To see the other types of publications on this topic, follow the link: Atomic volume.

Journal articles on the topic 'Atomic volume'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Atomic volume.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Xie, Youqing, and Xiaodong Zhang. "Atomic volumes and volume functions for Ag−Cu alloys." Science in China Series E: Technological Sciences 41, no. 2 (April 1998): 157–68. http://dx.doi.org/10.1007/bf02919678.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Trömel, M. "Atomvolumen, Packungsdichte der Atome und chemische Bindung in nichtmetallischen Elementen." Acta Crystallographica Section B Structural Science 63, no. 4 (July 17, 2007): 532–36. http://dx.doi.org/10.1107/s0108768107022604.

Full text
Abstract:
The atomic volume of crystalline elements is largely determined by the packing density of atoms in the respective modification. The determination of packing density is improved by assuming that the atomic distances depend on bond valences according to Pauling's equation. With the additional assumption of equal valence in different modifications, the experimental atomic volume of an element in any given structure is reduced to its volume in close-packed structures, e.g. f.c.c. The ratio of this reduced atomic volume and the experimental atomic volume is a measure of packing density. Reduced atomic volumes of C, Si, Ge, P, As, S and Se, as calculated from different modifications, correspond in most cases to within less than ±1% for each element, even if calculated from extremely different structures like diamond and buckminsterfullerene in the case of carbon, or from numerous modifications of sulfur with annular molecules of different sizes. Exceptions (graphite, white phosphorus, tin and selenium) indicate deviating valences.
APA, Harvard, Vancouver, ISO, and other styles
3

Politzer, Peter, Ping Jin, and Jane S. Murray. "Atomic polarizability, volume and ionization energy." Journal of Chemical Physics 117, no. 18 (November 8, 2002): 8197–202. http://dx.doi.org/10.1063/1.1511180.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Kollár, J., L. Vitos, and H. L. Skriver. "Anomalous atomic volume of α-Pu." Physical Review B 55, no. 23 (June 15, 1997): 15353–55. http://dx.doi.org/10.1103/physrevb.55.15353.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Bhattacharjee, S. "Haloethanes, geometric volume and atomic contribution method." Computers & Chemistry 18, no. 4 (December 1994): 419–29. http://dx.doi.org/10.1016/0097-8485(94)80036-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Tagawa, Masahito, Kumiko Yokota, Nobuo Ohmae, and Hiroshi Kinoshita. "Volume Diffusion of Atomic Oxygen in α-SiO2 Protective Coating." High Performance Polymers 12, no. 1 (March 2000): 53–63. http://dx.doi.org/10.1088/0954-0083/12/1/305.

Full text
Abstract:
The minimum thickness for an amorphous silicon dioxide (α-SiO2) protective coating required to prevent volume diffusion of atomic oxygen in a low Earth orbit (LEO) was evaluated by measuring the oxide thickness formed on Si(001) wafers in a hyperthermal atomic oxygen beam. The thickness of oxide film was measured by x-ray photoelectron spectroscopy. The diffusion length of atomic oxygen in α-SiO2 at temperatures between 297 K and 493 K, where exterior surfaces of a spacecraft may be heated in LEO, shows temperature and flux dependences, i.e. the diffusion length of atomic oxygen increases with increasing temperature and beam flux. It was also demonstrated that the atomic oxygen fluence is not a primary factor of the diffusion length since the oxide growth obeys a parabolic law. The ground-based testing condition to evaluate performances of protective coatings are also discussed, based on the experimental data obtained in the experiments.
APA, Harvard, Vancouver, ISO, and other styles
7

Pagano, Damiano, Claudine Gorse, and Mario Capitelli. "Atomic wall recombination and volume negative ion production." Review of Scientific Instruments 77, no. 3 (March 2006): 03A505. http://dx.doi.org/10.1063/1.2162858.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

McMaster, T. J. "Chromosome classification by atomic force microscopy volume measurement." Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures 14, no. 2 (March 1, 1996): 1438. http://dx.doi.org/10.1116/1.589115.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Pavlik, Philip, Robert C. Rittenhouse, Martin J. Rose, and William F. Coleman. "Abstract for Volume VB, Number 1. Atomic Spectroscopy." Journal of Chemical Education 69, no. 2 (February 1992): 129. http://dx.doi.org/10.1021/ed069p129.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Ellner, M. "Partial atomic volumes of early transition metals in A10 metal-based solid solutions." International Journal of Materials Research 95, no. 5 (May 1, 2004): 345–51. http://dx.doi.org/10.1515/ijmr-2004-0073.

Full text
Abstract:
Abstract Unit-cell parameters were measured for the A10 metal-based (= Ni, Pd, Pt) solid solutions (Pearson symbol cF4, space group Fm 3 ¯ ${\bar 3}$ m, Cu type) with the A5 transition metals (V, Nb, Ta) in the whole range of homogeneity. Composition dependence both of the average atomic volume and the enthalpy of formation were investigated for the A10 metal-based solid solutions with the A5 and A6 transition metals; the partial atomic volume and the partial molar enthalpy of formation of the A5 and A6 transition were determined for the A10-rich terminal phases. Among the systems investigated, only the solid solutions Pd(Cr) and Pt(Cr) show partial atomic volume of chromium larger than the chromium atomic volume. In the nickel-based solid solution, the partial atomic volume of chromium is equal to the chromium atomic volume. The partial molar enthalpy of formation for the A5 transition metals – in the nickel, palladium and platinum-rich alloys – shows larger negative values than values evaluated for the A6 transition metals. An enlarged investigation of the data available in literature shows that both the relative volume change, resulting from the dissolving process of the early transition metals in the A10 metal-based alloys, and the negative values of the partial molar enthalpy of formation increase with decreasing number of the 3d and 4d electrons of the early transition elements. For the 5d quasihomologous transition metals, this observation is valid for the A4 . . . A6 elements as well.
APA, Harvard, Vancouver, ISO, and other styles
11

Fatemi, Mohammad, and Zahra Ghorbannezhad. "Estimation of the volume of distribution of some pharmacologically important compounds from their structural descriptors." Journal of the Serbian Chemical Society 76, no. 7 (2011): 1003–14. http://dx.doi.org/10.2298/jsc101104091f.

Full text
Abstract:
Quantitative structure-activity relationship (QSAR) approaches were used to estimate the volume of distribution (Vd) using an artificial neural network (ANN). The data set consisted of the volume of distribution of 129 pharmacologically important compounds, i.e., benzodiazepines, barbiturates, NSAIDs, tricyclic anti-depressants and some antibiotics, such as betalactams, tetracyclines and quinolones. The descriptors, which were selected by stepwise variable selection methods, were: the Moriguchi octanol-water partition coefficient; the 3D-MoRSEsignal 30, weighted by atomic van der Waals volumes; the fragmentbased polar surface area; the d COMMA2 value, weighted by atomic masses; the Geary autocorrelation, weighted by the atomic Sanderson electronegativities; the 3D-MoRSE - signal 02, weighted by atomic masses, and the Geary autocorrelation - lag 5, weighted by the atomic van der Waals volumes. These descriptors were used as inputs for developing multiple linear regressions (MLR) and artificial neural network models as linear and non-linear feature mapping techniques, respectively. The standard errors in the estimation of Vd by the MLR model were: 0.104, 0.103 and 0.076 and for the ANN model: 0.029, 0.087 and 0.082 for the training, internal and external validation test, respectively. The robustness of these models were also evaluated by the leave-5-out cross validation procedure, that gives the statistics Q2 = 0.72 for the MLR model and Q2 = 0.82 for the ANN model. Moreover, the results of the Y-randomization test revealed that there were no chance correlations among the data matrix. In conclusion, the results of this study indicate the applicability of the estimation of the Vd value of drugs from their structural molecular descriptors. Furthermore, the statistics of the developed models indicate the superiority of the ANN over the MLR model.
APA, Harvard, Vancouver, ISO, and other styles
12

Schumacher, G., R. C. Birtcher, and L. E. Rehn. "Mass density of glassy Pd80Si20 during low-temperature light ion irradiation." Journal of Materials Research 16, no. 10 (October 2001): 2788–92. http://dx.doi.org/10.1557/jmr.2001.0383.

Full text
Abstract:
Changes in mass density of amorphous Pd80Si20 were monitored in situ during irradiation with He2+ and H+ ions at temperatures below 100 K and during subsequent thermal treatment. The mass density decreased with increasing ion fluence and exponentially approached a saturation value of −1.2%, corresponding to a recombination volume of 190 atomic volumes. The initial swelling rate was 2.3 atomic volumes/displaced atom. The mass density of the irradiated material increased during subsequent thermal treatment, and the irradiation-induced decrease of the mass density recovered completely at room temperature.
APA, Harvard, Vancouver, ISO, and other styles
13

Akahama, Y., K. Kamiue, N. Okawa, S. Kawaguchi, N. Hirao, and Y. Ohishi. "Volume compression of period 4 elements: Zn, Ge, As, and Se above 200 GPa: Ordering of atomic volume by atomic number." Journal of Applied Physics 129, no. 2 (January 14, 2021): 025901. http://dx.doi.org/10.1063/5.0033721.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Landa, Alexander, John E. Klepeis, Robert E. Rudd, Kyle J. Caspersen, and David A. Young. "Analytic Binary Alloy Volume–Concentration Relations and the Deviation from Zen’s Law." Applied Sciences 11, no. 13 (July 5, 2021): 6231. http://dx.doi.org/10.3390/app11136231.

Full text
Abstract:
Alloys expand or contract as concentrations change, and the resulting relationship between atomic volume and alloy content is an important property of the solid. While a well-known approximation posits that the atomic volume varies linearly with concentration (Zen’s law), the actual variation is more complicated. Here we use the apparent size of the solute (solvent) atom and the elasticity to derive explicit analytical expressions for the atomic volume of binary solid alloys. Two approximations, continuum and terminal, are proposed. Deviations from Zen’s law are studied for 22 binary alloy systems.
APA, Harvard, Vancouver, ISO, and other styles
15

Gleiter, Herbert. "Nanoglasses: A Way to Solid Materials with Tunable Atomic Structures and Properties." Materials Science Forum 584-586 (June 2008): 41–48. http://dx.doi.org/10.4028/www.scientific.net/msf.584-586.41.

Full text
Abstract:
Recently, a new class of materials - called nanoglasses - with a glassy structure was synthesized. The novel feature of these materials is that the atomic structure in the entire volume of the material as well as the density of the material can be tuned. Nanoglasses are generated by introducing interfaces into metallic glasses on a nanometer scale. Interfaces in these nanoglasses delocalize upon annealing, so that the free volume associated with these interfaces spreads throughout the volume of the glass. This delocalization changes the atomic structure and the density of the glass throughout the volume. In fact, by controlling the spacing between the interfaces introduced into the glass as well as the degree of the delocalization (by modifying the annealing time and/or annealing temperature), the atomic structures as well as the density (and hence all structure/density dependent properties) of nanoglasses may be controlled. A comparable tuning of the atomic structure/density of crystalline materials is not conceivable, because defects in crystals do not delocalize upon annealing.
APA, Harvard, Vancouver, ISO, and other styles
16

Neethirajan, Suresh, Tamaki Hirose, Junichi Wakayama, Kazumi Tsukamoto, Hiroko Kanahara, and Shigeru Sugiyama. "Karyotype Analysis of Buckwheat Using Atomic Force Microscopy." Microscopy and Microanalysis 17, no. 4 (July 13, 2011): 572–77. http://dx.doi.org/10.1017/s1431927611000481.

Full text
Abstract:
AbstractKaryotype analysis and classification of buckwheat chromosomes were performed without chemical banding or staining using atomic force microscopy (AFM). Fagopyrum esculentum (common buckwheat) and Fagopyrum tartaricum (Tartarian buckwheat) chromosomes were isolated from root tissues using an enzymatic maceration technique and spread over a glass substrate. Air-dried chromosomes had a surface with ridges, and the height of common and tartary buckwheat were approximately 350 and 150 nm. Volumes of metaphase sets of buckwheat chromosomes were calculated using three-dimensional AFM measurements. Chromosomes were morphologically characterized by the size, volume, arm lengths, and ratios. The calculated volumes of the F. esculentum and F. tartaricum chromosomes were in the ranges of 1.08–2.09 μm3 and 0.49–0.78 μm3, respectively. The parameters such as the relative arm length, centromere position, and the chromosome volumes measured using AFM provide accurate karyomorphological classification by avoiding the subjective inconsistencies in banding patterns of conventional methods. The karyotype evolutionary trend indicates that F. esculentum is an ancient species compared to F. tartaricum. This is the first report of a cytological karyotype of buckwheat using AFM.
APA, Harvard, Vancouver, ISO, and other styles
17

Gu, Shengyang, Han Zhao, Yafei Wei, Dong Wang, and Xiankang Dou. "Atomic Oxygen SAO, AO and QBO in the Mesosphere and Lower Thermosphere Based on Measurements from SABER on TIMED during 2002–2019." Atmosphere 13, no. 4 (March 23, 2022): 517. http://dx.doi.org/10.3390/atmos13040517.

Full text
Abstract:
Using version 1.0 of TIMED/SABER nighttime O(3P) density data in the mesosphere and lower thermosphere (MLT) retrieved from 2.0 and 1.6 μm radiances, we conducted a study on the semiannual oscillation (SAO), annual oscillation (AO) and quasi-biennial oscillation (QBO) of the atomic oxygen volume mixing ratio at 96 km, from 40° S to 40° N, for 2002–2019. We first analyzed the altitude profiles of the atomic oxygen volume mixing ratio and kinetic temperature, and chose to study the daily average of the atomic oxygen volume mixing ratio at 96 km. For the analysis of SAO and AO, we fitted two sinusoidal functions with periods of 6 and 12 months to the daily mean atomic oxygen volume mixing ratio to obtain the annual and semiannual amplitude. The SAO amplitudes had two peaks of 1.68 × 10−3 and 1.63 × 10-3 at about 25° S and 25° N, and displayed a clear hemispheric symmetry. The AO amplitude increased with the latitude and showed distinct minima (valleys) of 3.36 × 10−4 around the equator, as well as a clear hemispheric asymmetry. The correlation coefficient between the atomic oxygen volume mixing ratio QBO with equatorial stratospheric QBO was higher in the tropics than the mid latitudes.
APA, Harvard, Vancouver, ISO, and other styles
18

Wiese, W. L. "Working Group 2: Atomic Transition Probabilities." Transactions of the International Astronomical Union 20, no. 1 (1988): 117–23. http://dx.doi.org/10.1017/s0251107x00007069.

