Journal articles on the topic 'K Surface'

To see the other types of publications on this topic, follow the link: K Surface.

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 'K Surface.'

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

Nasiedkin, D. B., M. O. Nazarchuk, A. G. Grebenyuk, L. F. Sharanda, and Yu V. Plyuto. "Quantum chemical simulation of MoO3 dispergation on hydroxylated SiO2 surface." Surface 13(28) (December 30, 2021): 75–83. http://dx.doi.org/10.15407/surface.2021.13.075.

Full text
Abstract:
Метою даної роботи є оцінка енергетичної сприятливості утворення різних молібдатних груп (≡Si‑O‑)2Mo(=O)2 та =Si(‑O‑)2Mo(=O)2 під час термічно ініційованого диспергування MoO3 на гідроксильованій поверхні SiO2. Для цього було здійснено квантовохімічне моделювання реакції O12Si10(OH)16 + MoO3 = O12Si10(OH)14O2MoO2 + H2O в температурному інтервалі 300–1100 K із використанням обмеженого методу Хартрі-Фока (наближення ЛКАО) з валентним базисом SBKJC (Stevens-Basch-Krauss-Jasien-Cundari). Кластер O12Si10(OH)16, який являє собою структурний фрагмент кристала β‑кристобаліту, був використаний як модель високогідроксильованої поверхні кремнезему. Ми розглянули дві структури молібдатних груп (≡Si‑O‑)2Mo(=O)2, прикріплених до кремнеземного кластера O12Si10(OH)16 через силанольні групи. Молібдатні групи (Etot ‑584.60147 Hartree), прикріплені до кремнеземного кластера через віддалені силанольні групи, виявляються більш енергетично вигідними, ніж молібдатні групи (Etot ‑584.56565 Hartree), прикріплені до кремнеземного кластера через сусідні силанольні групи. Енергія молібдатних груп =Si(‑O‑)2Mo(=O)2 (Etot ‑584.48399 Hartree), прикріплених до кремнеземного кластера O12Si10(OH)16 через силандіольні групи, менш енергетично вигідні в порівнянні з подібними групами, прикріпленими через силанольні групи, через більше напруження кута між зв’язками. Знайдено, що реакція O12Si10(OH)16 + MoO3 = O12Si10(OH)14O2MoO2 + H2O в температурному інтервалі 300–1100 K, змодельована шляхом квантовохімічних розрахунків, свідчить, що процес диспергування MoO3 на гідроксильованій поверхні SiO2 є енергетично вигідним. Експ The aim of the present work is to evaluate the energetic favourability of the formation of different molybdate species (≡Si‑O‑)2Mo(=O)2 and =Si(‑O‑)2Mo(=O)2 during the thermally induced MoO3 dispergation on hydroxylated SiO2 surface. In order to do this a quantum chemical modelling of the reaction O12Si10(OH)16 + MoO3 = O12Si10(OH)14O2MoO2 + H2O within the temperature interval of 300–1100 K was undertaken using the Restricted Hartree-Fock method (the LCAO approximation) with the SBKJC (Stevens-Basch-Krauss-Jasien-Cundari) valence basis set. The cluster O12Si10(OH)16 which represents a structural fragment of a β‑cristobalite crystal was used in this work as a model of highly hydroxylated silica surface. We considered two structures of molybdate (≡Si‑O‑)2Mo(=O)2 species attached to O12Si10(OH)16 silica cluster via silanol groups. Molybdate species (Etot ‑584.60147 Hartree) attached to silica cluster via distant silanols appeared more energetically favourable than molybdate species (Etot ‑584.56565 Hartree) attached to silica cluster via nearby silanols. The energy of molybdate =Si(‑O‑)2Mo(=O)2 species (Etot ‑584.48399 Hartree) attached to O12Si10(OH)16 silica cluster via silanediol group is less favourable energetically in comparison with those attached via silanol groups because of higher bond angle straining. The reaction O12Si10(OH)16 + MoO3 = O12Si10(OH)14O2MoO2 + H2O in the temperature interval of 300–1100 K which simulates by quantum chemical calculations the dispergation of MoO3 on hydroxylated SiO2 surface was found to be energetically favourable. The experimentally optimised temperature of ca. 800 K required for dispergation of MoO3 on hydroxylated SiO2 surface is determined by MoO3 evaporation and transportation via the gas phase. ериментальна оптимальна температура (близько 800 K), потрібна для диспергування MoO3 на гідроксильованій поверхні SiO2, визначається випаровуванням та перенесенням MoO3 в газовій фазі.
APA, Harvard, Vancouver, ISO, and other styles
2

Garbuz, V. V., V. A. Petrova, T. A. Silinskaya, T. F. Lobunets, O. I. Bykov, V. B. Muratov, T. M. Terentyeva, et al. "Specific surface area, crystallite size and thermokinetic of oxide formation γ → α-Al2O3 nano powders at 570 – 1470 K." Surface 12(27) (December 30, 2020): 146–52. http://dx.doi.org/10.15407/surface.2020.12.146.

Full text
Abstract:
Powders where the γ≈α-Al2O3-nano phases are the priority precursors for catalysts for heterogeneous catalysis with the maximum content of surface 5-coordinated Al centers for Pt attachment. Hydrogenated nano powders (~8 nm) of γ-, γ '-, θ-, κ-Al2O3 soluble in hydrochloric acid were obtained from the processing of aluminum boride powders with an icosahedral structure. Samples, which underwent a step-by-step and single heating of 50-100K heat treatment for 2 hours at temperatures of 570-1470K, were received in quantity of 34. The specific surface area of SВET, m2g-1 was measured by the thermal nitrogen desorption express method of gas chromatography through the GC-1 device. X-ray (phase and coherent), fluorescence and phase chemical-analytical evaluation of the samples were performed. The thermokinetic characteristics of the processes are calculated using the exponential Arrhenius law. Dimensional characteristics of crystallites (10.4-48 nm); specific surface area of powders (213-8.6 m2g-1, SВET); thermokinetic parameters of α-Al2O3 crystallite growth process (V α-Al2O3 - 1.44 10-3 - 6.67 10-3 nm s-1; E α-Al2O3 = 38.7±2.1kJ mol-1; A0 = 0.16±0.0 s-1 along the temperature line 1220-1470K were determined and calculated. The process of dehydration of two OH-groups occurs in the region 570-720K Ea H2O ↑ = 30.5 ± 0.5 kJ mol-1 A0 = 1.33±0.3 s-1. The last group of OH at temperatures of 820 -1070К and a rate of 2.13 10-4 - 4.93 10-4 mol s-1 Ea H2O ↑ = 13.2 ± 0.8 kJ mol-1 A0 = 16.9 ± 0.9 s-1. The activation energy of the phase transition is Ea., γ → α-Al2O3 = 23.9 ± 1.0 kJ mol-1 A0 = 2.01 ± 0.72 s-1 (770-970K) and Ea., γ → α-Al2O3 = 83.5 ± 0.8 kJ mol-1 A0 =(2,05±0,95) 103 s-1 (1070-1170K). It agrees well with the known heat of conversion Eа, γ→α-Al2O3 = 85 kJ mol-1. The TK of γ≈α-Al2O3-nano phases is at 1170K.
APA, Harvard, Vancouver, ISO, and other styles
3

Han, Sang-Eon. "The k-fundamental group of a closed k-surface." Information Sciences 177, no. 18 (September 2007): 3731–48. http://dx.doi.org/10.1016/j.ins.2007.02.031.

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

Han, Sang-Eon. "Fixed Point Theory for Digital k-Surfaces and Some Remarks on the Euler Characteristics of Digital Closed Surfaces." Mathematics 7, no. 12 (December 16, 2019): 1244. http://dx.doi.org/10.3390/math7121244.

Full text
Abstract:
The present paper studies the fixed point property (FPP) for closed k-surfaces. We also intensively study Euler characteristics of a closed k-surface and a connected sum of closed k-surfaces. Furthermore, we explore some relationships between the FPP and Euler characteristics of closed k-surfaces. After explaining how to define the Euler characteristic of a closed k-surface more precisely, we confirm a certain consistency of the Euler characteristic of a closed k-surface and a continuous analog of it. In proceeding with this work, for a simple closed k-surface in Z 3 , say S k , we can see that both the minimal 26-adjacency neighborhood of a point x ∈ S k , denoted by M k ( x ) , and the geometric realization of it in R 3 , denoted by D k ( x ) , play important roles in both digital surface theory and fixed point theory. Moreover, we prove that the simple closed 18-surfaces M S S 18 and M S S 18 ′ do not have the almost fixed point property (AFPP). Consequently, we conclude that the triviality or the non-triviality of the Euler characteristics of simple closed k-surfaces have no relationships with the FPP in digital topology. Using this fact, we correct many errors in many papers written by L. Boxer et al.
APA, Harvard, Vancouver, ISO, and other styles
5

Bogatyrov, V. M., M. V. Borysenko, M. V. Galaburda, and O. I. Oranska. "Synthesis and properties of nanocomposites based on zinc phosphate and fumed silica." Surface 12(27) (December 30, 2020): 179–92. http://dx.doi.org/10.15407/surface.2020.12.179.

Full text
Abstract:
The aim of the work was to synthesize nanocomposites based on pyrogenic silica and zinc phosphate by a simple method without using a large amount of solvent and to study the characteristics and properties of the obtained materials. The dual systems of zinc phosphate/pyrogenic silica with the different ratio of components were synthesized via mechanical grinding in a porcelain drum ball mill of fumed silica (Orysyl A-380), zinc acetate (Zn(CH3COO)2·2H2O) and phosphoric acid with distilled water, followed by air-drying in an oven at 125 °C (2 h) and calcination in a muffle oven at 450 °C for 2 h. The zinc phosphate content was 0.1, 0.2, and 0.3 mmol per 1 g of SiO2. The control sample (ZP-K) was synthesized by thermal treatment of the precipitate, formed after mixing on a magnetic stirrer an aqueous solution of zinc acetate with the addition dropwise of phosphoric acid, without the use of SiO2. X-ray diffraction studies of the nanocomposites confirmed the formation of the crystalline phase of Zn3(PO4)2·4H2O (orthorhombic modification) both in the silica-containing and control ZP-K samples after air drying at 125 °C, while heat treatment at 450 °C leaded to the formation of the anhydrous monoclinic Zn3(PO4)2 phase. The content of the zinc phosphate in the dual composites was 0.1, 0.2, and 0.3 mmol per 1 g of SiO2. The IR spectra of the nanocomposites indicated the presence of absorption bands in the range of 3760-3600 cm-1, which were attributed to the unequal structural ‒OH groups of silicon and phosphorus atoms. It was found that the presence of zinc phosphate on the SiO2 surface does not cause the chemical interaction with silica during heat treatment of composites in air even at 900-1000 °C. It was shown that the ability of Zn3(PO4)2/SiO2 composites to adsorb water vapor decreases with increasing amount of modifying compound. The effect of the obtained phosphorus-containing nanocomposite on the thermal stability of an alkyd polymer matrix was considered.
APA, Harvard, Vancouver, ISO, and other styles
6

Pal, S. K., K. Takimoto, E. Aizenman, and E. S. Levitan. "Apoptotic surface delivery of K+ channels." Cell Death & Differentiation 13, no. 4 (November 4, 2005): 661–67. http://dx.doi.org/10.1038/sj.cdd.4401792.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

McAllister, I. W. "Surface current density K: an introduction." IEEE Transactions on Electrical Insulation 26, no. 3 (June 1991): 416–17. http://dx.doi.org/10.1109/14.85112.

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

Ferris, Daniel P., and Claire T. Farley. "Interaction of leg stiffness and surface stiffness during human hopping." Journal of Applied Physiology 82, no. 1 (January 1, 1997): 15–22. http://dx.doi.org/10.1152/jappl.1997.82.1.15.

Full text
Abstract:
Ferris, Daniel P., and Claire T. Farley. Interaction of leg stiffness and surface stiffness during human hopping. J. Appl. Physiol. 82(1): 15–22, 1997.—When mammals run, the overall musculoskeletal system behaves as a single linear “leg spring.” We used force platform and kinematic measurements to determine whether leg spring stiffness ( k leg) is adjusted to accommodate changes in surface stiffness ( k surf) when humans hop in place, a good experimental model for examining adjustments to k leg in bouncing gaits. We found that k leg was greatly increased to accommodate surfaces of lower stiffnesses. The series combination of k leg and k surf[total stiffness ( k tot)] was independent of k surf at a given hopping frequency. For example, when humans hopped at a frequency of 2 Hz, they tripled their k leg on the least stiff surface ( k surf = 26.1 kN/m; k leg = 53.3 kN/m) compared with the most stiff surface ( k surf = 35,000 kN/m; k leg = 17.8 kN/m). Values for k tot were not significantly different on the least stiff surface (16.7 kN/m) and the most stiff surface (17.8 kN/m). Because of the k leg adjustment, many aspects of the hopping mechanics (e.g., ground-contact time and center of mass vertical displacement) remained remarkably similar despite a >1,000-fold change in k surf. This study provides insight into how k leg adjustments can allow similar locomotion mechanics on the variety of terrains encountered by runners in the natural world.
APA, Harvard, Vancouver, ISO, and other styles
9

Ungerer, Marietjie J., David Santos-Carballal, Abdelaziz Cadi-Essadek, Cornelia G. C. E. van Sittert, and Nora H. de Leeuw. "Interaction of SO2 with the Platinum (001), (011), and (111) Surfaces: A DFT Study." Catalysts 10, no. 5 (May 18, 2020): 558. http://dx.doi.org/10.3390/catal10050558.

Full text
Abstract:
Given the importance of SO2 as a pollutant species in the environment and its role in the hybrid sulphur (HyS) cycle for hydrogen production, we carried out a density functional theory study of its interaction with the Pt (001), (011), and (111) surfaces. First, we investigated the adsorption of a single SO2 molecule on the three Pt surfaces. On both the (001) and (111) surfaces, the SO2 had a S,O-bonded geometry, while on the (011) surface, it had a co-pyramidal and bridge geometry. The largest adsorption energy was obtained on the (001) surface (Eads = −2.47 eV), followed by the (011) surface (Eads = −2.39 and −2.28 eV for co-pyramidal and bridge geometries, respectively) and the (111) surface (Eads = −1.85 eV). When the surface coverage was increased up to a monolayer, we noted an increase of Eads/SO2 for all the surfaces, but the (001) surface remained the most favourable overall for SO2 adsorption. On the (111) surface, we found that when the surface coverage was θ > 0.78, two neighbouring SO2 molecules reacted to form SO and SO3. Considering the experimental conditions, we observed that the highest coverage in terms of the number of SO2 molecules per metal surface area was (111) > (001) > (011). As expected, when the temperature increased, the surface coverage decreased on all the surfaces, and gradual desorption of SO2 would occur above 500 K. Total desorption occurred at temperatures higher than 700 K for the (011) and (111) surfaces. It was seen that at 0 and 800 K, only the (001) and (111) surfaces were expressed in the morphology, but at 298 and 400 K, the (011) surface was present as well. Taking into account these data and those from a previous paper on water adsorption on Pt, it was evident that at temperatures between 400 and 450 K, where the HyS cycle operates, most of the water would desorb from the surface, thereby increasing the SO2 concentration, which in turn may lead to sulphur poisoning of the catalyst.
APA, Harvard, Vancouver, ISO, and other styles
10

SPELLER, S., M. SCHLEBERGER, H. FRANKE, C. MÜLLER, and W. HEILAND. "SURFACE MELTING AND SURFACE ROUGHENING OF Pb(110) STUDIED BY LOW ENERGY ION SCATTERING." Modern Physics Letters B 08, no. 08n09 (April 20, 1994): 491–503. http://dx.doi.org/10.1142/s0217984994000522.

Full text
Abstract:
The Pb(110) surface undergoes two phase transitions. At about 400 K a roughening transition is observed. At 580 K, i.e. about 20 K below the bulk melting point, surface melting is found, The surface develops point defects at rather low temperatures. The roughening is connected with the generation of steps or the reduction of terrace size. There is also evidence for anisotropy of the roughening transition. Low energy ion scattering experiments in the temperature range from 160 to 590 K are used to study the structural changes of the Pb(110) surface.
APA, Harvard, Vancouver, ISO, and other styles
11

Zhao, Gang, Qiang Chang, Xia Zhang, Donghui Quan, Yong Zhang, and Xiao-Hu Li. "Effect of surface H2 on molecular hydrogen formation on interstellar grains." Monthly Notices of the Royal Astronomical Society 512, no. 3 (March 11, 2022): 3137–48. http://dx.doi.org/10.1093/mnras/stac655.

Full text
Abstract:
ABSTRACT We investigate how the existence of hydrogen molecules on grain surfaces may affect H2 formation efficiency in diffuse and translucent clouds. Hydrogen molecules are able to reduce the desorption energy of H atoms on grain surfaces in models. The detailed microscopic Monte Carlo method is used to perform model simulations. We found that the impact of the existence of H2 on H2 formation efficiency strongly depends on the diffusion barriers of H2 on grain surfaces. Diffuse cloud models that do not consider surface H2 predict that H atom recombination efficiency is above 0.5 over a grain temperature (T) range 10 and 14 K. The adopted H2 diffusion barriers in diffuse cloud models that consider surface H2 are 80${{\ \rm per\ cent}}$ H2 desorption energies so that H2 can be trapped in stronger binding sites. Depending on model parameters, these diffuse cloud models predict that the recombination efficiency is between nearly 0 and 0.5 at 10 ≤T≤ 14 K. Translucent cloud model results show that H2 formation efficiency is not affected by the existence of surface H2 if the adopted average H2 diffusion barrier on grain surfaces is low (194 K) so that H2 can diffuse rapidly on grain surfaces. However, the recombination efficiency can drop to below 0.002 atT≥ 10 K if higher average H2 diffusion barrier is used (255 K) in translucent cloud models.
APA, Harvard, Vancouver, ISO, and other styles
12

Itchkawitz, B. S., A. P. Baddorf, H. L. Davis, and E. W. Plummer. "Shear displacement of the K(110) surface." Physical Review Letters 68, no. 16 (April 20, 1992): 2488–91. http://dx.doi.org/10.1103/physrevlett.68.2488.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

Roushon, S. K. "K–theory of virtually poly-surface groups." Algebraic & Geometric Topology 3, no. 1 (February 8, 2003): 103–16. http://dx.doi.org/10.2140/agt.2003.3.103.

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

Chiarello, G., A. Cupolillo, A. Amoddeo, L. S. Caputi, L. Papagno, and E. Colavita. "K-induced surface vibrations on Ni(111)." Physical Review B 52, no. 7 (August 15, 1995): 4752–55. http://dx.doi.org/10.1103/physrevb.52.4752.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Garcias, F., M. Barranco, M. Pi, and J. Navarro. "Dipole surface plasmon in K+N clusters." Solid State Communications 84, no. 9 (December 1992): 905–9. http://dx.doi.org/10.1016/0038-1098(92)90456-j.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Lu, Hongliang, and David G. L. Wang. "Surface embedding of (n,k)-extendable graphs." Discrete Applied Mathematics 179 (December 2014): 163–73. http://dx.doi.org/10.1016/j.dam.2014.08.003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

HERMANN, K. "GENERAL METHOD TO SELECT CONVENIENT LATTICE VECTORS ON A SURFACE WITH ARBITRARY MILLER INDICES." Surface Review and Letters 04, no. 05 (October 1997): 1063–69. http://dx.doi.org/10.1142/s0218625x97001310.

Full text
Abstract:
Many theoretical methods dealing with electronic and structural properties of single crystal surfaces rely on a convenient description of the surface and bulk periodicity at the same time. This can be achieved by using surface adapted lattice vectors [Formula: see text], where vectors [Formula: see text] define (h k l) net planes parallel to the surface while [Formula: see text] connects adjacent (h k l) net planes. For selected low index (h k l) surfaces of common crystals the construction of appropriate lattice vectors may be trivial. However, the general problem of determining a lattice basis adapted to a surface orientation which is described by Miller indices (h k l) in a general crystal lattice is more involved. In this paper we show that such bases, [Formula: see text], can be uniquely determined by linear transformations from the bulk lattice basis [Formula: see text]. The transformations depend on Miller indices (h k l) but not on the lattice type and can be quantified by number-theoretical methods. Thus, they are numerically stable and can be easily implemented in computational algorithms dealing with surfaces of most general crystals.
APA, Harvard, Vancouver, ISO, and other styles
18

Ee Von, Lau, Lee Jun Rong, and Mohamed Ismail Harun. "Forced Convective Heat Transfer of Ionic Wind on Different Surface Conditions." Applied Mechanics and Materials 793 (September 2015): 445–49. http://dx.doi.org/10.4028/www.scientific.net/amm.793.445.

Full text
Abstract:
The heat transfer enhancement of the forced convection due to ionic wind over different surface conditions including a smooth, rough ruface and a source array of rectangular blocks surface (representing electronic components) was studied. Under laminar flow, the highest heat transfer rate of 0.0736 W/m2.K per minute was observed for the source array surface. The average heat transfer coefficient during steady state of ionic cooling on smooth, rough and source array surfaces were observed to be 19.144 W/m2.K, 18.736 W/m2.K and 21.126 W/m2.K respectively. The heat transfer properties of ionic wind are similar to moving air, generating high heat transfer coefficient and Nusselt number on source array surface due to recirculation eddies.
APA, Harvard, Vancouver, ISO, and other styles
19

Maehara, Y., H. Kawanowa, and Y. Gotoh. "Surface structures of K on Mo(110) surface investigated by RHEED." Surface Science 600, no. 18 (September 2006): 3575–80. http://dx.doi.org/10.1016/j.susc.2006.02.051.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Franco Jr., A., and S. G. Roberts. "Surface mechanical analyses by Hertzian indentation." Cerâmica 50, no. 314 (June 2004): 94–108. http://dx.doi.org/10.1590/s0366-69132004000200005.

Full text
Abstract:
Hertzian indentation technique was used to measure surface flaw sizes on polished dense polycrystalline alumina specimens with grain sizes G = 1.2, 3.8 and 14.1 µm. Two surfaces finishes were studied: well-polished (Syton) and coarse-polished (45 µm diamond paste). Flaw sizes depended on the surface finish and increased with increasing grain size. Fracture toughness (K Ic) for each material (relating to the propagation of flaws of a few µm depth) was determined from the minimum fracture load in a series of Hertzian tests. K Ic values were 3.58, 3.45 and 2.96 MPam½ for G=1.2, 3.8 and 14.1 µm, respectively. Fracture toughness values were also determined by Vickers indentation over a range of loads; the K Ic values determined from the Hertzian tests were consistent with the trends in K Ic with crack size from the Vickers indentation tests.
APA, Harvard, Vancouver, ISO, and other styles
21

XIE, X. P., M. H. LIANG, Z. M. CHOO, and S. LI. "A COMPARATIVE SIMULATION STUDY OF SILICON (001) SURFACE RECONSTRUCTION USING DIFFERENT INTERATOMIC POTENTIALS." Surface Review and Letters 08, no. 05 (October 2001): 471–75. http://dx.doi.org/10.1142/s0218625x01001397.

Full text
Abstract:
We have performed a comparative study of Si (001) surface reconstruction employing molecular dynamics simulation using the interatomic potentials of Stillinger–Weber, Tersoff and Bazant–Kaxiras. Simulations were carried out for temperatures at 300 K and 1000 K using each of these three potentials. At 300 K, the three potentials were found to generate surface features comprising mainly the simple (2 × 1) reconstruction. At 1000 K, more complex reconstruction similar to the p (2 × 2) and c (2 × 2) patterns was observed on the surfaces of Stillinger–Weber and Tersoff crystals while the surface generated on Bazant–Kaxiras crystal is characterized by disorderliness with no identifiable pattern of reconstruction.
APA, Harvard, Vancouver, ISO, and other styles
22

PAPAGEORGOPOULOS, C. A., M. KAMARATOS, D. C. PAPAGEORGOPOULOS, W. JAEGERMANN, C. PETTENKOFER, and O. HENRION. "ADSORPTION OF Br2 ON Na-INTERCALATED LAYERED COMPOUNDS: BROMINE-INDUCED DEINTERCALATION." Surface Review and Letters 04, no. 02 (April 1997): 237–44. http://dx.doi.org/10.1142/s0218625x97000237.

Full text
Abstract:
This work refers to a study of the behavior of Br2 adsorbed on Na-intercalated 1T-TaSe2(0001) and TiSe2(0001) surfaces at 100 K and during subsequent warming up to 300 K. The investigation was performed in UHV with LEED and photoemission with synchrotron radiation measurements. At 100 K, the bromine forms molecular multilayers on Na/TaSe2 . The Br2 layer, close to the substrate, interacts weakly with Na near the surface, forming [Formula: see text]. At 300 K, part of the Br 2 overlayer desorbs while the remainder on both 1T-TaSe2 and Tise2 surfaces interacts strongly with Na, leading to deintercalation of Na to the surface in the tendency to form NaBr. The formation of NaBr overlayers causes the transition of the incommensurate [Formula: see text] surface structure of 1T-TaSe2 (0001) to 1×1. ESD from the Br/Na/1T-TaSe2 surface causes the restoration of the [Formula: see text].
APA, Harvard, Vancouver, ISO, and other styles
23

Coupaye-Gerard, B., J. B. Zuckerman, P. Duncan, A. Bortnik, D. I. Avery, S. A. Ernst, and T. R. Kleyman. "Delivery of newly synthesized Na(+)-K(+)-ATPase to the plasma membrane of A6 epithelia." American Journal of Physiology-Cell Physiology 272, no. 6 (June 1, 1997): C1781—C1789. http://dx.doi.org/10.1152/ajpcell.1997.272.6.c1781.

Full text
Abstract:
Na(+)-K(+)-ATPase is localized to the basolateral cell surface of most epithelial cells. Conflicting results regarding the intracellular trafficking of Na(+)-K(+)-ATPase in Madin-Darby canine kidney cells have been reported, with delivery to both apical and basolateral membranes or exclusively to the basolateral cell surface. We examined the delivery and steady-state distribution of Na(+)-K(+)-ATPase in the amphibian epithelial cell line A6 using an antibody raised against Na(+)-K(+)-ATPase alpha-subunit and sulfo-N-hydroxysuccinimidobiotin to tag cell surface proteins. The steady-state distribution of the Na(+)-K(+)-ATPase was basolateral, as confirmed by immunocytochemistry. Delivery of newly synthesized Na(+)-K(+)-ATPase to the cell surface was examined using [35S]methionine and [35S]cysteine in a pulse-chase protocol. After a 20-min pulse, the alpha-subunit and core glycosylated beta-subunit were present at both apical and basolateral cell surfaces. The alpha-subunit and core glycosylated beta-subunit delivered to the apical cell surface were degraded within 2 h. Mature alpha/beta-heterodimer was found almost exclusively at the basolateral surface after a 1- to 24-h chase. These data suggest that immature Na(+)-K(+)-ATPase alpha-subunit and core glycosylated beta-subunits are not retained in the endoplasmic reticulum of A6 cells and apparently lack sorting signals. Mature Na(+)-K(+)-ATPase is targeted to the basolateral surface, suggesting that basolateral targeting of the protein is conformation dependent.
APA, Harvard, Vancouver, ISO, and other styles
24

Hundekar, Prateek, Swastik Basu, Xiulin Fan, Lu Li, Anthony Yoshimura, Tushar Gupta, Varun Sarbada, et al. "In situ healing of dendrites in a potassium metal battery." Proceedings of the National Academy of Sciences 117, no. 11 (March 2, 2020): 5588–94. http://dx.doi.org/10.1073/pnas.1915470117.

Full text
Abstract:
The use of potassium (K) metal anodes could result in high-performance K-ion batteries that offer a sustainable and low-cost alternative to lithium (Li)-ion technology. However, formation of dendrites on such K-metal surfaces is inevitable, which prevents their utilization. Here, we report that K dendrites can be healed in situ in a K-metal battery. The healing is triggered by current-controlled, self-heating at the electrolyte/dendrite interface, which causes migration of surface atoms away from the dendrite tips, thereby smoothening the dendritic surface. We discover that this process is strikingly more efficient for K as compared to Li metal. We show that the reason for this is the far greater mobility of surface atoms in K relative to Li metal, which enables dendrite healing to take place at an order-of-magnitude lower current density. We demonstrate that the K-metal anode can be coupled with a potassium cobalt oxide cathode to achieve dendrite healing in a practical full-cell device.
APA, Harvard, Vancouver, ISO, and other styles
25

Seguin, Brian, Yi-chao Chen, and Eliot Fried. "Bridging the gap between rectifying developables and tangent developables: a family of developable surfaces associated with a space curve." Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 477, no. 2246 (February 2021): 20200617. http://dx.doi.org/10.1098/rspa.2020.0617.

Full text
Abstract:
There are two familiar constructions of a developable surface from a space curve. The tangent developable is a ruled surface for which the rulings are tangent to the curve at each point and relative to this surface the absolute value of the geodesic curvature κ g of the curve equals the curvature κ . The alternative construction is the rectifying developable. The geodesic curvature of the curve relative to any such surface vanishes. We show that there is a family of developable surfaces that can be generated from a curve, one surface for each function k that is defined on the curve and satisfies | k | ≤ κ , and that the geodesic curvature of the curve relative to each such constructed surface satisfies κ g = k .
APA, Harvard, Vancouver, ISO, and other styles
26

Nohman, A. K. H., and M. I. Zaki. "Nitrogen Adsorption and Surface Texture of High-Area Manganese Oxides Derived from Potassium-Contaminated NH4MnO4." Adsorption Science & Technology 11, no. 1 (March 1994): 1–14. http://dx.doi.org/10.1177/026361749401100101.

Full text
Abstract:
Nitrogen sorption isotherms have been determined at 77 K on pre-characterised MnOx bulk samples obtained by calcination at 423–873 K of K+-contaminated NH4MnO4. The isotherms have been analysed for estimating the surface area and pore structure of the test samples adopting the αs-method and standard adsorption data determined on a local reference material (calcination product of the permanganate at 1273 K for 20 h). The most prominent result of the present investigation is that the presence of K+ leads to the formation of a channel-structured KMn8O16 bulk phase at 573 K and the following consequent textural modifications: (i) a considerable increase in the surface area to 205 m2/g, (ii) the construction of a uniform pore system consisting of slit-shaped pores with narrow entrances, and (iii) the evolution of energetically uniform internal surfaces of considerable area (= 92 m2/g). The results also reveal that the test surfaces generally exhibit a low heat towards N2 adsorption.
APA, Harvard, Vancouver, ISO, and other styles
27

Marlow, F. "Second harmonic generation at metal surfaces: influence of surface charge." Surface Science Letters 254, no. 1-3 (August 1991): L493—L496. http://dx.doi.org/10.1016/0167-2584(91)90016-k.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Zolotarenko, Ol D., E. P. Rudakova, N. Y. Akhanova, M. Ualkhanova, An D. Zolotarenko, D. V. Shchur, M. T. Gabdullin, et al. "Synthesis of carbon nanostructures using cheap grades of graphite." SURFACE 14(29) (December 30, 2022): 113–31. http://dx.doi.org/10.15407/surface.2022.14.113.

Full text
Abstract:
In the work, carbon nanostructures (CNS) were synthesized on a plasma chemical plant using graphite electrodes SIGE (Special Impregnated Graphite Electrodes) and FGDG-7 (Fine-grained dense graphite) in a helium environment. In the experiments, it was established that graphite electrodes of the SIGE brand are suitable for the synthesis of CNS by the electric arc plasma chemical method. In addition, the experiments indicate that SIGE graphite in electric arc synthesis in a gas environment allows the creation of centimeter composite rods (deposits), where the core consists of graphene sheets rolled into nanotubes that can withstand extremely high temperatures (>4000 K). Studies using scanning microscopy have shown that the synthetic deposit of SIGE graphite can be divided into blocks, which is important for its use in high voltage stations because it is possible to prepare deposits of the required length without mechanical impact and without violating the integrity of its structure. The structure of the synthesized carbon materials was studied by scanning and transmission electron microscopy and it was shown that carbon nanotubes are formed during the evaporation of SIGE brand graphite even without the use of a catalyst. Experiments have confirmed that the mass yield of wall fullerene-containing carbon black during the evaporation of SIGE grade graphite significantly exceeds the results obtained during the evaporation of FGDG-7 grade graphite electrodes. Such results make SIGE graphite more productive for the synthesis of expensive carbon nanoproducts (fullerenes and fullerene-like structures) by the electric arc method. It was also recorded that during the synthesis of carbon nanostructures, single-walled carbon nanotubes are formed, which have a positive charge and are deposited in the form of a core on the surface of the cathode electrode under the action of an electromagnetic field.
APA, Harvard, Vancouver, ISO, and other styles
29

Wang, Xinye, Min Chen, Changqi Liu, Changsheng Bu, Jubing Zhang, Chuanwen Zhao, and Yaji Huang. "Typical Gaseous Semi-Volatile Metals Adsorption by Meta-Kaolinite: A DFT Study." International Journal of Environmental Research and Public Health 15, no. 10 (September 30, 2018): 2154. http://dx.doi.org/10.3390/ijerph15102154.

Full text
Abstract:
Kaolinite can be used as in-furnace adsorbent to capture gaseous semi-volatile metals during combustion, incineration, or gasification processes for the purposes of toxic metals emission control, ash deposition/slagging/corrosion inhibition, ultrafine particulate matter emission control, and so on. In this work, the adsorptions of typical heavy metals (Pb and Cd) and typical alkali metals (Na and K) by meta-kaolinite were investigated by the DFT calculation. The adsorption energies followed the sequence of NaOH-Si surface > KOH-Si surface > PbO-Al surface ≈ CdO-Al surface ≈ NaOH-Al surface > KOH-Al surface > NaCl-Al surface ≈ Na-Si surface > Na-Al surface > KCl-Al surface > Pb-Al surface > PbCl2-Al surface > CdCl2-Al surface ≈ K-Si surface ≈ PbCl-Al surface > K-Al surface > CdCl-Al surface > NaCl-Si surface > KCl-Si surface > Cd-Al surface. Si surface was found available to the adsorptions of Na, K, and their compounds, although it was invalid to the adsorptions of Pb, Cd, and their compounds. The interactions between adsorbates and surfaces were revealed. Furthermore, the discussion of combining with the experimental data was applied to the subject validity of calculation results and the effect of chlorine on adsorption and the effect of reducing atmosphere on adsorption.
APA, Harvard, Vancouver, ISO, and other styles
30

Krätzel, Ekkehard, and Sabine Hoeppner. "The Number of Lattice Points inside and on the Surface ∣t1∣k+∣t2∣k+ … +∣t1∣k = x." Mathematische Nachrichten 163, no. 1 (1993): 257–68. http://dx.doi.org/10.1002/mana.19931630121.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Rut'kov E. V. and Gall N. R. "Growth and stability of surface chemical compounds in interactionbetween Be and (1010) Re surface." Physics of the Solid State 64, no. 1 (2022): 130. http://dx.doi.org/10.21883/pss.2022.01.52499.190.

Full text
Abstract:
It is shown that the adsorption of Be on Re (1010) in the temperature rangeof 850-950 K leads to the formation of specific adsorption states --- surfacechemical compounds (SC) of ReBe stoichiometry with a concentration ofadsorbed NBe atoms of ~1.4·1015 cm-2. A multilayer film of beryllium (3-4 layers) is destroyed upon heating, and at 900 K all Be atoms leave the surface into the bulk of rhenium, except those that are part of the SC; atoms from the SC, in turn, actively dissolve at T>1050-1150 K. This corresponds to a decrease in the activation energy of dissolution upon the formation of SC from about 3.3 to 2.7 eV. Thermal desorption of beryllium takes place only at T>2100 K due to the emergence of Be atoms dissolved in the bulk of the metal onto the surface. Keywords: beryllium, rhenium, surface chemical compounds, dissolution, thermal desorption\
APA, Harvard, Vancouver, ISO, and other styles
32

Payares, Gilberto, Diane J. McLaren, W. H. Evans, and S. R. Smithers. "Changes in the surface antigen profile ofSchistosoma mansoniduring maturation from cercaria to adult worm." Parasitology 91, no. 1 (August 1985): 83–99. http://dx.doi.org/10.1017/s0031182000056535.

Full text
Abstract:
Antigenic proteins on the surfaces of different developmental stages ofSchistosoma mansoniwere radio-iodinated by the lodogen-catalysed method and identified by immunoprecipitation with a panel of antisera. The sera comprised specific immune serum from mice harbouring a chronic schistosome infection or vaccinated with γ-irradiated cercariae; serum from rabbits immunized with adult schistosome tegumental outer membranes or a partially purified Mr32K glycoprotein from adult worm membranes; and a monoclonal antibody recognizing an Mr20 K antigen on the surface of schistosomula. The Mr38–32 K glycoproteins were the major antigens identified in surface-labelled cercariae and their probable association with the glycocalyx is discussed. Schistosomula transformed from cercariae either mechanically or by penetration of host skinin vitro, expressed a similar pattern of surface antigens to that identified for cercariae, but low molecular weight antigens of Mr20, 17 and 15 K were also detected. The Mr38–32 K glycoproteins, although present on newly transformed schistoso mula, were progressively replaced with time, by a single dominant glycoprotein (Mr32 K) expressing identical epitopes to those on the Mr 38–32 K complex. Moreover, the data confirm that the Mr32 K glycoprotein persists on the tegument afterin vivomaturation and is conserved, together with Mr20 and 15 K antigens, through to the adult stage. New antigens (Mr97 and 25 K) were also detected duringin vivomaturation and were present in late-stage schistosomes recovered from infected hosts. In addition, the enzyme alkaline phosphatase is expressed on the surfaces of 3-week-old liver worms as a dominant antigen (Mr65 K); this feature may be related to nutritional and/or physiological processes in the tegument of this metabolically active stage of development.
APA, Harvard, Vancouver, ISO, and other styles
33

Roop, B., P. M. Blass, X. ‐L Zhou, and J. M. White. "Interactions of CO and surface K: Negligible CO adsorption on K/Ag(111)." Journal of Chemical Physics 90, no. 1 (January 1989): 608–9. http://dx.doi.org/10.1063/1.456465.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Hashizume, Tomihiro, Mitsuhiro Katayama, Dong-Ryul Jeon, Masakazu Aono, and Toshio Sakurai. "The Absolute Coverage of K on the Si(111)-3×1-K Surface." Japanese Journal of Applied Physics 32, Part 2, No. 9A (September 1, 1993): L1263—L1265. http://dx.doi.org/10.1143/jjap.32.l1263.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Johri, G. K., R. Malaroda, G. Iernetti, P. Ciuti, L. Carpenedo, and L. Sandri. "Water Surface Tension in the Temperature Range from 288 K to 304 K." Physics and Chemistry of Liquids 17, no. 2 (September 1987): 153–59. http://dx.doi.org/10.1080/00319108708078549.

Full text
APA, Harvard, Vancouver, ISO, and other styles
36

QUASSOWSKI, S., and K. HERMANN. "THEORY OF ALKALI-INDUCED RECONSTRUCTION OF THE Cu(100) SURFACE." Surface Review and Letters 04, no. 06 (December 1997): 1209–14. http://dx.doi.org/10.1142/s0218625x97001553.

Full text
Abstract:
LEED experiments show that Li adsorbed at Cu(100) surfaces at room temperature induces a (2×1) missing row substrate reconstruction while adsorption at lower temperatures, T=180 K , results in an unreconstructed Cu (100)+c(2× 2)-Li overlayer structure. Substrate reconstruction has not been observed for Na or for K adsorption. In order to study the specific reconstruction behavior of the Li adsorbate ab initio DFT calculations have been performed on Cu(100) + Ad, Ad = Li, Na, K, systems at coverages Θ Ad =0.25–0.5 with and without reconstruction. The calculations show that the (2×1)MR-reconstructed surface lies energetically above the ideal (1×1) surface by 0.2 eV per unit cell. However, alkali binding is stronger in the MR geometry as compared to that of the ideal surface where the increase in bond strength becomes smaller in going from Li to Na to K . As a result, the MR-reconstructed and the overlayer adsorbate systems are energetically very close for Cu (100) + Li while for Na and K the overlayer geometry is always favored.
APA, Harvard, Vancouver, ISO, and other styles
37

Smith, Keith T. "Bennu ejects material from its surface." Science 366, no. 6470 (December 5, 2019): 1209.11–1211. http://dx.doi.org/10.1126/science.366.6470.1209-k.

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

Umurhan, Orkan M., William M. Grundy, Michael K. Bird, Ross Beyer, James T. Keane, Ivan R. Linscott, Samuel Birch, et al. "A Near-surface Temperature Model of Arrokoth." Planetary Science Journal 3, no. 5 (May 1, 2022): 110. http://dx.doi.org/10.3847/psj/ac5d3d.

Full text
Abstract:
Abstract A near-surface thermal model for Arrokoth is developed based on the recently released 105 facet model of the body. This thermal solution takes into account Arrokoth’s surface reradiation back onto itself. The solution method exploits Arrokoth’s periodic orbital character to develop a thermal response using a time-asymptotic solution method, which involves a Fourier transform solution of the heat equation, an approach recently used by others. We display detailed thermal solutions assuming that Arrokoth’s near-surface material’s thermal inertia  = 2.5 W/m−2 K−1 s1/2. We predict that at New Horizons’ encounter with Arrokoth, its encounter hemisphere surface temperatures were ∼57–59 K in its polar regions, 30–40 K in its equatorial zones, and 11–13 K for its winter hemisphere. Arrokoth’s orbitally averaged temperatures are around 30–35 K in its polar regions and closer to 40 K near its equatorial zones. Thermal reradiation from the surrounding surface amounts to less than 5% of the total energy budget, while the total energy ensconced into and exhumed out of Arrokoth’s interior via thermal conduction over one orbit is about 0.5% of the total energy budget. As a generalized application of this thermal modeling together with other Kuiper Belt object origins considerations, we favor the interpretation that New Horizons’ REX instrument’s 29 ± 5 K brightness temperature measurement is consistent with Arrokoth’s near-surface material being made of sub-to-few-millimeter-size tholin-coated amorphous H2O ice grains with 1 W/m−2 K−1 s1/2 <  < 10–20 W/m−2 K−1 s1/2 and which are characterized by an X-band emissivity in the range 0.9 and 1.
APA, Harvard, Vancouver, ISO, and other styles
39

Moresco, F., M. Rocca, T. Hildebrandt, V. Zielasek, and M. Henzler. "Influence of surface interband transitions on surface plasmon dispersion: K/Ag(110)." Europhysics Letters (EPL) 43, no. 4 (August 15, 1998): 433–38. http://dx.doi.org/10.1209/epl/i1998-00377-0.

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

Rutigliano, Maria, and Fernando Pirani. "Scattering of N2 Molecules from Silica Surfaces: Effect of Polymorph and Surface Temperature." Molecules 27, no. 21 (November 2, 2022): 7445. http://dx.doi.org/10.3390/molecules27217445.

Full text
Abstract:
The inelastic scattering of N2 molecules from silica surfaces, taken at 100 K, has been investigated by adopting a semiclassical collision model in conjunction with the appropriate treatment of the long-range interaction forces. Such forces promote the formation of the precursor state that controls all basic elementary processes occurring at the gas–surface interphase. The probabilities for the different elementary surface processes triggered by quartz are determined and compared with those recently obtained for another silica polymorph (cristobalite). In addition, the final roto-vibrational distributions of N2 molecules undergoing inelastic scattering have been characterized. N2 molecules, impinging on both considered surfaces in low-medium vibrational states, preserve the initial vibrational state, while those inelastically scattered are rotationally excited and translationally colder. The surface temperature effect, investigated by raising the temperature itself from 100 K up to 1000 K, emerges more sharply for the cristobalite polymorph, mainly for the molecules impinging in the ground roto-vibrational state and with low collision energies.
APA, Harvard, Vancouver, ISO, and other styles
41

Ţǎlu, Ştefan, Sebastian Stach, Shahoo Valedbagi, S. Mohammad Elahi, and Reza Bavadi. "Surface morphology of titanium nitride thin films synthesized by DC reactive magnetron sputtering." Materials Science-Poland 33, no. 1 (March 1, 2015): 137–43. http://dx.doi.org/10.1515/msp-2015-0010.

Full text
Abstract:
AbstractIn this paper the influence of temperature on the 3-D surface morphology of titanium nitride (TiN) thin films synthesized by DC reactive magnetron sputtering has been analyzed. The 3-D morphology variation of TiN thin films grown on p-type Si (100) wafers was investigated at four different deposition temperatures (473 K, 573 K, 673 K, 773 K) in order to evaluate the relation among the 3-D micro-textured surfaces. The 3-D surface morphology of TiN thin films was characterized by means of atomic force microscopy (AFM) and fractal analysis applied to the AFM data. The 3-D surface morphology revealed the fractal geometry of TiN thin films at nanometer scale. The global scale properties of 3-D surface geometry were quantitatively estimated using the fractal dimensions D, determined by the morphological envelopes method. The fractal dimension D increased with the substrate temperature variation from 2.36 (at 473 K) to 2.66 (at 673 K) and then decreased to 2.33 (at 773 K). The fractal analysis in correlation with the averaged power spectral density (surface) yielded better quantitative results of morphological changes in the TiN thin films caused by substrate temperature variations, which were more precise, detailed, coherent and reproducible. It can be inferred that fractal analysis can be easily applied for the investigation of morphology evolution of different film/substrate interface phases obtained using different thin-film technologies.
APA, Harvard, Vancouver, ISO, and other styles
42

Chaabouni, H., S. Diana, T. Nguyen, and F. Dulieu. "Thermal desorption of formamide and methylamine from graphite and amorphous water ice surfaces." Astronomy & Astrophysics 612 (April 2018): A47. http://dx.doi.org/10.1051/0004-6361/201731006.

Full text
Abstract:
Context. Formamide (NH2CHO) and methylamine (CH3NH2) are known to be the most abundant amine-containing molecules in many astrophysical environments. The presence of these molecules in the gas phase may result from thermal desorption of interstellar ices. Aims. The aim of this work is to determine the values of the desorption energies of formamide and methylamine from analogues of interstellar dust grain surfaces and to understand their interaction with water ice. Methods. Temperature programmed desorption (TPD) experiments of formamide and methylamine ices were performed in the sub-monolayer and monolayer regimes on graphite (HOPG) and non-porous amorphous solid water (np-ASW) ice surfaces at temperatures 40–240 K. The desorption energy distributions of these two molecules were calculated from TPD measurements using a set of independent Polanyi–Wigner equations. Results. The maximum of the desorption of formamide from both graphite and ASW ice surfaces occurs at 176 K after the desorption of H2O molecules, whereas the desorption profile of methylamine depends strongly on the substrate. Solid methylamine starts to desorb below 100 K from the graphite surface. Its desorption from the water ice surface occurs after 120 K and stops during the water ice sublimation around 150 K. It continues to desorb from the graphite surface at temperatures higher than160 K. Conclusions. More than 95% of solid NH2CHO diffuses through the np-ASW ice surface towards the graphitic substrate and is released into the gas phase with a desorption energy distribution Edes = 7460–9380 K, which is measured with the best-fit pre-exponential factor A = 1018 s−1. However, the desorption energy distribution of methylamine from the np-ASW ice surface (Edes = 3850–8420 K) is measured with the best-fit pre-exponential factor A = 1012 s−1. A fraction of solid methylamine monolayer of roughly 0.15 diffuses through the water ice surface towards the HOPG substrate. This small amount of methylamine desorbs later with higher binding energies (5050–8420 K) that exceed that of the crystalline water ice (Edes = 4930 K), which is calculated with the same pre-exponential factor A = 1012 s−1. The best wetting ability of methylamine compared to H2O molecules makes CH3NH2 molecules a refractory species for low coverage. Other binding energies of astrophysical relevant molecules are gathered and compared, but we could not link the chemical functional groups (amino, methyl, hydroxyl, and carbonyl) with the binding energy properties. Implications of these high binding energies are discussed.
APA, Harvard, Vancouver, ISO, and other styles
43

Nastasi, M., J. P. Hirvonen, T. R. Jervis, G. M. Pharr, and W. C. Oliver. "Surface mechanical properties of C implanted Ni." Journal of Materials Research 3, no. 2 (April 1988): 226–32. http://dx.doi.org/10.1557/jmr.1988.0226.

Full text
Abstract:
Nickel foils, 165 μm thick, have been carbon implanted at 300 K with 2, 3, and 4.2 × 1017 C ions/cm2 and implanted with a two-step implantation of 2.1 × 1017 C/cm2 at 300 K followed by 2.1 × 1017 C/cm2 at 77 K. All implantations performed at 300 K result in the formation of the metastablc phase Ni3C while the two-step implantation produces an amorphous Ni/C alloy. Surface mechanical property studies showed that both the surface hardness and wear properties are correlated with chemistry (carbon dose), and that the friction coefficient is additionally dependent on the surface microstructure. It was found that both the wear rate and coefficient of friction were reduced as the volume fracion of Ni3C increased. At the highest dose implanted 4.2 × 1017 C/cm7, the coefficient of friction was found to be lower for the sample implanted half at 300 K and half at 77 K and possessing an amorphous structure compared to the sample implanted entirely at 300 K and possessing a crystalline Ni3C structure. Increases in the surface hardness were also observed with increasing carbon content, with the greatest hardness observed in samples implanted to a total dose of 4.2 × 1017 C/cm2. The hardness at this dose was not dependent on the implant conditions or the metastable phase formed.
APA, Harvard, Vancouver, ISO, and other styles
44

null, null. "Surface Embedding of Non-Bipartite $k$-Extendable Graphs." Annals of Applied Mathematics 38, no. 1 (June 2022): 1–24. http://dx.doi.org/10.4208/aam.oa-2021-0008.

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

Palmer, A. J. "Delta-k-lidar sensing of the ocean surface." Applied Optics 31, no. 21 (July 20, 1992): 4275. http://dx.doi.org/10.1364/ao.31.004275.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Krauss, R. H., R. G. Hellier, and J. C. McDaniel. "Surface temperature imaging below 300 K using La_2O_2S:Eu." Applied Optics 33, no. 18 (June 20, 1994): 3901. http://dx.doi.org/10.1364/ao.33.003901.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Riddell, Fraser. "Pamela K. Gilbert. Victorian Skin: Surface, Self, History." Review of English Studies 71, no. 300 (December 2, 2019): 591–94. http://dx.doi.org/10.1093/res/hgz139.

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

Reinarz, Jonathan. "Pamela K. Gilbert, Victorian Skin: Surface, Self, History." Social History of Medicine 33, no. 2 (January 18, 2020): 675–76. http://dx.doi.org/10.1093/shm/hkz136.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Rojas, K. I., A. R. Villagracia, and N. B. Arboleda. "H2adsorption on K decorated germanene surface: anab-initioinvestigation." Materials Research Express 3, no. 11 (November 10, 2016): 115015. http://dx.doi.org/10.1088/2053-1591/3/11/115015.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Ossowski, T., and A. Kiejna. "Low-coverage K adsorption on Mg(0001) surface." Surface Science 566-568 (September 2004): 983–88. http://dx.doi.org/10.1016/j.susc.2004.06.040.

Full text
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