共查询到20条相似文献,搜索用时 15 毫秒
1.
A. Suzuki E. Ohtani K. Funakoshi H. Terasaki T. Kubo 《Physics and Chemistry of Minerals》2002,29(3):159-165
The viscosity of albite (NaAlSi3O8) melt was measured at high pressure by the in situ falling-sphere method using a high-resolution X-ray CCD camera and a large-volume
multianvil apparatus installed at SPring-8. This system enabled us to conduct in situ viscosity measurements more accurately
than that using the conventional technique at pressures of up to several gigapascals and viscosity in the order of 100 Pa s. The viscosity of albite melt is 5.8 Pa s at 2.6 GPa and 2.2 Pa s at 5.3 GPa and 1973 K. Experiments at 1873 and 1973
K show that the decrease in viscosity continues to 5.3 GPa. The activation energy for viscosity is estimated to be 316(8)
kJ mol−1 at 3.3 GPa. Molecular dynamics simulations suggest that a gradual decrease in viscosity of albite melt at high pressure may
be explained by structural changes such as an increase in the coordination number of aluminum in the melt.
Received: 6 January 2001 / Accepted: 27 August 2001 相似文献
2.
We present Raman and infrared spectra of gypsum to 21 GPa at 300 K. Our measurements encompass the internal modes of the
(SO4)−4 group that lie between 400 and 1150 cm−1, hydroxyl-stretching vibrations between 3200 and 3600 cm−1, and a libration and bending vibrations of the molecular H2O group. All vibrations of the sulfate group have positive pressure shifts, while the hydroxyl-stretching and -bending vibrations
have a mixture of positive and negative pressure shifts: the effect of pressure on the hydrogen bonding of the water molecule
thus appears to be complex. Near 5 GPa, the two infrared-active bending vibrations of the water molecule coalesce, and the
morphology of the hydroxyl-stretching region of the spectrum shifts dramatically. This behavior is consistent with a pressure-induced
phase transition in gypsum in the vicinity of 5–6 GPa, which is observed to be reversible on decompression to zero pressure.
The spectral observations are consistent with the onset of increased disorder in the position of the water molecule in gypsum:
the sulfate vibrations are largely unaffected by this transition. The Raman-active symmetric stretch of the sulfate group
undergoes an apparent splitting near 4 GPa, which is interpreted to be produced by Fermi resonance with an overtone of the
symmetric bending vibration. The average mode Grüneisen parameter of the 20 vibrational modes we sample is less than 0.05,
in contrast to the bulk thermal Grüneisen parameter of 1.20. Accordingly, the vibrations of both water and sulfate units within
gypsum are highly insensitive to volumetric compaction. Therefore, in spite of the changes in the bonding of the water unit
near 5 GPa, metastably compressed gypsum maintains strongly bound molecular-like units to over 20 GPa at 300 K.
Received: 31 July 2000 / Accepted: 5 April 2001 相似文献
3.
In order to clarify Al2O3 content and phase stability of aluminous CaSiO3-perovskite, high-pressure and high-temperature transformations of Ca3Al2Si3O12 garnet (grossular) were studied using a MA8-type high-pressure apparatus combined with synchrotron radiation. Recovered samples
were examined by analytical transmission electron microscopy. At pressures of 23–25 GPa and temperatures of 1000–1600 K, grossular
garnet decomposed into a mixture of aluminum-bearing Ca-perovskite and corundum, although a metastable perovskite with grossular
composition was formed when the heating duration was not long enough at 1000 K. On release of pressure, this aluminum-bearing
CaSiO3-perovskite transformed to the “LiNbO3-type phase” and/or amorphous phase depending on its Al2O3 content. The structure of this LiNbO3-type phase is very similar to that of LiNbO3 but is not identical. CaSiO3-perovskite with 8 to 25 mol% Al2O3 was quenched to alternating lamellae of amorphous layer and LiNbO3-type phase. On the other hand, a quenched product from CaSiO3-perovskite with less than 6 mol% consisted only of amorphous phase. Most of the inconsistencies amongst previous studies
could be explained by the formation of perovskite with grossular composition, amorphous phase, and the LiNbO3-type phase.
Received: 11 April 2001 / Accepted: 5 July 2002 相似文献
4.
F. Farges 《Physics and Chemistry of Minerals》2001,28(9):619-629
Fe–K edge XAFS spectra (pre-edge, XANES and EXAFS) were collected for eight grandidierites from Madagascar and Zimbabwe,
as well as for Fe(II) and Fe(III) model compounds (staurolite, siderite, enstatite, berlinite, yoderite, acmite, and andradite).
The pre-edge spectra for these samples are consistent with dominantly 5-coordinated ferrous iron. The analysis of the XANES
and EXAFS spectra confirms that Fe(II) substitutes for Mg(II) in grandidierite, with a slight expansion of the local structure
around Mg by ∼2%. In addition, ferric iron was also detected in some samples [5–10 mol% of the total Fe or 500–1100 ppm Fe(III)].
Based on theoretical calculations of the EXAFS region, Fe(III) appears to be located in the 5-coordinated sites of Mg(II)
or in the most distorted 6-coordinated sites of Al (depending on the sample studied). Special attention is therefore required
when using grandidierite as a model for ferrous iron in C3v geometry, because of the possible presence of an extra contribution related to Fe(III). This additional contribution enhances
significantly the Fe–K pre-edge integrated area [+40% for 1000 ppm Fe(III)]. Therefore, only a few grandidierite samples can
be used as a robust structural model for the study of the Fe(II) coordination in glasses and melts.
Received: 26 June 2000 / Accepted: 19 February 2001 相似文献
5.
A seeping sea-floor in a Ria environment: Ria de Vigo (NW Spain) 总被引:3,自引:0,他引:3
The occurrence of gas accumulations in the Ría de Vigo (NW Spain) have been characterized by the authors in previous research.
Pockmarks frequently appear on seismic and sonar records, covering ca. 45% of the sea-floor of the study area, which indicates
that gas expulsion is not an uncommon phenomena in the coastal Ría environment. Here we report the occurrence of gas seepage
for the first time along the coast of NW Spain. Side-scan sonar, echo-sounder and high-resolution seismic techniques, were
used for mapping gas-expulsion features. Some expulsion pockmarks sit over elongated features that represent bottom marks
created by anthropogenic activity. Thus, these anthropogenic sites may act as preferential venting zones for gas, as well
as being potential hazards on a muddy sea-bed such as that of the Ría where gas accumulates just below the surface of the
sea-floor.
Received: 25 May 1998 · Accepted: 20 November 1998 相似文献
6.
H. Graetsch 《Physics and Chemistry of Minerals》2001,28(5):313-321
The crystal structure of intermediate incommensurate tridymite was refined at 150 °C from powder data. Upon cooling from
above 220 °C, the basic structure with space group symmetry C2221 is gradually distorted from orthorhombic to monoclinic symmetry. With decreasing temperature, the monoclinic angle γ smoothly
opens up to 90.3°, while a displacive modulation with temperature-dependent wavelength develops. The 3 + 1 dimensional superspace
group of the incommensurate phase is C1121(αβ0). The modulation mainly consists of two sinusoidal transverse displacement waves for the silicon atoms coupled to rotations
of the rigid SiO4/2 tetrahedra. The wave vector is r=0.1192(1)a* − 0.0043(1)b* at 150 °C. Below 150 °C tridymite discontinuously transforms to another orthorhombic phase and the modulation partially
locks in at the wave vector r
1=1/3a*. Simultaneously, an additional incommensurate modulation with r
2= 0.0395(1)b* − 0.3882(1)c* is formed. The two-dimensional modulation does not vary significantly with the temperature.
Received: 13 September 2000 / Accepted: 29 January 2001 相似文献
7.
Variations of Raman spectra of hydroxyl-clinohumite were studied up to ∼370 kbar at room temperature, and in the range 81–873 K
at atmospheric pressure. With the exception of the symmetric OH-stretch bands, the Raman frequencies of all bands were observed
to increase monotonically with increasing pressure, and decrease with increasing temperature. This behavior is in line with
those observed for other humite members (norbergite and chondrodite) so far studied. The symmetric OH-stretching band shows
a mode softening with increasing pressure, and splits into two bands at either high pressure or low temperature. In the quasihydrostatic
experiment, the compression and decompression paths of one of the asymmetric OH-stretch bands form a hysteresis loop, but
the same behavior was not observed in the nonhydrostatic experiment. These results indicate that the two kinds of OH groups
in hydroxyl-clinohumite have nonequivalent movement paths on compression, and with one OH group experiencing a release of
spatial hindrance during compression. This behavior appears to be modified by shear stress. The same complication of the OH
groups was not observed in the temperature variation study. The pressure and temperature variations of the Raman frequencies
for the various vibrations involving the SiO4 tetrahedra and MgO6 octahedra below ∼1000 cm−1 for clinohumite behave similarly to other hydrous magnesium silicates. On the basis of the relationship between isothermal
bulk modulus and Raman data, it is suggested that the linear pressure dependences of vibrational frequencies of various Raman
bands reported in the literature are inadequate.
Received: 20 March 1999 / Revised, accepted: 24 August 1999 相似文献
8.
The stepwise dehydration process of the Ba-exchanged form of the zeolite phillipsite was studied by in situ synchrotron X-ray
powder diffraction. A series of structure refinements were performed using the Rietveld method on powder diffraction data
measured in the interval between 332 and 712 K. At 482 K, more than half of the water molecules were lost. The continuous
water loss causes the Ba cations to migrate inside the zeolite channels in order to achieve a stable coordination with the
framework oxygens.
The dehydration process was completed at 663 K, where a new, completely dehydrated stable phase was detected. The temperature
range of stability of this phase was more than 100 K, thanks to the stable coordination of the Ba cations with the framework
oxygens. This phase is the first example of completely dehydrated zeolite containing divalent (barium) cations.
Received: 8 January 2001 / Accepted: 1 November 2001 相似文献
9.
Incorporation of hydrocarbons (CH4, C2H6, C3H8, C4H10) in cordierite channels was experimentally studied at 700 °C and at pressures from 200 to 1000 MPa, (avoiding recrystallization).
The maximum concentrations of hydrocarbons determined by gas chromatographic analysis are CH4=78.3, C2H6=134, C2H4 + C2H6=26, C3H8=28, C4H10=32, C5H12=23, and C6H14=7 × 10−3 wt%). According to IR spectroscopy data, the channel forms of hydrocarbons differ from the forms on the surface. As a result
of interactions with the framework oxygen, normal hydrocarbons are converted to saturated oligomers and their fragments. Small
amounts of water molecules of the first and second types (up to 0.3 wt%) are formed in the same way. As pressure grows from
200 to 1000 MPa, the total content of structural hydrocarbons is nearly doubled. Special runs on cordierite saturation in
mixtures of hydrocarbons with water showed that low contents of hydrocarbons in natural cordierites correspond to their large
concentrations in fluid.
Received: 7 May 2001 / Accepted: 17 July 2001 相似文献
10.
11.
Geophysical methods, when integrated with soil chemical and hydrogeological methods, can be used to investigate groundwater
contamination. Direct current (DC) resistivity geo-electrical sounding and very-low-frequency electromagnetic (VLF-EM) data
were collected at an open waste site used by the municipality of the city of Isparta, Turkey. The groundwater at shallow depths
in alluvium, which is composed of gravel, sand and clay, were expected to be hazardously contaminated under and around the
open waste-disposal site, in which both household and industrial wastes are known to be disposed of improperly. In this study,
we mapped the spread of groundwater contamination using a VLF-EM method, which allows fast and inexpensive data collection.
The method complements the results of geo-electrical sounding. There is a good correlation between the results of the VLF-EM
and the DC-resistivity methods employed for the investigation of subsurface structure of the site, where soil chemical and
previous hydrogeological surveys have indicated high levels of chemical concentrations.
Received: 17 January 2000 · Accepted: 12 August 2000 相似文献
12.
A. Pavese 《Physics and Chemistry of Minerals》2002,29(1):43-51
The P–V–T equation of state (EoS) models of Birch–Murnaghan, Vinet and Poirier–Tarantola have been compared with one another and discussed
in the light of their ability to reproduce thermoelastic functions and parameters by means of fitting to pressure–volume–temperature
data artificially generated for spinel, corundum and forsterite. Numerical simulations relying upon semi-empirical potentials,
lattice dynamics and the quasiharmonic approximation have been used to generate P–V–T data. The results obtained indicate that all the P–V–T EoSs tested predict bulk modulus at ambient conditions with errors confined, at worst, within a few percent, and reproduce
correctly its dependence on temperature. The derivatives of the bulk modulus versus P and PT are less satisfactorily modelled. The bulk thermal expansion is determined by EoSs within a few percent error, but the deviations
increase significantly if the approximation of linear dependence of EoS on temperature is used (linearised thermal pressure
model).
Received: 30 January 2001 / Accepted: 16 June 2001 相似文献
13.
The nature of OH species in natural clear quartz was investigated by means of in-situ IR measurements over the temperature
range –185 to 1000 °C. Reversible thermal behavior of OH species was examined for a sample pre-heated to 1000 °C for 1 hour.
At room temperature, the IR spectrum of the quartz sample examined includes an intense absorption peak at 3379 cm–1 which has been assigned to an OH stretching vibration associated with Al substituting for Si (OH(Al)). The major spectral
changes of the OH(Al) bond involve a systematic shift of its peak position and a decrease in its integral absorbance with
temperature. A quasi-linear increase of the peak position from –185 to 400 °C is interpreted to be due to the change in the
vibrational frequency of OH(Al) with hydrogen bond (H bond) distance. At higher temperatures, the IR frequency shows only
a slight change, indicating a small influence of the H bond. On the other hand, the gradual decrease of the integral absorbance
of OH(Al) with temperature indicates a decrease of this defect’s molar absorptivity without any reduction in defect concentration.
This is interpreted to result from a decrease in dipole moment of OH(Al) with temperature. A sudden shift of the vibrational
frequency from 3396 to 3386 cm–1 between 550 and 560 °C and a constant value of the integral absorbance from 535 to 570 °C were considered to be related to
the change in H bond distance during the structural transformation of α-quartz to its β-form. The local environment of OH(Al)
begins to change at temperatures below 570 °C, where the crystallographic α–β transition occurs.
Received: 18 February 1998/ Accepted: 10 July 1998 相似文献
14.
T. Okada W. Utsumi H. Kaneko M. Yamakata O. Shimomura 《Physics and Chemistry of Minerals》2002,29(7):439-445
An experimental technique to make real-time observations at high pressure and temperature of the diamond-forming process
in candidate material of mantle fluids as a catalyst has been established for the first time. In situ X-ray diffraction experiments
using synchrotron radiation have been performed upon a mixture of brucite [Mg(OH)2] and graphite as starting material. Brucite decomposes into periclase (MgO) and H2O at 3.6 GPa and 1050 °C while no periclase is formed after the decomposition of brucite at 6.2 GPa and 1150 °C, indicating
that the solubility of the MgO component in H2O greatly increases with increasing pressure. The conversion of graphite to diamond in aqueous fluid has been observed at
7.7 GPa and 1835 °C. Time-dependent X-ray diffraction profiles for this transformation have been successfully obtained.
Received: 17 July 2001 / Accepted: 18 February 2002 相似文献
15.
Mobile sediment in an urbanizing karst aquifer: implications for contaminant transport 总被引:1,自引:0,他引:1
Here we investigate geochemical characteristics of sediment in different compartments of a karst aquifer and demonstrate
that mobile sediments in a karst aquifer can exhibit a wide range of properties affecting their contaminant transport potential.
Sediment samples were collected from surface streams, sinkholes, caves, wells, and springs of a karst aquifer (the Barton
Springs portion of the Edwards (Balcones Fault Zone) Aquifer, Central Texas) and their mineralogy, grain-size distribution,
organic carbon content, and specific surface area analyzed. Statistical analysis of the sediments separated the sampling sites
into three distinct groups: (1) streambeds, sinkholes, and small springs; (2) wells; and (3) caves. Sediments from the primary
discharge spring were a mix of these three groups. High organic carbon content and high specific surface area gives some sediments
an increased potential to transport contaminants; the volume of these sediments is likely to increase with continued urbanization
of the watershed.
Received: 13 April 1998 · Accepted: 6 October 1998 相似文献
16.
W. van Westrenen N. L. Allan J. D. Blundy M. Yu Lavrentiev B. R. Lucas J. A. Purton 《Physics and Chemistry of Minerals》2003,30(4):217-229
We have performed atomistic computer simulations on trace element incorporation into the divalent dodecahedral X-sites of
pyrope (Py — Mg3Al2Si3O12) – grossular (Gr — Ca3Al2Si3O12) solid solutions. An ionic model and the Mott–Littleton two-region approach to defect energies were used to calculate the
energetics of substitution by a range of divalent trace-elements and of charge-balanced substitution by trivalent ions in
the static limit. Results are compared with experimental high-temperature, high-pressure garnet-melt trace element partitioning
data obtained for the same garnet solid solution to refine our understanding of the factors controlling element partitioning
into solid solutions. Defect energies (U
def,f), relaxation (lattice strain) energies (U
rel), and solution energies (U
sol) were derived using two different approaches. One approach assumes the presence of one type of hybrid X-site with properties
intermediate between pure Mg and Ca sites, and the other assumes discrete Mg and Ca X-sites, and thus two distinct cation
sublattices. The hybrid model is shown to be inadequate, since it averages out local distortions in the garnet structure.
The discrete model results suggest trace elements are more soluble in Py50Gy50 than in either end-member compound. Physically this is due to small changes in size of the X-sites and the removal of unfavourable
interactions between third nearest neighbours of the same size. Surprisingly, depending on the local order, large trace element
cations may substitute for Mg2+ and small trace elements for Ca2+ in Py50Gr50. These solubilities provide an explanation for the anomalous trace-element partitioning behaviour along the pyrope–grossular
join observed experimentally.
Received: 27 January 2000 / Accepted: 14 February 2003 相似文献
17.
In polycrystalline aggregates of olivine with mean grain sizes above 35 μm plus a low basaltic melt fraction, both wetted
and melt-free grain boundaries are observed after equilibration times at high pressures and temperatures of between 15 and
25 days. In order to assess a possible dependence of the wetting behaviour on the relative orientation of neighbouring grains,
a SEM based technique, electron backscatter diffraction (EBSD), is used to determine grain orientations. From the grain orientations
relative orientations of neighbouring grains are calculated, which are expressed as misorientation axis/angle pairs. The distribution
of misorientation angles and axes of melt-free grain boundaries differ significantly from a purely random distribution, whereas
those of wetted grain boundaries are statistically indistinguishable from the random distribution. The relative orientation
of two neighbouring grains therefore influences the character of their common grain boundary. However, no clustering towards
special (coincident site lattice) misorientation axes is observed, with the inference that the energy differences between
special and general misorientations are too small to lead to the development of preferred misorientations during grain growth.
