The Pennsylvania State University The Graduate School College of Earth and Mineral Sciences THE EFFECT OF OXYGEN FUGACITY ON THE EQUILIBRIUM PARTITIONING OF LITHIUM BETWEEN OLIVINE AND CLINOPYROXENE A Thesis in Geosciences by James A. Deane Jr. © 2013 James A. Deane Jr. Submitted in Partial Fulfillment of the Requirements for the Degree of Master of Science August 2013 The thesis of James A. Deane Jr. was reviewed and approved* by the following: Maureen D. Feineman Assistant Professor of Geosciences Thesis Advisor David H. Eggler Emeritus Professor of Petrology Peter J. Heaney Professor of Geosciences James D. Kubicki Professor of Geoscience Chris J. Marone Professor of Geosciences Associate Department Head of Graduate Programs *Signatures are on file in the Graduate School ii ABSTRACT Mantle olivines and clinopyroxenes from xenoliths sampled worldwide commonly display disequilibrium distributions of lithium and its isotopes. While many theories of open- and closed- system redistribution of lithium have been proposed, none has been adequately demonstrated experimentally. Closed system redistribution of lithium between olivine and clinopyroxene has been proposed to occur in response to changing equilibrium partition coefficients driven by changes in temperature or oxygen fugacity. Previous experiments (Yakob et al., EPSL. 330, 11-21, 2012) have shown that there is no temperature dependence on Li partitioning between 700-1100°C. The goal of this thesis is to determine the potential effect of oxygen fugacity on lithium partitioning between olivine and clinopyroxene. To this end, a series of piston cylinder experiments was designed to fix the oxygen fugacity at various mantle values while natural olivine and diopside starting materials were equilibrated with a Li-bearing aqueous fluid. Oxygen fugacity was buffered using the solid state fO2 buffers Fayalite-MagnetiteQuartz, Magnetite-Wüstite, and Mo-MoO2. A silver capsule was employed to minimize the solubility of Fe in the capsule, thus keeping it in the experimental charge. Iron retention is critical because the mechanism being tested is coupled substitution of Li+ and Fe3+ for Mg2+ or Fe3+ in the olivine structure. Results from these experiments show there is no apparent relationship between oxygen fugacity and equilibrium partitioning of Li between olivine and clinopyroxene at upper mantle conditions. Over the range of fugacities investigated, DLiol/di was 1.7±0.4 regardless of fO2. The lack of a relationship between Li partitioning and oxygen fugacity, coupled with the previously determined lack of a relationship with temperature, suggests that closed system redistribution of Li between mantle phases is unlikely to explain the isotopic variability we see in xenoliths worldwide. Instead, open system interaction with metasomatic fluids or melts could be the cause. This implies that Li isotopes in mantle xenoliths are not necessarily representative of the ambient mantle from which they are sourced, and therefore limits the use of Li isotopes as a passive tracer of recycled crustal material in the mantle. However, Li isotopic fractionation during open-system processes related to xenolith entrainment and eruption is likely to be useful as a geospeedometer to determine the rates of these processes. iii TABLE OF CONTENTS List of Tables.....................................................................................................................vii List of Figures...................................................................................................................viii Acknowledgements.............................................................................................................xi 1 INTRODUCTION...................................................................................................1 1.1 Li in the Earth..............................................................................................1 1.2 Influence of Oxygen Fugacity on Li Partitioning........................................5 2 METHODS............................................................................................................11 2.1 Starting Materials.......................................................................................11 2.2 Experimental Methods...............................................................................11 2.3 Analytical Methods....................................................................................13 2.3.1 Scanning Electron Microscopy..................................................................13 2.3.2 Electron Microscopy..................................................................................14 2.3.3 Laser Ablation Inductively Coupled Mass Spectrometry..........................14 2.3.4 X-ray Diffraction of fO2 buffers................................................................15 3 RESULTS..............................................................................................................20 3.1 Experimental Results.................................................................................20 3.2 Run Product Compositions........................................................................20 3.3 Oxygen Fugacity Control...........................................................................21 3.4 Partition Coefficients.................................................................................21 4 DISCUSSION........................................................................................................35 4.1 Attainment of Equilibrium in Experiments................................................35 4.2 Oxygen Fugacity........................................................................................35 4.2.1 Oxygen Fugacity in the Mantle..................................................................37 4.3 Li Partitioning as a Function of fO2...........................................................38 4.4 Li Variability in Mantle Minerals..............................................................40 4.4.1 Equilibrium Isotope Fractionation.............................................................40 4.4.2 Open vs. Closed System Redistribution of Li............................................41 5 CONCLUSIONS...................................................................................................60 6 6.1 FAILED EXPERIMENTS.....................................................................................62 Methodology..........................................................................................................62 iv 6.2 Experimental Designs............................................................................................62 References..........................................................................................................................65 Appendix: Sample Assembly and Piston Cylinder Procedures.........................................71 v LIST OF TABLES Table 2.1.1 Amounts of starting materials and run conditions for all experiments used in this study........................................................................................................................16 Table 3.2.1 Major element concentrations of starting materials and experimental run products..............................................................................................................................22 Table 3.2.2 Trace element concentrations of starting materials and experimental run products..............................................................................................................................23 Table 3.4.1 Calculated Li partition coefficients (DLiol/cpx)............................................24 Table 4.4.1 Compilation of Li data from olivine and clinopyroxene mineral separates from xenoliths worldwide. Data from Aulbach and Rudnick (2009)1, Aulbach et al. (2008)2, Ionov and Seitz (2008)3, Jeffcoate et al. (2007)4, Kaliwoda et al. (2008)5, Magna et al. (2006, 2008)6,7, Ottolini et al. (2004)8, Rudnick and Ionov (2007)9, Seitz and Woodland (2000)10, Seitz et al. (2004)11, Tang et al. (2007b)12, and Woodland et al. (2004)13..............................................................................................................................44 Table A.1 Pressure calibration for the piston cylinder apparatus used in this study..78 vi LIST OF FIGURES Figure 1.1.1 δ7Li variability in various terrestrial reservoirs. Figure 1 from Tang et al., 2007a....................................................................................................................................7 Figure 1.1.2 The lithium geochemical cycle. References as follows: Seawater: Chan and Edmond, 1988; Tomascak et al., 1999; James and Palmer, 2000; Rainwater: Millot et al. 2010b; Continents: Teng et al., 2004; Mantle: Chan et al., 1992; Tomascak et al., 2008; Altered Oceanic Crust: Chan et al., 1992, 2002a; Subducted sediment: Chan et al., 2006; Arc lavas: Chan et al., 2002b; Tomascak et al., 2002; Moriguti et al., 2004; Serpentinized mantle: DeCitre et al., 2002; Benton et al., 2004; Rivers: Huh et al., 1998; Millot et al., 2010a; Hydrothermal fluids: Chan et al., 1993, 1994; Subduction reflux: Zhang et al., 1998; James et al., 1999, Chan and Kastner, 2000; Alteration and sediment uptake: Chan et al., 1992, 2002a, 2006; Plank and Langmuir, 1998; Tomascak et al., 2008......................................................................................................................................8 Figure 1.1.3 A compilation of olivine/clinopyroxene lithium partition coefficients (D) vs. isotopic fractionation factors (α) in mantle xenoliths worldwide. Figure 1 from Yakob et al., 2012 with data taken from Aulbach and Rudnick (2009), Aulbach et al. (2008), Ionov and Seitz (2008), Jeffcoate et al. (2007), Kaliwoda et al. (2008), Magna et al. (2006, 2008), Ottolini et al. (2004), Rudnick and Ionov (2007), Seitz and Woodland (2000), Seitz et al. (2004), Tang et al. (2007b), and Woodland et al. (2004)......................9 Figure 1.1.4 δ7Li variability between coexisting mantle phases. In most cases, olivine has significantly higher δ7Li than coexisting clinopyroxene. This inter-mineral variability remains unexplained. Modified from Figure 7 from Seitz et al. (2004) to include data from Rudnick and Ionov (2007) and Tang et al. (2007b)..................................................10 Figure 2.2.1 a) Photograph of the piston cylinder apparatus used to conduct the experiments presented in this thesis. b) Schematic of the piston cylinder showing the arrangement of parts..........................................................................................................17 Figure 2.2.2 Sample assembly used in the piston cylinder experiments. The figure is to scale...................................................................................................................................18 Figure 2.2.3 Schematic diagram of the double capsule experimental setup using two Ag capsules separated by a 50/50 Ag-Pd foil. Schematic diagram shows the theoretical progression of the experiment through time. All experiments were run at 900°C and 1 GPa.....................................................................................................................................19 Figure 3.1.1 Backscattered electron SEM images of run products from experiment TJ67. These crystals have not been mounted in epoxy and polished; rather they have been mounted on electrically conductive tape to image via SEM and show morphology of vii unpolished crystals. Image A shows multiple olivine crystals surrounded by quench material. Image B shows one large elongate clinopyroxene.............................................25 Figure 3.1.2 SEM backscattered electron image of run products from experiment TJ-52 buffered at Mo-MoO2. Sample was mounted in epoxy and polished with 600 grit siliconcarbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images...............................................26 Figure 3.1.3 SEM backscattered electron image of run products from experiment TJ-63 buffered at Magnetite-Wüstite. Sample was mounted in epoxy and polished with 600 grit silicon-carbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images................................27 Figure 3.1.4 SEM backscattered electron image of run products from experiment TJ-67 buffered at FMQ. Sample was mounted in epoxy and polished with 600 grit siliconcarbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images...............................................28 Figure 3.1.5 SEM backscattered electron image of run products from experiment TJ-68 which was unbuffered. Sample was mounted in epoxy and polished with 600 grit siliconcarbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images...............................................29 Figure 3.2.1 Typical time resolved laser ablation spectrum for standard BCR-2g used in sample-standard bracketing. Black vertical lines indicate the background differentiated from the analysis signal. LA-ICP-MS data was reduced with LasyBoy V3.01, a data reduction spreadsheet created by Joel Sparks at Boston University..................................30 Figure 3.3.1 XRD analysis of the fO2 buffer from TJ-52 is shown in black. The red dotted line represents an XRD spectrum for Mo and the blue dashed line represents an XRD spectrum for Mo-MoO2. TJ-52 shows clear peaks that overlap well with both phases, suggesting that the experimental buffer was successful........................................31 Figure 3.3.2 XRD analysis of the fO2 buffer from TJ-63 is shown in black. The red dashed line represents an XRD spectrum for magnetite. There is no confirmed wüstite in this buffer, but the broad increase in intensity from ~30-45° represents data from an amorphous phase with no defined crystal structure. This is being interpreted as the remnants of the wüstite present at run conditions which decomposed upon quenching as wüstite is not stable at earth surface conditions.................................................................32 Figure 3.3.3 XRD analysis of the fO2 buffer from TJ-67 is shown in black. The red dashed line represents an XRD spectrum for fayalite. The blue dash-dot line represents an XRD spectrum for magnetite. The green dotted line represents an XRD spectrum for viii quartz. While the strongest peaks in TJ-67 overlap with magnetite, there are overlapping peaks that only correspond to quartz and fayalite as well, suggesting that all three phases are present at the conclusion of the experiment and the buffer was successful.................33 Figure 3.4.1 Experimental results showing the measured Li partition coefficient between olivine and clinopyroxene as a function of fO2. Open symbol represents equilibrium partition coefficient from 3-day 900°C experiment presented by Yakob et al., 2012....................................................................................................................................34 Figure 4.1.1 Time series of piston cylinder equilibrium partitioning experiments from Yakob et al. (2012) showing the DLiol/cpx evolving over time. At 900°C the system approaches equilibrium with respect to Li partitioning between olivine and clinopyroxene within 24 hours. Figure 4 from Yakob et al., 2012..........................................................55 Figure 4.2.1 Solid state oxygen fugacity buffers rhenium-rhenium dioxide (Re-ReO2), fayalite-magnetite-quartz (FMQ), magnetite-wüstite (MW), and molybdenummolybdenum dioxide (Mo-MoO2) shown as a function of temperature. Vertical black line indicates experimental run conditions. Buffer curves calculated from: Mo-MoO2 = O’Neill, 1986; Magnetite-Wüstite & Fayalite-Magnetite-Quartz = Myers and Eugster, 1983....................................................................................................................................56 Figure 4.4.1 Experimentally derived DLiol/cpx showing that the olivine/clinopyroxene partition coefficient is ~2±0.5 regardless of temperature. Figure 6 from Yakob et al., 2012. Data from Ottolini et