Full text
Abstract:
The Data Center on Atomic Transition Probabilities at the National Bureau of Standards, Gaithersburg, Maryland, 20899, U.S.A. has continued its critical compilation work and maintains an up-to-date bibliographical data base. Work to revise and expand the existing NBS critical data compilations for the allowed and forbidden transitions in Fe-group elements, (Refs. A-D) has been completed. A single volume containing all these data for the Fe-group elements Sc to Ni is in press (Volume III of the NBS series of atomic transition probability tables) and is scheduled to be published in the near future, as a supplement to the Journal of Physical and Chemical Reference Data.
APA, Harvard, Vancouver, ISO, and other styles
19

Aicken, Fiona M., and Paul LA Popelier. "Atomic properties of selected biomolecules. Part 1. The interpretation of atomic integration errors." Canadian Journal of Chemistry 78, no. 4 (April 1, 2000): 415–26. http://dx.doi.org/10.1139/v00-026.

Full text
Abstract:
Reliable atomic properties can be obtained via the theory of "Atoms in Molecules" (AIM) via integration over a finite volume. These integrations are challenging because of the variety and complexity of the shape of the AIM atoms. In practice the integration of a large number of atoms (100-1000, sampled from many molecules) yields integration errors L(Ω) of varying magnitude. We prove that it is impossible to predict the size of an angular Gauss-Legendre grid (outside the β sphere) that guarantees a pre-set error. Hence it is incorrect to assume that a large grid (~23 000 angular grid points) will automatically yield a low L(Ω) value. The erratic relationship between the integration error and the grid size prompts a statistical interpretation of atomic integration, at a purely practical level. More importantly we have investigated the relationship between L(Ω) and seven atomic properties which include volume, energy, and the magnitudes of five electrostatic multipole moments. The electronic population (N(Ω)) and the volume (v(Ω)) of carbon is linearly correlated with L(Ω), enabling the interpolation or extrapolation of N(Ω) and v(Ω). Other properties of carbon and other atoms (N, O, and S) yield low correlation coefficients but occasionally trends can be observed. For example, we find that some properties are systematically underestimated if L(Ω) is negative. This work has led to an estimate of safe error bars of atomic properties for atoms occurring in biological molecules with reasonably sized integration grids. The most stable properties were found to be the energy and the population. Finally, we have observed that the influence of the grid orientation is less if L(Ω) is small, and that population and energy are the least affected.Key words: electron density, topology, atoms in molecules, atomic properties, amino acids.
APA, Harvard, Vancouver, ISO, and other styles
20

Kryzhevich, Dmitrij S., Aleksandr V. Korchuganov, and Konstantin P. Zolnikov. "Effect of Excess Atomic Volume on Crack Evolution in a Deformed Iron Single Crystal." Materials 14, no. 20 (October 15, 2021): 6124. http://dx.doi.org/10.3390/ma14206124.

Full text
Abstract:
This paper presents a molecular dynamics study of how the localization and transfer of excess atomic volume by structural defects affects the evolution and self-healing of nanosized cracks in bcc iron single crystals under different mechanical loading conditions at room temperature. It is shown that deformation is initially accompanied by a local growth of the atomic volume at the crack tips. The crack growth behavior depends on whether the excess atomic volume can be transferred by structural defects from the crack tips to the free surface or other interfaces. If an edge crack is oriented with respect to the loading direction so that dislocations are not emitted from its tip or only twins are emitted, then the sample undergoes a brittle-ductile fracture. The transfer of the excess atomic volume by dislocations from the crack tips prevents the opening of edge cracks and is an effective healing mechanism for nanocracks in a mechanically loaded material.
APA, Harvard, Vancouver, ISO, and other styles
21

Thompson, Avery. "Large-volume magnetic trap for studying ultracold atomic hydrogen." Scilight 2022, no. 5 (February 4, 2022): 051105. http://dx.doi.org/10.1063/10.0009561.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Das, V. M. "Atomic Genetics and Origin of the Universe- Volume-5." IOSR Journal of Research & Method in Education (IOSRJRME) 4, no. 5 (2014): 72–105. http://dx.doi.org/10.9790/7388-045172105.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

DAS, V. M. "Atomic Genetics and Origin of the Universe- Volume-8." IOSR Journal of Research & Method in Education (IOSRJRME) 4, no. 5 (2014): 57–98. http://dx.doi.org/10.9790/7388-04555798.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Smela, Elisabeth, and Nikolaj Gadegaard. "Volume Change in Polypyrrole Studied by Atomic Force Microscopy." Journal of Physical Chemistry B 105, no. 39 (October 2001): 9395–405. http://dx.doi.org/10.1021/jp004126u.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Gildenburg, V. B., V. A. Kostin, and I. A. Pavlichenko. "Resonances of surface and volume plasmons in atomic clusters." Physics of Plasmas 18, no. 9 (September 2011): 092101. http://dx.doi.org/10.1063/1.3626565.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Zhao, Feng, Yu-Kou Du, Jing-Kun Xu, and Shu-Feng Liu. "Determination of surfactant molecular volume by atomic force microscopy." Colloid Journal 68, no. 6 (December 2006): 784–87. http://dx.doi.org/10.1134/s1061933x06060172.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Flores, K. M., D. Suh, R. H. Dauskardt, P. Asoka-Kumar, P. A. Sterne, and R. H. Howell. "Characterization of Free Volume in a Bulk Metallic Glass Using Positron Annihilation Spectroscopy." Journal of Materials Research 17, no. 5 (May 2002): 1153–61. http://dx.doi.org/10.1557/jmr.2002.0171.

Full text
Abstract:
The free volume of metallic glasses has a significant effect on atomic relaxation processes, although a detailed understanding of the nature and distribution of free volume sites is currently lacking. Positron annihilation spectroscopy was employed to study free volume in a Zr–Ti–Ni–Cu–Be bulk metallic glass following plastic straining and cathodic charging with atomic hydrogen. Multiple techniques were used to show that strained samples had more open volume, while moderate hydrogen charging resulted in a free volume decrease. It was also shown that the free volume is associated with zirconium and titanium at the expense of nickel, copper, and beryllium. Plastic straining led to a slight chemical reordering.
APA, Harvard, Vancouver, ISO, and other styles
28

Račko, Dušan. "A Computational Model for Nano Scale Cavities in the Atomic Structure of Polymer Melt and Comparisons to PALS." Materials Science Forum 666 (December 2010): 15–20. http://dx.doi.org/10.4028/www.scientific.net/msf.666.15.

Full text
Abstract:
In the Present Contribution the Atomistic Structure of the Polymer Melt at 300 K Is Simulated by Means of Molecular Dynamics. the Agreement with an Experimental Density Is Obtained with a Deviation Lower than 1%. the Free Volume Is Analyzed in 1,000 Structures and 6.5 X 108 Cubic Å of Molecular Space. a Model for the Free Volume Cavities Is Proposed. in the Model the Size and Number of the Cavities Can Be Scaled by Three Parameters: Probe Radius, Cavity Depth and Cavity Threshold Volume. the Experimental Values of the Nano-Sized Cavity Volumes as Well as Ortho-Positronium Lifetimes Are Obtained, as Compared to Models with Cylindrical and Spherical Geometry. a Typical Value of the Number Density of Free Volume Cavities at 0.001 Å-3 Is Obtained. the Cavities Have Typically Elongated Shape with a Side-to-Length Ratio 1:2.
APA, Harvard, Vancouver, ISO, and other styles
29

Liu, Wei, Xiao-Gang Lu, Pascal Boulet, Marie-Christine Record, Hao Wang, and Qing-Miao Hu. "Influence of atomic mixing and atomic order on molar volume of the binary sigma phase." Intermetallics 98 (July 2018): 95–105. http://dx.doi.org/10.1016/j.intermet.2018.04.008.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Du, Bao Lei, and Qi Fei Jian. "An ISODATA Approach to the Estimation of Atomistic Definition for Continuum Stress." Materials Science Forum 909 (November 2017): 293–99. http://dx.doi.org/10.4028/www.scientific.net/msf.909.293.

Full text
Abstract:
The atomic stress tensor at a given continuum point is a spatial average value of some volume near the point. Recent progresses in multiscale modeling include the dealing of the optimal number and the size of these volumes. In this paper, we motivate the application of Iterative self-organizing data analysis technique algorithm to estimate volume numbers. The size of these space averaging volumes then could be got using Gaussian mixture model. Reduced computation complexity is offered by this method. Atomistic simulations are conducted to analyze the stress of a stone-wales defect graphene sheet to validate the method. Other multiscale values could also be determined using this method.
APA, Harvard, Vancouver, ISO, and other styles
31

Ellner, M., and I. Park. "Unit cell volumes of the silicon- and germanium-containing solid solutions based on the 3d bcc transition metals." International Journal of Materials Research 93, no. 11 (November 1, 2002): 1168–71. http://dx.doi.org/10.1515/ijmr-2002-0200.

Full text
Abstract:
Abstract The unit cell parameters and the unit cell volumes were investigated for Si- and Ge-containing solid solutions based on the 3d body-centred cubic transition metals. Excepting the high-temperature solid solution δ-Mn(Si), the solubility of Si increases with increasing number of 3d electrons. In contradiction to the large atomic volume of Si, which is substantially larger than that of the 3d transition metals, the unit cell volume of the terminal solid solutions decreases with increasing Si content. For solid solutions showing a large homogeneity range – α-Fe(Si) and α-Fe(Ge) –, the partial atomic volumes of Si and Ge were evaluated. Comparing the quasihomological Al- and Si-containing solid solutions α-Fe(Al) and α-Fe(Si), the higher bond energy occurs between atoms of Si and Fe than between Al and Fe atoms. Taking into consideration the homology of metalloids in the solid solutions α-Fe(Si) and α-Fe(Ge), the higher bond energy can also be expected between Si and Fe than between Ge and Fe atoms.
APA, Harvard, Vancouver, ISO, and other styles
32

Lousada, Cláudio M., and Pavel A. Korzhavyi. "Segregation of P and S Impurities to A Σ9 Grain Boundary in Cu." Metals 10, no. 10 (October 13, 2020): 1362. http://dx.doi.org/10.3390/met10101362.

Full text
Abstract:
The segregation of P and S to grain boundaries (GBs) in fcc Cu has implications in diverse physical-chemical properties of the material and this can be of particular high relevance when the material is employed in high performance applications. Here, we studied the segregation of P and S to the symmetric tilt Σ9 (22¯1¯) [110], 38.9° GB of fcc Cu. This GB is characterized by a variety of segregation sites within and near the GB plane, with considerable differences in both atomic site volume and coordination number and geometry. We found that the segregation energies of P and S vary considerably both with distance from the GB plane and sites within the GB plane. The segregation energy is significantly large at the GB plane but drops to almost zero at a distance of only ≈3.5 Å from this. Additionally, for each impurity there are considerable variations in energy (up to 0.6 eV) between segregation sites in the GB plane. These variations have origins both in differences in coordination number and atomic site volume with the effect of coordination number dominating. For sites with the same coordination number, up to a certain atomic site volume, a larger atomic site volume leads to a stronger segregation. After that limit in volume has been reached, a larger volume leads to weaker segregation. The fact that the segregation energy varies with such magnitude within the Σ9 GB plane may have implications in the accumulation of these impurities at these GBs in the material. Because of this, atomic-scale variations of concentration of P and S are expected to occur at the Σ9 GB center and in other GBs with similar features.
APA, Harvard, Vancouver, ISO, and other styles
33

Izrailov, E. K., and V. F. Ezhov. "A New Method for the Development of Frequency Standards for the Optical Wavelength Range of 243nm." Key Engineering Materials 295-296 (October 2005): 195–200. http://dx.doi.org/10.4028/www.scientific.net/kem.295-296.195.

Full text
Abstract:
A new method and devices for cooling of atomic hydrogen up to the values less than 100 µK aimed for the development of the optical frequency standard with significantly improved parameters (Δν/ν£1 x 10-16) is discussed. The method exploits the unique properties of atomic hydrogen such as atomic hydrogen does not condense at temperatures as low as 20µK and can not be heated by IR radiation in the absence of atom-wall collisions. Therefore, the most efficient and well-known gas cooling technique can be employed, namely, the adiabatic expansion of the volume occupied by the gas. This approach is used in a cryogenic gas expansion machine. It is suggested to use the adiabatic expansion of the volume of the magnetic atomic trap containing atomic gaseous hydrogen for embodiment of this idea.
APA, Harvard, Vancouver, ISO, and other styles
34

Kurita, Nobuaki, and Hiroshi Numakura. "Formation volume of atomic vacancies in body-centred cubic metals." International Journal of Materials Research 95, no. 10 (October 1, 2004): 876–79. http://dx.doi.org/10.1515/ijmr-2004-0164.

Full text
Abstract:
Abstract The formation volume of an atomic vacancy has been calculated by molecular statics simulation by the method originally proposed by Johnson and Brown (1962) for some transition metals of the body-centred cubic structure using several interatomic potentials. The values obtained are in satisfactory agreement with the analytic approximations proposed by Maysenhölder (1986). The formation volume is found to be correlated with Poisson’s ratio of the host crystal.
APA, Harvard, Vancouver, ISO, and other styles
35

Egami, T. "Atomic transport in amorphous metals." International Journal of Materials Research 93, no. 10 (October 1, 2002): 1071–76. http://dx.doi.org/10.1515/ijmr-2002-0183.

Full text
Abstract:
Abstract A novel mechanism of atomic transport in amorphous metallic alloys, or metallic glasses, is proposed based upon the fluctuations in the local structure. The proposed mechanism is very different from those in crystalline solids and from the free volume model, and is characterized by the bond-exchange action triggered by the local topological instability of the atomic environment. The implications of this mechanism on the liquid fragility and bulk metallic glass formation are discussed.
APA, Harvard, Vancouver, ISO, and other styles
36

Chen, Peiyu, Fabien Silly, Yingrui Zhao, and Martin R. Castell. "Transition volumes from multiply twinned particles to single crystals of supported Ag and Au nanoparticles." Applied Physics Letters 121, no. 6 (August 8, 2022): 061604. http://dx.doi.org/10.1063/5.0100156.