Received: 8 December 1997 / Revised, accepted: 6 April 1998 相似文献
18.
Pavese et al. (1999) examined cation partitioning vs. temperature in a synthetic spinel of composition (Mg0.70 Fe0.23
3+) Al1.97 O4 using structure data obtained from in situ neutron powder diffraction. After imposing assumptions on the site assignment
of vacancies and Fe3+, they assigned the remaining cations by applying least-squares minimization to chemical constraints on site-occupancy sums,
site-scattering, chemical composition, and thermal expansion of the octahedral site. Their proposed site assignments exhibit
a sharp discontinuity in occupancy fractions versus temperature, a necessary consequence of their assumptions on vacancy assignments.
In this paper we reexamine the cation partitioning of the same spinel using the constrained least-squares formulation of OccQP (Wright et al. 2000), which optimizes site occupancies without ad hoc assumptions. We obtain strikingly different results,
supporting the general view that spinel undergoes a lambda transition at ∼1000 K. For all observed parameters, the residuals
obtained with the OccQP assignments are lower than those obtained with the Pavese et al. assignments, in some cases by more than 1 order of magnitude.
Received: 05 April 2000 / Accepted: 19 October 2000 相似文献
19.
T. Hattori T. Matsuda T. Tsuchiya T. Nagai T. Yamanaka 《Physics and Chemistry of Minerals》1999,26(3):212-216
In order to confirm the possible existence of FeGeO3 perovskite, we have performed in situ X-ray diffraction measurements of FeGeO3 clinopyroxene at pressures up to 40 GPa at room temperature. The transition of FeGeO3 clinopyroxene into orthorhombic perovskite is observed at about 33GPa. The cell parameters of FeGeO3 perovskite are a=4.93(2) Å, b=5.06(6) Å, c=6.66(3) Å and V=166(3) Å3 at 40 GPa. On release of pressure, the perovskite phase transformed into lithium niobate structure. The previously reported decomposition process of clino-pyroxene into Fe2GeO4 (spinel)+GeO2 (rutile) or FeO (wüstite) +GeO2 (rutile) was not observed. This shows that the transition of pyroxene to perovskite is kinetically accessible compared to the decomposition processes under low-temperature pressurization. 相似文献
20.
The compression of synthetic pyrope Mg3Al2 (SiO4)3, almandine Fe3Al2(SiO4)3, spessartine Mn3Al2 (SiO4)3 grossular Ca3Al2(SiO4)3 and andradite Ca3Fe2 (SiO4)3 was studied by loading the crystals together in a diamond anvil cell. The unit-cell parameters were determined as a function
of pressure by X-ray diffraction up to 15 GPa using neon as a pressure transmitting medium. The unit-cell parameters of pyrope
and almandine were measured up to 33 and 21 GPa, respectively, using helium as a pressure medium. The bulk moduli, K
T
0, and their first pressure derivatives, K
T
0
′, were simultaneously determined for all five garnets by fitting the volume data to a third order Birch-Murnaghan equation
of state. Both parameters can be further constrained through a comparison of volume compressions between pairs of garnets,
giving for K
T
0 and K
T
0
′ 171(2) GPa and 4.4(2) for pyrope, 185(3) GPa and 4.2(3) for almandine, 189(1) GPa and 4.2 for spessartine, 175(1) GPa and
4.4 for grossular and 157(1) GPa and 5.1 for andradite, where the K
T
0
′ are fixed in the case of spessartine, grossular and andradite. Direct comparisons of the unit-cell volumes determined at
high pressures between pairs of garnets reveal anomalous compression behavior for Mg2+ in the 8-fold coordinated triangular dodecahedron in pyrope. This agrees with previous studies concerning the compression
behaviors of Mg2+ in 6-fold coordinated polyhedra at high pressures. The results show that simple bulk modulus–volume systematics are not obeyed
by garnets.
Received: 29 July 1998 / Revised, accepted: 7 April 1999 相似文献