Journal articles on the topic 'Surface energy'

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

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 'Surface energy.'

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

Shapochkina, I. V., T. Ye Korochkova, and V. M. Rozenbaum. "Symmetry properties of brownian motors with fluctuating periodic potential energy." Surface 9(24) (December 30, 2017): 57–68. http://dx.doi.org/10.15407/surface.2017.09.057.

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

Semchuk, O. Yu, and O. O. Havryliuk. "Absorption and relaxation of the laser pulse energy in substance (review)." Surface 9(24) (December 30, 2017): 118–35. http://dx.doi.org/10.15407/surface.2017.09.118.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Ip, S. W., and J. M. Toguri. "The equivalency of surface tension, surface energy and surface free energy." Journal of Materials Science 29, no. 3 (February 1994): 688–92. http://dx.doi.org/10.1007/bf00445980.

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

Pokytnyi, S. I., and A. D. Terets. "Exciton quasimolecules in nanosystems with semiconductor and dielectric colloidal quantum dots: a review." SURFACE 14(29) (December 30, 2022): 49–62. http://dx.doi.org/10.15407/surface.2022.14.049.

Full text
Abstract:
In review, deals with the theory of exciton quasimolecules (formed of spatially separated electrons and holes) in a nanosystems that consists of semiconductor and dielectric colloidal quantum dots (QDs) synthesized in a dielectric and semiconductor matrixs. It has been shown that the exciton quasimolecule formation is of the threshold character and possible in a nanosystem, where the distance D between the surfaces of QD is given by the condition (where and are some critical distance). We have shown that in such a nanoheterostructures acting as “exciton molecules” are the QDs with excitons localizing over their surfaces. The position of the quasimolecule state energy band depends both on the mean radius of the QDs, and the distance between their surfaces, which enables one to purposefully control it by varying these parameters of the nanostructure. It was found that the binding energy of singlet ground state of exciton quasimolecules, consisting of two semiconductor and dielectric QDs is a significant large values, larger than the binding energy of the biexciton in a semiconductor and dielectric single crystals almost two orders of magnitude. It is shown that the major contribution to tue binding energy of singlet ground state of exciton quasimolecule is made by the energy of the exchange interaction of electrons with holes and this contribution is much more substantial than the contribution of the energy of the Coulomb interaction between the electrons and holes. It is established that the position of the exciton quasimolecule energy band depends both on the mean radius of the QDs and the distance between their surfaces. It is shown that with increase in temperature above the threshold (), a transition can occur from the exciton quasimolecule to exciton state. It has been found that at a constant concentration of excitons (i.e. constant concentration of QD) and temperatures Т below , one can expect a new luminescence band shifted from the exciton band by the value of the exciton quasimolecule binding energy. This new band disappears at higher temperatures (). At a constant temperature below , an increase in exciton concentration (i.e. in QD concentration) brings about weakening of the exciton luminescence band and strengthening of the exciton quasimolecule. These exciton quasimolecules are of fundamental interest as new quasi-atomic colloidal nanostructures; they may also have practical value as new nanomaterials for nanooptoelectronics. The fact that the energy of the ground state singlet exciton quasimolecule is in the infrared range of the spectrum, presumably, allow the use of a quasimolecule to create new infrared sensors in biomedical research.
APA, Harvard, Vancouver, ISO, and other styles
5

Yurov, V. M., A. S. Baltabekov, V. Ch Laurinas, and S. A. Guchenko. "DIMENSIONAL EFFECTS AND SURFACE ENERGY OF FERROELECTRIC CRYSTALS." Eurasian Physical Technical Journal 16, no. 1 (June 14, 2019): 18–23. http://dx.doi.org/10.31489/2019no1/18-23.

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

Demianenko, E. M., M. I. Terets, S. V. Zhuravskyi, Yu I. Sementsov, V. V. Lobanov, V. S. Kuts, A. G. Grebenyuk, and M. T. Kartel. "Theoretical simulation of the interaction of Fe2 cluster with A N, B, Si-containing carbon graphene-like plane." SURFACE 14(29) (December 30, 2022): 37–48. http://dx.doi.org/10.15407/surface.2022.14.037.

Full text
Abstract:
Metal composites modified with various heteroatoms, such as N, B, Si, are used to obtain matrix composites with specified parameters with the strongest adhesive-cohesive bonds between metal atoms and a carbon nanoparticle. Such carbon nanoparticles functionalized with heteroatoms are promising for many metal composites. One of the interesting and promising metals as a matrix for such research work is iron. To predict the specifics of the interaction of iron with the surface of carbon nanomaterials supplemented with heteroatoms of different chemical structure, it is advisable to model such processes using quantum chemistry methods. The aim of the work was to find out the effect of temperature on the chemical interaction of iron clusters with native, boron-, silicon-, and nitrogen-containing graphene-like planes (GLP). The results of the calculations show that the highest value of the energy effect of the chemical interaction for the native graphene-like plane is +204.3 kJ/mol, in the case of calculations both by the B3LYP/6-31G(d,p) method and by the MP2/6-31G(d, p) (+370.7 kJ/mol). The lower value of the energy effect is found in the presence of nitrogen atoms in the composition of the graphene-like plane. This value is even lower for the interaction of iron dimers with a silicon-containing carbon nanocluster. The lowest values of the energy effect, calculated by both methods, are characteristic of the boron-containing graphene-like plane. In particular, for the B3LYP/6-31G(d,p) method, the value of the energy effect of the reaction is ‑210.5 kJ/mol, and for the MP2/6-31G(d,p) method this value is +16.6 kJ/mol. The presence of boron atoms in the composition of the nanocarbon matrix best contributes to the interaction with the iron nanocluster, regardless of the chosen research method. The dependence curves of the Gibbs free energy of the interaction of iron dimers with a graphene-like plane and its derivatives in all cases qualitatively correlate with similar energy effects. In addition, in all cases, the values of the Gibbs free energy increase with increasing temperature.
APA, Harvard, Vancouver, ISO, and other styles
7

