Группа авторов

Magma Redox Geochemistry


Скачать книгу

alt="equation"/>

      (1.49)equation

      (1.50)equation

      (1.51)equation

      All these equilibria have the interesting feature of displaying unitary activities for oxide component appearing as pure phases, such that their equilibrium constants simply describe the variations of O2 activity (aO2) with temperature:

      (1.52)equation

      or, by defining fugacity, with temperature and pressure:

      (1.53)equation

Schematic illustration of common solid oxygen buffers used in petrology and geochemistry. The lines represent the fugacity-temperature conditions where the phases coexist stably.

      However, when one of these mineral “buffers” is selected as a reference, the acting logfO2 can be given as a relative value, without temperature:

      A relative fO2 scale embodying temperature effects bears just a practical implication (tracking fO2 variations with respect to a reference) but not a real meaning about logfO2 evolution in igneous systems and related environments (e.g., Moretti and Steffansson, 2020). In particular, a common misconception was that in a system a given value of Δbuffer (e.g., ΔQFM = 0.5) could represent some kind of “magic number” characteristic of the whole “rock system” throughout its thermal and chemical evolution. In natural environments oxygen activity (hence fugacity) in fact varies to accommodate the compositional variations and the speciation state of the mineral/melt/fluid phases, and also when highly mobile volatile components are involved, such that fO2 can thus be fixed by factors that are external to the system object of the thermodynamic description.

      Similarly, the rock system evolution cannot be approximated by a unique FeII/FeIII ratio, that the system had when completely molten. Indeed, two rocks that have ideally crystallized along the QFM buffer may have different proportions of fayalite ad magnetite because of the crystallization style but different FeII/FeIII bulk ratios (Frost, 1991). However, in some systems, it is possible that the melt fixed the redox potential of the system via iron oxidation state, with the FeII/FeIII ratio approaching unity (e.g. Moretti et al., 2013).

      On the contrary, it is much less straightforward for fO2 estimates by oxidation states of Fe and/or S measured in glasses, as the oxybarometers derived by the study of synthetic quenched melts still suffer from too many empiric approaches. Because of their polymerized nature, silicate melts do not allow a precise distinction between solute and solvent like in aqueous solutions, where complexes and solvation shells can be easily defined in which covalence forces exhaust (see also Moretti et al., 2014). In fact, melt composition largely affects the ligand constitution and then the speciation state of redox‐sensitive elements. In the case of the FeII/FeIII ratio, the most common redox indicator for melts/glasses, the choice of components in reaction:

      over a large compositional range (e.g., from mafic to silicic) does not offer the possibility to find accurate and internally consistent expressions for the activity coefficients of oxide components γFeO and γFeO1.5 (with FeO1.5 conveniently replacing Fe2O3) that solve the reaction equilibrium constant: