Journal articles on the topic 'Interfacial Transport'

To see the other types of publications on this topic, follow the link: Interfacial Transport.

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 'Interfacial Transport.'

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

Blake, J. R. "Interfacial Transport Phenomena." Chemical Engineering Science 48, no. 6 (1993): 1182. http://dx.doi.org/10.1016/0009-2509(93)81051-v.

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

Kumar, S., and J. Y. Murthy. "Interfacial thermal transport between nanotubes." Journal of Applied Physics 106, no. 8 (October 15, 2009): 084302. http://dx.doi.org/10.1063/1.3245388.

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

Edwards, David A., Howard Brenner, Darsh T. Wasan, and Andrew M. Kraynik. "Interfacial Transport Processes and Rheology." Physics Today 46, no. 4 (April 1993): 63. http://dx.doi.org/10.1063/1.2808875.

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

Barnes, H. A. "Interfacial transport processes and rheology." Journal of Non-Newtonian Fluid Mechanics 46, no. 1 (January 1993): 123–24. http://dx.doi.org/10.1016/0377-0257(93)80009-z.

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

Anderson, J. L. "Colloid Transport by Interfacial Forces." Annual Review of Fluid Mechanics 21, no. 1 (January 1989): 61–99. http://dx.doi.org/10.1146/annurev.fl.21.010189.000425.

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

Van De Ven, T. G. M. "Interfacial transport processes and rheology." International Journal of Multiphase Flow 19, no. 2 (April 1993): 409–10. http://dx.doi.org/10.1016/0301-9322(93)90014-l.

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

Klingenberg, Daniel J. "Interfacial transport processes and rheology." Chemical Engineering Science 50, no. 6 (March 1995): 1069–70. http://dx.doi.org/10.1016/0009-2509(95)90141-8.

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

Yen, T. F., and George V. Chilingarian. "Interfacial transport processes and rheology." Journal of Petroleum Science and Engineering 10, no. 4 (April 1994): 351. http://dx.doi.org/10.1016/0920-4105(94)90025-6.

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

García-Mouton, Cristina, Mercedes Echaide, Luis A. Serrano, Guillermo Orellana, Fabrizio Salomone, Francesca Ricci, Barbara Pioselli, Davide Amidani, Antonio Cruz, and Jesús Pérez-Gil. "Beyond the Interface: Improved Pulmonary Surfactant-Assisted Drug Delivery through Surface-Associated Structures." Pharmaceutics 15, no. 1 (January 11, 2023): 256. http://dx.doi.org/10.3390/pharmaceutics15010256.

Full text
Abstract:
Pulmonary surfactant (PS) has been proposed as an efficient drug delivery vehicle for inhaled therapies. Its ability to adsorb and spread interfacially and transport different drugs associated with it has been studied mainly by different surface balance designs, typically interconnecting various compartments by interfacial paper bridges, mimicking in vitro the respiratory air–liquid interface. It has been demonstrated that only a monomolecular surface layer of PS/drug is able to cross this bridge. However, surfactant films are typically organized as multi-layered structures associated with the interface. The aim of this work was to explore the contribution of surface-associated structures to the spreading of PS and the transport of drugs. We have designed a novel vehiculization balance in which donor and recipient compartments are connected by a whole three-dimensional layer of liquid and not only by an interfacial bridge. By combining different surfactant formulations and liposomes with a fluorescent lipid dye and a model hydrophobic drug, budesonide (BUD), we observed that the use of the bridge significantly reduced the transfer of lipids and drug through the air–liquid interface in comparison to what can be spread through a fully open interfacial liquid layer. We conclude that three-dimensional structures connected to the surfactant interfacial film can provide an important additional contribution to interfacial delivery, as they are able to transport significant amounts of lipids and drugs during surfactant spreading.
APA, Harvard, Vancouver, ISO, and other styles
10

Machunsky, Stefanie, and Urs Alexander Peuker. "Liquid-Liquid Interfacial Transport of Nanoparticles." Physical Separation in Science and Engineering 2007 (January 8, 2007): 1–7. http://dx.doi.org/10.1155/2007/34832.