Koguchi, Hideo. "Adhesion Analysis Considering Surface Energy and Surface Stresses." Key Engineering Materials 297-300 (November 2005): 1736–41. http://dx.doi.org/10.4028/www.scientific.net/kem.297-300.1736.

Full text
Abstract:
A new formulation for an adhesive force between a substrate and an indenter is presented. The boundary condition taking into account surface stresses is used for the present analysis. The surface stress is originated from surface energy. A paraboloidal indenter is pressed to the substrate, and then adhesion occurs between both surfaces. Surface energy and surface stress will vary at the adhesion surface, and then the surfaces deform in a concave way. An attractive force occurs to keep the contact of two adhesion surfaces. In the present paper, an effect of surface stress on the adhesive force will be clarified.
APA, Harvard, Vancouver, ISO, and other styles
8

Filonenko, О. V., E. M. Demianenko, and V. V. Lobanov. "Quantum chemical modeling of orthophosphoric acid adsorption sites on hydrated anatase surface." Surface 12(27) (December 30, 2020): 20–35. http://dx.doi.org/10.15407/surface.2020.12.020.

Full text
Abstract:
Quantum chemical modeling of orthophosphoric acid adsorption sites on the hydrated surface of anatase was performed by the method of density functional theory (exchange-correlation functional PBE0, basis set 6-31 G(d,p)). The influence of the aqueous medium was taken into account within the framework of the continual solvent model. The work uses a cluster approach. The anatase surface is simulated by a neutral Ti(OH)4(H2O)2 cluster. The results of analysis of the geometry and energy characteristics of all the calculated complexes show that the highest interaction energy is inherent to the intermolecular complex of orthophosphoric acid and hydrated surface of anatase, where the oxygen atom of the phosphoryl group (О=Р≡) forms a hydrogen bond with a hydrogen atom of the coordinated water molecule of Ti(OH)4(H2O)2 cluster and two hydrogen atoms of the hydroxyl groups of the orthophosphoric acid molecule form two hydrogen bonds with two oxygen atoms of the titanol groups. The formation energy effect of this complex is -134.0 kJ/mol. The formation energy effect of the complex with separated charges by the proton transfer from the molecule H3PO4 to the Ti(OH)4(H2O)2 cluster with the formation of dihydrogen phosphate anion and the protonated form of the titanol group (º) is -131.1 kJ/mol, so indicating less thermodynamic probability of such intermolecular interaction. The smallest thermodynamic probability (-123.9 kJ/mol) of complexation between orthophosphoric acid and hydrated anatase surface where a water molecule moves from the coordination sphere of the titanium atom. The calculation results indicate a possible adsorption of the H3PO4 molecule in an aqueous solution on the hydrated anatase surface. Taking into account the effect of the solvent within the polarization continuum insignificantly changes the adsorption energy, which is -44.5 kJ/mol; for vacuum conditions this value is -49.0 kJ/mol.
APA, Harvard, Vancouver, ISO, and other styles
9

Frankcombe, Terry J., and Michael A. Collins. "Potential energy surfaces for gas-surface reactions." Physical Chemistry Chemical Physics 13, no. 18 (2011): 8379. http://dx.doi.org/10.1039/c0cp01843k.

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

Chibowski, Emil. "Apparent Surface Free Energy of Superhydrophobic Surfaces." Journal of Adhesion Science and Technology 25, no. 12 (January 2011): 1323–36. http://dx.doi.org/10.1163/016942411x555890.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Krupska, T. V., V. M. Gun'ko, I. S. Protsak, I. I. Gerashchenko, A. P. Golovan, N. Yu Klymenko, V. V. Turov, and M. T. Kartel. "Properties of composite systems based on polymethylsiloxane and silica in the water environment." Surface 12(27) (December 30, 2020): 100–136. http://dx.doi.org/10.15407/surface.2020.12.100.

Full text
Abstract:
The formation of a composite system based on equal amounts of hydrophobic, porous polymethylsiloxane and hydrophilic nanosilicon A-300 was studied. It is shown that during the formation of a composite system the specific surface of the material is significantly reduced, which is due to the close contact between hydrophobic and hydrophilic particles. When water is added to the composite system, in the process of homogenization under conditions of dosed mechanical loading, the effect of nanocoagulation is manifested – the formation of nanosized particles of hydrated silica inside the polymethylsiloxane matrix, recorded on TEM microphotographs. When measuring the value of the interfacial energy of PMS and PMS/A-300 composite by low-temperature 1H NMR spectroscopy, it was found that the effect of nanocoagulation is manifested in a decrease (compared to the original PMS) energy of water interaction with the surface of the composite obtained under small mechanical conditions. its growth when using high mechanical loads. In the process, the binding of water in heterogeneous systems containing PMS, pyrogenic nanosilica (A-300), water and surfactants – decamethoxine (DMT) was studied. Composite systems were created using metered mechanical loads. It is shown that when filling the interparticle gaps of PMS by the method of hydrosealing, the interphase energy of water in the interparticle gaps of hydrophobic PMS with the same hydration is twice the interfacial energy of water in hydrophilic silica A-300. This is due to the smaller linear dimensions of the interparticle gaps in PMS compared to A-300. In the composite system, A-300/PMS/DMT/H2O there are non-additive growth of binding energy of water, which is probably due to the formation, under the influence of mechanical stress in the presence of water, microheterogeneous areas consisting mainly of hydrophobic and hydrophilic components (microcoagulation). Thus, with the help of mechanical loads, you can control the adsorption properties of composite systems and create new materials with unique adsorption properties.
APA, Harvard, Vancouver, ISO, and other styles
12

Pokytnii, S. I., and A. D. Terets. "Theory of spatially indirect excitons in nanosystems containing double semiconductors quantum dots." Surface 15(30) (December 30, 2023): 23–33. http://dx.doi.org/10.15407/surface.2023.15.023.

Full text
Abstract:
In mini-review, deals with the theory of exciton quasimolecules in a nanosystem consisting of double quantum dots of germanium synthesized in a silicon matrix. An exciton quasimolecule was formed as a result of the interaction of two spatially indirect excitons. It is shown that, depending on the distance D between the surfaces of the quantum dots, spatially indirect excitons and of exciton quasimolecules was formedin the nanosystem.The binding energy of the singlet ground state of the exciton quasimolecule has been gigantic exceeding the binding energy of the biexciton in a silicon single crystal by almost two orders of magnitude. The emergence of a band of localized electron states in the band gap of the silicon matrix was found. This band of localized electron states appeared as a result of the splitting of electron levels in the chain of germanium quantum dots. The nature of formation in the Ge/Si heterostructures was analyzed depending on the distance D between the surfaces of QDs SIEs and of exciton quasimolecules.It was shown that the binding energy of the ground singlet state of an exciton quasimolecule was gigantic, exceeding the binding energy of a biexciton in a silicon single crystal by almost two orders of magnitude.The possibility of using quasimolecules of excitons to create elements of silicon infrared nanooptoelectronics, including new infrared sensors, was established. The emergence of a band of localized electron states in the band gap of the silicon matrix was found.In this case, the band of localized electron states appeared as a result of the splitting of electron levels in the chain of germanium QDs.It was shown that the movement of an electron along the zone of localized electron states in the linear chain of germanium QDs caused an increase in photoconductivity.The effect of increasing photoconductivity can make a significant contribution in the process of converting the energy of the optical range in photosynthesizing nanosystems.
APA, Harvard, Vancouver, ISO, and other styles
13

Demianenko, E. M., A. G. Grebenyuk, V. V. Lobanov, V. O. Gabovich, V. O. Pokrovskiy, and M. I. Terets. "Structure and formation energy of multiple protonated molecular ions of acridine yellow: quantum-chemical calculations." Surface 8(23) (December 30, 2016): 50–57. http://dx.doi.org/10.15407/surface.2016.08.050.

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

Semchuk, O. Yu, O. O. Havryliuk, and A. A. Biliuk. "Energy absorption of laser radiation by metal nanoparticles in the conditions of surface plasmon resonance." Surface 11(26) (December 30, 2019): 496–507. http://dx.doi.org/10.15407/surface.2019.11.496.

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

Grebenyuk, A. G. "Absorption color and spatial structure of lead monoxide nanoparticles." Surface 15(30) (December 30, 2023): 12–22. http://dx.doi.org/10.15407/surface.2023.15.012.

Full text
Abstract:
The results of theoretical modeling of the spatial structure and electronic absorption spectra of isolated small particles of polymorphic modifications of lead monoxide - litharge and massicot, which are characterized by the presence of hypercoordinated atoms, have been analyzed. The optical properties of many oligomers of lead oxide are systematically considered in the work. The results of quantum chemical calculations by the methods of electron density functional theory (DFT and TDDFT) on the spatial structure and energy characteristics, as well as the electronic absorption spectra of the considered models, are discussed. Theoretical results are compared with available experimental data. The calculation method used and the proposed cluster models with or without attached (chemisorbed) water molecules allow us to reproduce the spatial structure and energy characteristics of polymorphic modifications of lead oxide at a semi-quantitative level. It has been found that the addition of water molecules to models for α-lead oxide nanoparticles contributes to their stabilization. The calculated values of the cohesion energy for litharge nanoparticles are greater than those for massicot ones, which corresponds to the experimental data for lead monoxide crystals. The calculated electronic spectra of the litharge nanoparticle models are characterized by absorption bands which lie in the region that corresponds to the red color, and those of massicot relate to yellow; the corresponding values of the width of the energy gap for the litharge models are smaller than for the massicot ones, which is consistent with the experimental data for crystals. The results of calculations show that the presence of hydroxyl groups in nanoparticle models leads to a hypsochromic shift of absorption maxima, so that with a sufficient number of such groups, these species may lose their color.
APA, Harvard, Vancouver, ISO, and other styles
16

Danielewicz, Paweł. "Surface symmetry energy." Nuclear Physics A 727, no. 3-4 (November 2003): 233–68. http://dx.doi.org/10.1016/j.nuclphysa.2003.08.001.

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

Perutz, S., J. Wang, E. J. Kramer, C. K. Ober, and K. Ellis. "Synthesis and Surface Energy Measurement of Semi-Fluorinated, Low-Energy Surfaces†." Macromolecules 31, no. 13 (June 1998): 4272–76. http://dx.doi.org/10.1021/ma9700993.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Karpenko, O. S., V. V. Lobanov, and M. T. Kartel. "C1s core-level binding energy shift dependence from carbon atoms position in graphenenanoflakes C96 and polycyclic aromatic hydrocarbon C96H24: a dft study." SURFACE 14(29) (December 30, 2022): 63–77. http://dx.doi.org/10.15407/surface.2022.14.063.

Full text
Abstract:
The hexagon-shape graphene nanoflakes (GNFs) limited by zigzag edges only (with doubly and triply coordinated atoms) have unique increased reactivity. Despite the high systems symmetry (D6h) the Carbon atoms in GNFs occupy non-equivalent positions. Can such physical and chemical characteristics of GNFs, which depend of the atom position in the cluster, definition? This characteristic together with the simplicity of its calculation makes it possible to predict the properties of nanoflakes obtained from GNFs by introducing single and multiatomic vacancies into them or by replacing Carbon atoms with electron withdrawing and electron donating atoms. This characteristic includes the C1s core-level binding energy shifts, the maxima of which characterize the C atoms of a certain type. The proposed work is devoted to quantum chemical calculations of the electronic density of states (DOS) of pristine hexagon-shape GNF C96 (multiplicity, M=5), their saturated counterpart –polycyclic aromatic hydrocarbon(PAH) C96H24 (M=1) and their derivatives with one and two single vacancies in the ground electronic state (GES). All calculations were performed using the density functional theory (DFT) method with the involvement of the valence-split basis set 6-31G (d,p). Systems with open shells were considered using the UB3LYP exchange-correlation functional. The obtained spectra were fitted using Gaussian curve fitting program to determine the binding energy for each peak. The Gaussian function distribution of the theoretically calculated C1s core-level binding energy shifts of GNFs testified the presence of six peaks, each of which refers to a certain type of Carbon atoms. The C1s peak with the highest binding energy (-285.57 eV) is caused by contributions from the doubly coordinated edge cyclic chain (ECC) Carbon atoms. The C1s orbitals of the central hexagon (CHex) atoms and the first cyclic chain (FCC) atoms form delocalized molecular orbitals (MOs) in different parts of the cluster. The analogous spectrum of PAH C96H24 is slightly shifted to the region of lower binding energies and contains only two well-defined peaks. The peak with a higher binding energy (-284.36 eV) is generated by the 1s states of the CHex atoms and the atoms of the FCC, which are bounded to the CHex atoms. The electronic DOS difference in C1s core-level spectra of GNF C96 (M=5) and their saturated counterpart PAH C96H24 is established due to the presence of two weakly bounded π-systems in GNF and common conjugated system in PAH. The electronic DOS of defect-containing cluster C96-1(1) (M=3) (one CHex atom has been removed from the C96nanoflake) is generated by the C1s core-level atoms of the second cyclic chain (SCC), which are located at the different distances from the center of the nanoflake. The peak of the lowest intensity (-284.63 eV) appears in the spectrum as a reflection of the appearance of doubly coordinated Carbon atoms surrounding the single vacancy in the C96-1(1) nanoflake. The analysis of the electronic DOS of the C1s core-level spectrum of the C96-2(1) nanoflakeis shown, that doubly coordinated Carbon atoms, concentrated around two single vacancies, are essentially non-equivalent. If the MO with the lowest binding energy is localized on two of them – the MO with the highest binding energy is localized on the third atoms (one around each single vacancy). The electronic C1s core-level DOS spectrum of defect-containing molecular systems with one C96-1(1)H24 and two C96‑2(1)H24 single vacancies are similar to the analogous spectrum of PAH C96H24. In the first of them – one additional maximum appears due to C1s atoms surrounding the single vacancy. In the second – there are two additional maxima, each of which is generated by C1s core-level atoms adjacent to individual vacancies.
APA, Harvard, Vancouver, ISO, and other styles
19

Packham, D. E. "Surface energy, surface topography and adhesion." International Journal of Adhesion and Adhesives 23, no. 6 (January 2003): 437–48. http://dx.doi.org/10.1016/s0143-7496(03)00068-x.

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

Smith, Jeremy R., Jérôme Leveneur, and John V. Kennedy. "Design of intelligent surfaces for energy intensive processing industry." MATEC Web of Conferences 185 (2018): 00001. http://dx.doi.org/10.1051/matecconf/201818500001.

Full text
Abstract:
There are three different factors that can affect adhesion: the process fluid, the processing conditions and the surface of the processing equipment. Of these three factors, the surface properties of the processing equipment are the factor that offers the greatest opportunity for manipulation. The two key surface properties that have been identified to reduce adhesion are the surface energy and the surface topography. The surface energy of a material determines its degree of wettability and, a surface's affinity for water. In previous studies the surface energy of materials have been leveraged in order to create a surface with reduced levels of fouling through surface modification or the addition of polymer coatings with varying degrees of hydrophobicity. In addition, the topography of surfaces has been modified to reduce the level of particle adhesion. These modifications involve creating either a structured or random porous microstructure on the surface. Additional methods identified to reduce fouling include the application of liquid infused porous surfaces at low shear conditions and the use of non-contact heating through techniques such as microwave processing.
APA, Harvard, Vancouver, ISO, and other styles
21

Kazakova, O. O. "Quantum-chemical investigation of interactions in supramolecular systems: cholesterol - bile acids - silica in aqueous solutions." Surface 13(28) (December 30, 2021): 39–46. http://dx.doi.org/10.15407/surface.2021.13.039.