Full text
Abstract:
Shape changes of Ag and Au nanoparticles supported on single crystal reconstructed SrTiO3(001) and (111) substrates were investigated using scanning tunneling microscopy. Both metals nucleate as multiply twinned particles (MTPs) and transform into face-centered-cubic single crystals (SCs) beyond a critical volume. On SrTiO3(001)- c(4 × 2) the critical volumes are measured as 141 ± 51 nm3 for Ag and 107 ± 23 nm3 for Au, whereas on SrTiO3(111)–(4 × 4)+(6 × 6) the critical volumes are 53 ± 26 nm3 for Ag and 26 ± 40 nm3 for Au. A much larger transition volume was observed on SrTiO3(001)–(2 × 1), where Ag remains as MTPs up to 3400 nm3, while Au nucleates as atomic monolayers instead of MTPs. This work demonstrates the significant impact of small variations of the surface structure of the substrate on the MTP–SC transition volume.
APA, Harvard, Vancouver, ISO, and other styles
37

Pal, Snehanshu, K. Vijay Reddy, Tingting Yu, Jianwei Xiao, and Chuang Deng. "The spectrum of atomic excess free volume in grain boundaries." Journal of Materials Science 56, no. 19 (March 25, 2021): 11511–28. http://dx.doi.org/10.1007/s10853-021-06028-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Kurita, Nobuaki, and Hiroshi Numakura. "Formation volume of atomic vacancies in body-centred cubic metals." Zeitschrift für Metallkunde 95, no. 10 (October 2004): 876–79. http://dx.doi.org/10.3139/146.018045.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Wang, Yu-Guo, Wolfgang Kleemann, Theo Woike, and Rainer Pankrath. "Atomic force microscopy of domains and volume holograms inSr0.61Ba0.39Nb2O6:Ce3+." Physical Review B 61, no. 5 (February 1, 2000): 3333–36. http://dx.doi.org/10.1103/physrevb.61.3333.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Caspi, El’ad N., Haim Pinto, and Mordechai Melamud. "Local atomic volume dependence of Tb and Co magnetic moments." Journal of Applied Physics 87, no. 1 (January 2000): 416–18. http://dx.doi.org/10.1063/1.371877.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Daniels, R. S., and D. C. Wigfield. "Cold-vapor mercury atomic absorption spectrometry I. Reagent volume optimization." Science of The Total Environment 89, no. 3 (December 1989): 319–23. http://dx.doi.org/10.1016/0048-9697(89)90273-8.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Gould, Tim. "How polarizabilities and C6 coefficients actually vary with atomic volume." Journal of Chemical Physics 145, no. 8 (August 28, 2016): 084308. http://dx.doi.org/10.1063/1.4961643.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Sietsma, Jilt, and Barend J. Thijsse. "Characterization of free volume in atomic models of metallic glasses." Physical Review B 52, no. 5 (August 1, 1995): 3248–55. http://dx.doi.org/10.1103/physrevb.52.3248.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Gey, E., and K. Fiedler. "Methods in Computational Chemistry. Volume 5: Atomic and Molecular Properties." Zeitschrift für Physikalische Chemie 189, Part_1 (January 1995): 150–51. http://dx.doi.org/10.1524/zpch.1995.189.part_1.150a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Pyrz, Ryszard. "Atomistic/Continuum Transition – the Concept of Atomic Strain Tensor." Key Engineering Materials 312 (June 2006): 193–98. http://dx.doi.org/10.4028/www.scientific.net/kem.312.193.

Full text
Abstract:
A new atomic strain concept is formulated that allows calculation of continuum quantities directly within a discrete atomic (molecular) system. The concept is based on the Voronoi tessellation of the molecular system and calculation of atomic site strains, which connects continuum variables and atomic quantities when the later are averaged over a sufficiently large volume treated as a point of the continuum body. The atomic strain tensor is applied to investigate interfacial properties of polymer based nanocomposites.
APA, Harvard, Vancouver, ISO, and other styles
46

Valikova, Irina, and Andrei V. Nazarov. "Simulation of Pressure Effects on Self-Diffusion in BCC Metals." Defect and Diffusion Forum 277 (April 2008): 125–32. http://dx.doi.org/10.4028/www.scientific.net/ddf.277.125.

