Library

feed icon rss

Your email was sent successfully. Check your inbox.

An error occurred while sending the email. Please try again.

Proceed reservation?

Export
  • 1
    Electronic Resource
    Electronic Resource
    s.l. : American Chemical Society
    The @journal of physical chemistry 〈Washington, DC〉 96 (1992), S. 2801-2804 
    Source: ACS Legacy Archives
    Topics: Chemistry and Pharmacology , Physics
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 2
    Electronic Resource
    Electronic Resource
    [s.l.] : Nature Publishing Group
    Nature 325 (1987), S. 236-238 
    ISSN: 1476-4687
    Source: Nature Archives 1869 - 2009
    Topics: Biology , Chemistry and Pharmacology , Medicine , Natural Sciences in General , Physics
    Notes: [Auszug] The spectra were obtained for homogeneous, ordered plagioclases which had been purified from natural metamorphic and slowly cooled igneous rocks and had almost all been characterized by transmission electron microscopy, electron microprobe analysis, X-ray powder diffraction and high-temperature ...
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 3
    Electronic Resource
    Electronic Resource
    Springer
    Physics and chemistry of minerals 10 (1984), S. 256-260 
    ISSN: 1432-2021
    Source: Springer Online Journal Archives 1860-2000
    Topics: Chemistry and Pharmacology , Geosciences , Physics
    Notes: Abstract A number of previous investigations have examined the ordering behavior of magnesium cordierite using X-ray diffraction, transmission electron microscopy, infrared spectroscopy and solution calorimetry. In the present investigation, one series of samples from the above studies has been examined by Raman spectroscopy. Systematic modifications in the spectra with annealing time at 1,200° C are consistent with a continuous ordering of the average Al/Si distribution from 4 h to at least 64 h, and which may begin earlier. Spectral changes are first definitely observed when the ordered domains are around 100 Å across, suggesting that Raman spectroscopy is sensitive to this distance scale. The spectra of samples annealed at 1,200° C are compared with samples annealed at 1,400°; C where ordering proceeds much faster, and the possible use of Raman spectroscopy in characterization of Al/Si order in cordierite is discussed. Finally, the Raman spectrum of Mg2Al4Si5O18 with a stuffed β-quartz structure has been obtained. Comparison of its spectrum with that of cordierite glass suggests similar structures for both, which seem different to that of disordered cordierite.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 4
    Electronic Resource
    Electronic Resource
    Springer
    Contributions to mineralogy and petrology 71 (1980), S. 289-300 
    ISSN: 1432-0967
    Source: Springer Online Journal Archives 1860-2000
    Topics: Geosciences
    Notes: Abstract The free energy curves for simple binary solid solutions with limited miscibility or atomic ordering have been combined to predict the phase relations and exsolution mechanisms for a system in which both ordering and exsolution are possible. The nature of the ordering process affects which exsolution mechanisms may be used. If the ordering is second (or higher) order in character then continuous mechanisms predominate and a ‘conditional spinodal’ (Alien and Cahn, 1976) can be described which operates between ordered and disordered end members. For a first order case, the ordered phase can only precipitate a disordered phase by nucleation and growth. Microstructures in omphacites observed by transmission electron microscopy include exsolution lamellae and antiphase domains and the relations between them in selected specimens have been used to interpret the exsolution mechanisms which operated under geological conditions. It appears that most omphacites undergo cation ordering, and then remain homogeneous or exsolve a disordered pyroxene by spinodal decomposition. The predominance of continuous mechanisms has been used to indicate that the C2/c→P2/n transformation may be second (or higher) order in character. A possible phase diagram for jadeite-augite is presented. It is based on the idea that there should be limited miscibility between the disordered end members at low temperatures and that the cation ordering at intermediate compositions (omphacite) is superimposed on a solvus. It is adequate to explain many of the observed microstructures and fits with petrographic evidence of broad two phase fields between impure jadeite and omphacite and between omphacite and sodic augite. The effect of adding acmite is analogous to increasing temperature so that the phase relations for jadeite-acmite-augite can also be predicted.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 5
    Electronic Resource
    Electronic Resource
    Springer
    Contributions to mineralogy and petrology 78 (1982), S. 433-440 
    ISSN: 1432-0967
    Source: Springer Online Journal Archives 1860-2000
    Topics: Geosciences
    Notes: Abstract The kinetics of cation disordering in a natural ordered (P2/n) omphacite have been followed at P=18 and 30 kb, T= 750–1,260° C, for times of between 1.5 min and 16 days in a piston-cylinder apparatus. Time-temperature-transformation (TTT) analysis of the experimental data, using the presence or absence of the 11¯1 reflection in single crystal X-ray precession photographs to indicate the extent of reaction, yields an equilibrium order/disorder temperature (T ord) of 865±10° C, an activation enthalpy (1 bar) of 71±6 kcal mole−1 and an activation volume of 9±4 cm3 mole−1 (plus and minus figures represent the precision of a best fit between experimental data and TTT theory rather than absolute errors). The activation volume is consistent with a vacancy mechanism of cation diffusion. H2O, added in the form of oxalic acid, appears to speed the process up slightly. The overall transformation mechanism is continuous, involving neither the nucleation of a disordered phase nor a change in antiphase-domain distribution. This is consistent with both first- and non-first-order character for the C2/c⇋P2/n transformation, though a range of ordered states below T ord is indicated by the weakening of h+k=odd reflections. A simple extrapolation of the disordering rates to geological conditions leads to the first estimate of how long disordered omphacites would take to order in nature, ranging from less than one year at T≈800° C to more than 107 years at T〈350° C.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 6
    Electronic Resource
    Electronic Resource
    Springer
    Contributions to mineralogy and petrology 78 (1982), S. 441-451 
    ISSN: 1432-0967
    Source: Springer Online Journal Archives 1860-2000
    Topics: Geosciences
    Notes: Abstract Omphacites from a wide range of geological environments have been examined by transmission electron-microscopy. Their microstructures are sufficiently variable as to be potential indicators of thermal history for blueschist and eclogite metamorphism. In particular, the average size of equiaxed antiphase domains (APD's) arising from cation ordering appears to be a characteristic feature of each environment and increases in the sequence: Franciscan, blueschist (1) ≈ Turkey, blueschist (2) 〈 Guatemala, jadeitic blocks in serpentinite (3) 〈 Syros, blueschist (9) ≈ Red Wine Complex, Canada, amphibolite (1) 〈 Maksyutov Complex, Urals, blueschist (3) ≈ Zermatt-Saas, blueschist (5) ≈ Allalin, metagabbro (4) 〈 Tauern, eclogite (1) ≈ Franciscan, eclogite (5) 〈 Nybö, Norway, eclogite (2) (numbers in brackets indicate the number of hand specimens for which omphacite microstuctures are known). A relationship between APD size, annealing time and temperature has been derived by analogy with the known APD coarsening behaviour in other systems where: (APD size)n $$({\text{APD size)}}^{\text{n}} \propto {\text{e}}^{{\text{(}} - {\text{Q/RT)}}} \cdot {\text{ }}time{\text{.}}$$ . Most omphacites fit into a self-consistent scheme with n=8±2 if the activation energy (Q) is assumed to be that of cation disordering (75 kcal mole−1), available estimates of peak metamorphic temperature (T) are used, and a reasonable geological time-scale is taken as 104–108 years. According to this model, APD sizes are set in a relatively short interval of the total history of a rock when its temperature is close to its peak value. APD sizes are much more sensitive to temperature than to time and may be used as a geothermometer which has the advantage of not being reset by re-equilibration at low temperatures. Petrological implications arising from the model are that Allalin metagabbros were metamorphosed at a similar peak temperature to Zermatt-Saas blueschists, Franciscan eclogites reached higher temperatures than has been previously supposed and that the microstructures in some Sesia-Lanzo omphacites are consistent with a high temperature, pre-blueschist origin. Deviation from an ideal coarsening law with n=2 implies that the APD's are not simply stacking mistakes but have some associated structural or compositional modification locally. Excess titanium concentrated at APD's in Red Wine Complex omphacites may account for their anomalously low observed APD size.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 7
    Electronic Resource
    Electronic Resource
    Springer
    Contributions to mineralogy and petrology 87 (1984), S. 138-148 
    ISSN: 1432-0967
    Source: Springer Online Journal Archives 1860-2000
    Topics: Geosciences
    Notes: Abstract Taking account of the C¯1/I¯1 (Al/Si order/disorder) transformation at high temperatures in the albite-anorthite solid solution leads to a simple model for the mixing properties of the high structural state plagioclase feldspars. The disordered (C¯1) solid solution can be treated as ideal (constant activity coefficient) and, for anorthite-rich compositions, deviations from ideality can be ascribed to cation ordering. Values of the activity coefficient for anorthite in the C¯1 solid solution (γ An C¯1 ) are then controlled by the free energy difference between C¯1 and I¯1 anorthite at the temperature (T) of interest according to the relation: Δ¯G ord I¯1 ⇌C¯1 =RT ln γ An C¯1 . If the I¯1⇌C¯1 transformation in pure anorthite is treated, to a first approximation, as first order and the enthalpy and entropy of ordering are taken as 3.7±0.6 kcal/mole (extrapolated from calorimetric data) and 1.4–2.2 cal/mole (using an equilibrium order/disorder temperature for An100 of 2,000–2,250 K), a crude estimate of γ An C¯1 for all temperatures can be made. The activity coefficient of albite in the C¯1 solid solution (γ Ab C¯1 ) can be taken as 1.0. The possible importance of this model lies in its identification of the principal constraints on the mixing properties rather than in the actual values of γ An C¯1 and γ Ab C¯1 obtained. In particular it is recognised that γ An C¯1 depends critically on ordering in anorthite as well as, at lower temperatures, any ordering in the C¯1 solid solution. A brief review of activity-composition data, from published experiments involving ranges of plagioclase compositions and from the combined heats of mixing plus Al-avoidance entropy model (Newton et al. 1980), reveals some inconsistencies. The values of γ An C¯1 calculated using the approach of Newton et al. (1980), although consistent with Orville's (1972) ion exchange data, are slightly lower than values derived from experiments by Windom and Boettcher (1976) and Goldsmith (1982) or from ion-exchange experiments of Kotel'nikov et al. (1981). Based on the C¯1/I¯1 transformation model, values of γ An C¯1 〈1.0 are unlikely. Discrepancies between the experimental data sets are attributed to incomplete (non-equilibrium) Al/Si order attained during the experiments. It is suggested that the choice of activity coefficients remains somewhat subjective. The development of accurate mixing models would be greatly assisted by better thermodynamic data for ordering in pure anorthite and by more thorough characterisation of the state of order in plagioclase crystals used for phase equilibrium experiments.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 8
    Electronic Resource
    Electronic Resource
    Springer
    Physics and chemistry of minerals 16 (1989), S. 649-658 
    ISSN: 1432-2021
    Source: Springer Online Journal Archives 1860-2000
    Topics: Chemistry and Pharmacology , Geosciences , Physics
    Notes: Abstract Natural samples of K-feldspar representing various states of Al, Si order were characterised using X-ray methods, transmission electron microscopy, and Fourier transform infrared spectroscopy. Line profiles of infrared absorption bands were observed to show strong correlation with the degree of Al, Si order present. In particular, the absorption frequencies of the 540 cm−1 and 640 cm−1 bands were seen to vary by ca. 10 cm−1 between sanidine and microcline, with modulated samples respresenting intermediate behaviour. Linewidths of these modes also decrease by ca. 50% in this series. The experimental results are discussed within the framework of Hard Mode Infrared Spectroscopy (HMIS), and it is shown that the absorption frequencies vary with the short range order parameter τ = (4t1-1)2 and the symmetry breaking order parameter describing Al, Si order, Q od=(t1 0−t1 m)/Q od=(t1 0+t1 m), where t1 is the average Al occupancy on the T1 sites and t1 o and t1 m are the individual site occupancies of the T1 o and T1 m sites, respectively. The structural state of orthoclase is characterised by strain-induced modulations with large spatial variations of the modulation wavelength. No such modulations were observed in the degree of local Al, Si order. Sanidine shows mode hardening in excess of the extrapolated effect of symmetry breaking Al, Si order, which is presumably related to nonsymmetry breaking ordering between T1 and T2 sites and/or as yet unobserved short range order of the symmetry breaking ordering scheme. The possibility of an additional phase transition in K-feldspar at temperatures above 1300 K is discussed.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 9
    Electronic Resource
    Electronic Resource
    Springer
    Physics and chemistry of minerals 13 (1986), S. 119-139 
    ISSN: 1432-2021
    Source: Springer Online Journal Archives 1860-2000
    Topics: Chemistry and Pharmacology , Geosciences , Physics
    Notes: Abstract Approximately 125 hydrothermal annealing experiments have been carried out in an attempt to bracket the stability fields of different ordered structures within the plagioclase feldspar solid solution. Natural crystals were used for the experiments and were subjected to temperatures of ∼650°C to ∼1,000°C for times of up to 370 days at $$P_{{\text{H}}_{\text{2}} {\text{O}}} $$ =600 bars, or $$P_{{\text{H}}_{\text{2}} {\text{O}}} $$ =1,200 bars. The structural states of both parent and product materials were characterised by electron diffraction, with special attention being paid to the nature of type e and type b reflections (at h+k=(2n+1), l=(2n+1) positions). Structural changes of the type C $$\bar 1$$ → I $$\bar 1$$ , C $$\bar 1$$ → “e” structure, I $$\bar 1$$ → “e” and “e” structure → I $$\bar 1$$ have been followed. There are marked differences between the ordering behaviour of crystals with compositions on either side of the C $$\bar 1$$ ⇌ I $$\bar 1$$ transition line. In the composition range ∼ An50 to ∼ An70 the e structure appears to have a true field of stability relative to I $$\bar 1$$ ordering, and a transformation of the type I $$\bar 1$$ ⇌ e has been reversed. It is suggested that the e structure is the more stable ordered state at temperatures of ∼ 800°C and below. For compositions more albite-rich than ∼ An50 the upper temperature limit for long range e ordering is lower than ∼ 750°C, and there is no evidence for any I $$\bar 1$$ ordering. The evidence for a true stability field for “e” plagioclase, which is also consistent with calorimetric data, necessitates reanalysis both of the ordering behaviour of plagioclase crystals in nature and of the equilibrium phase diagram for the albite-anorthite system. Igneous crystals with compositions of ∼ An65, for example, probably follow a sequence of structural states C $$\bar 1$$ → I $$\bar 1$$ → e during cooling. The peristerite, Bøggild and Huttenlocher miscibility gaps are clearly associated with breaks in the albite, e and I $$\bar 1$$ ordering behaviour but their exact topologies will depend on the thermodynamic character of the order/disorder transformations.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
  • 10
    Electronic Resource
    Electronic Resource
    Springer
    Physics and chemistry of minerals 5 (1979), S. 119-131 
    ISSN: 1432-2021
    Source: Springer Online Journal Archives 1860-2000
    Topics: Chemistry and Pharmacology , Geosciences , Physics
    Notes: Abstract Antiphase domains in pigeonite arise from a displacive transformation whereas in omphacite they are due to cation ordering. Their properties and behaviour are quite distinct. Domains in pigeonite are elongate, coarsen by a homogeneous and rapid mechanism and have irregular boundaries which are susceptible to the segregation of cations. In omphacite they tend to be equiaxed with smoothly curving boundaries and they coarsen extremely slowly. A heterogeneous mode of coarsening has been postulated to explain some of the observed domain distributions. A brief review of antiphase textures in other minerals, in particular the type (b) and type (c) domains in calcic plagioclase, show many of these properties to be characteristic of the transformation by which they formed, i.e., displacive or order/disorder. It is suggested that displacive type antiphase boundaries have relatively low surface energies and that their migration requires only small atomic displacements locally. In contrast, the chemical interactions associated with mistakes in the cation ordering sequence at antiphase boundaries in an ordered phase give a relatively higher excess free energy. Furthermore their movement requires the exchange of cations between sites at the boundary and hence domain coarsening is slow. The factors which control the domain size in natural specimens have been evaluated qualitatively. For the purpose of comparing thermal histories only domains with a regular size distribution in homogeneous areas of crystal which are free of defects should be used.
    Type of Medium: Electronic Resource
    Library Location Call Number Volume/Issue/Year Availability
    BibTip Others were also interested in ...
Close ⊗
This website uses cookies and the analysis tool Matomo. More information can be found here...