Full text
Abstract:
Hypercholesterolemia significantly increases the risk of myocardial infarction associated with COVID-19. Along with pharmacological treatment, the possibility of the excretion of excess cholesterol from an organism by adsorption is also of great interest. The interaction of cholesterol with the surface of partially hydrophobized silica in aqueous solutions of bile acids was investigated by the PM7 method using the COSMO (COnductor-like Screening MOdel) solvation model. The distribution of electrostatic and hydrophobic potentials of molecules and complexes was calculated. The values of free Gibbs energy adsorption of bile acids on the surface of silica correlate with the distribution coefficients in the n-octanol-water system. The energy of interaction of cholesterol with bile acids affects its adsorption on silica. The stronger the bond of cholesterol with the molecules of bile acids, the less it is released from the primary micelles in solution and adsorbed on the surface.
APA, Harvard, Vancouver, ISO, and other styles
22

Grinko, А. M., А. V. Brichka, О. М. Bakalinska, and М. Т. Каrtel. "Application of nano cerium oxide in solid oxide fuel cells." Surface 12(27) (December 30, 2020): 231–50. http://dx.doi.org/10.15407/surface.2020.12.231.

Full text
Abstract:
This review is analyzed the state of modern literature on the nanoceria based materials application as components for solid oxide fuel cells. The principle of operation of fuel cells, their classification and the difference in the constructions of fuel cells are described. The unique redox properties of nanosized cerium oxide make this material promising for application as components for solid oxide fuel cells (SOFC). Because of high ionic conductivity, high coefficient of thermal expansion and low activation energy at relatively low temperatures, cerium-containing materials are widely used as a solid electrolyte. On the surface of nanosized CeO2 there many surface defects (which is determined by the concentration of oxygen vacancies) that lead to the electronic conductivity increases even at temperatures (300 - 700 °C). The concentration of surface defects can be increased by doping the surface of nanoceria by divalent and trivalent cations. The ionic and electrical properties of the obtained nanocomposites dependent from synthesis methods, ionic radii and concentration of doping cations. It is explained the effect of the transition in the size of cerium oxide particles in the nanoscale region on the concentration of surface defects and defects in the sample structure. Particular attention is paid to the effect of doping nanosized CeO2 by transition metal cations and lanthanides on the characteristics of the obtained material, namely, on the increase of concentration of surface defects due to the increase of oxygen vacancies. It is established that nanosized cerium oxide is used for the development and implementation of the main components of SOFC: electrolyte, anode and cathode. Advantages of using solid electrolytes based on nanosized cerium oxide over the classical electrolytes are listed. It was shown that doping of cerium oxide by double and triple cations lead to increase the ionic conductivity and reduces the activation energy and has a positive effect on its characteristics as a SOFC electrolyte. Composites, based on nanoscaled cerium oxide, are actively developed and studied for use as electrodes of solid oxide fuel cells. Cerium-containing anodes are resistant to the deposition of carbon and fuel impurities, increase the catalytic activity of solid oxide fuel cells, and compatible with other components. Nanosized cerium oxide particles are sprayed onto the cathode to prevent the cathode from interacting with the electrolyte. The prospects for the use of cerium-containing materials for the conversion of chemical energy of fuel into electrical energy are analyzed.
APA, Harvard, Vancouver, ISO, and other styles
23

Milionis, Athanasios, Despina Fragouli, Luigi Martiradonna, George C. Anyfantis, P. Davide Cozzoli, Ilker S. Bayer, and Athanassia Athanassiou. "Spatially Controlled Surface Energy Traps on Superhydrophobic Surfaces." ACS Applied Materials & Interfaces 6, no. 2 (January 3, 2014): 1036–43. http://dx.doi.org/10.1021/am404565a.

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

Turov, V. V., V. M. Gun’ko, T. V. Krupskaya, I. S. Protsak, L. S. Andriyko, A. I. Marinin, A. P. Golovan, N. V. Yelagina, and N. T. Kartel. "Interphase interactions of hydrophobic powders based on methilsilica in the water environment." Surface 12(27) (December 30, 2020): 53–99. http://dx.doi.org/10.15407/surface.2020.12.053.