al. (2009), Caciagli et al. (2011), Brenan et al. (1998), Yakob et al. (2012) and Blundy and Dalton (2000)......................................................................57 Figure 4.4.2 Northrop and Clayton (1966) state that a linear relationship should exist between ln(α)*103 vs. T-2*106 if isotopic fractionation exists as an equilibrium phenomenon. Plotting ln(α)*103 vs. T-2*106 for all available xenolith data shows that in most cases there is no observable relationship, and therefore the isotopic fractionation seen in xenoliths is not the product of equilibrium isotope fractionation. A particular suite from Rudnick and Ionov (2007), however, do seem to fit to a linear trend suggesting a potential relationship between isotopic fractionation and temperature. Data from Table 4.4.1....................................................................................................................................58 Figure 4.4.3 Apparent Li partition coefficients measured from mantle xenolith mineral separates worldwide show no relationship with temperature. Data from Table 4.4.1.......59 ix ACKNOWLEDGEMENTS The successes and failures of this research project proved to be huge test of patience and dedication to succeed, and I couldn’t have completed this project without the help from many wonderful people. First, I would like to thank my advisor, Dr. Maureen Feineman. She has always been supportive and helpful with matters both related and not to this project. Many experimental failures led to us talking about what to do next, and while we got frustrated, she never allowed us to turn our backs on this project. She has been a wonderful advisor and a great mentor for me to talk to about my past, present and future. I would also like to thank the members of my thesis committee who’ve helped me learn to think about scientific problems from many different angles. I’ve learned that focusing on one aspect of science is incredibly limiting, and true productivity can only happen when thinking about problems in many different ways. I’d like to specifically thank Dr. David Eggler, whose experimental expertise was particularly helpful. Whenever there were experimental failures or mechanical breakdowns, I would call Dave and he would come in and help me fix the machines or redesign our experimental approach. I’d like to thank Melanie Saffer, Nichole Wonderling and Mike Ackerson for their help with the geochemical analyses performed in this study. I’d also like to thank Bruce Watson, Dustin Trail and Jay Thomas for their help in designing the experimental setup used in this study. I would especially like to thank all of my friends and family, specifically my mom and dad, my brothers Ryan and Brandon, and my lovely wife-to-be Camille, who have all been incredibly supportive throughout the process of completing this research project. Their guidance and encouragement has truly been an inspiration in keeping me focused on finishing this project and completing graduate school. Thank you x 1. INTRODUCTION 1.1 Lithium in the Earth Lithium is a trace element found both on the surface and within the interior of the Earth. Because Li is very fluid mobile and found within oceanic basalts and sediments in isotopic ratios that contrast with those inherent to the mantle (Chan and Edmond 1988; Chan et al., 1994; Elderfield and Schultz 1996), there has been great hope over the past two decades that lithium, and in particular the isotopes of Li, can serve as a geochemical tracer of surface material recycled back into the mantle (Elliott et al., 2004; Seitz and Woodland, 2000; Kent and Rossman, 2002). Research into lithium behavior in the mantle has been an evolving field of geochemistry for the past 20 years, and there is still much to learn about how lithium and its isotopes behave in Earth systems. Lithium concentrations and isotopic ratios have been measured in whole rock samples, mineral separates, and analyses of intra-mineral zoning from all around the world (e.g. Chan et al., 1992; Chan et al., 1994; Coogan et al., 2005; Jeffcoate et al., 2007; Elliott et al., 2004; Zack et al., 2003; etc.), but they have not yet led to a full understanding of how Li behaves chemically in various environments, for example during recycling from the Earth’s surface into the mantle at subduction zones. Lithium is a light trace element with two stable isotopes, 6Li and 7Li (7.5% and 92.5% abundances, respectively). There is ~15% difference in mass between the two isotopes, which is relatively large, and 6Li and 7Li are strongly fractionated by mass dependent processes at low temperatures (Fritz and Whitworth, 1994; Pistiner and Henderson, 2003; Huh et al., 1998; Huh et al., 2001). Lithium is a very fluid mobile element and therefore can be mobilized and isotopically fractionated by weathering processes at the Earth’s surface (Pistiner and Henderson, 2003; Rudnick et al., 2004; Millot et al., 2010), circulation of hydrothermal fluids within spreading centers or volcanic systems (Chan et al., 1993), or during dehydration/hydration reactions within subduction zones (Chan et al., 2002; DeCitre et al., 2002; Zack et al., 2003; Leeman et al., 2004). The fluid mobility of Li contributes to its potential as a geochemical tracer for mass transfer within subduction zones. Lithium incorporated in altered oceanic crust and subducted sediment is released and carried by slab-derived fluids back up through the 1 overlying mantle wedge and into volcanic arcs (Chan et al. 1992; Leeman et al. 2004) and can be incorporated into mantle minerals such as olivine and pyroxene (Seitz and Woodland, 2000; Kent and Rossman, 2002; Rudnick and Ionov, 2007). This process should lead to a distinct chemical and isotopic signature in magmas and mantle peridotites derived from volcanic arcs (Elliott et al., 2004). Richter et al. (2003) showed that Li diffusion in silicate melt is 1-3 orders of magnitude faster than diffusion of almost all other major and trace elements. The only other species with comparable diffusion coefficients are light noble gases (He, Ne) and water. Because diffusion rates are mass-dependent, and because the two isotopes of lithium have such a large relative mass difference, significant isotope fractionation may occur during diffusion. This can result in Li isotopic fractionation of up to 10-25‰ in natural and experimental systems (Richter et al. 2003). Combining these two observations, rapid diffusion could cause Li to be highly fractionated on short timescales within minerals and rocks in natural samples. Thus Li zoning within phenocrysts or between minerals in mantle xenoliths has the potential to tell us how long a particular crystal took to cool below the closure temperature of Li diffusion (Coogan et al., 2005; Jeffcoate et al., 2007; Parkinson et al., 2007). Isotopically, Li is reported in terms of δ7Li [(7Li/ 6Lisample / 7Li/6LiL-SVEC − 1)103], and expressed in permil (‰) variation relative to the L-SVEC lithium carbonate standard reference material. The value of δ7Li varies widely in different Earth reservoirs (Fig. 1.1.1, from Tang et al., 2007a). Seawater, for example, has a very uniform δ7Li of ~ +33‰ (Chan and Edmond, 1988; Tomascak et al., 1999; James & Palmer, 2000) while unaltered Mid-Ocean Ridge Basalt (MORB), representative of Earth’s upper mantle has δ7Li values ranging from +2‰ to +7‰ (Chan et al., 1992; Seitz et al., 2004). Figure 1.1.2 shows the variability of Li isotopic signatures throughout the Earth’s hydrologic cycle, the crust and the upper mantle, and how these various reservoirs interact with each other. Many mantle xenoliths worldwide (Table 1.1.1; Figure 1.1.3) display expected “equilibrium” behavior with respect to Li distribution and isotope ratio between olivine and diopside. Nonetheless, there are many mantle xenolith mineral separates (e.g., Seitz 2 et al., 2004; Rudnick and Ionov, 2007; Tang et al., 2007b; Fig. 1.1.4) that demonstrate measurably different Li isotopic compositions. In the mantle xenoliths that display the most extreme isotopic difference between olivine and clinopyroxene, the partition coefficient (DLiol/di) decreases to values below 0.5 while the isotopic fractionation factor (ol/di) increases to values greater than 1.020, signifying that clinopyroxene has a greater Li concentration and lower δ7Li relative to olivine, representing an increase in 6Li in clinopyroxene relative to olivine (e.g., Rudnick and Ionov, 2007; Tang et al., 2007b; Fig. 1.1.3). Experiments have shown that the equilibrium partition coefficient for Li between olivine and clinopyroxene (D = [Li]ol/[Li]cpx) is between 1.5 and 2 (Yakob et al., 2012; Brenan et al., 1998; Blundy and Dalton 2000; Caciagli et al., 2011; Ottolini et al., 2009). At the high temperatures and pressures characteristic of the mantle, isotope fractionation between phases (e.g., olivine and diopside) at equilibrium is expected to be minimal. Equilibrium isotopic fractionation between two phases at high temperature (>200°C) has been shown to have a linear relationship between 1000*ln(α) (where α = isotopic fractionation factor, i.e.: (7Li/6Li)ol/(7Li/6Li)cpx) and 1/T2 (e.g., Northrop and Clayton, 1966). The lack of any linear relationship with temperature in worldwide mantle xenoliths suggests that isotopic variability does not represent equilibrium fractionation (Discussed in more detail in section 4; Table 4.1.1, Figures 4.1.1, 4.1.2). Also, Tomascak et al. (1999) showed that as the Kilauea Iki lava lake cooled, the fractional crystallization of olivine, clinopyroxene, and plagioclase resulted in no measurable isotopic fractionation of Li. Together, these observations suggest that at high temperature there is little Li isotopic fractionation between olivine and clinopyroxene. While it has been clearly shown that Li isotopic variability exists between olivine, clinopyroxene and orthopyroxene within individual mantle peridotite samples (Jeffcoate et al., 2007; Ionov and Seitz, 2008; Rudnick and Ionov 2007; Seitz et al., 2004; Aulbach and Rudnick 2009; Tang et al., 2007; Magna et al., 2006), experimental studies have not yet been able to confirm a mechanism that controls the variable Li distribution between coexisting phases. In the most general sense, disequilibrium behavior of Li could be the result of either closed-system or open-system processes. 3 1. Closed-system processes. In this category of hypotheses, the distribution of Li is dependent on some condition that changes prior to or during exhumation. For example, the Li partition coefficient between olivine and clinopyroxene could be temperature dependent. A xenolith or peridotite massif cools as it is exhumed. This change in temperature influences the partition coefficient such that Li becomes more compatible in clinopyroxene relative to olivine, resulting in Li moving from the olivine to the clinopyroxene (thus decreasing the Dol/cpx) via diffusion. As has been shown by Richter et al. (2003), 6Li diffuses faster than 7Li, so this process of diffusion could cause dynamic isotopic fractionation on the timescale of diffusion. The fractionation could then be locked into the minerals as they continue to cool beyond the Li closure temperature. Other variables, such as changes in oxygen fugacity, might likewise drive closed-system redistribution of lithium and its isotopes. 2. Open-system processes. In this school of interpretation, lithium from an external source infiltrates the mantle materials prior to or during exhumation. Many mantle rocks, including all xenoliths, interact with a metasomatic fluid or melt during their rise to the surface. This fluid or melt may have a distinctly different Li isotopic signature from the ambient mantle peridotites, and it may interact with each mantle mineral differently. For instance, if Li diffuses from the fluid or melt into the clinopyroxene faster than it diffuses into the olivine, different δ7Li values may be generated in clinopyroxene relative to olivine. The bottom line is that we do not yet have all the tools necessary to fully understand the large variability in measured Li concentrations and δ7Li values from mantle xenoliths worldwide. What is needed are experimental values of the partition coefficients, diffusion coefficients, and isotopic fractionation factors for Li between mantle phases, and an understanding of how these values are affected by environmental and geochemical factors so we can put together a model to help explain what we see in natural samples. 4 1.2 Influence of Oxygen Fugacity on Li Partitioning It has been proposed that Li partitioning between olivine and diopside is controlled by temperature, such that Li becomes more compatible in clinopyroxene relative to olivine as cooling proceeds. However, experimental evidence shows that DLiol/di remains close to 2 over a range of temperatures between 700˚C and 1400˚C (Yakob et al., 2012; Caciagli et al., 2011). Below 700˚C, diffusion of Li in clinopyroxene is too slow to accommodate dynamic redistribution (Coogan et al., 2005). Another variable that may control the Li partitioning between mantle phases is oxygen fugacity. The dominant species of iron in olivine is Fe2+, forming a solid solution with Mg2+ on the octahedral M1 and M2 sites where there is a random distribution of Fe atoms between the two sites (Spandler and O’Neill, 2010). It is possible, however, that under oxidizing conditions Fe3+ could serve as a coupled substitution partner with Li+. Coupled substitutions can preserve overall charge balance in the system, provided that the ionic radii of the substituting species are near enough to that of the main structural ion to not distort the crystal structure. Previous studies have shown that an increase in trivalent cations within the olivine structure can have a significant effect on the partitioning of monovalent cations (Grant and Wood, 2010). For example, an increase in Sc+3 within the system allows for increased Li within experimental olivines (Grant and Wood, 2010). In this case, Fe3+ and Li+ could substitute together into two M2 sites within the crystal (replacing Mg2+ or Fe2+), preserving the overall charge of the coupled sites (+4). Should more Fe3+ be present due to higher oxygen fugacity in the system, it is possible that more Li+ could be preferentially incorporated into the olivine. Although this is not the only mechanism by which Li can be incorporated into the mineral structures, it is hypothesized that fO2 could therefore play a role in in the partitioning of Li between coexisting olivine and clinopyroxene at mantle conditions. In a similar manner to the temperaturedependent partitioning hypothesis, if the partition coefficient is fO2 dependent, changing environmental fO2 conditions prior to or during exhumation could drive diffusion of lithium between olivine and diopside, resulting in kinetically induced isotopic fractionation. It has been proposed that the fO2 effect would be more significant in olivine relative to clinopyroxene, due to the greater Fe content in olivine (Caciagli, 2010). Should Li incorporation into mantle minerals scale directly with Fe+3 content, and 5 assuming the Fe+3/Fe+2 is controlled within the minerals by the oxygen fugacity, then olivine would be more strongly affected than clinopyroxene due to the respective iron contents. Oxygen fugacity effects on equilibrium partitioning of Li have not been tested experimentally for the olivine-clinopyroxene system. The goal of this research is to measure equilibrium distribution of Li between olivine and diopside over a range of oxygen fugacity values relevant to the upper mantle. 6 Figure 1.1.1. δ7Li variability in various terrestrial reservoirs. Figure 1 from Tang et al., 2007a 7 Figure 1.1.2. The lithium geochemical cycle. References as follows: Seawater: Chan and Edmond, 1988; Tomascak et al., 1999; James and Palmer, 2000; Rainwater: Millot et al. 2010b; Continents: Teng et al., 2004; Mantle: Chan et al., 1992; Tomascak et al., 2008; Altered Oceanic Crust: Chan et al., 1992, 2002a; Subducted sediment: Chan et al., 2006; Arc lavas: Chan et al., 2002b; Tomascak et al., 2002; Moriguti et al., 2004; Serpentinized mantle: DeCitre et al., 2002; Benton et al., 2004; Rivers: Huh et al., 1998; Millot et al., 2010a; Hydrothermal fluids: Chan et al., 1993, 1994; Subduction reflux: Zhang et al., 1998; James et al., 1999, Chan and Kastner, 2000; Alteration and sediment uptake: Chan et al., 1992, 2002a, 2006; Plank and Langmuir, 1998; Tomascak et al., 2008. 8 Figure 1.1.3. A compilation of olivine/clinopyroxene lithium partition coefficients (D) vs. isotopic fractionation factors (α) in mantle xenoliths worldwide. Figure 1 from Yakob et al., 2012 with data taken from Aulbach and Rudnick (2009), Aulbach et al. (2008), Ionov and Seitz (2008), Jeffcoate et al. (2007), Kaliwoda et al. (2008), Magna et al. (2006, 2008), Ottolini et al. (2004), Rudnick and Ionov (2007), Seitz and Woodland (2000), Seitz et al. (2004), Tang et al. (2007b), and Woodland et al. (2004). 9 15.0 - Seitz et al. 2004 - Tang et al. 2007 - Rudnick and Ionov 2007 10.0 δ7Li 5.0 0.0 -5.0 -10.0 -15.0 olivine clinopyroxene Figure 1.1.4. δ7Li variability between coexisting mantle phases. In most cases, olivine has significantly higher δ7Li than coexisting clinopyroxene. This inter-mineral variability remains unexplained. Modified from Figure 7 from Seitz et al. (2004) to include data from Rudnick and Ionov (2007) and Tang et al. (2007b). 10 2. METHODS 2.1 Starting Materials Natural crystals of forsteritic olivine (Mg1.8Fe0.2SiO4), from San Carlos, Arizona, and diopside (CaMgSi2O6) from DeKalb, New York, were used in this study as starting materials for the growth of olivine and clinopyroxene in the piston cylinder experiments discussed below. Crystals were broken with a hardened steel crusher. Clean, inclusion free, alteration free pieces were separated by hand under a stereoscopic microscope and cleaned with ultrapure Milli-Q water. These pieces were then powdered with an agate mortar and pestle while submerged in ethanol. Starting compositions for all materials used in this study can be found in Tables 3.2.1 and 3.2.2. Natural quartz (SiO2) from Brazil and synthetic anorthite (CaAl2Si2O8) were added to stabilize clinopyroxene in the run products by saturating the fluid with Si, Ca and Al. The mass of powders added to each experiment and oxygen fugacity buffer can be found in Table 2.1.1. The materials used to buffer the oxygen fugacity within the experiments include synthetic fayalite (Fe2SiO4), magnetite (Fe3O4), quartz (SiO2), wüstite (FeO), Mo (metal), and MoO2. 2.2 Experimental Methods An end-loaded piston cylinder (PC) apparatus in the High-Pressure Laboratory at the Pennsylvania State University was used to generate upper mantle pressure and temperature conditions in which to run the equilibrium partitioning experiments (Figure 2.2.1). The piston cylinder is made up of two large hydraulic rams; the end load ram controls the confining pressure, and the master ram controls the pressure on the sample assembly. The master ram is 9 inches in diameter. By using the relationship P = F/A, where P is pressure, F is force, and A is the area over which the force is applied, we can generate high pressures by stepping the applied force of the 9” ram down to a smaller area – in this case a ¾” tungsten carbide (WC) piston. The pressures generated in the sample assembly can be as much as 4-5 GPa, depending on the piston diameter used. The confining pressure is applied by a 10” diameter pressure plate, which is a series of two concentric tapered steel rings around a tungsten carbide core. When the end load 11 ram applies pressure to the face of the plate, that pressure is transferred laterally to the WC core in order to provide confining support for the sample assembly. The purpose of the confining pressure provided by the end load ram and pressure plate is to prevent lateral failure of the sample cell when high pressures are applied on the piston and sample assembly by the master ram. The experimental assembly (Figure 2.2.2) and method are adapted from Ayers et al. (1992) and Trail et al. (2012). The capsule is designed so one Ag capsule is inverted and set on top of a second Ag capsule, separated by a thin Ag-Pd foil. Silver capsules are used in these experiments because Ag does not alloy with Fe, such that mantle-like Mg#s (Mg/(Mg+Fe)) can be maintained in the run product olivines (~Fo90). Retention of Fe in the system is imperative, as the purpose of these experiments is to determine the effect of Fe3+ on Li partitioning in olivine. The lower capsule contains the experimental starting materials. The experimental charge consists of powdered San Carlos Olivine (~7.5 mg), DeKalb Diopside (~5 mg), synthetic anorthite (~1 mg) and natural Brazilian quartz (~1 mg) plus an aqueous solution containing 200 ppm Li and 100 ppm Ba in milli-Q water (~50 mg). The lower capsule is then loosely capped with the Ag-Pd foil. In the upper capsule the solid state fO2 buffer is mixed with water. The surface tension of the water in the relatively small capsule prevents the powders and fluid from escaping the capsule when it is inverted and placed on top of the Ag-Pd foil. Upon initial pressurization, the foil seals to the rim of the upper and lower capsules, preventing any fluid escape when the experiment is brought to run conditions. The solid-state fO2 buffers used in these experiments, from highest to lowest fO2, are fayalite-magnetite-quartz, magnetite-wüstite, and molybdenum-molybdenum dioxide. At 900°C, these oxygen fugacity buffers correspond to values of logfO2 of -12.5, -15 and -17 respectively. The range of fO2 values prescribed by these buffers spans the range of oxygen fugacities expected in the Earth’s mantle. The Ag-Pd foil is used to physically separate the solid state buffer assemblage from the experimental charge. Water is present on both sides of the foil, which is permeable to H2. This allows H2 formed by the dissociation of H2O during the run to move between chambers. If the buffer side of the capsule contains more free oxygen than the olivine-diopside capsule at run conditions, H2O on the experiment side will dissociate and H2 will diffuse through the foil to the 12 buffer side. If the buffer side contains less free oxygen, H2O in the buffer capsule will dissociate and H2 will diffuse through the foil to the experiment side (Figure 2.2.3). As long as H2O is present on both sides of the foil and all components of the solid-state buffer remain at the end of the experiment, then both sides of the capsule were fixed at the fO2 given by the buffer for the run conditions of interest. The sample capsule is placed inside a NaCl-pyrex assembly with a straight-walled graphite furnace and loaded into the piston cylinder apparatus. The capsule is initially “cold-sealed” at ~0.8 GPa, then the experiment is brought up to run conditions. All experiments were held at 1.0 GPa and 900˚C for ~72 hours. Experiments were quenched by manually cutting the power to the furnace while maintaining the flow of water through the cooling plates and coil; temperatures dropped below 300˚C in ~30 seconds. Detailed experimental procedures are given in Appendix 1. Subsequent to quenching, the cleaned sample capsule was weighed to ensure that no water was lost during the experiment. Upon piercing the capsule, water should bubble out if all the initial water was retained during the run. After opening, the capsule was left under a heat lamp overnight to evaporate the remaining water, leaving behind only solid material. This solid material was collected, weighed, cleaned in an ultrasonic bath, and labeled and stored in glass vials. An aliquot was mounted in epoxy in a ¼” stainless steel annulus. The mounts were then polished using 600 grit silicon carbide sand paper and 1 micron diamond paste in preparation for analysis. 2.3 Analytical Methods 2.3.1 Scanning Electron Microscope Experimental run products are initially analyzed on a FEI Quanta 200 Environmental Scanning Electron Microscope (ESEM) at the Materials Characterization Laboratory at Penn State University. Backscattered electron (BSE) images are captured to be used as maps for later analyses on other analytical instruments. Because the relative brightness of grains in a BSE image is a function of the average atomic mass of the material, olivine and clinopyroxene grains can usually be distinguished from each other 13 visibly. Individual energy dispersive x-ray spectroscopy (EDS) spot analyses are performed on specific grains of interest to obtain a semi-quantitative understanding of the composition of the material. Raw spectra were collected and converted to concentrations using the Oxford Instruments INCA software package. This helped to ensure that both Mg-rich olivine and diopside were successfully grown in each experiment. 2.3.2 Electron Microprobe Analysis Fully quantitative major element compositions of the run products were determined using a Cameca SX-100 Electron Probe Micro-analyzer (EPMA) at Rensselaer Polytechnic Institute. Polished samples were carbon coated and analyzed for Al, Ca, Cr, Fe, Mg, Mn, Ni, Si, and Ti using a focused electron beam with a current of 20 nA and an accelerating voltage of 15 kV. Natural diopside, tephroite, and kyanite, as well as synthetic fayalite and forsterite were used as standards, and x-ray peak intensities were converted to concentrations using the CIT-ZAF correction method. 2.3.3 Laser Ablation Inductively Coupled Plasma Mass Spectrometry Trace element abundances were determined by laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) using a Thermo X-series II Quadrupole ICPMS connected to a New Wave UP-213 (213 nm) Neodymium-Yttrium-Aluminum-Garnet (Nd-YAG) laser ablation system at the Laboratory for Isotopes and Metals in the Environment (LIME) facility within the Penn State Institutes of Energy and the Environment. A 30 µm diameter spot size was used to ensure that the laser analysis sampled only the interior of the ~50 µm diameter crystals, and not any surface material. Crystals were ablated with an output energy of 5-6 J/cm2 and a frequency of 20 Hz. Ablated material was carried by a stream of helium (He) gas, subsequently mixed with argon (Ar) gas before entering the Ar-plasma torch of the ICP-MS. Batch analyses of 15 samples were bracketed by standard analyses to correct for internal mass drift within the ICP-MS. Standards used in these analyses were USGS basalt glass standards BCR-2g and GSD-1g. The isotopes measured were 7Li, 25Mg, 14 26 Mg, 43Ca, 47Ti, 60Ni, and 137Ba. Raw counts were converted to concentrations using LasyBoy V3.01, a data reduction spreadsheet created by Joel Sparks at Boston University. Calibration lines were constructed from selected bracketing standards for each set of analyses, and elemental concentrations were calculated from backgroundcorrected counts-per-second of each isotope using 25Mg as a normalizing mass. To ensure that no surface material or fluid inclusions were incorporated during the analysis, barium (Ba) was added to the starting fluid. Barium is incompatible in the silicate phases in the experiments, and should remain in the fluid. Therefore, any Ba detected in the laser ablation analysis would have to originate from quenched material adhered to the surface of the crystals or fluid inclusions within individual crystals. Any analysis with Ba above detection limits was excluded from the results. 2.3.4 X-ray Diffraction of fO2 Buffers X-ray diffraction (XRD) analyses were performed on the post- experiment fO2 buffers to ensure that all solid phases were present at the conclusion of the experiment. The XRD analyses were performed by Nichole Wonderling at the Materials Characterization Lab at Penn State University using a PANalytical X’Pert Pro MPD Theta-Theta Diffractometer system with Cu K-alpha radiation (1.54059 Å). Phase identification was performed using Materials Data Incorporated (MDI) Jade 2010 software via comparison to the International Center for Diffraction Data (ICDDPDF) database. 15 Table 2.1.1 TJ-52 TJ-63 TJ-67 TJ-68 Amounts of starting materials and run conditions for all experiments used in this study. Experimental starting materials (mg) San Carlos DeKalb fO2Buffer Olivine Diopside CaAl2Si2O8 Mo-MoO2 189.5 133.5 --MW 7.7 4.8 1.1 FMQ 8.5 6.9 1.2 Unbuffered 8.0 5.6 1.0 SiO2 --0.8 1.0 0.9 Li-Ba fluid 12.0 41.1 67.7 60.7 fO2 Buffers (mg) MoFeOMoO2 Fe3O4 56.6 ----11.4 --------- SiO2 ----4.1 --- Fe2SiO4 ----4.8 --- Fe3O4 ----4.5 --- Li-Ba fluid 18.4 38.9 39.5 --- --- = not used for this experiment Mo-MoO2 mixed together in 4:1 ratio by mass FeO-Fe3O4 mixed together in 3:1 ratio by mass All experiments run at 900°C and 1 GPa 16 a b Figure 2.2.1. a) Photograph of the piston cylinder apparatus used to conduct the experiments presented in this thesis. b) Schematic of the piston cylinder showing the arrangement of parts 17 Ag-Pd foil Ag Capsules Ceramic Annulus and Lid 1 mm 3.9 mm 6.4 mm 6.4 mm 9.8 mm 44 mm 9.9 mm 2 mm 11.5 mm Solid Ceramic 15 mm 10 mm 10 mm Graphite Sleeve and Plug 42 mm Center Drilled Ceramic 1.7 mm 9.8 mm 44 mm 11.6 mm 13.9 mm Pyrex Sleeve Base Plug 12.7 mm 16 mm 1.7 mm 15.8 mm 9.8 mm 47 mm Pyrophyllite Sleeve 12.7 mm 16 mm 19 mm 44 mm 14 mm 1.6 mm Thermocouple 18.8 mm Insulator NaCl Sleeve 58 mm Lead Foil Figure 2.2.2. Sample assembly used in the piston cylinder experiments. The figure is to scale. 18 1 Mineral powders and fluid sealed in capsule 2 Minerals grow incorporating Li; fO2 buffer reaction progresses 2Fe3O4 + 3SiO2 M F Q Buffer powders Ag-Pd foil Experiment powders O2 H2 O2 O2 Li 3Fe2SiO4 + O2 Magnetite and quartz react to form fayalite and free oxygen (or vice versa). H 2 diffuses through the Ag-Pd foil to create water from excess free O 2, balancing the activity of O 2 in both capsules. H2 Li Li Silver capsule Figure 2.2.3. Schematic diagram of the double capsule experimental setup using two Ag capsules separated by a 50/50 Ag-Pd foil. Schematic diagram shows the theoretical progression of the experiment through time. All experiments run at 900°C and 1 GPa. 19 3. RESULTS 3.1 Experimental Results Successful experiments were characterized by the growth of both olivine and clinopyroxene 50 m in diameter, as well as retention of the fluid added at the beginning of the experiment. Within the oxygen fugacity buffer chamber, all components, including water, must be present at the conclusion of the experiment to be considered a successfully buffered experiment. The two chambers needed to remain physically isolated (that is, the Ag-Pd foil must have remained intact) for the experiment to be considered a success. Crystals were mounted in epoxy and analyzed using scanning electron microscopy (SEM) to identify the phases that were grown in the experiment. X-ray energy dispersive spectroscopy (EDS) analysis confirmed that both olivine and clinopyroxene were present in each experiment. Backscattered electron (BSE) images were taken and used as maps to help identify phases and show relative crustal sizes during subsequent analysis on the LA-ICP-MS (Figures 3.1.1- 3.1.5). Typical experimental crystal sizes ranged from 30-100 µm, with olivine typically growing larger than clinopyroxene. 3.2 Run Product Compositions Major element compositions of the run products and starting materials are presented in Table 3.2.1. A representation of a typical LA-ICP-MS spectrum showing the collection of a background and signal is shown in Figure 3.2.1. The olivine run products contain ~10 wt% FeO* (all iron calculated as FeO), confirming that the Ag capsule successfully preserved Fe within the system. Trace element (Li and Ba) compositions of the run products are presented in Table 3.2.2. Barium content was used to monitor for unintentional sampling of fluid inclusions or grain boundaries during analysis, and data were filtered to include only those samples with Ba content below detection limits. The Ba content of clinopyroxenes in experiment TJ-52 was above detection limits in all analyses. In this case, clinopyroxenes with < 10 ppm Ba were 20 included. Lithium contents ranged from ~10 to ~27 ppm in olivine, and from ~6 to ~13 ppm in diopside. 3.3 Oxygen Fugacity Control A solid-state oxygen fugacity buffer consumes “reactants” to generate “products”, including O2. These are reversible reactions, and can run in either direction to maintain fO2 in the capsule, such that the designation of which components are the reactants and which are the products is arbitrary. In order to confirm that the fO2 in the capsule was maintained at the buffer value for the given run conditions (shown in Fig. 2.2.4), however, it is critical that all reactants and products, including H2O, be present at the end of the experiment. The solid-state buffer materials were analyzed via x-ray diffraction (XRD) at the Materials Characterization Laboratory at Penn State University to confirm that all solid phases were indeed present at the end of the experiment. The results from the x-ray analyses showed that all necessary phases were present. TJ-52 showed peaks for Mo-metal and MoO2. TJ-63 showed major peaks for magnetite and an amorphous phase. Throughout the quenching process, wüstite becomes unstable and likely reverted to magnetite and a poorly-crystalline Fe phase. TJ-67 showed peaks for fayalite, magnetite and quartz. X-ray spectra from these buffers are presented in Figures 3.3.13.3.3. 3.4 Partition Coefficients The equilibrium partition coefficient (DLiol/cpx) is calculated by dividing the average Li concentration in the olivine by the average Li concentration in the diopside in each experiment. This calculation yields DLiol/cpx = 1.7±0.4 for all experiments (Table 3.4.1 and Figure 3.4.1). Partition coefficients are plotted against log fO2 in Figure 3.4.1. 21 Table 3.2.1. Major element concentrations of starting materials and experimental run products. fO2 buffer Mo-MoO2 Olivine MW FMQ Unbuffered Mo-MoO2 Clinopyroxene MW FMQ Unbuffered TJ-52 n = 10 TJ-63 n=7 TJ-67 n = 10 TJ-68 n=7 San Carlos Olivine* n = 17 TJ-52 n=9 TJ-63 n=7 TJ-67 n = 10 TJ-68 n=9 DeKalb Diopside* n = 14 49.67 (1.30) 41.47 (0.29) 40.32 (0.98) 40.76 (1.02) 40.91 (0.59) 54.63 (1.02) 54.99 (0.88) 54.63 (1.10) 54.88 (0.77) 54.54 (0.90) TiO2 0.00 (0.00) 0.00 (0.01) 0.00 (0.02) 0.00 (0.02) 0.00 (0.02) 0.01 (0.02) 0.02 (0.02) 0.02 (0.01) 0.02 (0.02) 0.02 (0.02) Al2O3 0.01 (0.02) 0.01 (0.02) 0.02 (0.04) 0.02 (0.03) 0.02 (0.03) 0.54 (0.22) 0.57 (0.19) 0.58 (0.06) 0.61 (0.10) 0.81 (1.18) Cr2O3 0.02 (0.04) 0.02 (0.02) 0.02 (0.03) 0.02 (0.03) 0.02 (0.02) 0.01 (0.02) 0.00 (0.03) 0.02 (0.02) 0.00 (0.02) 0.00 (0.02) FeO 9.11 (1.74) 8.67 (1.25) 9.63 (2.36) 9.07 (1.62) 8.75 (0.94) 0.89 (0.12) 0.92 (0.16) 0.91 (0.08) 0.89 (0.09) 1.01 (0.29) MnO 0.13 (0.06) 0.13 (0.04) 0.12 (0.03) 0.12 (0.04) 0.12 (0.04) 0.06 (0.04) 0.05 (0.04) 0.06 (0.04) 0.06 (0.04) 0.06 (0.03) MgO 49.67 (1.30) 49.05 (1.94) 49.13 (2.27) 50.04 (1.11) 49.70 (1.35) 18.06 (0.54) 17.71 (0.64) 17.42 (0.68) 17.49 (0.40) 17.59 (0.47) NiO 0.37 (0.04) 0.33 (0.12) 0.37 (0.10) 0.36 (0.11) 0.37 (0.16) 0.01 (0.02) 0.02 (0.03) 0.01 (0.04) 0.00 (0.03) 0.01 (0.03) CaO 0.07 (0.04) 0.11 (0.05) 0.10 (0.05) 0.11 (0.02) 0.13 (0.06) 25.62 (0.54) 25.52 (0.30) 25.33 (0.22) 25.50 (0.20) 25.31 (0.63) Na2O 0.00 (0.02) 0.02 (0.02) 0.02 (0.05) 0.01 (0.03) 0.04 (0.06) 0.37 (0.16) 0.33 (0.31) 0.44 (0.10) 0.44 (0.05) 0.47 (0.17) K2O 0.00 (0.00) 0.03 (0.01) 0.03 (0.02) 0.02 (0.01) 0.03 (0.02) 0.01 (0.00) 0.02 (0.02) 0.02 (0.01) 0.02 (0.01) 0.03 (0.03) TOTAL 100.5 99.8 99.8 100.5 100.1 100.2 100.2 99.4 99.9 99.9 SiO2 n = number of analyses 2σ uncertainties are shown in parentheses Data collected using the Cameca SX-100 Electron Probe Micro Analyzer at Rensselaer Polytechnic Institute * Data collected using the JEOL JXA-8900 Electron Probe Micro Analyzer at the University of Maryland All experiments run at 900°C and 1 GPa 22 Table 3.2.2. Trace element concentrations of starting materials and experimental run products fO2 Buffer [Li] (ppm) 1σ [Ba] (ppm) 1σ n Olivine San Carlos Olivine* TJ-52 TJ-63 TJ-67 TJ-68 Mo-MoO2 MW FMQ Unbuffered 1.80 20.6 10.6 9.59 26.8 0.10 2.0 3.7 2.03 6.6 BDL BDL BDL BDL BDL BDL BDL BDL BDL BDL 3 9 5 3 3 Clinopyroxene DeKalb Diopside* 5.10 0.40 BDL BDL 8 TJ-52 Mo-MoO2 11.1 2.8 9.40 1.30 3 TJ-63 MW 6.90 1.04 BDL BDL 3 TJ-67 FMQ 5.79 1.36 BDL BDL 3 TJ-68 Unbuffered 13.2 3.8 BDL BDL 4 All experiments run at 900°C and 1 GPa Trace element data collected using the Thermo X-Series II Quadrupole ICP-MS with New Wave UP-213 laser ablation system at PSU * Data collected using the Thermo-Finnigan Element II ICP-MS with New Wave UP-213 laser ablation system at the University of Maryland n = number of analyses BDL = below detection limit 23 Table 3.4.1. Calculated Li partition coefficients (DLiol/cpx) Sample fO2 Buffer D(ol/cpx) 1σ TJ-52 Mo-MoO2 1.9 0.5 TJ-63 MW 1.5 0.6 TJ-67 FMQ 1.7 0.5 TJ-68 Unbuffered 2.0 0.7 AVERAGE 1.7 0.2 All experiments run at 900°C and 1 GPa 24 B A Clinopyroxene Olivine Figure 3.1.1. Backscattered electron SEM images of run products from experiment TJ-67. These crystals have not been mounted in epoxy and polished; rather they have been mounted on electrically conductive tape to image via SEM and show morphology of unpolished crystals. Image A shows multiple olivine crystals surrounded by quench material. Image B shows one large elongate clinopyroxene. 25 Clinopyroxene Olivine Figure 3.1.2. SEM backscattered electron image of run products from experiment TJ-52 buffered at Mo-MoO2. Sample was mounted in epoxy and polished with 600 grit siliconcarbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images. 26 Olivine Clinopyroxene Figure 3.1.3. SEM backscattered electron image of run products from experiment TJ-63 buffered at Magnetite-Wüstite. Sample was mounted in epoxy and polished with 600 grit silicon-carbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images. 27 Clinopyroxene Olivine Figure 3.1.4. SEM backscattered electron image of run products from experiment TJ-67 buffered at FMQ. Sample was mounted in epoxy and polished with 600 grit siliconcarbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images. 28 Olivine Clinopyroxene Figure 3.1.5. SEM backscattered electron image of run products from experiment TJ-68 which was unbuffered. Sample was mounted in epoxy and polished with 600 grit silicon-carbide sandpaper and 1µm diamond paste. Clinopyroxene is typically more elongate and brighter than olivine in backscattered electron images. 29 BCR-2g 1,000,000.0 Signal Counts/Second Background Li 100,000.0 Mg Mg 10,000.0 Ca Ti Ni 1,000.0 100.0 0.000 Ba 10.000 20.000 30.000 40.000 50.000 60.000 Time (sec) Figure 3.2.1. Typical time resolved laser ablation spectrum for standard BCR-2g used in sample-standard bracketing. Black vertical lines indicate the background differentiated from the analysis signal. LA-ICP-MS data was reduced with LasyBoy V3.01, a data reduction spreadsheet created by Joel Sparks at Boston University. 30 20000 18000 16000 TJ-52 14000 Mo Intensity (counts) 12000 MoO2 10000 8000 6000 4000 2000 0 15 25 35 45 55 65 2-Θ (deg) Figure 3.3.1. XRD analysis of the fO2 buffer from TJ-52 is shown in black. The red dotted line represents an XRD spectrum for Mo and the blue dashed line represents an XRD spectrum for Mo-MoO2. TJ-52 shows clear peaks that overlap well with both phases, suggesting that the experimental buffer was successful. 31 14000 12000 TJ-63 10000 Magnetite Intensity (counts) 8000 6000 4000 2000 0 15 25 35 45 55 65 2-Θ (deg) Figure 3.3.2. XRD analysis of the fO2 buffer from TJ-63 is shown in black. The red dashed line represents an XRD spectrum for magnetite. There is no confirmed wüstite in this buffer, but the broad increase in intensity from ~30-45° represents data from an amorphous phase with no defined crystal structure. This is being interpreted as the remnants of the wüstite present at run conditions which decomposed upon quenching as wüstite is not stable below 550°C at 1 atm. 32 10000 9000 8000 TJ-67 Fayalite Magnetite Quartz 7000 Intensity (counts) 6000 5000 4000 3000 2000 1000 0 15 25 35 45 55 65 2-Θ (deg) Figure 3.3.3. XRD analysis of the fO2 buffer from TJ-67 is shown in black. The red dashed line represents an XRD spectrum for fayalite. The blue dash-dot line represents an XRD spectrum for magnetite. The green dotted line represents an XRD spectrum for quartz. While the strongest peaks in TJ-67 overlap with magnetite, there are overlapping peaks that only correspond to quartz and fayalite as well, suggesting that all three phases are present at the conclusion of the experiment and the buffer was successful. 33 3.0 DLi(ol/cpx) 2.5 2.0 Buffered experiments from this study 1.5 Unuffered experiment from this study 1.0 Average 900°C experiment from Yakob et al., 2012 0.5 0.0 -18.0 -16.0 -14.0 -12.0 -10.0 log fO2 Figure 3.4.1. Experimental results showing the measured Li partition coefficient between olivine and clinopyroxene as a function of fO2. Open symbol represents equilibrium partition coefficient from 3-day 900°C experiment presented by Yakob et al., 2012. It is assumed that the oxygen fugacity of the unbuffered experiments is near Ni-NiO, which has a logfO2 of ~-12 at 900°C. 34 4. DISCUSSION 4.1 Attainment of Equilibrium in Experiments Previous piston-cylinder experiments conducted in the High-Pressure Laboratory at Penn State (Yakob et al. 2012) demonstrated that Li partitioning between olivine and clinopyroxene achieves equilibrium on the order of hours to days. A time series conducted at 900°C and 1.5 GPa shows that the measured Li partition coefficient between olivine and clinopyroxene remains constant (within error) for experiments between 1 and 7 days in duration (Figure 4.1.1). Modeling of the diffusion of Li between olivine and clinopyroxene supports this claim. Using the published diffusion coefficients for Li in olivine (“fast path”; Dohmen et al., 2010) and clinopyroxene (Coogan et al., 2005), Li partitioning between olivine and clinopyroxene approaches equilibrium within crystals 50µm and smaller in 24 hours or less at 900°C. The experiments in this study were held at 900˚C for ~3 days, allowing ample time for crystals to achieve equilibrium with respect to Li partitioning. 4.2 Oxygen Fugacity In gaseous systems, fugacity is defined as the difference between the chemical potential of a real gas at a given pressure and temperature and the chemical potential of an ideal gas at temperature and a standard pressure of 1 bar. fi = exp{[µ(T,P) - µ(T,1 bar, pure)]/RT} Where: fi = fugacity of gas i μ = chemical potential R = gas constant T = temperature Fugacity is the pressure an ideal gas would have at the same chemical potential as a real gas. At high temperatures and low pressures we can think of fugacity as the partial pressure of a gas in a system. Oxygen fugacity can be then defined as the chemical 35 potential or partial pressure of oxygen within a system. The concept of oxygen fugacity (fO2) can be applied to solid, liquid and gaseous systems when we want to understand the chemical potential of oxygen, which is important in understanding the oxidation environment of a system, oxidation-reduction processes, speciation of ions with multiple valence states, liquidus temperatures in melts and crystallization sequences. Oxygen fugacity within solid systems is most easily thought of as the chemical potential or activity of oxygen within a system. Typically in solid systems, oxygen fugacity is reported as it relates to various solid state fO2 buffers such as fayalitemagnetite-quartz (FMQ) or magnetite-wüstite (MW) (Fig. 4.2.1). These buffer systems work because O2 is a component on one side of a chemical reaction. The three oxygen fugacity buffers used in this study were: Fayalite + Oxygen ↔ Quartz + Magnetite 3 Fe2SiO4 + O2 ↔ 3 SiO2 + 2 Fe3O4 Wüstite + Oxygen ↔ Magnetite 3 FeO + ½ O2 ↔ Fe3O4 Molybdenum + Oxygen ↔ Molybdenum dioxide Mo + O2 ↔ MoO2 Presuming all solid phases are present in the system (e.g. fayalite, magnetite and quartz), the chemical activity of O2 can be calculated at a given temperature and pressure. Generally speaking, any metal-oxide pair can be a solid-state fO2 buffer so long as it is calibrated with respect to pressure and temperature. Relatively speaking, “high fO2” environments are oxidizing and “low fO2” environments are reducing. This is important in any environment where multivalent cations, such as iron, are present because the oxidation environment plays a major control on the speciation of the cation. As described in the Experimental Methods section, the piston cylinder experiments presented here used San Carlos olivine and DeKalb diopside as starting materials; these have ~10% and <1% FeO by weight, respectively. The oxidation environment in which olivine and clinopyroxene crystals are grown from these starting materials dictates the Fe3+/Fe within the system, and in turn the amount of Fe3+ incorporated into the minerals. 36 4.2.1 Oxygen Fugacity in the Mantle The oxygen fugacity of the upper mantle has been calibrated as a function of Fe activity in spinel/orthopyroxene/olivine xenolith assemblages. Results have shown that the ambient fO2 of the upper mantle ranges from 2 log units above to 3 log units below the FMQ solid state buffer (Mattioli and Wood, 1986; O’Neill and Wall, 1987). Lee et al. (2005) developed a trace element method of determining the oxygen fugacity of various mantle environments by calibrating the V/Sc system with fO2; V and Sc are both non-fluid-mobile and behave very similarly in volcanic processes, which mean they are generally unfractionated during the genesis of mid-ocean ridge basalts (MORBs) and arc lavas. However, the speciation of V is redox sensitive while that of Sc is not. Therefore, V/Sc ratios give an opportunity to calculate the oxygen fugacity of the mantle while removing outside effects due to melt formation or metasomatism. Application of this method to natural basalts suggests that the oxygen fugacity of arc lavas and MORBs range from 1.25 log units below FMQ to 0.5 log units above FMQ (Lee et al. 2005). More recently, Cottrell and Kelley (2011) used µ-XANES (micro x-ray absorption near edge spectroscopy) to calculate the Fe+3/ΣFe of MORB glasses worldwide, showing that the oxygen fugacity of primary MORB melts is very near the FMQ buffer, in close agreement with the results from Lee et al. (2005). Beneath Archean cratons, the fO2 determined by Mössbauer spectroscopy of garnet and spinel from peridotite xenoliths shows the ambient fO2 decreases with depth to nearly 5 log units below FMQ (Woodland and Koch 2003). More recent experimental constraints on garnet oxythermobarometry in xenoliths shows that even the most reducing deep lithosphere environments have fO2 ~3.5 log units below the FMQ buffer, at least an order of magnitude more oxidizing than previously thought (Stagno et al. 2013). As depth increases through the mantle towards the core-mantle boundaries, oxygen fugacity continues to decrease to values below the iron-wüstite buffer, where Fe and Ni metal can be stable (Frost and McCammon 2008). Many different analytical and experimental methods have been applied to assess the oxygen fugacity of the interior of the Earth. In summary, in the upper mantle near mid-ocean ridges and within volcanic arcs, the fO2 is near the FMQ buffer (Mattioli and Wood, 1986; O’Neill and Wall, 1987; Lee et al., 2005; Cottrell and Kelley, 2011). 37 Deeper within the mantle, the ambient fO2 decreases to values as low as 5 log units below FMQ (Woodland and Koch, 2003; Stagno et al. 2013). The range of solid-state buffers (plus the unbuffered experiment) used in this study was chosen to cover the full range of fO2 values expected in the upper mantle. 4.3 Li Partitioning as a Function of fO2 The partitioning of Li between mantle phases could be influenced by the oxygen fugacity of the surrounding environment. Being a monovalent cation, Li does not partition into the crystal structure of olivine by direct substitution; rather, Li needs a coupled substitution partner to preserve overall charge balance. One potential mechanism allowing Li incorporation includes a coupled substitution with a trivalent cation, where Li and the trivalent cation occupy the M1 and M2 sites typically reserved for divalent cations within the structure. Fe and Mg typically make up the cations that occupy these M1 and M2 sites within the olivine crystal structure. Fe has two valence states, and oxygen fugacity effectively controls the Fe3+/Fe2+ within a system. A more oxidizing fO2 environment would result in an increase in the relative abundance of Fe3+ within the system, effectively allowing for more coupled substitution with monovalent cations, such as Li, into the olivine structure. Considering olivine within the mantle contains ~10 wt. % FeO, and clinopyroxene contains ~3.5 wt. % FeO, more Fe3+ can be incorporated in olivine relative to clinopyroxene. This is to say that an increasingly oxidizing environment could allow for more Li to be incorporated into olivine rather than clinopyroxene. As such, oxygen fugacity has been proposed as a possible controlling factor in the partitioning of Li between mantle phases. Li has been shown to be significantly affected by coupled substitution in olivine. Experiments have shown that an increase in trivalent cations, specifically Al, Sc and Ga, can lead to an increase in Li content of the olivine that scales with the increase in trivalent cations (Grant and Wood, 2010). Caciagli et al. (2011) also show that an increase in trivalent cations can show an increase in partitioning between mantle minerals. An experiment with significantly increased Yb resulted in an ol/cpx partition 38 coefficient of 10, attributed to coupled substitution of Li+ with Yb3+. In a more complex manner, Li has been linked to P concentrations in olivine (Mallmann et al. 2009). The modal abundance of P is significantly higher than Li in this study, so a direct charge balancing mechanism does not seem plausible. Mallmann et al. (2009) suggest that Li is being incorporated into vacancies within the olivine structure created by the distortions caused by P incorporation. Caciagli (2010) showed that varying oxygen fugacity may have an effect on Li partitioning between mantle phases. Two experiments were run with the intention of varying the oxygen fugacity of the system, one at Re-ReO2 (highly oxidizing, 5 log units > FMQ) and one at Mo-MoO2 (highly reducing, 5 log units < FMQ). The results from these experiments are not entirely clear, as no olivine was grown in the oxidizing experiment and the olivine/clinopyroxene partition coefficient determined in the reducing experiment was 0.7±0.4, slightly lower than in unbuffered equilibrium experiments (Caciagli, 2010) (Figure 4.4.1). This suggests that oxygen fugacity could possibly play a role in equilibrium partitioning of Li between mantle minerals, but more data at a range of oxygen fugacities is required. The results from our experiments, however, show no apparent relationship between oxygen fugacity and Li partitioning between olivine and clinopyroxene. We have studied a wide range of oxygen fugacities spanning nearly all fugacities relevant to the Earth’s mantle. The measured partition coefficient between olivine and clinopyroxene in our run products show that under all fO2 conditions tested, the equilibrium partition coefficient is 1.7±0.4 (Table 3.2.3). This, along with the results from Yakob et al. (2012), suggests that closed-system redistribution of Li between mantle minerals cannot explain the compositional and isotopic variability seen in mantle xenoliths from around the world. 39 4.4 Li Variability in Mantle Minerals The apparent disequilibrium distribution of Li and its isotopes is observed within some mantle peridotites collected from localities worldwide. As shown in Figure 1.1.3, the observed Li partition coefficients between olivine and clinopyroxene at some localities deviate from experimentally derived equilibrium partition coefficients determined from a variety of studies (Figure 4.4.1; Yakob et al., 2012; Ottolini et al., 2009; Caciagli et al., 2011; Brenan et al., 1998; Blundy and Dalton, 2000). In addition, the isotopic distribution of Li in many of the “non-equilibrium” samples shows significant variation. One trend, most evident in xenoliths from Tok, Siberia (Rudnick and Ionov, 2007) and Hannuoba, China (Tang et al., 2007b), shows increasing isotopic fractionation with a decreasing partition coefficient. In other words, the clinopyroxenes in these xenoliths are unusually enriched in isotopically light Li. Elevated concentrations of isotopically light Li have also been observed in clinopyroxenes hosted in mantle xenoliths from Kyushu, Japan, although the olivines in these samples were not analyzed (Nishio et al., 2002). 4.4.1 Equilibrium Isotope Fractionation The isotopic variability shown in some mantle xenoliths could be explained if the equilibrium isotopic fractionation factor between olivine and clinopyroxene (ol/di) at high temperature is greater than 1, such that 7Li is preferentially taken up by olivine relative to clinopyroxene, resulting in a higher δ7Li for olivine. Seitz et al. (2004) observed an apparent temperature dependence on isotopic fractionation as mantle xenolith samples from Siberia, Germany and Austria, with the highest equilibration temperatures (~1150°C) correlate with the lowest Δ7Liol-cpx (Δ7Liol-cpx = 1) (Δ7Liol-cpx = δ7Liol-δ7Licpx), while samples with lower equilibration temperatures (~850-900°C) show increased isotopic fractionation (Δ7Liol-cpx = 3.5-4). The idea that equilibrium isotopic fractionation takes place at high temperatures goes against thermodynamic convention, which suggests isotopic effects are minimized with increasing temperature. If the isotopic fractionation observed between mantle 40 minerals is an equilibrium effect related to temperature, at high temperatures (>200 °C) there should be a linear relationship between 1000*ln(α) vs. 106/T2 [where α = 7 Li/6Liol/7Li/6Licpx] (Northrop and Clayton 1966). A compilation of Li concentrations and isotopic ratios measured in bulk mineral separates or individual crystals from mantle xenoliths worldwide is presented in Table 4.4.1. The dataset reveals no apparent relationship between temperature and isotopic fractionation when 1000*ln(α) is plotted against 106/T2 (Figure 4.4.2). This suggests that the Li isotopic variability observed in mantle xenoliths is not representative of equilibrium fractionation. Furthermore, Tomascak et al. (1999) demonstrated a lack of measurable Li isotopic fractionation as olivine, clinopyroxene, and plagioclase sequentially crystallized out of a basaltic melt at temperatures between 1050 and >1200 ˚C at Kilauea Iki lava lake. Throughout the process of fractional crystallization, the isotopic composition of the residual melt remains unchanged. This suggests that there is no inherent isotopic fractionation of Li in silicate systems including olivine and diopside and therefore the isotopic fractionation factor is indeed close to 1 at temperatures above 1000°C. 4.4.2 Open vs. closed system redistribution Redistribution of Li between coexisting phases could result in the pattern displayed by mantle xenoliths that deviate from equilibrium behavior (Dol/cpx <1, αol/cpx>1). Diffusive re-distribution of Li between mantle minerals has interesting implications for isotopic fractionation. As has been shown by Richter et al. (2003), 6Li diffuses faster than 7Li, such that diffusion could cause dynamic isotopic fractionation on geologically short timescales by as much as 25‰. The fractionation could then be locked into the minerals as they continue to cool beyond the Li closure temperature. A driving force is required, but if Li is mobilized and allowed to move from olivine into clinopyroxene, the partition coefficient would decrease and the isotopic fractionation factor would increase. One potential driver for closed system Li redistribution is a temperature dependence of the Li partition coefficient between olivine and clinopyroxene. A xenolith or peridotite massif cools as it is exhumed. This change in 41 temperature might influence the partition coefficient such that Li becomes more compatible in clinopyroxene relative to olivine as temperature decreases, resulting in Li moving from the olivine to the clinopyroxene (thus decreasing the Dol/cpx) via diffusion. By analogy, Coogan et al. (2005) showed that the Li partition coefficient between clinopyroxene and plagioclase is dependent on temperature, and used this temperature dependence to develop a “geo-speedometer” to determine cooling rates for mid-ocean ridge basalts. Ionov and Seitz (2008) invoked this hypothesis to propose that eruption style, and therefore cooling rate, could be a mechanism for controlling Li distribution between mantle minerals. They observed greater Li isotopic fractionation in presumably rapidly cooled xenoliths sampled from explosive eruptions (tuff and scoria deposits) relative to xenoliths sampled from lava flows, where cooling rates would be substantially slower. The enhanced cooling rates in the xenoliths from explosive eruptions would allow the Li isotope ratios to be locked into the cooling crystals near the peak of fractionation, while the slower cooling rates would allow the minerals to approach equilibrium with one another prior to dropping below the closure temperature. Kaliwoda et al. (2008) invoked the same hypothesis, however in the opposite sense, to explain Li zoning in individual crystals in mantle xenoliths from Saudi Arabia. They observed a rim-ward increase in Li content of the olivines, while the rims of clinopyroxenes decreased their Li concentrations relative to the cores. They concluded that this was the result of temperature dependent partition coefficients changing such that Li moves from clinopyroxene to the olivine upon cooling. Yakob et al. (2012) experimentally tested whether the Li partition coefficient between olivine and clinopyroxene is temperature dependent. Piston cylinder experiments at 1.5 GPa and temperatures ranging from 700-1100°C all yielded DLiol/di = 2.0 0.2, suggesting that closed-system re-distribution of Li in response to changing temperature is not responsible for the observed isotopic fractionation between olivine and clinopyroxene in natural mantle xenoliths. Apparent DLiol/di measured from mineral separates of mantle xenoliths worldwide also shows no correlation with temperature (Figure 4.4.3). The experiments of Yakob et al. (2012), combined with the results from 42 this study, have failed to identify a suitable driving force for closed system redistribution of Li capable of generating the non-equilibrium Li signature seen in mantle xenoliths. This implies that the mechanism controlling the non-equilibrium Li signature seen in mantle xenoliths is an open system mechanism. Xenoliths can gain isotopically distinct Li by open-system interaction with a metasomatic fluid or melt prior to or during their rise to the surface. This fluid or melt may interact with each mantle mineral differently, depending on the relative diffusion rates of lithium in each phase. Rudnick and Ionov (2007) propose a few different mechanisms for interaction with an infiltrating fluid or melt. For example, if Li diffuses from the fluid or melt into the clinopyroxene faster than it diffuses into the olivine, on a short timescale such that equilibrium is not attained, the concentration of Li in clinopyroxene would increase relative to that in olivine. Likewise, the isotopic composition of the clinopyroxene would be relatively light due to faster diffusion of the lighter 6Li. Alternatively, if Li diffuses into olivine faster than clinopyroxene, the olivine could come to equilibrium with the fluid or melt, while the isotopically lighter clinopyroxene still captures kinetically induced isotopic fractionation. The final, and preferred, proposed mechanism of Rudnick and Ionov (2007), is that Li diffuses into clinopyroxene faster than olivine, with the added condition that the fluid:rock ratio is quite small. This would result in early uptake of isotopically light Li by clinopyroxene, which would draw down the 6Li in the fluid such that the remaining Li would become isotopically heavy. As time progresses, isotopically heavy Li would migrate into the olivines, allowing for greater diversity in the Li isotope composition of minerals coexisting within the same xenolith sample. This hypothesis has yet to be experimentally tested, although independent Li diffusion studies by Coogan et al. (2005) and Dohmen et al. (2010) suggest that Li diffusion could be up to three orders of magnitude faster in clinopyroxene relative to natural mantle olivine. 43 Table 4.4.1 Compilation of Li data from olivine and clinopyroxene mineral separates from xenoliths worldwide. Data from Aulbach and Rudnick (2009)1, Aulbach et al. (2008)2, Ionov and Seitz (2008)3, Jeffcoate et al. (2007)4, Kaliwoda et al. (2008)5, Magna et al. (2006, 2008)6,7, Ottolini et al. (2004)8, Rudnick and Ionov (2007)9, Seitz and Woodland (2000)10, Seitz et al. (2004)11, Tang et al. (2007b)12, and Woodland et al. (2004)13. --- = values not reported Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene 89-6611 Lashaine, Tanzania 0.7 --- --- --- --- --- --- 1090 89-6641 Lashaine, Tanzania --- --- 1.4 -2.5 --- --- --- 1230 89-669 1 Lashaine, Tanzania 2.7 --- 3.6 -3.2 0.75 --- --- 1040 89-6711 Lashaine, Tanzania 1.7 13.9 --- --- --- --- --- --- 89-6721 Lashaine, Tanzania 1.1 4.2 1.6 0.4 0.69 3.8 1.004 1110 89-6741 Lashaine, Tanzania 1.5 5 2.2 --- --- --- --- 1250 89-6751 Lashaine, Tanzania 0.6 --- --- --- --- --- --- 1240 89-6801 Lashaine, Tanzania 1.5 2.3 0.9 -1.5 1.67 3.8 1.004 1150 89-719 1 Lashaine, Tanzania 1.1 2.8 1.2 -3.9 0.92 6.7 1.007 1150 89-7721 Olmani, Tanzania 4.1 4.1 1.5 -2 2.73 6.1 1.006 950 89-7731 Olmani, Tanzania 0.7 3 --- --- --- --- --- 1080 89-7741 Olmani, Tanzania 3 3.3 --- --- --- --- --- --- 89-7761 Olmani, Tanzania 3.7 6.2 1.6 -0.4 2.31 6.6 1.007 1120 89-777 1 Olmani, Tanzania 3.5 5.7 1.9 -1.4 1.84 7.1 1.007 1120 89-7781 Olmani, Tanzania 2.6 9.3 --- --- --- --- --- --- 44 Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene LB-29 2 Labait, Tanzania 1.9 3.4 --- --- --- --- --- --- LB-312 Labait, Tanzania 1.8 3.3 0.9 1.5 2.00 1.8 1.002 --- KAT-172 Labait, Tanzania 3.4 4.7 2.2 -0.3 1.55 5 1.005 --- LB-42 Labait, Tanzania 2.3 4.7 1.2 -2.6 1.92 7.3 1.007 --- LB-452 Labait, Tanzania 2.4 4.7 2 0.1 1.20 4.6 1.005 --- LB-6 Labait, Tanzania 1.9 3.7 --- --- --- --- --- --- LB-172 Labait, Tanzania 2.7 4 --- --- --- --- --- --- LB-212 Labait, Tanzania 2.2 2.5 3.8 -4.9 0.58 7.4 1.007 --- LB-462 Labait, Tanzania 4.4 5.2 3 -6.7 1.47 11.9 1.012 --- LB-512 Labait, Tanzania 4.8 6.6 --- --- --- --- --- --- LB-592 2 Labait, Tanzania 3.2 6.6 --- --- --- --- --- --- LB-2 2 Labait, Tanzania 5 10.1 --- --- --- --- --- --- Av-13 Avacha, Siberia 1.35 --- --- --- --- --- --- 919 Av-23 Avacha, Siberia 1.59 3.25 --- --- --- --- --- 906 Av-33 Avacha, Siberia 1.42 -2.1 --- --- --- --- --- 946 Av-43 Avacha, Siberia 0.91 2.95 --- --- --- --- --- 920 Av-53 Avacha, Siberia 1.17 --- --- --- --- --- --- --- Av-6 3 Avacha, Siberia 1.75 4.71 --- --- --- --- --- 993 Av-73 Avacha, Siberia 1.26 --- --- --- --- --- --- --- 45 Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene Av-8 3 Avacha, Siberia 1.54 2.24 --- --- --- --- --- 978 Av-93 Avacha, Siberia 1.32 --- --- --- --- --- --- --- Av-103 Avacha, Siberia 1.21 --- --- --- --- --- --- --- Av-113 Avacha, Siberia 1.81 0.3 --- --- --- --- --- 930 Av-123 Avacha, Siberia 1.24 --- --- --- --- --- --- --- Av-13 3 Avacha, Siberia 1.4 --- --- --- --- --- --- --- Av-143 Avacha, Siberia 1.28 --- --- --- --- --- --- --- Av-153 Avacha, Siberia 1.37 1.53 --- --- --- --- --- 960 Av-163 Avacha, Siberia 1.24 0.86 --- --- --- --- --- 989 Av-173 Avacha, Siberia 1.33 --- --- --- --- --- --- --- 621-163 Vitim, Siberia 1.15 6.3 5.08 -17.9 0.23 24.2 1.025 1176 313-105 4 Vitim, Siberia 2.32 4.7 1.12 1.1 2.07 3.6 1.004 1034 313-1054 Vitim, Siberia 2.4 4.9 1.18 1 2.03 3.9 1.004 1034 313-1024 Vitim, Siberia 1.97 3.7 1.15 3.5 1.71 0.2 1.000 1053 314-564 Vitim, Siberia 1.87 3 1.3 6.6 1.44 -3.6 0.996 889 Mo-Z-14 Tariat, Mongolia 1.93 3 1.11 2.6 1.74 0.4 1.000 890 4230-164 Tariat, Mongolia 1.75 3.1 1.15 4.7 1.52 -1.6 0.998 867 8505-2 Dariganga, Mongolia 1.4 3 1.25 1.6 1.12 1.4 1.001 914 SC4 San Carlos, AZ 2.38 5.2 1.48 -7.8 1.61 13 1.013 1002 4 46 Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene 4 San Carlos, AZ 2.31 4.8 1.5 -8.7 1.54 13.5 1.014 1002 SC4 San Carlos, AZ 2.29 5 1.48 -8.1 1.55 13.1 1.013 1002 Th294 Icelandic basalt 1.25 3.3 --- --- --- --- --- --- AH4 Hawaiian basalt 1.42 3.5 --- --- --- --- --- --- 97KC06 (1960) 4 Hawaiian basalt 1.18 6.6 0.85 1.6 1.39 5 1.005 --- 4 Hawaiian basalt 1.28 6.5 0.78 1.8 1.64 4.7 1.005 --- SA84-32 cores5 Harrat Uwayrid, Saudi Arabia 1.382 --- 0.917 --- 1.51 --- --- 971 SA84-32 rims5 Harrat Uwayrid, Saudi Arabia 1.47 --- 0.666 --- 2.21 --- --- 988 SA84-123 cores5 Harrat Uwayrid, Saudi Arabia 1.296 --- 1.179 --- 1.10 --- --- 805 SA84-123 rims5 Harrat Uwayrid, Saudi Arabia 1.179 --- 0.62 --- 1.90 --- --- 844 SA84-124b cores5 Harrat Uwayrid, Saudi Arabia 1.481 --- 0.581 --- 2.55 --- --- 884 5 Harrat Uwayrid, Saudi Arabia 1.6 --- 0.522 --- 3.07 --- --- 884 SA84-128/1 cores5 Harrat Uwayrid, Saudi Arabia 1.563 --- 0.98 --- 1.59 --- --- 998 SA84-128/1 rims5 Harrat Uwayrid, Saudi Arabia 1.563 --- 0.853 --- 1.83 --- --- 1019 SA84-128/2 cores5 Harrat Uwayrid, Saudi Arabia 1.625 --- 1.003 --- 1.62 --- --- 1051 SA84-128/2 rims5 Harrat Uwayrid, Saudi Arabia 2.081 --- 0.833 --- 2.50 --- --- 1061 SA84-38 cores5 Harrat Uwayrid, Saudi Arabia 1.577 --- 1.393 --- 1.13 --- --- 874 5 Harrat Uwayrid, Saudi Arabia 1.569 --- 0.544 --- 2.88 --- --- 852 SA84-97 cores5 Harrat Uwayrid, Saudi Arabia 2.163 --- 1.19 --- 1.82 --- --- 799 SC 97KC05 (1955) SA84-124b rims SA84-38 rims 47 Sample Location Li (ppm) δ7Li Olivine SA84-97 rims 5 Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene Harrat Uwayrid, Saudi Arabia 2.281 --- 0.718 --- 3.18 --- --- 809 SA84-128/3 cores5 Harrat Uwayrid, Saudi Arabia 2.968 --- 1.936 --- 1.53 --- --- 987 SA84-128/3 rims5 Harrat Uwayrid, Saudi Arabia 3.373 --- 1.843 --- 1.83 --- --- 1028 SA84-91 cores5 Harrat Uwayrid, Saudi Arabia 3.883 --- 1.766 --- 2.20 --- --- 1039 SA84-91 rims5 Harrat Uwayrid, Saudi Arabia 3.823 --- 1.619 --- 2.36 --- --- 1060 5 Harrat Uwayrid, Saudi Arabia 1.923 --- 1.134 --- 1.70 --- --- 1122 SA84-166 rims5 Harrat Uwayrid, Saudi Arabia 2.029 --- 1.089 --- 1.86 --- --- 1118 SA84-50 cores5 Harrat Uwayrid, Saudi Arabia 1.934 --- 1.016 --- 1.90 --- --- --- SA84-50 rims5 Harrat Uwayrid, Saudi Arabia 1.577 --- 0.873 --- 1.81 --- --- --- SA84-63 cores5 Harrat Uwayrid, Saudi Arabia 3.865 --- 2.085 --- 1.85 --- --- 900 SA84-63 rims5 Harrat Uwayrid, Saudi Arabia 4.067 --- 2.451 --- 1.66 --- --- 981 San Carlos, AZ 1.6 3.64 0.52 3.32 3.08 0.32 1.000 --- 314-586 Vitim, Siberia 1.9 3.76 1.4 5.21 1.36 -1.45 0.999 --- 8520-096 Atsagin-Dush, Mongolia 1.5 3.55 0.82 5.96 1.83 -2.41 0.998 --- MPH 79/16 Tariat, Mongolia 2 3.75 1.4 2.57 1.43 1.18 1.001 --- 96-26 Kilbourne Hole, NM 1.7 3.56 --- --- --- --- --- --- Atsagin-Dush, Mongolia 1.74 5.21 1.11 4.34 1.57 0.87 1.001 --- BM90-6 Balmuccia, Italy 1.16 --- 1.18 --- 0.98 --- --- --- BM90-12L8 Balmuccia, Italy 1.1 --- 2.36 --- 0.47 --- --- --- SA84-166 cores SC 6 8250-157 8 48 Sample Location Li (ppm) δ7Li Olivine 8 Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene LH12 Lherz, France 1.75 --- 1.77 --- 0.99 --- --- --- Z20688 Zabargad Island 2.09 --- 3.9 --- 0.54 --- --- --- Z20708 Zabargad Island 2.73 --- 2.1 --- 1.30 --- --- --- Z20998 Zabargad Island 2 --- 1.75 --- 1.14 --- --- --- BR118 Massif Central, France 1.92 --- 1.13 --- 1.70 --- --- --- 8 Massif Central, France 1.67 --- 0.99 --- 1.69 --- --- --- RP688 Massif Central, France 1.87 --- 1.13 --- 1.65 --- --- --- RP708 Massif Central, France 1.88 --- 0.95 --- 1.98 --- --- --- Z78 Massif Central, France 2.26 --- 0.62 --- 3.65 --- --- --- Bt398 Massif Central, France 2.91 --- 1.45 --- 2.01 --- --- --- Szt10638 Carpatho-Pannonian Basin 1.2 --- 0.54 --- 2.22 --- --- --- 8 Carpatho-Pannonian Basin 1.55 --- 0.84 --- 1.85 --- --- --- Ib/kl8 Dreiser Weiher, Germany 1.34 --- 0.75 --- 1.79 --- --- --- D588 Dreiser Weiher, Germany 1.42 --- 0.82 --- 1.73 --- --- --- D508 Dreiser Weiher, Germany 1.36 --- 0.67 --- 2.03 --- --- --- 6--19 Tok, Siberia 1.3 8.1 3.3 -6.9 0.39 15 1.015 1010 6--29 Tok, Siberia 2 --- 2.3 --- 0.87 --- --- 1001 6--3 9 Tok, Siberia 1.5 11.8 6.7 -9.3 0.22 21.1 1.021 976 7--59 Tok, Siberia 1.6 11.9 4.8 -9.8 0.33 21.7 1.022 985 BR12 Bol082 49 Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene 8--5 9 Tok, Siberia 2.1 0.8 4.1 -12.9 0.51 13.7 1.014 1004 8--69 Tok, Siberia 1.5 8.1 4.3 -10.7 0.35 18.8 1.019 985 8--399 Tok, Siberia 1.8 8.3 3.2 -14.6 0.56 22.9 1.023 964 1--29 Tok, Siberia 4.2 0.6 4.4 -3 0.95 3.6 1.004 910 1--39 Tok, Siberia 4.6 --- 9.3 --- 0.49 --- --- 907 1--13 Tok, Siberia 3.1 --- 8.2 --- 0.38 --- --- 933 2--69 Tok, Siberia 3.6 --- 5.7 --- 0.63 --- --- 980 2--99 Tok, Siberia 3.1 --- 2.4 --- 1.29 --- --- 874 3--49 Tok, Siberia 4.3 --- 5.7 --- 0.75 --- --- 910 3--199 Tok, Siberia 2.6 --- 0.9 --- 2.89 --- --- 931 5--39 Tok, Siberia 2.9 --- 2.2 --- 1.32 --- --- 907 6--0 9 Tok, Siberia 2.3 --- --- --- --- --- --- 890 7--19 Tok, Siberia 3.7 --- 4.2 --- 0.88 --- --- 985 8--19 Tok, Siberia 2.3 --- 10 --- 0.23 --- --- 1005 8--29 Tok, Siberia 3 --- 9 --- 0.33 --- --- 976 8--39 Tok, Siberia 3.8 --- 1.7 --- 2.24 --- --- 956 8--79 Tok, Siberia 4.3 --- 6 --- 0.72 --- --- 968 8--8 9 Tok, Siberia 3.2 --- 8.6 --- 0.37 --- --- 955 8--119 Tok, Siberia 4.4 --- 6.7 --- 0.66 --- --- 957 9 50 Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene 8--31 9 Tok, Siberia 3 2.3 8.3 -7.3 0.36 9.6 1.010 887 8--509 Tok, Siberia 2.3 -1.7 2.3 -9.1 1.00 7.4 1.007 992 10--29 Tok, Siberia 4.3 --- 1.3 --- 3.31 --- --- 914 10--49 Tok, Siberia 2.8 --- 1.6 --- 1.75 --- --- 926 10--89 Tok, Siberia 3.2 --- 2.7 --- 1.19 --- --- 950 10--16 9 Tok, Siberia 1.1 3.3 1.6 -1.7 0.69 5 1.005 957 10--179 Tok, Siberia 1.5 2.1 1.4 -1.9 1.07 4 1.004 1011 10--199 Tok, Siberia 1.3 2.8 1.2 -2.1 1.08 4.9 1.005 951 2--19 Tok, Siberia 3.6 --- 11.5 --- 0.31 --- --- --- 2--29 Tok, Siberia 2.4 --- 3 --- 0.80 --- --- 909 2--39 Tok, Siberia 4 --- 3.7 --- 1.08 --- --- 908 2--4 9 Tok, Siberia 3.7 --- 8.4 --- 0.44 --- --- --- 2--109 Tok, Siberia 3.6 --- 8.7 --- 0.41 --- --- 1024 3--29 Tok, Siberia 3.2 2 1.6 -0.8 2.00 2.8 1.003 949 3--229 Tok, Siberia 3.4 --- 2 --- 1.70 --- --- --- 8--109 Tok, Siberia 3 --- 7.3 --- 0.41 --- --- 984 10--19 Tok, Siberia 2.7 --- 1.7 --- 1.59 --- --- 964 10--3 9 Tok, Siberia 3.6 --- 1.6 --- 2.25 --- --- 982 10--119 Tok, Siberia 5.5 --- 7.4 --- 0.74 --- --- 920 51 Sample Location Li (ppm) δ7Li Olivine 9 Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene 13--8 Sikhote-Alin 4.8 4.7 6.6 -6.3 0.73 11 1.011 905 8802-19 Sikhote-Alin 1.5 4 1.2 3 1.25 1 1.001 987 Ke 1906/110 Chyulu Hills, Kenya --- --- 18.67 --- --- --- --- 778 Ke 771/310 Chyulu Hills, Kenya --- --- 1.18 --- --- --- --- 878 MC 21/210 Cantal, France 1.18 --- 3.62 --- 0.33 --- --- 852 Chyulu Hills, Kenya 2.13 --- 0.96 --- 2.22 --- --- 912 MC 34/110 Deves, France 2.44 --- 1.36 --- 1.79 --- --- 940 Ke 50310 Chyulu Hills, Kenya --- --- 2.41 --- --- --- --- 974 Ke 1921/510 Chyulu Hills, Kenya --- --- 1.56 --- --- --- --- 977 Ke 1916/310 Chyulu Hills, Kenya --- --- 1.59 --- --- --- --- 1014 Vi 313-1010 Vitim, Siberia 1.28 --- 0.75 --- 1.71 --- --- 1050 San Carlos, AZ 1.58 --- 1 --- 1.58 --- --- 1052 MC 45/310 Vivarais, France 1.44 --- 0.92 --- 1.57 --- --- 1066 Pa110 Pali Aike, Chile 1.27 --- 0.74 --- 1.72 --- --- 1120 Mo 2210 Tariat, Mongolia 1.42 --- 0.79 --- 1.80 --- --- 1137 D 4510 Dreiser Weiher, Germany 1.23 --- 0.71 --- 1.73 --- --- 1147 Causses, France 1.26 --- 0.74 --- 1.70 --- --- 1180 Chyulu Hills, Kenya 1.1 --- 0.73 --- 1.51 --- --- 1346 Marsabit, Kenya 1.34 --- 0.71 --- 1.89 --- --- 1367 Ke 1928/3 SC 1 10 10 MC 49/110 Ke 514/1 10 Ke 604/610 52 Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene 10 Chyulu Hills, Kenya 1.61 --- 0.86 --- 1.87 --- --- 1363 Ke 1927/210 Chyulu Hills, Kenya 1.22 --- 0.61 --- 2.00 --- --- 1364 SC11 San Carlos, AZ 1.7 3.1 --- --- --- --- --- --- SC11 San Carlos, AZ 1.8 2.2 --- --- --- --- --- --- San Carlos, AZ 1.7 3.4 1.9 -0.8 0.89 4.2 1.004 1052 Vitim, Siberia 1.6 4.2 1 2.5 1.60 1.7 1.002 1050 Ka 16711 Kapfenstein, Austria 1.8 3.2 1.1 1.5 1.64 1.7 1.002 1040 Ia/17111 Dreiser Weiher, Germany 1.3 3.1 0.8 0.4 1.63 2.7 1.003 842 Ia/21111 Dreiser Weiher, Germany 1.8 1.4 1 -2.4 1.80 3.8 1.004 969 Ib/311 Dreiser Weiher, Germany 1.4 3 0.8 1.3 1.75 1.7 1.002 1185 Ib/811 Ke 1927/8 SC-111 11 Vi 313-10 Dreiser Weiher, Germany 1.6 4.5 1 2.8 1.60 1.7 1.002 1190 11 Dreiser Weiher, Germany 1.2 3.9 0.8 3 1.50 0.9 1.001 1152 Ib/K111 Dreiser Weiher, Germany 1 2.4 0.9 2.3 1.11 0.1 1.000 1121 Ib/5811 Dreiser Weiher, Germany 1.3 2.1 --- --- --- --- --- --- D112 Damaping, China 1.8 3.34 3.01 -8.25 0.60 11.59 1.012 --- D212 Damaping, China 1.66 6.39 3.49 -5.59 0.48 11.98 1.012 --- D312 Damaping, China 1.19 5.37 3.59 -5.32 0.33 10.69 1.011 --- D4 12 Damaping, China 1.51 5.35 2.41 -3.25 0.63 8.6 1.009 --- D512 Damaping, China 1.37 5.2 2.72 -7.54 0.50 12.74 1.013 --- Ib/24 53 Sample Location Li (ppm) δ7Li Olivine Li (ppm) δ7Li DLiol/cpx Δ7Liol-cpx αol/cpx T (°C) Clinopyroxene 70972 13 Victoria, Australia 4.75 --- 1.55 --- 3.06 --- --- --- 7098713 Victoria, Australia 2.87 --- 1.4 --- 2.05 --- --- --- 7100013 Victoria, Australia 3.13 --- 1.13 --- 2.77 --- --- --- 7100413 Victoria, Australia 2.67 --- 2.42 --- 1.10 --- --- --- 7100613 Victoria, Australia 4.05 --- 0.63 --- 6.43 --- --- --- 13 SH 45 Victoria, Australia 2.63 --- 1.28 --- 2.05 --- --- --- 7698913 Victoria, Australia 1.75 --- 0.89 --- 1.97 --- --- --- 7096513 Victoria, Australia 1.98 --- 0.59 --- 3.36 --- --- --- 7100313 Victoria, Australia 2.29 --- 0.98 --- 2.34 --- --- --- 7100813 Victoria, Australia 2.92 --- 0.71 --- 4.11 --- --- --- 7699313 Victoria, Australia 3.42 --- 1.42 --- 2.41 --- --- --- 76995 13 Victoria, Australia 1.34 --- 1.88 --- 0.71 --- --- --- 7699713 Victoria, Australia 1.73 --- 0.52 --- 3.33 --- --- --- SH 3513 Victoria, Australia 2.11 --- 1.27 --- 1.66 --- --- --- 7100113 Victoria, Australia 5.32 --- 1.13 --- 4.71 --- --- --- 7102313 Victoria, Australia 2.22 --- 1.35 --- 1.64 --- --- --- 7699113 Victoria, Australia 5.14 --- 1.18 --- 4.36 --- --- --- 13 Victoria, Australia 5.71 --- 0.6 --- 9.52 --- --- --- 76994 54 3.5 3 Dol.cpx 2.5 2 1.5 1 0.5 0 0 2 4 6 8 Run Duration (days) Figure 4.1.1. Time series of piston cylinder equilibrium partitioning experiments from Yakob et al. (2012) showing the DLiol/cpx evolving over time. At 900°C the system approaches equilibrium with respect to Li partitioning between olivine and clinopyroxene within 24 hours. Figure 4 from Yakob et al., 2012. 55 -5.00 -7.00 log fO2 -9.00 -11.00 -13.00 -15.00 FMQ -17.00 MW -19.00 Mo-MoO2 -21.00 700 900 1100 1300 1500 1700 Temperature (°C) Figure 4.2.1. Solid state oxygen fugacity buffers rhenium-rhenium dioxide (Re-ReO2), fayalite-magnetite-quartz (FMQ), magnetite-wüstite (MW), and molybdenum-molybdenum dioxide (Mo-MoO2) shown as a function of temperature. Vertical black line indicates experimental run conditions. Buffer curves calculated from: Mo-MoO2 = O’Neill, 1986; Magnetite-Wüstite & Fayalite-Magnetite-Quartz = Myers and Eugster, 1983. 56 5 Yakob et al. 2012 Ottolini et al. 2009 4 Brenan et al. 1998 Dol/cpx Blundy and Dalton 2000 3 Caciagli et al. 2011 2 1 0 600 700 800 900 1000 1100 1200 1300 1400 Temperature (C) Figure 4.4.1. Experimentally derived DLiol/cpx showing that the olivine/clinopyroxene partition coefficient is ~2±0.5 regardless of temperature. Figure 6 from Yakob et al., 2012. Data from Ottolini et al. (2009), Caciagli et al. (2011), Brenan et al. (1998), Yakob et al. (2012) and Blundy and Dalton (2000) 57 Isotopic Fractionation 30 25 ln(α)*103 20 15 10 5 0 -5 0.6 0.8 1 1.2 1.4 1.6 1.8 (1/T^2)*106 Figure 4.4.2. Northrop and Clayton (1966) state that a linear relationship should exist between ln(α)*103 vs. T-2*106 if isotopic fractionation exists as an equilibrium phenomenon. Plotting ln(α)*103 vs. T-2*106 for all available xenolith data shows that in most cases there is no observable relationship, and therefore the isotopic fractionation seen in xenoliths is not the product of equilibrium isotope fractionation. A particular suite from Rudnick and Ionov (2007), however, do seem to fit to a linear trend suggesting a potential relationship between isotopic fractionation and temperature. Data from Table 4.4.1 58 4 3.5 3 DLiol/cpx 2.5 2 1.5 1 0.5 0 750 950 1150 1350 Temperature (°C) Figure 4.4.3. Apparent Li partition coefficients measured from mantle xenolith mineral separates worldwide show no relationship with temperature. Data from Table 4.4.1. 59 5. CONCLUSIONS If equilibrium with respect to lithium distribution in mantle minerals is defined by DLiol/di = 2.0 0.5 and 7Liol-di = 0 5‰, then many, but not all, mantle xenoliths display non-equilibrium behavior with respect to Li partitioning and/or Li isotopic fractionation between coexisting mineral phases. The lack of any clear relationship between isotopic fractionation and temperature suggests that the isotopic variability seen in mantle xenoliths is not an equilibrium phenomenon, as isotopic effects should be minimized at high temperature, but rather a short lived kinetic effect resulting from the differing diffusivities of 6Li and 7Li. The driving force behind Li redistribution could be a result of closed or open system processes. One theory that has been invoked in the literature without experimental backing is that equilibrium elemental partitioning of Li between olivine and clinopyroxene changes as a function of temperature such that Li becomes more compatible in clinopyroxene as cooling progresses. The resulting closed-system diffusion of lithium from olivine to clinopyroxene would be accompanied by isotopic fractionation, as the lighter 6Li diffuses more rapidly than the heavier 7Li. While diffusion is in progress, then, clinopyroxene would become isotopically lighter and olivine would become isotopically heavier. This dynamic fractionation could be preserved in the minerals if the xenolith temperature drops below the closure temperature for Li diffusion before equilibration is complete. However, Yakob et al. (2012) demonstrated that the equilibrium partition coefficient for lithium between olivine and diopside is independent of temperature between 700˚C and 1100˚C, suggesting a temperature dependent partition coefficient is not the driver for Li redistribution. Another quasi-closed-system process that could drive Li diffusion between olivine and clinopyroxene is a change in the ambient oxygen fugacity. Substitution of Li+ for Mg2+ or Fe2+ in the olivine structure requires coupled substitution with a trivalent ion in order to maintain charge balance. One candidate for a substitution partner is Fe3+. Because Fe3+/Fe is a function of oxygen fugacity, a shift toward higher fO2 in the mantle would result in greater Fe3+ incorporation in olivine, and potentially greater opportunity for Li substitution in turn. By this logic, systems at higher fO2 should shift 60 toward higher DLiol/di and systems at lower fO2 should shift toward lower DLiol/di. I have shown through piston cylinder experiments using solid-state fO2 buffers that the partition coefficient for Li between olivine and clinopyroxene remains between 1.5-2.0 at 1.0 GPa and 900˚C, over a range of oxygen fugacity values representative of the upper mantle (~FMQ to ~5 log units below FMQ). This result suggests that closed-system redistribution of Li between phases is not a feasible method of generating the Li signatures observed in mantle xenoliths. Another possibility is that the Li signature is instead affected by open-system interaction with metasomatic fluids or melts prior to or during exhumation. This implies that the Li signature of many mantle xenoliths may not be representative of the ambient mantle, but is instead a mixture of the mantle and the metasomatic agent. The resetting of Li isotopic distribution in mantle xenoliths on short timescales prior to or during exhumation limits the applicability of stable Li isotopes as a passive tracer of recycled material in the mantle, but opens the possibility of using Li as a geo-speedometer to measure the timescales of short-lived processes such as magma migration and volcanic eruptions. 61 6. FAILED EXPERIMENTS 6.1 Methodology We wanted to develop an experimental setup to effectively control the oxygen fugacity within the sample chamber of our piston cylinder experiments while at high pressure and temperature for the duration of the experiment. Considering a piston cylinder experiment is effectively a closed system, the only experimental design that could control the oxygen fugacity was one that included a solid-state oxygen fugacity buffer which fixes the activity of O2 via a chemical reaction that occurs at run conditions. Assuming all necessary phases are present at the conclusion of the experiment, one can assume the activity of O2, and thus the oxygen fugacity was fixed within the buffer. Many experimental designs were constructed and tested until a successful design was chosen that controlled all important variables. 6.2 Experimental Designs The first few experimental designs attempted to use as similar a design as possible to that described by Yakob et al. (2012), where a Pt-lined Ni capsule was used to grow olivine and clinopyroxene at varying temperatures. The first design was an attempt to use a double capsuled experimental setup in the original Pt-lined Ni capsule. A fO2 capsule was created out of 3-mm gold tubing and put inside the experimental chamber. The tubing was crimped and welded on one end, filled with powder for an oxygen buffer and ultrapure Milli-Q water, crimped and welded shut. The capsule was weighed before and after adding every component to ensure material was retained in the capsule after each step. In these experiments, water is present on both sides of the small inner capsule, so H+ can become available via the breakdown of water and diffuse through the semipermeable gold walls to allow for chemical communication between the two chambers while physically isolating the buffer from the experiment. These first experiments had a few issues. For one, the amount of material that could be retained in the small inner buffer capsule was very small. The buffer being used in these first few experiments was Ni-NiO, and these buffer capsules were completely 62 full of NiO at the end of the experiment. Additionally, Ni escaped the inner capsules and creating bright green Ni-enriched olivine and clinopyroxene. We tried running experiments with the oxygen fugacity buffer in the bigger outside chamber and the actual experiment in the small Au capsule, however we still found large amounts of NiO in the run products. Before developing this experimental design further, we came to a revelation that the method had a major flaw. The experiments were being run in a Pt-lined Ni capsule. At run conditions, Fe alloys very strongly with Pt, thus we were producing Fe-free olivine and clinopyroxene. The premise behind our research was that the relative abundance of Fe3+ as controlled by the oxygen buffer was the controlling factor in Li partitioning between the mantle phases. If we remove all the Fe from the system, there are no major redox sensitive elements in the experiment, meaning that controlling the fO2 should not have any effect on Li partitioning. To address the Fe loss issue, we decided that the best method would be to run the experiments in a silver capsule with no noble-metal lining. Fe and Ag do not alloy at run conditions, so the Fe present in the starting materials should remain there throughout the experiment, and the oxygen fugacity buffer can effectively control the speciation of Fe. The first Ag capsule design was very similar to its predecessor. An inner Ag-Pd capsule was constructed in a similar manner as before to serve as the oxygen fugacity buffer capsule within the system. The Ag-Pd capsule was filled with fayalite, magnetite, quartz and ultrapure Milli-Q water, placed inside the larger Ag capsule with the experimental powders. This resulted in growing beautifully large brown fayalite crystals in the experiment. Fe from the buffer must have escaped the inner capsule, saturating the experimental charge with Fe and growing Fe-rich olivine. A new Ag capsule design was constructed where multiple holes were drilled as close together as possible into the Ag capsule, creating a multi-chambered Ag charge. The experiment and the oxygen fugacity buffer would go in separate chambers, with the hope that H+ could still diffuse through the much larger barrier separating the two chambers. It is unclear if the buffer worked in these experiments, because the experimental charge as a whole collapsed each time this design was used. The Ag walls 63 were too thin and weak to support the run conditions, and they would fail resulting in a loss of water to the outside assembly and a failed experiment. Finally, the inverted capsule technique described in the Methods section above was developed. This involved having two Ag capsules, one filled with the oxygen fugacity buffer and one filled with the experiment. The buffer capsule would be inverted and set on top of the experiment capsule, separated by a thin Ag-Pd foil. The foil allowed for the chemical communication between the two chambers while preventing physical contamination. 64 REFERENECES Aulbach, S., Rudnick, R.L. (2009) Origins of non-equilibrium lithium isotopic fractionation in xenolithic peridotite minerals: Examples from Tanzania. Chem. Geol. 258, 17-27 Aulbach, S., Rudnick, R.L, McDonough, W.F. (2008) Li-Sr-Nd isotope signatures of the plume and cratonic lithosphere mantle beneath the margin of the rifted Tanzanian craton (Labait) Contrib. Mineral. Petrol. 155, 79-92 Ayers, J.C., Brenan, J.B., Watson, E.B., Wark, D.A., Minarik, W.G. (1992) A new capsule technique for hydrothermal experiments using the piston-cylinder apparatus. American Mineralogist 77, 1080-1086. Benton, L.D., J. G. Ryan, and I. P. Savov. (2004). Lithium abundance and isotope systematics of forearc serpentinites, conical seamount, Mariana forearc; insights into the mechanics of slab-mantle exchange during subduction. Geochem, Geophys, Geosyst – G3, 5 (8). Blundy, J. and Dalton, J. (2000) Experimental comparison of trace element partitioning between clinopyroxene and melt in carbonate and silicate systems, and implications for mantle metasomatism. Contrib. Mineral. Petrol. 139, 356-371. Brenan, J.M., Neroda, E., Lundstrom, C.C., Shaw, H.F., Ryerson, F.J., Phinney, D.L. (1998a) Behavior of boron, beryllium, and lithium during melting and crystallization: Constraints from mineral-melt partitioning experiments. Geochim. Cosmochim. Acta 62, 2129-2141. Caciagli-Warman, N. (2010) Experimental constraints on lithium exchange between clinopyroxene, olivine and aqueous fluids at high pressures and temperatures. PhD thesis, University of Toronto Caciagli, N., Brenan, J.M., McDonough, W.F., Phinney, D. (2011) Mineral-fluid partitioning of lithium and implications for slab-mantle interaction. Chem. Geol. 280, 384-398. Chan L.H., Edmond J.M. (1988) Variation of lithium isotope composition in the marine environment: A preliminary report. Geochim Cosmochim Acta 52, 1711-1717 Chan, L.H., Edmond, J.M., Thompson, G., Gillis, K. (1992) Lithium isotopic composition of submarine basalts: implications for the lithium cycle in the oceans. Earth Planet. Sci. Lett. 108, 151-160. 65 Chan L.H., Edmond J.M., Thompson G. (1993) A lithium isotope study of hot springs and metabasalts from mid-ocean ridge hydrothermal systems. J. Geophys. Res. 98, 9653-9659 Chan, L.H., Gieskes J.M., You, C.F., Edmond, J.M. (1994) Lithiu isotope geochemistry of sediments and hydrothermal fluids of the Guaymas Basin, Gulf of California. Geochim. Cosmochim. Acta 58 4443-4454. Chan L.H., Kastner M. (2000) Lithium isotopic compositions of pore fluids and sediments in the Costa Rica subduction zone: implications for fluid processes and sediment contribution to the arc volcanoes. Earth Planet. Sci. Lett. 183, 275-290 Chan L.H., Alt J.C., Teagle D.A.H. (2002a) Lithium and lithium isotope profiles through the upper oceanic crust: a study of seawater-basalt exchange at ODP Sites 504B and 896A. Earth Planet. Sci. Lett. 201, 187-201 Chan L.H., Leeman W.P., Tonarini S., Singer B. (2002b) Lithium and boron isotopes in the Aleutian Islands: Contributions of marine sediments to island arc magmas. EOS Trans., Am. Geophys. Union 83, 1482 Chan, L.H., Leeman, W.P., Plank, T. (2006) Lithium isotopic composition of marine sediments. Geochem.Geophys. Geosyst. 7, Q06005, doi:10.1029/2005GC001202 Coogan, L.A., Kasemann, S.A., Chakraborty, S. (2005) Rates of hydrothermal cooling of new oceanic upper crust derived from lithium-geospeedometry. Earth Planet. Sci. Lett. 240, 415-424. Cottrell, E., Kelley, K.A., (2011) The oxidation state of Fe in MORB glasses and the oxygen fugacity of the upper mantle. Earth Planet. Sci. Lett. 305, 270-282 Decitre S., Deloule E., Reisberg L., James R., Agrinier P., Mével C. (2002) Behavior of lithium and its isotopes during serpentinization of oceanic peridotites. Geochem. Geophys. Geosyst. 3:10.1029/2001GC000178 Dohmen, R., Kasemann, S.A., Coogan, L., Chakraborty, S. (2010) Diffusion of Li in olivine. Part I: Experimental observations and a multi species diffusion model. Geochim. Cosmochim. Acta 74, 274-292. Elderfield H., Schultz A. (1996) Mid-ocean ridge hydrothermal fluxes and the chemical composition of the ocean. Ann. Rev. Earth Planet. Sci 24, 191-224 Elliott T., Jeffcoate A.B., Bouman C. (2004) The terrestrial Li isotope cycle: light-weight constraints on mantle convection. Earth Planet. Sci. Lett. 220, 231-245. 66 Fritz S.J., Whitworth T.M. (1994) Hyperfiltration-induced fractionation of lithium isotopes: Ramifications relating to the representativeness of aquifer sampling. Water Resour. Res. 30, 225-235 Frost, D.J., McCammon, C.A. (2008) The Redox State of Earth’s Mantle. Annu. Rev. Earth Planet. Sci. Lett. 36, 386-420. Grant, K.J., Wood, B.J. (2010) Experimental study of the incorporation of Li, Sc, Al and other trace elements into olivine. Geochim. Cosmochim. Acta. 74, 2412-2428 Huh Y., Chan L.H., Zhang L., Edmond J.M. (1998) Lithium and its isotopes in major world rivers: Implications for weathering and the oceanic budget. Geochim. Cosmochim. Acta. 62, 2039-2051 Huh Y., Chan L.H., Edmond J.M. (2001) Lithium isotopes as a probe of weathering processes: Orinoco River. Earth Planet. Sci. Lett. 194, 189-199 Ionov, D.A. and Seitz, H.M. (2008) Lithium abundances and isotopic compositions in mantle xenoliths from subduction and intra-plate settings: Mantle sources vs. eruption histories. Earth Planet. Sci. Lett. 266, 316-331. James R.H., Rudnicki M.D., Palmer M.R. (1999) The alkali element and boron geochemistry of the Escanaba Trough sediment-hosted hydrothermal system: The role of sediments. Earth Planet. Sci. Lett. 171, 157-169 James R.H., Palmer M.R. (2000a) The lithium isotope composition of international rock standards. Chem. Geol. 166, 319-326 Jamieson, H.E., Roeder, P.L., Grant, A.H. (1992) Olivine-Pyroxene-PtFe Alloy as an Oxygen Geobarometer. J. Geol 100, 138-145 Jeffcoate, A.B., Elliot, T., Kasemann, S.A., Ionov, D., Cooper, K., Brooker, R. (2007) Li isotope fractionation in peridotites and mafic melts. Geochim. Cosmochim. Acta 71, 202-218. 53 Kaliwoda, M., Ludwig, T., Altherr, R. (2008) A new SIMS study of Li, Be, B and δ7Li in mantle xenoliths from Harrat Uwayrid (Saudi Arabia). Lithos 106, 261-279 Kent, A.J.R., Rossman, G.R. (2002) Hydrogen, lithium, and boron in mantle-derived olivine: The role of coupled substitutions. Am. Min. 87, 1432-1436 Lee, C.-T., Leeman, W.P., Canil, D., Li, Z.-X.A., (2005) Similar V/Sc Systematics in MORB and Arc Basalts: Implications for the Oxygen Fugacities of their Mantle Source Regions. J. Pet. 46, 2313-2336 67 Leeman, W.P., Tonarini, S., Chan, L.H., Borg, L.E. (2004) Boron and lithium isotopic variations in a hot subduction zone – the southern Washington Cascades. Chem. Geol. 212, 101-124. Magna, T., Wiechert, U., Halliday, A.N. (2006) New constraints on the lithium isotope compositions of the Moon and terrestrial planets. Earth and Planet. Sci. Lett. 243, 336-353 Magna, T., Ionov, D.A., Oberli, F., Wiechert, U. (2008) Links between mantle metasomatism and lithium isotopes: Evidence from glass-bearing and cryptically metasomatized xenoliths from Mongolia. Earth Planet. Sci. Lett. 276, 214-222 Mallmann, G., O’Neill, H.St.C., Klemme, S. (2009) Heterogeneous distribution of phosphorous in olivine from otherwise well-equilibrated spinel peridotite xenoliths and its implications for the mantle geochemistry of lithium. Contrib. Mineral. Petrol. 158, 485-504 Mattioli, G.S., Wood, B.J. (1986) Upper mantle oxygen fugacity recorded by spinel lherzolites. Nature 322, 626-628 Millot, R., Vigier, N., Gaillardet, J., (2010a) Behaviour of lithium and its isotopes during weathering in the Mackenzie Basin, Canada. Geochim. Cosmochim. Acta. 74, 3897-3912 Millot, R., Petelet-Giraud, E., Guerrot, C., Negrel, P. (2010b) Multi-isotopic composition (δ7Li–δ11B–δD–δ18O) of rainwaters in France: Origin and spatio-temporal characterization. Appl. Geochem. 25, 1510-1524. Moriguti, T., T. Shibata, and E. Nakamura. (2004). Lithium, boron and lead isotope and trace element systematics of Quaternary basaltic volcanic rocks in northeastern Japan: mineralogical controls on slab-derived fluid composition. Chem. Geol., 212, 81- 100 Myers, J., Eugster, H.P. (1983) The System Fe-Si-O: Oxygen Buffer calibrations to 1,500K. Contrib. Mineral. Petrol. 82, 75-90 Northrop, D.A. and Clayton, R.N. (1966) Oxygen-isotope fractionations in systems containing dolomite. J. Geol. 74, 174-196. O’Neill, H.St.C. (1986) Mo-MoO2 (MOM) oxygen buffer and the free energy of formation of MoO2. Am. Min. 71, 1007-1010 O’Neill, H.St.C., Wall, V.J. (1987) The Olivine-Orthopyroxene-Spinel Oxygen Geobarometer, the Nickel Precipitation Curve, and the Oxygen Fugacity of the Earth’s Upper Mantle. J. Petrol. 28, 1169-1191 68 Ottolini, L., Le Fevre, B., Vannucci, R. (2004) Direct assessment of mantle boron and lithium contents and distribution by SIMS analyses of peridotite minerals. Earth and Planet. Sci. Lett. 228, 19-36 Ottolini, L., Laporte, D., Raffone, N., Devidal, J-L., Le Fevre, B., (2009) New experimental determination of Li and B partition coefficients during upper mantle partial melting. Contrib. Mineral. Petrol. 157, 313-325 Parkinson, I.J., Hammond, S.J., James, R.H., Rogers, N.W. (2007) High-temperature lithium isotope fractionation: Insights from lithium isotope diffusion in magmatic systems. Earth Planet. Sci. Lett. 257, 609-621. Plank, T., Langmuir, C.H. (1998) The chemical composition of subducting sediment and its consequences for the crust and mantle. Chem. Geol. 145, 325-394 Pistiner J., Henderson G.M. (2003) Lithium isotope fractionation during continental weathering processes. Earth Planet. Sci. Lett. 214, 327-339 Richter, F.M., Davis, A.M., DePaolo, D.J., Watson, E.B. (2003) Isotope fractionation by chemical diffusion between molten basalt and rhyolite. Geochim. Cosmochim. Acta 67, 3905-3923. Rudnick, R.L. and Ionov, D.A. (2007) Lithium elemental and isotopic disequilibrium in minerals from peridotite xenoliths from far-east Russia: Product of recent melt/fluid-rock reaction, Earth Planet. Sci. Lett. 256, 278-293. Rudnick, R.L., Tomascak, P.B., Njo, H.B., Gardner, L.R. (2004) Extreme lithium isotopic fractionation during continental weathering revealed in saprolites from South Carolina. Chem. Geol. 212, 45-57. Seitz, H-M. and Woodland, A.B. (2000) The distribution of lithium in peridotitic and pyroxenitic mantle lithologies – an indicator of magmatic and metasomatic processes. Chem. Geol. 166, 47-64. Seitz, H-M., Brey, G.P., Lahaye, Y., Durali, S., Weyer, S. (2004) Lithium isotopic signatures of peridotite xenoliths and isotopic fractionation at high temperatures between olivine and pyroxenes. Chem. Geol. 212, 163-177. Spandler, C., O’Neill, H.St.C. (2010) Diffusion and partition coefficients of minor and trace elements in San Carlos olivine at 1,300°C with some geochemical implications. Contrib. Mineral. Petrol. 159, 791-818 Stagno, V., Ojwang, D.O., McCammon, C.A., Frost, D.J. (2013) The oxidation state of the mantle and the extraction of carbon from Earth’s interior. Nature 493, 84-90 69 Tang, Y-J., Zhang, H-F., Ying, J-F. (2007a) Review of the lithium isotope system as a geochemical tracer. Internat. Geol. Rev. 49, 674-688. Tang et al., 2007b (Hannuoba xenoliths) Teng F.Z., McDonough W.F., Rudnick R.L., Dalpé C., Tomascak P.B., Gao S., Chappell B.W. (2004) Lithium isotopic composition and concentration of the upper continental crust. Geochim. Cosmochim. Acta. 68, 4167-4178 Tomascak, P.B., Tera, F., Helz, R.T., Walker, R.J. (1999a) The absence of lithium isotope fractionation during basalt differentiation: New measurements by multicollector sector ICP-MS. Geochim. Cosmochim. Acta 63, 907-910. 55 Tomascak P.B., Widom E., Benton L.D., Goldstein S.L., Ryan J.G. (2002) The control of lithium budgets in island arcs. Earth Planet Sci. Lett. 196, 227-238 Tomascak, P.B., Langmuir, C.H., le Roux, P.J., Shirey, S.B. (2008) Lithium isotopes in global mid-ocean ridge basalts. Geochim. Cosmochim. Acta 72, 1626-1637. Trail, D., Watson, E.B., Tailby, N.D. (2012) Ce and Eu anomalies in zircon as proxies for the oxidation state of magmas. Geochim. Cosmochim. Acta. 97, 70-87 Woodland, A.B., Koch, M. (2003) Variation in oxygen fugacity with depth in the upper mantle beneath the Kaapvaal craton, Southern Africa. Earth Planet. Sci. Lett. 214, 295-310 Woodland, A.B., Seitz, H.M., Yaxley, G.M. (2004) Varying behavior of Li in metasomatised spinel peridotite xenoliths from western Victoria, Australia. Lithos 75, 55-66. Yakob, J.L., Feineman, M.D., Deane Jr. J.A., Eggler, D.H., Penniston-Dorland, S.C. (2012) Lithium partitioning between olivine and diopside at upper mantle conditions: An experimental study. Earth Planet. Sci. Lett. 330, 11-21 Zack, T., Tomascak, P.B., Rudnick, R.L., Dalpé, C., McDonough, W.F. (2003) Extremely light Li in orogenic eclogites: The role of isotope fractionation during dehydration in subducted oceanic crust. Earth Planet. Sci. Lett. 208, 279-290. Zhang L., Chan L.H., Gieskes J.M. (1998) Lithium isotope geochemistry of pore waters from Ocean Drilling Program Sites 918 and 919, Irminger Basin. Geochim. Cosmochim. Acta. 62, 2437-2450 70 Appendix: Sample Assembly and Piston Cylinder Procedures The sample assembly consists of (Figure 2.2.2): 2 Ag capsules Ag-Pd foil Solid, center drilled, and annulus MgO ceramic pieces MgO ceramic disk NaCl tube Graphite tube Graphite disk Pyrex tube Lead foil Steel base plug Pyrophyllite ring Ceramic thermocouple insulator (“spaghetti”) Three pieces of “crushable” MgO ceramic is used to surround the Ag capsule. The bottom piece is solid ceramic, the middle piece resembles an annulus inside which the crucible sits, and the upper piece has a hole drilled through the center where the thermocouple can enter and sit directly on top of the sample capsule. To prevent the metal wires of the thermocouple from coming in contact with the Ag capsule, a thin disk of MgO ceramic is laid on top of the Ag capsule and MgO ceramic annulus to provide insulation. The four pieces of MgO ceramic are all made by hand on a lathe in the lab. The lathe allows us to shave the outside of the ceramic pieces to the correct diameter, if necessary, so they can fit snugly inside the rest of the sample assembly. The annulus ceramic piece is drilled with a stationary drill bit of the correct diameter while the ceramic piece spins in the lathe, and each individual piece is cut from a longer ceramic rod using a handsaw. 71 The sample capsule and ceramic insulators are then placed inside a graphite furnace. The height of the three ceramic pieces should be no longer than the height of the graphite because the furnace must make complete contact with the baseplug in order for current to be conducted uniformly to the furnace. A current is passed through the graphite, which has high electrical resistivity. Current is converted into heat according to the relationship P = RI2, where P is power (heat transfer in watts), R is electrical resistance (in ohms) and I is the current (in amperes). A graphite disk of the same diameter is carefully glued to the bottom of the graphite sleeve in order to allow nearly complete contact between the two graphite pieces. The graphite disk is in contact with the piston, which serves to ground the furnace. Surrounding the graphite is a Pyrex (glass) tube of the same height as the graphite sleeve and disk. The Pyrex tube serves to physically and electrically insulate the graphite sleeve from the rest of the system, so the electrical current can only pass through the graphite. The graphite disk is cemented to both the bottom of the graphite sleeve and the inside of the Pyrex tube, and is left to harden for a few moments before being inserted into the piston cylinder apparatus. A salt cell is used to transfer pressure to the sample assembly hydrostatically and with a minimum of frictional loss. The cell is made of NaCl using a custom die-press. NaCl powder is poured into the die, which has the shape of a ring, and pressurized so the powder recrystallizes to form a coherent tube of NaCl. These are prepared ahead of time and kept in an oven at 100°C to prevent atmospheric water from compromising the structure of the NaCl tube. The salt cell should be exactly the same length as the sample assembly. If the cell is too long, it can be ground to the appropriate length using coarse grit sandpaper. The salt cell is then wrapped in lead foil, which serves as a lubricant as well as a barrier to prevent salt from oozing into gaps or cracks in the tungsten carbide core of the pressure plate during an experiment. The lead foil wraps around the salt cell, but does not cover the entire circumference of the tube; a 2-3mm gap remains. The lead foil also wraps around part of the bottom end of the salt cell. Should the lead foil be wrapped completely around the bottom the salt cell, it could potentially come into contact with the graphite furnace, causing a short in the electrical circuit controlling the temperature of the system and resulting in an experimental failure. 72 Before loading an experiment into the piston cylinder apparatus, a thermocouple must be made in order to measure and control the temperature of the sample during the experiment. Pt wire is threaded through one of the holes of a two-holed piece of ceramic thermocouple insulator (“spaghetti”) 4.7 cm in length, while Pt/Rh(10%) wire is thread through the other. Upon exiting the ceramic, the two wires are lightly crimped and welded using an arc welder to form a bead of metal that rests on top of the thermocouple ceramic. The other ends of the wire are covered in insulating Teflon tubes to prevent them from contacting each other or any other metallic surface of the apparatus, which could cause an electrical short during the experiment. Before being inserted into the experiment, the freshly welded thermocouple must be annealed over a Bunsen-burner flame to heal the metal. When both the sample assembly and the thermocouple are prepared, the experiment is ready to be put together. First, a (insert correct number here) inch diameter tungsten carbide “pusher” is inserted into a guide ring. The guide ring serves to center the pusher on the master ram. The lower cooling plate is centered on the pusher by a snug-fitting brass sleeve. The cooling plate has input and output quick-release couplings where hoses can be attached to circulate cooling fluid (water) to cool the pressure vessel during the experiment. An upper cooling plate will later be aligned and these two are attached to one another by hoses so water can flow through and cool both the top and the bottom of the pressure vessel. A ¾-inch tungsten carbide piston is cleaned and placed on top of the pusher, inside the aligning brass sleeve, and centered using two steel guide rings. A sheet of Mylar with a ¾-inch center-punched hole is fit snugly around the piston, extending beyond the edges of the pressure vessel. The Mylar sheet serves as an insulator to prevent current from flowing through the pressure vessel instead of the graphite furnace, and also to prevent water from escaping out of the cooling plate. The core of the pressure vessel is then lubricated with a thin layer of Molykote and placed on top of the bottom cooling plate, centered on the piston. The inner bore has the same diameter as the piston, so it is important to lubricate the hole to prevent frictional forces from building up and resisting pressurization or causing fracturing or chipping of the brittle tungsten carbide. Surrounding the pressure plate is a cooling ring which is 73 connected by hoses to the upper and lower cooling plates. This ring serves to cool the pressure plate from the outside during the experiment. In order to load the sample assembly into the piston cylinder apparatus, first the lead-wrapped salt cell is inserted into the bore of the lubricated pressure plate core. The sample assembly is inserted carefully inside the salt cell. If using an unwelded capsule such as the one described above, it is particularly critical to avoid jarring or tiling the same assembly in order to avoid water loss from the capsule prior to cold-sealing. On top of the salt cell and sample assembly, but still inside the pressure plate bore, a steel base plug is inserted and surrounded by a pyrophyllite ring. The base plug serves as a conductor for the electrical current which flows through the graphite furnace, and as a soft metal that upon pressurization vertically, expands horizontally and seals the end of the pressure plate. The pyrophyllite ring insulates the electrical current from the pressure plate. The base plug and pyrophyllite ring are then aligned flush with the surface of the pressure plate by adjusting the master ram, and another Mylar sheet is centered on the surface. This Mylar sheet serves to prevent water leaks, but also electrical shorts as it electrically insulates the pressure plate from the upper cooling plate. A ½” hole is punched in the middle of the Mylar, which allows electricity to flow to the base plug. It is important to be sure that the hole is centered over the base plug such that no part of the pressure plate is exposed, as this could result in a short circuit. The upper cooling plate is then aligned using a series of thin metal rods with slightly varying lengths and stiffnesses. Holes in the center of the cooling plate, base plug, and upper MgO crushable ceramic piece have the same diameter as the thermocouple ceramic, such that they must all be perfectly aligned for the thermocouple to pass through and rest on top of the sample capsule. When the upper cooling plate is properly aligned, the thermocouple is inserted into the hole and pressed down until the entire thermocouple ceramic disappears from view. The insulated wires that emerge from the top are taped down into a groove cut into the upper cooling plate and led to the positive and negative leads that connect to the temperature control system. When the thermocouple is connected, the control system can read the temperature of the system 74 based on the ambient EMF generated by the temperature at the nose of the thermocouple. The EMF changes with temperature, resulting in changes in the voltage read at the control center. The voltage changes are calibrated with temperature based on the electrical properties of the individual thermocouple wires, and a specific temperature is displayed. A stack of steel spacers are placed on top of the upper cooling plate to fill void space between the cooling plate and the upper block of the PC frame. When all the plates are stacked, a 1-2mm gap between the stack and the upper block should remain. At this point, the sample assembly is complete and pressurization can begin. Initially the end load ram is raised to close the gap between the plate and the PC frame, after which point continued pumping pressurizes the pressure vessel, raising the confining pressure on the sample. The end load pressure is raised to the pressure necessary for the given experimental conditions (Table A1). Then, the master ram is raised, which pushes the piston up, causing the sample pressure to increase. The initial pressure increase takes significant time because the crushable components of the sample assembly, the ceramic, salt cell, and Pyrex tube, all have to break and compress before the pressure can increase. Initial pressurization is to about 0.8 GPa, where we assume the Pt lid and the Pt lined capsule have sealed together. This is important because when we heat the sample, an incomplete seal between the lid and capsule would allow fluid to escape, and the experiment would fail (although this may not be apparent until after the run was taken down). The water hoses are connected to the cooling plates and ring, draining to the sink, and the water is turned on in a fairly strong, steady stream. A failsafe switch will prevent power from being supplied to the furnace if the water flow is too low or interrupted. At this point, we begin bring up the temperature by turning on the power to the PC. High voltage from the wall-mounted electrical box is converted to high current using a step down transformer. The current is brought up slowly using a rheostat to avoid shocking the system. Current is applied from the bottom of the PC, through the master ram, pusher, piston, graphite furnace, base plug, upper cooling plate, metal stack and into the PC frame where the electrical circuit is completed. A Eurotherm temperature controller receives the temperature signal from the thermocouple and supplies current to 75 the furnace as necessary to maintain the temperature and the desired setpoint for the duration of the experiment. We maintain a stable temperature within 1°C throughout the duration of the experiment. Throughout the process of increasing the temperature to the set value, we continue to increase the pressure of the master ram until we reach the desired experimental pressure (Table A1). After a few minutes of fluctuations in pressure and temperature where minor manual adjustments to both may be necessary, the system reaches a steady state and the pressure and temperature remain constant for the duration of the experiment. Upon completion of the desired experiment duration, a kill switch is used to instantly cut off electrical power to the heating system, quenching the experiment. The temperature typically drops to < 300°C (well past the closure temperature for most high temperature geochemical exchange) in less than 30 seconds, and <100˚C in 60 seconds. This rapid cooling ensures that the chemical signature of the crystals grown during the experiment is representative of what was stable at the experimental run conditions. Once the sample has cooled to 30˚C, the water system is turned off and a hose is attached to one of the nozzles and air is forced through the cooling lines to clear out any remaining water. The pressure on the master ram and end load ram is released just enough to open a gap between the spacer stack and the PC frame without allowing any weight to be taken up by the piston. First, the stack of spacers is removed from the top cooling plate. The top cooling plate is jolted to the side, snapping the thermocouple ceramic at the cooling plate – pressure plate interface. The cooling plate is removed and cleaned, dried, and put away for storage. The Mylar sheet and cooling ring are both removed, and the pressure plate and lower cooling plate are lifted together to a bench where they can be separated. The pressure plate is then put below a hydraulic ram where a steel rod pushes the sample assembly, which is now stuck inside the pressure plate due to radial expansion, out of the bottom of the plate. Both the pressure plate and lower cooling plate are cleaned, dried, and put away. The remaining sample assembly is a condensed version of what was put into the system initially. The salt cell, Pyrex sleeve, graphite tube and ceramic pieces are 76 all delicately chipped away from the sample capsule using a small hammer, pliers, and/or wire cutters so that the sample capsule is all that remains. 77 Table A.1 Pressure derivation for the piston cylinder apparatus used in this study 1 inch Master Ram – Gauge pressure (bars) / 1 kbar sample pressure 10 End Load Ram – Gauge pressure (psi) / 1 kbar sample pressure 592 ¾ inch 5.625 333 ½ inch 2.5 148 Piston diameter 78
© Copyright 2025 Paperzz