Full text
Abstract:
The study presents the transfer of nanoparticles from the aqueous phase to the second nonmiscible nonaqueous liquid phase. The transfer is based on the sedimentation of the dispersed particles through a liquid-liquid interface. First, the colloidal aqueous dispersion is destabilised to flocculate the particles. The agglomeration is reversible and the flocs are large enough to sediment in a centrifugal field. The aqueous dispersion is laminated above the receiving organic liquid phase. When the particles start to penetrate into the liquid-liquid interface, the particle surface is covered with the stabilising surfactant. The sorption of the surfactant onto the surface of the primary particles leads to the disintegration of the flocs. This phase transfer process allows for a very low surfactant concentration within the receiving organic liquid, which is important for further application, that is, synthesis for polymer-nanocomposite materials. Furthermore, the phase transfer of the nanoparticles shows a high efficiency up to 100% yield. The particle size within the organosol corresponds to the primary particle size of the nanoparticles.
APA, Harvard, Vancouver, ISO, and other styles
11

Wang, Shaofei, and Liquan Chen. "Interfacial transport in lithium-ion conductors." Chinese Physics B 25, no. 1 (January 2016): 018202. http://dx.doi.org/10.1088/1674-1056/25/1/018202.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Whitaker, Stephen. "Interfacial Transport Phenomena (John C. Slattery)." SIAM Review 35, no. 1 (March 1993): 172–74. http://dx.doi.org/10.1137/1035040.

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

Chen, Bo, and Lifa Zhang. "Optimized couplers for interfacial thermal transport." Journal of Physics: Condensed Matter 27, no. 12 (March 5, 2015): 125401. http://dx.doi.org/10.1088/0953-8984/27/12/125401.

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

Mattes, Brenton L. "Possible interfacial superconducting transport in CuCl:Si." Physica C: Superconductivity and its Applications 162-164 (December 1989): 554–55. http://dx.doi.org/10.1016/0921-4534(89)91152-0.

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

Ling, Xiao, Mischa Bonn, Katrin F. Domke, and Sapun H. Parekh. "Correlated interfacial water transport and proton conductivity in perfluorosulfonic acid membranes." Proceedings of the National Academy of Sciences 116, no. 18 (April 15, 2019): 8715–20. http://dx.doi.org/10.1073/pnas.1817470116.

Full text
Abstract:
Water must be effectively transported and is also essential for maximizing proton conductivity within fuel-cell proton-exchange membranes (PEMs). Therefore, identifying relationships between PEM properties, water transport, and proton conductivity is essential for designing optimal PEMs. Here, we use coherent Raman spectroscopy to quantify real-time, in situ diffusivities of water subspecies, bulk-like and nonbulk-like (interfacial) water, in five different perfluorosulfonic acid (PFSA) PEMs. Although the PEMs were chemically diverse, water transport within them followed the same rule: Total water diffusivity could be represented by a linear combination of the bulk-like and interfacial water diffusivities. Moreover, the diffusivity of interfacial water was consistently larger than that of bulk-like water. These measurements of microscopic transport were combined with through-plane proton conductivity measurements to reveal the correlation between interfacial water transport and proton conductivity. Our results demonstrate the importance of maximizing the diffusivity and fractional contribution of interfacial water to maximize the proton conductivity in PFSA PEMs.
APA, Harvard, Vancouver, ISO, and other styles
16

Albaalbaki, Bashar, and Reghan J. Hill. "On molecular diffusion in nanostructured porous media: interfacial exchange kinetics and surface diffusion." Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences 468, no. 2146 (June 13, 2012): 3100–3120. http://dx.doi.org/10.1098/rspa.2012.0172.

Full text
Abstract:
Water-vapour transport in nanostructured composite materials is poorly understood because diffusion and interfacial exchange kinetics are coupled. We formulate an interfacial balance that couples diffusion in dispersed and continuous phases to adsorption, absorption and interfacial surface diffusion. This work is motivated by water-vapour transport in cellulose fibre-based barriers, but the model applies to nanostructured porous media such as catalysts, chromatography columns, nanocomposites, cementitious structures and biomaterials. The interfacial balance can be applied in an analytical or a computational framework to porous media with any microstructural geometry. Here, we explore its capabilities in a model porous medium: randomly dispersed solid spheres in a continuous (humid) gas. We elucidate the roles of equilibrium moisture uptake, solid, gas and surface diffusion coefficients, inclusion size and interfacial exchange kinetics on the effective diffusivity. We then apply the local model to predict water-vapour transport rates under conditions in which the effective diffusivity varies through the cross section of a dense, homogeneous membrane that is subjected to a finite moisture-concentration gradient. As the microstructural length scale decreases from micrometres to nanometres, interfacial exchange kinetics and surface diffusion produce a maximum in the tracer flux. This optimal flux is flanked, respectively, by interfacial-kinetic- and diffusion-limited transport at smaller and larger microscales.
APA, Harvard, Vancouver, ISO, and other styles
17

JIN, WEI QING, XIN AN LIANG, LI XIA CAI, ZHI LEI PAN, and KATSUO TSUKAMOTO. "INTERFACIAL MASS TRANSPORT IN OXIDE CRYSTAL GROWTH." International Journal of Modern Physics B 16, no. 01n02 (January 20, 2002): 122–27. http://dx.doi.org/10.1142/s0217979202009548.

Full text
Abstract:
A space high temperature in situ observation instrument (SHITISOI) is dedicated to visualize and record the whole growth process of oxide crystal in high temperature melts and solutions. Model experiments using transparent liquids such as KNbO 3, Li 2 B 4 O 7+ KNbO 3 were chosen to investigate effects of interfacial mass transport in oxide crystal growth. For the scaling of the coupled velocity, heat and concentration fields in KNbO 3 crystal growth, a rotating growth process was performed and the widths of interfacial concentration, heat and momentum transition zones (The "boundary layers") are obtained, which are 7.5 × 10-3, and 8.6 × 10-2 and 4.4 × 10-1 cm , respectively. Hence one can expect that interfacial concentration gradient will be confined to a narrow layer and in region of major concentration change at the interface. In order to study a mechanism based on the interfacial mass transport resulting from hydrodynamics, the growth of KNbO 3 grain in high temperature Li 2 B 4 O 7 and KNbO 3 solution was studied. The result shows that the pivotal feature in the KNbO 3 crystal growth is the initiated by KNbO 3 solute surface tension gradient which is caused by the slow diffusion of KNbO 3solutes. Direct comparison of the model predictions and experimental observed phenomena demonstratre the predicitive capability of this model.
APA, Harvard, Vancouver, ISO, and other styles
18

Subedi, Kashi N., Kishor Nepal, Chinonso Ugwumadu, Keerti Kappagantula, and D. A. Drabold. "Electronic transport in copper–graphene composites." Applied Physics Letters 122, no. 3 (January 16, 2023): 031903. http://dx.doi.org/10.1063/5.0137086.

Full text
Abstract:
We investigate electronic transport properties of copper–graphene (Cu–G) composites using a density-functional theory (DFT) framework. Conduction in composites is studied by varying the interfacial distance of copper/graphene/copper (Cu/G/Cu) interface models. Electronic conductivity of the models computed using the Kubo–Greenwood formula shows that the conductivity increases with decreasing Cu–G distance and saturates below a threshold Cu–G distance. The DFT-based Bader charge analysis indicates increasing charge transfer between Cu atoms at the interfacial layers and the graphene with decreasing Cu–G distance. The electronic density of states reveals increasing contributions from both copper and carbon atoms near the Fermi level with decreasing Cu–G interfacial distance. By computing the space-projected conductivity of the Cu/G/Cu models, we show that the graphene forms a bridge to the electronic conduction at small Cu–G distances, thereby enhancing the conductivity.
APA, Harvard, Vancouver, ISO, and other styles
19

Sullivan, Greg D., and E. John List. "On mixing and transport at a sheared density interface." Journal of Fluid Mechanics 273 (August 25, 1994): 213–39. http://dx.doi.org/10.1017/s0022112094001916.

Full text
Abstract:
Mixing and transport of a stratifying scalar are investigated at a density interface imbedded in a turbulent shear flow. Steady-state interfacial shear flows are generated in a laboratory water channel for layer Richardson numbers, Ri, between about 1 and 10. The flow field is made optically homogeneous, enabling the use of laser-induced fluorescence with photodiode array imaging to measure the concentration field at high resolution. False-colour images of the concentration field provide valuable insight into interfacial dynamics: when the local mean shear Richardson number, Ris, is less than about 0.40–0.45, interfacial mixing appears to be dominated by Kelvin–Helmholtz (K–H) instabilities; when Ris is somewhat larger than this, interfacial mixing appears to be dominated by shear-driven wave breaking. In both cases, vertical transport of mixed fluid from the interfacial region into adjacent turbulent layers is accomplished by large-scale turbulent eddies which impinge on the interface and scour fluid from its outer edges.Motivated by the experimental findings, a model for interfacial mixing and entrainment is developed. A local equilibrium is assumed in which the rate of loss of interfacial fluid by eddy scouring is balanced by the rate of production (local mixing) by interfacial instabilities and molecular diffusion. When a single layer is turbulent and entraining, the model results are as follows: in the molecular-diffusion-dominated regime, δ/h ∼ Pe−1/2 and E ∼ Ri−1Pe−1/2; in the wave-breaking-dominated regime, δ/h ∼ Ri−1/2 and E ∼ Ri−3/2; and in the K–H-dominated regime, δ/h ∼ Ri−1 and E ∼ Ri−2, where δ is the interface thickness, h is the boundary-layer thickness, Pe is the Péclet number, and E is the normalized entrainment velocity. In all three regimes, the maximum concentration anomaly, [gcy ]m ∼ Ri−1. When both layers are turbulent and entraining, E and δ depend on combinations of parameters from both layers.
APA, Harvard, Vancouver, ISO, and other styles
20

Ishii, M., S. S. Paranjape, S. Kim, and X. Sun. "Interfacial structures and interfacial area transport in downward two-phase bubbly flow." International Journal of Multiphase Flow 30, no. 7-8 (July 2004): 779–801. http://dx.doi.org/10.1016/j.ijmultiphaseflow.2004.04.009.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Qiu, Lin, Xiaohua Zhang, Zhixin Guo, and Qingwen Li. "Interfacial heat transport in nano-carbon assemblies." Carbon 178 (June 2021): 391–412. http://dx.doi.org/10.1016/j.carbon.2021.02.105.

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

Matoušek, Václav, and Jan Krupička. "Interfacial friction and transport in stratified flows." Proceedings of the Institution of Civil Engineers - Maritime Engineering 167, no. 3 (September 2014): 125–34. http://dx.doi.org/10.1680/maen.14.00005.

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

KATAOKA, Isao. "Development of Research on Interfacial Area Transport." Journal of Nuclear Science and Technology 47, no. 1 (January 2010): 1–19. http://dx.doi.org/10.1080/18811248.2010.9711923.

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

Moghadam, M. M., J. M. Rickman, M. P. Harmer, and H. M. Chan. "Orientational anisotropy and interfacial transport in polycrystals." Surface Science 646 (April 2016): 204–9. http://dx.doi.org/10.1016/j.susc.2015.06.011.

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

WU, H., P. S. LUO, X. L. DONG, B. XU, L. X. CAO, X. G. QIU, and B. R. ZHAO. "INTERFACIAL TRANSPORT IN YBa2Cu3O7-δ/La0.67Ca0.33MnO3 HETEROSTRUCTURES." International Journal of Modern Physics B 19, no. 01n03 (January 30, 2005): 499–502. http://dx.doi.org/10.1142/s0217979205028906.

Full text
Abstract:
YBa2Cu3O7-δ/La0.67Ca0.33MnO3 heterostructures were fabricated and patterned to study the interfacial transport between a ferromagnet (FM) and a dx2-y2-wave superconductor. Zero bias conductance peaks (ZBCPs) in the differential conductance spectra were observed in the interface where a YBCO layer shows a mixed c- and a-axis orientation. In contrast, a zero bias conductance dip (ZBCD) was observed in the interface with a purely c-axis oriented YBCO layer.
APA, Harvard, Vancouver, ISO, and other styles
26

Xing, Siyuan, Jia Jiang, and Tingrui Pan. "Interfacial microfluidic transport on micropatterned superhydrophobic textile." Lab on a Chip 13, no. 10 (2013): 1937. http://dx.doi.org/10.1039/c3lc41255e.

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

Smith, Bradford J., Sarah Lukens, Eiichiro Yamaguchi, and Donald P. Gaver III. "Lagrangian transport properties of pulmonary interfacial flows." Journal of Fluid Mechanics 705 (November 9, 2011): 234–57. http://dx.doi.org/10.1017/jfm.2011.391.

Full text
Abstract:
AbstractDisease states characterized by airway fluid occlusion and pulmonary surfactant insufficiency, such as respiratory distress syndrome, have a high mortality rate. Understanding the mechanics of airway reopening, particularly involving surfactant transport, may provide an avenue to increase patient survival via optimized mechanical ventilation waveforms. We model the occluded airway as a liquid-filled rigid tube with the fluid phase displaced by a finger of air that propagates with both mean and sinusoidal velocity components. Finite-time Lyapunov exponent (FTLE) fields are employed to analyse the convective transport characteristics, taking note of Lagrangian coherent structures (LCSs) and their effects on transport. The Lagrangian perspective of these techniques reveals flow characteristics that are not readily apparent by observing Eulerian measures. These analysis techniques are applied to surfactant-free velocity fields determined computationally, with the boundary element method, and measured experimentally with micro particle image velocimetry ($\ensuremath{\mu} $-PIV). We find that the LCS divides the fluid into two regimes, one advected upstream (into the thin residual film) and the other downstream ahead of the advancing bubble. At higher oscillatory frequencies particles originating immediately inside the LCS experience long residence times at the air–liquid interface, which may be conducive to surfactant transport. At high frequencies a well-mixed attractor region is identified; this volume of fluid cyclically travels along the interface and into the bulk fluid. The Lagrangian analysis is applied to velocity data measured with 0.01 mg ml−1 of the clinical pulmonary surfactant Infasurf in the bulk fluid, demonstrating flow field modifications with respect to the surfactant-free system that were not visible in the Eulerian frame.
APA, Harvard, Vancouver, ISO, and other styles
28

Hanratty, Thomas J. "Separated flow modelling and interfacial transport phenomena." Applied Scientific Research 48, no. 3-4 (October 1991): 353–90. http://dx.doi.org/10.1007/bf02008206.

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

Kang, Ju Hwan, Jin Hee Lee, Bright Walker, Jung Hwa Seo, and Gap Soo Chang. "Understanding interfacial energy structures in organic solar cells using photoelectron spectroscopy: A review." Journal of Applied Physics 132, no. 5 (August 7, 2022): 050701. http://dx.doi.org/10.1063/5.0091960.

Full text
Abstract:
Organic solar cells (OSCs) have received considerable attention as a promising clean energy-generating technology because of their low cost and great potential for large-scale commercial manufacturing. With significant advances in new charge-transport material design, interfacial engineering, and their operating conditions, power conversion efficiencies of OSCs have continued to increase. However, a fundamental understanding of charge carrier transport and especially how ionic moieties affect carrier transport is still lacking in OSCs. In this regard, photoelectron spectroscopy has provided valuable information about interfacial electronic structures. The interfacial electronic structure of OSC interlayers greatly impacts charge extraction and recombination, controls energy level alignment, guides active layer morphology, improves material’s compatibility, and plays a critical role in the resulting power conversion efficiency of OSCs. Interfacial engineering incorporating inorganic, organic, and hybrid materials can effectively enhance the performance of organic photovoltaic devices by reducing energy barriers for charge transport and injection while improving compatibility between metal oxides and donor–acceptor based active layers or transparent conducting electrodes. This article provides a review of recent developments in interfacial engineering underlying organic photovoltaic devices of donor–acceptor interfaces.
APA, Harvard, Vancouver, ISO, and other styles
30

Liang, Zhu, Wei Bu, Karl J. Schweighofer, David J. Walwark, Jeffrey S. Harvey, Glenn R. Hanlon, Daniel Amoanu, Cem Erol, Ilan Benjamin, and Mark L. Schlossman. "Nanoscale view of assisted ion transport across the liquid–liquid interface." Proceedings of the National Academy of Sciences 116, no. 37 (March 12, 2018): 18227–32. http://dx.doi.org/10.1073/pnas.1701389115.

Full text
Abstract:
During solvent extraction, amphiphilic extractants assist the transport of metal ions across the liquid–liquid interface between an aqueous ionic solution and an organic solvent. Investigations of the role of the interface in ion transport challenge our ability to probe fast molecular processes at liquid–liquid interfaces on nanometer-length scales. Recent development of a thermal switch for solvent extraction has addressed this challenge, which has led to the characterization by X-ray surface scattering of interfacial intermediate states in the extraction process. Here, we review and extend these earlier results. We find that trivalent rare earth ions, Y(III) and Er(III), combine with bis(hexadecyl) phosphoric acid (DHDP) extractants to form inverted bilayer structures at the interface; these appear to be condensed phases of small ion–extractant complexes. The stability of this unconventional interfacial structure is verified by molecular dynamics simulations. The ion–extractant complexes at the interface are an intermediate state in the extraction process, characterizing the moment at which ions have been transported across the aqueous–organic interface, but have not yet been dispersed in the organic phase. In contrast, divalent Sr(II) forms an ion–extractant complex with DHDP that leaves it exposed to the water phase; this result implies that a second process that transports Sr(II) across the interface has yet to be observed. Calculations demonstrate that the budding of reverse micelles formed from interfacial Sr(II) ion–extractant complexes could transport Sr(II) across the interface. Our results suggest a connection between the observed interfacial structures and the extraction mechanism, which ultimately affects the extraction selectivity and kinetics.
APA, Harvard, Vancouver, ISO, and other styles
31

Garcia-Castello, Nuria, Sergio Illera, Joan Daniel Prades, Stefano Ossicini, Albert Cirera, and Roberto Guerra. "Energetics and carrier transport in doped Si/SiO2quantum dots." Nanoscale 7, no. 29 (2015): 12564–71. http://dx.doi.org/10.1039/c5nr02616d.

Full text
Abstract:
For B or P substitutional doping in Si/SiO2quantum dots we indicate, respectively, interfacial and sub-interfacial sites as the most energetically-favored ones. B-doping enhances hole-current at a low voltage, while P-doping enhances electron-current at low and high voltage.
APA, Harvard, Vancouver, ISO, and other styles
32

Hazuku, Tatsuya, Yutaka Fukuhara, Tomoji Takamasa, and Takashi Hibiki. "ICONE19-44014 APPLICABILITY OF INTERFACIAL AREA TRANSPORT EQUATION TO BUBBLY TWO-PHASE FLOWS UNDER MICROGRAVIT." Proceedings of the International Conference on Nuclear Engineering (ICONE) 2011.19 (2011): _ICONE1944. http://dx.doi.org/10.1299/jsmeicone.2011.19._icone1944_8.

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

Abarzhi, Snezhana I., and Walter Gekelman. "Preface: Non-equilibrium transport, interfaces, and mixing in plasmas." Physics of Plasmas 29, no. 3 (March 2022): 032103. http://dx.doi.org/10.1063/5.0088600.

Full text
Abstract:
Non-equilibrium transport, interfaces, and interfacial mixing play an important role in plasmas in high and low energy density regimes, at astrophysical and at atomic scales, and in nature and technology. Examples include the instabilities and interfacial mixing in supernovae and in inertial confinement fusion, the particle-field interactions in magnetic fusion and in imploding Z-pinches, the downdrafts in stellar interiors and in the planetary magneto-convection, magnetic flux ropes and structures in the solar corona, and plasma thrusters and nano-fabrication. This Special Topic exposes the state-of-the-art research on non-equilibrium transport, interfaces, and interfacial mixing in plasmas, including theory, experiment, and simulations. The works were presented at the invited mini-conference “Non-equilibrium Transport, Interfaces and Mixing in Plasmas” at the 2019 Annual Meeting of the Division of Plasma Physics of the American Physical Society.
APA, Harvard, Vancouver, ISO, and other styles
34

Song, X. M., Z. G. Huang, M. Gao, D. Y. Chen, Z. Fan, and Z. Q. Ma. "Role of Interfacial Oxide Layer in MoOx/n-Si Heterojunction Solar Cells." International Journal of Photoenergy 2021 (April 22, 2021): 1–8. http://dx.doi.org/10.1155/2021/6623150.

Full text
Abstract:
Interfacial oxide layer plays a crucial role in a MoOx/ n -Si heterojunction (MSHJ) solar cell; however, the nature of this interfacial layer is not yet clarified. In this study, based on the experimental results, we theoretically analyzed the role of the interfacial oxide layer in the charge carrier transport of the MSHJ device. The interfacial oxide layer is regarded as two layers: a quasi p -type semiconductor interfacial oxide layer (SiOx(Mo))1 in which numerous negatively charged centers existed due to oxygen vacancies and molybdenum–ion-correlated ternary hybrids and a buffer layer (SiOx(Mo))2 in which the quantity of Si-O bonds was dominated by relatively good passivation. The thickness of (SiOx(Mo))1 and the thickness of (SiOx(Mo))2 were about 2.0 nm and 1.5 nm, respectively. The simulation results revealed that the quasi p -type layer behaved as a semiconductor material with a wide band gap of 2.30 eV, facilitating the transport of holes for negatively charged centers. Additionally, the buffer layer with an optical band gap of 1.90 eV played a crucial role in passivation in the MoOx/ n -Si devices. Furthermore, the negative charge centers in the interfacial layer had dual functions in both the field passivation and the tunneling processes. Combined with the experimental results, our model clarifies the interfacial physics and the mechanism of carrier transport for an MSHJ solar cell and provides an effective way to the high efficiency of MSHJ solar cells.
APA, Harvard, Vancouver, ISO, and other styles
35

Wen, Jiangang, and Philip L. F. Liu. "Mass transport of interfacial waves in a two-layer fluid system." Journal of Fluid Mechanics 297 (August 25, 1995): 231–54. http://dx.doi.org/10.1017/s0022112095003077.

Full text
Abstract:
Effects of viscous damping on mass transport velocity in a two-layer fluid system are studied. A temporally decaying small-amplitude interfacial wave is assumed to propagate in the fluids. The establishment and the decay of mean motions are considered as an initial-boundary-value problem. This transient problem is solved by using a Laplace transform with a numerical inversion. It is found that thin ‘second boundary layers’ are formed adjacent to the interfacial Stokes boundary layers. The thickness of these second boundary layers is of O(ε1/2) in the non-dimensional form, where ε is the dimensionless Stokes boundary layer thickness defined as $\epsilon = \hat{k}\hat{\delta}=\hat{k}(2\hat{v}/\hat{\sigma})^{1/2}$ for an interfacial wave with wave amplitude â, wavenumber $\hat{k}$ and frequency $\hat{\sigma}$ in a fluid with viscosity $\hat{v}$. Inside the second boundary layers there exists a strong steady streaming of O(α2ε−1/2), where $\alpha = \hat{k}\hat{a}$ is the surface wave slope. The mass transport velocity near the interface is much larger than that in a single-layer system, which is O(α2) (e.g. Longuet-Higgins 1953; Craik 1982). In the core regions outside the thin second boundary layers, the mass transport velocity is enhanced by the diffusion of the mean interfacial velocity and vorticity. Because of vertical diffusion and viscous damping of the mean interfacial vorticity, the ‘interfacial second boundary layers’ diminish as time increases. The mean motions eventually die out owing to viscous attenuation. The mass transport velocity profiles are very different from those obtained by Dore (1970, 1973) which ignored viscous attenuation. When a temporally decaying small-amplitude surface progressive wave is propagating in the system, the mean motions are found to be much less significant, O(α2).
APA, Harvard, Vancouver, ISO, and other styles
36

Guo, Jinyu, Paige Brimley, Matthew Liu, Elizabeth R. Corson, Wilson Smith, and William Abraham Tarpeh. "Unravelling Mass Transport Effects in Electrochemical Nitrate Reduction on Titanium." ECS Meeting Abstracts MA2022-02, no. 58 (October 9, 2022): 2197. http://dx.doi.org/10.1149/ma2022-02582197mtgabs.

Full text
Abstract:
Nitrate is a prevalent waterborne pollutant that threatens the health of humans and aquatic systems. By selectively producing ammonia, electrochemical nitrate reduction reaction (NO3RR) can directly transform nitrate pollutants into widely used commodity chemicals and fertilizers, thus balancing the nitrogen cycle while reducing energy consumption from the traditional Haber-Bosch process. Although research efforts to date have primarily focused on the development of efficient nitrate reduction catalysts, the electrolyte environment has been found to substantially influence NO3RR activity and selectivity. Specifically, the composition of the first ten nanometers of electrolyte away from the electrode surface dictates the immediate environment within which reactions take place. As has been demonstrated in the electrochemical CO2 reduction reaction, this interfacial reaction microenvironment comprised of near-surface electrolyte and electrocatalyst influences reaction activity and selectivity. In electrochemical NO3RR, where an anionic species reacts at a negatively charged electrode, the reaction rate can often be limited by the mass transport of the reactant nitrate. Additionally, mass transport phenomena define the electrolyte side of the reaction microenvironment by influencing other interfacial properties (i.e., electric potential, pH and ion concentrations). Therefore, we investigated the influence of mass transport phenomena on NO3RR activity and selectivity. In this study, we used a membrane-separated flow cell as a representative and translatable reactor and titanium foil as a common ammonia-selective NO3RR electrode. By varying the electrolyte flow rate, we controlled the mass transport phenomena in the flow cell and empirically determined the diffusion layer thickness. We found that NO3RR activity generally increased with increasing electrolyte flow rate, as predicted from the decreasing diffusion layer thickness, whereas NH3 selectivity slightly decreased, showing an activity-selectivity trade-off. To unravel the origins of the experimentally observed mass transport effects, simulations using the 1D generalized modified Poisson-Nernst-Planck (GMPNP) model were conducted to spatially resolve interfacial properties. Interfacial cation concentration and pH were the two properties that changed most significantly with electrolyte flow rate. Informed by simulation results, we controlled the bulk electrolyte composition to modify the interfacial cation concentration and pH and examine their impacts on NO3RR activity and selectivity. Nitrate removal rate decreased by half when the bulk cation concentration was lowered to 1/100 of the original bulk concentration but the product distribution was similar. Meanwhile, in phosphate buffered electrolytes with different bulk pH values, NH3 selectivity decreased as bulk pH increased, with NO3RR activity remaining largely unchanged. Therefore, we attributed the influence of mass transport on NO3RR selectivity to changes in the interfacial pH during reaction, likely as the result of the varying NO3RR activity under different flow rates. By comparing the nitrogen species product distribution in different electrolytes, we inferred that the interfacial pH increased rapidly to basic in non-buffered systems regardless of their bulk pH. In summary, we systematically investigated the underexplored effects of mass transport on NO3RR and found that it influenced the reaction activity and selectivity by influencing interfacial properties. This study underscores the importance of scrutinizing the interfacial electrolyte environment in electrocatalyst development and provides insight to optimize NH3 production by engineering the bulk electrolyte.
APA, Harvard, Vancouver, ISO, and other styles
37

Gerner, F. M., and C. L. Tien. "Axisymmetric Interfacial Condensation Model." Journal of Heat Transfer 111, no. 2 (May 1, 1989): 503–10. http://dx.doi.org/10.1115/1.3250705.

Full text
Abstract:
This paper employs a simple axisymmetric model to study interfacial condensation. Numerical results show the effects of interfacial forces, subcooling of the liquid, superheating of the vapor, and the presence of a noncondensable gas in the vapor. While pressure and shear stress play an important role in determining the flow fields, and hence interfacial mass and energy transport, interfacial mass fluxes do not. Subcooling of the liquid is the dominant mechanism in determining the interfacial condensation rate. While superheating of the vapor is insignificant, except near the critical point, a noncondensable gas in the vapor greatly reduces condensation.
APA, Harvard, Vancouver, ISO, and other styles
38

Hibiki, Takashi, Tatsuya Hazuku, Tomoji Takamasa, and Mamoru Ishii. "Interfacial-Area Transport Equation at Reduced-Gravity Conditions." AIAA Journal 47, no. 5 (May 2009): 1123–31. http://dx.doi.org/10.2514/1.38208.

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

Zhao, Qiao, Paul Majsztrik, and Jay Benziger. "Diffusion and Interfacial Transport of Water in Nafion." Journal of Physical Chemistry B 115, no. 12 (March 31, 2011): 2717–27. http://dx.doi.org/10.1021/jp1112125.

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

Hibiki, Takashi, and Mamoru Ishii. "Interfacial Area Transport Equations for Gas-Liquid Flow." Journal of Computational Multiphase Flows 1, no. 1 (January 2009): 1–22. http://dx.doi.org/10.1260/175748209787387089.

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

Narayanan, Shankar, Andrei G. Fedorov, and Yogendra K. Joshi. "Interfacial Transport of Evaporating Water Confined in Nanopores." Langmuir 27, no. 17 (September 6, 2011): 10666–76. http://dx.doi.org/10.1021/la201807a.

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

Tanioka, Akihiko, Mie Minagawa, Patricio Ramírez, and Salvador Mafé. "Interfacial Transport of Amino Acid trhough Charged Membranes." membrane 24, no. 2 (1999): 92–99. http://dx.doi.org/10.5360/membrane.24.92.

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

Zhou, Hangbo, and Gang Zhang. "General theories and features of interfacial thermal transport." Chinese Physics B 27, no. 3 (March 2018): 034401. http://dx.doi.org/10.1088/1674-1056/27/3/034401.

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

Weinbaum, S. "Interfacial Transport in Large and Small Blood Vessels." Applied Mechanics Reviews 43, no. 5S (May 1, 1990): S109—S118. http://dx.doi.org/10.1115/1.3120789.

Full text
Abstract:
In this paper we shall review some recent mathematical models which have led to new conceptual views of the ultrastructural pathways by which water, solutes and large molecules cross the endothelial interface between tissue and blood. In particular, we shall show how a sequence of models for the endothelium and underlying tissue in large arteries have finally led to the experimental discovery of the large pore via which LDL and other large molecules enter the artery wall and how a new three-dimensional model for the interendothelial cleft in capillaries might reconcile the several long standing paradoxes relating to the measured filtration and solute permeability coefficients in the transcapillary exchange of water and hydrophilic solutes in the microcirculation.
APA, Harvard, Vancouver, ISO, and other styles
45

Jiang, Tao, Xueqiang Zhang, Suresh Vishwanath, Xin Mu, Vasily Kanzyuba, Denis A. Sokolov, Sylwia Ptasinska, David B. Go, Huili Grace Xing, and Tengfei Luo. "Covalent bonding modulated graphene–metal interfacial thermal transport." Nanoscale 8, no. 21 (2016): 10993–1001. http://dx.doi.org/10.1039/c6nr00979d.

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

Chalopin, Yann, Ali Rajabpour, Haoxue Han, Yuxiang Ni, and Sebastian Volz. "EQUILIBRIUM MOLECULAR DYNAMICS SIMULATIONS ON INTERFACIAL PHONON TRANSPORT." Annual Review of Heat Transfer 17, N/A (2014): 147–76. http://dx.doi.org/10.1615/annualrevheattransfer.2014007292.

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

FAGAS, GIORGOS, GIANAURELIO CUNIBERTI, and KLAUS RICHTER. "Molecular Wire-Nanotube Interfacial Effects on Electron Transport." Annals of the New York Academy of Sciences 960, no. 1 (January 24, 2006): 216–24. http://dx.doi.org/10.1111/j.1749-6632.2002.tb03036.x.

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

Hibiki, Takashi, Mamoru Ishii, and Zheng Xiao. "Axial interfacial area transport of vertical bubbly flows." International Journal of Heat and Mass Transfer 44, no. 10 (May 2001): 1869–88. http://dx.doi.org/10.1016/s0017-9310(00)00232-5.

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

Kim, Youngjun, and Byoungnam Park. "Interfacial charge transport in MAPbI3 perovskite on ZnO." Results in Physics 13 (June 2019): 102207. http://dx.doi.org/10.1016/j.rinp.2019.102207.

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

Appetecchi, G. B., F. Croce, L. Persi, F. Ronci, and B. Scrosati. "Transport and interfacial properties of composite polymer electrolytes." Electrochimica Acta 45, no. 8-9 (January 2000): 1481–90. http://dx.doi.org/10.1016/s0013-4686(99)00363-1.

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