Full text
Abstract:
Using modern physicochemical research methods and quantum chemical modeling, the surface structure, morphological and adsorption characteristics, phase transitions in heterogeneous systems based on methylsilica and its mixtures with hydrophilic silica were studied. It is established that at certain concentrations of interfacial water, hydrophobic silica or their composites with hydrophilic silica form thermodynamically unstable systems in which energy dissipation can be carried out under the influence of external factors: increasing water concentration, mechanical loads and adsorption of air by hydrophobic component. When comparing the binding energies of water in wet powders of wettind-drying samples A-300 and AM-1, which had close values of bulk density (1 g/cm3) and humidity (1 g/g), close to 8 J/g. However, the hydration process of hydrophobic silica is accompanied by a decrease in entropy and the transition of the adsorbent-water system to a thermodynamically nonequilibrium state, which is easily fixed on the dependences of interfacial energy (S) on the amount of water in the system (h). It turned out that for pure AM-1 the interfacial energy of water increases in proportion to its amount in the interparticle gaps only in the case when h < 1 g/g. With more water, the binding energy decreases abruptly, indicating the transition of the system to a more stable state, which is characterized by the consolidation of clusters of adsorbed water and even the formation of a bulk phase of water. Probably there is a partial "collapse" of the interparticle gaps of hydrophobic particles AM-1 and the release of thermodynamically excess water. For mixtures of hydrophobic and hydrophilic silica, the maximum binding of water is shifted towards greater hydration. At AM1/A-300 = 1/1 the maximum is observed at h = 3g/g, and in the case of AM1/A-300 = 1/2 it is not reached even at h = 4 g/g. The study of the rheological properties of composite systems has shown that under the action of mechanical loads, the viscosity of systems decreases by almost an order of magnitude. However, after withstanding the load and then reducing the load to zero, the viscosity of the system increases again and becomes significantly higher than at the beginning of the study. That is, the obtained materials have high thixotropic properties. Thus, a wet powder that has all the characteristics of a solid after a slight mechanical impact is easily converted into a concentrated suspension with obvious signs of liquid.
APA, Harvard, Vancouver, ISO, and other styles
25

Wang, Dan, Zhili Hu, Gang Peng, and Yajun Yin. "Surface Energy of Curved Surface Based on Lennard-Jones Potential." Nanomaterials 11, no. 3 (March 9, 2021): 686. http://dx.doi.org/10.3390/nano11030686.

Full text
Abstract:
Although various phenomena have confirmed that surface geometry has an impact on surface energy at micro/nano scales, determining the surface energy on micro/nano curved surfaces remains a challenge. In this paper, based on Lennard-Jones (L-J) pair potential, we study the geometrical effect on surface energy with the homogenization hypothesis. The surface energy is expressed as a function of local principle curvatures. The accuracy of curvature-based surface energy is confirmed by comparing surface energy on flat surface with experimental results. Furthermore, the surface energy for spherical geometry is investigated and verified by the numerical experiment with errors within 5%. The results show that (i) the surface energy will decrease on a convex surface and increase on a concave surface with the increasing of scales, and tend to the value on flat surface; (ii) the effect of curvatures will be obvious and exceed 5% when spherical radius becomes smaller than 5 nm; (iii) the surface energy varies with curvatures on sinusoidal surfaces, and the normalized surface energy relates with the ratio of wave height to wavelength. The curvature-based surface energy offers new insights into the geometrical and scales effect at micro/nano scales, which provides a theoretical direction for designing NEMS/MEMS.
APA, Harvard, Vancouver, ISO, and other styles
26

Moreno Baqueiro Sansao, Bernardo, Jon J. Kellar, William M. Cross, Karen Schottler, and Albert Romkes. "Comparison of surface energy and adhesion energy of surface-treated particles." Powder Technology 384 (May 2021): 267–75. http://dx.doi.org/10.1016/j.powtec.2021.02.029.

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

Terebinska, M. I., O. I. Tkachuk, A. M. Datsyuk, O. V. Filonenko, and V. V. Lobanov. "Transformation of addimers >Ge=Ge<, >Ge=Si< AND >Si=Si< on the relaxed side of Si (001) (4 × 2)." Surface 13(28) (December 30, 2021): 66–74. http://dx.doi.org/10.15407/surface.2021.13.066.

Full text
Abstract:
By the method of density functional theory (B3LYP, 6-31G **) three types of displacements are calculated, namely oscillations as a whole, rotation and diffusion of dimers > Ge = Ge <, > Ge = Si < and > Si = Si <, which are formed on the crystalline surface of Si (001) (4×2) during the deposition of germanium atoms under conditions of molecular beam epitaxy. Calculations of angles of buckling of addimers are carried out. It is shown that when the addimers as a whole oscillate around the equilibrium position, the energy barriers are quite low, the highest of them occurs for a mixed addimer > Si = Ge <. Pure adders > Ge = Ge < and > Si = Si < oscillate between two degenerate states with an energy barrier of 0.042 and 0.014 eV, respectively. The structures of the transition state and the intermediate when the addimer > Ge = Ge < is moved between adjacent cells in the approximation of the constant bond length > Ge = Ge < are obtained. As calculations have shown, all transformations of surface dimers occur with relatively small activation energies, the numerical values of which agree satisfactorily with the results of STM experiments available in the literature.
APA, Harvard, Vancouver, ISO, and other styles
28

Park, Seong-Bum, Hyun-Je Sung, Dong-Min Shim, and Nack-Joo Kim. "Optimization of biomethane production by biogas upgrading process using response surface mothodolgy." Journal of Energy Engineering 23, no. 2 (June 30, 2014): 62–73. http://dx.doi.org/10.5855/energy.2014.23.2.062.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Yu, X. H., J. Rong, T. L. Fu, Z. L. Zhan, Z. Liu, and J. X. Liu. "Surface Tension and Surface Energy of Nanomaterials." Journal of Computational and Theoretical Nanoscience 12, no. 12 (December 1, 2015): 5318–22. http://dx.doi.org/10.1166/jctn.2015.4522.

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

Ding, Z.-J. "Self-energy in surface electron spectroscopy: II. Surface excitation on real metal surfaces." Journal of Physics: Condensed Matter 10, no. 8 (March 2, 1998): 1753–65. http://dx.doi.org/10.1088/0953-8984/10/8/010.

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

Serghei A, Baranov. "Surface energy for nanowire." Annals of Mathematics and Physics 5, no. 2 (July 8, 2022): 081–85. http://dx.doi.org/10.17352/amp.000043.

Full text
Abstract:
The theory of surface phenomena in the production of micro-and nanocylinder for important cases is considered. Analytical solution to Gibbs–Tolman–Koenig–Buff equation for nanowire surface is given. Analytical solutions to equations for case the cylindrical surface for the linear and nonlinear Van der Waals theory are analyzed. But for a nonlinear theory, this correspondence is absent.
APA, Harvard, Vancouver, ISO, and other styles
32

Fork, R. L., R. L. Laycock, W. W. Walker, S. T. Cole, S. D. Moultrie, J. Phillips, and C. Reinhardt. "Surface High-Energy Laser." Proceedings of the IEEE 93, no. 10 (October 2005): 1864–73. http://dx.doi.org/10.1109/jproc.2005.853551.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Ouyang, G., X. L. Li, X. Tan, and G. W. Yang. "Surface energy of nanowires." Nanotechnology 19, no. 4 (January 4, 2008): 045709. http://dx.doi.org/10.1088/0957-4484/19/04/045709.

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

Brooks, J. H., and M. S. Rahman. "Surface Energy of Wool." Textile Research Journal 56, no. 3 (March 1986): 164–71. http://dx.doi.org/10.1177/004051758605600303.

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

Métois, J. J., and P. Müller. "Absolute surface energy determination." Surface Science 548, no. 1-3 (January 2004): 13–21. http://dx.doi.org/10.1016/j.susc.2003.11.027.

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

Hopken, Jens, and Martin Moller. "Low-surface-energy polystyrene." Macromolecules 25, no. 5 (September 1992): 1461–67. http://dx.doi.org/10.1021/ma00031a016.

Full text
APA, Harvard, Vancouver, ISO, and other styles
37

Yudin, M., and B. D. Hughes. "Surface energy of solids." Physical Review B 49, no. 8 (February 15, 1994): 5638–42. http://dx.doi.org/10.1103/physrevb.49.5638.

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

Muchnick, Paul, and Arnold Russek. "The HeH2 energy surface." Journal of Chemical Physics 100, no. 6 (March 15, 1994): 4336–46. http://dx.doi.org/10.1063/1.466316.

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

Demianenko, E. M., M. I. Terets, V. V. Lobanov, Yu I. Sementsov, V. S. Kuts, and M. T. Kartel. "Quantum chemical investigation of the influence of the presence of a graphen-cluster on the energy of covalent bonds on the polyamid fragment in nanocomposite." Surface 11(26) (December 30, 2019): 484–95. http://dx.doi.org/10.15407/surface.2019.11.484.

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

Chhatre, Shreerang S., Jesus O. Guardado, Brian M. Moore, Timothy S. Haddad, Joseph M. Mabry, Gareth H. McKinley, and Robert E. Cohen. "Fluoroalkylated Silicon-Containing Surfaces−Estimation of Solid-Surface Energy." ACS Applied Materials & Interfaces 2, no. 12 (November 10, 2010): 3544–54. http://dx.doi.org/10.1021/am100729j.

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

Bian, Jianjun, Gangfeng Wang, and Xiqiao Feng. "Atomistic calculations of surface energy of spherical copper surfaces." Acta Mechanica Solida Sinica 25, no. 6 (December 2012): 557–61. http://dx.doi.org/10.1016/s0894-9166(12)60050-0.

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

Mohammadi, P., L. P. Liu, P. Sharma, and R. V. Kukta. "Surface energy, elasticity and the homogenization of rough surfaces." Journal of the Mechanics and Physics of Solids 61, no. 2 (February 2013): 325–40. http://dx.doi.org/10.1016/j.jmps.2012.10.010.

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

Kim, Soojin, Young-Hee Lee, Kyu Rang Kim, and Young-San Park. "Analysis of surface energy balance closure over heterogeneous surfaces." Asia-Pacific Journal of Atmospheric Sciences 50, S1 (October 6, 2014): 553–65. http://dx.doi.org/10.1007/s13143-014-0045-2.

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

Zebda, A., H. Sabbah, S. Ababou-Girard, F. Solal, and C. Godet. "Surface energy and hybridization studies of amorphous carbon surfaces." Applied Surface Science 254, no. 16 (June 2008): 4980–91. http://dx.doi.org/10.1016/j.apsusc.2008.01.147.

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

Hong, Suklyun. "Surface energy anisotropy of iron surfaces by carbon adsorption." Current Applied Physics 3, no. 5 (October 2003): 457–60. http://dx.doi.org/10.1016/j.cap.2003.07.001.

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

Strelko, V. V., and Yu I. Gorlov. "Influence of electronic states of nanographs in carbon microcrystallines on surface chemistry of activated charcoal varieties." Surface 13(28) (December 30, 2021): 15–38. http://dx.doi.org/10.15407/surface.2021.13.015.

Full text
Abstract:
In this paper, the nature of the chemical activity of pyrolyzed nanostructured carbon materials (PNCM), in particular active carbon (AC), in reactions of electron transfer considered from a single position, reflecting the priority role of paramagnetic centers and edge defunctionaled carbon atoms of carbon microcristallites (CMC) due to pyrolysis of precursors. Clusters in the form of polycyclic aromatic hydrocarbons with open (OES) and closed (CES) electronic shells containing terminal hydrogen atoms (or their vacancies) and different terminal functional groups depending on specific model reactions of radical recombination, combination, replacement and elimination were used to model of nanographenes (NG) and CM. Quantum-chemical calculations of molecular models of NG and CMC and heat effects of model reactions were performed in frames of the density functional theory (DFT) using extended valence-splitted basis 6-31G(d) with full geometry optimization of concrete molecules, ions, radicals and NG models. The energies of boundary orbitals were calculated by means of the restricted Hartry-Fock method for objects with closed (RHF) and open (ROHF) electronic shells. The total energies of small negative ions (HOO-, HO-) and anion-radical О2•‾) were given as the sum of calculated total energies of these compounds and their experimental electron affinities. The estimation of probability of considered chemical transformations was carried out on the base on the well-known Bell-Evans-Polyani principle about the inverse correlation of the thermal effects of reactions and its activation energies. It is shown that the energy gap ΔЕ (energy difference of boundary orbitals levels) in simulated nanographens should depend on a number of factors: the periphery structure of models, its size and shape, the number and nature of various structural defects, electronic states of NG. When considering possible chemical transformations on the AC surface, rectangular models of NG were used, for which the simple classification by type and number of edge structural elements of the carbon lattice was proposed. Quantum chemical calculations of molecular models of NG and CNC and the energy of model reactions in frames of DTF showed that the chemisorption of free radicals (3O2 and N•O), as recombination at free radical centers (FRC), should occur with significant heat effects. Such calculations give reason to believe that FRC play an important role in formation of the functional cover on the periphery of NG in CMC of studied materials. On the base of of cluster models of active carbon with OES new ideas about possible reactions mechanisms of radical-anion О2•‾ formation and decomposition of hydrogen peroxide on the surface of active carbon are offered. Explanation of increased activity of AC reduced by hydrogen in H2O2 decomposition is given. It is shown that these PNCM models, as first of all AC, allow to adequately describe their semiconductor nature and acid-base properties of such materials.
APA, Harvard, Vancouver, ISO, and other styles
47

Pawlak, Zenon. "The Amphoteric Nature of Phospholipid Bilayer Verified by the Surface Energy." Clinical Research Notes 3, no. 3 (April 30, 2022): 01–05. http://dx.doi.org/10.31579/2690-8816/065.

Full text
Abstract:
Background The surface energy the phospholipidic spherical membrane vs. pH, see Fig. 4A helped us to understand amphoteric character of cartilage surface (introduced together first time in literature) [4]. Methods The samples of articular cartilage were taken from the knees of an aged one to two-years-old ox. Friction coefficient (cartilage/cartilage) pair and the interfacial energy of a spherical lipid bilayer curves vs pH have “a bell-shaped curve” expressing amphoteric cartilage nature by the isoelectric point (IEP). Results It was found that the lowest energy values cover working conditions at pH 7.3 (±) 1.0 of natural joints, see Fig.4 A, B. The obtained “a bell-shaped curve” for (cartilage/cartilage) pair friction and spherical PLs bilayer is a novelty in this paper. Conclusions The observations have led to the conclusion that friction coefficient (cartilage/cartilage) pair and surface energy of a spherical lipid bilayer curves vs pH have (IEP) and expressing amphoteric surfaces character. The surface energy corresponds well with friction coefficient phospholipid amphoteric bilayer surface.
APA, Harvard, Vancouver, ISO, and other styles
48

Davidson, Peter L., Suzanne J. Wilson, David J. Chalmers, Barry D. Wilson, David Eager, and Andrew S. McIntosh. "Analysis of Energy Flow During Playground Surface Impacts." Journal of Applied Biomechanics 29, no. 5 (October 2013): 628–33. http://dx.doi.org/10.1123/jab.29.5.628.

Full text
Abstract:
The amount of energy dissipated away from or returned to a child falling onto a surface will influence fracture risk but is not considered in current standards for playground impact-attenuating surfaces. A two-mass rheological computer simulation was used to model energy flow within the wrist and surface during hand impact with playground surfaces, and the potential of this approach to provide insights into such impacts and predict injury risk examined. Acceleration data collected on-site from typical playground surfaces and previously obtained data from children performing an exercise involving freefalling with a fully extended arm provided input. The model identified differences in energy flow properties between playground surfaces and two potentially harmful surface characteristics: more energy was absorbed by (work done on) the wrist during both impact and rebound on rubber surfaces than on bark, and rubber surfaces started to rebound (return energy to the wrist) while the upper limb was still moving downward. Energy flow analysis thus provides information on playground surface characteristics and the impact process, and has the potential to identify fracture risks, inform the development of safer impact-attenuating surfaces, and contribute to development of new energy-based arm fracture injury criteria and tests for use in conjunction with current methods.
APA, Harvard, Vancouver, ISO, and other styles
49

Kashin, V. V., K. M. Shakirov, and A. I. Poshevneva. "Surface tension and surface energy in capillary theory." Steel in Translation 42, no. 2 (February 2012): 99–102. http://dx.doi.org/10.3103/s0967091212020088.

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

Liu, Feipeng P., Timothy G. Rials, and John Simonsen. "Relationship of Wood Surface Energy to Surface Composition." Langmuir 14, no. 2 (January 1998): 536–41. http://dx.doi.org/10.1021/la970573y.

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