Full text
Abstract:
This work is devoted to the study of the point defect diffusion features in metals. In particular, we propose the model, which allows calculating activation volumes that describe the influence of pressure on the diffusion processes in solids. Our model realizes a new approach that makes it possible to self-consistently determine atomic structure near defect and constants characterizing the displacement of atoms in an elastic matrix around computational cell. Also we take into consideration that the energy of perfect system and system with a defect differently depends on the outer pressure, and this gives an addition to the values of migration and formation volumes. This addition can comprise a considerable part of activation volume. Moreover, we take into account that the atomic jump is a momentary process and so we carry out only partial relaxation of the atomic structure in the vicinity of a defect. The formation and migration energies and formation and migration volumes have been calculated for vacancies, di-vacancies and interstitials in bcc iron and tungsten using pair and many-body potentials.
APA, Harvard, Vancouver, ISO, and other styles
47

Lee, Manhee, Bongsu Kim, QHwan Kim, JongGeun Hwang, Sangmin An, and Wonho Jhe. "Viscometry of single nanoliter-volume droplets using dynamic force spectroscopy." Physical Chemistry Chemical Physics 18, no. 39 (2016): 27684–90. http://dx.doi.org/10.1039/c6cp05896e.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Nomoto, S., and S. Shoji. "Optimized atomic absorption spectrophotometry of calcium in erythrocytes." Clinical Chemistry 33, no. 11 (November 1, 1987): 2004–7. http://dx.doi.org/10.1093/clinchem/33.11.2004.

Full text
Abstract:
Abstract We sought to establish optimum conditions for measuring calcium in erythrocytes by atomic absorption spectrophotometry. The conditions we selected are as follows. Wash one volume of fresh heparin-treated packed cells once with 30 volumes of isotonic buffered saline (pH 7.4) at a temperature somewhat exceeding 25 degrees C. Dilute the washed packed cells 10-fold with 12 mmol/L hydrochloric acid, and analyze the supernate for calcium. Measure the hematocrit of the washed packed cells, then analyze an aliquot of them for calcium, using a computer-readout type of flame or a non-flame atomic absorption spectrophotometer equipped with a pyrocoated graphite tube. The temperature program is 1000 degrees C for ashing [corrected] and 1800 degrees C for the atomizing cycle. Intraday and day-to-day reproducibility of the assay was 6.55% and 8.19%, respectively, at the mean concentration of calcium in the erythrocytes of healthy adults, which is 4.30 mumol/L.
APA, Harvard, Vancouver, ISO, and other styles
49

JOUDEH, B. R., M. K. AL-SUGHEIR, and H. B. GHASSIB. "SPIN-POLARIZED ATOMIC HYDROGEN IN THE STATIC FLUCTUATION APPROXIMATION." International Journal of Modern Physics B 19, no. 26 (October 20, 2005): 3985–4008. http://dx.doi.org/10.1142/s0217979205032474.

Full text
Abstract:
In this paper we use the so-called static fluctuation approximation (SFA) to calculate the thermodynamic properties of spin-polarized atomic hydrogen. This approximation is based on the replacement of the square of the local-field operator with its mean value. A closed set of nonlinear integral equations is derived for neutral many-bosonic systems. This set is solved numerically by an iteration method for two triplet-state potentials: a Morse- and Silvera-type potentials. It is found that the mean internal energy per unit volume, the pressure, the entropy per unit volume, and the specific heat per unit volume increase with temperature and decrease with spin polarization in the low-temperature region (<8 mK ); whereas the condensate fraction increases with an applied magnetic field. It is also found that these quantities are nearly independent of spin polarization for high temperatures (>0.1 K ), and that they are independent of the number density up to 10-3Å-3 in the low-temperature region.
APA, Harvard, Vancouver, ISO, and other styles
50

VARÃO, RÉGIS. "Center foliation: absolute continuity, disintegration and rigidity." Ergodic Theory and Dynamical Systems 36, no. 1 (August 11, 2014): 256–75. http://dx.doi.org/10.1017/etds.2014.53.

Full text
Abstract:
In this paper we address the issues of absolute continuity for the center foliation, as well as the disintegration on the non-absolute continuous case and rigidity of volume-preserving partially hyperbolic diffeomorphisms isotopic to a linear Anosov automorphism on $\mathbb{T}^{3}$. It is shown that the disintegration of volume on center leaves for these diffeomorphisms may be neither atomic nor Lebesgue, in contrast to the dichotomy (Lebesgue or atomic) obtained by Avila, Viana and Wilkinson [Absolute continuity, Lyapunov exponents and rigidity I: Geodesic flows. Preprint, 2012, arXiv:1110.2365v2] for perturbations of time-one of geodesic flow. In the case of atomic disintegration of volume on the center leaves of an Anosov diffeomorphism on $\mathbb{T}^{3}$, we show that it has to be one atom per leaf. Moreover, we show that not even a $C^{1}$ center foliation implies a rigidity result. However, for a volume-preserving partially hyperbolic diffeomorphism isotopic to a linear Anosov automorphism, assuming the center foliation is $C^{1}$ and transversely absolutely continuous with bounded Jacobians, we obtain smooth conjugacy to its linearization.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography