS -1.5 o Annealed at 250 C 3600s Annealed at 250 oC -2 ln[ Fe] 36000s (10 hrs) 0.2 57 511440s (>5 days) 57 [ Fe] 0.25 0.15 -2.5 -3 0.1 -3.5 0.05 0 0 (c) Fit to Gaussian diffusion solution 40 80 120 Depth (nm) -4 200 160 (d) Diffusion profiles: constant temperature 0 0.2 0.4 0.6 4 2 x 2 (nm x 10 ) 0.8 1 Diffusion profiles: constant anneal time 0.8 SPUTTER DEPOSITED FILMS 0.8 0.7 o % Fe 57 0.6 0.5 0.6 Fit D = 1.6 x 10-15 Fit D = 1.1 x 10-15 Fit D = 1.4 x 10-15 57 360 s 1200 s 3600 s 36000 s % Fe 0.7 0.4 Annealed 1 hour 0.5 275 oC 300 oC 315 oC Fit D = 3.5 x 10-16 Fit D = 1.1 x 10-15 Fit D = 1.4 x 10-15 0.4 0.3 0.3 0.2 0.2 0.1 0.1 0 0 SPUTTER DEPOSITED FILMS Annealed at 300 C 20 40 60 Depth (nm) 80 0 0 100 Figure C-3: Representative diffusion profiles of 50 Depth (nm) 100 57 Fe in thin film Fe1-x S, measured by secondary ion mass spectrometry (SIMS) for: (a, b) chemical vapor deposited (CVD) samples annealed at 250 o C for different lengths of time as indicated, fit to the Gaussian solution in Eq. (C.2); (c) sputter deposited samples annealed at 300 o C for different times, fit with the thin film error function solution in Eq. (C.1); (d) sputter deposited films annealed for 1 hour at different temperatures, as displayed. depost of tracer can be made, allowing more reasonble annealing experiments on the order of minutes-hours. 132 Surface oxidation of thin films reduces diffusion 1 185 oC 25 days Quartz vial 0.6 185 oC 5 days H2S furnace [ 57 Fe] 0.8 0.4 0.2 0 0 20 40 60 80 100 x (nm) Figure C-4: Oxidation of samples annealed in quartz vials. 57 Fe diffusion profiles measured for two identically-produced Fe1-x S samples annealed at 185 C inside an evacuated quartz vial, or under a dynamic H2 S-bearing envionment, as indicated. The quartz vial sample turned a blue lustre, whereas the furnace annealed sample retained its original silvery metallic surface. o 133 134 Bibliography [1] Faraday, M. Electricity, volume 2. Dover, NY, (reprinted 1965). [2] Uhlig, H. Passivity of Metals. The Electrochemical Society, Princeton, NJ, (1978). [3] Marcus, P. Corrosion mechanisms in theory and practice. CRC Press, 3rd edition, (2012). [4] Pourbaix, M. Atlas of electrochemical equilibria in aqueous solutions. Science, (1974). [5] OPEC. Share of world oil reserves (2011). [6] Eni. World Oil & Gas Review (2012). [7] http://www.eia.gov/todayinenergy/, (2012). [8] DeBruije, G., Skeates, C., Greenaway, R., Harrison, D., S., M. P., James, F. M., Ray, S., Riding, M., Temple, L., and Wutherich, K. Schlumberger Oilfield Review (2008). [9] Rickard, D., Schoonen, M. A. A., and Luther, G. W. Geochemical Transformations of Sedimentary Sulfur 612, 168–193 (1995). [10] Sun, W. and Nesic, S. NACE Corrosion 07655 (2007). [11] Anderko, A. and Shuler, P. Computers & Geosciences 23(6) (1997). [12] Pourbaix, A., Amalhay, M., and Singh, A. EUROCORR Norway (1997). [13] Ueda, M. Corrosion Engineering 44(3), 159–174 (1995). [14] Ning, J., Zheng, Y., Young, D., Brown, B., and Nesic, S. NACE Corrosion (Paper No. 2462) (2013). [15] Smith, S., Brown, B., and Sun, W. Corrosion 11081 (2011). [16] Sun, W. and Nesic, S. Corrosion 65(5), 291–307 (2009). [17] Ramanarayanan, T. A. and Smith, S. N. Corrosion 46(1), 66–74 (1990). [18] Kvarekval, J., Nyborg, R., and