Reversible displacive phase transition in [Ni(en)3]2+(NO3−)2: a potential temperature calibrant for area-detector diffractometers

2003 ◽  
Vol 36 (1) ◽  
pp. 141-145 ◽  
Author(s):  
L. J. Farrugia ◽  
P. Macchi ◽  
A. Sironi

The coordination complex [Ni(en)3]2+(NO{}_{3}^{- })2(en = 1,2-diaminoethane) undergoes a sharp reversible displacive phase transition at ∼109 K, changing space group fromP6322 above the transition temperature toP6522 below. The phase change is accompanied by a tripling of thecaxis on cooling, resulting in an easy detection of the transition in images from area-detector diffractometers. The transition has been followed using a Nonius KappaCCD and a Bruker SMART APEX CCD. Data sets were collected over the temperature range 100–113 K and integrated using the low-temperature orientation matrix. Reflections withl≠ 3nshow a smooth and rapid decrease in intensity to zero on warming from 106.5 to 111 K. The results are reproducible to within ±2 K in two laboratories and suggest that this compound may be useful as a liquid-nitrogen cryo-calibrant for diffraction instruments equipped with area detectors.

2012 ◽  
Vol 26 (32) ◽  
pp. 1250207 ◽  
Author(s):  
YONGKANG LUO ◽  
XIAO LIN ◽  
YUKE LI ◽  
QIAN TAO ◽  
LINJUN LI ◽  
...  

Parent compound of DyFeAsO was successfully synthesized by solid-state reaction under ambient pressure and superconductivity was induced by partial substitution of trivalent Dy 3+ ions with tetravalent Th 4+ in Dy 1-x Th x FeAsO . In the undoped parent compound, an anomaly in the resistivity appears around 140 K which corresponds to the structural phase transition and/or antiferromagnetic (AFM) order of the magnetic moments of Fe 2+ ions. At low temperature, another AFM order associated with the magnetic moments of Dy 3+ ions occurs at TN of 9.55 K. The AFM order around 140 K has significant influence on the transport properties, which can be interpreted by opening of partial gap on Fermi surface. Th doping suppresses the AFM order related to the Fe 2+ ions and the midpoint transition temperature [Formula: see text] of 49.3 K is observed for x = 0.3. Our results also indicate that the [ Ln 2 O 2]2+ layer has influence on the magnetism of [ Fe 2 As 2]2- layer.


2020 ◽  
Vol 32 (3) ◽  
pp. 305-310
Author(s):  
Wendi Liu ◽  
Yan Yang ◽  
Qunke Xia

Abstract. It has long been known that hydrogen impurities can be incorporated in the structure of nominally anhydrous minerals (NAMs) and substantially influence their physical properties. One of the geologically most prominent NAMs is feldspar. The hydrogen concentration in NAMs is usually expressed in parts per million of water by weight (ppm H2O wt.) In this paper, we use the term “hydrogen” for uniformity, except when we use “water” for describing its amount expressed as parts per million of H2O by weight. In our article (Liu et al., 2018), we carried out in situ high-temperature X-ray powder diffraction and Raman spectroscopic studies on three natural anorthoclase samples with similar Or (K-feldspar) contents (Ab67Or31An2, Ab66Or31An2, and Ab65Or33An3) and Al–Si disordering but contrasting water contents. The spectroscopic results suggested that the displacive phase transition temperature is higher for the nearly anhydrous anorthoclase sample than the anorthoclase samples with about 200 ppm water, and we thus concluded that hydrogen is another factor impacting the displacive phase transition temperature. We thank Kroll and Schmid-Beurmann for pointing out the weakness in our interpretation that hydrogen is a possible important factor (Kroll and Schmid-Beurmann, 2020). To clarify this issue, we conducted transmission electron microscopy (TEM) experiments on the three samples to check texture effects. The TEM studies indicated that the nearly anhydrous anorthoclase sample consists of two feldspar phases, a K-poor and a K-rich one, and that the K-poor area may be responsible for the higher displacive phase transition temperature. According to the observation that the temperature of redistribution of hydrogen is accordant with the displacive phase transition temperature, the effect of hydrogen could not be ruled out. Based on these results, it can be concluded that hydrogen may not be the sole possible factor, and it was a proposition more than a definitive proof for the moment. Natural feldspars are complex, and factors affecting displacive phase transitions are multiple (e.g., Salje et al., 1991; Harrison and Salje, 1994; Hayward and Salje, 1996; Dobrovolsky et al., 2017). Therefore, to further investigate hydrogen effects on displacive phase transition in feldspar, synthetic samples with pure chemical compositions and hydrogen species are necessary. In the following, we address each issue in the same order as in the comment by Kroll and Schmidt-Beurmann (2020).


2003 ◽  
Vol 10 (02n03) ◽  
pp. 519-524 ◽  
Author(s):  
Toshio Takahashi ◽  
Hiroo Tajiri ◽  
Kazushi Sumitani ◽  
Koichi Akimoto ◽  
Hiroshi Sugiyama ◽  
...  

The structure of the [Formula: see text] surface was studied at both room temperature and a low temperature of 50 K using grazing incidence X-ray diffraction. At low temperatures diffuse scattering was observed in addition to Bragg reflection. Least squares analyses for Bragg reflections using anisotropic Debye–Waller factors show that the structure at 50 K is consistent with an inequivalent triangle (IET) model, while the structure at room temperature is explained by a honeycomb-chained triangle (HCT) model with strong anisotropic Debye–Waller factors. From the temperature dependence of diffuse scattering, the phase transition temperature Tc and critical exponent β were determined to be about 150 K and 0.27. Some Bragg intensities showed discontinuous changes in their first derivatives at Tc. The results favor a displacive phase transition rather than an order–disorder one.


RSC Advances ◽  
2016 ◽  
Vol 6 (90) ◽  
pp. 86872-86879 ◽  
Author(s):  
Lanli Chen ◽  
Yuanyuan Cui ◽  
Siqi Shi ◽  
Bin Liu ◽  
Hongjie Luo ◽  
...  

The calculated oxygen-vacancy diffusion barrier indicates that the existence of oxygen-vacancy could stabilize the rutile phase at a low temperature.


2018 ◽  
Vol 30 (6) ◽  
pp. 1071-1081 ◽  
Author(s):  
Wendi Liu ◽  
Yan Yang ◽  
Qunke Xia ◽  
Yu Ye ◽  
Zhongping Wang ◽  
...  

2011 ◽  
Vol 25 (12n13) ◽  
pp. 1151-1155
Author(s):  
V. THANH NGO ◽  
N. A. VIET

Using the simple EPB model1 of DNA (the combined model of the pendulum model of Englander,2 and Peyrard–Bishop microscopic model3), we propose a bio (DNA) time measured by EPB pendulum. We have shown that the effective bio time has a thermodynamic origin which equals the ordinary (real) time in the range of room-body temperature. Furthermore, it flows slower at low temperature and much faster near the DNA melting phase transition temperature than real time.


Sign in / Sign up

Export Citation Format

Share Document