Choi, H. Corrosion 03339 (2003). [19] Smith, S. and Joosten, M. NACE Corrosion 06115 (2006). [20] Shoesmith, D., Bailey, M., and Ikeda, B. Electrochimica Acta 23(12), 1329 – 1339 (1978). [21] Nesic, S., Li, H., Huang, J., and Sormaz, D. NACE Corrosion 09572 (2009). [22] Zhang, L., Zhong, W., Yang, J., Gu, T., Xiao, X., and Lu, M. NACE Corrosion 11079 (2011). [23] Singer, M., Camacho, A., Brown, B., and Nesic, S. Corrosion 2010 10100 (2010). [24] Ennaoui, A., Fiechter, S., Pettenkofer, C., Alonsovante, N., Buker, K., Bronold, M., Hopfner, C., and Tributsch, H. Solar Energy Materials and Solar Cells 29(4), 289–370 (1993). [25] Verwey, E. Physica 2, 1059–1063 (1935). [26] Cabrera, N. and Mott, N. F. Reports on Progress in Physics 12(1), 163 (1949). [27] Fehlner, F. and Mott, N. 2(1), 59–99 (1970). [28] Wagner, C, S. W. Z. Phys. Chem. (B) 11, 163 (1930). [29] Mrowec, S. and Przybylski, K. 23(3-4), 107–139 (1985). [30] Danielewski, M., Mrowec, S., and Stoaosa, A. 17(1-2), 77–97 (1982). [31] Fryt, E. M., Smeltzer, W. W., and Kirkaldy, J. S. Journal of The Electrochemical Society 126(4), 673–683 (1979). [32] Macdonald, D. D. Electrochimica Acta 56(4), 1761–1772 (2011). [33] Macdonald, D. D. Pure and Applied Chemistry 71(6), 951–978 (1999). [34] Laycock, N., Noh, J., White, S., and Krouse, D. Corrosion Science 47(12), 3140–3177 (2005). [35] Turnbull, A., McCartney, L., and Zhou, S. Corrosion Science 48(8), 2084–2105 (2006). [36] Vedage, H., Ramanarayanan, T. A., Mumford, J. D., and Smith, S. N. Corrosion 49(2), 114–121 (1993). [37] Worrell, W. and Kaplan, H. Heterogeneous Kinetics at Elevated Temperatures. Plenum Press, New York, (1970). 135 [38] Sterten, A. Corrosion Science 14(6), 377–390 (1974). [39] Hobbins, R. Self-Diffusion of Iron in Single Crystals of Ferrous Sulfide and Magnetically Saturated Iron. PhD thesis, University of Delaware, (1970). [40] Condit, R., Hobbins, R., and Birchenall, C. 8(6), 409–455 (1974). [41] Marusak, L. A. and Mulay, L. N. Physical Review B 21, 238–244 (1980). [42] Powell, A. V., Vaqueiro, P., Knight, K. S., Chapon, L. C., and Sánchez, R. D. Physical Review B 70, 014415 (2004). [43] Wang, H. and Salveson, I. Phase Transitions 78(7-8), 547–567 (2005). [44] Hagemann, I. S., Huang, Q., Gao, X. P. A., Ramirez, A. P., and Cava, R. J. Physical Review Letters 86(5), 894–897 (2001). [45] Nakazawa, H. and Morimoto, N. Materials Research Bulletin 6(5), 345–357 (1971). [46] Elliot and Alexander. Acta Crystallographica Section B 66(3), 271–279 (2010). [47] de Villiers, J.P.R., L. D. B. M. American Mineralogist 94(10) (2009). [48] Francis, C. A. and Craig, J. R. American Mineralogist 61(1-2), 21–25 (1976). [49] Nakazawa, H., Morimoto, N., and Watanabe, E. In Electron Microscopy in Mineralogy, Wenk, H.-R., editor, 304–309. Springer Berlin Heidelberg (1976). [50] Fan, L. Studies of structures and phase transitions in pyrrhotite. PhD thesis, Iowa State University, (1997). [51] Li, F., Franzen, H. F., and Kramer, M. J. Journal of Solid State Chemistry 124(2), 264–271 (1996). [52] Pearce, C., Pattrick, R., and Vaughan, D. Reviews in Mineralogy and Geochemistry 61, 127–180 (2006). [53] Kissin, S. A. and Scott, S. D. Economic Geology 77(7), 1739–1754 (1982). [54] Schwarz, E. and Vaughan, D. Journal of geomagnetism and geoelectricity 24(4), 441–458 (1972). [55] Vaughan, D. J., Schwarz, E. J., and Owens, D. R. Economic Geology 66(8), 1131–1144 (1971). [56] Li, F. and Franzen, H. F. Journal of Solid State Chemistry 126(1), 108–120 (1996). [57] Yamamoto, A. and Nakazawa, H. Acta Crystallographica Section A 38(1), 79–86 (1982). [58] Neel, L. Journal de Physique et Le Radium 5, 241–265 (1944). [59] Lotgering, F. Phillips Research Reports 11, 190–217 (1956). [60] Yue, G. H., Yan, P. X., Fan, X. Y., Wang, M. X., Qu, D. M., Yan, D., and Liu, J. Z. Journal of Applied Physics 100(12), – (2006). [61] Nath, M., Choudhury, A., Kundu, A., and Rao, C. Adv. Mater. 15(24), 2098–2101 (2003). [62] Lyubutin, I. S., Lin, C.-R., Lu, S.-Z., Siao, Y.-J., Korzhetskiy, Y. V., Dmitrieva, T. V., Dubinskaya, Y. L., Pokatilov, V. S., and Konovalova, A. O. 13(10), 5507–5517– (2011). [63] Takayama, T. and Takagi, H. Applied Physics Letters 88(1), – (2006). [64] Townsend, M., Webster, A., Horwood, J., and Roux-Buisson, H. Journal of Physics and Chemistry of Solids 40(3), 183–189 (1979). [65] Arnold, R. G. Economic Geology 57(1), 72–90 (1962). [66] Arnold, R. G. Economic Geology 64(4), 405–419 (1969). [67] Walder, P. and Pelton, A. 26(1), 23–38 (2005). [68] Jesche, A., McCallum, R., Thimmaiah, S., Jacobs, J., Taufour, V., Kreyssig, A., Houk, R., Budanko, S., and Canfield, P. Nature Communications 5 (2014). [69] Loving, M. Understanding the magnetostructural transformation in FeRh thin films. PhD thesis, Northeastern University, (2014). [70] Balluffi, R.W., A. S. C. W. Kinetics of Materials. Wiley, (2005). [71] Metzler, R. and Klafter, J. Physics Reports 339(1), 1–77 (2000). [72] Jonscher, A. K. Journal of Physics D: Applied Physics 32(14), R57 (1999). [73] Walton, E. B. and VanVliet, K. J. Phys. Rev. E 74(6), 061901 (2006). [74] Kakalios, J., Street, R. A., and Jackson, W. B. Physical Review Letters 59(9), 1037–1040 (1987). [75] Klafter, J. and Shlesinger, M. F. Proceedings of the National Academy of Sciences 83(4), 848–851 (1986). [76] Palmer, R. G., Stein, D. L., Abrahams, E., and Anderson, P. W. Physical Review Letters 53(10), 958–961 (1984). [77] Egerton, R. Electron-energy loss spectroscopy in the electron microscope, 3rd Ed. Springer, (2011). [78] Sabioni, A., Huntz, A., Silva, F., and Jomard, F. Materials Science and Engineering: A 392(12), 254– 261 (2005). [79] http://www.nndc.bnl.gov/chart/. . 136 [80] Jacobson, A. J. Chemistry of Materials 22(3), 660–674 (2009). [81] Zhang, B. AIP Advances 4(1), – (2014). [82] Perez, R. and Weissmann, M. Journal of Physics: Condensed Matter 16(39), 7033 (2004). [83] Ding, H., Razumovskiy, V. I., and Asta, M. Acta Materialia 70(0), 130–136 (2014). [84] Perez, R., Torres, D., and Dyment, F. Applied Physics A 97(2), 381–385 (2009). [85] Nitta, H. and Iijima, Y. Philosophical Magazine Letters 85(10), 543–548 (2005). [86] Yang, J. and Goldstein, J. 35(6), 1681–1690 (2004). [87] Ruch, L., Sain, D. R., Yeh, H. L., and Girifalco, L. Journal of Physics and Chemistry of Solids 37(7), 649–653 (1976). [88] Pareek, V., Ramanarayanan, T., and Mumford, J. 46(3-4), 223–228 (1997). [89] Pareek, V., Ramanarayanan, T., Mumford, J., and Ozeckcin, A. 27(1-2), 11–25 (1994). [90] Pareek, V. K., Ramanarayanan, T. A., and Mumford, J. D. Journal of The Electrochemical Society 142(6), 1784–1788 (1995). [91] Birkholz, M., Lichtenberger, D., Hopfner, C., and Fiechter, S. Solar Energy Materials and Solar Cells 27(3), 243–251 (1992). [92] Pratt, A., Muir, I., and Nesbitt, H. Geochimica et Cosmochimica Acta 58(2), 827–841 (1994). [93] Gupta, R. P. and Sen, S. K. Phys. Rev. B 10(1), 71–77 (1974). [94] Gupta, R. P. and Sen, S. K. Physical Review B 12(1), 15–19 (1975). [95] Yuan, D. and Kröger, F. A. Journal of The Electrochemical Society 118(6), 841–846 (1971). [96] Pitzer, K. Thermodynamics. New York McGraw-Hill, (1995). [97] Wang, S., van der Heide, P., Chavez, C., Jacobson, A., and Adler, S. Solid State Ionics 156(12), 201– 208 (2003). [98] Kim, S., Wang, S., Chen, X., Yang, Y. L., Wu, N., Ignatiev, A., Jacobson, A. J., and Abeles, B. Journal of The Electrochemical Society 147(6), 2398–2406 (2000). [99] Kim, G., Wang, S., Jacobson, A., and Chen, C. Solid State Ionics 177(17), 1461–1467 (2006). [100] van der Haar, L. M., den Otter, M. W., Morskate, M., Bouwmeester, H. J. M., and Verweij, H. Journal of The Electrochemical Society 149(3), J41–J46 (2002). [101] ten Elshof, J. E., Lankhorst, M. H. R., and Bouwmeester, H. J. M. Journal of The Electrochemical Society 144(3), 1060–1067 (1997). [102] Egger, A. and Sitte, W. Solid State Ionics 258(0), 30–37 (2014). [103] Gopal, C. B. and Haile, S. M. Journal of Materials Chemistry A 2(7), 2405–2417 (2014). [104] Nakamura, T., Yashiro, K., Sato, K., and Mizusaki, J. Materials Chemistry and Physics 122(1), 250–258 (2010). [105] Chen, X., Wang, S., Yang, Y., Smith, L., Wu, N., Kim, B.-I., Perry, S., Jacobson, A., and Ignatiev, A. Solid State Ionics 146(3-4), 405–413 (2002). [106] Herbert, F., Krishnamoorthy, A., Van Vliet, K., and Yildiz, B. Surface Science 618(0), 53–61 (2013). [107] Bronold, M., Tomm, Y., and Jaegermann, W. Surface Science 314(3), L931–L936 (1994). [108] Krishnamoorthy, A., Herbert, F. W., Yip, S., Van Vliet, K., and Yildiz, B. Journal of Physics: Condensed Matter 25(4), 045004 (2012). [109] Yu, L. P., Lany, S., Kykyneshi, R., Jieratum, V., Ravichandran, R., Pelatt, B., Altschul, E., Platt, H. A. S., Wager, J. F., Keszler, D. A., and Zunger, A. Advanced Energy Materials 1(5), 748–753 (2011). [110] Zhang, Y. N., Hu, J., Law, M., and Wu, R. Q. Physical Review B 85(8) (2012). [111] Heine, V. Physical Review 138(6A), 1689 (1965). [112] Kronik, L. and Shapira, Y. Surface and Interface Analysis 31(10), 954–965 (2001). [113] Stroscio, J. A., Feenstra, R. M., and Fein, A. P. Physical Review Letters 57(20), 2579–2582 (1986). [114] Feenstra, R. M. Physical Review B 50(7), 4561–4570 (1994). [115] Feenstra, R. M., Dong, Y., Semtsiv, M. P., and Masselink, W. T. Nanotechnology 18(4) (2007). [116] Ishida, N., Sueoka, K., and Feenstra, R. M. Physical Review B 80(7) (2009). [117] Feenstra, R. M. Physical Review B 44(24), 13791–13794 (1991). [118] Maboudian, R., Pond, K., Bresslerhill, V., Wassermeier, M., Petroff, P. M., Briggs, G. A. D., and Weinberg, W. H. Surface Science 275(1-2), L662–L668 (1992). [119] Sabitova, A., Ebert, P., Lenz, A., Schaafhausen, S., Ivanova, L., Dahne, M., Hoffmann, A., DuninBorkowski, R. E., Forster, A., Grandidier, B., and Eisele, H. Applied Physics Letters 102(2), 021608–4 (2013). [120] Ebert, P., Schaafhausen, S., Lenz, A., Sabitova, A., Ivanova, L., Dahne, M., Hong, Y. L., Gwo, S., and 137 Eisele, H. Applied Physics Letters 98(6) (2011). [121] Ebert, P., Ivanova, L., and Eisele, H. Physical Review B 80(8), 085316 (2009). [122] Egan, C. K., Choubey, A., and Brinkman, A. W. Surface Science 604(19-20), 1825–1831 (2010). [123] Ivanova, L., Borisova, S., Eisele, H., Dahne, M., Laubsch, A., and Ebert, P. Applied Physics Letters 93(19) (2008). [124] Guevremont, J. M., Strongin, D. R., and Schoonen, M. A. A. American Mineralogist 83(11-12), 1246– 1255 (1998). [125] Guevremont, J. M., Elsetinow, A. R., Strongin, D. R., Bebie, J., and Schoonen, M. A. A. American Mineralogist 83(11-12), 1353–1356 (1998). [126] Guevremont, J., Strongin, D. R., Schoonen, M. A. A., and Bebie, J. Abstracts of Papers of the American Chemical Society 213, 140 (1997). [127] Murphy, R. and Strongin, D. R. Surface Science Reports 64(1), 1–45 (2009). [128] Uhlig, I., Szargan, R., Nesbitt, H. W., and Laajalehto, K. Applied Surface Science 179(1-4), 222–229 (2001). [129] Rosso, K. M. Molecular Modeling Theory: Applications in the Geosciences 42, 199–271 (2001). [130] Bebie, J. and Schoonen, M. A. Geochemical Transactions 1(1), 47 (2000). [131] Boehme, C. and Marx, D. Journal of the American Chemical Society 125(44), 13362–13363 (2003). [132] Nair, N. N., Schreiner, E., and Marx, D. Journal of the American Chemical Society 128(42), 13815– 13826 (2006). [133] Caban-Acevedo, M., Faber, M. S., Tan, Y. Z., Hamers, R. J., and Jin, S. Nano Letters 12(4), 1977–1982 (2012). [134] Hu, J., Zhang, Y. N., Law, M., and Wu, R. Q. Journal of the American Chemical Society 134(32), 13216–13219 (2012). [135] Sun, R. S., Chan, M. K. Y., Kang, S. Y., and Ceder, G. Physical Review B 84(3) (2011). [136] Sun, R. S., Chan, M. K. Y., and Ceder, G. Physical Review B 83(23) (2011). [137] Burton, J. D. and Tsymbal, E. Y. Physical Review Letters 107(16) (2011). [138] Bi, Y., Yuan, Y. B., Exstrom, C. L., Darveau, S. A., and Huang, J. S. Nano Letters 11(11), 4953–4957 (2011). [139] Yang, T. R., Yu, J. T., Huang, J. K., Chen, S. H., Tsay, M. Y., and Huang, Y. S. Journal of Applied Physics 77(4), 1710–1714 (1995). [140] Ho, C. H., Huang, Y. S., and Tiong, K. K. Journal of Alloys and Compounds 422(12), 321–327 (2006). [141] Tsay, M. Y., Huang, Y. S., and Chen, Y. F. Journal of Applied Physics 74(4), 2786–2789 (1993). [142] Wadia, C., Wu, Y., Gul, S., Volkman, S. K., Guo, J. H., and Alivisatos, A. P. Chemistry of Materials 21(13), 2568–2570 (2009). [143] Nesbitt, H. W., Bancroft, G. M., Pratt, A. R., and Scaini, M. J. American Mineralogist 83(9-10), 1067– 1076 (1998). [144] Schaufuss, A. G., Nesbitt, H. W., Kartio, I., Laajalehto, K., Bancroft, G. M., and Szargan, R. Surface Science 411(3), 321–328 (1998). [145] Leiro, J. A., Mattila, S. S., and Laajalehto, K. Surface Science 547(1-2), 157–161 (2003). [146] Mattila, S., Leiro, J. A., and Laajalehto, K. Applied Surface Science 212, 97–100 (2003). [147] Mattila, S., Leiro, J. A., and Heinonen, M. Surface Science 566, 1097–1101 (2004). [148] Andersson, K., Nyberg, M., Ogasawara, H., Nordlund, D., Kendelewicz, T., Doyle, C. S., Brown, G. E., Pettersson, L. G. M., and Nilsson, A. Physical Review B 70(19) (2004). [149] Rosso, K. M., Becker, U., and Hochella, M. F. American Mineralogist 85(10), 1428–1436 (2000). [150] Steinhagen, C., Harvey, T. B., Stolle, C. J., Harris, J., and Korgel, B. A. Journal of Physical Chemistry Letters 3(17), 2352–2356 (2012). [151] Rosso, K. M., Becker, U., and Hochella, M. F. American Mineralogist 84(10), 1535–1548 (1999). [152] Eggleston, C. M., Ehrhardt, J. J., and Stumm, W. American Mineralogist 81(9-10), 1036–1056 (1996). [153] Siebert, D. and Stocker, W. Physica Status Solidi A 134(1), K17–K20 (1992). [154] Willeke, G., Blenk, O., Kloc, C., and Bucher, E. Journal of Alloys and Compounds 178, 181–191 (1992). [155] Becker, U., Munz, A. W., Lennie, A. R., Thornton, G., and Vaughan, D. J. Surface Science 389(1-3), 66–87 (1997). [156] Chaturvedi, S., Katz, R., Guevremont, J., Schoonen, M. A. A., and Strongin, D. R. American Mineralogist 81(1-2), 261–264 (1996). [157] Feenstra, R. www.andrew.cmu.edu/user/feenstra/semitip_v6 (2011). [158] Feenstra, R. M. Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures 138 21(5), 2080–2088 (2003). [159] Dong, Y., Feenstra, R. M., Semtsiv, M. P., and Masselink, W. T. Journal of Applied Physics 103(7) (2008). [160] Gaan, S., He, G. W., Feenstra, R. M., Walker, J., and Towe, E. Journal of Applied Physics 108(11) (2010). [161] Tersoff, J. and Hamann, D. R. Physical Review B 31(2), 805–813 (1985). [162] Zhao, G. L., Callaway, J., and Hayashibara, M. Physical Review B 48(21), 15781–15786 (1993). [163] Hu, J., Zhang, Y. N., Law, M., and Wu, R. Q. Physical Review B 85(8) (2012). [164] Ebert, P. Surface Science Reports 33, 121–303 (1999). [165] Ebert, P., Heinrich, M., Simon, M., Urban, K., and Lagally, M. G. Physical Review B 51(15), 9696–9701 (1995). [166] Becker, U. and Rosso, K. M. American Mineralogist 86(7-8), 862–870 (2001). [167] Lazic, P., Armiento, R., Herbert, F. W., Chakraborty, R., Sun, R., Chan, M. K. Y., Hartman, K., Buonassisi, T., Yildiz, B., and G. Journal of Physics: Condensed Matter 25(46), 465801 (2013). [168] Limpinsel, M., Farhi, N., Berry, N., Lindemuth, J., Perkins, C. L., Lin, Q., and Law, M. Energy Environ. Sci. 7(6), 1974–1989 (2014). [169] Herbert, F., Krishnamoorthy, A., Ma, W., Van Vliet, K., and Yildiz, B. Electrochimica Acta 127(0), 416–426 (2014). [170] Macdonald, D. D. Journal of The Electrochemical Society 139(12), 3434–3449 (1992). [171] Ryan, M. P., Williams, D. E., Chater, R. J., Hutton, B. M., and McPhail, D. S. Nature 415(6873), 770–774 (2002). [172] Maurice, V., Klein, L. H., Strehblow, H.-H., and Marcus, P. J. Phys. Chem. C 111(44), 16351–16361 October (2007). [173] Jasperse, J. R. and Doherty, P. E. Philosophical Magazine 9(100), 635 (1964). [174] Zhang, Z., Pan, J. S., Zhang, J., and Tok, E. S. Physical Chemistry Chemical Physics 12(26), 7171–7183 (2010). [175] Massoud, T., Maurice, V., Wiame, F., Klein, L. H., Seyeux, A., and Marcus, P. Journal of the Electrochemical Society 159(8), C351–C356 (2012). [176] Marcus, P., Maurice, V., and Strehblow, H. H. Corrosion Science 50(9), 2698–2704 (2008). [177] Zhou, H., Fu, J., and Silver, R. M. The Journal of Physical Chemistry C 111(9), 3566–3574 (2007). [178] Lillard, R. S., Wang, G. F., and Baskes, M. I. Journal of The Electrochemical Society 153(9), B358–B364 (2006). [179] Maurice, V., Despert, G., Zanna, S., Bacos, M. P., and Marcus, P. Nature Materials 3(10), 687–691 (2004). [180] Birkholz, M., Fiechter, S., Hartmann, A., and Tributsch, H. Physical Review B 43(14), 11926–11936 (1991). [181] Fan, F. R. and Bard, A. J. Journal of Physical Chemistry 95(5), 1969–1976 (1991). [182] Herbert, F., Krishnamoorthy, A., Van Vliet, K., and Yildiz, B. In preparation (2014). [183] Kröger, F. A. The Chemistry of Imperfect Crystals. North Holland, Amsterdam, The Netherlands, (1964). [184] Hebert, K. R., Albu, S. P., Paramasivam, I., and Schmuki, P. Nature Mater 11(2), 162–166 (2012). [185] Seyeux, A., Maurice, V., and Marcus, P. Electrochemical and Solid-State Letters 12(10), C25–C27 (2009). [186] Yoshimi, K., Hanada, S., Haraguchi, T., Kato, H., Itoi, T., and Inoue, A. Materials Transactions 43(11), 2897–2902 (2002). [187] Uedono, A., Yamashita, Y., Tsutsui, T., Dordi, Y., Li, S., Oshima, N., and Suzuki, R. Journal of Applied Physics 111(10) (2012). [188] Namai, Y., Fukui, K. I., and Iwasawa, Y. Catalysis Today 85(2-4), 79–91 (2003). [189] Torbrugge, S., Reichling, M., Ishiyama, A., Morita, S., and Custance, O. Physical Review Letters 99(5) (2007). [190] Zhou, Y. T., Zhang, B., Zheng, S. J., Wang, J., San, X. Y., and Ma, X. L. Scientific Reports (2014). [191] Andersson, K. J., Ogasawara, H., Nordlund, D., Brown, G. E., and Nilsson, A. J. Phys. Chem. C (2014). [192] Frank, E. (2007). [193] Sakkopoulos, S., Vitoratos, E., and Argyreas, T. Journal of Physics and Chemistry of Solids 45(8-9), 923–928 (1984). [194] M.W. Chase, C.A. Davies, J. D. D. F. R. M. A. S. JANAF thermochemical tables, 3rd. ed., J. Phys. Chem. Ref. Data v. 14, Supplement No. I, p. 1-1856, (1985). 139 [195] Bemer, R. American Journal of Science 265(9), 773–785 (1967). [196] D. Wei, K. O.-A. Journal of Colloid Interface Science 174(2), 273–282 (1995). [197] O.M. Suleimenov, R. K. Geochimica et Cosmochimica Acta 61(24), 5187–5198 (1994). [198] Berry, N., Cheng, M., Perkins, C. L., Limpinsel, M., Hemminger, J. C., and Law, M. Advanced Energy Materials (2012). [199] Hopfner, C., Ellmer, K., Ennaoui, A., Pettenkofer, C., Fiechter, S., and Tributsch, H. Journal of Crystal Growth 151(34), 325–334 (1995). [200] Thomas, B., Ellmer, K., MÃijller, M., HÃűpfner, C., Fiechter, S., and Tributsch, H. Journal of Crystal Growth (1997). [201] Thomas, B., Cibik, T., Höpfner, C., Diesner, K., Ehlers, G., Fiechter, S., and Ellmer, K. 9(1), 61–64 (1998). [202] Almond, M. J., Redman, H., and Rice, D. A. J. Mater. Chem. 10(12), 2842–2846 (2000). [203] Chatzitheodorou, G., Fiechter, S., Konenkamp, R., Kunst, M., Jaegermann, W., and Tributsch, H. Materials Research Bulletin 21(12), 1481–1487 (1986). [204] Nagpal, P., Lindquist, N. C., Oh, S.-H., and Norris, D. J. Science 325(5940), 594–597 (2009). [205] Nam, S., Song, M., Kim, D.-H., Cho, B., Lee, H. M., Kwon, J.-D., Park, S.-G., Nam, K.-S., Jeong, Y., Kwon, S.-H., Park, Y. C., Jin, S.-H., Kang, J.-W., Jo, S., and Kim, C. S. Scientific Reports 4 (2014). [206] Vogel, N., Zieleniecki, J., and Koper, I. Nanoscale 4(13), 3820–3832 (2012). 140